Next Article in Journal
Three-Year Clinical Follow-Up of Children Intrauterine Exposed to Zika Virus
Next Article in Special Issue
The Antiviral Activities of Poly-ADP-Ribose Polymerases
Previous Article in Journal
Blood Counts, Biochemical Parameters, Inflammatory, and Immune Responses in Pigs Infected Experimentally with the African Swine Fever Virus Isolate Pol18_28298_O111
Previous Article in Special Issue
Host Components That Modulate the Disease Caused by hMPV
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Mammalian and Avian Host Cell Influenza A Restriction Factors

by
Joe McKellar
1,†,
Antoine Rebendenne
1,†,
Mélanie Wencker
2,
Olivier Moncorgé
1,* and
Caroline Goujon
1,*
1
Institut de Recherche en Infectiologie de Montpellier, CNRS, Université de Montpellier, CEDEX 5, 34293 Montpellier, France
2
Centre International de Recherche en Infectiologie, INSERM/CNRS/UCBL1/ENS de Lyon, 69007 Lyon, France
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Viruses 2021, 13(3), 522; https://doi.org/10.3390/v13030522
Submission received: 26 January 2021 / Revised: 12 March 2021 / Accepted: 15 March 2021 / Published: 22 March 2021
(This article belongs to the Special Issue Intrinsic Antiviral Factors)

Abstract

:
The threat of a new influenza pandemic is real. With past pandemics claiming millions of lives, finding new ways to combat this virus is essential. Host cells have developed a multi-modular system to detect incoming pathogens, a phenomenon called sensing. The signaling cascade triggered by sensing subsequently induces protection for themselves and their surrounding neighbors, termed interferon (IFN) response. This response induces the upregulation of hundreds of interferon-stimulated genes (ISGs), including antiviral effectors, establishing an antiviral state. As well as the antiviral proteins induced through the IFN system, cells also possess a so-called intrinsic immunity, constituted of antiviral proteins that are constitutively expressed, creating a first barrier preceding the induction of the interferon system. All these combined antiviral effectors inhibit the virus at various stages of the viral lifecycle, using a wide array of mechanisms. Here, we provide a review of mammalian and avian influenza A restriction factors, detailing their mechanism of action and in vivo relevance, when known. Understanding their mode of action might help pave the way for the development of new influenza treatments, which are absolutely required if we want to be prepared to face a new pandemic.

1. Introduction

Influenza viruses are segmented, negative, single stranded RNA viruses that are members of the Orthomyxoviridae family. Influenza viruses are classified into four genera: A, B, C and D (IAV, IBV, ICV and IDV) [1]. In this review, we will be focusing solely on IAV. IAV finds its reservoir in waterfowl, where it is generally asymptomatic [2], but it also circulates in poultry and certain mammalian species, including humans, pigs and horses [3]. IAV are classified based on their haemagglutinin (HA) and neuraminidase (NA) surface proteins. To date, 18 HA and 11 NA subtypes have been discovered, if we include the influenza A-like bat viruses although they have been shown to unexpectedly differ from avian and human IAV [4,5]. IAV is a seasonal virus and the causative agent of the flu, a usually mild illness causing fever, coughs, muscle aches and fatigue, but can become severe through the development of complications, such as pneumonia, possibly leading to death. This virus is known to undergo two different evolution processes: antigenic drift and antigenic shift. The IAV RNA polymerase has a low fidelity, consistently introducing novel mutations into the viral genome. Antigenic drift is a slow process where mutations are introduced into the HA and NA segments and accumulate over time. As a consequence, more recent circulating strains differ from those circulating during past seasons and can evade antibodies produced during those seasons if changes in the viral antigens HA and NA are sufficient. Antigenic shift is a more major event, where two different viruses infect the same cell and hybrid progeny viruses are produced, possessing a mix of parental viral segments, also called reassortment [6]. This can allow for the emergence of new subtypes, to which the human population is immunologically naïve, potentially leading to IAV pandemics (mainly due to new and/or different HA segments) [7]. Hence, four naturally occurring pandemics have arisen since the beginning of the 20th century. The infamous 1918 “Spanish flu” (named as such for being first reported in Spanish news outlets due to the fact that media of countries involved in World War I were under censure) was caused by a H1N1 virus that infected over a third of the world’s population, killing over 50 million. Other pandemics include the 1957 H2N2 “Asian flu” killing over 2 million, the 1968 H3N2 “Hong Kong flu” killing over 1 million and the more recent 2009 H1N1 “Swine flu” killing around 200,000 (the latter was, thankfully, milder than seasonal flu). Genetic analysis suggests that the 1918 H1N1 virus originated from direct transfer of a whole avian strain into humans [8]. In contrast, the two following pandemics (H2N2 and H3N2) occurred by reassortment between the circulating human strain at the time (H1N1 and H2N2, respectively) and avian influenza viruses. As for the virus responsible for the 2009 pandemic (pH1N1), it was a complicated reassortant whose combination of genomic segments were unique, but all segments of the virus were present in viruses found in swine [9,10]. Also noteworthy, but different to the previous naturally occurring pandemics, was that of 1977, also called the “Russian flu”, where the culprit was a virus that was identical to the H1N1 virus that had circulated in humans in the 1950’s [11]. The most probable explanation for the unlikely occurrence of reappearance of a strain identical to the one circulating two decades prior is that of the accidental release of this virus from a frozen source. Hence, IAV pandemics emerge in a seemingly random manner. To date there are only a few approved antiviral treatments against flu, to which IAV can easily escape [12,13,14,15], as well as vaccines against seasonal strains (currently pH1N1, H3N2 as well as influenza B lineages). Vaccines are efficient when fully matching the circulating strains, however, they take time to produce and would not be ready until months within a new pandemic. Therefore, we are currently not ready to face a severe influenza pandemic and the current COVID-19 crisis is a cruel reminder of the damage that can be caused by respiratory viruses with pandemic potential.
This review has the goal of detailing the main intrinsic and innate defenses humans and other hosts, such as ducks and chickens, possess to defend themselves from IAV infection. These defenses come in the form of restriction factors, which are antiviral proteins expressed by the host cell that restrict or impair viruses at various stages of their replication cycle. These restriction factors can be constitutively expressed and/or induced as a result of viral sensing and consequent signaling. Understanding these natural defenses could potentially unlock ideas that might help pave the way for new treatments, necessary in the uncertainty of future IAV pandemics.

2. Influenza A Virus Life Cycle

To understand the different restriction factors that inhibit IAV, a general understanding of the viral life cycle is necessary. When IAV (Figure 1) encounters a target cell, the HA surface glycoprotein interacts with a sialylated receptor on the cell plasma membrane [16,17]. Avian strains preferentially bind to α-2.3-linked sialic acids whereas human strains to α-2.6-linked sialic acids [18,19]. After attachment, the virus enters the cell through epsin1-dependent clathrin-mediated endocytosis, clathrin-independent endocytosis or through macropinocytosis [20,21,22]. Now inside the cell, the viral particle is localized in an endocytic vesicle. As this vesicle progresses through its lifecycle, acidification of the endosomal pH activates the surface M2 proton pump, acidifying the viral particle [23,24]. This decrease in pH fragilizes the matrix, allowing uncoating [25]. The acidification of the endosomal vesicle is also responsible for the multi-step fusogenic transformation of HA [26,27]. The HA fusion peptide then mediates the fusion of the cellular endosomal membrane together with the viral lipid envelope, allowing exit of the eight viral ribonucleoproteins (vRNPs) into the cytoplasm. vRNPs are constituted of a negative strand genomic RNA segment coated with viral nucleoprotein (NP) molecules and bound to the viral polymerase. Each vRNP acts as an autonomous transcription-replication unit once it reaches the cell nucleus. vRNPs enter the nucleus using the active importin α/β1 pathway mediated by the viral protein NP that possesses a Nuclear Localization Signal (NLS) that interacts with importin α [28,29].
Once the vRNPs are in the nucleus, viral messenger RNAs (mRNAs) are produced from the genomic viral RNA (vRNA) template by the resident viral RNA-dependent RNA polymerase (RdRp). This is a primed process, using primers from cellular mRNAs being transcribed by the cellular RNA polymerase II [30]. Polymerase Basic 2 (PB2), part of the tripartite viral polymerase along with Polymerase Acidic (PA) and Polymerase Basic 1 (PB1), binds the 5′ caps of cellular mRNAs. Following this, the PA endonuclease cleaves the 5′ end of the cellular mRNA, which releases a capped 9-17 nucleotide primer that is used to initiate the transcription of the viral mRNA using viral genomic RNA as a template [31]. This process is called cap-snatching. The viral mRNAs also possess polyA tails, just as cellular mRNAs, but that are produced by the viral polymerase. The newly produced mRNAs are exported into the cytoplasm and translated by cellular ribosomes, ending in the production of viral proteins. The viral polymerase subunits as well as NP, M1 and NS2/NEP proteins are then trafficked back to the nucleus. Occurring at the same time as mRNA synthesis, the viral polymerase is also responsible for vRNA replication. The viral polymerase transcribes the negative vRNA into positive complementary RNA (cRNA) in a non-primed process. The newly formed cRNA associates with novel NP, PB1, PB2 and PA molecules into complementary RNPs (cRNPs). Following this, the cRNPs undergo replication to create new negative vRNA copies which associate into novel vRNPs. For further details on the transcription and replication steps, see [32,33,34].
The newly formed vRNPs undergo export into the cytoplasm in a CRM1-dependent manner, mediated by a daisy-chain of vRNP-M1-NS2-CRM1 [35,36,37]. Once in the cytoplasm, vRNPs are next trafficked towards the plasma membrane in a Rab11-dependent process that relies on endoplasmic reticulum remodeling [38,39] and undergo assembly [40]. Ongoing questions in the lifecycle of IAV are how vRNPs are transferred to the plasma membrane, and how they organize themselves inside the viral particle into the classical “7+1” formation [41]. This highly coordinated process is followed by viral budding and egress. Release of the newly formed viral particle is catalyzed by NA that cleaves the sialic acids retaining the particle on the cell membrane.

3. Sensing and Interferon Response

3.1. Sensing and Interferon Response in Mammalian Species

The detection of viral infections by the innate immune system occurs through the recognition of pathogen-associated molecular patterns (PAMPs) by pattern recognition receptors (PRRs). PAMPs are either present in invading pathogens or are generated during infection. Viral PAMPs usually correspond to specific nucleic acid structures of the viral genome not found in host cells or viral replication intermediates such as double-stranded RNA (dsRNA). Distinct classes of PRRs are important for influenza virus recognition and induced inflammation (Figure 2): the Toll-Like receptors (TLR), the Retinoic acid-inducible gene I (RIG-I) receptor, the NOD-like receptor family member NOD-, LRR- and pyrin-containing 3 (NLRP3) inflammasome and the Z-DNA binding protein 1 (ZBP1) (recently reviewed in [42]). These PRRs share different subcellular localizations and act at different stages of the viral replication cycle. Indeed, NLRP3, RIG-I and ZBP1 are located in the cytoplasm, whereas TLR3, 7 and 8 are present within endosomes [42].
Activation of these PRRs triggers signaling cascades leading to the production and secretion of type I and III interferons (IFNs), pro-inflammatory cytokines, eicosanoids and chemokines, which can act in a paracrine and autocrine manner. In particular, type I and III IFNs are potent antiviral mediators leading to the expression of hundreds of IFN-stimulated genes (ISGs), which induce an antiviral state [43]. Type I and III IFNs are notably expressed after PAMP recognition by RIG-I and TLR3/7/8, shortly described hereafter.
RIG-I is expressed by all cell types and activated upon binding to short dsRNA and RNA harboring di- or tri-phosphorylated 5′ ends. Following RNA binding to RIG-I, the latter is ubiquitinylated by TRIM25 or RIPLET E3 ubiquitin ligases, which are required for RIG-I activation and allow its higher-order oligomerization [44]. Then, RIG-I interacts with mitochondrial antiviral-signaling protein (MAVS), promotes its oligomerization and the formation of a signaling hub called signalosome required for downstream signaling. RIG-I is considered as the main immune sensor for IAV and recognizes its genomic RNA [45]. RIG-I importance was confirmed in in vivo studies in a mouse model. Indeed, type I IFN production was impaired in Rig-I−/− mice infected by IAV, which correlated with delayed viral clearance [46]. Of note, mice are non-natural hosts for influenza, however, they are frequently used as models to study influenza-mediated pathogenesis given the large variety of genetic tools available.
TLR3 recognizes dsRNA and is expressed in various cell types including dendritic cells (DCs) and lung epithelial cells, the latter being target host cells of IAV. Numerous studies have shown the important role of TLR3 in IAV sensing. Indeed, in human bronchial epithelial cells, IAV infection leads to the activation of TLR3 and subsequent production of proinflammatory cytokines [47]. However, infection of Tlr3−/− mice with IAV resulted in higher viral production in lungs but also a diminished proinflammatory cytokine response, which is actually less detrimental for IAV infected mice [48]. These results suggest that defense mechanism against IAV requires a tight regulation of immune responses in order to control IAV replication without exacerbated inflammatory responses. In that context, TLR3, while promoting viral clearance, could also participate in IAV-mediated pathogenesis.
TLR7 is mainly expressed by plasmacytoid DCs (pDCs) and lung epithelial cells whereas TLR8 is expressed by immune cells such as monocytes, macrophages and DCs. Classically, binding of single-stranded RNA by TLR7/8 in endosomes leads to their activation and the induction of their downstream signaling pathway in a Myeloid differentiation primary response 88 (MyD88) adaptor-dependent manner (Figure 2). Mice deficient for Myd88 are more susceptible to IAV infection when sublethal doses where used for wild type animals [49]. However, the involvement of TLR7/8 in IAV immune sensing remains debated. Indeed, another study using Myd88−/− mice showed no requirement of MyD88 in protection against IAV infection using lethal doses of IAV even for the wild-type mice, which could explain the discrepancy between the two studies [48]. Moreover, pDCs, which are the professional IFN producers, were surprisingly shown to not impact the course of IAV infection in mice [50].
Following engagement of RIG-I and TLR3/7/8, several transcription factors such as IRF3, IRF7 (in pDCs or in cells pre-exposed to IFN), AP1 and NF-κB are activated and translocate to the nucleus, leading to the transcription of type I/III IFN genes. Type I and III IFNs are then secreted and act in a paracrine and autocrine manner. The type I IFN receptor (composed of IFNAR1 and IFNAR2 subunits) is ubiquitously expressed, whereas the type III IFN receptor (composed of IFNLR1 and IL10R2 subunits) is preferentially expressed on mucosal epithelial cells, liver cells and some myeloid cells [51]. Upon binding to their cognate receptors, type I and III IFNs activate a Janus kinase (JAK)/signal transducer and activator of transcription (STAT) pathway. This signaling cascade results in the transcription of several hundreds of ISGs, some having antiviral and immunomodulatory properties, and in the establishment of an antiviral state in infected and neighboring cells.
Type I and III IFNs are well known to potently inhibit IAV replication in humans and in mice, both in vitro and in vivo through the expression of ISGs (reviewed in [52]). Moreover, IAV infection has also been shown to be restricted by type I IFNs in vitro in bat cells [53]. In line with this, mice models where IFN signaling is abrogated or exogenous IFNs are administrated, demonstrated the important role of both type I [54,55,56,57,58] and type III [59], [60] IFNs in vivo in IAV clearance and in the control of influenza-induced pathogenesis. Type I and III IFNs induce a similar subset of ISGs, with type I IFN leading to a more rapid induction and decline of ISG expression [43,61]. Importantly, type III IFNs are produced earlier than type I IFNs following IAV infection of mice, which leads to viral inhibition without inducing inflammation [61,62].
Of note, the importance of type I and III IFNs antiviral activities was also evidenced in humans. Indeed, exogenous prophylactic administration of IFNα seemed to reduce the severity of influenza infection [63]. Moreover, in humans, carried inborn errors in genes involved in the type I/III IFN pathway such as TLR3 [64], IRF7 [65] or IRF9 [66] were reported in some life-threatening influenza patients. This highlights a critical role of type I and III IFNs in influenza disease severity and in the control of influenza replication.
Furthermore, the role of type I and III IFNs extends beyond a role in limiting viral replication. Indeed, type I and III IFNs are important to shape adaptative immune responses. Mice lacking type III IFN receptors show defective DC responses, ultimately preventing the optimal generation of effector CD8 T-cells [67,68], protective memory T-cell responses against IAV [67] and optimal antibody responses [68]. This effect of type III IFN appears indirect and depends, at least in part, on the secretion of thymic stromal lymphopoietin (TSLP) by microfold cells (M cells) in the upper airways [68], which promotes migration of resident CD103+ DCs to draining lymph nodes and germinal center reactions. Thus, type III IFNs via TSLP potentiate the establishment of adaptative immune responses [68] (reviewed in [61]).
Similarly, type I IFN favors memory CD8 T-cell cytolytic activity in an antigen-independent manner, by promoting the expression of granzyme B [69]. Furthermore, mice that lack STAT1, which fail to signal through all three types of IFN receptors, do not succeed to generate an IgG2a response and present a significant bias toward TH2 (T helper 2) differentiation and IgE response, which is at the origin of an exacerbated lung pathology [70].

3.2. Sensing and Interferon Response in Avian Species

The Galliformes order, which includes chickens, are characterized by the absence of RIG-I [71], which has been proposed to be at the origin of the high susceptibility of chickens to RNA viruses compared to ducks [72]. Despite the fact that RIG-I is the main sensor for IAV in mammalians, an IFN response is mounted in chickens infected by highly pathogenic influenza virus strains, potentially due to sensing of viral RNAs by RIG-I-like receptors MDA-5 and LGP2 [73,74,75]. Moreover, in chickens, only TLR3 and 7 are involved in the recognition of RNA viruses given that TLR8 is pseudogenized [76]. Type I and III IFNs are nevertheless produced in response to the activation of these PRRs, which inhibit IAV replication in vitro, in ovo and in vivo [77,78,79].

3.3. Viral Antagonism of Innate Immune Sensing by IAV

IAV has evolved various proteins such as NS1, PB1-F2 or PA-X to counteract the induction of innate immune responses either by inhibiting innate immune sensing or by inhibiting signaling downstream of type I or III IFN receptors (reviewed in [42,80]).
NS1 inhibits RIG-I-mediated immune sensing through several mechanisms. For instance, NS1, by interacting with dsRNA through its RNA binding domain, sequesters RIG-I ligands [81,82,83]. Furthermore, NS1 has been shown to interact with several proteins such as the E3 ubiquitin ligases TRIM25 or RIPLET, preventing RIG-I activation [84,85,86].
PB1-F2 is encoded from a +1 open reading frame (ORF) of the PB1 gene as a result of a leaky ribosomal scanning in some IAV strains. PB1-F2 plays pleiotropic effects, such as the enhancement of viral polymerase activity [87], the induction of cell death [88] or the modulation of innate immune signaling [89]. Concerning the latter, PB1-F2 has been shown to interact with the transmembrane domain of MAVS, leading to the dissipation of the mitochondrial membrane potential, which is essential for MAVS-mediated signaling. Moreover, association of PB1-F2 with MAVS could also impair its oligomerization at mitochondria, which is required for downstream signaling [90,91].
PA-X is encoded by IAV segment 3 via +1 ribosomal frameshifting, generated by ribosomal pausing on a rare CGU codon. PA-X shares the first 191 amino acids with PA, containing the endonuclease domain [92,93]. PA-X expression induces a general inhibition of host cell translation. This host cell shut-off is due to the degradation of host PolII-generated transcripts by the endonucleolytic domain in coordination with the 5′ to 3′-exonuclease Xrn1, leading to, among other effects, the inhibition of innate immune response [94,95,96,97].

4. IAV Restriction Factors

Restriction factors come in many forms but with the common effect of inhibiting viral replication. These proteins can act at various stages of the viral lifecycle and have also started emerging as being able to potentiate viral sensing and regulate antiviral signaling pathways. In this part of the review, we will focus on restriction factors that are able to inhibit IAV at various stages of its life cycle, be they constitutively expressed or induced following viral sensing. Figure 3 gives an overview of all restriction factors described here and the step they inhibit.

4.1. Restriction Factors Inhibiting Viral Entry

Viral entry into a host cell is an absolutely critical step in the life cycle of viruses, only possible if the adequate receptors are present at the cell surface. Interestingly, a number of restriction factors act at this step and are reviewed hereafter.

4.1.1. IFITMs: Major, Pan-Viral Restriction Factors Preventing Cytosolic Entry

Interferon-induced transmembrane proteins (IFITMs) are small transmembrane proteins [98,99] that are evolutionary conserved across vertebrates [100,101]. In humans, the IFITM family is composed of five genes (IFITM1; IFITM2; IFITM3; IFITM5; IFITM10), whereas in mice, two additional genes (Ifitm6; Ifitm7) have been described. Generally, vertebrate IFITM families can be divided into immunity-related IFITM (IR-IFITM), IFITM5 and IFITM10. Initially, IFITM3 was identified in a siRNA screen to identify host factors modulating IAV infection [102]. IFITM1/2/3 proteins have since been shown to exert an antiviral activity against a wide range of viruses (reviewed in [103]). IFITM proteins are thought to be type IV transmembrane proteins based on epitope mapping in cells [104,105,106] and preliminary structural studies [107]. IFITMs are mostly located in early and late endosomes, and lysosomes for IFITM2/3, as shown by immunofluorescence studies [108,109] and live cell imaging [110,111]. However, IFITM1 lacks the AP2 sorting motif (YXXΦ) at the N-terminus and is rather located at the plasma membrane. Both the antiviral mechanism and specificity of IFITMs are linked to their cellular localization, with IFITM1 being more active against viruses that enter the target cell at the plasma membrane, whereas IFITM2/3 preferentially inhibit viruses entering target cells through endocytosis. Of note, IFITM3 can be incorporated into nascent viral particles and prevent viral spread by inhibiting subsequent viral fusion with a new target cell, as shown in the case of HIV-1 [112,113].
IFITM proteins are known to prevent cytosolic entry of the targeted viruses and a number of hypotheses to explain their activity, mostly based on the study of IFITM3, have emerged (reviewed in [99]). Overexpression of IFITM3 leads to an expansion of late endocytic compartments (Rab7- and LAMP1-positive compartments) along with their over-acidification, which suggests that IFITM3 could have a role in controlling pH-dependent viral entry [114]. It has also been proposed that IFITM3 inhibits hemifusion and lipid mixing [115]. However, it was later shown that IFITM3 prevents IAV cytosolic entry without inhibiting hemifusion, by affecting the formation of the fusion pore [116]. Interestingly, the impact on the formation of the fusion pore was later confirmed and IFITM3-positive vesicles were shown to fuse with vesicles containing incoming virions before hemifusion. This phenomenon was specific to viruses restricted by IFITM3 only and leads to an increased rate of virus trafficking to late endocytic compartments [111,117].
IFITM proteins possess intramembrane (IMD) and transmembrane (TMD) domains separated by an intracellular loop (ICL) and variable N- and C-terminal domains. All these domains play an important role in anti-IAV activity (as reviewed in [103]). Indeed, the CD225 domain, which is highly conserved among them, comprises the IMD, the TMD and the ICL domains and is essential for antiviral activity [118]. Moreover, deletion of the N-terminal domain of IFITM3 resulted in an impaired ability to inhibit IAV [118]. Specific motifs have been shown to be important for the antiviral activity, namely the 20YXXΦ23 motif in IFITM2/3, which is important for correct subcellular localization [119] or the 81SVKS84 motif of IFITM3 in the CIL domain, which is essential for IAV inhibition [120]. IFITM proteins are highly modified by post-transcriptional modifications (PTMs), which modulate their antiviral activity. Indeed, murine and human IFITM3 have been reported to be S-palmitoylated on cysteines 71, 72 and 105, and this is important for antiviral activity [104,121]. S-palmitoylation is a reversible PTM, which increases IFITM3 hydrophobicity and affinity for cellular membranes and controls its clustering to membranes [121]. Furthermore, other PTMs have been shown to play a pivotal role in IFITM3 restriction of IAV. Indeed, phosphorylation of tyrosine 20 [122,123], ubiquitination of lysines 24, 83, 88 and 104 [104] and monomethylation of lysine 88 [122] have been shown to alter IFITM3 antiviral activity against IAV.
Importantly, in vivo studies in mice have confirmed that IFITM3 is a potent inhibitor of IAV infection [124,125]. Indeed, infection of Ifitm3−/− mice with different strains of IAV leads to a fulminant viral pneumonia and to death. More specifically, upon infection with the low-pathogenic A/X-31 (H3N2) IAV strain, mice showed little difference in virus replication in the lungs during the first 48 h of infection. However, lungs of Ifitm3−/− mice contained 10-fold higher levels of replicating virus than the WT mice at 6 days post-infection and this was associated with profound morbidity [124]. Analysis of immune cells recruited to lungs showed that the lack of IFITM3 resulted in a reduced proportion of CD4+ and CD8+ T-cells but an elevated proportion of neutrophils [124]. This was also associated with increased levels of pro-inflammatory cytokines such as TNF-α, IL-6, G-CSF and MCP-1 [124]. Moreover, extrapulmonary lesions, such as myocarditis, have also been reported upon IAV infection in mice, and IFITM3 expression in cardiac tissue has been shown to protect mice against such cardiac lesions [126]. Interestingly, IFITM3 also plays a pivotal role in protecting immune cells from IAV infection. Following IAV infection in mice, IFITM3 is upregulated in respiratory DCs, limiting viral load and apoptosis in those cells. In the absence of IFITM3, DCs also show impaired migration from lung mucosa to the draining lymph node, thus preventing antigen presentation to naïve CD8+ T-cells and the establishment of adaptive immunity [127]. Similarly, IFITM3 is transiently expressed in activated T-lymphocytes, following cognate antigen recognition or type I IFN stimulation. However, IFITM3 is mostly expressed in lung tissue-resident memory T cells (TRM), a cell population that ensures the first line of defense against pathogen re-encounter while being directly exposed to infection. In TRM cells, IFITM3 expression is maintained after primary IAV infection due to hypomethylation of its promoter, leading to enhanced survival after a new challenge with IAV [128]. Altogether, these observations indicate that IFITM3 plays a pivotal role in the establishment of both innate and adaptative immune responses.
As mentioned above, host genetics strongly impacts susceptibility to various infections. Two genetic variations within human IFITM3 have been described to be associated with increased IAV infection severity. Both are single nucleotide polymorphisms (SNPs), the most prevalent being the rs12252-C, mainly present in Asian populations [129]. Multiple studies, including two meta-analyses, found that rs12252-C is associated with severe outcomes following IAV infection [124,130,131,132,133] (meta-analyses: [134,135]). However, these results are controversial, as other studies showed no association between the polymorphism and the risk of severe influenza [136,137,138]. rs12252-C is a synonymous SNP which is located at a splice acceptor site, potentially leading to a putative truncated IFITM3 with a 21 amino acid deletion at its N-terminus, termed D21-IFITM3 [124]. Nevertheless, an in vitro study showed that D21-IFITM3 still restricts IAV infection [139], and studies performed in rs12252-C homozygous patients have failed to detect the truncated isoform of IFITM3 by high-throughput RNA-sequencing [140,141]. The second SNP associated with severe risk of influenza infection is rs34481144-A, mainly present in Caucasian populations [129]. This SNP is present in the 5′UTR region of IFITM3, in the promoter region. It modulates IFITM3 expression by governing CTCF binding to the gene promoter, thus inactivating IFITM3 transcription. However, in a recent Brazilian cohort, rs34481144-A was not associated with severity or mortality during IAV infection, nor with IFITM3 levels [138], highlighting the complexity of this type of analysis.
Considering the importance of several species in the zoonotic lifecycle of influenza viruses, numerous studies have elucidated roles of IFITM proteins in the restriction of IAV in avian species, swine and bats. Of note, the IR-IFITM gene cluster is conserved across mammals [142]. Indeed, it has been shown that five IFITM proteins are present in swine (two homologs of IFITM1, and one homolog of IFITM2, IFITM3 and IFITM5), which display antiviral activity against IAV in porcine cells [143]. The subcellular localization of swine IFITM proteins is similar to their human homologs [143]. In bats, IFITM3 has been shown to inhibit IAV cell entry and to be located in early/late endosomes [144]. Moreover, bat IFITM3 is S-palmitoylated on cysteines 71, 72 and 105, just as the human homolog, which is also important for its antiviral activity [145]. Finally, the repertoire of IR-IFITM is conserved between mammals and avian species. IFITM3 has been shown to restrict IAV in both chickens [146] and ducks [147], but no effect of duck IFITM1, IFITM2 and IFITM5 has been observed [147]. Following infection with both low and highly pathogenic avian influenza (LPAI/HPAI) strains, duck IFITM1/2/3 are highly upregulated in the lungs and the ileum from day 1 post-infection, whereas only IFITM3 was modestly upregulated in the ileum of chickens.

