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 15 毫秒
1.
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   

2.
The neuraminidase inhibitors zanamivir and oseltamivir are marketed for the treatment and prophylaxis of influenza and have been stockpiled by many countries for use in a pandemic. Although recent surveillance has identified a striking increase in the frequency of oseltamivir-resistant seasonal influenza A (H1N1) viruses in Europe, the United States, Oceania, and South Africa, to date there have been no reports of significant zanamivir resistance among influenza A (H1N1) viruses or any other human influenza viruses. We investigated the frequency of oseltamivir and zanamivir resistance in circulating seasonal influenza A (H1N1) viruses in Australasia and Southeast Asia. Analysis of 391 influenza A (H1N1) viruses isolated between 2006 and early 2008 from Australasia and Southeast Asia revealed nine viruses (2.3%) that demonstrated markedly reduced zanamivir susceptibility and contained a previously undescribed Gln136Lys (Q136K) neuraminidase mutation. The mutation had no effect on oseltamivir susceptibility but caused approximately a 300-fold and a 70-fold reduction in zanamivir and peramivir susceptibility, respectively. The role of the Q136K mutation in conferring zanamivir resistance was confirmed using reverse genetics. Interestingly, the mutation was not detected in the primary clinical specimens from which these mutant isolates were grown, suggesting that the resistant viruses either occurred in very low proportions in the primary clinical specimens or arose during MDCK cell culture passage. Compared to susceptible influenza A (H1N1) viruses, the Q136K mutant strains displayed greater viral fitness than the wild-type virus in MDCK cells but equivalent infectivity and transmissibility in a ferret model.Two classes of antiviral drugs are currently available for the treatment and prophylaxis of influenza, the adamantanes and the neuraminidase (NA) inhibitors (NAIs). The adamantanes were the first agents to be recognized to have anti-influenza virus activities as early as 1964 (2) although the rapid emergence of drug-resistant influenza virus strains has limited their clinical effectiveness (12). The NAIs, zanamivir (Relenza) and oseltamivir (Tamiflu), were the first drugs to be specifically designed as anti-influenza virus agents and have been available on the market in many countries since 1999. During oseltamivir clinical trials, 1 to 4% of treated adults (6) and 5 to 6% of treated children were found to shed resistant viruses (30) although more recent studies have reported resistance in 16 to 18% of viruses from oseltamivir-treated children (20, 29). In contrast to the frequency of resistance seen following oseltamivir treatment, only one occurrence of significant zanamivir resistance has been observed following zanamivir treatment. The zanamivir-resistant strain, an influenza B virus with an R152K NA mutation, was isolated from an immunocompromised patient undergoing prolonged zanamivir treatment (7).In addition to the analysis of influenza viruses isolated from patients undergoing either oseltamivir or zanamivir treatment, surveillance studies that analyze the NAI susceptibility of circulating viruses, predominantly from patients not undergoing NAI treatment, have also been conducted. Studies that have tested viruses isolated prior to the release of the NAIs (1996 to 1999) (23) and after the initiation of clinical use of these drugs (2000 to 2006) (16, 24) have found either no resistance or a very low frequency of resistance. In contrast, analysis of circulating seasonal influenza viruses from Europe during the 2007 to 2008 season revealed that 14% (59/437) of influenza A (H1N1) viruses had significantly decreased susceptibility to oseltamivir (21). Since this initial report, oseltamivir-resistant influenza A (H1N1) strains have spread throughout Europe (11) and have been detected at high frequencies in other countries including the United States (4), Japan (28), South Africa (1) and Oceania and Southeast Asia (17). These influenza A (H1N1) viruses have a mutation of histidine to tyrosine at residue 274 of the NA (N2 NA numbering; residue 275 by N1 NA numbering), which confers a high level of resistance to oseltamivir (10) but has no effect on susceptibility to zanamivir or to the adamantanes.Prior to May 2008, when the oseltamivir-resistant variants became the dominant influenza A (H1N1) strain in Oceania and Southeast Asia (17), NAI sensitivity monitoring conducted at the WHO Collaborating Centre for Reference and Research on Influenza, Melbourne, identified a number of influenza A (H1N1) viruses with reduced zanamivir susceptibility. These viruses contained a previously undescribed mutation at residue 136 of the NA. Here, we report on the detection of these mutant viruses from geographically distinct locations, the in vitro and in vivo fitness of the strains, and the finding that the mutant viruses appear to have been preferentially propagated during viral culture in Madin-Darby canine kidney (MDCK) cells.  相似文献   

