Next Article in Journal
Mouse Models for Filovirus Infections
Next Article in Special Issue
Human Leukocyte Antigen (HLA) Class I Down-Regulation by Human Immunodeficiency Virus Type 1 Negative Factor (HIV-1 Nef): What Might We Learn From Natural Sequence Variants?
Previous Article in Journal
Genetics-Based Classification of Filoviruses Calls for Expanded Sampling of Genomic Sequences
Previous Article in Special Issue
Herpesviruses Placating the Unwilling Host: Manipulation of the MHC Class II Antigen Presentation Pathway
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Evasion of Influenza A Viruses from Innate and Adaptive Immune Responses

Department of Virology, ErasmusMC, Dr. Molewaterplein 50, 3015 GE, Rotterdam, The Netherlands
*
Author to whom correspondence should be addressed.
Viruses 2012, 4(9), 1438-1476; https://doi.org/10.3390/v4091438
Submission received: 3 July 2012 / Revised: 10 August 2012 / Accepted: 22 August 2012 / Published: 3 September 2012
(This article belongs to the Special Issue Immune Evasion)

Abstract

:
The influenza A virus is one of the leading causes of respiratory tract infections in humans. Upon infection with an influenza A virus, both innate and adaptive immune responses are induced. Here we discuss various strategies used by influenza A viruses to evade innate immune responses and recognition by components of the humoral and cellular immune response, which consequently may result in reduced clearing of the virus and virus-infected cells. Finally, we discuss how the current knowledge about immune evasion can be used to improve influenza A vaccination strategies.

Graphical Abstract

1. Introduction

Influenza viruses belong to the family of orthomyxoviridae and are one of the leading causes of respiratory tract infections in humans [1]. Yearly, influenza viruses cause an estimated 3–5 million severe clinical infections and 250,000–500,000 fatal cases [2,3]. The genome of influenza A viruses consists of eight gene segments of single-stranded negative-sense RNA encoding 12 proteins: surface glycoproteins, hemaglutinin (HA) and neuraminidase (NA), two matrix proteins (M1 and M2), the nucleoprotein (NP), three polymerase complex proteins PB1, PB2 and PA, and four non-structural proteins NS1, NS2, PA-X and PB1-F2 [1,4].
Influenza A viruses are subdivided, based on their surface glycoproteins; currently 17 subtypes of HA and nine subtypes of NA are known [5,6,7]. Seasonal influenza A viruses of the H1N1 and H3N2 subtype and influenza B viruses cause yearly epidemics [2]. This can be attributed to their ability to escape recognition by virus specific antibodies due to antigenic drift (Figure 1). As a result, seasonal influenza vaccines need to be updated almost annually to match the new epidemic strains and to remain efficacious [8].
Figure 1. Antigenic drift and shift to escape immunity. The gradual accumulation of mutations, mainly in the highly variable globular head region of HA, causes the influenza virus to escape recognition by virus neutralizing antibodies and allows it to cause seasonal epidemic outbreaks. This phenomenon is called antigenic drift. The introduction of a novel subtype into the human population is called antigenic shift and may cause a pandemic outbreak in the naïve human population when the virus is efficiently transmitted from human to human, since antibodies directed against the novel subtype are absent. Past pandemic outbreaks were caused by exchange (re-assortment) of gene segments between two or more influenza strains, e.g., avian and human. However, recent studies in ferrets suggest that avian influenza viruses, like H5N1, could be directly transmitted from animal reservoirs into the human population, requiring only a small number of adaptive mutations [9] as indicated by the dotted line in this figure.
Figure 1. Antigenic drift and shift to escape immunity. The gradual accumulation of mutations, mainly in the highly variable globular head region of HA, causes the influenza virus to escape recognition by virus neutralizing antibodies and allows it to cause seasonal epidemic outbreaks. This phenomenon is called antigenic drift. The introduction of a novel subtype into the human population is called antigenic shift and may cause a pandemic outbreak in the naïve human population when the virus is efficiently transmitted from human to human, since antibodies directed against the novel subtype are absent. Past pandemic outbreaks were caused by exchange (re-assortment) of gene segments between two or more influenza strains, e.g., avian and human. However, recent studies in ferrets suggest that avian influenza viruses, like H5N1, could be directly transmitted from animal reservoirs into the human population, requiring only a small number of adaptive mutations [9] as indicated by the dotted line in this figure.
Viruses 04 01438 g001
Introduction in the human population of influenza A viruses with antigenically distinct HA molecules, including novel subtypes, is known as antigenic shift (Figure 1). When such an antigenically-distinct virus is transmitted efficiently from human to human, it may cause a pandemic influenza outbreak, since neutralizing antibodies to this virus are absent in the population at large [10]. Examples of pandemics include the Spanish flu of 1918 caused by a virus of the H1N1 subtype which killed 20–50 million people [11], the Asian flu of 1957 caused by a re-assorted H2N2 virus and the Hong Kong flu of 1968 caused by a re-assorted H3N2 virus. Each time, the pandemic virus replaced the subtype that circulated prior to the pandemic. In 1977, a virus of the H1N1 subtype was re-introduced, which did not result in a major pandemic, and this virus seasonally co-circulated with the H3N2 virus subtype until 2009. The first pandemic outbreak of the 21st century occurred in 2009 when a novel H1N1 re-assorted virus of swine origin was introduced into the human population [12]. In addition, other influenza virus subtypes are transmitted from animals (in particular swine and avian influenza viruses) to humans occasionally [13,14]. For example, in 1999, there were three isolated cases of influenza A/H9N2 virus infections in humans displaying mild symptoms only [15]. During an outbreak of highly pathogenic avian influenza A virus of the H7N7 subtype in the Netherlands in 2003, 89 human cases were reported of which one was fatal [16,17]. Larger is the impact of highly pathogenic avian influenza A/H5N1 viruses which have been transmitted on a regular basis from infected poultry to man since the first case was identified in 1997 in Hong Kong [18]. Since 2003, over 600 human cases have been reported, most of them suffering from severe pneumonia progressing to acute respiratory distress syndrome, 60% of the cases had a fatal outcome [19,20,21,22,23]. The reported case fatality rate most likely is an overestimate, since subclinical infections and mild cases are not reported [24]. So far, efficient human-to-human transmission has not been observed, although clusters of human cases have been reported [25,26,27]. Furthermore, recent studies have shown that, in principle, transmission of highly pathogenic H5N1 viruses amongst mammals is possible and that only a limited number of adaptive mutations are required for airborne transmission, emphasizing the pandemic potential of these viruses [9,28,29].
Without lifelong protection against seasonal influenza virus infections and the threat of possible future pandemics, it is of great importance to have insight in how immunity against influenza A infections is formed and how influenza A viruses manage to evade these immune responses. In this review, we describe the role of innate and adaptive immunity against influenza A virus infections and evasion strategies used by influenza A viruses to escape immunity. Finally, we briefly discuss the impact on influenza A vaccine development.

2. Innate Immunity

The primary targets for influenza viruses are the epithelial cells that line the respiratory tract and which initiate an antiviral immune response upon detection of the virus. The first line of defense is formed by the innate immune system, which is quick but lacks specificity and memory. Innate immunity is formed by physical barriers (e.g., mucus and collectins) and innate cellular immune responses [30]. Here, we outline several of the innate defense mechanisms directed against influenza A infections.

2.1. Sensing Of Influenza Virus Infection by Receptors of the Innate Immune System

Influenza A virus infection results in the recognition of pathogen-associated molecular patterns (PAMPs) by pattern recognition receptors (PRRs) that initiate antiviral signaling cascades, resulting in the production of interferons (IFNs), cytokines and chemokines [31]. Three main categories of PRRs are involved in recognition of an influenza A infection and the induction of an IFN response: Toll like receptors (TLRs), retinoic acid inducible gene-I (RIG-I) receptors and nucleotide oligomerization domain(NOD)-like receptor family pyrin domain containing 3 (NLRP3) [32,33,34].
The TLRs are the first receptors to recognize the influenza virus infection. TLR7 is an intracellular receptor that recognizes single stranded viral RNA (ssRNA) after the ribonucleoprotein complex has been degraded inside acidified endosomes [35,36]. TLR3 is another intracellular receptor that recognizes double stranded viral RNA (dsRNA) [37]. Other TLR receptors likely to sense an influenza virus infection are TLR2 and TLR4, which are present on the cell surface and recognize viral surface glycoproteins like influenza HA and NA [34,38,39,40]. At a later stage of infection, newly produced uncapped, 5'-triphosphates bearing viral RNAs are recognized by RIG-I receptors in the cytoplasm [41,42,43,44].
NLRP3 is part of the NLRP3 inflammasome and is activated by dsRNA which subsequently activates caspase I, resulting in the proteolytic maturation of IL-1β and IL-18 [45].
The signaling cascade of all activated TLRs, except for TLR3, starts with the activation of MyD88, which subsequently can activate tumor necrosis factor (TNF) receptor associated factor 6 (TRAF6), either directly or via IL-1R-associated kinase-1 (IRAK 1), eventually leading to the activation of mitogen-activated kinases (MAPKs) and nuclear factor kappa-light-chain-enhancer of activated B cells (NF-κB). TLR3 signaling cascade starts with the activation of TRIF (TIR-domain-containing adapter‑inducing interferon-β) eventually activating NF-κB and interferon regulatory factor 3 (IRF3). TLR4 can also signal via the TRIF dependent pathway by formation of a TRAM-TRIF complex.
Activation of RIG-I receptor by binding of viral 5'-triphosphates RNA or viral dsRNA to the repressor domain (RD) of RIG-I results in conformational changes exposing the caspase activation and recruitment domains (CARDs). These domains are ubiquitinated by IFN-inducible E3 ubiquitin ligase, tripartite motif 25 (TRIM25) [46]. RIG-I can then associate with mitochondrial antiviral signaling adaptor (MAVS; also known as IPS-1, VISA or Cardif), starting a signaling cascade that leads to the activation of IRF3 and NF-κB. The signaling cascades via TLRs and RIG-I receptors have been extensively reviewed by others [44,47,48,49] (Figure 2).
All these pathways eventually result in the transcription of proinflammatory cytokines, chemokines and IFNs that activate the antiviral response and the recruitment of neutrophils, activation of macrophages and maturation of dendritic cells (DCs) [31]. So far, three IFN types have been identified [50]. Type I IFNs include IFN-α and IFN-β which have an important role in limiting viral replication [51,52]. Type I interferons secreted by infected cells act on IFN-α/β receptors (IFN-α/βR) of the same cell or neighboring cells, activating an antiviral signaling cascade that involves phosphorylation of tyrosine kinase 2 (Tyk2) and Janus kinase 1 (Jak1), also called, “just another kinase 1”, which then phosphorylate signal transducer and activators of transcription (STAT) 1 and STAT2. Phosphorylated STAT1 and 2 combine with IRF9 to form ISGF3 (IFN-stimulated gene factor-3 transcription factor complex) which is responsible for the transcription of >300 genes that encode for e.g., antiviral proteins (Table 1) that establish an antiviral state in the cell that limits viral replication [49] (Figure 3). IFN-β acts through a positive feedback loop on the IFN-β receptor which activates IFN stimulated gene factor 3 (ISGF3), resulting in the expression of IRF-7. IRF-7 is phosphorylated in the presence of a viral infection and induces the expression of both IFN-α and IFN-β [53].
Figure 2. The RIG-I signaling pathway and inhibition by influenza A viruses (Figure adapted from [48]). By-products of viral replication are 5'-triphosphates ssRNA and dsRNA which can bind to the RIG-1 receptor, leading to conformational changes, causing exposure of the CARDS which are ubiqutinated by TRIM25. Subsequently, RIG-1 associates with MAVS and thereby starts a signaling cascade leading to activation of transcription factors IRF3, NF-κB and ATF-2/JunC, resulting in the transcription of IFN-β mRNA. Indicated in red are sites at which the influenza A virus interferes with this pathway, as explained in the text.
Figure 2. The RIG-I signaling pathway and inhibition by influenza A viruses (Figure adapted from [48]). By-products of viral replication are 5'-triphosphates ssRNA and dsRNA which can bind to the RIG-1 receptor, leading to conformational changes, causing exposure of the CARDS which are ubiqutinated by TRIM25. Subsequently, RIG-1 associates with MAVS and thereby starts a signaling cascade leading to activation of transcription factors IRF3, NF-κB and ATF-2/JunC, resulting in the transcription of IFN-β mRNA. Indicated in red are sites at which the influenza A virus interferes with this pathway, as explained in the text.
Viruses 04 01438 g002
Figure 3. The type I IFN signaling pathway and inhibition by influenza A viruses (Figure adapted from [48]). IFN-β produced by influenza virus-infected cells binds IFN receptors causing the phosphorylation of Tyk2 and Jak1. This is followed by binding and phosphorylation of STAT1 and STAT2 which subsequently form a complex with IRF9. This ISGF-3 complex acts as a transcription factor for >300 genes, several of which display an antiviral effect (see text). The expressed protein PKR is activated upon recognition of viral dsRNA, leading to inhibition of protein synthesis, including viral proteins. PKR is inhibited by the cellular protein P58IPK, however P58IPK activity is downregulated by binding cellular hsp40. The IRF7 protein is phosphorylated in the presence of influenza A virus, leading to activation of a positive feedback loop, causing the transcription of IFN-α and IFN-β. Indicated in red are mechanisms of the influenza A virus to interfere with this pathway, these interfering mechanisms are explained more extensively in the text.
Figure 3. The type I IFN signaling pathway and inhibition by influenza A viruses (Figure adapted from [48]). IFN-β produced by influenza virus-infected cells binds IFN receptors causing the phosphorylation of Tyk2 and Jak1. This is followed by binding and phosphorylation of STAT1 and STAT2 which subsequently form a complex with IRF9. This ISGF-3 complex acts as a transcription factor for >300 genes, several of which display an antiviral effect (see text). The expressed protein PKR is activated upon recognition of viral dsRNA, leading to inhibition of protein synthesis, including viral proteins. PKR is inhibited by the cellular protein P58IPK, however P58IPK activity is downregulated by binding cellular hsp40. The IRF7 protein is phosphorylated in the presence of influenza A virus, leading to activation of a positive feedback loop, causing the transcription of IFN-α and IFN-β. Indicated in red are mechanisms of the influenza A virus to interfere with this pathway, these interfering mechanisms are explained more extensively in the text.
Viruses 04 01438 g003
Table 1. IFN-induced antiviral proteins and their function.
Table 1. IFN-induced antiviral proteins and their function.
ProteinFunction Reference
MxA (Myxovirus resistance gene A)Inhibits viral replication by interfering with the viral ribonucleoprotein structure[54,55,56]
PKR (Protein kinase R)Limits viral replication by blocking general translation [57,58]
OAS (2'–5'oligoadenylate synthetase)Stops viral replication by means of activating RNAseL which results in degradation of viral and cellular RNA and eventually apoptosis of the virus infected cell[59,60]
ISG15 (IFN-stimulated gene 15)Regulates a number of IFN-stimulated proteins[61]
Viperin Inhibits viral release by interfering with viral budding[62]
TetherinInhibits formation of influenza virus particles[63,64]
IFITMsRestrict viral entry[65]
IFN-γ is the main type II IFN and contributes to the establishment of an effective adaptive cytotoxic T cell (CTL) response against the influenza virus infection. It regulates virus-specific CTL homeostasis in secondary lymph nodes and subsequent trafficking of CTLs to the site of infection [66]. Furthermore, IFN-γ plays an important role in memory CTL responses [67]. Type III IFNs, like IFN-λ, also control influenza A infections in the lung [68].

2.2. Macrophages

During homeostasis, alveolar macrophages exhibit a relatively quiescent state, producing only low levels of cytokines, and suppress the induction of innate and adaptive immunity [69,70]. CCL2, produced by infected epithelial cells during the initial phase of the influenza virus infection, attracts alveolar macrophages and monocytes via their CCR2 receptor [71,72,73]. Activated macrophages enhance their pro-inflammatory cytokine response, including IL-6 and TNF-α [74,75]. Alveolar macrophages have a direct role in limiting viral spread by phagocytosis of apoptotic infected cells [40,76,77] and by phagocyte-mediated opsonophagocytosis of influenza virus particles [78]. They are also involved in regulating the adaptive immune response. Depletion of alveolar macrophages prior to influenza virus infection led to a reduction of antibody titers and reduced numbers of virus-specific CTLs post-infection [77]. In contrast to these beneficial effects, alveolar macrophages also pose a negative effect, since their activation also results in the production of nitric oxide synthase 2 (NOS2) and TNF-α which contribute to the severe pathology that can be the result of an influenza virus infection [71,79,80].

2.3. Natural Killer Cells

Natural killer (NK) cells are cytotoxic lymphocytes of the innate immune system. They are able to lyse infected cells in a MHC class I independent manner via a direct or indirect mechanism of recognition. The sialylated NKp44 and NKp46 receptors are bound by the HA proteins expressed on the surface of influenza virus-infected cells [81]. This results in direct lysis of the infected cell [82,83]. It was shown that mice lacking the NKp46 receptor equivalent, NCR-1, displayed increased morbidity and mortality following influenza A infection [84]. NK cells with their CD16 receptor (FcγRIII) can bind to the Fc portion of antibodies bound to influenza virus-infected cells and mediate lysis of these cells. This process is known as antibody-dependent cell cytotoxicity (ADCC) [85,86,87].

