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 62 毫秒
1.
Positive strand viral replicases are membrane-bound complexes of viral and host proteins. The mechanism of viral replication and the role of host proteins are not well understood. To understand this mechanism, a viral replicase assay that utilizes extracts from dengue virus-infected mosquito (C6/36) cells and exogenous viral RNA templates is reported in this study. The 5'- and 3'-terminal regions (TR) of the template RNAs contain the conserved elements including the complementary (cyclization) motifs and stem-loop structures. RNA synthesis in vitro requires both 5'- and 3'-TR present in the same template molecule or when the 5'-TR RNA was added in trans to the 3'-untranslated region (UTR) RNA. However, the 3'-UTR RNA alone is not active. RNA synthesis occurs by elongation of the 3'-end of the template RNA to yield predominantly a double-stranded hairpin-like RNA product, twice the size of the template RNA. These results suggest that an interaction between 5'- and 3'-TR of the viral RNA that modulates the 3'-UTR RNA structure is required for RNA synthesis by the viral replicase. The complementary cyclization motifs of the viral genome also seem to play an important role in this interaction.  相似文献   

2.
3.
Viral replicases of many positive-strand RNA viruses are membrane-bound complexes of cellular and viral proteins that include viral RNA-dependent RNA polymerase (RdRP). The in vitro RdRP assay system that utilizes cytoplasmic extracts from dengue viral-infected cells and exogenous RNA templates was developed to understand the mechanism of viral replication in vivo. Our results indicated that in vitro RNA synthesis at the 3'-untranslated region (UTR) required the presence of the 5'-terminal region (TR) and the two cyclization (CYC) motifs suggesting a functional interaction between the TRs. In this study, using a psoralen-UV cross-linking method and an in vitro RdRP assay, we analyzed structural determinants for physical and functional interactions. Exogenous RNA templates that were used in the assays contained deletion mutations in the 5'-TR and substitution mutations in the 3'-stem-loop structure including those that would disrupt the predicted pseudoknot structure. Our results indicate that there is physical interaction between the 5'-TR and 3'-UTR that requires only the CYC motifs. RNA synthesis at the 3'-UTR, however, requires long range interactions involving the 5'-UTR, CYC motifs, and the 3'-stem-loop region that includes the tertiary pseudoknot structure.  相似文献   

4.
Replication of positive strand flaviviruses is mediated by the viral RNA-dependent RNA polymerases (RdRP). To study replication of dengue virus (DEN), a flavivirus family member, an in vitro RdRP assay was established using cytoplasmic extracts of DEN-infected mosquito cells and viral subgenomic RNA templates containing 5'- and 3'-terminal regions (TRs). Evidence supported that an interaction between the TRs containing conserved stem-loop, cyclization motifs, and pseudoknot structural elements is required for RNA synthesis. Two RNA products, a template size and a hairpin, twice that of the template, were formed. To isolate the function of the viral RdRP (NS5) from that of other host or viral factors present in the cytoplasmic extracts, the NS5 protein was expressed and purified from Escherichia coli. In this study, we show that the purified NS5 alone is sufficient for the synthesis of the two products and that the template-length RNA is the product of de novo initiation. Furthermore, the incubation temperature during initiation, but not elongation phase of RNA synthesis modulates the relative amounts of the hairpin and de novo RNA products. A model is proposed that a specific conformation of the viral polymerase and/or structure at the 3' end of the template RNA is required for de novo initiation.  相似文献   

