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 140 毫秒
1.
Brome mosaic virus (BMV) encodes two RNA replication proteins: 1a, which contains RNA capping and helicase-like domains, and 2a, which is related to polymerases. BMV 1a and 2a can direct virus-specific RNA replication in the yeast Saccharomyces cerevisiae, which reproduces the known features of BMV replication in plant cells. We constructed single amino acid point mutations at the predicted capping and helicase active sites of 1a and analyzed their effects on BMV RNA3 replication in yeast. The helicase mutants showed no function in any assays used: they were strongly defective in template recruitment for RNA replication, as measured by 1a-induced stabilization of RNA3, and they synthesized no detectable negative-strand or subgenomic RNA. Capping domain mutants divided into two groups. The first exhibited increased template recruitment but nevertheless allowed only low levels of negative-strand and subgenomic mRNA synthesis. The second was strongly defective in template recruitment, made very low levels of negative strands, and made no detectable subgenomes. To distinguish between RNA synthesis and capping defects, we deleted chromosomal gene XRN1, encoding the major exonuclease that degrades uncapped mRNAs. XRN1 deletion suppressed the second but not the first group of capping mutants, allowing synthesis and accumulation of large amounts of uncapped subgenomic mRNAs, thus providing direct evidence for the importance of the viral RNA capping function. The helicase and capping enzyme mutants showed no complementation. Instead, at high levels of expression, a helicase mutant dominantly interfered with the function of the wild-type protein. These results are discussed in relation to the interconnected functions required for different steps of positive-strand RNA virus replication.  相似文献   

2.
Brome mosaic virus (BMV), a positive-strand RNA virus, encodes two replication proteins: the 2a protein, which contains polymerase-like sequences, and the 1a protein, with N-terminal putative capping and C-terminal helicase-like sequences. These two proteins are part of a multisubunit complex which is necessary for viral RNA replication. We have previously shown that the yeast two-hybrid assay consistently duplicated results obtained from in vivo RNA replication assays and biochemical assays of protein-protein interaction, thus permitting the identification of additional interacting domains. We now map an interaction found to take place between two 1a proteins. Using previously characterized 1a mutants, a perfect correlation was found between the in vivo phenotypes of these mutants and their abilities to interact with wild-type 1a (wt1a) and each other. Western blot analysis revealed that the stabilities of many of the noninteracting mutant proteins were similar to that of wt1a. Deletion analysis of 1a revealed that the N-terminal 515 residues of the 1a protein are required and sufficient for 1a-1a interaction. This intermolecular interaction between the putative capping domain and itself was detected in another tripartite RNA virus, cucumber mosaic virus (CMV), suggesting that the 1a-1a interaction is a feature necessary for the replication of tripartite RNA viruses. The boundaries for various activities are placed in the context of the predicted secondary structures of several 1a-like proteins of members of the alphavirus-like superfamily. Additionally, we found a novel interaction between the putative capping and helicase-like portions of the BMV and CMV 1a proteins. Our cumulative data suggest a working model for the assembly of the BMV RNA replicase.  相似文献   

3.
4.
The cis-acting replication element (CRE) is a 61-nucleotide stem-loop RNA structure found within the coding sequence of poliovirus protein 2C. Although the CRE is required for viral RNA replication, its precise role(s) in negative- and positive-strand RNA synthesis has not been defined. Adenosine in the loop of the CRE RNA structure functions as the template for the uridylylation of the viral protein VPg. VPgpUpU(OH), the predominant product of CRE-dependent VPg uridylylation, is a putative primer for the poliovirus RNA-dependent RNA polymerase. By examining the sequential synthesis of negative- and positive-strand RNAs within preinitiation RNA replication complexes, we found that mutations that disrupt the structure of the CRE prevent VPg uridylylation and positive-strand RNA synthesis. The CRE mutations that inhibited the synthesis of VPgpUpU(OH), however, did not inhibit negative-strand RNA synthesis. A Y3F mutation in VPg inhibited both VPgpUpU(OH) synthesis and negative-strand RNA synthesis, confirming the critical role of the tyrosine hydroxyl of VPg in VPg uridylylation and negative-strand RNA synthesis. trans-replication experiments demonstrated that the CRE and VPgpUpU(OH) were not required in cis or in trans for poliovirus negative-strand RNA synthesis. Because these results are inconsistent with existing models of poliovirus RNA replication, we propose a new four-step model that explains the roles of VPg, the CRE, and VPgpUpU(OH) in the asymmetric replication of poliovirus RNA.  相似文献   