2.4. Dendritic Cells

Dendritic cells (DCs) are professional antigen presenting cells (APCs) which form an important bridge between the innate and the adaptive immune system. During an infection, DCs initiate adaptive immune responses by the presentation of viral antigens to naïve and memory T and B lymphocytes (Figure 4) [69,88,89]. At steady state conditions, DCs constantly survey the lungs for invading pathogens or foreign material [90]. Once the lungs are infected with influenza A virus, the DCs can acquire the antigens via two distinct mechanisms. The first route is by direct infection of DCs by influenza A virus [91,92]. Proteasomes in the cytosol degrade viral proteins into small peptides which are transported to the endoplasmic reticulum (ER) via TAP (transporter of antigen processing), where they are loaded to MHC class I molecules. These MHC class I peptide complexes are then transported via the Golgi complex onto the cell membrane where they can be recognized by virus-specific CD8+ cytotoxic T cells (CTLs) (Figure 5) [88,93]. The second mechanism of antigen acquisition by DCs is through phagocytosis of virus particles or apoptotic epithelial cells [88,94,95,96]. Viral proteins are then degraded into smaller peptides in endosomes/lysosomes and presented on the cell surface in MHC class II peptide complexes which can be recognized by CD4+ T helper cells. T helper cells assist B cells to proliferate and mature into antibody-producing plasma cells. Via this route of antigen acquisition, DCs can also present epitopes to CD8+ T cells. This is also known as cross-presentation. For presentation of viral antigens to virus-specific T cells, activated DCs migrate to the draining lymph nodes [90,97,98,99].

3. Adaptive Immunity

The second line of defense against influenza A virus infection is the adaptive immune response. Overall, this highly specific response is relatively slow upon first encounter with a pathogen. However, as a result of the formation of immunological memory, the response is faster and stronger after a second encounter with the same pathogen. The adaptive immune response consists of humoral (virus‑specific antibodies) and cellular (virus-specific CD4+ and CD8+ T cells) immunity. Here we summarize virus-specific recognition by components of the adaptive immune system and their contribution to clearance of influenza A virus infections.

3.1. Humoral Immunity

Influenza virus infection induces the production of influenza virus-specific antibodies by B cells [100,101,102,103]. In particular, antibodies directed to the viral HA and NA correlate with protective immunity.
Antibodies directed to the trimeric globular head of HA can afford sterilizing immunity to influenza virus infection. By binding to the HA receptor binding site located in this region they can block virus attachment to host cells and/or block receptor-mediated endocytosis [104,105,106]. However, most antibodies directed against HA are influenza virus strain-specific and fail to neutralize intrasubtypic drift variants and viruses of other subtypes [6,10,107,108,109]. This is mainly due to the high variability in the HA globular head (see below). Of interest, humoral immunity elicited after an influenza virus infection does provide long-lasting antibody mediated protection against the strains that resemble the infected strain. This was exemplified during the pandemic of 2009 caused by an influenza virus of the H1N1 subtype. Elderly people that were exposed to influenza A/H1N1 virus in the 1950s had antibodies which cross-reacted with the pandemic strain and were relatively spared from contracting infections and developing disease [109,110,111,112]. Recently, broadly reacting antibodies directed against the conserved stem region of HA have been identified [113,114,115,116,117,118,119,120].
Figure 4. Induction of humoral and cellular immunity. Induction of immune responses after a primary influenza A virus infection is indicated by solid arrows. The more rapid activation of virus-specific memory cell populations upon secondary encounter with an influenza A virus are indicated by dotted arrows.
Figure 4. Induction of humoral and cellular immunity. Induction of immune responses after a primary influenza A virus infection is indicated by solid arrows. The more rapid activation of virus-specific memory cell populations upon secondary encounter with an influenza A virus are indicated by dotted arrows.
Viruses 04 01438 g004
Antibodies to the other major glycoprotein, the viral NA, interfere with the last phase of the viral replication cycle and also exert protective immunity. NA is a sialydase and removes sialic acids from infected cells and budded virions, thereby facilitating efficient release and spread of newly formed viral particles [1]. Unlike HA-specific antibodies, NA-specific antibodies do not neutralize the virus. However, by inhibiting the enzymatic activity of NA, these antibodies limit the viral spread and thus shorten severity and duration of illness [121,122,123,124,125]. Furthermore, NA-specific antibodies may also contribute to clearance of virus-infected cells by facilitating ADCC [85].
In addition to HA and NA, influenza virus particles contain the minor glycoprotein M2. This tetrameric transmembrane protein has ion channel activity and plays an important role in unpacking the virus in the endosome [1]. A protective effect of M2-specific antibodies was first demonstrated in mice after passive transfer of M2-specific monoclonal antibodies [126,127]. M2-specific antibodies facilitate ADCC [87,128,129]. Since the M2 protein is highly conserved between various influenza A virus subtypes, M2-specific antibodies are likely to afford heterosubtypic immunity [130,131,132,133,134,135,136,137].
After infection, antibodies are also induced against other viral proteins, including NP [138]. Since NP is highly conserved between influenza A viruses, these antibodies could potentially contribute to heterosubtypic immunity. Although NP-specific antibodies are non-neutralizing, it was shown in mice that they contribute to protective immunity [139,140]. However, their mode of action is poorly understood, but may include ADCC of infected cells and opsonisation of NP, resulting in improved T cell responses [141,142].
After primary infection with influenza virus, serum antibodies of the IgM, IgA and IgG isotypes are induced, whereas after secondary responses, IgM responses are not observed [143]. IgM antibodies can neutralize the virus, but also activate the complement system [144,145]. In humans, virus-specific serum IgA responses seem indicative for recent infection with influenza virus [146,147]. Virus‑specific IgG antibodies correlate with long-lived protection, provided that the antibodies match the strains causing the infection [148,149,150]. In addition to serum antibodies, influenza virus infection also induces local mucosal sIgA antibody responses that protect airway epithelial cells from infection [149,151,152].
Young infants may be protected from influenza virus infection by maternal antibodies, when they match the incoming virus [153,154,155,156,157].

3.2. Cellular Immunity

Influenza A virus infection induces a cellular immune response, including virus-specific CD4+ T cells and CD8+ T cells. These cells play an important role in regulation of the immune response and viral clearance respectively.

3.2.1. CD4+ T Cells

CD4+ T cells are activated after recognition of viral epitopes associated with MHC class II molecules and interaction with co-stimulatory molecules on APCs. Depending on the cytokine milieu, activation of naïve CD4+ T cells can result in the differentiation into CD4+ T helper 1 cells (Th1) or Th2, which can be distinguished based on their cytokine expression profiles [158]. Th2 cells produce IL-4, IL-5 and IL-13 and promote the activation and differentiation of B cells, resulting in antibody production [159,160,161]. The antibody response is strengthened by the induction of antibody class switching and somatic hypermutations affecting the variable region of the antibody, resulting in affinity maturation of the influenza-specific antibodies [162,163,164,165]. Th1 cells produce IFN-γ and IL-2 and are mainly involved in promoting CTL responses [166,167], and are essential for the induction of memory CD8+ T cells [168,169,170]. Memory CD4+ T cells, induced after a primary influenza A virus infection, contribute to faster control of subsequent influenza A virus infections [171]. Lung-resident memory CD4+ T cells in particular have an important role in protection against secondary influenza A virus infections [172]. In addition to helper function, CD4+ T cells also display cytolytic activity [173,174]. It was shown that these cells play a role in protective immunity against influenza A virus infections in humans [175]. A more extensive review on the role of CD4+ T cells in heterosubtypic immunity can be found elsewhere [176].

3.2.2. CD8+ T Cells

Naïve CD8+ T cells are activated after recognition of viral epitopes associated with MHC class I molecules on APCs in the draining lymph nodes, and subsequently differentiate into CTLs. These CTLs migrate to the site of infection where they recognize and eliminate influenza virus infected cells and thereby prevent the production and spread of progeny virus [177]. Human influenza virus-specific CTLs are mainly directed against epitopes of the highly conserved internal viral proteins, like M1, NP, PA and PB2. Therefore, CTLs display a high degree of cross-reactivity with influenza A viruses of various subtypes [178,179,180,181,182,183,184]. T cell receptor (TCR) activation by a specific epitope-MHC class I complex results in a lytic response, mediated by the release of perforin and granzymes causing apoptosis of the infected cell [185,186,187]. Furthermore, proinflammatory cytokines are produced, like TNF-α, which also inhibit virus replication and enhance lytic activity [185,188,189,190,191]. Also, FasL expression is upregulated which promotes apoptosis of infected cells [187]. After infection, a pool of long-lived antigen-specific central memory and effector memory CD8+ T cells is formed, which form the basis for more rapid and stronger recall responses upon secondary infections [192,193,194,195,196,197,198,199,200,201].
Much of the current knowledge about the protective role of CD8+ T cells in influenza A virus infections has been obtained from mouse studies which showed that CD8+ T cells contribute to homo- and heterosubtypic immunity [202,203,204,205,206,207,208,209]. Evidence that CTLs protect against influenza in humans is sparse. A recent study indicated the presence of heterosubtypic memory CD4+ and CD8+ T cells against the 2009 pandemic H1N1 virus in naïve individuals [210]. It was shown that the extent of lytic activity of PBMC inversely correlated with the extent of virus shedding after experimental infection of subjects that lacked antibodies to the life-attenuated strain used for infection [211]. More circumstantial evidence for a protective role of CD8+ T cells in heterosubtypic influenza infections in humans stems from epidemiological studies. People who had a symptomatic influenza A infection with the H1N1 strain prior to the 1957 pandemic were partially protected from infection with the pandemic H2N2 strain [212,213]. A similar trend was found in isolated infections with the H5N1 [214].

3.2.3. Regulatory T Cells and Th17 Cells

In addition to the activation of virus-specific CD4+ T cells and CD8+ T cells, two other cell types are activated, namely forkhead box P3 (FOXP3)+ regulatory T cells (Tregs) and T helper 17 cells (Th17). Tregs play an important role in balancing the immune response during infection. They control CD4+ T helper cell and CTL responses in order to prevent immunopathology of infected tissues [191,215,216]. Th17 cells produce IL-6 which inhibits the effects of Tregs and therefore promote T helper responses [215]. Furthermore, Th17 have a role during influenza infections in counteracting secondary bacterial infections, e.g., S. aureus pneumonia. Influenza A virus infections may promote secondary bacterial pneumonia by suppressing Th17 cells in an type I IFN-dependent manner [217].

4. Evasion of the Antiviral Immune Response by Influenza Viruses

Immune pressure on influenza A viruses forces them to adopt strategies to evade immunity. Binding of influenza viral proteins to various components of the innate immune system leads to their inhibition (Figure 2 and Figure 3) [48], whereas a combination of immune pressure in the human population and the high mutation rate of the influenza A virus leads to the generation of new virus strains that escape the existing adaptive humoral and cellular immune responses (Figure 1 and Figure 5).

4.1. Escape from Innate Immunity

Influenza A viruses have adopted various strategies to evade the antiviral nature of the innate immune system.
In particular, the NS1 protein contributes to antagonizing the antiviral innate immune response. Cells infected with genetically modified influenza viruses with a non-functional NS1 gene displayed stronger IFN responses than cells infected with wild type virus. Viruses with a NS1 defect also display reduced virulence after infection of mice and pigs [218,219,220,221,222,223,224]. NS1 inhibits RIG-I receptor signaling by various means. The NS1 protein blocks the recognition of 5'-phosphorylated viral ssRNA by the RIG-I receptor [42]. More downstream of the RIG-I signaling pathway, NS1 prevents oligomerization of TRIM25 by interacting with the coiled coil domain, and so inhibits TRIM25-mediated RIG-I CARD ubiquitination which is essential for downstream signaling [225]. Finally, activation and nuclear translocation of IRF-3, NF-κB and ATF-2/c-Jun is also prevented by NS1 [226,227,228,229]. Hereby NS1 limits RIG-I mediated transcriptional activation of the IFN-β promoter [230,231]. NS1 also alters host cell gene expression by binding to CPSF30 (cleavage and polyadenylation specificity factor); it prevents polyadenylation of the 3' end of host pre-mRNA [232,233,234]. Furthermore, NS1 limits gene expression in general, interfering with the mRNA export machinery [235,236].
NS1 is not the only viral protein that restrains the innate immune system. Both influenza PB2 (especially variants containing an aspartic acid at position 9) and PB1-F2 (only variants containing a serine at position 66) limit the production of IFN-β through association with MAVS [237,238,239,240,241].
Viral proteins PB2, PB1 and PA form the influenza polymerase complex, the main function of which is viral RNA and mRNA synthesis. In addition, it is also involved in cap-snatching of host mRNAs and thereby reduces host cell gene expression including that of IFN-β [242,243,244,245,246].
The recently discovered PA-X viral protein is able to repress cellular gene expression, especially those genes involved in regulating the initiation of the cellular immune response [4].
As described above, influenza A virus infection leads to the production of antiviral PKR (Table 1). In order for PKR to limit viral replication, it first needs to be activated by viral dsRNA. PKR activation is under tight regulation of the cellular p58IPK protein which inhibits PKR activity, but is inactive when it forms a complex with heatshock protein 40 (hsp40) [247,248]. Binding of NP to the p58IPK-hsp40 complex releases p58IPK, and thereby NP inhibits the effects of PKR [249]. In contrast, the influenza M2 protein, which also binds the p58IPK-hsp40 complex, inhibits p58IPK release and thereby limits protein synthesis which eventually leads to host cell apoptosis, possibly enhancing viral particle release [250].
By encapsidating influenza A viral RNA, the NP protein is likely to reduce the formation of dsRNA, which could otherwise lead to activation of RIG-I signaling. Since most PRRs are located inside the cytoplasm, the nuclear replication strategy of the influenza A virus also prevents the recognition of viral RNA by cytosolic PRR.
In addition to limiting the production of type I IFNs, influenza A virus also disturbs type I IFN receptor signaling. Influenza A virus infection induces the expression of SOCS (suppressor of cytokine signaling) proteins which inhibit IFN α/β receptor signaling on the level of JAK/STAT activation [251,252].
Besides interfering with innate signaling, influenza A viruses are also able to counteract cells of the innate immune system. For example, influenza virus infection of monocytes impairs their ability to differentiate into mature DCs [253]. Furthermore, it was shown that NS1 can inhibit DC maturation, and so indirectly limit the induction of virus-specific CD8+ T cell responses [254]. The NK response elicited during an infection is also evaded by the influenza A virus [255]. The gradual mutation of glycosylation sites of influenza virus HA proteins leads to reduced NK recognition of the HA on virus-infected cells [256]. Downregulation of the ζ chain of NKp46 receptors by free HA proteins results in impaired signaling and thereby decreased cytotoxicity of NK cells [257]. Furthermore, influenza A virus can directly infect and kill NK cells [258].

4.2. Escaping the Humoral Immune Response

Various mechanisms contribute to immune evasion of influenza A viruses from the humoral immune response. Due to the lack of proofreading activity, the transcription of viral RNA by the viral RNA polymerase is error prone and results in mis-incorporation of nucleotides. As a result, quasi species of viruses are formed with random mutations in the genome. Under the selective pressure of antibodies that are present in the human population, induced after influenza virus infections and/or vaccination, variants are positively selected from the quasi species that have accumulated amino acid substitutions in the antigenic sites of HA that are recognized by virus-neutralizing antibodies. This phenomenon is known as antigenic drift and allows the virus to evade recognition by antibodies and to cause recurrent influenza epidemics yearly (Figure 1) [8,10,259].
Introduction of influenza A viruses of a novel antigenically distinct subtype into the human population is known as antigenic shift and may cause a pandemic outbreak, since neutralizing antibodies against the new virus strain are absent in the population at large (Figure 1) [10]. Introduction of antigenically distinct viruses can occur after zoonotic transmission. However, in most cases, pandemics were caused by viruses that had exchanged gene segments between human and avian or swine influenza A viruses [260,261]. For re-assortment to take place, cells need to be infected with two influenza A viruses simultaneously [262]. Since epithelial cells of the swine respiratory tract have receptors for both avian and human influenza A viruses, this species can serve as a mixing vessel for the emergence of re-assorted influenza A viruses [263,264,265,266,267]. The 1957 A/H2N2 pandemic was caused after re-assortment of human and avian influenza viruses, as was the 1968 A/H3N2 pandemic virus. The virus that caused the 2009 A/H1N1 pandemic emerged after multiple re-assortment events between avian, swine and human influenza A viruses [12,261,268,269,270,271]. The emergence of the 2009 pandemic strain highlights the importance of pigs as mixing vessels for influenza A viruses. A recent study suggests that besides the pig, the quail could also serve as a mixing vessel for emerging re-assorted influenza A viruses [272].
Of interest, functionally important and conserved sequences in the surface proteins, like the fusion peptide, are inaccessible for antibody recognition, since they are buried inside the protein [273]. Similar strategies to evade antibody recognition are shared by other viruses, like human immunodeficiency virus (HIV) [274,275].

