Open access peer-reviewed chapter

Host versus Virus: The Genetics in HCV Infection Leading to Treatment

Written By

Quratulain Maqsood, Maria Hussain and Aleena Sumrin

Submitted: 22 December 2022 Reviewed: 23 December 2022 Published: 17 February 2023

DOI: 10.5772/intechopen.1001050

From the Edited Volume

Hepatitis C - Recent Advances

Li Yang and Xingshun Qi

Chapter metrics overview

89 Chapter Downloads

View Full Metrics

Abstract

The spread of hepatitis C virus (HCV) infection is a worldwide crisis. Intricate host-viral interactions control the HCV infection’s natural course and treatment response according to new research. The patient’s HCV genotype is the best predictor of response to pegylated interferon plus ribavirin therapy. The most crucial viral factor in determining the efficacy of direct-acting antiviral therapy is the HCV genotype 1 subtype. In addition to baseline viral load and HCV genomic heterogeneity, these two factors are linked with the treatment response. In previous large genome-wide association studies, interferon3 gene polymorphisms have been shown to be linked with spontaneous clearance and treatment responsiveness. An inosine triphosphatase gene polymorphism has been shown to reduce the risk of anaemia and other side effects caused by the antiviral drug ribavirin. In HCV patients, a second genetic mutation in the three-gene patatin-like phospholipase domain is associated with hepatic steatosis and fibrosis. This study examined the effects of viral and host genetics on the course and results of HCV therapy while concentrating on the known viral and host variables linked to HCV patient outcomes. This will result in fresh concepts for individualising both preventative care and therapeutic treatment.

Keywords

  • HCV infection
  • HCV host interaction
  • chronic hepatitis
  • CD8+ and CD4+ T-cell
  • IFN
  • progenitor cells

1. Introduction

Hepatitis C virus infection (HCV) is the main cause of chronic hepatitis, which is estimated to impact 70 million individuals globally. Cirrhosis, hepatocellular carcinoma (HCC), and end-stage liver disease may all result from chronic HCV infection. After their first HCV exposure, only 20% of individuals spontaneously clear the virus; chronic hepatitis is often the result [1]. Adaptive immune responses are expected to have a significant role in the progression of HCV infection. Indeed, during spontaneous viral elimination, robust and broadly directed CD8+ and CD4+ T cell responses that are specific to the virus are produced and remain following HCV clearance [2]. Furthermore, there is mounting evidence that neutralising antibodies (nAb) may help in viral eradication. The emergence of multiple unique but closely similar viral variants is one strategy used by HCV to escape the adaptive immune response. Due to the complex population of HCV that circulates in vivo and is referred to be a “quasi-species” as a result of the high RNA-dependent RNA polymerase error rate and fast replication rate [3]. Adaptive immune responses favour the selection of persistent variant viruses that encode altered epitopes that, in this context, are either not recognised by T cells or antibodies at all or are only partly recognised by them. Additionally, several studies over the last 10 years have focussed on the functions of host genetic traits, such as favourable genotypes or alleles that may be a significant indication of clearance or be connected to a greater likelihood of sustained virological response (SVR) in treated patients [4]. In this chapter, we provide a summary of the knowledge about the role of host genetic factors in determining the course of HCV infection, as well as the role of antibodies and T cells in promoting in vivo HCV evolution [5].

Advertisement

2. Hepatitis C virus: an important virus for virology

In 1989, a virus that causes hepatitis (HCV) was first identified. In this year, conventional virology will give way to contemporary virology, which combines molecular science and biotechnology to research, characterise, and keep track of viruses. Traditional virology relied on the separation, growth, and biochemical study of viruses. HCV was the first infectious pathogen to be researched owing to molecular techniques, which are widely used to describe molecular components of HCV biology due to the virus’ difficulty in replicating in vitro. Nowadays, HCV is rewriting history. Direct-acting antivirals (DAAs), which totally remove infection in more than 90% of patients, have challenged the notion that antivirals can only inhibit viral reproduction and slow disease processes. IFN-free DAA regimens are quickly taking the place of poorly tolerated IFN-based treatment regimens, stabilising, and perhaps even reversing tissue degeneration in patients with severe disease [6]. The creation of host-directed antiviral medicines has also been aided by thorough research into how host factors influence HCV. This chapter provides information on various important elements of viral behaviour at the cellular and host phases with an emphasis on the fundamentals and also most recent clinical developments crucial in assessing susceptibility to DAAs. Additionally, it offers illustrations of modern methods for genotyping and tracking viral replication.