4.1.2. NCOA7-AS: A V-ATPase Regulator, Preventing Endocytosis-Mediated Viral Entry

The short isoform of Nuclear Receptor Coactivator 7 (NCOA7), NCOA7-Alternative-Start (NCOA7-AS, also called NCOA7-B), belongs to the TLDc (Tre2/Bub2/Cdc16 (TBC), lysin motif (LysM), domain catalytic)-containing family of proteins. Seven TLDc proteins are found in humans and mice, including Oxidation resistance (OXR) proteins, and NCOA7 short and long isoforms (AS and Full-Length, FL, respectively) [148]. TLDc proteins have been shown to play a protective role against oxidative stress, through an unknown mechanism [149] (reviewed in [148]). NCOA7-FL is a known interactor of the estrogen receptor, which translocates to the nucleus upon estradiol treatment, where it was suggested to act as a transcriptional coregulator [150]. NCOA7-AS does not seem to share this property and is uniquely upregulated by type I IFNs via an internal promoter [151]. Recently, NCOA7-AS has been shown to inhibit infection with IAV and other viruses entering the cell via endocytosis, by limiting viral entry into the cytoplasm of host cells [152]. Both NCOA7-FL and -AS have been shown to interact with several subunits of the vacuolar ATPase (V-ATPase), the proton pump responsible for endolysosomal acidification [152,153,154]. Through an as yet unelucidated mechanism, NCOA7-AS interaction with the V-ATPase leads to a higher acidification of the endolysosomal system, which increases antigen degradation and seems detrimental to IAV by, possibly, affecting the ability of HA to allow fusion [152]. Interestingly, NCOA7-AS acts independently of IFITM3 [152], but further work is now warranted to fully elucidate the NCOA7-AS molecular mechanism of action as well as its importance in vivo.

4.1.3. ZMPSTE24: An IFITM Cofactor, Which Also Acts Independently

Zinc-metallopeptidase STE24 (ZMPSTE24) is a zinc-metalloprotease containing seven transmembrane domains, which is constitutively expressed and localized at the membrane of cytoplasmic organelles as well as at the inner nuclear membrane [155]. In mammals, this protease is notably required for the maturation of lamin A, which is important for nuclear architecture. ZMPSTE24 has been recently identified as a restriction factor of a large number of viruses entering cells via endocytosis, among which IAV [155], [156]. However, ZMPSTE24 expression is not upregulated by type I IFNs or viral infection [155]. ZMPSTE24’s antiviral activity is independent of its protease activity, as shown by mutations of the conserved HEXXH catalytic motif [155]. Reminiscent of IFITM3 or NCOA7-AS, ZMPSTE24 inhibits cytosolic entry of IAV. Interestingly, ZMPSTE24 has been shown to interact with IFITM1/2/3 and to be required for IFITM antiviral activity. However, IFITM proteins are not necessary for ZMPSTE24′s antiviral activity, as overexpression of IFITM3 in ZMPSTE24−/− cells does not restore the restriction phenotype. This suggests that ZMPSTE24 acts downstream of IFITM proteins, as well as independently of IFITMs. Furthermore, in vivo studies in laboratory mice revealed that deletion of Zmpste24 leads to higher viral titers upon IAV infection, along with an increased production of pro-inflammatory cytokines and a higher mortality rate [155].

4.1.4. CH25H: A Restriction Factor In Vitro but an Enhancer of Inflammation In Vivo

Cholesterol-25-hydrolase (CH25H) is an enzyme induced by type I IFNs and which catalyzes the oxidation of cholesterol into 25-hydroxycholesterol (25HC), a soluble oxysterol. Both CH25H and 25HC are important regulators of cholesterol homeostasis. Shortly, CH25H and 25HC inhibit the transcription factor sterol regulatory element-binding protein (SREBP) and activate the transcription factor liver X receptor (LXR). These two transcriptional factors have a great importance in modulating cellular cholesterol homeostasis [157]. Moreover, 25HC also represses the expression of the HMGCR gene and induces the degradation of the 3-hydroxy-3-methyl-glutaryl-coenzyme A reductase (HMGCR) protein, which leads to an inhibition of the cholesterol biosynthesis pathway. CH25H and its product 25HC have been shown to be involved in the restriction of a large number of enveloped viruses by inhibiting the fusion between viral and endosomal membranes [157]. This antiviral activity was also observed in vitro for IAV both in canine MDCK cells (i.e., the cell line typically used for IAV amplification in vitro) [158] and in murine immortalized airway epithelial cells [159], suggesting that CH25H is a restriction factor for IAV. However, surprisingly, deletion of Ch25h in a mouse model of IAV infection is protective [159], and is associated with a decreased inflammatory-induced pathology. This discrepancy between in vitro and in vivo results may be correlated with a role of CH25H as an amplifier of inflammation, which surpasses its restriction activity [159].

4.1.5. B4GALNT2: A Factor with the Potential of Inhibiting Avian IAV Entry When Overexpressed

Genetic loss-of-function screens such as whole-genome CRISPR/Cas9 screens or RNAi screens have been widely used to identify host-dependency factors. However, little have focused on discovering factors that inhibit viral replication, until recently. Indeed, a new approach based on a whole-genome CRISPR/Cas9 activation screen recently led to the identification of β-1,4 N-acetylgalactosaminyltransferase 2 (B4GALNT2) as a pan-avian influenza virus inhibitor [160]. B4GALNT2 is a glycosyltransferase, which catalyzes the transfer of a N-acetylgalactosamine (GalNAc) to the penultimate galactose of α-2.3-linked sialic acids. Thus, without changing the relative amounts of sialic acids at the cell surface, B4GALNT2 prevents the interaction of HA with these receptors. The specificity of sialic acid targeted by this protein explains the spectrum of IAV strains inhibited, namely avian strains, which have a preference for α-2.3-linked sialic acids [160]. However, the importance of endogenous B4GALNT2 in the control of avian IAV has not been demonstrated by loss-of-function studies, which would be important to carry out. Moreover, as pigs are considered as mixing vessels, it would be interesting to study the role of B4GALNT2 in the inhibition of avian IAV replication in pigs.
In addition to B4GALNT2, this CRISPR/Cas9 activation screen identified two other potential influenza inhibitors: the transmembrane 9 superfamily member 2 (TM9SF2) and Ras Additionally, Rab Interactor 2 (RIN2) [160]. TM9SF2 is an endolysosomal protein whose role in the maturation of the endosome was postulated, while RIN2 is an interactor of the early endosome-associated protein Rab5. However, neither the molecular mechanisms involved in their restriction activity nor their physiological role have been further explored.

4.1.6. MUC1: A Decoy for IAV Viral Particles

Airway epithelial cells secrete mucus, which constitutes one of the first physical barriers to infection in the lungs, acting like a shield blocking the binding of pathogens to their cellular receptors. Proteins from the mucin glycoprotein family, and notably cell surface mucins (cs-mucins), are important components of the mucus. They contain a large extracellular domain which displays several oligosaccharides, and which is non-covalently linked to a transmembrane domain, allowing it to be shed from the cell surface. Moreover, the cytoplasmic tail of cs-mucins can activate signaling cascades upon phosphorylation and interaction with intracellular adaptor molecules [161,162,163,164,165]. MUC1 is a member of the cs-mucin family, is constitutively expressed and is upregulated by type I IFNs [166]. In addition to its shield role, MUC1 was suggested to be an important downregulator of infection-driven inflammation responses upon bacterial infections [167]. Interestingly, this role extends to IAV infection, as IAV-infected Muc1−/− mice display, more rapidly, high viral loads and a faster, enhanced inflammatory response than wild-type mice [168]. In vitro, IAV was found to interact with MUC1 and to trigger the shedding of its extracellular domain [168]. This secreted extracellular domain could then act as a decoy and bind IAV to limit cell infections. As cs-mucins are highly conserved across species, it could be interesting to explore the potential antiviral role of MUC1 in other species.

4.2. Restriction Factors Inhibiting Viral Genome Replication

The steps of nuclear import, transcription, replication, protein translation, vRNP assembly and nuclear export are targeted by a large number of restriction factors acting either by inhibiting viral replication and transcription, translation or by degrading viral components such as the subunits of the viral polymerase or vRNAs.

4.2.1. MX Dynamin-Like GTPases: Broad-Spectrum Antiviral Proteins with Possible Multiple Modes of Action

The family of large dynamin-like GTPases all share a similar three-dimensional structure and general organization: a N-terminal domain of various length and unknown structure, a globular head that sports the GTPase activity, a stalk domain as well as a middle domain (or bundle signaling element, BSE) that allows the flexibility of these mechanoenzymes [169]. Human Myxovirus resistance proteins 1 and 2, MX1 and MX2 (also named MxA and MxB) are members of this family and possess potent antiviral activity against a broad range of viruses [170]. MX1 is best known for its ability to inhibit IAV infection, but it can also inhibit a wide range of positive- or negative-single stranded or double stranded RNA viruses as well as certain DNA viruses [170], and MX2 inhibits HIV-1, Herpesviruses, HCV and HBV [171,172,173,174,175,176,177]. The MX genes were first discovered in mice (MmMx1), with the serendipitous observation that the A2G mice strain were resistant to IAV infection whereas most inbred laboratory strains, lacking a functional MmMx1 locus, were highly susceptible [178]. The subsequent creation of transgenic mice expressing a functional MmMx1 or human MX1 confirmed the importance in vivo of MX1 proteins [179,180]. MX1 is an ISG induced by type I or III IFN signaling [181], but it is commonly known that ectopic expression of MX1, without the induction of other ISGs, is sufficient for IAV restriction. Human MX1 is a cytoplasmic protein whereas mouse MmMx1 is mostly nuclear, which would probably hint at slightly different mechanisms of action for IAV restriction.
MX1 proteins possess intrinsic antiviral determinants that are essential for effective IAV restriction. Indeed, the presence of a functional GTPase domain [182], an intact BSE [183] and oligomerization via the stalk domain [184] are essential for correct anti-IAV activity. Other essential determinants are two loops localized at the end of the stalk: the L2 and L4 loops [185]. The L4 loop, which has been positively selected throughout evolution [186], harbors essential amino acids (F561 and G562) for its antiviral activity [187]. This is different from MX2, which only needs its N-terminal domain and an oligomerization domain for HIV-1 restriction [188,189,190]. Despite many years of intense study, the exact mechanism of action of MX proteins remains misunderstood. What is largely admitted nevertheless in the case of MX1 is that the main viral target in the case of IAV seems to be the viral NP protein, as it has been shown that mutations in NP can confer resistance to MX1 restriction [191,192,193,194]. This is reinforced by the fact that avian strains of IAV are more susceptible to inhibition than human strains and this has also been attributed to NP [180,195]. In the case of MmMx1, PB2 could also be an additional target [196,197]. MX1 proteins can form dimers and further oligomerize through numerous interfaces and have been theorized to form ring-like structures that could circle around vRNPs, possibly by interacting with NP (and potentially PB2) [196,198], but experimental data supporting this model are still lacking. In any case it seems that the antiviral mechanism of action of MX1 may act at several steps of the viral lifecycle, be it nuclear import of vRNPs, viral protein production or directly on replication itself [170,199,200,201]. To fully understand the fine details of MX1-induced IAV restriction, more studies are needed to confirm/contest the many hypotheses that exist in the literature.
Several studies have been performed to determine the importance of MX1 in humans through the search for polymorphisms within the human population [202,203,204]. G-G interface allelic variants were discovered, resulting in loss of GTPase activity, but were found in heterozygotes and did not show any dominant-negative effect on wild-type MX1 [202,203]. Several other variants were discovered in the stalk domain that resulted in loss of antiviral activity with dominant negative effects [203]. Unfortunately, no data of IAV infection history from homozygous carriers of these mutations are available. This aspect is nicely reviewed in [205].
Another interesting aspect of MX1 biology is the fact that mammalian MX1 proteins are active in cell types belonging to other species. For example, human MX1 expressed in mouse cells is active against IAV [206] and the opposite is also true for MmMx1 in human cells [196,207]. Transgenic mice expressing human MX1 are also protected against IAV infection [180,208,209] and have proven to be very useful models in the field. These observations may either suggest the existence of common crucial cofactors conserved between species, or that MX1 proteins are sufficient by themselves to inhibit IAV, but further in-depth studies are required to elucidate this.
A secondary, nonetheless important, role in IAV control was recently discovered for MX1. Indeed, MX1 was found to be an inflammasome sensor in human respiratory epithelial cells upon IAV infection, triggering IL-1β secretion [209]. The MX1 inflammasome formation was found to be dependent on MX1 oligomerization and also observed with cytoplasmic porcine MX1, but not nuclear MmMx1 [209], which may potentially be explained by the differential localization of these proteins. Transgenic mice expressing human MX1 induced inflammasome activation in respiratory epithelium upon IAV challenge [209], showing a potential relevance in vivo. Nevertheless, the authors showed that the MX1 inhibition of early steps of IAV replication was independent of the inflammasome [209], reinforcing the idea that MX1 plays multiple roles during IAV infection.
While MX1 from humans and mice as well as in other mammals such as rats, pigs and bats are potent anti-influenza A factors [210,211,212], in ducks and chickens, MX1 seems to be inactive against avian IAV. Indeed, chicken MX1 was shown to be devoid of antiviral activity against several low and highly pathogenic strains and to be non-essential for the IFN response in chickens [213,214,215,216]. This could be due to the fact that chicken MX1 does not possess a GTPase activity, which is known to be crucial for the anti-IAV activity of human and mouse MX1 [182,215]. Similar to chicken MX1, duck MX1 has not been found to possess antiviral activity against IAV [217]. It would be interesting to understand the molecular mechanisms that lead to this loss of function in poultry as MX1 seems to be a crucial IAV restriction factor, especially since the overexpression of human MX1 or MmMx1 in chicken cells is able to inhibit IAV replication [215].

4.2.2. GBP Dynamin-Like GTPases: Indirect and Direct Inhibitors

Guanylate binding proteins (GBPs) are also part of the large dynamin-like GTPases, like MX proteins, and have been shown to have a broad antimicrobial restriction activity, encompassing bacteria, protozoa and viruses [218]. The GBP family is composed of seven members which are ISGs induced by type I and type II IFNs [219]. Some of these proteins have been shown to inhibit IAV [220,221,222,223]. Indeed, a splice variant of human GBP3, hGBP-3ΔC, and to a much lower extent GBP1, were shown to possess anti-IAV activity [220]. Overexpression of hGBP-3ΔC potently decreased IAV replication and, conversely, silencing of GBP3 significantly increased IAV replication (by over 1 log). The overexpression of hGBP-3ΔC decreased IAV polymerase activity and the accumulation of viral RNA species, and GTP binding was found to be necessary for IAV restriction by hGBP-3ΔC [220]. Another study observed antiviral activity of GBP1 through overexpression experiments [221], however depletion experiments are still lacking to confirm the physiological role of GBP1 in inhibiting IAV replication. GBP1 antiviral activity was suggested to be antagonized by NS1, but only ectopic expression of NS1 was used to show this [221], therefore further evidence is needed. In cells from guinea-pigs, guinea-pig GBP1 has been proposed to play a role in RIG-I mediated restriction of IAV, as its knockdown dampened the IFN response as shown by reducing the induction of MX1 mRNA after infection [224]. In line with this, ectopic expression and silencing experiments showed that human GBP5 modestly inhibits IAV replication, and this was proposed to be due to an indirect effect, through the regulation of the IFN and inflammatory responses [222]. Interestingly, GBP2 and GBP5 were also shown to inhibit the furin protease and therefore to have a negative impact on infection by various viruses, including a modest effect on IAV [223]. On the contrary, GBP7 knockout out by CRISPR/Cas9 reduced IAV replication and its overexpression increased IAV replication [225]. This was found to be due to the suppression of the innate immune system by GBP7, defining it as a proviral factor [225]. In the future, it would be interesting to see if all these different observations could be recapitulated in Gbp KO mice. Finally, non-synonymous polymorphisms in human GBP2 have been linked to a reduced responsiveness to IFN treatment during HBV treatment [226], but no data on the relevance of GBP polymorphisms in IAV infection response have been reported.

4.2.3. TRIM Proteins: A Large Family of Antiviral Proteins Implicated in Innate Immune Signaling

The tripartite motif (TRIM) protein family contains more than 80 proteins in humans, which are involved in various cellular processes, including cell proliferation, autophagy and immunity [227]. They are characterized by a conserved domain organization at the N terminus (known as the TRIM or RBCC motif) composed of a catalytic RING domain which possesses an E3-ubiquitin ligase activity, one or two B-box domains and a coiled-coil dimerization (CCD) domain. They also contain C-terminal domains, which vary among TRIM proteins, for instance the SPRY domain or the PRY domain [227]. Various TRIM proteins have been shown to restrict IAV infection and act either by inducing the degradation of viral components or by inhibiting viral replication.
-
Role of TRIM22, TRIM41 and TRIM14 in the degradation of viral components:
TRIM22 has been shown to be upregulated by type I IFNs [228,229] or by IAV infection and to inhibit IAV [228]. TRIM22 knockdown increases IAV replication by more than one order of magnitude at low multiplicities of infection [228]. Moreover, TRIM22 is one of the players of the IFN-induced restriction of IAV as shown by the reduced inhibition observed once silenced [228]. Overexpression experiments have shown that TRIM22 interacts with NP and poly-ubiquitinates it via its RING domain, leading to NP degradation in a proteasome-dependent manner [228]. Interestingly, some H1N1 strains such as the H1N1 strains isolated in 1933 and 1934 (i.e., WSN and PR8) and pH1N1 were shown to be insensitive to TRIM22-mediated antiviral activity, whereas more recent seasonal H1N1 isolates were restricted by TRIM22 [230]. The presence of four R to K substitutions in NP differentiate the restricted strains from the non-restricted, potentially being the consequences of adaptation due to a sustained circulation in humans since 1918. Introduction of these mutations in resistant strains sensitizes them to TRIM22 restriction. This showed that adaptative mutations may sometimes surprisingly induce greater sensitivity to innate immune effectors [230].
Additional TRIM proteins inhibit IAV by interacting with NP, promoting its degradation. This is the case of TRIM41, which is notably not IFN-inducible [229,231]. TRIM41 interacts with NP via its C-terminal SPRY domain, mediates NP poly-ubiquitination via its RING domain, which in turn induces NP degradation by the proteasome [232]. Interestingly, TRIM14, a TRIM protein lacking a RING domain and upregulated by type I IFNs [229], has also been proposed to restrict IAV and to interact with NP [231]. This interaction, notably mediated by the PRYSPRY domain, alters NP stability by K48-linked polyubiquitination and NP-mediated proteasomal degradation. This suggests that the NP-TRIM14 interaction leads to the recruitment of a yet-to-define functional E3 ubiquitin ligase to mediate NP polyubiquitination [231].
TRIM32 was identified as a PB1 binding partner by mass spectrometry [233]. TRIM32 is constitutively expressed and located in the cytoplasm in normal conditions, but translocates to the nucleus upon IAV infection. TRIM32 overexpression inhibits IAV infection and its silencing increases IAV replication (by ~1 log) as measured by plaque assays. TRIM32 seems to inhibit IAV infection by inducing poly-ubiquitination of PB1, which leads to its proteasomal degradation and to the inhibition of the viral polymerase activity [233]. Further studies are still warranted to demonstrate the physiological role of TRIM32 in the control of IAV replication by using Trim32-deficient mice [234,235].
Finally, TRIM35, another ISG, which plays a general role in the induction of type I IFNs, leads to PB2 K48-linked polyubiquitination and to its proteasomal degradation [236]. Importantly, TRIM35 silencing has a significant impact on IAV replication in vitro (about 1 log) and Trim35−/− mice are more susceptible to IAV replication and induced death [236].
-
Role of TRIM25 and TRIM56 in the inhibition of viral genome replication:
TRIM25 is a protein that plays a fundamental role in RIG-I-dependent innate immune sensing of IAV and other viruses. Indeed, ubiquitination of RIG-I by the E3 ubiquitin ligase TRIM25 is necessary for RIG-I activation, thus leading to the production of type I and III IFNs following recognition of RNA ligands (reviewed in [237]). Interestingly, the NS1 proteins from all IAV strains can interact with TRIM25, however this interaction does not always translate to an inhibition of the IFN response [84,85,86]. Interestingly, an additional role of TRIM25 in the restriction of IAV has been recently unraveled [238]. Indeed, nuclear TRIM25 interacts with vRNPs and inhibits viral RNA synthesis independently of its ubiquitin ligase activity. More precisely, TRIM25 inhibits the onset of RNA elongation by preventing the movement of viral RNAs into the RdRp [238].
The overexpression of another TRIM protein, TRIM56, has been shown to restrict IAV infection (by 1 log) as measured by plaque assays [239]. Moreover, stable knockdown of TRIM56 led to a 6-fold increase of de novo infectious virion production. Interestingly, TRIM56 antiviral activity is independent of its ubiquitin ligase activity and overexpression of a 63-residue segment present in its C-terminal has the same antiviral activity as the full length TRIM56. Moreover, the authors showed that viral RNA synthesis was impeded in infected cells overexpressing TRIM56 [239]. However, an in vivo study demonstrated that Trim56−/− were not more susceptible to IAV infection than wild type mice, failing to validate a physiological role of TRIM56 in the control of IAV replication in vivo [240].
Of note, the validation of the antiviral activity of some of the TRIM proteins in vivo can be challenging, not because of the absence of mice models, but given the importance of many TRIM proteins in the activation or regulation of innate immune sensing pathways. Moreover, TRIM proteins have not been extensively studied in pigs, bats or birds in terms of diversity of the repertoire, conservation and antiviral activity compared to their human orthologs, studies which could nevertheless be of great interest.

