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.
C L Liao  M M Lai 《Journal of virology》1992,66(10):6117-6124
Mouse hepatitis virus (MHV), a coronavirus, has been shown to undergo a high frequency of RNA recombination both in tissue culture and in animal infection. So far, RNA recombination has been demonstrated only between genomic RNAs of two coinfecting viruses. To understand the mechanism of RNA recombination and to further explore the potential of RNA recombination, we studied whether recombination could occur between a replicating MHV RNA and transfected RNA fragments. We first used RNA fragments which represented the 5' end of genomic-sense sequences of MHV RNA for transfection. By using polymerase chain reaction amplification with two specific primers, we were able to detect recombinant RNAs which incorporated the transfected fragment into the 5' end of the viral RNA in the infected cells. Surprisingly, even the anti-genomic-sense RNA fragments complementary to the 5' end of MHV genomic RNA could also recombine with the MHV genomic RNAs. This observation suggests that RNA recombination can occur during both positive- and negative-strand RNA synthesis. Furthermore, the recombinant RNAs could be detected in the virion released from the infected cells even after several passages of virus in tissue culture cells, indicating that these recombinant RNAs represented functional virion RNAs. The crossover sites of these recombinants were detected throughout the transfected RNA fragments. However, when an RNA fragment with a nine-nucleotide (CUUUAUAAA) deletion immediately downstream of a pentanucleotide (UCUAA) repeat sequence in the leader RNA was transfected into MHV-infected cells, most of the recombinants between this RNA and the MHV genome contained crossover sites near this pentanucleotide repeat sequence. In contrast, when exogenous RNAs with the intact nine-nucleotide sequence were used in similar experiments, the crossover sites of recombinants in viral genomic RNA could be detected at more-downstream sites. This study demonstrated that recombination can occur between replicating MHV RNAs and RNA fragments which do not replicate, suggesting the potential of RNA recombination for genetic engineering.  相似文献   

2.
《Seminars in Virology》1997,8(2):101-111
Naturally occurring defective interfering RNAs have been found in 4 of 14 coronavirus species. They range in size from 2.2 kb to approximately 25 kb, or 80% of the 30-kb parent virus genome. The large DI RNAs do not in all cases appear to require helper virus for intracellular replication and it has been postulated that they may on their own function as agents of disease. Coronavirus DI RNAs appear to arise by internal deletions (through nonhomologous recombination events) on the virus genome or on DI RNAs of larger size by a polymerase strand-switching (copy-choice) mechanism. In addition to their use in the study of virus RNA replication and virus assembly, coronavirus DI RNAs are being used in a major way to study the mechanism of a high-frequency, site-specific RNA recombination event that leads to leader acquisition during virus replication (i.e., the leader fusion event that occurs during synthesis of subgenomic mRNAs, and the leader-switching event that can occur during DI RNA replication), a distinguishing feature of coronaviruses (and arteriviruses). Coronavirus DI RNAs are also being engineered as vehicles for the generation of targeted recombinants of the parent virus genome.  相似文献   

3.
Turnip yellow mosaic virus (TYMV) is a spherical plant virus that has a single 6.3 kb positive strand RNA as a genome. In this study, RNA1 sequence of Flock house virus (FHV) was inserted into the TYMV genome to test whether TYMV can accommodate and express another viral entity. In the resulting construct, designated TY-FHV, the FHV RNA1 sequence was expressed as a TYMV subgenomic RNA. Northern analysis of the Nicotiana benthamiana leaves agroinfiltrated with the TY-FHV showed that both genomic and subgenomic FHV RNAs were abundantly produced. This indicates that the FHV RNA1 sequence was correctly expressed and translated to produce a functional FHV replicase. Although these FHV RNAs were not encapsidated, the FHV RNA having a TYMV CP sequence at the 3’-end was efficiently encapsidated. When an eGFP gene was inserted into the B2 ORF of the FHV sequence, a fusion protein of B2-eGFP was produced as expected. [BMB Reports 2014; 47(6): 330-335]  相似文献   

