Next Article in Journal
Advances in Alphavirus and Flavivirus Research
Previous Article in Journal
A qPCR Assay for the Quantification of Selected Genotypic Variants of Spodoptera frugiperda Multiple Nucleopolyhedrovirus (Baculoviridae)
Previous Article in Special Issue
Pandemic Risk Assessment for Swine Influenza A Virus in Comparative In Vitro and In Vivo Models
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Markers and Mechanisms of Influenza A Virus Cross-Species Transmission and New Host Adaptation

1
School of Public Health (Shenzhen), Shenzhen Campus of Sun Yat-sen University, Shenzhen 518107, China
2
Guangzhou Eighth People’s Hospital, Guangzhou Medical University, Guangzhou 510440, China
3
Zhongshan School of Medicine, Sun Yat-sen University, Guangzhou 510080, China
4
College of Wildlife and Protected Area, Northeast Forestry University, Harbin 150040, China
5
Guangdong Provincial Highly Pathogenic Microorganism Science Data Center, Guangzhou 510080, China
6
Shenzhen Key Laboratory of Pathogenic Microbes and Biosafety, Shenzhen 518107, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2024, 16(6), 883; https://doi.org/10.3390/v16060883
Submission received: 16 April 2024 / Revised: 25 May 2024 / Accepted: 28 May 2024 / Published: 30 May 2024
(This article belongs to the Special Issue Animal and Human Respiratory Viruses—Causes of the Next Pandemic?)

Abstract

:
Influenza A viruses continue to be a serious health risk to people and result in a large-scale socio-economic loss. Avian influenza viruses typically do not replicate efficiently in mammals, but through the accumulation of mutations or genetic reassortment, they can overcome interspecies barriers, adapt to new hosts, and spread among them. Zoonotic influenza A viruses sporadically infect humans and exhibit limited human-to-human transmission. However, further adaptation of these viruses to humans may result in airborne transmissible viruses with pandemic potential. Therefore, we are beginning to understand genetic changes and mechanisms that may influence interspecific adaptation, cross-species transmission, and the pandemic potential of influenza A viruses. We also discuss the genetic and phenotypic traits associated with the airborne transmission of influenza A viruses in order to provide theoretical guidance for the surveillance of new strains with pandemic potential and the prevention of pandemics.

1. Introduction

Influenza viruses are segmented, capsular, negative-sense RNA viruses that are members of the Orthomyxoviridae family. They can be categorized into four groups, A, B, C, and D, according to the antigenic variations in nucleoprotein (NP) and matrix protein (M). Only influenza A viruses are capable of causing viral pandemics [1]. Influenza A viruses are divided into different subtypes based on the antigenicity of their hemagglutinin (HA) and neuraminidase (NA) membrane glycoproteins [2,3]. Wild aquatic birds are natural hosts of influenza A viruses [4]. However, through reassortment or adaptive mutations, some subtypes can overcome the host barrier and spread to poultry and mammals, including people, leading to occasional illnesses, epidemics, and pandemics [5].
Influenza A viruses continue to endanger human health throughout the world and have resulted in significant socio-economic losses. There are approximately 1 billion cases of seasonal influenza that occur annually, with 3–5 million of those cases being severe [6]. An influenza A virus is also a zoonotic pathogen with a multitude of hosts, allowing for interspecies transmission (birds → other mammals/humans), adaptation of new hosts, and the emergence of novel pandemics [7]. Properties that promote interspecies adaptation of viruses include extracellular properties (the stability of virions and the HA protein, virion morphology, balance of HA binding, and NA receptor-destroying activities) and intracellular properties (receptor-binding specificity by the HA protein, HA stability, polymerase efficiency, and interferon antagonism) [8]. Currently, there have been reports of cross-species transmission of avian influenza viruses such as H3N8, H5N6, H5N8, H10N3, and H7N4 [9,10,11,12,13]. The viruses have been observed in mammals by breaking the host barrier through multi-site adaptive changes and reassortment. “Antigenic drift” (accumulation of substitutions in the HA and NA proteins) and “antigenic shift” (recombination of gene segments) are responsible for the phenotypic diversity of influenza A viruses [14,15]. Four global pandemics have been triggered by influenza viruses of the H1, H2, and H3 subtypes in the last 100 years [16]. Unlike seasonal influenza, pandemic viruses result from antigenic switching and are unpredictable [17]. Pre-existing host antibodies offer little protection against these new viruses [18]. Furthermore, animals such as ferrets and pigs, which have both avian and human receptors, play a role as a mixing vessel for the reassortment and generation of the new influenza A virus’s phenotypes [19]. The 2009 pandemic H1N1 (pdmH1N1) was caused by a novel swine-origin recombinant virus, which contained gene segments from three different swine flu lineages [20]. The ability of zoonotic influenza viruses to become airborne should probably be considered an important determinant in crossing the barrier of animal-to-human transmission [21]. Airborne transmission is also the main mode of influenza virus transmission between humans [22]. It has been determined that all influenza viruses responsible for the four recorded pandemics had airborne transmission. This implies that acquiring airborne transmission ability is essential for influenza viruses to adapt to human hosts and have pandemic potential [22]. Consequently, we summarize genetic changes and mechanisms that may influence interspecific adaptation, cross-species transmission, and pandemic potential of influenza A viruses and discuss the phenotypic traits associated with airborne transmission of influenza A viruses. This facilitates the monitoring of the virus and the conduct of risk assessments, as well as the prevention of outbreaks of new pandemic influenza viruses.

2. Transmission Modes of the Influenza A Virus

Influenza viruses transmit through multiple modes, including contact (either direct or through a contaminated surface) and inhalation of expelled respiratory droplets or aerosols, the latter two being collectively referred to as airborne [22,23]. In previous studies, droplet diameter was frequently employed as a demarcation between droplet and aerosol transmission, most commonly 5 μm [24]. Droplet transmission (>5 μm) and aerosol transmission (≤5 μm) differ in terms of generation mode, aerodynamics, infectivity, and infection route. Droplets are usually produced by the patient through coughing, sneezing, or speaking, and due to gravity, have a restricted transmission distance and are suspended in the air for a brief period of time [25]. However, viruses are very efficiently spread through aerosols by the patient’s breathing alone [26]. Small aerosols are also more likely to penetrate deep into the alveoli and cause dramatic host responses [26,27]. In addition to the respiratory tract, the conjunctiva and digestive tract are also important routes of aerosol infection [28]. Aerosols may be the main mode of virus transmission in particular scenarios, such as poorly ventilated homes, hospital departments, and wards [29]. It is unclear how much each transmission route contributes to the overall viral propagation capacity, and further research is still required through extensive data collection and modeling. Recently, researchers have questioned the demarcation between droplets and aerosols because it is not based on the well-defined physical properties of droplets or their dynamics in a complex physical environment [30,31]. Viral transmissivity is a continuum that depends on numerous factors (gravitational settling rate, transport, and dispersion in a turbulent air jet, viral load and viral shedding, virus inactivation) that cannot be adequately characterized by a single droplet diameter. Therefore, an airborne transmission mode based on the physical properties of exhaled droplets and their interactions with the environment and human behavior can be used to replace the existing artificial dichotomy [30,32].
Both droplets and aerosols can facilitate efficient host-to-host transmission of influenza viruses. Therefore, it is crucial to comprehend the mechanisms of cross-species and airborne transmission of influenza A viruses and to explore the mutational pathways in the evolution of viruses [21]. Several requirements that we believe are important for influenza A viruses to become airborne are as follows: ① Influenza virions can attach to the appropriate cells in the upper respiratory tract and replicate efficiently. ② Virions are relatively stable and remain infectious while transiting between and within hosts. ③ Virions can be efficiently released outside the cell, leaving the upper respiratory tract as single particles and being expelled in large numbers. The requirements and corresponding mechanisms are summarized in Table 1.

3. Molecular Markers and Mechanisms of Influenza A Virus Cross-Species Transmission and New Host Adaptation

3.1. Acquisition of HA Human Receptor Binding Preference

The receptor-binding preference of influenza A viruses is critical for successful cross-species transmission of the virus from animal hosts to humans [33]. Studies have demonstrated that avian influenza viruses terminal sialic acid (SA) residues on glycans that are linked to the penultimate galactose through an α-2,3 linkage (SAα-2,3Gal), while human influenza viruses terminal sialic acid residues on glycans that are linked to the penultimate galactose through an α-2,6 linkage (SAα-2,6Gal) [34]. The expression and distribution of SA receptors in tissues may in part contribute to the host range and interspecific adaptation of virus infections [35]. In most avian species, avian-type receptors (SAα-2,3Gal) dominate in tracheal epithelial cells. In contrast, the human upper respiratory tract is rich in SA with α-2,6-linked carbohydrates and thus might be infected by human influenza viruses [36,37]. Both seasonal and pandemic influenza viruses have a high propensity for attaching to human receptors, according to a retrospective analysis by researchers. This preference contributes to virus attachment and efficient replication in the human upper respiratory tract [38]. The human lower respiratory tract contains SAα-2,3Gal, which could explain why avian influenza viruses are able to infect humans and cause a strong host response [39].
The receptor-binding site (RBS) of HA is formed by the 190 helix, the 130 loop, and the 220 loop [40]. RBS has a pocket-like shape with a sialic acid binding site, and amino acid changes in and around it are essential for the specificity of receptor binding and determining the host range of the virus [41]. The molecular mechanisms by which avian-specific HAs gain human receptor-binding avidity vary among different subtypes [42]. E190D and G225D for H1, G226L and G228S for H2 and H3 (HA residues here follow H3 numbering unless otherwise stated) not only confer α-2,6 binding and infection of humans but also successfully spread between human hosts [43,44,45,46]. The main reason for the shift in receptor binding preference is the widening of the binding pocket due to adaptive mutation [47]. It has been suggested that the G225D substitution induced a significant conformational change in the 220-loop, thereby altering the receptor binding specificity [48]. Similarly, the E190D can also expand the binding pocket by changing the conformation of the 220 loop [49]. The Q226L substitution provides a hydrophobic environment and prevents the avian receptor from binding in the trans conformation that is associated with tight binding [50]. The shift of the 220 loop is maximal when both Q226L and G228S substitutions are present [51]. The human H3N8 virus A/Henan/4-10/2022 haemagglutinin has a degenerative codon in position 228, which could be residue G or S [52]. As previously reported for H5N1 and H7N9 viruses, the novel virus could be quickly adapted to human receptor binding by dynamic substitution of key residues of viral proteins [53,54]. In contrast, the adaptive replacements needed for the H5, H6, and H7 subtypes of avian influenza strains to spread in human hosts remain elusive. Observational studies and reverse genetics studies may assist in the discovery of the corresponding genetic changes [55,56]. The two substitutions L129V and A134V, which were identified in a virus isolated from a human case, can change the HA receptor binding preference of H5N1 viruses from SAα-2,3Gal to SAα-2,3Gal and SAα-2,6Gal [55]. H5N1 virus A/duck/Egypt/10185SS/2010 (dkEgy10, clade 2.2.1) contains a Q226L mutation, while N224K, Q226L, N158D, and an L133a deletion were identified in A/chicken/Vietnam/NCVD-093/2008 (ckViet08, clade 7.2) [56]. In conclusion, the structural changes in the RBS of H5N1 viruses mainly focus on alteration of the length of the 130-loop, alteration of a combination of residue positions in the 130-loop and 220-loop, mutations in the 190-helix, and removal of a glycosylation sequon at position 158 [57]. The S123P substitution in HA increases the specificity of the H5N6 virus for human receptor binding. The I151T alteration and residue 129 deletion increase the virus’s ability to bind to human receptors and infect humans [58]. Similar mutations that confer human receptor binding preference on H1, H2, and H3 subtypes were introduced to H6N1 A/Taiwan/2013 viruses. It was found that the G225D mutation completely switches specificity to the human-type receptor [59]. For the H7 subtype, G186V and Q226L mutations may be important for increasing the human receptor binding preference of viruses [60,61]. The hydrophobic milieu surrounding the 220 loop, engendered by the alterations S138A, G186V, T221P, and Q226L, facilitates the attachment of H7N9 viruses to the α-2,6 sialic acid receptor [62]. Several alternative triamino acid mutations (V186G/K-K193T-G228S or V186N-N224K-G228S) have been shown to switch the receptor specificity of H7N9 HA from avian to human [63]. Furthermore, alterations in glycosylation near HA RBS would influence its affinity to receptors [64]. The loss of glycosylation at Asn91 has been demonstrated to affect human receptor binding of H1N1 A/South Carolina/1/18 (SC18) and a single amino acid D225G mutant of SC18 HA (referred to as NY18) by disrupting the network of inter-residue interactions in the RBS of SC18 and NY18 [65]. The N-glycans on HA may also cause steric hindrance near the HA–receptor binding domain, thereby enhancing the affinity to α2,3 sialic acid [66]. The occurrence of adaptive mutations and altered glycosylation in or near the HA receptor-binding site are significant factors that influence viral receptor binding preferences.
Studies of the receptor binding preference of HA in pandemic viruses (those that are able to achieve airborne transmission) revealed that all viruses exhibited high affinity binding to human receptors [67,68,69]. The precise molecular mechanism by which viruses gain airborne transmission remains unclear. It is hypothesized that respiratory droplets or aerosols, which contain few particles, require cells to expose sufficient high-affinity receptors in order to initiate infection and spread between hosts [69]. The relative binding affinity of HA to human receptors correlates with the transmission efficiency of avian influenza viruses in mammalian models such as ferrets. The experimental results of the transmission of influenza A virus H1 [70], H3 [71,72], H5 [73], H7 [74], and H9 [75] subtypes all support this view (Table 2). The acquisition of human receptor binding preference by the Tianjin H9N2 isolate is facilitated by the mutation G226L and seven glycosylation sites (GSs) in the HA protein. Transmission experiments indicated that the virus is efficiently transmitted between guinea pigs and ferrets through direct contact or an airborne route [75]. The two strains of H10N3 virus (A/chicken/Jiangsu/0104/2019 and A/chicken/Jiangsu/0110/2019), which possess the mutation G228S, exhibit dual receptor binding affinity and are highly mammalian-adapted. They can be transmitted between chickens through direct contact and efficiently between guinea pigs through direct contact and respiratory droplets [76] (Table 2). The shift in GS also modifies the transmissibility of the virus. Abdelwhab et al. (2016) discovered that the removal of glycosylation at sites 154 and 236 resulted in a reduction in viral release and a delay in mortality in exposed birds, reducing transmission of the virus from chicken to chicken [77]. The T160A substitution in the A/Vietnam/1203/2004 (H5N1) virus removes the N-linked glycosylation (NLG) sequence at position 158. The removal of GS works synergistically with Q226L or Q226L/G228S mutations to affect the receptor binding preference of the virus, thereby promoting the cross-species transmission of the virus [78,79]. However, possessing a human receptor binding preference is merely the initial stage for the virus to acquire pandemic potential. It needs to be combined with other adaptive traits to gain airborne capacity [76]. The HA G228S mutation increases the virus’s capacity to attach to human receptors. In conjunction with the PB2 E627K alteration, it facilitates the virus’s effective airborne transmission among ferrets [71]. In summary, mutations in RBS and glycosylation modifications around RBS are important factors affecting the receptor-binding specificity of the influenza A virus. Human receptor binding preferences are also important for viruses’ interspecific adaptation and the acquisition of airborne phenotypes.