Advertisement

3. HCV: life cycle and in vitro interactions between hosts and cells

Hepacivirus, Flavivirus, Pestivirus, and Pegivirus genera make up the huge group of encapsulated, single-stranded RNA viruses described as the Flaviviridae, which includes HCV [7]. This family of viruses, that includes several viruses spread via arthropods, is a developing public health affair [8]. Our understanding of numerous molecular pathways has been hampered by the difficulties in developing the in vitro replication model and the large network of cell surface particles that facilitate virus replication, which has contributed to poor understanding of the life cycle of HCV. While moving through the bloodstream, the HCV virion can either be a free particle or be held in check by host low-density lipoproteins [9]. It then adheres to the specific cell membrane and undergoes clathrin-mediated endocytosis to enter the cell. The viral genome, a 9.6 kb single-stranded RNA having positive charge, is exposed into the cytoplasm during endocytic division when the viral capsid is broken [10]. The RNA genome directly translates into a single polyprotein precursor with about 3000 amino acid residues at the rough endoplasmic reticulum (ER). The precursor is then broken down into 10 mature products by cellular and viral proteases. Following a Golgi-dependent secretory pathway, new virions are accumulated in ER-derived compartments and cleared via exocytosis [11]. At this point, the virus has reached maturity and is encased in endogenous lipoproteins, that, as we will discuss below, is thought to aid immune escape. HCV virions have a reduced buoyant density and a broad size range because of their adhesion to host lipoproteins and also the absence of readily recognisable surface characteristics (40–80 nm diameter) [12]. In addition to supporting immune evasion, virus persistence, and DAA resistance, this process may also make it more challenging to pinpoint the cellular receptors required for viral entry. Additionally, it appears that the kind of cell affects the use of receptors as well as infection by free particles or cell to cell transmission [13]. HCV trans-complemented particles and cell culture-derived HCV (HCVcc) are two of the most often employed approaches for examining viral replication in vitro (HCVTCP). HCV genotype 2a variant JFH1, which has been obtained from a Japanese person having fulminant hepatitis, is used to reproduce HCVcc in Huh-7, a human cell line originating from hepatocellular cancer [14]. HCVcc produces infectious viruses, making it possible to characterise the shape and biochemical properties of virion particles, identify some HCV entry factors, and assess the effectiveness of DAA using native or inter-genotype transgenic JFH1 variations. Pseudotyped HCV virions are created by packaging transfected cells with viral proteins provided by various designs for HCVTCP, that have been fully detailed elsewhere [15]. Although HCVTCP can potentially be acquired from any isolate, it can only be infected once and cannot spread. Despite generally being thought to facilitate HCV replication, Huh-7 cells vary from native hepatocytes in regards to restriction mechanisms, HCV receptor localisation, and the absence of cell polarity seen in hepatic tissue [16]. Nearly all in vitro approaches, including HCVTCP and HCVcc, make use of Huh-7 cells. As a result, in vitro replication of viral entrance, organisation, discharge, and cell-to-cell dispersion is incomplete. HCV can replicate well in hepG2 cell clones and hepatoma cells created from primary hepatocytes, which may help the researchers understand better how virus interacts with its host cells [17].

