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 984 毫秒
1.
Bats are hosts to a variety of viruses capable of zoonotic transmissions. Because of increased contact between bats, humans, and other animal species, the possibility exists for further cross-species transmissions and ensuing disease outbreaks. We describe here full and partial viral genomes identified using metagenomics in the guano of bats from California and Texas. A total of 34% and 58% of 390,000 sequence reads from bat guano in California and Texas, respectively, were related to eukaryotic viruses, and the largest proportion of those infect insects, reflecting the diet of these insectivorous bats, including members of the viral families Dicistroviridae, Iflaviridae, Tetraviridae, and Nodaviridae and the subfamily Densovirinae. The second largest proportion of virus-related sequences infects plants and fungi, likely reflecting the diet of ingested insects, including members of the viral families Luteoviridae, Secoviridae, Tymoviridae, and Partitiviridae and the genus Sobemovirus. Bat guano viruses related to those infecting mammals comprised the third largest group, including members of the viral families Parvoviridae, Circoviridae, Picornaviridae, Adenoviridae, Poxviridae, Astroviridae, and Coronaviridae. No close relative of known human viral pathogens was identified in these bat populations. Phylogenetic analysis was used to clarify the relationship to known viral taxa of novel sequences detected in bat guano samples, showing that some guano viral sequences fall outside existing taxonomic groups. This initial characterization of the bat guano virome, the first metagenomic analysis of viruses in wild mammals using second-generation sequencing, therefore showed the presence of previously unidentified viral species, genera, and possibly families. Viral metagenomics is a useful tool for genetically characterizing viruses present in animals with the known capability of direct or indirect viral zoonosis to humans.Bats belong to one of the most diverse, abundant, and widely distributed group of mammals. More than 1,100 bat species belong to the order of Chiroptera, representing approximately 20% of all mammalian species (54). Most bat species feed on insects and other arthropods, while others feed on fruit nectar, bird or mammal blood, and small vertebrates such as fish, frogs, mice, and birds (30). Of the 47 species of bats reported in the United States, most of them are insectivorous (http://www.batcon.org/).Bats are considered the natural reservoir of a large variety of zoonotic viruses causing serious human diseases such as lyssaviruses, henipaviruses, severe acute respiratory syndrome coronavirus, and Ebola virus (6, 38, 46, 59, 63, 65). Characteristics of bats, including their genetic diversity, broad geological distribution, gregarious habits, high population density, migratory habits, and long life span (30, 58), likely endow them with the ability to host diverse viruses, some of which are also able to infect humans and other mammals (41, 63).More than 80 virus species have been isolated or detected in bats using nucleic acid-based methods (6, 38, 59, 65). Viruses that have been recently discovered in bats include astroviruses, adeno-associated viruses (AAVs), adenoviruses, herpesviruses, and polyomavirus (8, 9, 13, 31, 32, 35, 37, 39, 40, 42, 61, 62, 68). For example, it was recently reported that a newly identified adenovirus isolated from bat guano was capable of infecting various vertebrate cell lines, including those of humans, monkeys, dogs, and pigs (35). With increasing human populations in previously wild areas, contact of bats with humans and with wild and domestic animals has increased, providing greater opportunities for cross-species transmissions of potentially pathogenic bat viruses. To better understand the range of viruses carried by bats, we undertook an initial characterization of the guano viromes of several common bat species in the United States.The development of massively parallel sequencing technology makes is possible to reveal uncultured viral assemblages within biological or environmental samples (11, 28). To date, this approach has been used to characterize viruses in equine feces (7), human blood (5), tissue (14), human feces (3, 4, 15, 45, 60, 67), and human respiratory secretions (64), which in turn has facilitated the discovery of many novel viruses (18, 20, 25, 33, 47, 50). In the present study, we analyzed the viruses present in guano from several bat species in California and Texas, using sequence-independent PCR amplification, pyrosequencing, and sequence similarity searches.  相似文献   

2.
The combinatorial nature of genetic recombination can potentially provide organisms with immediate access to many more positions in sequence space than can be reached by mutation alone. Recombination features particularly prominently in the evolution of a diverse range of viruses. Despite rapid progress having been made in the characterization of discrete recombination events for many species, little is currently known about either gross patterns of recombination across related virus families or the underlying processes that determine genome-wide recombination breakpoint distributions observable in nature. It has been hypothesized that the networks of coevolved molecular interactions that define the epistatic architectures of virus genomes might be damaged by recombination and therefore that selection strongly influences observable recombination patterns. For recombinants to thrive in nature, it is probably important that the portions of their genomes that they have inherited from different parents work well together. Here we describe a comparative analysis of recombination breakpoint distributions within the genomes of diverse single-stranded DNA (ssDNA) virus families. We show that whereas nonrandom breakpoint distributions in ssDNA virus genomes are partially attributable to mechanistic aspects of the recombination process, there is also a significant tendency for recombination breakpoints to fall either outside or on the peripheries of genes. In particular, we found significantly fewer recombination breakpoints within structural protein genes than within other gene types. Collectively, these results imply that natural selection acting against viruses expressing recombinant proteins is a major determinant of nonrandom recombination breakpoint distributions observable in most ssDNA virus families.Genetic recombination is a ubiquitous biological process that is both central to DNA repair pathways (10, 57) and an important evolutionary mechanism. By generating novel combinations of preexisting nucleotide polymorphisms, recombination can potentially accelerate evolution by increasing the population-wide genetic diversity upon which adaptive selection relies. Recombination can paradoxically also prevent the progressive accumulation of harmful mutations within individual genomes (18, 35, 53). Whereas its ability to defend high-fitness genomes from mutational decay possibly underlies the evolutionary value of sexuality in higher organisms, in many microbial species where pseudosexual genetic exchange is permissible among even highly divergent genomes, recombination can enable access to evolutionary innovations that would otherwise be inaccessible by mutation alone.Such interspecies recombination is fairly common in many virus families (8, 17, 27, 44, 82). It is becoming clear, however, that as with mutation events, most recombination events between distantly related genomes are maladaptive (5, 13, 38, 50, 63, 80). As genetic distances between parental genomes increase, so too does the probability of fitness defects in their recombinant offspring (16, 51). The viability of recombinants is apparently largely dependent on how severely recombination disrupts coevolved intragenome interaction networks (16, 32, 51). These networks include interacting nucleotide sequences that form secondary structures, sequence-specific protein-DNA interactions, interprotein interactions, and amino acid-amino acid interactions within protein three-dimensional folds.One virus family where such interaction networks appear to have a large impact on patterns of natural interspecies recombination are the single-stranded DNA (ssDNA) geminiviruses. As with other ssDNA viruses, recombination is very common among the species of this family (62, 84). Partially conserved recombination hot and cold spots have been detected in different genera (39, 81) and are apparently caused by both differential mechanistic predispositions of genome regions to recombination and natural selection disfavoring the survival of recombinants with disrupted intragenome interaction networks (38, 51).Genome organization and rolling circle replication (RCR)—the mechanism by which geminiviruses and many other ssDNA viruses replicate (9, 67, 79; see reference 24 for a review)—seem to have a large influence on basal recombination rates in different parts of geminivirus genomes (20, 33, 39, 61, 81). To initiate RCR, virion-strand ssDNA molecules are converted by host-mediated pathways into double-stranded “replicative-form” (RF) DNAs (34, 67). Initiated by a virus-encoded replication-associated protein (Rep) at a well-defined virion-strand replication origin (v-ori), new virion strands are synthesized on the complementary strand of RF DNAs (28, 73, 74) by host DNA polymerases. Virion-strand replication is concomitant with the displacement of old virion strands, which, once complete, yields covalently closed ssDNA molecules which are either encapsidated or converted into additional RF DNAs. Genome-wide basal recombination rates in ssDNA viruses are probably strongly influenced by the specific characteristics of host DNA polymerases that enable RCR. Interruption of RCR has been implicated directly in geminivirus recombination (40) and is most likely responsible for increased basal recombination rates both within genes transcribed in the opposite direction from that of virion-strand replication (40, 71) and at the v-ori (1, 9, 20, 69, 74).Whereas most ssDNA virus families replicate via either a rolling circle mechanism (the Nanoviridae, Microviridae, and Geminiviridae) (3, 23, 24, 31, 59, 67, 74) or a related rolling hairpin mechanism (the Parvoviridae) (25, 76), among the Circoviridae only the Circovirus genus is known to use RCR (45). Although the Gyrovirus genus (the other member of the Circoviridae) and the anelloviruses (a currently unclassified ssDNA virus group) might also use RCR, it is currently unknown whether they do or not (78). Additionally, some members of the Begomovirus genus of the Geminiviridae either have a second genome component, called DNA-B, or are associated with satellite ssDNA molecules called DNA-1 and DNA-Beta, all of which also replicate by RCR (1, 47, 68).Recombination is known to occur in the parvoviruses (19, 43, 70), microviruses (66), anelloviruses (40, 46), circoviruses (11, 26, 60), nanoviruses (30), geminivirus DNA-B components, and geminivirus satellite molecules (2, 62). Given that most, if not all, of these ssDNA replicons are evolutionarily related to and share many biological features with the geminiviruses (22, 31, 36), it is of interest to determine whether conserved recombination patterns observed in the geminiviruses (61, 81) are evident in these other groups. To date, no comparative analyses have ever been performed with different ssDNA virus families to identify, for example, possible influences of genome organization on recombination breakpoint distributions found in these viruses.Here we compare recombination frequencies and recombination breakpoint distributions in most currently described ssDNA viruses and satellite molecules and identify a number of sequence exchange patterns that are broadly conserved across this entire group.  相似文献   

3.