3.
4.
In less than 3 months after the first cases of swine origin 2009 influenza A (H1N1) virus infections were reported from Mexico, WHO declared a pandemic. The pandemic virus is antigenically distinct from seasonal influenza viruses, and the majority of human population lacks immunity against this virus. We have studied the activation of innate immune responses in pandemic virus-infected human monocyte-derived dendritic cells (DC) and macrophages. Pandemic A/Finland/553/2009 virus, representing a typical North American/European lineage virus, replicated very well in these cells. The pandemic virus, as well as the seasonal A/Brisbane/59/07 (H1N1) and A/New Caledonia/20/99 (H1N1) viruses, induced type I (alpha/beta interferon [IFN-α/β]) and type III (IFN-λ1 to -λ3) IFN, CXCL10, and tumor necrosis factor alpha (TNF-α) gene expression weakly in DCs. Mouse-adapted A/WSN/33 (H1N1) and human A/Udorn/72 (H3N2) viruses, instead, induced efficiently the expression of antiviral and proinflammatory genes. Both IFN-α and IFN-β inhibited the replication of the pandemic (H1N1) virus. The potential of IFN-λ3 to inhibit viral replication was lower than that of type I IFNs. However, the pandemic virus was more sensitive to the antiviral IFN-λ3 than the seasonal A/Brisbane/59/07 (H1N1) virus. The present study demonstrates that the novel pandemic (H1N1) influenza A virus can readily replicate in human primary DCs and macrophages and efficiently avoid the activation of innate antiviral responses. It is, however, highly sensitive to the antiviral actions of IFNs, which may provide us an additional means to treat severe cases of infection especially if significant drug resistance emerges.The novel swine origin 2009 influenza A (H1N1) virus was identified in April 2009, and it is currently causing the first influenza pandemic of the 21st century. The virus is a completely new reassortant virus (8, 38), and the majority of the human population does not have preexisting immunity against it. The case fatality rate of the current pandemic virus infection is still unclear, but it is estimated to be somewhat higher than that of seasonal influenza virus infections (8). In most cases, the pandemic 2009 A (H1N1) virus causes an uncomplicated respiratory tract illness with symptoms similar to those caused by seasonal influenza viruses. However, gastrointestinal symptoms atypical to seasonal influenza have been detected in a significant proportion of cases (4, 7, 35).The pandemic 2009 (H1N1) influenza A virus originates from a swine influenza A virus strain. It underwent multiple reassortment events in pigs and then transferred into the human population (8, 38). The new virus has gene segments from the North American triple-reassortant and Eurasian swine H1N1 viruses (8, 38). Sequence analysis of this new pandemic virus revealed that hemagglutinin (HA), NP, and NS gene segments are derived from the classical swine viruses, PB1 from human H3N2, and PB2 and PA from avian viruses within the triple-reassortant virus (8). In addition, the NA and M segments originate from the Eurasian swine virus lineage. The pandemic 2009 (H1N1) virus is genetically and antigenically distinct from previous seasonal human influenza A (H1N1) viruses. Thus, the current seasonal influenza vaccines are likely to give little, if any, protection against pandemic 2009 A (H1N1) virus infection (12, 14). However, some evidence indicates that people born early in the 20th century have cross-neutralizing antibodies against the pandemic 2009 A (H1N1) viruses (12, 14).At present, relatively little is known about the pathogenesis and transmission of the pandemic 2009 A (H1N1) virus in humans. Studies with ferrets revealed that the pandemic virus replicated better than seasonal H1N1 viruses in the respiratory tracts of the animals. This suggests that the virus is more pathogenic in ferrets than seasonal influenza viruses (19, 24). The respiratory tract is the primary infection site of all mammalian influenza viruses, and, indeed, the specific glycan receptors on the apical surface of the upper respiratory tract have been reported to bind HA of the 2009 A (H1N1) virus (19). In human lung tissue binding assays, 2009 A (H1N1) HA showed a glycan binding pattern similar to that of the HA from the pandemic 1918 A (H1N1) virus though its affinity to α2,6 glycans was much lower than that of the 1918 virus HA. The lower glycan binding properties of the pandemic 2009 A (H1N1) virus seemed to correlate with less-efficient transmission in ferrets compared to seasonal H1N1 viruses (19). According to another study with ferrets, the transmission of the pandemic 2009 A (H1N1) virus via respiratory droplets was as efficient as that of a seasonal A (H1N1) virus (24). It is clear that, besides experimental infections in animal models, analyses of the characters and pathogenesis of the pandemic 2009 A (H1N1) virus infection in humans are urgently needed.In the present study, we have focused on analyzing innate immune responses in primary human dendritic cells (DCs) and macrophages in response to an infection with one of the Finnish isolates of the pandemic 2009 A (H1N1) virus. DCs and macrophages reside beneath the epithelium of the respiratory organs, and these cells are thus potential targets for influenza viruses. From the epithelial cells influenza viruses spread in DCs and macrophages, which coordinate the development of an effective innate immune response against the virus (22, 34, 41). During influenza virus infection, DCs and macrophages secrete antiviral cytokines such as interferons (IFNs) and tumor necrosis factor alpha (TNF-α) (3, 13, 26). Moreover, DCs and macrophages activate virus-destroying NK cells and T cells with the cytokines they secrete and via direct cell-to-cell contacts (9, 29, 33, 37). Here we show that the pandemic (H1N1) virus infects and replicates very well in human monocyte-derived DCs and macrophages. The pandemic virus as well as two recent seasonal H1N1 viruses induced a relatively weak innate immune response in these cells, as evidenced by a poor expression of antiviral and proinflammatory cytokine genes. However, like seasonal influenza A viruses, the pandemic 2009 (H1N1) virus was extremely sensitive to the antiviral actions of type I IFNs (IFN-α/β). Interestingly, the pandemic 2009 (H1N1) virus was even more sensitive to antiviral IFN-λ3 than a seasonal A (H1N1) virus. Thus, IFNs may provide us with an additional means to combat severe pandemic influenza virus infections, especially if viral resistance against neuraminidase (NA) inhibitors begins to emerge.  相似文献   

