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 28 毫秒
1.
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   

2.
3.
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   

4.
The highly pathogenic avian influenza (HPAI) virus phenotype is restricted to influenza A viruses of the H5 and H7 hemagglutinin (HA) subtypes. To obtain more information on the apparent subtype-specific nature of the HPAI virus phenotype, a low-pathogenic avian influenza (LPAI) H6N1 virus was generated, containing an HPAI H5 RRRKKR↓G multibasic cleavage site (MBCS) motif in HA (the downward arrow indicates the site of cleavage). This insertion converted the LPAI virus phenotype into an HPAI virus phenotype in vitro and in vivo. The H6N1 virus with an MBCS displayed in vitro characteristics similar to those of HPAI H5 viruses, such as cleavage of HA0 (the HA protein of influenza A virus initially synthesized as a single polypeptide precursor) and virus replication in the absence of exogenous trypsin. Studies of chickens confirmed the HPAI phenotype of the H6N1 virus with an MBCS, with an intravenous pathogenicity index of 1.4 and systemic virus replication upon intranasal inoculation, the hallmarks of HPAI viruses. This study provides evidence that the subtype-specific nature of the emergence of HPAI viruses is not at the molecular, structural, or functional level, since the introduction of an MBCS resulted in a fully functional virus with an HPAI virus genotype and phenotype.Wild birds represent the natural reservoir of avian influenza A viruses in nature (43). Influenza A viruses are classified on the basis of the hemagglutinin (HA) and neuraminidase (NA) surface glycoproteins. In wild birds throughout the world, influenza A viruses representing 16 HA and 9 NA antigenic subtypes have been found in numerous combinations (also called subtypes, e.g., H1N1, H6N1) (12). Besides classification based on the antigenic properties of HA and NA, avian influenza A viruses can also be classified based on their pathogenic phenotype in chickens. Highly pathogenic avian influenza (HPAI) virus, an acute generalized disease of poultry in which mortality may be as high as 100%, is restricted to subtypes H5 and H7. Other avian influenza A virus subtypes are generally low-pathogenic avian influenza (LPAI) viruses that cause much milder, primarily respiratory disease in poultry, sometimes with loss of egg production (6).The HA protein of influenza A virus is initially synthesized as a single polypeptide precursor (HA0), which is cleaved into HA1 and HA2 subunits by host cell proteases. The mature HA protein mediates binding of the virus to host cells, followed by endocytosis and HA-mediated fusion with endosomal membranes (43). Influenza viruses of subtypes H5 and H7 may become highly pathogenic after introduction into poultry and cause outbreaks of HPAI. The switch from an LPAI phenotype to the HPAI phenotype of these H5 and H7 influenza A viruses is achieved by the introduction of basic amino acid residues into the HA0 cleavage site by substitution or insertion, resulting in the so-called multibasic cleavage site (MBCS), which facilitates systemic virus replication (4, 5, 14, 44). The cleavage of the HA0 of LPAI viruses is restricted to trypsin-like proteases which recognize the XXX(R/K)↓G cleavage motif, where the downward arrow indicates the site of cleavage. Replication of these LPAI viruses is therefore restricted to sites in the host where these enzymes are expressed, i.e., the respiratory and intestinal tract (32, 38). The introduction of an RX(R/K)R↓G or R(R/K)XR↓G minimal MBCS motif into the H5 and H7 subtype viruses facilitates the recognition and cleavage of the HA0 by ubiquitous proprotein convertases, such as furin (20, 32, 41, 45). H5 influenza A viruses with a minimal MBCS motif only have the highly pathogenic phenotype if the masking glycosylation site at position 11 in the HA is replaced by a nonglycosylation site. Otherwise, at least one additional basic amino acid has to be inserted to allow the shift from an LPAI virus phenotype to an HPAI virus phenotype to occur (15, 18, 21, 22, 28). No information is available on the minimal prerequisites of H7 influenza A viruses to become highly pathogenic, but all HPAI H7 viruses have at least 2 basic amino acid insertions in the HA0 cleavage site (22). HA0 with the MBCS is activated in a broad range of different host cells and therefore enables HPAI viruses to replicate systemically in poultry (46). To date, little is known about the apparent subtype-specific nature of the introduction of the MBCS into LPAI viruses and the evolutionary processes involved in the emergence of HPAI viruses. When an MBCS was introduced in a laboratory-adapted strain of influenza virus, A/Duck/Ukraine/1/1963 (H3N8), it did not result in a dramatic change in pathogenic phenotype (35). Here, the effect of the introduction of an MBCS into a primary LPAI H6N1 virus, A/Mallard/Sweden/81/2002, is described. The introduction of an MBCS resulted in trypsin-independent replication in vitro and enhanced pathogenesis in a chicken model. Understanding the basis of the HA subtype specificity of the introduction of an MBCS into avian influenza viruses will lead to a better understanding of potential molecular restrictions involved in emergence of HPAI outbreaks.  相似文献   

