Abstract
Plants and plant pathogens are subject to continuous co-evolutionary pressure for dominance, and the outcomes of these interactions can substantially impact agriculture and food security1–3. In virus– plant interactions, one of the major mechanisms for plant antiviral immunity relies on RNA silencing, which is often suppressed by co-evolving virus suppressors, thus enhancing viral pathogenicity in susceptible hosts1. In addition, plants use the nucleotide-binding and leucine-rich repeat (NB-LRR) domain-containing resistance proteins, which recognize viral effectors to activate effector-triggered immunity in a defence mechanism similar to that employed in non-viral infections2,3. Unlike most eukaryotic organisms, plants are not known to activate mechanisms of host global translation suppression to fight viruses1,2. Here we demonstrate in Arabidopsis that the constitutive activation of NIK1, a leucine-rich repeat receptor-like kinase (LRR-RLK) identified as a virulence target of the begomovirus nuclear shuttle protein (NSP)4–6, leads to global translation suppression and translocation of the downstream component RPL10 to the nucleus, where it interacts with a newly identified MYB-like protein, L10-INTERACTING MYB DOMAIN-CONTAINING PROTEIN (LIMYB), to downregulate translational machinery genes fully. LIMYB overexpression represses ribosomal protein genes at the transcriptional level, resulting in protein synthesis inhibition, decreased viral messenger RNA association with polysome fractions and enhanced tolerance to begomovirus. By contrast, the loss of LIMYB function releases the repression of translation-related genes and increases susceptibility to virus infection. Therefore, LIMYB links immune receptor LRR-RLK activation to global translation suppression as an antiviral immunity strategy in plants.
NIK1 was first identified as a virulence target of the begomovirus NSP5,6. For begomoviruses, a group of single-stranded DNA viruses that infect major crops, the success of infection relies not only on viral suppressors of RNA silencing4 but also on the viral inhibitor, NSP, of the immune receptor, NIK1 (ref. 5). The NIK1 protein belongs to the same LRR-RLK subfamily as the well-characterized PAMP recognition co-receptor BRI1-ASSOCIATED RECEPTOR KINASE 1 (BAK1)7,8. NIK1 is involved in plant antiviral immunity5, whereas BAK1is required for plant immunity against bacteria, fungi and oomycetes through its interactions with multiple PAMP-recognition LRR-RLKs9. We have previously demonstrated that the activation of NIK1 kinase is induced by the phosphorylation of Thr 474 within the activation (A)-loop10,11 (Supplementary Discussion 1). Apart from the identification of RPL10 as a downstream effector in NIK1-mediated antiviral immunity12,13, mechanistic knowledge of the signalling pathway is lacking, and the molecular nature of the defence response remains unclear. In this study, we replaced the normal NIK1 receptor with the NIK1 phosphomimetic gain-of-function mutant T474D11 in transgenic Arabidopsis lines to understand the molecular basis of the NIK1-mediated defence mechanism (Extended Data Fig. 1a–c). Transgenic lines possessing the gain-of-function mutant T474D in the nik1 knockout background10 were challenged with infectious clones of the Arabidopsis-infecting begomovirus cabbage leaf curl virus (CaLCuV)10,11. Then, we compared the virus-induced and T474D-induced transcriptomes at 10 days post-inoculation (dpi). A global cluster analysis of the expressed sequences among the mock-treated and infected wild-type (Col-0), NIK1 and T474D lines (Supplementary Table 1) revealed that the transcriptomes of the infected wild-type and mock-inoculated T474D lines were most closely related; these samples clustered together with a high bootstrap probability and a high approximately unbiased P value (Fig. 1a), which suggests that the NIK1-mediated response and the response to begomovirus infection share similar mechanisms. These transcriptomes differed greatly from the NIK1 mock-inoculated transcriptome, indicating that virus infection activates the NIK1-mediated response. Moreover, the gain-of-function T474D mutant might be activated in a constitutive manner that allows it to support a sustained NIK1-mediated response, in contrast with the expression of the intact NIK1 receptor in the nik1 genetic background. The transcriptome from NIK1-complemented lines clustered with the Col-0 mock-inoculated transcriptome.
We also employed these transgenic lines to assess the T474D-induced global variation in gene expression. Gene enrichment analyses of immune system category genes indicated that ectopic expression of T474D did not activate typical viral defences, such as salicylic acid signalling or virus-induced gene silencing (Supplementary Table 2, Extended Data Fig. 2a, b and Supplementary Discussion 2). Among the differentially expressed genes, we observed an overrepresentation of translational-machinery-related genes, which largely predominated the downregulated gene list (Extended Data Fig. 3a, red spots; Supplementary Tables 2 and 3). Using enrichment analysis, these downregulated genes included ribosomal genes and other components of the protein synthesis machinery. Therefore, T474D ectopic expression downregulates components of the translational machinery, suggesting that the constitutive activation of NIK1 might influence translation. To confirm that protein synthesis was impaired by constitutive activation of NIK1 in the T474D lines, we labelled leaf proteins in vivo with [35S]Met in the control and nik1 plants, as well as in the NIK1- (ref. 11), G473V/T474A- (ref. 11) and T474D-expressing lines (Fig. 1b and Extended Data Fig. 3b). The one-proportion statistical test indicated a significant decrease (12.8% for T474D line 4 (T474D-4) and 13% for T474D-6; P < 0.05) in the amount of newly synthesized protein in T474D-expressing leaves compared with wild-type and NIK1-expressing leaves. We also demonstrated that dexamethasone-inducible T474D expression for 8 h led to a higher inhibition of de novo protein synthesis in the transgenic lines (Fig. 1c and Extended Data Fig. 3c–e). In the dexamethasone-inducible lines, the expression of T474D significantly reduced polysome (PS) and monosome (NPS) fractions (12% total reduction) to a similar extent as it reduced PS and NPS RNA (13% reduction; Extended Data Fig. 4a–c). The loading of host mRNA (RBCS, Arabidopsis thaliana (At)WWP1, S13 and S39 genes) in actively translating PS fractions was significantly reduced in T474D-overexpressing lines compared to the wild-type line, although to a different extent (Fig. 1d, Extended Data Fig. 4d and Supplementary Discussion 3). Therefore, the activation of NIK1 reduces global levels of translation, but the effect may not be the same for all mRNAs. This downregulation of cytosolic translation might at least partially underlie the molecular mechanisms involved in NIK1-mediated antiviral defences.