5.
The central portion of the brome mosaic virus (BMV) 2a protein represents the most conserved element among the related RNA replication components of a large group of positive-strand RNA viruses of humans, animals, and plants. To characterize the functions of the 2a protein, mutations were targeted to a conserved portion of the 2a gene, resulting in substitutions between amino acids 451 and 484. After the temperature profile of wild-type BMV RNA replication was defined, RNA replication by nine selected mutants was tested in barley protoplasts at permissive (24 degrees C) and nonpermissive (34 degrees C) temperatures. Four mutants did not direct RNA synthesis at either temperature. Various levels of temperature-sensitive (ts) replication occurred in the remaining five mutants. For two ts mutants, no viral RNA synthesis was detected at 34 degrees C, while for two others, an equivalent reduction in positive- and negative-strand RNA accumulation was observed. For one mutant, positive-strand accumulation was preferentially reduced over negative-strand accumulation at 34 degrees C. Moreover, this mutant and another displayed preferential suppression of genomic over subgenomic RNA accumulation at both 24 and 34 degrees C. The combination of phenotypes observed suggests that the 2a protein may play a role in the differential initiation of specific classes of viral RNA in addition to a previously suggested role in RNA elongation.  相似文献   

6.
We report that protein 2C, the putative nucleoside triphosphatase/helicase protein of poliovirus, is required for the initiation of negative-strand RNA synthesis. Preinitiation RNA replication complexes formed upon the translation of poliovirion RNA in HeLa S10 extracts containing 2 mM guanidine HCI, a reversible inhibitor of viral protein 2C. Upon incubation in reactions lacking guanidine, preinitiation RNA replication complexes synchronously initiated and elongated negative-strand RNA molecules, followed by the synchronous initiation and elongation of positive-strand RNA molecules. The immediate and exclusive synthesis of negative-strand RNA upon the removal of guanidine demonstrates that guanidine specifically blocks the initiation of negative-strand RNA synthesis. Readdition of guanidine HCl to reactions synchronously elongating nascent negative-strand RNA molecules did not prevent their continued elongation and completion. In fact, readdition of guanidine HCl to reactions containing preinitiation complexes elongating nascent negative-strand RNA molecules had no effect on subsequent positive-strand RNA synthesis initiation or elongation. Thus, the guanidine-inhibited function of viral protein 2C was not required for the elongation of negative-strand RNA molecules, the initiation of positive-strand RNA molecules, or the elongation of positive-strand RNA molecules. The guanidine-inhibited function of viral protein 2C is required only immediately before or during the initiation of negative-strand RNA synthesis. We suggest that guanidine may block an irreversible structural maturation of protein 2C and/or RNA replication complexes necessary for the initiation of RNA replication.  相似文献   

7.
8.
The replication of positive-strand RNA viruses involves not only viral proteins but also multiple cellular proteins and intracellular membranes. In both plant cells and the yeast Saccharomyces cerevisiae, brome mosaic virus (BMV), a member of the alphavirus-like superfamily, replicates its RNA in endoplasmic reticulum (ER)-associated complexes containing viral 1a and 2a proteins. Prior to negative-strand RNA synthesis, 1a localizes to ER membranes and recruits both positive-strand BMV RNA templates and the polymerase-like 2a protein to ER membranes. Here, we show that BMV RNA replication in S. cerevisiae is markedly inhibited by a mutation in the host YDJ1 gene, which encodes a chaperone Ydj1p related to Escherichia coli DnaJ. In the ydj1 mutant, negative-strand RNA accumulation was inhibited even though 1a protein associated with membranes and the positive-strand RNA3 replication template and 2a protein were recruited to membranes as in wild-type cells. In addition, we found that in ydj1 mutant cells but not wild-type cells, a fraction of 2a protein accumulated in a membrane-free but insoluble, rapidly sedimenting form. These and other results show that Ydj1p is involved in forming BMV replication complexes active in negative-strand RNA synthesis and suggest that a chaperone system involving Ydj1p participates in 2a protein folding or assembly into the active replication complex.  相似文献   