5.
Hepatitis C virus (HCV) is the major etiological agent of non-A, non-B post-transfusion hepatitis. Its genome, a (+)-stranded RNA molecule of approximately 9.4 kb, encodes a large polyprotein that is processed by viral and cellular proteases into at least nine different viral polypeptides. As with other (+)-strand RNA viruses, the replication of HCV is thought to proceed via the initial synthesis of a complementary (-) RNA strand, which serves, in turn, as a template for the production of progeny (+)-strand RNA molecules. An RNA-dependent RNA polymerase has been postulated to be involved in both of these steps. Using the heterologous expression of viral proteins in insect cells, we present experimental evidence that an RNA-dependent RNA polymerase is encoded by HCV and that this enzymatic activity is the function of the 65 kDa non-structural protein 5B (NS5B). The characterization of the HCV RNA-dependent RNA polymerase product revealed that dimer-sized hairpin-like RNA molecules are generated in vitro, indicating that NS5B-mediated RNA polymerization proceeds by priming on the template via a 'copy-back' mechanism. In addition, the purified HCV NS5B protein was shown to perform RNA- or DNA oligonucleotide primer-dependent RNA synthesis on templates with a blocked 3' end or on homopolymeric templates. These results represent a first important step towards a better understanding of the life cycle of the HCV.  相似文献   

6.
The brome mosaic virus (BMV) RNA-dependent RNA polymerase (RdRp) directs template-specific synthesis of (-)-strand genomic and (+)-strand subgenomic RNAs in vitro. Although the requirements for (-)-strand RNA synthesis have been characterized previously, the mechanism of subgenomic RNA synthesis has not. Mutational analysis of the subgenomic promoter revealed that the +1 cytidylate and the +2 adenylate are important for RNA synthesis. Unlike (-)-strand RNA synthesis, which required only a high GTP concentration, subgenomic RNA synthesis required high concentrations of both GTP and UTP. Phylogenetic analysis of the sequences surrounding the initiation sites for subgenomic and genomic (+)-strand RNA synthesis in representative members of the alphavirus-like superfamily revealed that the +1 and +2 positions are highly conserved as a pyrimidine-adenylate. GDP and dinucleotide primers were able to more efficiently stimulate (-)-strand synthesis than subgenomic synthesis under conditions of limiting GTP. Oligonucleotide products of 6-, 7-, and 9-nt were synthesized and released by RdRp in 3-20-fold molar excess to full-length subgenomic RNA. Termination of RNA synthesis by RdRp was not induced by template sequence alone. Our characterization of the stepwise mechanism of subgenomic and (-)-strand RNA synthesis by RdRp permits comparisons to the mechanism of DNA-dependent RNA synthesis.  相似文献   

7.
Flaviviruses require complementarity between the 5' and 3' ends of the genome for RNA replication. For mosquito-borne flaviviruses, the cyclization sequences (CS) and upstream of AUG region (UAR) elements at the genomic termini are necessary for viral RNA replication, and a third motif, the downstream of AUG region (DAR), was recently designated for dengue virus. The 3' DAR sequence is also part of a hairpin (HP-3'SL), and both complementarity between 5' and 3' DAR motifs and formation of the HP-3'SL in the absence of the 5' end are conserved among mosquito-borne flaviviruses. Using West Nile virus as a model, we demonstrate that 5'-3' DAR complementarity and HP-3'SL formation are essential for viral RNA replication.  相似文献   

8.
Qi Y  Ding B 《The Plant cell》2003,15(11):2566-2577
The wide variety of RNAs produced in the nucleus must be localized correctly to perform their functions. However, the mechanism of this localization is poorly understood. We report here the differential subnuclear localization of RNA strands of opposite polarity derived from the replicating Potato spindle tuber viroid (PSTVd). During replication, (+)- and (-)-strand viroid RNAs are produced. We found that in infected cultured cells and plants, the (-)-strand RNA was localized in the nucleoplasm, whereas the (+)-strand RNA was localized in the nucleolus as well as in the nucleoplasm with distinct spatial patterns. Furthermore, the presence of the (+)-PSTVd in the nucleolus caused the redistribution of a small nucleolar RNA. Our results support a model in which (1) the synthesis of the (-)- and (+)-strands of PSTVd RNAs occurs in the nucleoplasm, (2) the (-)-strand RNA is anchored in the nucleoplasm, and (3) the (+)-strand RNA is transported selectively into the nucleolus. Our results imply that the eukaryotic cell has a machinery that recognizes and localizes the opposite strands of an RNA, which may have broad ramifications in the RNA regulation of gene expression and the infection cycle of pathogenic RNAs and in the development of RNA-based methods to control gene expression as well as pathogen infection.  相似文献   