4.3. Escaping the Cellular Immune Response

Viruses have adopted various strategies to evade recognition by virus-specific T cells. For example, viruses with a large DNA genome (e.g., herpes viruses) can encode proteins that interfere with various steps in the antigen processing and presentation pathways [276]. Most RNA viruses, including influenza viruses, have relatively small genomes and limited coding capacity. However, they can evade recognition by T cells through their high mutation rates and the selective pressure exerted by virus‑specific T cells.
Relatively more non-synonymous mutations are observed in the CTL epitope regions of influenza virus NP than in the rest of the protein, indicating that CTL epitopes are under selective pressure [277]. However, mutations flanking CTL epitopes may also affect the liberation of antigenic peptides from the protein by the proteasome or transport by TAP into the ER (Figure 5) [278,279,280], as was demonstrated for HIV [281].
Amino acid substitutions inside CTL epitopes may affect presentation of the epitope in different ways. Amino acid substitutions at an anchor residue may result in complete loss of the epitope, since it may no longer bind to its corresponding MHC class I molecule (Figure 5) [282,283,284,285,286]. Mutations at TCR contact residues can affect recognition by specific T cells, since the epitope no longer matches the specificity of the TCR (Figure 5) [283,284,287,288]. These types of mutations have been observed in escape mutants of viruses that chronically infect their host, like HIV-1 [289,290]. Both types of amino acid substitutions also have been observed during the evolution of seasonal A/H3N2 influenza viruses [283,285,286]. An example of a mutation at an anchor residue is the R384G substitution in the HLA‑B*2705 restricted NP383-391 epitope [285]. This substitution considerably affected the human virus-specific CTL response in vitro [282]. It is remarkable that this mutation reached fixation rapidly, despite the fact that it is recognized by a minority of human subjects only. This could be explained by strong intra-host advantages and founder effects in a theoretical model [291].
An example of amino acid variation at TCR residues includes that observed in the HLA-B*3501 restricted NP418-426 epitope [287,288]. Variation in this epitope displays signs of antigenic drift [292], and dictates the specificity of the CTL response to this epitope and also forms an explanation for cross‑reactivity of CTL against contemporary viruses with historic strains [293]. Of interest, functional constraints may limit variation in CTL. For example, the M158-66 epitope from the matrix protein is highly conserved, despite its immunodominant nature and its restriction by HLA-A*0201, which has a high prevalence in the human population. Influenza viruses do not tolerate mutations in this epitope without loss of viral fitness [277,294].
Figure 5. MHC class I presentation of influenza A virus epitopes and viral escape. This figure represents a virus-infected cell and the presentation of viral epitopes by MHC class I molecules. The virus can escape recognition by virus specific CTLs by: (1) Mutations in TCR contact residues of CTL epitopes in order to prevent recognition of the epitope MHC class I complex by specific CTLs, or (2) mutating the anchor residues of the CTL epitope which prevents binding of the epitope to MHC class I molecules. Furthermore, mutations outside the CTL epitope may affect antigen processing by the proteasome or transport via the TAP respectively (3 and 4).
Figure 5. MHC class I presentation of influenza A virus epitopes and viral escape. This figure represents a virus-infected cell and the presentation of viral epitopes by MHC class I molecules. The virus can escape recognition by virus specific CTLs by: (1) Mutations in TCR contact residues of CTL epitopes in order to prevent recognition of the epitope MHC class I complex by specific CTLs, or (2) mutating the anchor residues of the CTL epitope which prevents binding of the epitope to MHC class I molecules. Furthermore, mutations outside the CTL epitope may affect antigen processing by the proteasome or transport via the TAP respectively (3 and 4).
Viruses 04 01438 g005
The R384G mutation in the NP383-391 epitope was also detrimental to viral fitness and was only tolerated in the presence of functionally compensatory co-mutations [295,296]. At present, it is unknown if influenza viruses can also accumulate mutations flanking CTL epitopes in order to prevent efficient processing and presentation of these epitopes. Of interest, amino acid variation in the HA of influenza A/H3N2 viruses was also associated with escape from recognition by CD4+ T cells, but not with escape from recognition by antibodies [297].

5. Implications for Vaccine Development

5.1. Current Influenza Vaccines

Currently used seasonal influenza vaccines are predominantly inactivated vaccine preparations. Their use aims at the induction of strain-specific antibodies that match the epidemic strains [298,299]. Although these vaccines are considered safe and efficacious, they also have some drawbacks that are addressed by new vaccine technologies [300]. As described above, antigenic drift of influenza viruses allows the seasonal viruses to escape the neutralizing activity of antibodies induced by previous infections or vaccination. Therefore, the vaccine fails to afford life-long protection and needs to be updated almost annually [301]. Furthermore, the production of the vaccine takes several months, so the recommendation for the vaccine strains of the upcoming influenza season is made months in advance [301]. In most influenza seasons, the predicted vaccine strains match the epidemic strains. Occasionally however, a predicted influenza vaccine strain does not match the circulating strain, resulting in suboptimal protection afforded by the vaccine [302,303,304]. In the event of an emerging pandemic outbreak, the time needed to produce and distribute a pandemic influenza vaccine is also a major drawback [305,306,307]. Seasonal influenza vaccines do not afford protection against pandemic strains of novel subtypes, since the vaccine-induced antibodies do not cross-react and cross-reactive CD8+ T cell responses are induced inefficiently.
Alternatively, cold-adapted live-attenuated seasonal influenza vaccines are used [308,309,310,311]. The advantage of live-attenuated vaccines is that they also elicit cellular immune responses [312,313] and mucosal immunity [310]. More recently, live-attenuated influenza vaccines have been developed by disrupting NS1 activity through deletion or truncation of the NS1 gene [221,223,224,228,314]. Hence, the virus is unable to interfere with the innate immune response.

5.2. Novel Vaccines

Ideally, seasonal vaccines are used that induce broad-protective immunity against drift variants and potentially pandemic viruses of novel subtypes. Currently-used inactivated vaccines may even interfere with the induction of cell-mediated immunity otherwise induced by natural infections, especially in young children that are still immunologically naïve to influenza viruses. In this way, inactivated vaccines can hamper the development of heterosubtypic immunity [315,316,317,318,319]. Thus, the development of vaccines that induce broadly neutralizing antibodies and preferably long-lasting heterosubtypic CTL responses is desirable.
Since viral proteins like NP and M1 are highly conserved, they are likely targets for the induction of cross-reactive T cell responses [320]. Induction of efficient CTL responses depends on effective endogenous antigen processing and presentation by MHC class I. This requires effective delivery of viral proteins into the cytosol. Several cytosolic delivery vaccine candidates are now under investigation, including DNA vaccines, recombinant viral vectors, ISCOMS and virosomes, some of which already have made it into clinical trials [321,322,323,324,325,326,327]. Also, the induction of cross-reactive antibodies has attracted attention in recent years; antibodies directed against the more conserved stem region of the HA molecule are of especial interest. In contrast to the subtype-specific antibodies induced against the globular head of HA, these HA stem-specific antibodies display broad-neutralizing activity against multiple influenza virus subtypes [113,114]. Using this stem region as an immunogen, broadly protective antibody responses could be induced [117,328].
Also, the ectodomain of the M2 protein is highly conserved and antibodies induced against this region afford protection against challenge infection [133,136,137,329,330]. The mode of action is not neutralization per se, since M2 is a minor antigen on virus particles. However, since it is also expressed on the surface of infected cells, ADCC is probably responsible for the protective effect of these antibodies [135].
Thus, vaccines that induce both humoral and cell-mediated immune responses directed to conserved regions of influenza A virus, in addition to strain-specific antibodies, are likely to afford protective immunity to a large variety of influenza A viruses, including drift variants and viruses of novel subtypes.

6. Concluding Remarks

Our knowledge of the complexity of interaction between host immune responses and variable pathogens, like influenza viruses, has increased tremendously. Insight has been obtained on how influenza A viruses evade recognition by components of the immune system. Although these insights have helped us to decrease both morbidity and mortality, mainly by developing effective seasonal vaccines, there are still gaps in our understanding that provide room for improvement for the development of more broadly protective vaccines. The induction of CTL responses to conserved epitopes, preferably those that are under functional constraints, may be a venue to develop broadly protective vaccines. Current research also focuses on cross-neutralizing antibody responses directed to more conserved regions of the surface proteins. If these regions are also under functional constraints, vaccines aimed at the induction of antibodies against these regions would suffer to a lesser extent in terms of variability in the target region.
Collectively, the new developments may counteract the variable nature of influenza viruses and yield vaccines that afford protection against both seasonal influenza viruses and future pandemic strains.

