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 78 毫秒
1.
The occurrence of genome segment reassortment between two antigenically related orbiviruses was demonstrated in cattle. Individual virus clones were isolated by cell culture plaque assays directly from the blood of a calf infected with two serotypes of bluetongue virus. The majority (89%) of progeny viruses isolated from the calf represented reassortant viruses. A minimum of six genome segments participated in reassortment, with 16 unique reassortant constellations being identified. Such genome segment reassortment between unique, though antigenically related, orbiviruses has undoubtedly played a major role in generating the extensive phenotypic and genotypic diversity that is characteristic of this serogroup.  相似文献   

2.
Six orbiviruses, Corriparta, Wallal, Eubenangee, D'Aguilar, Warrego and Mitchell River, were inactivated within 30 min in solutions of pH less than 6. All but Wallal virus resisted treatment with ether at 1 degrees for 22 h but only Corriparta virus resisted treatment with chloroform at 1 degrees for 22 h. The genomes of the orbiviruses, including Corriparta virus, after electrophoresis in acrylamide gels separated into basically similar patterns which were distinct from the pattern of reovirus RNA.  相似文献   

3.
Viruses belonging to the species Wallal virus and Warrego virus of the genus Orbivirus were identified as causative agents of blindness in marsupials in Australia during 1994/5. Recent comparisons of nucleotide (nt) and amino acid (aa) sequences have provided a basis for the grouping and classification of orbivirus isolates. However, full-genome sequence data are not available for representatives of all Orbivirus species. We report full-genome sequence data for three additional orbiviruses: Wallal virus (WALV); Mudjinabarry virus (MUDV) and Warrego virus (WARV). Comparisons of conserved polymerase (Pol), sub-core-shell ‘T2’ and core-surface ‘T13’ proteins show that these viruses group with other Culicoides borne orbiviruses, clustering with Eubenangee virus (EUBV), another orbivirus infecting marsupials. WARV shares <70% aa identity in all three conserved proteins (Pol, T2 and T13) with other orbiviruses, consistent with its classification within a distinct Orbivirus species. Although WALV and MUDV share <72.86%/67.93% aa/nt identity with other orbiviruses in Pol, T2 and T13, they share >99%/90% aa/nt identities with each other (consistent with membership of the same virus species - Wallal virus). However, WALV and MUDV share <68% aa identity in their larger outer capsid protein VP2(OC1), consistent with membership of different serotypes within the species - WALV-1 and WALV-2 respectively.  相似文献   

4.
<正>Dear Editor,The genus Orbivirus,within the family Reoviridae,includes 22 virus species(King et al.,2011).They are distributed globally,but are particularly prevalent in Europe,Asia,and Africa.In addition,they can be transmitted by ticks or other hematophagous insect vectors,including Culicoides,mosquitoes,and sandflies(Belaga-  相似文献   

5.
Four rotavirus SA11 temperature-sensitive (ts) mutants and seven rotavirus RRV ts mutants, isolated at the National Institutes of Health (NIH) and not genetically characterized, were assigned to reassortment groups by pairwise crosses with the SA11 mutant group prototypes isolated and characterized at Baylor College of Medicine (BCM). Among the NIH mutants, three of the RRV mutants and all four SA11 mutants contained mutations in single reassortment groups, and four RRV mutants contained mutations in multiple groups. One NIH mutant [RRVtsK(2)] identified the previously undefined 11th reassortment group (K) expected for rotavirus. Three NIH single mutant RRV viruses, RRVtsD(7), RRVtsJ(5), and RRVtsK(2), were in reassortment groups not previously mapped to genome segments. These mutants were mapped using classical genetic methods, including backcrosses to demonstrate reversion or suppression in reassortants with incongruent genotype and temperature phenotype. Once located to specific genome segments by genetic means, the mutations responsible for the ts phenotype were identified by sequencing. The reassortment group K mutant RRVtsK(2) maps to genome segment 9 and has a Thr280Ileu mutation in the capsid surface glycoprotein VP7. The group D mutant RRVtsD(7) maps to segment 5 and has a Leu140Val mutation in the nonstructural interferon (IFN) antagonist protein NSP1. The group J mutant RRVtsJ(5) maps to segment 11 and has an Ala182Gly mutation affecting only the NSP5 open reading frame. Rotavirus ts mutation groups are now mapped to 9 of the 11 rotavirus genome segments. Possible segment locations of the two remaining unmapped ts mutant groups are discussed.  相似文献   