4.2.4. ZAP and ZFP36L1: Antiviral Zinc Finger Proteins

Zinc finger Antiviral Protein (ZAP, also named ZC3HAV1 for Zinc Finger CCCH-type containing Antiviral 1) is an IFN-induced antiviral protein that inhibits a large variety of viruses (reviewed in [241]). ZAP is expressed as four isoforms (S, M, L and XL) [242]. However, ZAP-S and ZAP-L are the predominantly expressed isoforms. ZAP-S and ZAP-M isoforms are upregulated by type I IFNs, ZAP-S being the most upregulated [242]. The main two isoforms of human ZAP protein (ZAP-S and ZAP-L), arising from alternative splicing, differ only by the presence or the absence of a poly ADP-ribose polymerase (PARP)-like domain at the C-terminus. Both isoforms have a N-terminal domain containing four CCCH-type zinc fingers as well as a large central domain containing a TiPARP homology domain (TPH) and a WWE domain. The four zinc fingers are responsible for the RNA binding [243]. A zinc ion is coordinated via three cysteines and one histidine in each zinc finger, although their folding and protein sequences are different.
ZAP has been shown to inhibit viral replication by repressing the translation or promoting the degradation of viral mRNAs. To mediate this antiviral activity, ZAP recognizes and interacts directly with specific RNA elements enriched in CpG dinucleotides [243,244]. ZAP-L silencing modestly increased IAV replication in human cells, as measured by plaque assays [245]. Ectopic expression of tagged viral proteins and ZAP-L showed that ZAP-L interacts with PB2 and PA subunits of IAV polymerase in a PARP domain-dependent manner [245]. ADP-ribosylated PA and PB2 are associated with ZAP-L and then ubiquitinated, which leads to their proteasomal degradation. The viral protein PB1 interacts with ZAP-L and counteracts its antiviral activity. Indeed, PA and PB2 binding sites on ZAP-L are in close proximity to their PB1 binding sites, thus leading to the dissociation of PA and PB2 from ZAP-L [245]. Furthermore, the ability of ZAP-S (which lacks the PARP domain) to inhibit IAV has also been shown by overexpression experiments [246]. However, the relevance of these findings is questionable, as Zc3hav1 knockout had no impact on IAV replication in murine embryonic fibroblasts (MEFs) [246]. ZAP-S overexpression inhibits viral protein expression by binding to viral mRNAs, promoting their degradation and inhibiting their translation [246]. The viral protein NS1 seems able to antagonize overexpressed ZAP-S through an unknown mechanism [246]. Nevertheless, it is tempting to speculate that NS1 might antagonize ZAP-S through TRIM25 inhibition given that the TRIM25-mediated ubiquitination of ZAP-S seems critical for its antiviral activity and that some NS1 mutants fail to do so [246]. No current validation of ZAP importance in vivo to restrict IAV replication has been published. However, Zc3hav1−/− mice have been previously described in the literature and could be used to serve that purpose [247,248].
ZAP appears to be evolutionary conserved in animals [249,250,251]. However, little is known about the antiviral activity of ZAP orthologues in pigs, bats or in avian species. Analysis of the antiviral activity and the binding specificity of ZAP from different avian species, showed that ZAP from aquatic birds exhibited a broader antiviral activity, potentially due to a lesser selectivity for CG-enriched RNA elements [250]. It could be of great interest to analyze the restriction of IAV by ZAP from different species, given that ZAP appears to be an important selection pressure shaping genome composition and could greatly influence cross-species transmission.
Another CCCH-type zinc finger protein, ZFP36L1 has been recently discovered to harbor anti-influenza activity [252]. ZFP36L1 was shown to be induced upon TNF-α treatment and by viral infection suggesting a potential role of this protein in the host cell antiviral defense. However, ZFP36L1 is not IFN-inducible [166]. Overexpression of ZFP36L1 decreased viral titer (around 1 log), as measured by plaque assays and decreased M1, M2, NS1, NS2 and HA protein levels, but not mRNA levels. Conversely, knockdown of this protein increased viral replication (around 1 log), and this was rescued by complementation. The proposed mechanism of action is that ZFP36L1 decreases M1 and NS2 protein levels, inhibiting vRNP nuclear export. This is supported by the fact that NP seemed to be trapped in the nucleus under overexpression conditions. Zfp36l1−/− mice have been generated in other studies [253] and could be used to validate this restriction factor in vivo.

4.2.5. IFITs: Friends or Foes?

The IFN-induced proteins with tetratricopeptide repeats (IFITs) are an RNA binding family of ISGs that are among the most upregulated during antiviral signaling. There are four IFIT proteins coded in humans: IFIT1 (or ISG56), IFIT2 (or ISG54), IFIT3 (or ISG60) and IFIT5 (or ISG58). These proteins are known to inhibit a wide range of viruses [254]. The first antiviral activity attributed to these proteins was the ability to bind eIF3 subunits and decrease cap-dependent translation efficiency [255]. Later, IFITs were found to bind non-self 5′-ppp or 2′-O unmethylated RNA [256,257,258] as well as to viral proteins [259]. Additionally, of note is the fact that IFITs can associate with each other which can modulate their functions (reviewed in [260]). In the case of IAV, data is quite conflicting. Indeed, the individual silencing of IFIT1, 2 and 3 was reported to increase IAV replication, as measured by a viral PolI reporter expression system after infection, suggesting the IFIT1/2/3 heterotrimer complex could be responsible for this antiviral activity [256]. However, interestingly, another more recent study, using both CRISPR/Cas9 knockout and overexpression experiments, showed that human and murine IFIT1 are not major restriction factors of IAV [261]. The authors also showed that IFIT1 actually has a low binding affinity for 5′-ppp. Moreover, in vivo studies have shown that Ifit1−/− mice do not show any differences compared to WT mice in regard to IAV infection, indicating that murine IFIT1 does not have a major role in restriction or pathogenesis of IAV in vivo [261]. In addition, a major study recently showed that IFIT2 is actually a cofactor for IAV [262]. Indeed, IFIT2 was shown to be repurposed by IAV to become a proviral effector promoting the translation of viral mRNAs [262]. Therefore, further studies are required to understand the role, if any, of mammalian IFITs as restriction factors of IAV in vitro and in vivo.
An avian IFIT (avIFIT) protein in ducks was found to resemble the structure of mammalian IFIT5. Upon testing, the authors showed that chicken cells stably expressing duck IFIT inhibited IAV infection (by 1 log) [263]. They also showed that avIFIT was able to bind to NP and also enhance the IFN response. Chicken IFIT5 (chIFIT5) may also harbor antiviral properties as it has been shown to be able to interact with 5′-ppp-containing viral RNAs [264]. The same group later showed that transgenic chickens stably expressing chIFIT5 had marked resistance to H5N1 infection [265], while another study showed that CRISPR/Cas9 chIFIT KO cells allowed for higher IAV replication (around 1 log) [263]. Further studies are needed to understand the molecular mechanisms at play for avian IFIT IAV inhibition and the differences between the avian and mammalian proteins.

4.2.6. OAS-Family Proteins: NS1-Counteracted dsRNA Binding Inhibitors

Humans possess four genes coding for the 2′,5′-oligoadenylate (2-5A) synthetase (OAS) proteins: OAS1, 2, 3 and OAS-like (OASL), that are IFN-inducible [166,266]. Once activated by dsRNA, OAS1, 2 and 3 are able to synthesize 2-5A which bind to monomeric RNaseL, inducing its dimerization and activation. Active RNaseL then cleaves cellular or viral ssRNA and can induce autophagy and apoptosis, which is detrimental to viral replication [267]. The small RNAs resulting from RNaseL cleavage of ssRNA have been also shown to activate RIG-I signaling, leading to IFN production [268,269]. OASL lacks the ability to produce 2-5A, but has nevertheless been shown to enhance RIG-I sensing [270].
The OAS/RNaseL system is able to inhibit a wide number of RNA viruses [271], but in the case of IAV, it has been shown that NS1 protects from the effects of this pathway [272,273]. Indeed, a virus encoding a dsRNA-binding deficient NS1 protein was sensitive to the RNaseL pathway [272], suggesting that NS1 dsRNA binding capability counteracts OAS dsRNA recognition, preventing RNaseL activation and subsequent ssRNA degradation. OAS3 has been shown to have a higher affinity for dsRNA compared to OAS1 and OAS2 and also to be the main producer of 2-5A in IAVΔNS1 infected cells, as OAS3 KO abolished 2-5A production and raised viral titer [273]. The importance of OAS proteins in vivo has not been studied (of note, Rnasel−/−, but not Oas−/− mice exist [274]), but they are likely to have little impact seeing as OAS proteins are antagonized by NS1.

4.2.7. PKR and NF90: An Example of Host–Virus Coevolution

Protein kinase R (PKR) is a serine/threonine kinase that possesses two N-terminal dsRNA-binding motifs as well as an effector C-terminal kinase domain [275]. PKR is constitutively expressed and slightly upregulated by IFN treatment [166]. During viral infection, PKR recognizes dsRNA which induces its dimerization and the subsequent phosphorylation of the eukaryotic translation initiation factor 2 (eIF2α) at position serine 51 [276]. This has the consequence of shutting down cellular translation, therefore inhibiting the replication of numerous viruses [277]. However, IAV seems to have evolved countermeasures. Indeed, in IAV infected cells, there is an increase of P58IPK which can inhibit PKR, thereby promoting viral protein production and efficient replication [278]. NS1 has also been shown to play a protective role against PKR by binding to viral dsRNA blocking recognition by PKR [279,280,281,282]. In line with this, IAV with defective or missing NS1 are more susceptible to the effects of PKR [279,280,281]. NS1 mediated inhibition of PKR has been mapped to two essential residues in the NS1 N-terminal domain: R35 and R46 [282]. Of note, mice lacking PKR are more susceptible to IAV infection, suggesting a role in vivo for this antiviral protein [280,283].
In chickens, PKR expression is induced following H5N1 inoculation, but this induction is not sufficient to induce resistance [216]. Mallard duck’s PKR anti-IAV phenotype has yet to be studied, but it has been shown to be upregulated during IAV infection and to miss the N-terminal second RNA binding domain compared to human PKR, which could affect its function [284].
Interestingly, a cellular countermeasure to the NS1-mediated PKR inhibition has recently been discovered: the protein Nuclear factor 90 (NF90), another dsRNA-binding protein [285,286] able to inhibit different viruses [287]. NF90 does not seem to be induced by IFN [166]. NF90 interacts with both NS1 and PKR, reducing the impact of NS1′s PKR inhibition [288]. This is an interesting example of coevolution between viruses and host cells. NF90 was also shown to exert antiviral activity against NS1-deficient IAV by interacting with the C-terminus of PKR and upregulating PKR phosphorylation (knockdown experiments in 293T cells) [289]. Of note, knockdown and overexpression experiments showed that NF90 also independently but modestly inhibits IAV by interfering with viral polymerase activity, possibly through an interaction with NP [290,291]. Unfortunately, Nf90−/− mice do not seem to be viable [292], so exploring the extent of NF90 importance in IAV infection in vivo will be complicated.

4.2.8. MOV10 and DDX21: Antiviral Cellular Helicases

Helicases are enzymes that unwind double-stranded nucleic acids, and are classified into three super-families and two small families based on conserved motifs (reviewed in [293]). Some of these helicases have proven to possess antiviral activity against many different viruses [294]. Here, we describe two helicases that have been shown to inhibit IAV.
Moloney leukemia virus 10 (MOV10) is a 5′ to 3′ RNA helicase from the UPF1-like family [295]. It has been shown to play roles in miRNA metabolism, mRNA stabilization and translation, retroelement inhibition and viral restriction [296,297,298,299] and is partially upregulated by type I and II interferon treatment [166], [300]. MOV10 has also been proposed to be an IAV restriction factor. Indeed, MOV10 was identified by mass spectrometry as an IAV vRNP binding partner, through an interaction with NP [301]. MOV10 silencing increased IAV replication (by ~1 log) and conversely, MOV10 overexpression decreased IAV replication [301,302]. MOV10 seems to inhibit IAV polymerase activity by preventing NP nuclear import [301,302] and by possibly sequestering incoming vRNPs inside P-Bodies [302]. Of note, NS1 was proposed to counteract MOV10 activity [302]. MOV10 has also been shown to inhibit other viruses by enhancing the production of IFN in a RIG-I/MAVS independent manner [300], and it would be interesting to determine whether this indirect activity play also a role in IAV inhibition. Mov10−/− mice are not viable [303], rendering difficult the in vivo investigation of the importance of this protein in IAV infection control.
DDX21 is part of the DEAD (Asp–Glu–Ala–Asp)-box family of RNA helicases which are RNA-binding helicases that play fundamental roles in RNA metabolism [304]. DDX21 has been shown to resolve R loops (i.e., DNA:RNA hybrids), promoting genomic stability [305]. It has also been shown to be slightly upregulated by type II IFN [166] and to harbor antiviral activity against RNA viruses [306]. DDX21 silencing increases IAV replication in vitro as shown by plaque assays [307]. DDX21 binds to PB1, preventing the assembly of the polymerase, thus inhibiting replication [307]. However, NS1 is able to attenuate this restriction by binding DDX21 and suppressing its antiviral activity [307]. Its physiological importance has not been demonstrated yet due to the lack of Ddx21−/− mice.

4.2.9. HDACs: Indirect Inhibitors

Histone Deacetylases (HDACs) are involved in the deacetylation of acetylated proteins [308]. These proteins are classified into four classes: class I, class II, class III and class IV HDACs, with only some of the members being upregulated after IFN treatment, such as HDAC4 and HDAC6 with type II IFN [166]. Recent publications have suggested roles for different members of this family in IAV replication, either as dependency factors or as inhibitors.
Silencing experiments showed that HDAC1, a class I HDAC, modestly increases viral infection, whereas HDAC8 seems to act as a viral cofactor [309,310]. The modest anti-IAV effect of HDAC1 was linked to STAT1 phosphorylation regulation and ISG induction [310]. A similar role was reported for HDAC2 [311].
Concerning class II HDACs, HDAC6 was proposed to be essential for IAV uncoating [312], however, HDAC6 silencing has an overall positive impact on IAV replication as measured by plaque assays [313]. Pharmacological inhibition of HDAC6 increased trafficking of viral components to the plasma membrane and viral release, a mechanism probably involving acetylated microtubules [313]. More recently, Hdac6−/− mice were shown to be slightly more susceptible to influenza infection, which was actually correlated with a downregulation of several innate immune pathways [314]. Further clarification on the impact and the role of HDAC6 is therefore required. Another class II HDAC, HDAC4, has also been proposed to harbor a modest anti-influenza activity [315].
Class III HDACs, also known as sirtuins, have been proposed to modestly inhibit influenza virus infection through depletion experiments, without further characterization [316].
Finally, a class IV HDAC member, HDAC11, has also been shown to exhibit anti-influenza activity, probably by regulation of the IFN response [317].
The roles of HDACs in the host cell influenza defense are often modest and indirect, with some members playing proviral rather than antiviral roles. More studies would therefore be necessary to understand the molecular mechanisms and the importance in vivo of HDACs on IAV infection.

4.2.10. ANP32s: Major Cofactors and Potential Minor Inhibitors

The mammalian acidic nuclear phosphoprotein 32kDa (ANP32) family consists of five members: ANP32A, B, C, D and E, but there is an ongoing debate over the true existence of C and D as bone fide expressed genes. These proteins all contain an N-terminal leucine-rich repeat (LRR) region, a central region and a C-terminal unstructured low-complexity acidic region (LCAR) region. ANP32 proteins are known to carry out various roles such as chromatin regulation, phosphatase regulation, involvement in apoptosis and intracellular transport [318]. ANP32A was initially discovered to be a major co-factor of IAV determining species specificity [319], mediating the assembly of the IAV replicase [320].
Identification of ANP32A as a major factor in IAV life cycle and in host restriction/adaptation was a breakthrough. It was known for a long time that the polymerase from classical avian influenza A viruses did not function efficiently in human cells and that the PB2 subunit, and in particular residue 627, contributed to this host restriction [321,322]. The lack of a cellular cofactor for the avian polymerase in mammalian cells was previously suggested to be responsible for this restriction [323]. Chicken ANP32A was indeed identified through an elegant genetic screen as an essential cofactor allowing an avian-origin PB2 protein to function in mammalian cells [319]. In addition, ANP32A was shown to be crucial for IAV replication in both birds and mammals, but avian-adapted polymerases are not able to efficiently use mammalian ANP32A. A 33 amino acid motif is missing from the mammalian version of ANP32A in comparison to avian ANP32A proteins and introduction of this motif in mammalian ANP32A allows avian-origin PB2 to function in mammalian cells [319]. Interestingly, ostriches and other ratites lack this region, which elegantly explains the previous observation that ostriches behave more like mammalian species regarding avian IAV PB2 adaptation (i.e., acquisition of the mammalian signature PB2 E627K) [324]. Later, both ANP32A and ANP32B were shown to be essential for IAV replication in human cells, but either of them being redundant, as cells that lacked one were still able to support the polymerase, but cells lacking both could not [325].
Another study, which confirmed the importance of ANP32A and ANP32B as host cofactors, highlighted that other members of ANP32 family such as ANP32C, ANP32D and ANP32E have an opposite, negative effect on IAV replication [326]. Indeed, overexpressing human ANP32C or ANP32D together with ANP32A had a negative effect on the replication of IAV (possessing either the avian PB2-627E signature or the human PB2-627K signature) compared to ANP32A overexpression alone [326]. ANP32E only had a negative effect for avian PB2-627E signature IAV, but knockdown of ANP32E did not show any effect [326]. Of note, Anp32e−/− mice have been produced and are viable [327], but seeing as a mice KO model of ANP32A did not show any significant effect on IAV infection [328] we could imagine a similar outcome for Anp32e−/− mice. Nevertheless, more studies must be performed to assess the mechanisms through which ANP32C, ANP32D and ANP32E might impair viral polymerase activity.
Chickens also possess ANP32s but with the particularity that chicken ANP32A contains a splicing site, which allows for the production of two splice variants, ANP32A-33 and ANP32A-29, of which ANP32A-33 is more potent [329,330]. A team recently made the discovery that SRSF10 is able to bind to a cis-regulatory element, promoting the production of ANP32A-29 transcripts and decreased that of ANP32A-33 transcripts. This had a negative impact on avian IAV polymerase activity but not mammalian strains, as mammals do not produce the two ANP32A-29 and ANP32A-33 splice variants [331].

4.2.11. ISG15: ISGylation-Mediated Inhibition of NS1 Activities

Interferon-stimulated gene product 15 (ISG15) is one of the most upregulated ISGs [332]. It is a ubiquitin-like protein known to be conjugated to lysine residues of proteins in a process called ISGylation. This process involves the E1 activating enzyme Ube1L, E2 conjugating enzyme UbcH8 and E3 ligase enzymes Herc5 or TRIM25/EFP. The effects on cellular proteins of ISGylation are many (reviewed in [333]), such as the enhancement of antiviral signaling through IRF3 ISGylation to give an example [334]. Many viral proteins have been discovered as targets of ISGylation [333]. The first evidence of ISG15 impacting IAV replication was that Isg15 deficient mice were shown to display increased susceptibility to IAV infection [335,336]. It is thought that NS1 is the target of ISG15′s antiviral activity, as NS1 has been shown to be a target of ISGylation, which prevents a number of its essential functions [337,338]. Further studies on ISG15 are therefore required to fully understand its impact on IAV infection.

4.2.12. ISG20: An Exonuclease Affecting IAV Replication When Overexpressed

IFN-stimulated gene 20 kDa protein (ISG20) is a 3′ to 5′ exonuclease with specificity for ssRNA and is induced by interferon signaling [339]. This protein has been shown to inhibit a wide range of viruses, including IAV [340,341]. ISG20 overexpression was shown to decrease IAV replication [342,343], and an IFN-inducible, long-non-coding RNA sharing most of its sequence with ISG20, Lnc-ISG20, and regulating ISG20 expression, was shown to also impact IAV replication [344]. The effect of overexpressed ISG20 was shown to be dependent on its exonuclease activity and attributed to an interaction with NP, as shown by immunoprecipitation using overexpressed ISG20 and viral infection [343]. Although not confirmed for IAV, the antiviral effect of ISG20 against some viruses has also been shown to be attributed to an upregulation of the type I IFN response [345] or to an inhibition of viral translation [341]. Evidence of a physiological role of ISG20 against IAV, using depletion experiments, as well as in vivo experiments are still lacking. Of note, Isg20−/− mice are available [341] and could be used to this end.

4.3. Restriction Factors Affecting Viral Assembly, Egress or Maturation

To date, only a few host restriction factors have been shown to interfere with the late stages of IAV replication cycle. The late stages comprise viral assembly, egress and maturation of newly generated virions.

4.3.1. Tetherin/BST-2: A General Inhibitor of Viral Release with a Controversial Activity on IAV

Bone marrow stromal antigen 2 (BST-2), also known as Tetherin, is encoded by an ISG [346]. This protein is a type II transmembrane protein with a unique topology, possessing a N-terminal transmembrane region, an ectodomain and a C-terminal glycosylphosphatidylinositol (GPI) anchor [346]. In humans, BST-2 is located at the trans-Golgi network (TGN), in endosomes and within lipid rafts. Its name, Tetherin, was given due to its ability to restrict viral release, by retaining viruses at the cell membrane like an anchor [347,348]. However, the anti-IAV activity of Tetherin remains quite controversial. Indeed, some studies describe an inhibition of IAV (or of IAV-derived virus-like particles) egress, [349,350,351,352,353], while some others refute an inhibitory role [354,355,356,357]. Interestingly, not all IAV strains could be restricted by Tetherin, suggesting a strain-specificity that could, in part, explain the controversial results concerning its restricting activity [353,358]. The inhibitory activity of Tetherin was proposed to be counteracted by some viral proteins, like M2 [356] or NA [354,355,358], but a role of M2 was also excluded [350]. Finally, in vivo studies performed on Bst2-deficient mice revealed that they were not more susceptible to IAV infection than wild-type mice, which would be in favor of the absence of antiviral activity of Tetherin, at least against the strain used in this study [351]. Further studies, using other strains, would be required to definitely conclude the absence of impact of Tetherin on IAV replication in vivo.
Orthologs of BST-2 are present in a high number of species, including mammalian species, but also birds, alligators, turtles and some fish [359]. Even if there is often no amino acid sequence homology between the orthologs from different species, they share the same unique topology. Only a few of these orthologs were tested for their ability to restrict influenza viruses. Chicken encodes an ortholog of BST-2 (chBST-2) [360,361] and to date, chBST-2 antiviral activity was described and characterized for only one virus, the avian sarcoma and leukosis virus (ASLV) [361]. Further studies are required to determine if IAV is restricted by chBST-2.

4.3.2. Viperin: An IAV Inhibitor When Overexpressed

Virus inhibitory protein, endoplasmic reticulum-associated, IFN inducible (viperin) is an ISG that was shown to inhibit a wide range of viruses through diverse mechanisms [362,363,364,365]. The first evidence for viperin inhibiting IAV was through overexpression experiments that showed that the inhibition was by disrupting lipid rafts in vitro, therefore inhibiting viral release [366]. Further studies confirmed an antiviral phenotype through overexpression in vitro [354,367]. However, Viperin−/− mice did not show an enhanced viral load or lung damage, implying that viperin does not have an anti-IAV role in vivo [367].

4.3.3. PAI-1: An Inhibitory Protein Having an Effect in the Extracellular Media

Plasminogen activator inhibitor-1 (PAI-1) is part of the serine protease inhibitor (SERPIN) family and is coded by the SERPINE1 gene. The main role of this protein is to regulate the activation of plasminogen, the inactive precursor of plasmin, a serine protease involved in clearing of fibrin blood clots, by inhibiting both the urokinase-type plasminogen activator (u-PA) and the tissue-type plasminogen activator (t-PA) [368]. A surprising role for this protein in the inhibition of IAV maturation was recently discovered. Indeed, an image-based screen to identify novel ISGs regulating the late stages of IAV infection led to the identification of secreted PAI-1 as able to inhibit glycoprotein cleavage therefore reducing the infectivity of newly produced IAV particles [369]. Serpine1−/− mice presented a slightly increased susceptibility to IAV infection as well as a more severe disease phenotype. Interestingly, human genetic variability in SERPINE1 may also influence virus spread and disease severity [369]. PAI-1 is therefore the first ISG to be described as having an inhibitory role against IAV in the extracellular media.

4.4. Additional Factors Inhibiting IAV

Other proteins have been shown to inhibit IAV, but very little is currently known with respect to their potential mechanism of action. A few of these factors are listed in the following section.
Phospholipid scramblase 1 (PLSCR1) has been shown to play a role in the inhibition of a number of viruses [370,371]. For IAV, a single study has suggested that overexpressed PLSCR1 can inhibit IAV replication (by 2 log) and that KO increased it (by 1 log) [372]. Overexpressed PLSCR1 can form a complex with NP and members of the importin α family and its proposed mechanism of action would be the inhibition of NP entry into the nucleus [372].
The knock-out of SERTA domain containing 3 (SERTAD3), an IFN-inducible transcription factor, was shown to have a 3 to 4-fold positive impact on IAV replication in human cells, and, conversely, its overexpression negatively affected replication [373]. Sertad3−/− mice were also somewhat more susceptible to IAV infection [373]. The potential mechanism of inhibition of SERTAD3 was suggested to be the inhibition of the polymerase as overexpressed SERTAD3 interacted with PA, PB1 and PB2, preventing their assembly [373].
Eukaryotic translation elongation factor 1 delta (eEF1D) is a subunit of the eEF1 complex, which is involved in translation elongation and other non-canonical processes. eEF1D knock-out has been shown to improve IAV replication (by 1 log) and, conversely, its overexpression caused a decrease in replication (by 1 log) [374]. This has been linked to an interaction between eEF1D and vRNP components, leading to a decrease in nuclear import, rather than its effect on viral translation as could have been expected [374,375].
Finally, the desmosome component Plakophilin 2 (PKP2) was discovered through an IAV comparative interatomic study to be a potential PB1 interactor [156]. Overexpressed PKP2 decreased infection (by 3- to 4-fold) and siRNA experiments increased infection levels (by 2- to 3-fold) [156]. The suggested mechanism of action is through the disruption the IAV polymerase complex by competing with PB2 [156].