5.
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   

6.
7.
Isolation of human subtype H3N2 influenza viruses in embryonated chicken eggs yields viruses with amino acid substitutions in the hemagglutinin (HA) that often affect binding to sialic acid receptors. We used a glycan array approach to analyze the repertoire of sialylated glycans recognized by viruses from the same clinical specimen isolated in eggs or cell cultures. The binding profiles of whole virions to 85 sialoglycans on the microarray allowed the categorization of cell isolates into two groups. Group 1 cell isolates displayed binding to a restricted set of α2-6 and α2-3 sialoglycans, whereas group 2 cell isolates revealed receptor specificity broader than that of their egg counterparts. Egg isolates from group 1 showed binding specificities similar to those of cell isolates, whereas group 2 egg isolates showed a significantly reduced binding to α2-6- and α2-3-type receptors but retained substantial binding to specific O- and N-linked α2-3 glycans, including α2-3GalNAc and fucosylated α2-3 glycans (including sialyl Lewis x), both of which may be important receptors for H3N2 virus replication in eggs. These results revealed an unexpected diversity in receptor binding specificities among recent H3N2 viruses, with distinct patterns of amino acid substitution in the HA occurring upon isolation and/or propagation in eggs. These findings also suggest that clinical specimens containing viruses with group 1-like receptor binding profiles would be less prone to undergoing receptor binding or antigenic changes upon isolation in eggs. Screening cell isolates for appropriate receptor binding properties might help focus efforts to isolate the most suitable viruses in eggs for production of antigenically well-matched influenza vaccines.Influenza A viruses are generally isolated and propagated in embryonated chicken eggs or in cultures of cells of mammalian origin. Human influenza viruses were previously noted to acquire mutations in the hemagglutinin (HA) gene upon isolation and culture in the allantoic sac of embryonated chicken eggs (herein simply referred to as “eggs”) compared to the sequences of those isolated in mammalian cell substrates (herein referred to as “cells”) (29, 30, 44, 53, 58). These mutations resulted in amino acid substitutions that were found to mediate receptor specificity changes and improved viral replication efficiency in eggs (37). In general, cell-grown viruses are assumed to be more similar than their egg-grown counterparts to the viruses present in respiratory secretions (30, 56). Since their emergence in 1968, influenza A (H3N2) viruses have evolved and adapted to the human host while losing their ability to be efficiently isolated and replicate in eggs, particularly after 1992 (37, 42, 48). The rate of isolation of H3N2 clinical specimens after inoculation into eggs can be up to ∼30 times lower than that in mammalian cell cultures, highlighting the strong selective pressure for the emergence of sequence variants (77).Virtually all influenza vaccines for human use were licensed decades ago by national regulatory authorities, which used a product manufactured from influenza viruses isolated and propagated exclusively in eggs; therefore, cell culture isolates have been unacceptable for this purpose (41, 71). The antigen composition of influenza vaccines requires frequent updates (every 2 years, on