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 31 毫秒
1.
Jonczyk M  Pathak KB  Sharma M  Nagy PD 《Virology》2007,362(2):320-330
Plus-strand RNA virus replication takes place on distinct membranous surfaces in infected cells via the assembly of viral replicase complexes involving multiple viral and host proteins. One group of tombusviruses, such as Tomato bushy stunt virus (TBSV), replicate on the surfaces of peroxisomal membranes in plant and yeast cells. Surprisingly, previous genome-wide screen performed in yeast demonstrated that a TBSV replicon RNA replicated as efficiently in yeast defective in peroxisome biogenesis as in the wt yeast (Panavas et al., Proc Natl Acad Sci U S A, 2005). To further test how the lack of peroxisomes could affect tombusvirus replication, we used yeast cells missing either PEX3 or PEX19 genes, which are absolutely essential for peroxisome biogenesis. Confocal microscopy-based approach revealed that the wild-type tombusvirus p33 replication protein accumulated in the endoplasmic reticulum (ER) in pex3Delta or pex19Delta yeast, suggesting that tombusvirus replication could take place on the surface of ER membrane. The activities of the isolated tombusvirus replicase preparations from wt, pex3Delta or pex19Delta yeasts were comparable, demonstrating that the assembly of the replicase was as efficient in the ER as in the authentic subcellular environments. The generation/accumulation of tombusvirus recombinants was similar in wt, pex3Delta and pex19Delta yeasts, suggesting that the rate of mistakes occurring during tombusvirus replication is comparable in the presence or absence of peroxisomes. Overall, this work demonstrates that a tombusvirus, relying on the wt replication proteins, can efficiently replicate on an alternative intracellular membrane. This suggests that RNA viruses might have remarkable flexibility for using various host membranes for their replication.  相似文献   

2.
Shapka N  Stork J  Nagy PD 《Virology》2005,343(1):65-78
Replication of the nonsegmented, plus-stranded RNA genome of Cucumber necrosis tombusvirus (CNV) requires two essential overlapping viral-coded replication proteins, the p33 replication co-factor and the p92 RNA-dependent RNA polymerase. In this paper, we demonstrate that p33 is phosphorylated in vivo and in vitro by a membrane-bound plant kinase. Phosphorylation of p33 was also demonstrated in vitro by using purified protein kinase C. The related p28 replication protein of Turnip crinkle virus was also found to be phosphorylated in vivo, suggesting that posttranslational modification of replication proteins is a general feature among members of the large Tombusviridae family. Based on in vitro studies with purified recombinant p33, we show evidence for phosphorylation of threonine and serine residues adjacent to the essential RNA-binding site in p33. Phosphorylation-mimicking aspartic acid mutations rendered p33 nonfunctional in plant protoplasts and in yeast, a model host. Comparable mutations within the prereadthrough portion of p92 did not abolish replication. The nonphosphorylation-mimicking alanine mutants of CNV were able to replicate in plant protoplasts and in yeast, albeit with reduced efficiency when compared to the wild type. These alanine mutants also showed altered subgenomic RNA synthesis and a reduction in the ratio between plus- and minus-strand RNAs produced during CNV infection. These findings suggest that phosphorylation of threonine/serine residues adjacent to the essential RNA-binding site in the auxiliary p33 protein likely plays a role in viral RNA replication and subgenomic RNA synthesis during tombusvirus infections.  相似文献   

3.
Panavas T  Stork J  Nagy PD 《Virology》2006,352(1):110-120
Plus-stranded RNA viruses replicate efficiently in infected hosts producing numerous copies of the viral RNA. One of the long-standing mysteries in RNA virus replication is the occurrence and possible role of the double-stranded (ds)RNA formed between minus- and plus-strands. Using the partially purified Cucumber necrosis virus (CNV) replicase from plants and the recombinant RNA-dependent RNA polymerase (RdRp) of Turnip crinkle virus (TCV), in this paper, we demonstrate that both CNV replicase and the related TCV RdRp can utilize dsRNA templates to produce viral plus-stranded RNA in vitro. Sequence and structure of the dsRNA around the plus-strand initiation site had a significant effect on initiation, suggesting that initiation on dsRNA templates is a rate-limiting step. In contrast, the CNV replicase could efficiently synthesize plus-strand RNA on partial dsRNAs that had the plus-strand initiation promoter "exposed", suggesting that the polymerase activity of CNV replicase is strong enough to unwind extended dsRNA regions in the template during RNA synthesis. Based on the in vitro data, we propose that dsRNA forms might have functional roles during tombus- and carmovirus replication and the AU-rich nature of the terminus could be important for opening the dsRNA structure around the plus-strand initiation promoter for tombus- and carmoviruses and possibly many other positive-strand RNA viruses.  相似文献   