Advertisement

4. Natural history of infection

Annual bacterial instances of up to 4 million cases are possible, while 130–200 million people worldwide are thought to have chronic HCV infections. Many of these afflicted people do not know whether they are infected or not. HCV mostly spreads through percutaneous contact with infected blood [18]. Many people acquired HCV before receiving a diagnosis as a result of poor injection techniques, infected blood, or infected blood products. In many developed countries, those who use injections or intranasal medications today suffer from the majority of infectious infections [19]. Particularly noteworthy, one of those often at risk of sexual transmission is HIV-positive men who participate in male-on-male sexual activity. Vertical transmission, transmission through piercings or tattoos, and other strategies are other strategies induced hepatic fibrosis and swelling may appear progressively over time. Cirrhosis, advanced liver disease, and HCC are long-term consequences [20, 21]. According to a large meta-analysis, over 20 years of infection, the likelihood of having cirrhosis rose from 7 to 18–41%. Men, ageing, drinking alcohol, and HIV co-infection are risk factors for fibrosis, cirrhosis, and HCC [22]. Although 15–30% of infected people may, the majority do not display symptoms during acute disease. It is challenging to anticipate how many people with acute HCV infections will be cured of the disease without therapy because the majority of these infections are subclinical. Between 20 and 50% of individuals are thought to have spontaneous clearance within the initial 6 months of exposure; this is impacted by genetic factors, race, age, sexual orientation, and chronic diseases like HIV [23].

Advertisement

5. HCV and host interaction

The lipid-centric HCV virus can enter suitable host cells thanks to its 2 envelope glycoproteins, E2 and E1. To enter the cell, the 2 glycoproteins link with CD81 and the other external membrane proteins that are occludin, claudin-1, and the epidermal growth factor receptor. Clathrin-mediated endocytosis, where the nucleocapsid is expelled into the cytoplasm, is required for the virus to replicate within the target cell [24]. The host’s immune system has access to the HCV genome after nucleocapsid is liberated into the cytoplasm. HCV proteins are translated via an internal ribosome binding domain that detects positive-strand RNA (IRES). During ER-related processing, HCV transforms a large polyprotein into structural and nonstructural proteins [25]. This is done by cellular and viral proteases. HCV RNA is unwound and stabilised in a reintegration complex with the help of NS5B and NS3 helicase regions, which control HCV replication. The development of “membranous web” structures, which serve as HCV replication chambers, is aided by NS4B [26]. Numerous host factors also promote HCV replication, such as Cyclophilin A, which interfaces with NS5A and NS5B to stimulate multiplication, and microRNA-122, which attaches to IRES to improve translation efficiency [27]. For assembly and release, HCV also makes advantage of fatty acid pathways and the creation of extremely low-density lipoprotein (VLDL). The HCV life cycle is shown in Figure 1 with an emphasis on the critical phases of Virus replication, such as HCV adherence and entry into the host organism, HCV RNA reproduction, viral assembling and release, and HCV RNA translation to generate a giant polyprotein that is translated into 10 HCV proteins [28].

Figure 1.

Replication of hepatitis C virus inside the host cell [29]. (Licence Number: 5445310176889).

Advertisement

6. Host response and outcomes of the HCV infection

6.1 Innate immunity

When cells within the liver produce a variety of IFN-stimulated genes (ISGs), which restrict HCV replication and dissemination, the rate of rise drastically lowers. HCV has an exponential “ramp-up” stage of replication quickly after being infected in hepatic foci [30]. Adaptive immunity is sparked by innate immunity, which guards against HCV infection. The mitochondrial antiviral signalling (MAVS) proteins are induced by an interaction between the HCV RNA and the retinoic acid-inducible gene I [31]. Interferon-gamma (IFN-γ) production is started when dual RNA attaches to the Toll-like receptor-3 and signals through an adapter that contains a TIR domain (TRIF). Both procedures aid in the movement of IRF3 and NFB into the nucleus [32]. They encourage the production of pro-inflammatory cytokines and proinflammatory cytokines to mobilise and excite immune cells, as well as IFN and ISG to inhibit viral replication HCV’s NS3-4A protease targets MAVS and TRIF to penetrate them, preventing the induction of IFN. Innate antiviral defences may be affected by HCV in a number of different ways [33]. ISG expression shows that even in HCV-infected hepatocytes with MAVS breakdown, these pathways do not completely suppress innate immunity. After HCV infection, hepatocytes selectively express IFN-γ. Progenitor cells, granulocytes, and other nonparenchymal cells can recognise viral molecular sequences and participate in IFN and cytokine secretion and response without increasing HCV replication [34].

