Virology and Emerging Diseases - Sci Forschen

Full Text

Review Article
Evasion of Host Innate Immunity by Emerging Viruses: Antagonizing Host RIG-I Pathways

  Ferran MC1      Skuse GR1*   

1Thomas H Gosnell School of Life Sciences, Rochester Institute of Technology, Rochester, New York, USA

*Corresponding author: Gary R Skuse, Thomas H Gosnell School of Life Sciences, Rochester Institute of Technology, 153 Lomb Memorial Drive, Rochester, NY 14623, USA; E-mail: grssbi@rit.edu


Abstract

Viruses confront a seemingly dichotomous relationship with their host cells. They must overcome host defenses in order to complete their infectious cycles and generate new viruses yet the host must remain healthy and hospitable for that to take place. Shortly after infection, the RIGI-like receptors (RLRs) within the cytoplasm of the infected cell recognize foreign motifs present in the pathogen. The host responds by activating a signaling pathway that leads to activation of cellular transcription factors, including the NF-κB and interferon regulatory factor 3 (IRF3), that are necessary for induction of the type 1 interferon genes. Many viruses subdue components of the host innate immune system to facilitate viral replication. Viruses with single stranded RNA genomes that possess double stranded replication intermediates, 5’ triphosphates or 5’ diphosphates along with other secondary recognition motifs including length express proteins that either hide their dsRNA from detection by RLRs, interact with RIG-I directly, or interfere with components of the RIG-I pathway with the ultimate goal of evading innate immunity. In every case the end result is that the host antiviral defense system is crippled and viral propagation can proceed. In this review we focus on the eight emerging viruses most likely to cause major epidemics, including Arenaviruses, Bunyaviruses, Coronaviruses, Filoviruses and Paramyxoviruses, as identified by the World Health Organization in 2016. Once fully understood, the mechanisms employed by viruses to evade host cell immunity may serve as effective targets for a variety of antiviral agents.

Keywords

RIG-I; Interferon; IRF3; RNA viruses; MAVS


Introduction

Upon viral infection host pathogen recognition receptors, including the Toll-like receptors (TLRs) and RIG-I-like receptors (RLRs), detect the presence of foreign motifs referred to as pathogen-associated molecular patterns (PAMPs) and activate a signaling pathway that ultimately leads to the induction and expression of the type 1 interferons (IFN). This newly produced IFN establishes an antiviral state in surrounding cells that prevents virus replication. Therefore, induction of IFN gene expression and the activation of subsequent IFN signaling pathways is crucial to the ability of a host cell to mount an innate immune response [1]. To counteract these powerful antiviral responses many viruses have evolved elegant, and often multi-pronged, mechanisms by which they evade the innate immune response [2]. There has been a tremendous amount of research done to understand how different viruses block induction of the IFN gene by either preventing recognition by RLRs or suppressing the signaling pathways they activate. One well-studied member of the RLR family is the retinoic acid–inducible gene-1 (RIG-I). This cytoplasmic receptor primarily detects 5′ppp-RNA molecules with short secondary motifs of dsRNA or ssRNA [3,4]. In contrast, another cytoplasmic RLR referred to as MDA5 recognizes longer dsRNA motifs so that each RLR recognizes different viruses based on their respective PAMPs [5]. Following binding of viral RNAs, RIG-I and MDA5 interact with the mitochondrial membrane bound adaptor molecule MAVS (mitochondrial antiviral signaling protein, also referred to as IPS-1, VISA, or CARDIF), which activates two kinase complexes. The IκBKinaseɛ/ TANK Binding Kinase 1 (IKKɛ/TBK1) phosphorylate the transcription factors, interferon regulatory factors (IRF), IRF3 and IRF7, which then form homodimers or heterodimers, enter the nucleus and initiate transcription of IFNα/β. For clarity, it is worth mentioning that the type I interferons include a subgroup of interferon proteins that include IFNα/β. While IRF3 is constitutively expressed in most cells, IRF7 is an interferon stimulated gene (ISG) that is typically expressed at low levels but can be induced several-fold in response to IFN signaling. Therefore, it is thought that IRF3 mediates transcription of the majority of early IFN expression. The IKKα/IKKβ/IKKγ kinase complex phosphorylates IκBα, targeting this repressor protein of nuclear factor kappa B (NF-κB) for degradation. Following secretion outside of the initially infected cell, the IFN protein is recognized by target cells and initiates their IFN signaling pathways [1,6]. Ultimately this leads to the expression or upregulation of hundreds of ISGs, including IFN, pro-apoptotic factors, and cytokines which establish an antiviral state in surrounding cells [1,7].

This review will focus on how select RNA viruses evade the innate immune response. Specifically, we will focus on how the top eight emerging viruses, as identified by the World Health Organization [8], suppress RIGI-mediated induction of the IFN antiviral response as shown in Figure 1. In order to provide perspective, we also include information about how vesicular stomatitis virus (VSV), a well-studied non-human pathogen, evades the host immune response. VSV serves as a model for how nonhuman pathogenic RNA viruses act in manners both similar to and different from the other emerging viruses. Taken together, the diversity of mechanisms employed by these pathogens to circumvent host defenses is remarkable. The similarities as well as the differences are striking.

Figure 1: Targeting of the RIG-I signaling pathway by emerging virusesUpon activation by cytoplasmic RNA, RIG-I is activated and interacts with MAVS. This initiates downstream signaling events that activate IRF3 and NF-κB, and ultimately results in induction of the IFNα/β gene. Many components in this pathway are inhibited by viral proteins, thereby suppressing the IFN response and enabling viral replication to occur. Viruses depicted above include Rift Valley fever virus (RVFV), Crimean-Congo hemorrhagic fever virus (CCHFV), ebolavirus (EBOV), Marburgh virus (MARV), Lassa fever virus (LASFV), Nipah virus (NiV), severe acute respiratory syndrome coronavirus (SARS-CoV), Middle East respiratory syndrome virus coronavirus (MERS-CoV) and vesicular stomatitis virus (VSV).

Rift Valley Fever Virus and Crimean-Congo Hemorrhagic Fever Virus

Members of the Bunyaviridae family that are listed in the 2016 WHO list of emerging viruses include the zoonotic arthropod-borne Rift Valley fever virus (RVFV) and the Crimean-Congo fever virus (CCHFV). Both of these viruses carry a tripartite negative sense RNA genome [9] and can cause severe disease in humans, including fulminating hemorrhagic fever [10,11].There are currently no prophylactic or therapeutic treatments available for these viruses [9]. The pathogenicity of these viruses is largely attributed to the ability of the multifunctional nonstructural protein NSs to inhibit global host cell transcription and to antagonize the IFN system [9,12-14].

