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 46 毫秒
1.
The 5′ end of the genomic RNA of rubella virus (RUB) contains a 14-nucleotide (nt) single-stranded leader (ss-leader) followed by a stem-and-loop structure [5′(+)SL] (nt 15 to 65), the complement of which at the 3′ end of the minus-strand RNA [3′(−)SL] has been proposed to function as a promoter for synthesis of genomic plus strands. A second intriguing feature of the 5′ end of the RUB genomic RNA is the presence of a short (17 codons) open reading frame (ORF) located between nt 3 and 54; the ORF encoding the viral nonstructural proteins (NSPs) initiates at nt 41 in an alternate translational frame. To address the functional significance of these features, we compared the 5′-terminal sequences of six different strains of RUB, with the result that the short ORF is preserved (although the coding sequence is not conserved) as is the stem part of both the 5′(+)SL and 3′(−)SL, while the upper loop part of both structures varies. Next, using Robo302, an infectious cDNA clone of RUB, we introduced 31 different mutations into the 5′-terminal noncoding region, and their effects on virus replication and macromolecular synthesis were examined. This mutagenesis revealed that the short ORF is not essential for virus replication. The AA dinucleotide at nt 2 and 3 is of critical importance since point mutations and deletions that altered or removed both of these nucleotides were lethal. None of the other mutations within either the ss-leader or the 5′(+)SL [and accordingly within the 3′(−)SL], including deletions of up to 15 nt from the 5′(+)SL and three different multiple-point mutations that lead to destabilization of the 5′(+)SL, were lethal. Some of the mutations within both ss-leader and the 5′(+)SL resulted in viruses that grew to lower titers than the wild-type virus and formed opaque and/or small plaques; in general mutations within the stem had a more profound effect on viral phenotype than did mutations in either the ss-leader or upper loop. Mutations in the 5′(+)SL, but not in the ss-leader, resulted in a significant reduction in NSP synthesis, indicating that this structure is important for efficient translation of the NSP ORF. In contrast, viral plus-strand RNA synthesis was unaffected by the 5′(+)SL mutations as well as the ss-leader mutations, which argues against the proposed function of the 3′(−)SL as a promoter for initiation of the genomic plus-strand RNA.  相似文献   

2.
The cis-acting genomic RNA requirements for the assembly of vesicular stomatitis virus (VSV) ribonucleocapsids into infectious particles were investigated. Using a biological assay based on particle infectivity, we demonstrated that subgenomic replicons that contained all four possible combinations of the natural genomic termini, the 3′ leader (Le) and 5′ trailer (Tr) regions, were replication competent; however, a 3′ copyback replicon (3′CB), containing the natural 3′ terminus but having the 5′ Tr replaced by a sequence complementary to the 3′ Le for 46 nucleotides, was unable to assemble infectious particles, despite efficient replication. When a copy of Tr was inserted 51 nucleotides from the 5′ end of 3′CB, infectious particles were produced. However, analysis of the replication products of these particles showed that the 51 nucleotides which corresponded to the Le complement sequences at the 5′ terminus were removed during RNA replication, thus restoring the wild-type 5′ Tr to the exact 5′ terminus. These data showed that a cis-acting signal was necessary for assembly of VSV RNAs into infectious particles and that this signal was supplied by Tr when located at the 5′ end. The regions within Tr required for assembly were analyzed by a series of deletions and exchanges for Le complement sequences, which demonstrated that the 5′ terminal 29 nucleotides of Tr allowed assembly of infectious particles but that the 5′ terminal 22 nucleotides functioned poorly. Deletions in Tr also altered the balance between negative- and positive-strand genomic RNA and affected levels of replication. RNAs that retained fewer than 45 but at least 22 nucleotides of the 5′ terminus could replicate but were impaired in RNA replication, and RNAs that retained only 14 nucleotides of the 5′ terminus were severely reduced in ability to replicate. These data define the VSV Tr as a position-dependent, cis-acting element for the assembly of RNAs into infectious particles, and they delineate RNA sequences that are essential for negative-strand RNA synthesis. These observations are consistent with, and offer an explanation for, the absence of 3′ copyback defective interfering particles in nature.  相似文献   

