Topic Highlight Open Access
Copyright ©The Author(s) 2015. Published by Baishideng Publishing Group Inc. All rights reserved.
World J Gastroenterol. Oct 14, 2015; 21(38): 10739-10748
Published online Oct 14, 2015. doi: 10.3748/wjg.v21.i38.10739
Immune and non-immune responses to hepatitis C virus infection
Jiaren Sun, Ricardo Rajsbaum, MinKyung Yi, Department of Microbiology and Immunology, Institute for Human Infections and Immunity, University of Texas Medical Branch, Galveston, TX 77555-1070, United States
Author contributions: Sun J, Rajsbaum R and Yi M analyzed the literature and wrote the manuscript.
Supported by NIH AI109100 (to Sun J), R01AI110358 (to Yi M); and UTMB Institute for Human Infection and Immunity and UT System Rising STAR Award (to Rajsbaum R).
Conflict-of-interest statement: The authors have no conflict of interest to report.
Open-Access: This article is an open-access article which was selected by an in-house editor and fully peer-reviewed by external reviewers. It is distributed in accordance with the Creative Commons Attribution Non Commercial (CC BY-NC 4.0) license, which permits others to distribute, remix, adapt, build upon this work non-commercially, and license their derivative works on different terms, provided the original work is properly cited and the use is non-commercial. See: http://creativecommons.org/licenses/by-nc/4.0/
Correspondence to: Jiaren Sun, MD, PhD, Department of Microbiology and Immunology, Institute for Human Infections and Immunity, University of Texas Medical Branch, 301 University Blvd, Galveston, TX 77555-1070, United States. jisun@utmb.edu
Telephone: +1-409-7470186 Fax: +1-409-7476869
Received: May 18, 2015
Peer-review started: May 20, 2015
First decision: June 23, 2015
Revised: July 6, 2015
Accepted: September 14, 2015
Article in press: September 14, 2015
Published online: October 14, 2015

Abstract

The host innate and adaptive immune systems are involved in nearly every step of hepatitis C virus (HCV) infection. In patients, the outcome is determined by a series of complex host-virus interactions, whether it is a natural infection or results from clinical intervention. Strong and persistent CD8+ and CD4+ T-cell responses are critical in HCV clearance, as well as cytokine-induced factors that can directly inhibit virus replication. Newly available direct-acting antivirals (DAAs) are very effective in viral clearance in patients. DAA treatment may further result in the down-regulation of programmed death-1, leading to rapid restoration of HCV-specific CD8+ T cell functions. In this review, we focus on recent studies that address the host responses critical for viral clearance and disease resolution. Additional discussion is devoted to the prophylactic vaccine development as well as to current efforts aimed at understanding the host innate responses against HCV infection. Current theories on how the ubiquitin system and interferon-stimulated genes may affect HCV replication are also discussed.

Key Words: Hepatitis C virus, Hepatitis, T cell, Direct-acting antiviral, Innate immune response

Core tip: Hepatitis C virus (HCV) is an etiologic agent that can cause severe liver diseases, including chronic hepatitis, liver fibrosis, liver cirrhosis, and hepatocellular carcinoma. Although newly available direct-acting antivirals (DAAs) are very effective in viral clearance in patients, it remains unclear as to how many of the world’s infected individuals will benefit from the new DAAs. In this review, we focus on recent studies that address the host responses critical for viral clearance and disease resolution. Additional discussion is devoted to the prophylactic vaccine development and innate responses against HCV infection. Current theories on how the ubiquitin system and interferon-stimulated genes may affect HCV replication are also discussed.



INTRODUCTION

Hepatitis C virus (HCV) is an etiologic agent that can cause severe liver diseases, including chronic hepatitis, liver fibrosis, liver cirrhosis, and hepatocellular carcinoma. Since the first discovery of this virus in 1989 by Choo et al[1], much has been learned about viral replication mechanisms and detailed functions of the viral proteins involved in these processes. Such knowledge has accelerated the development of direct-acting antivirals that could cure HCV infections, with much higher efficiency in shortening treatment duration compared to the traditional interferon-α-based therapy[2-6]. On the other hand, the vaccine development front seems to have lagged behind. This is certainly not due to the lack of effort to understand the HCV-induced immunity, but rather to the unique challenges in understanding effective immune responses against HCV.

The first challenge is the virus itself, which could successfully establish a chronic infection in about 80% of infected persons by effectively modulating innate and adaptive immune responses[7,8]. The second challenge is the difficulty in identifying and obtaining samples from acutely infected patients who successfully eliminated the virus due to the lack of distinct symptoms during acute phase of HCV infection. The third challenge is the lack of suitable small animal models that can recapitulate the HCV infection-mediated immune responses in humans. Although chimpanzee model has been the best immunocompetent animal model of HCV infection, with the recent National Institutes of Health (NIH) moratorium of usage of chimpanzee in HCV research, this challenge just got worse. Of note, there are continuing efforts to develop other immunocompetent animal model system[9,10]. Despite these difficulties, new information that could help us eventually control HCV-mediated immune dysregulation keeps emerging.

The goal of this review is to summarize the most up-to-date knowledge regarding both innate antiviral and adaptive immune responses that are available in the literature, to define the successful host responses that could contribute to HCV elimination. In addition, we discuss the implications of effective anti-HCV therapy on HCV-mediated immune modulation and vaccine development.

ADAPTIVE IMMUNE RESPONSES IN HEPATITIS C

Although HCV is capable of interfering with a wide range of host physiological processes, it is readily detected by the host sensing machinery, followed by the triggering of innate cellular responses[11,12], including the production of type I interferons (IFN-I) and the activation of downstream, antiviral target genes. However, despite these responses, HCV continues to replicate in the liver in the incubation phase. The adaptive immune response to HCV infection develops over several weeks[13]. Although the reasons for this delay are not understood, it is clear that the magnitude, diversity, and quality of the adaptive immune responses are the determinants of the outcome[13]. While this acute immune response has the potential to clear the viral infection, it is unsuccessful at least 50% of the time, and the virus has a very strong propensity to cause chronic infections. Development of chronicity is marked by a dramatic decrease in the activity of CD8+ cytotoxic T lymphocytes (CTL) and CD4+ Th cells in the liver without achieving viral clearance. Interestingly, the T cell dysfunction seems to be restricted to HCV-specific CD8+ T cells, since influenza-specific CD8+ cells were functional in chronic HCV patients[12].

T cell dysregulation in chronic infection

The effector functions of virus-specific CD8+ and CD4+ T cells are critical in viral clearance and disease resolution[14]. Interestingly, virus clearance and disease progression are seemingly mediated by phenotypically distinctive CD8+ CTL populations. The CD8+ CTLs participating in virus clearance expressed high levels of IFN-γ, but low levels of the activation marker CD38 (IFN-γhi CD38lo)[12,13]. In contrast, the CTL involved in liver injury are commonly IFN-γlo CD38hi, and the frequency of these cells tends to increase with the inflammation score (Figure 1, left panel)[15]. Although CTLs can exert limited antiviral activity, they are unable to keep pace with the evolution of HCV. As a result, HCV rapidly accumulates escape mutations in its genome, and persistent viremia ensues[16]. In studies with large cohorts of chronic subjects and spontaneous resolvers, adequate help from CD4+ T cells was found to be essential to promoting immune protection[17]. Among resolvers, the HCV epitopes are presented by multiple alleles of major histocompatibility complex II molecules, and nonstructural (NS) protein-directed CD4+ T-cell responses are associated with high levels of IL-2 and IFN-γ[18]. On the other hand, HCV persistence is associated with a high frequency of CD4+ regulatory T cells (Treg) that could directly suppress HCV-specific CTL in patients[19]. The resolution of disease is usually associated with the loss of expression of programmed death-1 (PD-1) and decreased functional suppression[17,20]. In addition to Treg cells, high expression of inhibitory receptor-PD-1 on CD4+ and CD8+ T cells also contributes to viral persistence as well as to failure of antiviral therapy (Figure 1, left panel)[21,22].