5. Conclusions and Perspectives

As we have seen throughout this review, host cells possess a wide array of restriction factors, belonging to different families of proteins, that can act as one of the first barriers of the intrinsic and innate immunity against influenza virus infection. These antiviral proteins target multiple steps of the viral life cycle, be it entry, replication, maturation, assembly, egress or even have an effect in the extracellular media once the virus has budded. Type I and III IFNs are important players in the innate immunity to control IAV replication in vivo in humans, avian species and mice. In other mammalian species such as pigs and bats, this importance has only been shown in vitro. The protection induced by type I/III IFNs is due to the expression of several hundreds of ISGs. Indeed, the vast majority of the antiviral proteins described in this review are induced by type I and III IFNs, showing the strong relationship between these factors and the IFN response. However, in the host–virus arms race and in order to overcome the antagonism of the IFN response, viruses have evolved proteins, such as NS1, PB1-F2 or PA-X, to counteract either the induction of the IFN response or to directly inhibit the restriction factors (e.g., the NS1-induced inhibition of PKR) and, thus, to promote viral replication.
Despite extensive studies and decades of investigation on some anti-IAV factors, such as MX1 or IFITM3, mechanistic details are still lacking to fully understand the restriction they impose on IAV. Moreover, the importance of most restriction factors in vivo or their role in the adaptive immunity are still unknown or debated for the vast majority. Given that some restriction factors play a role in both antiviral restriction and induction/regulation of immune responses, results obtained in KO mice might be challenging in their interpretation. Of note, the type I/III IFN response is important for the innate immunity but also crucially shapes adaptative immune responses. In line with this, recent studies have shown that some of these factors such as IFITM3 are versatile actors at the interplay between innate and adaptive immunity. Indeed, these antiviral factors seem to modulate dendritic cells and antiviral CD4+ and CD8+ T cell-responses. However, these mechanisms are still not fully understood. The characterization of the immunomodulatory roles of these antiviral proteins is essential to bridge the understanding between innate and adaptative immunity and is an expanding area of research that is essential to fully apprehend protective immune responses against IAV.
In this review, we also focused on the restriction factors present in other species of interest for IAV, namely pigs, bats and avian species. It is important to keep in mind that most of IAV strains circulating in humans arise from zoonotic introductions of avian-derived strains. As a consequence, after cross-species transmission and spill-over, avian-derived strains are not perfectly adapted to replicate in humans, where the viral polymerase activity is strongly impaired, missing crucial cofactors. There is also a lack of adaptation concerning the antagonism imposed by cellular restriction factors. Despite the poor genome annotation of poultry species, such as ducks, chickens and turkeys, or bats and pigs compared to that of humans, the vast majority of these proteins appear to be evolutionary conserved. Even if experimental data on these conserved proteins in all of these species are still lacking, recent studies pinpoint the conservation of the antiviral activity of some antiviral proteins such as MX1, IFITs or IFITM3. However, future studies are required to investigate whether each restriction factor described here are conserved in all these species of interest and whether their antiviral functions are similar to those of their mammalian counterparts.
With available treatments few and far between, focusing on understanding the molecular mechanisms of action of these known antiviral effectors is crucial to pave the way for the development of new therapeutic strategies against influenza. A particular focus should also be applied to the discovery of new antiviral inhibitors for the same reasons, taking advantage of state-of-the-art technologies, such as CRISPR/Cas9 screens, to lengthen the list of known host cell influenza inhibitors. Of note, results from large-scale siRNA and CRISPR screens tend to have little overlap, which is expected as screens can differ largely due to many variables such as cell types, viral strains and siRNA/sgRNA libraries used as well as multiple experimental variables which are harder to control. Therefore, it is of continued interest to perform these screens anew to discover as of yet unknown restriction factors.
The uncertainty as to when the next influenza pandemic will occur, with a vivid reminder being the current global situation due to the COVID-19 pandemic, should be initiative enough to push research forward in this field. We are not ready for the next influenza pandemic and the solution may well reside within our cells.

Author Contributions

Conceptualization, J.M. and A.R.; writing—original draft preparation, J.M. and A.R.; writing—review and editing, C.G., O.M., M.W., J.M. and A.R.; supervision, C.G. and O.M., funding acquisition, C.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded the Institut National de la Santé et de la Recherche Médicale (INSERM) (to C.G. and to M.W.), the Agence Nationale de la Recherche (ANR, reference ANR-20-CE15-0010; to C.G. and to M.W.), by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (ERC Starting Grant, grant agreement no. 759226) (to C.G.), the ATIP-Avenir programme (to C.G.) and institutional funds from the Centre National de la Recherche Scientifique (CNRS) and Montpellier University (to C.G.), 3-year PhD studentships from the Ministry of Higher Education and Research (to A.R. and to J.M.).