Although RIG-I is activated upon recognition of RVFV RNA [15], IFN production is delayed in RVFV-infected animal models [13]. Several studies have demonstrated that the NSs protein utilizes several mechanisms to block IFN-β gene expression during early RVFV infection [13,16,17]. NSs was found to directly target IFN- β gene expression through its interaction with the cellular repressor protein Sin3A-associated protein 30 (SAP30), a subunit of the Sin3A/nuclear receptor co-repressor (NCoR)/histone deacetylase repressor complex. NSs simultaneously interacts with YY1, a transcription factor that regulates IFN-β gene expression [18]. YY1 directs the SAP30-NSs-YY1 complex to the IFN- β promoter site to form a multi protein repression complex on the promoter, which inhibits induction of the IFN-β gene [17]. RVFV NSs also indirectly down regulates IFN-β gene expression by shutting-off global host gene transcription by sequestering the p44 and XPD subunits of the TFIIH basal transcription factor [19]. NSs also inhibits host transcription by promoting the degradation of the TFIIH p62 subunit [20].

Similarly, IFN production and secretion is delayed during CCHFV infection [21,22]. A virally encoded protease processes the CCHFV genome to include a 5′ monophosphate (5′p) end [23], rather than the 5′ppp and 5′pp ends strongly recognized by RIG-I [24]. Therefore it was proposed that due to this modification CCHFV RNA is not sensed by RIG-I [23,25]. However, recently it was established that RIG-I does mediate an IFN response to CCHFV [26]. In fact, immuno stimulatory RNA (isRNA) was isolated from infected cells as well as from virion preparations, and RIG-I co-immuno precipitation resulted in the isolation of CCHFV isRNA from infected cell lysates. These findings indicate that RIG-I signaling is critical to the activation of an antiviral response to CCHFV infection [26].

While the CCHFV protein that antagonizes RIG-I-dependent IFN production has not yet been identified, the viral L protein has been suggested as a potential candidate. In addition to functioning as the viral RNA dependent-RNA polymerase, the CCHFV L protein is a cysteine protease that contains a viral homologue to the ovarian tumor protease domain (OTU) [27], which allows the removal of conjugated poly-ubiquitin (Ub) and interferon-induced Ub-like protein (ISG15) from target proteins [28,29]. Viral proteases which contain this domain evade ubiquitin- and ISG15-dependent innate immune responses [27,30], therefore it is possible that the CCHFV OTU directly antagonizes the innate immune response. More research must be done to determine if the CCHFV OTU blocks RIG-I signaling and to identify which proteins in the RIG-I pathway are targeted for OTU-dependent de-conjugation of Ub and ISG15.

Ebola and Marburg Viruses

Ebolavirus (EBOV) and Marburg virus (MARV) are members of the Filoviridae family that infect primates. They can cause hemorrhagic fever and are among the most virulent pathogens known, with case fatality rates reaching 90% during some outbreaks [31]. Mortality is swift and follows the shock and subsequent multi-organ failure that results from hemorrhagic complications [32]. This virulence is attributed to virally encoded proteins that antagonize the ability of the host to mount an effective innate immune response, leading to uncontrolled virus replication. It has been demonstrated that EBOV VP24 and the MARV VP40 inhibit the IFN signaling pathway [33,34]. As this occurs during the later phase of the IFN response it will not be discussed further herein.

In addition to its function as a polymerase cofactor and its role in viral assembly, the EBOV VP35 (eVP35) and MARV VP35 (mVP35) suppress innate immunity by targeting multiple steps in the RIG-Idependent induction of IFN gene expression [35,36]. Both eVP35 and mVP35 bind dsRNA [37] through a basic amino acid motif located in the highly conserved C-terminal IFN-inhibitory domain (IID). This binding sequesters the dsRNA from RIG-I surveillance and therefore prevents IFN production. The IID domain interacts with dsRNA in a sequencespecific manner and was demonstrated to be essential for VP35-mediated inhibition of IFN production [38-41]. By binding to viral dsRNA, eVP35 inhibited activation of the IFN-β promoter normally induced by overexpression of RIG-I, MAVS, IKKε and TBK1 [37]. Mutation of dsRNA-binding residues led to a decrease in dsRNA binding [37,42]. 

Comparison of the crystal structures of eVP35 and mVP35 IIDs bound to dsRNA revealed that eVP35 interacts with both the phosphodiester backbone and caps the ends of dsRNA [40,43], while mVP35 was found to interact with the dsRNA backbone only [41]. Edwards and coworkers established that eVP35 was able to more strongly inhibit RLR signaling than mVP35. This correlated with induction of a more robust IFN response in MARV-infected cells as compared to EBOV-infected cells. These functional differences between eVP35 and mVP35 mapped to IID. Therefore the binding mode of both viral VP35s with dsRNA plays a significant role in the magnitude of the IFN response in filoviral-infected cells [44].

While VP35 has been shown to bind synthetic dsRNA molecules introduced in vitro [45], direct evidence that VP35 binds isRNA to limit RIG-I activation was lacking. Utilizing a Sendai virus (SeV) infection model and deep sequencing of purified eVP35-bound RNAs, Dilley and coworkers demonstrated that the SeV defective interfering (DI) RNA, a known activator of RIG-I, is the is RNA bound by eVP35 proteins in infected cells. Mutation of basic residues in the IID domain that were required for dsRNA binding and inhibition of IFN destroyed the ability of eVP35 to bind the SeV DI RNA. In addition, select host RNAs were preferentially bound by  wild type eVP35  in cell culture. These findings support the contention that VP35 binds viral isRNA to block the RIG-I pathway and thereby evade the IFN response [45]. VP35 also inhibits IFN production by targeting the RLR pathways in a dsRNA bindingindependent manner by interacting with key components of the RIG-I pathway.

The IID was critical for the ability of eVP35 and mVP35 to block IRF3 phosphorylation and activation by over expression of IKKɛ and TBK1, the kinases that activate this transcription factor [38,46]. In contrast, these viral proteins did not inhibit IFN-β promoter activation induced by expression of a constitutively active form of IRF3 [41,47]. Interestingly, eVP35 was found to target and bind to the N-terminal domain of both IKKɛ and TBK1 and was subsequently phosphorylated by these kinases. Overexpression of eVP35 and its interaction with IKKɛ and TBK1, sequesters them and impairs their normal interactions with IRF3, IRF7, and MAVS, and decreases the kinase activity in cells transfected with IKKɛ [47]. Taken together, these findings indicate that VP35 can act as a decoy substrate for the TBK1-IKKε complex, thereby impairing IRF3 phosphorylation through its normal interaction with TBK1 and IKKɛ [37,47].