5.
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   

6.
Novel swine-origin influenza viruses of the H1N1 subtype were first detected in humans in April 2009. As of 12 August 2009, 180,000 cases had been reported globally. Despite the fact that they are of the same antigenic subtype as seasonal influenza viruses circulating in humans since 1977, these viruses continue to spread and have caused the first influenza pandemic since 1968. Here we show that a pandemic H1N1 strain replicates in and transmits among guinea pigs with similar efficiency to that of a seasonal H3N2 influenza virus. This transmission was, however, partially disrupted when guinea pigs had preexisting immunity to recent human isolates of either the H1N1 or H3N2 subtype and was fully blocked through daily intranasal administration of interferon to either inoculated or exposed animals. Our results suggest that partial immunity resulting from prior exposure to conventional human strains may blunt the impact of pandemic H1N1 viruses in the human population. In addition, the use of interferon as an antiviral prophylaxis may be an effective way to limit spread in at-risk populations.A pandemic of novel swine-origin influenza virus (H1N1) is developing rapidly. As of 12 August 2009, nearly 180,000 cases had been reported to the WHO from around the globe (36). Sustained human-to-human transmission has furthermore been observed in multiple countries, prompting the WHO to declare a public health emergency of international concern and to raise the pandemic alert level to phase 6 (7).Swine are a natural host of influenza viruses, and although sporadic incidences of human infection with swine influenza viruses occur (8, 9, 14, 29, 35), human-to-human transmission is rare. H1N1 influenza viruses have likely circulated in swine since shortly after the 1918 human influenza pandemic (38). From the 1930s, when a swine influenza virus was first isolated, to the late 1990s, this classical swine lineage has remained relatively stable antigenically (34). In the late 1990s, however, genetic reassortment between a human H3N2 virus, a North American avian virus, and a classical swine influenza virus produced a triple reassortant virus, which subsequently spread among North American swine (34). Further reassortment events involving human influenza viruses led to the emergence in pigs of triple reassortants of the H1N1 and H1N2 subtypes (34). None of these swine viruses have demonstrated the potential for sustained human-to-human transmission.The swine-origin influenza viruses now emerging in the human population possess a previously uncharacterized constellation of eight genes (28). The NA and M segments derive from a Eurasian swine influenza virus lineage, having entered pigs from the avian reservoir around 1979, while the HA, NP, and NS segments are of the classical swine lineage and the PA, PB1, and PB2 segments derive from the North American triple reassortant swine lineage (13). This unique combination of genetic elements (segments from multiple swine influenza virus lineages, some of them derived from avian and human influenza viruses) may account for the improved fitness of pandemic H1N1 viruses, relative to that of previous swine isolates, in humans.Several uncertainties remain about how this outbreak will develop over time. Although the novel H1N1 virus has spread over a broad geographical area, the number of people known to be infected remains low in many countries, which could be due, at least in part, to the lack of optimal transmission of influenza viruses outside the winter season; thus, it is unclear at this point whether the new virus will become established in the long term. Two major factors will shape the epidemiology of pandemic H1N1 viruses in the coming months and years: the intrinsic transmissibility of the virus and the degree of protection offered by previous exposure to seasonal human strains. Initial estimates of the reproductive number (R0) have been made based on the epidemiology of the virus to date and suggest that its rate of spread is intermediate between that of seasonal flu and that of previous pandemic strains (3, 11). However, more precise estimates of R0 will depend on better surveillance data in the future. The transmission phenotype of pandemic H1N1 viruses in a ferret model was also recently reported and was found to be similar to (16, 27) or less efficient (25) than that of seasonal H1N1 strains. The reason for this discrepancy in the ferret model is unclear.Importantly, in considering the human population, the impact of immunity against seasonal strains on the transmission potential of pandemic H1N1 viruses is not clear. According to conventional wisdom, an influenza virus must be of a hemagglutinin (HA) subtype which is novel to the human population in order to cause a pandemic (18, 38). Analysis of human sera collected from individuals with diverse influenza virus exposure histories has indicated that in those born in the early part of the 20th century, neutralizing activity against A/California/04/09 (Cal/04/09) virus is often present (16). Conversely, serological analyses of ferret postinfection sera (13) and human pre- and postvaccination sera (4a) revealed that neutralizing antibodies against recently circulating human H1N1 viruses do not react with pandemic H1N1 isolates. These serological findings may explain the relatively small number of cases seen to date in individuals greater than 65 years of age (6). Even in the absence of neutralizing antibodies, however, a measure of immune protection sufficient to dampen transmission may be present in a host who has recently experienced seasonal influenza (10). If, on the other hand, transmission is high and immunity is low, then pandemic H1N1 strains will likely continue to spread rapidly through the population. In this situation, a range of pharmaceutical interventions will be needed to dampen the public health impact of the pandemic.Herein we used the guinea pig model (4, 21-24, 26, 30) to assess the transmissibility of the pandemic H1N1 strains Cal/04/09 and A/Netherlands/602/09 (NL/602/09) relative to that of previous human and swine influenza viruses. To better mimic the human situation, we then tested whether the efficiency of transmission is decreased by preexisting immunity to recent human H1N1 or H3N2 influenza viruses. Finally, we assessed the efficacy of intranasal treatment with type I interferon (IFN) in limiting the replication and transmission of pandemic H1N1 viruses.  相似文献   

7.
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   

8.
Highly pathogenic avian influenza A viruses of the H5N1 subtype continue to circulate in poultry, and zoonotic transmissions are reported frequently. Since a pandemic caused by these highly pathogenic viruses is still feared, there is interest in the development of influenza A/H5N1 virus vaccines that can protect humans against infection, preferably after a single vaccination with a low dose of antigen. Here we describe the induction of humoral and cellular immune responses in ferrets after vaccination with a cell culture-derived whole inactivated influenza A virus vaccine in combination with the novel adjuvant CoVaccine HT. The addition of CoVaccine HT to the influenza A virus vaccine increased antibody responses to homologous and heterologous influenza A/H5N1 viruses and increased virus-specific cell-mediated immune responses. Ferrets vaccinated once with a whole-virus equivalent of 3.8 μg hemagglutinin (HA) and CoVaccine HT were protected against homologous challenge infection with influenza virus A/VN/1194/04. Furthermore, ferrets vaccinated once with the same vaccine/adjuvant combination were partially protected against infection with a heterologous virus derived from clade 2.1 of H5N1 influenza viruses. Thus, the use of the novel adjuvant CoVaccine HT with cell culture-derived inactivated influenza A/H5N1 virus antigen is a promising and dose-sparing vaccine approach warranting further clinical evaluation.Since the first human case of infection with a highly pathogenic avian influenza A virus of the H5N1 subtype in 1997 (9, 10, 37), hundreds of zoonotic transmissions have been reported, with a high case-fatality rate (10, 44). Since these viruses continue to circulate among domestic birds and human cases are regularly reported, it is feared that they will adapt to their new host or exchange gene segments with other influenza A viruses, become transmissible from human to human, and cause a new pandemic. Recently, a novel influenza A virus of the H1N1 subtype emerged. This virus, which originated from pigs, was transmitted between humans efficiently, resulting in the first influenza pandemic of the 21st century (8, 45). Although millions of people have been inoculated with the (H1N1)2009 virus, the case-fatality rate was relatively low compared to that for infections with the H5N1 viruses (11, 31). However, the unexpected pandemic caused by influenza A/H1N1(2009) viruses has further highlighted the importance of rapid availability of safe and effective pandemic influenza virus vaccines. Other key issues for the development of pandemic influenza A virus vaccines include optimal use of the existing (limited) capacity for production of viral antigen and effectiveness against viruses that are antigenically distinct. Ideally, a single administration of a low dose of antigen would be sufficient to induce protective immunity against the homologous strain and heterologous antigenic variant strains. However, since the population at large will be immunologically naïve to a newly introduced virus, high doses of antigen are required to induce protective immunity in unprimed subjects (23, 36). The use of safe and effective adjuvants in pandemic influenza virus vaccines is considered a dose-sparing strategy. Clinical trials evaluating candidate inactivated influenza A/H5N1 virus vaccines showed that the use of adjuvants can increase their immunogenicity and broaden the specificity of the induced antibody responses (2, 7, 19, 23, 27, 36, 41). These research efforts have resulted in the licensing of adjuvanted vaccines against seasonal and pandemic influenza viruses (17). The protective efficacy of immune responses induced with candidate influenza A/H5N1 virus vaccines was demonstrated in ferrets after two immunizations (1, 22, 24, 25) or after a single immunization. The latter was achieved with a low dose of antigen in combination with the adjuvant Iscomatrix (26).Recently, a novel adjuvant that consists of a sucrose fatty acid sulfate ester (SFASE) immobilized on the oil droplets of a submicrometer emulsion of squalane in water has been developed (4). It has been demonstrated that the addition of this novel adjuvant, called CoVaccine HT, to multiple antigens increased the immune response to these antigens in pigs and horses and was well tolerated in both species (4, 16, 40). Furthermore, it was shown that the use of CoVaccine HT increased the virus-specific antibody responses in mice and ferrets after vaccination with a cell culture-derived whole inactivated influenza A/H5N1 virus vaccine (5, 13). One of the mode of actions of CoVaccine HT is the activation of antigen-presenting cells such as dendritic cells, most likely through Toll-like receptor 4 (TLR4) signaling (5).In the present study, we evaluated the protective potential of CoVaccine HT-adjuvanted cell culture-derived whole inactivated influenza A/H5N1 virus (WIV) vaccine in the ferret model, which is considered the most suitable animal model for the evaluation of candidate influenza virus vaccines (6, 14, 15). To this end, ferrets were vaccinated once or twice with various antigen doses with or without the adjuvant to test whether dose sparing could be achieved. The use of CoVaccine HT increased virus-specific antibody responses and T cell responses. A single administration of 3.8 μg hemagglutinin (HA) of WIV NIBRG-14 vaccine preparation in combination with CoVaccine HT conferred protection against challenge infection with the homologous highly pathogenic A/H5N1 virus strain A/VN/1194/04 and partial protection against infection with a heterologous, antigenically distinct strain, A/IND/5/05. Therefore, it was concluded that the use of CoVaccine HT in inactivated influenza virus vaccines induced protective virus-specific humoral and cell-mediated immune responses and that it could be suitable as adjuvant in (pre)pandemic A/H5N1 virus vaccines. Further clinical testing of these candidate vaccines seems to be warranted.  相似文献   