Human fecal matter contains a large number of viruses, and current bacterial indicators used for monitoring water quality do not correlate with the presence of pathogenic viruses. Adenoviruses and enteroviruses have often been used to identify fecal pollution in the environment; however, other viruses shed in fecal matter may more accurately detect fecal pollution. The purpose of this study was to develop a baseline understanding of the types of viruses found in raw sewage. PCR was used to detect adenoviruses, enteroviruses, hepatitis B viruses, herpesviruses, morbilliviruses, noroviruses, papillomaviruses, picobirnaviruses, reoviruses, and rotaviruses in raw sewage collected throughout the United States. Adenoviruses and picobirnaviruses were detected in 100% of raw sewage samples and 25% and 33% of final effluent samples, respectively. Enteroviruses and noroviruses were detected in 75% and 58% of raw sewage samples, respectively, and both viral groups were found in 8% of final effluent samples. This study showed that adenoviruses, enteroviruses, noroviruses, and picobirnaviruses are widespread in raw sewage. Since adenoviruses and picobirnaviruses were detected in 100% of raw sewage samples, they are potential markers of fecal contamination. Additionally, this research uncovered previously unknown sequence diversity in human picobirnaviruses. This baseline understanding of viruses in raw sewage will enable educated decisions to be made regarding the use of different viruses in water quality assessments.Millions of viruses and bacteria are excreted in human fecal matter (5, 17, 82), and current methods of sewage treatment do not always effectively remove these organisms (74, 76-78). The majority of treated wastewater, as well as untreated sewage, drains into the marine environment (1) and has the potential to threaten environmental (e.g., nutrients and chemicals) (45) and public (e.g., pathogen exposure via swimming and seafood consumption) (1, 24, 28, 29, 33, 44, 57, 63) health. Currently, the U.S. Environmental Protection Agency (EPA) mandates the use of bacterial indicators such as fecal coliforms and enterococci to assess water quality (75). Although monitoring of these bacteria is simple and inexpensive, it has been shown that fecal-associated bacteria are not ideal indicators of fecal pollution.Since fecal-associated bacteria are able to live in sediments in the absence of fecal pollution (18, 32, 55), their resuspension into the water column can result in false-positive results and mask correlations between their concentrations and the extent of recent fecal pollution. Another unfavorable characteristic of current bacterial indicators is their inability to predict or correlate with the presence of pathogenic viruses (25, 40, 41, 64, 80). Human-pathogenic viruses associated with feces are generally more robust than enteric bacteria and are not as easily eliminated by current methods of wastewater treatment (43, 80). For example, adenoviruses are more resilient to tertiary wastewater treatment and UV disinfection than are bacterial indicators of fecal pollution (74). Since bacterial indicators cannot accurately depict the risks to human health from fecal pollution, several studies have proposed the use of a viral indicator of wastewater contamination (35, 41, 61).While it is impractical to monitor the presence of all viral pathogens related to wastewater pollution, the development of an accurate viral indicator of sewage contamination is needed for enhanced water quality monitoring. Enteric viruses (including viruses belonging to the families Adenoviridae, Caliciviridae, Picornaviridae, and Reoviridae) are transmitted via the fecal-oral route and are known to be abundant in raw sewage. These viruses have been used to identify fecal pollution in coastal environments throughout the world (27, 35, 39, 40, 48, 50, 56, 57, 63, 64, 67-69, 71, 80). To determine which viruses are effective indicators of fecal pollution, it is first necessary to establish a broad, baseline understanding of the many diverse groups of eukaryotic viruses in raw sewage. Several studies have identified adenoviruses, noroviruses, reoviruses, rotaviruses, and other enteroviruses (e.g., polioviruses, coxsackie viruses, and echoviruses) in raw sewage in Australia, Europe, and South Africa (30, 47, 58, 76-78). However, no broad baseline data on the presence of eukaryotic viruses in raw sewage in the United States currently exist.This study determined the presence of 10 viral groups (adenoviruses, enteroviruses, hepatitis B viruses, herpesviruses, morbilliviruses, noroviruses, papillomaviruses, picobirnaviruses, reoviruses, and rotaviruses) in raw sewage samples collected throughout the United States. All viral groups that were detected in raw sewage were then examined further to determine if they were also present in final treated wastewater effluent. These 10 viral groups were chosen because of their potential to be transmitted via the fecal-oral route, suggesting that they might be found in raw sewage. Many of these viruses (excluding adenoviruses, enteroviruses, noroviruses, reoviruses, and rotaviruses) have not been studied in sewage despite their likely presence. Picobirnaviruses have been detected in individual fecal samples (12, 70, 79, 82); however, their presence has never been analyzed in collective waste, nor have they been proposed to be potential markers of fecal pollution. This study identified potential viral indicators of fecal pollution and will have important applications to water quality monitoring programs throughout the country.  相似文献   

4.
5.
6.
