Next Article in Journal
Epidemiological Aspects of Equid Herpesvirus-Associated Myeloencephalopathy (EHM) Outbreaks
Next Article in Special Issue
The Enigma of Lymphocyte Apoptosis in the Response to Influenza Virus Infection
Previous Article in Journal
Impact of Hepatitis B Virus Infection, Non-alcoholic Fatty Liver Disease, and Hepatitis C Virus Co-infection on Liver-Related Death among People Tested for Hepatitis B Virus in British Columbia: Results from a Large Longitudinal Population-Based Cohort Study
Previous Article in Special Issue
Cellular PSMB4 Protein Suppresses Influenza A Virus Replication through Targeting NS1 Protein
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Understanding the Role of HLA Class I Molecules in the Immune Response to Influenza Infection and Rational Design of a Peptide-Based Vaccine

by
A. K. M. Muraduzzaman
,
Patricia T. Illing
,
Nicole A. Mifsud
* and
Anthony W. Purcell
*
Infection and Immunity Program, Department of Biochemistry and Molecular Biology, Biomedicine Discovery Institute, Monash University, Clayton, VIC 3800, Australia
*
Authors to whom correspondence should be addressed.
Viruses 2022, 14(11), 2578; https://doi.org/10.3390/v14112578
Submission received: 8 October 2022 / Revised: 4 November 2022 / Accepted: 18 November 2022 / Published: 21 November 2022
(This article belongs to the Special Issue Influenza A Viruses: New Insights in 2022)

Abstract

:
Influenza A virus is a respiratory pathogen that is responsible for regular epidemics and occasional pandemics that result in substantial damage to life and the economy. The yearly reformulation of trivalent or quadrivalent flu vaccines encompassing surface glycoproteins derived from the current circulating strains of the virus does not provide sufficient cross-protection against mismatched strains. Unlike the current vaccines that elicit a predominant humoral response, vaccines that induce CD8+ T cells have demonstrated a capacity to provide cross-protection against different influenza strains, including novel influenza viruses. Immunopeptidomics, the mass spectrometric identification of human-leukocyte-antigen (HLA)-bound peptides isolated from infected cells, has recently provided key insights into viral peptides that can serve as potential T cell epitopes. The critical elements required for a strong and long-living CD8+ T cell response are related to both HLA restriction and the immunogenicity of the viral peptide. This review examines the importance of HLA and the viral immunopeptidome for the design of a universal influenza T-cell-based vaccine.
Keywords:
influenza; HLA; vaccine

1. Introduction

Whilst Hippocrates described influenza disease in 412 BC (reviewed in [1]), the first influenza A virus (IAV) pandemic, known as the Russian Flu, was not reported until 1889 [2]. This was followed by the isolation of the virus half a century later in 1933 [3]. Since then, IAV has affected a significant percentage of the world’s population through regular seasonal epidemics and occasional pandemics, causing millions of deaths [4]. The best way to limit influenza’s medical, social, and economic burdens is through annual vaccination, as the World Health Organization (WHO) recommended in their Global Action Plan [5]. Unfortunately, the current antibody-inducing trivalent or quadrivalent flu vaccines have several limitations, including their inadequate boosting of cellular immunity via cytotoxic T cell (CTL) responses [6,7,8,9,10]. Due to antigenic drift, which generates genetic changes or mutations in cell surface glycoproteins, hemagglutinin (HA), and neuraminidase (NA), the current influenza vaccines sometimes fail to produce neutralizing antibodies against seasonal or emerging influenza virus strains [11,12,13]. The United States Centers for Disease Control and Prevention (US-CDC) vaccine effectiveness study showed that the efficacy of the current strain-based vaccines can vary from as low as 10% in an unmatched season to 60% in the case of a very well-matched vaccine [14]. The currently available trivalent (composed of surface glycoproteins from two IAVs and one influenza B virus (IBV)) and quadrivalent (consisting of surface glycoproteins from two IAVs and two IBVs) vaccines require annual reformulation and administration to circumvent the effects of antigenic drift. In addition to humans, IAV can infect a wide range of animals, including pigs, horses, marine mammals, cats, dogs, and birds. Due to this wide host range, antigenically novel strains or IAV subtypes can emerge [15], triggering global pandemics, as observed in the last century (1918 Spanish flu, 1957 Asian flu, 1968 Hong Kong flu) and, more recently, the 2009 H1N1 pandemic (pdmH1N1) [2,16].
There is considerable evidence to suggest that, in both humans [17,18,19] and mice [20,21,22], CD8+ T cells are important in providing protection against influenza through viral clearance, thereby limiting the disease severity. Therefore, T-cell-inducing vaccines, particularly CD8+ CTLs, targeting the highly conserved internal proteins, may be an effective alternative for the provision of universal protection against influenza virus subtypes [23,24]. To stimulate the T cells, viral antigens are processed by the infected host cells and presented as short peptides by HLA molecules encoded within the human major histocompatibility complex (MHC) [25]. The human MHC is the most polymorphic gene system [26], and the encoded HLA allotypes display a diverse peptide array to the T cells for immune recognition (termed the immunopeptidome). Importantly, HLA polymorphisms have been shown to influence peptide binding specificity, T cell receptor (TCR) interactions with the peptide/HLA complex, and the nature of the ensuing anti-viral immune response. Studies shown that some HLA alleles, such as HLA-B*27, that present a conserved epitope from the p24 HIV capsid protein are protective, resulting in a significantly improved survival by maintaining high CD4 counts and a low-level viral load without antiretroviral therapy [27]. Conversely, some other HLA alleles, such as HLA-B*35, are associated with rapid progression and a high viral load [28]. HLA associations have also been well reported for chronic infectious diseases, including hepatitis B and hepatitis C [29,30,31]. However, associations with the influenza virus pathogenesis are still emerging. This review focuses on the identification of the contributions of classical HLA class I (HLA-I) allotypes to the immunoregulation of influenza’s pathogenesis and implications for the development of a universal T cell-inducing vaccine.

2. Influenza: An Unpredictable Virus with a High Pandemic Potential

Influenza viruses belong to the Orthomyxoviridae family and are classified into four subtypes: IAV, IBV, influenza C (ICV), and influenza D (IDV) [32,33,34]. All influenza viruses are enveloped and contain a negative-sense, single-stranded segmented RNA genome [32]. Although IAV and IBV virions contain eight viral RNA (vRNA) segments, ICV [35] and IDV [34] contain seven vRNA segments. A schematic representation of IAV/IBV is shown in Figure 1A. The influenza virion has three subviral components. The first is an envelope lipid bilayer incorporating three virally encoded transmembrane proteins, HA, NA, and the proton channel protein, matrix-2 (M2). Secondly, a matrix layer underneath the envelope, composed of M1, interacts with the cytoplasmic tails of HA, NA, and M2 on the outer side and vRNA and nucleoprotein (NP) on the inner side of this layer [36,37]. Thirdly, the viral core consists of helical ribonucleoprotein complex (RNP) segments containing the negative-stranded genomic vRNAs, NP, the nuclear export protein (NEP) and a 3-polymerase (3P) complex. The 3P complex is heterotrimeric, being composed of the polymerase basic protein 1 (PB1), polymerase basic protein 2 (PB2), and polymerase acidic protein (PA) [36,37,38,39,40].
Among the four subtypes of the influenza virus, IAV is the most important public health consideration in regard to seasonal epidemics and potential pandemic threats. In addition to humans, IAV can infect a large number of animal species, including ducks, chickens, pigs, whales, horses, and seals [41,42]. Different combinations of IAV HA (1–18) and NA (1–11) proteins can potentially give rise to 198 subtypes, although only 131 subtypes have been naturally isolated [41]. Different IAV subtypes, such as H5N1, H7N9, and H9N2, circulating in aquatic birds are called avian influenza viruses. These avian strains sometimes switch between hosts to form new lineages and may emerge as antigenically novel human IAVs that can promote pandemics [43]. Although human infections with avian influenza viruses are rare, this can occur following exposure to infected poultry, either via direct contact with chickens or by visiting a live poultry market [44,45,46,47]. In the last century, there have been four instances where influenza viruses with genes originating from avian or swine reservoirs emerged, either with or without reassortment with human influenza viruses, resulting in a pandemic wave [48]. As the IAV reservoirs cannot be eradicated from wild aquatic birds, another influenza pandemic is inevitable, and based on past experience, it is widely believed that the currently circulating avian influenza virus H5N1 or H7N9 may be the culprit subtype.
Influenza viruses are constantly evolving via two different mechanisms known as antigenic drift and antigenic shift (Figure 1B). Firstly, antigenic drift is the result of minor changes or mutations due to error-prone RNA replication, which occurs in both IAV and IBV as the virus replicates. Secondly, novel influenza viruses with pandemic potential can arise through the antigenic shift resulting from abrupt changes in, or the reassortment of, gene segments between avian or swine influenza viruses, which cross the species barrier and infect humans [49]. Due to these rapid antigenic drifts or shifts, vaccines against surface glycoproteins may be ineffective against seasonal or novel influenza viruses [9,50,51,52,53,54]. In contrast, influenza virus internal proteins (i.e., NP, PA, PB, NS and M) are highly conserved across the different subtypes [49]. Importantly, studies have shown that CD8+ T cells can recognize antigenic peptides derived from conserved internal proteins, including NP, M1, and PB1 [55,56,57,58]. A recent study has shown that memory CD8+ T cells specific to a conserved PB1 (A2-PB1413–422) peptide cross-reacted with homologous sequences from IAV, IBV, and ICV, offering a degree of cross-strain immunity [59]. It is recognized that the epitopes derived from conserved internal proteins are likely to elicit broader immunity against a spectrum of influenza strains so as to promote lifelong protection [60]. Thus, T cells against these conserved proteins have the potential to provide heterosubtypic protection against influenza disease and are of substantial interest for the design of next-generation universal vaccines.