To examine whether the constitutive activation of NIK1 was effective at controlling begomovirus infection, the transgenic lines were inoculated with CaLCuV DNA-A and DNA-B. The wild-type plants displayed typical symptoms of CaLCuV infection, whereas the symptoms in the T474D-expressing lines were greatly attenuated (Fig. 1e and Extended Data Fig. 5b). The symptomless CaLCuV infections of the T474D-expressing lines were associated with a delayed course of infection (Extended Data Fig. 5a) and a lower accumulation of viral DNA in the systemically infected leaves (Fig. 1f).
Because T474D expression caused downregulation of protein synthesis, we determined whether viral RNA translation was impaired in the T474D-expressing lines. We examined viral RNA transcripts in actively translating PS fractions prepared from infected leaves at 10 dpi (Extended Data Fig. 4e–g), when the total viral mRNA accumulation in the Col-0 and T474D-expressing lines was not very dissimilar (Fig. 1d and Extended Data Fig. 5c). We observed a significant reduction in the PS loading of viral mRNA in the T474D infected leaves compared with the wild-type infected leaves (Fig. 1d, g), indicating that the begomovirus was not capable of sustaining high levels of viral mRNA translation in the T474D-expressing lines.
Our data indicate that the translational suppression induced by NIK1 constitutive activation is associated with the downregulation of translational-machinery-related genes. Hence, NIK1-mediated nucleocytoplasmic trafficking of RPL10 may be linked to the regulation of gene expression. The previously identified extraribosomal functions of RPL10 associated with transcription factor regulation14–16 may serve as potential targets for assessing this hypothesis. We used the two-hybrid system to search for RPL10 nuclear partners and isolated a MYB-domain-containing transcription factor, which was designated as LIMYB (Extended Data Fig. 6a–e). As a putative transcription factor, LIMYB localized in the nucleus of transiently or stably transformed plant cells (Extended Data Fig. 7a–i) and interacted with RPL10 in the nuclei of plant cells (Fig.2a and Extended Data Fig.6f, g). We further demonstrated that RPL10 and LIMYB interact in vivo using co-immunoprecipitation assays (Fig. 2b, c).
The function of LIMYB in NIK1-mediated antiviral signalling was examined using several different approaches. We first demonstrated that LIMYB, RPL10 and NIK1 are co-expressed in several organs (Extended Data Fig. 8). Then, we identified transfer DNA (T-DNA) insertion mutants (limyb-32 and limyb-82) in the LIMYB gene (Extended Data Fig.1e, f). We also prepared LIMYB-overexpressing lines (Extended Data Fig. 1g) and determined the LIMYB-induced global variation in gene expression compared to Col-0 leaves. Remarkably, the overexpression of LIMYB resulted in a downregulation of translational-machinery-related genes similar to that induced by T474D expression (Extended Data Fig. 1h, red spots). We selected five ribosomal protein (RP) genes to confirm the deep-sequencing results for the LIMYB-overexpressing lines by quantitative polymerase chain reaction with reverse transcription (qRT–PCR) (Fig. 3a). LIMYB overexpression downregulated the expression of the selected RP genes but not of the unrelated gene (AtWWP1; Extended Data Fig. 9c). Conversely, in the limyb-32 line, the RP genes were up-regulated (Fig. 3 and Extended Data Fig. 9a, d). Expression of LIMYB in limyb lines restored the wild-type expression of the RP genes (Extended Data Fig. 9b). As in the T474D-expressing lines, protein synthesis was slightly but significantly reduced in the LIMYB-overexpressing lines (Fig. 3c and Extended Data Fig. 4h). Because RPL10 functions in NIK1-mediated antiviral signalling and interacts with LIMYB, we examined whether RPL10 also controls the expression of RP genes. In RPL10-overexpressing lines (Extended Data Fig. 1i, j), the expression of RP genes but not of the unrelated AtWWP1 gene was downregulated (Fig. 3d and Extended Data Fig.9e). We then examined whether LIMYB, a putative transcription factor, binds to an RP promoter (RPL18) in vivo. We performed chromatin immunoprecipitation (ChIP) experiments in which a 150 base pair (bp) RPL18 promoter was amplified from the precipitated DNA of LIMYB-expressing tissues but not of wild-type tissues (Fig. 3e). ChIP-qPCR showed that the 150-bp promoter fragment was significantly enriched in samples precipitated by anti-green fluorescent protein (GFP) antibody but not in samples pulled down from wild-type lines (Fig. 3f). These results suggest that LIMYB may function as a DNA-binding protein that associates with RP promoters in vivo.
To provide further evidence for the regulation of RP target genes by LIMYB and RPL10, we performed a luciferase transactivation assay in agro-infiltrated leaves. Consistent with the gene expression profile, LIMYB and RPL10 specifically repressed the L28e, S13A and L18E promoters but not an unrelated ubiquitin promoter (Fig. 3g, h). The co-transfection of Arabidopsis protoplasts with both LIMYB and RPL10 promoted increased repression of the L18E promoter compared to regulation of the gene reporter by individual expression of the transfactors. Collectively, these results indicate that LIMYB and RPL10 function as transcriptional repressors of common RP genes and that both transfactors are required for full regulation (Fig. 3h). Therefore, LIMYB and RPL10 may coordinately regulate common target promoters.