6.1.1 The function of adaptive immune system in the development of infection

Innate antiviral responses rarely aid in the complete eradication of infection, despite the fact that they can restrict HCV multiplication and transmission in the absence of the host’s natural defensive mechanism. Blood transaminase levels rise as viral loads decrease, a sign of hepatocyte cell death. An infection is deemed chronic when it persists for more than 6 months. Adaptive immunity initiates a cell-mediated reaction that targets several HCV epitopes and generates strong, broadly reactive neutralising antibodies to end infections (bNAbs) [35]. To lessen the chance of viral immunological escape, T cells concentrate on a variety of epitopes. HCV’s ineffective reproductive strategy promotes quick evolution, immunological responses, and the selection of undetectable variations. Certain immunological escape mutations are undesirable because they decrease the virus’s survival [36]. A second indication of successful anti-HCV immunity is the maintenance of polyfunctional T-lymphocyte activity. Whatever the outcome, HCV-selective CD4+ T lymphocytes are necessary for CD8+ T lymphocytes (and other immune cells) to function. By encouraging CD8+ T-lymphocyte survival, multiplication, and antiviral activity, HCV-specific CD4+ T-lymphocytes contribute to the healing of infection. To halt viral replication, effector T-lymphocytes assemble in the liver wherein they release the cytokines IFN and TNF, kill infected cells, and destroy infected tissues [37]. Key viral neutralising targets or stable domains necessary for hepatocyte infection are concealed by glycans, lipoproteins, and non-conserved decoy domains. BNAbs bind the necessary domains in place of the decoys. BNAbs may aid in eradication because HCV must continuously infect various target cells to retain even an established infection. T-lymphocytes in recurring HCV infection may concentrate on a narrower spectrum of epitopes; an initially broad response typically narrows [38]. This is in contrast towards the extensive and persistent T-lymphocyte responses documented in the remission of HCV infection. Thus, immunological escape requires less alteration of the viral code. Immune-mediated selection for variations that evade CD8+ T cell identification is typically seen in chronic HCV infection. When viral antigenic patterns are intact, HCV-specific T-lymphocyte responses are characterised by a gradual function loss. Unrelated to epitope escape, a mechanism prevents CD4+ T-lymphocyte responses [39]. Without support from CD4+ T cells, CD8+ T cells cease growing, show signs of tiredness, and lose their capacity for communication. Damage to liver tissue may result from infiltrating inflammatory T cells, that are not always HCV-specific. Not to mention, individuals with latent infection have neutralising antibodies, though they may not manifest straight away and might be isolate-specific, often target hypervariable epitopes with increasing immunological escape potential [40].

6.1.2 The influence of the host’s genetics on the infection’s results

IFN-stimulated genes are linked to the collapse of IFN-based antiviral therapy, and individuals with these alleles are more probable to still have HCV infection. Increased ISG activity in the affected liver also suggests a bad outlook for IFN-based HCV therapy [41]. The frame-shifted IFN-4 gene in the IFN-β locus polymorphisms group suppresses the expression of IFN-4 protein as shown in Figure 2. IFN-4 may influence hepatocyte IFN responsiveness through negative feedback mechanisms, according to a theory [42]. IFN-4, on the other hand, may promote ongoing innate immune activation and impede the development of adaptive immunological responses. It is unclear if IFN-4 controls only innate immunity or also influences adaptive immunity. At the HLA locus, there are further significant polymorphisms [43].

Figure 2.

The interferon lamda 3–4 (IFNL) gene, IFN 1, and two loci on chromosome 19q13 are represented, and these regions include important single nucleotide polymorphisms (SNPs) implicated in the elimination of HCV.