3.2. Changes in HA Stability

A critical stage in the infection phase of the influenza virus is the membrane fusion process mediated by HA. An acidic endosomal environment triggers irreversible conformational changes in the HA protein, leading to viral fusion with the endosomal membrane and injecting nucleic acid into the host cells [85,86]. The optimal pH for membrane fusion varies from host to host [87]. The reasons why HA stability affects the interspecific adaptation and transmission of influenza A viruses are as follows: Firstly, increased HA stability boosts the environmental persistence of influenza viruses and prolongs the half-life of virus transmission between hosts in the external environment [88]. This has been proven by influenza virus transmission tunnel (IVTT) experiments [89]. Secondly, HA with moderate acid stability is indispensable to early conformational changes of HA and prevents virus inactivation before entering mammalian cells [8]. The virus has to pass through the human nasal airway epithelium, which is known to be mildly acidic, before entering the human cells [90]. Finally, HA stability affects membrane fusion sites and triggers varying host responses through distinct pathways, which in turn regulate the infectivity and transmissibility of the virus [90,91]. It is worth noting that the fusion pH of human-adapted subtypes is in general lower than that of their avian counterparts [92]. HA with moderate acid stability is necessary to support airborne transmission (the primary mode of human-to-human transmission) of the virus among ferrets [8].
The stability of HA is largely influenced by electrostatic interactions and van der Waals interactions at domains (RBS, HA1, HA2, and fusion peptide pocket) and subunit interfaces (HA1-HA1 interfaces, HA1-HA2 interfaces) [93,94,95,96]. By regulating these interactions, adaptive mutations affect structural reorganization under viral membrane fusion, thereby changing HA stability [97,98]. Amino acid substitutions, including T642H (HA2 numbering position 64, H1 numbering position HA407), V662H (HA2 numbering position 66, H1 numbering position HA409), T642H, and V662H double mutations, were introduced in the HA1/HA2 interaction region of the A/WSN/33 (H1N1) virus. These substitutions alter the pH of the virus-activated HA by affecting molecular interactions at the HA1 and HA2 interface by a pH of 5.4–5.6, like the wild-type strain, while the V662H and double mutant strains have a lower pH of approximately 5.1–5.3 [99]. Similarly, the HA2-E47K mutation diminishes the membrane fusion pH of the 2009 H1N1 pandemic (pdmH1N1) virus from 5.4 to 5.0. This is due to the formation of a salt bridge between the HA1–E21 residue and the HA2-K47 residue, which stabilizes the HA trimer structure [94]. Several reviews have summarized adaptive mutations that affect the stability of HA acid in influenza A viruses of the H1, H3, H5, and H7 subtypes [96,100]. Key amino acid residues and charged residues located in and around RBS, the HA1-HA1 interface, the HA1–HA2 interface, and the fusion peptide region may play a key role in virus adaptation to a new host and interspecific transmission.
Numerous investigations have confirmed the strong relationship between HA acid stability and the transmission capacity of viruses. For example, in mice and ferrets, the swine influenza virus (pH1N1) A/Tennessee/1–560/2009 was generated via reverse genetics to contain the HA1-Y17H mutation (pH 6.0), which reduces pathogenicity and eliminates airborne transmission capacity. After adaptive culture in ferrets, the loss-of-function virus acquires the HA1-H17Y and HA2-R106K restoration mutations. These two mutations decline the activation pH to 5.3 and restore the virus’s airborne capacity [101]. In another study, the A/Perth/16/2009 (H3N2) after cell culture exhibits mutations at positions 78 and 212 of the HA head (G78D and T212I). The virus’s membrane fusion pH increases from 5.5 (wild type) to 5.8 (mutant). The higher replication titers of mutant viruses in cell culture but lower airborne efficiency in a ferret model suggest that the membrane fusion pH of H3N2 viruses is related to the airborne efficiency of viruses [102]. Hu et al. (2020) compared the transmission capacity of two groups of swine H1N1 influenza viruses (gamma-clade) based on their HA acid stability. The study revealed that strains A/swine/Illinois/2A-1213-G15/2013 (G15) and A/swine/Illinois/2B-0314-P4/2014 (P4), which have moderate stability of HA, gain the ability to be airborne among ferrets. In contrast, viruses with a metastable HA protein (A/swine/Illinois/2E-0113-P19/2013 (P19) pH 5.9) outcompeted and lost the airborne phenotype [103] (Table 3). In conclusion, HA acid stability is an important factor in the airborne transmission of viruses. In addition, the virus-host pH feedback pathway indicates that the virus induces alterations in respiratory tract pH during host infection, and viruses with metastable HA are inactivated extracellularly. Subsequently, the virus increases its infectious dose, and the extracellular pH further decreases, leading to continued virus deactivation. The negative feedback pathway suggests that hosts may select virus strains containing HA proteins with optimal pH stability [104]. It is noteworthy that when an HA mutation associated with increased human receptor binding preference affects HA stability, the virus may require additional compensatory mutations to counteract the decreased stability caused by the mutation [105,106]. This proposes that a precise balance of mutations associated with various functions in HA may be necessary for efficient virus transmission in mammals.

3.3. Functional Balance between the Activities of HA Binding and NA Cleavage

HA and NA of influenza viruses recognize the same host molecule SA and exhibit complementary roles during replication [113]. HA has binding activity and is responsible for binding to the sialic acid to allow virus internalization. NA is a sialidase that mediates the hydrolysis of the link between the SA bound to HA and the adjacent sugar. The cleavage activity of NA prevents the aggregation of nascent virus particles at the surface of the infected cell and allows for viral release [114,115]. In addition to the traditional virion release activity, NA has also been proposed to carry a second SIA-binding site (2SBS) and be involved in the viral entry phase [116]. The 2SBS consists of 370 loops (residues 366–373), 400 loops (residues 399–404), and 430 loops (residues 430–433), which contain residues that interact with SIA (S367, S370, S372, N400, W403, and K432 (N2 numbering)) [117]. It was found that the 2SBS enhances the activity of catalytic sites and affects the HA-NA balance [118]. In addition, HA-NA functional balance affects virus attachment. HA binds to abundant sialylated mucins in the mucus layer, and without NA activity, the viruses may be trapped before they reach the site on the epithelial cell membrane [119,120,121]. NA cleavage activity, which is associated with viral motility in the mucus layer, allows the virion to penetrate the sialylated mucus layer and attach to host cells [122,123]. The sialoglycan repertoire of a host, which varies between species, has a complex composition and distribution of decoy and functional receptors [124]. Influenza A virus’s NA receptor-destroying activity and HA receptor-binding affinity need to be balanced with the host receptor repertoire. When influenza A viruses cross the species barrier, the HA-NA functional balance needs to be restored for optimal viral adaptation [125]. Therefore, the balance between HA binding and NA cleavage activity plays an important role in overcoming host barriers and adapting to new hosts [113]. The optimal functional balance allows viruses to penetrate the sialylated mucus layer and attach to cells, and they can be released from cells after assembly [126].
Any alteration in HA or in NA may cause a modification of the functional HA-NA balance (including mutations affecting HA and NA activity, the addition or removal of glycosylation at key locations, the length of the NA stalk, virus particle morphology, and the NA receptor binding) [118,127,128,129,130]. The HA D222G/E/N substitution increases the binding intensity of influenza A (H1N1) pdm09 virus SAα-2,6, but does not confer the corresponding NA activity, thus disrupting the HA-NA balance. This may be the reason why D222G/E/N viruses are less adaptable to human hosts and cannot transmit efficiently [127]. Mutations affecting the virus particle morphology have been mainly mapped to matrix protein 1 (M1), and these mutations may be related to HA-NA balance and viral transmission ability [128]. Kong et al. found that the M1 D156E substitution increases the proportion of filamentous morphology while decreasing the replicative capacity of the H7N9 A/Anhui/1/2013 (AH/1) virus in cells, thereby abolishing its transmissibility among guinea pigs [131]. The M1 P41A mutation in A/swine/Spain/53207/04 (H1N1) (SPN04) virus reduced the number and length of filamentous morphology and decreased the NA activity of the virus. The enhanced transmissibility of the SPN04 virus may be related to the adjustment of the HA-NA balance [132]. In addition, longer filaments would be able to extend directly through the mucous layer, facilitating airborne transmission of the virus [128]. NLG of HA has been reported to affect its receptor binding and immune response, thereby contributing to the immune escape and virulence of influenza A viruses [66,133]. The NLG of NA impacts its structure, activity, specificity, and thermostability. It was discovered that H5N6 viruses lose a potential n-linked glycosylation site at position 154 and gain an NA mutation at V202I, which affects the balance of HA-NA function, replication, stability, and pathogenicity of the viruses [134]. In addition, it has been found that loss of NLG attenuates viral budding and replication [135,136]. The addition or removal of NLG from the H7 HA head disrupts the HA-NA balance in the viruses, leading to a reduction in viral fitness. This necessitates corresponding changes in N9 NA to restore the balance. Influenza viruses with functional balance can achieve higher transmission efficiencies [129]. Due to the opposite effects of HA and NA in an infection cycle, simultaneous changes in glycosylation patterns should be regarded as a characteristic predicting future pandemic outbreaks of viruses [137]. A shortened NA stalk mainly exists in H5 and H7 subtypes of avian influenza viruses and is also an important factor in the adaptation of waterfowl influenza viruses to poultry [130,138]. The NA stalk length can affect the NA activity of H9N2 G1 lineage virus (8-amino acid-deficient mutants), thereby balancing their HA and NA functions and regulating the entry of viruses into host cells. This effect is species-specific and influences the host range of the virus [139]. It has been demonstrated that the highly pathogenic avian influenza (HPAI) H5N1 viruses with a short stalk exhibit reduced respiratory droplet transmission between ferrets, possibly due to the diminished capacity of viruses with a short-stalk NA to penetrate mucus and deaggregate virions [140]. The 2SBS is an important determinant of the HA-NA-receptor balance. Its effect on HA-NA balance and viral replication depends on the receptor-binding properties of HA [141]. The 2SBS is conserved in most avian virus subtypes but is lost in mammalian variants [118]. Mutations in 2SBS and HA in avian influenza viruses may become human pandemic influenza viruses by adjusting the HA-NA balance [116]. On the one hand, mutations in HA precede and drive mutations in NA that allow optimal viral replication and spread in humans. On the other hand, substitutions in the 2SBS of NA may precede or drive compensatory mutations in HA, decreasing receptor binding to avian-type receptors and increasing human receptor binding affinity [142,143]. In pandemic influenza viruses, the rapid loss of 2SBS binding function may help the virus restore HA-NA stability in a new host [144].
One of the most significant features of airborne influenza A viruses is the high level of individual particle shedding. It is directly affected by the functional balance of HA-NA [22]. Although the mechanism by which the virus acquires an airborne phenotype remains unclear, numerous studies have confirmed the significance of HA-NA functional balance in relation to airborne transmission [69,145,146]. The HA R149K substitution enhances the human receptor binding affinity of triple reassortant swine (TRsw) H1N1 influenza virus A/swine/North Carolina/18161/2002 (NC/02) and its ability to bind to the nasal turbinates of ferrets, thereby promoting contact transmission of the virus. Nevertheless, an airborne phenotype for NC/02 was not established until the introduction of the NA and M segments of the H1N1 pdm09 virus, which yielded NA that matched the highly active HA [145]. Furthermore, the HA-NA balance is crucial for the effective and long-lasting spread of viruses from person to person. The TRsw-like A/swine/Hong Kong/915/04 (sw915) (H1N2) virus exhibits comparable receptor binding specificity and affinity to the pandemic H1N1 virus A/HK/415742/09 (HK415742), with variation only in the NA segment of their internal genes. Introducing the NA segment from HK415742 into sw915 results in balanced HA-NA activity and a developed respiratory droplet transmission of the virus [69]. However, although the A/Iowa/CEID23/2005 (Iowa05) (H1N1) virus has successfully crossed the host barrier to infect humans, the mismatch in HA-NA activity has made it difficult for the virus to spread from person to person and cause a pandemic [146]. In conclusion, the functional balance of HA-NA has been demonstrated to be closely associated with the regulation of host range, cross-species transmission, and airborne transmission of viruses. Further research is warranted to investigate the mechanism by which the functional balance of HA-NA directly impacts the airborne transmission of viruses.