9.
The replication enzyme of RNA viruses must preferentially recognize their RNAs in an environment that contains an abundance of cellular RNAs. The factors responsible for specific RNA recognition are not well understood, in part because viral RNA synthesis takes place within enzyme complexes associated with modified cellular membrane compartments. Recombinant RNA-dependent RNA polymerases (RdRps) from the human norovirus and the murine norovirus (MNV) were found to preferentially recognize RNA segments that contain the promoter and a short template sequence for subgenomic RNA synthesis. Both the promoter and template sequence contribute to stable RdRp binding, accurate initiation of the subgenomic RNAs and efficient RNA synthesis. Using a method that combines RNA crosslinking and mass spectrometry, residues near the template channel of the MNV RdRp were found to contact the hairpin RNA motif. Mutations in the hairpin contact site in the MNV RdRp reduced MNV replication and virus production in cells. This work demonstrates that the specific recognition of the norovirus subgenomic promoter is through binding by the viral RdRp.  相似文献   

10.
Kwon SJ  Rao AL 《Journal of virology》2012,86(9):5204-5220
Despite overwhelming interest in the impact exerted by recombination during evolution of RNA viruses, the relative contribution of the polarity of inoculum templates remains poorly understood. Here, by agroinfiltrating Nicotiana benthamiana leaves, we show that brome mosaic virus (BMV) replicase is competent to initiate positive-strand [(+)-strand] synthesis on an ectopically expressed RNA3 negative strand [(-) strand] and faithfully complete the replication cycle. Consequently, we sought to examine the role of RNA polarity in BMV recombination by expressing a series of replication-defective mutants of BMV RNA3 in (+) or (-) polarity. Temporal analysis of progeny sequences revealed that the genetic makeup of the primary recombinant pool is determined by the polarity of the inoculum template. When the polarity of the inoculum template was (+), the recombinant pool that accumulated during early phases of replication was a mixture of nonhomologous recombinants. These are longer than the inoculum template length, and a nascent 3' untranslated region (UTR) of wild-type (WT) RNA1 or RNA2 was added to the input mutant RNA3 3' UTR due to end-to-end template switching by BMV replicase during (-)-strand synthesis. In contrast, when the polarity of the inoculum was (-), the progeny contained a pool of native-length homologous recombinants generated by template switching of BMV replicase with a nascent UTR from WT RNA1 or RNA2 during (+)-strand synthesis. Repair of a point mutation caused by polymerase error occurred only when the polarity of the inoculum template was (+). These results contribute to the explanation of the functional role of RNA polarity in recombination mediated by copy choice mechanisms.  相似文献   

11.
L-A is a 4.6-kilobase double-stranded RNA virus of Saccharomyces cerevisiae. The in vitro L-A replication reaction ((-)-strand synthesis) requires an internal site 400 bases from the 3' end in addition to the 3'-terminal 30 nucleotides of the L-A (+)-single-stranded RNA. Elimination of the internal site reduces the template activity 5-10-fold. Here we investigate how the internal site can stimulate the replication reaction which starts at the 3' end of the template. When these two sites are split into two distinct RNA molecules, the internal site can no longer stimulate replication (no trans-activation). However, establishment of an intermolecular hydrogen bonding between these RNAs restored the replication-enhancing activity of the internal site. This result is consistent with a model wherein L-A's RNA polymerase interacts first with the internal site and then with the 3' end site by either looping or by a local dissociation-reassociation mechanism. These results, however, clearly eliminate anchored tracking and sliding models which require continuity of the RNA molecule between these two cis sites.  相似文献   