9.
To determine the relative fitness of oseltamivir-resistant strains compared to susceptible wild-type viruses, we combined mathematical modeling and statistical techniques with a novel in vivo “competitive-mixtures” experimental model. Ferrets were coinfected with either pure populations (100% susceptible wild-type or 100% oseltamivir-resistant mutant virus) or mixed populations of wild-type and oseltamivir-resistant influenza viruses (80%:20%, 50%:50%, and 20%:80%) at equivalent infectivity titers, and the changes in the relative proportions of those two viruses were monitored over the course of the infection during within-host and over host-to-host transmission events in a ferret contact model. Coinfection of ferrets with mixtures of an oseltamivir-resistant R292K mutant A(H3N2) virus and a R292 oseltamivir-susceptible wild-type virus demonstrated that the R292K mutant virus was rapidly outgrown by the R292 wild-type virus in artificially infected donor ferrets and did not transmit to any of the recipient ferrets. The competitive-mixtures model was also used to investigate the fitness of the seasonal A(H1N1) oseltamivir-resistant H274Y mutant and showed that within infected ferrets the H274Y mutant virus was marginally outgrown by the wild-type strain but demonstrated equivalent transmissibility between ferrets. This novel in vivo experimental method and accompanying mathematical analysis provide greater insight into the relative fitness, both within the host and between hosts, of two different influenza virus strains compared to more traditional methods that infect ferrets with only pure populations of viruses. Our statistical inferences are essential for the development of the next generation of mathematical models of the emergence and spread of oseltamivir-resistant influenza in human populations.The neuraminidase (NA) inhibitors are a class of influenza antiviral drugs that are specifically designed to inhibit the enzymatic function of the NA, thereby preventing normal viral replication. Since 1999, two NA inhibitors (NAIs), oseltamivir (Tamiflu) and zanamivir (Relenza), have been shown to be effective for the treatment and prophylaxis of patients infected with not only seasonal influenza, but also highly pathogenic A(H5N1) and the newly emerged A(H1N1) pandemic virus. Prior to 2007, resistance to this class of drugs was considered relatively uncommon, particularly in comparison with the other class of influenza antivirals, the adamantanes, which readily select for viral resistance in treated patients. During early clinical trials, oseltamivir resistance was detected in only 1 to 2% of adults (14) and 5 to 6% of children (33) under treatment, although later studies detected resistance in up to 18% of oseltamivir-treated children (16). In contrast, resistance following zanamivir treatment is rare, with only one reported case observed in an immunocompromised patient (6). Influenza viruses that develop resistance to these drugs typically contain mutations in the NA which, either directly or indirectly, alter the shape of the NA enzymatic site, thereby reducing the ability of the drugs to bind to this specific pocket. One of the most commonly observed mutations in oseltamivir-resistant A(H3N2) viruses is an arginine-to-lysine mutation at residue 292 (R292K) of the NA, while the predominant NA mutation in oseltamivir-resistant A(H1N1) viruses is a histidine-to-tyrosine mutation at residue 274 (H274Y) (N2 NA amino acid numbering, equivalent to residue 275 based on N1 numbering). Both of these mutations have an indirect impact on drug binding, as they affect the ability of the glutamic acid residue at position 276 to reorientate, as required for slow binding by oseltamivir (3). Many mutations that cause NAI resistance also cause reduced NA enzyme activity and, consequently, can compromise viral fitness.Previous studies have demonstrated that viruses with an R292K NA mutation demonstrated compromised growth in vitro (36) and in ferrets were significantly less infectious and did not transmit (9). The replication and transmission fitness of the H274Y mutation has also been studied previously. An H274Y mutant A(H1N1) strain isolated from a patient under oseltamivir treatment demonstrated compromised growth in cell culture compared to a wild-type (WT) virus (13), although a strain carrying the same mutation selected in vitro was found to replicate as well as the wild type (32). The infectivity and transmissibility of an H274Y mutant were found to be restricted in ferrets (13), although a second study demonstrated that transmission of the mutant virus between ferrets was possible, but required a greater viral dose of the mutant compared to the wild type (10). These results suggest that resistant virus variants with the same NA mutation may differ in replication or transmission fitness depending on other viral components. Nevertheless, based on these data and the viral fitness of other resistant mutants, it was believed that NAI-resistant viruses were unlikely to spread throughout the community due to their compromised viral fitness in the absence of drug selective pressure. This was proven incorrect during the Northern Hemisphere 2007-2008 influenza season, when large numbers of oseltamivir-resistant seasonal A(H1N1) viruses with an H274Y mutation were detected in patients who had not been treated with oseltamivir (4, 24). The mutant strain continued to spread to the Southern Hemisphere, such that by late 2008 virtually all circulating seasonal A(H1N1) viruses were oseltamivir resistant (11). The rapid global spread of this strain clearly suggested that the oseltamivir-resistant seasonal A(H1N1) virus had fitness equivalent to or greater than that of the previous oseltamivir-sensitive A(H1N1) strain. The reasons for enhanced viral fitness in this strain, when previous studies demonstrated that the acquisition of an H274Y mutation led to reduced viral fitness, remain unclear but probably involve compensatory mutations or reassortment events which may have improved the hemagglutinin (HA)/NA balance, allowing efficient transmission (5, 26).Experimental methods have been developed to assess the relative fitness of NAI-resistant strains compared with respective wild-type viruses, both in vitro and in vivo. Ferrets have been considered the most appropriate model animal for influenza research, and fitness studies have assessed variables such as minimum dose required to achieve infection, duration of viral shedding, and levels of viral load to allow comparisons between viruses. The guinea pig model has also been previously used to assess the viral fitness of influenza viruses, particularly in comparing the transmissibility of strains via either the contact or aerosol route (2). As an alternative to these traditional approaches, we have investigated a methodology that involves coinfection of ferrets with a mixture of two influenza viruses. Daily monitoring of changes in the relative proportion of those viruses over the course of the infection allows determination of the relative replication fitness of the viruses. Monitoring of recipient ferrets exposed to the infected ferrets enables the relative transmissibility of the viruses (henceforth, the relative transmission fitness) to be determined. In this study, the “competitive-mixtures” methodology was used to assess the relative replication and transmission fitness of an oseltamivir-resistant R292K mutant A(H3N2) virus compared with an oseltamivir-sensitive A(H3N2) wild-type strain and also to asses the relative replication and transmission fitness of an oseltamivir-resistant H274Y seasonal A(H1N1) mutant compared with an oseltamivir-sensitive A(H1N1) wild-type strain. Quantitative estimates for the replication fitness of mutant viruses were determined using a simple mathematical model of within-host viral replication and mixed-effects statistical tests. Transmission fitness was evaluated by application of a graphical technique that demonstrated the relationship between the proportion of mutant virus in the infectee ferrets as a function of the proportion of mutant virus in the infector ferrets.Inferences drawn from the statistical analyses presented here are essential for the refinement of existing mathematical models that simulate the spread of influenza in the human population and model the deployment of antiviral agents. These models are designed to assess the likely impact of different antiviral agent deployment strategies to control pandemic influenza (18, 21, 35). At present, data on the probability of emergence of NAI-resistant strains, the relative transmission fitness of these strains, and the probability of an individual''s infection reverting to an NAI-sensitive strain in the absence of ongoing selective pressure are severely limited. In consequence, human population-level models of influenza spread must make gross assumptions on the likely characteristics of NAI-resistant strains. Data such as those presented here will be used to inform new models of drug deployment and result in improved pandemic policy advice (20, 23).  相似文献   