average) to closely match their antigenic properties to the most prevalent circulating antigenic drift variant viruses (51). The limited availability of H3N2 viruses isolated in eggs has on one or more occasions delayed vaccine composition updates and may have reduced the efficacy of vaccination against new antigenically drifted viruses (3, 34, 37).Entry of influenza viruses into host cells is mediated by HA, which binds to sialic acid containing glycoconjugates on the surface of epithelial cells in the upper respiratory tract (2, 13). The nature of the linkage between sialic acid and the vicinal sugar (usually galactose) varies in different host species and tissues and may therefore determine whether an influenza virus binds to and infects avian or human cells (40, 46, 59, 62, 72-75). Human influenza viruses preferentially bind to α2-6-linked sialic acids, and avian viruses predominantly bind to α2-3-linked sialic acids (59). Previous studies with chicken embryo chorioallantoic membranes revealed differential lectin binding, suggesting that α2-3-linked but not α2-6-linked sialosides are present on the epithelial cells (28). Human H3N2 viruses isolated in cell culture were reported to bind with a high affinity to α2-6-linked sialosides, while viruses isolated in eggs often had increased specificity for α2-3-linked sialosides (19, 20, 28). The functional classification of avian and mammalian influenza virus receptors is further complicated since in vitro and tissue-binding assays have led to new working hypotheses involving glycan chain length, topology, and the composition of the inner fragments of the carbohydrate chain as additional receptor specificity determinants (9, 17, 65, 66, 82). However, the significance of these in vitro properties remains unknown, since the structures of the natural sialosides on host cells that are used for infectious virus entry are undefined.The techniques most widely used to study the interactions of the influenza virus with host cell receptors employ animal cells in various assay formats (36, 57, 59, 64, 69). To overcome the problems of cell-based techniques, new assays that rely on labeled sialyl-glycoproteins or polymeric sialoglycans have been developed (18). However, these assays are limited by having only a few glycans available in polymeric form and offer low throughput. In contrast, glycan microarrays can assess virus binding to multiple well-defined glycans simultaneously. Previous work with influenza live or β-propiolactone (BPL)-inactivated virions as well as recombinantly produced HAs revealed a good correlation with receptor specificity compared to that achieved by other methods of analysis (4, 11, 57, 58, 65-68).Here we have compared paired isolates derived in eggs or cell cultures from the single clinical specimen to better understand their receptor binding specificity and its implications for vaccine production. We examined the differences in the sequences of the HAs between egg- and cell-grown isolates and analyzed their receptor binding profiles using glycan microarrays. Sequence analysis of the HA and glycan binding results revealed two distinct groups of viruses, with many egg isolates showing unexpectedly reduced levels of binding to α2-3 and α2-6 sialosides compared to the levels for the viruses isolated in mammalian cells. Furthermore, these studies highlighted that specific glycans may be important for H3N2 virus growth in eggs.  相似文献   

8.
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   

9.
Since dendritic cells may play a key role in defense against influenza virus infection, we examined the effects of recombinant hemagglutinin (HA) proteins derived from mouse-adapted H1N1 (A/WSN/1933), swine-origin 2009 pandemic H1N1 (A/Texas/05/2009), and highly pathogenic avian influenza H5N1 (A/Thailand/KAN-1/2004) viruses on mouse myeloid dendritic cells (mDCs). The results reveal that tumor necrosis factor alpha (TNF-α), interleukin-12 (IL-12) p70, and major histocompatibility complex class II (MHC-II) expression was increased in mDCs after treatment with recombinant HA proteins of H1N1 and H5N1. The specificity of recombinant HA treatments for mDC activation was diminished after proteinase K digestion. HA apparently promotes mDC maturation by enhancing CD40 and CD86 expression and suppressing endocytosis. No significant differences in mDC activation were observed among recombinant proteins of H1N1 and H5N1. The stimulation of mDCs by HA proteins of H1N1 and H5N1 was completely MyD88 dependent. These findings may provide useful information for the development of more-effective influenza vaccines.Influenza viruses trigger seasonal epidemics or pandemics of contagious diseases with mild to severe consequences in human and poultry populations worldwide (28). Members of the Orthomyxoviridae family, influenza viruses consist of single-stranded, eight-segment, negative-sense genomic RNAs, helical viral ribonucleoprotein (RNP) complexes (RNA segments, NP, PB2, PB1, and PA) and four viral envelope proteins (hemagglutinin [HA], neuraminidase [NA], and M1 and M2 matrix proteins). Type A influenza viruses are further classified into various serotypes based on the antigenic characteristics of HA and NA glycoproteins (14).In 2009, a swine-origin H1N1 strain emerged from the genetic reassortants of existing human, avian, and swine influenza viruses, resulting in a global pandemic marked by symptoms more severe than those associated with seasonal influenza virus (3, 24). According to comparative pathology in macaque monkeys, H5N1 induces greater cytokinemia, tissue damage, and interference with immune regulatory mechanisms than H1N1 infection (2). The HA spike protein of influenza virus is believed to play important roles in viral receptor binding, fusion, transmission, host range restriction, virulence, and pathogenesis (13, 27-30).Dendritic cells (DCs), considered the most potent professional antigen-presenting cells, serve as links between innate and adaptive immunity (31). Upon encountering microbial pathogens, endogenous danger signals, or inflammatory mediators, DCs mature and elicit rapid and short-lived innate immune responses before migrating to secondary lymphoid organs and enhancing adaptive immunity (17). Two major subsets of DCs are recognized in mice and humans: (i) myeloid DCs (mDCs, also called conventional DCs), which participate most directly in antigen presentation and activation of naïve T cells, and (ii) plasmacytoid DCs (pDCs), which produce type I interferons in response to viral infection (16, 42) and are also capable of inducing immunotolerance under some conditions (9). mDCs and pDCs also comprise different heterologous subsets, with unique localizations, phenotypes, and functions (36). Due to their key role in immune regulation, DCs have been developed for immunotherapeutic agents or prophylactic or therapeutic vaccines for cancer, infectious diseases, and immune system-related diseases (32, 34).DCs are essential in controlling the innate and adaptive immune responses against influenza virus infection (21). Viral RNA is recognized by various pattern recognition receptors (PRRs), including RIG-I-like receptors (RLRs), Toll-like receptors (TLRs), and nucleotide oligomerization domain (NOD)-like receptors (NLRs). TLRs play an especially important role in detecting virus invasion and activating DCs (18, 35). However, the mechanisms causing DC activation and maturation in response to influenza viruses are not clear. HA has been described as playing an important role in modulating influenza virus virulence and host immune responses (29). In this study, we examined the effects of several recombinant HA proteins (rHAs) derived from rHA of H1N1 (rH1HA) (A/WSN/1933) and (A/Texas/05/2009) and rHA of H5N1 (rH5HA) (A/Thailand/KAN-1/2004) viruses on the activation and maturation of the mDC subset derived from mouse bone marrow.  相似文献   