12.
Replication of plus-strand RNA viruses depends on recruited host factors that aid several critical steps during replication. Several of the co-opted host factors bind to the viral RNA, which plays multiple roles, including mRNA function, as an assembly platform for the viral replicase (VRC), template for RNA synthesis, and encapsidation during infection. It is likely that remodeling of the viral RNAs and RNA-protein complexes during the switch from one step to another requires RNA helicases. In this paper, we have discovered a second group of cellular RNA helicases, including the eIF4AIII-like yeast Fal1p and the DDX5-like Dbp3p and the orthologous plant AtRH2 and AtRH5 DEAD box helicases, which are co-opted by tombusviruses. Unlike the previously characterized DDX3-like AtRH20/Ded1p helicases that bind to the 3′ terminal promoter region in the viral minus-strand (−)RNA, the other class of eIF4AIII-like RNA helicases bind to a different cis-acting element, namely the 5′ proximal RIII(−) replication enhancer (REN) element in the TBSV (−)RNA. We show that the binding of AtRH2 and AtRH5 helicases to the TBSV (−)RNA could unwind the dsRNA structure within the RIII(−) REN. This unique characteristic allows the eIF4AIII-like helicases to perform novel pro-viral functions involving the RIII(−) REN in stimulation of plus-strand (+)RNA synthesis. We also show that AtRH2 and AtRH5 helicases are components of the tombusvirus VRCs based on co-purification experiments. We propose that eIF4AIII-like helicases destabilize dsRNA replication intermediate within the RIII(−) REN that promotes bringing the 5′ and 3′ terminal (−)RNA sequences in close vicinity via long-range RNA-RNA base pairing. This newly formed RNA structure promoted by eIF4AIII helicase together with AtRH20 helicase might facilitate the recycling of the viral replicases for multiple rounds of (+)-strand synthesis, thus resulting in asymmetrical viral replication.  相似文献   

13.
RNase footprinting and nitrocellulose filter binding assays were previously used to map one major and two minor binding sites for the cell protein eEF1A on the 3'(+) stem-loop (SL) RNA of West Nile virus (WNV) (3). Base substitutions in the major eEF1A binding site or adjacent areas of the 3'(+) SL were engineered into a WNV infectious clone. Mutations that decreased, as well as ones that increased, eEF1A binding in in vitro assays had a negative effect on viral growth. None of these mutations affected the efficiency of translation of the viral polyprotein from the genomic RNA, but all of the mutations that decreased in vitro eEF1A binding to the 3' SL RNA also decreased viral minus-strand RNA synthesis in transfected cells. Also, a mutation that increased the efficiency of eEF1A binding to the 3' SL RNA increased minus-strand RNA synthesis in transfected cells, which resulted in decreased synthesis of genomic RNA. These results strongly suggest that the interaction between eEF1A and the WNV 3' SL facilitates viral minus-strand synthesis. eEF1A colocalized with viral replication complexes (RC) in infected cells and antibody to eEF1A coimmunoprecipitated viral RC proteins, suggesting that eEF1A facilitates an interaction between the 3' end of the genome and the RC. eEF1A bound with similar efficiencies to the 3'-terminal SL RNAs of four divergent flaviviruses, including a tick-borne flavivirus, and colocalized with dengue virus RC in infected cells. These results suggest that eEF1A plays a similar role in RNA replication for all flaviviruses.  相似文献   

14.
The terminal half of the 5' untranslated region (UTR) in the (+)-strand RNA genome of tomato bushy stunt virus was analyzed for possible roles in viral RNA replication. Computer-aided thermodynamic analysis of secondary structure, phylogenetic comparisons for base-pair covariation, and chemical and enzymatic solution structure probing were used to analyze the 78 nucleotide long 5'-terminal sequence. The results indicate that this sequence adopts a branched secondary structure containing a three-helix junction core. The T-shaped domain (TSD) formed by this terminal sequence is closed by a prominent ten base-pair long helix, termed stem 1 (S1). Deletion of either the 5' or 3' segment forming S1 (coordinates 1-10 or 69-78, respectively) in a model subviral RNA replicon, i.e. a prototypical defective interfering (DI) RNA, reduced in vivo accumulation levels of this molecule approximately 20-fold. Compensatory-type mutational analysis of S1 within this replicon revealed a strong correlation between formation of the predicted S1 structure and efficient DI RNA accumulation. RNA decay studies in vivo did not reveal any notable changes in the physical stabilities of DI RNAs containing disrupted S1s, thus implicating RNA replication as the affected process. Further investigation revealed that destabilization of S1 in the (+)-strand was significantly more detrimental to DI RNA accumulation than (-)-strand destabilization, therefore S1-mediated activity likely functions primarily via the (+)-strand. The essential role of S1 in DI RNA accumulation prompted us to examine the 5'-proximal secondary structure of a previously identified mutant DI RNA, RNA B, that lacks the 5' UTR but is still capable of low levels of replication. Mutational analysis of a predicted S1-like element present within a cryptic 5'-terminal TSD confirmed the importance of the former in RNA B accumulation. Collectively, these data support a fundamental role for the TSD, and in particular its S1 subelement, in tombusvirus RNA replication.  相似文献   