10.
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   

11.
Many novel reassortant influenza viruses of the H9N2 genotype have emerged in aquatic birds in southern China since their initial isolation in this region in 1994. However, the genesis and evolution of H9N2 viruses in poultry in eastern China have not been investigated systematically. In the current study, H9N2 influenza viruses isolated from poultry in eastern China during the past 10 years were characterized genetically and antigenically. Phylogenetic analysis revealed that these H9N2 viruses have undergone extensive reassortment to generate multiple novel genotypes, including four genotypes (J, F, K, and L) that have never been recognized before. The major H9N2 influenza viruses represented by A/Chicken/Beijing/1/1994 (Ck/BJ/1/94)-like viruses circulating in poultry in eastern China before 1998 have been gradually replaced by A/Chicken/Shanghai/F/1998 (Ck/SH/F/98)-like viruses, which have a genotype different from that of viruses isolated in southern China. The similarity of the internal genes of these H9N2 viruses to those of the H5N1 influenza viruses isolated from 2001 onwards suggests that the Ck/SH/F/98-like virus may have been the donor of internal genes of human and poultry H5N1 influenza viruses circulating in Eurasia. Experimental studies showed that some of these H9N2 viruses could be efficiently transmitted by the respiratory tract in chicken flocks. Our study provides new insight into the genesis and evolution of H9N2 influenza viruses and supports the notion that some of these viruses may have been the donors of internal genes found in H5N1 viruses.Wild birds, including wild waterfowls, gulls, and shorebirds, are the natural reservoirs for influenza A viruses, in which they are thought to be in evolutionary stasis (2, 33). However, when avian influenza viruses are transmitted to new hosts such as terrestrial poultry or mammals, they evolve rapidly and may cause occasional severe systemic infection with high morbidity (20, 29). Despite the fact that avian influenza virus infection occurs commonly in chickens, it is unable to persist for a long period of time due to control efforts and/or a failure of the virus to adapt to new hosts (29). In the past 20 years, greater numbers of outbreaks in poultry have occurred, suggesting that the avian influenza virus can infect and spread in aberrant hosts for an extended period of time (5, 14-16, 18, 32).During the past 10 years, H9N2 influenza viruses have become panzootic in Eurasia and have been isolated from outbreaks in poultry worldwide (3, 5, 11, 14-16, 18, 24). A great deal of previous studies demonstrated that H9N2 influenza viruses have become established in terrestrial poultry in different Asian countries (5, 11, 13, 14, 18, 21, 24, 35). In 1994, H9N2 viruses were isolated from diseased chickens in Guangdong province, China, for the first time (4), and later in domestic poultry in other provinces in China (11, 16, 18, 35). Two distinct H9N2 virus lineages represented by A/Chicken/Beijing/1/94 (H9N2) and A/Quail/Hong Kong/G1/98 (H9N2), respectively, have been circulating in terrestrial poultry of southern China (9). Occasionally these viruses expand their host range to other mammals, including pigs and humans (6, 17, 22, 34). Increasing epidemiological and laboratory findings suggest that chickens may play an important role in expanding the host range for avian influenza virus. Our systematic surveillance of influenza viruses in chickens in China showed that H9N2 subtype influenza viruses continued to be prevalent in chickens in mainland China from 1994 to 2008 (18, 19, 36).Eastern China contains one metropolitan city (Shanghai) and five provinces (Jiangsu, Zhejiang, Anhui, Shandong, and Jiangxi), where domestic poultry account for approximately 50% of the total poultry population in China. Since 1996, H9N2 influenza viruses have been isolated regularly from both chickens and other minor poultry species in our surveillance program in the eastern China region, but their genetic diversity and the interrelationships between H9N2 influenza viruses and different types of poultry have not been determined. Therefore, it is imperative to explore the evolution and properties of these viruses. The current report provides insight into the genesis and evolution of H9N2 influenza viruses in eastern China and presents new evidence for the potential crossover between H9N2 and H5N1 influenza viruses in this region.  相似文献   