6.
In early 2009, new swine-origin influenza A (H1N1) virus emerged in Mexico and the United States. The emerging influenza virus had made global influenza pandemic for nearly one year. To every emerging pathogen, exploring the origin sources is vital for viral control and clearance. Influenza virus is different from other virus in that it has 8 segments, making the segment reassortment a main drive in virus evolution. In exploring reassortment evolution origins of a newly emerging influenza virus, integrated comparing of the origin sources of all the segments is necessary. If some segments have high homologous with one parental strain, lower homologous with another parental strain, while other segments are reverse, can we proposed that this emerging influenza virus may re-assort from the two parental strains. Here we try to explore the multilevel reassortment evolution origins of 2009 H1N1 influenza virus using this method. By further validating the fidelity of this strategy, this method might be useful in judging the reassortment origins of newly emerging influenza virus.  相似文献   

7.
The evolutionary success of La Crosse virus (LACV, family Bunyaviridae) is due to its ability to adapt to changing conditions through intramolecular genetic changes and segment reassortment. Vertical transmission of LACV in mosquitoes increases the potential for segment reassortment. Studies were conducted to determine if segment reassortment was occurring in naturally infected Aedes triseriatus from Wisconsin and Minnesota in 2000, 2004, 2006 and 2007. Mosquito eggs were collected from various sites in Wisconsin and Minnesota. They were reared in the laboratory and adults were tested for LACV antigen by immunofluorescence assay. RNA was isolated from the abdomen of infected mosquitoes and portions of the small (S), medium (M) and large (L) viral genome segments were amplified by RT-PCR and sequenced. Overall, the viral sequences from 40 infected mosquitoes and 5 virus isolates were analyzed. Phylogenetic and linkage disequilibrium analyses revealed that approximately 25% of infected mosquitoes and viruses contained reassorted genome segments, suggesting that LACV segment reassortment is frequent in nature.  相似文献   

8.
Coinfection of a cell by two different strains of a segmented virus can give rise to a “reassortant” with phenotypic characteristics that might differ from those of the parental strains. Bluetongue virus (BTV) is a double-stranded RNA (dsRNA) segmented virus and the cause of bluetongue, a major infectious disease of livestock. BTV exists as at least 26 different serotypes (BTV-1 to BTV-26). Prompted by the isolation of a field reassortant between BTV-1 and BTV-8, we systematically characterized the process of BTV reassortment. Using a reverse genetics approach, our study clearly indicates that any BTV-1 or BTV-8 genome segment can be rescued in the heterologous “backbone.” To assess phenotypic variation as a result of reassortment, we examined viral growth kinetics and plaque sizes in in vitro experiments and virulence in an experimental mouse model of bluetongue disease. The monoreassortants generated had phenotypes that were very similar to those of the parental wild-type strains both in vitro and in vivo. Using a forward genetics approach in cells coinfected with BTV-1 and BTV-8, we have shown that reassortants between BTV-1 and BTV-8 are generated very readily. After only four passages in cell culture, we could not detect wild-type BTV-1 or BTV-8 in any of 140 isolated viral plaques. In addition, most of the isolated reassortants contained heterologous VP2 and VP5 structural proteins, while only 17% had homologous VP2 and VP5 proteins. Our study has shown that reassortment in BTV is very flexible, and there is no fundamental barrier to the reassortment of any genome segment. Given the propensity of BTV to reassort, it is increasingly important to have an alternative classification system for orbiviruses.  相似文献   