Acknowledgments

We would like to thank Jason Long for critical reading of parts of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Asha, K.; Kumar, B. Emerging Influenza D Virus Threat: What We Know so Far! J. Clin. Med. 2019, 8, 192. [Google Scholar] [CrossRef] [Green Version]
  2. Webster, R.G.; Bean, W.J.; Gorman, O.T.; Chambers, T.M.; Kawaoka, Y. Evolution and ecology of influenza A viruses. Microbiol. Rev. 1992, 56, 152–179. [Google Scholar] [CrossRef]
  3. Parrish, C.R.; Murcia, P.R.; Holmes, E.C. Influenza Virus Reservoirs and Intermediate Hosts: Dogs, Horses, and New Possibilities for Influenza Virus Exposure of Humans. J. Virol. 2015, 89, 2990–2994. [Google Scholar] [CrossRef] [Green Version]
  4. Tong, S.; Zhu, X.; Li, Y.; Shi, M.; Zhang, J.; Bourgeois, M.; Yang, H.; Chen, X.; Recuenco, S.; Gomez, J.; et al. New World Bats Harbor Diverse Influenza A Viruses. PLoS Pathog. 2013, 9, e1003657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Ciminski, K.; Schwemmle, M. Bat-Borne Influenza A Viruses: An Awakening. Cold Spring Harb. Perspect. Med. 2021, 11, a038612. [Google Scholar] [CrossRef] [Green Version]
  6. Krammer, F.; Smith, G.J.D.; Fouchier, R.A.M.; Peiris, M.; Kedzierska, K.; Doherty, P.C.; Palese, P.; Shaw, M.L.; Treanor, J.; Webster, R.G.; et al. Influenza. Nat. Rev. Dis. Prim. 2018, 4, 3. [Google Scholar] [CrossRef] [PubMed]
  7. Auladell, M.; Jia, X.; Hensen, L.; Chua, B.; Fox, A.; Nguyen, T.H.O.; Doherty, P.C.; Kedzierska, K. Recalling the Future: Immunological Memory Toward Unpredictable Influenza Viruses. Front. Immunol. 2019, 10, 1400. [Google Scholar] [CrossRef]
  8. Reid, A.H.; Taubenberger, J.K.; Fanning, T.G. Evidence of an absence: The genetic origins of the 1918 pandemic influenza virus. Nat. Rev. Genet. 2004, 2, 909–914. [Google Scholar] [CrossRef]
  9. Morens, D.M.; Taubenberger, J.K.; Fauci, A.S. The Persistent Legacy of the 1918 Influenza Virus. N. Engl. J. Med. 2009, 361, 225–229. [Google Scholar] [CrossRef] [Green Version]
  10. 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] [CrossRef] [Green Version]
  11. Nakajima, K.; Desselberger, U.; Palese, P. Recent human influenza A (H1N1) viruses are closely related genetically to strains isolated in 1950. Nat. Cell Biol. 1978, 274, 334–339. [Google Scholar] [CrossRef]
  12. Bloom, J.D.; Gong, L.I.; Baltimore, D. Permissive Secondary Mutations Enable the Evolution of Influenza Oseltamivir Resistance. Science 2010, 328, 1272–1275. [Google Scholar] [CrossRef] [Green Version]
  13. Bright, R.A.; Shay, D.K.; Shu, B.; Cox, N.J.; Klimov, A.I. Adamantane Resistance Among Influenza A Viruses Isolated Early During the 2005-2006 Influenza Season in the United States. JAMA 2006, 295, 891–894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Goldhill, D.H.; te Velthuis, A.J.W.; Fletcher, R.A.; Langat, P.; Zambon, M.; Lackenby, A.; Barclay, W.S. The mechanism of resistance to favipiravir in influenza. Proc. Natl. Acad. Sci. USA 2018, 115, 11613–11618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Omoto, S.; Speranzini, V.; Hashimoto, T.; Noshi, T.; Yamaguchi, H.; Kawai, M.; Kawaguchi, K.; Uehara, T.; Shishido, T.; Naito, A.; et al. Characterization of influenza virus variants induced by treatment with the endonuclease inhibitor baloxavir marboxil. Sci. Rep. 2018, 8, 9633. [Google Scholar] [CrossRef]
  16. Weis, W.; Brown, J.H.; Cusack, S.; Paulson, J.C.; Skehel, J.J.; Wiley, D.C. Structure of the influenza virus haemagglutinin complexed with its receptor, sialic acid. Nat. Cell Biol. 1988, 333, 426–431. [Google Scholar] [CrossRef] [PubMed]
  17. Gottschalk, A. On the Mechanism Underlying Initiation of Influenza Virus Infection. Ergeb. Mikrobiol. Immun. Exp. Ther. 1959, 32, 1–22. [Google Scholar] [CrossRef]
  18. Matrosovich, M.; Gambaryan, A.; Teneberg, S.; Piskarev, V.; Yamnikova, S.; Lvov, D.; Robertson, J.; Karlsson, K.-A. Avian Influenza A Viruses Differ from Human Viruses by Recognition of Sialyloligosaccharides and Gangliosides and by a Higher Conservation of the HA Receptor-Binding Site. Virology 1997, 233, 224–234. [Google Scholar] [CrossRef] [Green Version]
  19. Gamblin, S.J.; Skehel, J.J. Influenza Hemagglutinin and Neuraminidase Membrane Glycoproteins. J. Biol. Chem. 2010, 285, 28403–28409. [Google Scholar] [CrossRef] [Green Version]
  20. Rust, M.J.; Lakadamyali, M.; Zhang, F.; Zhuang, X. Assembly of endocytic machinery around individual influenza viruses during viral entry. Nat. Struct. Mol. Biol. 2004, 11, 567–573. [Google Scholar] [CrossRef] [Green Version]
  21. Chen, C.; Zhuang, X. Epsin 1 is a cargo-specific adaptor for the clathrin-mediated endocytosis of the influenza virus. Proc. Natl. Acad. Sci. USA 2008, 105, 11790–11795. [Google Scholar] [CrossRef] [Green Version]
  22. De Vries, E.; Tscherne, D.M.; Wienholts, M.J.; Cobos-Jiménez, V.; Scholte, F.; García-Sastre, A.; Rottier, P.J.M.; De Haan, C.A.M. Dissection of the Influenza A Virus Endocytic Routes Reveals Macropinocytosis as an Alternative Entry Pathway. PLoS Pathog. 2011, 7, e1001329. [Google Scholar] [CrossRef]
  23. Helenius, A. Unpacking the incoming influenza virus. Cell 1992, 69, 577–578. [Google Scholar] [CrossRef]
  24. Pinto, L.H.; Lamb, R.A. The M2 Proton Channels of Influenza A and B Viruses. J. Biol. Chem. 2006, 281, 8997–9000. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Li, S.; Sieben, C.; Ludwig, K.; Höfer, C.T.; Chiantia, S.; Herrmann, A.; Eghiaian, F.; Schaap, I.A. pH-Controlled Two-Step Uncoating of Influenza Virus. Biophys. J. 2014, 106, 1447–1456. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Skehel, J.J.; Wiley, D.C. Receptor Binding and Membrane Fusion in Virus Entry: The Influenza Hemagglutinin. Annu. Rev. Biochem. 2000, 69, 531–569. [Google Scholar] [CrossRef] [PubMed]
  27. Xu, R.; Wilson, I.A. Structural Characterization of an Early Fusion Intermediate of Influenza Virus Hemagglutinin. J. Virol. 2011, 85, 5172–5182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Cros, J.F.; García-Sastre, A.; Palese, P. An Unconventional NLS is Critical for the Nuclear Import of the Influenza A Virus Nucleoprotein and Ribonucleoprotein. Traffic 2005, 6, 205–213. [Google Scholar] [CrossRef] [PubMed]
  29. Eisfeld, A.J.; Neumann, G.; Kawaoka, Y. At the centre: Influenza A virus ribonucleoproteins. Nat. Rev. Genet. 2015, 13, 28–41. [Google Scholar] [CrossRef] [Green Version]
  30. 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]
  31. Dias, A.; Bouvier, D.; Crépin, T.; McCarthy, A.A.; Hart, D.J.; Baudin, F.; Cusack, S.; Ruigrok, R.W.H. The cap-snatching endonuclease of influenza virus polymerase resides in the PA subunit. Nat. Cell Biol. 2009, 458, 914–918. [Google Scholar] [CrossRef]
  32. Velthuis, A.J.W.T.; Fodor, A.J.W.T.V.E. Influenza virus RNA polymerase: Insights into the mechanisms of viral RNA synthesis. Nat. Rev. Genet. 2016, 14, 479–493. [Google Scholar] [CrossRef] [Green Version]
  33. Pflug, A.; Lukarska, M.; Resa-Infante, P.; Reich, S.; Cusack, S. Structural insights into RNA synthesis by the influenza virus transcription-replication machine. Virus Res. 2017, 234, 103–117. [Google Scholar] [CrossRef]
  34. Peacock, T.P.; Sheppard, C.M.; Staller, E.; Barclay, W.S. Host Determinants of Influenza RNA Synthesis. Annu. Rev. Virol. 2019, 6, 215–233. [Google Scholar] [CrossRef]
  35. Huang, S.; Chen, J.; Chen, Q.; Wang, H.; Yao, Y.; Chen, Z.; Chen, J.; Chen, J. A Second CRM1-Dependent Nuclear Export Signal in the Influenza A Virus NS2 Protein Contributes to the Nuclear Export of Viral Ribonucleoproteins. J. Virol. 2012, 87, 767–778. [Google Scholar] [CrossRef] [Green Version]
  36. Cros, J.F.; Palese, P. Trafficking of viral genomic RNA into and out of the nucleus: Influenza, Thogoto and Borna disease viruses. Virus Res. 2003, 95, 3–12. [Google Scholar] [CrossRef]
  37. Brunotte, L.; Flies, J.; Bolte, H.; Reuther, P.; Vreede, F.; Schwemmle, M. The Nuclear Export Protein of H5N1 Influenza A Viruses Recruits Matrix 1 (M1) Protein to the Viral Ribonucleoprotein to Mediate Nuclear Export. J. Biol. Chem. 2014, 289, 20067–20077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Bruce, E.A.; Digard, P.; Stuart, A.D. The Rab11 Pathway Is Required for Influenza A Virus Budding and Filament Formation. J. Virol. 2010, 84, 5848–5859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Martin, I.F.D.C.; Fournier, G.; Sachse, M.; Pizarro-Cerda, J.; Risco, C.; Naffakh, N. Influenza virus genome reaches the plasma membrane via a modified endoplasmic reticulum and Rab11-dependent vesicles. Nat. Commun. 2017, 8, 1–12. [Google Scholar] [CrossRef] [Green Version]
  40. Lakdawala, S.S.; Fodor, E.; Subbarao, K. Moving On Out: Transport and Packaging of Influenza Viral RNA into Virions. Annu. Rev. Virol. 2016, 3, 411–427. [Google Scholar] [CrossRef]
  41. Haralampiev, I.; Prisner, S.; Nitzan, M.; Schade, M.; Jolmes, F.; Schreiber, M.; Loidolt-Krüger, M.; Jongen, K.; Chamiolo, J.; Nilson, N.; et al. Selective flexible packaging pathways of the segmented genome of influenza A virus. Nat. Commun. 2020, 11, 1–13. [Google Scholar] [CrossRef]
  42. Malik, G.; Zhou, Y. Innate Immune Sensing of Influenza a Virus. Viruses 2020, 12, 755. [Google Scholar] [CrossRef]
  43. Lazear, H.M.; Schoggins, J.W.; Diamond, M.S. Shared and Distinct Functions of Type I and Type III Interferons. Immunity 2019, 50, 907–923. [Google Scholar] [CrossRef] [PubMed]
  44. Oshiumi, H. Recent Advances and Contradictions in the Study of the Individual Roles of Ubiquitin Ligases That Regulate RIG-I-Like Receptor-Mediated Antiviral Innate Immune Responses. Front. Immunol. 2020, 11, 1296. [Google Scholar] [CrossRef] [PubMed]
  45. Rehwinkel, J.; Tan, C.P.; Goubau, D.; Schulz, O.; Pichlmair, A.; Bier, K.; Robb, N.; Vreede, F.; Barclay, W.; Fodor, E.; et al. RIG-I Detects Viral Genomic RNA during Negative-Strand RNA Virus Infection. Cell 2010, 140, 397–408. [Google Scholar] [CrossRef] [Green Version]
  46. Kandasamy, M.; Suryawanshi, A.; Tundup, S.; Perez, J.T.; Schmolke, M.; Manicassamy, S.; Manicassamy, B. RIG-I Signaling Is Critical for Efficient Polyfunctional T Cell Responses during Influenza Virus Infection. PLoS Pathog. 2016, 12, e1005754. [Google Scholar] [CrossRef]
  47. Le Goffic, R.; Pothlichet, J.; Vitour, D.; Fujita, T.; Meurs, E.; Chignard, M.; Si-Tahar, M. Cutting Edge: Influenza A virus activates TLR3-dependent inflammatory and RIG-I-dependent antiviral responses in human lung epithelial cells. J. Immunol. 2007, 178, 3368–3372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Le Goffic, R.; Balloy, V.; Lagranderie, M.; Alexopoulou, L.; Escriou, N.; Flavell, R.; Chignard, M.; Si-Tahar, M. Detrimental Contribution of the Toll-Like Receptor (TLR)3 to Influenza A Virus–Induced Acute Pneumonia. PLoS Pathog. 2006, 2, e53. [Google Scholar] [CrossRef] [PubMed]
  49. Seo, S.-U.; Kwon, H.-J.; Song, J.-H.; Byun, Y.-H.; Seong, B.L.; Kawai, T.; Akira, S.; Kweon, M.-N. MyD88 Signaling Is Indispensable for Primary Influenza A Virus Infection but Dispensable for Secondary Infection. J. Virol. 2010, 84, 12713–12722. [Google Scholar] [CrossRef] [Green Version]
  50. Wolf, A.I.; Buehler, D.; Hensley, S.E.; Cavanagh, L.L.; Wherry, E.J.; Kastner, P.; Chan, S.; Weninger, W. Plasmacytoid Dendritic Cells Are Dispensable during Primary Influenza Virus Infection. J. Immunol. 2009, 182, 871–879. [Google Scholar] [CrossRef] [Green Version]
  51. Wack, A.; Terczyńska-Dyla, E.; Hartmann, R. Guarding the frontiers: The biology of type III interferons. Nat. Immunol. 2015, 16, 802–809. [Google Scholar] [CrossRef]
  52. Wu, W.; Metcalf, J.P. The Role of Type I IFNs in Influenza: Antiviral Superheroes or Immunopathogenic Villains? J. Innate Immun. 2020, 12, 437–447. [Google Scholar] [CrossRef]
  53. Zhang, Q.; Zeng, L.-P.; Zhou, P.; Irving, A.T.; Li, S.; Shi, Z.-L.; Wang, L.-F. IFNAR2-dependent gene expression profile induced by IFN-α in Pteropus alecto bat cells and impact of IFNAR2 knockout on virus infection. PLoS ONE 2017, 12, e0182866. [Google Scholar] [CrossRef] [Green Version]
  54. Shepardson, K.M.; Larson, K.; Johns, L.L.; Stanek, K.; Cho, H.; Wellham, J.; Henderson, H.; Rynda-Apple, A. IFNAR2 Is Required for Anti-influenza Immunity and Alters Susceptibility to Post-influenza Bacterial Superinfections. Front. Immunol. 2018, 9, 2589. [Google Scholar] [CrossRef]
  55. Seo, S.-U.; Kwon, H.-J.; Ko, H.-J.; Byun, Y.-H.; Seong, B.L.; Uematsu, S.; Akira, S.; Kweon, M.-N. Type I Interferon Signaling Regulates Ly6Chi Monocytes and Neutrophils during Acute Viral Pneumonia in Mice. PLoS Pathog. 2011, 7, e1001304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Koerner, I.; Kochs, G.; Kalinke, U.; Weiss, S.; Staeheli, P. Protective Role of Beta Interferon in Host Defense against Influenza A Virus. J. Virol. 2006, 81, 2025–2030. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. García-Sastre, A.; Durbin, R.K.; Zheng, H.; Palese, P.; Gertner, R.; Levy, D.E.; Durbin, J.E. The Role of Interferon in Influenza Virus Tissue Tropism. J. Virol. 1998, 72, 8550–8558. [Google Scholar] [CrossRef] [PubMed]
  58. Davidson, S.; Maini, M.K.; Wack, A. Disease-Promoting Effects of Type I Interferons in Viral, Bacterial, and Coinfections. J. Interf. Cytokine Res. 2015, 35, 252–264. [Google Scholar] [CrossRef] [PubMed]
  59. Klinkhammer, J.; Schnepf, D.; Ye, L.; Schwaderlapp, M.; Gad, H.H.; Hartmann, R.; Garcin, D.; Mahlakõiv, T.; Staeheli, P. IFN-λ prevents influenza virus spread from the upper airways to the lungs and limits virus transmission. eLife 2018, 7, e33354. [Google Scholar] [CrossRef] [PubMed]
  60. Mordstein, M.; Kochs, G.; Dumoutier, L.; Renauld, J.-C.; Paludan, S.R.; Klucher, K.; Staeheli, P. Interferon-λ Contributes to Innate Immunity of Mice against Influenza A Virus but Not against Hepatotropic Viruses. PLoS Pathog. 2008, 4, e1000151. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Ye, L.; Schnepf, D.; Staeheli, P. Interferon-λ orchestrates innate and adaptive mucosal immune responses. Nat. Rev. Immunol. 2019, 19, 614–625. [Google Scholar] [CrossRef] [PubMed]
  62. Galani, I.E.; Triantafyllia, V.; Eleminiadou, E.-E.; Koltsida, O.; Stavropoulos, A.; Manioudaki, M.; Thanos, D.; Doyle, S.E.; Kotenko, S.V.; Thanopoulou, K.; et al. Interferon-λ Mediates Non-redundant Front-Line Antiviral Protection against Influenza Virus Infection without Compromising Host Fitness. Immunity 2017, 46, 875–890.e6. [Google Scholar] [CrossRef] [PubMed]
  63. Bennett, A.L.; Smith, D.W.; Cummins, M.J.; Jacoby, P.A.; Cummins, J.M.; Beilharz, M.W. Low-dose oral interferon alpha as prophylaxis against viral respiratory illness: A double-blind, parallel controlled trial during an influenza pandemic year. Influenza Other Respir. Viruses 2013, 7, 854–862. [Google Scholar] [CrossRef]
  64. Lim, H.K.; Huang, S.X.; Chen, J.; Kerner, G.; Gilliaux, O.; Bastard, P.; Dobbs, K.; Hernandez, N.; Goudin, N.; Hasek, M.L.; et al. Severe influenza pneumonitis in children with inherited TLR3 deficiency. J. Exp. Med. 2019, 216, 2038–2056. [Google Scholar] [CrossRef] [PubMed]
  65. Ciancanelli, M.J.; Huang, S.X.L.; Luthra, P.; Garner, H.; Itan, Y.; Volpi, S.; Lafaille, F.G.; Trouillet, C.; Schmolke, M.; Albrecht, R.A.; et al. Life-threatening influenza and impaired interferon amplification in human IRF7 deficiency. Science 2015, 348, 448–453. [Google Scholar] [CrossRef] [Green Version]
  66. Hernandez, N.; Melki, I.; Jing, H.; Habib, T.; Huang, S.S.; Danielson, J.; Kula, T.; Drutman, S.; Belkaya, S.; Rattina, V.; et al. Life-threatening influenza pneumonitis in a child with inherited IRF9 deficiency. J. Exp. Med. 2018, 215, 2567–2585. [Google Scholar] [CrossRef] [PubMed]
  67. Hemann, E.A.; Green, R.; Turnbull, J.B.; Langlois, R.A.; Savan, R.; Gale, M. Interferon-λ modulates dendritic cells to facilitate T cell immunity during infection with influenza A virus. Nat. Immunol. 2019, 20, 1035–1045. [Google Scholar] [CrossRef]
  68. Ye, L.; Schnepf, D.; Becker, J.; Ebert, K.; Tanriver, Y.; Bernasconi, V.; Gad, H.H.; Hartmann, R.; Lycke, N.; Staeheli, P. Interferon-λ enhances adaptive mucosal immunity by boosting release of thymic stromal lymphopoietin. Nat. Immunol. 2019, 20, 593–601. [Google Scholar] [CrossRef]
  69. Kohlmeier, J.E.; Cookenham, T.; Roberts, A.D.; Miller, S.C.; Woodland, D.L. Type I Interferons Regulate Cytolytic Activity of Memory CD8+ T Cells in the Lung Airways during Respiratory Virus Challenge. Immunity 2010, 33, 96–105. [Google Scholar] [CrossRef] [Green Version]
  70. Durbin, J.E.; Fernandez-Sesma, A.; Lee, C.-K.; Rao, T.D.; Frey, A.B.; Moran, T.M.; Vukmanovic, S.; García-Sastre, A.; Levy, D.E. Type I IFN Modulates Innate and Specific Antiviral Immunity. J. Immunol. 2000, 164, 4220–4228. [Google Scholar] [CrossRef] [Green Version]
  71. Karpala, A.J.; Stewart, C.; McKay, J.; Lowenthal, J.W.; Bean, A.G.D. Characterization of Chicken Mda5 Activity: Regulation of IFN-β in the Absence of RIG-I Functionality. J. Immunol. 2011, 186, 5397–5405. [Google Scholar] [CrossRef]
  72. Barber, M.R.W.; Aldridge, J.R.; Webster, R.G.; Magor, K.E. Association of RIG-I with innate immunity of ducks to influenza. Proc. Natl. Acad. Sci. USA 2010, 107, 5913–5918. [Google Scholar] [CrossRef] [Green Version]
  73. Liniger, M.; Summerfield, A.; Zimmer, G.; McCullough, K.C.; Ruggli, N. Chicken Cells Sense Influenza A Virus Infection through MDA5 and CARDIF Signaling Involving LGP2. J. Virol. 2011, 86, 705–717. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Hayashi, T.; Watanabe, C.; Suzuki, Y.; Tanikawa, T.; Uchida, Y.; Saito, T. Chicken MDA5 Senses Short Double-Stranded RNA with Implications for Antiviral Response against Avian Influenza Viruses in Chicken. J. Innate Immun. 2013, 6, 58–71. [Google Scholar] [CrossRef] [PubMed]
  75. Uchikawa, E.; Lethier, M.; Malet, H.; Brunel, J.; Gerlier, D.; Cusack, S. Structural Analysis of dsRNA Binding to Anti-viral Pattern Recognition Receptors LGP2 and MDA5. Mol. Cell 2016, 62, 586–602. [Google Scholar] [CrossRef] [Green Version]
  76. Philbin, V.J.; Iqbal, M.; Boyd, Y.; Goodchild, M.J.; Beal, R.K.; Bumstead, N.; Young, J.; Smith, A.L. Identification and characterization of a functional, alternatively spliced Toll-like receptor 7 (TLR7) and genomic disruption of TLR8 in chickens. Immunology 2005, 114, 507–521. [Google Scholar] [CrossRef] [PubMed]
  77. Xia, C.; Liu, J.; Wu, Z.G.; Lin, C.Y.; Wang, M. The Interferon-α Genes from Three Chicken Lines and Its Effects on H9N2 Influenza Viruses. Anim. Biotechnol. 2004, 15, 77–88. [Google Scholar] [CrossRef]
  78. Penski, N.; Härtle, S.; Rubbenstroth, D.; Krohmann, C.; Ruggli, N.; Schusser, B.; Pfann, M.; Reuter, A.; Gohrbandt, S.; Hundt, J.; et al. Highly Pathogenic Avian Influenza Viruses Do Not Inhibit Interferon Synthesis in Infected Chickens but Can Override the Interferon-Induced Antiviral State. J. Virol. 2011, 85, 7730–7741. [Google Scholar] [CrossRef] [Green Version]
  79. Reuter, A.; Soubies, S.; Härtle, S.; Schusser, B.; Kaspers, B.; Staeheli, P.; Rubbenstroth, D.; Garcia-Sastre, A. Antiviral Activity of Lambda Interferon in Chickens. J. Virol. 2013, 88, 2835–2843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Hale, B.G.; Randall, R.E.; Ortín, J.; Jackson, D. The multifunctional NS1 protein of influenza A viruses. J. Gen. Virol. 2008, 89, 2359–2376. [Google Scholar] [CrossRef]
  81. Jureka, A.S.; Kleinpeter, A.B.; Tipper, J.L.; Harrod, K.S.; Petit, C.M. The influenza NS1 protein modulates RIG-I activation via a strain-specific direct interaction with the second CARD of RIG-I. J. Biol. Chem. 2020, 295, 1153–1164. [Google Scholar] [CrossRef]
  82. Mibayashi, M.; Martínez-Sobrido, L.; Loo, Y.-M.; Cárdenas, W.B.; Gale, M.; García-Sastre, A. Inhibition of Retinoic Acid-Inducible Gene I-Mediated Induction of Beta Interferon by the NS1 Protein of Influenza A Virus. J. Virol. 2006, 81, 514–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Guo, Z.; Chen, L.-M.; Zeng, H.; Gomez, J.A.; Plowden, J.; Fujita, T.; Katz, J.M.; Donis, R.O.; Sambhara, S. NS1 Protein of Influenza A Virus Inhibits the Function of Intracytoplasmic Pathogen Sensor, RIG-I. Am. J. Respir. Cell Mol. Biol. 2007, 36, 263–269. [Google Scholar] [CrossRef] [Green Version]
  84. Gack, M.U.; Albrecht, R.A.; Urano, T.; Inn, K.-S.; Huang, I.-C.; Carnero, E.; Farzan, M.; Inoue, S.; Jung, J.U.; García-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] [CrossRef] [PubMed] [Green Version]
  85. Rajsbaum, R.; Albrecht, R.A.; Wang, M.K.; Maharaj, N.P.; Versteeg, G.A.; Nistal-Villán, E.; García-Sastre, A.; Gack, M.U. Species-Specific Inhibition of RIG-I Ubiquitination and IFN Induction by the Influenza A Virus NS1 Protein. PLoS Pathog. 2012, 8, e1003059. [Google Scholar] [CrossRef] [PubMed]
  86. Koliopoulos, M.G.; Lethier, M.; Van Der Veen, A.G.; Haubrich, K.; Hennig, J.; Kowalinski, E.; Stevens, R.V.; Martin, S.R.; de Sousa, C.R.; Cusack, S.; et al. Molecular mechanism of influenza A NS1-mediated TRIM25 recognition and inhibition. Nat. Commun. 2018, 9, 1–13. [Google Scholar] [CrossRef]
  87. Mazur, I.; Anhlan, D.; Mitzner, D.; Wixler, L.; Schubert, U.; Ludwig, S. The proapoptotic influenza A virus protein PB1-F2 regulates viral polymerase activity by interaction with the PB1 protein. Cell. Microbiol. 2008, 10, 1140–1152. [Google Scholar] [CrossRef]
  88. Chen, W.; Calvo, P.A.; Malide, D.; Gibbs, J.; Schubert, U.; Bacik, I.; Basta, S.; O’Neill, R.; Schickli, J.; Palese, P.; et al. A novel influenza A virus mitochondrial protein that induces cell death. Nat. Med. 2001, 7, 1306–1312. [Google Scholar] [CrossRef] [PubMed]
  89. Leymarie, O.; Jouvion, G.; Hervé, P.-L.; Chevalier, C.; Lorin, V.; Lecardonnel, J.; Da Costa, B.; Delmas, B.; Escriou, N.; Le Goffic, R. Kinetic Characterization of PB1-F2-Mediated Immunopathology during Highly Pathogenic Avian H5N1 Influenza Virus Infection. PLoS ONE 2013, 8, e57894. [Google Scholar] [CrossRef] [PubMed]
  90. Varga, Z.T.; Ramos, I.; Hai, R.; Schmolke, M.; García-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]
  91. Varga, Z.T.; Grant, A.; Manicassamy, B.; Palese, P. Influenza Virus Protein PB1-F2 Inhibits the Induction of Type I Interferon by Binding to MAVS and Decreasing Mitochondrial Membrane Potential. J. Virol. 2012, 86, 8359–8366. [Google Scholar] [CrossRef] [Green Version]
  92. Firth, A.E.; Jagger, B.W.; Wise, H.M.; Nelson, C.C.; Parsawar, K.; Wills, N.M.; Napthine, S.; Taubenberger, J.K.; Digard, P.; Atkins, J.F. Ribosomal frameshifting used in influenza A virus expression occurs within the sequence UCC_UUU_CGU and is in the +1 direction. Open Biol. 2012, 2, 120109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. 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] [Green Version]
  94. Khaperskyy, D.A.; Schmaling, S.; Larkins-Ford, J.; McCormick, C.; Gaglia, M.M. Selective Degradation of Host RNA Polymerase II Transcripts by Influenza A Virus PA-X Host Shutoff Protein. PLoS Pathog. 2016, 12, e1005427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Hayashi, T.; Chaimayo, C.; McGuinness, J.; Takimoto, T. Critical Role of the PA-X C-Terminal Domain of Influenza A Virus in Its Subcellular Localization and Shutoff Activity. J. Virol. 2016, 90, 7131–7141. [Google Scholar] [CrossRef] [Green Version]
  96. Khaperskyy, D.A.; McCormick, C. Timing Is Everything: Coordinated Control of Host Shutoff by Influenza A Virus NS1 and PA-X Proteins. J. Virol. 2015, 89, 6528–6531. [Google Scholar] [CrossRef] [Green Version]
  97. Levene, R.E.; Gaglia, M.M. Host Shutoff in Influenza A Virus: Many Means to an End. Viruses 2018, 10, 475. [Google Scholar] [CrossRef] [Green Version]
  98. Bailey, C.C.; Zhong, G.; Huang, I.-C.; Farzan, M. IFITM-Family Proteins: The Cell’s First Line of Antiviral Defense. Annu. Rev. Virol. 2014, 1, 261–283. [Google Scholar] [CrossRef] [Green Version]
  99. Shi, G.; Schwartz, O.; Compton, A.A. More than meets the I: The diverse antiviral and cellular functions of interferon-induced transmembrane proteins. Retrovirology 2017, 14, 1–11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Zhang, Z.; Liu, J.; Li, M.; Yang, H.; Zhang, C. Evolutionary Dynamics of the Interferon-Induced Transmembrane Gene Family in Vertebrates. PLoS ONE 2012, 7, e49265. [Google Scholar] [CrossRef] [Green Version]
  101. Siegrist, F.; Ebeling, M.; Certa, U. The Small Interferon-Induced Transmembrane Genes and Proteins. J. Interf. Cytokine Res. 2011, 31, 183–197. [Google Scholar] [CrossRef]