3. Host Immunity to Influenza

When an individual is infected with influenza virus, both the innate and adaptive arms of the host immune system are engaged to combat and clear the pathogenic assault. The innate immune system is the first defense line that recognizes the viral components called pathogen-associated molecular patterns (PAMPs) by the host pathogen recognition receptors (PRRs). For example, in influenza virus infection, the retinoic-acid-inducible gene-I protein (RIG-I) recognizes the viral RNA in the cytosol and induces the production of pro-inflammatory cytokines and type I IFN. Toll-like receptor (TLR) 3 and TLR7 recognize double-stranded and single-stranded RNA (dsRNA and ssRNA) and many more PRRs which target different proteins of the influenza virus, also leading to the activation of innate immune signaling and the induction of cytokine and antiviral molecule expression [61,62,63]. However, the professional antigen-presenting cells, e.g., dendritic cells (DCs), link the innate and adaptive immune responses during the IAV infection [64]. After infection with IAV, the conventional DCs (cDCs) migrate from the lungs to the lymph nodes and present antigens derived from IAV with major histocompatibility complex (MHC) class I molecules for the purpose of recognition by virus-specific CD8+ CTLs [65,66]. In addition, viral proteins degraded in endosomes/lysosomes are associated with the MHC class II molecule and are presented on the cell membrane for their recognition by CD4+ T helper (Th) cells that leads to B cell proliferation and maturation into antibody-producing plasma cells [67,68]. The humoral immune response is mediated by B cells that produce neutralizing antibodies targeting viral surface proteins, such as HA and NA. These antibodies can also induce antibody-dependent, cell-mediated cytotoxicity or complement activation by directly binding to the surface of infected cells [52,69]. The cellular immune response is primarily mediated by CD8+ T cells, which play a crucial role in the surveillance, detection, and elimination of virus-infected cells (Figure 2) [70,71,72]. Indeed, T cells have been shown to play an essential role in the clearance of many respiratory viruses [73,74,75,76].
Animal studies have demonstrated the protective role of CD8+ T cells against IAV infection. When influenza virus-specific CD8+ T cells were injected into naïve mice, they showed an increased survival and decreased weight loss after being challenged with a lethal dose of virus, confirming the role of CTL in viral clearance and protection [77,78]. In contrast, T cell-deficient mice (athymic mice bred on a BALB/c background) or transgenic mice lacking MHC-I showed delayed viral clearance and increased mortality [71,79]. Most importantly, heterosubtypic immunity, wherein host immune responses to highly conserved peptides across all the influenza virus subtypes take place, confers protection against infection and disease caused by different subtypes. Although the evidence regarding humans is limited, T cell-mediated heterosubtypic immunity against influenza viruses has been confirmed in murine studies. When conserved NP-peptide-specific T cell clones (DbNP366 epitope) or H1N1-virus-specific T cell clones were transferred into naïve mice and then challenged with different IAV subtypes, these mice showed the early recruitment of T cells in the lung tissues, leading to early viral clearance and increased survival [70,80,81,82]. CTL heterosubtypic immunity has also been confirmed in guinea pigs, in which priming with the 2009 pandemic H1N1 virus conferred protection against the 2013 avian H7N9 IAV [83].
In humans, the control of influenza infection by the CD8+ T cells was first described by McMichael et al., who demonstrated that the IAV titers and clinical symptom severity were inversely correlated with the magnitude of the pre-existing anti-influenza CD8+ T cells [84]. Furthermore, studies found that amino acid changes in the major viral NP CTL epitopes (HLA-B*27:05-restricted epitope NP383–391 (SRYWAIRTR) and the HLA-B*08-restricted epitope NP380–388 (ELRSRYWAI)) [85] can rapidly accumulate in vivo, suggesting that CD8+ CTLs exert selection pressure on the virus [85,86,87,88,89].
The first evidence of heterosubtypic protection against a pandemic strain was described by Epstein in the Cleveland family study during the 1957 H2N2 pandemic [90]. The results showed that adults who had laboratory-confirmed H1N1 seasonal influenza virus infections between 1950 and 1957 were three times less likely to have symptomatic laboratory-confirmed pandemic H2N2 influenza virus compared to those who were not previously infected with the seasonal H1N1 influenza virus [90]. Recently, during the 2009 H1N1 influenza virus pandemic, individuals with pre-existing IAV-specific CD8+ T cells showed a decreased risk of fever, fewer influenza-like symptoms of illness, reduced illness severity scores, and the absence of viral shedding [17]. In addition, the flu watch cohort study also observed a positive correlation between pre-existing IAV-specific CD8+ T cells and less symptomatic IAV infections during three seasonal epidemics and the 2009 H1N1 pandemic, which was independent of baseline antibodies, confirming the heterosubtypic protection delivered by CD8+ T cells in influenza virus infection [19]. Many other studies have also described the potential cross-reactive response of the CD8+ T cells to different influenza virus strains, including the highly pathogenic H7N9 and H5N1 avian influenza viruses [18,56,82,91,92,93,94,95]. Indeed, H7N9-infected patients with early H7N9-specific, IFN-γ-producing CD8+ T cell populations recovered more rapidly from severe infection. In contrast, patients with delayed or absent CD8+ T cell responses had increased morbidity and mortality [18]. These studies, collectively, confirm the role of the CD8+ T cells in influenza recovery.

4. Immunodominance and MHC Genotype in the Antiviral Response

Immunodominance is a central feature of immune responsiveness to viruses. It refers to the phenomenon whereby the antiviral CTL population tends to be dominated by T cells restricted to a small number of antigenic MHC-peptide complexes (i.e., presenting immunodominant epitopes) among the available MHC peptides [96]. Indeed, immunodominant epitopes have been shown to play a critical role in eliminating virus-infected cells and contributing to the development of a memory T cell pool so as to combat subsequent encounters [96,97,98,99]. The mechanism of immunodominance is complex, involving features associated with (i) epitope abundance, (ii) the proteasomal processing of antigens, with immunoproteasomes playing an important role in the establishment of the immunodominance hierarchy of CD8+ T cells [100], (iii) IFN-γ production [101], (iv) TCR repertoire and affinity [102], and (v) the structural features of the antigen (disulfide bonds, protease-processing sites), all of which contribute to the overall CTL response [103,104,105,106,107,108,109]. A number of studies demonstrated that the co-expression of different MHC (or HLA) allotypes influences both the magnitude and specificity of the CTL response to influenza infection. Peripheral blood mononuclear cells (PBMCs) isolated from an individual co-expressing HLA-A2 and HLA-B37 were stimulated with either infectious virus, influenza A (A/PR8 (H1N1)), or synthetic peptides, HLA-A2-restricted M57–68 and HLA-B37-restricted NP335–349, and only an HLA-B37-restricted CTL response was observed [110]. Although the immunodominance of the HLA-A2–restricted response to the conserved IAV M57-68 peptide has been well-established [111,112], the immunodominance of NP335–349 over the HLA-A2-restricted M57–68 was observed in individuals co-expressing HLA-A2 and B37. In two separate studies, Boon et al. observed that the magnitude of influenza-A-virus-specific CTL responses depended on the HLA-A and -B alleles expressed [112,113]. Here, the frequency of HLA-B8-restricted NP380–388-specific CTLs were 3-fold lower in HLA-B27+ carriers compared to HLA-B27- individuals (n = 5). Additionally, HLA-A1-restricted NP44–52-specific CTLs were 3-fold higher in individuals (n = 4) that co-expressed HLA-A1, HLA-A2, HLA-B8, and HLA-B35, which indicated that the overall CTL response was influenced by an individual’s complete HLA genotype [112]. In their second study, it was demonstrated that both HLA-B*08:01 and HLA-A*01:01 influenza-A-virus-specific CTL responses were much higher in HLA-B*27:05+ and HLA-B*35:01+ individuals [113]. Indeed, in the presence of HLA-B*27:05, the magnitude of the HLA-B*08:01-restricted response was even lower [113]. The effects of different HLA genotypes on immunodominant peptide presentation was examined by Wu et al. [57]. After comparing the immunogenicity towards the influenza M58–66 epitope in healthy HLA-A*02:01+ donors, they reported that the response was only immunodominant in three out of eight individuals. The remaining five HLA-A2+ carriers demonstrated a prominent CD8+ T cell response to the NP380–388 epitope that was restricted to HLA-B*08:01+ [57]. Indeed, HLA-B allotypes have been shown to significantly (p = 0.012) generate a more robust polyfunctional (via the production of IFN-γ and IL-2) CD8+ T cell response to different viral antigens, encompassing HIV-1, CMV, EBV, and the influenza virus, compared to HLA-A allotypes [114].
Additionally, Quiñones-Parra et al. demonstrated that HLA-A*03:01 NP265–273, HLA-B*08:01 NP225–233, HLA-B*18:01 NP219–226, HLA-B*27:05 NP383–391, and HLA-B*57:01 NP199–207 epitopes were capable of eliciting robust CTL responses to any IAV strain in individuals carrying these HLA allotypes. Contrary to this, limited CTL responses were observed in the ethnicities in which HLA-A*01:01, HLA-A*24:02, HLA-A*68:01, and HLA-B*15:01 are prominent [115]. Moreover, HLA-B*37:01 showed promising results in presenting different NP peptide variants capable of producing robust CTL reactivity to protect against H1N1, pH1N1, H3N2, H5N1, H7N9, and possibly other novel IAV strains [116]. Furthermore, HLA-A*02:01 PB1413–421 can exhibit universal cross-reactivity across IAV, IBV, and ICV, and HLA-A*01:01 PB1591–599 and HLA-B*37:01 NP338–345 peptides could exhibit cross-reactive CD8+ T cell responses in IAV and IBV infections [59].
A recent study reported that serologically defined HLA-A2 homozygous lymphocytes, in contrast to heterozygous lymphocytes, did not synthesize detectable influenza virus RNA or protein after exposure to the virus. Meanwhile, the HLA-A2 homozygous lymphocytes could internalize the infectious viruses, and they were not susceptible to lysis by autologous virus-specific CTLs. Similar findings were also observed in the case of HLA-A1, -A11, and -B44 homozygous lymphocytes, suggesting that individuals carrying homozygous HLA class I allotypes may possess lymphocytes that are not susceptible to influenza virus infection and CTL lysis, which is a unique HLA-associated lymphocyte defense mechanism against influenza virus [117]. Therefore, the combination of HLA allotypes expressed by an individual may impact peptide presentation, the modulation of surface expression, and the competition for overlapping peptides (this can lead to a decrease in the ability to generate a CTL response). Moreover, HLA carriage can also shape the activation and proliferation of T cells during infection [112,118,119,120,121,122,123,124].