Expression of wild-type eVP35 also interferes with the ability of RIG-I to interact with PACT, a cellular dsRNA binding protein that is an essential coactivator of RIG-I [48]. Mutations in the eVP35 IID domain prevented eVP35-PACT binding and limited the ability of eVP35 to inhibit PACTmediated activation of RIG-I. Cells in which PACT had been knocked down were defective for IFN induction and were insensitive to eVP35 activity [49].

It has been shown that TLR and RIG-I signaling covalently conjugates SUMO molecules to both IRF3 and IRF7 and this modification was correlated with reduced IFN transcription [50]. In addition, physical interaction of eVP35 with IRF3 and IRF7 led to their sumoylation. This modification inhibited the transcriptional activity of these IRFs and the downstream expression from the IFN-β promoter [51].

Lassa Fever Virus

Like other members of the Arenaviridae family, Lassa fever virus (LASFV) is an enveloped negative-sense RNA virus that carries a bisegmented genome [52]. LASFV is endemic in several West African countries where there are between 300,000-500,000 cases annually. This virus can cause fatal hemorrhagic fever in humans, resulting in approximately 5,000 deaths per year [53-56]. The pathogenesis of LASFV is associated with the ability of this virus to specifically target dendritic and endothelial cells [57,58]. In addition, LASFV is able to suppress the induction of host IFNs.

While the 5′-ppp dsRNA associated with the LASFV genome activates the RIG-I pathway [23], the virally encoded protein, NP, was identified as an IFN antagonist [59-61]. By inhibiting IRF3 phosphorylation, the multifunctional NP suppresses IFN induction [60,62]. This function of the LASFV NP is dependent on its intrinsic 3′–5′ exoribonuclease (ExoN) activity, which digests free dsRNA and thereby prevents RIG-I recognition of that non-cellular nucleic acid [63,64]. Mutations in the exoribonuclease active site dramatically reduced this activity and abrogated the ability of the LASFV NP to inhibit viral- or synthetic polyI:C-induced activation of the IFNα/β promoter in vitro [63-65]. Importantly, residues essential for NP-mediated IFN inhibition are highly conserved among all arenaviruses, indicating that this function too is conserved across all members of this viral family [63,65,66]. A robust, RIG-I dependent, innate immune response was activated in cells infected with a recombinant LASFV in which the ExoN function was abolished. These results correlate with earlier in vitro studies and underscore the essential role of the NP exonuclease activity in suppression of innate immunity during LASFV infection [67].

This same region within the NP protein was found to antagonize induction of IFN gene expression by inhibiting the nuclear translocation and transcriptional activity of NF-κB [68] and by blocking the autocatalytic activity of IKKɛ. By binding to the kinase domain of IKKɛ, NP inhibited the ability of the kinase to phosphorylate, and therefore activate IRF3. This NP-IKKɛ interaction also prevented IKKɛ from interacting with MAVS, thereby blocking the RIG-I pathway [69]. Interestingly, mutation of the same NP residues that are critical for its 3′–5′ exoribonuclease activity perturbed the interaction of NP with IKKɛ[69].

Nipah Virus

Nipah virus (NiV), also identified as an emerging virus, is a lethal pathogen that causes death in up to 70% of infected humans [70]. This virus infects both bats and humans but most likely originated in the former [71]. While other paramyxoviruses, such as Hendra virus, also use bats as a natural reservoir they do not all infect both bats and humans [72]. In fact, Hendra virus and NiV may be the only two and they are both lethal in humans [73]. One study suggested that bat to human transmission, and therefore the risk of human infection, is increased in those individuals who drink tree sap [71]. Other studies have elucidated the mechanisms employed by NiV to evade host innate immune responses.

When Pteropus vampyrus bat kidney (PVK) cells are infected with the related avian Newcastle disease virus (NDV), Glennon and coworkers observed an increase in expression of the genes encoding IFN, the GMCSF and IL-2 inhibitory factor I (GIF-I) and MDA5, among others [74]. In contrast, when those same cells are infected with NiV these genes are not upregulated, suggesting that NiV, perhaps uniquely, antagonizes expression of these host genes to facilitate viral replication. Suppression of IFN expression is most likely achieved by the viral accessory proteins V, W and C [75]. Similar responses involving the viral C protein have been observed in cells infected with measles virus [76]. In that system the suppression is most likely achieved by a combined mechanism that includes suppression of Janus Kinase 1 (jak1) phosphorylation and associated effects of the viral C protein [77]. The diversity observed in the ways different paramyxoviruses suppress host antiviral responses suggests that not only are their biological differences interesting but potential therapeutic approaches must be targeted to specific viral pathogens.

Severe Acute Respiratory Syndrome and Middle East Respiratory Syndrome Viruses

The Severe Acute Respiratory Syndrome Corona Virus (SARS-CoV) was first identified in 2002 in China as the causative agent in those affected individuals presenting with respiratory complications after exposure to a single health care worker [78]. Within eleven weeks of the first incidence in neighboring Hong Kong, the virus had spread to at least 27 countries or distinct political entities with nearly one fourth of the reported cases occurring among health care workers [79]. A wide range of fatality rates have been reported and not surprisingly they vary by location and they decrease over time [80,81]. The Middle East Respiratory Syndrome coronavirus (MERS-CoV) is another highly pathogenic member of this family. This lethal virus appears to be carried by Dromedary camels and is transmitted directly from them to humans [82]. When discovered in 2012 the virus displayed a nearly 37% mortality rate [82].

Patients with severe SARS disease displayed dysregulated IFN, ISGs and cytokine responses [83]. Similarly, MERS-CoV-infected cells exhibited reduced IFN and cytokine expression, blocked IRF3-mediated induction of the IFN response and upregulation of RIG-I, IRFs and other genes associated with innate immunity [84-86]. Taken together, these findings strongly suggest that the extreme virulence of SARS-CoV and MERSCoV is related to their ability to evade the host innate immune response.