References and Notes

  1. Palese, P.; Shaw, M.L. Orthomyxoviridae: The viruses and their replication. In Fields Virology, 5th ed; Lippincott Williams & Wilkins, a Wolters Kluwer Business: Philadelphia, PA, USA, 2007; Volume 2, pp. 1647–1689. [Google Scholar]
  2. WHO. Influenza (seasonal) fact sheet No 211. Available online: http://www.who.int/mediacentre/factsheets/fs211/en/index.html (accessed on 31 March 2012).
  3. Stohr, K. Influenza—WHO cares. Lancet Infect. Dis. 2002, 2, 517. [Google Scholar] [CrossRef]
  4. Jagger, B.W.; Wise, H.M.; Kash, J.C.; Walters, K.A.; Wills, N.M.; Xiao, Y.L.; Dunfee, R.L.; Schwartzman, L.M.; Ozinsky, A.; Bell, G.L.; et al. An overlapping protein-coding region in influenza a virus segment 3 modulates the host response. Science 2012, 337, 199–204. [Google Scholar] [CrossRef]
  5. Fouchier, R.A.; Munster, V.; Wallensten, A.; Bestebroer, T.M.; Herfst, S.; Smith, D.; Rimmelzwaan, G.F.; Olsen, B.; Osterhaus, A.D. Characterization of a novel influenza A virus hemagglutinin subtype (H16) obtained from black-headed gulls. J. Virol. 2005, 79, 2814–2822. [Google Scholar]
  6. WHO. A revision of the system of nomenclature for influenza viruses: A WHO memorandum. Bull World Health Organ. 1980, 58, 585–591.
  7. Tong, S.; Li, Y.; Rivailler, P.; Conrardy, C.; Castillo, D.A.; Chen, L.M.; Recuenco, S.; Ellison, J.A.; Davis, C.T.; York, I.A.; et al. A distinct lineage of influenza A virus from bats. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 4269–4274. [Google Scholar]
  8. Smith, D.J.; Lapedes, A.S.; de Jong, J.C.; Bestebroer, T.M.; Rimmelzwaan, G.F.; Osterhaus, A.D.; Fouchier, R.A. Mapping the antigenic and genetic evolution of influenza virus. Science 2004, 305, 371–376. [Google Scholar]
  9. Herfst, S.; Schrauwen, E.J.A.; Linster, M.; Chutinimitkul, S.; de Wit, E.; Munster, V.J.; Sorrell, E.M.; Bestebroer, T.M.; Burke, D.F.; Smith, D.J.; et al. Airborne transmission of influenza A/H5N1 virus between ferrets. Science 2012, 336, 1534–1541. [Google Scholar]
  10. De Jong, J.C.; Rimmelzwaan, G.F.; Fouchier, R.A.; Osterhaus, A.D. Influenza virus: A master of metamorphosis. J. Infect. 2000, 40, 218–228. [Google Scholar] [CrossRef]
  11. Johnson, N.P.; Mueller, J. Updating the accounts: Global mortality of the 1918-1920 "Spanish" influenza pandemic. Bull. Hist. Med. 2002, 76, 105–115. [Google Scholar] [CrossRef]
  12. Garten, R.J.; Davis, C.T.; Russell, C.A.; Shu, B.; Lindstrom, S.; Balish, A.; Sessions, W.M.; Xu, X.; Skepner, E.; Deyde, V.; et al. Antigenic and genetic characteristics of swine-origin 2009 A(H1N1) influenza viruses circulating in humans. Science 2009, 325, 197–201. [Google Scholar]
  13. de Wit, E.; Kawaoka, Y.; de Jong, M.D.; Fouchier, R.A. Pathogenicity of highly pathogenic avian influenza virus in mammals. Vaccine 2008, 26, D54–D58. [Google Scholar] [CrossRef]
  14. Kuiken, T.; Holmes, E.C.; McCauley, J.; Rimmelzwaan, G.F.; Williams, C.S.; Grenfell, B.T. Host species barriers to influenza virus infections. Science 2006, 312, 394–397. [Google Scholar]
  15. Lin, Y.P.; Shaw, M.; Gregory, V.; Cameron, K.; Lim, W.; Klimov, A.; Subbarao, K.; Guan, Y.; Krauss, S.; Shortridge, K.; et al. Avian-to-human transmission of H9N2 subtype influenza A viruses: Relationship between H9N2 and H5N1 human isolates. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 9654–9658. [Google Scholar]
  16. Fouchier, R.A.; Schneeberger, P.M.; Rozendaal, F.W.; Broekman, J.M.; Kemink, S.A.; Munster, V.; Kuiken, T.; Rimmelzwaan, G.F.; Schutten, M.; Van Doornum, G.J.; et al. Avian influenza A virus (H7N7) associated with human conjunctivitis and a fatal case of acute respiratory distress syndrome. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 1356–1361. [Google Scholar]
  17. Koopmans, M.; Wilbrink, B.; Conyn, M.; Natrop, G.; van der Nat, H.; Vennema, H.; Meijer, A.; van Steenbergen, J.; Fouchier, R.; Osterhaus, A.; et al. Transmission of H7N7 avian influenza A virus to human beings during a large outbreak in commercial poultry farms in the Netherlands. Lancet 2004, 363, 587–593. [Google Scholar]
  18. de Jong, J.C.; Claas, E.C.; Osterhaus, A.D.; Webster, R.G.; Lim, W.L. A pandemic warning? Nature 1997, 389, 554. [Google Scholar]
  19. de Jong, M.D.; Bach, V.C.; Phan, T.Q.; Vo, M.H.; Tran, T.T.; Nguyen, B.H.; Beld, M.; Le, T.P.; Truong, H.K.; Nguyen, V.V.; et al. Fatal avian influenza A (H5N1) in a child presenting with diarrhea followed by coma. New Engl. J. Med. 2005, 352, 686–691. [Google Scholar]
  20. Gambotto, A.; Barratt-Boyes, S.M.; de Jong, M.D.; Neumann, G.; Kawaoka, Y. Human infection with highly pathogenic H5N1 influenza virus. Lancet 2008, 371, 1464–1475. [Google Scholar] [CrossRef]
  21. Abdel-Ghafar, A.N.; Chotpitayasunondh, T.; Gao, Z.; Hayden, F.G.; Nguyen, D.H.; de Jong, M.D.; Naghdaliyev, A.; Peiris, J.S.; Shindo, N.; Soeroso, S.; et al. Update on avian influenza A (H5N1) virus infection in humans. New Engl. J. Med. 2008, 358, 261–273. [Google Scholar] [CrossRef]
  22. Beigel, J.H.; Farrar, J.; Han, A.M.; Hayden, F.G.; Hyer, R.; de Jong, M.D.; Lochindarat, S.; Nguyen, T.K.; Nguyen, T.H.; Tran, T.H.; et al. Avian influenza A (H5N1) infection in humans. New Engl. J. Med. 2005, 353, 1374–1385. [Google Scholar] [CrossRef]
  23. WHO. Cumulative number of confirmed human cases for avian influenza A(H5N1) reported to WHO, 2003-2012. Available online: http://www.who.int/influenza/human_animal_interface/avian_influenza/EN_GIP_20120326CumulativeNumberH5N1cases.pdf (accessed on 1 April 2012).
  24. Wang, T.T.; Parides, M.K.; Palese, P. Seroevidence for H5N1 influenza infections in humans: Meta-analysis. Science 2012, 335, 1463. [Google Scholar] [CrossRef]
  25. Kandun, I.N.; Wibisono, H.; Sedyaningsih, E.R.; Yusharmen; Hadisoedarsuno, W.; Purba, W.; Santoso, H.; Septiawati, C.; Tresnaningsih, E.; Heriyanto, B.; et al. Three Indonesian clusters of H5N1 virus infection in 2005. New Engl. J. Med. 2006, 355, 2186–2194. [Google Scholar]
  26. Ungchusak, K.; Auewarakul, P.; Dowell, S.F.; Kitphati, R.; Auwanit, W.; Puthavathana, P.; Uiprasertkul, M.; Boonnak, K.; Pittayawonganon, C.; Cox, N.J.; et al. Probable person-to-person transmission of avian influenza A (H5N1). New Engl. J. Med. 2005, 352, 333–340. [Google Scholar] [CrossRef]
  27. Wang, H.; Feng, Z.; Shu, Y.; Yu, H.; Zhou, L.; Zu, R.; Huai, Y.; Dong, J.; Bao, C.; Wen, L.; et al. Probable limited person-to-person transmission of highly pathogenic avian influenza A (H5N1) virus in China. Lancet 2008, 371, 1427–1434. [Google Scholar]
  28. 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]
  29. Russell, C.A.; Fonville, J.M.; Brown, A.E.X.; Burke, D.F.; Smith, D.L.; James, S.L.; Herfst, S.; van Boheemen, S.; Linster, M.; Schrauwen, E.J.; et al. The potential for respiratory droplet transmissible A/H5N1 influenza virus to evolve in a mammalian host. Science 2012, 336, 1541–1547. [Google Scholar]
  30. Holt, P.G.; Strickland, D.H.; Wikstrom, M.E.; Jahnsen, F.L. Regulation of immunological homeostasis in the respiratory tract. Nat. Rev. 2008, 8, 142–152. [Google Scholar] [CrossRef]
  31. Sanders, C.J.; Doherty, P.C.; Thomas, P.G. Respiratory epithelial cells in innate immunity to influenza virus infection. Cell Tissue Res. 2011, 343, 13–21. [Google Scholar]
  32. Blasius, A.L.; Beutler, B. Intracellular toll-like receptors. Immunity 2010, 32, 305–315. [Google Scholar] [CrossRef]
  33. Pang, I.K.; Iwasaki, A. Inflammasomes as mediators of immunity against influenza virus. Trends Immunol. 2011, 32, 34–41. [Google Scholar]
  34. Takeuchi, O.; Akira, S. Innate immunity to virus infection. Immunol. Rev. 2009, 227, 75–86. [Google Scholar]
  35. Lund, J.M.; Alexopoulou, L.; Sato, A.; Karow, M.; Adams, N.C.; Gale, N.W.; Iwasaki, A.; Flavell, R.A. Recognition of single-stranded RNA viruses by Toll-like receptor 7. Proc. Natl. Acad. Sci. U. S. A. 2004, 101, 5598–5603. [Google Scholar]
  36. Diebold, S.S.; Kaisho, T.; Hemmi, H.; Akira, S.; Reis e Sousa, C. Innate antiviral responses by means of TLR7-mediated recognition of single-stranded RNA. Science 2004, 303, 1529–1531. [Google Scholar] [CrossRef]
  37. Guillot, L.; Le Goffic, R.; Bloch, S.; Escriou, N.; Akira, S.; Chignard, M.; Si-Tahar, M. Involvement of toll-like receptor 3 in the immune response of lung epithelial cells to double-stranded RNA and influenza A virus. J. Biol. Chem. 2005, 280, 5571–5580. [Google Scholar]
  38. Imai, Y.; Kuba, K.; Neely, G.G.; Yaghubian-Malhami, R.; Perkmann, T.; van Loo, G.; Ermolaeva, M.; Veldhuizen, R.; Leung, Y.H.; Wang, H.; et al. Identification of oxidative stress and Toll-like receptor 4 signaling as a key pathway of acute lung injury. Cell 2008, 133, 235–249. [Google Scholar] [CrossRef]
  39. Kurt-Jones, E.A.; Popova, L.; Kwinn, L.; Haynes, L.M.; Jones, L.P.; Tripp, R.A.; Walsh, E.E.; Freeman, M.W.; Golenbock, D.T.; Anderson, L.J.; et al. Pattern recognition receptors TLR4 and CD14 mediate response to respiratory syncytial virus. Nat. Immunol. 2000, 1, 398–401. [Google Scholar] [CrossRef]
  40. Hashimoto, Y.; Moki, T.; Takizawa, T.; Shiratsuchi, A.; Nakanishi, Y. Evidence for phagocytosis of influenza virus-infected, apoptotic cells by neutrophils and macrophages in mice. J. Immunol. 2007, 178, 2448–2457. [Google Scholar]
  41. Hornung, V.; Ellegast, J.; Kim, S.; Brzozka, K.; Jung, A.; Kato, H.; Poeck, H.; Akira, S.; Conzelmann, K.K.; Schlee, M.; et al. 5'-Triphosphate RNA is the ligand for RIG-I. Science 2006, 314, 994–997. [Google Scholar]
  42. Pichlmair, A.; Schulz, O.; Tan, C.P.; Naslund, T.I.; Liljestrom, P.; Weber, F.; Reis e Sousa, C. RIG-I-mediated antiviral responses to single-stranded RNA bearing 5'-phosphates. Science 2006, 314, 997–1001. [Google Scholar] [CrossRef]
  43. Kato, H.; Takeuchi, O.; Sato, S.; Yoneyama, M.; Yamamoto, M.; Matsui, K.; Uematsu, S.; Jung, A.; Kawai, T.; Ishii, K.J.; et al. Differential roles of MDA5 and RIG-I helicases in the recognition of RNA viruses. Nature 2006, 441, 101–105. [Google Scholar]
  44. Loo, Y.M.; Gale, M., Jr. Immune signaling by RIG-I-like receptors. Immunity 2011, 34, 680–692. [Google Scholar]
  45. Kanneganti, T.D.; Body-Malapel, M.; Amer, A.; Park, J.H.; Whitfield, J.; Franchi, L.; Taraporewala, Z.F.; Miller, D.; Patton, J.T.; Inohara, N.; et al. Critical role for Cryopyrin/Nalp3 in activation of caspase-1 in response to viral infection and double-stranded RNA. J. Biol. Chem. 2006, 281, 36560–36568. [Google Scholar]
  46. Gack, M.U.; Shin, Y.C.; Joo, C.H.; Urano, T.; Liang, C.; Sun, L.; Takeuchi, O.; Akira, S.; Chen, Z.; Inoue, S.; et al. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature 2007, 446, 916–920. [Google Scholar]
  47. Kawai, T.; Akira, S. Toll-like receptors and their crosstalk with other innate receptors in infection and immunity. Immunity 2011, 34, 637–650. [Google Scholar] [CrossRef]
  48. Hale, B.G.; Albrecht, R.A.; Garcia-Sastre, A. Innate immune evasion strategies of influenza viruses. Future Microbiol. 2010, 5, 23–41. [Google Scholar]
  49. Randall, R.E.; Goodbourn, S. Interferons and viruses: An interplay between induction, signalling, antiviral responses and virus countermeasure. J. Gen. Virol. 2008, 89, 1–47. [Google Scholar]
  50. Chelbi-Alix, M.K.; Wietzerbin, J. Interferon, a growing cytokine family: 50 years of interferon research. Biochimie 2007, 89, 713–718. [Google Scholar] [CrossRef]
  51. Van Hoeven, N.; Belser, J.A.; Szretter, K.J.; Zeng, H.; Staeheli, P.; Swayne, D.E.; Katz, J.M.; Tumpey, T.M. Pathogenesis of 1918 pandemic and H5N1 influenza virus infections in a guinea pig model: Antiviral potential of exogenous alpha interferon to reduce virus shedding. J. Virol. 2009, 83, 2851–2861. [Google Scholar]
  52. Szretter, K.J.; Gangappa, S.; Belser, J.A.; Zeng, H.; Chen, H.; Matsuoka, Y.; Sambhara, S.; Swayne, D.E.; Tumpey, T.M.; Katz, J.M. Early control of H5N1 influenza virus replication by the type I interferon response in mice. J. Virol. 2009, 83, 5825–5834. [Google Scholar]
  53. Sato, M.; Hata, N.; Asagiri, M.; Nakaya, T.; Taniguchi, T.; Tanaka, N. Positive feedback regulation of type I IFN genes by the IFN-inducible transcription factor IRF-7. FEBS Lett. 1998, 441, 106–110. [Google Scholar]
  54. von der Malsburg, A.; Abutbul-Ionita, I.; Haller, O.; Kochs, G.; Danino, D. Stalk domain of the dynamin-like MxA GTPase protein mediates membrane binding and liposome tubulation via the unstructured L4 loop. J. Biol. Chem. 2011, 286, 37858–37865. [Google Scholar]
  55. Haller, O.; Gao, S.; von der Malsburg, A.; Daumke, O.; Kochs, G. Dynamin-like MxA GTPase: Structural insights into oligomerization and implications for antiviral activity. J. Biol. Chem. 2010, 285, 28419–28424. [Google Scholar]
  56. Haller, O.; Kochs, G. Interferon-induced mx proteins: Dynamin-like GTPases with antiviral activity. Traffic 2002, 3, 710–717. [Google Scholar] [CrossRef]
  57. Pindel, A.; Sadler, A. The role of protein kinase R in the interferon response. J. Interferon Cytokine Res. 2011, 31, 59–70. [Google Scholar] [CrossRef]
  58. Garcia, M.A.; Gil, J.; Ventoso, I.; Guerra, S.; Domingo, E.; Rivas, C.; Esteban, M. Impact of protein kinase PKR in cell biology: From antiviral to antiproliferative action. Microbiol. Mol. Biol. Rev. 2006, 70, 1032–1060. [Google Scholar]
  59. Chakrabarti, A.; Jha, B.K.; Silverman, R.H. New insights into the role of RNase L in innate immunity. J. Interferon Cytokine Res. 2011, 31, 49–57. [Google Scholar] [CrossRef]
  60. Silverman, R.H. Viral encounters with 2',5'-oligoadenylate synthetase and RNase L during the interferon antiviral response. J. Virol. 2007, 81, 12720–12729. [Google Scholar] [CrossRef]
  61. Lenschow, D.J.; Lai, C.; Frias-Staheli, N.; Giannakopoulos, N.V.; Lutz, A.; Wolff, T.; Osiak, A.; Levine, B.; Schmidt, R.E.; Garcia-Sastre, A.; et al. IFN-stimulated gene 15 functions as a critical antiviral molecule against influenza, herpes, and Sindbis viruses. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 1371–1376. [Google Scholar]
  62. Wang, X.; Hinson, E.R.; Cresswell, P. The interferon-inducible protein viperin inhibits influenza virus release by perturbing lipid rafts. Cell Host Microbe 2007, 2, 96–105. [Google Scholar]
  63. Watanabe, R.; Leser, G.P.; Lamb, R.A. Influenza virus is not restricted by tetherin whereas influenza VLP production is restricted by tetherin. Virology 2011, 417, 50–56. [Google Scholar]
  64. Yondola, M.A.; Fernandes, F.; Belicha-Villanueva, A.; Uccelini, M.; Gao, Q.; Carter, C.; Palese, P. Budding capability of the influenza virus neuraminidase can be modulated by tetherin. J. Virol. 2011, 85, 2480–2491. [Google Scholar] [CrossRef]
  65. Brass, A.L.; Huang, I.C.; Benita, Y.; John, S.P.; Krishnan, M.N.; Feeley, E.M.; Ryan, B.J.; Weyer, J.L.; van der Weyden, L.; Fikrig, E.; et al. The IFITM proteins mediate cellular resistance to influenza A H1N1 virus, West Nile virus, and dengue virus. Cell 2009, 139, 1243–1254. [Google Scholar]
  66. Turner, S.J.; Olivas, E.; Gutierrez, A.; Diaz, G.; Doherty, P.C. Disregulated influenza A virus-specific CD8+ T cell homeostasis in the absence of IFN-gamma signaling. J. Immunol. 2007, 178, 7616–7622. [Google Scholar]
  67. Bot, A.; Bot, S.; Bona, C.A. Protective role of gamma interferon during the recall response to influenza virus. J. Virol. 1998, 72, 6637–6645. [Google Scholar]
  68. Mordstein, M.; Kochs, G.; Dumoutier, L.; Renauld, J.C.; Paludan, S.R.; Klucher, K.; Staeheli, P. Interferon-lambda contributes to innate immunity of mice against influenza A virus but not against hepatotropic viruses. PLoS Pathog. 2008, 4, e1000151. [Google Scholar] [CrossRef]
  69. McGill, J.; Heusel, J.W.; Legge, K.L. Innate immune control and regulation of influenza virus infections. J. Leukoc. Biol. 2009, 86, 803–812. [Google Scholar]
  70. Snelgrove, R.J.; Goulding, J.; Didierlaurent, A.M.; Lyonga, D.; Vekaria, S.; Edwards, L.; Gwyer, E.; Sedgwick, J.D.; Barclay, A.N.; Hussell, T. A critical function for CD200 in lung immune homeostasis and the severity of influenza infection. Nat. Immunol. 2008, 9, 1074–1083. [Google Scholar]
  71. Lin, K.L.; Suzuki, Y.; Nakano, H.; Ramsburg, E.; Gunn, M.D. CCR2+ monocyte-derived dendritic cells and exudate macrophages produce influenza-induced pulmonary immune pathology and mortality. J. Immunol. 2008, 180, 2562–2572. [Google Scholar]