3.4. Variations in the Activity of RNA-Dependent RNA Polymerase

RNA-dependent RNA polymerase (RdRp) is a heterotrimeric complex consisting of polymerase basic protein 1 (PB1), polymerase basic protein 2 (PB2), and polymerase acidic protein (PA) [147]. PB1 is the core subunit of RNA synthesis. It facilitates the assembly and synthesis of RdRp and catalyzes RNA polymerization. PB2 binds to PB1 and NP, assisting in the assembly of the viral ribonucleoprotein (vRNP) and capturing the 5′ cap of the nascent host-capped RNA during transcription. It adopts a different conformation to maintain the replication and transcription steps. PA binds to PB1 and plays an important role in endonuclease activity during transcription. All subunits contribute to the formation of entry and exit channels for template RNA, nucleoside triphosphates (NTPs), and products [148,149]. Avian influenza viruses require adaptive mutations or reassortment to enlarge polymerase activity to adapt to human or other mammalian hosts. The PB2 subunit contains well-known mammalian adaptation sites, such as the E627K and D701N mutations, which strengthen viral polymerase activity, replication, pathogenicity, and transmissibility [150,151]. The temperature-dependent growth of the influenza A virus is related to the virus host range [152]. Avian influenza A virus is known to replicate in the intestinal tract of infected birds at a temperature close to 42 °C. The PB2 E627K mutation allows the virus to adapt to the human host and replicate at higher levels in the human respiratory tract (33 °C) [153]. The host factor acidic leucine-rich nuclear phosphoprotein 32 kDa (ANP32) protein displays differential capacity to support viral polymerase activity across different species [154]. Mammals only express the shorter ANP32 protein, which does not efficiently support avian polymerase. However, the presence of PB2 E627K can compensate for the impaired interactions between different host polymerases [155,156]. In addition, it has been demonstrated that the adaptation of the viral polymerase to importin-α plays an important role in interspecies transmission of influenza virus [157]. Although PB2 E627K and D701N mutations caused the same shift to importin-α7 specificity, the adaptation mechanisms are different. The D701N mutation promoted binding of PB2 to importin-α1 and exposure of a nuclear localization signal of PB2, thereby enhancing importin-α mediated nuclear transport in mammalian cells. However, the E627K mutation only enhances the binding of importin-α without affecting the nuclear entry of PB2 [157,158]. Furthermore, changes to the residues in the PA and PB1 subunits have a comparable impact on the virulence, reproduction, and transmission of viruses. For example, by boosting viral polymerase activity, the PA A343S/D347E mutation markedly boosted the virulence of the H5N1 virus in mice and thus elevated the risk of infection in people [159]. Similarly, the PA K356R alteration increases the H9N2 virus’s polymerase activity, which raises viral transcription and export levels in human A549 cells and results in a serious infection in mice [160] (Table 4). Additionally, the PB1 K577E mutation enhances the polymerase activity of the H9N2 virus in human cells and boosts its pathogenicity in mice [161].
Numerous investigations have shown that increased polymerase activity is associated with the acquisition of an airborne phenotype of the virus [162,163]. Effective replication of upper respiratory viruses mediated by polymerase is essential for the airborne transmission of avian influenza viruses among mammals [164]. Experimental data support a strong correlation between the number of influenza viral RNA-containing particles released into the air and airborne transmissibility among ferrets [165]. We speculate that high-titer viral particles are expelled as droplets or aerosols when the donor coughs, sneezes, or even breathes. Once viruses enter the respiratory tract of the recipient host, they replicate efficiently and multiply rapidly, affecting transmissibility and increasing the potential for airborne transmission. Researchers also investigated the impact and corresponding mechanism of polymerase mutation sites on mammalian transmission models [160,162,166]. When replicating within ferrets, recombinant H7N9 viruses with the PB2 E627K and E701N mutations exhibit increased viral polymerase activity and are effectively spread through respiratory droplets [167]. Meng et al. (2022) discovered that four mutations (V100I, N321K, I330V, and A639T) in PA have a synergistic effect on the pathogenicity and airborne transmission of influenza viruses A/swine/Liaoning/265/2017 (LN265) (H1N1). The proposed molecular mechanism suggests that the V100I mutation in PA augments the cleavage activity of the nucleic acid endonuclease. Furthermore, the N321K and I330V mutations develop the binding capacity of vRNA, resulting in the development of the transcriptional efficiency and replication capacity of the virus [168] (Table 4). Similar to Meng et al.’s study, H9N2 avian influenza viruses with the PA K356R mutation demonstrated increased accumulation of PA nuclei, improved viral polymerase activity, and boosted transcriptional replication. The possible molecular mechanisms are as follows: the mutation at site 356, which is in loops 350–355 of the PA-C structural domain, may affect nucleic acid endonuclease activity, protease activity, or interaction with other PA proteins, thereby affecting viral replication and transmission [160]. Less research has been performed recently on the molecular mechanisms of mutation sites in PB1. However, being a fundamental component of RNA synthesis, PB1 is required for viral replication. The wild avian influenza virus A/Mallard/Inner Mongolia/T222/2018 (H3N8) and the virus containing PB1 S524G (rT222-S524G) were saved by reverse genetics. The virus with the PB1 S524G mutation exhibits heightened replication efficiency, increased polymerase activity, and enhanced airborne transmission among ferrets [163]. Moreover, researchers found that the histone protein H1, H2, encoded by HIST1H1C, regulates the ability of influenza viruses to replicate. And specific substitutions on PB2 regulate viral replication by affecting the expression and modification of HIST1H1C [169].
In summary, the improvement of viral polymerase activity facilitates the adaptation of avian influenza viruses to mammalian hosts and enhances their transmission potential, potentially contributing to the acquisition of airborne transmission capability.
Table 4. Sites of polymerase mutation that affect the replication and transmission of influenza A viruses.
Table 4. Sites of polymerase mutation that affect the replication and transmission of influenza A viruses.
HostViral SubtypesSubunit
(PA/PB1/PB2)
Polymerase Mutation SiteImpact on Viral Replication and TransmissibilityReference
MouseH9N2PAK356REnhanced viral replication in mice[160]
Guinea pigH1N1PB2D309NVirus can be transmitted between guinea pigs by direct contact and has an increased replication capacity[166]
H7N9PB2V292I K627EThe ability of airborne transmission between guinea pigs is lost[131]
H9N2PB2R340K A588VIndividual mutations enable guinea pigs to acquire contact transmissibility, and combined mutations facilitate the virus in acquiring airborne transmissibility[170]
FerretH1N1PAV100I N321K
I330V A639T
The virus is transmitted efficiently between ferrets through respiratory droplets and replicates with higher efficiency[168]
H7N1PB2T81IThe virus can be transmitted through the air between ferrets when combined with other mutations[170]
H3N8PB1S524GThe virus is transmitted efficiently between ferrets through respiratory droplets and replicates with higher efficiency[163]

3.5. Reassortment

The genome of the influenza A virus is composed of eight single-stranded, negative-sense RNA segments. Co-infection by different strains in the same host allows for reassortment, resulting in enhanced viral diversity and rapid evolution [171]. It has to be noted that among the four influenza pandemics in the past, three were caused by reassortment [172]. The phenomenon of segment mismatch (including RNA mismatch and protein mismatch between co-infecting viruses) results in the production of progeny viruses with fitness defects [173]. Conversely, reassortant viruses with high levels of genetic compatibility may enhance host adaptability, pathogenicity, and transmissibility and even acquire airborne capability. A reassortant H7N6 virus with internal genes likely derived from H7N9 and H5N6 exhibited comparable binding affinity for both avian-like and human-like receptors and displayed efficient airborne transmission among guinea pigs [174]. The novel reassortant H10N3 virus, whose internal genes are well matched and derived from the H9N2 virus, is transmitted between guinea pigs through direct contact and respiratory droplets [76]. In addition, it has been found that the matching mechanism of HA, NA, and internal gene segments (such as PB2-PB1-PA-NP) is crucial to determining whether influenza A viruses will emerge and successfully transmit among humans [146,175]. The development of reverse genetics techniques has facilitated the assessment of the effects of a single gene on viral phenotypic traits such as virus replication, host range, and transmissibility [176]. The introduction of the HA gene from non-airborne transmissible H7N9 into the genome of the airborne transmissible H9N2 virus decreased acid and heat stability and completely eliminated airborne transmission among chickens. The results demonstrated that the HA gene influenced the airborne phenotypes of H7N9 and H9N2 avian influenza viruses among chickens [177]. The PA gene of the influenza A virus A/swine/Shandong/07/2011 (SD07) (H1N1) was introduced into the avian influenza virus A/Chicken/Shandong/01/2008 (SD01) (H9N2) to obtain the reassortant virus rSD01-PA. rSD01-PA virus lost the ability of airborne transmission among SPF chickens but was still able to infect guinea pigs through direct contact. The findings indicate that the PA gene plays a crucial role in the replication and airborne transmission of the H9N2 avian influenza virus [178]. In conclusion, reassortment is a significant factor in enabling novel host adaptation and cross-species transmission (including airborne transmission, the characteristic of pandemic viruses) of influenza A viruses.

3.6. Innate Immune Responses of the Host and Immune Evasion Mechanisms of the Virus

The innate immune system serves as a rapidly responsive first line of defense against influenza A viruses, which helps to restrict viral replication at an early stage and prevent further viral spread [179]. Upon viral infection, pathogen-associated molecular patterns (PAMPs) are recognized by pattern recognition receptors (PRRs) [180]. Activation of these PRRs triggers signaling pathways, leading to the production of interferon (IFN) and proinflammatory cytokines, followed by the expression of interferon-stimulated genes (ISGs) and recruitment of innate immune cells [181]. Many ISG genes have antiviral activity, one of the most potent of these genes is the human myxovirus resistance protein A (MXA). MXA targets the NP of MXA-sensitive viruses and inhibits the transcriptional and replicative functions of viruses. It is an important barrier for cross-species transmission of influenza viruses [182,183]. The tripartite-motif-containing (TRIM) protein TRIM56 has been identified as an intrinsic host restriction factor of influenza A and B viruses. It impedes influenza virus infection by hindering viral RNA synthesis via its C-terminal-tail portion [184]. Moreover, research has demonstrated that eosinophils can initiate a self-preservation response during influenza A virus infection, survive the virus infection, and participate in the antiviral response of the host. The host’s innate immune system restricts the replication and infection of influenza A viruses [185]. Consequently, viruses have to counteract host antiviral activities in order to replicate in host cells. Viral non-structure proteins, nonstructural protein 1 (NS1), and PA-X (a fusion protein), play an important role in the evasion of the host innate immune response by the virus [186]. The NS1 protein counteracts the host antiviral response through a variety of mechanisms, such as inhibiting interferon regulatory factor 3 (IRF3) and nuclear factor kappa-B (NF-κβ) transcription factors, impairing IFN and ISG production [187]. During viral infection, NS1 and PA-X mediate the shutoff of host protein expression and inhibit cellular antiviral responses [188]. Studies have shown that multiple mutations associated with the express ability of host genes have appeared in NS1 and PA-X, and they are most likely related to host adaptation [189]. Therefore, it is necessary to monitor adaptive mutations in the NS1 and PA-X proteins of influenza A viruses.

4. Conclusions

Adaptive mutations and reassortment can affect a variety of characteristics of influenza viruses, altering their adaptation, and transmissibility. These characteristics include receptor binding preference, HA stability, HA-NA functional balance, and polymerase activity. Numerous factors influence the choice of adaptation pathways that viruses follow to obtain the desired phenotype. These include the virus’s subtype and strain, compensatory mutations, frequent viral exchange in mixed hosts (like pigs and ferrets), and phenotypic selection for protein function. In summary, a variety of factors influence viral phenotypes, and there exist several routes to phenotype acquisition that consider the features of the virus itself, the interactions between various viruses inside the host, and the type of host.