We also demonstrated that T474D expression downregulated the RP genes (Fig. 3i) but not AtWWP1 (Extended Data Fig. 9f), confirming the RNA-sequencing data (Extended Data Fig. 3a). As an additional control, we showed that the double-mutant NIK1 G4743V/T474A12, which does not complement the enhanced susceptibility phenotype of the nik1-null alleles (Extended Data Fig. 5a), also does not affect RP genes (Extended Data Fig. 9g). Nevertheless, the loss of LIMYB function prevented the T474D-mediated downregulation of the RP genes in both the limyb-32 transgenic lines that stably expressed T474D (Fig. 3j and Extended Data Fig. 9h, i) and the limyb-32 protoplasts that transiently expressed T474D (Extended Data Fig. 10). These results genetically link LIMYB to the translation suppression portion of the NIK1-mediated antiviral response.
We predicted that if the suppression of host translation was the basis for the begomovirus-tolerant phenotype of the T474D-expressing lines, LIMYB overexpression would also be effective against CaLCuV and the loss of LIMYB function would further debilitate the plant upon begomovirus infection. Col-0, limyb-32, limyb-82 and LIMYB-overexpressing lines were inoculated with infectious clones of CaLCuV DNA-A and DNA-B (Fig. 3k). Both Col-0 and limyb lines developed typical CaLCuV symptoms, although to different extents. The disease symptoms varied in severity from extreme stunting and leaf distortion with severe chlorosis in the limyb-32 and limyb-82 leaves to mild leaf distortion and moderate chlorosis in Col-0. The course of infection in the limyb leaves was accelerated compared to that in the Col-0 plants (Extended Data Fig. 5d) and the limyb-32 and limyb-82 systemic leaves accumulated higher levels of viral DNA than Col-0 (Fig. 3l). Therefore, the loss of LIMYB function recapitulated the enhanced begomovirus susceptibility phenotype of the nik1-null alleles, as would be expected from a downstream component of the NIK1-mediated antiviral defence. By contrast, the LIMYB-overexpressing lines did not develop symptoms, displayed delayed infection and accumulated a lower level of viral DNA in their systemic leaves, resembling the tolerant phenotype of the T474D-overexpressing lines. We also found that LIMYB overexpression did not induce salicylic-acid-signalling marker genes or typical defence responses to viral infection (Extended Data Fig. 2c and Supplementary Table 2). Collectively, these results further support the notion that the inhibition of host translation observed in the T474D and LIMYB lines may be an effective mechanism exploited by plant cells to fight begomovirus infection. Therefore, the demonstration that immune-receptor-mediated defence signalling controls translation in plant cells may represent a new paradigm for antiviral defences in plants (Supplementary Discussion 4).
METHODS
Plasmid constructs
The clone pK7F-NIK1T474D was described previously11. This clone harbours a GFP gene fused in frame after the last codon of the respective mutant cDNA under the control of the CaMV 35S promoter. In the mutant cDNA T474D, the threonine residue at position 474 within the activation loop of NIK1 was replaced with an aspartate residue. The clones rpL10AST-pDONR201 (pUFV608), rpL10ANS-pDONR201 (pUFV609), rpL10AST-pDONR207 (pUFV900) and rpL10ANS-pDONR207 (pUFV901), which harbour the RPL10A coding region either with (ST) or without (NS) a translational stop codon inserted into the entry vectors pDONR201 or pDONR207, have also been previously described12. The RPL10A coding region was transferred from these entry clones to the yeast expression vectors pDEST32 and pDEST22, generating the clones pBD-RPL10A (pUFV1422) and pAD-RPL10 (pUFV785) as GAL4 binding domain (BD) or activation domain (AD) fusions. For BiFC, the RPL10A coding region was transferred to the vectors SPY-NE-GW and SPY-CE-GW. The resulting clones, pSPYNE-RPL10A (pUFV1652) and pSPYCE-RPL10A (pUFV1653), harboured the RPL10A coding region fused to either the N-terminal (NE) region of YFP or the YFP C terminus (CE), respectively, under the control of the 35S promoter. The clones pK7F-L10 and pYFP-L10, which harbour RPL10A fused in frame with GFP after its last codon or with YFP before its first codon under the control of the 35S promoter, respectively, have been previously described10.
The clone pYFP-NLS-L10, which harbours a YFP–L10 fusion with a nuclear localization signal (NLS) under the control of the 35S promoter, was constructed by first inserting the amplified coding region of RPL10A into the SacI site of the pGR vector. The resulting clone, pGR-L10, which contains the RPL10A coding region fused to a glucocorticoid receptor (GR) domain with an NLS at the 5′ end, was then used as the template for the amplification of the NLS-containing L10 fusion with specific primers. The resulting product was inserted by recombination into pDONR207 and transferred to 35S-YFP-casseteA-Nos-pCAMBIA1300, generating pYFP-NLS-L10.
For the induction of T474D expression, the mutant cDNA T474D was transferred by recombination from the pDON201-T474D entry vector to the destination vector pBAV150 (ref. 17). The resulting clone pBAV150-NIK1 T474D–GFP harbours the mutant T474D open reading frame (ORF) under the control of a dexamethasone-inducible promoter and fused to a C-terminal GFP tag.
Arabidopsis growth conditions and transformation
The Columbia (Col-0) ecotype of A. thaliana was used as the wild-type control, and the nik1 knockout line10 was used for plant transformation. The nik1 lines10 were transformed with pK7F-NIK1T474D or pBAV150- NIK1 T474D-GFP using the floral dip method. Two independently transformed lines expressing the T474D transgene (T474D-4 and T474D-6) were selected for the infection assays. The transgenic lines ectopically expressing NIK1–GFP (NIK1-5 and NIK1-8) or an inactive kinase, the double mutant of NIK1, and the G473V/T474A–GFP dead kinase (G473V/T474A-10 and G473V/T474A-8) have been previously described11,12 (Supplementary Table 1).
RT–PCR and real-time RT–PCR analyses
Total RNA was extracted from Arabidopsis leaves using TRIzol (Invitrogen). RT–PCR assays were performed using gene-specific primers, as previously described18. For gene expression analysis by qRT–PCR, gene-specific primers were designed using Primer Express 3.0 (Life Technologies). Real-time RT–PCR assays were performed on an ABI7500 instrument (Life Technologies) using the SYBR Green PCR Master Mix (Life Technologies). The amplification reactions were performed as follows: 2 min at 50 °C, 10 min at 95 °C and 40 cycles of 94 °C for 15 s and 60 °C for 1 min. The variation in gene expression was quantified using the comparative Ct method (2−ΔΔCt), and absolute gene expression was quantified using the 2−ΔCt method. The values were normalized to endogenous actin and ubiquitin as control genes11.