  72. Dawson, T.C.; Beck, M.A.; Kuziel, W.A.; Henderson, F.; Maeda, N. Contrasting effects of CCR5 and CCR2 deficiency in the pulmonary inflammatory response to influenza A virus. Am. J. Pathol. 2000, 156, 1951–1959. [Google Scholar] [CrossRef]
  73. Herold, S.; von Wulffen, W.; Steinmueller, M.; Pleschka, S.; Kuziel, W.A.; Mack, M.; Srivastava, M.; Seeger, W.; Maus, U.A.; Lohmeyer, J. Alveolar epithelial cells direct monocyte transepithelial migration upon influenza virus infection: Impact of chemokines and adhesion molecules. J. Immunol. 2006, 177, 1817–1824. [Google Scholar]
  74. van Riel, D.; Leijten, L.M.; van der Eerden, M.; Hoogsteden, H.C.; Boven, L.A.; Lambrecht, B.N.; Osterhaus, A.D.; Kuiken, T. Highly pathogenic avian influenza virus H5N1 infects alveolar macrophages without virus production or excessive TNF-alpha induction. PLoS Pathog. 2011, 7, e1002099. [Google Scholar]
  75. Becker, S.; Quay, J.; Soukup, J. Cytokine (tumor necrosis factor, IL-6, and IL-8) production by respiratory syncytial virus-infected human alveolar macrophage. J. Immunol. 1991, 147, 4307–4312. [Google Scholar]
  76. Tumpey, T.M.; Garcia-Sastre, A.; Taubenberger, J.K.; Palese, P.; Swayne, D.E.; Pantin-Jackwood, M.J.; Schultz-Cherry, S.; Solorzano, A.; van Rooijen, N.; Katz, J.M.; et al. Pathogenicity of influenza viruses with genes from the 1918 pandemic virus: Functional roles of alveolar macrophages and neutrophils in limiting virus replication and mortality in mice. J. Virol. 2005, 79, 14933–14944. [Google Scholar]
  77. Kim, H.M.; Lee, Y.W.; Lee, K.J.; Kim, H.S.; Cho, S.W.; van Rooijen, N.; Guan, Y.; Seo, S.H. Alveolar macrophages are indispensable for controlling influenza viruses in lungs of pigs. J. Virol. 2008, 82, 4265–4274. [Google Scholar]
  78. Huber, V.C.; Lynch, J.M.; Bucher, D.J.; Le, J.; Metzger, D.W. Fc receptor-mediated phagocytosis makes a significant contribution to clearance of influenza virus infections. J. Immunol. 2001, 166, 7381–7388. [Google Scholar]
  79. Jayasekera, J.P.; Vinuesa, C.G.; Karupiah, G.; King, N.J. Enhanced antiviral antibody secretion and attenuated immunopathology during influenza virus infection in nitric oxide synthase-2-deficient mice. J. Gen. Virol. 2006, 87, 3361–3371. [Google Scholar] [CrossRef]
  80. Peper, R.L.; van Campen, H. Tumor necrosis factor as a mediator of inflammation in influenza A viral pneumonia. Microb. Pathog. 1995, 19, 175–183. [Google Scholar] [CrossRef]
  81. Mendelson, M.; Tekoah, Y.; Zilka, A.; Gershoni-Yahalom, O.; Gazit, R.; Achdout, H.; Bovin, N.V.; Meningher, T.; Mandelboim, M.; Mandelboim, O.; et al. NKp46 O-glycan sequences that are involved in the interaction with hemagglutinin type 1 of influenza virus. J. Virol. 2010, 84, 3789–3797. [Google Scholar]
  82. Arnon, T.I.; Lev, M.; Katz, G.; Chernobrov, Y.; Porgador, A.; Mandelboim, O. Recognition of viral hemagglutinins by NKp44 but not by NKp30. Eur. J. Immunol. 2001, 31, 2680–2689. [Google Scholar]
  83. Mandelboim, O.; Lieberman, N.; Lev, M.; Paul, L.; Arnon, T.I.; Bushkin, Y.; Davis, D.M.; Strominger, J.L.; Yewdell, J.W.; Porgador, A. Recognition of haemagglutinins on virus-infected cells by NKp46 activates lysis by human NK cells. Nature 2001, 409, 1055–1060. [Google Scholar]
  84. Gazit, R.; Gruda, R.; Elboim, M.; Arnon, T.I.; Katz, G.; Achdout, H.; Hanna, J.; Qimron, U.; Landau, G.; Greenbaum, E.; et al. Lethal influenza infection in the absence of the natural killer cell receptor gene Ncr1. Nat. Immunol. 2006, 7, 517–523. [Google Scholar]
  85. Hashimoto, G.; Wright, P.F.; Karzon, D.T. Antibody-dependent cell-mediated cytotoxicity against influenza virus-infected cells. J. Infect. Dis. 1983, 148, 785–794. [Google Scholar]
  86. Sun, P.D. Structure and function of natural-killer-cell receptors. Immunol. Res. 2003, 27, 539–548. [Google Scholar] [CrossRef]
  87. Jegerlehner, A.; Schmitz, N.; Storni, T.; Bachmann, M.F. Influenza A vaccine based on the extracellular domain of M2: Weak protection mediated via antibody-dependent NK cell activity. J. Immunol. 2004, 172, 5598–5605. [Google Scholar]
  88. Banchereau, J.; Steinman, R.M. Dendritic cells and the control of immunity. Nature 1998, 392, 245–252. [Google Scholar] [CrossRef]
  89. Guermonprez, P.; Valladeau, J.; Zitvogel, L.; Thery, C.; Amigorena, S. Antigen presentation and T cell stimulation by dendritic cells. Annu. Rev. Immunol. 2002, 20, 621–667. [Google Scholar] [CrossRef]
  90. GeurtsvanKessel, C.H.; Lambrecht, B.N. Division of labor between dendritic cell subsets of the lung. Mucosal Immunol. 2008, 1, 442–450. [Google Scholar] [CrossRef]
  91. Bhardwaj, N.; Bender, A.; Gonzalez, N.; Bui, L.K.; Garrett, M.C.; Steinman, R.M. Influenza virus-infected dendritic cells stimulate strong proliferative and cytolytic responses from human CD8+ T cells. J. Clin. Investig. 1994, 94, 797–807. [Google Scholar]
  92. Hamilton-Easton, A.; Eichelberger, M. Virus-specific antigen presentation by different subsets of cells from lung and mediastinal lymph node tissues of influenza virus-infected mice. J. Virol. 1995, 69, 6359–6366. [Google Scholar]
  93. Yewdell, J.W.; Reits, E.; Neefjes, J. Making sense of mass destruction: Quantitating MHC class I antigen presentation. Nat. Rev. 2003, 3, 952–961. [Google Scholar]
  94. Mori, I.; Komatsu, T.; Takeuchi, K.; Nakakuki, K.; Sudo, M.; Kimura, Y. In vivo induction of apoptosis by influenza virus. J. Gen. Virol. 1995, 76, 2869–2873. [Google Scholar] [CrossRef]
  95. Braciale, T.J.; Sun, J.; Kim, T.S. Regulating the adaptive immune response to respiratory virus infection. Nat. Rev. 2012, 12, 295–305. [Google Scholar] [CrossRef]
  96. Kim, T.S.; Braciale, T.J. Respiratory dendritic cell subsets differ in their capacity to support the induction of virus-specific cytotoxic CD8+ T cell responses. PLoS One 2009, 4, e4204. [Google Scholar]
  97. Norbury, C.C.; Malide, D.; Gibbs, J.S.; Bennink, J.R.; Yewdell, J.W. Visualizing priming of virus-specific CD8+ T cells by infected dendritic cells in vivo. Nat. Immunol. 2002, 3, 265–271. [Google Scholar] [CrossRef]
  98. Heer, A.K.; Harris, N.L.; Kopf, M.; Marsland, B.J. CD4+ and CD8+ T cells exhibit differential requirements for CCR7-mediated antigen transport during influenza infection. J. Immunol. 2008, 181, 6984–6994. [Google Scholar]
  99. Legge, K.L.; Braciale, T.J. Accelerated migration of respiratory dendritic cells to the regional lymph nodes is limited to the early phase of pulmonary infection. Immunity 2003, 18, 265–277. [Google Scholar] [CrossRef]
  100. Baumgarth, N.; Herman, O.C.; Jager, G.C.; Brown, L.E.; Herzenberg, L.A.; Chen, J. B-1 and B-2 cell-derived immunoglobulin M antibodies are nonredundant components of the protective response to influenza virus infection. J. Exp. Med. 2000, 192, 271–280. [Google Scholar]
  101. Baumgarth, N.; Tung, J.W.; Herzenberg, L.A. Inherent specificities in natural antibodies: A key to immune defense against pathogen invasion. Springer Semin. Immunopathol. 2005, 26, 347–362. [Google Scholar] [CrossRef]
  102. Potter, C.W.; Oxford, J.S. Determinants of immunity to influenza infection in man. Br. Med. Bull. 1979, 35, 69–75. [Google Scholar]
  103. Waffarn, E.E.; Baumgarth, N. Protective B cell responses to flu—No fluke! J. Immunol. 2011, 186, 3823–3829. [Google Scholar] [CrossRef]
  104. de Jong, J.C.; Palache, A.M.; Beyer, W.E.; Rimmelzwaan, G.F.; Boon, A.C.; Osterhaus, A.D. Haemagglutination-inhibiting antibody to influenza virus. Dev. Biologicals 2003, 115, 63–73. [Google Scholar]
  105. Knossow, M.; Skehel, J.J. Variation and infectivity neutralization in influenza. Immunology 2006, 119, 1–7. [Google Scholar] [CrossRef]
  106. Wilson, I.A.; Cox, N.J. Structural basis of immune recognition of influenza virus hemagglutinin. Annu. Rev. Immunol. 1990, 8, 737–771. [Google Scholar] [CrossRef]
  107. Kilbourne, E.D. Influenza as a problem in immunology. J. Immunol. 1978, 120, 1447–1452. [Google Scholar]
  108. Potter, C.W.; Oxford, J.S.; Shore, S.L.; McLaren, C.; Stuart-Harris, C. Immunity to influenza in ferrets. I. Response to live and killed virus. Br. J. Exp. Pathol. 1972, 53, 153–167. [Google Scholar]
  109. Yu, X.; Tsibane, T.; McGraw, P.A.; House, F.S.; Keefer, C.J.; Hicar, M.D.; Tumpey, T.M.; Pappas, C.; Perrone, L.A.; Martinez, O.; et al. Neutralizing antibodies derived from the B cells of 1918 influenza pandemic survivors. Nature 2008, 455, 532–536. [Google Scholar]
  110. Ikonen, N.; Strengell, M.; Kinnunen, L.; Osterlund, P.; Pirhonen, J.; Broman, M.; Davidkin, I.; Ziegler, T.; Julkunen, I. High frequency of cross-reacting antibodies against 2009 pandemic influenza A(H1N1) virus among the elderly in Finland. Euro Surveill. 2010, 15, pii=19478. [Google Scholar]
  111. Reed, C.; Katz, J.M. Serological surveys for 2009 pandemic influenza A H1N1. Lancet 2010, 375, 1062–1063. [Google Scholar] [CrossRef]
  112. Hancock, K.; Veguilla, V.; Lu, X.; Zhong, W.; Butler, E.N.; Sun, H.; Liu, F.; Dong, L.; DeVos, J.R.; Gargiullo, P.M.; et al. Cross-reactive antibody responses to the 2009 pandemic H1N1 influenza virus. New Engl. J. Med. 2009, 361, 1945–1952. [Google Scholar]
  113. Ekiert, D.C.; Bhabha, G.; Elsliger, M.A.; Friesen, R.H.; Jongeneelen, M.; Throsby, M.; Goudsmit, J.; Wilson, I.A. Antibody recognition of a highly conserved influenza virus epitope. Science 2009, 324, 246–251. [Google Scholar]
  114. Ekiert, D.C.; Friesen, R.H.; Bhabha, G.; Kwaks, T.; Jongeneelen, M.; Yu, W.; Ophorst, C.; Cox, F.; Korse, H.J.; Brandenburg, B.; et al. A highly conserved neutralizing epitope on group 2 influenza A viruses. Science 2011, 333, 843–850. [Google Scholar]
  115. Sui, J.; Hwang, W.C.; Perez, S.; Wei, G.; Aird, D.; Chen, L.M.; Santelli, E.; Stec, B.; Cadwell, G.; Ali, M.; et al. Structural and functional bases for broad-spectrum neutralization of avian and human influenza A viruses. Nat. Struct. Mol. Biol. 2009, 16, 265–273. [Google Scholar]
  116. Wang, T.T.; Tan, G.S.; Hai, R.; Pica, N.; Petersen, E.; Moran, T.M.; Palese, P. Broadly protective monoclonal antibodies against H3 influenza viruses following sequential immunization with different hemagglutinins. PLoS Pathog. 2010, 6, e1000796. [Google Scholar] [CrossRef]
  117. Steel, J.; Lowen, A.C.; Wang, T.T.; Yondola, M.; Gao, Q.; Haye, K.; Garcia-Sastre, A.; Palese, P. Influenza virus vaccine based on the conserved hemagglutinin stalk domain. mBio 2010, 1, e00018–10. [Google Scholar]
  118. Throsby, M.; van den Brink, E.; Jongeneelen, M.; Poon, L.L.; Alard, P.; Cornelissen, L.; Bakker, A.; Cox, F.; van Deventer, E.; Guan, Y.; et al. Heterosubtypic neutralizing monoclonal antibodies cross-protective against H5N1 and H1N1 recovered from human IgM+ memory B cells. PLoS One 2008, 3, e3942. [Google Scholar]
  119. Wang, T.T.; Tan, G.S.; Hai, R.; Pica, N.; Ngai, L.; Ekiert, D.C.; Wilson, I.A.; Garcia-Sastre, A.; Moran, T.M.; Palese, P. Vaccination with a synthetic peptide from the influenza virus hemagglutinin provides protection against distinct viral subtypes. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 18979–18984. [Google Scholar]
  120. Okuno, Y.; Isegawa, Y.; Sasao, F.; Ueda, S. A common neutralizing epitope conserved between the hemagglutinins of influenza A virus H1 and H2 strains. J. Virol. 1993, 67, 2552–2558. [Google Scholar]
  121. Bosch, B.J.; Bodewes, R.; de Vries, R.P.; Kreijtz, J.H.; Bartelink, W.; van Amerongen, G.; Rimmelzwaan, G.F.; de Haan, C.A.; Osterhaus, A.D.; Rottier, P.J. Recombinant soluble, multimeric HA and NA exhibit distinctive types of protection against pandemic swine-origin 2009 A(H1N1) influenza virus infection in ferrets. J. Virol. 2010, 84, 10366–10374. [Google Scholar]
  122. Johansson, B.E.; Bucher, D.J.; Kilbourne, E.D. Purified influenza virus hemagglutinin and neuraminidase are equivalent in stimulation of antibody response but induce contrasting types of immunity to infection. J. Virol. 1989, 63, 1239–1246. [Google Scholar]
  123. Johansson, B.E.; Grajower, B.; Kilbourne, E.D. Infection-permissive immunization with influenza virus neuraminidase prevents weight loss in infected mice. Vaccine 1993, 11, 1037–1039. [Google Scholar] [CrossRef]
  124. Kilbourne, E.D.; Pokorny, B.A.; Johansson, B.; Brett, I.; Milev, Y.; Matthews, J.T. Protection of mice with recombinant influenza virus neuraminidase. J. Infect. Dis. 2004, 189, 459–461. [Google Scholar] [CrossRef]
  125. Schulman, J.L.; Khakpour, M.; Kilbourne, E.D. Protective effects of specific immunity to viral neuraminidase on influenza virus infection of mice. J. Virol. 1968, 2, 778–786. [Google Scholar]
  126. Treanor, J.J.; Tierney, E.L.; Zebedee, S.L.; Lamb, R.A.; Murphy, B.R. Passively transferred monoclonal antibody to the M2 protein inhibits influenza A virus replication in mice. J. Virol. 1990, 64, 1375–1377. [Google Scholar]
  127. Zebedee, S.L.; Lamb, R.A. Influenza A virus M2 protein: Monoclonal antibody restriction of virus growth and detection of M2 in virions. J. Virol. 1988, 62, 2762–2772. [Google Scholar]
  128. Mozdzanowska, K.; Maiese, K.; Furchner, M.; Gerhard, W. Treatment of influenza virus-infected SCID mice with nonneutralizing antibodies specific for the transmembrane proteins matrix 2 and neuraminidase reduces the pulmonary virus titer but fails to clear the infection. Virology 1999, 254, 138–146. [Google Scholar] [CrossRef]
  129. El Bakkouri, K.; Descamps, F.; de Filette, M.; Smet, A.; Festjens, E.; Birkett, A.; van Rooijen, N.; Verbeek, S.; Fiers, W.; Saelens, X. Universal vaccine based on ectodomain of matrix protein 2 of influenza A: Fc receptors and alveolar macrophages mediate protection. J. Immunol. 2011, 186, 1022–1031. [Google Scholar] [CrossRef]
  130. Ebrahimi, S.M.; Tebianian, M. Influenza A viruses: Why focusing on M2e-based universal vaccines. Virus Genes 2011, 42, 1–8. [Google Scholar] [CrossRef]
  131. Fiers, W.; de Filette, M.; Birkett, A.; Neirynck, S.; Min Jou, W. A "universal" human influenza A vaccine. Virus Res. 2004, 103, 173–176. [Google Scholar]
  132. Fiers, W.; de Filette, M.; El Bakkouri, K.; Schepens, B.; Roose, K.; Schotsaert, M.; Birkett, A.; Saelens, X. M2e-based universal influenza A vaccine. Vaccine 2009, 27, 6280–6283. [Google Scholar] [CrossRef]
  133. Neirynck, S.; Deroo, T.; Saelens, X.; Vanlandschoot, P.; Jou, W.M.; Fiers, W. A universal influenza A vaccine based on the extracellular domain of the M2 protein. Nat. Med. 1999, 5, 1157–1163. [Google Scholar]
  134. Wang, Y.; Zhou, L.; Shi, H.; Xu, H.; Yao, H.; Xi, X.G.; Toyoda, T.; Wang, X.; Wang, T. Monoclonal antibody recognizing SLLTEVET epitope of M2 protein potently inhibited the replication of influenza A viruses in MDCK cells. Biochem. Biophys. Res. Comm. 2009, 385, 118–122. [Google Scholar] [CrossRef]
  135. Fu, T.M.; Freed, D.C.; Horton, M.S.; Fan, J.; Citron, M.P.; Joyce, J.G.; Garsky, V.M.; Casimiro, D.R.; Zhao, Q.; Shiver, J.W.; et al. Characterizations of four monoclonal antibodies against M2 protein ectodomain of influenza A virus. Virology 2009, 385, 218–226. [Google Scholar]
  136. Schotsaert, M.; de Filette, M.; Fiers, W.; Saelens, X. Universal M2 ectodomain-based influenza A vaccines: Preclinical and clinical developments. Expert Rev. Vaccine 2009, 8, 499–508. [Google Scholar] [CrossRef]
  137. Tompkins, S.M.; Zhao, Z.S.; Lo, C.Y.; Misplon, J.A.; Liu, T.; Ye, Z.; Hogan, R.J.; Wu, Z.; Benton, K.A.; Tumpey, T.M.; et al. Matrix protein 2 vaccination and protection against influenza viruses, including subtype H5N1. Emerg. Infect. Dis. 2007, 13, 426–435. [Google Scholar] [CrossRef]
  138. Sukeno, N.; Otsuki, Y.; Konno, J.; Yamane, N.; Odagiri, T.; Arikawa, J.; Ishida, N. Anti-nucleoprotein antibody response in influenza A infection. Tohoku J. Exp. Med. 1979, 128, 241–249. [Google Scholar] [CrossRef]
  139. Lamere, M.W.; Moquin, A.; Lee, F.E.; Misra, R.S.; Blair, P.J.; Haynes, L.; Randall, T.D.; Lund, F.E.; Kaminski, D.A. Regulation of antinucleoprotein IgG by systemic vaccination and its effect on influenza virus clearance. J. Virol. 2011, 85, 5027–5035. [Google Scholar] [CrossRef]