Author Contributions

Conceptualization, L.L. and X.G.; methodology, Q.A. and Y.Z.; investigation, Q.A. and H.Y.; data curation, H.Y. and C.L.; writing—original draft preparation, X.G.; writing—review and editing, X.G., Y.Z., C.L., L.L. and J.Q.; supervision, J.Q. and L.L.; project administration, L.L. and J.Q.; funding acquisition, J.Q. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by Science and Technology Planning Project of Guangdong Province, China (2022B1111010004); Basic Research Program of Guangzhou Municipal Science and Technology Bureau (SL2023A03J01074), Science and Technology Planning Project of Guangdong Province, China (2021B1212040017).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Asha, K.; Kumar, B. Emerging Influenza D Virus Threat: What We Know so Far! J. Clin. Med. 2019, 8, 192. [Google Scholar] [CrossRef]
  2. Lloren, K.K.S.; Lee, T.; Kwon, J.J.; Song, M.S. Molecular Markers for Interspecies Transmission of Avian Influenza Viruses in Mammalian Hosts. Int. J. Mol. Sci. 2017, 18, 2706. [Google Scholar] [CrossRef] [PubMed]
  3. Yang, W.; Schountz, T.; Ma, W. Bat Influenza Viruses: Current Status and Perspective. Viruses 2021, 13, 547. [Google Scholar] [CrossRef] [PubMed]
  4. Yoon, S.W.; Webby, R.J.; Webster, R.G. Evolution and ecology of influenza A viruses. Curr. Top. Microbiol. Immunol. 2014, 385, 359–375. [Google Scholar] [PubMed]
  5. Wille, M.; Holmes, E.C. The Ecology and Evolution of Influenza Viruses. Cold Spring Harb. Perspect. Med. 2020, 10, a038489. [Google Scholar] [CrossRef] [PubMed]
  6. WHO Influenza (Seasonal). Available online: https://www.who.int/news-room/fact-sheets/detail/influenza-(seasonal) (accessed on 27 November 2023).
  7. Griffin, E.F.; Tompkins, S.M. Fitness Determinants of Influenza A Viruses. Viruses 2023, 15, 1959. [Google Scholar] [CrossRef] [PubMed]
  8. Russell, C.J.; Hu, M.; Okda, F.A. Influenza Hemagglutinin Protein Stability, Activation, and Pandemic Risk. Trends Microbiol. 2018, 26, 841–853. [Google Scholar] [CrossRef] [PubMed]
  9. Zhao, T.; Li, Y.; Xu, M.; Wang, W.; Li, S.; Cao, X.; Fuxiang, N.; Wang, Y.; Li, Y.; Zhang, H.; et al. High proportion of H3 avian influenza virus circulating in chickens—An increasing threat to public health. J. Infect. 2023, 87, 153–155. [Google Scholar] [CrossRef]
  10. Pan, M.; Gao, R.; Lv, Q.; Huang, S.; Zhou, Z.; Yang, L.; Li, X.; Zhao, X.; Zou, X.; Tong, W.; et al. Human infection with a novel, highly pathogenic avian influenza A (H5N6) virus: Virological and clinical findings. J. Infect. 2016, 72, 52–59. [Google Scholar] [CrossRef] [PubMed]
  11. Tong, X.C.; Weng, S.S.; Xue, F.; Wu, X.; Xu, T.M.; Zhang, W.H. First human infection by a novel avian influenza A(H7N4) virus. J. Infect. 2018, 77, 249–257. [Google Scholar] [CrossRef] [PubMed]
  12. Pyankova, O.G.; Susloparov, I.M.; Moiseeva, A.A.; Kolosova, N.P.; Onkhonova, G.S.; Danilenko, A.V.; Vakalova, E.V.; Shendo, G.L.; Nekeshina, N.N.; Noskova, L.N.; et al. Isolation of clade 2.3.4.4b A(H5N8), a highly pathogenic avian influenza virus, from a worker during an outbreak on a poultry farm, Russia, December 2020. Eurosurveillance 2021, 26, 2100439. [Google Scholar] [CrossRef] [PubMed]
  13. WHO Human Infection with Avian Influenza A(H5) Viruses. Avian Influenza Weekly Update Number 884 from 24 February 2023. Available online: https://www.who.int/docs/default-source/wpro—documents/emergency/surveillance/avian-influenza/ai_20230224.pdf?sfvrsn=5f006f99_111 (accessed on 7 March 2023).
  14. Wong, K.H.; Lal, S.K. Alternative antiviral approaches to combat influenza A virus. Virus Genes 2023, 59, 25–35. [Google Scholar] [CrossRef] [PubMed]
  15. Abente, E.J.; Santos, J.; Lewis, N.S.; Gauger, P.C.; Stratton, J.; Skepner, E.; Anderson, T.K.; Rajao, D.S.; Perez, D.R.; Vincent, A.L. The Molecular Determinants of Antibody Recognition and Antigenic Drift in the H3 Hemagglutinin of Swine Influenza A Virus. J. Virol. 2016, 90, 8266–8280. [Google Scholar] [CrossRef] [PubMed]
  16. Harrington, W.N.; Kackos, C.M.; Webby, R.J. The evolution and future of influenza pandemic preparedness. Exp. Mol. Med. 2021, 53, 737–749. [Google Scholar] [CrossRef] [PubMed]
  17. Sautto, G.A.; Ross, T.M. Hemagglutinin consensus-based prophylactic approaches to overcome influenza virus diversity. Vet. Ital. 2019, 55, 195–201. [Google Scholar] [PubMed]
  18. Dunning, J.; Thwaites, R.S.; Openshaw, P.J.M. Seasonal and pandemic influenza: 100 years of progress, still much to learn. Mucosal Immunol. 2020, 13, 566–573. [Google Scholar] [CrossRef] [PubMed]
  19. Sun, H.; Li, F.; Liu, Q.; Du, J.; Liu, L.; Sun, H.; Li, C.; Liu, J.; Zhang, X.; Yang, J.; et al. Mink is a highly susceptible host species to circulating human and avian influenza viruses. Emerg. Microbes Infect. 2021, 10, 472–480. [Google Scholar] [CrossRef]
  20. Mena, I.; Nelson, M.I.; Quezada-Monroy, F.; Dutta, J.; Cortes-Fernández, R.; Lara-Puente, J.H.; Castro-Peralta, F.; Cunha, L.F.; Trovão, N.S.; Lozano-Dubernard, B.; et al. Origins of the 2009 H1N1 influenza pandemic in swine in Mexico. Elife 2016, 5, e16777. [Google Scholar] [CrossRef] [PubMed]
  21. Reperant, L.A.; Kuiken, T.; Osterhaus, A.D. Adaptive pathways of zoonotic influenza viruses: From exposure to establishment in humans. Vaccine 2012, 30, 4419–4434. [Google Scholar] [CrossRef] [PubMed]
  22. Sorrell, E.M.; Schrauwen, E.J.A.; Linster, M.; De Graaf, M.; Herfst, S.; Fouchier, R.A.M. Predicting ‘airborne’ influenza viruses: (trans-) mission impossible? Curr. Opin. Virol. 2011, 1, 635–642. [Google Scholar] [CrossRef]
  23. Brankston, G.; Gitterman, L.; Hirji, Z.; Lemieux, C.; Gardam, M. Transmission of influenza A in human beings. Lancet Infect. Dis. 2007, 7, 257–265. [Google Scholar] [CrossRef] [PubMed]
  24. Morawska, L.; Johnson, G.R.; Ristovski, Z.D.; Hargreaves, M.; Mengersen, K.; Corbett, S.; Chao, C.Y.H.; Li, Y.; Katoshevski, D. Size distribution and sites of origin of droplets expelled from the human respiratory tract during expiratory activities. J. Aerosol Sci. 2009, 40, 256–269. [Google Scholar] [CrossRef]
  25. Liu, L.; Wei, J.; Li, Y.; Ooi, A. Evaporation and dispersion of respiratory droplets from coughing. Indoor Air 2017, 27, 179–190. [Google Scholar] [CrossRef] [PubMed]
  26. Scheuch, G. Breathing Is Enough: For the Spread of Influenza Virus and SARS-CoV-2 by Breathing Only. J. Aerosol Med. Pulm. Drug Deliv. 2020, 33, 230–234. [Google Scholar] [CrossRef] [PubMed]
  27. Cowling, B.J.; Ip, D.K.; Fang, V.J.; Suntarattiwong, P.; Olsen, S.J.; Levy, J.; Uyeki, T.M.; Leung, G.M.; Malik Peiris, J.S.; Chotpitayasunondh, T.; et al. Aerosol transmission is an important mode of influenza A virus spread. Nat. Commun. 2013, 4, 1935. [Google Scholar] [CrossRef] [PubMed]
  28. Cáceres, C.J.; Rajao, D.S.; Perez, D.R. Airborne Transmission of Avian Origin H9N2 Influenza A Viruses in Mammals. Viruses 2021, 13, 1919. [Google Scholar] [CrossRef] [PubMed]
  29. Lv, J.; Gao, J.; Wu, B.; Yao, M.; Yang, Y.; Chai, T.; Li, N. Aerosol Transmission of Coronavirus and Influenza Virus of Animal Origin. Front. Vet. Sci. 2021, 8, 572012. [Google Scholar] [CrossRef] [PubMed]
  30. Drossinos, Y.; Weber, T.P.; Stilianakis, N.I. Droplets and aerosols: An artificial dichotomy in respiratory virus transmission. Health Sci. Rep. 2021, 4, e275. [Google Scholar] [CrossRef] [PubMed]
  31. Le Sage, V.; Lowen, A.C.; Lakdawala, S.S. Block the Spread: Barriers to Transmission of Influenza Viruses. Annu. Rev. Virol. 2023, 10, 347–370. [Google Scholar] [CrossRef] [PubMed]
  32. Bahl, P.; Doolan, C.; de Silva, C.; Chughtai, A.A.; Bourouiba, L.; MacIntyre, C.R. Airborne or Droplet Precautions for Health Workers Treating Coronavirus Disease 2019? J. Infect. Dis. 2022, 225, 1561–1568. [Google Scholar] [CrossRef] [PubMed]
  33. Imai, M.; Kawaoka, Y. The role of receptor binding specificity in interspecies transmission of influenza viruses. Curr. Opin. Virol. 2012, 2, 160–167. [Google Scholar] [CrossRef] [PubMed]
  34. Zhao, C.; Pu, J. Influence of Host Sialic Acid Receptors Structure on the Host Specificity of Influenza Viruses. Viruses 2022, 14, 2141. [Google Scholar] [CrossRef] [PubMed]
  35. Shinya, K.; Ebina, M.; Yamada, S.; Ono, M.; Kasai, N.; Kawaoka, Y. Avian flu: Influenza virus receptors in the human airway. Nature 2006, 440, 435–436. [Google Scholar] [CrossRef] [PubMed]
  36. Imai, M.; Takada, K.; Kawaoka, Y. Receptor-Binding Specificity of Influenza Viruses. Methods Mol. Biol. 2022, 2556, 79–96. [Google Scholar] [PubMed]
  37. Kuchipudi, S.V.; Nelli, R.; White, G.A.; Bain, M.; Chang, K.C.; Dunham, S. Differences in influenza virus receptors in chickens and ducks: Implications for interspecies transmission. J. Mol. Genet. Med. 2009, 3, 143–151. [Google Scholar] [CrossRef] [PubMed]
  38. Richard, M.; Fouchier, R.A. Influenza A virus transmission via respiratory aerosols or droplets as it relates to pandemic potential. FEMS Microbiol. Rev. 2016, 40, 68–85. [Google Scholar] [CrossRef] [PubMed]
  39. Neumann, G.; Kawaoka, Y. Host range restriction and pathogenicity in the context of influenza pandemic. Emerg. Infect. Dis. 2006, 12, 881–886. [Google Scholar] [CrossRef] [PubMed]
  40. Liu, J.; Stevens, D.J.; Haire, L.F.; Walker, P.A.; Coombs, P.J.; Russell, R.J.; Gamblin, S.J.; Skehel, J.J. Structures of receptor complexes formed by hemagglutinins from the Asian Influenza pandemic of 1957. Proc. Natl. Acad. Sci. USA 2009, 106, 17175–17180. [Google Scholar] [CrossRef] [PubMed]
  41. Gamblin, S.J.; Vachieri, S.G.; Xiong, X.; Zhang, J.; Martin, S.R.; Skehel, J.J. Hemagglutinin Structure and Activities. Cold Spring Harb. Perspect. Med. 2021, 11, a038638. [Google Scholar] [CrossRef] [PubMed]
  42. Shi, Y.; Wu, Y.; Zhang, W.; Qi, J.; Gao, G.F. Enabling the ‘host jump’: Structural determinants of receptor-binding specificity in influenza A viruses. Nat. Rev. Microbiol. 2014, 12, 822–831. [Google Scholar] [CrossRef]
  43. Matrosovich, M.; Tuzikov, A.; Bovin, N.; Gambaryan, A.; Klimov, A.; Castrucci, M.R.; Donatelli, I.; Kawaoka, Y. Early alterations of the receptor-binding properties of H1, H2, and H3 avian influenza virus hemagglutinins after their introduction into mammals. J. Virol. 2000, 74, 8502–8512. [Google Scholar] [CrossRef] [PubMed]
  44. Carbone, V.; Schneider, E.K.; Rockman, S.; Baker, M.; Huang, J.X.; Ong, C.; Cooper, M.A.; Yuriev, E.; Li, J.; Velkov, T. Molecular Characterisation of the Haemagglutinin Glycan-Binding Specificity of Egg-Adapted Vaccine Strains of the Pandemic 2009 H1N1 Swine Influenza A Virus. Molecules 2015, 20, 10415–10434. [Google Scholar] [CrossRef] [PubMed]
  45. Kutter, J.S.; Linster, M.; de Meulder, D.; Bestebroer, T.M.; Lexmond, P.; Rosu, M.E.; Richard, M.; de Vries, R.P.; Fouchier, R.A.M.; Herfst, S. Continued adaptation of A/H2N2 viruses during pandemic circulation in humans. J. Gen. Virol. 2023, 104, 001881. [Google Scholar] [CrossRef] [PubMed]
  46. Liu, M.; Bakker, A.S.; Narimatsu, Y.; van Kuppeveld, F.J.M.; Clausen, H.; de Haan, C.A.M.; de Vries, E. H3N2 influenza A virus gradually adapts to human-type receptor binding and entry specificity after the start of the 1968 pandemic. Proc. Natl. Acad. Sci. USA 2023, 120, e2304992120. [Google Scholar] [CrossRef]