SARS-CoV may hide its dsRNA from detection by RIG-I by replicating in “inner vesicles” within the lumen of a virus-induced reticulo vesicular network of modified endoplasmic reticulum (ER) membranes. The viral replicase (composed of the nsp3, nsp5, and nsp8 proteins) as well as the viral genomic RNA co-localize to these double membrane vesicles (DMVs), providing evidence that SARS-CoV replicates in this membrane network. The interior of these DMVs label for SARS-CoV dsRNA, therefore this virus forms DMVs to coordinate its replication and also hide replicating RNA from RLRs. The nsp4 viral replication protein appears to direct this membrane rearrangement, as its mutation alters assembly of these DMVs [87]. Interestingly, a similar phenomenon was observed in MERS-CoV-infected cells [88], indicating that at least two coronaviruses hide their dsRNA inside DMVs, avoiding detection by the host [89]. The SARS-CoV nucleocapsid (N) protein may suppress IFN production via a similar mechanism. Studies indicate that the N protein suppresses IFN signaling by targeting an early step in the pathway [90,91] and binds to dsRNA [51,92]. Therefore the N protein likely plays a key role in blocking the innate immune response [91] by shielding dsRNA from recognition by RIG-I. The SARS nsp14 protein contains a 3’-5’ exoribonuclease domain, therefore this protein may function to limit the IFN response by degrading viral dsRNA replication intermediates. Indirect support for this notion comes from studies of the LASFV encoded NP which contains a similar exonuclease domain. Mutation of critical residues within this domain abrogated the ability of LASFV NP to inhibit induction of the IFNα/β promoter [63-65]. While it is conceivable that the SARS-CoV nsp14 protein suppresses the IFN response by degrading dsRNA, further work is required to determine if this is indeed the case. Nevertheless, it is interesting that similar approaches are employed by viruses from different families. In this case an arenavirus and a coronavirus.

Several other proteins encoded by SARS-CoV antagonize the RIG-I signaling pathway. For example, the ORF9b protein suppresses innate immunity by targeting mitochondria and MAVS/TRAF3/TRAF6. Expression of ORF9b altered the mitochondrial morphology and subcellular localization of MAVS. The presence of ORF9b also led to the ubiquitination and degradation of MAVS, accompanied by a loss of TRAF3 and TRAF6, two key components of the RIG-I signaling pathway [93]. The SARS-CoV ORF3b and ORF6 proteins limit RLR-mediated induction of IFN. ORF3 localized to the mitochondrial outer membrane and may therefore inhibit MAVS at the mitochondria or at a point downstream of MAVS [90,94]. In contrast, ORF6 localized primarily to the ER and Golgi apparatus and may disrupt the ER/Golgi transport necessary for the IFN response [90]. The SARS-CoV M protein inhibits induction of IFN by binding to TRAF3 and impeding the formation of a TRAF- ·TANK·TBK1/IKKɛ complex, thereby inhibiting TBK1/IKKɛ-dependent activation of IRF3 and IRF7 [95]. Finally, the papain–like protease (PLP) domain of the SARS-CoV nsp3 protein interacts with STING and disrupts the dimerization and activation of this adaptor molecule. Inactive STING is unable to recruit MAVS to the TBK1-IKKɛ complex, therefore these kinases do not phosphorylate IRF3 and IFN gene expression is not induced. The PLP domain of nsp3 also disrupts NF-κB signaling, possibly by a similar mechanism [96] and it expresses a deubiquitinating activity that removes Ub from key components of the pathway, including RIG-I, STING, TBK1 and IRF3 [96,97]. 

Expression of the MERS-CoV ORF4b antagonizes the host IFNα/β expression that is normally upregulated in response to viral infection [98]. The accessory protein encoded by ORF4b, termed p4b, acts in both the cytoplasm and the nucleus [99]. Interestingly, Yang and coworkers demonstrated that in the cytoplasm p4b binds to TBK1 and IKKɛ, thereby suppressing molecular interactions between MAVS and IKKɛ, while inhibiting the phosphorylation of IRF3 [98]. When in the nucleus, the same protein inhibits the IRF3 and IRF7 induced expression of IFN-β. However, ablation of the protein’s nuclear localization signal eliminated its ability to inhibit IFN- β expression but not the IFN- β expression induced by RIG-I, TBK-1, MAVS, MDA5 and IKKɛ. This suggests that p4b employs multiple approaches to inhibit IFN- β in both the cytoplasm and the nucleus, no doubt contributing to the observed viral pathogenicity. Interestingly, the MERS-CoV M protein is able to interact with TRAF3 which hampers the TRAF3-TBK1 interaction and therefore leads to a decrease in IRF3 activation. The N-terminal transmembrane domain of the MERS-CoV M protein is sufficient for interaction with TRAF3 [100], which is similar to what has been shown for the SARS-CoV M protein [101].

Vesicular Stomatitis Virus

While not on the WHO list of emerging viruses, VSV is a well-studied member of the Rhabdoviridae with a host range that includes insects, cattle, horses and pigs, and it serves as an excellent model system to study the interplay between viruses and the IFN responses of their hosts. The absence of IFN induction in wild type virus infected cells is thought to result from the presence of one or more virally encoded IFN suppressors that presumably are defective in IFN-inducing viruses [102]. One of these suppressors is the matrix (M) protein which is crucial for many of the cytotoxic effects associated with VSV infection, including the downregulation of global host gene expression [39,78,103] and inhibition of the nuclear-cytoplasmic transport of host mRNAs [9,11,104]. The M protein has been shown to inhibit host transcription [39,103] and suppress IFN-β gene expression in the absence of other viral components [78]. Therefore, several researchers have proposed that VSV evades the IFN response by an M-mediated “shut-off ” of host gene expression. In support of this hypothesis there is a strong correlation between the virus’s ability to inhibit host gene expression and its ability to suppress IFN expression.

Wild type VSV rapidly inhibits host RNA and protein synthesis and is a poor inducer, or non-inducer, of IFN [22]. In contrast, the VSV mutant strain T1026R1 [103], which contains a single amino acid mutation at position 51 (M51R) of the M protein [105], is delayed in its ability to inhibit host RNA and protein synthesis [106] and is an excellent inducer of IFN [21,30]. A recent study indicates that the M protein either in the context of viral infection or when expressed alone is able to block viralmediated activation of NF-κB by targeting a step in the canonical NF-κB pathway, and the M51R mutation abrogates this function [107]. These results imply that the VSV M protein encodes two suppressors of IFN gene expression; the well-described ability to inhibit host gene expression as well as the ability to suppress induction of the IFN-β promoter by specifically interfering with the NF-κB pathway. This is similar to the molecular strategies used by the RVFV NSs protein, which inhibits IFN gene expression indirectly by inhibiting global host transcription and directly by forming a multiprotein repression complex on the IFN gene promoter.

Conclusion and Recommendation

Many of the emerging viruses discussed herein are lethal to humans. While VSV is not lethal, it serves as a well-studied model of virus infection and host immune detection and has revealed mechanisms of host innate immune evasion that are seen in other viruses. Interestingly, even within families of viruses the approaches used by the individual viruses to thwart host innate immune surveillance vary. In contrast, some approaches are shared among viruses of different families. Taken together, this tangled story of host immune evasion by disparate RNA viruses makes the prospect of using a single therapeutic approach impossible. Therefore it is imperative that we better understand the specific interactions between virally-encoded proteins and those of their hosts in order to develop lifesaving therapies.