5. HLA Association with Influenza Disease Susceptibility: Population-Based Studies

Population studies examining the HLA-associated influenza pathogenesis provide critical information regarding disease protection or susceptibility. Global indigenous population studies have reported higher incidences of morbidity and mortality during both seasonal and pandemic influenza virus infections than in non-indigenous populations [125,126,127,128,129,130,131,132,133,134,135,136]. The lesser degree of HLA diversity exhibited by indigenous communities is likely to have arisen from an evolutionary bottleneck that established a small ancestral pool with little HLA diversity. This evolutionary effect is amplified by limited outbreeding to other populations, which may explain the low prevalence of protective HLA (e.g., only ~15% HLA-A2 among the Indigenous peoples of Australia) against influenza virus infection [135]. To analyze the preferential antigen binding and capability of the T cell responses of different HLA alleles, Hertz et al. introduced a computational prediction tool called the “targeting efficiency”. This tool analyzes the Spearman correlation coefficient of a given HLA molecule binding to the conserved regions of a protein [137]. When an HLA preferentially binds the conserved regions, it provides a positive score, and a negative score indicates the preferential binding of variable regions. To determine whether the targeting efficiency is correlated with the CD8+ T cell response in influenza virus infection, thirteen pH1N1-positive individuals were recruited in the 2009–2010 influenza season and ranked based on their average HLA-A and HLA-B allele targeting efficiency scores over the entire pH1N1 proteome. The enzyme-linked immunosorbent spot (ELISpot) was used to detect IFN-γ secretion patterns from PBMCs and showed that the HLA targeting efficiency scores were correlated with IFN-γ spot-forming units (SFUs) in regard to pH1N1 (r = 0.54, p = 0.054) [134]. To assess the broader implications of the targeting efficiency on the population level, 95 of the most prevalent HLA class I alleles against a wide variety of influenza virus subtypes, including 2009 pH1N1, previously circulating H1N1 and H3N2 subtypes, and H5N1 and 1918 H1N1 subtypes, were computed. The study identified three HLA alleles (HLA-A*24, HLA-A*68, HLA-A*39) which had a low targeting efficiency score and, hence, were positively correlated with pH1N1 mortality. The high prevalence of HLA-A*24 in the Indigenous peoples of Australia (where HLA-A*02:01 (17%), HLA-A*11:01 (16%), HLA-A*24:02 (22%), and HLA-A*34:01 (24%) account for 79% of the HLA-A allotypes) may underpin the vulnerability of this population in an influenza virus pandemic or even cases of seasonal influenza [135]. Another study aiming to examine the susceptibility to avian influenza virus (H7N9) among different ethnicities identified six conserved epitopes from the matrix and NP proteins of H7N9 influenza virus, which are restricted to HLA-A*03:01(NP265+), HLA-A*02:01(M158+), HLA-B*27:05 (NP383+), HLA-B*57:01(NP199+), HLA-B*18:01(NP219+), and HLA-B*08:01(NP225+), that displayed robust CD8+ T cell responses. Unfortunately, the identified HLA alleles that produced robust CD8+ T cell responses were found at a very low frequency in the Alaskan and Australian Indigenous populations (16%) compared to higher frequency in other ethnicities (African, 37%; European descent, 57%; Oriental, 37%; Amerindian, 36%). Thus, the influenza disease severity could therefore show an ethnic bias in the face of an H7N9 pandemic [115].
Different HLAs may also be more efficient in presenting specific influenza virus proteins. Most studies have shown that the NP, M, and PA are the sources of most T cell epitopes in both humans and mice [58,111,138,139,140]. This result may have some bias, as these influenza proteins were mostly studied in the context of the most prevalent HLA alleles, such as HLA-A2. Even then, when Hensen L et al. identified the immunopeptides of IAV by LC-MS/MS restricted to HLA-A*24, they discovered that the majority (62%) of the peptides are from PB1 and PB2, with none derived from the NA or M proteins [141]. These findings highlight the importance of the including PB1- and PB2-derived peptides in a universal influenza vaccine design so as to cover vulnerable global Indigenous populations [141].
In a case–control study in the Mexican-Mestizo population, a lower frequency of HLA-A*24:02:01 in patients compared to asymptomatic contacts suggested a possible protective effect [142], which conflicts with the low targeting efficiency of HLA-A*24 described by Hertz et al. [134]. The same study also identified the HLA-A*68:01:01 as a susceptible allele, as it was found only in the influenza-positive patient group [142], which, again, corelates with the low HLA targeting efficiency score of HLA-A*68:01 but not the closely related allotype, HLA-A*68:02 [134]. The examination of the HLA-B allotypes in the same Mexican population revealed HLA-B*39:06:02 and HLA-B*51:01:05 as susceptible allotypes, which were observed at a high frequency in individuals who experienced the 2009 influenza A H1N1 [142]. Additionally, a Northeast Indian study reported that, during the 2009 pandemic, the HLA-A*11 and HLA-B*35 allotypes conferred susceptibility to IAV (H1N1) [143]. In contrast to these influenza-susceptible alleles, different studies have shown that HLA-A*01:01, HLA-A*02:01, HLA-A*03:01, HLA-B*08:01, HLA-B*18:01, HLA-B*27:05, HLA-B*37:01, and HLA-B*57:01 present highly conserved influenza viral peptides and were associated with universal protective CD8+ T cell-mediated immunity [59,115,144]. Although not statistically significant, a higher prevalence of HLA-C*03:02 (3.98% in the influenza positive group compared to 1% in the influenza contact group) was observed in the Mexican IAV H1N1/09 patient group. In contrast, HLA-C*03:03 (1.44% patients vs. 5.44% contact), HLA-C*03:04 (4.34% patients vs. 10.66% contact), and HLA-C*07:01 (5.07% patients vs. 9.33% contact) were reported at lower frequencies in the patient group [142]. There is minimal information regarding the influenza viral epitopes presented by HLA-C, with only 30 entries in the Immune Epitope Database (IEDB) [145]. Interestingly, compared to the known GILGFVFTL–HLA-A*02:01 structure (PDB 2VLL), the structure of GILGFVFTL–HLA-C*08:01 (PDB 4NT6) demonstrated near-identical conformations and elicited CTL responses in both HLA-A*02- and HLA-C*08-positive individuals [141]. This finding provided evidence for the degenerate antigenicity between HLA products derived from separate loci and could provide new insights for the future development of universal IAV vaccines.

6. Towards a Universal T Cell Epitope Vaccine

The most recent pandemic of severe Acute Respiratory Syndrome Coronavirus 2 (SARS-CoV-2) and the rapid emergence of novel variants have reminded us how crucial it is to possess an effective vaccine when the treatment options are limited. The ideal and most effective vaccine is likely to employ both cellular and humoral immune responses. Unfortunately, the appearance of the variants of viruses such as SARS-CoV2, IAV, or some persisting viral infections, such as HIV and hepatitis C, that evade vaccines or pre-existing antiviral immunity has often raised the issue of the use of antibody-inducing vaccines alone. The T cells play a critical role in the duration and cross-reactivity of vaccines. Moreover, pre-existing T cell immunity is associated with the decreased severity of infectious diseases.
The most significant advantage of a T cell-inducing influenza vaccine is based on its cross-protective capability against different strains, engendered by conserved epitopes. Substantial progress has been made in the identification of conserved peptides that generate cross-protective CTL responses across all three influenza virus subtypes [59] or individual peptides that can generate a robust immune response in vulnerable indigenous populations [141]. Still, there is a considerable knowledge gap in regard to the formulation of a universal vaccine due to the complexity of HLA diversity and restricted T cell responses.
In the last ten years, more than 20 T cell-based vaccines have been in various stages of clinical development, with the majority of studies focusing on NP and M1 epitopes using different vaccine delivery platforms (Table 1). Notably, most of the vaccines have shown a promising safety and efficacy via the induction of CD8+ T cell responses. However, no candidate vaccine has proven to be effective in HLA-diversified populations. Although some data are available regarding specific HLA (especially HLA-A2), there is limited information on the magnitude, quality, and breadth of influenza virus-specific T cell responses in African, Latin American, and Asian populations. Therefore, we must map the conserved antigenic epitopes across the different HLAs prevalent in diverse populations to design a genuinely universal influenza vaccine. It is likely that there will need to develop different formulations so as to address the HLA diversity across global populations, including the highly vulnerable indigenous people.

Author Contributions

Conceptualization, A.K.M.M. and N.A.M.; writing—original draft preparation, A.K.M.M. and N.A.M.; writing—review and editing, A.K.M.M., P.T.I., N.A.M. and A.W.P. All authors have read and agreed to the published version of the manuscript.

Funding

Funding acquisition, Australian National Health and Medical Research Council (NHMRC) project grant (#1122524) and Principal Research Fellowship (#1137739) both awarded to A.W.P.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

Influenza A virus (IAV), influenza B virus (IBV), influenza C virus (ICV), cytotoxic T cell (CTL), human leukocyte antigen (HLA), hemagglutinin (HA), polymerase basic 1 (PB1), polymerase basic 2 (PB2), nucleoprotein (NP), major histocompatibility complex (MHC), neuraminidase (NA)