H2 influenza viruses have not circulated in humans since 1968, and therefore a large segment of the population would likely be susceptible to infection should H2 influenza viruses reemerge. The development of an H2 pandemic influenza virus vaccine candidate should therefore be considered a priority in pandemic influenza preparedness planning. We selected a group of geographically and temporally diverse wild-type H2 influenza viruses and evaluated the kinetics of replication and compared the ability of these viruses to induce a broadly cross-reactive antibody response in mice and ferrets. In both mice and ferrets, A/Japan/305/1957 (H2N2), A/mallard/NY/1978 (H2N2), and A/swine/MO/2006 (H2N3) elicited the broadest cross-reactive antibody responses against heterologous H2 influenza viruses as measured by hemagglutination inhibition and microneutralization assays. These data suggested that these three viruses may be suitable candidates for development as live attenuated H2 pandemic influenza virus vaccines.Influenza pandemics occur when a novel influenza virus enters a population with little preexisting immunity (36). During the pandemics of the last century, novel influenza viruses were introduced either directly from an avian reservoir (34) or were the result of reassortment between contemporaneously circulating human, avian, and swine influenza viruses (5, 29, 36). Due to the lack of preexisting immunity to the novel virus, morbidity and mortality rates are typically higher than in epidemics caused by seasonal influenza viruses (4).Although pandemic preparedness planning has largely focused on the highly pathogenic H5 and H7 avian influenza virus subtypes, the recent emergence of the 2009 pandemic H1N1 viruses underscores the need to consider other influenza virus subtypes as well. Of the 16 hemagglutinin (HA) influenza A virus subtypes that have been identified to date, H1, H2, and H3 have been known to cause influenza pandemics (7, 27), suggesting that these viruses are capable of sustained transmission and can cause disease in humans. While the H1 and H3 subtypes have cocirculated in humans since 1977, H2 influenza viruses have not circulated in humans since 1968 (36) and therefore a large segment of the population would likely be susceptible to infection should H2 influenza viruses reemerge. The 1957 H2 pandemic virus was a reassortant that derived the HA, neuraminidase (NA), and PB1 genes from an avian virus and the remaining gene segments from the circulating H1N1 virus (15, 30). As H2 subtype viruses continue to circulate in avian reservoirs worldwide (12, 17, 18, 22, 33), they remain a potential pandemic threat. The development of an H2 influenza virus vaccine candidate should therefore be considered a priority in future pandemic influenza preparedness planning.Given the low likelihood that a previously selected vaccine virus will exactly match the pandemic virus, the ability to elicit a broadly cross-reactive antibody response to antigenically distinct viruses within a subtype is an important consideration in the selection of a pandemic influenza vaccine candidate. Previous studies have examined the ability of inactivated H2 influenza viruses to provide cross-protection against mouse-adapted variants of reassortant human viruses and an avian H2 influenza virus from 1978 (9, 14). Given the potential for live attenuated influenza virus vaccines to confer a great breadth of heterologous cross-protection (1, 2, 6, 35), we recently conducted a study evaluating cold-adapted A/Ann Arbor/6/1960 (AA CA), an H2 influenza virus used as the backbone of the seasonal live attenuated influenza A virus vaccine currently licensed in the United States (3). However, as H2 influenza virus continues to circulate widely and appear in migratory birds (10, 24, 26), in poultry markets (20), and in swine (21), with evidence of interregional gene transmission (19, 22), a more extensive evaluation of recent isolates may be warranted in the selection of a potential H2 pandemic vaccine candidate.H2 influenza viruses fall into three main lineages: a human lineage, a North American avian lineage, and a Eurasian avian lineage (29). In addition to viruses whose replicative ability in mammals has previously been established (11, 21, 23, 25), we selected a group of geographically and temporally diverse H2 influenza viruses from each lineage. We evaluated the kinetics of replication of each of these viruses in mice and ferrets and compared the abilities of these viruses to induce a broadly cross-reactive antibody response to determine which of these viruses would be suitable for further development as an H2 pandemic influenza vaccine candidate.  相似文献   

7.
Understanding the mechanisms underlying potential altered susceptibility to human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) individuals and the later clinical consequences of breakthrough infection can provide insight into strategies to control HIV-1 with an effective vaccine. From our Seattle ES cohort, we identified one individual (LSC63) who seroconverted after over 2 years of repeated unprotected sexual contact with his HIV-1-infected partner (P63) and other sexual partners of unknown HIV-1 serostatus. The HIV-1 variants infecting LSC63 were genetically unrelated to those sequenced from P63. This may not be surprising, since viral load measurements in P63 were repeatedly below 50 copies/ml, making him an unlikely transmitter. However, broad HIV-1-specific cytotoxic T-lymphocyte (CTL) responses were detected in LSC63 before seroconversion. Compared to those detected after seroconversion, these responses were of lower magnitude and half of them targeted different regions of the viral proteome. Strong HLA-B27-restricted CTLs, which have been associated with disease control, were detected in LSC63 after but not before seroconversion. Furthermore, for the majority of the protein-coding regions of the HIV-1 variants in LSC63 (except gp41, nef, and the 3′ half of pol), the genetic distances between the infecting viruses and the viruses to which he was exposed through P63 (termed the exposed virus) were comparable to the distances between random subtype B HIV-1 sequences and the exposed viruses. These results suggest that broad preinfection immune responses were not able to prevent the acquisition of HIV-1 infection in LSC63, even though the infecting viruses were not particularly distant from the viruses that may have elicited these responses.Understanding the mechanisms of altered susceptibility or control of human immunodeficiency virus type 1 (HIV-1) infection in highly exposed seronegative (ES) persons may provide invaluable information aiding the design of HIV-1 vaccines and therapy (9, 14, 15, 33, 45, 57, 58). In a cohort of female commercial sex workers in Nairobi, Kenya, a small proportion of individuals remained seronegative for over 3 years despite the continued practice of unprotected sex (12, 28, 55, 56). Similarly, resistance to HIV-1 infection has been reported in homosexual men who frequently practiced unprotected sex with infected partners (1, 15, 17, 21, 61). Multiple factors have been associated with the resistance to HIV-1 infection in ES individuals (32), including host genetic factors (8, 16, 20, 37-39, 44, 46, 47, 49, 59, 63), such as certain HLA class I and II alleles (41), as well as cellular (1, 15, 26, 55, 56), humoral (25, 29), and innate immune responses (22, 35).Seroconversion in previously HIV-resistant Nairobi female commercial sex workers, despite preexisting HIV-specific cytotoxic T-lymphocyte (CTL) responses, has been reported (27). Similarly, 13 of 125 ES enrollees in our Seattle ES cohort (1, 15, 17) have become late seroconverters (H. Zhu, T. Andrus, Y. Liu, and T. Zhu, unpublished observations). Here, we analyze the virology, genetics, and immune responses of HIV-1 infection in one of the later seroconverting subjects, LSC63, who had developed broad CTL responses before seroconversion.  相似文献   

8.