Figure 1
Figure 1 Direct-acting antiviral therapy and immune function restoration in hepatitis C virus infection. In chronic hepatitis C infection, T cells having a narrow repertoire of TCRs mount a weak response to HCV antigens, and their effector functions are often impaired. Many CD8+ and CD4+ T cells express low levels of IFN-γ and IL-2 accompanied by up-regulation of PD-1 molecules in the liver. Development of T regulatory cells and compromised dendritic cell functions also contribute to T cell functional impairment. Recent data suggest that IFN-free direct-acting antivirals not only clear HCV in the majority of patients, but also result in the down-regulation of PD-1, leading to rapid restoration of virus-specific CD8+ T cell functions in patients[28]. HCV: Hepatitis C virus; IFN: Interferon; IL: Interleukin; PD-1: Program death-1; TCR: T cell receptor.
T cell responses in immunomodulatory and direct-acting antiviral therapies

Until recently, the mainstay of treatment for chronic HCV infection had been pegylated interferon and ribavirin for HCV genotype 1 infection. This treatment resulted in a sustained virologic response (SVR) in 50%-80% of patients with HCV genotype 1 infection (higher SVR among those with genotypes 2 or 3 infections). There has been a clear demonstration that IFN-α, even when it achieved viral clearance, was not able to rescue antiviral T cells from exhaustion due to its pleiotropic effects on T cells[23-25]. To improve the efficacy of hepatitis C therapy, one critical question is whether antiviral T cells are permanently blemished by the sustained expression of PD-1 or other tolerating mechanisms; or alternatively, the removal of IFN-α would promote host recovery and the expansion of T cells with antiviral functions[26,27]. In a recent study involving a clinical cohort undergoing IFN-free therapy, a combination of Faldaprevir (a protease inhibitor) and Deleobuvir (a non-nucleoside polymerase inhibitor) cleared HCV in the majority of patients[28]. Furthermore, viral antigen reduction in response to direct-acting antiviral (DAA) treatment resulted in the down-regulation of PD-1, leading to rapid restoration of HCV-specific CD8+ T cell functions in patients (Figure 1, right panel)[28]. While this study unveils an unexpected benefit of these two DAAs, it is interesting to see whether other DAAs with or without ribavirin are also able to reverse exhaustion and restore full functions of HCV-specific T cells.

HCV PROPHYLACTIC VACCINES

HCV infection is a major public health problem in the world. Although DAAs are a significant advancement for an HCV treatment, it remains unclear as to how many of the world’s infected individuals will benefit from the new DAAs. For instance, the high cost of these DAAs will undoubtedly deny their access to low income countries and may also result in selective use in middle and even high income counties[29]. Thus there will likely still be a substantial number of HCV cases, including those individuals who are not screened and unaware of their infection. Additionally, DAAs are not efficacious in those having already developed advanced hepatic cirrhosis, carcinoma and liver failures. Importantly, a recent study showed that persistently HCV-infected chimpanzees cured with DAA maintained narrowly focused stable CD8+ T cell repertoires that were incapable of preventing persistent infection following HCV re-challenge[30]. Thus, vaccination may be necessary even for those individuals who have cleared virus following DAA treatment. For these reasons, an effective prophylactic HCV vaccine remains a critical instrument to halt the global HCV epidemic. Current human and chimpanzee studies suggest that a prophylactic vaccine inducing both protective T cells and broadly-neutralizing antibody (bnAb) responses is important for HCV control and thus highly desirable characteristics for the future vaccine candidates[29].

Circulating HCV is genetically diverse, and therefore a broadly effective vaccine must target conserved T- and B-cell epitopes of the virus. Several prophylactic vaccine candidates based on different strategies and viral targets have been developed in the last two decades. Vaccines aimed to target conserved T cell epitopes are shown to induce vigorous and broadly directed CD4+ and CD8+ T cell responses, and these are well underway in clinical development[31,32]. Despite the molecular trickeries employed by HCV, a number of bnAbs have been identified. The protective role of bnAbs against HCV infection has been demonstrated in chimpanzees, highlighting the possibility of developing a broadly effective vaccine by inducing bnAbs[33,34]. The HCV viral RNA genome encodes two structural envelope glycoproteins, E1 and E2. Although neutralizing antibodies (nAbs) to E1 have also been isolated, E2 is the main target of nAbs in HCV-infected patients[35]. In the last decade, significant discoveries of bnAbs and their structural analysis with antigenic epitopes have been made in the HIV-1 vaccine field; now a similar trend has begun to emerge in the HCV vaccine field. Recently, researchers have combined the latest findings in HCV structural biology and cutting-edge technologies in protein design and next-generation sequencing of Ab repertoires to facilitate HCV immunogen design for the induction of bnAbs in vaccination[35]. The linear HCV epitopes will be grated onto protein scaffolds, which allow epitope presentation in their bnAb-bound conformations. These studies demonstrate the feasibility of generating a highly potent antibody formulation against multiple, conserved neutralizing epitopes on HCV.

HCV AND INNATE IMMUNE RESPONSES

It is now clear that HCV infection induces innate responses capable of limiting virus replication to some extent. However, HCV is still able to establish chronic infections by escaping immune responses. Similar to other viruses, HCV encodes pathogen-associated molecular patterns (PAMPs), which are recognized by the host pattern-recognition receptors (PRRs). Members of the endosomal Toll-like receptor (TLR) family and the cytoplasmic Retinoic acid inducible gene (RIG-I)-like receptors (RLRs) can also recognize HCV PAMPs (Figure 2A). Pathogen recognition by PRRs results in activation of downstream signaling pathways leading to the production of pro-inflammatory cytokines, chemokines, IFN-I and type-III IFN (IFN-λ)[36]. IFNs elicit their antiviral activity through the up-regulation of many IFN-I-stimulated genes (ISGs), which act as direct effectors of the antiviral response[37]. HCV-encoded PAMPs recognized by the RIG-I sensor include the polyuridine motif of the HCV genome 3’ non-translated region and its replication intermediate, which binds RIG-I through the 5’ terminal triphosphate on the viral RNA. This signaling induces IFN-I and antiviral ISGs in the liver in vivo[38]. HCV PAMPs that are recognized by RIG-I are also produced by cleavage of viral NS5B region of HCV RNA by the IFN-inducible host endoribonuclease RNase L, releasing small structured RNAs with 5’-hydroxyl (5’-OH) and 3’-monophosphoryl (3’-p) groups[39]. Binding of the HCV PAMP induces a conformational change in RIG-I and subsequent ubiquitination by the E3-ubiquitin ligase TRIM25 (Figure 2A). This process and interaction with the chaperone protein 14-3-3e promote recruitment of activated RIG-I to the adaptor protein mitochondria antiviral signaling protein (MAVS) that is anchored to the mitochondria and also located in an intracellular membrane network at the peroxisomes and on mitochondrial- associated membranes[40-43]. Subsequently, different signaling partners are known to be recruited to MAVS, resulting in the activation of the IκB (IKK) and IKK-related kinases[44,45], TBK1 and IKKε, which phosphorylate the transcription factors IRF3 and IRF7 required for IFN-I production as well as IFN-λ[46-48].