  47. Joseph, U.; Su, Y.C.; Vijaykrishna, D.; Smith, G.J. The ecology and adaptive evolution of influenza A interspecies transmission. Influenza Other Respir. Viruses 2017, 11, 74–84. [Google Scholar] [CrossRef] [PubMed]
  48. Zhang, W.; Shi, Y.; Qi, J.; Gao, F.; Li, Q.; Fan, Z.; Yan, J.; Gao, G.F. Molecular basis of the receptor binding specificity switch of the hemagglutinins from both the 1918 and 2009 pandemic influenza A viruses by a D225G substitution. J. Virol. 2013, 87, 5949–5958. [Google Scholar] [CrossRef] [PubMed]
  49. Glaser, L.; Stevens, J.; Zamarin, D.; Wilson, I.A.; García-Sastre, A.; Tumpey, T.M.; Basler, C.F.; Taubenberger, J.K.; Palese, P. A single amino acid substitution in 1918 influenza virus hemagglutinin changes receptor binding specificity. J. Virol. 2005, 79, 11533–11536. [Google Scholar] [CrossRef] [PubMed]
  50. Xiong, X.; McCauley, J.W.; Steinhauer, D.A. Receptor binding properties of the influenza virus hemagglutinin as a determinant of host range. Curr. Top. Microbiol. Immunol. 2014, 385, 63–91. [Google Scholar] [PubMed]
  51. Xu, R.; McBride, R.; Paulson, J.C.; Basler, C.F.; Wilson, I.A. Structure, receptor binding, and antigenicity of influenza virus hemagglutinins from the 1957 H2N2 pandemic. J. Virol. 2010, 84, 1715–1721. [Google Scholar] [CrossRef]
  52. Yang, R.; Sun, H.; Gao, F.; Luo, K.; Huang, Z.; Tong, Q.; Song, H.; Han, Q.; Liu, J.; Lan, Y.; et al. Human infection of avian influenza A H3N8 virus and the viral origins: A descriptive study. Lancet Microbe 2022, 3, e824–e834. [Google Scholar] [CrossRef] [PubMed]
  53. Liu, W.J.; Li, J.; Zou, R.; Pan, J.; Jin, T.; Li, L.; Liu, P.; Zhao, Y.; Yu, X.; Wang, H.; et al. Dynamic PB2-E627K substitution of influenza H7N9 virus indicates the in vivo genetic tuning and rapid host adaptation. Proc. Natl. Acad. Sci. USA 2020, 117, 23807–23814. [Google Scholar] [CrossRef] [PubMed]
  54. Min, J.Y.; Santos, C.; Fitch, A.; Twaddle, A.; Toyoda, Y.; DePasse, J.V.; Ghedin, E.; Subbarao, K. Mammalian adaptation in the PB2 gene of avian H5N1 influenza virus. J. Virol. 2013, 87, 10884–10888. [Google Scholar] [CrossRef] [PubMed]
  55. Auewarakul, P.; Suptawiwat, O.; Kongchanagul, A.; Sangma, C.; Suzuki, Y.; Ungchusak, K.; Louisirirotchanakul, S.; Lerdsamran, H.; Pooruk, P.; Thitithanyanont, A.; et al. An avian influenza H5N1 virus that binds to a human-type receptor. J. Virol. 2007, 81, 9950–9955. [Google Scholar] [CrossRef] [PubMed]
  56. Zhu, X.; Viswanathan, K.; Raman, R.; Yu, W.; Sasisekharan, R.; Wilson, I.A. Structural Basis for a Switch in Receptor Binding Specificity of Two H5N1 Hemagglutinin Mutants. Cell Rep. 2015, 13, 1683–1691. [Google Scholar] [CrossRef] [PubMed]
  57. Tharakaraman, K.; Raman, R.; Viswanathan, K.; Stebbins, N.W.; Jayaraman, A.; Krishnan, A.; Sasisekharan, V.; Sasisekharan, R. Structural determinants for naturally evolving H5N1 hemagglutinin to switch its receptor specificity. Cell 2013, 153, 1475–1485. [Google Scholar] [CrossRef]
  58. Guo, F.; Luo, T.; Pu, Z.; Xiang, D.; Shen, X.; Irwin, D.M.; Liao, M.; Shen, Y. Increasing the potential ability of human infections in H5N6 avian influenza A viruses. J. Infect. 2018, 77, 349–356. [Google Scholar] [CrossRef] [PubMed]
  59. de Vries, R.P.; Tzarum, N.; Peng, W.; Thompson, A.J.; Ambepitiya Wickramasinghe, I.N.; de la Pena, A.T.T.; van Breemen, M.J.; Bouwman, K.M.; Zhu, X.; McBride, R.; et al. A single mutation in Taiwanese H6N1 influenza hemagglutinin switches binding to human-type receptors. EMBO Mol. Med. 2017, 9, 1314–1325. [Google Scholar] [CrossRef] [PubMed]
  60. Xu, R.; de Vries, R.P.; Zhu, X.; Nycholat, C.M.; McBride, R.; Yu, W.; Paulson, J.C.; Wilson, I.A. Preferential recognition of avian-like receptors in human influenza A H7N9 viruses. Science 2013, 342, 1230–1235. [Google Scholar] [CrossRef] [PubMed]
  61. Ramos, I.; Krammer, F.; Hai, R.; Aguilera, D.; Bernal-Rubio, D.; Steel, J.; García-Sastre, A.; Fernandez-Sesma, A. H7N9 influenza viruses interact preferentially with α2,3-linked sialic acids and bind weakly to α2,6-linked sialic acids. J. Gen. Virol. 2013, 94 Pt 11, 2417–2423. [Google Scholar] [CrossRef] [PubMed]
  62. Xu, Y.; Peng, R.; Zhang, W.; Qi, J.; Song, H.; Liu, S.; Wang, H.; Wang, M.; Xiao, H.; Fu, L.; et al. Avian-to-Human Receptor-Binding Adaptation of Avian H7N9 Influenza Virus Hemagglutinin. Cell Rep. 2019, 29, 2217–2228.e5. [Google Scholar] [CrossRef] [PubMed]
  63. Peng, W.; Bouwman, K.M.; McBride, R.; Grant, O.C.; Woods, R.J.; Verheije, M.H.; Paulson, J.C.; de Vries, R.P. Enhanced Human-Type Receptor Binding by Ferret-Transmissible H5N1 with a K193T Mutation. J. Virol. 2018, 92, e02016-17. [Google Scholar] [CrossRef] [PubMed]
  64. Alymova, I.V.; York, I.A.; Air, G.M.; Cipollo, J.F.; Gulati, S.; Baranovich, T.; Kumar, A.; Zeng, H.; Gansebom, S.; McCullers, J.A. Glycosylation changes in the globular head of H3N2 influenza hemagglutinin modulatereceptor binding without affecting virus virulence. Sci. Rep. 2016, 6, 36216. [Google Scholar] [CrossRef] [PubMed]
  65. Jayaraman, A.; Koh, X.; Li, J.; Raman, R.; Viswanathan, K.; Shriver, Z.; Sasisekharan, R. Glycosylation at Asn91 of H1N1 haemagglutinin affects binding to glycan receptors. Biochem. J. 2012, 444, 429–435. [Google Scholar] [CrossRef]
  66. Wang, C.-C.; Chen, J.-R.; Tseng, Y.-C.; Hsu, C.-H.; Hung, Y.-F.; Chen, S.-W.; Chen, C.-M.; Khoo, K.-H.; Cheng, T.-J.; Cheng, Y.-S.E.; et al. Glycans on influenza hemagglutinin affect receptor binding and immune response. Proc. Natl. Acad. Sci. USA 2009, 106, 18137–18142. [Google Scholar] [CrossRef] [PubMed]
  67. Tumpey, T.M.; Maines, T.R.; Van Hoeven, N.; Glaser, L.; Solórzano, A.; Pappas, C.; Cox, N.J.; Swayne, D.E.; Palese, P.; Katz, J.M.; et al. A two-amino acid change in the hemagglutinin of the 1918 influenza virus abolishes transmission. Science 2007, 315, 655–659. [Google Scholar] [CrossRef] [PubMed]
  68. Pappas, C.; Viswanathan, K.; Chandrasekaran, A.; Raman, R.; Katz, J.M.; Sasisekharan, R.; Tumpey, T.M. Receptor specificity and transmission of H2N2 subtype viruses isolated from the pandemic of 1957. PLoS ONE 2010, 5, e11158. [Google Scholar]
  69. Yen, H.L.; Liang, C.H.; Wu, C.Y.; Forrest, H.L.; Ferguson, A.; Choy, K.T.; Jones, J.; Wong, D.D.; Cheung, P.P.; Hsu, C.H.; et al. Hemagglutinin-neuraminidase balance confers respiratory-droplet transmissibility of the pandemic H1N1 influenza virus in ferrets. Proc. Natl. Acad. Sci. USA 2011, 108, 14264–14269. [Google Scholar] [CrossRef] [PubMed]
  70. Srinivasan, A.; Viswanathan, K.; Raman, R.; Chandrasekaran, A.; Raguram, S.; Tumpey, T.M.; Sasisekharan, V.; Sasisekharan, R. Quantitative biochemical rationale for differences in transmissibility of 1918 pandemic influenza A viruses. Proc. Natl. Acad. Sci. USA 2008, 105, 2800–2805. [Google Scholar] [CrossRef] [PubMed]
  71. Sun, H.; Li, H.; Tong, Q.; Han, Q.; Liu, J.; Yu, H.; Song, H.; Qi, J.; Li, J.; Yang, J.; et al. Airborne transmission of human-isolated avian H3N8 influenza virus between ferrets. Cell 2023, 186, 4074–4084.e11. [Google Scholar] [CrossRef] [PubMed]
  72. Karlsson, E.A.; Ip, H.S.; Hall, J.S.; Yoon, S.W.; Johnson, J.; Beck, M.A.; Webby, R.J.; Schultz-Cherry, S. Respiratory transmission of an avian H3N8 influenza virus isolated from a harbour seal. Nat. Commun. 2014, 5, 4791. [Google Scholar] [CrossRef]
  73. Linster, M.; van Boheemen, S.; de Graaf, M.; Schrauwen, E.J.A.; Lexmond, P.; Mänz, B.; Bestebroer, T.M.; Baumann, J.; van Riel, D.; Rimmelzwaan, G.F.; et al. Identification, characterization, and natural selection of mutations driving airborne transmission of A/H5N1 virus. Cell 2014, 157, 329–339. [Google Scholar] [CrossRef] [PubMed]
  74. Zhu, H.; Wang, D.; Kelvin, D.J.; Li, L.; Zheng, Z.; Yoon, S.W.; Wong, S.S.; Farooqui, A.; Wang, J.; Banner, D.; et al. Infectivity, transmission, and pathology of human-isolated H7N9 influenza virus in ferrets and pigs. Science 2013, 341, 183–186. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, X.; Li, Y.; Jin, S.; Wang, T.; Sun, W.; Zhang, Y.; Li, F.; Zhao, M.; Sun, L.; Hu, X.; et al. H9N2 influenza virus spillover into wild birds from poultry in China bind to human-type receptors and transmit in mammals via respiratory droplets. Transbound. Emerg. Dis. 2022, 69, 669–684. [Google Scholar] [CrossRef] [PubMed]
  76. Liu, K.; Ding, P.; Pei, Y.; Gao, R.; Han, W.; Zheng, H.; Ji, Z.; Cai, M.; Gu, J.; Li, X.; et al. Emergence of a novel reassortant avian influenza virus (H10N3) in Eastern China with high pathogenicity and respiratory droplet transmissibility to mammals. Sci. China Life Sci. 2022, 65, 1024–1035. [Google Scholar] [CrossRef] [PubMed]
  77. Abdelwhab, E.M.; Veits, J.; Tauscher, K.; Ziller, M.; Grund, C.; Hassan, M.K.; Shaheen, M.; Harder, T.C.; Teifke, J.; Stech, J.; et al. Progressive glycosylation of the haemagglutinin of avian influenza H5N1 modulates virus replication, virulence and chicken-to-chicken transmission without significant impact on antigenic drift. J. Gen. Virol. 2016, 97, 3193–3204. [Google Scholar] [CrossRef]
  78. Kim, J.I.; Park, M.S. N-linked glycosylation in the hemagglutinin of influenza A viruses. Yonsei Med. J. 2012, 53, 886–893. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, W.; Lu, B.; Zhou, H.; Suguitan, A.L., Jr.; Cheng, X.; Subbarao, K.; Kemble, G.; Jin, H. Glycosylation at 158N of the hemagglutinin protein and receptor binding specificity synergistically affect the antigenicity and immunogenicity of a live attenuated H5N1 A/Vietnam/1203/2004 vaccine virus in ferrets. J. Virol. 2010, 84, 6570–6577. [Google Scholar] [CrossRef] [PubMed]
  80. Liu, K.; Guo, Y.; Zheng, H.; Ji, Z.; Cai, M.; Gao, R.; Zhang, P.; Liu, X.; Xu, X.; Wang, X.; et al. Enhanced pathogenicity and transmissibility of H9N2 avian influenza virus in mammals by hemagglutinin mutations combined with PB2-627K. Virol. Sin. 2023, 38, 47–55. [Google Scholar] [CrossRef]
  81. Wang, Z.; Yang, H.; Chen, Y.; Tao, S.; Liu, L.; Kong, H.; Ma, S.; Meng, F.; Suzuki, Y.; Qiao, C.; et al. A Single-Amino-Acid Substitution at Position 225 in Hemagglutinin Alters the Transmissibility of Eurasian Avian-Like H1N1 Swine Influenza Virus in Guinea Pigs. J. Virol. 2017, 91, e00800-17. [Google Scholar] [CrossRef] [PubMed]
  82. Yu, Y.; Wu, M.; Cui, X.; Xu, F.; Wen, F.; Pan, L.; Li, S.; Sun, H.; Zhu, X.; Lin, J.; et al. Pathogenicity and transmissibility of current H3N2 swine influenza virus in Southern China: A zoonotic potential. Transbound. Emerg. Dis. 2022, 69, 2052–2064. [Google Scholar] [CrossRef] [PubMed]
  83. Yao, X.Y.; Lian, C.Y.; Lv, Z.H.; Zhang, X.L.; Shao, J.W. Emergence of a novel reassortant H5N6 subtype highly pathogenic avian influenza virus in farmed dogs in China. J. Infect. 2023, 87, e70–e72. [Google Scholar] [CrossRef] [PubMed]
  84. Li, X.; Shi, J.; Guo, J.; Deng, G.; Zhang, Q.; Wang, J.; He, X.; Wang, K.; Chen, J.; Li, Y.; et al. Genetics, receptor binding property, and transmissibility in mammals of naturally isolated H9N2 Avian Influenza viruses. PLoS Pathog. 2014, 10, e1004508. [Google Scholar] [CrossRef] [PubMed]
  85. Russell, C.J. Acid-induced membrane fusion by the hemagglutinin protein and its role in influenza virus biology. Curr. Top. Microbiol. Immunol. 2014, 385, 93–116. [Google Scholar] [PubMed]
  86. Jakubcová, L.; Hollý, J.; Varečková, E. The role of fusion activity of influenza A viruses in their biological properties. Acta Virol. 2016, 60, 121–135. [Google Scholar] [CrossRef] [PubMed]