  140. Carragher, D.M.; Kaminski, D.A.; Moquin, A.; Hartson, L.; Randall, T.D. A novel role for non-neutralizing antibodies against nucleoprotein in facilitating resistance to influenza virus. J. Immunol. 2008, 181, 4168–4176. [Google Scholar]
  141. Bodewes, R.; Osterhaus, A.D.; Rimmelzwaan, G.F. Targets for the induction of protective immunity against influenza a viruses. Viruses 2010, 2, 166–188. [Google Scholar] [CrossRef]
  142. Sambhara, S.; Kurichh, A.; Miranda, R.; Tumpey, T.; Rowe, T.; Renshaw, M.; Arpino, R.; Tamane, A.; Kandil, A.; James, O.; et al. Heterosubtypic immunity against human influenza A viruses, including recently emerged avian H5 and H9 viruses, induced by FLU-ISCOM vaccine in mice requires both cytotoxic T-lymphocyte and macrophage function. Cell. Immunol. 2001, 211, 143–153. [Google Scholar]
  143. Rimmelzwaan, G.F.; Baars, M.; van Beek, R.; van Amerongen, G.; Lovgren-Bengtsson, K.; Claas, E.C.; Osterhaus, A.D. Induction of protective immunity against influenza virus in a macaque model: Comparison of conventional and iscom vaccines. J. Gen. Virol. 1997, 78, 757–765. [Google Scholar]
  144. Fernandez Gonzalez, S.; Jayasekera, J.P.; Carroll, M.C. Complement and natural antibody are required in the long-term memory response to influenza virus. Vaccine 2008, 26, I86–I93. [Google Scholar]
  145. Jayasekera, J.P.; Moseman, E.A.; Carroll, M.C. Natural antibody and complement mediate neutralization of influenza virus in the absence of prior immunity. J. Virol. 2007, 81, 3487–3494. [Google Scholar]
  146. Voeten, J.T.; Groen, J.; van Alphen, D.; Claas, E.C.; de Groot, R.; Osterhaus, A.D.; Rimmelzwaan, G.F. Use of recombinant nucleoproteins in enzyme-linked immunosorbent assays for detection of virus-specific immunoglobulin A (IgA) and IgG antibodies in influenza virus A- or B-infected patients. J. Clin. Microbiol. 1998, 36, 3527–3531. [Google Scholar]
  147. Rothbarth, P.H.; Groen, J.; Bohnen, A.M.; de Groot, R.; Osterhaus, A.D. Influenza virus serology—A comparative study. J. Virol. Meth. 1999, 78, 163–169. [Google Scholar] [CrossRef]
  148. Koutsonanos, D.G.; del Pilar Martin, M.; Zarnitsyn, V.G.; Jacob, J.; Prausnitz, M.R.; Compans, R.W.; Skountzou, I. Serological memory and long-term protection to novel H1N1 influenza virus after skin vaccination. J. Infect. Dis. 2011, 204, 582–591. [Google Scholar] [CrossRef]
  149. Onodera, T.; Takahashi, Y.; Yokoi, Y.; Ato, M.; Kodama, Y.; Hachimura, S.; Kurosaki, T.; Kobayashi, K. Memory B cells in the lung participate in protective humoral immune responses to pulmonary influenza virus reinfection. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 2485–2490. [Google Scholar]
  150. Jones, P.D.; Ada, G.L. Persistence of influenza virus-specific antibody-secreting cells and B-cell memory after primary murine influenza virus infection. Cell. Immunol. 1987, 109, 53–64. [Google Scholar] [CrossRef]
  151. Armstrong, S.J.; Dimmock, N.J. Neutralization of influenza virus by low concentrations of hemagglutinin-specific polymeric immunoglobulin A inhibits viral fusion activity, but activation of the ribonucleoprotein is also inhibited. J. Virol. 1992, 66, 3823–3832. [Google Scholar]
  152. Mazanec, M.B.; Coudret, C.L.; Fletcher, D.R. Intracellular neutralization of influenza virus by immunoglobulin A anti-hemagglutinin monoclonal antibodies. J. Virol. 1995, 69, 1339–1343. [Google Scholar]
  153. Zuccotti, G.; Pogliani, L.; Pariani, E.; Amendola, A.; Zanetti, A. Transplacental antibody transfer following maternal immunization with a pandemic 2009 influenza A(H1N1) MF59-adjuvanted vaccine. JAMA 2010, 304, 2360–2361. [Google Scholar]
  154. Bodewes, R.; de Mutsert, G.; van der Klis, F.R.; Ventresca, M.; Wilks, S.; Smith, D.J.; Koopmans, M.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Prevalence of antibodies against seasonal influenza A and B viruses in children in Netherlands. Clin. Vaccine Immunol. 2011, 18, 469–476. [Google Scholar]
  155. Mbawuike, I.N.; Six, H.R.; Cate, T.R.; Couch, R.B. Vaccination with inactivated influenza A virus during pregnancy protects neonatal mice against lethal challenge by influenza A viruses representing three subtypes. J. Virol. 1990, 64, 1370–1374. [Google Scholar]
  156. Hwang, S.D.; Shin, J.S.; Ku, K.B.; Kim, H.S.; Cho, S.W.; Seo, S.H. Protection of pregnant mice, fetuses and neonates from lethality of H5N1 influenza viruses by maternal vaccination. Vaccine 2010, 28, 2957–2964. [Google Scholar]
  157. Steinhoff, M.C.; Omer, S.B.; Roy, E.; Arifeen, S.E.; Raqib, R.; Altaye, M.; Breiman, R.F.; M, B.B.S.K. Influenza immunization in pregnancy—Antibody responses in mothers and infants. New Engl. J. Med. 2010, 362, 1644–1646. [Google Scholar]
  158. Zhu, J.; Paul, W.E. Peripheral CD4+ T-cell differentiation regulated by networks of cytokines and transcription factors. Immunol. Rev. 2010, 238, 247–262. [Google Scholar]
  159. Lamb, J.R.; Woody, J.N.; Hartzman, R.J.; Eckels, D.D. In vitro influenza virus-specific antibody production in man: Antigen-specific and HLA-restricted induction of helper activity mediated by cloned human T lymphocytes. J. Immunol. 1982, 129, 1465–1470. [Google Scholar]
  160. Okoye, I.S.; Wilson, M.S. CD4+ T helper 2 cells—Microbial triggers, differentiation requirements and effector functions. Immunology 2011, 134, 368–377. [Google Scholar] [CrossRef]
  161. Eichelberger, M.C.; Wang, M.L.; Allan, W.; Webster, R.G.; Doherty, P.C. Influenza virus RNA in the lung and lymphoid tissue of immunologically intact and CD4-depleted mice. J. Gen. Virol. 1991, 72, 1695–1698. [Google Scholar] [CrossRef]
  162. Justewicz, D.M.; Doherty, P.C.; Webster, R.G. The B-cell response in lymphoid tissue of mice immunized with various antigenic forms of the influenza virus hemagglutinin. J. Virol. 1995, 69, 5414–5421. [Google Scholar]
  163. Kamperschroer, C.; Dibble, J.P.; Meents, D.L.; Schwartzberg, P.L.; Swain, S.L. SAP is required for Th cell function and for immunity to influenza. J. Immunol. 2006, 177, 5317–5327. [Google Scholar]
  164. Scherle, P.A.; Gerhard, W. Functional analysis of influenza-specific helper T cell clones in vivo. T cells specific for internal viral proteins provide cognate help for B cell responses to hemagglutinin. J. Exp. Med. 1986, 164, 1114–1128. [Google Scholar] [CrossRef]
  165. Schonbeck, U.; Libby, P. The CD40/CD154 receptor/ligand dyad. Cell. Mol. Life Sci. 2001, 58, 4–43. [Google Scholar] [CrossRef]
  166. Zhu, J.; Paul, W.E. Heterogeneity and plasticity of T helper cells. Cell Res. 2010, 20, 4–12. [Google Scholar]
  167. Mosmann, T.R.; Cherwinski, H.; Bond, M.W.; Giedlin, M.A.; Coffman, R.L. Two types of murine helper T cell clone. I. Definition according to profiles of lymphokine activities and secreted proteins. J. Immunol. 1986, 136, 2348–2357. [Google Scholar]
  168. Belz, G.T.; Wodarz, D.; Diaz, G.; Nowak, M.A.; Doherty, P.C. Compromised influenza virus-specific CD8(+)-T-cell memory in CD4(+)-T-cell-deficient mice. J. Virol. 2002, 76, 12388–12393. [Google Scholar] [CrossRef]
  169. Deliyannis, G.; Jackson, D.C.; Ede, N.J.; Zeng, W.; Hourdakis, I.; Sakabetis, E.; Brown, L.E. Induction of long-term memory CD8(+) T cells for recall of viral clearing responses against influenza virus. J. Virol. 2002, 76, 4212–4221. [Google Scholar]
  170. Riberdy, J.M.; Christensen, J.P.; Branum, K.; Doherty, P.C. Diminished primary and secondary influenza virus-specific CD8(+) T-cell responses in CD4-depleted Ig(-/-) mice. J. Virol. 2000, 74, 9762–9765. [Google Scholar]
  171. Strutt, T.M.; McKinstry, K.K.; Dibble, J.P.; Winchell, C.; Kuang, Y.; Curtis, J.D.; Huston, G.; Dutton, R.W.; Swain, S.L. Memory CD4+ T cells induce innate responses independently of pathogen. Nat. Med. 2010, 16 551p following 564, 558–564. [Google Scholar]
  172. Teijaro, J.R.; Turner, D.; Pham, Q.; Wherry, E.J.; Lefrancois, L.; Farber, D.L. Cutting edge: Tissue-retentive lung memory CD4 T cells mediate optimal protection to respiratory virus infection. J. Immunol. 2011, 187, 5510–5514. [Google Scholar]
  173. Brown, D.M.; Dilzer, A.M.; Meents, D.L.; Swain, S.L. CD4 T cell-mediated protection from lethal influenza: Perforin and antibody-mediated mechanisms give a one-two punch. J. Immunol. 2006, 177, 2888–2898. [Google Scholar]
  174. Graham, M.B.; Braciale, V.L.; Braciale, T.J. Influenza virus-specific CD4+ T helper type 2 T lymphocytes do not promote recovery from experimental virus infection. J. Exp. Med. 1994, 180, 1273–1282. [Google Scholar]
  175. Wilkinson, T.M.; Li, C.K.; Chui, C.S.; Huang, A.K.; Perkins, M.; Liebner, J.C.; Lambkin-Williams, R.; Gilbert, A.; Oxford, J.; Nicholas, B.; et al. Preexisting influenza-specific CD4+ T cells correlate with disease protection against influenza challenge in humans. Nat. Med. 2012, 18, 274–280. [Google Scholar]
  176. McKinstry, K.K.; Strutt, T.M.; Swain, S.L. Hallmarks of CD4 T cell immunity against influenza. J. Intern. Med. 2011, 269, 507–518. [Google Scholar] [CrossRef]
  177. Nakanishi, Y.; Lu, B.; Gerard, C.; Iwasaki, A. CD8(+) T lymphocyte mobilization to virus-infected tissue requires CD4(+) T-cell help. Nature 2009, 462, 510–513. [Google Scholar]
  178. Assarsson, E.; Bui, H.H.; Sidney, J.; Zhang, Q.; Glenn, J.; Oseroff, C.; Mbawuike, I.N.; Alexander, J.; Newman, M.J.; Grey, H.; et al. Immunomic analysis of the repertoire of T-cell specificities for influenza A virus in humans. J. Virol. 2008, 82, 12241–12251. [Google Scholar]
  179. Bednarek, M.A.; Sauma, S.Y.; Gammon, M.C.; Porter, G.; Tamhankar, S.; Williamson, A.R.; Zweerink, H.J. The minimum peptide epitope from the influenza virus matrix protein. Extra and intracellular loading of HLA-A2. J. Immunol. 1991, 147, 4047–4053. [Google Scholar]
  180. Boon, A.C.; de Mutsert, G.; Graus, Y.M.; Fouchier, R.A.; Sintnicolaas, K.; Osterhaus, A.D.; Rimmelzwaan, G.F. The magnitude and specificity of influenza A virus-specific cytotoxic T-lymphocyte responses in humans is related to HLA-A and -B phenotype. J. Virol. 2002, 76, 582–590. [Google Scholar] [CrossRef]
  181. Gotch, F.; McMichael, A.; Smith, G.; Moss, B. Identification of viral molecules recognized by influenza-specific human cytotoxic T lymphocytes. J. Exp. Med. 1987, 165, 408–416. [Google Scholar]
  182. Jameson, J.; Cruz, J.; Ennis, F.A. Human cytotoxic T-lymphocyte repertoire to influenza A viruses. J. Virol. 1998, 72, 8682–8689. [Google Scholar]
  183. Kreijtz, J.H.; de Mutsert, G.; van Baalen, C.A.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Cross-recognition of avian H5N1 influenza virus by human cytotoxic T-lymphocyte populations directed to human influenza A virus. J. Virol. 2008, 82, 5161–5166. [Google Scholar]
  184. Lee, L.Y.; Ha do, L.A.; Simmons, C.; de Jong, M.D.; Chau, N.V.; Schumacher, R.; Peng, Y.C.; McMichael, A.J.; Farrar, J.J.; Smith, G.L.; et al. Memory T cells established by seasonal human influenza A infection cross-react with avian influenza A (H5N1) in healthy individuals. J. Clin. Investig. 2008, 118, 3478–3490. [Google Scholar]
  185. Andrade, F. Non-cytotoxic antiviral activities of granzymes in the context of the immune antiviral state. Immunol. Rev. 2010, 235, 128–146. [Google Scholar]
  186. Moffat, J.M.; Gebhardt, T.; Doherty, P.C.; Turner, S.J.; Mintern, J.D. Granzyme A expression reveals distinct cytolytic CTL subsets following influenza A virus infection. Eur. J. Immunol. 2009, 39, 1203–1210. [Google Scholar] [CrossRef]
  187. Topham, D.J.; Tripp, R.A.; Doherty, P.C. CD8+ T cells clear influenza virus by perforin or Fas-dependent processes. J. Immunol. 1997, 159, 5197–5200. [Google Scholar]
  188. La Gruta, N.L.; Turner, S.J.; Doherty, P.C. Hierarchies in cytokine expression profiles for acute and resolving influenza virus-specific CD8+ T cell responses: Correlation of cytokine profile and TCR avidity. J. Immunol. 2004, 172, 5553–5560. [Google Scholar]
  189. Metkar, S.S.; Menaa, C.; Pardo, J.; Wang, B.; Wallich, R.; Freudenberg, M.; Kim, S.; Raja, S.M.; Shi, L.; Simon, M.M.; et al. Human and mouse granzyme A induce a proinflammatory cytokine response. Immunity 2008, 29, 720–733. [Google Scholar]
  190. van Domselaar, R.; Bovenschen, N. Cell death-independent functions of granzymes: Hit viruses where it hurts. Rev. Med. Virol. 2011, 21, 301–314. [Google Scholar]
  191. La Gruta, N.L.; Kedzierska, K.; Stambas, J.; Doherty, P.C. A question of self-preservation: Immunopathology in influenza virus infection. Immunol. Cell Biol. 2007, 85, 85–92. [Google Scholar]
  192. Woodland, D.L.; Hogan, R.J.; Zhong, W. Cellular immunity and memory to respiratory virus infections. Immunol. Res. 2001, 24, 53–67. [Google Scholar] [CrossRef]
  193. Kedl, R.M.; Mescher, M.F. Qualitative differences between naive and memory T cells make a major contribution to the more rapid and efficient memory CD8+ T cell response. J. Immunol. 1998, 161, 674–683. [Google Scholar]
  194. Lalvani, A.; Brookes, R.; Hambleton, S.; Britton, W.J.; Hill, A.V.; McMichael, A.J. Rapid effector function in CD8+ memory T cells. J. Exp. Med. 1997, 186, 859–865. [Google Scholar] [CrossRef]
  195. Zimmermann, C.; Prevost-Blondel, A.; Blaser, C.; Pircher, H. Kinetics of the response of naive and memory CD8 T cells to antigen: Similarities and differences. Eur. J. Immunol. 1999, 29, 284–290. [Google Scholar] [CrossRef]
  196. DiSpirito, J.R.; Shen, H. Quick to remember, slow to forget: Rapid recall responses of memory CD8+ T cells. Cell Res. 2010, 20, 13–23. [Google Scholar]
  197. Chang, J.T.; Palanivel, V.R.; Kinjyo, I.; Schambach, F.; Intlekofer, A.M.; Banerjee, A.; Longworth, S.A.; Vinup, K.E.; Mrass, P.; Oliaro, J.; et al. Asymmetric T lymphocyte division in the initiation of adaptive immune responses. Science 2007, 315, 1687–1691. [Google Scholar] [CrossRef]
  198. Hikono, H.; Kohlmeier, J.E.; Takamura, S.; Wittmer, S.T.; Roberts, A.D.; Woodland, D.L. Activation phenotype, rather than central- or effector-memory phenotype, predicts the recall efficacy of memory CD8+ T cell. J. Exp. Med. 2007, 204, 1625–1636. [Google Scholar]
  199. Sallusto, F.; Lenig, D.; Forster, R.; Lipp, M.; Lanzavecchia, A. Two subsets of memory T lymphocytes with distinct homing potentials and effector functions. Nature 1999, 401, 708–712. [Google Scholar]
  200. van Gisbergen, K.P.; Klarenbeek, P.L.; Kragten, N.A.; Unger, P.P.; Nieuwenhuis, M.B.; Wensveen, F.M.; ten Brinke, A.; Tak, P.P.; Eldering, E.; Nolte, M.A.; et al. The costimulatory molecule CD27 maintains clonally diverse CD8(+) T cell responses of low antigen affinity to protect against viral variants. Immunity 2011, 35, 97–108. [Google Scholar]
  201. Wherry, E.J.; Teichgraber, V.; Becker, T.C.; Masopust, D.; Kaech, S.M.; Antia, R.; von Andrian, U.H.; Ahmed, R. Lineage relationship and protective immunity of memory CD8 T cell subsets. Nat. Immunol. 2003, 4, 225–234. [Google Scholar]
  202. Bender, B.S.; Croghan, T.; Zhang, L.; Small, P.A., Jr. Transgenic mice lacking class I major histocompatibility complex-restricted T cells have delayed viral clearance and increased mortality after influenza virus challenge. J. Exp. Med. 1992, 175, 1143–1145. [Google Scholar]
  203. Graham, M.B.; Braciale, T.J. Resistance to and recovery from lethal influenza virus infection in B lymphocyte-deficient mice. J. Exp. Med. 1997, 186, 2063–2068. [Google Scholar]
  204. Hillaire, M.L.; van Trierum, S.E.; Kreijtz, J.H.; Bodewes, R.; Geelhoed-Mieras, M.M.; Nieuwkoop, N.J.; Fouchier, R.A.; Kuiken, T.; Osterhaus, A.D.; Rimmelzwaan, G.F. Cross-protective immunity against influenza pH1N1 2009 viruses induced by seasonal influenza A (H3N2) virus is mediated by virus-specific T-cells. J. Gen. Virol. 2011, 92, 2339–2349. [Google Scholar]
  205. Kreijtz, J.H.; Bodewes, R.; van Amerongen, G.; Kuiken, T.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Primary influenza A virus infection induces cross-protective immunity against a lethal infection with a heterosubtypic virus strain in mice. Vaccine 2007, 25, 612–620. [Google Scholar]
  206. Kreijtz, J.H.; Bodewes, R.; van den Brand, J.M.; de Mutsert, G.; Baas, C.; van Amerongen, G.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Infection of mice with a human influenza A/H3N2 virus induces protective immunity against lethal infection with influenza A/H5N1 virus. Vaccine 2009, 27, 4983–4989. [Google Scholar]