Isolation of human subtype H3N2 influenza viruses in embryonated chicken eggs yields viruses with amino acid substitutions in the hemagglutinin (HA) that often affect binding to sialic acid receptors. We used a glycan array approach to analyze the repertoire of sialylated glycans recognized by viruses from the same clinical specimen isolated in eggs or cell cultures. The binding profiles of whole virions to 85 sialoglycans on the microarray allowed the categorization of cell isolates into two groups. Group 1 cell isolates displayed binding to a restricted set of α2-6 and α2-3 sialoglycans, whereas group 2 cell isolates revealed receptor specificity broader than that of their egg counterparts. Egg isolates from group 1 showed binding specificities similar to those of cell isolates, whereas group 2 egg isolates showed a significantly reduced binding to α2-6- and α2-3-type receptors but retained substantial binding to specific O- and N-linked α2-3 glycans, including α2-3GalNAc and fucosylated α2-3 glycans (including sialyl Lewis x), both of which may be important receptors for H3N2 virus replication in eggs. These results revealed an unexpected diversity in receptor binding specificities among recent H3N2 viruses, with distinct patterns of amino acid substitution in the HA occurring upon isolation and/or propagation in eggs. These findings also suggest that clinical specimens containing viruses with group 1-like receptor binding profiles would be less prone to undergoing receptor binding or antigenic changes upon isolation in eggs. Screening cell isolates for appropriate receptor binding properties might help focus efforts to isolate the most suitable viruses in eggs for production of antigenically well-matched influenza vaccines.Influenza A viruses are generally isolated and propagated in embryonated chicken eggs or in cultures of cells of mammalian origin. Human influenza viruses were previously noted to acquire mutations in the hemagglutinin (HA) gene upon isolation and culture in the allantoic sac of embryonated chicken eggs (herein simply referred to as “eggs”) compared to the sequences of those isolated in mammalian cell substrates (herein referred to as “cells”) (29, 30, 44, 53, 58). These mutations resulted in amino acid substitutions that were found to mediate receptor specificity changes and improved viral replication efficiency in eggs (37). In general, cell-grown viruses are assumed to be more similar than their egg-grown counterparts to the viruses present in respiratory secretions (30, 56). Since their emergence in 1968, influenza A (H3N2) viruses have evolved and adapted to the human host while losing their ability to be efficiently isolated and replicate in eggs, particularly after 1992 (37, 42, 48). The rate of isolation of H3N2 clinical specimens after inoculation into eggs can be up to ∼30 times lower than that in mammalian cell cultures, highlighting the strong selective pressure for the emergence of sequence variants (77).Virtually all influenza vaccines for human use were licensed decades ago by national regulatory authorities, which used a product manufactured from influenza viruses isolated and propagated exclusively in eggs; therefore, cell culture isolates have been unacceptable for this purpose (41, 71). The antigen composition of influenza vaccines requires frequent updates (every 2 years, on average) to closely match their antigenic properties to the most prevalent circulating antigenic drift variant viruses (51). The limited availability of H3N2 viruses isolated in eggs has on one or more occasions delayed vaccine composition updates and may have reduced the efficacy of vaccination against new antigenically drifted viruses (3, 34, 37).Entry of influenza viruses into host cells is mediated by HA, which binds to sialic acid containing glycoconjugates on the surface of epithelial cells in the upper respiratory tract (2, 13). The nature of the linkage between sialic acid and the vicinal sugar (usually galactose) varies in different host species and tissues and may therefore determine whether an influenza virus binds to and infects avian or human cells (40, 46, 59, 62, 72-75). Human influenza viruses preferentially bind to α2-6-linked sialic acids, and avian viruses predominantly bind to α2-3-linked sialic acids (59). Previous studies with chicken embryo chorioallantoic membranes revealed differential lectin binding, suggesting that α2-3-linked but not α2-6-linked sialosides are present on the epithelial cells (28). Human H3N2 viruses isolated in cell culture were reported to bind with a high affinity to α2-6-linked sialosides, while viruses isolated in eggs often had increased specificity for α2-3-linked sialosides (19, 20, 28). The functional classification of avian and mammalian influenza virus receptors is further complicated since in vitro and tissue-binding assays have led to new working hypotheses involving glycan chain length, topology, and the composition of the inner fragments of the carbohydrate chain as additional receptor specificity determinants (9, 17, 65, 66, 82). However, the significance of these in vitro properties remains unknown, since the structures of the natural sialosides on host cells that are used for infectious virus entry are undefined.The techniques most widely used to study the interactions of the influenza virus with host cell receptors employ animal cells in various assay formats (36, 57, 59, 64, 69). To overcome the problems of cell-based techniques, new assays that rely on labeled sialyl-glycoproteins or polymeric sialoglycans have been developed (18). However, these assays are limited by having only a few glycans available in polymeric form and offer low throughput. In contrast, glycan microarrays can assess virus binding to multiple well-defined glycans simultaneously. Previous work with influenza live or β-propiolactone (BPL)-inactivated virions as well as recombinantly produced HAs revealed a good correlation with receptor specificity compared to that achieved by other methods of analysis (4, 11, 57, 58, 65-68).Here we have compared paired isolates derived in eggs or cell cultures from the single clinical specimen to better understand their receptor binding specificity and its implications for vaccine production. We examined the differences in the sequences of the HAs between egg- and cell-grown isolates and analyzed their receptor binding profiles using glycan microarrays. Sequence analysis of the HA and glycan binding results revealed two distinct groups of viruses, with many egg isolates showing unexpectedly reduced levels of binding to α2-3 and α2-6 sialosides compared to the levels for the viruses isolated in mammalian cells. Furthermore, these studies highlighted that specific glycans may be important for H3N2 virus growth in eggs.  相似文献   

9.
The Asian H5N1 highly pathogenic avian influenza (HPAI) viruses have been increasing in pathogenicity in diverse avian species since 1996 and are now widespread in Asian, European, and African countries. To better understand the basis of the increased pathogenicity of recent Asian H5N1 HPAI viruses in chickens, we compared the fevers and mean death times (MDTs) of chickens infected with the Asian H5N1 A/chicken/Yamaguchi/7/04 (CkYM7) strain with those infected with the H5N1 Duck/Yokohama/aq10/03 (DkYK10) strain, using a wireless thermosensor. Asian H5N1 CkYM7 caused peracute death in chickens before fever could be induced, whereas DkYK10 virus induced high fevers and had a long MDT. Real-time PCR analyses of cytokine mRNA expressions showed that CkYM7 quickly induced antiviral and proinflammatory cytokine mRNA expressions at 24 h postinfection (hpi) that suddenly decreased at 32 hpi. In contrast, these cytokine mRNA expressions increased at 24 hpi in the DkYK10 group, but decreased from 48 hpi onward to levels similar to those resulting from infection with the low-pathogenicity H5N2 A/chicken/Ibaraki/1/2004 strain. Sequential titrations of viruses in lungs, spleens, and kidneys demonstrated that CkYM7 replicated rapidly and efficiently in infected chickens and that the viral titers were more than twofold higher than those of DkYK10. CkYM7 preferentially and efficiently replicated in macrophages and vascular endothelial cells, while DkYK10 grew moderately in macrophages. These results indicate that the increased pathogenicity in chickens of the recent Asian H5N1 HPAI viruses may be associated with extremely rapid and high replication of the virus in macrophages and vascular endothelial cells, which resulted in disruption of the thermoregulation system and innate immune responses.Since the first detection of the Asian lineage of highly pathogenic avian influenza (HPAI) virus (H5N1) in southern China in 1996, H5N1 virus infection in birds has continued for 13 years in Asia, acquiring pathogenicity not only in birds but also in mammals. In 1997, the H5N1 Hong Kong isolates caused illness and death in a variety of terrestrial birds and even in humans (9, 37, 48, 49). In 2001, emerging H5N1 Hong Kong isolates were more pathogenic to chickens and the mean death time (MDT) was 2 days without any prior clinical signs (12). In 2003 to 2004, the H5N1 epizootic occurred simultaneously in East Asian countries (22, 30). The 2003/2004 H5N1 isolates caused death in taxonomically diverse avian species, including domestic ducks (46, 47, 51), and humans (7, 55). Furthermore, the first indication of wild aquatic bird involvement occurred at recreational parks in Hong Kong in late 2002 to 2003 (46), and then migratory aquatic bird die-off occurred in 2005 at Qinghai Lake in China (6, 24). The broad host spectrum and increased pathogenicity of H5N1 viruses to diverse bird species raise serious concerns about