  87. Okamatsu, M.; Motohashi, Y.; Hiono, T.; Tamura, T.; Nagaya, K.; Matsuno, K.; Sakoda, Y.; Kida, H. Is the optimal pH for membrane fusion in host cells by avian influenza viruses related to host range and pathogenicity? Arch. Virol. 2016, 161, 2235–2242. [Google Scholar] [CrossRef] [PubMed]
  88. Poulson, R.L.; Tompkins, S.M.; Berghaus, R.D.; Brown, J.D.; Stallknecht, D.E. Environmental Stability of Swine and Human Pandemic Influenza Viruses in Water under Variable Conditions of Temperature, Salinity, and pH. Appl. Environ. Microbiol. 2016, 82, 3721–3726. [Google Scholar] [CrossRef] [PubMed]
  89. Singanayagam, A.; Zhou, J.; Elderfield, R.A.; Frise, R.; Ashcroft, J.; Galiano, M.; Miah, S.; Nicolaou, L.; Barclay, W.S. Characterising viable virus from air exhaled by H1N1 influenza-infected ferrets reveals the importance of haemagglutinin stability for airborne infectivity. PLoS Pathog. 2020, 16, e1008362. [Google Scholar] [CrossRef] [PubMed]
  90. Man, W.H.; de Steenhuijsen Piters, W.A.; Bogaert, D. The microbiota of the respiratory tract: Gatekeeper to respiratory health. Nat. Rev. Microbiol. 2017, 15, 259–270. [Google Scholar] [CrossRef] [PubMed]
  91. Hensen, L.; Matrosovich, T.; Roth, K.; Klenk, H.D.; Matrosovich, M. HA-Dependent Tropism of H5N1 and H7N9 Influenza Viruses to Human Endothelial Cells Is Determined by Reduced Stability of the HA, Which Allows the Virus To Cope with Inefficient Endosomal Acidification and Constitutively Expressed IFITM3. J. Virol. 2019, 94, e01223-19. [Google Scholar] [CrossRef] [PubMed]
  92. Galloway, S.E.; Reed, M.L.; Russell, C.J.; Steinhauer, D.A. Influenza HA subtypes demonstrate divergent phenotypes for cleavage activation and pH of fusion: Implications for host range and adaptation. PLoS Pathog. 2013, 9, e1003151. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, W.; Shi, Y.; Lu, X.; Shu, Y.; Qi, J.; Gao, G.F. An airborne transmissible avian influenza H5 hemagglutinin seen at the atomic level. Science 2013, 340, 1463–1467. [Google Scholar] [CrossRef] [PubMed]
  94. Cotter, C.R.; Jin, H.; Chen, Z. A single amino acid in the stalk region of the H1N1pdm influenza virus HA protein affects viral fusion, stability and infectivity. PLoS Pathog. 2014, 10, e1003831. [Google Scholar] [CrossRef]
  95. Mair, C.M.; Meyer, T.; Schneider, K.; Huang, Q.; Veit, M.; Herrmann, A. A histidine residue of the influenza virus hemagglutinin controls the pH dependence of the conformational change mediating membrane fusion. J. Virol. 2014, 88, 13189–13200. [Google Scholar] [CrossRef] [PubMed]
  96. Mair, C.M.; Ludwig, K.; Herrmann, A.; Sieben, C. Receptor binding and pH stability—How influenza A virus hemagglutinin affects host-specific virus infection. Biochim. Biophys. Acta (BBA)-Biomembr. 2014, 1838, 1153–1168. [Google Scholar] [CrossRef]
  97. Koerner, I.; Matrosovich, M.N.; Haller, O.; Staeheli, P.; Kochs, G. Altered receptor specificity and fusion activity of the haemagglutinin contribute to high virulence of a mouse-adapted influenza A virus. J. Gen. Virol. 2012, 93 Pt 5, 970–979. [Google Scholar] [CrossRef] [PubMed]
  98. Reed, M.L.; Bridges, O.A.; Seiler, P.; Kim, J.K.; Yen, H.L.; Salomon, R.; Govorkova, E.A.; Webster, R.G.; Russell, C.J. The pH of activation of the hemagglutinin protein regulates H5N1 influenza virus pathogenicity and transmissibility in ducks. J. Virol. 2010, 84, 1527–1535. [Google Scholar] [CrossRef] [PubMed]
  99. Jakubcová, L.; Vozárová, M.; Hollý, J.; Tomčíková, K.; Fogelová, M.; Polčicová, K.; Kostolanský, F.; Fodor, E.; Varečková, E. Biological properties of influenza A virus mutants with amino acid substitutions in the HA2 glycoprotein of the HA1/HA2 interaction region. J. Gen. Virol. 2019, 100, 1282–1292. [Google Scholar] [CrossRef] [PubMed]
  100. Di Lella, S.; Herrmann, A.; Mair, C.M. Modulation of the pH Stability of Influenza Virus Hemagglutinin: A Host Cell Adaptation Strategy. Biophys. J. 2016, 110, 2293–2301. [Google Scholar] [CrossRef] [PubMed]
  101. Russier, M.; Yang, G.; Rehg, J.E.; Wong, S.S.; Mostafa, H.H.; Fabrizio, T.P.; Barman, S.; Krauss, S.; Webster, R.G.; Webby, R.J.; et al. Molecular requirements for a pandemic influenza virus: An acid-stable hemagglutinin protein. Proc. Natl. Acad. Sci. USA 2016, 113, 1636–1641. [Google Scholar] [CrossRef] [PubMed]
  102. Le Sage, V.; Kormuth, K.A.; Nturibi, E.; Lee, J.M.; Frizzell, S.A.; Myerburg, M.M.; Bloom, J.D.; Lakdawala, S.S. Cell-Culture Adaptation of H3N2 Influenza Virus Impacts Acid Stability and Reduces Airborne Transmission in Ferret Model. Viruses 2021, 13, 719. [Google Scholar] [CrossRef]
  103. Hu, M.; Yang, G.; DeBeauchamp, J.; Crumpton, J.C.; Kim, H.; Li, L.; Wan, X.F.; Kercher, L.; Bowman, A.S.; Webster, R.G.; et al. HA stabilization promotes replication and transmission of swine H1N1 gamma influenza viruses in ferrets. Elife 2020, 9, e56236. [Google Scholar] [CrossRef] [PubMed]
  104. Okda, F.A.; Perry, S.S.; Webby, R.J.; Russell, C.J. Interplay between H1N1 influenza a virus infection, extracellular and intracellular respiratory tract pH, and host responses in a mouse model. PLoS ONE 2021, 16, e0251473. [Google Scholar] [CrossRef] [PubMed]
  105. Imai, M.; Watanabe, T.; Hatta, M.; Das, S.C.; Ozawa, M.; Shinya, K.; Zhong, G.; Hanson, A.; Katsura, H.; Watanabe, S.; et al. Experimental adaptation of an influenza H5 HA confers respiratory droplet transmission to a reassortant H5 HA/H1N1 virus in ferrets. Nature 2012, 486, 420–428. [Google Scholar] [CrossRef] [PubMed]
  106. Hanson, A.; Imai, M.; Hatta, M.; McBride, R.; Imai, H.; Taft, A.; Zhong, G.; Watanabe, T.; Suzuki, Y.; Neumann, G.; et al. Identification of Stabilizing Mutations in an H5 Hemagglutinin Influenza Virus Protein. J. Virol. 2015, 90, 2981–2992. [Google Scholar] [CrossRef] [PubMed]
  107. Xu, N.; Wang, X.; Cai, M.; Tang, X.; Yang, W.; Lu, X.; Liu, X.; Hu, S.; Gu, M.; Hu, J.; et al. Mutations in HA and PA affect the transmissibility of H7N9 avian influenza virus in chickens. Vet. Microbiol. 2023, 287, 109910. [Google Scholar] [CrossRef] [PubMed]
  108. Zhong, L.; Wang, X.; Li, Q.; Liu, D.; Chen, H.; Zhao, M.; Gu, X.; He, L.; Liu, X.; Gu, M.; et al. Molecular mechanism of the airborne transmissibility of H9N2 avian influenza A viruses in chickens. J. Virol. 2014, 88, 9568–9578. [Google Scholar] [CrossRef] [PubMed]
  109. Russier, M.; Yang, G.; Marinova-Petkova, A.; Vogel, P.; Kaplan, B.S.; Webby, R.J.; Russell, C.J. H1N1 influenza viruses varying widely in hemagglutinin stability transmit efficiently from swine to swine and to ferrets. PLoS Pathog. 2017, 13, e1006276. [Google Scholar] [CrossRef] [PubMed]
  110. Hu, M.; Kackos, C.; Banoth, B.; Ojha, C.R.; Jones, J.C.; Lei, S.; Li, L.; Kercher, L.; Webby, R.J.; Russell, C.J. Hemagglutinin destabilization in H3N2 vaccine reference viruses skews antigenicity and prevents airborne transmission in ferrets. Sci. Adv. 2023, 9, eadf5182. [Google Scholar] [CrossRef] [PubMed]
  111. Herfst, S.; Zhang, J.; Richard, M.; McBride, R.; Lexmond, P.; Bestebroer, T.M.; Spronken, M.I.J.; de Meulder, D.; van den Brand, J.M.; Rosu, M.E.; et al. Hemagglutinin Traits Determine Transmission of Avian A/H10N7 Influenza Virus between Mammals. Cell Host Microbe 2020, 28, 602–613. [Google Scholar] [CrossRef] [PubMed]
  112. Sun, X.; Belser, J.A.; Pulit-Penaloza, J.A.; Brock, N.; Kieran, T.J.; Zeng, H.; Pappas, C.; Tumpey, T.M.; Maines, T.R. A naturally occurring HA-stabilizing amino acid (HA1-Y17) in an A(H9N2) low-pathogenic influenza virus contributes to airborne transmission. mBio 2024, 15, e0295723. [Google Scholar] [CrossRef] [PubMed]
  113. Gaymard, A.; Le Briand, N.; Frobert, E.; Lina, B.; Escuret, V. Functional balance between neuraminidase and haemagglutinin in influenza viruses. Clin. Microbiol. Infect. 2016, 22, 975–983. [Google Scholar] [CrossRef] [PubMed]
  114. Ohuchi, M.; Asaoka, N.; Sakai, T.; Ohuchi, R. Roles of neuraminidase in the initial stage of influenza virus infection. Microbes Infect. 2006, 8, 1287–1293. [Google Scholar] [CrossRef] [PubMed]
  115. Russell, R.J.; Haire, L.F.; Stevens, D.J.; Collins, P.J.; Lin, Y.P.; Blackburn, G.M.; Hay, A.J.; Gamblin, S.J.; Skehel, J.J. The structure of H5N1 avian influenza neuraminidase suggests new opportunities for drug design. Nature 2006, 443, 45–49. [Google Scholar] [CrossRef] [PubMed]
  116. Du, W.; de Vries, E.; van Kuppeveld, F.J.M.; Matrosovich, M.; de Haan, C.A.M. Second sialic acid-binding site of influenza A virus neuraminidase: Binding receptors for efficient release. FEBS J. 2021, 288, 5598–5612. [Google Scholar] [CrossRef] [PubMed]
  117. Air, G.M. Influenza neuraminidase. Influenza Other Respir. Viruses 2012, 6, 245–256. [Google Scholar] [CrossRef] [PubMed]
  118. Du, W.; Dai, M.; Li, Z.; Boons, G.J.; Peeters, B.; van Kuppeveld, F.J.M.; de Vries, E.; de Haan, C.A.M. Substrate Binding by the Second Sialic Acid-Binding Site of Influenza A Virus N1 Neuraminidase Contributes to Enzymatic Activity. J. Virol. 2018, 92, e01243-18. [Google Scholar] [CrossRef] [PubMed]
  119. Iseli, A.N.; Pohl, M.O.; Glas, I.; Gaggioli, E.; Martínez-Barragán, P.; David, S.C.; Schaub, A.; Luo, B.; Klein, L.K.; Bluvshtein, N.; et al. The neuraminidase activity of influenza A virus determines the strain-specific sensitivity to neutralization by respiratory mucus. J. Virol. 2023, 97, e0127123. [Google Scholar] [CrossRef] [PubMed]
  120. Kaler, L.; Engle, E.M.; Iverson, E.; Boboltz, A.; Ignacio, M.A.; Rife, M.; Scull, M.A.; Duncan, G.A. Mucus physically restricts influenza A viral particle access to the epithelium. bioRxiv 2023. [Google Scholar] [CrossRef]
  121. Yang, X.; Steukers, L.; Forier, K.; Xiong, R.; Braeckmans, K.; Van Reeth, K.; Nauwynck, H. A beneficiary role for neuraminidase in influenza virus penetration through the respiratory mucus. PLoS ONE 2014, 9, e110026. [Google Scholar] [CrossRef] [PubMed]
  122. de Vries, E.; Du, W.; Guo, H.; de Haan, C.A.M. Influenza A Virus Hemagglutinin-Neuraminidase-Receptor Balance: Preserving Virus Motility. Trends Microbiol. 2020, 28, 57–67. [Google Scholar] [CrossRef]
  123. Guo, H.; Rabouw, H.; Slomp, A.; Dai, M.; van der Vegt, F.; van Lent, J.W.M.; McBride, R.; Paulson, J.C.; de Groot, R.J.; van Kuppeveld, F.J.M.; et al. Kinetic analysis of the influenza A virus HA/NA balance reveals contribution of NA to virus-receptor binding and NA-dependent rolling on receptor-containing surfaces. PLoS Pathog. 2018, 14, e1007233. [Google Scholar] [CrossRef] [PubMed]
  124. Jia, N.; Barclay, W.S.; Roberts, K.; Yen, H.L.; Chan, R.W.; Lam, A.K.; Air, G.; Peiris, J.S.; Dell, A.; Nicholls, J.M.; et al. Glycomic characterization of respiratory tract tissues of ferrets: Implications for its use in influenza virus infection studies. J. Biol. Chem. 2014, 289, 28489–28504. [Google Scholar] [CrossRef] [PubMed]
  125. Lai, J.C.C.; Karunarathna, H.; Wong, H.H.; Peiris, J.S.M.; Nicholls, J.M. Neuraminidase activity and specificity of influenza A virus are influenced by haemagglutinin-receptor binding. Emerg. Microbes Infect. 2019, 8, 327–338. [Google Scholar] [CrossRef] [PubMed]
  126. Cohen, M.; Zhang, X.Q.; Senaati, H.P.; Chen, H.W.; Varki, N.M.; Schooley, R.T.; Gagneux, P. Influenza A penetrates host mucus by cleaving sialic acids with neuraminidase. Virol. J. 2013, 10, 321. [Google Scholar] [CrossRef] [PubMed]