Figure 2
Figure 2 Innate immune response to hepatitis C virus and viral escape mechanisms. A: RIG-I recognizes HCV PAMPs including HCV 5’-triphosphate RNA and small structured RNAs with 5'-hydroxyl (5'-OH) and 3'-monophosphoryl (3'-p) groups, which are cleaved by the host endoribonuclease RNase L. Binding of the HCV PAMP induces a conformational change in RIG-I and subsequent ubiquitination by TRIM25 and Riplet. RIG-I is then recruited to MAVS via mitochondria associated membranes (MAMs) and the chaperone protein 14-3-3e. Subsequently, TBK1 and IKKε phosphorylate IRF3 and IRF7 for IFN-I production as well as IFN-λ, and pro-inflammatory cytokines via NF-κB activation. The NS3-4A of HCV inhibits Riplet-dependent activation of RIG-I; B: IFN-I and IFN-λ are recognized by the IFNA receptor and IL28/IL-10 receptor respectively. Both trigger downstream signaling through phosphorylation of STAT1 and STAT2. Together, STAT1, STAT2 and IRF9 form the ISGF3 complex, which translocates to the nucleus for induction of antiviral ISGs. The tyrosine kinases JAK1 and TYK2, phosphorylate tyrosine 701 (Y701) on STAT1. In addition, phosphorylation of S708 on STAT1 by the IKKε kinase (activated by unanchored polyubiquitin chains) is also required for ISG induction. It is currently unknown whether IFN-λ stimulation also results in IKKε activation. miR-373 is up-regulated in HCV-infected cells and inhibits JAK1 and IRF9 function resulting in reduced STAT1 phosphorylation. TRIM22, which is induced by IFN-I, inhibits HCV replication probably by a mechanism involving ubiquitination of viral NS5. ISG15, another IFN-I inducible protein can inhibit HCV replication by ISGylation of viral NS5 rendering unstable. ISG15 has also been proposed to have pro-viral roles. The core protein of HCV associates with STAT1 and promotes its degradation. RIG-I: Retinoic acid inducible gene; HCV: Hepatitis C virus; PAMP: Pathogen-associated molecular pattern; IFN: Interferon; IL: Interleukin; ISGF: Interferon-stimulated gene factor; STAT: Signal transducers and activators of transcription; NF-κB: Nuclear factor-kappa B; IKK: Inhibitor of NF-κB kinase.
Deregulation of innate anti-viral signaling pathways during HCV replication

Both IFN-I and IFN-λ are produced upon innate recognition of viruses, although differential expression has been found in tissues such as the brain upon viral infection[49]. Although signaling occurs through different receptors, both IFN-I and IFN-λ can trigger downstream signaling through phosphorylation of signal transducers and activators of transcription 1 (STAT1) and STAT2, suggesting that both IFN-I and IFN-λ signaling result in induction of the same ISGs. Together, STAT1, STAT2 and IRF9 form the interferon-stimulated gene factor 3 (ISGF3) complex, which is essential for induction of ISGs (Figure 2B)[50]. The tyrosine kinases JAK1 and TYK2, which are both activated by IFN-I and IFN-λ, phosphorylate tyrosine 701 (Y701) on STAT1[51]. In addition, phosphorylation of S708 on STAT1 by the IKKε kinase is also required for the efficient induction of all ISGs in response to IFN-I[52]; however, it is currently unknown whether IFN-λ stimulation also results in activation of IKKε and STAT1-S708 phosphorylation. Furthermore, activation of IKKε during IFN-I signaling also requires binding to unanchored lysine-48 (K48)-linked polyubiquitin chains, which are not covalently attached to any protein[53]. Some evidence suggests that HCV inhibits STAT1 function. For example, by using a microRNA array in human hepatocytes infected with HCV, it was shown that miR-373 is up-regulated in HCV-infected cells (Figure 2B). This microRNA targeted JAK1 and IRF9 and reduced phosphorylation of STAT1. Consistent with this observation, knockdown of miR-373 resulted in the reduction of HCV RNA replication[54]. Furthermore, the core protein of HCV associates with STAT1 and promotes its degradation[55]. Immune cell populations and hepatocytes from HCV+ patients have reduced STAT1 and STAT3 proteins. Furthermore, STAT3 was preferentially ubiquitinated and targeted for proteosomal degradation in the presence of HCV[56]. These observations may explain in part the hypo-responsiveness observed in some patients to IFN-α treatment.

Since IFN-α/β and IFN-λ trigger similar signaling pathways, they are predicted to induce the same group of ISGs. However, in contrast to the widely expressed IFN-I receptor in different cell types, the expression of the IL-10R2 subunit of the IFN-λ receptor appears to be restricted to cell types and tissues, particularly on epithelial cells[49]. This raised the possibility that IFN-λ administration may be a better option to the common therapy for treating HCV infection with pegylated IFN-α (Peg-IFN-α) and ribavirin. It is now known that Peg-IFN-α treatment is only effective in a fraction of HCV-infected individuals with the presence of some side effects. Thus far Peg-IFN-λ has shown promising results[57,58]. Besides the tissue/cell type-specific expression of the IL-10R2 receptor, Peg-IFN-λ may be more effective since, unlike IFN-α treatment[59], Peg-IFN-λ treatment did not lead to refractoriness of JAK-STAT signaling following multiple stimulations[60]. IFN-λ treatment also results in a faster reduction in viral load as compared to those with IFN-α[61]. In addition, HCV infection appears to down-regulate the expression of IFN-I receptor, in contrast to sustained expression of the IL-28 receptor subunit. Global transcriptome analysis in hepatocytes indicated that IFN-λ stimulation prolonged the expression of various ISGs that are potentially beneficial to antiviral defense mechanisms[62].

In addition to inhibition of STAT-dependent innate immune responses, HCV has also been shown to inhibit TLR3 signaling and IFN-λ production in human hepatoma cell lines, with subsequent reduction in the antiviral ISG56, MxA and OAS-1. NS3/4A, NS5A and NS5B had the ability to inhibit poly I:C-induced IFN-λ1 expression in Huh7 cells[63,64].

Role of ubiquitin and ubiquitin-like molecules in HCV replication

Ubiquitination of proteins is a post-translational process, which has been demonstrated to regulate not only protein stability but also various steps of the signaling pathways in immune regulation and cytokine production[65]. Viruses have adapted to antagonize innate immune responses by using different mechanisms including manipulating the ubiquitin system for their own advantage[66]. For example, it was recently proposed that the Influenza virion carries unanchored polyubiquitin chains that promote virus uncoating by utilizing the host aggresome machinery[67]. At the same time these polyubiquitin chains that are released in the cytoplasm may function as a mechanism to alert the innate cellular response early upon virus entry to the cell[68]. Whether HCV utilizes unanchored polyubiquitin chains for replication remains to be tested. Nevertheless, HCV also has been shown to target the ubiquitin system to escape the innate immune response. HCV NS3-4A proteases inhibit the E3-ubiquitin ligase Riplet, which, together with TRIM25, is required for efficient ubiquitination and activation of RIG-I (Figure 2A)[69]. The ubiquitin system is also utilized by the host to restrict virus replication, and, in fact, may be one of the mechanisms by which IFN therapy limits virus replication in patients. For example the E3-ubiquitin ligase TRIM22, which is highly induced by IFN-I[70-72], may be associated with responsiveness to Peg-IFN-α-2a/RBV combination therapy[73]. A possible mechanistic explanation for these findings is supported by data showing that TRIM22 overexpression inhibits HCV replication and knockdown reduces IFN-induced anti-HCV activity. In addition, TRIM22 appears to promote ubiquitination of HCV-NS5A[74]. Although this study did not elucidate the functional effects of NS5A ubiquitanation, ubiquitin ligases as well as de-ubiquitinating enzymes are emerging as important proteins in controlling HCV replication. In particular, a recent study using an RNA interference (RNAi) screen found a few ubiquitin ligases to be important in HCV replication; the E2-conjugating enzyme UbE2J1, is involved in viral RNA replication; USP11, a de-ubiquitinating enzyme is involved in HCV IRES-mediated translation. In addition, TRIM42, another member of the E3-ubiquitin ligase family of proteins, and UbE2M, another E2 ubiquitin-conjugating enzyme, are also involved at early stages of viral post-entry[75].