References

  1. Randall RE, Goodbourn S (2008) Interferons and viruses: an interplay between induction, signalling, antiviral responses and virus countermeasures. J Gen Virol 89: 1-47. [Ref.]
  2. Bowie AG, Unterholzner L (2008) Viral evasion and subversion of pattern-recognition receptor signalling. Nat Rev Immunol 8: 911-922. [Ref.]
  3. Pichlmair A, Schulz O, Tan CP, Näslund TI, Liljeström P, et al. (2006) RIG-I-mediated antiviral responses to single-stranded RNA bearing 5’-phosphates. Science 314: 997-1001. [Ref.]
  4. Baum A, García-Sastre A (2011) Differential recognition of viral RNA by RIG-I. Virulence 2: 166-169. [Ref.]
  5. Kato H, Takeuchi O, Mikamo-Satoh E, Hirai R, Kawai T, et al. (2008) Length-dependent recognition of double-stranded ribonucleic acids by retinoic acid-inducible gene-I and melanoma differentiationassociated gene 5. J Exp Med 205: 1601-1610. [Ref.]
  6. Isaacs A, Lindenmann J (1957) Virus interference. I. The interferon. Proc R Soc Lond B Biol Sci 147: 258-267. [Ref.]
  7. Akira S (2006) TLR signaling. Curr Top Microbiol Immunol 311: 1-16. [Ref.]
  8. WHO (2016) Top 8 emerging diseases likely to cause major epidemics. World Economic Forum, World health Organization, Geneva, Switzerland. [Ref.]
  9. Walter CT, Barr JN (2011) Recent advances in the molecular and cellular biology of bunyaviruses. J Gen Virol 92: 2467-2484. [Ref.]
  10. Ikegami T (2012) Molecular biology and genetic diversity of Rift Valley fever virus. Antiviral Res 95: 293-310. [Ref.]
  11. Keshtkar-Jahromi M, Kuhn JH, Christova I, Bradfute SB, Jahrling PB, et al. (2011) Crimean-Congo hemorrhagic fever: current and future prospects of vaccines and therapies. Antiviral Res 90: 85-92. [Ref.]
  12. Bridgen A, Weber F, Fazakerley JK, Elliott RM (2001) Bunyamwera bunyavirus nonstructural protein NSs is a nonessential gene product that contributes to viral pathogenesis. Proc Natl Acad Sci U S A 98: 664-669. [Ref.]
  13. Elliott RM, Weber F (2009) Bunyaviruses and the type I interferon system. Viruses 1: 1003-1021. [Ref.]
  14. Blakqori G, Delhaye S, Habjan M, Blair CD, Sánchez-Vargas I, et al. (2007) La Crosse bunyavirus nonstructural protein NSs serves to suppress the type I interferon system of mammalian hosts. J Virol 81: 4991-4999. [Ref.]
  15. Weber M, Gawanbacht A, Habjan M, Rang A, Borner C, et al. (2013) Incoming RNA virus nucleocapsids containing a 5’-triphosphorylated genome activate RIG-I and antiviral signaling. Cell Host Microbe 13: 336-346. [Ref.]
  16. Bouloy M, Janzen C, Vialat P, Khun H, Pavlovic J, et al. (2001) Genetic evidence for an interferon-antagonistic function of rift valley fever virus nonstructural protein NSs. J Virol 75: 1371-1377. [Ref.]
  17. Le May N, Mansuroglu Z, Léger P, Josse T, Blot G, et al. (2008) A SAP30 complex inhibits IFN-beta expression in Rift Valley fever virus infected cells. PLoS Pathog 4: e13. [Ref.]
  18. Huang NE, Lin CH, Lin YS, Yu WC (2003) Modulation of YY1 activity by SAP30. Biochem Biophys Res Commun 306: 267-275. [Ref.]
  19. Le May N, Dubaele S, Proietti De Santis L, Billecocq A, Bouloy M (2004) TFIIH transcription factor, a target for the Rift Valley hemorrhagic fever virus. Cell 116: 541-550. [Ref.]
  20. Kalveram B, Lihoradova O, Ikegami T (2011) NSs protein of rift valley fever virus promotes posttranslational downregulation of the TFIIH subunit p62. J Virol 85: 6234-6243. [Ref.]
  21. Andersson I, Karlberg H, Mousavi-Jazi M, Martínez-Sobrido L, Weber F (2008) Crimean-Congo hemorrhagic fever virus delays activation of the innate immune response. J Med Virol 80: 1397-1404. [Ref.]
  22. Weber F, Mirazimi A (2008) Interferon and cytokine responses to Crimean Congo hemorrhagic fever virus; an emerging and neglected viral zonoosis. Cytokine Growth Factor Rev 19: 395-404. [Ref.]
  23. Habjan M, Andersson I, Klingström J, Schümann M, Martin A, et al. (2008) Processing of genome 5’ termini as a strategy of negativestrand RNA viruses to avoid RIG-I-dependent interferon induction. PLoS One 3: e2032. [Ref.]
  24. Goubau D, Schlee M, Deddouche S, Pruijssers AJ, Zillinger T, et al. (2014) Antiviral immunity via RIG-I-mediated recognition of RNA bearing 5’-diphosphates. Nature 514: 372-375. [Ref.]
  25. Wang Y, Ludwig J, Schuberth C, Goldeck M, Schlee M, et al. (2010) Structural and functional insights into 5’-ppp RNA pattern recognition by the innate immune receptor RIG-I. Nat Struct Mol Biol 17: 781-787. [Ref.]
  26. Spengler JR, Patel JR, Chakrabarti AK, Zivcec M, García-Sastre A, et al. (2015) RIG-I Mediates an Antiviral Response to Crimean-Congo Hemorrhagic Fever Virus. J Virol 89: 10219-10229. [Ref.]
  27. Frias-Staheli N, Giannakopoulos NV, Kikkert M, Taylor SL, Bridgen A, et al. (2007) Ovarian tumor domain-containing viral proteases evade ubiquitin- and ISG15-dependent innate immune responses. Cell Host Microbe 2: 404-416. [Ref.]
  28. Capodagli GC, McKercher MA, Baker EA, Masters EM, Brunzelle JS, et al. (2011) Structural Analysis of a Viral Ovarian Tumor Domain Protease from the Crimean-Congo Hemorrhagic Fever Virus in Complex with Covalently Bonded Ubiquitin. J Virol 85: 3621-3630. [Ref.]
  29. James TW, Frias-Staheli N, Bacik JP, Levingston Macleod JM, Khajehpour M, et al. (2011) Structural basis for the removal of ubiquitin and interferon-stimulated gene 15 by a viral ovarian tumor domaincontaining protease. Proc Natl Acad Sci U S A 108: 2222-2227. [Ref.]