the worldwide spread of the virus by migratory birds.According to the international criteria, HPAI viruses are defined by over 75% mortality in 4- to 8-week-old chickens following an intravenous pathogenicity test or an intravenous pathogenicity index (IVPI) of more than 1.2 in 6-week-old chickens (34); however, there are some variations in pathogenicity intensity among the HPAI viruses in chickens (1, 3, 5, 12, 15, 28, 31, 48, 50-52, 57). Most of the HPAI viruses that were isolated before 1996 cause severe clinical signs (e.g., ruffled feathers, depression, labored breathing, and neurological signs) and severe gross lesions (e.g., head and face edema, cyanosis, subcutaneous hemorrhages in combs and leg shanks, and necrosis of combs and wattles) in chickens (1, 3, 15, 31, 50, 52, 57). These viruses usually kill chickens 3 to 6 days after intranasal inoculation. On the other hand, the recently emerged Asian H5N1 HPAI viruses are more virulent and kill chickens within 1 to 2 days without causing typical clinical signs and gross lesions (5, 12, 27, 33, 48, 51), although some Asian H5N1 viruses, such as A/Goose/Guangdong/2/96 (23), A/goose/Hong Kong/437-10/99 (17), and A/chicken/Indonesia/7/03 (58), were less virulent. To successfully control HPAI in poultry, it is important to better understand the mechanisms of increased pathogenicity of recent H5N1 HPAI viruses in chickens.The Asian H5N1 HPAI virus has another important characteristic, which is its capability of crossing host-species barriers. It was reported that the H5N1 virus can infect and cause death in mammals such as mice (5, 9, 12, 14, 29), cats (21), tigers (2), ferrets (11, 26), monkeys (40), and humans (7, 49, 55). High-level inductions of proinflammatory cytokines in mammals infected with the H5N1 viruses, referred to as “cytokine storms,” have been hypothesized to contribute to the severity of pathological changes and ultimate death (4, 7, 13, 45, 55). Cytokine and chemokine dysregulation was detected in clinical cases of H5N1-infected humans (8, 13, 36) and also in monkeys experimentally infected with the H1N1 Spanish flu strain (20). In a mouse model, lymphocyte apoptosis and cytokine dysregulation have been proposed to contribute to the severity of the disease caused by H5N1 (56). Investigations with transgenic mice deficient in each cytokine gene suggest that tumor necrosis factor alpha (TNF-α) may contribute to morbidity and interleukin-1 (IL-1) may be important for virus clearance (53). However, mice deficient in TNF-α or IL-6 succumb to infection with H5N1, and cytokine inhibition treatment does not prevent death (42), suggesting that therapies targeting the virus rather than cytokines may be preferable. Thus, the significance of elevated proinflammatory cytokine responses in the pathogenesis of H5N1-infected mammals requires further studies.In contrast, little is known about proinflammatory cytokine responses and their roles in pathogenicity in chickens infected with HPAI viruses, including the recent Asian H5N1 viruses. It was reported that infection with an HPAI virus results in upregulation of gene expression of gamma interferon (IFN-γ) and inducible nitric oxide synthase (58). However, the roles of proinflammatory cytokines in disease severity and outcomes in chickens infected systemically with HPAI viruses are largely unknown. The less-virulent Asian H5N1 virus, which causes severe clinical signs and gross lesions in chickens (17, 23, 27, 58), would be a valuable tool for investigating the role of proinflammatory cytokines in chickens infected with HPAI viruses, as well as for exploring the pathogenesis of the more-virulent Asian H5N1 HPAI virus, because of the antigenic and molecular similarities between them.In this study, we compared the pathogenicities in chickens of the less-virulent and more-virulent Asian H5N1 HPAI viruses based on MDT, fever, cytokine responses, and viral replication. Our results suggest that the shift in the Asian H5N1 virus to increased virulence may be associated with efficient and rapid replication of the virus in chickens, accompanied by early destruction of host immune responses and followed by peracute death before fever can be induced. Finally, we discuss candidate genes that may account for the high pathogenicity of Asian H5N1 HPAI viruses in chickens.  相似文献   

10.
11.
Avian H7 influenza viruses have been responsible for poultry outbreaks worldwide and have resulted in numerous cases of human infection in recent years. The high rate of conjunctivitis associated with avian H7 subtype virus infections may represent a portal of entry for avian influenza viruses and highlights the need to better understand the apparent ocular tropism observed in humans. To study this, mice were inoculated by the ocular route with viruses of multiple subtypes and degrees of virulence. We found that in contrast to human (H3N2 and H1N1) viruses, H7N7 viruses isolated from The Netherlands in 2003 and H7N3 viruses isolated from British Columbia, Canada, in 2004, two subtypes that were highly virulent for poultry, replicated to a significant titer in the mouse eye. Remarkably, an H7N7 virus, as well as some avian H5N1 viruses, spread systemically following ocular inoculation, including to the brain, resulting in morbidity and mortality of mice. This correlated with efficient replication of highly pathogenic H7 and H5 subtypes in murine corneal epithelial sheets (ex vivo) and primary human corneal epithelial cells (in vitro). Influenza viruses were labeled to identify the virus attachment site in the mouse cornea. Although we found abundant H7 virus attachment to corneal epithelial tissue, this did not account for the differences in virus replication as multiple subtypes were able to attach to these cells. These findings demonstrate that avian influenza viruses within H7 and H5 subtypes are capable of using the eye as a portal of entry.Highly pathogenic avian influenza (HPAI) H5N1 viruses, which have resulted in over 420 documented cases of human infection to date, have generally caused acute, often severe and fatal, respiratory illness (1, 50). While conjunctivitis following infection with H5N1 or human influenza viruses has been rare, most human infections associated with H7 subtype viruses have resulted in ocular and not respiratory disease (1, 9, 37, 38). Infrequent reports of human conjunctivitis infection following exposure to H7 influenza viruses date from 1977, predominantly resulting from laboratory or occupational exposure (21, 40, 48). However, in The Netherlands in 2003, more than 80 human infections with H7N7 influenza virus occurred among poultry farmers and cullers amid widespread outbreaks of HPAI in domestic poultry; the majority of these human infections resulted in conjunctivitis (14, 20). Additionally, conjunctivitis was documented in the two human infections resulting from an H7N3 outbreak in British Columbia, Canada, in 2004, as well as in H7N3- and H7N2-infected individuals in the United Kingdom in 2006 and 2007, respectively (13, 18, 29, 46, 51). The properties that contribute to an apparent ocular tropism of some influenza viruses are currently not well understood (30).Host cell glycoproteins bearing sialic acids (SAs) are the cellular receptors for influenza viruses and can be found on epithelial cells within both the human respiratory tract and ocular tissue (26, 31, 41). Both respiratory and ocular tissues additionally secrete sialylated mucins that function in pathogen defense and protection of the epithelial surface (5, 11, 22). Within the upper respiratory tract, α2-6-linked SAs (the preferred receptor for human influenza viruses) predominate on epithelial cells (26). While α2-3-linked SAs are also present to a lesser degree on respiratory epithelial cells, this linkage is more abundantly expressed on secreted mucins (2). In contrast, α2-3-linked SAs (the preferred receptor for avian influenza viruses) are found on corneal and conjunctival epithelial cells of the human eye (31, 41), while secreted ocular mucins are abundantly composed of α2-6 SAs (5). It has been suggested that avian influenza viruses are more suited to infect the ocular surface due to their general α2-3-linked SA binding preference, but this has not been demonstrated experimentally (30).The mouse model has been used previously to study the role of ocular exposure to respiratory viruses (6, 39). In mice, ocular inoculation with an H3N2 influenza virus resulted in virus replication in nasal turbinates and lung (39), whereas ocular infection with respiratory syncytial virus (RSV) resulted in detectable virus titers in the eye and lung (6). These studies have revealed that respiratory viruses are not limited to the ocular area following inoculation at this site. However, the ability of influenza viruses to replicate specifically within ocular tissue has not been examined.Despite repeated instances of conjunctivitis associated with H7 subtype infections in humans, the reasons for this apparent ocular tropism have not been studied extensively. Here, we present a murine model to study the ability of human and avian influenza viruses to cause disease by the ocular route. We found that highly pathogenic H7 and H5 influenza viruses were capable of causing a systemic and lethal infection in mice following ocular inoculation. These highly pathogenic viruses, unlike human H3N2 and H1N1 viruses, replicated to significant titers in the mouse corneal epithelium and primary human corneal epithelial cells (HCEpiCs). Identification of viruses well suited to infecting the ocular surface is the first step in better understanding the ability of influenza viruses of multiple subtypes to use this tissue as a portal of entry.  相似文献   

12.