Similar to posttranslational covalent modification by ubiquitin, the small ubiquitin-like modifier (SUMO) can also be covalently attached to protein lysines post-translationally, and regulates many different cellular processes. HCV also utilizes this cellular process for its own advantage; the NS5A protein of HCV is SUMOylated, resulting in increased protein stability by inhibiting ubiquitination[76].

ISG15, another ubiquitin-like molecule that can be covalently attached to lysine residues of proteins post-translationally, has been shown to play both positive and negative roles during HCV infection. ISG15 is a highly IFN-inducible gene that has antiviral functions against many viruses[77]. Accordingly, overexpression of ISG15- and ISG15-conjugation enzymes resulted in inhibition of HCV replication. Furthermore, HCV-NS5A protein was ISGylated, and this appeared to decrease NS5A stability[78]. However, in contrast to this study, other studies identified ISG15 as a pro-HCV host factor promoting HCV replication[79,80]. ISG15 may have different effects depending on the cell type or tissue expression. If indeed ISG15 acts as a pro-viral factor, its high induction by IFN treatment may help explain why in some patients IFN-based treatment results in persistent HCV infection. Additional evidence that ISG15 plays a role in HCV replication comes from studies on USP18, which specifically cleaves ISG15 from its cellular targets[81]. USP18 knockout mice are hypersensitive to IFN, with prolonged Jak/Stat signaling[82]. Expression of USP18 is increased in the liver biopsy specimens of patients who do not respond to IFN-α therapy, and siRNA knockdown of USP18 in human cells increases the ability of IFN to inhibit HCV replication as well as to increase cellular protein ISGylation and prolonged STAT1 phosphorylation, and a general enhancement of IFN-stimulated gene expression[83].

To demonstrate that in vivo knockdown of ISG15 may be used therapeutically to inhibit HCV replication, Real et al[84] used lipid nanoparticles to deliver siRNA specific to ISG15. The treatment resulted in specific reduction of ISG15 expression in the liver in vivo, resulting in reduced responses to IFN treatment. This also resulted in a reduction in HCV replication, supporting the role of ISG15 as a pro-viral factor. In addition, ISG15 knockdown revealed five potential candidates as pro-viral factors that depend on ISG15 expression. In particular, knockdown of the ISG15-dependent heterogeneous nuclear ribonucleoprotein K (HnrnpK) also resulted in decreased levels of HCV replication[84].

HCV infection is also known to induce the ubiquitin-dependent degradation of some cellular proteins including the retinoblastoma tumor suppressor protein by viral NS5B[85,86], and the suppressor of cytokine signaling 3 (SOCS3), which is a negative regulator of the JAK-STAT pathway[87]. HCV viral proteins have also been described to be ubiquitinated and degraded by the proteasome through both ubiquitin-dependent and independent mechanisms (review by Shoji et al[88]).

IFN-stimulated genes as antiviral factors to HCV

For decades Peg-IFN-α has been one of the most important therapeutics to control HCV infection in patients, although the exact mechanisms of viral inhibition have remained unclear. It is well established that IFN treatment will induce a large number of ISGs with known antiviral functions, but which ISGs and how they act against HCV remained largely unknown until recently. Some of these ISGs have direct or indirect anti-HCV functions and have been reviewed recently (see Horner et al[43]). These include ADAR, DDIT4, DDX58 (RIG-I), DDX60, EIF2AK2 (PKR), GBP1, IFI44L, IFI6, IFIT1, IFIT3, IFITM3, IRF1, IRF7, ISG12, ISG20, MAP3K14 (NIK), MOV10, MS4A4A, MX1 (MxA), NOS2, NT5C3, OAS1, OASL, PLSCR1, RNASEL, RSAD2 (viperin), SSBP3, and TRIM14[43]. Many of these genes were found by using lentiviral vectors expressing 389 selected ISGs in Huh-7.5 cells, a RIG-I-defective derivative of Huh-7 cells. Although most of the ISGs showed some degree of inhibition of HCV replication, RIG-I, MDA5, IRF1 and IRF7, which are genes involved in signaling to produce IFN-I, were the strongest inhibitors of HCV[37]. Other studies showed the IFN-induced transmembrane protein 1 (IFITM1) as an inhibitor of HCV[89,90]. In addition, another screen identified several antiviral ISGs induced by IFN-α and IFN-γ using an RNAi[91]. IFITM1 is highly induced by both IFN-I and IFN-γ has been shown to inhibit different viruses including West Nile virus, Influenza, HIV and HCV[89,92]. IFITM1 accumulates at hepatic tight junctions in HCV-infected human patient liver during IFN therapy and interacts with the HCV co-receptors CD81 and occludin, blocking viral entry[90]. ISG56 was also shown to inhibit HCV replication[89]. Additional ISGs have been reported to inhibit HCV. ISG20, and PKR are reported to inhibit HCV RNA synthesis[93]. Recently, the Cholesterol-25-hydroxylase (CH25H), a 31.6-kDa endoplasmic reticulum-associated enzyme that catalyzes oxidation of cholesterol to 25-hydroxycholesterol (25HC), was also shown to inhibit HCV. CH25H is an ISG that is induced in many tissues upon in vivo exposure to TLR ligands and IFN stimulations[94]. 25HC has also been reported to possess anti-HCV activity[95]. CH25H can interact with the NS5A protein of HCV and inhibit its dimer formation, which is essential for HCV replication[96].

In summary, although in recent years there have been great advances in our understanding of the anti-HCV functions of ISGs, many of these studies still fail to take into consideration the physiological conditions in which the virus replicates, as well as relevant immune cell types that are localized in the liver. Furthermore, it remains unclear as to how to induce these genes with exogenous treatments, or how to deliver lentiviral vectors containing specific ISG as potential antiviral treatments. Thus, additional studies are required using novel in vivo models combined with biochemical methods to identify the molecular mechanisms of antiviral functions.

CONCLUSION

The recent availability of highly effective DAAs against HCV infection brings hope for HCV eradication. However, initial reports suggest that DAA alone may not be enough to achieve this goal. Instead, HCV eradication will ultimately require the boosting of favorable innate and adaptive immune responses and ultimately vaccine development. Based on our knowledge of antiviral immune responses, the raising of effective antiviral responses against HCV will require agents or vaccine candidates that promote innate antiviral signaling and enhance both CD4 and CD8 responses effectively without inducing exhausted phenotypes and bnAb that could neutralize multiple genotypes of HCV. There is no doubt that effective immune-modulators against HCV infection will be available someday as a result of our continued efforts to understand HCV-induced immune regulation.