  207. Taylor, P.M.; Askonas, B.A. Influenza nucleoprotein-specific cytotoxic T-cell clones are protective in vivo. Immunology 1986, 58, 417–420. [Google Scholar]
  208. Hillaire, M.L.; Osterhaus, A.D.; Rimmelzwaan, G.F. Induction of virus-specific cytotoxic T lymphocytes as a basis for the development of broadly protective influenza vaccines. J. Biomed. Biotechnol. 2011, 2011, 939860. [Google Scholar]
  209. Grebe, K.M.; Yewdell, J.W.; Bennink, J.R. Heterosubtypic immunity to influenza A virus: Where do we stand? Microb. Infect. 2008, 10, 1024–1029. [Google Scholar] [CrossRef]
  210. Sridhar, S.; Begom, S.; Bermingham, A.; Ziegler, T.; Roberts, K.L.; Barclay, W.S.; Openshaw, P.; Lalvani, A. Predominance of heterosubtypic IFN-gamma-only-secreting effector memory T cells in pandemic H1N1 naive adults. Eur. J. Immunol. 2012. [Google Scholar]
  211. McMichael, A.J.; Gotch, F.M.; Noble, G.R.; Beare, P.A. Cytotoxic T-cell immunity to influenza. New Engl. J. Med. 1983, 309, 13–17. [Google Scholar]
  212. Epstein, S.L. Prior H1N1 influenza infection and susceptibility of Cleveland Family Study participants during the H2N2 pandemic of 1957: An experiment of nature. J. Infect. Dis. 2006, 193, 49–53. [Google Scholar]
  213. Slepushkin, A.N. The effect of a previous attack of A1 influenza on susceptibility to A2 virus during the 1957 outbreak. Bull. World Health Organ. 1959, 20, 297–301. [Google Scholar]
  214. Smallman-Raynor, M.; Cliff, A.D. Avian influenza A (H5N1) age distribution in humans. Emerg. Infect. Dis. 2007, 13, 510–512. [Google Scholar] [CrossRef]
  215. Campbell, D.J.; Koch, M.A. Phenotypical and functional specialization of FOXP3+ regulatory T cells. Nat. Rev. 2011, 11, 119–130. [Google Scholar] [CrossRef]
  216. Surls, J.; Nazarov-Stoica, C.; Kehl, M.; Casares, S.; Brumeanu, T.D. Differential effect of CD4+Foxp3+ T-regulatory cells on the B and T helper cell responses to influenza virus vaccination. Vaccine 2010, 28, 7319–7330. [Google Scholar] [CrossRef]
  217. Kudva, A.; Scheller, E.V.; Robinson, K.M.; Crowe, C.R.; Choi, S.M.; Slight, S.R.; Khader, S.A.; Dubin, P.J.; Enelow, R.I.; Kolls, J.K.; et al. Influenza A inhibits Th17-mediated host defense against bacterial pneumonia in mice. J. Immunol. 2011, 186, 1666–1674. [Google Scholar] [CrossRef]
  218. Garcia-Sastre, A.; Egorov, A.; Matassov, D.; Brandt, S.; Levy, D.E.; Durbin, J.E.; Palese, P.; Muster, T. Influenza A virus lacking the NS1 gene replicates in interferon-deficient systems. Virology 1998, 252, 324–330. [Google Scholar] [CrossRef]
  219. Talon, J.; Salvatore, M.; O'Neill, R.E.; Nakaya, Y.; Zheng, H.; Muster, T.; Garcia-Sastre, A.; Palese, P. Influenza A and B viruses expressing altered NS1 proteins: A vaccine approach. Proc. Natl. Acad. Sci. U. S. A. 2000, 97, 4309–4314. [Google Scholar]
  220. Falcon, A.M.; Fernandez-Sesma, A.; Nakaya, Y.; Moran, T.M.; Ortin, J.; Garcia-Sastre, A. Attenuation and immunogenicity in mice of temperature-sensitive influenza viruses expressing truncated NS1 proteins. J. Gen. Virol. 2005, 86, 2817–2821. [Google Scholar] [CrossRef]
  221. Donelan, N.R.; Basler, C.F.; Garcia-Sastre, A. A recombinant influenza A virus expressing an RNA-binding-defective NS1 protein induces high levels of beta interferon and is attenuated in mice. J. Virol. 2003, 77, 13257–13266. [Google Scholar] [CrossRef]
  222. Solorzano, A.; Webby, R.J.; Lager, K.M.; Janke, B.H.; Garcia-Sastre, A.; Richt, J.A. Mutations in the NS1 protein of swine influenza virus impair anti-interferon activity and confer attenuation in pigs. J. Virol. 2005, 79, 7535–7543. [Google Scholar]
  223. Richt, J.A.; Garcia-Sastre, A. Attenuated influenza virus vaccines with modified NS1 proteins. Curr. Top. Microbiol. Immunol. 2009, 333, 177–195. [Google Scholar]
  224. Park, H.J.; Ferko, B.; Byun, Y.H.; Song, J.H.; Han, G.Y.; Roethl, E.; Egorov, A.; Muster, T.; Seong, B.; Kweon, M.N.; et al. Sublingual immunization with a live attenuated influenza a virus lacking the nonstructural protein 1 induces broad protective immunity in mice. PLoS One 2012, 7, e39921. [Google Scholar]
  225. Gack, M.U.; Albrecht, R.A.; Urano, T.; Inn, K.S.; Huang, I.C.; Carnero, E.; Farzan, M.; Inoue, S.; Jung, J.U.; Garcia-Sastre, A. Influenza A virus NS1 targets the ubiquitin ligase TRIM25 to evade recognition by the host viral RNA sensor RIG-I. Cell Host Microbe 2009, 5, 439–449. [Google Scholar]
  226. Ludwig, S.; Wang, X.; Ehrhardt, C.; Zheng, H.; Donelan, N.; Planz, O.; Pleschka, S.; Garcia-Sastre, A.; Heins, G.; Wolff, T. The influenza A virus NS1 protein inhibits activation of Jun N-terminal kinase and AP-1 transcription factors. J. Virol. 2002, 76, 11166–11171. [Google Scholar]
  227. Talon, J.; Horvath, C.M.; Polley, R.; Basler, C.F.; Muster, T.; Palese, P.; Garcia-Sastre, A. Activation of interferon regulatory factor 3 is inhibited by the influenza A virus NS1 protein. J. Virol. 2000, 74, 7989–7996. [Google Scholar] [CrossRef]
  228. Wang, X.; Li, M.; Zheng, H.; Muster, T.; Palese, P.; Beg, A.A.; Garcia-Sastre, A. Influenza A virus NS1 protein prevents activation of NF-kappaB and induction of alpha/beta interferon. J. Virol. 2000, 74, 11566–11573. [Google Scholar]
  229. Ruckle, A.; Haasbach, E.; Julkunen, I.; Planz, O.; Ehrhardt, C.; Ludwig, S. The NS1 protein of influenza A virus blocks RIG-I mediated activation of the noncanonical NF-kappaB pathway and p52/RelB dependent gene expression in lung epithelial cells. J. Virol. 2012, 86, 10211–7. [Google Scholar] [CrossRef]
  230. Mibayashi, M.; Martinez-Sobrido, L.; Loo, Y.M.; Cardenas, W.B.; Gale, M., Jr.; Garcia-Sastre, A. Inhibition of retinoic acid-inducible gene I-mediated induction of beta interferon by the NS1 protein of influenza A virus. J. Virol. 2007, 81, 514–524. [Google Scholar]
  231. Opitz, B.; Rejaibi, A.; Dauber, B.; Eckhard, J.; Vinzing, M.; Schmeck, B.; Hippenstiel, S.; Suttorp, N.; Wolff, T. IFNbeta induction by influenza A virus is mediated by RIG-I which is regulated by the viral NS1 protein. Cell. Microbiol. 2007, 9, 930–938. [Google Scholar]
  232. Nemeroff, M.E.; Barabino, S.M.; Li, Y.; Keller, W.; Krug, R.M. Influenza virus NS1 protein interacts with the cellular 30 kDa subunit of CPSF and inhibits 3'end formation of cellular pre-mRNAs. Mol. Cell 1998, 1, 991–1000. [Google Scholar] [CrossRef]
  233. Noah, D.L.; Twu, K.Y.; Krug, R.M. Cellular antiviral responses against influenza A virus are countered at the posttranscriptional level by the viral NS1A protein via its binding to a cellular protein required for the 3' end processing of cellular pre-mRNAS. Virology 2003, 307, 386–395. [Google Scholar]
  234. Das, K.; Ma, L.C.; Xiao, R.; Radvansky, B.; Aramini, J.; Zhao, L.; Marklund, J.; Kuo, R.L.; Twu, K.Y.; Arnold, E.; et al. Structural basis for suppression of a host antiviral response by influenza A virus. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 13093–13098. [Google Scholar]
  235. Qian, X.Y.; Alonso-Caplen, F.; Krug, R.M. Two functional domains of the influenza virus NS1 protein are required for regulation of nuclear export of mRNA. J. Virol. 1994, 68, 2433–2441. [Google Scholar]
  236. Satterly, N.; Tsai, P.L.; van Deursen, J.; Nussenzveig, D.R.; Wang, Y.; Faria, P.A.; Levay, A.; Levy, D.E.; Fontoura, B.M. Influenza virus targets the mRNA export machinery and the nuclear pore complex. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 1853–1858. [Google Scholar]
  237. Varga, Z.T.; Ramos, I.; Hai, R.; Schmolke, M.; Garcia-Sastre, A.; Fernandez-Sesma, A.; Palese, P. The influenza virus protein PB1-F2 inhibits the induction of type I interferon at the level of the MAVS adaptor protein. PLoS Pathog. 2011, 7, e1002067. [Google Scholar] [CrossRef]
  238. Conenello, G.M.; Tisoncik, J.R.; Rosenzweig, E.; Varga, Z.T.; Palese, P.; Katze, M.G. A single N66S mutation in the PB1-F2 protein of influenza A virus increases virulence by inhibiting the early interferon response in vivo. J. Virol. 2011, 85, 652–662. [Google Scholar]
  239. Dudek, S.E.; Wixler, L.; Nordhoff, C.; Nordmann, A.; Anhlan, D.; Wixler, V.; Ludwig, S. The influenza virus PB1-F2 protein has interferon antagonistic activity. Biol. Chem. 2011, 392, 1135–1144. [Google Scholar]
  240. Graef, K.M.; Vreede, F.T.; Lau, Y.F.; McCall, A.W.; Carr, S.M.; Subbarao, K.; Fodor, E. The PB2 subunit of the influenza virus RNA polymerase affects virulence by interacting with the mitochondrial antiviral signaling protein and inhibiting expression of beta interferon. J. Virol. 2010, 84, 8433–8445. [Google Scholar] [CrossRef]
  241. Iwai, A.; Shiozaki, T.; Kawai, T.; Akira, S.; Kawaoka, Y.; Takada, A.; Kida, H.; Miyazaki, T. Influenza A virus polymerase inhibits type I interferon induction by binding to interferon beta promoter stimulator 1. J. Biol. Chem. 2010, 285, 32064–32074. [Google Scholar]
  242. Guilligay, D.; Tarendeau, F.; Resa-Infante, P.; Coloma, R.; Crepin, T.; Sehr, P.; Lewis, J.; Ruigrok, R.W.; Ortin, J.; Hart, D.J.; et al. The structural basis for cap binding by influenza virus polymerase subunit PB2. Nat. Struct. Mol. Biol. 2008, 15, 500–506. [Google Scholar]
  243. Sugiyama, K.; Obayashi, E.; Kawaguchi, A.; Suzuki, Y.; Tame, J.R.; Nagata, K.; Park, S.Y. Structural insight into the essential PB1-PB2 subunit contact of the influenza virus RNA polymerase. EMBO J. 2009, 28, 1803–1811. [Google Scholar] [CrossRef]
  244. Dias, A.; Bouvier, D.; Crepin, T.; McCarthy, A.A.; Hart, D.J.; Baudin, F.; Cusack, S.; Ruigrok, R.W. The cap-snatching endonuclease of influenza virus polymerase resides in the PA subunit. Nature 2009, 458, 914–918. [Google Scholar]
  245. Plotch, S.J.; Bouloy, M.; Ulmanen, I.; Krug, R.M. A unique cap(m7GpppXm)-dependent influenza virion endonuclease cleaves capped RNAs to generate the primers that initiate viral RNA transcription. Cell 1981, 23, 847–858. [Google Scholar] [CrossRef]
  246. Yuan, P.; Bartlam, M.; Lou, Z.; Chen, S.; Zhou, J.; He, X.; Lv, Z.; Ge, R.; Li, X.; Deng, T.; et al. Crystal structure of an avian influenza polymerase PA(N) reveals an endonuclease active site. Nature 2009, 458, 909–913. [Google Scholar]
  247. Goodman, A.G.; Smith, J.A.; Balachandran, S.; Perwitasari, O.; Proll, S.C.; Thomas, M.J.; Korth, M.J.; Barber, G.N.; Schiff, L.A.; Katze, M.G. The cellular protein P58IPK regulates influenza virus mRNA translation and replication through a PKR-mediated mechanism. J. Virol. 2007, 81, 2221–2230. [Google Scholar]
  248. Melville, M.W.; Hansen, W.J.; Freeman, B.C.; Welch, W.J.; Katze, M.G. The molecular chaperone hsp40 regulates the activity of P58IPK, the cellular inhibitor of PKR. Proc. Natl. Acad. Sci. U. S. A. 1997, 94, 97–102. [Google Scholar]
  249. Sharma, K.; Tripathi, S.; Ranjan, P.; Kumar, P.; Garten, R.; Deyde, V.; Katz, J.M.; Cox, N.J.; Lal, R.B.; Sambhara, S.; et al. Influenza A virus nucleoprotein exploits Hsp40 to inhibit PKR activation. PLoS One 2011, 6, e20215. [Google Scholar]
  250. Guan, Z.; Liu, D.; Mi, S.; Zhang, J.; Ye, Q.; Wang, M.; Gao, G.F.; Yan, J. Interaction of Hsp40 with influenza virus M2 protein: Implications for PKR signaling pathway. Protein Cell 2010, 1, 944–955. [Google Scholar] [CrossRef]
  251. Pauli, E.K.; Schmolke, M.; Wolff, T.; Viemann, D.; Roth, J.; Bode, J.G.; Ludwig, S. Influenza A virus inhibits type I IFN signaling via NF-kappaB-dependent induction of SOCS-3 expression. PLoS Pathog. 2008, 4, e1000196. [Google Scholar] [CrossRef]
  252. Pothlichet, J.; Chignard, M.; Si-Tahar, M. Cutting edge: Innate immune response triggered by influenza A virus is negatively regulated by SOCS1 and SOCS3 through a RIG-I/IFNAR1-dependent pathway. J. Immunol. 2008, 180, 2034–2038. [Google Scholar]
  253. Boliar, S.; Chambers, T.M. A new strategy of immune evasion by influenza A virus: Inhibition of monocyte differentiation into dendritic cells. Vet. Immunol. Immunopathol. 2010, 136, 201–210. [Google Scholar] [CrossRef]
  254. Fernandez-Sesma, A.; Marukian, S.; Ebersole, B.J.; Kaminski, D.; Park, M.S.; Yuen, T.; Sealfon, S.C.; Garcia-Sastre, A.; Moran, T.M. Influenza virus evades innate and adaptive immunity via the NS1 protein. J. Virol. 2006, 80, 6295–6304. [Google Scholar]
  255. Guo, H.; Kumar, P.; Malarkannan, S. Evasion of natural killer cells by influenza virus. J. Leukoc. Biol. 2011, 89, 189–194. [Google Scholar] [CrossRef]
  256. Owen, R.E.; Yamada, E.; Thompson, C.I.; Phillipson, L.J.; Thompson, C.; Taylor, E.; Zambon, M.; Osborn, H.M.; Barclay, W.S.; Borrow, P. Alterations in receptor binding properties of recent human influenza H3N2 viruses are associated with reduced natural killer cell lysis of infected cells. J. Virol. 2007, 81, 11170–11178. [Google Scholar]
  257. Mao, H.; Tu, W.; Liu, Y.; Qin, G.; Zheng, J.; Chan, P.L.; Lam, K.T.; Peiris, J.S.; Lau, Y.L. Inhibition of human natural killer cell activity by influenza virions and hemagglutinin. J. Virol. 2010, 84, 4148–4157. [Google Scholar] [CrossRef]
  258. Mao, H.; Tu, W.; Qin, G.; Law, H.K.; Sia, S.F.; Chan, P.L.; Liu, Y.; Lam, K.T.; Zheng, J.; Peiris, M.; et al. Influenza virus directly infects human natural killer cells and induces cell apoptosis. J. Virol. 2009, 83, 9215–9222. [Google Scholar]
  259. Rambaut, A.; Pybus, O.G.; Nelson, M.I.; Viboud, C.; Taubenberger, J.K.; Holmes, E.C. The genomic and epidemiological dynamics of human influenza A virus. Nature 2008, 453, 615–619. [Google Scholar]
  260. McHardy, A.C.; Adams, B. The role of genomics in tracking the evolution of influenza A virus. PLoS Pathog. 2009, 5, e1000566. [Google Scholar]
  261. Sorrell, E.M.; Schrauwen, E.J.; Linster, M.; De Graaf, M.; Herfst, S.; Fouchier, R.A. Predicting 'airborne' influenza viruses: (Trans-) mission impossible? Curr. Opin. Virol. 2011, 1, 635–642. [Google Scholar] [CrossRef]
  262. Bodewes, R.; Nieuwkoop, N.J.; Verburgh, R.J.; Fouchier, R.A.; Osterhaus, A.; Rimmelzwaan, G.F. The use of influenza A viruses expressing reporter genes to assess the frequency of double infections in vitro. J. Gen. Virol. 2012, 93, 1645–1648. [Google Scholar]
  263. Ito, T.; Couceiro, J.N.; Kelm, S.; Baum, L.G.; Krauss, S.; Castrucci, M.R.; Donatelli, I.; Kida, H.; Paulson, J.C.; Webster, R.G.; et al. Molecular basis for the generation in pigs of influenza A viruses with pandemic potential. J. Virol. 1998, 72, 7367–7373. [Google Scholar]
  264. Ma, W.; Lager, K.M.; Vincent, A.L.; Janke, B.H.; Gramer, M.R.; Richt, J.A. The role of swine in the generation of novel influenza viruses. Zoonoses Public Health 2009, 56, 326–337. [Google Scholar] [CrossRef]
  265. Suzuki, Y.; Ito, T.; Suzuki, T.; Holland, R.E., Jr.; Chambers, T.M.; Kiso, M.; Ishida, H.; Kawaoka, Y. Sialic acid species as a determinant of the host range of influenza A viruses. J. Virol. 2000, 74, 11825–11831. [Google Scholar] [CrossRef]
  266. Claas, E.C.; Kawaoka, Y.; de Jong, J.C.; Masurel, N.; Webster, R.G. Infection of children with avian-human reassortant influenza virus from pigs in Europe. Virology 1994, 204, 453–457. [Google Scholar]
  267. Pensaert, M.; Ottis, K.; Vandeputte, J.; Kaplan, M.M.; Bachmann, P.A. Evidence for the natural transmission of influenza A virus from wild ducts to swine and its potential importance for man. Bull. World Health Organ. 1981, 59, 75–78. [Google Scholar]
  268. Kilbourne, E.D. Influenza pandemics of the 20th century. Emerg. Infect. Dis. 2006, 12, 9–14. [Google Scholar] [CrossRef]
  269. Smith, G.J.; Bahl, J.; Vijaykrishna, D.; Zhang, J.; Poon, L.L.; Chen, H.; Webster, R.G.; Peiris, J.S.; Guan, Y. Dating the emergence of pandemic influenza viruses. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 11709–11712. [Google Scholar]
  270. Scholtissek, C.; Rohde, W.; Von Hoyningen, V.; Rott, R. On the origin of the human influenza virus subtypes H2N2 and H3N2. Virology 1978, 87, 13–20. [Google Scholar] [CrossRef]
  271. Smith, G.J.; Vijaykrishna, D.; Bahl, J.; Lycett, S.J.; Worobey, M.; Pybus, O.G.; Ma, S.K.; Cheung, C.L.; Raghwani, J.; Bhatt, S.; et al. Origins and evolutionary genomics of the 2009 swine-origin H1N1 influenza A epidemic. Nature 2009, 459, 1122–1125. [Google Scholar]
  272. Thontiravong, A.; Kitikoon, P.; Wannaratana, S.; Tantilertcharoen, R.; Tuanudom, R.; Pakpinyo, S.; Sasipreeyajan, J.; Oraveerakul, K.; Amonsin, A. Quail as a potential mixing vessel for the generation of new reassortant influenza A viruses. Vet. Microbiol. 2012, in press. [Google Scholar]