3.
Feline calicivirus (FCV), a member of the Caliciviridae, produces its major structural protein as a precursor polyprotein from a subgenomic-sized mRNA. In this study, we show that the proteinase responsible for processing this precursor into the mature capsid protein is encoded by the viral genome at the 3′-terminal portion of open reading frame 1 (ORF1). Protein expression studies of either the entire or partial ORF1 indicate that the proteinase is active when expressed either in in vitro translation or in bacterial cells. Site-directed mutagenesis was used to characterize the proteinase Glu-Ala cleavage site in the capsid precursor, utilizing an in vitro cleavage assay in which mutant precursor proteins translated from cDNA clones were used as substrates for trans cleavage by the proteinase. In general, amino acid substitutions in the P1 position (Glu) of the cleavage site were less well tolerated by the proteinase than those in the P1′ position (Ala). The precursor cleavage site mutations were introduced into an infectious cDNA clone of the FCV genome, and transfection of RNA derived from these clones into feline kidney cells showed that efficient cleavage of the capsid precursor by the virus-encoded proteinase is a critical determinant in the growth of the virus.  相似文献   

4.

Background

Positive strand RNA viruses rely heavily on host cell RNA binding proteins for various aspects of their life cycle. Such proteins interact with sequences usually present at the 5′ or 3′ extremities of the viral RNA genome, to regulate viral translation and/or replication. We have previously reported that the well characterized host RNA binding protein polypyrimidine tract binding protein (PTB) interacts with the 5′end of the feline calicivirus (FCV) genomic and subgenomic RNAs, playing a role in the FCV life cycle.

Principal Findings

We have demonstrated that PTB interacts with at least two binding sites within the 5′end of the FCV genome. In vitro translation indicated that PTB may function as a negative regulator of FCV translation and this was subsequently confirmed as the translation of the viral subgenomic RNA in PTB siRNA treated cells was stimulated under conditions in which RNA replication could not occur. We also observed that PTB redistributes from the nucleus to the cytoplasm during FCV infection, partially localizing to viral replication complexes, suggesting that PTB binding may be involved in the switch from translation to replication. Reverse genetics studies demonstrated that synonymous mutations in the PTB binding sites result in a cell-type specific defect in FCV replication.

Conclusions

Our data indicates that PTB may function to negatively regulate FCV translation initiation. To reconcile this with efficient virus replication in cells, we propose a putative model for the function of PTB in the FCV life cycle. It is possible that during the early stages of infection, viral RNA is translated in the absence of PTB, however, as the levels of viral proteins increase, the nuclear-cytoplasmic shuttling of PTB is altered, increasing the cytoplasmic levels of PTB, inhibiting viral translation. Whether PTB acts directly to repress translation initiation or via the recruitment of other factors remains to be determined but this may contribute to the stimulation of viral RNA replication via clearance of ribosomes from viral RNA.  相似文献   

5.
6.
Caliciviruses use reinitiation of translation governed by a ‘termination upstream ribosomal binding site’ (TURBS) for expression of their minor capsid protein VP2. Mutation analysis allowed to identify sequences surrounding the translational start/stop site of the feline calicivirus (FCV) that fine tune reinitiation frequency. A selection of these changes was introduced into the infectious FCV cDNA clone to check the influence of altered VP2 levels on virus replication. In addition, full length constructs were established that displayed a conformation, in which VP2 expression occurred under control of a duplicated subgenomic promoter. Viable viruses recovered from such constructs revealed a rather broad range of VP2 expression levels but comparable growth kinetics showing that caliciviruses can tolerate gross changes of the VP2 expression level.  相似文献   

7.
Open reading frame 2 (ORF2) of the feline calicivirus (FCV) genome encodes a capsid precursor that is posttranslationally processed to release the mature capsid protein (VP1) and a small protein of 124 amino acids, designated the leader of the capsid (LC). To investigate the role of the LC protein in the virus life cycle, mutations and deletions were introduced into the LC coding region of an infectious FCV cDNA clone. Three cysteine residues that are conserved among all vesivirus LC sequences were found to be critical for the recovery of FCV with a characteristic cytopathic effect in feline kidney cells. A cell-rounding phenotype associated with the transient expression of wild-type and mutagenized forms of the LC correlated with the cytopathic and growth properties of the corresponding engineered viruses. The host cellular protein annexin A2 was identified as a binding partner of the LC protein, consistent with a role for the LC in mediating host cell interactions that alter the integrity of the cell and enable virus spread.  相似文献   