ACKNOWLEDGMENTS

The authors wish to express gratitude to their trainees, as well as Ms. Mardelle Susman for assistance with manuscript preparation.

Footnotes

P- Reviewer: Takahashi T S- Editor: Ma YJ L- Editor: A E- Editor: Ma S

References
1.  Choo QL, Kuo G, Weiner AJ, Overby LR, Bradley DW, Houghton M. Isolation of a cDNA clone derived from a blood-borne non-A, non-B viral hepatitis genome. Science. 1989;244:359-362.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Feld JJ. Interferon-free strategies with a nucleoside/nucleotide analogue. Semin Liver Dis. 2014;34:37-46.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Rupp D, Bartenschlager R. Targets for antiviral therapy of hepatitis C. Semin Liver Dis. 2014;34:9-21.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Scheel TK, Rice CM. Understanding the hepatitis C virus life cycle paves the way for highly effective therapies. Nat Med. 2013;19:837-849.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Welzel TM, Dultz G, Zeuzem S. Interferon-free antiviral combination therapies without nucleosidic polymerase inhibitors. J Hepatol. 2014;61:S98-S107.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Yau AH, Yoshida EM. Hepatitis C drugs: the end of the pegylated interferon era and the emergence of all-oral interferon-free antiviral regimens: a concise review. Can J Gastroenterol Hepatol. 2014;28:445-451.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Abdel-Hakeem MS, Shoukry NH. Protective immunity against hepatitis C: many shades of gray. Front Immunol. 2014;5:274.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Rehermann B. Hepatitis C virus versus innate and adaptive immune responses: a tale of coevolution and coexistence. J Clin Invest. 2009;119:1745-1754.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Billerbeck E, de Jong Y, Dorner M, de la Fuente C, Ploss A. Animal models for hepatitis C. Curr Top Microbiol Immunol. 2013;369:49-86.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  Scull MA, Shi C, de Jong YP, Gerold G, Ries M, von Schaewen M, Donovan BM, Labitt RN, Horwitz JA, Gaska JM. Hepatitis C virus infects rhesus macaque hepatocytes and simianized mice. Hepatology. 2015;62:57-67.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Su AI, Pezacki JP, Wodicka L, Brideau AD, Supekova L, Thimme R, Wieland S, Bukh J, Purcell RH, Schultz PG. Genomic analysis of the host response to hepatitis C virus infection. Proc Natl Acad Sci USA. 2002;99:15669-15674.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Thimme R, Bukh J, Spangenberg HC, Wieland S, Pemberton J, Steiger C, Govindarajan S, Purcell RH, Chisari FV. Viral and immunological determinants of hepatitis C virus clearance, persistence, and disease. Proc Natl Acad Sci USA. 2002;99:15661-15668.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Thimme R, Chang KM, Pemberton J, Sette A, Chisari FV. Degenerate immunogenicity of an HLA-A2-restricted hepatitis B virus nucleocapsid cytotoxic T-lymphocyte epitope that is also presented by HLA-B51. J Virol. 2001;75:3984-3987.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Thimme R, Wieland S, Steiger C, Ghrayeb J, Reimann KA, Purcell RH, Chisari FV. CD8(+) T cells mediate viral clearance and disease pathogenesis during acute hepatitis B virus infection. J Virol. 2003;77:68-76.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Wang J, Holmes TH, Cheung R, Greenberg HB, He XS. Expression of chemokine receptors on intrahepatic and peripheral lymphocytes in chronic hepatitis C infection: its relationship to liver inflammation. J Infect Dis. 2004;190:989-997.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Grakoui A, Shoukry NH, Woollard DJ, Han JH, Hanson HL, Ghrayeb J, Murthy KK, Rice CM, Walker CM. HCV persistence and immune evasion in the absence of memory T cell help. Science. 2003;302:659-662.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Smyk-Pearson S, Golden-Mason L, Klarquist J, Burton JR, Tester IA, Wang CC, Culbertson N, Vandenbark AA, Rosen HR. Functional suppression by FoxP3+CD4+CD25(high) regulatory T cells during acute hepatitis C virus infection. J Infect Dis. 2008;197:46-57.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Schulze zur Wiesch J, Lauer GM, Day CL, Kim AY, Ouchi K, Duncan JE, Wurcel AG, Timm J, Jones AM, Mothe B. Broad repertoire of the CD4+ Th cell response in spontaneously controlled hepatitis C virus infection includes dominant and highly promiscuous epitopes. J Immunol. 2005;175:3603-3613.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Sugimoto K, Ikeda F, Stadanlick J, Nunes FA, Alter HJ, Chang KM. Suppression of HCV-specific T cells without differential hierarchy demonstrated ex vivo in persistent HCV infection. Hepatology. 2003;38:1437-1448.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Bowen DG, Shoukry NH, Grakoui A, Fuller MJ, Cawthon AG, Dong C, Hasselschwert DL, Brasky KM, Freeman GJ, Seth NP. Variable patterns of programmed death-1 expression on fully functional memory T cells after spontaneous resolution of hepatitis C virus infection. J Virol. 2008;82:5109-5114.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Golden-Mason L, Palmer B, Klarquist J, Mengshol JA, Castelblanco N, Rosen HR. Upregulation of PD-1 expression on circulating and intrahepatic hepatitis C virus-specific CD8+ T cells associated with reversible immune dysfunction. J Virol. 2007;81:9249-9258.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Golden-Mason L, Klarquist J, Wahed AS, Rosen HR. Cutting edge: programmed death-1 expression is increased on immunocytes in chronic hepatitis C virus and predicts failure of response to antiviral therapy: race-dependent differences. J Immunol. 2008;180:3637-3641.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Abdel-Hakeem MS, Bédard N, Badr G, Ostrowski M, Sékaly RP, Bruneau J, Willems B, Heathcote EJ, Shoukry NH. Comparison of immune restoration in early versus late alpha interferon therapy against hepatitis C virus. J Virol. 2010;84:10429-10435.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 44]  [Cited by in F6Publishing: 48]  [Article Influence: 3.4]  [Reference Citation Analysis (0)]
24.  Barnes E, Gelderblom HC, Humphreys I, Semmo N, Reesink HW, Beld MG, van Lier RA, Klenerman P. Cellular immune responses during high-dose interferon-alpha induction therapy for hepatitis C virus infection. J Infect Dis. 2009;199:819-828.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 34]  [Cited by in F6Publishing: 36]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
25.  Missale G, Pilli M, Zerbini A, Penna A, Ravanetti L, Barili V, Orlandini A, Molinari A, Fasano M, Santantonio T. Lack of full CD8 functional restoration after antiviral treatment for acute and chronic hepatitis C virus infection. Gut. 2012;61:1076-1084.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 46]  [Cited by in F6Publishing: 48]  [Article Influence: 4.0]  [Reference Citation Analysis (0)]
26.  Odorizzi PM, Wherry EJ. Immunology. An interferon paradox. Science. 2013;340:155-156.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 49]  [Cited by in F6Publishing: 50]  [Article Influence: 4.5]  [Reference Citation Analysis (0)]