12.
The clinical impact of the 2009 pandemic influenza A(H1N1) virus (pdmH1N1) has been relatively low. However, amino acid substitution D222G in the hemagglutinin of pdmH1N1 has been associated with cases of severe disease and fatalities. D222G was introduced in a prototype pdmH1N1 by reverse genetics, and the effect on virus receptor binding, replication, antigenic properties, and pathogenesis and transmission in animal models was investigated. pdmH1N1 with D222G caused ocular disease in mice without further indications of enhanced virulence in mice and ferrets. pdmH1N1 with D222G retained transmissibility via aerosols or respiratory droplets in ferrets and guinea pigs. The virus displayed changes in attachment to human respiratory tissues in vitro, in particular increased binding to macrophages and type II pneumocytes in the alveoli and to tracheal and bronchial submucosal glands. Virus attachment studies further indicated that pdmH1N1 with D222G acquired dual receptor specificity for complex α2,3- and α2,6-linked sialic acids. Molecular dynamics modeling of the hemagglutinin structure provided an explanation for the retention of α2,6 binding. Altered receptor specificity of the virus with D222G thus affected interaction with cells of the human lower respiratory tract, possibly explaining the observed association with enhanced disease in humans.In April 2009, the H1N1 influenza A virus of swine origin was detected in humans in North America (9, 12, 42). Evidence for its origin came from analyses of the viral genome, with six gene segments displaying the closest resemblance to American “triple-reassortant” swine viruses and two to “Eurasian-lineage” swine viruses (13, 42). After this first detection in humans, the virus spread rapidly around the globe, starting the first influenza pandemic of the 21st century. The 2009 pandemic influenza A(H1N1) virus (pdmH1N1) has been relatively mild, with a spectrum of disease ranging from subclinical infections or mild upper respiratory tract illness to sporadic cases of severe pneumonia and acute respiratory distress syndrome (3, 11, 27, 29, 30, 37). Overall, the case-fatality rate during the start of the pandemic was not significantly higher than in seasonal epidemics in most countries. However, a marked difference was observed in the case-fatality rate in specific age groups, with seasonal influenza generally causing highest mortality in elderly and immunocompromised individuals, and the pdmH1N1 affecting a relatively large proportion of (previously healthy) young individuals (3, 11, 27, 29, 30, 37).Determinants of influenza A virus virulence have been mapped for a wide variety of zoonotic and pandemic influenza viruses to the polymerase genes, hemagglutinin (HA), neuraminidase (NA), and nonstructural protein 1 (NS1). Such virulence-associated substitutions generally facilitate more efficient replication in humans via improved interactions with host cell factors. Since most of these virulence-associated substitutions were absent in the earliest pdmH1N1s, it has been speculated that the virus could acquire some of these mutations, potentially resulting in the emergence of more pathogenic viruses. Such virulence markers could be acquired by gene reassortment with cocirculating influenza A viruses, or by mutation. The influenza virus polymerase genes, in particular PB2, have been shown to be important determinants of the virulence of the highly pathogenic avian influenza (HPAI) H5N1 and H7N7 viruses and the transmission of the 1918 H1N1 Spanish influenza virus (17, 26, 34, 51). One of the most commonly identified virulence markers to date is E627K in PB2. The glutamic acid (E) residue is generally found in avian influenza viruses, while human viruses have a lysine (K), and this mutation was described as a determinant of host range in vitro (48). Given that all human and many zoonotic influenza viruses of the last century contained 627K, it was surprising that the pdmH1N1 had 627E. In addition, an aspartate (D)-to-asparagine (N) substitution at position 701 (D701N) of PB2 has previously been shown to expand the host range of avian H5N1 virus to mice and humans and to increase virus transmission in guinea pigs (26, 46). Like E627K, D701N was absent in the genome of pdmH1N1. Thus, the pdmH1N1 was the first known human pandemic virus with 627E and 701D, and it has been speculated that pdmH1N1 could mutate into a more virulent form by acquiring one of these mutations or both. Recently, it was shown that neither E627K nor D701N in PB2 of pdmH1N1 increased its virulence in ferrets and mice (18). The PB1-F2 protein has previously also been associated with high pathogenicity of the 1918 H1N1 and HPAI H5N1 viruses (8). The PB1-F2 protein of the pdmH1N1 is truncated due to premature stop codons. However, restoration of the PB1-F2 reading frame did not result in viruses with increased virulence (15). The NS1 protein of pdmH1N1 is also truncated due to a stop codon and, as a result, does not contain a PDZ ligand domain that is involved in cell-signaling pathways and has been implicated in the pathogenicity of 1918 H1N1 and HPAI H5N1 viruses (5, 8, 21). Surprisingly, restoration of a full-length version of the NS1 gene did not result in increased virulence in animal models (16). Mutations affecting virulence and host range have further frequently been mapped to hemagglutinin (HA) and neuraminidase (NA) in relation to their interaction with α2,3- or α2,6-linked sialic acids (SAs), the virus receptors on host cells (17, 32, 35, 50). The HA gene of previous pandemic viruses incorporated substitutions that allow efficient attachment to α2,6-SAs—the virus receptor on human cells—compared to ancestral avian viruses that attach more efficiently to α2,3-SAs (35, 47, 50).To search for mutations of potential importance to public health, numerous laboratories performed genome sequencing of pdmH1N1s, resulting in the real-time accumulation of information on emergence of potential virulence markers. Of specific interest were reports on amino acid substitutions from aspartic acid (D) to glycine (G) at position 222 (position 225 in H3) in HA of pdmH1N1. This substitution was observed in a fatal case of pdmH1N1 infection in June 2009 in the Netherlands (M. Jonges et al., unpublished data). Between July and December 2009, viruses from 11 (18%) of 61 cases with severe disease outcome in Norway have also been reported to harbor the D222G substitution upon direct sequencing of HA in clinical specimens. Such mutant viruses were not observed in any of 205 mild cases investigated, and the frequency of detection of this mutation was significantly higher in severe cases than in mild cases (23). In Hong Kong, the D222G substitution was detected in 12.5% (6) and 4.1% (31) of patients with severe disease and in 0% of patients with mild disease, in two different studies without prior propagation in embryonated chicken eggs. In addition to Norway and Hong Kong, the mutation has been detected in Brazil, Japan, Mexico, Ukraine, and the United States (56). Thus, D222G in HA could be the first identified “virulence marker” of pdmH1N1. pdmH1N1 with D222G in HA have not become widespread in the population, although they were detected in several countries. However, D222G in HA is of special interest, since it has also been described as the single change in HA between two strains of the “Spanish” 1918 H1N1 virus that differed in receptor specificity (47). Furthermore, upon propagation in embryonated chicken eggs, pdmH1N1 can acquire the mutation rapidly, presumably because it results in virus adaptation to avian (α2,3-SAs) receptors (49). The presence of the substitution in pdmH1N1s in the human population and its potential association with more severe disease prompted us to test its effect on pdmH1N1 receptor binding, replication, antigenic properties, and pathogenesis and transmission in animal models.  相似文献   