8.
RNA isolated from southern bean mosaic virions contains, in small amount, a subgenomic RNA (molecular weight, 0.38 × 106) that serves in vitro as an mRNA for southern bean mosaic virus coat protein. The RNA has a 5′-linked protein indistinguishable from the protein linked to the 5′ end of full-length genomic RNA. Its base sequence, determined to 91 bases from the 3′ end, is identical to the 3′-terminal sequence of the genomic RNA. The results suggest that the coat protein messenger sequence exists as a “silent” cistron near the 3′ end of the genomic RNA.  相似文献   

9.
A previous analysis of naturally occurring defective interfering (DI) RNA genomes of the prototypic paramyxovirus simian virus 5 (SV5) indicated that 113 bases at the 3′ terminus of the antigenome were sufficient to direct RNA encapsidation and replication. A nucleotide sequence alignment of the antigenomic 3′-terminal 113 bases of members of the Rubulavirus genus of the Paramyxoviridae family identified two regions of sequence identity: bases 1 to 19 at the 3′ terminus (conserved region I [CRI]) and a more distal region consisting of antigenome bases 73 to 90 (CRII) that was contained within the 3′ coding region of the L protein gene. To determine whether these regions of the antigenome were essential for SV5 RNA replication, a reverse genetics system was used to analyze the replication of copyback DI RNA analogs that contained a foreign gene (GL, encoding green fluorescence protein) flanked by 113 5′-terminal bases and various amounts of SV5 3′-terminal antigenomic sequences. Results from a deletion analysis showed that efficient encapsidation and replication of SV5-GL DI RNA analogs occurred when the 90 3′-terminal bases of the SV5 antigenomic RNA were retained, but replication was reduced ~5- to 14-fold in the case of truncated antigenomes that lacked the 3′-end CRII sequences. A chimeric copyback DI RNA containing the 3′-terminal 98 bases including the CRI and CRII sequences from the human parainfluenza virus type 2 (HPIV2) antigenome in place of the corresponding SV5 sequences was efficiently replicated by SV5 cDNA-derived components. However, replication was reduced ~20-fold for a truncated SV5-HPIV2 chimeric RNA that lacked the HPIV2 CRII sequences between antigenome bases 72 and 90. Progressive deletions of 6 to 18 bases in the region located between the SV5 antigenomic CRI and CRII segments (3′-end nucleotides 21 to 38) resulted in a ~25-fold decrease in SV5-GL RNA synthesis. Surprisingly, replication was restored to wild-type levels when these length alterations between CRI and CRII were corrected by replacing the deleted bases with nonviral sequences. Together, these data suggest that a functional SV5 antigenomic promoter requires proper spacing between an essential internal region and the 3′ terminus. A model is presented for the structure of the 3′ end of the SV5 antigenome which proposes that positioning of CRI and CRII along the same face of the helical nucleocapsid is an essential feature of a functional antigenomic promoter.  相似文献   

10.
Many questions regarding the initiation of replication and translation of the segmented, double-stranded RNA genome of infectious bursal disease virus (IBDV) remain to be solved. Computer analysis shows that the non-polyadenylated extreme 3′-untranslated regions (UTRs) of the coding strand of both genomic segments are able to fold into a single stem–loop structure. To assess the determinants for a functional 3′-UTR, we mutagenized the 3′-UTR stem–loop structure of the B-segment. Rescue of infectious virus from mutagenized cDNA plasmids was impaired in all cases. However, after one passage, the replication kinetics of these viruses were restored. Sequence analysis revealed that additional mutations had been acquired in most of the stem–loop structures, which compensated the introduced ones. A rescued virus with a modified stem–loop structure containing four nucleotide substitutions, but preserving its overall secondary structure, was phenotypically indistinguishable from wild-type virus, both in vitro (cell culture) and in vivo (chickens, natural host). Sequence analysis showed that the modified stem–loop structure of this virus was fully preserved after four serial passages. Apparently, it is the stem–loop structure and not the primary sequence that is the functional determinant in the 3′-UTRs of IBDV.  相似文献   