15.
16.
Host factors are recruited into viral replicase complexes to aid replication of plus-strand RNA viruses. In this paper, we show that deletion of eukaryotic translation elongation factor 1Bgamma (eEF1Bγ) reduces Tomato bushy stunt virus (TBSV) replication in yeast host. Also, knock down of eEF1Bγ level in plant host decreases TBSV accumulation. eEF1Bγ binds to the viral RNA and is one of the resident host proteins in the tombusvirus replicase complex. Additional in vitro assays with whole cell extracts prepared from yeast strains lacking eEF1Bγ demonstrated its role in minus-strand synthesis by opening of the structured 3' end of the viral RNA and reducing the possibility of re-utilization of (+)-strand templates for repeated (-)-strand synthesis within the replicase. We also show that eEF1Bγ plays a synergistic role with eukaryotic translation elongation factor 1A in tombusvirus replication, possibly via stimulation of the proper positioning of the viral RNA-dependent RNA polymerase over the promoter region in the viral RNA template.These roles for translation factors during TBSV replication are separate from their canonical roles in host and viral protein translation.  相似文献   

17.
The segmented double-stranded (ds) RNA genome of the rotaviruses is replicated asymmetrically, with viral mRNA serving as the template for the synthesis of minus-strand RNA. Previous studies with cell-free replication systems have shown that the highly conserved termini of rotavirus gene 8 and 9 mRNAs contain cis-acting signals that promote the synthesis of dsRNA. Based on the location of the cis-acting signals and computer modeling of their secondary structure, the ends of the gene 8 or 9 mRNAs are proposed to interact in cis to form a modified panhandle structure that promotes the synthesis of dsRNA. In this structure, the last 11 to 12 nucleotides of the RNA, including the cis-acting signal that is essential for RNA replication, extend as a single-stranded tail from the panhandled region, and the 5′ untranslated region folds to form a stem-loop motif. To understand the importance of the predicted secondary structure in minus-strand synthesis, mutations were introduced into viral RNAs which affected the 3′ tail and the 5′ stem-loop. Analysis of the RNAs with a cell-free replication system showed that, in contrast to mutations which altered the structure of the 5′ stem-loop, mutations which caused complete or near-complete complementarity between the 5′ end and the 3′ tail significantly inhibited (≥10-fold) minus-strand synthesis. Likewise, incubation of wild-type RNAs with oligonucleotides which were complementary to the 3′ tail inhibited replication. Despite their replication-defective phenotype, mutant RNAs with complementary 5′ and 3′ termini were shown to competitively interfere with the replication of wild-type mRNA and to bind the viral RNA polymerase VP1 as efficiently as wild-type RNA. These results indicate that the single-strand nature of the 3′ end of rotavirus mRNA is essential for efficient dsRNA synthesis and that the specific binding of the RNA polymerase to the mRNA template is required but not sufficient for the synthesis of minus-strand RNA.  相似文献   

18.
Terminal adenylation in the synthesis of RNA by Q beta replicase   总被引:10,自引:0,他引:10  
We investigated the apparent requirement that Q beta replicase must add a nontemplated adenosine to the 3' end of newly synthesized RNA strands. We used abbreviated MDV-1 (+)-RNA templates that lacked either 62 or 63 nucleotides at their 5' end in Q beta replicase reactions. The MDV-1 (-)-RNA strands synthesized from these abbreviated (+)-strand templates were released from the replication complex, yet they did not possess a nontemplated 3'-terminal adenosine. These results imply that, despite observations that all naturally occurring RNAs synthesized by Q beta replicase possess a nontemplated 3'-adenosine, the addition of an extra adenosine is not an obligate step for the release of completed strands. Since the abbreviated templates lacked a normal 5' end, it is probable that a particular sequence at the 5' end of the template is required for terminal adenylation to occur.  相似文献   

19.
Emara MM  Liu H  Davis WG  Brinton MA 《Journal of virology》2008,82(21):10657-10670
Previous data showed that the cellular proteins TIA-1 and TIAR bound specifically to the West Nile virus 3' minus-strand stem-loop [WNV3'(-)SL] RNA (37) and colocalized with flavivirus replication complexes in WNV- and dengue virus-infected cells (21). In the present study, the sites on the WNV3'(-)SL RNA required for efficient in vitro T-cell intracellular antigen-related (TIAR) and T-cell intracellular antigen-1 (TIA-1) protein binding were mapped to short AU sequences (UAAUU) located in two internal loops of the WNV3'(-)SL RNA structure. Infectious clone RNAs with all or most of the binding site nucleotides in one of the 3' (-)SL loops deleted or substituted did not produce detectable virus after transfection or subsequent passage. With one exception, deletion/mutation of a single terminal nucleotide in one of the binding sequences had little effect on the efficiency of protein binding or virus production, but mutation of a nucleotide in the middle of a binding sequence reduced both the in vitro protein binding efficiency and virus production. Plaque size, intracellular genomic RNA levels, and virus production progressively decreased with decreasing in vitro TIAR/TIA-1 binding activity, but the translation efficiency of the various mutant RNAs was similar to that of the parental RNA. Several of the mutant RNAs that inefficiently interacted with TIAR/TIA-1 in vitro rapidly reverted in vivo, indicating that they could replicate at a low level and suggesting that an interaction between TIAR/TIA-1 and the viral 3'(-)SL RNA is not required for initial low-level symmetric RNA replication but instead facilitates the subsequent asymmetric amplification of genome RNA from the minus-strand template.  相似文献   

20.
The replicase activity of rotavirus open cores has been used to study the synthesis of (-) strand RNA from viral (+) strand RNA in a cell-free replication system. The last 7 nt of the (+) strand RNA, 5'-UGUGACC-3', are highly conserved and are necessary for efficient (-) strand synthesis in vitro. Characterization of the cell-free replication system revealed that the addition of NaCl inhibited (-) strand synthesis. By preincubating open cores with (+) strand RNA and ATP, CTP, and GTP prior to the addition of NaCl and UTP, the salt-sensitive step was overcome. Thus, (-) strand initiation, but not elongation, was a salt-sensitive process in the cell-free system. Further analysis of the requirements for initiation showed that preincubating open cores and the (+) strand RNA with GTP or UTP, but not with ATP or CTP, allowed (-) strand synthesis to occur in the presence of NaCl. Mutagenesis suggested that in the presence of GTP, (-) strand synthesis initiated at the 3'-terminal C residue of the (+) strand template, whereas in the absence of GTP, an aberrant initiation event occurred at the third residue upstream from the 3' end of the (+) strand RNA. During preincubation with GTP, formation of the dinucleotides pGpG and ppGpG was detected; however, no such products were made during preincubation with ATP, CTP, or UTP. Replication assays showed that pGpG, but not GpG, pApG, or ApG, served as a specific primer for (-) strand synthesis and that the synthesis of pGpG may occur by a template-independent process. From these data, we conclude that initiation of rotavirus (-) strand synthesis involves the formation of a ternary complex consisting of the viral RNA-dependent RNA polymerase, viral (+) strand RNA, and possibly a 5'-phosphorylated dinucleotide, that is, pGpG or ppGpG.  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