首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 437 毫秒
1.
The identification of cellular proteins associated with virus replicase complexes is crucial to our understanding of virus-host interactions, influencing the host range, replication, and virulence of viruses. A previous in vitro study has demonstrated that partially purified Bamboo mosaic virus (BaMV) replicase complexes can be employed for the replication of both BaMV genomic and satellite BaMV (satBaMV) RNAs. In this study, we investigated the BaMV and satBaMV 3' untranslated region (UTR) binding proteins associated with these replicase complexes. Two cellular proteins with molecular masses of ~35 and ~55 kDa were specifically cross-linked with RNA elements, whereupon the ~35-kDa protein was identified as the glycolytic enzyme glyceraldehyde 3-phosphate dehydrogenase (GAPDH). Gel mobility shift assays confirmed the direct interaction of GAPDH with the 3' UTR sequences, and competition gel shift analysis revealed that GAPDH binds preferentially to the positive-strand BaMV and satBaMV RNAs over the negative-strand RNAs. It was observed that the GAPDH protein binds to the pseudoknot poly(A) tail of BaMV and stem-loop-C poly(A) tail of satBaMV 3' UTR RNAs. It is important to note that knockdown of GAPDH in Nicotiana benthamiana enhances the accumulation of BaMV and satBaMV RNA; conversely, transient overexpression of GAPDH reduces the accumulation of BaMV and satBaMV RNA. The recombinant GAPDH principally inhibits the synthesis of negative-strand RNA in exogenous RdRp assays. These observations support the contention that cytosolic GAPDH participates in the negative regulation of BaMV and satBaMV RNA replication.  相似文献   

2.

Background

Satellite RNAs (satRNAs), virus parasites, are exclusively associated with plant virus infection and have attracted much interest over the last 3 decades. Upon virus infection, virus-specific small interfering RNAs (vsiRNAs) are produced by dicer-like (DCL) endoribonucleases for anti-viral defense. The composition of vsiRNAs has been studied extensively; however, studies of satRNA-derived siRNAs (satsiRNAs) or siRNA profiles after satRNA co-infection are limited. Here, we report on the small RNA profiles associated with infection with Bamboo mosaic virus (BaMV) and its two satellite RNAs (satBaMVs) in Nicotiana benthamiana and Arabidopsis thaliana.

Methodology/Principal Findings

Leaves of N. benthamiana or A. thaliana inoculated with water, BaMV alone or co-inoculated with interfering or noninterfering satBaMV were collected for RNA extraction, then large-scale Solexa sequencing. Up to about 20% of total siRNAs as BaMV-specific siRNAs were accumulated in highly susceptible N. benthamiana leaves inoculated with BaMV alone or co-inoculated with noninterfering satBaMV; however, only about 0.1% of vsiRNAs were produced in plants co-infected with interfering satBaMV. The abundant region of siRNA distribution along BaMV and satBaMV genomes differed by host but not by co-infection with satBaMV. Most of the BaMV and satBaMV siRNAs were 21 or 22 nt, of both (+) and (−) polarities; however, a higher proportion of 22-nt BaMV and satBaMV siRNAs were generated in N. benthamiana than in A. thaliana. Furthermore, the proportion of non-viral 24-nt siRNAs was greatly increased in N. benthamiana after virus infection.

Conclusions/Significance

The overall composition of vsiRNAs and satsiRNAs in the infected plants reflect the combined action of virus, satRNA and different DCLs in host plants. Our findings suggest that the structure and/or sequence demands of various DCLs in different hosts may result in differential susceptibility to the same virus. DCL2 producing 24-nt siRNAs under biotic stresses may play a vital role in the antiviral mechanism in N. benthamiana.  相似文献   