10.
Adaptation of influenza A viruses to a new host species usually involves the mutation of one or more of the eight viral gene segments, and the molecular basis for host range restriction is still poorly understood. To investigate the molecular changes that occur during adaptation of a low-pathogenic avian influenza virus subtype commonly isolated from migratory birds to a mammalian host, we serially passaged the avirulent wild-bird H5N2 strain A/Aquatic bird/Korea/W81/05 (W81) in the lungs of mice. The resulting mouse-adapted strain (ma81) was highly virulent (50% mouse lethal dose = 2.6 log10 50% tissue culture infective dose) and highly lethal. Nonconserved mutations were observed in six viral genes (those for PB2, PB1, PA, HA, NA, and M). Reverse genetic experiments substituting viral genes and mutations demonstrated that the PA gene was a determinant of the enhanced virulence in mice and that a Thr-to-Iso substitution at position 97 of PA played a key role. In growth kinetics studies, ma81 showed enhanced replication in mammalian but not avian cell lines; the PA97I mutation in strain W81 increased its replicative fitness in mice but not in chickens. The high virulence associated with the PA97I mutation in mice corresponded to considerably enhanced polymerase activity in mammalian cells. Furthermore, this characteristic mutation is not conserved among avian influenza viruses but is prevalent among mouse-adapted strains, indicating a host-dependent mutation. To our knowledge, this is the first study that the isoleucine residue at position 97 in PA plays a key role in enhanced virulence in mice and is implicated in the adaptation of avian influenza viruses to mammalian hosts.Migratory waterfowl are the natural reservoir of influenza A viruses (11, 53). The viruses replicate efficiently in their natural hosts but replicate poorly if at all in other species (53). However, these viruses can undergo adaptation or genetic reassortment to infect other hosts (43, 44, 53), including humans. Since 1997, the World Health Organization has documented more than 400 laboratory-confirmed cases of human infection with H5N1 avian influenza virus (54).The molecular basis of influenza virus host range restriction and adaptation to a new host species is poorly understood. Mutations associated with cross-species adaptation are thought to be associated with increased virulence (30). Therefore, studies in animal models have attempted to identify the viral molecular determinants of virulence in specific hosts. Reverse genetics (Rg) methods have also identified genetic differences that affect virus virulence and host range, including changes in the viral internal proteins. Experimental infection of mouse lungs is an effective approach for understanding influenza virus virulence and adaptation (reviewed by A. C. Ward [51]). To acquire virulence in mice, influenza A viruses usually must adapt to these hosts over several consecutive generations (serial passages) in the lungs or brain (1, 25, 30). Previous studies have found that the acquisition of virulence during adaptation in the mouse model is associated with mutations in the HA, NP, NA, M, and NS genes and one or more polymerase genes (2, 3, 18, 36, 42, 51). The polymerase basic protein 2 (PB2) gene is a particularly well-characterized polymerase subunit (7, 23, 40, 46). The PB1 and polymerase acidic protein (PA) genes have been implicated in mouse lung virulence (5, 18, 36, 39, 49) but have shown no evidence of having acquired mutations during adaptation (52). However, the many studies conducted to date have focused mainly on highly pathogenic avian influenza (HPAI) viruses such as the H1N1, H5N1, and H7N7 subtypes (7, 23, 48, 50).Various low-pathogenic avian influenza (LPAI) viruses are considered to be potential genetic contributors to the next pandemic strain. Lee et al. (2009) recently reported the presence of avian-like LPAI H5N2 viruses in a number of Korean swine and proposed that the efficient transmissibility of the swine-adapted H5N2 virus could facilitate spread of the virus. They suggested that this adapted virus could potentially serve as a model for pandemic outbreaks of HPAI (e.g., H5N1 and H7N7) virus or could become a pandemic strain itself (21). These findings prompted our interest in the adaptation of an LPAI virus often harbored by wild migratory birds of South Korea. In our ongoing surveillance from 2004 to 2008, approximately 27% of the viruses isolated were of the H5N2 subtype (unpublished data). Studies show that influenza viruses with different genetic backgrounds can acquire different mutations during adaptation in mice. Therefore, we sought to determine whether this common H5N2 virus (nonlethal in mice) would undergo changes different from those observed in highly virulent viruses during adaptation in mice. Wild-bird influenza virus strain A/Aquatic bird/Korea/W81/05 (W81) was adapted in mice over 11 passages and became highly virulent. To identify molecular determinants of this adaptation and altered virulence, we used Rg-generated recombinant viruses to compare the parental and mouse-adapted strains. Here we show that the PA subunit of the polymerase complex, independently of PB2, contributed to adaptation and increased virulence in our mammalian model.  相似文献   