This study investigated whether transmissible H5 subtype human-avian reassortant viruses could be generated in vivo. To this end, ferrets were coinfected with recent avian H5N1 (A/Thailand/16/04) and human H3N2 (A/Wyoming/3/03) viruses. Genotype analyses of plaque-purified viruses from nasal secretions of coinfected ferrets revealed that approximately 9% of recovered viruses contained genes from both progenitor viruses. H5 and H3 subtype viruses, including reassortants, were found in airways extending toward and in the upper respiratory tract of ferrets. However, only parental H5N1 genotype viruses were found in lung tissue. Approximately 34% of the recovered reassortant viruses possessed the H5 hemagglutinin (HA) gene, with five unique H5 subtypes recovered. These H5 reassortants were selected for further studies to examine their growth and transmissibility characteristics. Five H5 viruses with representative reassortant genotypes showed reduced titers in nasal secretions of infected ferrets compared to the parental H5N1 virus. No transmission by direct contact between infected and naïve ferrets was observed. These studies indicate that reassortment between H5N1 avian influenza and H3N2 human viruses occurred readily in vivo and furthermore that reassortment between these two viral subtypes is likely to occur in ferret upper airways. Given the relatively high incidence of reassortant viruses from tissues of the ferret upper airway, it is reasonable to conclude that continued exposure of humans and animals to H5N1 alongside seasonal influenza viruses increases the risk of generating H5 subtype reassortant viruses that may be shed from upper airway secretions.Highly pathogenic avian influenza (HPAI) viruses of the H5N1 subtype have caused devastating outbreaks in avian species during the past decade. After emerging in the Guangdong province of China in 1996, H5N1 viruses have extended their geographic distribution from Asia into Europe and Africa (45, 51). Sporadic transmission of H5N1 viruses from infected birds to humans has resulted in over 380 laboratory-confirmed infections and a case fatality rate of ∼60% since 2003 (48). Currently circulating H5N1 viruses lack the ability to undergo efficient and sustained transmission among humans although instances of limited human-to-human transmission have been reported (13, 41). If H5N1 viruses were to acquire genetic changes that confer efficient transmissibility among humans, then another pandemic would likely occur.The pandemics of 1957 and 1968 highlight the importance of genetic reassortment between avian and human influenza viruses as a mechanism for the generation of human pandemic strains (15, 46, 47). The structural separation of the influenza virus genome into eight independent genes allows formation of hybrid progeny viruses during coinfections. The 1957 H2N2 and 1968 H3N2 pandemic viruses acquired the hemagglutinin (HA) and PB1 genes, with or without the neuraminidase (NA) gene, respectively, from an avian virus progenitor (14, 33). The remaining genes of these pandemic reassortants were derived from a contemporary human virus (14, 33). The host species in which such human pandemic strains were generated by reassortment between human and avian viruses is not known. However, coinfection of the same cell with both human and avian viruses must have occurred, even though human and avian influenza viruses have preferences for different sialic acid receptor structures present on cell surface glycoproteins and glycolipids (20, 30). The HA of human viruses preferentially binds α(2,6)-linked sialic acids while that of avian viruses preferentially bind α(2,3)-linked sialic acids (3, 12). Cells possessing both of these receptors could support coinfection of avian and human viruses, leading to reassortment.Human respiratory tract epithelial cells can possess surface glycans with α(2,3)- and α(2,6)-linked sialic acids and as such represent a potential host for the generation of avian-human reassortant viruses (24, 35). The general distribution of surface α(2,3)- and α(2,6)-linked sialic acids varies among cells of the human upper and lower respiratory tracts, which are anatomically separated by the larynx. Recent studies have shown that α(2,3)-linked sialic acids are present in tissues of the human lower respiratory tract (i.e., lung alveolar cells) (24, 35) as well as tissues of the human upper respiratory tract (24). Consistent with these findings, HPAI H5N1 viruses have been shown to attach to and infect tissues belonging to the lower respiratory tract (i.e., trachea, bronchi, and lung) (5, 25, 35, 40, 42, 43) as well as tissues belonging to the upper respiratory tract (i.e., nasopharyngeal, adenoid, and tonsillar) (25). Glycans with α(2,6)-linked sialic acids are more widespread on epithelial cells of the upper airways than lung alveoli (24, 35). In accordance, human seasonal influenza viruses preferentially attach to and infect cells of the upper respiratory tract (6, 25, 35, 43). If cells with both types of receptors are present in the human respiratory tract, simultaneous infection of a person with both human and avian viruses could generate reassortant viruses.Although viruses derived by reassortment between avian H5N1 and human H3N2 progenitors have been generated in vitro (17), reassortment between these avian and human strains in a coinfected mammalian host has not been shown. Furthermore, our knowledge of the genetic and phenotypic repertoire of such reassortants generated in vivo and their potential for transmission to uninfected hosts is limited (2, 17). In the present study, we used the ferret model to better understand the generation of reassortant viruses in a host coinfected with contemporary avian (H5N1) and human (H3N2) viruses and the extent to which such reassortants replicate and transmit from animal to animal. The domestic ferret (Mustela putoris) serves as an ideal small-animal model for influenza because ferrets are susceptible to human and avian influenza viruses, including HPAI H5N1 viruses, and reflect the relative transmissibility of human and avian influenza viruses in humans (9, 17, 18, 31, 36, 39, 53). Our study revealed that coinfection of ferrets reproducibly generated reassortant viruses that could be recovered from tissues within and extending toward the upper respiratory tract. Although H5 reassortant viruses were recovered from the upper airways, they displayed no transmissibility to contact ferrets, suggesting that additional functional changes are required for these viral subtypes to become pandemic within human populations.  相似文献   

13.
14.
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   

15.
16.