13.
14.
The spread of the recently emerged, highly pathogenic H5N1 avian influenza virus has raised concern. Preclinical studies suggest that passive immunotherapy could be a new form of treatment for H5N1 virus infection. Here, a neutralizing monoclonal antibody (MAb) against the hemagglutinin (HA) of the influenza A/chicken/Hatay/2004 H5N1 virus, MAb 9F4, was generated and characterized. MAb 9F4 binds both the denatured and native forms of HA. It was shown to recognize the HA proteins of three heterologous strains of H5N1 viruses belonging to clades 1, 2.1, and 2.2, respectively. By use of lentiviral pseudotyped particles carrying HA on the surface, MAb 9F4 was shown to effectively neutralize the homologous strain, Hatay04, and another clade 1 strain, VN04, at a neutralization titer of 8 ng/ml. Furthermore, MAb 9F4 also neutralized two clade 2 viruses at a neutralizing titer of 40 ng/ml. The broad cross-neutralizing activity of MAb 9F4 was confirmed by its ability to neutralize live H5N1 viruses of clade 2.2.2. Epitope-mapping analysis revealed that MAb 9F4 binds a previously uncharacterized epitope below the globular head of the HA1 subunit. Consistently, this epitope is well conserved among the different clades of H5N1 viruses. MAb 9F4 does not block the interaction between HA and its receptor but prevents the pH-mediated conformational change of HA. MAb 9F4 was also found to be protective, both prophylactically and therapeutically, against a lethal viral challenge of mice. Taken together, our results showed that MAb 9F4 is a neutralizing MAb that binds a novel and well-conserved epitope in the HA1 subunit of H5N1 viruses.The highly pathogenic avian influenza A subtype H5N1 virus was first isolated from geese in Guangdong province, China, in 1996 (44). Since 2003, the H5N1 strains have caused major morbidity and mortality in poultry populations across Asia, Europe, and Africa (3, 25). In 1997, the virus was transmitted from chickens to humans in Hong Kong, causing 18 reported cases of illness, including 6 deaths (6, 7, 37). As of September 2009, there were 442 confirmed human infections in 15 countries, with an alarming fatality rate of 59% (42). Although occurrences of human H5N1 infection are sporadic and rare, its rapid dissemination, the ongoing evolution of the avian H5N1 virus, and the absence of anti-H5N1 herd immunity in humans raise concerns regarding a possible H5N1 influenza pandemic (2, 4, 13). Since human infections are associated with severe disease and high mortality, the consequences of a pandemic could be catastrophic.Current strategies against influenza include vaccination and antiviral drug treatment (1). Due to the existence of multiple antigenic clades and subclades of the H5N1 virus, the difficulty of predicting the major strain that may cause the next pandemic is the main obstacle to current vaccine development. Moreover, resistance to M2 ion channel inhibitors (rimantidine and amantidine) has been reported in H5N1 isolates (1, 5), and the neuraminidase inhibitors (oseltamivir and zanamivir) require higher doses and prolonged treatment (45), and resistance has been reported in children (21). Passive immunotherapy is now increasingly used to treat numerous human infectious diseases (28, 33). Convalescent-phase blood and serum products were used to improve clinical outcomes for severely ill influenza patients during the 1918 influenza pandemic (27). Promising results with mouse models using a neutralizing monoclonal antibody (MAb) for H5N1 influenza treatment (17, 26) and a report of the recovery of an H5N1 virus-infected patient after treatment with convalescent-phase plasma (47) indicate that MAbs could be a potential treatment against H5N1 viruses.The hemagglutinin (HA) protein is one of the two major surface glycoproteins on the envelope of influenza A virus, with 16 distinct types identified in the avian species. The HA protein is responsible for receptor binding to host cells and for viral entry and is therefore the primary target of neutralizing antibodies (Abs) (35). It is a homotrimer, with each subunit made up of two disulfide-linked polypeptides, HA1 and HA2. Structurally, each subunit consists of a membrane-proximal helix-rich stem structure and a membrane-distal receptor binding globular domain (35).In this study, we describe a MAb, named MAb 9F4, raised against the recombinant baculovirus-expressed HA protein of A/chicken/Hatay/2004 H5N1 virus. Its neutralizing property was investigated, and epitope mapping was performed. The MAb 9F4 binding site was found to lie outside previously characterized antigenic sites in the HA protein. This epitope is well conserved among the different clades of H5N1 viruses, consistent with the cross-neutralizing activity of MAb 9F4. The mode of inhibition was also investigated, and MAb 9F4 was found to mediate postattachment neutralization in a dose-dependent manner. Finally, the protective ability of MAb 9F4 was also evaluated in a mouse model, and it was shown to protect against lethal H5N1 challenge both prophylactically and therapeutically. Taken together, the data could provide new information for the design of an H5N1 vaccine, and MAb 9F4 may be a possible candidate for use in passive immunotherapy.  相似文献   