References

  1. Mammas, I.N.; Spandidos, D.A. Paediatric Virology in the Hippocratic Corpus. Exp. Ther. Med. 2016, 12, 541–549. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hill, E.M.; Tildesley, M.J.; House, T. Evidence for history-dependence of influenza pandemic emergence. Sci. Rep. 2017, 7, 43623. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Burnet, F.M. Portraits of viruses: Influenza virus A. Intervirology 1979, 11, 201–214. [Google Scholar] [CrossRef] [PubMed]
  4. Hsieh, Y.-C.; Wu, T.-Z.; Liu, D.-P.; Shao, P.-L.; Chang, L.-Y.; Lu, C.-Y.; Lee, C.-Y.; Huang, F.-Y.; Huang, L.-M. Influenza pandemics: Past, present and future. J. Formos. Med. Assoc. 2006, 105, 1–6. [Google Scholar] [CrossRef] [Green Version]
  5. World Health Organization. The Global Action Plan for Influenza Vaccines: Report of the Ninth Meeting of the Advisory Group of the WHO Global Action Plan for Influenza Vaccines; World Health Organization: Geneva, Switzerland, 2014. [Google Scholar]
  6. McMichael, A.; Gotch, F.; Cullen, P.; Askonas, B.; Webster, R. The human cytotoxic T cell response to influenza A vaccination. Clin. Exp. Immunol. 1981, 43, 276. [Google Scholar]
  7. Zambon, M.C. Epidemiology and pathogenesis of influenza. J. Antimicrob. Chemother. 1999, 44, 3–9. [Google Scholar] [CrossRef]
  8. Monto, A. Epidemiology and virology of influenza illness. Am. J. Manag. Care 2000, 6, S255–S264. [Google Scholar]
  9. Treanor, J. Influenza vaccine—Outmaneuvering antigenic shift and drift. N. Engl. J. Med. 2004, 350, 218–220. [Google Scholar] [CrossRef] [Green Version]
  10. 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] [CrossRef] [Green Version]
  11. Chan, M.C.; Wang, M.H.; Chen, Z.; Hui, D.S.; Kwok, A.K.; Yeung, A.C.; Liu, K.M.; Yeoh, Y.K.; Lee, N.; Chan, P.K. Frequent genetic mismatch between vaccine strains and circulating seasonal influenza viruses, Hong Kong, China, 1996–2012. Emerg. Infect. Dis. 2018, 24, 1825. [Google Scholar] [CrossRef]
  12. Skowronski, D.M.; Sabaiduc, S.; Leir, S.; Rose, C.; Zou, M.; Murti, M.; Dickinson, J.A.; Olsha, R.; Gubbay, J.B.; Croxen, M.A. Paradoxical clade-and age-specific vaccine effectiveness during the 2018/19 influenza A (H3N2) epidemic in Canada: Potential imprint-regulated effect of vaccine (I-REV). Eurosurveillance 2019, 24, 1900585. [Google Scholar] [CrossRef] [Green Version]
  13. Flannery, B.; Kondor, R.J.G.; Chung, J.R.; Gaglani, M.; Reis, M.; Zimmerman, R.K.; Nowalk, M.P.; Jackson, M.L.; Jackson, L.A.; Monto, A.S. Spread of antigenically drifted influenza A (H3N2) viruses and vaccine effectiveness in the United States during the 2018–2019 season. J. Infect. Dis. 2020, 221, 8–15. [Google Scholar] [CrossRef] [PubMed]
  14. Centers for Disease Control and Prevention. CDC Seasonal Flu Vaccine Effectiveness Studies; Centers for Disease Control and Prevention: Guangzhou, China, 2019. [Google Scholar]
  15. Taubenberger, J.K.; Kash, J.C. Influenza virus evolution, host adaptation, and pandemic formation. Cell Host Microbe 2010, 7, 440–451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Kawaoka, Y.; Krauss, S.; Webster, R.G. Avian-to-human transmission of the PB1 gene of influenza A viruses in the 1957 and 1968 pandemics. J. Virol. 1989, 63, 4603–4608. [Google Scholar] [CrossRef] [Green Version]
  17. Sridhar, S.; Begom, S.; Bermingham, A.; Hoschler, K.; Adamson, W.; Carman, W.; Bean, T.; Barclay, W.; Deeks, J.J.; Lalvani, A. Cellular immune correlates of protection against symptomatic pandemic influenza. Nat. Med. 2013, 19, 1305–1312. [Google Scholar] [CrossRef]
  18. Wang, Z.; Wan, Y.; Qiu, C.; Quiñones-Parra, S.; Zhu, Z.; Loh, L.; Tian, D.; Ren, Y.; Hu, Y.; Zhang, X.; et al. Recovery from severe H7N9 disease is associated with diverse response mechanisms dominated by CD8+ T cells. Nat. Commun. 2015, 6, 6833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Hayward, A.C.; Wang, L.; Goonetilleke, N.; Fragaszy, E.B.; Bermingham, A.; Copas, A.; Dukes, O.; Millett, E.R.; Nazareth, I.; Nguyen-Van-Tam, J.S.; et al. Natural T Cell-mediated Protection against Seasonal and Pandemic Influenza. Results of the Flu Watch Cohort Study. Am. J. Respir. Crit. Care Med. 2015, 191, 1422–1431. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  20. Valkenburg, S.A.; Gras, S.; Guillonneau, C.; La Gruta, N.L.; Thomas, P.G.; Purcell, A.W.; Rossjohn, J.; Doherty, P.C.; Turner, S.J.; Kedzierska, K. Protective efficacy of cross-reactive CD8+ T cells recognising mutant viral epitopes depends on peptide-MHC-I structural interactions and T cell activation threshold. PLoS Pathog. 2010, 6, e1001039. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Grant, E.J.; Josephs, T.M.; Valkenburg, S.A.; Wooldridge, L.; Hellard, M.; Rossjohn, J.; Bharadwaj, M.; Kedzierska, K.; Gras, S. Lack of Heterologous Cross-reactivity toward HLA-A*02:01 Restricted Viral Epitopes Is Underpinned by Distinct αβT Cell Receptor Signatures. J. Biol. Chem. 2016, 291, 24335–24351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Quiñones-Parra, S.M.; Clemens, E.B.; Wang, Z.; Croom, H.A.; Kedzierski, L.; McVernon, J.; Vijaykrishna, D.; Kedzierska, K. A Role of Influenza Virus Exposure History in Determining Pandemic Susceptibility and CD8+ T Cell Responses. J. Virol. 2016, 90, 6936–6947. [Google Scholar] [CrossRef] [Green Version]
  23. 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] [CrossRef] [PubMed]
  24. Thomas, P.G.; Keating, R.; Hulse-Post, D.J.; Doherty, P.C. Cell-mediated protection in influenza infection. Emerg. Infect. Dis. 2006, 12, 48. [Google Scholar] [CrossRef] [PubMed]
  25. Daniels, M.A.; Jameson, S.C. Critical role for CD8 in T cell receptor binding and activation by peptide/major histocompatibility complex multimers. J. Exp. Med. 2000, 191, 335–346. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Mungall, A.J.; Palmer, S.; Sims, S.; Edwards, C.; Ashurst, J.; Wilming, L.; Jones, M.; Horton, R.; Hunt, S.; Scott, C. The DNA sequence and analysis of human chromosome 6. Nature 2003, 425, 805–811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Bailey, J.R.; Williams, T.M.; Siliciano, R.F.; Blankson, J.N. Maintenance of viral suppression in HIV-1-infected HLA-B*57+ elite suppressors despite CTL escape mutations. J. Exp. Med. 2006, 203, 1357–1369. [Google Scholar] [CrossRef] [Green Version]
  28. Tomiyama, H.; Miwa, K.; Shiga, H.; Moore, Y.I.; Oka, S.; Iwamoto, A.; Kaneko, Y.; Takiguchi, M. Evidence of presentation of multiple HIV-1 cytotoxic T lymphocyte epitopes by HLA-B*3501 molecules that are associated with the accelerated progression of AIDS. J. Immunol. 1997, 158, 5026–5034. [Google Scholar]
  29. Deeks, S.G.; Walker, B.D. Human immunodeficiency virus controllers: Mechanisms of durable virus control in the absence of antiretroviral therapy. Immunity 2007, 27, 406–416. [Google Scholar] [CrossRef] [Green Version]
  30. Fitzmaurice, K.; Petrovic, D.; Ramamurthy, N.; Simmons, R.; Merani, S.; Gaudieri, S.; Sims, S.; Dempsey, E.; Freitas, E.; Lea, S.; et al. Molecular footprints reveal the impact of the protective HLA-A*03 allele in hepatitis C virus infection. Gut 2011, 60, 1563–1571. [Google Scholar] [CrossRef] [Green Version]
  31. Zaunders, J.; van Bockel, D. Innate and Adaptive Immunity in Long-Term Non-Progression in HIV Disease. Front. Immunol. 2013, 4, 95. [Google Scholar] [CrossRef] [Green Version]
  32. Ritchey, M.B.; Palese, P.; Kilbourne, E.D. RNAs of influenza A, B, and C viruses. J. Virol. 1976, 18, 738–744. [Google Scholar] [CrossRef] [Green Version]
  33. Hause, B.M.; Collin, E.A.; Liu, R.; Huang, B.; Sheng, Z.; Lu, W.; Wang, D.; Nelson, E.A.; Li, F. Characterization of a novel influenza virus in cattle and Swine: Proposal for a new genus in the Orthomyxoviridae family. mBio 2014, 5, e00031-14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hause, B.M.; Ducatez, M.; Collin, E.A.; Ran, Z.; Liu, R.; Sheng, Z.; Armien, A.; Kaplan, B.; Chakravarty, S.; Hoppe, A.D.; et al. Isolation of a novel swine influenza virus from Oklahoma in 2011 which is distantly related to human influenza C viruses. PLoS Pathog. 2013, 9, e1003176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lamb, R.A. The influenza virus RNA segments and their encoded proteins. In Genetics of Influenza Viruses; Springer: Berlin/Heidelberg, Germany, 1983; pp. 21–69. [Google Scholar]
  36. Nayak, D.P.; Balogun, R.A.; Yamada, H.; Zhou, Z.H.; Barman, S. Influenza virus morphogenesis and budding. Virus Res. 2009, 143, 147–161. [Google Scholar] [CrossRef]
  37. Rossman, J.S.; Lamb, R.A. Influenza virus assembly and budding. Virology 2011, 411, 229–236. [Google Scholar] [CrossRef] [Green Version]
  38. Hutchinson, E.C.; von Kirchbach, J.C.; Gog, J.R.; Digard, P. Genome packaging in influenza A virus. J. Gen. Virol. 2010, 91, 313–328. [Google Scholar] [CrossRef] [PubMed]
  39. Nayak, D.P.; Hui, E.K.-W.; Barman, S. Assembly and budding of influenza virus. Virus Res. 2004, 106, 147–165. [Google Scholar] [CrossRef] [PubMed]
  40. Noda, T.; Kawaoka, Y. Structure of influenza virus ribonucleoprotein complexes and their packaging into virions. Rev. Med. Virol. 2010, 20, 380–391. [Google Scholar] [CrossRef]
  41. CDC. Transmission of Avian Influenza a Viruses between Animals and People. Available online: https://www.cdc.gov/flu/avianflu/virus-transmission.htm (accessed on 7 January 2020).
  42. Taubenberger, J.K.; Morens, D.M. Influenza: The once and future pandemic. Public Health Rep. 2010, 125 (Suppl. S3), 16–26. [Google Scholar] [CrossRef] [Green Version]
  43. Nasreen, S.; Khan, S.U.; Luby, S.P.; Gurley, E.S.; Abedin, J.; Zaman, R.U.; Sohel, B.M.; Rahman, M.; Hancock, K.; Levine, M.Z.; et al. Highly pathogenic Avian Influenza A(H5N1) virus infection among workers at live bird markets, Bangladesh, 2009–2010. Emerg. Infect. Dis. 2015, 21, 629–637. [Google Scholar] [CrossRef]
  44. Plague, F.; Aviaire, G. Avian influenza. Virol. J. 2006, 5, 54. [Google Scholar]