3.
4.
Plant virus‐based gene‐silencing vectors have been extensively and successfully used to elucidate functional genomics in plants. However, only limited virus‐induced gene‐silencing (VIGS) vectors can be used in both monocot and dicot plants. Here, we established a dual gene‐silencing vector system based on Bamboo mosaic virus (BaMV) and its satellite RNA (satBaMV). Both BaMV and satBaMV vectors could effectively silence endogenous genes in Nicotiana benthamiana and Brachypodium distachyon. The satBaMV vector could also silence the green fluorescent protein (GFP) transgene in GFP transgenic N. benthamiana. GFP transgenic plants co‐agro‐inoculated with BaMV and satBaMV vectors carrying sulphur and GFP genes, respectively, could simultaneously silence both genes. Moreover, the silenced plants could still survive with the silencing of genes essential for plant development such as heat‐shock protein 90 (Hsp90) and Hsp70. In addition, the satBaMV‐ but not BaMV‐based vector could enhance gene‐silencing efficiency in newly emerging leaves of N. benthamiana deficient in RNA‐dependant RNA polymerase 6. The dual gene‐silencing vector system of BaMV and satBaMV provides a novel tool for comparative functional studies in monocot and dicot plants.  相似文献   

5.
Satellite RNA of Bamboo mosaic virus (satBaMV), a single-stranded mRNA type satellite encoding a protein of 20 kDa (P20), depends on the helper BaMV for replication and encapsidation. Two satBaMV isolates, BSF4 and BSL6, exhibit distinctly differential phenotypes in Nicotiana benthamiana plants when coinoculated with BaMV RNA. BSL6 significantly reduces BaMV RNA replication and suppresses the BaMV-induced symptoms, whereas BSF4 does not. By studies with chimeric satBaMVs generated by exchanging the components between BSF4 and BSL6, the genetic determinants responsible for the downregulation of BaMV replication and symptom expression were mapped at the 5' untranslated region (UTR) of BSL6. The 5' UTR of BSL6 alone is sufficient to diminish BaMV RNA replication when the 5' UTR is inserted in cis into the BaMV expression vector or when coinoculation with mutants that block the synthesis of P20 protein takes place. Further, the 5' UTR of natural satBaMV isolates contains one hypervariable (HV) region which folds into a conserved apical hairpin stem-loop (AHSL) structure (W. B. Yeh, Y. H. Hsu, H. C. Chen, and N. S. Lin, Virology 330:105-115, 2004). Interchanges of AHSL segment of HV regions between BSF4 and BSL6 led to the ability of chimeric satBaMV to interfere with BaMV replication and symptom expression. The conserved secondary structure within the HV region is a potent determinant of the downregulation of helper virus replication.  相似文献   

6.
7.
Genomes and antigenomes of many positive-strand RNA viruses contain 3′-poly(A) and 5′-poly(U) tracts, respectively, serving as mutual templates. Mechanism(s) controlling the length of these homopolymeric stretches are not well understood. Here, we show that in coxsackievirus B3 (CVB3) and three other enteroviruses the poly(A) tract is ~80–90 and the poly(U) tract is ~20 nt-long. Mutagenesis analysis indicate that the length of the CVB3 3′-poly(A) is determined by the oriR, a cis-element in the 3′-noncoding region of viral RNA. In contrast, while mutations of the oriR inhibit initiation of (−) RNA synthesis, they do not affect the 5′-poly(U) length. Poly(A)-lacking genomes are able to acquire genetically unstable AU-rich poly(A)-terminated 3′-tails, which may be generated by a mechanism distinct from the cognate viral RNA polyadenylation. The aberrant tails ensure only inefficient replication. The possibility of RNA replication independent of oriR and poly(A) demonstrate that highly debilitated viruses are able to survive by utilizing ‘emergence’, perhaps atavistic, mechanisms.  相似文献   

