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 46 毫秒
1.
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   

2.
3.
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   

4.
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   

5.
Highly pathogenic A/H5N1 avian influenza (HPAI H5N1) viruses have seriously affected the Nigerian poultry industry since early 2006. Previous studies have identified multiple introductions of the virus into Nigeria and several reassortment events between cocirculating lineages. To determine the spatial, evolutionary, and population dynamics of the multiple H5N1 lineages cocirculating in Nigeria, we conducted a phylogenetic analysis of whole-genome sequences from 106 HPAI H5N1 viruses isolated between 2006 and 2008 and representing all 25 Nigerian states and the Federal Capital Territory (FCT) reporting outbreaks. We identified a major new subclade in Nigeria that is phylogenetically distinguishable from all previously identified sublineages, as well as two novel reassortment events. A detailed analysis of viral phylogeography identified two major source populations for the HPAI H5N1 virus in Nigeria, one in a major commercial poultry area (southwest region) and one in northern Nigeria, where contact between wild birds and backyard poultry is frequent. These findings suggested that migratory birds from Eastern Europe or Russia may serve an important role in the introduction of HPAI H5N1 viruses into Nigeria, although virus spread through the movement of poultry and poultry products cannot be excluded. Our study provides new insight into the genesis and evolution of H5N1 influenza viruses in Nigeria and has important implications for targeting surveillance efforts to rapidly identify the spread of the virus into and within Nigeria.Since its emergence in 1996 in Guangdong, China, highly pathogenic avian influenza virus of the H5N1 subtype (HPAI H5N1 virus) has disseminated widely across Asia, Europe, and Africa, infecting a range of domestic and wild avian species and sporadically spilling over into humans and other mammals (4, 35). Over time, the HPAI H5N1 virus has diversified into multiple phylogenetically distinct lineages, classified as clades 0 to 9 according to the unified nomenclature system (33). The H5N1 lineage currently circulating in central Asia, the Middle East, Europe, and Africa is referred to as clade 2.2 (33) and has also been described as “EMA” or Qinghai-like in previous publications (4, 17, 27). This clade originated in April 2005 during a large outbreak of a phylogenetically distinct H5N1 virus among wild bird populations at Qinghai Lake in western China (4, 17) and rapidly spread west through central Asia and Europe, eventually reaching Africa in 2006 (27). Clade 2.2 has further diversified, forming the genetic third-order clade 2.2.1 (32) and three genetically distinct sublineages (I, II, and III) (2, 19, 28), all of which are found in Africa.Since 2006 HPAI H5N1 viruses belonging to clade 2.2 have disseminated across multiple countries in western, eastern, and northern Africa: Egypt, Niger, Cameroon, Sudan, Burkina Faso, Djibouti, Ivory Coast, Ghana, Togo, Benin, and Nigeria (2). With a large poultry industry, estimated at 140 million birds (11), Nigeria has experienced several major outbreaks of HPAI H5N1 virus, posing a serious threat to food security and public health in Africa. The first case of HPAI H5N1 virus in Nigeria (sublineage I) occurred in January 2006 in the state of Kaduna, and the virus subsequently was detected in Ghana, Burkina Faso, Ivory Coast, and Sudan (2). In February 2006 sublineage II was reported in Nigeria, and it disseminated widely across the country during 2006 and 2007, also appearing in Togo (2). Clade 2.2.1, which has been prevalent in Egypt, Israel, and the Gaza Strip from 2006 to 2008, was also detected in Nigeria in 2006 (10).By the end of 2007, outbreaks of HPAI H5N1 virus in Nigeria appeared to have been successfully controlled by measures such as “stamping out with compensation,” restrictions on movement of poultry, and enhanced surveillance (13). However, in July 2008 new cases of HPAI H5N1 from a sublineage never previously detected in Africa (sublineage III) were registered in the Nigerian states of Kano and Katsina and in live bird markets in Gombe and Kebbi states (13, 21). Hence, Nigeria is the only African country where viruses belonging to clade 2.2.1 and to three different sublineages (I, II, and III) of clade 2.2 have all been detected. At least three different reassortment events between sublineages have been documented in Nigeria. Salzberg et al. identified the first reassortant strain (which we refer to as “R1”), in which four genome segments (hemagglutinin [HA], NP, NS, and PB1) belong to sublineage I and the other four segments (NA, MP, PA, and PB2) are derived from sublineage II (27). Subsequently, phylogenetic analysis showed that a 2007 reassortant strain (which we refer to as “R3”) contained the HA and NS segments from sublineage I and the other six segments from sublineage II (19, 22). Another reassortant virus (which we refer to as “R5”) contained only the NS gene segment from sublineage I, while the other seven segments were derived from sublineage II (22).Although the genetic diversity of the Nigerian HPAI H5N1 virus population has been well characterized, including multiple introductions of the virus into Nigeria and several reassortment events, little is known about the evolutionary and population growth dynamics of the virus within Nigeria. Particularly understudied are the spatial movements of individual sublineages among Nigeria''s vast poultry population. To explore the spatial, evolutionary, and population dynamics of the multiple H5N1 lineages cocirculating in Nigeria, we conducted a phylogenetic analysis of whole-genome sequences from 106 HPAI H5N1 viruses isolated between 2006 and 2008 and representing all 25 Nigerian states and the Federal Capital Territory (FCT) reporting outbreaks. Using the exact date and location of collection for each viral isolate, we inferred from their phylogenetic relationships the directionality of viral gene flow among Nigerian states and identified critical regions that are likely to serve as key sources for the H5N1 virus in Nigeria.  相似文献   