11.
12.
Sequences and structures within the terminal genomic regions of plus-strand RNA viruses are targets for the binding of host proteins that modulate functions such as translation, RNA replication, and encapsidation. Using murine norovirus 1 (MNV-1), we describe the presence of long-range RNA-RNA interactions that were stabilized by cellular proteins. The proteins potentially responsible for the stabilization were selected based on their ability to bind the MNV-1 genome and/or having been reported to be involved in the stabilization of RNA-RNA interactions. Cell extracts were preincubated with antibodies against the selected proteins and used for coprecipitation reactions. Extracts treated with antibodies to poly(C) binding protein 2 (PCBP2) and heterogeneous nuclear ribonucleoprotein (hnRNP) A1 significantly reduced the 5′-3′ interaction. Both PCBP2 and hnRNP A1 recombinant proteins stabilized the 5′-3′ interactions and formed ribonucleoprotein complexes with the 5′ and 3′ ends of the MNV-1 genomic RNA. Mutations within the 3′ complementary sequences (CS) that disrupt the 5′-3′-end interactions resulted in a significant reduction of the viral titer, suggesting that the integrity of the 3′-end sequence and/or the lack of complementarity with the 5′ end is important for efficient virus replication. Small interfering RNA-mediated knockdown of PCBP2 or hnRNP A1 resulted in a reduction in virus yield, confirming a role for the observed interactions in efficient viral replication. PCBP2 and hnRNP A1 induced the circularization of MNV-1 RNA, as revealed by electron microscopy. This study provides evidence that PCBP2 and hnRNP A1 bind to the 5′ and 3′ ends of the MNV-1 viral RNA and contribute to RNA circularization, playing a role in the virus life cycle.  相似文献   

13.
Genomes of RNA viruses encounter a continual threat from host cellular ribonucleases. Therefore, viruses have evolved mechanisms to protect the integrity of their genomes. To study the mechanism of 3′-end repair in dengue virus-2 in mammalian cells, a series of 3′-end deletions in the genome were evaluated for virus replication by detection of viral antigen NS1 and by sequence analysis. Limited deletions did not cause any delay in the detection of NS1 within 5 d. However, deletions of 7–10 nucleotides caused a delay of 9 d in the detection of NS1. Sequence analysis of RNAs from recovered viruses showed that at early times, virus progenies evolved through RNA molecules of heterogeneous lengths and nucleotide sequences at the 3′ end, suggesting a possible role for terminal nucleotidyl transferase activity of the viral polymerase (NS5). However, this diversity gradually diminished and consensus sequences emerged. Template activities of 3′-end mutants in the synthesis of negative-strand RNA in vitro by purified NS5 correlate well with the abilities of mutant RNAs to repair and produce virus progenies. Using the Mfold program for RNA structure prediction, we show that if the 3′ stem–loop (3′ SL) structure was abrogated by mutations, viruses eventually restored the 3′ SL structure. Taken together, these results favor a two-step repair process: non-template-based nucleotide addition followed by evolutionary selection of 3′-end sequences based on the best-fit RNA structure that can support viral replication.  相似文献   

14.
Alphaviruses are mosquito-transmitted RNA viruses that cause important diseases in both humans and livestock. Sindbis virus (SIN), the type species of the alphavirus genus, carries a 11.7-kb positive-sense RNA genome which is capped at its 5′ end and polyadenylated at its 3′ end. The 3′ nontranslated region (3′NTR) of the SIN genome carries many AU-rich motifs, including a 19-nucleotide (nt) conserved element (3′CSE) and a poly(A) tail. This 3′CSE and the adjoining poly(A) tail are believed to regulate the synthesis of negative-sense RNA and genome replication in vivo. We have recently demonstrated that the SIN genome lacking the poly(A) tail was infectious and that de novo polyadenylation could occur in vivo (K. R. Hill, M. Hajjou, J. Hu, and R. Raju, J. Virol. 71:2693–2704, 1997). Here, we demonstrate that the 3′-terminal 29-nt region of the SIN genome carries a signal for possible cytoplasmic polyadenylation. To further investigate the polyadenylation signals within the 3′NTR, we generated a battery of mutant genomes with mutations in the 3′NTR and tested their ability to generate infectious virus and undergo 3′ polyadenylation in vivo. Engineered SIN genomes with terminal deletions within the 19-nt 3′CSE were infectious and regained their poly(A) tail. Also, a SIN genome carrying the poly(A) tail but lacking a part or the entire 19-nt 3′CSE was also infectious. Sequence analysis of viruses generated from these engineered SIN genomes demonstrated the addition of a variety of AU-rich sequence motifs just adjacent to the poly(A) tail. The addition of AU-rich motifs to the mutant SIN genomes appears to require the presence of a significant portion of the 3′NTR. These results indicate the ability of alphavirus RNAs to undergo 3′ repair and the existence of a pathway for the addition of AU-rich sequences and a poly(A) tail to their 3′ end in the infected host cell. Most importantly, these results indicate the ability of alphavirus replication machinery to use a multitude of AU-rich RNA sequences abutted by a poly(A) motif as promoters for negative-sense RNA synthesis and genome replication in vivo. The possible roles of cytoplasmic polyadenylation machinery, terminal transferase-like enzymes, and the viral polymerase in the terminal repair processes are discussed.  相似文献   

15.
The 3′-terminal nucleotides (nt) of West Nile virus (WNV) genomic RNA form a penultimate 16-nt small stem-loop (SSL) and an 80-nt terminal stem-loop (SL). These RNA structures are conserved in divergent flavivirus genomes. A previous in vitro study using truncated WNV 3′ RNA structures predicted a putative tertiary interaction between the 5′ side of the 3′-terminal SL and the loop of the SSL. Although substitution or deletion of the 3′ G (nt 87) within the SSL loop, which forms the only G-C pair in the predicted tertiary interaction, in a WNV infectious clone was lethal, a finding consistent with the involvement in a functionally relevant pseudoknot interaction, extensive mutagenesis of nucleotides in the terminal SL did not identify a cis-acting pairing partner for this SSL 3′ G. However, both the sequence and the structural context of two adjacent base pairs flanked by symmetrical internal loops in the 3′-terminal SL were shown to be required for efficient viral RNA replication. Nuclear magnetic resonance analysis confirmed the predicted SSL and SL structures but not the tertiary interaction. The SSL was previously reported to contain one of three eEF1A binding sites, and G87 in the SSL loop was shown to be involved in eEF1A binding. The nucleotides at the bottom part of the 3′-terminal SL switch between 3′ RNA-RNA and 3′-5′ RNA-RNA interactions. The data suggest that interaction of the 3′ SL RNA with eEF1A at three sites and a unique metastable structural feature may participate in regulating structural changes in the 3′-terminal SL.  相似文献   

16.
17.
18.
The flavivirus genome is a positive-stranded ~11-kb RNA including 5′ and 3′ noncoding regions (NCR) of approximately 100 and 400 to 600 nucleotides (nt), respectively. The 3′ NCR contains adjacent, thermodynamically stable, conserved short and long stem-and-loop structures (the 3′-SL), formed by the 3′-terminal ~100 nt. The nucleotide sequences within the 3′-SL are not well conserved among species. We examined the requirement for the 3′-SL in the context of dengue virus type 2 (DEN2) replication by mutagenesis of an infectious cDNA copy of a DEN2 genome. Genomic full-length RNA was transcribed in vitro and used to transfect monkey kidney cells. A substitution mutation, in which the 3′-terminal 93 nt constituting the wild-type (wt) DEN2 3′-SL sequence were replaced by the 96-nt sequence of the West Nile virus (WN) 3′-SL, was sublethal for virus replication. An analysis of the growth phenotypes of additional mutant viruses derived from RNAs containing DEN2-WN chimeric 3′-SL structures suggested that the wt DEN2 nucleotide sequence forming the bottom half of the long stem and loop in the 3′-SL was required for viability. One 7-bp substitution mutation in this domain resulted in a mutant virus that grew well in monkey kidney cells but was severely restricted in cultured mosquito cells. In contrast, transpositions of and/or substitutions in the wt DEN2 nucleotide sequence in the top half of the long stem and in the short stem and loop were relatively well tolerated, provided the stem-loop secondary structure was conserved.  相似文献   

19.
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   

20.
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