11.
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   

12.
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   

13.
14.
The hemagglutinin (HA) surface glycoprotein promotes influenza virus entry and is the key protective antigen in natural immunity and vaccines. The HA protein is a trimeric envelope glycoprotein consisting of a globular receptor-binding domain (HA-RBD) that is inserted into a membrane fusion-mediating stalk domain. Similar to other class I viral fusion proteins, the fusogenic stalk domain spontaneously refolds into its postfusion conformation when expressed in isolation, consistent with this domain being trapped in a metastable conformation. Using X-ray crystallography, we show that the influenza virus HA-RBD refolds spontaneously into its native, immunogenic structure even when expressed in an unglycosylated form in Escherichia coli. In the 2.10-Å structure of the HA-RBD, the receptor-binding pocket is intact and its conformational epitopes are preserved. Recombinant HA-RBD is immunogenic and protective in ferrets, and the protein also binds with specificity to sera from influenza virus-infected humans. Overall, the data provide a structural basis for the rapid production of influenza vaccines in E. coli. From an evolutionary standpoint, the ability of the HA-RBD to refold spontaneously into its native conformation suggests that influenza virus acquired this domain as an insertion into an ancestral membrane-fusion domain. The insertion of independently folding domains into fusogenic stalk domains may be a common feature of class I viral fusion proteins.The genetic drift of seasonal influenza viruses and the occasional emergence of pandemic strains represent a continuing and serious burden on human health. Pandemic influenza viruses arise at irregular intervals, can infect up to 50% or more of the population, and vary in disease severity. Most notably, the H1N1 Spanish influenza pandemic of 1918 killed an estimated 20 to 50 million people worldwide, and the 1957 H2N2 Asian flu and 1968 H3N2 Hong Kong flu pandemics killed between 0.5 and 1 million people in the United States alone (30). The ongoing danger of influenza was recently emphasized by the emergence of the novel H1N1 pandemic virus from Mexico in April of 2009. The urgent need to speed up vaccine production was highlighted by this outbreak because over 340,000 confirmed cases and 4,100 deaths had occurred worldwide during the 6 months that were necessary to produce a vaccine using current procedures (39).As the major surface antigen of influenza A viruses, the hemagglutinin (HA) envelope glycoprotein is the primary source of natural immunity and the key target in vaccination. However, changes in the antigenic sites of the HA protein due to antigenic drift result in lost or diminished immunity acquired from previous infection or vaccination (35). This necessitates the production of new vaccines against seasonal influenza viruses each year. The HA protein also plays a central role in the emergence of human pandemic influenza viruses. There are 16 known antigenic subtypes of HA proteins in influenza A viruses (H1 through H16), and a pandemic occurs when an influenza virus that has an HA protein to which most of the population lacks immunity acquires the ability to be efficiently transmitted from person to person.The HA protein has multiple roles in the virus life cycle, notably receptor binding and membrane fusion. The protein is synthesized as a single precursor protein, HA0, that trimerizes and becomes glycosylated in the endoplasmic reticulum as it traffics to the cell surface (33). The HA protein contains multiple disulfide bonds and is cleaved into a mature form consisting of two subunits, HA1 and HA2 (9, 18). HA2 and the N- and C-terminal portions of HA1 form a membrane-proximal stalk that mediates membrane fusion during viral entry (40). A receptor-binding domain (HA-RBD) forms the distal head of the molecule and is inserted into the HA1 subunit. During virus entry, the HA-RBD engages sialic acid-containing receptors on the surface of the host cell, and the virion is subsequently internalized by endocytosis (33). Structurally and functionally, the HA-RBD is a member of the lectin superfamily, and the specificity of the binding pocket contributes to the host range of influenza viruses. For example, α(2,6)-containing sialosides are typically preferred by the HA protein from human viruses and α(2,3) sialosides by the HA proteins from avian viruses (13, 28). Upon triggering by the low-pH environment of endosomes, the HA protein undergoes an irreversible conformational change (6, 40) during which the intact HA-RBDs dissociate from the stalk of the trimer (3, 14, 19, 21). This observation, together with the manner in which the lectin-like domain is inserted as a folded module into the full-length HA protein, led us to hypothesize that the HA-RBD is able to adopt its native structure in isolation. Proper folding of the isolated HA-RBD into its native immunogenic structure has important therapeutic implications because the domain contains all of the known HA antigenic epitopes responsible for antibody recognition (5), and producing a protein-based influenza vaccine composed of isolated HA-RBD would dramatically speed up vaccine development during the early stages of a pandemic.In a recently published report, a construct of the 2009 pandemic H1N1 HA protein that encompasses the HA-RBD, designated HA63-286-RBD, was expressed in Escherichia coli as inclusion bodies, refolded and purified, and used as a vaccine to produce immunity in ferrets (2). In this report, we show that this construct behaves as a stable, structured protein in solution, can be readily crystallized, and indeed adopts a structure that is virtually indistinguishable from that in the H1N1 HA protein ectodomain (41).  相似文献   