9.
Preliminary characterization of naturally occurring temperature-sensitive bovine rotavirus mutants, generated via gene reassortment, indicated that the product of gene segment 10 can influence virus adsorption to MA-104 cells. Cell-free translation of this gene segment in a rabbit reticulocyte system was then utilized to identify a polypeptide with a molecular weight of ca. 21,500 as the primary translation product. This identification provides a basis for further studies of rotavirus assembly and cell interactions which may, in turn, aid in elucidating the replicative strategy of the virus.  相似文献   

10.
Genetic exchange by a process of genome-segment ‘reassortment’ represents an important mechanism for evolutionary change in all viruses with segmented genomes, yet in many cases a detailed understanding of its frequency and biological consequences is lacking. We provide a comprehensive assessment of reassortment in bluetongue virus (BTV), a globally important insect-borne pathogen of livestock, during recent outbreaks in Europe. Full-genome sequences were generated and analysed for over 150 isolates belonging to the different BTV serotypes that have emerged in the region over the last 5 decades. Based on this novel dataset we confirm that reassortment is a frequent process that plays an important and on-going role in evolution of the virus. We found evidence for reassortment in all ten segments without a significant bias towards any particular segment. However, we observed biases in the relative frequency at which particular segments were associated with each other during reassortment. This points to selective constraints possibly caused by functional relationships between individual proteins or genome segments and genome-wide epistatic interactions. Sites under positive selection were more likely to undergo amino acid changes in newly reassorted viruses, providing additional evidence for adaptive dynamics as a consequence of reassortment. We show that the live attenuated vaccines recently used in Europe have repeatedly reassorted with field strains, contributing to their genotypic, and potentially phenotypic, variability. The high degree of plasticity seen in the BTV genome in terms of segment origin suggests that current classification schemes that are based primarily on serotype, which is determined by only a single genome segment, are inadequate. Our work highlights the need for a better understanding of the mechanisms and epidemiological consequences of reassortment in BTV, as well as other segmented RNA viruses.  相似文献   

11.

Background

All viruses in the family Bunyaviridae possess a tripartite genome, consisting of a small, a medium, and a large RNA segment. Bunyaviruses therefore possess considerable evolutionary potential, attributable to both intramolecular changes and to genome segment reassortment. Hantaviruses (family Bunyaviridae, genus Hantavirus) are known to cause human hemorrhagic fever with renal syndrome or hantavirus pulmonary syndrome. The primary reservoir host of Sin Nombre virus is the deer mouse (Peromyscus maniculatus), which is widely distributed in North America. We investigated the prevalence of intramolecular changes and of genomic reassortment among Sin Nombre viruses detected in deer mice in three western states.

Methods

Portions of the Sin Nombre virus small (S) and medium (M) RNA segments were amplified by RT-PCR from kidney, lung, liver and spleen of seropositive peromyscine rodents, principally deer mice, collected in Colorado, New Mexico and Montana from 1995 to 2007. Both a 142 nucleotide (nt) amplicon of the M segment, encoding a portion of the G2 transmembrane glycoprotein, and a 751 nt amplicon of the S segment, encoding part of the nucleocapsid protein, were cloned and sequenced from 19 deer mice and from one brush mouse (P. boylii), S RNA but not M RNA from one deer mouse, and M RNA but not S RNA from another deer mouse.

Results

Two of 20 viruses were found to be reassortants. Within virus sequences from different rodents, the average rate of synonymous substitutions among all pair-wise comparisons (πs) was 0.378 in the M segment and 0.312 in the S segment sequences. The replacement substitution rate (πa) was 7.0 × 10-4 in the M segment and 17.3 × 10-4 in the S segment sequences. The low πa relative to πs suggests strong purifying selection and this was confirmed by a Fu and Li analysis. The absolute rate of molecular evolution of the M segment was 6.76 × 10-3 substitutions/site/year. The absolute age of the M segment tree was estimated to be 37 years. In the S segment the rate of molecular evolution was 1.93 × 10-3 substitutions/site/year and the absolute age of the tree was 106 years. Assuming that mice were infected with a single Sin Nombre virus genotype, phylogenetic analyses revealed that 10% (2/20) of viruses were reassortants, similar to the 14% (6/43) found in a previous report.