  45. Wong, S.S.; Yuen, K.-y. Avian influenza virus infections in humans. Chest 2006, 129, 156–168. [Google Scholar] [CrossRef] [PubMed]
  46. Harder, T.; Buda, S.; Hengel, H.; Beer, M.; Mettenleiter, T. Poultry food products—A source of avian influenza virus transmission to humans? Clin. Microbiol. Infect. 2016, 22, 141–146. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. World Health Organization. Influenza (Avian and Other Zoonotic). Available online: https://www.who.int/news-room/fact-sheets/detail/influenza-(avian-and-other-zoonotic) (accessed on 4 March 2021).
  48. Kilbourne, E.D. Influenza Pandemics of the 20th Century. Emerg. Infect. Dis. 2006, 12, 9–14. [Google Scholar] [CrossRef] [PubMed]
  49. Webster, R.; Laver, W.; Air, G.; Schild, G. Molecular mechanisms of variation in influenza viruses. Nature 1982, 296, 115–121. [Google Scholar] [CrossRef]
  50. Carrat, F.; Flahault, A. Influenza vaccine: The challenge of antigenic drift. Vaccine 2007, 25, 6852–6862. [Google Scholar] [CrossRef]
  51. Chen, J.; Deng, Y.-M. Influenza virus antigenic variation, host antibody production and new approach to control epidemics. Virol. J. 2009, 6, 30. [Google Scholar] [CrossRef] [Green Version]
  52. Couch, R. An overview of serum antibody responses to influenza virus antigens. Dev. Biol. 2003, 115, 25–30. [Google Scholar]
  53. Gross, P.A.; Hermogenes, A.W.; Sacks, H.S.; Lau, J.; Levandowski, R.A. The efficacy of influenza vaccine in elderly persons: A meta-analysis and review of the literature. Ann. Intern. Med. 1995, 123, 518–527. [Google Scholar] [CrossRef]
  54. Nypaver, C.; Dehlinger, C.; Carter, C. Influenza and Influenza Vaccine: A Review. J. Midwifery Women’s Health 2021, 66, 45–53. [Google Scholar] [CrossRef]
  55. 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] [CrossRef] [Green Version]
  56. 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] [CrossRef] [PubMed] [Green Version]
  57. Wu, C.; Zanker, D.; Valkenburg, S.; Tan, B.; Kedzierska, K.; Zou, Q.M.; Doherty, P.C.; Chen, W. Systematic identification of immunodominant CD8+ T-cell responses to influenza A virus in HLA-A2 individuals. Proc. Natl. Acad. Sci. USA 2011, 108, 9178–9183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Grant, E.; Wu, C.; Chan, K.F.; Eckle, S.; Bharadwaj, M.; Zou, Q.M.; Kedzierska, K.; Chen, W. Nucleoprotein of influenza A virus is a major target of immunodominant CD8+ T-cell responses. Immunol. Cell Biol. 2013, 91, 184–194. [Google Scholar] [CrossRef] [PubMed]
  59. Koutsakos, M.; Illing, P.T.; Nguyen, T.H.O.; Mifsud, N.A.; Crawford, J.C.; Rizzetto, S.; Eltahla, A.A.; Clemens, E.B.; Sant, S.; Chua, B.Y.; et al. Human CD8+ T cell cross-reactivity across influenza A, B and C viruses. Nat. Immunol. 2019, 20, 613–625. [Google Scholar] [CrossRef]
  60. Wieczorek, M.; Abualrous, E.T.; Sticht, J.; Álvaro-Benito, M.; Stolzenberg, S.; Noé, F.; Freund, C. Major Histocompatibility Complex (MHC) Class I and MHC Class II Proteins: Conformational Plasticity in Antigen Presentation. Front. Immunol. 2017, 8, 292. [Google Scholar] [CrossRef]
  61. Kumagai, Y.; Takeuchi, O.; Kato, H.; Kumar, H.; Matsui, K.; Morii, E.; Aozasa, K.; Kawai, T.; Akira, S. Alveolar macrophages are the primary interferon-α producer in pulmonary infection with RNA viruses. Immunity 2007, 27, 240–252. [Google Scholar] [CrossRef] [Green Version]
  62. Kawai, T.; Akira, S. The role of pattern-recognition receptors in innate immunity: Update on Toll-like receptors. Nat. Immunol. 2010, 11, 373. [Google Scholar] [CrossRef]
  63. 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] [Green Version]
  64. Summerfield, A.; McCullough, K.C. Dendritic Cells in Innate and Adaptive Immune Responses against Influenza Virus. Viruses 2009, 1, 1022–1034. [Google Scholar] [CrossRef] [Green Version]
  65. Hintzen, G.; Ohl, L.; del Rio, M.L.; Rodriguez-Barbosa, J.I.; Pabst, O.; Kocks, J.R.; Krege, J.; Hardtke, S.; Förster, R. Induction of tolerance to innocuous inhaled antigen relies on a CCR7-dependent dendritic cell-mediated antigen transport to the bronchial lymph node. J. Immunol. 2006, 177, 7346–7354. [Google Scholar] [CrossRef] [Green Version]
  66. GeurtsvanKessel, C.H.; Willart, M.A.; van Rijt, L.S.; Muskens, F.; Kool, M.; Baas, C.; Thielemans, K.; Bennett, C.; Clausen, B.E.; Hoogsteden, H.C.; et al. Clearance of influenza virus from the lung depends on migratory langerin+CD11b- but not plasmacytoid dendritic cells. J. Exp. Med. 2008, 205, 1621–1634. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Couch, R.B.; Kasel, J.A. Immunity to influenza in man. Annu. Rev. Microbiol. 1983, 37, 529–549. [Google Scholar] [CrossRef] [PubMed]
  68. Katz, J.M.; Plowden, J.; Renshaw-Hoelscher, M.; Lu, X.; Tumpey, T.M.; Sambhara, S. Immunity to influenza. Immunol. Res. 2004, 29, 113–124. [Google Scholar] [CrossRef]
  69. Sui, J.; Hwang, W.C.; Perez, S.; Wei, G.; Aird, D.; Chen, L.-M.; Santelli, E.; Stec, B.; Cadwell, G.; Ali, M. Structural and functional bases for broad-spectrum neutralization of avian and human influenza A viruses. Nat. Struct. Mol. Biol. 2009, 16, 265. [Google Scholar] [CrossRef] [PubMed]
  70. Taylor, P.M.; Askonas, B.A. Influenza nucleoprotein-specific cytotoxic T-cell clones are protective in vivo. Immunology 1986, 58, 417–420. [Google Scholar]
  71. 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] [CrossRef]
  72. 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]
  73. McDermott, D.S.; Weiss, K.A.; Knudson, C.J.; Varga, S.M. Central role of dendritic cells in shaping the adaptive immune response during respiratory syncytial virus infection. Future Virol. 2011, 6, 963–973. [Google Scholar] [CrossRef] [Green Version]
  74. Hufford, M.M.; Kim, T.S.; Sun, J.; Braciale, T.J. The effector T cell response to influenza infection. Influenza Pathog. Control.-Vol. II 2014, 386, 423–455. [Google Scholar]
  75. Kim, T.H.; Lee, H.K. Differential roles of lung dendritic cell subsets against respiratory virus infection. Immune Netw. 2014, 14, 128. [Google Scholar] [CrossRef] [Green Version]
  76. Rosendahl Huber, S.; van Beek, J.; de Jonge, J.; Luytjes, W.; van Baarle, D. T cell responses to viral infections—Opportunities for Peptide vaccination. Front. Immunol. 2014, 5, 171. [Google Scholar] [CrossRef] [PubMed]
  77. Yap, K.L.; Ada, G.L.; McKenzie, I.F. Transfer of specific cytotoxic T lymphocytes protects mice inoculated with influenza virus. Nature 1978, 273, 238–239. [Google Scholar] [CrossRef] [PubMed]
  78. Hamada, H.; Bassity, E.; Flies, A.; Strutt, T.M.; Garcia-Hernandez Mde, L.; McKinstry, K.K.; Zou, T.; Swain, S.L.; Dutton, R.W. Multiple redundant effector mechanisms of CD8+ T cells protect against influenza infection. J. Immunol. 2013, 190, 296–306. [Google Scholar] [CrossRef] [Green Version]
  79. Wells, M.A.; Albrecht, P.; Ennis, F.A. Recovery from a viral respiratory infection. I. Influenza pneumonia in normal and T-deficient mice. J. Immunol. 1981, 126, 1036–1041. [Google Scholar] [PubMed]
  80. Kedzierska, K.; Turner, S.J.; Doherty, P.C. Conserved T cell receptor usage in primary and recall responses to an immunodominant influenza virus nucleoprotein epitope. Proc. Natl. Acad. Sci. USA 2004, 101, 4942–4947. [Google Scholar] [CrossRef] [Green Version]
  81. Flynn, K.J.; Belz, G.T.; Altman, J.D.; Ahmed, R.; Woodland, D.L.; Doherty, P.C. Virus-specific CD8+ T cells in primary and secondary influenza pneumonia. Immunity 1998, 8, 683–691. [Google Scholar] [CrossRef] [Green Version]
  82. Nguyen, H.H.; Moldoveanu, Z.; Novak, M.J.; van Ginkel, F.W.; Ban, E.; Kiyono, H.; McGhee, J.R.; Mestecky, J. Heterosubtypic immunity to lethal influenza A virus infection is associated with virus-specific CD8+ cytotoxic T lymphocyte responses induced in mucosa-associated tissues. Virology 1999, 254, 50–60. [Google Scholar] [CrossRef] [Green Version]
  83. Wiersma, L.C.; Vogelzang-van Trierum, S.E.; Kreijtz, J.H.; van Amerongen, G.; van Run, P.; Ladwig, M.; Banneke, S.; Schaefer, H.; Fouchier, R.A.; Kuiken, T.; et al. Heterosubtypic immunity to H7N9 influenza virus in isogenic guinea pigs after infection with pandemic H1N1 virus. Vaccine 2015, 33, 6977–6982. [Google Scholar] [CrossRef]
  84. McMichael, A.J.; Gotch, F.M.; Noble, G.R.; Beare, P.A. Cytotoxic T-cell immunity to influenza. N. Engl. J. Med. 1983, 309, 13–17. [Google Scholar] [CrossRef]
  85. 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]
  86. 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] [PubMed] [Green Version]
  87. Berkhoff, E.; De Wit, E.; Geelhoed-Mieras, M.; Boon, A.; Symons, J.; Fouchier, R.; Osterhaus, A.; Rimmelzwaan, G. Functional constraints of influenza A virus epitopes limit escape from cytotoxic T lymphocytes. J. Virol. 2005, 79, 11239–11246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Berkhoff, E.; Geelhoed-Mieras, M.; Verschuren, E.; Van Baalen, C.; Gruters, R.; Fouchier, R.; Osterhaus, A.; Rimmelzwaan, G. The loss of immunodominant epitopes affects interferon-γ 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] [PubMed]
  89. Valkenburg, S.A.; Quiñones-Parra, S.; Gras, S.; Komadina, N.; McVernon, J.; Wang, Z.; Halim, H.; Iannello, P.; Cole, C.; Laurie, K.; et al. Acute emergence and reversion of influenza A virus quasispecies within CD8+ T cell antigenic peptides. Nat. Commun. 2013, 4, 2663. [Google Scholar] [CrossRef] [Green Version]
  90. 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] [CrossRef] [PubMed]
  91. 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] [CrossRef] [Green Version]
  92. 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] [CrossRef] [Green Version]
  93. 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. USA 2010, 107, 12599–12604. [Google Scholar] [CrossRef]
  94. Tu, W.; Mao, H.; Zheng, J.; Liu, Y.; Chiu, S.S.; Qin, G.; Chan, P.L.; Lam, K.T.; Guan, J.; Zhang, L.; et al. Cytotoxic T lymphocytes established by seasonal human influenza cross-react against 2009 pandemic H1N1 influenza virus. J. Virol. 2010, 84, 6527–6535. [Google Scholar] [CrossRef] [Green Version]