  127. Casalegno, J.S.; Ferraris, O.; Escuret, V.; Bouscambert, M.; Bergeron, C.; Linès, L.; Excoffier, T.; Valette, M.; Frobert, E.; Pillet, S.; et al. Functional balance between the hemagglutinin and neuraminidase of influenza A(H1N1)pdm09 HA D222 variants. PLoS ONE 2014, 9, e104009. [Google Scholar] [CrossRef] [PubMed]
  128. Dadonaite, B.; Vijayakrishnan, S.; Fodor, E.; Bhella, D.; Hutchinson, E.C. Filamentous influenza viruses. J. Gen. Virol. 2016, 97, 1755–1764. [Google Scholar] [CrossRef]
  129. Park, S.; Lee, I.; Kim, J.I.; Bae, J.Y.; Yoo, K.; Kim, J.; Nam, M.; Park, M.; Yun, S.H.; Cho, W.I.; et al. Effects of HA and NA glycosylation pattern changes on the transmission of avian influenza A(H7N9) virus in guinea pigs. Biochem. Biophys. Res. Commun. 2016, 479, 192–197. [Google Scholar] [CrossRef] [PubMed]
  130. Matrosovich, M.; Zhou, N.; Kawaoka, Y.; Webster, R. The surface glycoproteins of H5 influenza viruses isolated from humans, chickens, and wild aquatic birds have distinguishable properties. J. Virol. 1999, 73, 1146–1155. [Google Scholar] [CrossRef] [PubMed]
  131. Kong, H.; Ma, S.; Wang, J.; Gu, C.; Wang, Z.; Shi, J.; Deng, G.; Guan, Y.; Chen, H. Identification of Key Amino Acids in the PB2 and M1 Proteins of H7N9 Influenza Virus That Affect Its Transmission in Guinea Pigs. J. Virol. 2019, 94, e01180-19. [Google Scholar] [CrossRef] [PubMed]
  132. Campbell, P.J.; Kyriakis, C.S.; Marshall, N.; Suppiah, S.; Seladi-Schulman, J.; Danzy, S.; Lowen, A.C.; Steel, J. Residue 41 of the Eurasian avian-like swine influenza a virus matrix protein modulates virion filament length and efficiency of contact transmission. J. Virol. 2014, 88, 7569–7577. [Google Scholar] [CrossRef]
  133. Liao, H.Y.; Hsu, C.H.; Wang, S.C.; Liang, C.H.; Yen, H.Y.; Su, C.Y.; Chen, C.H.; Jan, J.T.; Ren, C.T.; Chen, C.H.; et al. Differential receptor binding affinities of influenza hemagglutinins on glycan arrays. J. Am. Chem. Soc. 2010, 132, 14849–14856. [Google Scholar] [CrossRef]
  134. Gu, M.; Zhao, Y.; Ge, Z.; Li, Y.; Gao, R.; Wang, X.; Hu, J.; Liu, X.; Hu, S.; Peng, D.; et al. Effects of HA2 154 deglycosylation and NA V202I mutation on biological property of H5N6 subtype avian influenza virus. Vet. Microbiol. 2022, 266, 109353. [Google Scholar] [CrossRef] [PubMed]
  135. Bao, D.; Xue, R.; Zhang, M.; Lu, C.; Ma, T.; Ren, C.; Zhang, T.; Yang, J.; Teng, Q.; Li, X.; et al. N-Linked Glycosylation Plays an Important Role in Budding of Neuraminidase Protein and Virulence of Influenza Viruses. J. Virol. 2021, 95, e02042-20. [Google Scholar] [CrossRef] [PubMed]
  136. Wu, C.-Y.; Lin, C.-W.; Tsai, T.-I.; Lee, C.-C.D.; Chuang, H.-Y.; Chen, J.-B.; Tsai, M.-H.; Chen, B.-R.; Lo, P.-W.; Liu, C.-P.; et al. Influenza A surface glycosylation and vaccine design. Proc. Natl. Acad. Sci. USA 2017, 114, 280–285. [Google Scholar] [CrossRef] [PubMed]
  137. Kim, P.; Jang, Y.H.; Kwon, S.B.; Lee, C.M.; Han, G.; Seong, B.L. Glycosylation of Hemagglutinin and Neuraminidase of Influenza A Virus as Signature for Ecological Spillover and Adaptation among Influenza Reservoirs. Viruses 2018, 10, 183. [Google Scholar] [CrossRef] [PubMed]
  138. Munier, S.; Larcher, T.; Cormier-Aline, F.; Soubieux, D.; Su, B.; Guigand, L.; Labrosse, B.; Cherel, Y.; Quéré, P.; Marc, D.; et al. A genetically engineered waterfowl influenza virus with a deletion in the stalk of the neuraminidase has increased virulence for chickens. J. Virol. 2010, 84, 940–952. [Google Scholar] [CrossRef]
  139. Arai, Y.; Elgendy, E.M.; Daidoji, T.; Ibrahim, M.S.; Ono, T.; Sriwilaijaroen, N.; Suzuki, Y.; Nakaya, T.; Matsumoto, K.; Watanabe, Y. H9N2 Influenza Virus Infections in Human Cells Require a Balance between Neuraminidase Sialidase Activity and Hemagglutinin Receptor Affinity. J. Virol. 2020, 94, e01210-20. [Google Scholar] [CrossRef] [PubMed]
  140. Blumenkrantz, D.; Roberts, K.L.; Shelton, H.; Lycett, S.; Barclay, W.S. The short stalk length of highly pathogenic avian influenza H5N1 virus neuraminidase limits transmission of pandemic H1N1 virus in ferrets. J. Virol. 2013, 87, 10539–10551. [Google Scholar] [CrossRef]
  141. Du, W.; Guo, H.; Nijman, V.S.; Doedt, J.; van der Vries, E.; van der Lee, J.; Li, Z.; Boons, G.J.; van Kuppeveld, F.J.M.; de Vries, E.; et al. The 2nd sialic acid-binding site of influenza A virus neuraminidase is an important determinant of the hemagglutinin-neuraminidase-receptor balance. PLoS Pathog. 2019, 15, e1007860. [Google Scholar] [CrossRef] [PubMed]
  142. Du, W.; Wolfert, M.A.; Peeters, B.; van Kuppeveld, F.J.M.; Boons, G.J.; de Vries, E.; de Haan, C.A.M. Mutation of the second sialic acid-binding site of influenza A virus neuraminidase drives compensatory mutations in hemagglutinin. PLoS Pathog. 2020, 16, e1008816. [Google Scholar] [CrossRef] [PubMed]
  143. Dai, M.; McBride, R.; Dortmans, J.C.F.M.; Peng, W.; Bakkers, M.J.G.; de Groot, R.J.; van Kuppeveld, F.J.M.; Paulson, J.C.; de Vries, E.; de Haan, C.A.M. Mutation of the Second Sialic Acid-Binding Site, Resulting in Reduced Neuraminidase Activity, Preceded the Emergence of H7N9 Influenza A Virus. J. Virol. 2017, 91, e00049-17. [Google Scholar] [CrossRef] [PubMed]
  144. Gambaryan, A.S.; Matrosovich, M.N. What adaptive changes in hemagglutinin and neuraminidase are necessary for emergence of pandemic influenza virus from its avian precursor? Biochemistry 2015, 80, 872–880. [Google Scholar] [CrossRef]
  145. Yoon, S.W.; Chen, N.; Ducatez, M.F.; McBride, R.; Barman, S.; Fabrizio, T.P.; Webster, R.G.; Haliloglu, T.; Paulson, J.C.; Russell, C.J.; et al. Changes to the dynamic nature of hemagglutinin and the emergence of the 2009 pandemic H1N1 influenza virus. Sci. Rep. 2015, 5, 12828. [Google Scholar] [CrossRef] [PubMed]
  146. Xu, R.; Zhu, X.; McBride, R.; Nycholat, C.M.; Yu, W.; Paulson, J.C.; Wilson, I.A. Functional balance of the hemagglutinin and neuraminidase activities accompanies the emergence of the 2009 H1N1 influenza pandemic. J. Virol. 2012, 86, 9221–9232. [Google Scholar] [CrossRef] [PubMed]
  147. Te Velthuis, A.J.; Fodor, E. Influenza virus RNA polymerase: Insights into the mechanisms of viral RNA synthesis. Nat. Rev. Microbiol. 2016, 14, 479–493. [Google Scholar] [CrossRef]
  148. Lo, C.Y.; Tang, Y.S.; Shaw, P.C. Structure and Function of Influenza Virus Ribonucleoprotein. Subcell. Biochem. 2018, 88, 95–128. [Google Scholar]
  149. Fodor, E.; Te Velthuis, A.J.W. Structure and Function of the Influenza Virus Transcription and Replication Machinery. Cold Spring Harb. Perspect. Med. 2020, 10, a038398. [Google Scholar] [CrossRef] [PubMed]
  150. Alkie, T.N.; Cox, S.; Embury-Hyatt, C.; Stevens, B.; Pople, N.; Pybus, M.J.; Xu, W.; Hisanaga, T.; Suderman, M.; Koziuk, J.; et al. Characterization of neurotropic HPAI H5N1 viruses with novel genome constellations and mammalian adaptive mutations in free-living mesocarnivores in Canada. Emerg. Microbes Infect. 2023, 12, 2186608. [Google Scholar] [CrossRef] [PubMed]
  151. Bordes, L.; Vreman, S.; Heutink, R.; Roose, M.; Venema, S.; Pritz-Verschuren, S.B.E.; Rijks, J.M.; Gonzales, J.L.; Germeraad, E.A.; Engelsma, M.; et al. Highly Pathogenic Avian Influenza H5N1 Virus Infections in Wild Red Foxes (Vulpes vulpes) Show Neurotropism and Adaptive Virus Mutations. Microbiol. Spectr. 2023, 11, e0286722. [Google Scholar] [CrossRef]
  152. Massin, P.; van der Werf, S.; Naffakh, N. Residue 627 of PB2 is a determinant of cold sensitivity in RNA replication of avian influenza viruses. J. Virol. 2001, 75, 5398–5404. [Google Scholar] [CrossRef]
  153. Aggarwal, S.; Dewhurst, S.; Takimoto, T.; Kim, B. Biochemical impact of the host adaptation-associated PB2 E627K mutation on the temperature-dependent RNA synthesis kinetics of influenza A virus polymerase complex. J. Biol. Chem. 2011, 286, 34504–34513. [Google Scholar] [CrossRef] [PubMed]
  154. Nilsson-Payant, B.E.; tenOever, B.R.; Te Velthuis, A.J.W. The Host Factor ANP32A Is Required for Influenza A Virus vRNA and cRNA Synthesis. J. Virol. 2022, 96, e0209221. [Google Scholar] [CrossRef] [PubMed]
  155. Long, J.S.; Giotis, E.S.; Moncorgé, O.; Frise, R.; Mistry, B.; James, J.; Morisson, M.; Iqbal, M.; Vignal, A.; Skinner, M.A.; et al. Species difference in ANP32A underlies influenza A virus polymerase host restriction. Nature 2016, 529, 101–104. [Google Scholar] [CrossRef] [PubMed]
  156. Peacock, T.P.; Swann, O.C.; Salvesen, H.A.; Staller, E.; Leung, P.B.; Goldhill, D.H.; Zhou, H.; Lillico, S.G.; Whitelaw, C.B.A.; Long, J.S.; et al. Swine ANP32A Supports Avian Influenza Virus Polymerase. J. Virol. 2020, 94, e00132-20. [Google Scholar] [CrossRef] [PubMed]
  157. Resa-Infante, P.; Jorba, N.; Zamarreño, N.; Fernández, Y.; Juárez, S.; Ortín, J. The host-dependent interaction of alpha-importins with influenza PB2 polymerase subunit is required for virus RNA replication. PLoS ONE 2008, 3, e3904. [Google Scholar] [CrossRef] [PubMed]
  158. Gabriel, G.; Czudai-Matwich, V.; Klenk, H.D. Adaptive mutations in the H5N1 polymerase complex. Virus Res. 2013, 178, 53–62. [Google Scholar] [CrossRef] [PubMed]
  159. Zhong, G.; Le, M.Q.; Lopes, T.J.S.; Halfmann, P.; Hatta, M.; Fan, S.; Neumann, G.; Kawaoka, Y. Mutations in the PA Protein of Avian H5N1 Influenza Viruses Affect Polymerase Activity and Mouse Virulence. J. Virol. 2018, 92, e01557-17. [Google Scholar] [CrossRef] [PubMed]
  160. Xu, G.; Zhang, X.; Gao, W.; Wang, C.; Wang, J.; Sun, H.; Sun, Y.; Guo, L.; Zhang, R.; Chang, K.C.; et al. Prevailing PA Mutation K356R in Avian Influenza H9N2 Virus Increases Mammalian Replication and Pathogenicity. J. Virol. 2016, 90, 8105–8114. [Google Scholar] [CrossRef] [PubMed]
  161. Kamiki, H.; Matsugo, H.; Kobayashi, T.; Ishida, H.; Takenaka-Uema, A.; Murakami, S.; Horimoto, T. A PB1-K577E Mutation in H9N2 Influenza Virus Increases Polymerase Activity and Pathogenicity in Mice. Viruses 2018, 10, 653. [Google Scholar] [CrossRef] [PubMed]
  162. Lina, L.; Saijuan, C.; Chengyu, W.; Yuefeng, L.; Shishan, D.; Ligong, C.; Kangkang, G.; Zhendong, G.; Jiakai, L.; Jianhui, Z.; et al. Adaptive amino acid substitutions enable transmission of an H9N2 avian influenza virus in guinea pigs. Sci. Rep. 2019, 9, 19734. [Google Scholar] [CrossRef] [PubMed]
  163. Zhang, X.; Li, Y.; Jin, S.; Zhang, Y.; Sun, L.; Hu, X.; Zhao, M.; Li, F.; Wang, T.; Sun, W.; et al. PB1 S524G mutation of wild bird-origin H3N8 influenza A virus enhances virulence and fitness for transmission in mammals. Emerg. Microbes Infect. 2021, 10, 1038–1051. [Google Scholar] [CrossRef]
  164. Schrauwen, E.J.; Fouchier, R.A. Host adaptation and transmission of influenza A viruses in mammals. Emerg. Microbes Infect. 2014, 3, e9. [Google Scholar] [CrossRef] [PubMed]
  165. Lakdawala, S.S.; Lamirande, E.W.; Suguitan, A.L., Jr.; Wang, W.; Santos, C.P.; Vogel, L.; Matsuoka, Y.; Lindsley, W.G.; Jin, H.; Subbarao, K. Eurasian-origin gene segments contribute to the transmissibility, aerosol release, and morphology of the 2009 pandemic H1N1 influenza virus. PLoS Pathog 2011, 7, e1002443. [Google Scholar] [CrossRef] [PubMed]
  166. Hu, J.; Hu, Z.; Wei, Y.; Zhang, M.; Wang, S.; Tong, Q.; Sun, H.; Pu, J.; Liu, J.; Sun, Y. Mutations in PB2 and HA are crucial for the increased virulence and transmissibility of H1N1 swine influenza virus in mammalian models. Vet. Microbiol. 2022, 265, 109314. [Google Scholar] [CrossRef]