  273. Han, T.; Marasco, W.A. Structural basis of influenza virus neutralization. Ann. New York Acad. Sci. 2011, 1217, 178–190. [Google Scholar]
  274. Kwong, P.D.; Doyle, M.L.; Casper, D.J.; Cicala, C.; Leavitt, S.A.; Majeed, S.; Steenbeke, T.D.; Venturi, M.; Chaiken, I.; Fung, M.; et al. HIV-1 evades antibody-mediated neutralization through conformational masking of receptor-binding sites. Nature 2002, 420, 678–682. [Google Scholar]
  275. Wyatt, R.; Kwong, P.D.; Desjardins, E.; Sweet, R.W.; Robinson, J.; Hendrickson, W.A.; Sodroski, J.G. The antigenic structure of the HIV gp120 envelope glycoprotein. Nature 1998, 393, 705–711. [Google Scholar]
  276. Horst, D.; Verweij, M.C.; Davison, A.J.; Ressing, M.E.; Wiertz, E.J. Viral evasion of T cell immunity: Ancient mechanisms offering new applications. Curr. Opin. Immunol. 2011, 23, 96–103. [Google Scholar]
  277. Berkhoff, E.G.; de Wit, E.; Geelhoed-Mieras, M.M.; Boon, A.C.; Symons, J.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Functional constraints of influenza A virus epitopes limit escape from cytotoxic T lymphocytes. J. Virol. 2005, 79, 11239–11246. [Google Scholar]
  278. Del Val, M.; Schlicht, H.J.; Ruppert, T.; Reddehase, M.J.; Koszinowski, U.H. Efficient processing of an antigenic sequence for presentation by MHC class I molecules depends on its neighboring residues in the protein. Cell 1991, 66, 1145–1153. [Google Scholar] [CrossRef]
  279. Eisenlohr, L.C.; Yewdell, J.W.; Bennink, J.R. Flanking sequences influence the presentation of an endogenously synthesized peptide to cytotoxic T lymphocytes. J. Exp. Med. 1992, 175, 481–487. [Google Scholar]
  280. Neisig, A.; Roelse, J.; Sijts, A.J.; Ossendorp, F.; Feltkamp, M.C.; Kast, W.M.; Melief, C.J.; Neefjes, J.J. Major differences in transporter associated with antigen presentation (TAP)-dependent translocation of MHC class I-presentable peptides and the effect of flanking sequences. J. Immunol. 1995, 154, 1273–1279. [Google Scholar]
  281. Milicic, A.; Price, D.A.; Zimbwa, P.; Booth, B.L.; Brown, H.L.; Easterbrook, P.J.; Olsen, K.; Robinson, N.; Gileadi, U.; Sewell, A.K.; et al. CD8+ T cell epitope-flanking mutations disrupt proteasomal processing of HIV-1 Nef. J. Immunol. 2005, 175, 4618–4626. [Google Scholar]
  282. Berkhoff, E.G.; Boon, A.C.; Nieuwkoop, N.J.; Fouchier, R.A.; Sintnicolaas, K.; Osterhaus, A.D.; Rimmelzwaan, G.F. A mutation in the HLA-B*2705-restricted NP383-391 epitope affects the human influenza A virus-specific cytotoxic T-lymphocyte response in vitro. J. Virol. 2004, 78, 5216–5222. [Google Scholar]
  283. Berkhoff, E.G.; Geelhoed-Mieras, M.M.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Assessment of the extent of variation in influenza A virus cytotoxic T-lymphocyte epitopes by using virus-specific CD8+ T-cell clones. J. Gen. Virol. 2007, 88, 530–535. [Google Scholar] [CrossRef]
  284. Price, G.E.; Ou, R.; Jiang, H.; Huang, L.; Moskophidis, D. Viral escape by selection of cytotoxic T cell-resistant variants in influenza A virus pneumonia. J. Exp. Med. 2000, 191, 1853–1867. [Google Scholar] [CrossRef]
  285. Rimmelzwaan, G.F.; Boon, A.C.; Voeten, J.T.; Berkhoff, E.G.; Fouchier, R.A.; Osterhaus, A.D. Sequence variation in the influenza A virus nucleoprotein associated with escape from cytotoxic T lymphocytes. Virus Res. 2004, 103, 97–100. [Google Scholar] [CrossRef]
  286. Voeten, J.T.; Bestebroer, T.M.; Nieuwkoop, N.J.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Antigenic drift in the influenza A virus (H3N2) nucleoprotein and escape from recognition by cytotoxic T lymphocytes. J. Virol. 2000, 74, 6800–6807. [Google Scholar] [CrossRef]
  287. Berkhoff, E.G.; Geelhoed-Mieras, M.M.; Verschuren, E.J.; van Baalen, C.A.; Gruters, R.A.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. The loss of immunodominant epitopes affects interferon-gamma production and lytic activity of the human influenza virus-specific cytotoxic T lymphocyte response in vitro. Clin. Exp. Immunol. 2007, 148, 296–306. [Google Scholar] [CrossRef]
  288. Boon, A.C.; de Mutsert, G.; Graus, Y.M.; Fouchier, R.A.; Sintnicolaas, K.; Osterhaus, A.D.; Rimmelzwaan, G.F. Sequence variation in a newly identified HLA-B35-restricted epitope in the influenza A virus nucleoprotein associated with escape from cytotoxic T lymphocytes. J. Virol. 2002, 76, 2567–2572. [Google Scholar]
  289. Cale, E.M.; Bazick, H.S.; Rianprakaisang, T.A.; Alam, S.M.; Letvin, N.L. Mutations in a dominant Nef epitope of simian immunodeficiency virus diminish TCR: Epitope peptide affinity but not epitope peptide:MHC class I binding. J. Immunol. 2011, 187, 3300–3313. [Google Scholar] [CrossRef]
  290. Huet, S.; Nixon, D.F.; Rothbard, J.B.; Townsend, A.; Ellis, S.A.; McMichael, A.J. Structural homologies between two HLA B27-restricted peptides suggest residues important for interaction with HLA B27. Int. Immunol. 1990, 2, 311–316. [Google Scholar] [CrossRef]
  291. Gog, J.R.; Rimmelzwaan, G.F.; Osterhaus, A.D.; Grenfell, B.T. Population dynamics of rapid fixation in cytotoxic T lymphocyte escape mutants of influenza A. Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 11143–11147. [Google Scholar]
  292. Boon, A.C.; de Mutsert, G.; van Baarle, D.; Smith, D.J.; Lapedes, A.S.; Fouchier, R.A.; Sintnicolaas, K.; Osterhaus, A.D.; Rimmelzwaan, G.F. Recognition of homo- and heterosubtypic variants of influenza A viruses by human CD8+ T lymphocytes. J. Immunol. 2004, 172, 2453–2460. [Google Scholar]
  293. Gras, S.; Kedzierski, L.; Valkenburg, S.A.; Laurie, K.; Liu, Y.C.; Denholm, J.T.; Richards, M.J.; Rimmelzwaan, G.F.; Kelso, A.; Doherty, P.C.; et al. Cross-reactive CD8+ T-cell immunity between the pandemic H1N1-2009 and H1N1-1918 influenza A viruses. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 12599–12604. [Google Scholar]
  294. Berkhoff, E.G.; de Wit, E.; Geelhoed-Mieras, M.M.; Boon, A.C.; Symons, J.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Fitness costs limit escape from cytotoxic T lymphocytes by influenza A viruses. Vaccine 2006, 24, 6594–6596. [Google Scholar] [CrossRef]
  295. Rimmelzwaan, G.F.; Berkhoff, E.G.; Nieuwkoop, N.J.; Fouchier, R.A.; Osterhaus, A.D. Functional compensation of a detrimental amino acid substitution in a cytotoxic-T-lymphocyte epitope of influenza a viruses by comutations. J. Virol. 2004, 78, 8946–8949. [Google Scholar] [CrossRef]
  296. Rimmelzwaan, G.F.; Berkhoff, E.G.; Nieuwkoop, N.J.; Smith, D.J.; Fouchier, R.A.; Osterhaus, A.D. Full restoration of viral fitness by multiple compensatory co-mutations in the nucleoprotein of influenza A virus cytotoxic T-lymphocyte escape mutants. J. Gen. Virol. 2005, 86, 1801–1805. [Google Scholar] [CrossRef]
  297. Berkhoff, E.G.; Geelhoed-Mieras, M.M.; Jonges, M.; Smith, D.J.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. An amino acid substitution in the influenza A virus hemagglutinin associated with escape from recognition by human virus-specific CD4+ T-cells. Virus Res. 2007, 126, 282–287. [Google Scholar] [CrossRef]
  298. Cox, R.J.; Brokstad, K.A.; Ogra, P. Influenza virus: Immunity and vaccination strategies. Comparison of the immune response to inactivated and live, attenuated influenza vaccines. Scand. J. Immunol. 2004, 59, 1–15. [Google Scholar] [CrossRef]
  299. Thomas, P.G.; Keating, R.; Hulse-Post, D.J.; Doherty, P.C. Cell-mediated protection in influenza infection. Emerg. Infect. Dis. 2006, 12, 48–54. [Google Scholar] [CrossRef]
  300. Kreijtz, J.H.; Osterhaus, A.D.; Rimmelzwaan, G.F. Vaccination strategies and vaccine formulations for epidemic and pandemic influenza control. Hum. Vaccine. 2009, 5, 126–135. [Google Scholar] [CrossRef]
  301. Russell, C.A.; Jones, T.C.; Barr, I.G.; Cox, N.J.; Garten, R.J.; Gregory, V.; Gust, I.D.; Hampson, A.W.; Hay, A.J.; Hurt, A.C.; et al. Influenza vaccine strain selection and recent studies on the global migration of seasonal influenza viruses. Vaccine 2008, 26, D31–D34. [Google Scholar]
  302. de Jong, J.C.; Beyer, W.E.; Palache, A.M.; Rimmelzwaan, G.F.; Osterhaus, A.D. Mismatch between the 1997/1998 influenza vaccine and the major epidemic A(H3N2) virus strain as the cause of an inadequate vaccine-induced antibody response to this strain in the elderly. J. Med. Virol. 2000, 61, 94–99. [Google Scholar] [CrossRef]
  303. Salzberg, S. The contents of the syringe. Nature 2008, 454, 160–161. [Google Scholar]
  304. Jin, H.; Zhou, H.; Liu, H.; Chan, W.; Adhikary, L.; Mahmood, K.; Lee, M.S.; Kemble, G. Two residues in the hemagglutinin of A/Fujian/411/02-like influenza viruses are responsible for antigenic drift from A/Panama/2007/99. Virology 2005, 336, 113–119. [Google Scholar] [CrossRef]
  305. WHO. Evolution of H5N1 avian influenza viruses in Asia. Emerg. Infect. Dis. 2005, 11, 1515–1521.
  306. Pada, S.; Tambyah, P.A. Overview/reflections on the 2009 H1N1 pandemic. Microb. Infect. 2011, 13, 470–478. [Google Scholar] [CrossRef]
  307. Rockman, S.; Brown, L. Pre-pandemic and pandemic influenza vaccines. Hum. Vaccine. 2010, 6, 792–801. [Google Scholar] [CrossRef]
  308. Belshe, R.B.; Nichol, K.L.; Black, S.B.; Shinefield, H.; Cordova, J.; Walker, R.; Hessel, C.; Cho, I.; Mendelman, P.M. Safety, efficacy, and effectiveness of live, attenuated, cold-adapted influenza vaccine in an indicated population aged 5-49 years. Clin. Infect. Dis. 2004, 39, 920–927. [Google Scholar] [CrossRef]
  309. Mossad, S.B. Demystifying FluMist, a new intranasal, live influenza vaccine. Cleve. Clin. J. Med. 2003, 70, 801–806. [Google Scholar] [CrossRef]
  310. Beyer, W.E.; Palache, A.M.; de Jong, J.C.; Osterhaus, A.D. Cold-adapted live influenza vaccine versus inactivated vaccine: Systemic vaccine reactions, local and systemic antibody response, and vaccine efficacy. A meta-analysis. Vaccine 2002, 20, 1340–1353. [Google Scholar] [CrossRef]
  311. Rudenko, L.; van den Bosch, H.; Kiseleva, I.; Mironov, A.; Naikhin, A.; Larionova, N.; Bushmenkov, D. Live attenuated pandemic influenza vaccine: Clinical studies on A/17/California/2009/38 (H1N1) and licensing of the Russian-developed technology to WHO for pandemic influenza preparedness in developing countries. Vaccine 2011, 29, A40–A44. [Google Scholar]
  312. Hoft, D.F.; Babusis, E.; Worku, S.; Spencer, C.T.; Lottenbach, K.; Truscott, S.M.; Abate, G.; Sakala, I.G.; Edwards, K.M.; Creech, C.B.; et al. Live and inactivated influenza vaccines induce similar humoral responses, but only live vaccines induce diverse T-cell responses in young children. J. Infect. Dis. 2011, 204, 845–853. [Google Scholar] [CrossRef]
  313. He, Q.; Martinez-Sobrido, L.; Eko, F.O.; Palese, P.; Garcia-Sastre, A.; Lyn, D.; Okenu, D.; Bandea, C.; Ananaba, G.A.; Black, C.M.; et al. Live-attenuated influenza viruses as delivery vectors for Chlamydia vaccines. Immunology 2007, 122, 28–37. [Google Scholar] [CrossRef]
  314. Pica, N.; Langlois, R.A.; Krammer, F.; Margine, I.; Palese, P. NS1-truncated live attenuated virus vaccine provides robust protection to aged mice from viral challenge. J. Virol. 2012. [Google Scholar]
  315. Bodewes, R.; Kreijtz, J.H.; Rimmelzwaan, G.F. Yearly influenza vaccinations: A double-edged sword? Lancet Infect. Dis. 2009, 9, 784–788. [Google Scholar] [CrossRef]
  316. Bodewes, R.; Fraaij, P.L.; Geelhoed-Mieras, M.M.; van Baalen, C.A.; Tiddens, H.A.; van Rossum, A.M.; van der Klis, F.R.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Annual vaccination against influenza virus hampers development of virus-specific CD8(+) T cell immunity in children. J. Virol. 2011, 85, 11995–12000. [Google Scholar]
  317. Bodewes, R.; Kreijtz, J.H.; Baas, C.; Geelhoed-Mieras, M.M.; de Mutsert, G.; van Amerongen, G.; van den Brand, J.M.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Vaccination against human influenza A/H3N2 virus prevents the induction of heterosubtypic immunity against lethal infection with avian influenza A/H5N1 virus. PLoS One 2009, 4, e5538. [Google Scholar]
  318. Bodewes, R.; Kreijtz, J.H.; Geelhoed-Mieras, M.M.; van Amerongen, G.; Verburgh, R.J.; van Trierum, S.E.; Kuiken, T.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Vaccination against seasonal influenza A/H3N2 virus reduces the induction of heterosubtypic immunity against influenza A/H5N1 virus infection in ferrets. J. Virol. 2011, 85, 2695–2702. [Google Scholar]
  319. Bodewes, R.; Kreijtz, J.H.; Hillaire, M.L.; Geelhoed-Mieras, M.M.; Fouchier, R.A.; Osterhaus, A.D.; Rimmelzwaan, G.F. Vaccination with whole inactivated virus vaccine affects the induction of heterosubtypic immunity against influenza virus A/H5N1 and immunodominance of virus-specific CD8+ T-cell responses in mice. J. Gen. Virol. 2010, 91, 1743–1753. [Google Scholar]
  320. Heiny, A.T.; Miotto, O.; Srinivasan, K.N.; Khan, A.M.; Zhang, G.L.; Brusic, V.; Tan, T.W.; August, J.T. Evolutionarily conserved protein sequences of influenza a viruses, avian and human, as vaccine targets. PLoS One 2007, 2, e1190. [Google Scholar]
  321. Rimmelzwaan, G.F.; Fouchier, R.A.; Osterhaus, A.D. Influenza virus-specific cytotoxic T lymphocytes: A correlate of protection and a basis for vaccine development. Curr. Opin. Biotechnol. 2007, 18, 529–536. [Google Scholar]
  322. Daemen, T.; de Mare, A.; Bungener, L.; de Jonge, J.; Huckriede, A.; Wilschut, J. Virosomes for antigen and DNA delivery. Adv. Drug Deliv. Rev. 2005, 57, 451–463. [Google Scholar] [CrossRef]
  323. Berthoud, T.K.; Hamill, M.; Lillie, P.J.; Hwenda, L.; Collins, K.A.; Ewer, K.J.; Milicic, A.; Poyntz, H.C.; Lambe, T.; Fletcher, H.A.; et al. Potent CD8+ T-cell immunogenicity in humans of a novel heterosubtypic influenza A vaccine, MVA-NP+M1. Clin. Infect. Dis. 2011, 52, 1–7. [Google Scholar]
  324. Liniger, M.; Zuniga, A.; Naim, H.Y. Use of viral vectors for the development of vaccines. Expet Rev. Vaccine 2007, 6, 255–266. [Google Scholar] [CrossRef]
  325. Ulmer, J.B. Influenza DNA vaccines. Vaccine 2002, 20, S74–S76. [Google Scholar] [CrossRef]
  326. Rimmelzwaan, G.F.; Nieuwkoop, N.; Brandenburg, A.; Sutter, G.; Beyer, W.E.; Maher, D.; Bates, J.; Osterhaus, A.D. A randomized, double blind study in young healthy adults comparing cell mediated and humoral immune responses induced by influenza ISCOM vaccines and conventional vaccines. Vaccine 2000, 19, 1180–1187. [Google Scholar] [CrossRef]
  327. Sambhara, S.; Woods, S.; Arpino, R.; Kurichh, A.; Tamane, A.; Underdown, B.; Klein, M.; Lovgren Bengtsson, K.; Morein, B.; Burt, D. Heterotypic protection against influenza by immunostimulating complexes is associated with the induction of cross-reactive cytotoxic T lymphocytes. J. Infect. Dis. 1998, 177, 1266–1274. [Google Scholar]
  328. Lipatov, A.S.; Gitelman, A.K.; Smirnov Yu, A. Prevention and treatment of lethal influenza A virus bronchopneumonia in mice by monoclonal antibody against haemagglutinin stem region. Acta Virol. 1997, 41, 337–340. [Google Scholar]
  329. Fan, J.; Liang, X.; Horton, M.S.; Perry, H.C.; Citron, M.P.; Heidecker, G.J.; Fu, T.M.; Joyce, J.; Przysiecki, C.T.; Keller, P.M.; et al. Preclinical study of influenza virus A M2 peptide conjugate vaccines in mice, ferrets, and rhesus monkeys. Vaccine 2004, 22, 2993–3003. [Google Scholar]
  330. Song, J.M.; Wang, B.Z.; Park, K.M.; Van Rooijen, N.; Quan, F.S.; Kim, M.C.; Jin, H.T.; Pekosz, A.; Compans, R.W.; Kang, S.M. Influenza virus-like particles containing M2 induce broadly cross protective immunity. PLoS One 2011, 6, e14538. [Google Scholar]

Share and Cite

MDPI and ACS Style

Van de Sandt, C.E.; Kreijtz, J.H.C.M.; Rimmelzwaan, G.F. Evasion of Influenza A Viruses from Innate and Adaptive Immune Responses. Viruses 2012, 4, 1438-1476. https://doi.org/10.3390/v4091438

AMA Style

Van de Sandt CE, Kreijtz JHCM, Rimmelzwaan GF. Evasion of Influenza A Viruses from Innate and Adaptive Immune Responses. Viruses. 2012; 4(9):1438-1476. https://doi.org/10.3390/v4091438

Chicago/Turabian Style

Van de Sandt, Carolien E., Joost H. C. M. Kreijtz, and Guus F. Rimmelzwaan. 2012. "Evasion of Influenza A Viruses from Innate and Adaptive Immune Responses" Viruses 4, no. 9: 1438-1476. https://doi.org/10.3390/v4091438

Article Metrics

Back to TopTop