  95. Hillaire, M.L.B.; Vogelzang-van Trierum, S.E.; Kreijtz, J.; de Mutsert, G.; Fouchier, R.A.M.; Osterhaus, A.; Rimmelzwaan, G.F. Human T-cells directed to seasonal influenza A virus cross-react with 2009 pandemic influenza A (H1N1) and swine-origin triple-reassortant H3N2 influenza viruses. J. Gen. Virol. 2013, 94, 583–592. [Google Scholar] [CrossRef] [Green Version]
  96. Yewdell, J.W. Confronting complexity: Real-world immunodominance in antiviral CD8+ T cell responses. Immunity 2006, 25, 533–543. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Julkunen, I.; Melén, K.; Nyqvist, M.; Pirhonen, J.; Sareneva, T.; Matikainen, S. Inflammatory responses in influenza A virus infection. Vaccine 2000, 19, S32–S37. [Google Scholar] [CrossRef]
  98. Yewdell, J.W.; Bennink, J.R. Immunodominance in major histocompatibility complex class I–restricted T lymphocyte responses. Annu. Rev. Immunol. 1999, 17, 51–88. [Google Scholar] [CrossRef]
  99. Yewdell, J.W.; Del Val, M. Immunodominance in TCD8+ responses to viruses: Cell biology, cellular immunology, and mathematical models. Immunity 2004, 21, 149–153. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Chen, W.; Norbury, C.C.; Cho, Y.; Yewdell, J.W.; Bennink, J.R. Immunoproteasomes shape immunodominance hierarchies of antiviral CD8+ T cells at the levels of T cell repertoire and presentation of viral antigens. J. Exp. Med. 2001, 193, 1319–1326. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Rodriguez, F.; Harkins, S.; Slifka, M.K.; Whitton, J.L. Immunodominance in virus-induced CD8+ T-cell responses is dramatically modified by DNA immunization and is regulated by gamma interferon. J. Virol. 2002, 76, 4251–4259. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Ritmahan, W.; Kesmir, C.; Vroomans, R.M.A. Revealing factors determining immunodominant responses against dominant epitopes. Immunogenetics 2020, 72, 109–118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Maric, M.; Arunachalam, B.; Phan, U.T.; Dong, C.; Garrett, W.S.; Cannon, K.S.; Alfonso, C.; Karlsson, L.; Flavell, R.A.; Cresswell, P. Defective antigen processing in GILT-free mice. Science 2001, 294, 1361–1365. [Google Scholar] [CrossRef]
  104. Dai, G.; Carmicle, S.; Steede, N.K.; Landry, S.J. Structural basis for helper T-cell and antibody epitope immunodominance in bacteriophage T4 Hsp10. Role of disordered loops. J. Biol. Chem. 2002, 277, 161–168. [Google Scholar] [CrossRef] [Green Version]
  105. Manoury, B.; Mazzeo, D.; Fugger, L.; Viner, N.; Ponsford, M.; Streeter, H.; Mazza, G.; Wraith, D.C.; Watts, C. Destructive processing by asparagine endopeptidase limits presentation of a dominant T cell epitope in MBP. Nat. Immunol. 2002, 3, 169–174. [Google Scholar] [CrossRef]
  106. Sercarz, E.E.; Maverakis, E. Mhc-guided processing: Binding of large antigen fragments. Nat. Rev. Immunol. 2003, 3, 621–629. [Google Scholar] [CrossRef]
  107. Lazarski, C.A.; Chaves, F.A.; Jenks, S.A.; Wu, S.; Richards, K.A.; Weaver, J.M.; Sant, A.J. The kinetic stability of MHC class II:peptide complexes is a key parameter that dictates immunodominance. Immunity 2005, 23, 29–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Delamarre, L.; Couture, R.; Mellman, I.; Trombetta, E.S. Enhancing immunogenicity by limiting susceptibility to lysosomal proteolysis. J. Exp. Med. 2006, 203, 2049–2055. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Moon, J.J.; Chu, H.H.; Pepper, M.; McSorley, S.J.; Jameson, S.C.; Kedl, R.M.; Jenkins, M.K. Naive CD4+ T cell frequency varies for different epitopes and predicts repertoire diversity and response magnitude. Immunity 2007, 27, 203–213. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Martinon, F.; Gomard, E.; Hannoun, C.; Lévy, J.P. In vitro human cytotoxic T cell responses against influenza A virus can be induced and selected by synthetic peptides. Eur. J. Immunol. 1990, 20, 2171–2176. [Google Scholar] [CrossRef]
  111. McMichael, A.J.; Gotch, F.M.; Rothbard, J. HLA B37 determines an influenza A virus nucleoprotein epitope recognized by cytotoxic T lymphocytes. J. Exp. Med. 1986, 164, 1397–1406. [Google Scholar] [CrossRef] [Green Version]
  112. 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] [Green Version]
  113. Boon, A.C.; De Mutsert, G.; Fouchier, R.A.; Sintnicolaas, K.; Osterhaus, A.D.; Rimmelzwaan, G.F. Preferential HLA usage in the influenza virus-specific CTL response. J. Immunol. 2004, 172, 4435–4443. [Google Scholar] [CrossRef]
  114. Harari, A.; Cellerai, C.; Bellutti Enders, F.; Köstler, J.; Codarri, L.; Tapia, G.; Boyman, O.; Castro, E.; Gaudieri, S.; James, I.; et al. Skewed association of polyfunctional antigen-specific CD8 T cell populations with HLA-B genotype. Proc. Natl. Acad. Sci. USA 2007, 104, 16233–16238. [Google Scholar] [CrossRef] [Green Version]
  115. Quinones-Parra, S.; Grant, E.; Loh, L.; Nguyen, T.H.; Campbell, K.A.; Tong, S.Y.; Miller, A.; Doherty, P.C.; Vijaykrishna, D.; Rossjohn, J.; et al. Preexisting CD8+ T-cell immunity to the H7N9 influenza A virus varies across ethnicities. Proc. Natl. Acad. Sci. USA 2014, 111, 1049–1054. [Google Scholar] [CrossRef] [Green Version]
  116. Grant, E.J.; Josephs, T.M.; Loh, L.; Clemens, E.B.; Sant, S.; Bharadwaj, M.; Chen, W.; Rossjohn, J.; Gras, S.; Kedzierska, K. Broad CD8+ T cell cross-recognition of distinct influenza A strains in humans. Nat. Commun. 2018, 9, 5427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Ochoa, E.E.; Huda, R.; Scheibel, S.F.; Nichols, J.E.; Mock, D.J.; El-Daher, N.; Domurat, F.M.; Roberts, N.J., Jr. HLA-associated protection of lymphocytes during influenza virus infection. Virol. J. 2020, 17, 128. [Google Scholar] [CrossRef]
  118. O’Neill, H.C.; McKenzie, I.F. Quantitative variation in H-2-antigen expression. I. Estimation of H-2K and H-2D expression in different strains of mice. Immunogenetics 1980, 11, 225–239. [Google Scholar] [CrossRef] [PubMed]
  119. Deng, H.; Apple, R.; Clare-Salzler, M.; Trembleau, S.; Mathis, D.; Adorini, L.; Sercarz, E. Determinant capture as a possible mechanism of protection afforded by major histocompatibility complex class II molecules in autoimmune disease. J. Exp. Med. 1993, 178, 1675–1680. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Rowland-Jones, S.L.; Powis, S.H.; Sutton, J.; Mockridge, I.; Gotch, F.M.; Murray, N.; Hill, A.B.; Rosenberg, W.M.; Trowsdale, J.; McMichael, A.J. An antigen processing polymorphism revealed by HLA-B8-restricted cytotoxic T lymphocytes which does not correlate with TAP gene polymorphism. Eur. J. Immunol. 1993, 23, 1999–2004. [Google Scholar] [CrossRef] [PubMed]
  121. Burrows, S.R.; Silins, S.L.; Moss, D.J.; Khanna, R.; Misko, I.S.; Argaet, V.P. T cell receptor repertoire for a viral epitope in humans is diversified by tolerance to a background major histocompatibility complex antigen. J. Exp. Med. 1995, 182, 1703–1715. [Google Scholar] [CrossRef] [Green Version]
  122. Tussey, L.G.; Rowland-Jones, S.; Zheng, T.S.; Androlewicz, M.J.; Cresswell, P.; Frelinger, J.A.; McMichael, A.J. Different MHC class I alleles compete for presentation of overlapping viral epitopes. Immunity 1995, 3, 65–77. [Google Scholar] [CrossRef] [Green Version]
  123. Tourdot, S.; Gould, K.G. Competition between MHC class I alleles for cell surface expression alters CTL responses to influenza A virus. J. Immunol. 2002, 169, 5615–5621. [Google Scholar] [CrossRef]
  124. Day, E.B.; Charlton, K.L.; La Gruta, N.L.; Doherty, P.C.; Turner, S.J. Effect of MHC class I diversification on influenza epitope-specific CD8+ T cell precursor frequency and subsequent effector function. J. Immunol. 2011, 186, 6319–6328. [Google Scholar] [CrossRef] [Green Version]
  125. Herda, P.S. Disease and the colonial narrative-The 1918 influenza pandemic in Western Polynesia. N. Z. J. Hist. 2000, 34, 133–144. [Google Scholar]
  126. Curson, P.; McCracken, K. An Australian perspective of the 1918–1919 influenza pandemic. New South Wales Public Health Bull. 2006, 17, 103–107. [Google Scholar]
  127. Ahmed, R.; Oldstone, M.B.; Palese, P. Protective immunity and susceptibility to infectious diseases: Lessons from the 1918 influenza pandemic. Nat. Immunol. 2007, 8, 1188–1193. [Google Scholar] [CrossRef] [PubMed]
  128. Castrodale, L. Deaths Related to 2009 Pandemic Influenza A (H1N1) among American Indian/Alaska Natives---12 States; 2009. Available online: https://www.cdc.gov/mmwr/preview/mmwrhtml/mm5848a1.htm (accessed on 21 July 2022).
  129. Kelly, H.; Mercer, G.; Cheng, A. Quantifying the risk of pandemic influenza in pregnancy and indigenous people in Australia in 2009. Eurosurveillance 2009, 14, 19441. [Google Scholar] [CrossRef]
  130. Flint, S.M.; Davis, J.S.; Su, J.Y.; Oliver-Landry, E.P.; Rogers, B.A.; Goldstein, A.; Thomas, J.H.; Parameswaran, U.; Bigham, C.; Freeman, K. Disproportionate impact of pandemic (H1N1) 2009 influenza on Indigenous people in the Top End of Australia’s Northern Territory. Med. J. Aust. 2010, 192, 617–622. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  131. Verrall, A.; Norton, K.; Rooker, S.; Dee, S.; Olsen, L.; Tan, C.E.; Paull, S.; Allen, R.; Blackmore, T.K. Hospitalizations for pandemic (H1N1) 2009 among Maori and Pacific Islanders, New Zealand. Emerg. Infect. Dis. 2010, 16, 100. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Markey, P.; Su, J.-Y.; Wattiaux, A.; Trauer, J.; Krause, V. H1N1 2009 pandemic influenza in Indigenous Australians. Microbiol. Aust. 2011, 32, 36–38. [Google Scholar] [CrossRef] [Green Version]
  133. Lenzi, L.; Mello, A.; Silva, L.d.; Grochocki, M.; Pontarolo, R. Pandemic influenza A (H1N1) 2009: Risk factors for hospitalization. J. Bras. Pneumol. 2012, 38, 57–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Hertz, T.; Oshansky, C.M.; Roddam, P.L.; DeVincenzo, J.P.; Caniza, M.A.; Jojic, N.; Mallal, S.; Phillips, E.; James, I.; Halloran, M.E. HLA targeting efficiency correlates with human T-cell response magnitude and with mortality from influenza A infection. Proc. Natl. Acad. Sci. USA 2013, 110, 13492–13497. [Google Scholar] [CrossRef] [Green Version]
