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.
Cytosolic chaperones are a diverse group of ubiquitous proteins that play central roles in multiple processes within the cell, including protein translation, folding, intracellular trafficking, and quality control. These cellular proteins have also been implicated in the replication of numerous viruses, although the full extent of their involvement in viral replication is unknown. We have previously shown that the heat shock protein 40 (hsp40) chaperone encoded by the yeast YDJ1 gene facilitates RNA replication of flock house virus (FHV), a well-studied and versatile positive-sense RNA model virus. To further explore the roles of chaperones in FHV replication, we examined a panel of 30 yeast strains with single deletions of cytosolic proteins that have known or hypothesized chaperone activity. We found that the majority of cytosolic chaperone deletions had no impact on FHV RNA accumulation, with the notable exception of J-domain-containing hsp40 chaperones, where deletion of APJ1 reduced FHV RNA accumulation by 60%, while deletion of ZUO1, JJJ1, or JJJ2 markedly increased FHV RNA accumulation, by 4- to 40-fold. Further studies using cross complementation and double-deletion strains revealed that the contrasting effects of J domain proteins were reproduced by altering expression of the major cytosolic hsp70s encoded by the SSA and SSB families and were mediated in part by divergent effects on FHV RNA polymerase synthesis. These results identify hsp70 chaperones as critical regulators of FHV RNA replication and indicate that cellular chaperones can have both positive and negative regulatory effects on virus replication.The compact genomes of viruses relative to those of other infectious agents restrict their ability to encode all proteins required to complete their replication cycles. To circumvent this limitation, viruses often utilize cellular factors or processes to complete essential steps in replication. One group of cellular proteins frequently targeted by viruses are cellular chaperones, which include a diverse set of heat shock proteins (hsps) that normally facilitate cellular protein translation, folding, trafficking, and degradation (18, 64). The connection between viruses and cellular chaperones was originally identified in bacteria, where the Escherichia coli hsp40 and hsp70 homologues, encoded by dnaJ and dnaK, respectively, were identified as bacterial genes essential for bacteriophage λ DNA replication (62). Research over the past 30 years has further revealed the importance of cellular chaperones in viral replication, such that the list of virus-hsp connections is now quite extensive and includes viruses from numerous families with diverse genome structures (4, 6, 7, 16, 19, 20, 23, 25, 40, 41, 44, 51, 54, 60). These studies have demonstrated the importance of cellular chaperones in multiple steps of the viral life cycle, including entry, viral protein translation, genome replication, encapsidation, and virion release. However, the list of virus-hsp connections is likely incomplete. Further studies to explore this particular host-pathogen interaction will shed light on virus replication mechanisms and pathogenesis, and potentially highlight targets for novel antiviral agents.To study the role of cellular chaperones in the genome replication of positive-sense RNA viruses, we use flock house virus (FHV), a natural insect pathogen and well-studied member of the Nodaviridae family. The FHV life cycle shares many common features with other positive-sense RNA viruses, including the membrane-specific targeting and assembly of functional RNA replication complexes (37, 38), the exploitation of various cellular processes and host factors for viral replication (5, 23, 60), and the induction of large-scale membrane rearrangements (24, 28, 38, 39). FHV virions contain a copackaged bipartite genome consisting of RNA1 (3.1 kb) and RNA2 (1.4 kb), which encode protein A, the viral RNA-dependent RNA polymerase, and the structural capsid protein precursor, respectively (1). During active genome replication, FHV produces a subgenomic RNA3 (0.4 kb), which encodes the RNA interference inhibitor protein B2 (12, 29, 32). These viral characteristics make FHV an excellent model system to study many aspects of positive-sense RNA virus biology.In addition to the benefits of a simple genome, FHV is able to establish robust RNA replication in a wide variety of genetically tractable eukaryotic hosts, including Drosophila melanogaster (38), Caenorhabditis elegans (32), and Saccharomyces cerevisiae (46). The budding yeast S. cerevisiae has been an exceptionally useful model host to study the mechanisms of viral RNA replication complex assembly and function with FHV (31, 37, 39, 45, 53, 55, 56, 60) as well as other positive-sense RNA viruses (11). The facile genetics of S. cerevisiae, along with the vast array of well-defined cellular and molecular tools and techniques, make it an ideal eukaryotic host for the identification of cellular factors required for positive-sense RNA virus replication. Furthermore, readily available yeast libraries with deletions and regulated expression of individual proteins have led to the completion of several high-throughput screens to provide a global survey of host factors that impact virus replication (26, 42, 52). An alternative approach with these yeast libraries that reduces the inherently high false-negative rates associated with high-throughput screens is to focus on a select set of host genes associated with a particular cellular pathway, process, or location previously implicated in virus replication.We have utilized such a targeted approach and focused on examining the impact of cytosolic chaperones on FHV RNA replication. Previously, we have shown that the cellular chaperone hsp90 facilitates protein A synthesis in Drosophila cells (5, 23), and the hsp40 encoded by the yeast YDJ1 gene facilitates FHV RNA replication in yeast, in part through effects on both protein A accumulation and function (60). In this report, we further extend these observations by examining FHV RNA accumulation in a panel of yeast strains with deletions of known or hypothesized cytosolic chaperones. We demonstrate that cytosolic chaperones can have either suppressive or enhancing effects on FHV RNA accumulation. In particular, related hsp70 members encoded by the SSA and SSB yeast chaperone families have marked and dramatically divergent effects on both genomic and subgenomic RNA accumulation and viral polymerase synthesis. These results highlight the complexities of the host-pathogen interactions that influence positive-sense RNA virus replication and identify the hsp70 family of cytosolic chaperones as key regulators of FHV replication.  相似文献   

2.
Hepatitis E virus (HEV) is a causative agent of acute hepatitis, and it is the sole member of the genus Hepevirus in the family Hepeviridae. The open reading frame 1 (ORF1) protein of HEV encodes nonstructural polyprotein with putative domains for methyltransferase, cysteine protease, helicase and RNA-dependent RNA polymerase. It is not yet known whether ORF1 functions as a single protein with multiple domains or is processed to form separate functional units. On the basis of amino acid conserved motifs, HEV helicase has been grouped into helicase superfamily 1 (SF-1). In order to examine the RNA helicase activity of the NTPase/helicase domain of HEV, the region (amino acids 960 to 1204) was cloned and expressed as histidine-tagged protein in Escherichia coli (HEV Hel) and purified. HEV Hel exhibited NTPase and RNA unwinding activities. Enzyme hydrolyzed all rNTPs efficiently, dATP and dCTP with moderate efficiency, while it showed less hydrolysis of dGTP and dTTP. Enzyme showed unwinding of only RNA duplexes with 5′ overhangs showing 5′-to-3′ polarity. We also expressed and purified two HEV Hel mutants. Helicase mutant I, with substitution in the nucleotide-binding motif I (GKS to GAS), showed 30% ATPase activity. Helicase mutant II, with substitutions in the Mg2+ binding motif II (DEAP to AAAP), showed 50% ATPase activity. Both mutants completely lost ability to unwind RNA duplexes with 5′ overhangs. These findings represent the first report demonstrating NTPase/RNA helicase activity of the helicase domain of HEV ORF1.Viruses with single-strand positive-sense RNA genomes represent the largest class of viruses, which includes numerous pathogens of humans, plants, and animals. In these viruses, RNA replication occurs through negative-strand RNA intermediate, which may also act as the template for synthesis of subgenomic RNAs in some viruses. During replication, various nonstructural proteins remain associated with the viral polymerase in a small compartmentalized replisome. Most of the other accessory proteins are obtained from the cellular machinery.Helicase seems to be essential for RNA replication by many positive-sense RNA viruses (19). Many positive-strand RNA viruses encode their own RNA helicases and besides RNA-dependent RNA polymerase, helicase is the most conserved viral sequence in these viruses. It has been shown by direct mutagenesis studies in poliovirus (26, 39), alphaviruses (31), brome mosaic virus (2, 41), nidoviruses (40), and flaviviruses (15) that helicase functions are essential for viral replication. In addition, it may be involved in RNA translocation, genome packaging, protection of RNA at the replication center, modulating RNA-protein interactions, etc.Helicases are classified into six superfamilies, SF-1 to SF-6 (11, 35), and can be classified further into subfamilies, A (3′→5′) or B (5′→3′) depending on their unwinding directionality. Classic helicases (exhibiting both NTPase and unwinding activities) are referred to as subtype α, while translocases (with no unwinding activity) are referred to as subtype β (35). SF-1 and SF-2 constitute largest of these superfamilies with seven signature motifs (I, Ia, II, III, IV, V, and VI), which form core of the enzyme. Although these motifs are not comparable between SF-1 and SF-2, universal features of core domains include (i) conserved residues involved in binding and hydrolysis of the NTP and (ii) an arginine finger that plays a key role in energy coupling.Hepatitis E virus (HEV) is a nonenveloped virus in the genus Hepevirus of the family Hepeviridae. Hepatitis E is an important public health disease in many developing countries and is also endemic in some industrialized countries (8). Infection by HEV has a known association with increased mortality during pregnancy (22, 23). HEV has a positive-sense RNA genome of ∼7.2 kb, consisting of a 5′ noncoding region (5′NCR) of 27 to 35 nucleotides (nt), followed by three open reading frames (ORFs)—ORF1, ORF2, and ORF3—and a 3′NCR of 65 to 74 nt, ending with a poly(A) tail of variable length (37). The 5′ end has m7G cap (18). ORF1 is known to encode for the viral nonstructural polyprotein with a proposed molecular mass of ∼186 kDa (3). Based on protein sequence homology, the ORF1 polyprotein is proposed to contain four putative domains indicative of methyltransferase, papain-like cysteine protease, RNA helicase (Hel), and RNA-dependent RNA polymerase (RdRp) (24). ORF2 encodes the major structural protein (capsid protein), which has N-terminal signal peptide and three glycosylation sites and is translocated across the endoplasmic reticulum (ER). ORF2 protein associates with the 5′ end of the viral RNA, suggesting its regulatory role in the virus replication (36, 37, 44, 45). ORF3 encodes a protein which gets phosphorylated by the cellular mitogen activated protein kinase and is associated with cellular membranes and cytoskeleton fractions (43).HEV belongs to an “alpha-like” supergroup of positive-sense single-stranded RNA (+ssRNA) viruses with conserved motifs of replication-related proteins in the ORF1, with typical signature sequences homologous with the other members of the family (11, 12, 13). ORF1 of HEV encodes additional domains such as the Y domain, papainlike protease, “proline-rich hinge,” and the X domain. Methyltransferase (25), RdRp (1), and X domain (binding to poly-ADP-ribose) (9) in ORF1 have been characterized, whereas the functions of the other domains are yet to be identified. Intracellularly expressed RdRp localizes itself in the ER membranes (30), suggesting that HEV replicates probably in ER in the cytosolic compartment of the cells. It is still unknown whether ORF1 polyprotein undergoes cleavages to form separate functional units of the replication machinery or functions as a single protein with multiple functional domains.The putative RNA helicase of HEV contains all of the seven conserved segments typical of the SF-1 helicase (12, 13). Putative SF-1 helicases are extremely widespread among +ssRNA viruses. Based on sequence comparisons, such helicases have been identified in a variety of plant virus families, as well as in animal viruses such as alphavirus, rubivirus, hepatitis E virus, and coronavirus (11). When compared to other +ssRNA viral helicases belonging to SF-1, HEV helicase showed the highest overall similarity with the helicase of beet necrotic yellow vein virus, a plant furovirus. HEV helicase was speculated to have N-terminal NTPase and C-terminal RNA-binding domains (24). A major obstacle in studying HEV replication has been lack of cell culture system. We report here experimental verification of the helicase activity of the recombinant helicase domain protein of HEV.  相似文献   

3.
4.
5.
6.
In their vertebrate hosts, arboviruses such as Semliki Forest virus (SFV) (Togaviridae) generally counteract innate defenses and trigger cell death. In contrast, in mosquito cells, following an early phase of efficient virus production, a persistent infection with low levels of virus production is established. Whether arboviruses counteract RNA interference (RNAi), which provides an important antiviral defense system in mosquitoes, is an important question. Here we show that in Aedes albopictus-derived mosquito cells, SFV cannot prevent the establishment of an antiviral RNAi response or prevent the spread of protective antiviral double-stranded RNA/small interfering RNA (siRNA) from cell to cell, which can inhibit the replication of incoming virus. The expression of tombusvirus siRNA-binding protein p19 by SFV strongly enhanced virus spread between cultured cells rather than virus replication in initially infected cells. Our results indicate that the spread of the RNAi signal contributes to limiting virus dissemination.In animals, RNA interference (RNAi) was first described for Caenorhabditis elegans (27). The production or introduction of double-stranded RNA (dsRNA) in cells leads to the degradation of mRNAs containing homologous sequences by sequence-specific cleavage of mRNAs. Central to RNAi is the production of 21- to 26-nucleotide small interfering RNAs (siRNAs) from dsRNA and the assembly of an RNA-induced silencing complex (RISC), followed by the degradation of the target mRNA (23, 84). RNAi is a known antiviral strategy of plants (3, 53) and insects (21, 39, 51). Study of Drosophila melanogaster in particular has given important insights into RNAi responses against pathogenic viruses and viral RNAi inhibitors (31, 54, 83, 86, 91). RNAi is well characterized for Drosophila, and orthologs of antiviral RNAi genes have been found in Aedes and Culex spp. (13, 63).Arboviruses, or arthropod-borne viruses, are RNA viruses mainly of the families Bunyaviridae, Flaviviridae, and Togaviridae. The genus Alphavirus within the family Togaviridae contains several mosquito-borne pathogens: arboviruses such as Chikungunya virus (16) and equine encephalitis viruses (88). Replication of the prototype Sindbis virus and Semliki Forest virus (SFV) is well understood (44, 71, 74, 79). Their genome consists of a positive-stranded RNA with a 5′ cap and a 3′ poly(A) tail. The 5′ two-thirds encodes the nonstructural polyprotein P1234, which is cleaved into four replicase proteins, nsP1 to nsP4 (47, 58, 60). The structural polyprotein is encoded in the 3′ one-third of the genome and cleaved into capsid and glycoproteins after translation from a subgenomic mRNA (79). Cytoplasmic replication complexes are associated with cellular membranes (71). Viruses mature by budding at the plasma membrane (35).In nature, arboviruses are spread by arthropod vectors (predominantly mosquitoes, ticks, flies, and midges) to vertebrate hosts (87). Little is known about how arthropod cells react to arbovirus infection. In mosquito cell cultures, an acute phase with efficient virus production is generally followed by the establishment of a persistent infection with low levels of virus production (9). This is fundamentally different from the cytolytic events following arbovirus interactions with mammalian cells and pathogenic insect viruses with insect cells. Alphaviruses encode host response antagonists for mammalian cells (2, 7, 34, 38).RNAi has been described for mosquitoes (56) and, when induced before infection, antagonizes arboviruses and their replicons (1, 4, 14, 15, 29, 30, 32, 42, 64, 65). RNAi is also functional in various mosquito cell lines (1, 8, 43, 49, 52). In the absence of RNAi, alphavirus and flavivirus replication and/or dissemination is enhanced in both mosquitoes and Drosophila (14, 17, 31, 45, 72). RNAi inhibitors weakly enhance SFV replicon replication in tick and mosquito cells (5, 33), posing the questions of how, when, and where RNAi interferes with alphavirus infection in mosquito cells.Here we use an A. albopictus-derived mosquito cell line to study RNAi responses to SFV. Using reporter-based assays, we demonstrate that SFV cannot avoid or efficiently inhibit the establishment of an RNAi response. We also demonstrate that the RNAi signal can spread between mosquito cells. SFV cannot inhibit cell-to-cell spread of the RNAi signal, and spread of the virus-induced RNAi signal (dsRNA/siRNA) can inhibit the replication of incoming SFV in neighboring cells. Furthermore, we show that SFV expression of a siRNA-binding protein increases levels of virus replication mainly by enhancing virus spread between cells rather than replication in initially infected cells. Taken together, these findings suggest a novel mechanism, cell-to-cell spread of antiviral dsRNA/siRNA, by which RNAi limits SFV dissemination in mosquito cells.  相似文献   

7.
8.
Plant viruses move through plasmodesmata (PD) either as nucleoprotein complexes (NPCs) or as tubule-guided encapsidated particles with the help of movement proteins (MPs). To explore how and why MPs specialize in one mechanism or the other, we tested the exchangeability of MPs encoded by DNA and RNA virus genomes by means of an engineered alfalfa mosaic virus (AMV) system. We show that Caulimoviridae (DNA genome virus) MPs are competent for RNA virus particle transport but are unable to mediate NPC movement, and we discuss this restriction in terms of the evolution of DNA virus MPs as a means of mediating DNA viral genome entry into the RNA-trafficking PD pathway.Following virus entry and replication, successful infection of a host requires viral spread to distal parts of the organism through the vascular tissue. In plants, virus movement involves mostly symplastic trafficking of different viral components through the connections of plasmodesmata (PD) (13). With this aim, plant viruses encode one or more movement proteins (MPs), which allow viral genomes to cross the host cell wall by altering the size exclusion limit (SEL) or the structure of PD (6, 11). Plant viruses have evolved distinct mechanisms to move their genomes within the host. These mechanisms can be grouped into two general strategies: one in which the genome is transported in the form of a nucleoprotein complex (NPC) and another in which nucleic acids are encapsidated and move as virus particles. In both cases, besides altering PD SEL, MPs are involved either in NPC assembly or in forming tubules traversing modified PD and helping transport of either NPC or virions to the neighboring cell. Within these two major strategies, there exists a wide range of variability in terms of the number and type of viral and host proteins helping MPs to mediate virus spread within the host (11).In spite of such variability, several different MPs have been classified into a 30K superfamily; these MPs, from 20 genera including both RNA and DNA genome viruses, are structurally related to the 30-kDa MP of Tobacco mosaic virus (TMV), independent of the movement strategy followed (14). Members of this family have a common core of predicted secondary structure elements (α-helices and β-elements) containing a nucleic acid binding domain. Distinct MPs belong to this family, including several tubule-forming MPs, although these are phylogenetically separated from the other members (14). Thus, 30K superfamily MPs are closely related, and some of them are functionally interchangeable in the viral context (2, 20). In particular, MPs from five distinct genera with an RNA genome can successfully replace the corresponding gene of Alfalfa mosaic virus (AMV) (19), indicating that one or more basic and fundamental movement properties might be associated with the common 30K structural core.Among all known plant viruses, only three viral families have evolved a DNA genome: Geminiviridae, Caulimoviridae, and Nanoviridae (6). One possible explanation for this restriction is that endogenous cell-to-cell transport via PD is specialized to use RNA as the communication and signaling molecule (12). To circumvent this restriction, and to allow the efficient exploitation of endogenous transport machineries, DNA genome viruses have evolved appropriate mechanisms involving their MPs. Interestingly, Begomovirus and Caulimovirus MPs also belong to the 30K superfamily discussed above (14). The MP encoded by Cauliflower mosaic virus (CaMV), the type member of Caulimoviridae, forms tubules that guide the movement of encapsidated virus via an indirect MP-virion interaction (16, 21), whereas geminivirus MPs selectively bind their genomes and transport them as NPCs (6, 9, 17). In this study, we investigated the evolutionary convergence of MPs encoded by DNA and RNA viruses by testing their exchangeability in the viral context.  相似文献   

9.
10.
Only a few archaeal viruses have been subjected to detailed structural analyses. Major obstacles have been the extreme conditions such as high salinity or temperature needed for the propagation of these viruses. In addition, unusual morphotypes of many archaeal viruses have made it difficult to obtain further information on virion architectures. We used controlled virion dissociation to reveal the structural organization of Halorubrum pleomorphic virus 1 (HRPV-1) infecting an extremely halophilic archaeal host. The single-stranded DNA genome is enclosed in a pleomorphic membrane vesicle without detected nucleoproteins. VP4, the larger major structural protein of HRPV-1, forms glycosylated spikes on the virion surface and VP3, the smaller major structural protein, resides on the inner surface of the membrane vesicle. Together, these proteins organize the structure of the membrane vesicle. Quantitative lipid comparison of HRPV-1 and its host Halorubrum sp. revealed that HRPV-1 acquires lipids nonselectively from the host cell membrane, which is typical of pleomorphic enveloped viruses.In recent years there has been growing interest in viruses infecting hosts in the domain Archaea (43). Archaeal viruses were discovered 35 years ago (52), and today about 50 such viruses are known (43). They represent highly diverse virion morphotypes in contrast to the vast majority (96%) of head-tail virions among the over 5,000 described bacterial viruses (1). Although archaea are widespread in both moderate and extreme environments (13), viruses have been isolated only for halophiles and anaerobic methanogenes of the kingdom Euryarchaeota and hyperthermophiles of the kingdom Crenarchaeota (43).In addition to soil and marine environments, high viral abundance has also been detected in hypersaline habitats such as salterns (i.e., a multipond system where seawater is evaporated for the production of salt) (19, 37, 50). Archaea are dominant organisms at extreme salinities (36), and about 20 haloarchaeal viruses have been isolated to date (43). The majority of these are head-tail viruses, whereas electron microscopic (EM) studies of highly saline environments indicate that the two other described morphotypes, spindle-shaped and round particles, are the most abundant ones (19, 37, 43). Thus far, the morphological diversity of the isolated haloarchaeal viruses is restricted compared to viruses infecting hyperthermophilic archaea, which are classified into seven viral families (43).All of the previously described archaeal viruses have a double-stranded DNA (dsDNA) genome (44). However, a newly characterized haloarchaeal virus, Halorubrum pleomorphic virus 1 (HRPV-1), has a single-stranded DNA (ssDNA) genome (39). HRPV-1 and its host Halorubrum sp. were isolated from an Italian (Trapani, Sicily) solar saltern. Most of the studied haloarchaeal viruses lyse their host cells, but persistent infections are also typical (40, 44). HRPV-1 is a nonlytic virus that persists in the host cells. In liquid propagation, nonsynchronous infection cycles of HRPV-1 lead to continuous virus production until the growth of the host ceases, resulting in high virus titers in the growth medium (39).The pleomorphic virion of HRPV-1 represents a novel archaeal virus morphotype constituted of lipids and two major structural proteins VP3 (11 kDa) and VP4 (65 kDa). The genome of HRPV-1 is a circular ssDNA molecule (7,048 nucleotides [nt]) containing nine putative open reading frames (ORFs). Three of them are confirmed to encode structural proteins VP3, VP4, and VP8, which is a putative ATPase (39). The ORFs of the HRPV-1 genome show significant similarity, at the amino acid level, to the minimal replicon of plasmid pHK2 of Haloferax sp. (20, 39). Furthermore, an ∼4-kb region, encoding VP4- and VP8-like proteins, is found in the genomes of two haloarchaea, Haloarcula marismortui and Natronomonas pharaonis, and in the linear dsDNA genome (16 kb) of spindle-shaped haloarchaeal virus His2 (39). The possible relationship between ssDNA virus HRPV-1 and dsDNA virus His2 challenges the classification of viruses, which is based on the genome type among other criteria (15, 39).HRPV-1 is proposed to represent a new lineage of pleomorphic enveloped viruses (39). A putative representative of this lineage among bacterial viruses might be L172 of Acholeplasma laidlawii (14). The enveloped virion of L172 is pleomorphic, and the virus has a circular ssDNA genome (14 kb). In addition, the structural protein pattern of L172 with two major structural proteins, of 15 and 53 kDa, resembles that of HRPV-1.The structural approach has made it possible to reveal relationships between viruses where no sequence similarity can be detected. It has been realized that several icosahedral viruses infecting hosts in different domains of life share common virion architectures and folds of their major capsid proteins. These findings have consequences for the concept of the origin of viruses. A viral lineage hypothesis predicts that viruses within the same lineage may have a common ancestor that existed before the separation of the cellular domains of life (3, 5, 8, 26). Currently, limited information is available on the detailed structures of viruses infecting archaea. For example, the virion structures of nontailed icosahedral Sulfolobus turreted icosahedral virus (STIV) and SH1 have been determined (21, 23, 46). However, most archaeal viruses represent unusual, sometimes nonregular, morphotypes (43), which makes it difficult to apply structural methods that are based on averaging techniques.A biochemical approach, i.e., controlled virion dissociation, gives information on the localization and interaction of virion components. In the present study, controlled dissociation was used to address the virion architecture of HRPV-1. A comparative lipid analysis of HRPV-1 and its host was also carried out. Our results show that the unique virion type is composed of a flexible membrane decorated with the glycosylated spikes of VP4 and internal membrane protein VP3. The circular ssDNA genome resides inside the viral membrane vesicle without detected association to any nucleoproteins.  相似文献   

11.
Despite their importance as agents of emerging disease, the time scale and evolutionary processes that shape the appearance of new viral species are largely unknown. To address these issues, we analyzed intra- and interspecific evolutionary processes in the Luteoviridae family of plant RNA viruses. Using the coat protein gene of 12 members of the family, we determined their phylogenetic relationships, rates of nucleotide substitution, times to common ancestry, and patterns of speciation. An associated multigene analysis enabled us to infer the nature of selection pressures and the genomic distribution of recombination events. Although rates of evolutionary change and selection pressures varied among genes and species and were lower in some overlapping gene regions, all fell within the range of those seen in animal RNA viruses. Recombination breakpoints were commonly observed at gene boundaries but less so within genes. Our molecular clock analysis suggested that the origin of the currently circulating Luteoviridae species occurred within the last 4 millennia, with intraspecific genetic diversity arising within the last few hundred years. Speciation within the Luteoviridae may therefore be associated with the expansion of agricultural systems. Finally, our phylogenetic analysis suggested that viral speciation events tended to occur within the same plant host species and country of origin, as expected if speciation is largely sympatric, rather than allopatric, in nature.Although RNA viruses are the most common agents of emerging disease, key aspects of their evolution are still only partly understood. This is of both academic and practical importance, as virus evolution may compromise disease control strategies, including the rapid generation of genotypes that are able to evade host immune responses or of those that are resistant to antivirals or crop genetic resistance (20, 34, 47).Most of our knowledge of the rapidity of RNA virus evolution comes from the study of animal viruses, for which estimates of rates of nucleotide substitution normally fall within 1 order of magnitude of 1 × 10−3 nucleotide substitutions per site per year (subs/site/year) and largely reflect the background mutation rate (10, 13, 29, 37, 53). Equivalent studies on plant RNA viruses have reported more heterogeneous rates. Early studies suggested that some plant RNA viruses evolved more slowly than RNA viruses that infect animals. For example, estimates of the nucleotide substitution rate in the range of ∼1 × 10−6 to 1 × 10−8 subs/site/year have been obtained for Turnip yellow mosaic virus (4, 23) and some tobamoviruses (21, 25). In contrast, more recent estimates using Bayesian coalescent methods applied to sequences with known dates of sampling and allowing for rate variation among lineages have reported substitution rates in the same range as those of animal RNA viruses (22, 63) and therefore suggest relatively high rates of mutation, as expected, given the intrinsically error-prone nature of RNA replication (15, 65). As well as the differences in how these rates are estimated, a reasonable biological explanation for such a diversity of rate estimates is that they are increased in the short term due to the presence of mutational polymorphisms but lower in the long term because any such deleterious mutations would have then been removed by purifying selection (17, 22). In particular, severe population bottlenecks at transmission would allow deleterious mutations to rise to a high frequency due to strong genetic drift. Such effects make it dangerous to extrapolate long-term rates of evolutionary change from the analysis of intraspecific sequence data (31). Differences in the strength of adaptive evolution could also cause rate heterogeneity, including such processes as competition for susceptible individuals and the colonization of new host species (22, 65).Although there is a growing body of data on intraspecific evolutionary processes in plant RNA viruses, including rates of nucleotide substitution, there has been a general neglect of long-term evolutionary patterns, including the determinants of viral speciation. Exceptions are recent analyses of the Potyviridae and the Sobemovirus, which associated viral speciation with the development of agriculture (17, 24). Although RNA viruses reproduce asexually, it is informative to consider as analogies the two major forms of speciation used in studies of sexually reproducing eukaryotes: allopatric speciation, in which reproductive isolation follows geographic separation, and sympatric speciation, in which reproductive isolation occurs within an interbreeding population (67). In the context of RNA viruses, allopatric speciation can be thought of as the genetic diversification that occurs when viruses jump to new host species and thereafter evolve independently, as is commonly associated with the process of viral “emergence.” In contrast, sympatric speciation would occur when viruses diversify within a single host species, perhaps by exploiting different cell types (34). Despite the importance of these processes for our understanding of the macroevolution of RNA viruses, their respective roles are currently unknown.To better understand the nature of long-term evolutionary processes in plant RNA viruses, we undertook an extensive molecular evolutionary analysis of the family Luteoviridae, a heterogeneous family of plant viruses divided into three genera, Luteovirus, Polerovirus, and Enamovirus, containing five, nine, and one classified species, respectively, as well as a number of unclassified species (18). The Luteoviridae possess positive-sense single-stranded RNA genomes of 5,600 to 6,000 nucleotides (nt). These genomes can harbor five or six open reading frames (ORFs). 5′-proximal partially overlapping ORF1 and -2 encode proteins P1 and P2, which are involved in virus replication. Low-frequency −1 ribosomal frameshifting in the overlapping region results in the P1-P2 fusion RNA-dependent RNA polymerase protein (RdRp). ORF3 encodes the coat protein (CP) and completely contains ORF4, which is not found in Enamovirus and is needed for virus movement in the plant (the movement protein, MP). ORF5, which is necessary for aphid transmission (6, 27, 49) and is also involved in virus movement (57) and Luteoviridae phloem limitation (58), is translated through in-frame read-through of the ORF3 stop codon, existing as a read-through domain (RTD) fused to the CP. Members of the genus Polerovirus have an extra ORF0 in the 5′ end of the genome partially overlapping ORF1. Its translation product (P0) acts as a repressor of the RNA-silencing plant defense response (44, 59). Finally, some Luteovirus species have an additional ORF6 with an unknown function in the 3′ end of the genome (18, 48, 70). As a consequence of this particular genomic organization, approximately one-third of the Polerovirus genome, and a smaller fraction in the Luteovirus genome, is composed of overlapping regions.Due to their agronomic importance, gene sequence data, together with information on host range and geographical distribution, are available for a relatively large number of members of the family Luteoviridae. However, to date the only luteovirus for which rates of evolutionary change have been estimated is Barley yellow dwarf virus (BYDV). In this case, an analysis of substitution rates based on viral RNA extracted from herbarium specimens produced estimates of between 6.2 × 10−4 and 9.7 × 10−5 subs/site/year (43). Similarly, only one estimate of the point at which genetic diversity arose in the family Luteoviridae has been obtained, i.e., approximately 9,000 years ago, and therefore it is perhaps associated with the rise of agriculture (17). However, only a limited number of Luteoviridae species and sequences were included in this analysis. No studies have yet considered the mechanisms of speciation in the family Luteoviridae.The family Luteoviridae also represents a useful data set to study two other evolutionary phenomena: the pattern and determinants of recombination, which appears to be commonplace within the family Luteoviridae (26, 49, 51, 70, 71), and the differing evolutionary dynamics in genes with overlapping reading frames. There are contrasting hypotheses as to why overlapping reading frames are so commonly used in RNA viruses. According to one view, gene overlapping maximizes the genetic information in smaller genomes (1, 39). Alternatively, it has been suggested that gene overlap generates mutational robustness (i.e., the ability to preserve phenotypes despite the genomic mutational load) at the population level (2, 16, 42). Under the latter hypothesis, gene overlapping generates hypersensitivity to deleterious mutations, as these affect more than one gene. Although this hypersensitivity reduces the capacity of each individual to buffer mutation effects, it represents a selective advantage for wild-type genotypes, which then bolsters robustness at the population level (16, 42). As a consequence of this elevated burden of deleterious mutation, RNA viruses with larger proportions of their genomes present as overlapping reading frames are expected to exhibit lower rates of nucleotide substitution (41, 50). Such a rate reduction has been observed in many animal DNA and RNA viruses (for example, see references 35, 36, 55, 77, and 78), although only a few studies have considered plant RNA viruses in this context (28, 61).  相似文献   

12.
Genetic bottlenecks facilitate the fixation and extinction of variants in populations, and viral populations are no exception to this theory. To examine the existence of genetic bottlenecks in cell-to-cell movement of plant RNA viruses, we prepared constructs for Soil-borne wheat mosaic virus RNA2 vectors carrying two different fluorescent proteins, yellow fluorescent protein (YFP) and cyan fluorescent protein (CFP). Coinoculation of host plant leaves with the two RNA2 vectors and the wild-type RNA1 showed separation of the two vector RNA2s, mostly within seven to nine cell-to-cell movements from individual initially coinfected cells. Our statistical analysis showed that the number of viral RNA genomes establishing infection in adjacent cells after the first cell-to-cell movement from an initially infected cell was 5.97 ± 0.22 on average and 5.02 ± 0.29 after the second cell-to-cell movement. These results indicate that plant RNA viruses may generally face narrow genetic bottlenecks in every cell-to-cell movement. Furthermore, our model suggests that, rather than suffering from fitness losses caused by the bottlenecks, the plant RNA viruses are utilizing the repeated genetic bottlenecks as an essential element of rapid selection of their adaptive variants in trans-acting genes or elements to respond to host shifting and changes in the growth conditions of the hosts.Plant RNA viruses change their genomes so rapidly that variant viruses with altered biological properties are often found after prolonged growth of infected plants or after serial mechanical inoculations (26, 33). Furthermore, inoculation of less-fit artificial mutants produces revertants or pseudo-revertants even after short infection times (12, 14). The rapid evolution of plant RNA viral genomes is achieved not only by high mutation rates due to error-prone replication by the nonproofreading viral RNA-dependent RNA polymerase (19) but also by rapid selection and strong genetic drift. Generally, narrow genetic bottlenecks facilitate the fixation and extinction of variants in populations (15), and viral populations are no exception to this theory.Plant RNA viruses are known to face many narrow genetic bottlenecks during their life cycles (23). The life cycles of most plant RNA viruses are as follows: After replicating in cells, viruses move from cell to cell through plasmodesmata, which connect the cytoplasms of adjacent cells separated by cell walls in plant tissue. Following the establishment of infection in cells and cell-to-cell movements, the viruses expand their infected regions, spreading to the veins and moving through the vascular system and infecting the plant systemically. Some plant RNA viruses are transmitted through the seeds or via mechanical injuries, but most are transmitted from plant to plant by biological vectors such as insects, nematodes, and fungi. Previous studies have found that genetic bottlenecks occur during the transfer from lower leaves to upper leaves in systemic infections of Wheat streak mosaic virus (WSMV) (11), Tobacco mosaic virus (TMV) (24), and Cucumber mosaic virus (CMV) (18) and during the transfer from one tiller to another tiller of WSMV (11). Vector transmissions were also shown to act as genetic bottlenecks for WSMV (11), CMV (1, 3), and Potato virus Y (PVY) (20). With the exception of PVY, the typical method for detecting genetic bottlenecks has been to observe the spatial separation of closely related strains or artificial synonymous mutants inoculated as mixed populations: the narrower the genetic bottleneck, the more frequently the spatial separation should be observed. Using this idea with mathematical analyses, WSMV was estimated to infect a new tiller starting with four genomes (9), TMV was estimated to infect the upper leaves starting with 10 genomes (24), and CMV was estimated to infect a new plant starting with one to two particles after aphid transmission (3). Studies of PVY using sets of host plant cultivars with or without resistance genes and mixed strains of viruses with or without resistance-breaking abilities also estimated the number of virus particles transmitted by an aphid vector to be 0.5 to 3.2 on average (20).However, genetic bottlenecks in cell-to-cell movement of viruses have not been well characterized, although these occurrences are likely (11) and have been expected to be important for understanding the life cycle and population dynamics of plant RNA viruses. The size of genetic bottlenecks in cell-to-cell movement can be referred to as “multiplicity of infection (MOI) in plant tissue colonization,” and only a recent study showing that the estimated MOI of TMV is between 6 and 1 to 2 (10) indicates the occurrence and the size of genetic bottlenecks in cell-to-cell movement of a plant RNA virus. In this paper, we also show the occurrence of narrow genetic bottlenecks during cell-to-cell movement of a plant RNA virus, Soil-borne wheat mosaic virus (SBWMV, type species of the genus Furovirus), by observing the spatial separation of RNA2 vectors carrying different fluorescent proteins, yellow fluorescent protein (YFP) and cyan fluorescent protein (CFP). Both of the fluorescent proteins were expressed as fusion proteins to the N-terminal nuclear localization signal (NLS) peptide from Simian virus 40 (SV40) large T antigen, which enabled us to observe and count the infected cells accurately using nuclear fluorescence. Numerical data were analyzed to estimate the size of bottlenecks. We also carried out a simulation to show that, due to the narrow genetic bottlenecks, rapid selection occurs even on trans-acting elements in plant RNA virus genomes, overcoming the negative effect of complementation among adaptive and defective genomes in each intracellular population. We discuss the possible roles of the bottlenecks in the life cycle and evolution mechanisms of plant RNA viruses.  相似文献   

13.
The multiplicity of infection (MOI), i.e., the number of virus genomes that infect a cell, is a key parameter in virus evolution, as it determines processes such as genetic exchange among genomes, selection intensity on viral genes, epistatic interactions, and the evolution of multipartite viruses. In fact, the MOI level is equivalent to the virus ploidy during genome expression. Nevertheless, there are few experimental estimates of MOI, particularly for viruses with eukaryotic hosts. Here we estimate the MOI of Tobacco mosaic virus (TMV) in its systemic host, Nicotiana benthamiana. The progress of infection of two TMV genotypes, differently tagged with the green or red fluorescent proteins GFP and RFP, was monitored by determining the number of leaf cell protoplasts that showed GFP, RFP, or GFP and RFP fluorescence at different times postinoculation. This approach allowed the quantitative analysis of the kinetics of infection and estimation of the generation time and the number of infection cycles required for leaf colonization. MOI levels were estimated from the frequency of cells infected by only TMV-GFP or TMV-RFP. The MOI was high, but it changed during the infection process, decreasing from an initial level of about 6 to a final one of 1 to 2, with most infection cycles occurring at the higher MOI levels. The decreasing MOI can be explained by mechanisms limiting superinfection and/or by genotype competition within double-infected cells, which was shown to occur in coinfected tobacco protoplasts. To our knowledge, this is the first estimate of MOI during virus colonization of a eukaryotic host.Virus evolution has been a very active area of research in the last few decades, as viruses are both important pathogens of humans, animals, and plants and good models to experimentally test hypotheses on parasite evolution or, more generally, central questions on evolutionary biology (11, 12, 21, 36). Considerable efforts have been devoted to modeling the evolution of viral populations. However, contrasting the theoretical models with reality may be hindered by limited experimental information on important parameters of the virus life cycle. The multiplicity of infection (MOI), i.e., the number of virus particles or genomes that may infect a cell, is a key parameter in many models of virus evolution (5, 6, 14, 15, 37, 38, 39, 52, 53, 57, 61) for which experimental estimates are scant.When a cell is coinfected by different viral genomes, competition may lead to decreased fitness of individual genotypes in comparison with their fitness in single infections (15, 31, 40). Thus, limiting coinfection may result in a selective advantage for viruses (58), which have developed mechanisms to prevent superinfection of previously infected cells (51, 60). On the other hand, infection of a cell by more than one virus genome is a prerequisite for two central phenomena in virus genetics to take place: recombination and complementation of defective mutants. Recombination between viral strains during replication in the same cell and complementation of defective mutants have been extensively documented for viruses infecting prokaryotes, animals, and plants (2, 25, 56), indicating that there must be some degree of coinfection and, hence, that the MOI must be higher than one in at least some infected cells. However, estimates of MOI in the natural hosts of viruses are surprisingly scarce in spite of this parameter''s relevance: values of about 2 to 3 have been reported for different DNA or RNA bacteriophages (26, 41, 51, 58), and a value of 4 to 5 was reported for Autographa californica nuclear polyhedrosis virus infecting larvae of the moth Tricoplusia ni (3), to our knowledge, the only estimate for a virus in its eukaryotic host. We are not aware of estimates reported for viruses infecting mammals or plants, although a MOI of about 3 can be inferred from the number of proviral copies of HIV in spleen cells of infected patients (29). This paucity of data may be due to the technical difficulty of directly measuring MOI, particularly within a eukaryotic host. Genetic approaches may provide valid alternatives for estimating MOI levels (3, 58), and here, the MOI of a plant virus is estimated through the analysis of the relative frequencies of two genotypes during the process of host colonization.Host colonization by plant-infecting viruses has been known for a long time to be a two-step phenomenon. First, colonization proceeds slowly from the initially infected cells to their neighbors by way of the cytoplasmic connections called plasmodesmata, a process known as cell-to-cell movement. After infection thus reaches the cells in the vasculature, the second step, known as long-distance or systemic movement, occurs as viruses move faster to distant organs through the vascular tissue, the phloem in most cases (59). As a result of these processes, the virus population within the infected plant may be strongly structured. Analyses of different viruses in different host plant species have shown that systemic movement causes population bottlenecks that may be severe (16, 28, 32, 34, 46), resulting in differences in the genetic composition of the virus subpopulations in different systemically infected organs. No analysis of population bottlenecks during cell-to-cell movement has been reported, but data indicate that the virus population within a leaf has a strong spatial structure with a separate distribution of different genotypes in different leaf areas. These reports derive from analyses of viruses that differ in genomic organization and gene expression strategies in different host plant species (9, 10, 23, 55); they indicate that a separate distribution of viral genotypes within the infected leaf is a general phenomenon and suggest limitation of coinfection. Data on the spatial exclusion of virus genotypes within the infected leaf are in apparent contradiction with the abundant evidence of recombination and complementation of defective mutants, which has been widely documented for plant viruses (19, 44, 50, 62). It should be pointed out that all reports on the spatial exclusion of virus genotypes in an infected leaf derive from microscopy observations, mostly at late times after infection of the tissue. No information is available on the kinetics of leaf colonization by viruses, and current data do not allow the estimation of MOI.In this report, we estimate the MOI of a plant RNA virus, Tobacco mosaic virus (TMV), in its systemic host, Nicotiana benthamiana. For this, we have reexamined the process of virus colonization by monitoring the progress of infection of two TMV genotypes in inoculated and in systemically infected leaves. The two TMV genotypes differed in the expression of fluorescent tags, either the green fluorescent protein (GFP) from Aequorea victoria (42, 43) or a red fluorescent protein (RFP) from Discosoma sp. (49). The expression of GFP and RFP allowed the precise quantification of the number of cells infected by either one or both TMV genotypes, and these data allowed the estimation of genotype frequencies and of MOI. The results show evidence of strong spatial structure of the virus population, with most cells being infected by either TMV-GFP or TMV-RFP alone and only a small fraction of cells being double infected. The kinetics of the single and double infections show that the MOI changes with time, decreasing as colonization progresses and therefore suggesting that exclusion mechanisms operate at later times after infection.  相似文献   

14.
Recently, complete replication of hepatitis C virus (HCV) in tissue culture was established using the JFH1 isolate. To analyze determinants of HCV genome packaging and virion assembly, we developed a system that supports particle production based on trans-packaging of subgenomic viral RNAs. Using JFH1 helper viruses, we show that subgenomic JFH1 replicons lacking the entire core to NS2 coding region are efficiently encapsidated into infectious virus-like particles. Similarly, chimeric helper viruses with heterologous structural proteins trans-package subgenomic JFH1 replicons. Like authentic cell culture-produced HCV (HCVcc) particles, these trans-complemented HCV particles (HCVTCP) penetrate target cells in a CD81 receptor-dependent fashion. Since HCVTCP production was limited by competition between the helper and subgenomic RNA and to avoid contamination of HCVTCP stocks with helper viruses, we created HCV packaging cells. These cells encapsidate various HCV replicons with high efficiency, reaching infectivity titers up to 106 tissue culture infectious doses 50 per milliliter. The produced particles display a buoyant density comparable to HCVcc particles and can be propagated in the packaging cell line but support only a single-round infection in naïve cells. Together, this work demonstrates that subgenomic HCV replicons are assembly competent, thus excluding cis-acting RNA elements in the core-to-NS2 genomic region essential for RNA packaging. The experimental system described here should be helpful to decipher the mechanisms of HCV assembly and to identify RNA elements and viral proteins involved in particle formation. Similar to other vector systems of plus-strand RNA viruses, HCVTCP may prove valuable for gene delivery or vaccination approaches.Hepatitis C virus (HCV) is an enveloped plus-strand RNA virus of the genus Hepacivirus within the family Flaviviridae (34). The HCV genome is approximately 9.6 kb in length and consists of a single open reading frame encoding a polyprotein of ca. 3,000 amino acids and nontranslated regions (NTRs) located at the 5′ and 3′ termini. These NTRs are highly structured RNA segments encompassing critical cis-active RNA elements essential for genome replication and RNA translation (31). Viral proteins are expressed in a cap-independent manner by means of an internal ribosome entry site (IRES) located in the 5′ NTR. Co- and posttranslational cleavages liberate 10 viral proteins: core; envelope protein 1 (E1) and E2, representing the structural proteins that constitute the virion; a small membrane-associated ion channel protein designated p7 that is essential for virus assembly (16, 22, 43, 57); and six nonstructural (NS) proteins (NSs 2, 3, 4A, 4B, 5A, and 5B). HCV proteins NS3 to NS5B are both necessary and sufficient to establish membrane-bound replication complexes catalyzing RNA replication (13, 36). More recent data indicate that the NS2 protease that catalyzes cleavage at the NS2-NS3 site in addition participates in assembly and release of infectious viruses (22). Finally, ribosomal frame-shifting and internal translation initiation yield a group of additional proteins designated ARFP (alternative reading frame protein) or core+1 proteins. However, their function for the HCV replication cycle is currently not known.One hallmark of HCV is its high propensity to establish a persistent infection, which frequently causes progressive morbidity ranging from hepatic fibrosis to cirrhosis and hepatocellular carcinoma (20). Despite considerable progress in the treatment of HCV infection, the currently available therapy (a combination of pegylated interferon alpha with ribavirin) is not well tolerated and is efficacious in only ca. 50% of patients infected with the most prevalent genotype 1 (38). Therapeutic or prophylactic vaccines are not available. For these reasons and with currently ca. 170 million persistently infected individuals, HCV infection represents a considerable global health problem necessitating pertinent basic and applied research efforts.In recent years three major advances enabled analysis of the HCV replication cycle in tissue culture. First, Lohmann and colleagues developed subgenomic HCV replicons (36). These autonomously replicating RNA molecules carry all the genetic elements necessary for self-replication (the NTRs and NS3 to NS5B), including a selectable marker or a reporter gene in place of the viral structural proteins, and an internal IRES for expression of the HCV replicase genes (reviewed in reference 45). Second, HCV pseudotype particles, i.e., retroviral particles surrounded by an envelope carrying HCV E1-E2 complexes in place of their cognate envelope proteins, were established (3, 21). As these particles carry functional HCV glycoprotein complexes on their surface, HCV pseudotype particles have been instrumental for the analysis of E1-E2 receptor interactions and the early events of HCV infection (reviewed in reference 2). Finally, in 2005 fully permissive cell culture systems which are based on the JFH1 clone were described (33, 66, 72). This isolate replicates with unprecedented efficiency in transfected Huh7 human hepatoma cells and produces particles infectious both in vitro and in vivo, thus providing a model system reproducing the complete HCV replication cycle.Use of these novel models has considerably expanded our knowledge of viral and host cell factors involved in HCV replication (for a recent review, see reference 59). It is now known that similar to virtually all other plus-strand RNA viruses, HCV induces intracellular membrane alterations and replicates its genome in conjunction with vesicular membrane structures, the so-called “membranous web” (10, 13). Presumably as a consequence of this specific, rather secluded architecture of the membrane-associated replication machinery, all viral proteins involved in RNA replication, with the exception of NS5A function in cis, cannot be complemented in trans (1). Restricted trans-complementation of viral replicase proteins has been observed for other plus-strand RNA viruses as well, thus indicating that a rather “closed” replication machinery is a shared feature of these viruses (15, 27, 60). In contrast, for a number of plus-strand RNA viruses from diverse virus families like Picornaviridae (poliovirus), Alphaviridae (Sindbis virus, Semliki Forest virus, and Venezuelan equine encephalitis virus), Coronaviridae (human coronavirus E229), and Flaviviridae (tick-borne encephalitis virus, Kunjin virus, West Nile virus, and yellow fever virus), assembly of progeny viruses can be achieved when structural proteins are expressed in trans and independent from the RNA molecule that encodes the replicase proteins. Similarly, Miyanari recently reported that HCV genomes with lethal mutations in core protein can be rescued by ectopic expression of functional core protein (39). This flexibility has been extensively used to create viral vectors for gene delivery as well as viral vector-based immunization approaches (32, 48, 49, 61, 68) (for a recent review on alphaviral vectors, the most frequently used among plus strand RNA vectors, see reference 37). In these systems the viral genome region encoding the structural proteins is replaced by a transgene. The resulting defective vector genomes are capable of RNA replication but due to the lack of structural proteins are unable to produce progeny virus particles. This defect is rescued by expression of the structural proteins in trans via helper viruses (28, 55) or, in some cases, by DNA constructs stably expressed in packaging cell lines (17). The resulting virus-like particles are infectious but support only single-round infection and are unable to spread, thus improving the safety of the viral transduction system.Given the success of plus-strand RNA vector technology for basic and applied clinical research, in this study we developed a trans-complementation system for HCV that provided new insights into the basic principles of HCV particle assembly.  相似文献   

15.
Intermolecular recombination between the genomes of closely related RNA viruses can result in the emergence of novel strains with altered pathogenic potential and antigenicity. Although recombination between flavivirus genomes has never been demonstrated experimentally, the potential risk of generating undesirable recombinants has nevertheless been a matter of concern and controversy with respect to the development of live flavivirus vaccines. As an experimental system for investigating the ability of flavivirus genomes to recombine, we developed a “recombination trap,” which was designed to allow the products of rare recombination events to be selected and amplified. To do this, we established reciprocal packaging systems consisting of pairs of self-replicating subgenomic RNAs (replicons) derived from tick-borne encephalitis virus (TBEV), West Nile virus (WNV), and Japanese encephalitis virus (JEV) that could complement each other in trans and thus be propagated together in cell culture over multiple passages. Any infectious viruses with intact, full-length genomes that were generated by recombination of the two replicons would be selected and enriched by end point dilution passage, as was demonstrated in a spiking experiment in which a small amount of wild-type virus was mixed with the packaged replicons. Using the recombination trap and the JEV system, we detected two aberrant recombination events, both of which yielded unnatural genomes containing duplications. Infectious clones of both of these genomes yielded viruses with impaired growth properties. Despite the fact that the replicon pairs shared approximately 600 nucleotides of identical sequence where a precise homologous crossover event would have yielded a wild-type genome, this was not observed in any of these systems, and the TBEV and WNV systems did not yield any viable recombinant genomes at all. Our results show that intergenomic recombination can occur in the structural region of flaviviruses but that its frequency appears to be very low and that therefore it probably does not represent a major risk in the use of live, attenuated flavivirus vaccines.RNA viruses are able to undergo rapid genetic changes in order to adapt to new hosts or environments. Although much of this flexibility is due to the error-prone nature of the RNA-dependent RNA polymerase, which generates an array of different point mutations within the viral population (23), recombination is also a common and important mechanism for generating viral diversity (18, 31, 42, 58). Recombination occurs when the RNA-dependent RNA polymerase switches templates during replication, an event that is favored when both templates share identical or very similar sequences. Three types of RNA recombination have been identified: homologous recombination occurs at sites with exact sequence matches; aberrant homologous recombination requires sequence homology, but crossover occurs either upstream or downstream of the site of homology, resulting in a duplication or deletion; and nonhomologous (or illegitimate) recombination is independent of sequence homology (31, 42).When the same cell is infected by viruses of two different strains, or even different species, recombination between their genomic RNAs can potentially lead to the emergence of new pathogens. A case in point is the emergence of Western equine encephalitis virus, a member of the genus Alphavirus, family Togaviridae, which arose by homologous recombination between Eastern equine encephalitis virus and Sindbis virus (14).Some mammalian RNA viruses can recombine at a frequency that is detectable in experimental settings (1, 2, 55), and phylogenetic analysis of partial or complete genome sequences suggests that RNA recombination is a widespread phenomenon. Naturally occurring recombinant viruses have been identified in almost every family of positive-stranded RNA viruses (31, 58).Flaviviruses are members of the genus Flavivirus, family Flaviviridae, a family that also includes the genera Pestivirus and Hepacivirus. Several of the flaviviruses are important human pathogens, such as Japanese encephalitis virus (JEV), West Nile virus (WNV), the dengue viruses, yellow fever virus, and tick-borne encephalitis virus (TBEV).Although there has never been a report of a pathogenic flavivirus strain arising due to recombination involving attenuated vaccine strains (39), the urgent necessity to develop tetravalent vaccines containing all four serotypes of dengue virus—two such vaccines are currently undergoing clinical testing (45)—has recently brought the recombination issue to the forefront of discussion among researchers, regulators, and vaccine producers (39). It has been suggested that recombination, either between the strains present in a multivalent vaccine or between an attenuated vaccine strain and a wild-type strain, could lead to the emergence of new viruses with unpredictable properties (49).So far, recombination between flavivirus genomes has not been demonstrated directly in the laboratory. However, phylogenetic analysis of partial genome sequences available in the GenBank database has suggested that homologous recombination may have occurred between closely related strains of dengue virus (20, 52, 54, 59). An experimental approach for assessing the ability of flavivirus genomes to recombine is therefore urgently needed.Flavivirus virions are composed of a single-stranded, positive-sense RNA genome that, together with the capsid protein C, forms the viral nucleocapsid. The nucleocapsid is covered by a lipid envelope containing the surface glycoproteins prM and E. These glycoproteins drive budding at the membrane of the endoplasmic reticulum during the assembly stage and mediate entry of the virus into host cells (41). Replicons, defined as self-replicating, noninfectious RNA molecules, can be generated by deleting parts or all of the region coding for the structural proteins C, prM, and E from the viral genome but maintaining all seven of the nonstructural proteins and the flanking noncoding sequences, which are required in cis for RNA replication (25). By providing the missing structural protein components in trans, replicons can be packaged into virus-like particles that are capable of a single round of infection (10, 15, 24, 47).Typically, researchers developing novel replicating vaccines, especially ones that involve multiple components, make an effort to come up with strategies to prevent recombination, for example by “wobbling” codons, i.e., replacing codons in homologous regions with synonymous ones encoding the same amino acid but consisting of a different nucleotide triplet (50, 57). In this study, in order to assess the propensity of flavivirus genomes to recombine, we took an opposite approach, establishing a “recombination trap” that favors the selection and sensitive detection of recombination products. This system takes advantage of the ability of replicon pairs containing deletions in their structural protein genes to complement each other in trans and thus be propagated together in cell culture, and by passage at limiting dilutions, it allows infectious RNA genomes arising by recombination between the two replicons to be preferentially selected.Using the recombination trap, we have now obtained the first direct evidence of recombination between flavivirus genomes in the laboratory. Aberrant homologous recombination was observed twice with JEV replicons, resulting in viruses with unnatural gene arrangements and reduced growth properties compared to those of wild-type JEV. No infectious recombinants of any kind were obtained when TBEV or WNV replicons were used. Interestingly, we never detected a fully infectious wild-type genome arising by homologous recombination in any of these systems. The results of this study show that the propensity of flavivirus genomes to recombine in the region coding for the structural proteins appears to be quite low, suggesting that recombination does not represent a major risk in the use of live, attenuated flavivirus vaccines.  相似文献   

16.
17.
Like other RNA viruses, coxsackievirus B5 (CVB5) exists as circulating heterogeneous populations of genetic variants. In this study, we present the reconstruction and characterization of a probable ancestral virion of CVB5. Phylogenetic analyses based on capsid protein-encoding regions (the VP1 gene of 41 clinical isolates and the entire P1 region of eight clinical isolates) of CVB5 revealed two major cocirculating lineages. Ancestral capsid sequences were inferred from sequences of these contemporary CVB5 isolates by using maximum likelihood methods. By using Bayesian phylodynamic analysis, the inferred VP1 ancestral sequence dated back to 1854 (1807 to 1898). In order to study the properties of the putative ancestral capsid, the entire ancestral P1 sequence was synthesized de novo and inserted into the replicative backbone of an infectious CVB5 cDNA clone. Characterization of the recombinant virus in cell culture showed that fully functional infectious virus particles were assembled and that these viruses displayed properties similar to those of modern isolates in terms of receptor preferences, plaque phenotypes, growth characteristics, and cell tropism. This is the first report describing the resurrection and characterization of a picornavirus with a putative ancestral capsid. Our approach, including a phylogenetics-based reconstruction of viral predecessors, could serve as a starting point for experimental studies of viral evolution and might also provide an alternative strategy for the development of vaccines.The group B coxsackieviruses (CVBs) (serotypes 1 to 6) were discovered in the 1950s in a search for new poliovirus-like viruses (33, 61). Infections caused by CVBs are often asymptomatic but may occasionally result in severe diseases of the heart, pancreas, and central nervous system (99). CVBs are small icosahedral RNA viruses belonging to the Human enterovirus B (HEV-B) species within the family Picornaviridae (89). In the positive single-stranded RNA genome, the capsid proteins VP1 to VP4 are encoded within the P1 region, whereas the nonstructural proteins required for virus replication are encoded within the P2 and P3 regions (4). The 30-nm capsid has an icosahedral symmetry and consists of 60 copies of each of the four structural proteins. The VP1, VP2, and VP3 proteins are surface exposed, whereas the VP4 protein lines the interior of the virus capsid (82). The coxsackievirus and adenovirus receptor (CAR), a cell adhesion molecule of the immunoglobulin superfamily, serves as the major cell surface attachment molecule for all six serotypes of CVB (5, 6, 39, 60, 98). Some strains of CVB1, CVB3 and CVB5 also interact with the decay-accelerating factor (DAF) (CD55), a member of the family of proteins that regulate the complement cascade. However, the attachment of CVBs to DAF alone does not permit the infection of cells (6, 7, 59, 85).Picornaviruses exist as genetically highly diverse populations within their hosts, referred to as quasispecies (20, 57). This genetic plasticity enables these viruses to adapt rapidly to new environments, but at the same time, it may compromise the structural integrity and enzymatic functionality of the virus. The selective constraints imposed on the picornavirus genome are reflected in the different regions used for different types of evolutionary studies. The highly conserved RNA-dependent RNA polymerase (3Dpol) gene is used to establish phylogenetic relationships between more-distantly related viruses (e.g., viruses belonging to different genera) (38), whereas the variable genomic sequence encoding the VP1 protein is used for the classification of serotypes (13, 14, 69, 71, 72).In 1963, Pauling and Zuckerkandl proposed that comparative analyses of contemporary protein sequences can be used to predict the sequences of their ancient predecessors (73). Experimental reconstruction of ancestral character states has been applied to evolutionary studies of several different proteins, e.g., galectins (49), G protein-coupled receptors (52), alcohol dehydrogenases (95), rhodopsins (15), ribonucleases (46, 88, 110), elongation factors (32), steroid receptors (10, 96, 97), and transposons (1, 45, 87). In the field of virology, reconstructed ancestral or consensus protein sequences have been used in attempts to develop vaccine candidates for human immunodeficiency virus type 1 (21, 51, 66, 81) but rarely to examine general phenotypic properties.In this study, a CVB5 virus with a probable ancestral virion (CVB5-P1anc) was constructed and characterized. We first analyzed in detail the evolutionary relationships between structural genes of modern CVB5 isolates and inferred a time scale for their evolutionary history. An ancestral virion sequence was subsequently inferred by using a maximum likelihood (ML) method. This sequence was then synthesized de novo, cloned into a replicative backbone of an infectious CVB5 cDNA clone, and transfected into HeLa cells. The hypothetical CVB5-P1anc assembled into functional virus particles that displayed phenotypic properties similar to those of contemporary clinical isolates. This is the first report describing the reconstruction and characterization of a fully functional picornavirus with a putative ancestral capsid.  相似文献   

18.
Previously we described Tomato bushy stunt virus (TBSV) vectors, which retained their capsid protein gene and were engineered with magnesium chelatase (ChlH) and phytoene desaturase (PDS) gene sequences from Nicotiana benthamiana. Upon plant infection, these vectors eventually lost the inserted sequences, presumably as a result of recombination. Here, we modified the same vectors to also contain the plant miR171 or miR159 target sequences immediately 3′ of the silencing inserts. We inoculated N. benthamiana plants and sequenced recombinant RNAs recovered from noninoculated upper leaves. We found that while some of the recombinant RNAs retained the microRNA (miRNA) target sites, most retained only the 3′ 10 and 13 nucleotides of the two original plant miRNA target sequences, indicating in planta miRNA-guided RNA-induced silencing complex cleavage of the recombinant TBSV RNAs. In addition, recovered RNAs also contained various fragments of the original sequence (ChlH and PDS) upstream of the miRNA cleavage site, suggesting that the 3′ portion of the miRNA-cleaved TBSV RNAs served as a template for negative-strand RNA synthesis by the TBSV RNA-dependent RNA polymerase (RdRp), followed by template switching by the RdRp and continued RNA synthesis resulting in loss of nonessential nucleotides.Several plant viruses have been developed as tools for various biotechnology applications, including expression platforms for protein production in plants (1, 2, 6) and as gene silencing systems as part of reverse genetics approaches toward understanding host plant gene function (4, 5). For both of these applications, nonviral sequences conferring the desired function are cloned into the virus genome in order to be expressed during replication in plants. One advantage of using viruses engineered with nonviral sequences is flexibility in manipulating these “extra” sequences, which are not essential for viral replication or movement (1). However, recombinant viruses also tend to lose these sequences, causing instability at the insertion site and resulting in loss of function of the recombinant viral vector. The relatively high error rates of viral replicases (7, 14, 24) and the propensity for recombination events (9) contribute to the instability often seen with some viral vector systems.Recombination events in RNA viruses typically result in joining of two noncontiguous RNA segments (16). These could be sequences from two separate RNA molecules or distant regions of the same molecule. Retention by viruses of favorable sequences is selection driven and eliminates sequences that are unnecessary or negatively affect fitness (11, 31), hence making recombination critical to virus evolution (13, 29). Although phylogenetic analyses predict that recombination events have affected evolution for essentially all groups of RNA viruses (3), some viruses appear to be more prone to recombination than others. For example, plant-infecting supergroup II viruses of the family Tombusviridae appear to undergo frequent recombination, as is supported by the many well-characterized defective-interfering (DI) RNAs of Tomato bushy stunt virus (TBSV) (10, 30). The TBSV DI RNAs, derived entirely from the parental viral RNA, are not replication competent alone and depend on the parent virus to replicate them in trans. Recent developments of in vitro systems (19, 21) have further enhanced dissection of recombination mechanisms giving rise to TBSV DI RNAs.Of the proposed mechanisms for viral recombination (12, 20), the copy choice or template-switching mechanism is the most widely reported (8, 16, 18). This occurs when the viral replicase and its attached nascent polynucleotide chain switches viral RNA templates (or jumps locations on the same template) when making cRNA. Some properties for preferred donor/acceptor sites (sequences in the RNA molecule at which viral replicase switches from one template to another) are known for various RNA viruses (3, 20, 27), suggesting that recombination is not entirely random.The previously described TBSV vectors were efficient silencers of host genes but only while the inserted sequences were retained (23). Thus, optimizing viral vectors requires a better understanding of factors responsible for recombination and consequent loss of insert sequences. In order to address possible recombination mechanisms, we used previously characterized sequence-specific microRNA (miRNA)-guided cleavage determinants as parts of our TBSV vectors. We introduced the miRNA target site sequences for miR171 or miR159 3′ to the silencing inserts of our TBSV vectors (23). After plant inoculation, we analyzed recombinant virus sequences, determining specific recombination patterns, and demonstrated miRNA-mediated recombination events in vivo for the recombinant TBSV vectors. We also showed miRNA-guided RNA-induced silencing complex (RISC)-mediated cleavage for recombinant TBSV RNA and evidence supporting the TBSV RNA-dependent RNA polymerase (RdRp) switching templates during cRNA synthesis.  相似文献   

19.
White root rot, caused by the ascomycete Rosellinia necatrix, is a devastating disease worldwide, particularly in fruit trees in Japan. Here we report on the biological and molecular properties of a novel bipartite double-stranded RNA (dsRNA) virus encompassing dsRNA-1 (8,931 bp) and dsRNA-2 (7,180 bp), which was isolated from a field strain of R. necatrix, W779. Besides the strictly conserved 5′ (24 nt) and 3′ (8 nt) terminal sequences, both segments show high levels of sequence similarity in the long 5′ untranslated region of approximately 1.6 kbp. dsRNA-1 and -2 each possess two open reading frames (ORFs) named ORF1 to -4. Although the protein encoded by 3′-proximal ORF2 on dsRNA-1 shows sequence identities of 22 to 32% with RNA-dependent RNA polymerases from members of the families Totiviridae and Chrysoviridae, the remaining three virus-encoded proteins lack sequence similarities with any reported mycovirus proteins. Phylogenetic analysis showed that the W779 virus belongs to a separate clade distinct from those of other known mycoviruses. Purified virions ∼50 nm in diameter consisted of dsRNA-1 and -2 and a single major capsid protein of 135 kDa, which was shown by peptide mass fingerprinting to be encoded by dsRNA-1 ORF1. We developed a transfection protocol using purified virions to show that the virus was responsible for reduction of virulence and mycelial growth in several host strains. These combined results indicate that the W779 virus is a novel bipartite dsRNA virus with potential for biological control (virocontrol), named Rosellinia necatrix megabirnavirus 1 (RnMBV1), that possibly belongs to a new virus family.Viruses are found ubiquitously in major groups of filamentous fungi (40), and an increasing number of novel mycoviruses are being reported (3, 36). Mycoviruses with RNA genomes are now classified into 10 families, of which four accommodate double-stranded RNA (dsRNA) viruses and the remaining six comprise single-stranded RNA (ssRNA) viruses (23). While many ssRNA mycoviruses, like hypoviruses and endornaviruses, do not produce particles, dsRNA virus genomes, whether undivided (the family Totiviridae) or divided (11 or 12 segments for the family Reoviridae, 4 segments for the family Chrysoviridae, and 2 segments for the family Partitiviridae), are encapsidated in rigid particles. Most mycoviruses are considered to cause cryptic infections, while some cause phenotypic alterations that include hypovirulence and debilitation. However, the lack of artificial introduction methods for most mycoviruses has greatly hampered progress in exploring mycovirus-host interactions (23, 40). Thus, a virus etiology of altered fungal phenotypes was established only for a limited number of examples, including hypovirus-C. parasitica and mycoreovirus-C. parasitica.White root rot is one of the most devastating diseases of perennial crops worldwide, particularly highly valued fruits in Japan like apple, Japanese pear, and grapevine. The causal fungus, Rosellinia necatrix, is an ascomycete with a wide range of host plants of >197 species spanning 50 families (31) and is difficult to control by conventional methods, as is often the case for soilborne pathogens. Fungicide application, though it may be effective, is labor-intensive and raises environmental concerns, while cultural practices may not be effective. Successful biocontrol of chestnut blight disease in Europe with hypovirulent strains (25, 38) inspired a group of Japanese researchers to conduct an extensive search of a large collection of >1,000 field fungal isolates for mycoviruses that might serve as virocontrol agents. Virocontrol or virological control refers to one form of biological control utilizing viruses that infect organisms pathogenic to useful organisms (23). Approximately 20% of the collected isolates of R. necatrix were found to be dsRNA positive and presumed to be infected by mycoviruses (4, 29). Agarose gel profiles of dsRNAs suggested infections by members in the families Totiviridae, Partitiviridae, Reoviridae, and Chrysoviridae, as well as unassigned viruses (S. Kanematsu and A. Sasaki, unpublished results). Among those dsRNAs, the genomic segments of Mycoreovirus 3 (MyRV3) (55) and Rosellinia necatrix partitivirus 1 (RnPV1) (44) were well characterized. However, many other dsRNAs remain uncharacterized.Artificial virion introduction protocols, which are often unavailable for mycoviruses, have been developed for specific viruses infecting the white root rot fungus. Using a polyethylene glycol (PEG)-mediated method, as established for MyRV1 and MyRV2 infecting C. parasitaca (27, 28), RnPV1 and MyRV3 were shown to be infectious as particles (45, 46). Subsequently, the cause-effect relationship was established: MyRV3 was demonstrated to confer hypovirulence (attenuated virulence) on an isogenic strain and a few vegetatively incompatible virulent strains of R. necatrix (33, 45), while RnPV1 was shown to be associated with symptomless infection. Protoplast fusion is also available for introduction of partitiviruses and uncharacterized viruses into recipient fungal strains that are vegetatively incompatible with virus-containing ones (A. Sasaki, unpublished results). Furthermore, DNA transformation systems are available for foreign gene expression in R. necatrix (33, 42). These technical advances have made the R. necatrix-mycovirus systems attractive for studies of virus-host interactions and virocontrol (23, 37).R. necatrix strain W779 was isolated by Ikeda et al. (29, 30) from soil in Ibaraki Prefecture as a dsRNA-positive strain that had yet to be characterized. Here we describe the purification and biological and molecular properties of a novel virus isolated from W779. Particles ∼50 nm in diameter isolated from strain W779 consist of two dsRNA elements termed dsRNA-1 and -2 of approximately 9 and 7 kbp and a major protein of 135 kDa encoded by one of two open reading frames (ORFs) on dsRNA-1. Importantly, purified virus particles were shown to be infectious and confer hypovirulence on vegetatively incompatible fungal strains. The two dsRNA segments share the conserved terminal sequences at both ends, and both possess extremely long (>1.6 kb) 5′ untranslated regions (UTRs) similar to each other, two ORFs, and relatively short 3′ UTRs. The 3′-proximal ORF of dsRNA-1 encodes an RNA-dependent RNA polymerase (RdRp) showing low levels (22 to 32%) of sequence identity to those of members of the families Totiviridae and Chrysoviridae. A phylogenetic analysis with RdRp sequences revealed that the W779 virus is placed into a separate clade from the recognized virus families. These attributes indicate that dsRNA-1 and -2 represent the genome segments of a novel bipartite virus, designated Rosellinia necatrix megabirnavirus 1 (RnMBV1), with virolocontrol agent potential. We propose the establishment of a new family, Megabirnaviridae, to accommodate RnMBV1 as the type species.  相似文献   

20.
During the search for haloarchaeal viruses, we isolated and characterized a new pleomorphic lipid-containing virus, Haloarcula hispanica pleomorphic virus 1 (HHPV-1), that infects the halophilic archaeon Haloarcula hispanica. The virus contains a circular double-stranded DNA genome of 8,082 bp in size. The organization of the genome shows remarkable synteny and amino acid sequence similarity to the genome and predicted proteins of the halovirus HRPV-1, a pleomorphic single-stranded DNA virus that infects a halophilic archaeon Halorubrum sp. Analysis of the two halovirus sequences, as well as the entire nucleotide sequence of the 10.8-kb pHK2-plasmid and a 12.6-kb chromosomal region in Haloferax volcanii, allows us to suggest a new group of closely related viruses with genomes of either single-stranded or double-stranded DNA. Currently, closely related viruses are considered to have the same genome type. Our observation clearly contradicts this categorization and indicates that we should reconsider the way we classify viruses. Our results also provide a new example of related viruses where the viral structural proteins have not diverged as much as the proteins associated with genome replication. This result further strengthens the proposal for higher-order classification to be based on virion architecture rather than on genome type or replication mechanism.Metagenomic studies have increased the amount of information on the nucleotide sequence space in our environment. It has also increased our awareness of the viral abundance and diversity not recognized before (16, 24, 26). Along with this new information, we have learned to acknowledge the significance of viruses in the evolution and behavior of other organisms (55). To reveal the dynamics and molecular interactions in the interplay between a particular virus and its host, however, isolation of single viruses and their hosts is needed. Even though a number of viruses pathogenic to humans, domestic animals, and plants, as well as some bacteriophages, have been studied in great detail, much of the diversity of the archaeal viruses has remained unknown. By the year 2007 only 44 archaeal viruses had been described (2). That embraces less than 1% of all reported viruses. Although the diversity among these few isolated archaeal viruses is considerable, a head-and-tail morphology is prevalent among isolated viruses infecting euryarchaeal cells. In contrast, viruses of Crenarchaeota are diverse and often unusual with no viruses having a head-tail morphology (53).Archaeal haloviruses infect euryarchaeal hosts living in environments up to saturated salt. This makes them an interesting group of viruses that reside in a very restricted habitat. In samples taken from high salt environments, the Dead Sea and Spanish solar salterns, viral morphotypes most often observed were spindle-shaped, head-and-tail or tailless icosahedral particles (25, 31, 47). Isolated haloviruses, however, do not seem to reflect the proportions of different morphotypes found in the nature as nearly all of the isolates possess a head-and-tail morphology (2). Molecular level studies on only two spindle-shaped (10, 11) and one tailless icosahedral particle have been carried out (37, 51). Virus-like particles of other morphologies have also been observed in high-salt environments (47), but only one additional morphotype has been described in detail (50). This recently isolated lipid containing halovirus, HRPV-1, is the first archaeal virus containing a single-stranded DNA (ssDNA) genome (50). It infects Halorubrum sp. and has a pleomorphic appearance with glycosylated spike structures protruding from its external membrane (49, 50).The evolution of prokaryotic viral genome sequences is very fast (18), and the assessment of viral relationships using homology of the genome sequences applies only to closely related viruses (17, 19). Current higher-order classification of viruses is based on the host organism, the nature of the genome (RNA/DNA, single stranded versus double stranded) and the virion morphology. Recently, a higher-order clustering of virus families has been proposed based on common principles of virion architectures as well as on the fold of the major capsid protein (1, 6, 12, 13, 42). Consequently, major capsid proteins most probably belong to the vertically inherited viral “self” (4), whereas proteins involved in replication of the viral genome can be swapped by horizontal exchange (21, 63). The proposal is based on observations that structurally related viruses have been found to infect organisms that reside in all three domains of life.We have isolated a new pleomorphic virus infecting Haloarcula hispanica (Har. hispanica pleomorphic virus 1 [HHPV-1]). Here, we determine the molecular constituents of HHPV-1 and its genetic relatedness to other archaeal viruses and putative proviruses. Sequence homology and gene order (synteny) shows distinct genomic regions shared between four genetic elements separating replication, virus assembly, and integration functions. Surprisingly, in spite of the close relatedness of HRPV-1 and HHPV-1, the genome types of these two viruses differ (ssDNA and dsDNA, respectively).  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