  30. Zhao C, Collins MN, Hsiang TY, Krug RM (2013) Interferon-induced ISG15 pathway: an ongoing virus-host battle. Trends Microbiol 21: 181-186. [Ref.]
  31. Mire CE, Geisbert JB, Marzi A, Agans KN, Feldmann H, et al. (2013) Vesicular stomatitis virus-based vaccines protect nonhuman primates against Bundibugyo ebolavirus. PLoS Negl Trop Dis 7: e2600. [Ref.]
  32. Connor JH, McKenzie MO, Parks GD, Lyles DS (2007) Antiviral activity and RNA polymerase degradation following Hsp90 inhibition in a range of negative strand viruses. Virology 362: 109-119. [Ref.]
  33. Zhang AP, Abelson DM, Bornholdt ZA, Liu T, Woods VL, et al. (2012) The ebolavirus VP24 interferon antagonist: know your enemy. Virulence 3: 440-445. [Ref.]
  34. Mateo M, Reid SP, Leung LW, Basler CF, Volchkov VE (2010) Ebolavirus VP24 binding to karyopherins is required for inhibition of interferon signaling. J Virol 84: 1169-1175.
  35. Basler CF, Amarasinghe GK (2009) Evasion of interferon responses by Ebola and Marburg viruses. J Interferon Cytokine Res 29: 511-520. [Ref.]
  36. Ramanan P, Shabman RS, Brown CS, Amarasinghe GK, Basler CF, et al. (2011) Filoviral immune evasion mechanisms. Viruses 3: 1634- 1649. [Ref.]
  37. Cárdenas WB, Loo YM, Gale M Jr, Hartman AL, Kimberlin CR, et al. (2006) Ebola virus VP35 protein binds double-stranded RNA and inhibits alpha/beta interferon production induced by RIG-I signaling. J Virol 80: 5168-5178. [Ref.]
  38. Hartman AL, Towner JS, Nichol ST (2004) A C-terminal basic amino acid motif of Zaire ebolavirus VP35 is essential for type I interferon antagonism and displays high identity with the RNA-binding domain of another interferon antagonist, the NS1 protein of influenza A virus. Virology 328: 177-184. [Ref.]
  39. Leung DW, Ginder ND, Fulton DB, Nix J, Basler CF, et al. (2009) Structure of the Ebola VP35 interferon inhibitory domain. Proc Natl Acad Sci U S A 106: 411-416. [Ref.]
  40. Leung DW, Prins KC, Borek DM, Farahbakhsh M, Tufariello JM, et al. (2010) Structural basis for dsRNA recognition and interferon antagonism by Ebola VP35. Nat Struct Mol Biol 17: 165-172. [Ref.]
  41. Ramanan P, Edwards MR, Shabman RS, Leung DW, Endlich-Frazier AC, et al. (2012) Structural basis for Marburg virus VP35-mediated immune evasion mechanisms. Proc Natl Acad Sci U S A 109: 20661- 20666. [Ref.]
  42. Zinzula L, Esposito F, Pala D, Tramontano E (2012) dsRNA binding characterization of full length recombinant wild type and mutants Zaire ebolavirus VP35. Antiviral Res 93: 354-363. [Ref.]
  43. Kimberlin CR, Bornholdt ZA, Li S, Woods VL, MacRae IJ, et al. (2010) Ebolavirus VP35 uses a bimodal strategy to bind dsRNA for innate immune suppression. Proc Natl Acad Sci U S A 107: 314-319. [Ref.]
  44. Edwards MR, Liu G, Mire CE, Sureshchandra S, Luthra P, et al. (2016) Differential Regulation of Interferon Responses by Ebola and Marburg Virus VP35 Proteins. Cell Rep 14: 1632-1640. [Ref.]
  45. Dilley KA, Voorhies AA, Luthra P, Puri V, Stockwell TB, et al. (2017) The Ebola virus VP35 protein binds viral immunostimulatory and host RNAs identified through deep sequencing. PLoS One 12: e0178717. [Ref.]
  46. Hartman AL, Dover JE, Towner JS, Nichol ST (2006) Reverse genetic generation of recombinant Zaire Ebola viruses containing disrupted IRF-3 inhibitory domains results in attenuated virus growth in vitro and higher levels of IRF-3 activation without inhibiting viral transcription or replication. J Virol 80: 6430-6440. [Ref.]
  47. Prins KC, Cárdenas WB, Basler CF (2009) Ebola virus protein VP35 impairs the function of interferon regulatory factor-activating kinases IKKepsilon and TBK-1. J Virol 83: 3069-3077. [Ref.]
  48. Lui PY, Wong LR, Ho TH, Au SWN, Chan CP, et al. (2017) PACT Facilitates RNA-Induced Activation of MDA5 by Promoting MDA5 Oligomerization. J Immunol 199: 1846-1855. [Ref.]
  49. Luthra P, Ramanan P, Mire CE, Weisend C, Tsuda Y, et al. (2013) Mutual antagonism between the Ebola virus VP35 protein and the RIG-I activator PACT determines infection outcome. Cell Host Microbe 14: 74-84. [Ref.]
  50. Kubota T, Matsuoka M, Chang TH, Tailor P, Sasaki T, et al. (2008) Virus infection triggers SUMOylation of IRF3 and IRF7, leading to the negative regulation of type I interferon gene expression. J Biol Chem 283: 25660-25670. [Ref.]
  51. Chang TH, Kubota T, Matsuoka M, Jones S, Bradfute SB, et al. (2009) Ebola Zaire virus blocks type I interferon production by exploiting the host SUMO modification machinery. PLoS Pathog 5: e1000493. [Ref.]
  52. Lan S, McLay L, Aronson J, Ly H, Liang Y (2008) Genome comparison of virulent and avirulent strains of the Pichinde arenavirus. Arch Virol 153: 1241-1250. [Ref.]
  53. Briese T, Paweska JT, McMullan LK, Hutchison SK, Street C, et al. (2009) Genetic detection and characterization of Lujo virus, a new hemorrhagic fever-associated arenavirus from southern Africa. PLoS Pathog 5: e1000455. [Ref.]
  54. Charrel RN, Coutard B, Baronti C, Canard B, Nougairede A, et al. (2011) Arenaviruses and hantaviruses: from epidemiology and genomics to antivirals. Antiviral Res 90: 102-114. [Ref.]
  55. Ogbu O, Ajuluchukwu E, Uneke CJ (2007) Lassa fever in West African sub-region: an overview. J Vector Borne Dis 44: 1-11. [Ref.]
  56. Charrel RN, de Lamballerie X, Emonet S (2008) Phylogeny of the genus Arenavirus. Curr Opin Microbiol 11: 362-368. [Ref.]
  57. Mahanty S, Hutchinson K, Agarwal S, McRae M, Rollin PE, et al. (2003) Cutting edge: impairment of dendritic cells and adaptive immunity by Ebola and Lassa viruses. J Immunol 170: 2797-2801. [Ref.]