6.
7.
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   

8.
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   

9.
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   

10.
11.
12.
13.
14.
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   

15.
16.
The continuous circulation of the highly pathogenic avian influenza (HPAI) H5N1 virus has been a cause of great concern. The possibility of this virus acquiring specificity for the human influenza A virus receptor, α2,6-linked sialic acids (SA), and being able to transmit efficiently among humans is a constant threat to human health. Different studies have described amino acid substitutions in hemagglutinin (HA) of clinical HPAI H5N1 isolates or that were introduced experimentally that resulted in an increased, but not exclusive, binding of these virus strains to α2,6-linked SA. We introduced all previously described amino acid substitutions and combinations thereof into a single genetic background, influenza virus A/Indonesia/5/05 HA, and tested the receptor specificity of these 27 mutant viruses. The attachment pattern to ferret and human tissues of the upper and lower respiratory tract of viruses with α2,6-linked SA receptor preference was then determined and compared to the attachment pattern of a human influenza A virus (H3N2). At least three mutant viruses showed an attachment pattern to the human respiratory tract similar to that of the human H3N2 virus. Next, the replication efficiencies of these mutant viruses and the effects of three different neuraminidases on virus replication were determined. These data show that influenza virus A/Indonesia/5/05 potentially requires only a single amino acid substitution to acquire human receptor specificity, while at the same time remaining replication competent, thus suggesting that the pandemic threat posed by HPAI H5N1 is far from diminished.Influenza A virus is a negative-strand RNA virus with a segmented genome within the family of Orthomyxoviridae. Influenza A viruses are divided into subtypes based on the surface glycoproteins hemagglutinin (HA) and neuraminidase (NA). Currently, 16 subtypes of HA and 9 subtypes of NA have been identified in the natural reservoir of all influenza A viruses, wild aquatic birds (24). Occasionally, viruses from this reservoir cross the species barrier into mammals, including humans. When animal influenza viruses are introduced in humans, the spread of the virus is generally limited but may on occasion result in sustained human-to-human transmission. Three influenza A virus subtypes originating from the wild bird reservoir—H1, H2, and H3—have formed stable lineages in humans, starting off with a pandemic and subsequently causing yearly influenza epidemics. In the 20th century, three such pandemics have occurred, in 1918 (H1N1), 1957 (H2N2), and 1968 (H3N2). In 2009, the swine-origin H1N1 virus caused the first influenza pandemic of the 21st century (23).Efficient human-to-human transmission is a prerequisite for any influenza A virus to become pandemic. Currently, the determinants of efficient human-to-human transmission are not completely understood. However, it is believed that a switch of receptor specificity from α2,3-linked sialic acids (SA), used by avian influenza A viruses, to α2,6-linked SA, used by human influenza viruses, is essential (6, 17, 31). It has been shown that the difference in receptor use between avian and human influenza A viruses combined with the distribution of the avian and human virus receptors in the human respiratory tract results in a different localization of virus attachment (26, 33-35). Human viruses attach more abundantly to the upper respiratory tract and trachea, whereas avian viruses predominantly attach to the lower respiratory tract (5, 33-35). Theoretically, the increased presence of virus in the upper respiratory tract, due to the specificity of human influenza A viruses for α2,6-linked SA, could facilitate efficient transmission.Since 1997, highly pathogenic avian influenza (HPAI) H5N1 virus has been circulating in Southeast Asia and has spread westward to Europe, the Middle East, and Africa, resulting in outbreaks of HPAI H5N1 virus in poultry and wild birds and sporadic human cases of infection in 15 different countries (38). The widespread, continuous circulation of the HPAI H5N1 strain has spiked fears that it may acquire specificity for α2,6-linked SA, potentially resulting in a pandemic. Given the currently high case fatality rate of HPAI H5N1 virus infection in humans of ca. 60%, the effect of such a pandemic on the human population could be devastating. In recent years, several amino acid substitutions in HA of HPAI H5N1 viruses have been described, either in virus isolates from patients or introduced experimentally, that increased the binding of the HPAI H5N1 HA to α2,6-linked SA (1, 2, 10, 14, 16, 29, 39, 40). However, none of the described substitutions conferred a full switch of receptor specificity from α2,3-linked SA to α2,6-linked SA and the substitutions were described in virus strains of different geographical origins. Furthermore, it is unknown whether these substitutions led to increased attachment of the virus to cells of the upper respiratory tract, the primary site of replication of human influenza A viruses.Here, we have introduced all of the 21 previously described amino acid substitutions or combinations thereof that changed the receptor specificity of HPAI H5N1 virus strains and six additional combinations not previously described, into HA of influenza virus A/Indonesia/5/05 (IND05). Indonesia is the country that has the highest cumulative number of human cases of HPAI H5N1 virus infection (38). The receptor specificity of 27 mutant H5N1 viruses was determined and the attachment pattern of a subset of these viruses to tissues of the respiratory tract of ferret and human was determined and compared to the attachment pattern of human influenza A virus (H3N2). Subsequently, the role of NA in efficient replication of these mutant viruses was investigated. The data presented here show that receptor specificity of HA of the IND05 virus can be changed by introducing a single amino acid substitution in the receptor-binding domain, resulting in replication competent viruses that attach abundantly to the human upper respiratory tract.  相似文献   