  135. Clemens, E.B.; Grant, E.J.; Wang, Z.; Gras, S.; Tipping, P.; Rossjohn, J.; Miller, A.; Tong, S.Y.; Kedzierska, K. Towards identification of immune and genetic correlates of severe influenza disease in Indigenous Australians. Immunol. Cell Biol. 2016, 94, 367–377. [Google Scholar] [CrossRef]
  136. Li-Kim-Moy, J.; Yin, J.K.; Patel, C.; Beard, F.H.; Chiu, C.; Macartney, K.K.; McIntyre, P.B. Australian vaccine preventable disease epidemiological review series: Influenza. Commun. Dis. Intell. 2016, 40, E482–E495. [Google Scholar]
  137. Hertz, T.; Nolan, D.; James, I.; John, M.; Gaudieri, S.; Phillips, E.; Huang, J.C.; Riadi, G.; Mallal, S.; Jojic, N. Mapping the landscape of host-pathogen coevolution: HLA class I binding and its relationship with evolutionary conservation in human and viral proteins. J. Virol. 2011, 85, 1310–1321. [Google Scholar] [CrossRef] [PubMed]
  138. Oukka, M.; Manuguerra, J.C.; Livaditis, N.; Tourdot, S.; Riche, N.; Vergnon, I.; Cordopatis, P.; Kosmatopoulos, K. Protection against lethal viral infection by vaccination with nonimmunodominant peptides. J. Immunol. 1996, 157, 3039–3045. [Google Scholar] [PubMed]
  139. Marshall, D.R.; Turner, S.J.; Belz, G.T.; Wingo, S.; Andreansky, S.; Sangster, M.Y.; Riberdy, J.M.; Liu, T.; Tan, M.; Doherty, P.C. Measuring the diaspora for virus-specific CD8+ T cells. Proc. Natl. Acad. Sci. USA 2001, 98, 6313–6318. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Valkenburg, S.A.; Josephs, T.M.; Clemens, E.B.; Grant, E.J.; Nguyen, T.H.; Wang, G.C.; Price, D.A.; Miller, A.; Tong, S.Y.; Thomas, P.G.; et al. Molecular basis for universal HLA-A*0201-restricted CD8+ T-cell immunity against influenza viruses. Proc. Natl. Acad. Sci. USA 2016, 113, 4440–4445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Hensen, L.; Illing, P.T.; Bridie Clemens, E.; Nguyen, T.H.O.; Koutsakos, M.; van de Sandt, C.E.; Mifsud, N.A.; Nguyen, A.T.; Szeto, C.; Chua, B.Y.; et al. CD8+ T cell landscape in Indigenous and non-Indigenous people restricted by influenza mortality-associated HLA-A*24:02 allomorph. Nat. Commun. 2021, 12, 2931. [Google Scholar] [CrossRef] [PubMed]
  142. Falfan-Valencia, R.; Narayanankutty, A.; Resendiz-Hernandez, J.M.; Perez-Rubio, G.; Ramirez-Venegas, A.; Nava-Quiroz, K.J.; Bautista-Felix, N.E.; Vargas-Alarcon, G.; Castillejos-Lopez, M.D.J.; Hernandez, A. An Increased Frequency in HLA Class I Alleles and Haplotypes Suggests Genetic Susceptibility to Influenza A (H1N1) 2009 Pandemic: A Case-Control Study. J. Immunol. Res. 2018, 2018, 3174868. [Google Scholar] [CrossRef]
  143. Dutta, M.; Dutta, P.; Medhi, S.; Borkakoty, B.; Biswas, D. Polymorphism of HLA class I and class II alleles in influenza A(H1N1)pdm09 virus infected population of Assam, Northeast India. J. Med. Virol. 2018, 90, 854–860. [Google Scholar] [CrossRef]
  144. IEDB. Immune Epitope Database. Available online: https://www.iedb.org/ (accessed on 8 April 2021).
  145. Choo, J.A.; Liu, J.; Toh, X.; Grotenbreg, G.M.; Ren, E.C. The immunodominant influenza A virus M158-66 cytotoxic T lymphocyte epitope exhibits degenerate class I major histocompatibility complex restriction in humans. J. Virol. 2014, 88, 10613–10623. [Google Scholar] [CrossRef] [Green Version]
  146. Antrobus, R.D.; Lillie, P.J.; Berthoud, T.K.; Spencer, A.J.; McLaren, J.E.; Ladell, K.; Lambe, T.; Milicic, A.; Price, D.A.; Hill, A.V.; et al. A T cell-inducing influenza vaccine for the elderly: Safety and immunogenicity of MVA-NP+M1 in adults aged over 50 years. PLoS ONE 2012, 7, e48322. [Google Scholar] [CrossRef]
  147. Antrobus, R.D.; Berthoud, T.K.; Mullarkey, C.E.; Hoschler, K.; Coughlan, L.; Zambon, M.; Hill, A.V.; Gilbert, S.C. Coadministration of seasonal influenza vaccine and MVA-NP+M1 simultaneously achieves potent humoral and cell-mediated responses. Mol. Ther. 2014, 22, 233–238. [Google Scholar] [CrossRef] [Green Version]
  148. 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] [CrossRef] [PubMed]
  149. Powell, T.J.; Peng, Y.; Berthoud, T.K.; Blais, M.E.; Lillie, P.J.; Hill, A.V.; Rowland-Jones, S.L.; McMichael, A.J.; Gilbert, S.C.; Dong, T. Examination of influenza specific T cell responses after influenza virus challenge in individuals vaccinated with MVA-NP+M1 vaccine. PLoS ONE 2013, 8, e62778. [Google Scholar] [CrossRef] [PubMed]
  150. Lillie, P.J.; Berthoud, T.K.; Powell, T.J.; Lambe, T.; Mullarkey, C.; Spencer, A.J.; Hamill, M.; Peng, Y.; Blais, M.E.; Duncan, C.J.; et al. Preliminary assessment of the efficacy of a T-cell-based influenza vaccine, MVA-NP+M1, in humans. Clin. Infect. Dis. 2012, 55, 19–25. [Google Scholar] [CrossRef] [PubMed]
  151. Swayze, H.; Allen, J.; Folegatti, P.; Yu, L.M.; Gilbert, S.; Hill, A.; Ellis, C.; Butler, C.C. A phase IIb study to determine the safety and efficacy of candidate INfluenza Vaccine MVA-NP+M1 in combination with licensed Ina CTivated infl Uenza vaccine in adult S aged 65 years and above (INVICTUS): A study protocol. F1000Res 2019, 8, 719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Vaccitech. Vaccitech Universal Flu Vaccine Passes Phase 2b Clinical Development Milestones. Available online: https://www.vaccitech.co.uk/vaccitech-universal-flu-vaccine-passes-phase-2b-clinical-development-milestones/ (accessed on 12 June 2021).
  153. Atsmon, J.; Caraco, Y.; Ziv-Sefer, S.; Shaikevich, D.; Abramov, E.; Volokhov, I.; Bruzil, S.; Haima, K.Y.; Gottlieb, T.; Ben-Yedidia, T. Priming by a novel universal influenza vaccine (Multimeric-001)-a gateway for improving immune response in the elderly population. Vaccine 2014, 32, 5816–5823. [Google Scholar] [CrossRef] [PubMed]
  154. Lowell, G.H.; Ziv, S.; Bruzil, S.; Babecoff, R.; Ben-Yedidia, T. Back to the future: Immunization with M-001 prior to trivalent influenza vaccine in 2011/12 enhanced protective immune responses against 2014/15 epidemic strain. Vaccine 2017, 35, 713–715. [Google Scholar] [CrossRef]
  155. BiondVax Pharmaceuticals Ltd. A Pivotal Trial to Assess the Safety and Clinical Efficacy of the M-001 as a Standalone Universal Flu Vaccine. Available online: https://ClinicalTrials.gov/show/NCT03450915 (accessed on 19 June 2022).
  156. AVIR Green Hills Biotechnology AG. Dose Finding Study of Single Dose GHB11L1 in Healthy Adults. Available online: https://ClinicalTrials.gov/show/NCT01078701 (accessed on 19 June 2022).
  157. Pleguezuelos, O.; Robinson, S.; Stoloff, G.A.; Caparrós-Wanderley, W. Synthetic Influenza vaccine (FLU-v) stimulates cell mediated immunity in a double-blind, randomised, placebo-controlled Phase I trial. Vaccine 2012, 30, 4655–4660. [Google Scholar] [CrossRef]
  158. Pleguezuelos, O.; Dille, J.; de Groen, S.; Oftung, F.; Niesters, H.G.M.; Islam, M.A.; Næss, L.M.; Hungnes, O.; Aldarij, N.; Idema, D.L.; et al. Immunogenicity, Safety, and Efficacy of a Standalone Universal Influenza Vaccine, FLU-v, in Healthy Adults: A Randomized Clinical Trial. Ann. Intern. Med. 2020, 172, 453–462. [Google Scholar] [CrossRef]
  159. Hull, S.; Ghosh, R. Study of VGX-3400X, H5N1 Avian Influenza Virus DNA Plasmid + Electroporation in Healthy Adults. Available online: https://ClinicalTrials.gov/show/NCT01142362 (accessed on 19 June 2022).
  160. Francis, J.N.; Bunce, C.J.; Horlock, C.; Watson, J.M.; Warrington, S.J.; Georges, B.; Brown, C.B. A novel peptide-based pan-influenza A vaccine: A double blind, randomised clinical trial of immunogenicity and safety. Vaccine 2015, 33, 396–402. [Google Scholar] [CrossRef] [Green Version]
  161. Mischler, R.; Metcalfe, I.C. Inflexal V a trivalent virosome subunit influenza vaccine: Production. Vaccine 2002, 20 (Suppl. S5), B17–B23. [Google Scholar] [CrossRef]
  162. Chichester, J.A.; Jones, R.M.; Green, B.J.; Stow, M.; Miao, F.; Moonsammy, G.; Streatfield, S.J.; Yusibov, V. Safety and immunogenicity of a plant-produced recombinant hemagglutinin-based influenza vaccine (HAI-05) derived from A/Indonesia/05/2005 (H5N1) influenza virus: A phase 1 randomized, double-blind, placebo-controlled, dose-escalation study in healthy adults. Viruses 2012, 4, 3227–3244. [Google Scholar] [CrossRef]
  163. Mössler, C.; Groiss, F.; Wolzt, M.; Wolschek, M.; Seipelt, J.; Muster, T. Phase I/II trial of a replication-deficient trivalent influenza virus vaccine lacking NS1. Vaccine 2013, 31, 6194–6200. [Google Scholar] [CrossRef] [PubMed]
  164. Treanor, J.J.; Chu, L.; Essink, B.; Muse, D.; El Sahly, H.M.; Izikson, R.; Goldenthal, K.L.; Patriarca, P.; Dunkle, L.M. Stable emulsion (SE) alone is an effective adjuvant for a recombinant, baculovirus-expressed H5 influenza vaccine in healthy adults: A Phase 2 trial. Vaccine 2017, 35, 923–928. [Google Scholar] [CrossRef] [PubMed]
  165. Antrobus, R.D.; Coughlan, L.; Berthoud, T.K.; Dicks, M.D.; Hill, A.V.; Lambe, T.; Gilbert, S.C. Clinical assessment of a novel recombinant simian adenovirus ChAdOx1 as a vectored vaccine expressing conserved Influenza A antigens. Mol. Ther. 2014, 22, 668–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Liebowitz, D.; Lindbloom, J.D.; Brandl, J.R.; Garg, S.J.; Tucker, S.N. High titre neutralising antibodies to influenza after oral tablet immunisation: A phase 1, randomised, placebo-controlled trial. Lancet Infect. Dis. 2015, 15, 1041–1048. [Google Scholar] [CrossRef]
  167. Kolhatkar, N.; Gottlieb, K.; Kasparek, K.; Hodgson, K.; Tucker, S.; Liebowitz, D. Influenza Vaccination via Oral Tablet is Protective and Induces a Unique Mucosal Immune Response. In Open Forum Infectious Diseases; Oxford University Press: Oxford, UK, 1947; p. S561. [Google Scholar]
  168. Thomas, D.N. A(H7N9) VLP Antigen Dose-Ranging Study with Matrix-M1™ Adjuvant. Available online: https://ClinicalTrials.gov/show/NCT02078674 (accessed on 19 June 2022).
  169. Feldman, R.A.; Fuhr, R.; Smolenov, I.; Mick Ribeiro, A.; Panther, L.; Watson, M.; Senn, J.J.; Smith, M.; Almarsson, Ö.; Pujar, H.S.; et al. mRNA vaccines against H10N8 and H7N9 influenza viruses of pandemic potential are immunogenic and well tolerated in healthy adults in phase 1 randomized clinical trials. Vaccine 2019, 37, 3326–3334. [Google Scholar] [CrossRef] [PubMed]
  170. NIAID. Safety and Immunogenicity Study of an Influenza Vaccination Strategy Including a H3N2 M2SR Prime Followed by a Seasonal Quadrivalent Inactivated Vaccine Boost in a Pediatric Population 9–17 Years Old. Available online: https://ClinicalTrials.gov/show/NCT03553940 (accessed on 19 June 2022).
  171. Eurocine Vaccines, A.B. Study to Assess the Safety, Tolerability and Immune Response Following Vaccination with Immunose™ FLU. Available online: https://ClinicalTrials.gov/show/NCT02998996 (accessed on 19 June 2022).
  172. Novavax. Evaluation of the Safety and Immunogenicity of a Recombinant Trivalent Nanoparticle Influenza Vaccine with Matrix M-1 Adjuvant (NanoFlu). Available online: https://ClinicalTrials.gov/show/NCT03293498 (accessed on 19 June 2022).
  173. Novavax. Phase 3 Pivotal Trial of NanoFlu™ in Older Adults. Available online: https://ClinicalTrials.gov/show/NCT04120194 (accessed on 19 June 2022).
  174. Bernstein, D.I.; Guptill, J.; Naficy, A.; Nachbagauer, R.; Berlanda-Scorza, F.; Feser, J.; Wilson, P.C.; Solórzano, A.; Van der Wielen, M.; Walter, E.B.; et al. Immunogenicity of chimeric haemagglutinin-based, universal influenza virus vaccine candidates: Interim results of a randomised, placebo-controlled, phase 1 clinical trial. Lancet Infect. Dis. 2020, 20, 80–91. [Google Scholar] [CrossRef] [Green Version]
  175. Chen, G.L. Influenza HA Ferritin Vaccine, Alone or in Prime-Boost Regimens with an Influenza DNA Vaccine in Healthy Adults. Available online: https://ClinicalTrials.gov/show/NCT03186781 (accessed on 19 June 2022).
  176. Kanekiyo, M.; Wei, C.J.; Yassine, H.M.; McTamney, P.M.; Boyington, J.C.; Whittle, J.R.; Rao, S.S.; Kong, W.P.; Wang, L.; Nabel, G.J. Self-assembling influenza nanoparticle vaccines elicit broadly neutralizing H1N1 antibodies. Nature 2013, 499, 102–106. [Google Scholar] [CrossRef] [PubMed]
  177. Leroux-Roels, I.; Waerlop, G.; Tourneur, J.; De Boever, F.; Maes, C.; Bruhwyler, J.; Guyon-Gellin, D.; Moris, P.; Del Campo, J.; Willems, P.; et al. Randomized, Double-Blind, Reference-Controlled, Phase 2a Study Evaluating the Immunogenicity and Safety of OVX836, A Nucleoprotein-Based Influenza Vaccine. Front. Immunol. 2022, 13, 852904. [Google Scholar] [CrossRef]
Figure 1. Influenza virus structure and antigenic drift and shift in influenza infections. (A) Typical influenza virus spheroidal form showing all the viral components. HA, NA, and M2 proteins form the envelope, with M1 forming a critical bridge between the viral envelope and the vRNP core, which consists of helical RNP segments containing negative-stranded genomic vRNAs and NP, along with the 3-P complex. (B) Over time, small changes associated with antigenic drift can accumulate and result in viruses that are antigenically different from their ancestors. The current vaccines that induce strain-specific antibodies do not afford complete protection, and as a result, we observe yearly epidemic outbreaks of seasonal influenza. Antigenic shift can result in a new influenza A subtype which is either generated via an intermediate host where there is, for example, gene exchange between human and avian influenza viruses, forming a new strain capable of infecting humans, or, sometimes, influenza strains from other species can directly infect humans by crossing the species barrier (e.g., in the case of H5N1). When antigenic shift occurs, most people have little or no immunity against the new virus, thereby enabling pandemics. Image created with BioRender.com. (accessed on 14 September 2022).
Figure 1. Influenza virus structure and antigenic drift and shift in influenza infections. (A) Typical influenza virus spheroidal form showing all the viral components. HA, NA, and M2 proteins form the envelope, with M1 forming a critical bridge between the viral envelope and the vRNP core, which consists of helical RNP segments containing negative-stranded genomic vRNAs and NP, along with the 3-P complex. (B) Over time, small changes associated with antigenic drift can accumulate and result in viruses that are antigenically different from their ancestors. The current vaccines that induce strain-specific antibodies do not afford complete protection, and as a result, we observe yearly epidemic outbreaks of seasonal influenza. Antigenic shift can result in a new influenza A subtype which is either generated via an intermediate host where there is, for example, gene exchange between human and avian influenza viruses, forming a new strain capable of infecting humans, or, sometimes, influenza strains from other species can directly infect humans by crossing the species barrier (e.g., in the case of H5N1). When antigenic shift occurs, most people have little or no immunity against the new virus, thereby enabling pandemics. Image created with BioRender.com. (accessed on 14 September 2022).
Viruses 14 02578 g001
Figure 2. T cell recognition of influenza virus-infected cells. Upon infection, respiratory epithelial cells and professional antigen-presenting cells (APCs), such as dendritic cells (DCs), process the viral proteins and present them as peptides in complex with MHC-I on the cell surface. Specific T cell receptors (TCR) expressed on the CD8+ T cells then recognize these peptide/MHC-I complexes. Upon engagement, the CD8+ T cells become activated and secrete cytotoxic granules (i.e., granzymes, perforin) targeting the infected cell to induce its death and produce proinflammatory cytokines (i.e., IFNs, TNF, and IL2). Image created with BioRender.com. (accessed on 14 September 2022).
Figure 2. T cell recognition of influenza virus-infected cells. Upon infection, respiratory epithelial cells and professional antigen-presenting cells (APCs), such as dendritic cells (DCs), process the viral proteins and present them as peptides in complex with MHC-I on the cell surface. Specific T cell receptors (TCR) expressed on the CD8+ T cells then recognize these peptide/MHC-I complexes. Upon engagement, the CD8+ T cells become activated and secrete cytotoxic granules (i.e., granzymes, perforin) targeting the infected cell to induce its death and produce proinflammatory cytokines (i.e., IFNs, TNF, and IL2). Image created with BioRender.com. (accessed on 14 September 2022).
Viruses 14 02578 g002
Table 1. Universal vaccine candidates in clinical trials (December 2021).
Table 1. Universal vaccine candidates in clinical trials (December 2021).
Vaccine NameTargetPlatformTargeted Age GroupAdjuvantPhase and Major FindingsReference
MVA-NP+M1NP, M1Viral vector50 and overNonePhase I:
  • Safe and induces NP- and M1-specific CD8+IFNγ+ T cells
[146,147,148]
Phase IIa:
  • Increased T-cell-killing capacity
Reduced duration of virus shedding
[149,150]
Phase IIb:
  • Completed. Result to be published
[151,152]
M-001HA, NP, M1Recombinant protein65 and overMontanide ISA-51Phase I/II:
  • Well-tolerated and safe
  • Enhanced seroconversion
  • Significant elevation in influenza-specific CMI
[153]
Phase II:
  • Induced broad protective T cells effective against different strains
[154]
Phase III:
  • Results submitted on 16th June 2021
[155,156]
FLU-vNP, M1, M2Peptide-based18–40Montanide ISA-51Phase I:
  • Safe and well-tolerated
  • IFN-γ responses > 2-fold
[157]
Phase II:
  • Well-tolerated
  • IFN-γ 38.2-fold
  • TNF-α 7.0-fold
  • IL-2 1.7-fold
[158]
VGX-3400XHA, NA, NPDNAUnknownNonePhase I:
  • Not available
[159]
FP-01.1NP, M1, P1, P2Peptide-based22–55NonePhase I
  • Acceptable safety and tolerability
  • Generated robust anti-viral T cell responses
[160,161,162,163,164]
ChAdOx1-NP+M1NP. M1Viral vector18–50NonePhase I:
  • Highly immunogenic and well-tolerated
[165]
VXA-A1.1HAViral Vector18–49TLR Agonist (dsRNA)Phase I:
  • Well-tolerated
  • 4-fold HAI
[166]
Phase II:
  • 48% of participants were protected
[167]
Avian influenza virus VLPHA NAViral vector VLPPhase I:
  • Not available
[168]
MER4101Whole virusAttenuated virus Phase I:
  • Ongoing
VAL-506440HALipid nanoparticle18–64 Phase I:
  • H10N8–78.3% (HAI) and 87.0% (MN)
  • H7N9–96.3% (HAI) and 100% (MN)
[169]
M2SRM2Attenuated virus Phase I:
  • Ongoing
[170]
Immunose FluHA Whole virusSplit virion Phase I/II:
  • Completed
  • Results not yet published
[171]
NanoFlu Phase I/II:
  • Completed
  • Results not yet published
[172]
Phase III:
  • Ongoing
[173]
cH8/1N1, H5/1N1 HA stalkAttenuated virus18–39ASO3APhase I:
  • High-level cross-reactive serum IgG antibodies.
[174]
VRCFLUDNA081-00-VPHAFerritin Nanoparticle Phase I:
  • Ongoing
[175]
OVX836NPRecombinant protein18–65 Phase I:
  • 10-fold HAI titre
[176]
Phase II:
  • Ongoing
[177]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Muraduzzaman, A.K.M.; Illing, P.T.; Mifsud, N.A.; Purcell, A.W. Understanding the Role of HLA Class I Molecules in the Immune Response to Influenza Infection and Rational Design of a Peptide-Based Vaccine. Viruses 2022, 14, 2578. https://doi.org/10.3390/v14112578

AMA Style

Muraduzzaman AKM, Illing PT, Mifsud NA, Purcell AW. Understanding the Role of HLA Class I Molecules in the Immune Response to Influenza Infection and Rational Design of a Peptide-Based Vaccine. Viruses. 2022; 14(11):2578. https://doi.org/10.3390/v14112578

Chicago/Turabian Style

Muraduzzaman, A. K. M., Patricia T. Illing, Nicole A. Mifsud, and Anthony W. Purcell. 2022. "Understanding the Role of HLA Class I Molecules in the Immune Response to Influenza Infection and Rational Design of a Peptide-Based Vaccine" Viruses 14, no. 11: 2578. https://doi.org/10.3390/v14112578

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

Article Metrics

Back to TopTop