Conclusion

Age estimates from both segments suggest that Sin Nombre virus has evolved within the past 37–106 years. The rates of evolutionary changes reported here suggest that Sin Nombre virus M and S segment reassortment occurs frequently in nature.  相似文献   

12.
Reassortment is fundamental to the evolution of influenza viruses and plays a key role in the generation of epidemiologically significant strains. Previous studies indicate that reassortment is restricted by segment mismatch, arising from functional incompatibilities among components of two viruses. Additional factors that dictate the efficiency of reassortment remain poorly characterized. Thus, it is unclear what conditions are favorable for reassortment and therefore under what circumstances novel influenza A viruses might arise in nature. Herein, we describe a system for studying reassortment in the absence of segment mismatch and exploit this system to determine the baseline efficiency of reassortment and the effects of infection dose and timing. Silent mutations were introduced into A/Panama/2007/99 virus such that high-resolution melt analysis could be used to differentiate all eight segments of the wild-type and the silently mutated variant virus. The use of phenotypically identical parent viruses ensured that all progeny were equally fit, allowing reassortment to be measured without selection bias. Using this system, we found that reassortment occurred efficiently (88.4%) following high multiplicity infection, suggesting the process is not appreciably limited by intracellular compartmentalization. That co-infection is the major determinant of reassortment efficiency in the absence of segment mismatch was confirmed with the observation that the proportion of viruses with reassortant genotypes increased exponentially with the proportion of cells co-infected. The number of reassortants shed from co-infected guinea pigs was likewise dependent on dose. With 106 PFU inocula, 46%–86% of viruses isolated from guinea pigs were reassortants. The introduction of a delay between infections also had a strong impact on reassortment and allowed definition of time windows during which super-infection led to reassortment in culture and in vivo. Overall, our results indicate that reassortment between two like influenza viruses is efficient but also strongly dependent on dose and timing of the infections.  相似文献   

13.
Wild populations of the critically endangered woylie (Bettongia penicillata) recently declined by 90% in southwest Western Australia. Increased predation is the leading hypothesis for decline, but disease may be playing a role increasing susceptibility to predation. To explore this possibility, we surveyed woylie populations in the wild, in captivity and in a predator-free sanctuary for exposure to, and infection with, four known pathogens of macropods: herpesviruses, Wallal and Warrego orbiviruses, and Toxoplasma gondii. Our study found two of 68 individuals positive for neutralizing antibodies against known macropodid alphaherpesviruses. Three of 45 individuals were PCR positive for a herpesvirus that was shown to be a novel gammaherpesvirus or a new strain/variant of Potoroid Herpesvirus 1. Further sequence information is required to definitively determine its correct classification. There was no evidence of antibodies to orbivirus Wallal and Warrego serogroups, and all serological samples tested for T. gondii were negative. This is the first report of PCR and serological detection of herpesviruses in the woylie. Positive individuals did not demonstrate clinical signs of herpesviral diseases; therefore, the clinical significance of herpesviruses to wild woylie populations remains unclear. Further monitoring for herpesvirus infections will be important to inform disease risk analysis for this virus and determine temporal trends in herpesvirus activity that may relate to population health and conservation outcomes.  相似文献   

14.
A high particle to infectivity ratio is a feature common to many RNA viruses, with ~90–99% of particles unable to initiate a productive infection under low multiplicity conditions. A recent publication by Brooke et al. revealed that, for influenza A virus (IAV), a proportion of these seemingly non-infectious particles are in fact semi-infectious. Semi-infectious (SI) particles deliver an incomplete set of viral genes to the cell, and therefore cannot support a full cycle of replication unless complemented through co-infection. In addition to SI particles, IAV populations often contain defective-interfering (DI) particles, which actively interfere with production of infectious progeny. With the aim of understanding the significance to viral evolution of these incomplete particles, we tested the hypothesis that SI and DI particles promote diversification through reassortment. Our approach combined computational simulations with experimental determination of infection, co-infection and reassortment levels following co-inoculation of cultured cells with two distinct influenza A/Panama/2007/99 (H3N2)-based viruses. Computational results predicted enhanced reassortment at a given % infection or multiplicity of infection with increasing semi-infectious particle content. Comparison of experimental data to the model indicated that the likelihood that a given segment is missing varies among the segments and that most particles fail to deliver ≥1 segment. To verify the prediction that SI particles augment reassortment, we performed co-infections using viruses exposed to low dose UV. As expected, the introduction of semi-infectious particles with UV-induced lesions enhanced reassortment. In contrast to SI particles, inclusion of DI particles in modeled virus populations could not account for observed reassortment outcomes. DI particles were furthermore found experimentally to suppress detectable reassortment, relative to that seen with standard virus stocks, most likely by interfering with production of infectious progeny from co-infected cells. These data indicate that semi-infectious particles increase the rate of reassortment and may therefore accelerate adaptive evolution of IAV.  相似文献   

15.
Two seronegative sheep were infected intravenously with 10(9) PFU each of bluetongue virus (BTV) serotype 10 and BTV serotype 17. One animal experienced a mild bluetongue-like disease, and both experienced a short-duration viremia and developed neutralizing immune responses to both virus serotypes. Progeny virus was isolated from venous blood from each animal by using conditions in which reassortment could not have occurred during isolation. Electropherotypes were determined for the progeny viruses from the infected sheep, yielding strikingly similar results for the two animals. In both sheep, serotype 10 dominated among the progeny, accounting for 92% of the progeny. Serotype 17 was rarely isolated and accounted for 3% of the progeny analyzed. The remaining 5% of the progeny clones were reassortant and derived genome segments from both serotypes 10 and 17. Analysis of the parental origin of genome segments in the small number of reassortant progeny analyzed suggested that selection of specific genome segments may have occurred in the infected sheep. These data indicate that reassortment of genome segments occurs, at low frequency, in sheep mixedly infected with BTV.  相似文献   

16.
Severe fever with thrombocytopenia syndrome virus (SFTSV) was a novel tick-borne bunyavirus that caused hemorrhagic fever with a high fatality rate in East Asia. In this study we analyzed the complete genome sequences of 122 SFTSV strains to determine the phylogeny, evolution and reassortment of the virus. We revealed that the evolutionary rate of three genome segments were different, with highest in the S segment and lowest in the L segment. The SFTSV strains were phylogenetically classified into 5 lineages (A, B, C, D and E) with each genome segment. SFTSV strains from China were classified in all 5 lineages, strains from South Korea were classified into 3 lineages (A, D, and E), and all strains from Japan were classified in only linage E. Using the average evolutionary rate of the three genome segments, we found that the extant SFTSV originated 20–87 years ago in the Dabie Mountain area in central China. The viruses were then transmitted to other areas of China, Japan and South Korea. We also found that six SFTSV strains were reassortants. Selection pressure analysis suggested that SFTSV was under purifying selection according to the four genes (RNA-dependent RNA polymerase, glycoprotein, nucleocapsid protein, non-structural protein), and two sites (37, 1033) of glycoproteins were identified as being under strong positive selection. We concluded that SFTSV originated in central China and spread to other places recently and the virus was under purifying selection with high frequency of reassortment.  相似文献   

17.
The genetic structure of natural bacteriophage populations is poorly understood. Recent metagenomic studies suggest that phage biogeography is characterized by frequent migration. Using virus samples mostly isolated in Southern California, we recently showed that very little population structure exists in segmented RNA phage of the Cystoviridae family due to frequent segment reassortment (sexual genetic mixis) between unrelated virus individuals. Here we use a larger genetic dataset to examine the structure of Cystoviridae phage isolated from three geographic locations in Southern New England. We document extensive natural variation in the physical sizes of RNA genome segments for these viruses. In addition, consistent with earlier findings, our phylogenetic analyses and calculations of linkage disequilibrium (LD) show no evidence of within‐segment recombination in wild populations. However, in contrast to the prior study, our analysis finds that reassortment of segments between individual phage plays a lesser role among cystoviruses sampled in New England, suggesting that the evolutionary importance of genetic mixis in Cystoviridae phage may vary according to geography. We discuss possible explanations for these conflicting results across the studies, such as differing local ecology and its impact on phage growth, and geographic differences in selection against hybrid phage genotypes.  相似文献   