17.
In this study, we show that the highly pathogenic H5N1 avian influenza virus (AIV) (A/crow/Kyoto/53/04 and A/chicken/Egypt/CL6/07) induced apoptosis in duck embryonic fibroblasts (DEF). In contrast, apoptosis was reduced among cells infected with low-pathogenic AIVs (A/duck/HK/342/78 [H5N2], A/duck/HK/820/80 [H5N3], A/wigeon/Osaka/1/01 [H7N7], and A/turkey/Wisconsin/1/66 [H9N2]). Thus, we investigated the molecular mechanisms of apoptosis induced by H5N1-AIV infection. Caspase-dependent and -independent pathways contributed to the cytopathic effects. We further showed that, in the induction of apoptosis, the hemagglutinin of H5N1-AIV played a major role and its cleavage sequence was not critical. We also observed outer membrane permeabilization and loss of the transmembrane potential of the mitochondria of infected DEF, indicating that mitochondrial dysfunction was caused by the H5N1-AIV infection. We then analyzed Ca2+ dynamics in the infected cells and demonstrated an increase in the concentration of Ca2+ in the cytosol ([Ca2+]i) and mitochondria ([Ca2+]m) after H5N1-AIV infection. Regardless, gene expression important for regulating Ca2+ efflux from the endoplasmic reticulum did not significantly change after H5N1-AIV infection. These results suggest that extracellular Ca2+ may enter H5N1-AIV-infected cells. Indeed, EGTA, which chelates extracellular free Ca2+, significantly reduced the [Ca2+]i, [Ca2+]m, and apoptosis induced by H5N1-AIV infection. In conclusion, we identified a novel mechanism for influenza A virus-mediated cell death, which involved the acceleration of extracellular Ca2+ influx, leading to mitochondrial dysfunction and apoptosis. These findings may be useful for understanding the pathogenesis of H5N1-AIV in avian species as well as the impact of Ca2+ homeostasis on influenza A virus infection.Avian influenza viruses (AIVs) are classified as highly or low-pathogenic AIVs (HPAIVs or LPAIVs, respectively) based on their pathogenicity in chickens (1). HPAIVs cause systemic infections and high mortality in chickens (28), whereas poultry are asymptomatic or develop mild respiratory problems and/or intestinal illness after LPAIV infection (49). Hemagglutinin (HA) cleavability is a critical determinant of AIV pathogenicity in avian species (61). Other determinants, such as nonstructural (NS) protein and neuraminidase (NA) protein, reportedly regulate the virulence of AIVs (9, 29, 44). However, waterfowl, known as the natural host for AIVs, do not usually have any symptoms during an HPAIV infection (21), whereas they show neurologic symptoms and death after infection with some of the recently emerged HPAIVs, such as the Asian H5N1 virus (11, 46, 62). Thus, the entire mechanism responsible for the pathogenicity of the AIVs is not yet known. Unknown cellular and viral factors probably underlie the pathogenesis of HPAIVs in avian species, especially waterfowl.The alveolar epithelial cells (66) or vascular endothelial cells (32) of human patients and chickens infected by H5N1-AIV show apoptosis. Other reports suggest that apoptosis of these cells is essential for the development of acute lung injury in mice and acute respiratory distress syndrome in humans (39), which is often observed in H5N1-AIV-infected patients. Therefore, it is necessary to evaluate whether apoptosis is critical for the pathogenesis of H5N1-AIV in vivo and to understand the molecular mechanisms of the apoptotic cell death induced by H5N1-AIV infection.Ca2+ is a key regulator of cell survival, and the breakdown of Ca2+ homeostasis, due to sustained elevations in Ca2+ inside cells, triggers programmed cell death involving apoptosis (24). Indeed, disruption of Ca2+ homeostasis plays a key role in apoptosis during the pathogenic process of several types of viral infections, including those with human immunodeficiency virus (HIV), hepatitis C virus, and human T-cell leukemia virus type 1 (3, 4, 31, 57). In addition, the HIV gp120 envelope protein induces neuronal cell death through Ca2+ dysregulation, even in the absence of viral particles (25).In this study, we used duck embryonic fibroblasts (DEF) to elucidate the molecular mechanisms of the apoptotic cell death induced by H5N1-AIV. We show here that H5N1-AIV infection triggered extracellular Ca2+ influx and that this alteration in the concentration of Ca2+ inside the cells subsequently induced mitochondrial dysfunction and led to apoptotic cell death. In addition, we demonstrate that H5N1-HA was a critical viral factor for inducing apoptosis.  相似文献   

18.
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   

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

20.
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