9.
One of the distinguishing features of the alphaviruses is a sequential processing of the nonstructural polyproteins P1234 and P123. In the early stages of the infection, the complex of P123+nsP4 forms the primary replication complexes (RCs) that function in negative-strand RNA synthesis. The following processing steps make nsP1+P23+nsP4, and later nsP1+nsP2+nsP3+nsP4. The latter mature complex is active in positive-strand RNA synthesis but can no longer produce negative strands. However, the regulation of negative- and positive-strand RNA synthesis apparently is not the only function of ns polyprotein processing. In this study, we developed Sindbis virus mutants that were incapable of either P23 or P123 cleavage. Both mutants replicated in BHK-21 cells to levels comparable to those of the cleavage-competent virus. They continuously produced negative-strand RNA, but its synthesis was blocked by the translation inhibitor cycloheximide. Thus, after negative-strand synthesis, the ns proteins appeared to irreversibly change conformation and formed mature RCs, in spite of the lack of ns polyprotein cleavage. However, in the cells having no defects in alpha/beta interferon (IFN-alpha/beta) production and signaling, the cleavage-deficient viruses induced a high level of type I IFN and were incapable of causing the spread of infection. Moreover, the P123-cleavage-deficient virus was readily eliminated, even from the already infected cells. We speculate that this inability of the viruses with unprocessed polyprotein to productively replicate in the IFN-competent cells and in the cells of mosquito origin was an additional, important factor in ns polyprotein cleavage development. In the case of the Old World alphaviruses, it leads to the release of nsP2 protein, which plays a critical role in inhibiting the cellular antiviral response.  相似文献   

10.
We had previously demonstrated that a cellular protein specifically interacts with the 3' end of poliovirus negative-strand RNA. We now report the identity of this protein as heterogeneous nuclear ribonucleoprotein (hnRNP) C1/C2. Formation of an RNP complex with poliovirus RNA was severely impaired by substitution of a lysine, highly conserved among vertebrates, with glutamine in the RNA recognition motif (RRM) of recombinant hnRNP C1, suggesting that the binding is mediated by the RRM in the protein. We have also shown that in a glutathione S-transferase (GST) pull-down assay, GST/hnRNP C1 binds to poliovirus polypeptide 3CD, a precursor to the viral RNA-dependent RNA polymerase, 3D(pol), as well as to P2 and P3, precursors to the nonstructural proteins. Truncation of the auxiliary domain in hnRNP C1 (C1DeltaC) diminished these protein-protein interactions. When GST/hnRNP C1DeltaC was added to in vitro replication reactions, a significant reduction in RNA synthesis was observed in contrast to reactions supplemented with wild-type fusion protein. Indirect functional depletion of hnRNP C from in vitro replication reactions, using poliovirus negative-strand cloverleaf RNA, led to a decrease in RNA synthesis. The addition of GST/hnRNP C1 to the reactions rescued RNA synthesis to near mock-depleted levels. Furthermore, we demonstrated that poliovirus positive-strand and negative-strand RNA present in cytoplasmic extracts prepared from infected HeLa cells coimmunoprecipitated with hnRNP C1/C2. Our findings suggest that hnRNP C1 has a role in positive-strand RNA synthesis in poliovirus-infected cells, possibly at the level of initiation.  相似文献   

11.
Pokeweed antiviral protein (PAP) is a glycosidase of plant origin that has been shown to depurinate some viral RNAs in vitro. We have demonstrated previously that treatment of Brome mosaic virus (BMV) RNAs with PAP inhibited their translation in a cell-free system and decreased their accumulation in barley protoplasts. In the current study, we map the depurination sites on BMV RNA3 and describe the mechanism by which replication of the viral RNA is inhibited by depurination. Specifically, we demonstrate that the viral replicase exhibited reduced affinity for depurinated positive-strand RNA3 compared with intact RNA3, resulting in less negative-strand product. This decrease was due to depurination within the intergenic region of RNA3, between ORF3 and 4, and distant from the 3′ terminal core promoter required for initiation of negative-strand RNA synthesis. Depurination within the intergenic region alone inhibited the binding of the replicase to full-length RNA3, whereas depurination outside the intergenic region permitted the replicase to initiate negative-strand synthesis; however, elongation of the RNA product was stalled at the abasic nucleotide. These results support a role of the intergenic region in controlling negative-strand RNA synthesis and contribute new insight into the effect of depurination by PAP on BMV replication.  相似文献   

12.
13.
Vlot AC  Laros SM  Bol JF 《Journal of virology》2003,77(20):10790-10798
RNAs 1 and 2 of the tripartite genome of alfalfa mosaic virus encode the replicase proteins P1 and P2, respectively, whereas RNA 3 encodes the movement protein and coat protein. Transient expression of wild-type (wt) and mutant viral RNAs and proteins by agroinfiltration of plant leaves was used to study cis- and trans-acting functions of the helicase-like domain in P1 and the polymerase-like domain in P2. Three mutations in conserved motifs of the helicase-like domain of P1 affected one or more steps leading to synthesis of minus-strand RNAs 1, 2, and 3. In leaves containing transiently expressed P1 and P2, replication of wt but not mutant RNA 1 was observed. Apparently, the transiently expressed P1 could not complement the defect in replication of the RNA 1 mutant. Moreover, the transiently expressed wt replicase supported replication of RNA 2, but this replication was blocked in trans by coexpression of mutant RNA 1. However, expression of mutant RNA 1 did not interfere with the replication of RNA 3 by the wt replicase. Similarly, a mutation in the GDD motif encoded by RNA 2 could not be complemented in trans and affected the replication of RNA 1 by a wt replicase, while replication of RNA 3 remained unaffected. In competition assays, the transient wt replicase preferentially replicated RNA 3 over RNAs 1 and 2. The results indicate that one or more functions of P1 and P2 act in cis and point to the existence of a mechanism that coordinates the replication of RNAs 1 and 2.  相似文献   

14.
15.
During the past year, genetic studies of picornaviruses, vastly facilitated by the application of infectious picornaviral cDNAs and RNAs, have contributed to our understanding of the function of individual picornavirus polypeptides and to the genetic processes that operate in these small RNA viruses. Especially notable were the demonstrations that the RNA-dependent RNA polymerase may have a function in RNA synthesis as an uncleaved precursor polypeptide, and that a mutation in the polymerase can be complemented in trans, in contrast to data obtained from previously studied polymerase mutants. A new in vitro system, in which positive-strand synthesis, negative-strand synthesis and RNA packaging were all observed, will facilitate further studies into the mechanism of these processes.  相似文献   

16.
A defining feature of alfalfa mosaic virus (AMV) and ilarviruses [type virus: tobacco streak virus (TSV)] is that, in addition to genomic RNAs, viral coat protein is required to establish infection in plants. AMV and TSV coat proteins, which share little primary amino acid sequence identity, are functionally interchangeable in RNA binding and initiation of infection. The lysine-rich amino-terminal RNA binding domain of the AMV coat protein lacks previously identified RNA binding motifs. Here, the AMV coat protein RNA binding domain is shown to contain a single arginine whose specific side chain and position are crucial for RNA binding. In addition, the putative RNA binding domain of two ilarvirus coat proteins, TSV and citrus variegation virus, is identified and also shown to contain a crucial arginine. AMV and ilarvirus coat protein sequence alignment centering on the key arginine revealed a new RNA binding consensus sequence. This consensus may explain in part why heterologous viral RNA-coat protein mixtures are infectious.  相似文献   

17.
Smith RM  Walton CM  Wu CH  Wu GY 《Journal of virology》2002,76(19):9563-9574
The 3'-terminal sequences of hepatitis C virus (HCV) positive- and negative-strand RNAs contribute cis-acting functions essential for viral replication. The secondary structure and protein-binding properties of these highly conserved regions are of interest not only for the further elucidation of HCV molecular biology, but also for the design of antisense therapeutic constructs. The RNA structure of the positive-strand 3' untranslated region has been shown previously to influence binding by various host and viral proteins and is thus thought to promote HCV RNA synthesis and genome stability. Recent studies have attributed analogous functions to the negative-strand 3' terminus. We evaluated the HCV negative-strand secondary structure by enzymatic probing with single-strand-specific RNases and thermodynamic modeling of RNA folding. The accessibility of both 3'-terminal sequences to hybridization by antisense constructs was evaluated by RNase H cleavage mapping in the presence of combinatorial oligodeoxynucleotide libraries. The mapping results facilitated identification of antisense oligodeoxynucleotides and a 10-23 deoxyribozyme active against the positive-strand 3'-X region RNA in vitro.  相似文献   