RNA sequencing and data analysis
The transgenic and wild-type lines were infected at the seven-leaf stage with CaLCuV, as described later. After 10 dpi or after 21 dpi, total RNA from systemically infected leaves, as diagnosed by PCR, and mock-inoculated leaves from wild-type, 35S∷NIK1-5 and 35S∷T474D-4 lines was isolated using TRIzol (Invitrogen). For the RNA-sequencing experiments, we used two biological replicates of a pool of ten plants at 10 dpi, when we detected high levels of viral DNA in systemic leaves but symptoms had not yet developed, or at 21 dpi, when symptoms were visible. The RNA-sequencing data were obtained using an Illumina Hi-seq 2000. The paired-end 100-bp protocol was used with the following quality filter parameters: 5 bases trimmed at the 3′ and 5′ ends of the reads, a minimum average quality of 30 phred score. The data were stored in a comma-separated value spreadsheet file, and differential gene expression (DGE) analysis was performed using the R/Bioconductor package edger19. The raw data were normalized using the TMM normalization factor20. The dispersion was estimated by the tagwise edger parameter. Differential expression was determined using the cut-off P value of 0.1 adjusted by the false discovery rate (FDR). The read mapping process was executed using the Bowtie program21 with the cDNA data set retrieved from The Arabidopsis Information Resource (TAIR) database (http://www.arabidopsis.org), tenth release. Gene ontology classification was performed using the R/Bioconductor packages GSEABase and GOstats. Clustering analysis was performed using the R package pvclust (Hierarchical Clustering with P-Values via Multiscale Bootstrap Resampling) using Ward’s method22, and heatmaps were generated using gplots. The results were stored in a relational database created in PostgreSQL, and a web interface was created using PHP to allow the database to be accessed and navigated (http://arabidopsisnik.inctipp.ufv.br).
In vivo labelling of leaf proteins
Arabidopsis seedlings (300 mg) were incubated with 1 ml of nutrient solution containing 30 μg ml−1 dexamathosone, 10 μg ml−1 cycloheximide, 10 μg ml−1 puromycin or 25 μg ml−1 chloramphenicol for 1 h, 4 h or 8 h. After the incubation period, 20 μCi of [35S]Met (EasyTag Protein Labelling Mix, [35S]-, 2 mCi (74 MBq), Perkin Elmer) was added for 1 h at room temperature. To quantitate the incorporation of [35S]Met into protein, aliquots of protein extracts were placed in 10% (w/v) TCA and incubated on ice for 30 min. The samples were filtered onto glass microfibre filters, and the filters were washed three times with 5 ml cold 5% (w/v) TCA and two times with 5 ml 95% ethanol. After drying, the filters were counted with a scintillation counter.
PS fractionation
PSs were fractionated from 500 mg of 15-day-old Arabidopsis seedlings over sucrose gradients as described23. Fractions were collected manually from the top, and total RNA was extracted with phenol/chloroform/isoamyl alcohol, precipitated with isopropanol, and treated with DNase I. The specific transcripts were examined by northern blot analysis. For the infection assays, PSs from wild type, NIK1-4-overexpressing and T474D-6-overexpressing lines were isolated 10 dpi with infectious CaLCuV clones.
PS gradients of dexamethasone-inducible T474D seedlings were prepared from 8-day-old seedlings, treated or not with 30 μg ml−1 dexamathosone. After 8 h of dexamethasone treatment, the seedlings were dried, frozen and ground in liquid nitrogen. Cytoplasmic extracts were prepared from 350 mg of powder that were re-suspended in the ice-cold extraction buffer. Cell debris was removed by centrifugation and 0.6 ml of the supernadant were loaded onto a 4.5 ml 15–55% sucrose density gradient and separated by ultracentrifugation. Fractions (16) of 310 μl were collected from the top. Fractions 2–7 correspond to density regions of complexes ≤80S (NPSs) and fractions 8–15 correspond to density regions containing small and large polysomes (PSs). Quantitation of NPSs and PSs from the absorbance profile was performed as described previously24. The quantitation of total RNA from the combined NPS and PS fractions was performed as described previously25. To quantify mRNA loading into PSs, total RNA isolated from combined NPS (2–7) and PS (9–15) fractions was extracted using Trizol (Invitrogen), precipitated with isopropanol and quantified by qRT–PCR, as described earlier. qRT–PCR on each fraction from the PS gradient of dexamethasone-inducible T474D seedlings was also performed for quantitation of the distribution of AtWWP1 and S39 transcripts after induction of T474D expression. Each sample was amplified with 1 cycle at 50 °C for 2 min, 95 °C for 10 min and 40 cycles at 95 °C for 15 s and 60 °C for 1 min. Values were normalized to actin.
CaLCuV inoculation and viral DNA accumulation
A. thaliana plants at the seven-leaf stage were inoculated with plasmids containing partial tandem repeats of CaLCuV DNA-A and DNA-B by biolistic delivery26, and the course of infection was monitored as described previously10. We used an attenuated form of the virus in which the coat protein ORF in the CaLCuV DNA-A was interrupted by the introduction of a stop codon at amino acid position 47. The inoculated plants were transferred to a growth chamber and examined for symptom development (leaf necrosis, chlorosis, leaf epinasty, leaf curly, young leaf death and stunted growth). Total nucleic acids were extracted from systemically infected leaves, and viral DNA was detected by PCR with DNA-B CaLCuV-specific primers (566CLCVBFBR1v, 5′-GGCGTGGGGTATCTTACTC-3′, and 1253CLCVBRBR1c, 5′-GACATAGCATCGGACATCC-3′) as well as the actin-specific primers as an endogenous control. In each experiment, 20 plants of each line (Col-0, nik1 and nik1-expressing NIK1 mutant proteins) were inoculated with 2 μg of tandemly repeated DNA-A plus DNA-B per plant. The course of infection was examined using data from three independent experiments.