27.  Welsh RM, Bahl K, Marshall HD, Urban SL. Type 1 interferons and antiviral CD8 T-cell responses. PLoS Pathog. 2012;8:e1002352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 129]  [Cited by in F6Publishing: 134]  [Article Influence: 11.2]  [Reference Citation Analysis (0)]
28.  Martin B, Hennecke N, Lohmann V, Kayser A, Neumann-Haefelin C, Kukolj G, Böcher WO, Thimme R. Restoration of HCV-specific CD8+ T cell function by interferon-free therapy. J Hepatol. 2014;61:538-543.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Baumert TF, Fauvelle C, Chen DY, Lauer GM. A prophylactic hepatitis C virus vaccine: a distant peak still worth climbing. J Hepatol. 2014;61:S34-S44.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 60]  [Cited by in F6Publishing: 65]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
30.  Tarr AW, Urbanowicz RA, Ball JK. The role of humoral innate immunity in hepatitis C virus infection. Viruses. 2012;4:1-27.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 35]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
31.  Folgori A, Capone S, Ruggeri L, Meola A, Sporeno E, Ercole BB, Pezzanera M, Tafi R, Arcuri M, Fattori E. A T-cell HCV vaccine eliciting effective immunity against heterologous virus challenge in chimpanzees. Nat Med. 2006;12:190-197.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Barnes E, Folgori A, Capone S, Swadling L, Aston S, Kurioka A, Meyer J, Huddart R, Smith K, Townsend R. Novel adenovirus-based vaccines induce broad and sustained T cell responses to HCV in man. Sci Transl Med. 2012;4:115ra1.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 321]  [Cited by in F6Publishing: 317]  [Article Influence: 26.4]  [Reference Citation Analysis (0)]
33.  Ray R, Meyer K, Banerjee A, Basu A, Coates S, Abrignani S, Houghton M, Frey SE, Belshe RB. Characterization of antibodies induced by vaccination with hepatitis C virus envelope glycoproteins. J Infect Dis. 2010;202:862-866.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 80]  [Article Influence: 5.7]  [Reference Citation Analysis (0)]
34.  de Jong YP, Dorner M, Mommersteeg MC, Xiao JW, Balazs AB, Robbins JB, Winer BY, Gerges S, Vega K, Labitt RN. Broadly neutralizing antibodies abrogate established hepatitis C virus infection. Sci Transl Med. 2014;6:254ra129.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 175]  [Cited by in F6Publishing: 177]  [Article Influence: 19.7]  [Reference Citation Analysis (0)]
35.  Kong L, Giang E, Nieusma T, Kadam RU, Cogburn KE, Hua Y, Dai X, Stanfield RL, Burton DR, Ward AB. Hepatitis C virus E2 envelope glycoprotein core structure. Science. 2013;342:1090-1094.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 314]  [Cited by in F6Publishing: 311]  [Article Influence: 28.3]  [Reference Citation Analysis (0)]
36.  Thomson EC, Smith JA, Klenerman P. The natural history of early hepatitis C virus evolution; lessons from a global outbreak in human immunodeficiency virus-1-infected individuals. J Gen Virol. 2011;92:2227-2236.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 18]  [Cited by in F6Publishing: 20]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
37.  Schoggins JW, Rice CM. Interferon-stimulated genes and their antiviral effector functions. Curr Opin Virol. 2011;1:519-525.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 859]  [Cited by in F6Publishing: 888]  [Article Influence: 74.0]  [Reference Citation Analysis (0)]
38.  Saito T, Owen DM, Jiang F, Marcotrigiano J, Gale M. Innate immunity induced by composition-dependent RIG-I recognition of hepatitis C virus RNA. Nature. 2008;454:523-527.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 555]  [Cited by in F6Publishing: 557]  [Article Influence: 34.8]  [Reference Citation Analysis (0)]
39.  Malathi K, Saito T, Crochet N, Barton DJ, Gale M, Silverman RH. RNase L releases a small RNA from HCV RNA that refolds into a potent PAMP. RNA. 2010;16:2108-2119.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 110]  [Cited by in F6Publishing: 107]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
40.  Liu HM, Loo YM, Horner SM, Zornetzer GA, Katze MG, Gale M. The mitochondrial targeting chaperone 14-3-3ε regulates a RIG-I translocon that mediates membrane association and innate antiviral immunity. Cell Host Microbe. 2012;11:528-537.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 173]  [Cited by in F6Publishing: 158]  [Article Influence: 13.2]  [Reference Citation Analysis (0)]
41.  Saito T, Hirai R, Loo YM, Owen D, Johnson CL, Sinha SC, Akira S, Fujita T, Gale M. Regulation of innate antiviral defenses through a shared repressor domain in RIG-I and LGP2. Proc Natl Acad Sci USA. 2007;104:582-587.  [PubMed]  [DOI]  [Cited in This Article: ]
42.  Gack MU, Shin YC, Joo CH, Urano T, Liang C, Sun L, Takeuchi O, Akira S, Chen Z, Inoue S. TRIM25 RING-finger E3 ubiquitin ligase is essential for RIG-I-mediated antiviral activity. Nature. 2007;446:916-920.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Horner SM, Gale M. Regulation of hepatic innate immunity by hepatitis C virus. Nat Med. 2013;19:879-888.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 229]  [Cited by in F6Publishing: 225]  [Article Influence: 20.5]  [Reference Citation Analysis (0)]
44.  Kawai T, Takahashi K, Sato S, Coban C, Kumar H, Kato H, Ishii KJ, Takeuchi O, Akira S. IPS-1, an adaptor triggering RIG-I- and Mda5-mediated type I interferon induction. Nat Immunol. 2005;6:981-988.  [PubMed]  [DOI]  [Cited in This Article: ]
45.  Yamamoto M, Sato S, Hemmi H, Hoshino K, Kaisho T, Sanjo H, Takeuchi O, Sugiyama M, Okabe M, Takeda K. Role of adaptor TRIF in the MyD88-independent toll-like receptor signaling pathway. Science. 2003;301:640-643.  [PubMed]  [DOI]  [Cited in This Article: ]
46.  Hemmi H, Takeuchi O, Sato S, Yamamoto M, Kaisho T, Sanjo H, Kawai T, Hoshino K, Takeda K, Akira S. The roles of two IkappaB kinase-related kinases in lipopolysaccharide and double stranded RNA signaling and viral infection. J Exp Med. 2004;199:1641-1650.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Lee HC, Narayanan S, Park SJ, Seong SY, Hahn YS. Transcriptional regulation of IFN-λ genes in hepatitis C virus-infected hepatocytes via IRF-3·IRF-7·NF-κB complex. J Biol Chem. 2014;289:5310-5319.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 32]  [Cited by in F6Publishing: 35]  [Article Influence: 3.5]  [Reference Citation Analysis (0)]
48.  Sharma S, tenOever BR, Grandvaux N, Zhou GP, Lin R, Hiscott J. Triggering the interferon antiviral response through an IKK-related pathway. Science. 2003;300:1148-1151.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Sommereyns C, Paul S, Staeheli P, Michiels T. IFN-lambda (IFN-lambda) is expressed in a tissue-dependent fashion and primarily acts on epithelial cells in vivo. PLoS Pathog. 2008;4:e1000017.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 578]  [Cited by in F6Publishing: 610]  [Article Influence: 38.1]  [Reference Citation Analysis (0)]
50.  Platanias LC. Mechanisms of type-I- and type-II-interferon-mediated signalling. Nat Rev Immunol. 2005;5:375-386.  [PubMed]  [DOI]  [Cited in This Article: ]
51.  Shuai K, Ziemiecki A, Wilks AF, Harpur AG, Sadowski HB, Gilman MZ, Darnell JE. Polypeptide signalling to the nucleus through tyrosine phosphorylation of Jak and Stat proteins. Nature. 1993;366:580-583.  [PubMed]  [DOI]  [Cited in This Article: ]
52.  Tenoever BR, Ng SL, Chua MA, McWhirter SM, García-Sastre A, Maniatis T. Multiple functions of the IKK-related kinase IKKepsilon in interferon-mediated antiviral immunity. Science. 2007;315:1274-1278.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Rajsbaum R, Versteeg GA, Schmid S, Maestre AM, Belicha-Villanueva A, Martínez-Romero C, Patel JR, Morrison J, Pisanelli G, Miorin L. Unanchored K48-linked polyubiquitin synthesized by the E3-ubiquitin ligase TRIM6 stimulates the interferon-IKKε kinase-mediated antiviral response. Immunity. 2014;40:880-895.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 118]  [Article Influence: 11.8]  [Reference Citation Analysis (0)]
54.  Mukherjee A, Di Bisceglie AM, Ray RB. Hepatitis C virus-mediated enhancement of microRNA miR-373 impairs the JAK/STAT signaling pathway. J Virol. 2015;89:3356-3365.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 69]  [Cited by in F6Publishing: 68]  [Article Influence: 7.6]  [Reference Citation Analysis (0)]
55.  Lin W, Choe WH, Hiasa Y, Kamegaya Y, Blackard JT, Schmidt EV, Chung RT. Hepatitis C virus expression suppresses interferon signaling by degrading STAT1. Gastroenterology. 2005;128:1034-1041.  [PubMed]  [DOI]  [Cited in This Article: ]
56.  Stevenson NJ, Bourke NM, Ryan EJ, Binder M, Fanning L, Johnston JA, Hegarty JE, Long A, O’Farrelly C. Hepatitis C virus targets the interferon-α JAK/STAT pathway by promoting proteasomal degradation in immune cells and hepatocytes. FEBS Lett. 2013;587:1571-1578.  [PubMed]  [DOI]  [Cited in This Article: ]
57.  Muir AJ, Shiffman ML, Zaman A, Yoffe B, de la Torre A, Flamm S, Gordon SC, Marotta P, Vierling JM, Lopez-Talavera JC. Phase 1b study of pegylated interferon lambda 1 with or without ribavirin in patients with chronic genotype 1 hepatitis C virus infection. Hepatology. 2010;52:822-832.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 192]  [Cited by in F6Publishing: 181]  [Article Influence: 12.9]  [Reference Citation Analysis (0)]
58.  Ramos EL. Preclinical and clinical development of pegylated interferon-lambda 1 in chronic hepatitis C. J Interferon Cytokine Res. 2010;30:591-595.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Sarasin-Filipowicz M, Wang X, Yan M, Duong FH, Poli V, Hilton DJ, Zhang DE, Heim MH. Alpha interferon induces long-lasting refractoriness of JAK-STAT signaling in the mouse liver through induction of USP18/UBP43. Mol Cell Biol. 2009;29:4841-4851.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 145]  [Cited by in F6Publishing: 141]  [Article Influence: 9.4]  [Reference Citation Analysis (0)]
60.  Makowska Z, Duong FH, Trincucci G, Tough DF, Heim MH. Interferon-β and interferon-λ signaling is not affected by interferon-induced refractoriness to interferon-α in vivo. Hepatology. 2011;53:1154-1163.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 85]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
61.  Muir AJ, Arora S, Everson G, Flisiak R, George J, Ghalib R, Gordon SC, Gray T, Greenbloom S, Hassanein T. A randomized phase 2b study of peginterferon lambda-1a for the treatment of chronic HCV infection. J Hepatol. 2014;61:1238-1246.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 105]  [Cited by in F6Publishing: 107]  [Article Influence: 10.7]  [Reference Citation Analysis (0)]
62.  Friborg J, Ross-Macdonald P, Cao J, Willard R, Lin B, Eggers B, McPhee F. Impairment of type I but not type III IFN signaling by hepatitis C virus infection influences antiviral responses in primary human hepatocytes. PLoS One. 2015;10:e0121734.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 8]  [Cited by in F6Publishing: 8]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
63.  Li K, Foy E, Ferreon JC, Nakamura M, Ferreon AC, Ikeda M, Ray SC, Gale M, Lemon SM. Immune evasion by hepatitis C virus NS3/4A protease-mediated cleavage of the Toll-like receptor 3 adaptor protein TRIF. Proc Natl Acad Sci USA. 2005;102:2992-2997.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Wang Y, Li J, Wang X, Ye L, Zhou Y, Thomas RM, Ho W. Hepatitis C virus impairs TLR3 signaling and inhibits IFN-λ 1 expression in human hepatoma cell line. Innate Immun. 2014;20:3-11.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 16]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
65.  Oudshoorn D, Versteeg GA, Kikkert M. Regulation of the innate immune system by ubiquitin and ubiquitin-like modifiers. Cytokine Growth Factor Rev. 2012;23:273-282.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 23]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
66.  Rajsbaum R, García-Sastre A. Viral evasion mechanisms of early antiviral responses involving regulation of ubiquitin pathways. Trends Microbiol. 2013;21:421-429.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 53]  [Article Influence: 4.8]  [Reference Citation Analysis (0)]
67.  Banerjee I, Miyake Y, Nobs SP, Schneider C, Horvath P, Kopf M, Matthias P, Helenius A, Yamauchi Y. Influenza A virus uses the aggresome processing machinery for host cell entry. Science. 2014;346:473-477.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 174]  [Cited by in F6Publishing: 181]  [Article Influence: 18.1]  [Reference Citation Analysis (0)]
68.  Rajsbaum R, García-Sastre A. Virology. Unanchored ubiquitin in virus uncoating. Science. 2014;346:427-428.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 13]  [Cited by in F6Publishing: 16]  [Article Influence: 1.6]  [Reference Citation Analysis (0)]
69.  Oshiumi H, Miyashita M, Matsumoto M, Seya T. A distinct role of Riplet-mediated K63-Linked polyubiquitination of the RIG-I repressor domain in human antiviral innate immune responses. PLoS Pathog. 2013;9:e1003533.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 163]  [Cited by in F6Publishing: 159]  [Article Influence: 14.5]  [Reference Citation Analysis (0)]
70.  Barr SD, Smiley JR, Bushman FD. The interferon response inhibits HIV particle production by induction of TRIM22. PLoS Pathog. 2008;4:e1000007.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 212]  [Cited by in F6Publishing: 218]  [Article Influence: 13.6]  [Reference Citation Analysis (0)]
71.  Carthagena L, Bergamaschi A, Luna JM, David A, Uchil PD, Margottin-Goguet F, Mothes W, Hazan U, Transy C, Pancino G. Human TRIM gene expression in response to interferons. PLoS One. 2009;4:e4894.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 191]  [Cited by in F6Publishing: 206]  [Article Influence: 13.7]  [Reference Citation Analysis (0)]
72.  Tissot C, Mechti N. Molecular cloning of a new interferon-induced factor that represses human immunodeficiency virus type 1 long terminal repeat expression. J Biol Chem. 1995;270:14891-14898.  [PubMed]  [DOI]  [Cited in This Article: ]