15.
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   

16.
Highly pathogenic H5N1 influenza viruses continue to cause concern, even though currently circulating strains are not efficiently transmitted among humans. For efficient transmission, amino acid changes in viral proteins may be required. Here, we examined the amino acids at positions 627 and 701 of the PB2 protein. A direct analysis of the viral RNAs of H5N1 viruses in patients revealed that these amino acids contribute to efficient virus propagation in the human upper respiratory tract. Viruses grown in culture or eggs did not always reflect those in patients. These results emphasize the importance of the direct analysis of original specimens.Given the continued circulation of highly pathogenic H5N1 avian influenza viruses and their sporadic transmission to humans, the threat of a pandemic persists. However, for H5N1 influenza viruses to be efficiently transmitted among humans, amino acid substitutions in the avian viral proteins may be necessary.Two positions in the PB2 protein affect the growth of influenza viruses in mammalian cells (3, 11, 18): the amino acid at position 627 (PB2-627), which in most human influenza viruses is lysine (PB2-627Lys) and most avian viruses is glutamic acid (PB2-627Glu), and the amino acid at position 701. PB2-627Lys is associated with the efficient replication (16) and high virulence (5) of H5N1 viruses in mice. Moreover, an H7N7 avian virus isolated from a fatal human case of pneumonia possessed PB2-627Lys, whereas isolates from a nonfatal human case of conjunctivitis and from chickens during the same outbreak possessed PB2-627Glu (2).The amino acid at position 701 in PB2 is important for the high pathogenicity of H5N1 viruses in mice (11). Most avian influenza viruses possess aspartic acid at this position (PB2-701Asp); however, A/duck/Guangxi/35/2001 (H5N1), which is highly virulent in mice (11), possesses asparagine at this position (PB2-701Asn). PB2-701Asn is also found in equine (4) and swine (15) viruses, as well as some H5N1 human isolates (7, 9). Thus, both amino acids appear to be markers for the adaptation of H5N1 viruses in humans (1, 3, 17).Massin et al. (13) reported that the amino acid at PB2-627 affects viral RNA replication in cultured cells at low temperatures. Recently, we demonstrated that viruses, including those of the H5N1 subtype, with PB2-627Lys (human type) grow better at low temperatures in cultured cells than those with PB2-627Glu (avian type) (6). This association between the PB2 amino acid and temperature-dependent growth correlates with the body temperatures of hosts; the human upper respiratory tract is at a lower temperature (around 33°C) than the lower respiratory tract (around 37°C) and the avian intestine, where avian influenza viruses usually replicate (around 41°C). The ability to replicate at low temperatures may be crucial for viral spread among humans via sneezing and coughing by being able to grow in the upper respiratory organs. Therefore, the Glu-to-Lys mutation in PB2-627 is an important step for H5N1 viruses to develop pandemic potential.However, there is no direct evidence that the substitutions of PB2-627Glu with PB2-627Lys and PB2-701Asp with PB2-701Asn occur during the replication of H5N1 avian influenza viruses in human respiratory organs. Therefore, here, we directly analyzed the nucleotide sequences of viral genes from several original specimens collected from patients infected with H5N1 viruses.  相似文献   

17.
18.
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   

19.
While few children and young adults have cross-protective antibodies to the pandemic H1N1 2009 (pdmH1N1) virus, the illness remains mild. The biological reasons for these epidemiological observations are unclear. In this study, we demonstrate that the bulk memory cytotoxic T lymphocytes (CTLs) established by seasonal influenza viruses from healthy individuals who have not been exposed to pdmH1N1 can directly lyse pdmH1N1-infected target cells and produce gamma interferon (IFN-γ) and tumor necrosis factor alpha (TNF-α). Using influenza A virus matrix protein 1 (M158-66) epitope-specific CTLs isolated from healthy HLA-A2+ individuals, we further found that M158-66 epitope-specific CTLs efficiently killed both M158-66 peptide-pulsed and pdmH1N1-infected target cells ex vivo. These M158-66-specific CTLs showed an effector memory phenotype and expressed CXCR3 and CCR5 chemokine receptors. Of 94 influenza A virus CD8 T-cell epitopes obtained from the Immune Epitope Database (IEDB), 17 epitopes are conserved in pdmH1N1, and more than half of these conserved epitopes are derived from M1 protein. In addition, 65% (11/17) of these epitopes were 100% conserved in seasonal influenza vaccine H1N1 strains during the last 20 years. Importantly, seasonal influenza vaccination could expand the functional M158-66 epitope-specific CTLs in 20% (4/20) of HLA-A2+ individuals. Our results indicated that memory CTLs established by seasonal influenza A viruses or vaccines had cross-reactivity against pdmH1N1. These might explain, at least in part, the unexpected mild pdmH1N1 illness in the community and also might provide some valuable insights for the future design of broadly protective vaccines to prevent influenza, especially pandemic influenza.Since its first identification in North America in April 2009, the novel pandemic H1N1 2009 (pdmH1N1) virus has been spreading in humans worldwide, giving rise to the first pandemic in the 21st century (13, 18). The pdmH1N1 virus contains a unique gene constellation, with its NA and M gene segments being derived from the Eurasian swine lineage while the other gene segments originated from the swine triple-reassortant H1N1 lineage. The triple-reassortant swine viruses have in turn derived the HA, NP, and NS gene segments from the classical swine lineage (20). The 1918 pandemic virus gave rise to both the seasonal influenza H1N1 and the classical swine H1N1 virus lineages (41). Evolution in different hosts during the subsequent 90 years has led to increasing antigenic differences between recent seasonal H1N1 viruses and swine H1 viruses (42). Thus, younger individuals have no antibodies that cross neutralize pdmH1N1, while those over 65 years of age are increasingly likely to have cross-neutralizing antibodies to pdmH1N1 (10, 25).Currently available seasonal influenza vaccines do not induce cross-reactive antibodies against this novel virus in any age group (10, 25). In animal models, it has been shown that pdmH1N1 replicated more efficiently and caused more severe pathological lesions than the current seasonal influenza virus (28). However, most patients with pdmH1N1 virus infection show a mild illness comparable to seasonal influenza (9, 42). The incidence of severe cases caused by pdmH1N1 was not significantly higher than that caused by human seasonal influenza viruses (43). These findings imply that seasonal influenza A virus-specific memory T cells preexisting in previously infected individuals may have cross-protection to this novel pdmH1N1.Cross-reactivity of influenza A virus-specific T-cell immunity against heterosubtypic strains which are serologically distinct has been demonstrated (5, 29, 33, 47). Humans who have not been exposed to avian influenza A (H5N1) virus do have cross-reactive memory CD4 and CD8 T cells to a wide range of H5N1 peptides (33, 47). More recently, one study also showed that some seasonal influenza A virus-specific memory T cells in individuals without exposure to prior pdmH1N1 infection can recognize pdmH1N1 (24). However, the results in most of these studies were determined by the gamma interferon (IFN-γ) responses to influenza virus peptides. Although the recalled IFN-γ response is commonly used to detect memory CD4 and CD8 T cells, the activated T cells that bind major histocompatibility complex (MHC)-presented peptide are not necessarily capable of lysing the target cells (6). In addition, the peptides, but not the whole virus, may not be able to fully represent the human cross-response against the virus as a whole. Therefore, in addition to cytokine production, the demonstration of direct antigen-specific cytotoxicity of cytotoxic T lymphocytes (CTLs) against both peptide-pulsed and virus-infected target cells is needed for better understanding of human CTL responses against pdmH1N1 virus.In this study, using bulk memory CTLs and epitope-specific CTLs established by seasonal influenza A viruses and epitope-specific peptide from healthy individuals, respectively, we evaluated their cross-cytotoxicity and cytokine responses to pdmH1N1. We also examined the expression of chemokine receptors CXCR3 and CCR5, which could help CTLs to migrate to the site of infection. In addition, to understand whether the seasonal influenza vaccines have benefit for people who have not been exposed to pdmH1N1, we further examined the ability of seasonal influenza vaccines to induce the conserved M158-66 epitope-specific CTLs in HLA-A2-seropositive healthy individuals.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号