4.
Jaag HM  Stork J  Nagy PD 《Virology》2007,368(2):388-404
Previous genome-wide screens identified over 100 host genes whose deletion/down-regulation affected tombusvirus replication and 32 host genes that affected tombusvirus RNA recombination in yeast, a model host for replication of Tomato bushy stunt virus (TBSV). Down-regulation of several of the identified host genes affected the accumulation levels of p33 and p92(pol) replication proteins, raising the possibility that these host factors could be involved in the regulation of the amount of viral replication proteins and, thus, they are indirectly involved in TBSV replication and recombination. To test this model, we developed a tightly regulated expression system for recombinant p33 and p92(pol) replication proteins in yeast. We demonstrate that high accumulation level of p33 facilitated efficient viral RNA replication, while the effect of p33 level on RNA recombination was less pronounced. On the other hand, high level of p92(pol) accumulation promoted TBSV RNA recombination more efficiently than RNA replication. As predicted, Rpb11p, which is part of the polII complex, affected the accumulation levels of p33 and p92(pol) as well as altered RNA replication and recombination. An in vitro assay with the tombusvirus replicase further supported that Rpb11p affects TBSV replication and recombination only indirectly, via regulating p33 and p92(pol) levels. In contrast, the mechanism by which Rpt4p endopeptidase/ATPase and Mps1p threonine/tyrosine kinase affect TBSV recombination is different from that proposed for Rpb11p. We propose a model that the concentration (molecular crowding) of replication proteins within the viral replicase is a factor affecting viral replication and recombination.  相似文献   

5.
Panaviene Z  Baker JM  Nagy PD 《Virology》2003,308(1):191-205
Two of the five viral-coded proteins of tombusviruses, which are small, nonsegmented, plus-stranded RNA viruses of plants, are required for replication in infected cells. These replicase proteins, namely, p33 and p92, of cucumber necrosis virus are expressed directly from the genomic RNA via a readthrough mechanism. Their overlapping domains contain an arginine/proline-rich RNA-binding motif (termed RPR, which has the sequence RPRRRP). Site-directed mutagenesis of p33 expressed in Escherichia coli, followed by a gel shift assay, defined two of the four arginines as required for efficient RNA binding in vitro. In vivo testing of 19 RPR motif mutants revealed that the RPR motif, and therefore the ability to bind RNA, is important for the replication of tombusviruses and their associated defective interfering (DI) RNAs. Mutation within the RPR motif also affected the ratio of subgenomic versus genomic RNAs in infected cells. To test whether the RPR motif is essential for the function of either p33 or p92 in replication, we used a two-component system developed by, J. Virol. 5845-5851), in which p92 was expressed from the genomic RNA of a tombusvirus, while p33 was expressed from a DI RNA. The protoplast experiments with the two-component system revealed that the RPR motif is essential for the replication function of both proteins. Interestingly, mutations within the RPR motif of p33 and p92 had different effects on RNA replication, suggesting different roles for the RNA-binding motifs of these proteins in tombusvirus replication.  相似文献   

6.
7.
Cheng CP  Panavas T  Luo G  Nagy PD 《Virology》2005,341(1):107-121
The viral RNA plays multiple roles during replication of RNA viruses, serving as a template for complementary RNA synthesis and facilitating the assembly of the viral replicase complex. These roles are coordinated by cis-acting regulatory elements, such as promoters and replication enhancers (REN). To test if these RNA elements can be used by related viral RNA-dependent RNA polymerases (RdRp), we compared the potential stimulatory effects of homologous and heterologous REN elements on complementary RNA synthesis and template-switching by the tombus- (Cucumber necrosis virus, CNV), carmovirus (Turnip crinkle virus, TCV) and hepatitis C virus (HCV) RdRps in vitro. The CNV RdRp selectively utilized its cognate REN, while discriminating against the heterologous TCV REN. On the contrary, RNA synthesis by the TCV RdRp was stimulated by the TCV REN and the heterologous tombusvirus REN with comparable efficiency. The heterologous REN elements also promoted in vitro template-switching by the TCV and HCV RdRps. Based on these observations, we propose that REN elements could facilitate intervirus recombination and post-recombinational amplification of new recombinant viruses.  相似文献   

8.
9.
Stork J  Panaviene Z  Nagy PD 《Virology》2005,343(1):79-92
Tombusviruses, which are small plus-strand RNA viruses of plants, require the viral-coded p33 replication co-factor for template selection and recruitment into replication in infected cells. As presented in the accompanying paper [Shapka, N., Stork, J., Nagy, P.D., 2005. Phosphorylation of the p33 replication protein of Cucumber necrosis tombusvirus adjacent to the RNA binding site affects viral RNA replication. J. Virol. 343, 65-78.], p33 can be phosphorylated in vitro at serine and threonine residues adjacent to its arginine-proline-rich RNA binding site. To test the effect of phosphorylation on p33 function, in this paper, we used phosphorylation-mimicking aspartic acid mutants of Cucumber necrosis virus (CNV) p33 and in-vitro-phosphorylated p33 in gel mobility shift experiments. We found that phosphorylation inhibited the ability of p33 to bind to the viral RNA. In contrast, the nonphosphorylation-mimicking alanine mutants of p33 bound to viral RNA as efficiently as the nonphosphorylated wild type p33 did. In vitro assays with purified CNV replicase preparations revealed that phosphorylation-mimicking mutants of p33 did not support the assembly of functional CNV replicase complexes in yeast, a model host. Based on these results, we propose that the primary function of reversible phosphorylation of p33 is to regulate the RNA binding capacity of p33, which could affect the assembly of new viral replicase complexes, recruitment of the viral RNA template into replication and/or release of viral RNA from replication. Thus, phosphorylation of p33 might help in switching the role of the viral RNA from replication to other processes, such as viral RNA encapsidation and cell-to-cell movement in infected hosts.  相似文献   

10.
Akira Mine  Masanori Kaido 《Virology》2010,407(2):213-224
Red clover necrotic mosaic virus (RCNMV), a positive-sense RNA virus with a bipartite genome, encodes p27 and p88 replicase proteins that are required for viral RNA replication and suppression of RNA silencing. In this study, we indentified domains in p27 and p88 responsible for their protein-protein interactions using in vitro pull-down assays with the purified recombinant proteins. Coimmunoprecipitation analysis in combination with blue-native polyacrylamide gel electrophoresis using mutated p27 proteins showed that both p27-p27 and p27-p88 interactions are essential for the formation of the 480-kDa complex, which has RCNMV-specific RNA-dependent RNA polymerase activity. Furthermore, we found a good correlation between the accumulated levels of the 480-kDa complex and replication levels and the suppression of RNA silencing activity. Our results indicate that interactions between RCNMV replicase proteins play an essential role in viral RNA replication and in suppressing RNA silencing via the 480-kDa replicase complex assembly.  相似文献   

11.
Hyodo K  Mine A  Iwakawa HO  Kaido M  Mise K  Okuno T 《Virology》2011,413(2):300-309
The specific recognition of genomic RNAs by viral replicase proteins is a key regulatory step during the early replication process in positive-strand RNA viruses. In this study, we characterized the RNA-binding activity of the auxiliary replicase protein p27 of Red clover necrotic mosaic virus (RCNMV), which has a bipartite genome consisting of RNA1 and RNA2. Aptamer pull-down assays identified the amino acid residues of p27 involved in its specific interaction with RNA2. The RNA-binding activity of p27 correlated with its activity in recruiting RNA2 to membranes. We also identified the amino acids required for the formation of the 480-kDa replicase complex, a key player of RCNMV RNA replication. These amino acids are not involved in the functions of p27 that bind viral RNA or replicase proteins, suggesting an additional role for p27 in the assembly of the replicase complex. Our results demonstrate that p27 has multiple functions in RCNMV replication.  相似文献   

12.
Yi Jiang  Elena Serviene 《Virology》2010,404(1):96-105
Maintenance of genome integrity is of major importance for plus-stranded RNA viruses that are vulnerable to degradation by host ribonucleases or to replicase errors. We demonstrate that short truncations at the 5′ end of a model Tomato bushy stunt virus (TBSV) RNA could be repaired during replication in yeast and plant cells. Although the truncations led to the loss of important cis-regulatory elements, the genome repair mechanisms led to the recovery of promoter and enhancer-like sequences in 92% of TBSV progeny. Using in vitro approaches, we demonstrate that the repaired TBSV RNAs are replication-competent. We propose three different mechanisms for genome repair: initiation of RNA synthesis from internal sequences and addition of nonviral nucleotides by the tombusvirus replicase; and via RNA recombination. The ability to repair cis-sequences makes the tombusvirus genome more flexible, which could be beneficial to increase the virus fitness and adaptation to new hosts.  相似文献   

13.
14.
Daniel Barajas 《Virology》2010,397(2):358-368
Post-translational modifications of viral replication proteins could be widespread phenomena during the replication of plus-stranded RNA viruses. In this article, we identify two lysines in the tombusvirus p33 replication co-factor involved in ubiquitination and show that the same lysines are also important for the p33 to interact with the host Vps23p ESCRT-I factor. We find that the interaction of p33 with Vps23p is also affected by a “late-domain”-like sequence in p33. The combined mutations of the two lysines and the late-domain-like sequences in p33 reduced replication of a replicon RNA of Tomato bushy stunt virus in yeast model host, in plant protoplasts, and plant leaves, suggesting that p33-Vps23p ESCRT protein interaction affects tombusvirus replication. Using ubiquitin-mimicking p33 chimeras, we demonstrate that high level of p33 ubiquitination is inhibitory for TBSV replication. These findings argue that optimal level of p33 ubiquitination plays a regulatory role during tombusvirus infections.  相似文献   

15.
N Kovalev  D Barajas  PD Nagy 《Virology》2012,432(2):470-484
Recruited host factors aid replication of plus-strand RNA viruses. In this paper, we show that Dbp2 DEAD-box helicase of yeast, which is a homolog of human p68 DEAD-box helicase, directly affects replication of Tomato bushy stunt virus (TBSV). We demonstrate that Dbp2 binds to the 3′-end of the viral minus-stranded RNA and enhances plus-strand synthesis by the viral replicase in a yeast-based cell-free TBSV replication assay. In vitro data with wt and an ATPase-deficient Dbp2 mutant indicate that Dbp2 unwinds local secondary structures at the 3′-end of the TBSV (−)RNA. We also show that Dbp2 complements the replication deficiency of TBSV in yeast containing reduced amount of Ded1 DEAD-box helicase, another host factor involved in TBSV replication, suggesting that Dbp2 and Ded1 helicases play redundant roles in TBSV replication. We also show that the orthologous AtRH20 DEAD-box helicase from Arabidopsis can increase tombusvirus replication in vitro and in yeast.  相似文献   

16.
Many (+)-strand RNA viruses use subgenomic (SG) RNAs as messengers for protein expression, or to regulate their viral life cycle. Three different mechanisms have been described for the synthesis of SG RNAs. The first mechanism involves internal initiation on a (−)-strand RNA template and requires an internal SGP promoter. The second mechanism makes a prematurely terminated (−)-strand RNA which is used as template to make the SG RNA. The third mechanism uses discontinuous RNA synthesis while making the (−)-strand RNA templates. Most SG RNAs are translated into structural proteins or proteins related to pathogenesis: however other SG RNAs regulate the transition between translation and replication, function as riboregulators of replication or translation, or support RNA-RNA recombination. In this review we discuss these functions of SG RNAs and how they influence viral replication, translation and recombination.  相似文献   

17.
Yi Jiang 《Virology》2010,396(1):10-20
Previous genome-wide screens identified > 100 host genes affecting tombusvirus replication using yeast model host. One of those factors was Nsr1p (nucleolin), which is an abundant RNA-binding shuttle protein involved in rRNA maturation and ribosome assembly. We find that overexpression of Nsr1p in yeast or in Nicotiana benthamiana inhibited the accumulation of tombusvirus RNA by ∼10-fold. Regulated overexpression of Nsr1p revealed that Nsr1p should be present at the beginning of viral replication for efficient inhibition, suggesting that Nsr1p inhibits an early step in the replication process. In vitro experiments revealed that Nsr1p binds preferably to the 3′ UTR in the viral RNA. The purified recombinant Nsr1p inhibited the in vitro replication of the viral RNA in a yeast cell-free assay when preincubated with the viral RNA before the assay. These data support the model that Nsr1p/nucleolin inhibits tombusvirus replication by interfering with the recruitment of the viral RNA for replication.  相似文献   

18.
Pathak KB  Sasvari Z  Nagy PD 《Virology》2008,379(2):294-305
Replication of Tomato bushy stunt virus (TBSV) RNA takes place on the cytosolic membrane surface of peroxisomes in plants and in yeast, a model host. To identify the host proteins involved in assisting the peroxisomal localization of the tombusvirus p33 replication protein, we tested if p33 could bind directly to yeast proteins involved in peroxisomal transport in vitro. This work has led to the demonstration of Pex19p-p33 interaction via pull-down and co-purification experiments. Pex19p was also detected in the tombusvirus replicase after protein cross-linking, suggesting that Pex19p transiently binds to the replicase as could be expected from a transporter. To validate the importance of Pex19p-p33 interaction in TBSV replication in yeast, we re-targeted Pex19p to the mitochondria, which resulted in the re-distribution of a large fraction of p33 to the mitochondria. The expression of the mitochondrial-targeted Pex19p inhibited TBSV RNA accumulation by 2-4-fold in vivo and reduced the in vitro activity of the tombusvirus replicase by 80%. These data support the model that Pex19p is a cellular transporter for localization of p33 replication protein to the host peroxisomal membranes.  相似文献   

19.
The Sindbis virus RNA-dependent RNA polymerase (nsP4) is responsible for the replication of the viral RNA genome. In infected cells, nsP4 is localized in a replication complex along with the other viral non-structural proteins. nsP4 has been difficult to homogenously purify from infected cells due to its interactions with the other replication proteins and the fact that its N-terminal residue, a tyrosine, causes the protein to be rapidly turned over in cells. We report the successful expression and purification of Sindbis nsP4 in a bacterial system, in which nsP4 is expressed as an N-terminal SUMO fusion protein. After purification the SUMO tag is removed, resulting in the isolation of full-length nsP4 possessing the authentic N-terminal tyrosine. This purified enzyme is able to produce minus-strand RNA de novo from plus-strand templates, as well as terminally add adenosine residues to the 3′ end of an RNA substrate. In the presence of the partially processed viral replicase polyprotein, P123, purified nsP4 is able to synthesize discrete template length minus-strand RNA products. Mutations in the 3′ CSE or poly(A) tail of viral template RNA prevent RNA synthesis by the replicase complex containing purified nsP4, consistent with previously reported template requirements for minus-strand RNA synthesis. Optimal reaction conditions were determined by investigating the effects of time, pH, and the concentrations of nsP4, P123 and magnesium on the synthesis of RNA.  相似文献   

20.
Kim KH  Rümenapf T  Strauss EG  Strauss JH 《Virology》2004,323(1):153-163
Alphavirus nonstructural proteins are translated as a polyprotein that is ultimately cleaved into four mature proteins called nsP1, nsP2, nsP3, and nsP4 from their order in the polyprotein. The role of this nonstructural polyprotein, of cleavage intermediates, and of mature proteins in synthesis of Semliki Forest virus (SFV) RNA has been studied using mutants unable to cleave one or more of the sites in the nonstructural polyprotein or that had the arginine sense codon between nsP3 and nsP4 changed to an opal termination codon. The results were compared with those obtained for Sindbis virus (SINV), which has a naturally occurring opal codon between nsP2 and nsP3. We found that (1) an active nonstructural protease in nsP2 is required for RNA synthesis. This protease is responsible for all three cleavages in the nonstructural polyprotein. (2) Cleavage between nsP3 and nsP4 (the viral RNA polymerase) is required for RNA synthesis by SFV. (3) SFV mutants that are able to produce only polyprotein P123 and nsP4 synthesize minus-strand RNA early after infection as efficiently as SF wild type but are defective in the synthesis of plus-strand RNA. The presence of sense or opal following nsP3 did not affect this result. At 30 degrees C, they give rise to low yields of virus after a delay, but at 39 degrees C, they are nonviable. (4) SFV mutants that produce nsP1, P23, nsP4, as well as the precursor P123 are viable but produce an order of magnitude less virus than wild type at 30 degrees C and two orders of magnitude less virus at 39 degrees C. The ratio of subgenomic mRNA to genomic RNA is much reduced in these mutants relative to the parental viruses. (5) At 30 degrees C, the variants containing an opal codon grow as well as or slightly better than the corresponding virus with a sense codon. At 39 degrees C, however, the opal variants produce significantly more virus. These results support the conclusion that SFV and SINV, and by extension all alphaviruses, regulate their RNA synthesis in the same fashion after infection. P123 and nsP4 form a minus-strand replicase that synthesizes plus-strand RNA only inefficiently, especially at the higher temperatures found in mammals and birds. A replicase containing nsP1, P23, and nsP4 can make both plus and minus strands, but prefers the promoter for genomic plus sense RNA to that for subgenomic mRNA. The fully cleaved replicase can make only plus-strand RNA, and prefers the promoter for subgenomic mRNA to that for genomic RNA. Alphaviruses alternate between infection of hematophagous arthropods and higher vertebrates. Although the infection of higher vertebrates is acute and often accompanied by disease, continuing transmission of the virus in nature requires that infection of arthropods be persistent and relatively asymptomatic. We propose that this mechanism for control of RNA synthesis evolved to moderate the pathogenicity of the viruses in their arthropod hosts.  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