8.
Positive-strand RNA viruses use diverse mechanisms to regulate viral and host gene expression for ensuring their efficient proliferation or persistence in the host. We found that a small viral noncoding RNA (0.4 kb), named SR1f, accumulated in Red clover necrotic mosaic virus (RCNMV)-infected plants and protoplasts and was packaged into virions. The genome of RCNMV consists of two positive-strand RNAs, RNA1 and RNA2. SR1f was generated from the 3′ untranslated region (UTR) of RNA1, which contains RNA elements essential for both cap-independent translation and negative-strand RNA synthesis. A 58-nucleotide sequence in the 3′ UTR of RNA1 (Seq1f58) was necessary and sufficient for the generation of SR1f. SR1f was neither a subgenomic RNA nor a defective RNA replicon but a stable degradation product generated by Seq1f58-mediated protection against 5′→3′ decay. SR1f efficiently suppressed both cap-independent and cap-dependent translation both in vitro and in vivo. SR1f trans inhibited negative-strand RNA synthesis of RCNMV genomic RNAs via repression of replicase protein production but not via competition of replicase proteins in vitro. RCNMV seems to use cellular enzymes to generate SR1f that might play a regulatory role in RCNMV infection. Our results also suggest that Seq1f58 is an RNA element that protects the 3′-side RNA sequences against 5′→3′ decay in plant cells as reported for the poly(G) tract and stable stem-loop structure in Saccharomyces cerevisiae.  相似文献   

9.
10.
Bamboo mosaic virus (BaMV) satellite RNA (satBaMV) depends on BaMV for its replication and encapsidation. SatBaMV-encoded P20 protein is an RNA-binding protein that facilitates satBaMV systemic movement in co-infected plants. Here, we examined phosphorylation of P20 and its regulatory functions. Recombinant P20 (rP20) was phosphorylated by host cellular kinase(s) in vitro, and matrix-assisted laser desorption/ionization time-of-flight mass spectrometry and mutational analyses revealed Ser-11 as the phosphorylation site. The phosphor-mimic rP20 protein interactions with satBaMV-translated mutant P20 were affected. In overlay assay, the Asp mutation at S11 (S11D) completely abolished the self-interaction of rP20 and significantly inhibited the interaction with both the WT and S11A rP20. In chemical cross-linking assays, S11D failed to oligomerize. Electrophoretic mobility shift assay and subsequent Hill transformation analysis revealed a low affinity of the phospho-mimicking rP20 for satBaMV RNA. Substantial modulation of satBaMV RNA conformation upon interaction with nonphospho-mimic rP20 in circular dichroism analysis indicated formation of stable satBaMV ribonucleoprotein complexes. The dissimilar satBaMV translation regulation of the nonphospho- and phospho-mimic rP20 suggests that phosphorylation of P20 in the ribonucleoprotein complex converts the translation-incompetent satBaMV RNA to messenger RNA. The phospho-deficient or phospho-mimicking P20 mutant of satBaMV delayed the systemic spread of satBaMV in co-infected Nicotiana benthamiana with BaMV. Thus, satBaMV likely regulates the formation of satBaMV RNP complex during co-infection in planta.  相似文献   

11.
Song BH  Yun SI  Choi YJ  Kim JM  Lee CH  Lee YM 《RNA (New York, N.Y.)》2008,14(9):1791-1813
Tertiary or higher-order RNA motifs that regulate replication of positive-strand RNA viruses are as yet poorly understood. Using Japanese encephalitis virus (JEV), we now show that a key element in JEV RNA replication is a complex RNA motif that includes a string of three discontinuous complementary sequences (TDCS). The TDCS consists of three 5-nt-long strands, the left (L) strand upstream of the translation initiator AUG adjacent to the 5′-end of the genome, and the middle (M) and right (R) strands corresponding to the base of the Flavivirus-conserved 3′ stem–loop structure near the 3′-end of the RNA. The three strands are arranged in an antiparallel configuration, with two sets of base-pairing interactions creating L-M and M-R duplexes. Disrupting either or both of these duplex regions of TDCS completely abolished RNA replication, whereas reconstructing both duplex regions, albeit with mutated sequences, fully restored RNA replication. Modeling of replication-competent genomes recovered from a large pool of pseudorevertants originating from six replication-incompetent TDCS mutants suggests that both duplex base-pairing potentials of TDCS are required for RNA replication. In all cases, acquisition of novel sequences within the 3′M-R duplex facilitated a long-range RNA–RNA interaction of its 3′M strand with either the authentic 5′L strand or its alternative (invariably located upstream of the 5′ initiator), thereby restoring replicability. We also found that a TDCS homolog is conserved in other flaviviruses. These data suggest that two duplex base-pairings defined by the TDCS play an essential regulatory role in a key step(s) of Flavivirus RNA replication.  相似文献   