18.
C C Kao  P Ahlquist 《Journal of virology》1992,66(12):7293-7302
Brome mosaic virus is a positive-strand RNA virus whose RNA replication requires viral protein 1a, which has putative helicase and capping functions, and 2a, which has putative polymerase function. Since domains of related sequence are conserved in a wide range of plus-strand RNA viruses, analysis of 1a and 2a function should have applicability to many other viruses. We have recently demonstrated that 1a and 2a form a complex in vivo and in vitro. Using immune coprecipitation and mutant polypeptides made in reticulocyte lysates, we have now mapped both the 1a and 2a domains necessary for complex formation. The sequences needed to bind 2a map to the carboxy-terminal helicase-like domain of 1a. Truncated polypeptides containing this domain were able to bind to 2a, while several small insertions in the helicase-like domain disrupted binding. The sequence required for binding 1a lies within a 115-residue subset of the 2a N-terminal segment preceding the polymerase-like domain. Truncations or fusion polypeptides containing this segment can bind 1a. We also determined that highly purified 2a protein made in insect cells can form a complex with highly purified 1a helicase-like domain made in Escherichia coli, suggesting that no other factor is required to mediate 1a-2a interaction. Previous genetic analyses of 1a and 2a are consistent with this mapping and show that the newly defined 1a and 2a binding regions are required for RNA synthesis. The locations of these interacting regions are discussed with regard to models of viral replication and the evolution of positive-strand RNA virus genomes.  相似文献   

19.
Steil BP  Barton DJ 《Journal of virology》2008,82(19):9400-9408
Initiation of RNA synthesis by RNA-dependent RNA polymerases occurs when a phosphodiester bond is formed between the first two nucleotides in the 5′ terminus of product RNA. The concentration of initiating nucleoside triphosphates (NTPi) required for RNA synthesis is typically greater than the concentration of NTPs required for elongation. VPg, a small viral protein, is covalently attached to the 5′ end of picornavirus negative- and positive-strand RNAs. A cis-acting replication element (CRE) within picornavirus RNAs serves as a template for the uridylylation of VPg, resulting in the synthesis of VPgpUpUOH. Mutations within the CRE RNA structure prevent VPg uridylylation. While the tyrosine hydroxyl of VPg can prime negative-strand RNA synthesis in a CRE- and VPgpUpUOH-independent manner, CRE-dependent VPgpUpUOH synthesis is absolutely required for positive-strand RNA synthesis. As reported herein, low concentrations of UTP did not support negative-strand RNA synthesis when CRE-disrupting mutations prevented VPg uridylylation, whereas correspondingly low concentrations of CTP or GTP had no negative effects on the magnitude of CRE-independent negative-strand RNA synthesis. The experimental data indicate that CRE-dependent VPg uridylylation lowers the Km of UTP required for viral RNA replication and that CRE-dependent VPgpUpUOH synthesis was required for efficient negative-strand RNA synthesis, especially when UTP concentrations were limiting. By lowering the concentration of UTP needed for the initiation of RNA replication, CRE-dependent VPg uridylylation provides a mechanism for a more robust initiation of RNA replication.  相似文献   

20.
T Furuya  M M Lai 《Journal of virology》1993,67(12):7215-7222
The termini of viral genomic RNA and its complementary strand are important in the initiation of viral RNA replication, which probably involves both viral and cellular proteins. To detect the possible cellular proteins involved in the replication of mouse hepatitis virus RNA, we performed RNA-protein binding studies with RNAs representing both the 5' and 3' ends of the viral genomic RNA and the 3' end of the negative-strand complementary RNA. Gel-retardation assays showed that both the 5'-end-positive- and 3'-end-negative-strand RNA formed an RNA-protein complex with cellular proteins from the uninfected cells. UV cross-linking experiments further identified a 55-kDa protein bound to the 5' end of the positive-strand viral genomic RNA and two proteins 35 and 38 kDa in size bound to the 3' end of the negative-strand cRNA. The results of the competition assay confirmed the specificity of this RNA-protein binding. No proteins were found to bind to the 3' end of the viral genomic RNA under the same conditions. The binding site of the 55-kDa protein was mapped within the 56-nucleotide region from nucleotides 56 to 112 from the 5' end of the positive-strand RNA, and the 35- and 38-kDa proteins bound to the complementary region on the negative-strand RNA. The 38-kDa protein was detected only in DBT cells but was not detected in HeLa or COS cells, while the 35-kDa protein was found in all three cell types. The juxtaposition of the different cellular proteins on the complementary sites near the ends of the positive- and negative-strand RNAs suggests that these proteins may interact with each other and play a role in mouse hepatitis virus RNA replication.  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