  58. Baize S, Kaplon J, Faure C, Pannetier D, Georges-Courbot MC, et al. (2004) Lassa virus infection of human dendritic cells and macrophages is productive but fails to activate cells. J Immunol 172: 2861-2869. [Ref.]
  59. Carnec X, Baize S, Reynard S, Diancourt L, Caro V, et al. (2011) Lassa virus nucleoprotein mutants generated by reverse genetics induce a robust type I interferon response in human dendritic cells and macrophages. J Virol 85: 12093-12097. [Ref.]
  60. Martinez-Sobrido L, Giannakas P, Cubitt B, García-Sastre A, de la Torre JC (2007) Differential inhibition of type I interferon induction by arenavirus nucleoproteins. J Virol 81: 12696-12703. [Ref.]
  61. Müller S, Geffers R, Günther S (2007) Analysis of gene expression in Lassa virus-infected HuH-7 cells. J Gen Virol 88: 1568-1575.
  62. Martínez-Sobrido L, Zúñiga EI, Rosario D, García-Sastre A, de la Torre JC (2006) Inhibition of the type I interferon response by the nucleoprotein of the prototypic arenavirus lymphocytic choriomeningitis virus. J Virol 80: 9192-9199. [Ref.]
  63. Qi X, Lan S, Wang W, Schelde LM, Dong H, et al. (2010) Cap binding and immune evasion revealed by Lassa nucleoprotein structure. Nature 468: 779-783. [Ref.]
  64. Hastie KM, Kimberlin CR, Zandonatti MA, MacRae IJ, Saphire EO (2011) Structure of the Lassa virus nucleoprotein reveals a dsRNAspecific 3’ to 5’ exonuclease activity essential for immune suppression. Proc Natl Acad Sci U S A 108: 2396-2401. [Ref.]
  65. Jiang X, Huang Q, Wang W, Dong H, Ly H, et al. (2013) Structures of arenaviral nucleoproteins with triphosphate dsRNA reveal a unique mechanism of immune suppression. J Biol Chem 288: 16949-16959. [Ref.]
  66. Harmon B, Kozina C, Maar D, Carpenter TS, Branda CS, et al. (2013) Identification of critical amino acids within the nucleoprotein of Tacaribe virus important for anti-interferon activity. J Biol Chem 288: 8702-8711. [Ref.]
  67. Reynard S, Russier M, Fizet A, Carnec X, Baize S (2014) Exonuclease domain of the Lassa virus nucleoprotein is critical to avoid RIG-I signaling and to inhibit the innate immune response. J Virol 88: 13923- 13927. [Ref.]
  68. Rodrigo WW, Ortiz-Riaño E, Pythoud C, Kunz S, de la Torre JC, et al. (2012) Arenavirus nucleoproteins prevent activation of nuclear factor kappa B. J Virol 86: 8185-8197.
  69. Pythoud C, Rodrigo WW, Pasqual G, Rothenberger S, MartínezSobrido L, et al. (2012) Arenavirus nucleoprotein targets interferon regulatory factor-activating kinase IKKε. J Virol 86: 7728-7738. [Ref.]
  70. Audsley MD, Moseley GW (2013) Paramyxovirus evasion of innate immunity: Diverse strategies for common targets. World J Virol 2: 57-70. [Ref.]
  71. Gurley ES, Hegde ST, Hossain K, Sazzad HMS, Hossain MJ, et al. (2017) Convergence of Humans, Bats, Trees, and Culture in Nipah Virus Transmission, Bangladesh. Emerging Infectious Diseases 23: 1446-1453. [Ref.]
  72. Mc Michael L, Edson D, Smith C, Mayer D, Smith I, et al. ( 2017) Physiological stress and Hendra virus in flying-foxes (Pteropus spp.), Australia. PLoS One 12: e0182171. [Ref.]
  73. Marsh GA, Wang LF (2012) Hendra and Nipah viruses: why are they so deadly? Curr Opin Virol 2: 242-247. [Ref.]
  74. Glennon NB, Jabado O, Lo MK, Shaw ML (2015) Transcriptome Profiling of the Virus-Induced Innate Immune Response in Pteropus vampyrus and Its Attenuation by Nipah Virus Interferon Antagonist Functions. J Virol 89: 7550-7566. [Ref.]
  75. Park MS, Shaw ML, Muñoz-Jordan J, Cros JF, Nakaya T, et al. (2003) Newcastle disease virus (NDV)-based assay demonstrates interferonantagonist activity for the NDV V protein and the Nipah virus V, W, and C proteins. J Virol 77: 1501-1511. [Ref.]
  76. Shaffer JA, Bellini WJ, Rota PA (2003) The C protein of measles virus inhibits the type I interferon response. Virology 315: 389-397. [Ref.]
  77. Yokota S, Saito H, Kubota T, Yokosawa N, Amano K, et al. (2003) Measles virus suppresses interferon-alpha signaling pathway: suppression of Jak1 phosphorylation and association of viral accessory proteins, C and V, with interferon-alpha receptor complex. Virology 306: 135-146. [Ref.]
  78. Ksiazek TG, Erdman D, Goldsmith CS, Zaki SR, Peret T, et al. (2003) A novel coronavirus associated with severe acute respiratory syndrome. N Engl J Med 348: 1953-1966. [Ref.]
  79. Cherry JD (2004) The chronology of the 2002-2003 SARS mini pandemic. Paediatr Respir Rev 5: 262-269. [Ref.]
  80. Fung WK, Yu PL (2003) SARS case-fatality rates. CMAJ 169: 277- 278. [Ref.]
  81. Suresh MR, Bhatnagar PK, Das D (2008) Molecular targets for diagnostics and therapeutics of severe acute respiratory syndrome (SARS-CoV). J Pharm Pharm Sci 11: 1s-13s. [Ref.]
  82. Zumla A, Hui DS, Perlman S (2015) Middle East respiratory syndrome. Lancet 386: 995-1007. [Ref.]
  83. Totura AL, Baric RS (2012) SARS coronavirus pathogenesis: host innate immune responses and viral antagonism of interferon. Curr Opin Virol 2: 264-275. [Ref.]
  84. Zielecki F, Weber M, Eickmann M, Spiegelberg L, Zaki AM, et al. (2013) Human cell tropism and innate immune system interactions of human respiratory coronavirus EMC compared to those of severe acute respiratory syndrome coronavirus. J Virol 87: 5300-5304. [Ref.]
  85. Josset L, Menachery VD, Gralinski LE, Agnihothram S, Sova P, et al. (2013) Cell host response to infection with novel human coronavirus EMC predicts potential antivirals and important differences with SARS coronavirus. MBio 4: e00165-13. [Ref.]