  102. 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] [CrossRef] [Green Version]
  103. Zhao, X.; Li, J.; Winkler, C.A.; An, P.; Guo, J.-T. IFITM Genes, Variants, and Their Roles in the Control and Pathogenesis of Viral Infections. Front. Microbiol. 2019, 9, 3228. [Google Scholar] [CrossRef] [Green Version]
  104. Yount, J.S.; Karssemeijer, R.A.; Hang, H.C. S-Palmitoylation and Ubiquitination Differentially Regulate Interferon-induced Transmembrane Protein 3 (IFITM3)-mediated Resistance to Influenza Virus. J. Biol. Chem. 2012, 287, 19631–19641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Bailey, C.C.; Kondur, H.R.; Huang, I.-C.; Farzan, M. Interferon-induced Transmembrane Protein 3 Is a Type II Transmembrane Protein. J. Biol. Chem. 2013, 288, 32184–32193. [Google Scholar] [CrossRef] [Green Version]
  106. Weston, S.; Czieso, S.; White, I.J.; Smith, S.E.; Kellam, P.; Marsh, M. A Membrane Topology Model for Human Interferon Inducible Transmembrane Protein 1. PLoS ONE 2014, 9, e104341. [Google Scholar] [CrossRef]
  107. Ling, S.; Zhang, C.; Wang, W.; Cai, X.; Yu, L.; Wu, F.; Zhang, L.; Tian, C. Combined approaches of EPR and NMR illustrate only one transmembrane helix in the human IFITM3. Sci. Rep. 2016, 6, 24029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Chesarino, N.M.; McMichael, T.M.; Yount, J.S. Regulation of the trafficking and antiviral activity of IFITM3 by post-translational modifications. Future Microbiol. 2014, 9, 1151–1163. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Perreira, J.M.; Chin, C.R.; Feeley, E.M.; Brass, A.L. IFITMs Restrict the Replication of Multiple Pathogenic Viruses. J. Mol. Biol. 2013, 425, 4937–4955. [Google Scholar] [CrossRef] [PubMed]
  110. Peng, T.; Hang, H.C. Site-Specific Bioorthogonal Labeling for Fluorescence Imaging of Intracellular Proteins in Living Cells. J. Am. Chem. Soc. 2016, 138, 14423–14433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Spence, J.S.; He, R.; Hoffmann, H.-H.; Das, T.; Thinon, E.; Rice, C.M.; Peng, T.; Chandran, K.; Hang, H.C. IFITM3 directly engages and shuttles incoming virus particles to lysosomes. Nat. Chem. Biol. 2019, 15, 259–268. [Google Scholar] [CrossRef] [PubMed]
  112. Compton, A.A.; Bruel, T.; Porrot, F.; Mallet, A.; Sachse, M.; Euvrard, M.; Liang, C.; Casartelli, N.; Schwartz, O. IFITM Proteins Incorporated into HIV-1 Virions Impair Viral Fusion and Spread. Cell Host Microbe 2014, 16, 736–747. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Tartour, K.; Appourchaux, R.; Gaillard, J.; Nguyen, X.-N.; Durand, S.; Turpin, J.; Beaumont, E.; Roch, E.; Berger, G.; Mahieux, R.; et al. IFITM proteins are incorporated onto HIV-1 virion particles and negatively imprint their infectivity. Retrovirology 2014, 11, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Feeley, E.M.; Sims, J.S.; John, S.P.; Chin, C.R.; Pertel, T.; Chen, L.-M.; Gaiha, G.D.; Ryan, B.J.; Donis, R.O.; Elledge, S.J.; et al. IFITM3 Inhibits Influenza A Virus Infection by Preventing Cytosolic Entry. PLoS Pathog. 2011, 7, e1002337. [Google Scholar] [CrossRef] [Green Version]
  115. Li, K.; Markosyan, R.M.; Zheng, Y.-M.; Golfetto, O.; Bungart, B.; Li, M.; Ding, S.; Chen, L.; Liang, C.; Lee, J.C.; et al. IFITM Proteins Restrict Viral Membrane Hemifusion. PLoS Pathog. 2013, 9, e1003124. [Google Scholar] [CrossRef]
  116. Desai, T.M.; Marin, M.; Chin, C.R.; Savidis, G.; Brass, A.L.; Melikyan, G.B. IFITM3 Restricts Influenza A Virus Entry by Blocking the Formation of Fusion Pores following Virus-Endosome Hemifusion. PLoS Pathog. 2014, 10, e1004048. [Google Scholar] [CrossRef] [Green Version]
  117. Suddala, K.C.; Lee, C.C.; Meraner, P.; Marin, M.; Markosyan, R.M.; Desai, T.M.; Cohen, F.S.; Brass, A.L.; Melikyan, G.B. Interferon-induced transmembrane protein 3 blocks fusion of sensitive but not resistant viruses by partitioning into virus-carrying endosomes. PLoS Pathog. 2019, 15, e1007532. [Google Scholar] [CrossRef] [Green Version]
  118. John, S.P.; Chin, C.R.; Perreira, J.M.; Feeley, E.M.; Aker, A.M.; Savidis, G.; Smith, S.E.; Elia, A.E.H.; Everitt, A.R.; Vora, M.; et al. The CD225 Domain of IFITM3 Is Required for both IFITM Protein Association and Inhibition of Influenza A Virus and Dengue Virus Replication. J. Virol. 2013, 87, 7837–7852. [Google Scholar] [CrossRef] [Green Version]
  119. Jia, R.; Xu, F.; Qian, J.; Yao, Y.; Miao, C.; Zheng, Y.-M.; Liu, S.-L.; Guo, F.; Geng, Y.; Qiao, W.; et al. Identification of an endocytic signal essential for the antiviral action of IFITM3. Cell. Microbiol. 2014, 16, 1080–1093. [Google Scholar] [CrossRef]
  120. Wrensch, F.; Winkler, M.; Pöhlmann, S. IFITM Proteins Inhibit Entry Driven by the MERS-Coronavirus Spike Protein: Evidence for Cholesterol-Independent Mechanisms. Viruses 2014, 6, 3683–3698. [Google Scholar] [CrossRef] [Green Version]
  121. Yount, J.S.; Moltedo, B.; Yang, Y.-Y.; Charron, G.; Moran, T.M.; López, C.B.; Hang, H.C. Palmitoylome profiling reveals S-palmitoylation–dependent antiviral activity of IFITM3. Nat. Chem. Biol. 2010, 6, 610–614. [Google Scholar] [CrossRef]
  122. Chesarino, N.M.; McMichael, T.M.; Hach, J.C.; Yount, J.S. Phosphorylation of the Antiviral Protein Interferon-inducible Transmembrane Protein 3 (IFITM3) Dually Regulates Its Endocytosis and Ubiquitination. J. Biol. Chem. 2014, 289, 11986–11992. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Jia, R.; Pan, Q.; Ding, S.; Rong, L.; Liu, S.-L.; Geng, Y.; Qiao, W.; Liang, C. The N-Terminal Region of IFITM3 Modulates Its Antiviral Activity by Regulating IFITM3 Cellular Localization. J. Virol. 2012, 86, 13697–13707. [Google Scholar] [CrossRef] [Green Version]
  124. Everitt, A.R.; The GenISIS Investigators; Clare, S.; Pertel, T.; John, S.P.; Wash, R.S.; Smith, S.E.; Chin, C.R.; Feeley, E.M.; Sims, J.S.; et al. IFITM3 restricts the morbidity and mortality associated with influenza. Nat. Cell Biol. 2012, 484, 519–523. [Google Scholar] [CrossRef] [Green Version]
  125. Bailey, C.C.; Huang, I.-C.; Kam, C.; Farzan, M. Ifitm3 Limits the Severity of Acute Influenza in Mice. PLoS Pathog. 2012, 8, e1002909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Kenney, A.D.; McMichael, T.M.; Imas, A.; Chesarino, N.M.; Zhang, L.; Dorn, L.E.; Wu, Q.; Alfaour, O.; Amari, F.; Chen, M.; et al. IFITM3 protects the heart during influenza virus infection. Proc. Natl. Acad. Sci. USA 2019, 116, 18607–18612. [Google Scholar] [CrossRef] [Green Version]
  127. Infusini, G.; Smith, J.M.; Yuan, H.; Pizzolla, A.; Ng, W.C.; Londrigan, S.L.; Haque, A.; Reading, P.C.; Villadangos, J.A.; Wakim, L.M. Respiratory DC Use IFITM3 to Avoid Direct Viral Infection and Safeguard Virus-Specific CD8+ T Cell Priming. PLoS ONE 2015, 10, e0143539. [Google Scholar] [CrossRef] [PubMed]
  128. Wakim, L.M.; Gupta, N.; Mintern, J.D.; Villadangos, A.J. Enhanced survival of lung tissue-resident memory CD8+ T cells during infection with influenza virus due to selective expression of IFITM3. Nat. Immunol. 2013, 14, 238–245. [Google Scholar] [CrossRef]
  129. Wellington, D.; Laurenson-Schafer, H.; Abdel-Haq, A.; Dong, T. IFITM3: How genetics influence influenza infection demographically. Biomed. J. 2019, 42, 19–26. [Google Scholar] [CrossRef]
  130. Zhang, Y.-H.; Zhao, Y.; Liu, N.; Peng, Y.-C.; Giannoulatou, E.; Jin, R.-H.; Yan, H.-P.; Wu, H.; Liu, J.-H.; Wang, D.-Y.; et al. Interferon-induced transmembrane protein-3 genetic variant rs12252-C is associated with severe influenza in Chinese individuals. Nat. Commun. 2013, 4, 1418. [Google Scholar] [CrossRef] [Green Version]
  131. Mehrbod, P.; Eybpoosh, S.; Fotouhi, F.; Targhi, H.S.; Mazaheri, V.; Farahmand, B. Association of IFITM3 rs12252 polymorphisms, BMI, diabetes, and hypercholesterolemia with mild flu in an Iranian population. Virol. J. 2017, 14, 218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Pan, Y.; Yang, P.; Dong, T.; Zhang, Y.; Shi, W.; Peng, X.; Cui, S.; Zhang, D.; Lu, G.; Liu, Y.; et al. IFITM3 Rs12252-C Variant Increases Potential Risk for Severe Influenza Virus Infection in Chinese Population. Front. Cell Infect. Microbiol. 2017, 7, 294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Wang, Z.; Zhang, A.; Wan, Y.; Liu, X.; Qiu, C.; Xi, X.; Ren, Y.; Wang, J.; Dong, Y.; Bao, M.; et al. Early hypercytokinemia is associated with interferon-induced transmembrane protein-3 dysfunction and predictive of fatal H7N9 infection. Proc. Natl. Acad. Sci. USA 2014, 111, 769–774. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Yang, X.; Tan, B.; Zhou, X.; Xue, J.; Zhang, X.; Wang, P.; Shao, C.; Li, Y.; Li, C.; Xia, H.; et al. Interferon-Inducible Transmembrane Protein 3 Genetic Variant rs12252 and Influenza Susceptibility and Severity: A Meta-Analysis. PLoS ONE 2015, 10, e0124985. [Google Scholar] [CrossRef]
  135. Xuan, Y.; Wang, L.N.; Li, W.; Zi, H.R.; Guo, Y.; Yan, W.J.; Chen, X.B.; Wei, P.M. IFITM3 rs12252 T>C polymorphism is associated with the risk of severe influenza: A meta-analysis. Epidemiol. Infect. 2015, 143, 2975–2984. [Google Scholar] [CrossRef]
  136. Kim, Y.-C.; Jeong, B.-H. No Correlation of the Disease Severity of Influenza A Virus Infection with the rs12252 Polymorphism of the Interferon-Induced Transmembrane Protein 3 Gene. Intervirology 2017, 60, 69–74. [Google Scholar] [CrossRef]
  137. David, S.; Correia, V.; Antunes, L.; Faria, R.; Ferrão, J.; Faustino, P.; Nunes, B.; Maltez, F.; Lavinha, J.; De Andrade, H.R. Population genetics of IFITM3 in Portugal and Central Africa reveals a potential modifier of influenza severity. Immunogenetics 2018, 70, 169–177. [Google Scholar] [CrossRef]
  138. Martins, J.S.C.; Oliveira, M.L.A.; Garcia, C.C.; Siqueira, M.M.; Matos, A.R. Investigation of Human IFITM3 Polymorphisms rs34481144A and rs12252C and Risk for Influenza A(H1N1)pdm09 Severity in a Brazilian Cohort. Front. Cell. Infect. Microbiol. 2020, 10, 352. [Google Scholar] [CrossRef]
  139. Williams, D.E.J.; Wu, W.-L.; Grotefend, C.R.; Radic, V.; Chung, C.; Chung, Y.-H.; Farzan, M.; Huang, I.-C. IFITM3 Polymorphism rs12252-C Restricts Influenza A Viruses. PLoS ONE 2014, 9, e110096. [Google Scholar] [CrossRef] [Green Version]
  140. Randolph, A.G.; Yip, W.-K.; Allen, E.K.; Rosenberger, C.M.; Agan, A.A.; Ash, S.A.; Zhang, Y.; Bhangale, T.R.; Finkelstein, D.; Cvijanovich, N.Z.; et al. Evaluation of IFITM3 rs12252 Association With Severe Pediatric Influenza Infection. J. Infect. Dis. 2017, 216, 14–21. [Google Scholar] [CrossRef]
  141. Makvandi-Nejad, S.; Laurenson-Schafer, H.; Wang, L.; Wellington, D.; Zhao, Y.; Jin, B.; Qin, L.; Kite, K.; Moghadam, H.K.; Song, C.; et al. Lack of Truncated IFITM3 Transcripts in Cells Homozygous for the rs12252-C Variant That is Associated With Severe Influenza Infection. J. Infect. Dis. 2018, 217, 257–262. [Google Scholar] [CrossRef] [PubMed]
  142. Smith, J.; Smith, N.; Yu, L.; Paton, I.R.; Gutowska, M.W.; Forrest, H.L.; Danner, A.F.; Seiler, J.P.; Digard, P.; Webster, R.G.; et al. A comparative analysis of host responses to avian influenza infection in ducks and chickens highlights a role for the interferon-induced transmembrane proteins in viral resistance. BMC Genom. 2015, 16, 574. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Lanz, C.; Yángüez, E.; Andenmatten, D.; Stertz, S. Swine Interferon-Inducible Transmembrane Proteins Potently Inhibit Influenza A Virus Replication. J. Virol. 2014, 89, 863–869. [Google Scholar] [CrossRef] [Green Version]
  144. Benfield, C.T.O.; Smith, S.E.; Wright, E.; Wash, R.S.; Ferrara, F.; Temperton, N.J.; Kellam, P. Bat and pig IFN-induced transmembrane protein 3 restrict cell entry by influenza virus and lyssaviruses. J. Gen. Virol. 2015, 96, 991–1005. [Google Scholar] [CrossRef] [Green Version]
  145. Benfield, C.T.; MacKenzie, F.; Ritzefeld, M.; Mazzon, M.; Weston, S.; Tate, E.W.; Teo, B.H.; E Smith, S.; Kellam, P.; Holmes, E.C.; et al. Bat IFITM3 restriction depends on S-palmitoylation and a polymorphic site within the CD225 domain. Life Sci. Alliance 2020, 3, e201900542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Smith, S.E.; Gibson, M.S.; Wash, R.S.; Ferrara, F.; Wright, E.; Temperton, N.; Kellam, P.; Fife, M. Chicken Interferon-Inducible Transmembrane Protein 3 Restricts Influenza Viruses and LyssavirusesIn Vitro. J. Virol. 2013, 87, 12957–12966. [Google Scholar] [CrossRef] [Green Version]
  147. Blyth, G.A.D.; Chan, W.F.; Webster, R.G.; Magor, K.E. Duck Interferon-Inducible Transmembrane Protein 3 Mediates Restriction of Influenza Viruses. J. Virol. 2015, 90, 103–116. [Google Scholar] [CrossRef] [Green Version]
  148. Finelli, M.J.; Oliver, P.L. TLDc proteins: New players in the oxidative stress response and neurological disease. Mamm. Genome 2017, 28, 395–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Finelli, M.J.; Sanchez-Pulido, L.; Liu, K.X.; Davies, K.E.; Oliver, P.L. The Evolutionarily Conserved Tre2/Bub2/Cdc16 (TBC), Lysin Motif (LysM), Domain Catalytic (TLDc) Domain Is Neuroprotective against Oxidative Stress. J. Biol. Chem. 2016, 291, 2751–2763. [Google Scholar] [CrossRef] [Green Version]
  150. Shao, W.; Halachmi, S.; Brown, M. ERAP140, a Conserved Tissue-Specific Nuclear Receptor Coactivator. Mol. Cell. Biol. 2002, 22, 3358–3372. [Google Scholar] [CrossRef] [Green Version]
  151. Yu, L.; Croze, E.; Yamaguchi, K.D.; Tran, T.; Reder, A.T.; Litvak, V.; Volkert, M.R. Induction of a Unique Isoform of theNCOA7Oxidation Resistance Gene by Interferon β-1b. J. Interf. Cytokine Res. 2015, 35, 186–199. [Google Scholar] [CrossRef] [Green Version]
  152. Doyle, T.; Moncorgé, O.; Bonaventure, B.; Pollpeter, D.; Lussignol, M.; Tauziet, M.; Apolonia, L.; Catanese, M.-T.; Goujon, C.; Malim, M.H. The interferon-inducible isoform of NCOA7 inhibits endosome-mediated viral entry. Nat. Microbiol. 2018, 3, 1369–1376. [Google Scholar] [CrossRef]
  153. Merkulova, M.; Păunescu, T.G.; Azroyan, A.; Marshansky, V.; Breton, S.; Brown, D. Mapping the H+ (V)-ATPase interactome: Identification of proteins involved in trafficking, folding, assembly and phosphorylation. Sci. Rep. 2015, 5, 14827. [Google Scholar] [CrossRef] [PubMed]
  154. Castroflorio, E.; Hoed, J.D.; Svistunova, D.; Finelli, M.J.; Cebrian-Serrano, A.; Corrochano, S.; Bassett, A.R.; Davies, B.; Oliver, P.L. The Ncoa7 locus regulates V-ATPase formation and function, neurodevelopment and behaviour. Cell. Mol. Life Sci. 2020, 1–22. [Google Scholar] [CrossRef]
  155. Fu, B.; Wang, L.; Li, S.; Dorf, M.E. ZMPSTE24 defends against influenza and other pathogenic viruses. J. Exp. Med. 2017, 214, 919–929. [Google Scholar] [CrossRef] [PubMed]
  156. Wang, L.; Fu, B.; Li, W.; Patil, G.; Liu, L.; Dorf, M.E.; Li, S. Comparative influenza protein interactomes identify the role of plakophilin 2 in virus restriction. Nat. Commun. 2017, 8, 13876. [Google Scholar] [CrossRef]
  157. Zhao, J.; Chen, J.; Li, M.; Chen, M.; Sun, C. Multifaceted Functions of CH25H and 25HC to Modulate the Lipid Metabolism, Immune Responses, and Broadly Antiviral Activities. Viruses 2020, 12, 727. [Google Scholar] [CrossRef]
  158. Blanc, M.; Hsieh, W.Y.; Robertson, K.A.; Kropp, K.A.; Forster, T.; Shui, G.; Lacaze, P.; Watterson, S.; Griffiths, S.J.; Spann, N.J.; et al. The Transcription Factor STAT-1 Couples Macrophage Synthesis of 25-Hydroxycholesterol to the Interferon Antiviral Response. Immunity 2013, 38, 106–118. [Google Scholar] [CrossRef] [Green Version]
  159. Gold, E.S.; Diercks, A.H.; Podolsky, I.; Podyminogin, R.L.; Askovich, P.S.; Treuting, P.M.; Aderem, A. 25-Hydroxycholesterol acts as an amplifier of inflammatory signaling. Proc. Natl. Acad. Sci. USA 2014, 111, 10666–10671. [Google Scholar] [CrossRef] [Green Version]
  160. Heaton, B.E.; Kennedy, E.M.; Dumm, R.E.; Harding, A.T.; Sacco, M.T.; Sachs, D.; Heaton, N.S. A CRISPR Activation Screen Identifies a Pan-avian Influenza Virus Inhibitory Host Factor. Cell Rep. 2017, 20, 1503–1512. [Google Scholar] [CrossRef] [Green Version]
  161. Meerzaman, D.; Shapiro, P.S.; Kim, K.C. Involvement of the MAP kinase ERK2 in MUC1 mucin signaling. Am. J. Physiol. Cell. Mol. Physiol. 2001, 281, L86–L91. [Google Scholar] [CrossRef]
  162. Mukherjee, P.; Tinder, T.L.; Basu, G.D.; Gendler, S.J. MUC1 (CD227) interacts with lck tyrosine kinase in Jurkat lymphoma cells and normal T cells. J. Leukoc. Biol. 2004, 77, 90–99. [Google Scholar] [CrossRef] [PubMed]
  163. Rahn, J.J.; Shen, Q.; Mah, B.K.; Hugh, J.C. MUC1 Initiates a Calcium Signal after Ligation by Intercellular Adhesion Molecule-1. J. Biol. Chem. 2004, 279, 29386–29390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Wang, H.; Lillehoj, E.P.; Kim, K.C. MUC1 tyrosine phosphorylation activates the extracellular signal-regulated kinase. Biochem. Biophys. Res. Commun. 2004, 321, 448–454. [Google Scholar] [CrossRef] [PubMed]
  165. Lillehoj, E.P.; Kim, H.; Chun, E.Y.; Kim, K.C. Pseudomonas aeruginosastimulates phosphorylation of the airway epithelial membrane glycoprotein Muc1 and activates MAP kinase. Am. J. Physiol. Cell. Mol. Physiol. 2004, 287, L809–L815. [Google Scholar] [CrossRef] [Green Version]
  166. Rusinova, I.; Forster, S.; Yu, S.; Kannan, A.; Masse, M.; Cumming, H.; Chapman, R.; Hertzog, P.J. INTERFEROME v2.0: An updated database of annotated interferon-regulated genes. Nucleic Acids Res. 2012, 41, D1040–D1046. [Google Scholar] [CrossRef]
  167. Ng, G.Z.; Menheniott, T.R.; Every, A.L.; Stent, A.; Judd, L.M.; Chionh, Y.T.; Dhar, P.; Komen, J.C.; Giraud, A.S.; Wang, T.C.; et al. The MUC1 mucin protects againstHelicobacter pyloripathogenesis in mice by regulation of the NLRP3 inflammasome. Gut 2016, 65, 1087–1099. [Google Scholar] [CrossRef] [PubMed]
  168. McAuley, J.L.; Corcilius, L.; Tan, H.-X.; Payne, R.J.; A McGuckin, M.; E Brown, L. The cell surface mucin MUC1 limits the severity of influenza A virus infection. Mucosal Immunol. 2017, 10, 1581–1593. [Google Scholar] [CrossRef] [Green Version]
  169. Ramachandran, R.; Schmid, S.L. The dynamin superfamily. Curr. Biol. 2018, 28, R411–R416. [Google Scholar] [CrossRef] [Green Version]
  170. Haller, O.; Staeheli, P.; Schwemmle, M.; Kochs, G. Mx GTPases: Dynamin-like antiviral machines of innate immunity. Trends Microbiol. 2015, 23, 154–163. [Google Scholar] [CrossRef]
  171. Goujon, C.; Moncorgé, O.; Bauby, H.; Doyle, T.; Ward, C.C.; Schaller, T.; Hué, S.; Barclay, W.S.; Schulz, R.; Malim, M.H. Human MX2 is an interferon-induced post-entry inhibitor of HIV-1 infection. Nat. Cell Biol. 2013, 502, 559–562. [Google Scholar] [CrossRef]
  172. Kane, M.; Yadav, S.S.; Bitzegeio, J.; Kutluay, S.B.; Zang, T.; Wilson, S.J.; Schoggins, J.W.; Rice, C.M.; Yamashita, M.; Hatziioannou, T.; et al. MX2 is an interferon-induced inhibitor of HIV-1 infection. Nat. Cell Biol. 2013, 502, 563–566. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  173. Liu, Z.; Pan, Q.; Ding, S.; Qian, J.; Xu, F.; Zhou, J.; Cen, S.; Guo, F.; Liang, C. The Interferon-Inducible MxB Protein Inhibits HIV-1 Infection. Cell Host Microbe 2013, 14, 398–410. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Schilling, M.; Bulli, L.; Weigang, S.; Graf, L.; Naumann, S.; Patzina, C.; Wagner, V.; Bauersfeld, L.; Goujon, C.; Hengel, H.; et al. Human MxB Protein Is a Pan-herpesvirus Restriction Factor. J. Virol. 2018, 92, e01056-18. [Google Scholar] [CrossRef] [Green Version]
  175. Crameri, M.; Bauer, M.; Caduff, N.; Walker, R.; Steiner, F.; Franzoso, F.D.; Gujer, C.; Boucke, K.; Kucera, T.; Zbinden, A.; et al. MxB is an interferon-induced restriction factor of human herpesviruses. Nat. Commun. 2018, 9, 1–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Yi, D.-R.; An, N.; Liu, Z.-L.; Xu, F.-W.; Raniga, K.; Li, Q.-J.; Zhou, R.; Wang, J.; Zhang, Y.-X.; Zhou, J.-M.; et al. Human MxB Inhibits the Replication of Hepatitis C Virus. J. Virol. 2018, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Wang, Y.-X.; Niklasch, M.; Liu, T.; Wang, Y.; Shi, B.; Yuan, W.; Baumert, T.F.; Yuan, Z.; Tong, S.; Nassal, M.; et al. Interferon-inducible MX2 is a host restriction factor of hepatitis B virus replication. J. Hepatol. 2020, 72, 865–876. [Google Scholar] [CrossRef]
  178. Haller, O.; Arnheiter, H.; Pavlovic, J.; Staeheli, P. The Discovery of the Antiviral Resistance GeneMx: A Story of Great Ideas, Great Failures, and Some Success. Annu. Rev. Virol. 2018, 5, 33–51. [Google Scholar] [CrossRef]
  179. Kolb, E.; Laine, E.; Strehler, D.; Staeheli, P. Resistance to influenza virus infection of Mx transgenic mice expressing Mx protein under the control of two constitutive promoters. J. Virol. 1992, 66, 1709–1716. [Google Scholar] [CrossRef] [Green Version]
  180. Deeg, C.M.; Hassan, E.; Mutz, P.; Rheinemann, L.; Götz, V.; Magar, L.; Schilling, M.; Kallfass, C.; Nürnberger, C.; Soubies, S.; et al. In vivo evasion of MxA by avian influenza viruses requires human signature in the viral nucleoprotein. J. Exp. Med. 2017, 214, 1239–1248. [Google Scholar] [CrossRef]
  181. Holzinger, D.; Jorns, C.; Stertz, S.; Boisson-Dupuis, S.; Thimme, R.; Weidmann, M.; Casanova, J.-L.; Haller, O.; Kochs, G. Induction of MxA Gene Expression by Influenza A Virus Requires Type I or Type III Interferon Signaling. J. Virol. 2007, 81, 7776–7785. [Google Scholar] [CrossRef] [Green Version]
  182. Pitossi, F.; Blank, A.; Schröder, A.; Schwarz, A.; Hüssi, P.; Schwemmle, M.; Pavlovic, J.; Staeheli, P. A functional GTP-binding motif is necessary for antiviral activity of Mx proteins. J. Virol. 1993, 67, 6726–6732. [Google Scholar] [CrossRef] [Green Version]
  183. Ponten, A.; Sick, C.; Weeber, M.; Haller, O.; Kochs, G. Dominant-negative mutants of human MxA protein: Domains in the carboxy-terminal moiety are important for oligomerization and antiviral activity. J. Virol. 1997, 71, 2591–2599. [Google Scholar] [CrossRef] [Green Version]
  184. Gao, S.; Von Der Malsburg, A.; Paeschke, S.; Behlke, J.; Haller, O.; Kochs, G.; Daumke, O. Structural basis of oligomerization in the stalk region of dynamin-like MxA. Nat. Cell Biol. 2010, 465, 502–506. [Google Scholar] [CrossRef] [Green Version]
  185. 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] [CrossRef] [Green Version]
  186. Mitchell, P.S.; Patzina, C.; Emerman, M.; Haller, O.; Malik, H.S.; Kochs, G. Evolution-Guided Identification of Antiviral Specificity Determinants in the Broadly Acting Interferon-Induced Innate Immunity Factor MxA. Cell Host Microbe 2012, 12, 598–604. [Google Scholar] [CrossRef] [Green Version]
  187. Patzina, C.; Haller, O.; Kochs, G. Structural Requirements for the Antiviral Activity of the Human MxA Protein against Thogoto and Influenza A Virus. J. Biol. Chem. 2014, 289, 6020–6027. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Goujon, C.; Moncorgé, O.; Bauby, H.; Doyle, T.; Barclay, W.S.; Malim, M.H. Transfer of the Amino-Terminal Nuclear Envelope Targeting Domain of Human MX2 Converts MX1 into an HIV-1 Resistance Factor. J. Virol. 2014, 88, 9017–9026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Goujon, C.; Greenbury, R.A.; Papaioannou, S.; Doyle, T.; Malim, M.H. A Triple-Arginine Motif in the Amino-Terminal Domain and Oligomerization Are Required for HIV-1 Inhibition by Human MX2. J. Virol. 2015, 89, 4676–4680. [Google Scholar] [CrossRef] [Green Version]
  190. Dicks, M.D.J.; Goujon, C.; Pollpeter, D.; Betancor, G.; Apolonia, L.F.S.; Bergeron, J.R.C.; Malim, M.H. Oligomerization Requirements for MX2-Mediated Suppression of HIV-1 Infection. J. Virol. 2015, 90, 22–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Zimmermann, P.; Mänz, B.; Haller, O.; Schwemmle, M.; Kochs, G. The Viral Nucleoprotein Determines Mx Sensitivity of Influenza A Viruses. J. Virol. 2011, 85, 8133–8140. [Google Scholar] [CrossRef] [Green Version]
  192. Mänz, B.; Dornfeld, D.; Götz, H.; Zell, R.; Zimmermann, P.; Haller, O.; Kochs, G.; Schwemmle, M. Pandemic Influenza A Viruses Escape from Restriction by Human MxA through Adaptive Mutations in the Nucleoprotein. PLoS Pathog. 2013, 9, e1003279. [Google Scholar] [CrossRef]