Mature glycoprotein spikes are inserted in the Lassa virus envelope and consist of the distal subunit GP-1, the transmembrane-spanning subunit GP-2, and the signal peptide, which originate from the precursor glycoprotein pre-GP-C by proteolytic processing. In this study, we analyzed the oligomeric structure of the viral surface glycoprotein. Chemical cross-linking studies of mature glycoprotein spikes from purified virus revealed the formation of trimers. Interestingly, sucrose density gradient analysis of cellularly expressed glycoprotein showed that in contrast to trimeric mature glycoprotein complexes, the noncleaved glycoprotein forms monomers and oligomers spanning a wide size range, indicating that maturation cleavage of GP by the cellular subtilase SKI-1/S1P is critical for formation of the correct oligomeric state. To shed light on a potential relation between cholesterol and GP trimer stability, we performed cholesterol depletion experiments. Although depletion of cholesterol had no effect on trimerization of the glycoprotein spike complex, our studies revealed that the cholesterol content of the viral envelope is important for the infectivity of Lassa virus. Analyses of the distribution of viral proteins in cholesterol-rich detergent-resistant membrane areas showed that Lassa virus buds from membrane areas other than those responsible for impaired infectivity due to cholesterol depletion of lipid rafts. Thus, derivation of the viral envelope from cholesterol-rich membrane areas is not a prerequisite for the impact of cholesterol on virus infectivity.Lassa virus (LASV) is a member of the family Arenaviridae, of which Lymphocytic choriomeningitis virus (LCMV) is the prototype. Arenaviruses comprise more than 20 species, divided into the Old World and New World virus complexes (19). The Old World arenaviruses include the human pathogenic LASV strains, Lujo virus, which was first identified in late 2008 and is associated with an unprecedented high case fatality rate in humans, the nonhuman pathogenic Ippy, Mobala, and Mopeia viruses, and the recently described Kodoko virus (10, 30, 49). The New World virus complex contains, among others, the South American hemorrhagic fever-causing viruses Junín virus, Machupo virus, Guanarito virus, Sabiá virus, and the recently discovered Chapare virus (22).Arenaviruses contain a bisegmented single-stranded RNA genome encoding the polymerase L, matrix protein Z, nucleoprotein NP, and glycoprotein GP. The bipartite ribonucleoprotein of LASV is surrounded by a lipid envelope derived from the plasma membrane of the host cell. The matrix protein Z has been identified as a major budding factor, which lines the interior of the viral lipid membrane, in which GP spikes are inserted (61, 75). The glycoprotein is synthesized as precursor protein pre-GP-C and is cotranslationally cleaved by signal peptidase into GP-C and the signal peptide, which exhibits unusual length, stability, and topology (3, 27, 28, 33, 70, 87). Moreover, the arenaviral signal peptide functions as trans-acting maturation factor (2, 26, 33). After processing by signal peptidase, GP-C of both New World and Old World arenaviruses is cleaved by the cellular subtilase subtilisin kexin isozyme-1/site-1 protease (SKI-1/S1P) into the distal subunit GP-1 and the membrane-anchored subunit GP-2 within the secretory pathway (5, 52, 63). For LCMV, it has been shown that GP-1 subunits are linked to each other by disulfide bonds and are noncovalently connected to GP-2 subunits (14, 24, 31). GP-1 is responsible for binding to the host cell receptor, while GP-2 mediates fusion between the virus envelope and the endosomal membrane at low pH due to a bipartite fusion peptide near the amino terminus (24, 36, 44). Sequence analysis of the LCMV GP-2 ectodomain revealed two heptad repeats that most likely form amphipathic helices important for this process (34, 86).In general, viral class I fusion proteins have triplets of α-helical structures in common, which contain heptad repeats (47, 73). In contrast, class II fusion proteins are characterized by β-sheets that form dimers in the prefusion status and trimers in the postfusion status (43). The class III fusion proteins are trimers that, unlike class I fusion proteins, were not proteolytically processed N-terminally of the fusion peptide, resulting in a fusion-active membrane-anchored subunit (39, 62). Previous studies with LCMV described a tetrameric organization of the glycoprotein spikes (14), while more recent data using a bacterially expressed truncated ectodomain of the LCMV GP-2 subunit pointed toward a trimeric spike structure (31). Due to these conflicting data regarding the oligomerization status of LCMV GP, it remains unclear to which class of fusion proteins the arenaviral glycoproteins belong.The state of oligomerization and the correct conformation of viral glycoproteins are crucial for membrane fusion during virus entry. The early steps of infection have been shown for several viruses to be dependent on the cholesterol content of the participating membranes (i.e., either the virus envelope or the host cell membrane) (4, 9, 15, 20, 21, 23, 40, 42, 53, 56, 76, 78, 79). In fact, it has been shown previously that entry of both LASV and LCMV is susceptible to cholesterol depletion of the target host cell membrane using methyl-β-cyclodextrin (MβCD) treatment (64, 71). Moreover, cholesterol not only plays an important role in the early steps during entry in the viral life cycle but also is critical in the virus assembly and release process. Several viruses of various families, including influenza virus, human immunodeficiency virus type 1 (HIV-1), measles virus, and Ebola virus, use the ordered environment of lipid raft microdomains. Due to their high levels of glycosphingolipids and cholesterol, these domains are characterized by insolubility in nonionic detergents under cold conditions (60, 72). Recent observations have suggested that budding of the New World arenavirus Junin virus occurs from detergent-soluble membrane areas (1). Assembly and release from distinct membrane microdomains that are detergent soluble have also been described for vesicular stomatitis virus (VSV) (12, 38, 68). At present, however, it is not known whether LASV requires cholesterol in its viral envelope for successful virus entry or whether specific membrane microdomains are important for LASV assembly and release.In this study, we first investigated the oligomeric state of the premature and mature LASV glycoprotein complexes. Since it has been shown for several membrane proteins that the oligomerization and conformation are dependent on cholesterol (58, 59, 76, 78), we further analyzed the dependence of the cholesterol content of the virus envelope on glycoprotein oligomerization and virus infectivity. Finally, we characterized the lipid membrane areas from which LASV is released.  相似文献   

17.
18.
The highly pathogenic H5N1 avian influenza virus emerged from China in 1996 and has spread across Eurasia and Africa, with a continuous stream of new cases of human infection appearing since the first large-scale outbreak among migratory birds at Qinghai Lake. The role of wild birds, which are the natural reservoirs for the virus, in the epidemiology of the H5N1 virus has raised great public health concern, but their role in the spread of the virus within the natural ecosystem of free-ranging terrestrial wild mammals remains unclear. In this study, we investigated H5N1 virus infection in wild pikas in an attempt to trace the circulation of the virus. Seroepidemiological surveys confirmed a natural H5N1 virus infection of wild pikas in their native environment. The hemagglutination gene of the H5N1 virus isolated from pikas reveals two distinct evolutionary clades, a mixed/Vietnam H5N1 virus sublineage (MV-like pika virus) and a wild bird Qinghai (QH)-like H5N1 virus sublineage (QH-like pika virus). The amino acid residue (glutamic acid) at position 627 encoded by the PB2 gene of the MV-like pika virus was different from that of the QH-like pika virus; the residue of the MV-like pika virus was the same as that of the goose H5N1 virus (A/GS/Guangdong [GD]/1/96). Further, we discovered that in contrast to the MV-like pika virus, which is nonpathogenic to mice, the QH-like pika virus is highly pathogenic. To mimic the virus infection of pikas, we intranasally inoculated rabbits, a species closely related to pikas, with the H5N1 virus of pika origin. Our findings first demonstrate that wild pikas are mammalian hosts exposed to H5N1 subtype avian influenza viruses in the natural ecosystem and also imply a potential transmission of highly pathogenic avian influenza virus from wild mammals into domestic mammalian hosts and humans.Highly pathogenic avian influenza (HPAI) is an extremely infectious, systemic viral disease that causes a high rate of mortality in birds. HPAI H5N1 viruses are now endemic in avian populations in Southeast Asia and have repeatedly been transmitted to humans (9, 14, 27). Since 2003, the H5N1 subtype has been reported in 391 human cases of influenza and has caused 247 human deaths in 15 countries, leading to greater than 60% mortality among infected individuals (38). Although currently incapable of sustained human-to-human transmission, H5N1 viruses undoubtedly pose a serious threat to public health, as well as to the global economy. Hence, preparedness for such a threat is a global priority (36).Wild birds are considered to be natural reservoirs for influenza A viruses (6, 18, 21, 35, 37). Of the 144 type A influenza virus hemagglutinin-neuraminidase (HA-NA) combinations, 103 have been found in wild birds (5, 7, 17, 37). Since the first HPAI outbreak among migratory wild birds appeared at Qinghai Lake in western China in May 2005 (3, 16, 25, 34, 41), HPAI viruses of the H5N1 subtype have been isolated from poultry throughout Eurasia and Africa. The continued occurrence of human cases has created a situation that could facilitate a pandemic emergence. There is heightened concern that wild birds are a reservoir for influenza A viruses that switch hosts and stably adapt to mammals, including horses, swine, and humans (11, 19, 22, 37).Despite the recent expansion of avian influenza virus (AIV) surveillance and genomic data (5, 17, 20, 21, 33, 40), fundamental questions remain concerning the ecology and evolution of these viruses. Little is known about how terrestrial wild mammals within their natural ecological systems affect HPAI H5N1 epidemiology or about the virus''s effects on public health, though there are many reports of natural and experimental H5N1 virus infection in animals belonging to the taxonomic orders Carnivora (12, 13, 15, 28, 29) and Artiodactyla (15). Herein, we provide the results of our investigation into H5N1 virus infection in wild pikas (Ochotona curzoniae of the order Lagomorpha) within their natural ecological setting. We describe our attempt to trace the circulation of H5N1 viruses and to characterize pika H5N1 influenza virus (PK virus).  相似文献   

19.