12.
We compared infection of Nicotiana benthamiana plants by the positive-sense RNA viruses Cucumber mosaic virus (CMV), Potato virus Y (PVY), and by a Potato virus X (PVX) vector, the latter either unaltered or expressing the CMV 2b protein or the PVY HCPro suppressors of silencing, at 25°C vs. 30°C, or at standard (~401 parts per million, ppm) vs. elevated (970 ppm) CO2 levels. We also assessed the activities of their suppressors of silencing under those conditions. We found that at 30°C, accumulation of the CMV isolate and infection symptoms remained comparable to those at 25°C, whereas accumulation of the PVY isolate and those of the three PVX constructs decreased markedly, even when expressing the heterologous suppressors 2b or HCPro, and plants had either very attenuated or no symptoms. Under elevated CO2 plants grew larger, but contained less total protein/unit of leaf area. In contrast to temperature, infection symptoms remained unaltered for the five viruses at elevated CO2 levels, but viral titers in leaf disks as a proportion of the total protein content increased in all cases, markedly for CMV, and less so for PVY and the PVX constructs. Despite these differences, we found that neither high temperature nor elevated CO2 prevented efficient suppression of silencing by their viral suppressors in agropatch assays. Our results suggest that the strength of antiviral silencing at high temperature or CO2 levels, or those of the viral suppressors that counteract it, may not be the main determinants of the observed infection outcomes.  相似文献   

13.
We have previously reported that the NS3 helicase (N3H) and NS5B-to-3′X (N5BX) regions are important for the efficient replication of hepatitis C virus (HCV) strain JFH-1 and viral production in HuH-7 cells. In the current study, we investigated the relationships between HCV genome replication, virus production, and the structure of N5BX. We found that the Q377R, A450S, S455N, R517K, and Y561F mutations in the NS5B region resulted in up-regulation of J6CF NS5B polymerase activity in vitro. However, the activation effects of these mutations on viral RNA replication and virus production with JFH-1 N3H appeared to differ. In the presence of the N3H region and 3′ untranslated region (UTR) of JFH-1, A450S, R517K, and Y561F together were sufficient to confer HCV genome replication activity and virus production ability to J6CF in cultured cells. Y561F was also involved in the kissing-loop interaction between SL3.2 in the NS5B region and SL2 in the 3′X region. We next analyzed the 3′ structure of HCV genome RNA. The shorter polyU/UC tracts of JFH-1 resulted in more efficient RNA replication than J6CF. Furthermore, 9458G in the JFH-1 variable region (VR) was responsible for RNA replication activity because of its RNA structures. In conclusion, N3H, high polymerase activity, enhanced kissing-loop interactions, and optimal viral RNA structure in the 3′UTR were required for J6CF replication in cultured cells.  相似文献   

14.
15.
16.
The small size of RNA virus genomes (2-to-32 kb) has been attributed to high mutation rates during replication, which is thought to lack proof-reading. This paradigm is being revisited owing to the discovery of a 3′-to-5′ exoribonuclease (ExoN) in nidoviruses, a monophyletic group of positive-stranded RNA viruses with a conserved genome architecture. ExoN, a homolog of canonical DNA proof-reading enzymes, is exclusively encoded by nidoviruses with genomes larger than 20 kb. All other known non-segmented RNA viruses have smaller genomes. Here we use evolutionary analyses to show that the two- to three-fold expansion of the nidovirus genome was accompanied by a large number of replacements in conserved proteins at a scale comparable to that in the Tree of Life. To unravel common evolutionary patterns in such genetically diverse viruses, we established the relation between genomic regions in nidoviruses in a sequence alignment-free manner. We exploited the conservation of the genome architecture to partition each genome into five non-overlapping regions: 5′ untranslated region (UTR), open reading frame (ORF) 1a, ORF1b, 3′ORFs (encompassing the 3′-proximal ORFs), and 3′ UTR. Each region was analyzed for its contribution to genome size change under different models. The non-linear model statistically outperformed the linear one and captured >92% of data variation. Accordingly, nidovirus genomes were concluded to have reached different points on an expansion trajectory dominated by consecutive increases of ORF1b, ORF1a, and 3′ORFs. Our findings indicate a unidirectional hierarchical relation between these genome regions, which are distinguished by their expression mechanism. In contrast, these regions cooperate bi-directionally on a functional level in the virus life cycle, in which they predominantly control genome replication, genome expression, and virus dissemination, respectively. Collectively, our findings suggest that genome architecture and the associated region-specific division of labor leave a footprint on genome expansion and may limit RNA genome size.  相似文献   