73.  Sadeghi F, Bokharaei-Salim F, Salehi-Vaziri M, Monavari SH, Alavian SM, Salimi S, Vahabpour R, Keyvani H. Associations between human TRIM22 gene expression and the response to combination therapy with Peg-IFNα-2a and ribavirin in Iranian patients with chronic hepatitis C. J Med Virol. 2014;86:1499-1506.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 19]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
74.  Yang C, Zhao X, Sun D, Yang L, Chong C, Pan Y, Chi X, Gao Y, Wang M, Shi X. Interferon alpha (IFNα)-induced TRIM22 interrupts HCV replication by ubiquitinating NS5A. Cell Mol Immunol. 2015;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 65]  [Cited by in F6Publishing: 66]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
75.  Li Q, Zhang YY, Chiu S, Hu Z, Lan KH, Cha H, Sodroski C, Zhang F, Hsu CS, Thomas E. Integrative functional genomics of hepatitis C virus infection identifies host dependencies in complete viral replication cycle. PLoS Pathog. 2014;10:e1004163.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 78]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
76.  Lee HS, Lim YS, Park EM, Baek SH, Hwang SB. SUMOylation of nonstructural 5A protein regulates hepatitis C virus replication. J Viral Hepat. 2014;21:e108-e117.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 15]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
77.  Morales DJ, Lenschow DJ. The antiviral activities of ISG15. J Mol Biol. 2013;425:4995-5008.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 163]  [Article Influence: 14.8]  [Reference Citation Analysis (0)]
78.  Kim MJ, Yoo JY. Inhibition of hepatitis C virus replication by IFN-mediated ISGylation of HCV-NS5A. J Immunol. 2010;185:4311-4318.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 42]  [Cited by in F6Publishing: 44]  [Article Influence: 3.1]  [Reference Citation Analysis (0)]
79.  Chen L, Sun J, Meng L, Heathcote J, Edwards AM, McGilvray ID. ISG15, a ubiquitin-like interferon-stimulated gene, promotes hepatitis C virus production in vitro: implications for chronic infection and response to treatment. J Gen Virol. 2010;91:382-388.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 82]  [Cited by in F6Publishing: 87]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
80.  Broering R, Zhang X, Kottilil S, Trippler M, Jiang M, Lu M, Gerken G, Schlaak JF. The interferon stimulated gene 15 functions as a proviral factor for the hepatitis C virus and as a regulator of the IFN response. Gut. 2010;59:1111-1119.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 81]  [Article Influence: 5.8]  [Reference Citation Analysis (0)]
81.  Malakhov MP, Malakhova OA, Kim KI, Ritchie KJ, Zhang DE. UBP43 (USP18) specifically removes ISG15 from conjugated proteins. J Biol Chem. 2002;277:9976-9981.  [PubMed]  [DOI]  [Cited in This Article: ]
82.  Malakhova OA, Yan M, Malakhov MP, Yuan Y, Ritchie KJ, Kim KI, Peterson LF, Shuai K, Zhang DE. Protein ISGylation modulates the JAK-STAT signaling pathway. Genes Dev. 2003;17:455-460.  [PubMed]  [DOI]  [Cited in This Article: ]
83.  Randall G, Chen L, Panis M, Fischer AK, Lindenbach BD, Sun J, Heathcote J, Rice CM, Edwards AM, McGilvray ID. Silencing of USP18 potentiates the antiviral activity of interferon against hepatitis C virus infection. Gastroenterology. 2006;131:1584-1591.  [PubMed]  [DOI]  [Cited in This Article: ]
84.  Real CI, Megger DA, Sitek B, Jahn-Hofmann K, Ickenstein LM, John MJ, Walker A, Timm J, Kuhlmann K, Eisenacher M. Identification of proteins that mediate the pro-viral functions of the interferon stimulated gene 15 in hepatitis C virus replication. Antiviral Res. 2013;100:654-661.  [PubMed]  [DOI]  [Cited in This Article: ]
85.  Munakata T, Liang Y, Kim S, McGivern DR, Huibregtse J, Nomoto A, Lemon SM. Hepatitis C virus induces E6AP-dependent degradation of the retinoblastoma protein. PLoS Pathog. 2007;3:1335-1347.  [PubMed]  [DOI]  [Cited in This Article: ]
86.  Munakata T, Nakamura M, Liang Y, Li K, Lemon SM. Down-regulation of the retinoblastoma tumor suppressor by the hepatitis C virus NS5B RNA-dependent RNA polymerase. Proc Natl Acad Sci USA. 2005;102:18159-18164.  [PubMed]  [DOI]  [Cited in This Article: ]
87.  Starr R, Willson TA, Viney EM, Murray LJ, Rayner JR, Jenkins BJ, Gonda TJ, Alexander WS, Metcalf D, Nicola NA. A family of cytokine-inducible inhibitors of signalling. Nature. 1997;387:917-921.  [PubMed]  [DOI]  [Cited in This Article: ]
88.  Shoji I. Roles of the two distinct proteasome pathways in hepatitis C virus infection. World J Virol. 2012;1:44-50.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 14]  [Cited by in F6Publishing: 14]  [Article Influence: 1.2]  [Reference Citation Analysis (0)]
89.  Raychoudhuri A, Shrivastava S, Steele R, Kim H, Ray R, Ray RB. ISG56 and IFITM1 proteins inhibit hepatitis C virus replication. J Virol. 2011;85:12881-12889.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 126]  [Cited by in F6Publishing: 131]  [Article Influence: 10.1]  [Reference Citation Analysis (0)]
90.  Wilkins C, Woodward J, Lau DT, Barnes A, Joyce M, McFarlane N, McKeating JA, Tyrrell DL, Gale M. IFITM1 is a tight junction protein that inhibits hepatitis C virus entry. Hepatology. 2013;57:461-469.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 122]  [Cited by in F6Publishing: 123]  [Article Influence: 11.2]  [Reference Citation Analysis (0)]
91.  Metz P, Dazert E, Ruggieri A, Mazur J, Kaderali L, Kaul A, Zeuge U, Windisch MP, Trippler M, Lohmann V. Identification of type I and type II interferon-induced effectors controlling hepatitis C virus replication. Hepatology. 2012;56:2082-2093.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 110]  [Article Influence: 9.2]  [Reference Citation Analysis (0)]
92.  Brass AL, Huang IC, Benita Y, John SP, Krishnan MN, Feeley EM, Ryan BJ, Weyer JL, van der Weyden L, Fikrig E. The IFITM proteins mediate cellular resistance to influenza A H1N1 virus, West Nile virus, and dengue virus. Cell. 2009;139:1243-1254.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 945]  [Cited by in F6Publishing: 988]  [Article Influence: 70.6]  [Reference Citation Analysis (0)]
93.  Jiang D, Guo H, Xu C, Chang J, Gu B, Wang L, Block TM, Guo JT. Identification of three interferon-inducible cellular enzymes that inhibit the replication of hepatitis C virus. J Virol. 2008;82:1665-1678.  [PubMed]  [DOI]  [Cited in This Article: ]
94.  Park K, Scott AL. Cholesterol 25-hydroxylase production by dendritic cells and macrophages is regulated by type I interferons. J Leukoc Biol. 2010;88:1081-1087.  [PubMed]  [DOI]  [Cited in This Article: ]
95.  Ye J, Wang C, Sumpter R, Brown MS, Goldstein JL, Gale M. Disruption of hepatitis C virus RNA replication through inhibition of host protein geranylgeranylation. Proc Natl Acad Sci USA. 2003;100:15865-15870.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Chen Y, Wang S, Yi Z, Tian H, Aliyari R, Li Y, Chen G, Liu P, Zhong J, Chen X. Interferon-inducible cholesterol-25-hydroxylase inhibits hepatitis C virus replication via distinct mechanisms. Sci Rep. 2014;4:7242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 68]  [Cited by in F6Publishing: 78]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]