Viral DNA accumulation was measured by qPCR. The reactions were prepared in a final volume of 10 μl using the Fast SYBR Green Master Mix (Life Technologies) according to the manufacturer’s instructions and analysed on a 7500 Real Time PCR System. Virus-specific primers were designed using Primer Express 3.0 (Life Technologies) and tested by conventional PCR using plasmids containing the complete DNA-B of each virus (106 copies per reaction). The genomic copies of CaLCuV were normalized against an internal control (18S rRNA) to consider template input variation between tubes. CaLCuV DNA was amplified with primers B-Fwd (5′-GGGCCTGGGCCTGTTAGT-3′) and B-Rvs (5′-ACGGAAGATGGGAGAGGAAGA-3′). In this case, the genomic unit refers to one copy of the DNA-B of CaLCuV. PCR reactions were run in parallel with primers 18S-Fwd (5′-TAATTTGCGCGCCTGCTGCC-3′) and 18S-Rvs (5′-TGTGCTGGCGACGCATCATT-3′) for the reference plant gene 18SRNA. Standard curves were obtained by regression analysis of the Ct values of each of the three replicates of a given dilution as a function of the log of the amount of DNA in each dilution. Standard curves containing viral DNA and host DNA served as references. Each sample was analysed in triplicate from at least two biological replicates.
Two-hybrid screening
The yeast reporter strain MaV203 (MATαleu2-3,112 trp1-901 his3200 ade2-101 gal4 gal80 SPAL10∷URA3 GAL1∷lacZ HIS3UAS GAL1∷HIS3-LYS2 can1R cyh2R; Trp- Leu- Ura-) was transformed sequentially with pBD-L10 and 25 μg of the pEXAD502 cDNA library previously prepared27 and approximately 5 × 106 transformants were screened for histidine protoprophy an β-galactosidase activity, as described27,28.
LIMYB-based plasmid constructs
The At5g05800 cDNA (LIMYB), which was isolated based on its ability to bind to rpL10A in yeast, was amplified by PCR from the cDNA library vector using AT5G05800-specific primers, re-amplified with the primers AttB1-Fwd and AttB1-Rvs, and cloned by recombination into the entry vectors pDONR201 and pDONR207. The resulting products were then transferred to different destination vectors for expression in yeast (pDEST32 and pDEST 22) and plants (pK7FWG2, 35S-YFP-casseteA-Nos-pCAMBIA1300, pSPYNEGW and pSPYCEGW). The resulting clones, pBD-At5g05800ST (pUFV1903) and pAD-At5g05800ST (pUFV1480), enabled the expression of AT5G05800 (LIMYB) in yeast as GAL4 binding domain (BD) or activation domain (AD) fusions, respectively. The clone pAt5g05800-GFP (pUFV1395) harboured the AT5G05800 (LIMYB) coding region fused to the GFP N terminus under the control of the 35S promoter, whereas in pYFP-At5g05800 (pUFV1886), the At5g05800 coding region was fused to the YFP C terminus. In the resulting clones, pSPYNE-At5g05800 (pUFV1658) and pSPYCE-At5g05800 (pUFV1657), the At5g05800 cDNA was linked to the YFP NE and the YFP CE, respectively, under the control of the 35S promoter.
The HA epitope was fused to the At5g05800 C terminus using the triple Gateway system, which consisted of the vector p5′, in which the 35S (2×) promoter had been previously cloned, the clone pUFV1378 (At5g05800 cDNA without a translational stop codon in pDONR201) and the vector p3′, in which a HA (6×) epitope had been previously inserted. The At5g05800 cDNA was transferred by recombination into the destination vector pK7M34GW along with 2×35S promoter and the HA tail, yielding the clone p2×35S-At5g05800-6HA (pUFV1984), which enabled the expression of At5g05800 fused in frame to HA under the control of the 35S promoter in plants. The same triple Gateway system was used to place the LIMYB fused in frame with Cherry or GFP under the control of the LIMYB promoter.
Transient expression in N. benthamiana leaves
To determine the subcellular localization of the proteins, N. benthamiana leaves were agro-infiltrated with pAt5g05800-GFP, pYFP-At5g05800 and pYFP-NAC81 (ref. 29), while for the co-immunoprecipitation assays, they were agro-infiltrated with pL10-GFP and p2×35S-At5g05800-6HA. Agrobacterium tumefaciens strain GV3101 was used for agroinfiltration. Agrobacterium infiltration was performed in the leaves of 3-week-old N. benthamiana, as previously described12.
Confocal microscopy
Approximately 72 h after agro-infiltration, 1-cm2 leaf ex-plants were excised, and the GFP and YFP fluorescence patterns were examined in epidermal cells with a ×40 or ×60 oil immersion objective and a Zeiss inverted LSM510 META laser scanning microscope equipped with an argon laser and a helium laser as excitation sources, as described27.
Co-immunoprecipitation assays
The co-immunoprecipitation assay was performed using the μMACSTM Epitope Tag Protein Isolation Kit (MACS/Miltenyi Biotec) according to the manufacturer’s instructions. Total protein extracts were prepared from N. benthamiana leaves that had been agro-infiltrated with pL10-GFP and/or p2×35S-At5g05800-6HA. At 48 h after infiltration, 200 mg of leaves were homogenized with 1 ml of lysis buffer (50 mM Tris-HCl pH 8.0, 1% (v/v) Nonidet P-40) and incubated for 2 h with anti-GFP magneticbeads (MACS/Miltenyi Biotec) at 4 °C undergentle agitation. After extensive washing of the magnetic beads, the bound proteins were eluted using 50 ml of elution buffer pre-warmed to 95 °C. The immunoprecipitated proteins were separated by 10% (w/v) SDS–PAGE and immunoblotted with anti-HA (Miltenyi Biotec, catalogue number 130-091-972) or anti-GFP (Miltenyi Biotec, catalogue number 130-091-833) monoclonal antibodies. The reacting antibodies were detected using Signal West Pico Chemiluminescent Substrate (Thermo Scientific) according to the manufacturer’s instructions.