15.
16.
17.
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   

18.
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   

19.
The events and mechanisms that lead to interspecies transmission of, and host adaptation to, influenza A virus are unknown; however, both surface and internal proteins have been implicated. Our previous report highlighted the role that Japanese quail play as an intermediate host, expanding the host range of a mallard H2N2 virus, A/mallard/Potsdam/178-4/83 (H2N2), through viral adaptation. This quail-adapted virus supported transmission in quail and increased its host range to replicate and be transmitted efficiently in chickens. Here we report that of the six amino acid changes in the quail-adapted virus, a single change in the hemagglutinin (HA) was crucial for transmission in quail, while the changes in the polymerase genes favored replication at lower temperatures than those for the wild-type mallard virus. Reverse genetic analysis indicated that all adaptive mutations were necessary for transmission in chickens, further implicating quail in extending this virus to terrestrial poultry. Adaptation of the quail-adapted virus in chickens resulted in the alteration of viral tropism from intestinal shedding to shedding and transmission via the respiratory tract. Sequence analysis indicated that this chicken-adapted virus maintained all quail-adaptive mutations, as well as an additional change in the HA and, most notably, a 27-amino-acid deletion in the stalk region of neuraminidase (NA), a genotypic marker of influenza virus adaptation to chickens. This stalk deletion was shown to be responsible for the change in virus tropism from the intestine to the respiratory tract.Of the 16 known hemagglutinin (HA) subtypes, only 3 (H1, H2, and H3) have established stable lineages in humans. The H2N2 virus caused a pandemic in 1957 and circulated in the human population until reassortment of the H2N2 virus with an avian H3 virus resulted in the H3N2 pandemic of 1968 (36). Since then, H2N2 viruses have been absent from the human population; however, the H2 subtype has been repeatedly isolated in wild-bird surveillance, and its HA has been found to be antigenically similar to the H2 pandemic virus HA (23, 25, 36). An H2 influenza virus containing human-like receptor specificity was recently isolated as an H2N3 avian-swine reassortant. This virus caused disease and was transmitted in swine and ferrets (24), indicating that this subtype continues to circulate and mutate and can cross the species barrier to mammals. The repeat introduction of a novel H1N1 pandemic this past year (12, 37) highlights the need to understand the mechanisms of introduction, adaptation, and transmission of avian H2N2 influenza viruses in terrestrial birds and potentially mammalian species.Our previous study built on reports that Japanese quail (Coturnix coturnix) play an important role as an intermediate host in the adaptation of avian influenza viruses to land-based birds (38). Japanese quail are typically more susceptible to aquatic influenza viruses than other terrestrial poultry. These viruses establish infection in the respiratory tract, and shedding occurs via aerosol (2, 19, 26, 34, 38, 43). Quail have been implicated in the transmission of avian influenza viruses, such as H5N1 and H9N2 viruses, which have crossed the species barrier to infect humans (9, 14, 15, 22, 28). The susceptibility of quail to multiple subtypes and their role in interspecies transmission led to their removal from live-bird markets in Hong Kong in 2000; however, they continue to be an integral part of live-bird markets throughout the world. Their role as potential intermediate hosts requires further study to identify important molecular markers in the adaptation via quail of avian viruses to other terrestrial poultry, and possibly to humans.The molecular determinants of the host range and pathogenesis of influenza A viruses have been linked to multiple regions of the 11 genes, most notably those encoding the viral surface glycoproteins (HA and neuraminidase [NA]) and the polymerase proteins (PB2, PB1, PA, and NP). However, a comprehensive map of the various determinants remains incomplete, and the molecular mechanisms involved are unclear. In our previous report, we demonstrated that through the use of quail as an intermediate host, a mallard H2N2 influenza virus, A/mallard/Potsdam/178-4/83 (mall/178), which in its wild-type (wt) form was unable to be transmitted in quail or to establish an efficient infection in chickens, was able, in its adapted form (qa-mall/178), not only to be transmitted to sentinel quail but also to replicate to efficient levels in the chicken intestinal tract and to be transmitted to sentinel cagemates via the fecal-oral route. This adaptation was the result of six serial passages of lung homogenates in quail that led to six amino acid changes in four genes (38). Here we present data confirming the role that Japanese quail play in the transmission of this mall/178 H2N2 virus in land-based birds. Reverse genetics studies confirmed that the amino acid changes produced during the adaptation in quail were necessary for the infection of chickens with this virus and for its transmission in chickens. Further adaptation of the qa-mall/178 H2N2 virus in chickens, aimed at establishing replication in the respiratory tract, resulted in a deletion in the stalk region of the NA, which supported replication in the chicken trachea and lung. This 27-amino-acid deletion in the stalk region of the N2 NA is characteristic of the adaptation of aquatic influenza viruses to domestic poultry, particularly chickens (3, 5, 29). Our work indicates that through the use of quail as an intermediate host, this mallard H2N2 virus is able to further adapt within an additional terrestrial poultry species, potentially improving its chances of expanding its host range further.  相似文献   

20.
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   

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