  193. Riegger, D.; Hai, R.; Dornfeld, D.; Mänz, B.; Leyva-Grado, V.; Sánchez-Aparicio, M.T.; Albrecht, R.A.; Palese, P.; Haller, O.; Schwemmle, M.; et al. The Nucleoprotein of Newly Emerged H7N9 Influenza A Virus Harbors a Unique Motif Conferring Resistance to Antiviral Human MxA. J. Virol. 2014, 89, 2241–2252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Ashenberg, O.; Padmakumar, J.; Doud, M.B.; Bloom, J.D. Deep mutational scanning identifies sites in influenza nucleoprotein that affect viral inhibition by MxA. PLoS Pathog. 2017, 13, e1006288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Dittmann, J.; Stertz, S.; Grimm, D.; Steel, J.; García-Sastre, A.; Haller, O.; Kochs, G. Influenza A Virus Strains Differ in Sensitivity to the Antiviral Action of Mx-GTPase. J. Virol. 2008, 82, 3624–3631. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Verhelst, J.; Parthoens, E.; Schepens, B.; Fiers, W.; Saelens, X. Interferon-Inducible Protein Mx1 Inhibits Influenza Virus by Interfering with Functional Viral Ribonucleoprotein Complex Assembly. J. Virol. 2012, 86, 13445–13455. [Google Scholar] [CrossRef] [Green Version]
  197. Strandén, A.M.; Staeheli, P.; Pavlovic, J. Function of the Mouse Mx1 Protein Is Inhibited by Overexpression of the PB2 Protein of Influenza Virus. Virology 1993, 197, 642–651. [Google Scholar] [CrossRef]
  198. Gao, S.; Von Der Malsburg, A.; Dick, A.; Faelber, K.; Schröder, G.F.; Haller, O.; Kochs, G.; Daumke, O. Structure of Myxovirus Resistance Protein A Reveals Intra- and Intermolecular Domain Interactions Required for the Antiviral Function. Immunity 2011, 35, 514–525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Xiao, H.; Killip, M.J.; Staeheli, P.; Randall, R.E.; Jackson, D. The Human Interferon-Induced MxA Protein Inhibits Early Stages of Influenza A Virus Infection by Retaining the Incoming Viral Genome in the Cytoplasm. J. Virol. 2013, 87, 13053–13058. [Google Scholar] [CrossRef] [Green Version]
  200. Götz, V.; Magar, L.; Dornfeld, D.; Giese, S.; Pohlmann, A.; Höper, D.; Kong, B.-W.; Jans, D.A.; Beer, M.; Haller, O.; et al. Influenza A viruses escape from MxA restriction at the expense of efficient nuclear vRNP import. Sci. Rep. 2016, 6, 23138. [Google Scholar] [CrossRef] [Green Version]
  201. Pavlovic, J.; Haller, O.; Staeheli, P. Human and mouse Mx proteins inhibit different steps of the influenza virus multiplication cycle. J. Virol. 1992, 66, 2564–2569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Duc, T.T.T.; Farnir, F.; Michaux, C.; Desmecht, D.; Cornet, A. Detection of new biallelic polymorphisms in the human MxA gene. Mol. Biol. Rep. 2012, 39, 8533–8538. [Google Scholar] [CrossRef] [PubMed]
  203. Graf, L.; Dick, A.; Sendker, F.; Barth, E.; Marz, M.; Daumke, O.; Kochs, G. Effects of allelic variations in the human myxovirus resistance protein A on its antiviral activity. J. Biol. Chem. 2018, 293, 3056–3072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Tazi-Ahnini, R.; di Giovine, F.S.; McDonagh, A.J.; Messenger, A.G.; Amadou, C.; Cox, A.; Duff, G.W.; Cork, M.J. Structure and polymorphism of the human gene for the interferon-induced p78 protein (MX1): Evidence of association with alopecia areata in the Down syndrome region. Qual. Life Res. 2000, 106, 639–645. [Google Scholar] [CrossRef]
  205. Haller, O.; Kochs, G. Mx genes: Host determinants controlling influenza virus infection and trans-species transmission. Qual. Life Res. 2019, 139, 695–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Pavlovic, J.; Zürcher, T.; Haller, O.; Staeheli, P. Resistance to influenza virus and vesicular stomatitis virus conferred by expression of human MxA protein. J. Virol. 1990, 64, 3370–3375. [Google Scholar] [CrossRef] [Green Version]
  207. Verhelst, J.; Spitaels, J.; Nürnberger, C.; De Vlieger, D.; Ysenbaert, T.; Staeheli, P.; Fiers, W.; Saelens, X. Functional Comparison of Mx1 from Two Different Mouse Species Reveals the Involvement of Loop L4 in the Antiviral Activity against Influenza A Viruses. J. Virol. 2015, 89, 10879–10890. [Google Scholar] [CrossRef] [Green Version]
  208. Pavlovic, J.; Arzet, A.H.; Hefti, H.P.; Frese, M.; Rost, D.; Ernst, B.; Kolb, E.; Staeheli, P.; Haller, O. Enhanced virus resistance of transgenic mice expressing the human MxA protein. J. Virol. 1995, 69, 4506–4510. [Google Scholar] [CrossRef] [Green Version]
  209. Lee, S.; Ishitsuka, A.; Noguchi, M.; Hirohama, M.; Fujiyasu, Y.; Petric, P.P.; Schwemmle, M.; Staeheli, P.; Nagata, K.; Kawaguchi, A. Influenza restriction factor MxA functions as inflammasome sensor in the respiratory epithelium. Sci. Immunol. 2019, 4, eaau4643. [Google Scholar] [CrossRef]
  210. Palm, M.; Garigliany, M.-M.; Cornet, F.; Desmecht, D. Interferon-inducedSus scrofaMx1 blocks endocytic traffic of incoming influenza A virus particles. Vet. Res. 2010, 41, 29. [Google Scholar] [CrossRef] [Green Version]
  211. Meier, E.; Kunz, G.; Haller, O.; Arnheiter, H. Activity of rat Mx proteins against a rhabdovirus. J. Virol. 1990, 64, 6263–6269. [Google Scholar] [CrossRef] [Green Version]
  212. Fuchs, J.; Hölzer, M.; Schilling, M.; Patzina, C.; Schoen, A.; Hoenen, T.; Zimmer, G.; Marz, M.; Weber, F.; Müller, M.A.; et al. Evolution and Antiviral Specificities of Interferon-Induced Mx Proteins of Bats against Ebola, Influenza, and Other RNA Viruses. J. Virol. 2017, 91, e00361-17. [Google Scholar] [CrossRef] [Green Version]
  213. Benfield, C.T.O.; Lyall, J.W.; Kochs, G.; Tiley, L.S. Asparagine 631 Variants of the Chicken Mx Protein Do Not Inhibit Influenza Virus Replication in Primary Chicken Embryo Fibroblasts or In Vitro Surrogate Assays. J. Virol. 2008, 82, 7533–7539. [Google Scholar] [CrossRef] [Green Version]
  214. Bernasconi, D.; Schultz, U.; Staeheli, P. The Interferon-Induced Mx Protein of Chickens Lacks Antiviral Activity. J. Interf. Cytokine Res. 1995, 15, 47–53. [Google Scholar] [CrossRef] [PubMed]
  215. Schusser, B.; Reuter, A.; Von Der Malsburg, A.; Penski, N.; Weigend, S.; Kaspers, B.; Staeheli, P.; Härtle, S. Mx Is Dispensable for Interferon-Mediated Resistance of Chicken Cells against Influenza A Virus. J. Virol. 2011, 85, 8307–8315. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Daviet, S.; Van Borm, S.; Habyarimana, A.; Ahanda, M.-L.E.; Morin, V.; Oudin, A.; Berg, T.V.D.; Zoorob, R. Induction of Mx and PKR Failed to Protect Chickens from H5N1 Infection. Viral Immunol. 2009, 22, 467–472. [Google Scholar] [CrossRef] [PubMed]
  217. Bazzigher, L.; Schwarz, A.; Staeheli, P. No Enhanced Influenza Virus Resistance of Murine and Avian Cells Expressing Cloned Duck Mx Protein. Virology 1993, 195, 100–112. [Google Scholar] [CrossRef]
  218. Tretina, K.; Park, E.-S.; Maminska, A.; MacMicking, J.D. Interferon-induced guanylate-binding proteins: Guardians of host defense in health and disease. J. Exp. Med. 2019, 216, 482–500. [Google Scholar] [CrossRef] [Green Version]
  219. Ngo, C.C.; Man, S.M. Mechanisms and functions of guanylate-binding proteins and related interferon-inducible GTPases: Roles in intracellular lysis of pathogens. Cell. Microbiol. 2017, 19, e12791. [Google Scholar] [CrossRef] [Green Version]
  220. Nordmann, A.; Wixler, L.; Boergeling, Y.; Wixler, V.; Ludwig, S. A new splice variant of the human guanylate-binding protein 3 mediates anti-influenza activity through inhibition of viral transcription and replication. FASEB J. 2011, 26, 1290–1300. [Google Scholar] [CrossRef] [PubMed]
  221. Zhu, Z.; Shi, Z.; Yan, W.; Wei, J.; Shao, N.; Deng, X.; Wang, S.; Li, B.; Tong, G.; Ma, Z. Nonstructural Protein 1 of Influenza A Virus Interacts with Human Guanylate-Binding Protein 1 to Antagonize Antiviral Activity. PLoS ONE 2013, 8, e55920. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Feng, J.; Cao, Z.; Wang, L.; Wan, Y.; Peng, N.; Wang, Q.; Chen, X.; Zhou, Y.; Zhu, Y. Inducible GBP5 Mediates the Antiviral Response via Interferon-Related Pathways during Influenza A Virus Infection. J. Innate Immun. 2017, 9, 419–435. [Google Scholar] [CrossRef] [PubMed]
  223. Braun, E.; Hotter, D.; Koepke, L.; Zech, F.; Groß, R.; Sparrer, K.M.; Müller, J.A.; Pfaller, C.K.; Heusinger, E.; Wombacher, R.; et al. Guanylate-Binding Proteins 2 and 5 Exert Broad Antiviral Activity by Inhibiting Furin-Mediated Processing of Viral Envelope Proteins. Cell Rep. 2019, 27, 2092–2104.e10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Zhang, K.; Xu, W.W.; Zhang, Z.; Liu, J.; Sun, L.; Sun, W.; Jiao, P.; Sang, X.; Ren, Z.; Yu, Z.; et al. The innate immunity of guinea pigs against highly pathogenic avian influenza virus infection. Oncotarget 2017, 8, 30422–30437. [Google Scholar] [CrossRef] [PubMed]
  225. Feng, M.; Zhang, Q.; Wu, W.; Chen, L.; Gu, S.; Ye, Y.; Zhong, Y.; Huang, Q.; Liu, S. Inducible Guanylate-Binding Protein 7 Facilitates Influenza A Virus Replication by Suppressing Innate Immunity via NF-κB and JAK-STAT Signaling Pathways. J. Virol. 2021, 95. [Google Scholar] [CrossRef] [PubMed]
  226. King, J.K.; Yeh, S.; Lin, M.; Liu, C.; Lai, M.; Kao, J.; Chen, D.; Chen, P. Genetic polymorphisms in interferon pathway and response to interferon treatment in hepatitis B patients: A pilot study. Hepatology 2002, 36, 1416–1424. [Google Scholar] [CrossRef]
  227. Hatakeyama, S. TRIM Family Proteins: Roles in Autophagy, Immunity, and Carcinogenesis. Trends Biochem. Sci. 2017, 42, 297–311. [Google Scholar] [CrossRef]
  228. Di Pietro, A.; Kajaste-Rudnitski, A.; Oteiza, A.; Nicora, L.; Towers, G.J.; Mechti, N.; Vicenzi, E. TRIM22 Inhibits Influenza A Virus Infection by Targeting the Viral Nucleoprotein for Degradation. J. Virol. 2013, 87, 4523–4533. [Google Scholar] [CrossRef] [Green Version]
  229. Carthagena, L.; Bergamaschi, A.; Luna, J.M.; David, A.; Uchil, P.D.; Margottin-Goguet, F.; Mothes, W.; Hazan, U.; Transy, C.; Pancino, G.; et al. Human TRIM Gene Expression in Response to Interferons. PLoS ONE 2009, 4, e4894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  230. Pagani, I.; Di Pietro, A.; Oteiza, A.; Ghitti, M.; Mechti, N.; Naffakh, N.; Vicenzi, E. Mutations Conferring Increased Sensitivity to Tripartite Motif 22 Restriction Accumulated Progressively in the Nucleoprotein of Seasonal Influenza A (H1N1) Viruses between 1918 and 2009. mSphere 2018, 3, e00110-18. [Google Scholar] [CrossRef] [Green Version]
  231. Wu, X.; Wang, J.; Wang, S.; Wu, F.; Chen, Z.; Li, C.; Cheng, G.; Qin, F.X.-F. Inhibition of Influenza A Virus Replication by TRIM14 via Its Multifaceted Protein–Protein Interaction With NP. Front. Microbiol. 2019, 10, 344. [Google Scholar] [CrossRef] [Green Version]
  232. Patil, G.; Zhao, M.; Song, K.; Hao, W.; Bouchereau, D.; Wang, L.; Li, S. TRIM41-Mediated Ubiquitination of Nucleoprotein Limits Influenza A Virus Infection. J. Virol. 2018, 92, 15. [Google Scholar] [CrossRef] [Green Version]
  233. Fu, B.; Wang, L.; Ding, H.; Schwamborn, J.C.; Li, S.; Dorf, M.E. TRIM32 Senses and Restricts Influenza A Virus by Ubiquitination of PB1 Polymerase. PLoS Pathog. 2015, 11, e1004960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Kudryashova, E.; Wu, J.; Havton, L.A.; Spencer, M.J. Deficiency of the E3 ubiquitin ligase TRIM32 in mice leads to a myopathy with a neurogenic component. Hum. Mol. Genet. 2009, 18, 1353–1367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Nicklas, S.; Otto, A.; Wu, X.; Miller, P.; Stelzer, S.; Wen, Y.; Kuang, S.; Wrogemann, K.; Patel, K.; Ding, H.; et al. TRIM32 Regulates Skeletal Muscle Stem Cell Differentiation and Is Necessary for Normal Adult Muscle Regeneration. PLoS ONE 2012, 7, e30445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Sun, N.; Jiang, L.; Ye, M.; Wang, Y.; Wang, G.; Wan, X.; Zhao, Y.; Wen, X.; Liang, L.; Ma, S.; et al. TRIM35 mediates protection against influenza infection by activating TRAF3 and degrading viral PB2. Protein Cell 2020, 11, 894–914. [Google Scholar] [CrossRef]
  237. Choudhury, N.R.; Heikel, G.; Michlewski, G. TRIM25 and its emerging RNA-binding roles in antiviral defense. Wiley Interdiscip. Rev. RNA 2020, 11, e1588. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  238. Meyerson, N.R.; Zhou, L.; Guo, Y.R.; Zhao, C.; Tao, Y.J.; Krug, R.M.; Sawyer, S.L. Nuclear TRIM25 Specifically Targets Influenza Virus Ribonucleoproteins to Block the Onset of RNA Chain Elongation. Cell Host Microbe 2017, 22, 627–638.e7. [Google Scholar] [CrossRef] [Green Version]
  239. Liu, B.; Li, N.L.; Shen, Y.; Bao, X.; Fabrizio, T.; Elbahesh, H.; Webby, R.J.; Li, K. The C-Terminal Tail of TRIM56 Dictates Antiviral Restriction of Influenza A and B Viruses by Impeding Viral RNA Synthesis. J. Virol. 2016, 90, 4369–4382. [Google Scholar] [CrossRef] [Green Version]
  240. Seo, G.J.; Kim, C.; Shin, W.-J.; Sklan, E.H.; Eoh, H.; Jung, J.U. TRIM56-mediated monoubiquitination of cGAS for cytosolic DNA sensing. Nat. Commun. 2018, 9, 1–13. [Google Scholar] [CrossRef] [Green Version]
  241. Chemudupati, M.; Kenney, A.D.; Bonifati, S.; Zani, A.; McMichael, T.M.; Wu, L.; Yount, J.S. From APOBEC to ZAP: Diverse mechanisms used by cellular restriction factors to inhibit virus infections. Biochim. Biophys. Acta Bioenerg. 2019, 1866, 382–394. [Google Scholar] [CrossRef]
  242. Li, M.M.H.; Aguilar, E.G.; Michailidis, E.; Pabon, J.; Park, P.; Wu, X.; De Jong, Y.P.; Schneider, W.M.; Molina, H.; Rice, C.M.; et al. Characterization of Novel Splice Variants of Zinc Finger Antiviral Protein (ZAP). J. Virol. 2019, 93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  243. Meagher, J.L.; Takata, M.; Gonçalves-Carneiro, D.; Keane, S.C.; Rebendenne, A.; Ong, H.; Orr, V.K.; Macdonald, M.R.; Stuckey, J.A.; Bieniasz, P.D.; et al. Structure of the zinc-finger antiviral protein in complex with RNA reveals a mechanism for selective targeting of CG-rich viral sequences. Proc. Natl. Acad. Sci. USA 2019, 116, 24303–24309. [Google Scholar] [CrossRef]
  244. Takata, M.A.; Gonçalves-Carneiro, D.; Zang, T.M.; Soll, S.J.; York, A.; Blanco-Melo, D.; Bieniasz, P.D. CG dinucleotide suppression enables antiviral defence targeting non-self RNA. Nat. Cell Biol. 2017, 550, 124–127. [Google Scholar] [CrossRef] [PubMed]
  245. Liu, C.-H.; Zhou, L.; Chen, G.; Krug, R.M. Battle between influenza A virus and a newly identified antiviral activity of the PARP-containing ZAPL protein. Proc. Natl. Acad. Sci. USA 2015, 112, 14048–14053. [Google Scholar] [CrossRef] [Green Version]
  246. Tang, Q.; Wang, X.; Gao, G. The Short Form of the Zinc Finger Antiviral Protein Inhibits Influenza A Virus Protein Expression and Is Antagonized by the Virus-Encoded NS1. J. Virol. 2016, 91. [Google Scholar] [CrossRef] [Green Version]
  247. Lee, H.; Komano, J.; Saitoh, Y.; Yamaoka, S.; Kozaki, T.; Misawa, T.; Takahama, M.; Satoh, T.; Takeuchi, O.; Yamamoto, N.; et al. Zinc-finger antiviral protein mediates retinoic acid inducible gene I-like receptor-independent antiviral response to murine leukemia virus. Proc. Natl. Acad. Sci. USA 2013, 110, 12379–12384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Wang, X.; Li, M.M.H.; Zhao, J.; Li, S.; Macdonald, M.R.; Rice, C.M.; Gao, X.; Gao, G. Sindbis Virus Can Exploit a Host Antiviral Protein To Evade Immune Surveillance. J. Virol. 2016, 90, 10247–10258. [Google Scholar] [CrossRef] [Green Version]
  249. Kerns, J.A.; Emerman, M.; Malik, H.S. Positive Selection and Increased Antiviral Activity Associated with the PARP-Containing Isoform of Human Zinc-Finger Antiviral Protein. PLoS Genet. 2008, 4, e21. [Google Scholar] [CrossRef]
  250. Gonçalves-Carneiro, D.; Takata, M.A.; Ong, H.; Shilton, A.; Bieniasz, P.D. Origin and evolution of the Zinc Finger Antiviral Protein. Microbiology 2020. [Google Scholar] [CrossRef]
  251. Goossens, K.E.; Karpala, A.J.; Ward, A.; Bean, A.G. Characterisation of chicken ZAP. Dev. Comp. Immunol. 2014, 46, 373–381. [Google Scholar] [CrossRef]
  252. Lin, R.-J.; Huang, C.-H.; Liu, P.-C.; Lin, I.-C.; Huang, Y.-L.; Chen, A.-Y.; Chiu, H.-P.; Shih, S.-R.; Lin, L.-H.; Lien, S.-P.; et al. Zinc finger protein ZFP36L1 inhibits influenza A virus through translational repression by targeting HA, M and NS RNA transcripts. Nucleic Acids Res. 2020, 48, 7371–7384. [Google Scholar] [CrossRef] [PubMed]
  253. Son, Y.-O.; Kim, H.-E.; Choi, W.-S.; Chun, C.-H.; Chun, J.-S. RNA-binding protein ZFP36L1 regulates osteoarthritis by modulating members of the heat shock protein 70 family. Nat. Commun. 2019, 10, 77. [Google Scholar] [CrossRef]
  254. Diamond, M.S.; Farzan, M. The broad-spectrum antiviral functions of IFIT and IFITM proteins. Nat. Rev. Immunol. 2012, 13, 46–57. [Google Scholar] [CrossRef] [PubMed]
  255. Guo, J.; Peters, K.L.; Sen, G.C. Induction of the Human Protein P56 by Interferon, Double-Stranded RNA, or Virus Infection. Virology 2000, 267, 209–219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Pichlmair, A.; Lassnig, C.; Eberle, C.-A.; Górna, M.W.; Baumann, C.L.; Burkard, T.R.; Bürckstümmer, T.; Stefanovic, A.; Krieger, S.; Bennett, K.L.; et al. IFIT1 is an antiviral protein that recognizes 5′-triphosphate RNA. Nat. Immunol. 2011, 12, 624–630. [Google Scholar] [CrossRef]
  257. Habjan, M.; Hubel, P.; Lacerda, L.; Benda, C.; Holze, C.; Eberl, C.H.; Mann, A.; Kindler, E.; Gil-Cruz, C.; Ziebuhr, J.; et al. Sequestration by IFIT1 Impairs Translation of 2′O-unmethylated Capped RNA. PLoS Pathog. 2013, 9, e1003663. [Google Scholar] [CrossRef] [Green Version]
  258. Kimura, T.; Katoh, H.; Kayama, H.; Saiga, H.; Okuyama, M.; Okamoto, T.; Umemoto, E.; Matsuura, Y.; Yamamoto, M.; Takeda, K. Ifit1 Inhibits Japanese Encephalitis Virus Replication through Binding to 5′ Capped 2′-O Unmethylated RNA. J. Virol. 2013, 87, 9997–10003. [Google Scholar] [CrossRef] [Green Version]
  259. Terenzi, F.; Saikia, P.; Sen, G.C. Interferon-inducible protein, P56, inhibits HPV DNA replication by binding to the viral protein E1. EMBO J. 2008, 27, 3311–3321. [Google Scholar] [CrossRef] [Green Version]
  260. Mears, H.V.; Sweeney, T.R. Better together: The role of IFIT protein–protein interactions in the antiviral response. J. Gen. Virol. 2018, 99, 1463–1477. [Google Scholar] [CrossRef]
  261. Pinto, A.K.; Williams, G.D.; Szretter, K.J.; White, J.P.; Proença-Módena, J.L.; Liu, G.; Olejnik, J.; Brien, J.D.; Ebihara, H.; Mühlberger, E.; et al. Human and Murine IFIT1 Proteins Do Not Restrict Infection of Negative-Sense RNA Viruses of the Orthomyxoviridae, Bunyaviridae, and Filoviridae Families. J. Virol. 2015, 89, 9465–9476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Tran, V.; Ledwith, M.P.; Thamamongood, T.; Higgins, C.A.; Tripathi, S.; Chang, M.W.; Benner, C.; García-Sastre, A.; Schwemmle, M.; Boon, A.C.M.; et al. Influenza virus repurposes the antiviral protein IFIT2 to promote translation of viral mRNAs. Nat. Microbiol. 2020, 5, 1490–1503. [Google Scholar] [CrossRef]
  263. Rong, E.; Hu, J.; Yang, C.; Chen, H.; Wang, Z.; Liu, X.; Liu, W.; Lu, C.; He, P.; Wang, X.; et al. Broad-spectrum antiviral functions of duck interferon-induced protein with tetratricopeptide repeats (AvIFIT). Dev. Comp. Immunol. 2018, 84, 71–81. [Google Scholar] [CrossRef]
  264. Santhakumar, D.; Rohaim, M.A.M.S.; Hussein, H.A.; Hawes, P.; Ferreira, H.L.; Behboudi, S.; Iqbal, M.; Nair, V.; Arns, C.W.; Munir, M. Chicken Interferon-induced Protein with Tetratricopeptide Repeats 5 Antagonizes Replication of RNA Viruses. Sci. Rep. 2018, 8, 1–20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Rohaim, M.A.; Santhakumar, D.; El Naggar, R.F.; Iqbal, M.; Hussein, H.A.; Munir, M. Chickens Expressing IFIT5 Ameliorate Clinical Outcome and Pathology of Highly Pathogenic Avian Influenza and Velogenic Newcastle Disease Viruses. Front. Immunol. 2018, 9. [Google Scholar] [CrossRef]
  266. Marié, I.; Svab, J.; Robert, N.; Galabru, J.; Hovanessian, A.G. Differential expression and distinct structure of 69- and 100-kDa forms of 2-5A synthetase in human cells treated with interferon. J. Biol. Chem. 1990, 265, 18601–18607. [Google Scholar] [CrossRef]
  267. Silverman, R.H.; Weiss, S.R. Viral Phosphodiesterases That Antagonize Double-Stranded RNA Signaling to RNase L by Degrading 2-5A. J. Interf. Cytokine Res. 2014, 34, 455–463. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Malathi, K.; Saito, T.; Crochet, N.; Barton, D.J.; Gale, M.; Silverman, R.H. RNase L releases a small RNA from HCV RNA that refolds into a potent PAMP. RNA 2010, 16, 2108–2119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  269. Malathi, K.; Dong, B.; Gale, M.; Silverman, R.H. Small self-RNA generated by RNase L amplifies antiviral innate immunity. Nature 2007, 448, 816–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Zhu, J.; Zhang, Y.; Ghosh, A.; Cuevas, R.A.; Forero, A.; Dhar, J.; Ibsen, M.S.; Schmid-Burgk, J.L.; Schmidt, T.; Ganapathiraju, M.K.; et al. Antiviral Activity of Human OASL Protein Is Mediated by Enhancing Signaling of the RIG-I RNA Sensor. Immunity 2014, 40, 936–948. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Drappier, M.; Michiels, T. Inhibition of the OAS/RNase L pathway by viruses. Curr. Opin. Virol. 2015, 15, 19–26. [Google Scholar] [CrossRef] [PubMed]
  272. Min, J.-Y.; Krug, R.M. The primary function of RNA binding by the influenza A virus NS1 protein in infected cells: Inhibiting the 2′-5′ oligo (A) synthetase/RNase L pathway. Proc. Natl. Acad. Sci. USA 2006, 103, 7100–7105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Li, Y.; Banerjee, S.; Wang, Y.; Goldstein, S.A.; Dong, B.; Gaughan, C.; Silverman, R.H.; Weiss, S.R. Activation of RNase L is dependent on OAS3 expression during infection with diverse human viruses. Proc. Natl. Acad. Sci. USA 2016, 113, 2241–2246. [Google Scholar] [CrossRef] [Green Version]
  274. Zhou, A.; Paranjape, J.; Brown, T.L.; Nie, H.; Naik, S.; Dong, B.; Chang, A.; Trapp, B.; Fairchild, R.; Colmenares, C.; et al. Interferon action and apoptosis are defective in mice devoid of 2′,5′-oligoadenylate-dependent RNase L. EMBO J. 1997, 16, 6355–6363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  275. Meurs, E.; Chong, K.; Galabru, J.; Thomas, N.B.; Kerr, I.M.; Williams, B.R.; Hovanessian, A.G. Molecular cloning and characterization of the human double-stranded RNA-activated protein kinase induced by interferon. Cell 1990, 62, 379–390. [Google Scholar] [CrossRef]
  276. Dar, A.C.; Dever, T.E.; Sicheri, F. Higher-Order Substrate Recognition of eIF2α by the RNA-Dependent Protein Kinase PKR. Cell 2005, 122, 887–900. [Google Scholar] [CrossRef] [Green Version]
  277. Dauber, B.; Wolff, T. Activation of the Antiviral Kinase PKR and Viral Countermeasures. Viruses 2009, 1, 523–544. [Google Scholar] [CrossRef] [Green Version]
  278. 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. 2006, 81, 2221–2230. [Google Scholar] [CrossRef] [Green Version]
  279. Lu, Y.; Wambach, M.; Katze, M.G.; Krug, R.M. Binding of the Influenza Virus NS1 Protein to Double-Stranded RNA Inhibits the Activation of the Protein Kinase That Phosphorylates the eIF-2 Translation Initiation Factor. Virology 1995, 214, 222–228. [Google Scholar] [CrossRef] [Green Version]
  280. Bergmann, M.; Garcia-Sastre, A.; Carnero, E.; Pehamberger, H.; Wolff, K.; Palese, P.; Muster, T. Influenza Virus NS1 Protein Counteracts PKR-Mediated Inhibition of Replication. J. Virol. 2000, 74, 6203–6206. [Google Scholar] [CrossRef] [Green Version]
  281. Hatada, E.; Saito, S.; Fukuda, R. Mutant Influenza Viruses with a Defective NS1 Protein Cannot Block the Activation of PKR in Infected Cells. J. Virol. 1999, 73, 2425–2433. [Google Scholar] [CrossRef] [Green Version]
  282. Schierhorn, K.L.; Jolmes, F.; Bespalowa, J.; Saenger, S.; Peteranderl, C.; Dzieciolowski, J.; Mielke, M.; Budt, M.; Pleschka, S.; Herrmann, A.; et al. Influenza A Virus Virulence Depends on Two Amino Acids in the N-Terminal Domain of Its NS1 Protein To Facilitate Inhibition of the RNA-Dependent Protein Kinase PKR. J. Virol. 2017, 91, e00198-17. [Google Scholar] [CrossRef] [Green Version]
  283. Balachandran, S.; Roberts, P.C.; Brown, E.L.; Truong, H.; Pattnaik, A.K.; Archer, D.R.; Barber, G.N. Essential Role for the dsRNA-Dependent Protein Kinase PKR in Innate Immunity to Viral Infection. Immunity 2000, 13, 129–141. [Google Scholar] [CrossRef] [Green Version]
  284. Fleming-Canepa, X.; Aldridge, J.R.; Canniff, L.; Kobewka, M.; Jax, E.; Webster, R.G.; Magor, K.E. Duck innate immune responses to high and low pathogenicity H5 avian influenza viruses. Vet. Microbiol. 2019, 228, 101–111. [Google Scholar] [CrossRef]
  285. Kao, P.N.; Chen, L.; Brock, G.; Ng, J.; Kenny, J.; Smith, A.J.; Corthesy, B. Cloning and expression of cyclosporin A- and FK506-sensitive nuclear factor of activated T-cells: NF45 and NF90. J. Biol. Chem. 1994, 269, 20691–20699. [Google Scholar] [CrossRef]
  286. Saunders, L.R.; Perkins, D.J.; Balachandran, S.; Michaels, R.; Ford, R.; Mayeda, A.; Barber, G.N. Characterization of Two Evolutionarily Conserved, Alternatively Spliced Nuclear Phosphoproteins, NFAR-1 and -2, That Function in mRNA Processing and Interact with the Double-stranded RNA-dependent Protein Kinase, PKR. J. Biol. Chem. 2001, 276, 32300–32312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  287. Patiño, C.; Haenni, A.-L.; Urcuqui-Inchima, S. NF90 isoforms, a new family of cellular proteins involved in viral replication? Biochimie 2015, 108, 20–24. [Google Scholar] [CrossRef]