  86. Kindler E, Jonsdottir HR, Muth D, Hamming OJ, Hartmann R, et al. (2013) Efficient replication of the novel human betacoronavirus EMC on primary human epithelium highlights its zoonotic potential. MBio 4: e00611-12. [Ref.]
  87. Clementz MA, Kanjanahaluethai A, O Brien TE, Baker SC (2008) Mutation in murine coronavirus replication protein nsp4 alters assembly of double membrane vesicles. Virology 375: 118-129. [Ref.]
  88. de Wilde AH, Raj VS, Oudshoorn D, Bestebroer TM, van Nieuwkoop S, et al. (2013) MERS-coronavirus replication induces severe in vitro cytopathology and is strongly inhibited by cyclosporin A or interferon-α treatment. J Gen Virol 94: 1749-1760. [Ref.]
  89. Knoops K, Kikkert M, Worm SH, Zevenhoven-Dobbe JC, van der Meer Y, et al. (2008) SARS-coronavirus replication is supported by a reticulovesicular network of modified endoplasmic reticulum. PLoS Biol 6: e226. [Ref.]
  90. Kopecky-Bromberg SA, Martínez-Sobrido L, Frieman M, Baric RA, Palese P (2007) Severe acute respiratory syndrome coronavirus open reading frame (ORF) 3b, ORF 6, and nucleocapsid proteins function as interferon antagonists. J Virol 81: 548-557. [Ref.]
  91. Lu X, Pan J, Tao J, Guo D (2011) SARS-CoV nucleocapsid protein antagonizes IFN-β response by targeting initial step of IFN-β induction pathway, and its C-terminal region is critical for the antagonism. Virus Genes 42: 37-45. [Ref.]
  92. Tang TK, Wu MP, Chen ST, Hou MH, Hong MH, et al. (2005) Biochemical and immunological studies of nucleocapsid proteins of severe acute respiratory syndrome and 229E human coronaviruses. Proteomics 5: 925-937. [Ref.]
  93. Shi CS, Qi HY, Boularan C, Huang NN, Abu-Asab M, et al. (2014) SARS-coronavirus open reading frame-9b suppresses innate immunity by targeting mitochondria and the MAVS/TRAF3/TRAF6 signalosome. J Immunol 193: 3080-3089. [Ref.]
  94. Freundt EC, Yu L, Goldsmith CS, Welsh S, Cheng A, et al. (2010) The open reading frame 3a protein of severe acute respiratory syndromeassociated coronavirus promotes membrane rearrangement and cell death. J Virol 84: 1097-1109. [Ref.]
  95. Siu KL, Kok KH, Ng MH, Poon VK, Yuen KY, et al. (2009) Severe acute respiratory syndrome coronavirus M protein inhibits type I interferon production by impeding the formation of TRAF3.TANK. TBK1/IKKepsilon complex. J Biol Chem 284: 16202-16209. [Ref.]
  96. Frieman M, Ratia K, Johnston RE, Mesecar AD, Baric RS (2009) Severe acute respiratory syndrome coronavirus papain-like protease ubiquitin-like domain and catalytic domain regulate antagonism of IRF3 and NF-kappaB signaling. J Virol 83: 6689-6705. [Ref.]
  97. Clementz MA, Chen Z, Banach BS, Wang Y, Sun L, et al. (2010) Deubiquitinating and interferon antagonism activities of coronavirus papain-like proteases. J Virol 84: 4619-4629. [Ref.]
  98. Yang Y, Ye F, Zhu N, Wang W, Deng Y, et al. (2015) Middle East respiratory syndrome coronavirus ORF4b protein inhibits type I interferon production through both cytoplasmic and nuclear targets. Sci Rep 5: 17554. [Ref.]
  99. Matthews KL, Coleman CM, van der Meer Y, Snijder EJ, Frieman MB (2014) The ORF4b-encoded accessory proteins of Middle East respiratory syndrome coronavirus and two related bat coronaviruses localize to the nucleus and inhibit innate immune signalling. J Gen Virol 95: 874-882.
  100. Liu Y, Olagnier D, Lin R (2016) Host and Viral Modulation of RIG-IMediated Antiviral Immunity. Front Immunol 7: 662. [Ref.]
  101. Siu KL, Chan CP, Kok KH, Chiu-Yat Woo P, Jin DY (2014) Suppression of innate antiviral response by severe acute respiratory syndrome coronavirus M protein is mediated through the first transmembrane domain. Cell Mol Immunol 11: 141-149. [Ref.]
  102. Marcus PI, Sekellick MJ (1987) Interferon induction by viruses. XV. Biological characteristics of interferon induction-suppressing particles of vesicular stomatitis virus. J Interferon Res 7: 269-284. [Ref.]
  103. Stanners CP, Francoeur AM, Lam T (1977) Analysis of VSV mutant with attenuated cytopathogenicity: mutation in viral function, P, for inhibition of protein synthesis. Cell 11: 273-281. [Ref.]
  104. Petersen JM, Her LS, Varvel V, Lund E, Dahlberg JE (2000) The matrix protein of vesicular stomatitis virus inhibits nucleocytoplasmic transport when it is in the nucleus and associated with nuclear pore complexes. Mol Cell Biol 20: 8590-8601. [Ref.]
  105. Ferran MC, Lucas-Lenard JM (1997) The vesicular stomatitis virus matrix protein inhibits transcription from the human beta interferon promoter. J Virol 71: 371-377. [Ref.]
  106. Dunigan DD, Lucas-Lenard JM (1983) Two transcription products of the vesicular stomatitis virus genome may control L-cell protein synthesis. J Virol 45: 618-626. [Ref.]
  107. Varble AJ, Ried CD, Hammond WJ, Marquis KA, Woodruff MC, et al. (2016) The vesicular stomatitis virus matrix protein inhibits NF-κB activation in mouse L929 cells. Virology 499: 99-104. [Ref.]

Download Provisional PDF Here

 

Article Information

Aritcle Type: Review Article

Citation: Ferran MC, Skuse GR (2017) Evasion of Host Innate Immunity by Emerging Viruses: Antagonizing Host RIG-I Pathways. J Emerg Dis Virol 3(3): doi http://dx.doi.org/10.16966/2473-1846.135

Copyright: © 2017 Skuse GR, et al. This is an open-access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.

Publication history: 

  • Received date: 11 Sep 2017

  • Accepted date: 05 Oct 2017

  • Published date: 10 Oct 2017