18.
Bluetongue virus (BTV) is a midge-borne member of the genus Orbivirus that causes an eponymous debilitating livestock disease of great agricultural impact and which has expanded into Europe in recent decades. Reassortment among the ten segments comprising the double-stranded (ds) RNA genome of BTV has played an important role in generating the epidemic strains of this virus in Europe. In this study, we investigated the dynamics of BTV genome segment evolution utilizing time-structured data sets of complete sequences from four segments, totalling 290 sequences largely sampled from ruminant hosts. Our analysis revealed that BTV genome segments generally evolve under strong purifying selection and at substitution rates that are generally lower (mean rates of ~0.5–7 × 10−4 nucleotide substitutions per site, per year) than vector-borne positive-sense viruses with single-strand (ss) RNA genomes. These also represent the most robust estimates of the nucleotide substitution rate in a dsRNA virus generated to date. Additionally, we determined that patterns of geographic structure and times to most recent common ancestor differ substantially between each segment, including a relatively recent origin for the diversity of segment 10 within the past millennium. Together, these findings demonstrate the effect of reassortment to decouple the evolutionary dynamics of BTV genome segments.  相似文献   

19.
Although the ferret model has been extensively used to study pathogenesis and transmission of influenza viruses, little has been done to determine whether ferrets are a good surrogate animal model to study influenza virus reassortment. It has been previously shown that the pandemic 2009 H1N1 (H1N1pdm) virus was able to transmit efficiently in ferrets. In coinfection studies with either seasonal H1N1 or H3N2 strains (H1N1s or H3N2s, respectively), the H1N1pdm virus was able to outcompete these strains and become the dominant transmissible virus. However, lack of reassortment could have been the result of differences in the cell or tissue tropism of these viruses in the ferret. To address this issue, we performed coinfection studies with recombinant influenza viruses carrying the surface genes of a seasonal H3N2 strain in the background of an H1N1pdm strain and vice versa. After serial passages in ferrets, a dominant H1N2 virus population was obtained with a constellation of gene segments, most of which, except for the neuraminidase (NA) and PB1 segments, were from the H1N1pdm strain. Our studies suggest that ferrets recapitulate influenza virus reassortment events. The H1N2 virus generated through this process resembles similar viruses that are emerging in nature, particularly in pigs.  相似文献   

20.
Reassortments and point mutations are two major contributors to diversity of Influenza A virus; however, the link between these two processes is unclear. It has been suggested that reassortments provoke a temporary increase in the rate of amino acid changes as the viral proteins adapt to new genetic environment, but this phenomenon has not been studied systematically. Here, we use a phylogenetic approach to infer the reassortment events between the 8 segments of influenza A H3N2 virus since its emergence in humans in 1968. We then study the amino acid replacements that occurred in genes encoded in each segment subsequent to reassortments. In five out of eight genes (NA, M1, HA, PB1 and NS1), the reassortment events led to a transient increase in the rate of amino acid replacements on the descendant phylogenetic branches. In NA and HA, the replacements following reassortments were enriched with parallel and/or reversing replacements; in contrast, the replacements at sites responsible for differences between antigenic clusters (in HA) and at sites under positive selection (in NA) were underrepresented among them. Post-reassortment adaptive walks contribute to adaptive evolution in Influenza A: in NA, an average reassortment event causes at least 2.1 amino acid replacements in a reassorted gene, with, on average, 0.43 amino acid replacements per evolving post-reassortment lineage; and at least ∼9% of all amino acid replacements are provoked by reassortments.  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