BiFC
N. benthamiana leaves were agro-infiltrated as previously described for the transient expression experiments using the following combinations of recombinant plasmids: pSPYNE-At5g05800 + pSPYCE-rpL10A; pSPYCE-At5g05800 + pSPYNE-rpL10A; pSPYNE-At5g05800 + pSPYCE empty vector; pSPYNE-rpL10A + pSPYCE empty vector; pSPYCE-At5g05800 + pSPYNE empty vector; pSPYCE-rpL10A + pSPYNE empty vector. After incubation for 72 h, the leaf sectors were examined by confocal microscopy. YFP was excited at 514 nm using an argon laser, and YFP emission was detected using a 560–615 nm filter. The stability of the CE and NE YFP regions was monitored by immunoblotting of transfected leaf protein extracts with a polyclonal anti-YFP serum.
LIMYB- orRPL10A-expressing lines and T474D-expressinglimyb lines
Arabidopsis Col-0 was transformed with pAt5g05800-GFP, generating the transgenic line LIMYB-1; with pYFP-At5g05800, generating the transgenic lines LIMYB-2 and LiMYB-3; and with pYFP-NLS-L10, generating the transgenic lines RPL10-1, RPL10-2 and RPL10-3. Homozygous seeds of the T-DNA insertion limyb mutants (Salk_032054C and Salk_082995C) were obtained from the Arabidopsis Biological Resource Center. The limyb-32 mutant was transformed with the gain-of-function mutant T474D, generating the transgenic lines limyb/T474-1, limyb/T474D-2 and limyb/T474D-3. Primary transformants were selected using the appropriate antibiotic (50 μg ml−1 kanamycinor 30 μg ml−1 hygromycin), and the stable incorporation of the transgene in the plant genome was evaluated by PCR with gene-specific primers. The expression of the transgenes was monitored by qRT–PCR.
LIMYB promoter construct and GUS histochemical assay
Approximately 1 kb of the 5′ flanking sequences of NIK1, RPL10A and At5g05800 (LIMYB) were amplified from Arabidopsis genomic DNA using Taq Platinum and specific oligonucleotides and inserted into the cloning vector pCR8/GW/TOPO (Invitrogen). The promoter sequences were then transferred by recombination into the destination vector pMDC162. The resulting clones, pNIK1-MDC162, pL10A-MDC162 and pAt5g05800-MDC162 (pUFV 1892), harboured the respective promoter sequences of the three genes fused to the β-glucuronidase (GUS) reporter gene and were used to transform Arabidopsis Col-0 plants. GUS expression was assayed histochemically.
Luciferase reporter gene assay
The At1g29970 (RPL18A) promoter was isolated from Arabidopsis DNA by PCR and cloned into pDONR-P4-P1R, generating pUFV2155. The clone LUCF-term-pDON221 (pUFV 2132) harbours the firefly luciferase cDNA in the entry vector pDON221, whereas the clone 2×35S-RLUCF-pDONR-P2R-P3 (pUFV 2131) contains Renilla luciferase cDNA under the control of a 2× 35S promoter. The RPL18A promoter was transferred from pUFV2155 to the destination vector pK7m34GW (pUFV 1918), along with 2×35S-RLUCF-pDONR-P2R-P3 (pUFV 2131) and LUCF-term-pDON221 (pUFV 2132), by triple recombination. The resulting clone, prAt1g29970-Lucif-term-2X35S RLucif pH7M34GW (pUVF2231), harboured the firefly luciferase cDNA under the control of the rpL18A promoter, as well as the Renilla luciferase cDNA under the control of a 2× 35S promoter. The same procedure was used to clone the firefly luciferase cDNA under the control of the RPL28e, RPS13A and ubiquitin promoters for luciferase transactivation assays.
N. benthamiana leaves were agro-infiltrated with A. tumefaciens GV3101 strains carrying the following combinations of DNA constructs: At1g29970-Lucif-term-2×35S RLucif pH7M34GW; At1g29970-Lucif-term-2X35S RLucif pH7M34GW + AT5G05800NS-pK7FWG2; At1g29970-Lucif-term-2X35S RLucif pH7M34GW + AT5G05800NS-pK7FWG2 + rpL10ANS-pK7FWG2; and At1g29970-Lucif-term-2×35S RLucif pH7M34GW + rpL10ANS-pK7FWG2. Forty-eight hours after infiltration, 200 mg of leaf tissue was harvested for total protein extraction. Luciferase activity was assayed with the Dual-Luciferase Reporter Assay System (Promega) according to the manufacturer’s instructions.
Protoplast preparation fromA. thaliana leaves and transient expression assays
Protoplasts were prepared from 15-day-old Arabidopsis seedlings as previously described30, except that digestion was initiated for 30 s under vacuum and then prolonged for 6 h with agitation at 80 r.p.m. The transient expression assays were performed by electroporating (250 V, 250 μF) 10 mg of the expression cassette DNA and 30 μg of the sheared salmon sperm DNA into 2 × 105–5 × 106 protoplasts in a final volume of 0.8 ml. The protoplasts were diluted into 8 ml of MS medium supplemented with 0.2 mg ml−1 2,4-dichlorophenoxyacetic acid and 0.8 M mannitol at pH 5.7. After 36 h of incubation in the dark, the protoplasts were washed with 0.8 M mannitol plus 20 mM MES at pH 5.7 and frozen in liquid N2 until further use.
Statistical analyses
All statistical analyses (including gene expression clustering, qRT–PCR and protein synthesis data) were performed in R.
Cluster analysis
Cluster analysis of the RNA-sequencing data was used to classify the treatments (mock-inoculated and infected wild type, T474D and NIK1) according to similarities in the profiles of genome-wide expression at 10 and 21 dpi. The uncertainty of clustering results caused by sampling variations was verified by the probability-based cluster analysis, which was implemented using the pvclust package of R software. In this context, the bootstrap (BP) and approximately unbiased (AU) probability are used to validate the reported cluster. The BP and AU values are the percentage that a given cluster appears in the bootstrap andmultiscale bootstrap replicates, respectively.