17.
The tertiary structure in the 3′-untranslated region (3′-UTR) of Bamboo mosaic virus (BaMV) RNA is known to be involved in minus-strand RNA synthesis. Proteins found in the RNA-dependent RNA polymerase (RdRp) fraction of BaMV-infected leaves interact with the radio labeled 3′-UTR probe in electrophoretic mobility shift assays (EMSA). Results derived from the ultraviolet (UV) cross-linking competition assays suggested that two cellular factors, p43 and p51, interact specifically with the 3′-UTR of BaMV RNA. p43 and p51 associate with the poly(A) tail and the pseudoknot of the BaMV 3′-UTR, respectively. p51-containing extracts specifically down-regulated minus-strand RNA synthesis when added to in vitro RdRp assays. LC/MS/MS sequencing indicates that p43 is a chloroplast phosphoglycerate kinase (PGK). When the chloroplast PKG levels were knocked down in plants, using virus-induced gene silencing system, the accumulation level of BaMV coat protein was also reduced.  相似文献   

18.
The accumulation of heat shock protein 70 (Hsp70) generally occurs in plants infected with viruses. However, the effect of Hsp70 accumulation on plant viral infection and pathogenesis remains elusive. In this study, the expression of six Hsp70 genes was found to be induced by the four diverse RNA viruses, Tobacco mosaic virus, Potato virus X (PVX), Cucumber mosaic virus and Watermelon mosaic virus, in Nicotiana benthamiana. Heat treatment enhanced the accumulation and systemic infection of these viruses. Similar results were obtained for viral infection in plants heterologously expressing an Arabidopsis cytoplasmic Hsp70 through either a PVX vector or Agrobacterium infiltration. In contrast, viral infection was compromised in cytoplasmic NbHsp70c‐1 gene‐silenced plants. These data demonstrate that the cytoplasmic Hsp70s can enhance the infection of N. benthamiana by diverse viruses.  相似文献   

19.
20.
The core protein of hepatitis C virus (HCV) has been shown previously to act as a potent nucleic acid chaperone in vitro, promoting the dimerization of the 3′-untranslated region (3′-UTR) of the HCV genomic RNA, a process probably mediated by a small, highly conserved palindromic RNA motif, named DLS (dimer linkage sequence) [G. Cristofari, R. Ivanyi-Nagy, C. Gabus, S. Boulant, J. P. Lavergne, F. Penin and J. L. Darlix (2004) Nucleic Acids Res., 32, 2623–2631]. To investigate in depth HCV RNA dimerization, we generated a series of point mutations in the DLS region. We find that both the plus-strand 3′-UTR and the complementary minus-strand RNA can dimerize in the presence of core protein, while mutations in the DLS (among them a single point mutation that abolished RNA replication in a HCV subgenomic replicon system) completely abrogate dimerization. Structural probing of plus- and minus-strand RNAs, in their monomeric and dimeric forms, indicate that the DLS is the major if not the sole determinant of UTR RNA dimerization. Furthermore, the N-terminal basic amino acid clusters of core protein were found to be sufficient to induce dimerization, suggesting that they retain full RNA chaperone activity. These findings may have important consequences for understanding the HCV replicative cycle and the genetic variability of the virus.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号