  288. Li, T.; Li, X.; Zhu, W.; Wang, H.; Mei, L.; Wu, S.; Lin, X.; Han, X. NF90 is a novel influenza A virus NS1-interacting protein that antagonizes the inhibitory role of NS1 on PKR phosphorylation. FEBS Lett. 2016, 590, 2797–2810. [Google Scholar] [CrossRef]
  289. Wen, X.; Huang, X.; Mok, B.W.-Y.; Chen, Y.; Zheng, M.; Lau, S.-Y.; Wang, P.; Song, W.; Jin, D.-Y.; Yuen, K.-Y.; et al. NF90 Exerts Antiviral Activity through Regulation of PKR Phosphorylation and Stress Granules in Infected Cells. J. Immunol. 2014, 192, 3753–3764. [Google Scholar] [CrossRef] [PubMed]
  290. Bortz, E.; Westera, L.; Maamary, J.; Steel, J.; Albrecht, R.A.; Manicassamy, B.; Chase, G.; Martínez-Sobrido, L.; Schwemmle, M.; García-Sastre, A. Host- and Strain-Specific Regulation of Influenza Virus Polymerase Activity by Interacting Cellular Proteins. mBio 2011, 2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Wang, P.; Song, W.; Mok, B.W.-Y.; Zhao, P.; Qin, K.; Lai, A.; Smith, G.J.D.; Zhang, J.; Lin, T.; Guan, Y.; et al. Nuclear Factor 90 Negatively Regulates Influenza Virus Replication by Interacting with Viral Nucleoprotein. J. Virol. 2009, 83, 7850–7861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  292. Shi, L.; Godfrey, W.R.; Lin, J.; Zhao, G.; Kao, P.N. NF90 regulates inducible IL-2 gene expression in T cells. J. Exp. Med. 2007, 204, 971–977. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  293. Kwong, A.D.; Rao, B.G.; Jeang, K.-T. Viral and cellular RNA helicases as antiviral targets. Nat. Rev. Drug Discov. 2005, 4, 845–853. [Google Scholar] [CrossRef] [PubMed]
  294. Ranji, A.; Boris-Lawrie, K. RNA helicases. RNA Biol. 2010, 7, 775–787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  295. Gregersen, L.H.; Schueler, M.; Munschauer, M.; Mastrobuoni, G.; Chen, W.; Kempa, S.; Dieterich, C.; Landthaler, M. MOV10 Is a 5′ to 3′ RNA Helicase Contributing to UPF1 mRNA Target Degradation by Translocation along 3′ UTRs. Mol. Cell 2014, 54, 573–585. [Google Scholar] [CrossRef] [Green Version]
  296. Kenny, P.J.; Zhou, H.; Kim, M.; Skariah, G.; Khetani, R.S.; Drnevich, J.; Arcila, M.L.; Kosik, K.S.; Ceman, S. MOV10 and FMRP Regulate AGO2 Association with MicroRNA Recognition Elements. Cell Rep. 2014, 9, 1729–1741. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Goodier, J.L.; Cheung, L.E.; Kazazian, H.H., Jr. MOV10 RNA Helicase Is a Potent Inhibitor of Retrotransposition in Cells. PLoS Genet. 2012, 8, e1002941. [Google Scholar] [CrossRef] [Green Version]
  298. Zheng, Y.-H.; Jeang, K.-T.; Tokunaga, K. Host restriction factors in retroviral infection: Promises in virus-host interaction. Retrovirology 2012, 9, 112. [Google Scholar] [CrossRef] [Green Version]
  299. Schoggins, J.W.; Wilson, S.J.; Panis, M.; Murphy, M.Y.; Jones, C.T.; Bieniasz, P.D.; Rice, C.M. A diverse range of gene products are effectors of the type I interferon antiviral response. Nat. Cell Biol. 2011, 472, 481–485. [Google Scholar] [CrossRef]
  300. Cuevas, R.A.; Ghosh, A.; Wallerath, C.; Hornung, V.; Coyne, C.B.; Sarkar, S.N. MOV10 Provides Antiviral Activity against RNA Viruses by Enhancing RIG-I-MAVS-Independent IFN Induction. J. Immunol. 2016, 196, 3877–3886. [Google Scholar] [CrossRef] [Green Version]
  301. Zhang, J.; Huang, F.; Tan, L.; Bai, C.; Chen, B.; Liu, J.; Liang, J.; Liu, C.; Zhang, S.; Lu, G.; et al. Host Protein Moloney Leukemia Virus 10 (MOV10) Acts as a Restriction Factor of Influenza A Virus by Inhibiting the Nuclear Import of the Viral Nucleoprotein. J. Virol. 2016, 90, 3966–3980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Li, J.; Hu, S.; Xu, F.; Mei, S.; Liu, X.; Yin, L.; Zhao, F.; Zhao, X.; Sun, H.; Xiong, Z.; et al. MOV10 sequesters the RNP of influenza A virus in the cytoplasm and is antagonized by viral NS1 protein. Biochem. J. 2019, 476, 467–481. [Google Scholar] [CrossRef]
  303. Skariah, G.; Seimetz, J.; Norsworthy, M.; Lannom, M.C.; Kenny, P.J.; Elrakhawy, M.; Forsthoefel, C.; Drnevich, J.; Kalsotra, A.; Ceman, S. Mov10 suppresses retroelements and regulates neuronal development and function in the developing brain. BMC Biol. 2017, 15, 54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  304. Linder, P.; Jankowsky, E. From unwinding to clamping—The DEAD box RNA helicase family. Nat. Rev. Mol. Cell Biol. 2011, 12, 505–516. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  305. Song, C.; Hotz-Wagenblatt, A.; Voit, R.; Grummt, I. SIRT7 and the DEAD-box helicase DDX21 cooperate to resolve genomic R loops and safeguard genome stability. Genes Dev. 2017, 31, 1370–1381. [Google Scholar] [CrossRef]
  306. Taschuk, F.; Cherry, S. DEAD-Box Helicases: Sensors, Regulators, and Effectors for Antiviral Defense. Viruses 2020, 12, 181. [Google Scholar] [CrossRef] [Green Version]
  307. Chen, G.; Liu, C.-H.; Zhou, L.; Krug, R.M. Cellular DDX21 RNA Helicase Inhibits Influenza A Virus Replication but Is Counteracted by the Viral NS1 Protein. Cell Host Microbe 2014, 15, 484–493. [Google Scholar] [CrossRef] [Green Version]
  308. Haberland, M.; Montgomery, R.L.; Olson, E.N. The many roles of histone deacetylases in development and physiology: Implications for disease and therapy. Nat. Rev. Genet. 2009, 10, 32–42. [Google Scholar] [CrossRef] [PubMed]
  309. Yamauchi, Y.; Boukari, H.; Banerjee, I.; Sbalzarini, I.F.; Horvath, P.; Helenius, A. Histone Deacetylase 8 Is Required for Centrosome Cohesion and Influenza A Virus Entry. PLoS Pathog. 2011, 7, e1002316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  310. Nagesh, P.T.; Husain, M. Influenza A Virus Dysregulates Host Histone Deacetylase 1 That Inhibits Viral Infection in Lung Epithelial Cells. J. Virol. 2016, 90, 4614–4625. [Google Scholar] [CrossRef] [Green Version]
  311. Nagesh, P.T.; Hussain, M.; Galvin, H.D.; Husain, M. Histone Deacetylase 2 Is a Component of Influenza A Virus-Induced Host Antiviral Response. Front. Microbiol. 2017, 8, 1315. [Google Scholar] [CrossRef]
  312. Banerjee, I.; Miyake, Y.; Nobs, S.P.; Schneider, C.; Horvath, P.; Kopf, M.; Matthias, P.; Helenius, A.; Yamauchi, Y. Influenza A virus uses the aggresome processing machinery for host cell entry. Science 2014, 346, 473–477. [Google Scholar] [CrossRef] [PubMed]
  313. Husain, M.; Cheung, C.-Y. Histone Deacetylase 6 Inhibits Influenza A Virus Release by Downregulating the Trafficking of Viral Components to the Plasma Membrane via Its Substrate, Acetylated Microtubules. J. Virol. 2014, 88, 11229–11239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  314. Zanin, M.; DeBeauchamp, J.; Vangala, G.; Webby, R.J.; Husain, M. Histone Deacetylase 6 Knockout Mice Exhibit Higher Susceptibility to Influenza A Virus Infection. Viruses 2020, 12, 728. [Google Scholar] [CrossRef]
  315. Galvin, H.D.; Husain, M. Influenza A virus-induced host caspase and viral PA-X antagonize the antiviral host factor, histone deacetylase 4. J. Biol. Chem. 2019, 294, 20207–20221. [Google Scholar] [CrossRef]
  316. Koyuncu, E.; Budayeva, H.G.; Miteva, Y.V.; Ricci, D.P.; Silhavy, T.J.; Shenk, T.; Cristea, I.M. Sirtuins Are Evolutionarily Conserved Viral Restriction Factors. mBio 2014, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Nutsford, A.N.; Galvin, H.D.; Ahmed, F.; Husain, M. The Class IV human deacetylase, HDAC11, exhibits anti-influenza A virus properties via its involvement in host innate antiviral response. Cell Microbiol. 2019, 21, e12989. [Google Scholar] [CrossRef]
  318. Reilly, P.T.; Yu, Y.; Hamiche, A.; Wang, L. Cracking the ANP32 whips: Important functions, unequal requirement, and hints at disease implications. BioEssays 2014, 36, 1062–1071. [Google Scholar] [CrossRef] [Green Version]
  319. Long, J.S.; Giotis, E.S.; Moncorgé, O.; Frise, R.; Mistry, B.; James, J.; Morisson, M.; Iqbal, M.; Vignal, A.; Skinner, M.A.; et al. Species difference in ANP32A underlies influenza A virus polymerase host restriction. Nat. Cell Biol. 2016, 529, 101–104. [Google Scholar] [CrossRef] [Green Version]
  320. Carrique, L.; Fan, H.; Walker, A.P.; Keown, J.R.; Sharps, J.; Staller, E.; Barclay, W.S.; Fodor, E.; Grimes, J.M. Host ANP32A mediates the assembly of the influenza virus replicase. Nat. Cell Biol. 2020, 587, 638–643. [Google Scholar] [CrossRef]
  321. Almond, J.W. A single gene determines the host range of influenza virus. Nat. Cell Biol. 1977, 270, 617–618. [Google Scholar] [CrossRef]
  322. Subbarao, E.K.; London, W.; Murphy, B.R. A single amino acid in the PB2 gene of influenza A virus is a determinant of host range. J. Virol. 1993, 67, 1761–1764. [Google Scholar] [CrossRef] [Green Version]
  323. Moncorgé, O.; Long, J.S.; Cauldwell, A.V.; Zhou, H.; Lycett, S.J.; Barclay, W.S. Investigation of Influenza Virus Polymerase Activity in Pig Cells. J. Virol. 2013, 87, 384–394. [Google Scholar] [CrossRef] [Green Version]
  324. Shinya, K.; Makino, A.; Ozawa, M.; Kim, J.H.; Sakai-Tagawa, Y.; Ito, M.; Le, Q.M.; Kawaoka, Y. Ostrich Involvement in the Selection of H5N1 Influenza Virus Possessing Mammalian-Type Amino Acids in the PB2 Protein. J. Virol. 2009, 83, 13015–13018. [Google Scholar] [CrossRef] [Green Version]
  325. Staller, E.; Sheppard, C.M.; Neasham, P.J.; Mistry, B.; Peacock, T.P.; Goldhill, D.H.; Long, J.S.; Barclay, W.S. ANP32 Proteins Are Essential for Influenza Virus Replication in Human Cells. J. Virol. 2019, 93, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Park, Y.H.; Chungu, K.; Bin Lee, S.; Woo, S.J.; Cho, H.Y.; Lee, H.J.; Rengaraj, D.; Lee, J.-H.; Song, C.-S.; Lim, J.M.; et al. Host-Specific Restriction of Avian Influenza Virus Caused by Differential Dynamics of ANP32 Family Members. J. Infect. Dis. 2020, 221, 71–80. [Google Scholar] [CrossRef]
  327. Reilly, P.T.; Afzal, S.; Wakeham, A.; Haight, J.; You-Ten, A.; Zaugg, K.; Dembowy, J.; Young, A.; Mak, T.W. Generation and Characterization of the Anp32e-Deficient Mouse. PLoS ONE 2010, 5, e13597. [Google Scholar] [CrossRef] [Green Version]
  328. Beck, S.; Zickler, M.; Dos Reis, V.P.; Günther, T.; Grundhoff, A.; Reilly, P.T.; Mak, T.W.; Stanelle-Bertram, S.; Gabriel, G. ANP32B Deficiency Protects Mice From Lethal Influenza A Virus Challenge by Dampening the Host Immune Response. Front. Immunol. 2020, 11, 450. [Google Scholar] [CrossRef] [PubMed]
  329. Baker, S.F.; Ledwith, M.P.; Mehle, A. Differential Splicing of ANP32A in Birds Alters Its Ability to Stimulate RNA Synthesis by Restricted Influenza Polymerase. Cell Rep. 2018, 24, 2581–2588.e4. [Google Scholar] [CrossRef] [Green Version]
  330. Domingues, P.; Eletto, D.; Magnus, C.; Turkington, H.L.; Schmutz, S.; Zagordi, O.; Lenk, M.; Beer, M.; Stertz, S.; Hale, B.G. Profiling host ANP32A splicing landscapes to predict influenza A virus polymerase adaptation. Nat. Commun. 2019, 10, 1–12. [Google Scholar] [CrossRef] [PubMed]
  331. Fang, A.; Bi, Z.; Ye, H.; Yan, L. SRSF10 inhibits the polymerase activity and replication of avian influenza virus by regulating the alternative splicing of chicken ANP32A. Virus Res. 2020, 286, 198063. [Google Scholar] [CrossRef]
  332. Der, S.D.; Zhou, A.; Williams, B.R.G.; Silverman, R.H. Identification of genes differentially regulated by interferon, or using oligonucleotide arrays. Proc. Natl. Acad. Sci. USA 1998, 95, 15623–15628. [Google Scholar] [CrossRef] [Green Version]
  333. Dzimianski, J.V.; Scholte, F.E.; Bergeron, É.; Pegan, S.D. ISG15: It’s Complicated. J. Mol. Biol. 2019, 431, 4203–4216. [Google Scholar] [CrossRef]
  334. Shi, H.-X.; Yang, K.; Liu, X.; Liu, X.-Y.; Wei, B.; Shan, Y.-F.; Zhu, L.-H.; Wang, C. Positive Regulation of Interferon Regulatory Factor 3 Activation by Herc5 via ISG15 Modification. Mol. Cell. Biol. 2010, 30, 2424–2436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Lenschow, D.J.; Lai, C.; Frias-Staheli, N.; Giannakopoulos, N.V.; Lutz, A.; Wolff, T.; Osiak, A.; Levine, B.; Schmidt, R.E.; García-Sastre, A.; et al. From the cover: IFN-stimulated gene 15 functions as a critical antiviral molecule against influenza, herpes, and Sindbis viruses. Proc. Natl. Acad. Sci. USA 2007, 104, 1371–1376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Morales, D.J.; Monte, K.; Sun, L.; Struckhoff, J.J.; Agapov, E.; Holtzman, M.J.; Stappenbeck, T.S.; Lenschow, D.J. Novel Mode of ISG15-Mediated Protection against Influenza A Virus and Sendai Virus in Mice. J. Virol. 2014, 89, 337–349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Zhao, C.; Hsiang, T.-Y.; Kuo, R.-L.; Krug, R.M. ISG15 conjugation system targets the viral NS1 protein in influenza A virus–infected cells. Proc. Natl. Acad. Sci. USA 2010, 107, 2253–2258. [Google Scholar] [CrossRef] [Green Version]
  338. Tang, Y.; Zhong, G.; Zhu, L.; Liu, X.; Shan, Y.; Feng, H.; Bu, Z.; Chen, H.; Wang, C. Herc5 Attenuates Influenza A Virus by Catalyzing ISGylation of Viral NS1 Protein. J. Immunol. 2010, 184, 5777–5790. [Google Scholar] [CrossRef] [Green Version]
  339. Nguyen, L.H.; Espert, L.; Mechti, N.; Wilson, D.M. The Human Interferon- and Estrogen-RegulatedISG20/HEM45Gene Product Degrades Single-Stranded RNA and DNA in Vitro. Biochemistry 2001, 40, 7174–7179. [Google Scholar] [CrossRef]
  340. Zheng, Z.; Wang, L.; Pan, J. Interferon-stimulated gene 20-kDa protein (ISG20) in infection and disease: Review and outlook. Intractable Rare Dis. Res. 2017, 6, 35–40. [Google Scholar] [CrossRef] [Green Version]
  341. Wu, N.; Nguyen, X.-N.; Wang, L.; Appourchaux, R.; Zhang, C.; Panthu, B.; Gruffat, H.; Journo, C.; Alais, S.; Qin, J.; et al. The interferon stimulated gene 20 protein (ISG20) is an innate defense antiviral factor that discriminates self versus non-self translation. PLoS Pathog. 2019, 15, e1008093. [Google Scholar] [CrossRef]
  342. Espert, L.; Degols, G.; Gongora, C.; Blondel, D.; Williams, B.R.; Silverman, R.H.; Mechti, N. ISG20, a New Interferon-induced RNase Specific for Single-stranded RNA, Defines an Alternative Antiviral Pathway against RNA Genomic Viruses. J. Biol. Chem. 2003, 278, 16151–16158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  343. Qu, H.; Li, J.; Yang, L.; Sun, L.; Liu, W.; He, H. Influenza A Virus-induced expression of ISG20 inhibits viral replication by interacting with nucleoprotein. Virus Genes 2016, 52, 759–767. [Google Scholar] [CrossRef]
  344. Chai, W.; Li, J.; Shangguan, Q.; Liu, Q.; Li, X.; Qi, D.; Tong, X.; Liu, W.; Ye, X. Lnc-ISG20 Inhibits Influenza A Virus Replication by Enhancing ISG20 Expression. J. Virol. 2018, 92, e00539-18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  345. Weiss, C.M.; Trobaugh, D.W.; Sun, C.; Lucas, T.M.; Diamond, M.S.; Ryman, K.D.; Klimstra, W.B. The Interferon-Induced Exonuclease ISG20 Exerts Antiviral Activity through Upregulation of Type I Interferon Response Proteins. mSphere 2018, 3, e00209-18. [Google Scholar] [CrossRef] [Green Version]
  346. Simon, V.; Bloch, N.; Landau, N.R. Intrinsic host restrictions to HIV-1 and mechanisms of viral escape. Nat. Immunol. 2015, 16, 546–553. [Google Scholar] [CrossRef] [PubMed]
  347. Neil, S.J.D.; Zang, T.; Bieniasz, P.D. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nat. Cell Biol. 2008, 451, 425–430. [Google Scholar] [CrossRef] [Green Version]
  348. Perez-Caballero, D.; Zang, T.; Ebrahimi, A.; McNatt, M.W.; Gregory, D.A.; Johnson, M.C.; Bieniasz, P.D. Tetherin Inhibits HIV-1 Release by Directly Tethering Virions to Cells. Cell 2009, 139, 499–511. [Google Scholar] [CrossRef] [Green Version]
  349. Winkler, M.; Bertram, S.; Gnirß, K.; Nehlmeier, I.; Gawanbacht, A.; Kirchhoff, F.; Ehrhardt, C.; Ludwig, S.; Kiene, M.; Moldenhauer, A.-S.; et al. Influenza A Virus Does Not Encode a Tetherin Antagonist with Vpu-Like Activity and Induces IFN-Dependent Tetherin Expression in Infected Cells. PLoS ONE 2012, 7, e43337. [Google Scholar] [CrossRef] [Green Version]
  350. Bruce, E.A.; Abbink, T.E.; Wise, H.M.; Rollason, R.; Galao, R.P.; Banting, G.; Neil, S.J.; Digard, P. Release of filamentous and spherical influenza A virus is not restricted by tetherin. J. Gen. Virol. 2012, 93, 963–969. [Google Scholar] [CrossRef]
  351. Londrigan, S.L.; Tate, M.D.; Job, E.R.; Moffat, J.M.; Wakim, L.M.; Gonelli, C.A.; Purcell, D.F.J.; Brooks, A.G.; Villadangos, J.A.; Reading, P.C.; et al. Endogenous Murine BST-2/Tetherin Is Not a Major Restriction Factor of Influenza A Virus Infection. PLoS ONE 2015, 10, e0142925. [Google Scholar] [CrossRef] [PubMed]
  352. 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] [CrossRef] [PubMed] [Green Version]
  353. Gnirß, K.; Zmora, P.; Blazejewska, P.; Winkler, M.; Lins, A.; Nehlmeier, I.; Gärtner, S.; Moldenhauer, A.-S.; Hofmann-Winkler, H.; Wolff, T.; et al. Tetherin Sensitivity of Influenza A Viruses Is Strain Specific: Role of Hemagglutinin and Neuraminidase. J. Virol. 2015, 89, 9178–9188. [Google Scholar] [CrossRef] [Green Version]
  354. Mangeat, B.; Cavagliotti, L.; Lehmann, M.; Gers-Huber, G.; Kaur, I.; Thomas, Y.; Kaiser, L.; Piguet, V. Influenza Virus Partially Counteracts Restriction Imposed by Tetherin/BST-2. J. Biol. Chem. 2012, 287, 22015–22029. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  355. Leyva-Grado, V.H.; Hai, R.; Fernandes, F.; Belicha-Villanueva, A.; Carter, C.; Yondola, M.A. Modulation of an Ectodomain Motif in the Influenza A Virus Neuraminidase Alters Tetherin Sensitivity and Results in Virus Attenuation In Vivo. J. Mol. Biol. 2014, 426, 1308–1321. [Google Scholar] [CrossRef] [Green Version]
  356. Hu, S.; Yin, L.; Mei, S.; Li, J.; Xu, F.; Sun, H.; Liu, X.; Cen, S.; Liang, C.; Li, A.; et al. BST-2 restricts IAV release and is countered by the viral M2 protein. Biochem. J. 2017, 474, 715–730. [Google Scholar] [CrossRef] [PubMed]
  357. Yi, E.; Oh, J.; Giao, N.Q.; Oh, S.; Park, S.-H. Enhanced production of enveloped viruses in BST-2-deficient cell lines. Biotechnol. Bioeng. 2017, 114, 2289–2297. [Google Scholar] [CrossRef] [PubMed]
  358. 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] [PubMed] [Green Version]
  359. Blanco-Melo, D.; Venkatesh, S.; Bieniasz, P.D. Origins and Evolution of tetherin, an Orphan Antiviral Gene. Cell Host Microbe 2016, 20, 189–201. [Google Scholar] [CrossRef] [Green Version]
  360. Heusinger, E.; Kluge, S.F.; Kirchhoff, F.; Sauter, D. Early Vertebrate Evolution of the Host Restriction Factor Tetherin. J. Virol. 2015, 89, 12154–12165. [Google Scholar] [CrossRef] [Green Version]
  361. Krchlíková, V.; Fábryová, H.; Hron, T.; Young, J.M.; Koslová, A.; Hejnar, J.; Strebel, K.; Elleder, D. Antiviral Activity and Adaptive Evolution of Avian Tetherins. J. Virol. 2020, 94, 01. [Google Scholar] [CrossRef]
  362. Chin, K.-C.; Cresswell, P. Viperin (cig5), an IFN-inducible antiviral protein directly induced by human cytomegalovirus. Proc. Natl. Acad. Sci. USA 2001, 98, 15125–15130. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  363. Panayiotou, C.; Lindqvist, R.; Kurhade, C.; Vonderstein, K.; Pasto, J.; Edlund, K.; Upadhyay, A.S.; Överby, A.K. Viperin Restricts Zika Virus and Tick-Borne Encephalitis Virus Replication by Targeting NS3 for Proteasomal Degradation. J. Virol. 2018, 92, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  364. Tang, H.-B.; Lu, Z.-L.; Wei, X.-K.; Zhong, T.-Z.; Zhong, Y.-Z.; Ouyang, L.-X.; Luo, Y.; Xing, X.-W.; Liao, F.; Peng, K.-K.; et al. Viperin inhibits rabies virus replication via reduced cholesterol and sphingomyelin and is regulated upstream by TLR4. Sci. Rep. 2016, 6, 30529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  365. Xu, C.; Feng, L.; Chen, P.; Li, A.; Guo, S.; Jiao, X.; Zhang, C.; Zhao, Y.; Jin, X.; Zhong, K.; et al. Viperin inhibits classical swine fever virus replication by interacting with viral nonstructural 5A protein. J. Med. Virol. 2020, 92, 149–160. [Google Scholar] [CrossRef] [PubMed]
  366. 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] [CrossRef] [PubMed] [Green Version]
  367. Tan, K.S.; Olfat, F.; Phoon, M.C.; Hsu, J.P.; Howe, J.L.C.; Seet, J.E.; Chin, K.C.; Chow, V.T.K. In vivo and in vitro studies on the antiviral activities of viperin against influenza H1N1 virus infection. J. Gen. Virol. 2012, 93, 1269–1277. [Google Scholar] [CrossRef]
  368. Cesari, M.; Pahor, M.; Incalzi, R.A. Review: Plasminogen Activator Inhibitor-1 (PAI-1): A Key Factor Linking Fibrinolysis and Age-Related Subclinical and Clinical Conditions. Cardiovasc. Ther. 2010, 28, e72–e91. [Google Scholar] [CrossRef] [Green Version]
  369. Dittmann, M.; Hoffmann, H.-H.; Scull, M.A.; Gilmore, R.H.; Bell, K.L.; Ciancanelli, M.; Wilson, S.J.; Crotta, S.; Yu, Y.; Flatley, B.; et al. A Serpin Shapes the Extracellular Environment to Prevent Influenza A Virus Maturation. Cell 2015, 160, 631–643. [Google Scholar] [CrossRef] [Green Version]
  370. Yang, J.; Zhu, X.; Liu, J.; Ding, X.; Han, M.; Hu, W.; Wang, X.; Zhou, Z.; Wang, S. Inhibition of Hepatitis B virus replication by Phospholipid scramblase 1 in vitro and in vivo. Antivir. Res. 2012, 94, 9–17. [Google Scholar] [CrossRef]
  371. Kusano, S.; Eizuru, Y. Human phospholipid scramblase 1 interacts with and regulates transactivation of HTLV-1 Tax. Virology 2012, 432, 343–352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  372. Luo, W.; Zhang, J.; Liang, L.; Wang, G.; Li, Q.; Zhu, P.; Zhou, Y.; Li, J.; Zhao, Y.; Sun, N.; et al. Phospholipid scramblase 1 interacts with influenza A virus NP, impairing its nuclear import and thereby suppressing virus replication. PLoS Pathog. 2018, 14, e1006851. [Google Scholar] [CrossRef]
  373. Sun, N.; Li, C.; Li, X.-F.; Deng, Y.-Q.; Jiang, T.; Zhang, N.-N.; Zu, S.; Zhang, R.-R.; Li, L.; Chen, X.; et al. Type-IInterferon-Inducible SERTAD3 Inhibits Influenza A Virus Replication by Blocking the Assembly of Viral RNA Polymerase Complex. Cell Rep. 2020, 33, 108342. [Google Scholar] [CrossRef] [PubMed]
  374. Gao, Q.; Yang, C.; Ren, C.; Zhang, S.; Gao, X.; Jin, M.; Chen, H.; Ma, W.; Zhou, H. Eukaryotic Translation Elongation Factor 1 Delta Inhibits the Nuclear Import of the Nucleoprotein and PA-PB1 Heterodimer of Influenza A Virus. J. Virol. 2020, 95. [Google Scholar] [CrossRef]
  375. Tafforeau, L.; Chantier, T.; Pradezynski, F.; Pellet, J.; Mangeot, P.E.; Vidalain, P.-O.; Andre, P.; Rabourdin-Combe, C.; Lotteau, V. Generation and Comprehensive Analysis of an Influenza Virus Polymerase Cellular Interaction Network. J. Virol. 2011, 85, 13010–13018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Annotated schematic representation of an IAV virion.
Figure 1. Annotated schematic representation of an IAV virion.
Viruses 13 00522 g001
Figure 2. Schematic representation of IAV sensing and subsequent IFN signaling.
Figure 2. Schematic representation of IAV sensing and subsequent IFN signaling.
Viruses 13 00522 g002
Figure 3. Schematic representation of IAV life cycle with the host cell restriction factors that target each step. Factors highlighted in green are those that are generally accepted or that have consequent data backing them, in orange factors that need further investigation to confirm their role as IAV restriction factors for instance in vivo (e.g., NCOA7-AS) and in grey, those for which very little data is available.
Figure 3. Schematic representation of IAV life cycle with the host cell restriction factors that target each step. Factors highlighted in green are those that are generally accepted or that have consequent data backing them, in orange factors that need further investigation to confirm their role as IAV restriction factors for instance in vivo (e.g., NCOA7-AS) and in grey, those for which very little data is available.
Viruses 13 00522 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

McKellar, J.; Rebendenne, A.; Wencker, M.; Moncorgé, O.; Goujon, C. Mammalian and Avian Host Cell Influenza A Restriction Factors. Viruses 2021, 13, 522. https://doi.org/10.3390/v13030522

AMA Style

McKellar J, Rebendenne A, Wencker M, Moncorgé O, Goujon C. Mammalian and Avian Host Cell Influenza A Restriction Factors. Viruses. 2021; 13(3):522. https://doi.org/10.3390/v13030522

Chicago/Turabian Style

McKellar, Joe, Antoine Rebendenne, Mélanie Wencker, Olivier Moncorgé, and Caroline Goujon. 2021. "Mammalian and Avian Host Cell Influenza A Restriction Factors" Viruses 13, no. 3: 522. https://doi.org/10.3390/v13030522

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