  167. Shi, J.; Deng, G.; Kong, H.; Gu, C.; Ma, S.; Yin, X.; Zeng, X.; Cui, P.; Chen, Y.; Yang, H.; et al. H7N9 virulent mutants detected in chickens in China pose an increased threat to humans. Cell Res. 2017, 27, 1409–1421. [Google Scholar] [CrossRef] [PubMed]
  168. Meng, F.; Yang, H.; Qu, Z.; Chen, Y.; Zhang, Y.; Zhang, Y.; Liu, L.; Zeng, X.; Li, C.; Kawaoka, Y.; et al. A Eurasian avian-like H1N1 swine influenza reassortant virus became pathogenic and highly transmissible due to mutations in its PA gene. Proc. Natl. Acad. Sci. USA 2022, 119, e2203919119. [Google Scholar] [CrossRef] [PubMed]
  169. Boergeling, Y.; Rozhdestvensky, T.S.; Schmolke, M.; Resa-Infante, P.; Robeck, T.; Randau, G.; Wolff, T.; Gabriel, G.; Brosius, J.; Ludwig, S. Evidence for a Novel Mechanism of Influenza Virus-Induced Type I Interferon Expression by a Defective RNA-Encoded Protein. PLoS Pathog. 2015, 11, e1004924. [Google Scholar] [CrossRef] [PubMed]
  170. Sutton, T.C.; Finch, C.; Shao, H.; Angel, M.; Chen, H.; Capua, I.; Cattoli, G.; Monne, I.; Perez, D.R. Airborne transmission of highly pathogenic H7N1 influenza virus in ferrets. J. Virol. 2014, 88, 6623–6635. [Google Scholar] [CrossRef] [PubMed]
  171. Ganti, K.; Bagga, A.; Carnaccini, S.; Ferreri, L.M.; Geiger, G.; Joaquin Caceres, C.; Seibert, B.; Li, Y.; Wang, L.; Kwon, T.; et al. Influenza A virus reassortment in mammals gives rise to genetically distinct within-host subpopulations. Nat. Commun. 2022, 13, 6846. [Google Scholar] [CrossRef] [PubMed]
  172. Li, C.; Chen, H. Enhancement of influenza virus transmission by gene reassortment. Curr. Top. Microbiol. Immunol. 2014, 385, 185–204. [Google Scholar] [PubMed]
  173. White, M.C.; Lowen, A.C. Implications of segment mismatch for influenza A virus evolution. J. Gen. Virol. 2018, 99, 3–16. [Google Scholar] [CrossRef] [PubMed]
  174. Zhao, Z.; Liu, L.; Guo, Z.; Zhang, C.; Wang, Z.; Wen, G.; Zhang, W.; Shang, Y.; Zhang, T.; Jiao, Z.; et al. A Novel Reassortant Avian H7N6 Influenza Virus Is Transmissible in Guinea Pigs via Respiratory Droplets. Front. Microbiol. 2019, 10, 18. [Google Scholar] [CrossRef] [PubMed]
  175. Octaviani, C.P.; Goto, H.; Kawaoka, Y. Reassortment between seasonal H1N1 and pandemic (H1N1) 2009 influenza viruses is restricted by limited compatibility among polymerase subunits. J. Virol. 2011, 85, 8449–8452. [Google Scholar] [CrossRef] [PubMed]
  176. Zhang, Y.; Zhang, Q.; Kong, H.; Jiang, Y.; Gao, Y.; Deng, G.; Shi, J.; Tian, G.; Liu, L.; Liu, J.; et al. H5N1 hybrid viruses bearing 2009/H1N1 virus genes transmit in guinea pigs by respiratory droplet. Science 2013, 340, 1459–1463. [Google Scholar] [CrossRef] [PubMed]
  177. Naiqing, X.; Tang, X.; Wang, X.; Cai, M.; Liu, X.; Lu, X.; Hu, S.; Gu, M.; Hu, J.; Gao, R.; et al. Hemagglutinin affects replication, stability and airborne transmission of the H9N2 subtype avian influenza virus. Virology 2024, 589, 109926. [Google Scholar] [CrossRef]
  178. Hao, M.; Han, S.; Meng, D.; Li, R.; Lin, J.; Wang, M.; Zhou, T.; Chai, T. The PA Subunit of the Influenza Virus Polymerase Complex Affects Replication and Airborne Transmission of the H9N2 Subtype Avian Influenza Virus. Viruses 2019, 11, 40. [Google Scholar] [CrossRef] [PubMed]
  179. Iwasaki, A.; Pillai, P.S. Innate immunity to influenza virus infection. Nat. Rev. Immunol. 2014, 14, 315–328. [Google Scholar] [CrossRef] [PubMed]
  180. Goraya, M.U.; Zaighum, F.; Sajjad, N.; Anjum, F.R.; Sakhawat, I.; Rahman, S.U. Web of interferon stimulated antiviral factors to control the influenza A viruses replication. Microb. Pathog. 2020, 139, 103919. [Google Scholar] [CrossRef] [PubMed]
  181. Shim, J.M.; Kim, J.; Tenson, T.; Min, J.Y.; Kainov, D.E. Influenza Virus Infection, Interferon Response, Viral Counter-Response, and Apoptosis. Viruses 2017, 9, 223. [Google Scholar] [CrossRef]
  182. Haller, O.; Kochs, G. Human MxA protein: An interferon-induced dynamin-like GTPase with broad antiviral activity. J. Interferon Cytokine Res. 2011, 31, 79–87. [Google Scholar] [CrossRef] [PubMed]
  183. Haller, O.; Kochs, G. Mx genes: Host determinants controlling influenza virus infection and trans-species transmission. Hum. Genet. 2020, 139, 695–705. [Google Scholar] [CrossRef] [PubMed]
  184. Liu, B.; Li, N.L.; Shen, Y.; Bao, X.; Fabrizio, T.; Elbahesh, H.; Webby, R.J.; Li, K. The C-Terminal Tail of TRIM56 Dictates Antiviral Restriction of Influenza A and B Viruses by Impeding Viral RNA Synthesis. J. Virol. 2016, 90, 4369–4382. [Google Scholar] [CrossRef] [PubMed]
  185. LeMessurier, K.S.; Rooney, R.; Ghoneim, H.E.; Liu, B.; Li, K.; Smallwood, H.S.; Samarasinghe, A.E. Influenza A virus directly modulates mouse eosinophil responses. J. Leukoc. Biol. 2020, 108, 151–168. [Google Scholar] [CrossRef] [PubMed]
  186. Nogales, A.; Martinez-Sobrido, L.; Topham, D.J.; DeDiego, M.L. Modulation of Innate Immune Responses by the Influenza A NS1 and PA-X Proteins. Viruses 2018, 10, 708. [Google Scholar] [CrossRef] [PubMed]
  187. Wang, L.; Fu, X.; Zheng, Y.; Zhou, P.; Fang, B.; Huang, S.; Zhang, X.; Chen, J.; Cao, Z.; Tian, J.; et al. The NS1 protein of H5N6 feline influenza virus inhibits feline beta interferon response by preventing NF-κB and IRF3 activation. Dev. Comp. Immunol. 2017, 74, 60–68. [Google Scholar] [CrossRef] [PubMed]
  188. Chaimayo, C.; Dunagan, M.; Hayashi, T.; Santoso, N.; Takimoto, T. Specificity and functional interplay between influenza virus PA-X and NS1 shutoff activity. PLoS Pathog. 2018, 14, e1007465. [Google Scholar] [CrossRef]
  189. Nogales, A.; Martinez-Sobrido, L.; Chiem, K.; Topham, D.J.; DeDiego, M.L. Functional Evolution of the 2009 Pandemic H1N1 Influenza Virus NS1 and PA in Humans. J. Virol. 2018, 92, e01206-18. [Google Scholar] [CrossRef] [PubMed]
Table 1. Requirements and corresponding mechanisms for cross-species transmission and adaptation of influenza A viruses to new hosts.
Table 1. Requirements and corresponding mechanisms for cross-species transmission and adaptation of influenza A viruses to new hosts.
RequirementsCorresponding Mechanisms
① Influenza virions can attach to the appropriate cells in the upper respiratory tract and replicate efficiently.Alteration in HA receptor binding specificity to facilitate viral infection.
Enhancement of viral polymerase activity to increase the number of viruses in the respiratory tract and efficiently releases them into the air.
② Virions are relatively stable and remain infectious while transiting between and within hosts.The occurrence of adaptive mutations to improve HA stability.
③ Virions can be efficiently released outside the cell, leaving the upper respiratory tract as single particles and being expelled in large numbers.Maintenance of the functional balance of HA-NA to produce single-particle viruses
①②③Reassortment of genes in different segments to help the virus be airborne.
Table 2. The HA mutation sites that affect HA receptor binding specificity and transmission of the influenza A virus.
Table 2. The HA mutation sites that affect HA receptor binding specificity and transmission of the influenza A virus.
HostViral SubtypesMutations in HAReceptor BindingImpact on Viral TransmissibilityReference
Guinea pigH9N2T187P + M227LEnhanced affinity for α-2,6 receptorsEffective transmission between guinea pigs by direct contact[80]
H1N1E225GReduced binding affinity forα-2,6 receptors.The ability of droplet transmission between guinea pigs is lost[81]
Guinea pig and ferretH9N2Q226LReceptor preference changed from α-2,3 to α-2,6 receptorsEffective transmission between guinea pigs through direct contact and airborne routes[75]
Guinea pig and pigH3N2E190D V226I G228SReceptor preference changed from α-2,3 to α-2,6 receptorsEffective transmission by contact[82]
Guinea pigH10N3G228SAcquisition of dual receptor affinity for α-2,3 and α-2,6Effective transmission between guinea pigs by contact and aerosols[76]
SealH3N8A134TDeveloped affinity for α-2,6 receptorsEffective transmission through respiratory droplets
(Transmission rates: 100%)
[72]
DogH5N6Q226L
Glycosylation deletion at locus 158
Increased affinity for α-2,6 receptorsCrosses the mammalian host barrier and is capable of infecting dogs[83]
FerretH3N8G228SEnlarged affinity for α-2,6 receptorsSimultaneous introduction of the HA G228S and PB2 E627K mutation sites can cause viral droplet transmission[71]
H5N1Q226S G228SReceptor preference changed from α-2,3 to α-2,6 receptorsEffective airborne transmission[73]
H9N2I155T H183N A190VImproved affinity for α-2,6 receptorsEffective airborne transmission[84]
Note: Glycine (G), Alanine (A), Valine (V), Leucine (L), Isoleucine (I), Phenylalanine (F), Proline (P), Threonine (T), Tryptophane (W), Serine (S), Tyrosine (Y), Cysteine (C), Methionine (M), Aspartic acid (D), Asparagine (N), Glutamic acid (E), Glutamine (Q), Lysine (K), Arginine (R), Histidine (H).
Table 3. The HA mutation sites that affect HA acid stability and transmission of the influenza A virus.
Table 3. The HA mutation sites that affect HA acid stability and transmission of the influenza A virus.
HostViral SubtypesMutations in HAChanges in Membrane Fusion pHImpact on Viral TransmissibilityReference
ChickenH7N9D167N (H7 numbering)Reduced HA stabilityCannot be transmitted between chickens by air[107]
H9N2K363R (H9 numbering)Reduced HA stabilityDecreased seroconversion rate, Decreased airborne transmission between chickens[108]
PigH1N1HA1-Y17H
HA2-R106K
5.5→6.0Efficiently transmitted between pigs by contact (Transmission rates: 100%)[109]
Pig→Ferret (interspecies transmission)H1N1HA1-Y17H
HA2-R106K
5.5→5.3Effective airborne transmission between pigs and ferrets (Transmission rates: 100%)[109]
Mouse and FerretH1N1HA1-H17Y
HA2-R106K
6.0→5.3Virus regains airborne capacity[101]
FerretH3N2G78D5.5→5.8Reduced airborne efficiency[102]
H1N1HA1-N210S
HA2-T117N
5.8→5.5
5.9→5.6
Improved airborne transmission between ferrets[103]
H3N2HA1-L194P<5.5→>5.5The ability of airborne transmission between ferrets is lost (Transmission rates: 100%→0%)[110]
H5N1H103Y (H5 numbering)≤5.6→≤5.5Simultaneous introduction of five mutation sites raises airborne efficiency[73]
H10N7T244I HA2-E74D5.7→5.2Transmission between ferrets by aerosols or respiratory droplets[111]
H9N2HA1-Y17H5.8→5.4Loss of airborne transmission, only through contact (Less efficient dissemination)[112]
Note: HA1 and HA2 numbering are the N termini of the HA2 glycoprotein after HA0 cleavage.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, X.; Zhou, Y.; Yan, H.; An, Q.; Liang, C.; Liu, L.; Qian, J. Molecular Markers and Mechanisms of Influenza A Virus Cross-Species Transmission and New Host Adaptation. Viruses 2024, 16, 883. https://doi.org/10.3390/v16060883

AMA Style

Guo X, Zhou Y, Yan H, An Q, Liang C, Liu L, Qian J. Molecular Markers and Mechanisms of Influenza A Virus Cross-Species Transmission and New Host Adaptation. Viruses. 2024; 16(6):883. https://doi.org/10.3390/v16060883

Chicago/Turabian Style

Guo, Xinyi, Yang Zhou, Huijun Yan, Qing An, Chudan Liang, Linna Liu, and Jun Qian. 2024. "Molecular Markers and Mechanisms of Influenza A Virus Cross-Species Transmission and New Host Adaptation" Viruses 16, no. 6: 883. https://doi.org/10.3390/v16060883

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop