Next Article in Journal
Synthesis and Evaluation of Novel Pyrazole Ethandiamide Compounds as Inhibitors of Human THP-1 Monocytic Cell Neurotoxicity
Next Article in Special Issue
Influenza A Hemagglutinin Passage Bias Sites and Host Specificity Mutations
Previous Article in Journal
The Aberrant Expression of the Mesenchymal Variant of FGFR2 in the Epithelial Context Inhibits Autophagy
Previous Article in Special Issue
Cellular Proteostasis During Influenza A Virus Infection—Friend or Foe?
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Viroporins in the Influenza Virus

School of Biological Sciences, Nanyang Technological University, 60 Nanyang Drive, Singapore 637551, Singapore
*
Author to whom correspondence should be addressed.
Cells 2019, 8(7), 654; https://doi.org/10.3390/cells8070654
Submission received: 19 May 2019 / Revised: 21 June 2019 / Accepted: 27 June 2019 / Published: 29 June 2019
(This article belongs to the Special Issue Host–Pathogen Interactions During Influenza Virus Infection)

Abstract

:
Influenza is a highly contagious virus that causes seasonal epidemics and unpredictable pandemics. Four influenza virus types have been identified to date: A, B, C and D, with only A–C known to infect humans. Influenza A and B viruses are responsible for seasonal influenza epidemics in humans and are responsible for up to a billion flu infections annually. The M2 protein is present in all influenza types and belongs to the class of viroporins, i.e., small proteins that form ion channels that increase membrane permeability in virus-infected cells. In influenza A and B, AM2 and BM2 are predominantly proton channels, although they also show some permeability to monovalent cations. By contrast, M2 proteins in influenza C and D, CM2 and DM2, appear to be especially selective for chloride ions, with possibly some permeability to protons. These differences point to different biological roles for M2 in types A and B versus C and D, which is also reflected in their sequences. AM2 is by far the best characterized viroporin, where mechanistic details and rationale of its acid activation, proton selectivity, unidirectionality, and relative low conductance are beginning to be understood. The present review summarizes the biochemical and structural aspects of influenza viroporins and discusses the most relevant aspects of function, inhibition, and interaction with the host.

1. Introduction

1.1. Influenza Viruses

Influenza viruses are enveloped, segmented, negative-sense RNA viruses belonging to the Orthomyxoviridae family. Four influenza virus types have been identified to date, classified based on their core proteins: A, B, C [1] and D [2,3]. Type A is known to infect humans, as well as porcine, bovine and canine species [4]. The main reservoir of the influenza A viruses are aquatic birds [5]. The seasonal flu caused by influenza A virus (IAV) in humans is a very contagious respiratory illness that is among the top ten leading causes of death in the United States and associated with high medical burden [6,7,8]. Sudden pandemics with high mortality rates are caused by the sporadic transmission of avian or swan influenza viruses to humans, as pre-immunity to these new strains is non-existent [6,7]. Influenza B virus (IBV), its close relative, is less severe but still capable of causing serious outbreaks, is responsible for seasonal influenza epidemics among humans [9,10], and accounts for half of the influenza diseases in recent years (www.cdc.gov). Types B (IBV) and C (ICV) infect humans and pigs [11], whereas type D (IDV) infects cattle and pigs [3]. In addition, IBV has also been reported to infect harbor seals [12]. Although IDV infection in cattle is usually asymptomatic, it can lead to disease in swine [2,13].

1.2. The Proteins in Influenza Viruses

In IAV, the viral envelope contains cholesterol-enriched lipid rafts and transmembrane glycoproteins hemagglutinin (HA) and neuraminidase (NA) [14], which are used to further classify IAV strains. A third membrane protein is matrix protein 2 (AM2 in IAV), present in lower abundance [15]. HA, NA, and M2 are type I, type II, and type III membrane proteins, respectively [16,17,18]. A layer of matrix protein 1 (M1) just underneath the membrane is the most abundant protein of the virus. M1 forms an internal coat, enclosing the viral ribonucleoproteins as well as three polymerase proteins (PA, PB1, and PB2) that form the viral RNA polymerase complex [19].
IAV and IBV have eight negative-sense RNA segments, each encoding one or two viral proteins. IAV encodes 10–11 proteins: HA, NA, NP, M1, M2, PA, PB1, PB2, NS1, NS2, and PB1-F2 [19]. IBV has the same proteins as IAV, except for the PB1-F2 protein, and has an additional surface glycoprotein NB that is unique to flu B. ICV and IDV consist of only seven gene segments each, and they do not encode envelope glycoproteins HA and NA. Instead, they carry only one glycoprotein, HEF (haemagglutinin–esterase fusion), which combines the functions of HA and NA [20].
M2, PB1-F2, and NB proteins are classified as viroporins, i.e., virus-encoded short polypeptides (approximately 60–120 amino acids) that have one, two, or even three [21] transmembrane (TM) α-helical domains. These polypeptides form oligomers of typically four to six monomers that permeabilize membranes to ions [22,23,24].

2. AM2 and BM2

2.1. AM2

The RNA segment 7 from IAV encodes both the M1 protein and AM2 proton channel via alternative splicing [25,26]. AM2 is 97-residues long and forms a channel assembled by the tetramerization of four left-handed TM helices stabilized by disulfide bridges [18,27]. The AM2 monomer consists of an N-terminal ectodomain (residues 1–24), an α-helical (TM) domain (residues 25–46), and a C-terminal cytoplasmic tail (CT) domain (residues 47–97) that contains a highly conserved amphipathic helix (APH) (residues 51–59) [25,28].
AM2 forms an acid-activated proton channel with an asymmetrical preference of proton conduction from the viral exterior (N-terminus) to the interior (C-terminus). The proton channel activity of AM2 is essential for two important roles. Upon viral entry into cells via sialic acid-containing receptor-mediated endocytosis, the low pH (5 to 6) of late endosomes activates the proton channel activity of AM2, which acidifies the viral interior [25] with subsequent detachment of M1 from the vRNP complex. This enables trafficking of vRNPs into the host cell nucleus for synthesis of mRNA and vRNA [29,30]. The channel activity of M2 is also crucial during viral maturation, where it equilibrates the pH across the trans-Golgi membrane to prevent the viral HA from undergoing premature conformational change while being transported to the plasma membrane of infected cells [31,32,33,34]. Conductance and inhibition studies have used the transmembrane domain of AM2 (AM2-TM), since in Xenopus oocytes AM2(21–51) has retained the proton selectivity and drug sensitivity observed in full-length AM2 protein [35].
Four main features characterize AM2-facilitated membrane permeabilization to protons: acid activation, proton selectivity, relative low proton conductance, and unidirectionality. These features are encoded in key pore-lining residues of the AM2 channel. AM2 has a conserved HxxxW functional motif in its TM domain (Figure 1), where His37 is responsible for proton selectivity and acid activation [36], and Trp41 ensures asymmetric proton conduction from the N-terminus to the C-terminus [37]. Other residues surrounding this motif also contribute to the dynamics and proton transfer equilibria in the channel.

2.2. BM2

In IBV, RNA segment 7 encodes both M1 protein and BM2 [38,39]. Like AM2, BM2 is a pH-activated proton channel [40] and has a similar monomeric and oligomeric organization as described above for AM2 [41,42]. Like AM2, a truncated peptide containing its TM, BM2(1–33), conducts protons when incorporated into artificial liposomes [42] and Xenopus oocytes, with similar conductance and proton-selectivity as that observed in full-length BM2 protein [43]. Despite these similarities, AM2 and BM2 share almost no sequence identity, with the exception of an HxxxW motif in the TM domain (see Figure 1), which may explain some of their common features. AM2 and BM2 also differ in post-translational modifications; while BM2 is only modified by phosphorylation [44], AM2 contains disulfide bonds and is palmitoylated and phosphorylated [45,46,47,48]. Like AM2, BM2 is essential for virus uncoating in the endosome and for pH equilibration between Golgi lumen and cytoplasm during virus protein transport [38]. However, while the AM2 ectodomain is important for its incorporation into virions [49], the BM2 has only a small ectodomain [41,42] (Figure 1).

2.3. Acid Activation Mechanism of AM2

At a high pH (e.g., 7 to 8), the AM2 channel is in a Cclosed conformation (Figure 2, left), where the side chains at the C-terminal end of the channel, including His37 and Trp41 tetrads, are tightly packed, and the pore is lined by alternating layers of side chains and well-ordered water clusters. The closed Trp41 tetrad dehydrates the His37 tetrad and raises the His37 deprotonation barrier, thus blocking proton conduction through the channel. This conformation has been observed by X-ray crystallography [50,51] as well as in both solution and solid-state NMR [52,53,54,55,56]. When the pH decreases to around 6, the His tetrad increases its protonation state to +2, and the channel becomes activated. Electrostatic repulsion with protons is lower due to the low charge state of the His37 tetrad, allowing proton permeation from the viral exterior. This asymmetry partly explains the rectification of proton flow observed experimentally. Protons rapidly diffuse to the His37 tetrad via an ordered water cluster, when the IAV particle is incorporated into the endosome, as it is surrounded by an acidic environment. At a low pH, the positive charge of the His37 tetrad increases, and the Trp41 gate and the C-terminal helices open and become more hydrated. This lowers the His37 deprotonation barrier, increasing proton conductance. Further reduction of pH expands the channel and increases pore water mobility, further increasing proton conductance [57].
The His37 tetrad can adopt four protonation states, one for each His protonation, where the first two protonations already occur at a high pH [58,59]. The transition between Cclosed to Copen state (Figure 2, right) is triggered by protonation of the third His37 at an acidic pH [58,60], with a pKa value that ranges from 4.9 to 6.6 depending on membrane composition [54,58,61,62,63,64]. At this point, protons can be released into the virus interior. This Copen form has been characterized by X-ray crystallography [50,51,55,58,65,66,67,68]. Deprotonation of the His37 tetrad triggers a change back to the Cclosed conformation [55,68].
Depending on the role of the His37 tetrad, several proton transport models have been suggested [35]. In the “shutter” mechanism [62], the His37 tetrad works as a gate. At a low pH (around pH 6, see above) the gate opens due to the electrostatic repulsion between the biprotonated, positively charged histidine residues. The excess proton is transferred via the Grotthuss mechanism (i.e., proton transfer through a network of hydrogen-bonded water molecules [69]), without changing the protonation state of His37. In the “shuttle” mechanism (e.g., [55]), the His37 tetrad changes protonation state at a low pH, while the proton is shuttled through the His37–Trp41 region.

2.4. Rate of Proton Conductance in AM2 and BM2

Although the proton conductance rate of AM2 is higher at a low pH [71,72], in the range of 101–104 protons s−1 [71,73,74], it is still orders of magnitude lower than expected for a water-filled pore of the size of the M2 channel without ionizable groups [75]. This low conductance may exist to prevent toxicity in the host and has been attributed to a rate-limiting step caused by His37 protonation–deprotonation [36,76]. However, the His37–water proton exchange rate appears to be too fast to account for this (~105 s−1) [54], and additional factors may account for the low rate of proton conduction [77,78,79,80]. Recent single-molecule fluorescence experiments using Trp41 as a probe have detected interhelical motion in AM2 on a microsecond timescale, which is consistent with the proton conduction rate, and which increased from a high to low pH with a transition midpoint in the range of those reported for the pKa of the His37 tetrad [81]. This is consistent with a range of studies that show that the AM2 tetramer is more dynamic at a low pH [70,76,77,82,83,84,85,86]. Thus, proton conductance through AM2 may be limited by the transition between Cclosed and Copen conformers. The structure and dynamics of water in the AM2 pore are also important because it is the essential medium for proton permeation. X-ray free-electron laser (XFEL) [68] and 2D infrared (2D-IR) spectroscopy [87] studies of AM2 have shown that, at a low pH, water molecules form a continuous hydrogen-bonded network spanning the channel length.
In BM2, although conductance at a low pH increases more than for AM2, proton conductance is still much lower than the theoretical conductance at a low pH. This suggests that conformational change of BM2 may also constitute a rate-limiting step [43]. The role of the two His residues, especially His19, is important since an H19C mutation abolishes proton conduction [88], whereas an H27A mutation can reduce proton transport by ∼25% in a liposome proton flux assay [42]. The proton-dissociation equilibrium constants (pKa values) of the His19 tetrad in BM2 TM domain in lipid bilayers were found to be one pH unit lower (i.e., more difficult to protonate) than the pKa values of the AM2-equivalent His37 tetrad [89,90]. This was attributed to the presence of a second titratable histidine, His27 (in the position occupied by Arg45 in AM2), separated by three residues from Trp23 (see Figure 1), forming a putative reverse WxxxH motif. This His27 residue was suggested to increase the proton-dissociation rate of His19 by stabilization of its neutral state. This has been confirmed by a two-electrode voltage clamp (TEVC) electrophysiological assay in Xenopus oocytes [43] that compared AM2 (A/Udorn/72) and BM2 (B/Lee/40) (both C-terminally fused to FLAG), where the apparent pK3 (3rd protonation state) value of the AM2 channel was about one pH unit higher than that of the BM2 channel, indicating that the latter is less easily protonated and requires a lower pH for activation. Despite this, BM2 was found to have a 10-fold higher conductance than AM2, resulting in an overall 60% higher specific conductance of the BM2 channel than the AM2 channel at a pH of 5.5, which is consistent with liposome flux assays, where the specific conductance of BM2(1–33) is about double that of AM2(18–60) [42].

2.5. Asymmetric Conductance in AM2 and BM2

The AM2 proton channel is inward rectified (i.e., it conducts more easily in the inward direction), from the N- to C-terminus. This is attributed to Trp41 and Asp44 [37,60]. The Trp41 tetrad forms a gate that physically occludes the C-terminal end of the pore, while Asp44 stabilizes the Trp41 gate via water-mediated hydrogen bonds. When oocytes are subjected to low pHin and high pHout conditions, no outward current in the AM2 proton channel is observed [60]. The W41F mutant of AM2 allows a reverse current when the internal pH was low [59]. The latter ssNMR study showed that the protonation and tautomeric equilibria at His37 were altered, suggesting a role for Trp41 in preventing C-terminal acid activation by His37.
Interestingly, the BM2 proton channel is slightly outwardly rectified [40]. In BM2, the corresponding residue of AM2-Asp44 is Gly26 (see alignment). Although mutation G26D in BM2 is unable to prevent this outward leakage in a pHin low/pHout high condition, the double mutant G26D/H27R (to mimic AM2 Asp44/Arg45) has achieved similar asymmetric proton flow (I–V curve inward rectified) to the AM2 channel [43]. Thus, asymmetric inward rectified proton flow can be restored by engineering these two residues of AM2 into the corresponding positions of the BM2 channel. Indeed, in the X-ray crystal structure of AM2(22–46), Asp44 participates in ionic interactions with Arg45 and also in water-mediated hydrogen bonding with Trp41 [50]. Conversely, AM2-Asp44 mutants D44N and D44C exhibit an outward current under a low pHin /high pHout condition, similar to BM2 [60].

3. CM2 and DM2

3.1. CM2

The influenza C virus M2 protein (CM2) was first reported by Hongo et al. [91] and was proposed to function as a voltage-activated chloride ion channel when expressed in Xenopus oocytes [92]. CM2 is encoded by RNA segment 6 (M gene) of the influenza C virus [20], which produces a p42 protein precursor, and upon cleavage by a peptidase, CM1 and CM2 are obtained [93,94]. Like AM2, CM2 is a type III membrane protein, with a 23-aa N-terminal extracellular domain, a 23-aa transmembrane domain and a 69-aa C-terminal cytoplasmic domain [95] (Figure 1). CM2 is abundantly expressed in virus-infected cells, and a small amount is incorporated into the virus particles [96]. CM2 forms disulfide-linked dimers and tetramers via evolutionarily conserved Cys1–Cys6–Cys20 [97] and is post-translationally modified by N-glycosylation (Asn11), palmitoylation (Cys65), and phosphorylation [95,96,98].
The structure of the TM domain of the CM2 protein (Tyr27–Val46) has been obtained by site-specific infrared dichroism and the molecular modeling [99] of isolate C/Ann Arbor/1/1950 [100]. In that model, side chains of L31, L34, M41 and L44 residues were predicted to be lumenally oriented, and the peptide formed a left-handed coiled-coil tetramer [99,101]. The transmembrane pore of CM2 was occluded by residue M41, and a motif of hydrophilic residues, T30, S33, T40 and Y43, was located at the outer surface of the CM2 channel.
CM2 protein is capable of modifying pH within the trans-Golgi network (TGN). Indeed, when CM2 was co-expressed with a pH-sensitive hemagglutinin of IAV, CM2 protein was able to protect the co-expressed HA against acid activation in the TGN [102], although its activity was much lower than that of the AM2 protein. Additionally, a chimeric AM2 protein containing the CM2 transmembrane domain could partially restore the infectious virus production of an M2-deficient influenza A virus [103], which suggests that CM2 can alter intracellular pH. Electrophysiological studies of CM2-expressing mouse erythroleukemia cells also identified proton permeability [97], and such permeability probably plays a role in the uncoating process of the influenza C virus. The glycoprotein HEF of the influenza C virus has the ability to cause low pH-dependent hemolysis and fusion [104], and the virion is presumably uncoated in the acidic endosomal compartment. The CM2 protein might, therefore, have a function similar to that of the M2 protein in virus uncoating. Nevertheless, the proton and chloride permeabilities have not been clearly dissected.
AM2 is 106- to 107-fold more permeable to H+ than to alkali metal cations Na+/K+ [71,105,106]. By contrast, CM2 is permeable to Cl- anions but not to cations, exhibits a milder response to pH [92,102,107] and has only a small proton channel activity that is insensitive to amantadine [92]. Those authors proposed that the CM2 protein is transported to the cell surface where it may lower the ionic strength just beneath the viral budding site by inducing the efflux of Cl- ions. This may be advantageous during virion assembly to promote the interaction of M1 with RNPs, since a slight increase in salt concentration may trigger the dissociation of the M1–RNP complex in ICV [108]. Indeed, deletion of CM2 has caused impaired packaging and uncoating in virus-like particles (VLPs) and recombinant influenza viruses [109].

3.2. DM2

A new type of influenza virus, known as type D, has recently been identified in cattle and pigs [3]. The proteome of IDV is closer to ICV than to IAV or IBV [2] and, correspondingly, CM2 and DM2 share similarities. DM2 was recently tested in Xenopus oocytes using the two-electrode voltage clamp (TEVC) method, and an induced inward current was observed similar to that of CM2, with similar reversal potential [110]. Neither CM2 nor DM2 contain an HxxxW motif (found in AM2 and BM2), but a “YxxxK” motif has been proposed instead [110], where Y and K would have a lumenal orientation (parallel to H and W in the HxxxW motif found in AM2 and BM2) (see alignment). Indeed, the gating voltage of CM2 and DM2 was shown to be affected by modifications at the YxxxK motif [110], which were proposed to be involved in cation–pi pairs between Y and K, and were proposed to be able to regulate ion flow. Overall, the significance of the difference in functional motifs between AM2 and BM2 versus CM2 and DM2 remains to be elucidated. The main features of viroporins in influenza are summarized in Table 1.

4. Other Influenza Viroporins

IAV and IBV encode other viroporins, PB1-F2 and NB, respectively. For the sake of completion, these two viroporins are briefly described here. PB1-F2 forms nonselective ion channels in planar lipid bilayers and microsomes [111], and is known to localize to the mitochondria of infected cells [112]. PB1-F2 can interact with two mitochondrial proteins—adenine nucleotide translocator (ANT3) and voltage-dependent anion channel 1 (VDAC1)—present in the inner and outer mitochondrial membranes, respectively, leading to the dissipation of mitochondria membrane potential [113,114]. PB1-F2 from pathogenic IAV can be incorporated into the phagolysosomal compartment to trigger NLRP3 inflammasome activation, inducing the secretion of pyrogenic cytokine IL-1β and thereby leading to severe pathophysiology [115]. In addition, the PB1-F2 of A/Puerto Rico/8/1934(H1N1) (PR8) has been reported to bind the mitochondrial antiviral signaling protein (MAVS), leading to impairment of IFN-β production [114]. However, the precise role of PB1-F2 in modulation of IAV-induced immunopathogenesis remains to be dissected.
NB protein is encoded by IBV at RNA segment 6 [116]. NB is modified by glycosylation and palmitoylation, with the latter being important for NB trafficking to the cell surface [117]. Purified NB has been found to form ion channels when incorporated into artificial lipid bilayers [118], and further studies on an NB-S20A mutant have resulted in altered proton permeability and channel gating [119]. However, in contrast to the BM2 protein, NB is not required for viral replication in cell culture [120]. IBV NB cannot modify pH within the TGN [107].

5. Influenza Viroporin Inhibition

The vast majority of viroporin inhibitors have been developed against the AM2 protein. This is not surprising since AM2 was the first viroporin discovered, has a well-established biological role in viral pathogenesis and is a proven drug target [31,38,121]. Amantadine (Amt, Symmetrel) is considered the first viroporin channel inhibitor and the second antiviral agent ever discovered [122]. It binds the pore of the AM2 channel in a hydrophilic pocket [56,123,124]. Together with its close derivative, rimantadine (Rim, Flumadine), they are the only licensed antiviral drugs that target viroporins [125] despite their modest affinity for AM2 (e.g., for Amt, IC50 = 16 μM). However, most currently circulating IAV strains are Amt- and Rim-resistant [126,127,128,129], and therefore the use of Amt and Rim has been discontinued in humans [130]. The focus of the current research is the Amt-resistant variants of AM2 [131,132,133,134]. Indeed, more than 95% of circulating influenza A viruses carry the S31N mutation in their M2 sequence, while about 1% have V27A mutations, and less than 0.2% carry rare mutations (L26F, A30T, G34E and L38F) [135]. The reader is referred to previous reviews (e.g., [136,137]) and a more recent one that is also general for viroporin inhibition [138]. Recently, the role of water in drug binding was highlighted by the demonstration that small molecules can enable potent inhibition by targeting key waters in hydrogen-bonded networks that are used to facilitate proton diffusion [51].
Although AM2 and BM2 are functional homologs, wild-type BM2 is insensitive to Amt and Rim [38,40], and no BM2 inhibitors have been identified to date. The lack of amantadine inhibition of BM2 may be explained by differences in lumen hydrophobicity in the two channels. For example, the BM2 structure solved by solution NMR in DHPC [42] shows a coiled-coil tetramer channel apparently in the Cclosed conformation, with three polar residues (Ser-9, Ser-12, and Ser-16) lining the pore and a bulky phenylalanine ring (Phe-5) protruding into the pore at the N-terminus and possibly blocking the BM2 channel pore. By contrast, the corresponding pore-lining residues in the AM2 channel are much more hydrophobic: Val-27, Ala-30 and Gly-34 [36,53]. Thus, the more hydrophilic channel of BM2 protein may not be able to accommodate a hydrophobic drug like amantadine. However, the poor inhibition by other small molecules is difficult to explain, since N-terminal interhelical separation of BM2, measured using fluorinated Phe-5, cannot prevent the access of small molecules to the lumen of the pore [90].
In addition to adamantanes and other small molecules, the AM2 proton channel can be also inhibited by both oxidized and reduced copper ions, Cu(II) and Cu(I). Biphasic inhibition observed in electrophysiological assays has suggested two Cu(II)-binding sites with different affinities [139]. Solid-state NMR studies have shown that the low-affinity site is non-specific and localized at the membrane surface, whereas the high-affinity site would be located inside the channel in between the four imidazole rings from His37 residues and the four indole rings from Trp41 [140]. Influenza A virus replication is affected by Cu(II), presumably through inhibiting the AM2 proton channel by the dissociated Cu(II) [141]. BM2 is also inhibited by Cu(II) in a biphasic manner [43], although it is less sensitive to Cu(II) inhibition. Therefore, Cu(II) complexes may also be toxic to influenza B viruses.

6. M2-Mediated Disruption of Ion Homeostasis

The channel activity of IAV M2 has been found to be sufficient for the activation of the NLRP3 inflammasome in influenza-infected cells [142]. AM2 also contributes to influenza pathogenesis by downregulating the expression and function of two host ion channels, the amiloride-sensitive epithelial sodium channels (ENaC) [143] and the cystic fibrosis transmembrane conductance regulator (CFTR) chloride channels [144]. These are present in apical membranes of lung epithelium and are important regulators for the absorption and secretion of fluids and electrolytes [145,146]. An impairment in their activity is associated with detrimental consequences, including pulmonary edema and rhinorrhea (see [147]). AM2 has been found to reduce the surface protein levels and channel activity of ENaC, probably via signal transduction mechanisms. M2 can upregulate the steady-state levels of reactive oxygen species, likely by decreasing mitochondrial membrane potential and activation of protein kinase C (PKC), leading to modification of ENaC to enhance its endocytosis and proteasomal degradation. Inhibition of ENaC appears to be mediated by the C-terminus of M2 [143].
CFTR is a cyclic AMP-activated chloride channel that mediates the regulation of the thickness and composition of lung epithelial lining fluids (ELFs). M2 reduces the activity of the CFTR channel by alkalinizing the secretory organelle pH, thereby targeting CFTR for destruction via a ubiquitin-dependent pathway. Several studies have demonstrated that the proton channel activity of M2 is required for inhibition of CFTR: (i) addition of M2 ion channel inhibitor amantadine rescued CFTR expression and activity in Xenopus oocytes expressing both M2 and CFTR channels, and (ii) co-expression of CFTR with M2 channel-defective mutants (M2-G34V and M2-V27F) did not alter the GlyH-101-sensitive currents [144].

7. Protein–Protein Interactions (PPIs)

Influenza viroporins are also involved in viral pathogenesis by engaging in critical interactions with viral proteins or disrupting normal host cellular pathways through coordinated interactions with host proteins. The reader is referred to a recent review on this subject [148] that provides an update on the characterization of the main PPIs for most viroporins, as well as the role of viroporins in these PPI interactions. In the next part of this review, we highlight and summarize the nature and outcome of some of these protein–protein interactions involving M2, and how the influenza virus may exploit these host cell factors to its advantage.

7.1. Modulation of Host Autophagy

In IAV, M2 plays an important role in evading host autophagy by interacting with autophagic proteins such as Beclin-1 and LC3 (microtubule-associated protein light chain 3). M2 blocks autophagosome fusion with lysosomes by interacting with Beclin-1 through the N-terminal 60 amino acids of M2, resulting in enhanced apoptotic cell death [149]. IAV is also able to subvert host autophagy by encoding a short linear motif (SLiM) in its M2 protein to mimic a highly conserved LC3-interacting region (LIR) motif, W/FxxI/L/V. This interaction hijacks LC3 to the plasma membrane to prevent autophagosome–lysosome fusion and is essential for IAV budding and transmission [150]. Mutations in M2 that abolish its binding to LC3 have led to reduced filamentous virion budding and stability in vitro [151], suggesting that the M2–LC3 interaction may also aid in viral transmission by enhancing virion stability.

7.2. Interplay with Host Defense

Influenza infection induces a number of host antiviral activities, including interferon (IFN) responses and activation of double-stranded RNA (dsRNA)-activated protein kinase (PKR) [152]. During the early stages of infection, host PKR activation is inhibited by IAV to permit viral replication, by (i) masking of viral dsRNA using its nonstructural protein 1 (NS1) to prevent PKR activation [153], and (ii) activation of the PKR inhibitor P58IPK through nucleoprotein (NP)-mediated dissociation of Hsp40 from the Hsp40–P58IPK complex [154]. During the late stages of infection, AM2 stabilizes the Hsp40–P58IPK complex, leading to PKR activation and subsequent cell death and viral release [155]. While BM2 was found to interact with Hsp40 through the C-terminal domain 1 (CTD1) of Hsp40 [155], the binding site in P58IPK is not known.
Tetherin (BST-2, bone stromal cell antigen 2) is an interferon-inducible antiviral host factor that restricts the release of many enveloped viruses by forming a proteinaceous link, tethering budding virions to the cell surface [156]. IAV has evolved several mechanisms to neutralize the activity of tetherin, one of which involves AM2 interaction with tetherin to downregulate its surface expression via the proteasomal pathway, probably mediated by AM2 extracellular and transmembrane domains [157].

7.3. Targeting by Host Restriction Factors

One identified binder of the cytoplasmic tail of M2 is annexin A6 (AnxA6) [158], a Ca2+/lipid-binding scaffold protein that interacts with lipid rafts and regulates cholesterol homeostasis. siRNA-silencing of AnxA6 expression has enhanced viral release, while its overexpression has resulted in reduced viral titer, suggesting that AnxA6 negatively regulates IAV infection [158]. It is speculated that AnxA6 may interact with the M2 cytoplasmic tail during membrane scission, although it remains to be elucidated how this host–pathogen interaction may contribute to the impairment of IAV budding and release.
Cyclin D3 is a key cell-cycle regulator of the G0/G1 phase that binds to the M2 cytoplasmic tail [159]. Like AnxA6, siRNA-knockdown of cyclin D3 led to increased virus titer. It has been proposed that cyclin D3 disrupts the interaction between IAV M1 and M2 by competing with M1 for binding to M2, consequently impairing the proper assembly of progeny virions. To counteract cyclin D3 activity, IAV is able to induce the relocalization of cyclin D3 from the nucleus to the cytosol, where it is targeted for proteasomal degradation.

7.4. Modulation of Viral Replication

The cytoplasmic tail region of IAV M2 encodes another host SLiM that interacts with the host caveolin-1 (Cav-1). Cav-1 is a cholesterol-binding raft-residing membrane protein with multiple roles in lipid and membrane trafficking, as well as in signal transduction [160]. Inhibition of Cav-1 correlates with decreased virus titers in IAV-infected cells [161]. A consensus Cav-1-binding motif (CBM), ΦxxxxΦxxΦ (where Φ represents an aromatic amino acid) [162], was found in the amphipathic helix of the M2 cytoplasmic tail (F47FKCIYRRF55) [161]. The M2/Cav-1 interaction has been found to be involved in modulating IAV replication [161], but it remains to be determined whether this interaction also affects the surface localization of M2 and other viral proteins.
Targeting the function of the host cellular Na,K-ATPase pump offers a promising antiviral treatment (recently reviewed in [163]), since infection by RNA viruses can affect the expression and activity of Na,K-ATPase. For instance, infection with IAV/H1N1 and IAV/H5N1 has led to the downregulation of Na,K-ATPase in alveolar epithelium [164]. Cardiac glycosides are classical inhibitors of Na,K-ATPases and have been reported to exhibit inhibitory effects on the influenza virus [165]. This action of cardiac glycosides on viral replication could be due to activation of signaling cascades, or by altering the concentration of intracellular ions. Influenza protein translation has been reported to be affected by cardiac glycosides such as ouabain, digoxin and lanatoside C, which express antiviral activity against influenza [166,167].
The β1 auxiliary subunit of the Na+/K+ ATPase, ATP1B1, was identified to interact with the cytoplasmic region of both AM2 and BM2, and a stable ATP1B1-knockdown cell line was able to suppress influenza virus A/WSN/33 replication [167]. Using deletion mutations, the authors narrowed down the interacting region to residues 28–48 in BM2, and residues 117–303 in ATP1B1, although the interaction between AM2 and ATP1B1 has not been characterized.

7.5. Modulation of Surface Expression

During the late stages of the IAV replication cycle, the vRNP complexes as well as viral envelope glycoproteins and structural proteins need to be targeted to the virus assembly sites. M2 has to be transported from the rough endoplasmic reticulum (RER) to the apical plasma membrane through the host secretory pathway [168,169]. This transport travelling is thought to be modulated by several cellular factors, including the ubiquitin protein ligase E3 component N-recognin 4 (UBR4), the vesicle transport protein Rab11 and the transport protein particle complex 6A delta (TRAPPC6AΔ).
UBR4 is a 600-kDa member of the UBR box-containing an N-recognin family with multiple roles, including targeting protein for ubiquitination and proteasomal degradation [170]. In the context of an IAV infection, UBR4 has been proposed to promote surface localization of the viral envelope proteins HA, NA, and M2. The M2 transmembrane and cytoplasmic tail regions have been found to be important for its interaction with host UBR4. Knockout of UBR4 has led to enhanced co-localization of M2 with the autophagosome marker DIRAS3 and reduced M2 cell surface expression [171]. Taken together, the interaction between M2 and host UBR4 may benefit IAV replication by (i) protecting M2 from host degradation, and (ii) promoting surface localization of M2 and other viral glycoproteins to facilitate efficient IAV budding and release.
The Rab11 pathway is known to be involved in the IAV budding process [172]. Rab11 is a small GTP-binding protein that traffics proteins and vesicles between the trans-Golgi network (TGN), recycling endosome and apical membrane [173,174,175]. Knockout of Rab11 has also led to a reduction of M2 surface levels [176], proposing a role for Rab11 in the apical delivery of M2.
TRAPPC6A and its truncated form TRAPPC6AΔ also interact with IAV M2 through a highly conserved leucine residue located at the M2 cytoplasmic tail (M2-L96) [177]. TRAPPC6A is a component of the multi-subunit transport protein particle (TRAPP) complex that mediates ER-to-Golgi transport [178]. siRNA knockdown of TRAPPC6AΔ increases M2 surface expression, while TRAPPC6AΔ overexpression has produced the opposite effect, suggesting that TRAPPC6AΔ slows down M2 apical trafficking. Two other subunits of the TRAPP complex, TRAPPC5 and TRAPPC9, were identified as binders to M2 in all of four independent influenza interactome screening campaigns [171,179,180,181]. It would be interesting to determine if these TRAPP complex subunits may cooperate in the modulation of the influenza life cycle.
Finally, the host factor Golgi-specific brefeldin A-resistant guanine nucleotide exchange factor 1 (GBF1) is proposed to play an important role in the trafficking of viral proteins to the cell surface during the late stages of virus replication [181]. The latter work reported that siRNA-mediated depletion of GBF1 correlates with (i) significant reduction in virus titers, (ii) reduced efficiency of VLP formation and (iii) altered intracellular localization of the viral envelope proteins HA and NA. However, while GBF1 has been reported to associate with M2, it is not completely understood how M2 contributes to the process.

8. Conclusions

In summary, most information available for influenza viroporins is centered around AM2, followed by BM2. For the latter, structural data are increasingly available, but protein–protein interaction data studies comparable to AM2 are lacking. Viroporins in ICV and IDV have attracted much less interest, concomitant with their lower virulence in humans and biomedical relevance. Nevertheless, their mode of action, and those of lesser known viroporins, can help to better understand ionic transport across membranes.

Author Contributions

Writing—original draft preparation, J.T. (Janet To) and J.T. (Jaume Torres); writing—review and editing, J.T. (Janet To) and J.T. (Jaume Torres); funding acquisition, J.T. (Jaume Torres).

Funding

Jaume Torres acknowledges the funding of the Singapore Ministry of Education Tier 1 grant RG134/16.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ritchey, M.B.; Palese, P.; Kilbourne, E.D. RNAs of influenza A, B, and C viruses. J. Virol. 1976, 18, 738–744. [Google Scholar] [PubMed]
  2. Hause, B.M.; Collin, E.A.; Liu, R.; Huang, B.; Sheng, Z.; Lu, W.; Wang, D.; Nelson, E.A.; Li, F. Characterization of a novel influenza virus in cattle and swine: Proposal for a new genus in the Orthomyxoviridae family. MBio 2014, 5. [Google Scholar] [CrossRef] [PubMed]
  3. Hause, B.M.; Ducatez, M.; Collin, E.A.; Ran, Z.; Liu, R.; Sheng, Z.; Armien, A.; Kaplan, B.; Chakravarty, S.; Hoppe, A.D.; et al. Isolation of a novel swine influenza virus from Oklahoma in 2011 which is distantly related to human influenza C viruses. PLoS Pathog. 2013, 9, e1003176. [Google Scholar] [CrossRef] [PubMed]
  4. 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] [PubMed]
  5. 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] [PubMed]
  6. Hay, A.J.; Gregory, V.; Douglas, A.R.; Yi, P.L. The evolution of human influenza viruses. Philos. Trans. R. Soc. B Biol. Sci. 2001, 356, 1861–1870. [Google Scholar] [CrossRef] [Green Version]
  7. Neumann, G.; Noda, T.; Kawaoka, Y. Emergence and pandemic potential of swine-origin H1N1 influenza virus. Nature 2009, 459, 931–939. [Google Scholar] [CrossRef] [Green Version]
  8. Molinari, N.A.M.; Ortega-Sanchez, I.R.; Messonnier, M.L.; Thompson, W.W.; Wortley, P.M.; Weintraub, E.; Bridges, C.B. The annual impact of seasonal influenza in the US: Measuring disease burden and costs. Vaccine 2007, 25, 5086–5096. [Google Scholar] [CrossRef]
  9. Webster, R.G.; Monto, A.S.; Braciale, T.J.; Lamb, R.A. Textbook of Influenza; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2013. [Google Scholar]
  10. Koutsakos, M.; Nguyen, T.H.; Barclay, W.S.; Kedzierska, K. Knowns and unknowns of influenza B viruses. Future Microbiol. 2016, 11, 119–135. [Google Scholar] [CrossRef]
  11. Ran, Z.; Shen, H.; Lang, Y.; Kolb, E.A.; Turan, N.; Zhu, L.; Ma, J.; Bawa, B.; Liu, Q.; Liu, H.; et al. Domestic pigs are susceptible to infection with influenza B viruses. J. Virol. 2015, 89, 4818–4826. [Google Scholar] [CrossRef]
  12. Osterhaus, A.D.M.E.; Rimmelzwaan, G.F.; Martina, B.E.E.; Bestebroer, T.M.; Fouchier, R.A.M. Influenza B virus in seals. Science 2000, 288, 1051–1053. [Google Scholar] [CrossRef] [PubMed]
  13. Ferguson, L.; Olivier, A.K.; Genova, S.; Epperson, W.B.; Smith, D.R.; Schneider, L.; Barton, K.; McCuan, K.; Webby, R.J.; Wan, X.F. Pathogenesis of influenza D virus in cattle. J. Virol. 2016, 90, 5636–5642. [Google Scholar] [CrossRef] [PubMed]
  14. Palese, P.; Shaw, M.L. Orthomyxoviridae: The viruses and their replication. Fields Virol. 2007, 1647–1689. [Google Scholar]
  15. Zebedee, S.L.; Lamb, R.A. Influenza A virus M2 protein: Monoclonal antibody restriction of virus growth and detection of M2 in virions. J. Virol. 1988, 62, 2762–2772. [Google Scholar] [PubMed]
  16. 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]
  17. Kundu, A.; Avalos, R.T.; Sanderson, C.M.; Nayak, D.P. Transmembrane domain of influenza virus neuraminidase, a type II protein, possesses an apical sorting signal in polarized MDCK cells. J. Virol. 1996, 70, 6508–6515. [Google Scholar] [PubMed]
  18. Sakaguchi, T.; Tu, Q.A.; Pinto, L.H.; Lamb, R.A. The active oligomeric state of the minimalistic influenza virus M-2 ion channel is a tetramer. Proc. Natl. Acad. Sci. USA 1997, 94, 5000–5005. [Google Scholar] [CrossRef] [PubMed]
  19. Fields, B.N.; Knipe, D.M.; Howley, P.M. Fields Virology; Wolters Kluwer Health/Lippincott Williams & Wilkins: Philadelphia, PA, USA, 2013. [Google Scholar]
  20. Muraki, Y.; Hongo, S. The molecular virology and reverse genetics of influenza C virus. Jpn. J. Infect. Dis. 2010, 63, 157–165. [Google Scholar]
  21. Castano-Rodriguez, C.; Honrubia, J.M.; Gutierrez-Alvarez, J.; DeDiego, M.L.; Nieto-Torres, J.L.; Jimenez-Guardeno, J.M.; Regla-Nava, J.A.; Fernandez-Delgado, R.; Verdia-Baguena, C.; Queralt-Martin, M.; et al. Role of Severe Acute Respiratory Syndrome Coronavirus Viroporins E, 3a, and 8a in Replication and Pathogenesis. MBio 2018, 9. [Google Scholar] [CrossRef] [Green Version]
  22. Carrasco, L. Modification of membrane permeability by animal viruses. Adv. Virus Res. 1995, 45, 61–112. [Google Scholar]
  23. Nieva, J.L.; Carrasco, L. Viroporins: Structures and functions beyond cell membrane permeabilization. Viruses 2015, 7, 5169–5171. [Google Scholar] [CrossRef] [Green Version]
  24. Hyser, J.M.; Estes, M.K. Pathophysiological Consequences of Calcium-Conducting Viroporins. Annu. Rev. Virol. 2015, 2, 473–496. [Google Scholar] [CrossRef] [PubMed]
  25. Lamb, R.A.; Zebedee, S.L.; Richardson, C.D. Influenza virus M2 protein is an integral membrane protein expressed on the infected-cell surface. Cell 1985, 40, 627–633. [Google Scholar] [CrossRef]
  26. Lamb, R.A.; Choppin, P.W. The gene structure and replication of influenza virus. Annu. Rev. Biochem. 1983, 52, 467–506. [Google Scholar] [CrossRef] [PubMed]
  27. Sugrue, R.J.; Hay, A.J. Structural characteristics of the M2 protein of influenza a viruses: Evidence that it forms a tetrameric channel. Virology 1991, 180, 617–624. [Google Scholar] [CrossRef]
  28. Hull, J.D.; Gilmore, R.; Lamb, R.A. Integration of a small integral membrane protein, M2, of influenza virus into the endoplasmic reticulum: Analysis of the internal signal-anchor domain of a protein with an ectoplasmic NH2 terminus. J. Cell Biol. 1988, 106, 1489–1498. [Google Scholar] [CrossRef] [PubMed]
  29. Helenius, A. Unpacking the incoming influenza virus. Cell 1992, 69, 577–578. [Google Scholar] [CrossRef]
  30. Martin, K.; Helenius, A. Transport of incoming influenza virus nucleocapsids into the nucleus. J. Virol. 1991, 65, 232–244. [Google Scholar] [CrossRef]
  31. Takeuchi, K.; Lamb, R.A. Influenza virus M2 protein ion channel activity stabilizes the native form of fowl plague virus hemagglutinin during intracellular transport. J. Virol. 1994, 68, 911–919. [Google Scholar] [Green Version]
  32. Ciampor, F.; Bayley, P.M.; Nermut, M.V.; Hirst, E.M.A.; Sugrue, R.J.; Hay, A.J. Evidence that the amantadine-induced, M2-mediated conversion of influenza A virus hemagglutinin to the low pH conformation occurs in an acidic trans golgi compartment. Virology 1992, 188, 14–24. [Google Scholar] [CrossRef]
  33. Grambas, S.; Hay, A.J. Maturation of influenza a virus hemagglutinin-Estimates of the pH encountered during transport and its regulation by the M2 protein. Virology 1992, 190, 11–18. [Google Scholar] [CrossRef]
  34. Sugrue, R.J.; Bahadur, G.; Zambon, M.C.; Hall-Smith, M.; Douglas, A.R.; Hay, A.J. Specific structural alteration of the influenza haemagglutinin by amantadine. EMBO J. 1990, 9, 3469–3476. [Google Scholar] [CrossRef] [PubMed]
  35. Ma, C.; Polishchuk, A.L.; Ohigashi, Y.; Stouffer, A.L.; Schön, A.; Magavern, E.; Jing, X.; Lear, J.D.; Freire, E.; Lamb, R.A.; et al. Identification of the functional core of the influenza A virus A/M2 proton-selective ion channel. Proc. Natl. Acad. Sci. USA 2009, 106, 12283–12288. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Pinto, L.H.; Dieckmann, G.R.; Gandhi, C.S.; Papworth, C.G.; Braman, J.; Shaughnessy, M.A.; Lear, J.D.; Lamb, R.A.; Degrado, W.F. A functionally defined model for the M2 proton channel of influenza A virus suggests a mechanism for its ion selectivity. Proc. Natl. Acad. Sci. USA 1997, 94, 11301–11306. [Google Scholar] [CrossRef] [PubMed]
  37. Tang, Y.; Zaitseva, F.; Lamb, R.A.; Pinto, L.H. The gate of the influenza virus M2 proton channel is formed by a single tryptophan residue. J. Biol. Chem. 2002, 277, 39880–39886. [Google Scholar] [CrossRef] [PubMed]
  38. 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]
  39. Hatta, M.; Goto, H.; Kawaoka, Y. Influenza B virus requires BM2 protein for replication. J. Virol. 2004, 78, 5576–5583. [Google Scholar] [CrossRef] [PubMed]
  40. Mould, J.A.; Paterson, R.G.; Takeda, M.; Ohigashi, Y.; Venkataraman, P.; Lamb, R.A.; Pinto, L.H. Influenza B virus BM2 protein has ion channel activity that conducts protons across membranes. Dev. Cell 2003, 5, 175–184. [Google Scholar] [CrossRef]
  41. Paterson, R.G.; Takeda, M.; Ohigashi, Y.; Pinto, L.H.; Lamb, R.A. Influenza B virus BM2 protein is an oligomeric integral membrane protein expressed at the cell surface. Virology 2003, 306, 7–17. [Google Scholar] [CrossRef] [Green Version]
  42. Wang, J.; Pielak, R.M.; McClintock, M.A.; Chou, J.J. Solution structure and functional analysis of the influenza B proton channel. Nat. Struct. Mol. Biol. 2009, 16, 1267–1271. [Google Scholar] [CrossRef]
  43. Ma, C.; Wang, J. Functional studies reveal the similarities and differences between AM2 and BM2 proton channels from influenza viruses. Biochim. Biophys. Acta Biomembr. 2018, 1860, 272–280. [Google Scholar] [CrossRef] [PubMed]
  44. Odagiri, T.; Hong, J.; Ohara, Y. The BM2 protein of influenza B virus is synthesized in the late phase of infection and incorporated into virions as a subviral component. J. Gen. Virol. 1999, 80, 2573–2581. [Google Scholar] [CrossRef] [PubMed]
  45. Sugrue, R.J.; Belshe, R.B.; Hay, A.J. Palmitoylation of the influenza a virus M2 protein. Virology 1990, 179, 51–56. [Google Scholar] [CrossRef]
  46. Holsinger, L.J.; Alams, R. Influenza virus M2 integral membrane protein is a homotetramer stabilized by formation of disulfide bonds. Virology 1991, 183, 32–43. [Google Scholar] [CrossRef]
  47. Thomas, J.M.; Stevens, M.P.; Percy, N.; Barclay, W.S. Phosphorylation of the M2 protein of influenza A virus is not essential for virus viability. Virology 1998, 252, 54–64. [Google Scholar] [CrossRef] [PubMed]
  48. Hutchinson, E.C.; Denham, E.M.; Thomas, B.; Trudgian, D.C.; Hester, S.S.; Ridlova, G.; York, A.; Turrell, L.; Fodor, E. Mapping the phosphoproteome of influenza A and B viruses by mass spectrometry. PLoS Pathog. 2012, 8, e1002993. [Google Scholar] [CrossRef]
  49. Park, E.K.; Castrucci, M.R.; Portner, A.; Kawaoka, Y. The M2 ectodomain is important for its incorporation into influenza A virions. J. Virol. 1998, 72, 2449–2455. [Google Scholar]
  50. Acharya, R.; Carnevale, V.; Fiorin, G.; Levine, B.G.; Polishchuk, A.L.; Balannik, V.; Samish, I.; Lamb, R.A.; Pinto, L.H.; DeGrado, W.F.; et al. Structure and mechanism of proton transport through the transmembrane tetrameric M2 protein bundle of the influenza A virus. Proc. Natl. Acad. Sci. USA 2010, 107, 15075–15080. [Google Scholar] [CrossRef] [Green Version]
  51. Thomaston, J.L.; Polizzi, N.F.; Konstantinidi, A.; Wang, J.; Kolocouris, A.; Degrado, W.F. Inhibitors of the M2 Proton Channel Engage and Disrupt Transmembrane Networks of Hydrogen-Bonded Waters. J. Am. Chem. Soc. 2018, 140, 15219–15226. [Google Scholar] [CrossRef]
  52. Schnell, J.R.; Chou, J.J. Structure and mechanism of the M2 proton channel of influenza A virus. Nature 2008, 451, 591–595. [Google Scholar] [CrossRef] [Green Version]
  53. Wang, J.; Wu, Y.; Ma, C.; Fiorin, G.; Wang, J.; Pinto, L.H.; Lamb, R.A.; Klein, M.L.; DeGrado, W.F. Structure and inhibition of the drug-resistant S31N mutant of the M2 ion channel of influenza A virus. Proc. Natl. Acad. Sci. USA 2013, 110, 1315–1320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Hu, F.; Schmidt-Rohr, K.; Hong, M. NMR detection of pH-dependent histidine-water proton exchange reveals the conduction mechanism of a transmembrane proton channel. J. Am. Chem. Soc. 2012, 134, 3703–3713. [Google Scholar] [CrossRef] [PubMed]
  55. Sharma, M.; Yi, M.; Dong, H.; Qin, H.; Peterson, E.; Busath, D.D.; Zhou, H.X.; Cross, T.A. Insight into the mechanism of the influenza A proton channel from a structure in a lipid bilayer. Science 2010, 330, 509–512. [Google Scholar] [CrossRef] [PubMed]
  56. Cady, S.D.; Schmidt-Rohr, K.; Wang, J.; Soto, C.S.; Degrado, W.F.; Hong, M. Structure of the amantadine binding site of influenza M2 proton channels in lipid bilayers. Nature 2010, 463, 689–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Liang, R.; Li, H.; Swanson, J.M.; Voth, G.A. Multiscale simulation reveals a multifaceted mechanism of proton permeation through the influenza A M2 proton channel. Proc. Natl. Acad. Sci. USA 2014, 111, 9396–9401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Hu, J.; Fu, R.; Nishimura, K.; Zhang, L.; Zhou, H.X.; Busath, D.D.; Vijayvergiya, V.; Cross, T.A. Histidines, heart of the hydrogen ion channel from influenza A virus: Toward an understanding of conductance and proton selectivity. Proc. Natl. Acad. Sci. USA 2006, 103, 6865–6870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Mandala, V.S.; Liao, S.Y.; Kwon, B.; Hong, M. Structural Basis for Asymmetric Conductance of the Influenza M2 Proton Channel Investigated by Solid-State NMR Spectroscopy. J. Mol. Biol. 2017, 429, 2192–2210. [Google Scholar] [CrossRef]
  60. Ma, C.; Fiorin, G.; Carnevale, V.; Wang, J.; Lamb, R.A.; Klein, M.L.; Wu, Y.; Pinto, L.H.; Degrado, W.F. Asp44 stabilizes the Trp41 gate of the M2 proton channel of influenza a virus. Structure 2013, 21, 2033–2041. [Google Scholar] [CrossRef]
  61. Pielak, R.M.; Chou, J.J. Kinetic analysis of the M2 proton conduction of the influenza virus. J. Am. Chem. Soc. 2010, 132, 17695–17697. [Google Scholar] [CrossRef]
  62. Okada, A.; Miura, T.; Takeuchi, H. Protonation of histidine and histidine-tryptophan interaction in the activation of the M2 ion channel from influenza A virus. Biochemistry 2001, 40, 6053–6060. [Google Scholar] [CrossRef]
  63. Colvin, M.T.; Andreas, L.B.; Chou, J.J.; Griffin, R.G. Proton association constants of his 37 in the influenza-A M218-60dimer-of-dimers. Biochemistry 2014, 53, 5987–5994. [Google Scholar] [CrossRef] [PubMed]
  64. Liao, S.Y.; Yang, Y.; Tietze, D.; Hong, M. The Influenza M2 Cytoplasmic Tail Changes the Proton-Exchange Equilibria and the Backbone Conformation of the Transmembrane Histidine Residue to Facilitate Proton Conduction. J. Am. Chem. Soc. 2015, 137, 6067–6077. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Stouffer, A.L.; Acharya, R.; Salom, D.; Levine, A.S.; Di Costanzo, L.; Soto, C.S.; Tereshko, V.; Nanda, V.; Stayrook, S.; DeGrado, W.F. Structural basis for the function and inhibition of an influenza virus proton channel. Nature 2008, 451, 596–599. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Thomaston, J.L.; Alfonso-Prieto, M.; Woldeyes, R.A.; Fraser, J.S.; Klein, M.L.; Fiorin, G.; DeGrado, W.F. High-resolution structures of the M2 channel from influenza A virus reveal dynamic pathways for proton stabilization and transduction. Proc. Natl. Acad. Sci. USA 2015, 112, 14260–14265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Thomaston, J.L.; DeGrado, W.F. Crystal structure of the drug-resistant S31N influenza M2 proton channel. Protein Sci. 2016, 1551–1554. [Google Scholar] [CrossRef] [PubMed]
  68. Thomaston, J.L.; Woldeyes, R.A.; Nakane, T.; Yamashita, A.; Tanaka, T.; Koiwai, K.; Brewster, A.S.; Barad, B.A.; Chen, Y.; Lemmin, T.; et al. XFEL structures of the influenza M2 proton channel: Room temperature water networks and insights into proton conduction. Proc. Natl. Acad. Sci. USA 2017, 114, 13357–13362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Agmon, N. The Grotthuss mechanism. Chem. Phys. Lett. 1995, 244, 456–462. [Google Scholar] [CrossRef]
  70. Liang, R.; Swanson, J.M.J.; Madsen, J.J.; Hong, M.; DeGrado, W.F.; Voth, G.A. Acid activation mechanism of the influenza A M2 proton channel. Proc. Natl. Acad. Sci. USA 2016, 113, E6955–E6964. [Google Scholar] [CrossRef] [Green Version]
  71. Lin, T.; Schroeder, C. Definitive assignment of proton selectivity and attoampere unitary current to the M2 ion channel protein of influenza A virus. J. Virol. 2001, 75, 3647–3656. [Google Scholar] [CrossRef]
  72. Chizhmakov, I.V.; Ogden, D.C.; Geraghty, F.M.; Hayhurst, A.; Skinner, A.; Betakova, T.; Hay, A.J. Differences in conductance of M2 proton channels of two influenza viruses at low and high pH. J. Physiol. 2003, 546, 427–438. [Google Scholar] [CrossRef]
  73. Mould, J.A.; Li, H.C.; Dudlak, C.S.; Lear, J.D.; Pekosz, A.; Lamb, R.A.; Pinto, L.H. Mechanism for proton conduction of the M2 ion channel of influenza A virus. J. Biol. Chem. 2000, 275, 8592–8599. [Google Scholar] [CrossRef] [PubMed]
  74. Moffat, J.C.; Vijayvergiya, V.; Gao, P.F.; Cross, T.A.; Woodbury, D.J.; Busath, D.D. Proton transport through influenza A virus M2 protein reconstituted in vesicles. Biophys. J. 2008, 94, 434–445. [Google Scholar] [CrossRef] [PubMed]
  75. Decoursey, T.E. Voltage-gated proton channels and other proton transfer pathways. Physiol. Rev. 2003, 83, 475–579. [Google Scholar] [CrossRef] [PubMed]
  76. Hu, F.; Luo, W.; Hong, M. Mechanisms of proton conduction and gating in influenza M2 proton channels from solid-state NMR. Science 2010, 330, 505–508. [Google Scholar] [CrossRef] [PubMed]
  77. Williams, J.K.; Zhang, Y.; Schmidt-Rohr, K.; Hong, M. PH-dependent conformation, dynamics, and aromatic interaction of the gating tryptophan residue of the influenza M2 proton channel from solid-state NMR. Biophys. J. 2013, 104, 1698–1708. [Google Scholar] [CrossRef] [PubMed]
  78. DiFrancesco, M.L.; Hansen, U.P.; Thiel, G.; Moroni, A.; Schroeder, I. Effect of cytosolic pH on inward currents reveals structural characteristics of the proton transport cycle in the influenza A protein M2 in cell-free membrane patches of Xenopus oocytes. PLoS ONE 2014, 9. [Google Scholar] [CrossRef]
  79. Zhou, H.X. A theory for the proton transport of the influenza virus M2 protein: Extensive test against conductance data. Biophys. J. 2011, 100, 912–921. [Google Scholar] [CrossRef] [PubMed]
  80. Markiewicz, B.N.; Lemmin, T.; Zhang, W.; Ahmed, I.A.; Jo, H.; Fiorin, G.; Troxler, T.; DeGrado, W.F.; Gai, F. Infrared and fluorescence assessment of the hydration status of the tryptophan gate in the influenza A M2 proton channel. PCCP 2016, 18, 28939–28950. [Google Scholar] [CrossRef] [Green Version]
  81. Lin, C.W.; Mensa, B.; Barniol-Xicota, M.; DeGrado, W.F.; Gai, F. Activation pH and Gating Dynamics of Influenza A M2 Proton Channel Revealed by Single-Molecule Spectroscopy. Angew. Chem. Int. Ed. Engl. 2017, 56, 5283–5287. [Google Scholar] [CrossRef]
  82. Miao, Y.; Fu, R.; Zhou, H.X.; Cross, T.A. Dynamic Short Hydrogen Bonds in Histidine Tetrad of Full-Length M2 Proton Channel Reveal Tetrameric Structural Heterogeneity and Functional Mechanism. Structure 2015, 23, 2300–2308. [Google Scholar] [CrossRef] [Green Version]
  83. Li, C.; Qin, H.; Gao, F.P.; Cross, T.A. Solid-state NMR characterization of conformational plasticity within the transmembrane domain of the influenza A M2 proton channel. Biochim. Biophys. Acta 2007, 1768, 3162–3170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Yi, M.; Cross, T.A.; Zhou, H.X. Conformational heterogeneity of the M2 proton channel and a structural model for channel activation. Proc. Natl. Acad. Sci. USA 2009, 106, 13311–13316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Khurana, E.; Dal Peraro, M.; DeVane, R.; Vemparala, S.; DeGrado, W.F.; Klein, M.L. Molecular dynamics calculations suggest a conduction mechanism for the M2 proton channel from influenza A virus. Proc. Natl. Acad. Sci. USA 2009, 106, 1069–1074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Wei, C.; Pohorille, A. M2 Proton Channel: Toward a Model of a Primitive Proton Pump. Orig. Life Evol. Biosph. 2015, 45, 241–248. [Google Scholar] [CrossRef] [PubMed]
  87. Ghosh, A.; Qiu, J.; DeGrado, W.F.; Hochstrasser, R.M. Tidal surge in the M2 proton channel, sensed by 2D IR spectroscopy. Proc. Natl. Acad. Sci. USA 2011, 108, 6115–6120. [Google Scholar] [CrossRef] [Green Version]
  88. Ma, C.; Soto, C.S.; Ohigashi, Y.; Taylor, A.; Bournas, V.; Glawe, B.; Udo, M.K.; DeGrado, W.F.; Lamb, R.A.; Pinto, L.H. Identification of the pore-lining residues of the BM2 ion channel protein of influenza B virus. J. Biol. Chem. 2008, 283, 15921–15931. [Google Scholar] [CrossRef] [PubMed]
  89. Williams, J.K.; Tietze, D.; Lee, M.; Wang, J.; Hong, M. Solid-State NMR Investigation of the Conformation, Proton Conduction, and Hydration of the Influenza B Virus M2 Transmembrane Proton Channel. J. Am. Chem. Soc. 2016, 138, 8143–8155. [Google Scholar] [CrossRef] [Green Version]
  90. Williams, J.K.; Shcherbakov, A.A.; Wang, J.; Hong, M. Protonation equilibria and pore-opening structure of the dual-histidine influenza B virus M2 transmembrane proton channel from solid-state NMR. J. Biol. Chem. 2017, 292, 17876–17884. [Google Scholar] [CrossRef] [Green Version]
  91. Hongo, S.; Sugawara, K.; Nishimura, H.; Muraki, Y.; Kitame, F.; Nakamura, K. Identification of a second protein encoded by influenza C virus RNA segment 6. J. Gen. Virol. 1994, 75, 3503–3510. [Google Scholar] [CrossRef]
  92. Hongo, S.; Ishii, K.; Mori, K.; Takashita, E.; Muraki, Y.; Matsuzaki, Y.; Sugawara, K. Detection of ion channel activity in Xenopus laevis oocytes expressing Influenza C virus CM2 protein. Arch. Virol. 2004, 149, 35–50. [Google Scholar] [CrossRef]
  93. Pekosz, A.; Lamb, R.A. Influenza C virus CM2 integral membrane glycoprotein is produced from a polypeptide precursor by cleavage of an internal signal sequence. Proc. Natl. Acad. Sci. USA 1998, 95, 13233–13238. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Hongo, S.; Sugawara, K.; Muraki, Y.; Matsuzaki, Y.; Takashita, E.; Kitame, F.; Nakamura, K. Influenza C virus CM2 protein is produced from a 374-amino-acid protein (P42) by signal peptidase cleavage. J. Virol. 1999, 73, 46–50. [Google Scholar] [PubMed]
  95. Pekosz, A.; Lamb, R.A. The CM2 protein of influenza C virus is an oligomeric integral membrane glycoprotein structurally analogous to influenza A virus M2 and influenza B virus NB proteins. Virology 1997, 237, 439–451. [Google Scholar] [CrossRef] [PubMed]
  96. Hongo, S.; Sugawara, K.; Muraki, Y.; Kitame, F.; Nakamura, K. Characterization of a second protein (CM2) encoded by RNA segment 6 of influenza C virus. J. Virol. 1997, 71, 2786–2792. [Google Scholar] [PubMed]
  97. Muraki, Y.; Okuwa, T.; Himeda, T.; Hongo, S.; Ohara, Y. Effect of Cysteine Mutations in the Extracellular Domain of CM2 on the Influenza C Virus Replication. PLoS ONE 2013, 8. [Google Scholar] [CrossRef] [PubMed]
  98. Tada, Y.; Hongo, S.; Muraki, Y.; Matsuzaki, Y.; Sugawara, K.; Kitame, F.; Nakamura, K. Phosphorylation of influenza C virus CM2 protein. Virus Res. 1998, 58, 65–72. [Google Scholar] [CrossRef]
  99. Kukol, A.; Arkin, I.T. Structure of the Influenza C virus CM2 protein transmembrane domain obtained by site-specific infrared dichroism and global molecular dynamics searching. J. Biol. Chem. 2000, 275, 4225–4229. [Google Scholar] [CrossRef] [PubMed]
  100. Muraki, Y.; Washioka, H.; Sugawara, K.; Matsuzaki, Y.; Takashita, E.; Hongo, S. Identification of an amino acid residue on influenza C virus M1 protein responsible for formation of the cord-like structures of the virus. J. Gen. Virol. 2004, 85, 1885–1893. [Google Scholar] [CrossRef] [PubMed]
  101. Torres, J.; Kukol, A.; Arkin, I.T. Mapping the energy surface of transmembrane helix-helix interactions. Biophys. J. 2001, 81, 2681–2692. [Google Scholar] [CrossRef]
  102. Betakova, T.; Hay, A.J. Evidence that the CM2 protein of influenza C virus can modify the pH of the exocytic pathway of transfected cells. J. Gen. Virol. 2007, 88, 2291–2296. [Google Scholar] [CrossRef]
  103. Stewart, S.M.; Pekosz, A. The influenza C virus CM2 protein can alter intracellular pH, and: Its transmembrane domain can substitute for that of the influenza a virus M2 protein and support infectious virus production. J. Virol. 2012, 86, 1277–1281. [Google Scholar] [CrossRef] [PubMed]
  104. Ohuchi, M.; Ohuchi, R.; Mifune, K. Demonstration of hemolytic and fusion activities of influenza C virus. J. Virol. 1982, 42, 1076–1079. [Google Scholar] [PubMed]
  105. Chizhmakov, I.V.; Geraghty, F.M.; Ogden, D.C.; Hayhurst, A.; Antoniou, M.; Hay, A.J. Selective proton permeability and pH regulation of the influenza virus M2 channel expressed in mouse erythroleukaemia cells. J. Physiol. 1996, 494, 329–336. [Google Scholar] [CrossRef] [PubMed]
  106. Mould, J.A.; Drury, J.E.; Fring, S.M.; Kaupp, U.B.; Pekosz, A.; Lamb, R.A.; Pinto, L.H. Permeation and activation of the M2 ion channel of influenza A virus. J. Biol. Chem. 2000, 275, 31038–31050. [Google Scholar] [CrossRef] [PubMed]
  107. Betáková, T.; Kollerová, E. pH modulating activity of ion channels of influenza A, B, and C viruses. Acta Virol. 2006, 50, 187–193. [Google Scholar] [PubMed]
  108. Zhirnov, O.P.; Grigoriev, V.B. Disassembly of influenza C viruses, distinct from that of influenza A and B viruses requires neutral-alkaline pH. Virology 1994, 200, 284–291. [Google Scholar] [CrossRef] [PubMed]
  109. Furukawa, T.; Muraki, Y.; Noda, T.; Takashita, E.; Sho, R.; Sugawara, K.; Matsuzaki, Y.; Shimotai, Y.; Hongo, S. Role of the CM2 protein in the influenza C virus replication cycle. J. Virol. 2011, 85, 1322–1329. [Google Scholar] [CrossRef]
  110. Kesinger, E.; Liu, J.; Jensen, A.; Chia, C.P.; Demers, A.; Moriyama, H. Influenza D virus M2 protein exhibits ion channel activity in Xenopus laevis oocytes. PLoS ONE 2018, 13, e0199227. [Google Scholar] [CrossRef]
  111. Henkel, M.; Mitzner, D.; Henklein, P.; Meyer-Almes, F.J.; Moroni, A.; DiFrancesco, M.L.; Henkes, L.M.; Kreim, M.; Kast, S.M.; Schubert, U.; et al. Proapoptotic influenza A virus protein PB1-F2 forms a nonselective ion channel. PLoS ONE 2010, 5. [Google Scholar] [CrossRef]
  112. 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]
  113. Zamarin, D.; García-Sastre, A.; Xiao, X.; Wang, R.; Palese, P. Influenza virus PB1-F2 protein induces cell death through mitochondrial ANT3 and VDAC1. PLoS Path. 2005, 1, 0040–0054. [Google Scholar] [CrossRef] [PubMed]
  114. 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] [PubMed]
  115. McAuley, J.L.; Tate, M.D.; MacKenzie-Kludas, C.J.; Pinar, A.; Zeng, W.; Stutz, A.; Latz, E.; Brown, L.E.; Mansell, A. Activation of the NLRP3 Inflammasome by IAV Virulence Protein PB1-F2 Contributes to Severe Pathophysiology and Disease. PLoS Path. 2013, 9. [Google Scholar] [CrossRef] [PubMed]
  116. Williams, M.A.; Lamb, R.A. Effects of mutations and deletions in a bicistronic mRNA on the synthesis of influenza B virus NB and NA glycoproteins. J. Virol. 1989, 63, 28–35. [Google Scholar] [PubMed]
  117. Demers, A.; Ran, Z.; Deng, Q.; Wang, D.; Edman, B.; Lu, W.; Li, F. Palmitoylation is required for intracellular trafficking of influenza B virus NB protein and efficient influenza B virus growth in vitro. J. Gen. Virol. 2014, 95, 1211–1220. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Sunstrom, N.A.; Premkumar, L.S.; Premkumar, A.; Ewart, G.; Cox, G.B.; Gage, P.W. Ion channels formed by NB, an influenza B virus protein. J. Membr. Biol. 1996, 150, 127–132. [Google Scholar] [CrossRef] [PubMed]
  119. Premkumar, A.; Ewart, G.D.; Cox, G.B.; Gage, P.W. An amino-acid substitution in the influenza-B NB protein affects ion-channel gating. J. Membr. Biol. 2004, 197, 135–143. [Google Scholar] [CrossRef]
  120. Hatta, M.; Kawaoka, Y. The NB protein of influenza B virus is not necessary for virus replication in vitro. J. Virol. 2003, 77, 6050–6054. [Google Scholar] [CrossRef]
  121. Pinto, L.H.; Holsinger, L.J.; Lamb, R.A. Influenza virus M2 protein has ion channel activity. Cell 1992, 69, 517–528. [Google Scholar] [CrossRef]
  122. Davies, W.L.; Grunert, R.R.; Haff, R.F.; McGahen, J.W.; Neumayer, E.M.; Paulshock, M.; Watts, J.C.; Wood, T.R.; Hermann, E.C.; Hoffmann, C.E. Antiviral Activity of 1-Adamantanamine (Amantadine). Science 1964, 144, 862–863. [Google Scholar] [CrossRef]
  123. Pielak, R.M.; Oxenoid, K.; Chou, J.J. Structural Investigation of Rimantadine Inhibition of the AM2-BM2 Chimera Channel of Influenza Viruses. Structure 2011, 19, 1655–1663. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Hay, A.J.; Wolstenholme, A.J.; Skehel, J.J.; Smith, M.H. The molecular basis of the specific anti-influenza action of amantadine. EMBO J. 1985, 4, 3021–3024. [Google Scholar] [CrossRef] [PubMed]
  125. Vanderlinden, E.; Naesens, L. Emerging Antiviral Strategies to Interfere with Influenza Virus Entry. Med. Res. Rev. 2014, 34, 301–339. [Google Scholar] [CrossRef] [PubMed]
  126. Deyde, V.M.; Xu, X.; Bright, R.A.; Shaw, M.; Smith, C.B.; Zhang, Y.; Shu, Y.; Gubareva, L.V.; Cox, N.J.; Klimov, A.I. Surveillance of resistance to adamantanes among influenza A(H3N2) and A(H1N1) viruses isolated worldwide. J. Infect. Dis. 2007, 196, 249–257. [Google Scholar] [CrossRef] [PubMed]
  127. 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]
  128. Bright, R.A.; Medina, M.J.; Xu, X.; Perez-Oronoz, G.; Wallis, T.R.; Davis, X.M.; Povinelli, L.; Cox, N.J.; Klimov, A.I. Incidence of adamantane resistance among influenza A (H3N2) viruses isolated worldwide from 1994 to 2005: A cause for concern. Lancet 2005, 366, 1175–1181. [Google Scholar] [CrossRef]
  129. Hayden, F.G.; De Jong, M.D. Emerging influenza antiviral resistance threats. J. Infect. Dis. 2011, 203, 6–10. [Google Scholar] [CrossRef] [PubMed]
  130. Fiore, A.E.; Fry, A.; Shay, D.; Gubareva, L.; Bresee, J.S.; Uyeki, T.M.; Centers for Disease Control Prevention. Antiviral agents for the treatment and chemoprophylaxis of influenza—Recommendations of the Advisory Committee on Immunization Practices (ACIP). MMWR Recomm. Rep. 2011, 60, 1–24. [Google Scholar]
  131. Li, F.; Ma, C.; DeGrado, W.F.; Wang, J. Discovery of Highly Potent Inhibitors Targeting the Predominant Drug-Resistant S31N Mutant of the Influenza A Virus M2 Proton Channel. J. Med. Chem. 2016, 59, 1207–1216. [Google Scholar] [CrossRef] [Green Version]
  132. Li, F.; Ma, C.; Hu, Y.; Wang, Y.; Wang, J. Discovery of Potent Antivirals against Amantadine-Resistant Influenza A Viruses by Targeting the M2-S31N Proton Channel. ACS Infect. Dis. 2016, 2, 726–733. [Google Scholar] [CrossRef] [Green Version]
  133. Ma, C.; Zhang, J.; Wang, J. Pharmacological Characterization of the Spectrum of Antiviral Activity and Genetic Barrier to Drug Resistance of M2-S31N Channel Blockers. Mol. Pharmacol. 2016, 90, 188–198. [Google Scholar] [CrossRef] [PubMed]
  134. Hu, Y.; Musharrafieh, R.; Ma, C.; Zhang, J.; Smee, D.F.; DeGrado, W.F.; Wang, J. An M2-V27A channel blocker demonstrates potent in vitro and in vivo antiviral activities against amantadine-sensitive and -resistant influenza A viruses. Antivir. Res. 2017, 140, 45–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Dong, G.; Peng, C.; Luo, J.; Wang, C.; Han, L.; Wu, B.; Ji, G.; He, H. Adamantane-resistant influenza a viruses in the world (1902-2013): Frequency and distribution of M2 gene mutations. PLoS ONE 2015, 10. [Google Scholar] [CrossRef] [PubMed]
  136. Du, J.; Cross, T.A.; Zhou, H.X. Recent progress in structure-based anti-influenza drug design. Drug Discov. Today 2012, 17, 1111–1120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Gu, R.X.; Liu, L.A.; Wei, D.Q. Structural and energetic analysis of drug inhibition of the influenza A M2 proton channel. Trends Pharmacol. Sci. 2013, 34, 571–580. [Google Scholar] [CrossRef]
  138. To, J.; Surya, W.; Torres, J. Targeting the Channel Activity of Viroporins. Adv. Prot. Chem. Struct. Biol. 2016, 104, 307–355. [Google Scholar]
  139. Gandhi, C.S.; Shuck, K.; Lear, J.D.; Dieckmann, G.R.; DeGrado, W.F.; Lamb, R.A.; Pinto, L.H. Cu(II) inhibition of the proton translocation machinery of the influenza A virus M 2 protein. J. Biol. Chem. 1999, 274, 5474–5482. [Google Scholar] [CrossRef]
  140. Su, Y.; Hu, F.; Hong, M. Paramagnetic Cu(II) for probing membrane protein structure and function: Inhibition mechanism of the influenza M2 proton channel. J. Am. Chem. Soc. 2012, 134, 8693–8702. [Google Scholar] [CrossRef]
  141. Gordon, N.A.; McGuire, K.L.; Wallentine, S.K.; Mohl, G.A.; Lynch, J.D.; Harrison, R.G.; Busath, D.D. Divalent copper complexes as influenza A M2 inhibitors. Antivir. Res. 2017, 147, 100–106. [Google Scholar] [CrossRef]
  142. Ichinohe, T.; Pang, I.K.; Iwasaki, A. Influenza virus activates inflammasomes via its intracellular M2 ion channel. Nat. Immunol. 2010, 11, 404–410. [Google Scholar] [CrossRef]
  143. Lazrak, A.; Iles, K.E.; Liu, G.; Noah, D.L.; Noah, J.W.; Matalon, S. Influenza virus M2 protein inhibits epithelial sodium channels by increasing reactive oxygen species. FASEB J. 2009, 23, 3829–3842. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Londino, J.D.; Lazrak, A.; Jurkuvenaite, A.; Collawn, J.F.; Noah, J.W.; Matalon, S. Influenza matrix protein 2 alters CFTR expression and function through its ion channel activity. J. Physiol. Lung Cell. Mol. Physiol. 2013, 304, L582–L592. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Matalon, S.; O’Brodovich, H. Sodium channels in alveolar epithelial cells: Molecular characterization, biophysical properties, and physiological significance. Annu. Rev. Physiol. 1999, 61, 627–661. [Google Scholar] [CrossRef] [PubMed]
  146. Rogan, M.P.; Stoltz, D.A.; Hornick, D.B. Cystic fibrosis transmembrane conductance regulator intracellular processing, trafficking, and opportunities for mutation-specific treatment. Chest 2011, 139, 1480–1490. [Google Scholar] [CrossRef] [PubMed]
  147. Londino, J.D.; Lazrak, A.; Collawn, J.F.; Bebok, Z.; Harrod, K.S.; Matalon, S. Influenza virus infection alters ion channel function of airway and alveolar cells: Mechanisms and physiological sequelae. J. Physiol. Lung Cell. Mol. Physiol. 2017, 313, L845–L858. [Google Scholar] [CrossRef]
  148. To, J.; Torres, J. Beyond Channel Activity: Protein-Protein Interactions Involving Viroporins. Sub-Cell. Biochem. 2018, 88, 329–377. [Google Scholar] [CrossRef]
  149. Gannagé, M.; Dormann, D.; Albrecht, R.; Dengjel, J.; Torossi, T.; Rämer, P.C.; Lee, M.; Strowig, T.; Arrey, F.; Conenello, G.; et al. Matrix Protein 2 of Influenza A Virus Blocks Autophagosome Fusion with Lysosomes. Cell Host Microbe 2009, 6, 367–380. [Google Scholar] [CrossRef] [Green Version]
  150. Bourmakina, S.V.; García-Sastre, A. Reverse genetics studies on the filamentous morphology of influenza a virus. J. Gen. Virol. 2003, 84, 517–527. [Google Scholar] [CrossRef]
  151. Beale, R.; Wise, H.; Stuart, A.; Ravenhill, B.J.; Digard, P.; Randow, F. A LC3-interacting motif in the influenza A virus M2 protein is required to subvert autophagy and maintain virion stability. Cell Host Microbe 2014, 15, 239–247. [Google Scholar] [CrossRef]
  152. Meurs, E.; Chong, K.; Galabru, J.; Thomas, N.S.B.; Kerr, I.M.; Williams, B.R.G.; 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]
  153. 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] [PubMed] [Green Version]
  154. Sharma, K.; Tripathi, S.; Ranjan, P.; Kumar, P.; Garten, R.; Deyde, V.; Katz, J.M.; Cox, N.J.; Lal, R.B.; Sambhara, S.; et al. Influenza a virus nucleoprotein exploits Hsp40 to inhibit PKR activation. PLoS ONE 2011, 6. [Google Scholar] [CrossRef] [PubMed]
  155. Guan, Z.; Liu, D.; Mi, S.; Zhang, J.; Ye, Q.; Wang, M.; Gao, G.F.; Yan, J. Interaction of Hsp40 with influenza virus M2 protein: Implications for PKR signaling pathway. Protein Cell 2010, 1, 944–955. [Google Scholar] [CrossRef] [PubMed]
  156. Le Tortorec, A.; Willey, S.; Neil, S.J.D. Antiviral inhibition of enveloped virus release by Tetherin/BST-2: Action and counteraction. Viruses 2011, 3, 520–540. [Google Scholar] [CrossRef] [PubMed]
  157. 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]
  158. Ma, H.; Kien, F.; Manière, M.; Zhang, Y.; Lagarde, N.; Tse, K.S.; Poon, L.L.M.; Nal, B. Human annexin A6 interacts with influenza a virus protein M2 and negatively modulates infection. J. Virol. 2012, 86, 1789–1801. [Google Scholar] [CrossRef] [PubMed]
  159. Fan, Y.; Mok, C.K.P.; Chan, M.C.W.; Zhang, Y.; Nal, B.; Kien, F.; Bruzzone, R.; Sanyal, S. Cell Cycle-independent Role of Cyclin D3 in Host Restriction of Influenza Virus Infection. J. Biol. Chem. 2017, 292, 5070–5088. [Google Scholar] [CrossRef] [Green Version]
  160. Rudick, M.; Anderson, R.G.W. Multiple functions of caveolin-1. J. Biol. Chem. 2002, 277, 41295–41298. [Google Scholar]
  161. Sun, L.; Hemgård, G.V.; Susanto, S.A.; Wirth, M. Caveolin-1 influences human influenza A virus (H1N1) multiplication in cell culture. Virol. J. 2010, 7. [Google Scholar] [CrossRef]
  162. Couet, J.; Li, S.; Okamoto, T.; Ikezu, T.; Lisanti, M.P. Identification of peptide and protein ligands for the caveolin- scaffolding domain. Implications for the interaction of caveolin with caveolae-associated proteins. J. Biol. Chem. 1997, 272, 6525–6533. [Google Scholar] [CrossRef]
  163. Amarelle, L.; Lecuona, E. The antiviral effects of na,K-ATPase inhibition: A minireview. Int. J. Mol. Sci. 2018, 19, 2154. [Google Scholar] [CrossRef] [PubMed]
  164. Chan, M.C.W.; Kuok, D.I.T.; Leung, C.Y.H.; Hui, K.P.Y.; Valkenburg, S.A.; Lau, E.H.Y.; Nicholls, J.M.; Fang, X.; Guan, Y.; Lee, J.W.; et al. Human mesenchymal stromal cells reduce influenza A H5N1-associated acute lung injury in vitro and in vivo. Proc. Natl. Acad. Sci. USA 2016, 113, 3621–3626. [Google Scholar] [CrossRef] [Green Version]
  165. Hoffmann, H.H.; Palese, P.; Shaw, M.L. Modulation of influenza virus replication by alteration of sodium ion transport and protein kinase C activity. Antivir. Res. 2008, 80, 124–134. [Google Scholar] [CrossRef] [PubMed]
  166. Dowall, S.D.; Bewley, K.; Watson, R.J.; Vasan, S.S.; Ghosh, C.; Konai, M.M.; Gausdal, G.; Lorens, J.B.; Long, J.; Barclay, W.; et al. Antiviral screening of multiple compounds against ebola virus. Viruses 2016, 8, 277. [Google Scholar] [CrossRef] [PubMed]
  167. Mi, S.F.; Li, Y.; Yan, J.H.; Gao, G.F. Na+/K+-ATPase β1 subunit interacts with M2 proteins of influenza A and B viruses and affects the virus replication. Sci. China Life. Sci. 2010, 53, 1098–1105. [Google Scholar] [CrossRef] [PubMed]
  168. Doms, R.W.; Lamb, R.A.; Rose, J.K.; Helenius, A. Folding and assembly of viral membrane proteins. Virology 1993, 193, 545–562. [Google Scholar] [CrossRef] [PubMed]
  169. Hughey, P.G.; Compans, R.W.; Zebedee, S.L.; Lamb, R.A. Expression of the influenza A virus M2 protein Is restricted to apical surfaces of polarized epithelial cells. J. Virol. 1992, 66, 5542–5552. [Google Scholar] [PubMed]
  170. Tasaki, T.; Mulder, L.C.F.; Iwamatsu, A.; Lee, M.J.; Davydov, I.V.; Varshavsky, A.; Muesing, M.; Kwon, Y.T. A family of mammalian E3 ubiquitin ligases that contain the UBR box motif and recognize N-degrons. Mol. Cell. Biol. 2005, 25, 7120–7136. [Google Scholar] [CrossRef]
  171. Tripathi, S.; Pohl, M.O.; Zhou, Y.; Rodriguez-Frandsen, A.; Wang, G.; Stein, D.A.; Moulton, H.M.; Dejesus, P.; Che, J.; Mulder, L.C.F.; et al. Meta- and Orthogonal Integration of Influenza “oMICs“ Data Defines a Role for UBR4 in Virus Budding. Cell Host Microbe 2015, 18, 723–735. [Google Scholar] [CrossRef]
  172. 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]
  173. Chen, W.; Feng, Y.; Chen, D.; Wandinger-Ness, A. Rab11 is required for trans-Golgi network-to-plasma membrane transport and a preferential target for GDP dissociation inhibitor. Mol. Biol. Cell 1998, 9, 3241–3257. [Google Scholar] [CrossRef] [PubMed]
  174. Ullrich, O.; Reinsch, S.; Urbé, S.; Zerial, M.; Parton, R.G. Rab11 regulates recycling through the pericentriolar recycling endosome. J. Cell Biol. 1996, 135, 913–924. [Google Scholar] [CrossRef] [PubMed]
  175. Urbé, S.; Huber, L.A.; Zerial, M.; Tooze, S.A.; Parton, R.G. Rab11, a small GTPase associated with both constitutive and regulated secretory pathways in PC12 cells. FEBS Lett. 1993, 334, 175–182. [Google Scholar] [CrossRef] [Green Version]
  176. Rossman, J.S.; Jing, X.; Leser, G.P.; Lamb, R.A. Influenza Virus M2 Protein Mediates ESCRT-Independent Membrane Scission. Cell 2010, 142, 902–913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Zhu, P.; Liang, L.; Shao, X.; Luo, W.; Jiang, S.; Zhao, Q.; Sun, N.; Zhao, Y.; Li, J.; Wang, J.; et al. Host cellular protein TRAPPC6AΔ interacts with influenza A virus M2 protein and regulates viral propagation by modulating M2 trafficking. J. Virol. 2017, 91. [Google Scholar] [CrossRef] [PubMed]
  178. Sacher, M.; Jiang, Y.; Barrowman, J.; Scarpa, A.; Burston, J.; Zhang, L.; Schieltz, D.; Yates Iii, J.R.; Abeliovich, H.; Ferro-Novick, S. TRAPP, a highly conserved novel complex on the cis-Golgi that mediates vesicle docking and fusion. EMBO J. 1998, 17, 2494–2503. [Google Scholar] [CrossRef] [PubMed]
  179. Heaton, N.S.; Moshkina, N.; Fenouil, R.; Gardner, T.J.; Aguirre, S.; Shah, P.S.; Zhao, N.; Manganaro, L.; Hultquist, J.F.; Noel, J.; et al. Targeting Viral Proteostasis Limits Influenza Virus, HIV, and Dengue Virus Infection. Immunity 2016, 44, 46–58. [Google Scholar] [CrossRef] [Green Version]
  180. 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. [Google Scholar] [CrossRef]
  181. Watanabe, T.; Kawakami, E.; Shoemaker, J.E.; Lopes, T.J.S.; Matsuoka, Y.; Tomita, Y.; Kozuka-Hata, H.; Gorai, T.; Kuwahara, T.; Takeda, E.; et al. Influenza virus-host interactome screen as a platform for antiviral drug development. Cell Host Microbe 2014, 16, 795–805. [Google Scholar] [CrossRef]
Figure 1. Sequence alignment of M2 viroporins in influenza. Influenza A M2 (AM2; strain A/Udorn/1972 H3N2), influenza B M2 (BM2; strain B/Taiwan/70061/2006), influenza C M2 (CM2; strain C/Ann Arbor/1/1950) and influenza D M2 (DM2; strain D/swine/Oklahoma/1334/2011). The predicted transmembrane regions are underlined. The functional motifs HxxxW (in AM2 and BM2) and YxxxK (in CM2 and DM2) are indicated in bold red font. Numbering corresponds to AM2. Sequences were retrieved from UniProt (www.uniprot.org).
Figure 1. Sequence alignment of M2 viroporins in influenza. Influenza A M2 (AM2; strain A/Udorn/1972 H3N2), influenza B M2 (BM2; strain B/Taiwan/70061/2006), influenza C M2 (CM2; strain C/Ann Arbor/1/1950) and influenza D M2 (DM2; strain D/swine/Oklahoma/1334/2011). The predicted transmembrane regions are underlined. The functional motifs HxxxW (in AM2 and BM2) and YxxxK (in CM2 and DM2) are indicated in bold red font. Numbering corresponds to AM2. Sequences were retrieved from UniProt (www.uniprot.org).
Cells 08 00654 g001
Figure 2. Acid activation mechanism of the AM2 channel. Left: At a high pH (e.g., 7 to 8), the AM2 channel adopts a Cclosed conformation. The closed Trp41 tetrad dehydrates the His37 tetrad and raises the His37 deprotonation barrier, thereby blocking proton conduction. The low charge state of the His37 tetrad at a high pH reduces the electrostatic repulsion with incoming protons, allowing proton permeation from the viral exterior. Right: At a low pH (below 6), the positive charge on the His37 tetrad increases and the Trp41 gate and C-terminal open and become more hydrated, lowering the His37 deprotonation barrier and increasing proton conductance, thereby leading to channel activation in the Copen conformation. Scheme adapted from [70].
Figure 2. Acid activation mechanism of the AM2 channel. Left: At a high pH (e.g., 7 to 8), the AM2 channel adopts a Cclosed conformation. The closed Trp41 tetrad dehydrates the His37 tetrad and raises the His37 deprotonation barrier, thereby blocking proton conduction. The low charge state of the His37 tetrad at a high pH reduces the electrostatic repulsion with incoming protons, allowing proton permeation from the viral exterior. Right: At a low pH (below 6), the positive charge on the His37 tetrad increases and the Trp41 gate and C-terminal open and become more hydrated, lowering the His37 deprotonation barrier and increasing proton conductance, thereby leading to channel activation in the Copen conformation. Scheme adapted from [70].
Cells 08 00654 g002
Table 1. Properties of influenza viroporins. PB1-F2 and NB are much less studied compared to M2 proteins, and although studies have been performed, the precise selectivity is not yet clear. The YxxxK motifs in CM2 and DM2 are predicted to be the functional motifs of these channels, although the exact mechanism of action remains to be elucidated.
Table 1. Properties of influenza viroporins. PB1-F2 and NB are much less studied compared to M2 proteins, and although studies have been performed, the precise selectivity is not yet clear. The YxxxK motifs in CM2 and DM2 are predicted to be the functional motifs of these channels, although the exact mechanism of action remains to be elucidated.
Influenza TypeHost RangeViroporinCoding RNATypical LengthFunctional MotifIon Conductance
AHumans, aquatic birds, porcine, bovine, canineAM2
PB1-F2
Segment 7
Segment 2
97 aa
90 aa
HxxxW
--
H+, K+
Ca2+, MVC
BHumans, pigs, harbor sealsBM2
NB
Segment 7
Segment 6
109 aa
100 aa
HxxxW
--
H+, K+
ND
CHumans, pigsCM2Segment 6115 aaYxxxKCl
DCattle, pigsDM2Segment 6152 aaYxxxKCl
Abbreviations: MVC, monovalent cation; ND, not determined.

Share and Cite

MDPI and ACS Style

To, J.; Torres, J. Viroporins in the Influenza Virus. Cells 2019, 8, 654. https://doi.org/10.3390/cells8070654

AMA Style

To J, Torres J. Viroporins in the Influenza Virus. Cells. 2019; 8(7):654. https://doi.org/10.3390/cells8070654

Chicago/Turabian Style

To, Janet, and Jaume Torres. 2019. "Viroporins in the Influenza Virus" Cells 8, no. 7: 654. https://doi.org/10.3390/cells8070654

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