The fitness of oseltamivir-resistant highly pathogenic H5N1 influenza viruses has important clinical implications. We generated recombinant human A/Vietnam/1203/04 (VN; clade 1) and A/Turkey/15/06 (TK; clade 2.2) influenza viruses containing the H274Y neuraminidase (NA) mutation, which confers resistance to NA inhibitors, and compared the fitness levels of the wild-type (WT) and resistant virus pairs in ferrets. The VN-H274Y and VN-WT viruses replicated to similar titers in the upper respiratory tract (URT) and caused comparable disease signs, and none of the animals survived. On days 1 to 3 postinoculation, disease signs caused by oseltamivir-resistant TK-H274Y virus were milder than those caused by TK-WT virus, and all animals survived. We then studied fitness by using a novel approach. We coinoculated ferrets with different ratios of oseltamivir-resistant and -sensitive H5N1 viruses and measured the proportion of clones in day-6 nasal washes that contained the H274Y NA mutation. Although the proportion of VN-H274Y clones increased consistently, that of TK-H274Y virus decreased. Mutations within NA catalytic (R292K) and framework (E119A/K, I222L, H274L, and N294S) sites or near the NA enzyme active site (V116I, I117T/V, Q136H, K150N, and A250T) emerged spontaneously (without drug pressure) in both pairs of viruses. The NA substitutions I254V and E276A could exert a compensatory effect on the fitness of VN-H274Y and TK-H274Y viruses. NA enzymatic function was reduced in both drug-resistant H5N1 viruses. These results show that the H274Y NA mutation affects the fitness of two H5N1 influenza viruses differently. Our novel method of assessing viral fitness accounts for both virus-host interactions and virus-virus interactions within the host.The neuraminidase (NA) inhibitors (orally administered oseltamivir and inhaled zanamivir) are currently an important class of antiviral drugs available for the treatment of seasonal and pandemic influenza. Although administration of NA inhibitors may significantly reduce influenza virus transmission, it risks the emergence of drug-resistant variants (16, 32). The impact of drug resistance would depend on the fitness (i.e., infectivity in vitro and virulence and transmissibility in vivo) of the resistant virus. If the resistance mutation only modestly reduces the virus'' biological fitness and does not impair its replication efficiency and transmissibility, the effectiveness of antiviral treatment can be significantly impaired. The unexpected natural emergence and spread of oseltamivir-resistant variants (carrying the H274Y NA amino acid substitution) among seasonal H1N1 influenza viruses of the A/Brisbane/59/07 lineage demonstrated that drug-resistant viruses can be highly fit and transmissible in humans (11, 22, 29), although the fitness of these variants is not completely understood. They are hypothesized to have lower NA receptor affinity and more-optimal NA and hemagglutinin (HA) functional balance than do wild-type (WT) viruses (38). Fortunately, oseltamivir-resistant variants have rarely been reported to occur among the novel pandemic H1N1 influenza viruses that emerged in April 2009; therefore, initial data suggest that currently circulating wild-type viruses possibly possess greater fitness than drug-resistant viruses (45), although only retrospective epidemiological data can provide a conclusive answer. The key questions are whether the risk posed by NA inhibitor-resistant viruses can be assessed experimentally and what the most reliable approach may be.All NA inhibitor-resistant influenza viruses characterized to date have contained specific mutations in the NA molecule. Clinically derived drug-resistant viruses have carried mutations that are NA subtype specific and differ in accordance with the NA inhibitor used (12, 35). The most commonly observed mutations are H274Y and N294S in the influenza A N1 NA subtype, E119A/G/D/V and R292K in the N2 NA subtype, and R152K and D198N in influenza B viruses (35, 36). The fitness of NA inhibitor-resistant viruses has been studied in vitro and in vivo. Many groups have assessed their replicative capacity in MDCK cells, but this assay system can yield anomalous results (49), particularly in the case of low-passage clinical isolates. The mismatch between virus specificity and cellular receptors can be overcome by using cell lines engineered to express human-like α-2,6-linked sialyl cell surface receptors (MDCK-SIAT1) (15, 34) or a novel cell culture-based system that morphologically and functionally recapitulates differentiated normal human bronchial epithelial (NHBE) cells (24). Investigations in vivo typically compare replication efficiencies, clinical signs, and transmissibility levels between oseltamivir-resistant viruses and the corresponding wild-type virus. Initial studies found that NA inhibitor-resistant influenza viruses were severely compromised in vitro and in animal models (6, 17, 26) and thus led to the idea that resistant viruses will unlikely have an impact on epidemic and pandemic influenza. However, clinically derived H1N1 virus with the H274Y NA mutation (18) and reverse genetics-derived H3N2 virus with the E119V NA mutation (46) were subsequently found to possess biological fitness and transmissibility similar to those of drug-sensitive virus in direct-contact ferrets. Recent studies in a guinea pig model showed that recombinant human H3N2 influenza viruses carrying either a single E119V NA mutation or the double NA mutation E119V-I222V were transmitted efficiently by direct contact but not by aerosol (5).There is limited information about the fitness of NA inhibitor-resistant H5N1 influenza viruses. Although they are not efficiently transmitted from human to human, their pandemic potential remains a serious public health concern because of their virulence in humans (1, 4, 7). H5N1 viruses isolated from untreated patients are susceptible to the NA inhibitors oseltamivir and zanamivir (21), although oseltamivir-resistant variants with the H274Y NA mutation have been reported to occur in five patients after (9, 30) or before (41) treatment with oseltamivir. The World Health Organization reported the isolation of two oseltamivir-resistant H5N1 viruses from an Egyptian girl and her uncle (44) after oseltamivir treatment. The virus was moderately resistant and possessed an N294S NA mutation. Preliminary evidence suggests that the resistance mutation existed before transmission of the virus from birds to the patients and thus before initiation of treatment (41). We previously showed that wild-type A/Vietnam/1203/04 (H5N1) influenza virus and recombinants carrying either the H274Y or the N294S NA mutation reached comparable titers in MDCK and MDCK-SIAT1 cells and caused comparable mortality rates among BALB/c mice (48). In contrast, clinically derived A/Hanoi/30408/05 (H5N1) influenza virus with the H274Y NA mutation reproduced to lower titers than the oseltamivir-sensitive virus in the lungs of inoculated ferrets (30).In a ferret model, we compared the fitness levels of two pairs of H5N1 viruses in the absence of selective drug pressure. One virus of each pair was the wild type, while the other carried the H274Y NA mutation conferring oseltamivir resistance. The two viruses used, A/Vietnam/1203/04 (HA clade 1) and A/Turkey/15/06 (HA clade 2.2), differ in their pathogenicity to ferrets. Virus fitness was evaluated by two approaches. Using the traditional approach, we compared clinical disease signs, relative inactivity indexes, weight and temperature changes, and virus replication levels in the upper respiratory tract (URT). We then used a novel competitive fitness approach in which we genetically analyzed individual virus clones after coinfection of ferrets with mixtures of oseltamivir-sensitive and -resistant H5N1 viruses; thus, we determined virus-virus interactions within the host. We observed no difference between the resistant and sensitive virus of each pair in clinical signs or virus replication in the URT; however, analysis of virus-virus interactions within the host showed that the H274Y NA mutation affected the fitness of the two viruses differently. The oseltamivir-resistant A/Vietnam/1203/04-like virus outgrew its wild-type counterpart, while the oseltamivir-resistant A/Turkey/15/06-like virus showed less fitness than its wild-type counterpart.  相似文献   

20.
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   

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