RNA-sequencing differential gene expression analysis
The edgeR package of R/bioconductor software was used to carry out the gene expression analysis. This package assumes negative binomial distribution for the read counts, and the used normalization is given by the TMM (trimmed mean of M value) method. The significance of differential gene expression was reported in terms of q values (FDR-adjusted P values).
Enrichment analysis
Gene set enrichment analysis (GSEA) was used to uncover biological processes associated with sets of differentially expressed genes instead of focusing on individual genes. This analysis was implemented in the GOstats package of R/bioconductor software by using the function GSEAGOHyperGParams, which uses the relationships among the gene ontology terms as extra information in the statistical inference of groups.
Percentage of protein synthesis analysis
Analysis of protein synthesis was based on the one-proportion one-sided (less) test using chi-squared distribution. The tested hypothesis was the 100% synthesized labelled protein for each treatment; thus, if a given treatment is significant, it indicates that the proportion (or percentage) is different from 100%.
Confidence interval analysis
Nonparametric bootstrap confidence intervals were used on our graphics to increase the accuracy of the confidence limits by an overlapping analysis31. This method was introduced as a nonparametric device for estimating standard errors and biases. It is an automatic algorithm for producing highly accurate confidence limits from a bootstrap distribution implemented with the ‘boot’ package in R software.
Phylogenetic analysis
MYB family sequences were retrieved from the Agris database (http://arabidopsis.med.ohio-state.edu). The alignment was performed by Maft aligner software32, and the tree was built by Fasttree33 software.
Extended Data
Supplementary Material
Acknowledgments
This research was financially supported by the following grants from Brazilian government agencies: Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) grants 573600/2008-2 and 470287/2011-0 (to E.P.B.F.) and FAPEMIG grant CBB-APQ-00070-09 (to E.P.B.F.); and by the US National Institutes of Health grant 5R01-GM94428 (to J.C.). O.J.B.B. was supported by a Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) graduate fellowship; C.Z., K.V.G.L., J.P.B.M., I.P.C., B.C.G. and P.A.B.R. were supported by CNPq graduate fellowships; K.J.T.N., W.A.P. and M.D. were supported by postdoctoral fellowships from CNPq. A.A.S. was supported by postdoctoral fellowships from CAPES; and M.A.C.S. was the recipient of an undergraduate scholarship from CNPq. J.C. is an Investigator of the Howard Hughes Medical Institute.
Footnotes
Online Content Methods, along with any additional Extended Data display items and Source Data, are available in the online version of the paper; references unique to these sections appear only in the online paper.
Supplementary Information is available in the online version of the paper.
Author Contributions C.Z. and J.P.B.M. co-wrote the manuscript and performed most of the experiments related to LIMYB isolation and characterization. K.V.G.L. performed the T474D-related experiments. K.J.T.N. performed qPCR for viral DNA. W.A.P. generated the RPL10 constructs and conducted plant transformation. O.J.B.B. and F.F.S. performed the bioinformatic analysis of the RNA-sequencing data and the statistical analysis of the data. M.D. performed the infectivity assays. B.C.G. conducted qRT–PCR. P.A.B.R. and I.P.C. performed the protein synthesis assays. V.A.P.L. performed complementation assays. G.S.-M. performed the bimolecular fluorescence complementation experiments. M.A.C.S. performed the tissue expression experiments. A.A.S. constructed the mutant proteins and designed the infectivity assays. J.C. conceived the experiments and edited the final draft. E.P.F.B. co-wrote the manuscript, designed the experiments and directed the project.
RNA-sequencing data have been deposited in the Gene Expression Omnibus under accession number GSE56922
The authors declare no competing financial interests.
Readers are welcome to comment on the online version of the paper.
References
- 1.Pumplin N, Voinnet O. RNA silencing suppression by plant pathogens: defence, counter-defence and counter-counter-defence. Nature Rev Microbiol. 2013;11:745–760. doi: 10.1038/nrmicro3120. [DOI] [PubMed] [Google Scholar]
- 2.Mandadi KK, Scholthof KB. Plant immune responses against viruses: how does a virus cause disease? Plant Cell. 2013;25:1489–1505. doi: 10.1105/tpc.113.111658. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3.Jones JD, Dangl JL. The plant immune system. Nature. 2006;444:323–329. doi: 10.1038/nature05286. [DOI] [PubMed] [Google Scholar]
- 4.Hanley-Bowdoin L, Bejarano ER, Robertson D, Mansoor S. Geminiviruses: masters at redirecting and reprogramming plant processes. Nature Rev Microbiol. 2013;11:777–788. doi: 10.1038/nrmicro3117. [DOI] [PubMed] [Google Scholar]
- 5.Santos AA, Lopes KVG, Apfta JAC, Fontes EPB. NSP-interacting kinase, NIK: a transducer of plant defence signalling. J Exp Bot. 2010;61:3839–3845. doi: 10.1093/jxb/erq219. [DOI] [PubMed] [Google Scholar]
- 6.Mariano AC, et al. Identification of a novel receptor- like protein kinase that interacts with a geminivirus nuclear shuttle protein. Virology. 2004;318:24–31. doi: 10.1016/j.virol.2003.09.038. [DOI] [PubMed] [Google Scholar]
- 7.Sakamoto T, et al. The tomato RLK superfamily: phylogeny and functional predictions about the role of the LRRII-RLK subfamily in antiviral defense. BMC Plant Biol. 2012;12:229. doi: 10.1186/1471-2229-12-229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Shiu SH, Bleecker AB. Receptor-like kinases from Arabidopsis form a monophyletic gene family related to animal receptor kinases. Proc Natl Acad Sci USA. 2001;98:10763–10768. doi: 10.1073/pnas.181141598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Chinchilla D, Shan L, He P, de Vries S, Kemmerling B. One for all: the receptor-associated kinase BAK1. Trends Plant Sci. 2009;14:535–541. doi: 10.1016/j.tplants.2009.08.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Fontes EPB, Santos AA, Luz DF, Waclawovsky AJ, Chory J. The geminivirus NSP acts as virulence factor to suppress an innate transmembrane receptor kinase-mediated defense signaling. Genes Dev. 2004;18:2545–2556. doi: 10.1101/gad.1245904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Santos AA, Carvalho CM, Florentino LH, Ramos JJO, Fontes EPB. Conserved threonine residues within the A-loop of the receptor NIK differentially regulate the kinase function required for antiviral signaling. PLoS ONE. 2009;4:e5781. doi: 10.1371/journal.pone.0005781. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Carvalho CM, et al. Regulated nuclear trafficking of rpL10A mediated by NIK1 represents a defense strategy of plant cells against virus. PLoS Pathog. 2008;4:e1000247. doi: 10.1371/journal.ppat.1000247. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Rocha CS, Santos AA, Machado JPB, Fontes EPB. The ribosomal protein L10/QM-like protein is a component of the NIK-mediated antiviral signaling. Virology. 2008;380:165–169. doi: 10.1016/j.virol.2008.08.005. [DOI] [PubMed] [Google Scholar]
- 14.Oh HS, Kwon H, Sun SK, Yang CH. QM, a putative tumor suppressor, regulates proto-oncogene c-Yes. J Biol Chem. 2002;277:36489–36498. doi: 10.1074/jbc.M201859200. [DOI] [PubMed] [Google Scholar]
- 15.Imafuku I, et al. Presenilin 1 suppresses the function of c-Jun homodimers via interaction with QM/Jif-1. J Cell Biol. 1999;147:121–134. doi: 10.1083/jcb.147.1.121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Monteclaro FS, Vogt PK. A Jun-binding protein related to a putative tumor suppressor. Proc Natl Acad Sci USA. 1993;90:6726–6730. doi: 10.1073/pnas.90.14.6726. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Vinatzer BA, et al. The type III effector repertoire of Pseudomona syringae pv. syringae B728a and its role in survival and disease on host and non-host plants. Mol Microbiol. 2006;62:26–44. doi: 10.1111/j.1365-2958.2006.05350.x. [DOI] [PubMed] [Google Scholar]
- 18.Delú-Filho N, et al. A sucrose binding protein homologue from soybean affects sucrose uptake in transgenic tobacco suspension-cultured cells. Plant Physiol Biochem. 2000;38:353–361. [Google Scholar]
- 19.Robinson MD, McCarthy DJ, Smyth GK. EdgeR: a Bioconductor package for differential expression analysis of digital gene expression data. Bioinformatics. 2010;26:139–140. doi: 10.1093/bioinformatics/btp616. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Robinson MD, Oshlack A. A scaling normalization method for differential expression analysis of RNA-seq data. Genome Biol. 2010;11:R25. doi: 10.1186/gb-2010-11-3-r25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Langmead B, et al. Ultrafast and memory-efficient alignment of short DNA sequences to the human genome. Genome Biol. 2009;10:R25. doi: 10.1186/gb-2009-10-3-r25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Ward JH., Jr Hierarchical grouping to optimize an objective function. J Am Stat Assoc. 1963;58:236–244. [Google Scholar]
- 23.Kim TH, Kim BH, Yahalom A, Chamovitz DA, von Arnim AG. Translational regulation via 59 mRNA leader sequences revealed by mutational analysis of the Arabidopsis translation initiation factor subunit eIF3h. Plant Cell. 2004;16:3341–3356. doi: 10.1105/tpc.104.026880. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Kawaguchi R, William AJ, Bray EA, Bailey-Serres J. Water deficit-induced translational control in Nicotiana tabacum. Plant Cell Environ. 2003;26:221–229. [Google Scholar]
- 25.Kawaguchi R, Girke T, Bray EA, Bailey-Serres J. Differential mRNA translation contributes to gene regulation under non-stress and dehydration stress conditions in Arabidopsis thaliana. Plant J. 2004;38:823–839. doi: 10.1111/j.1365-313X.2004.02090.x. [DOI] [PubMed] [Google Scholar]
- 26.Santos AA, Florentino LH, Pires ABL, Fontes EPB. Geminivirus: biolistic inoculation and molecular diagnosis. Methods Mol Biol. 2008;451:563–579. doi: 10.1007/978-1-59745-102-4_39. [DOI] [PubMed] [Google Scholar]
- 27.Florentino LH, et al. A PERK-like receptor kinase interacts with the geminivirus nuclear shuttle protein and potentiates viral infection. J Virol. 2006;80:6648–6656. doi: 10.1128/JVI.00173-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Carvalho CM, et al. A novel nucleocytoplasmic traffic GTPase identified as a functional target of the bipartite geminivirus nuclear shuttle protein. Plant J. 2008;55:869–880. doi: 10.1111/j.1365-313X.2008.03556.x. [DOI] [PubMed] [Google Scholar]
- 29.Pinheiro GL, et al. Complete inventory of soybean NAC transcription factors: sequence conservation and expression analysis uncover their distinct roles in stress response. Gene. 2009;444:10–23. doi: 10.1016/j.gene.2009.05.012. [DOI] [PubMed] [Google Scholar]
- 30.Costa MDL, et al. A new branch of endoplasmic reticulum-stress signaling and the osmotic signal converge on plant specific asparagine-rich proteins to promote cell death. J Biol Chem. 2008;283:20209–20219. doi: 10.1074/jbc.M802654200. [DOI] [PubMed] [Google Scholar]
- 31.DiCiccio TJ, Efron B. Bootstrap confidence intervals. Stat Sci. 1996;11:189–228. [Google Scholar]
- 32.Katoh K, Standley DM. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 2013;30:772–780. doi: 10.1093/molbev/mst010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Price MN, Dehal PS, Arkin AP. FastTree 2—approximately maximum-likelihood trees for large alignments. PLoS ONE. 2010;5:e9490. doi: 10.1371/journal.pone.0009490. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.