4.
Several mechanisms, including a high mutation rate and reassortment of genes, have been found to be responsible for the variability of influenza A viruses. RNA recombination would be another mechanism leading to genetic variation; however, recombination has only rarely been reported to occur in influenza viruses. During ribonucleoprotein transfection experiments designed to generate viable influenza viruses from in vitro-synthesized RNA, we discovered several viruses which must have originated from recombination events. The ribonucleoprotein transfection system may enhance the formation of viruses which result from jumping of the viral polymerase between RNAs or from ligation of different viral RNAs. Five different recombinant viruses are described. Two of these, REC1 and REC2, contain a neuraminidase (NA) gene whose defective polyadenylation signal has been repaired via intergenic recombination; 124 and 95 nucleotides have been added, respectively. Another virus, REC5, must have originated by multiple recombination events since it contains a mosaic gene with sequences derived from the NA gene of influenza A/WSN/33 virus and the matrix, polymerase protein PB1, and NA genes of influenza A/PR/8/34 virus.  相似文献   

5.
Brome mosaic virus (BMV) packages its genomic and subgenomic RNAs into three separate viral particles. BMV purified from barley, wheat, and tobacco have distinct relative abundances of the encapsidated RNAs. We seek to identify the basis for the host-dependent differences in viral RNA encapsidation. Sequencing of the viral RNAs revealed recombination events in the 3′ untranslated region of RNA1 of BMV purified from barley and wheat, but not from tobacco. However, the relative amounts of the BMV RNAs that accumulated in barley and wheat are similar and RNA accumulation is not sufficient to account for the difference in RNA encapsidation. Virions purified from barley and wheat were found to differ in their isoelectric points, resistance to proteolysis, and contacts between the capsid residues and the RNA. Mass spectrometric analyses revealed that virions from the three hosts had different post-translational modifications that should impact the physiochemical properties of the virions. Another major source of variation in RNA encapsidation was due to the purification of BMV particles to homogeneity. Highly enriched BMV present in lysates had a surprising range of sizes, buoyant densities, and distinct relative amounts of encapsidated RNAs. These results show that the encapsidated BMV RNAs reflect a combination of host effects on the physiochemical properties of the viral capsids and the enrichment of a subset of virions. The previously unexpected heterogeneity in BMV should influence the timing of the infection and also the host innate immune responses.  相似文献   

6.
The intricate interactions between viruses and hosts include an evolutionary arms race and adaptation that is facilitated by the ability of RNA viruses to evolve rapidly due to high frequency mutations and genetic RNA recombination. In this paper, we show evidence that the co-opted cellular DDX3-like Ded1 DEAD-box helicase suppresses tombusviral RNA recombination in yeast model host, and the orthologous RH20 helicase functions in a similar way in plants. In vitro replication and recombination assays confirm the direct role of the ATPase function of Ded1p in suppression of viral recombination. We also present data supporting a role for Ded1 in facilitating the switch from minus- to plus-strand synthesis. Interestingly, another co-opted cellular helicase, the eIF4AIII-like AtRH2, enhances TBSV recombination in the absence of Ded1/RH20, suggesting that the coordinated actions of these helicases control viral RNA recombination events. Altogether, these helicases are the first co-opted cellular factors in the viral replicase complex that directly affect viral RNA recombination. Ded1 helicase seems to be a key factor maintaining viral genome integrity by promoting the replication of viral RNAs with correct termini, but inhibiting the replication of defective RNAs lacking correct 5’ end sequences. Altogether, a co-opted cellular DEAD-box helicase facilitates the maintenance of full-length viral genome and suppresses viral recombination, thus limiting the appearance of defective viral RNAs during replication.  相似文献   

7.
8.
Equine arteritis virus (EAV), the prototype arterivirus, is an enveloped plus-strand RNA virus with a genome of approximately 13 kb. Based on similarities in genome organization and protein expression, the arteriviruses have recently been grouped together with the coronaviruses and toroviruses in the newly established order Nidovirales. Previously, we reported the construction of pEDI, a full-length cDNA copy of EAV DI-b, a natural defective interfering (DI) RNA of 5.6 kb (R. Molenkamp et al., J. Virol. 74:3156-3165, 2000). EDI RNA consists of three noncontiguous parts of the EAV genome fused in frame with respect to the replicase gene. As a result, EDI RNA contains a truncated replicase open reading frame (EDI-ORF) and encodes a truncated replicase polyprotein. Since some coronavirus DI RNAs require the presence of an ORF for their efficient propagation, we have analyzed the importance of the EDI-ORF in EDI RNA replication. The EDI-ORF was disrupted at different positions by the introduction of frameshift mutations. These were found either to block DI RNA replication completely or to be removed within one virus passage, probably due to homologous recombination with the helper virus genome. Using recombination assays based on EDI RNA and full-length EAV genomes containing specific mutations, the rates of homologous RNA recombination in the 3'- and 5'-proximal regions of the EAV genome were studied. Remarkably, the recombination frequency in the 5'-proximal region was found to be approximately 100-fold lower than that in the 3'-proximal part of the genome.  相似文献   

9.
RNA recombination plays an important role in the diversification and evolution of RNA viruses. Most of these events are believed to be mediated by an actively copying viral replicase switching from a donor template to an acceptor template, where it resumes synthesis. In addition, intramolecular replicase-mediated events (i.e., rearrangements) can lead to the generation of replicable deleted forms of a viral genome, termed defective interfering (DI) RNAs. To gain further insight into the recombination process, the effect of various primary and secondary structures on recombination site selection in vivo was examined using plant RNA tombusviruses. The effect of sequence identity and complementarity on deletion events that generate DI RNAs was also investigated. Our results suggest that (1) 5' termini and strong hairpin structures in donor templates represent preferred sites for recombinations, (2) junction sites in acceptor templates do not occur in double-stranded regions, (3) nucleotide homology can shift donor and acceptor recombination sites closer to regions of identity and, (4) both sequence identity and complementarity can direct deletion sites in DI RNAs. These results further define RNA determinants of tombusvirus RNA recombination and rearrangement.  相似文献   

10.
11.
Previously we demonstrated frequent homologous crossovers among molecules of the RNA3 segment in the tripartite brome mosaic bromovirus (BMV) RNA genome (A. Bruyere, M. Wantroba, S. Flasinski, A. Dzianott, and J. J. Bujarski, J. Virol. 74:4214-4219, 2000). To further our knowledge about mechanisms of viral RNA genome variability, in this paper we have studied homologous recombination in BMV RNA1 and RNA2 components during infection. We have found that basal RNA-RNA crossovers could occur within coding regions of both RNAs, although recombination frequencies slightly varied at different RNA sections. In all cases, the frequencies were much lower than the rate observed for the intercistronic recombination hot spot in BMV RNA3. Probability calculations accounted for at least one homologous crossover per RNA molecule per replication cycle. In addition, we have demonstrated an efficient repair of mutations within the conserved 3' and 5' noncoding regions, most likely due to error-prone BMV RNA replication. Overall, our data verify that homologous crossovers are common events a during virus life cycle, and we discuss their importance for viral RNA genetics.  相似文献   

12.
Shapka N  Nagy PD 《Journal of virology》2004,78(5):2288-2300
RNA recombination can be facilitated by recombination signals present in viral RNAs. Among such signals are short sequences with high AU contents that constitute recombination hot spots in Brome mosaic virus (BMV) and retroviruses. In this paper, we demonstrate that a defective interfering (DI) RNA, a model template associated with Tomato bushy stunt virus (TBSV), a tombusvirus, undergoes frequent recombination in plants and protoplast cells when it carries the AU-rich hot spot sequence from BMV. Similar to the situation with BMV, most of the recombination junction sites in the DI RNA recombinants were found within the AU-rich region. However, unlike BMV or retroviruses, where recombination usually occurred with precision between duplicated AU-rich sequences, the majority of TBSV DI RNA recombinants were imprecise. In addition, only one copy of the AU-rich sequence was essential to promote recombination in the DI RNA. The selection of junction sites was also influenced by a putative cis-acting element present in the DI RNA. We found that this RNA sequence bound to the TBSV replicase proteins more efficiently than did control nonviral sequences, suggesting that it might be involved in replicase "landing" during the template switching events. In summary, evidence is presented that a tombusvirus can use the recombination signal of BMV. This supports the idea that common AU-rich recombination signals might promote interviral recombination between unrelated viruses.  相似文献   

13.
Cheng CP  Nagy PD 《Journal of virology》2003,77(22):12033-12047
RNA recombination occurs frequently during replication of tombusviruses and carmoviruses, which are related small plus-sense RNA viruses of plants. The most common recombinants generated by these viruses are either defective interfering (DI) RNAs or chimeric satellite RNAs, which are thought to be generated by template switching of the viral RNA-dependent RNA polymerase (RdRp) during the viral replication process. To test if RNA recombination is mediated by the viral RdRp, we used either a purified recombinant RdRp of Turnip crinkle carmovirus or a partially purified RdRp preparation of Cucumber necrosis tombusvirus. We demonstrated that these RdRp preparations generated RNA recombinants in vitro. The RdRp-driven template switching events occurred between either identical templates or two different RNA templates. The template containing a replication enhancer recombined more efficiently than templates containing artificial sequences. We also observed that AU-rich sequences promote recombination more efficiently than GC-rich sequences. Cloning and sequencing of the generated recombinants revealed that the junction sites were located frequently at the ends of the templates (end-to-end template switching). We also found several recombinants that were generated by template switching involving internal positions in the RNA templates. In contrast, RNA ligation-based RNA recombination was not detected in vitro. Demonstration of the ability of carmo- and tombusvirus RdRps to switch RNA templates in vitro supports the copy-choice models of RNA recombination and DI RNA formation for these viruses.  相似文献   

14.
15.
16.
F Qu  T J Morris 《Journal of virology》1997,71(2):1428-1435
A protoplast infection assay has been used to reliably examine the viral RNA encapsidation of turnip crinkle virus (TCV). Analysis of the encapsidation of various mutant viral RNAs revealed that a 186-nucleotide (nt) region at the 3' end of the coat protein (CP) gene, with a bulged hairpin loop of 28 nt as its most essential element, was indispensable for TCV RNA encapsidation. When RNA fragments containing the 186-nt region were used to replace the CP gene of a different virus, tomato bushy stunt virus, the resulting chimeric viral RNAs were encapsidated into TCV virions. Furthermore, analysis of the encapsidated chimeric RNA species established that the RNA size was an important determinant of the TCV assembly process.  相似文献   

17.
The satellite RNA of tobacco ringspot virus (STobRV RNA) replicates and becomes encapsidated in association with tobacco ringspot virus. Previous results show that the infected tissue produces multimeric STobRV RNAs of both polarities. RNA that is complementary to encapsidated STobRV RNA, designated as having the (-) polarity, cleaves autolytically at a specific ApG bond. Purified autolysis products spontaneously join in a non-enzymic reaction. We report characteristics of this RNA ligation reaction: the terminal groups that react, the type of bond in the newly formed junction and the nucleotide sequence of the joined RNA. The nucleotide sequence of the ligated RNA shows that joining of the reacting RNAs restored an ApG bond. The junction ApG has a 3'-to-5' phosphodiester bond. Thus the net ligation reaction of STobRV (-)RNA is the precise reversal of autolysis. We discuss this new type of RNA ligation reaction and its implications for the formation of multimeric STobRV RNAs during replication.  相似文献   

18.
Picornavirus positive-strand RNAs are selectively encapsidated despite the coexistence of viral negative-strand RNAs and cellular RNAs in infected cells. However, the precise mechanism of the RNA encapsidation process in picornaviruses remains unclear. Here we report the first identification of an RNA element critical for encapsidation in picornaviruses. The 5' end of the genome of Aichi virus, a member of the family Picornaviridae, folds into three stem-loop structures (SL-A, SL-B, and SL-C, from the most 5' end). In the previous study, we constructed a mutant, termed mut6, by exchanging the seven-nucleotide stretches of the middle part of the stem in SL-A with each other to maintain the base pairings of the stem. mut6 exhibited efficient RNA replication and translation but formed no plaques. The present study showed that in cells transfected with mut6 RNA, empty capsids were accumulated, but few virions containing RNA were formed. This means that mut6 has a severe defect in RNA encapsidation. Site-directed mutational analysis indicated that as the mutated region was narrowed, the encapsidation was improved. As a result, the mutation of the 7 bp of the middle part of the stem in SL-A was required for abolishing the plaque-forming ability. Thus, the 5'-end sequence of the Aichi virus genome was shown to play an important role in encapsidation.  相似文献   

19.
Cis-acting RNA signals are required for replication of positive-strand viruses such as the picornaviruses. Although these generally have been mapped to the 5' and/or 3' termini of the viral genome, RNAs derived from human rhinovirus type 14 are unable to replicate unless they contain an internal cis-acting replication element (cre) located within the genome segment encoding the capsid proteins. Here, we show that the essential cre sequence is 83-96 nt in length and located between nt 2318-2413 of the genome. Using dicistronic RNAs in which translation of the P1 and P2-P3 segments of the polyprotein were functionally dissociated, we further demonstrate that translation of the cre sequence is not required for RNA replication. Thus, although it is located within a protein-coding segment of the genome, the cre functions as an RNA entity. Computer folds suggested that cre sequences could form a stable structure in either positive- or minus-strand RNA. However, an analysis of mutant RNAs containing multiple covariant and non-covariant nucleotide substitutions within these putative structures demonstrated that only the predicted positive-strand structure is essential for efficient RNA replication. The absence of detectable minus-strand synthesis from RNAs that lack the cre suggests that the cre is required for initiation of minus-strand RNA synthesis. Since a lethal 3' noncoding region mutation could be partially rescued by a compensating mutation within the cre, the cre appears to participate in a long-range RNA-RNA interaction required for this process. These data provide novel insight into the mechanisms of replication of a positive-strand RNA virus, as they define the involvement of an internally located RNA structure in the recognition of viral RNA by the viral replicase complex. Since internally located RNA replication signals have been shown to exist in several other positive-strand RNA virus families, these observations are potentially relevant to a wide array of related viruses.  相似文献   

20.
A duck hepatitis B virus (DHBV) genome cloned from a domestic duck from the People's Republic of China has been sequenced and exhibits no variation in sequences known to be important in viral replication or generation of gene products. Intrahepatic transfection of a dimer of this viral genome into ducklings did not result in viremia or any sign of virus infection, indicating that the genome was defective. Functional analysis of this mutant genome, performed by transfecting the DNA into a chicken hepatoma cell line capable of replicating wild-type virus, indicated that viral RNA is not encapsidated. However, virus core protein is made and can assemble into particles in the absence of encapsidation of viral nucleic acid. Using genetic approaches, it was determined that a change of cysteine to tyrosine in position 711 in the polymerase (P) gene C terminus led to this RNA-packaging defect. By site-directed mutagenesis, it was found that while substitution of Cys-711 with tryptophan also abolished packaging, substitution with methionine did not affect packaging or viral replication. Therefore, Cys-711, which is conserved in all published sequences of DHBV, may not be involved in a disulfide bridge structure essential to viral RNA packaging or replication. Our results, showing that a missense mutation in the region of the DHBV polymerase protein thought to be primarily the RNase H domain results in packaging deficiency, support the previous findings that multiple regions of the complex hepadnaviral polymerase protein may be required for viral RNA packaging.  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