6.1.3 Implications of the HVC diversity in transmission and pathogenesis

The onset of disease and viral evolution have been linked in studies utilising serially sampled HCV sequences. Cirrhosis, liver injury, and death can result from CHC, which can develop gradually and dependably or quickly. The results of the analysis of serial prospective study samples from hepatitis C patients associated with transfusions revealed a correlation between rapid illness progression and larger viral quasispecies variety and divergence, as well as higher percentage of synonymous substitution [44]. The mean substitution rates across all investigated viral covering sections were higher in instances with rapid infection progression over the course of the first 7 years of infection than in those with slower progression. Particularly for synonymous replacements, occurrences that advanced more quickly had a higher mean total number of changes per site [45]. These findings suggest that shorter viral generation times are associated with rapid disease onset, much like HIV-1. By doing a phylogenetic analysis on the full HVR1 amino acid sequence from every individual at different times in time, 2 topology sequences based on the progression of the disease were identified. Sequences from different eras blended together to form a mostly monophyletic community after hepatitis cases were settled [46]. Progressive hepatitis episodes showed a tendency for grouping over time and consistently longer branch lengths than instances of acute clearing hepatitis. This early neutralising restriction unexpectedly predicted the slow clinical progression of CHC. Patients with CHC who have cirrhosis have a 1–5% annual risk of developing HCC and a 3–6% yearly risk of hepatic decompensation after cirrhosis, both of which could necessitate an orthotopic liver transplant or lead to liver related death [47]. Additionally, both inside and outside the HVR1, the serum of HCC patients had a greater genetic diversity. On the partition of liver viruses, there is still a lot of disagreement. There was no proof of intra-hepatic E1/E2 quasispecies fragmentation when liver transplant recipients with end-stage liver disease were examined [48]. HCV enters cells by a complex mechanism involving a number of host proteins. In the viral life cycle, entry is a critical stage that may affect the evolution and diversity of the virus over time. The findings showed that there were 1–37 or more viruses that were transmitted and caused productive clinical infections [49].

6.1.4 HCV therapy

It is more challenging to develop vaccines, antiviral medications, and host defence awareness because of the diversity of quasispecies. Despite immunological activity, HCV frequently succeeds in getting past the host defence and maintaining a prolonged infection. There is general agreement that those who have CHC have compromised and lost their adaptive immunity [50]. CHC individuals who get antiviral therapy for viremia clearing; these individuals are still prone to infection after treatment has ended. The potential for re-infections from the same HCV subtype shows how difficult it is to develop a preventive HCV vaccination. Viral escape mutations are most common in the first 6 months following infection and can alter up to 50% of the CD8+ T cell-targeted epitopes [51]. Viral escape mutations are additionally uncommon during CHC, which might mean that T cell-mediated selective factors are not existent at this time. The way that specific CD8+ T lymphocytes are affected by HCV genetic mutations in terms of their capacity to reproduce the virus is essential. When the selected immunological pressure is released, the virus’ fitness declines, which results in balancing substitutions or their swift reversion [52]. The preserved virus genomic portions are frequently targeted when viruses associated with these HLAs are subjected to CD8+ T cell responses, and escaping mutations are not well tolerated. These conserved epitopes are the obvious targets for developing a potent HCV T cell-based vaccine. To create a potent HCV vaccine, the correlations of protective immunity to combat HCV variation must be clarified. Despite the challenges in developing an HCV preventative vaccine, antiviral research has been successful in treating the illness [53]. The cure rate for CHC has increased to almost 95% when direct-acting antiviral (DAA) medication was introduced in 2014 [54].

An important development in the treatment of HCV infection was the creation of the NS5B polymerase inhibitor sofosbuvir (SOF). Early chain termination happens as a result of the integration of SOF into newly synthesised viral RNA. NS5B’s conserved active area is the target of SOF, which is effective against all HCV genotypes and has a high resistance barrier. Although SOF tolerance is sometimes shown in vivo, tissue culture can produce it [55]. Those with the SOF resistance mutation have very little viral replication (NS5B at location S282T). After 12 weeks of treatment, SOF, pegylated interferon alfa-2a, and ribavirin produced an SVR rate of 90% in patients with genotypes 1 and 4 infection. Similar to this, 12 weeks of treatment with an oral SOF + ribavirin combination led to SVR rates of 95% and 82%, respectively, in people with genotypes 2 and 3, in both treatment-experienced and naive patients [56]. In 2014, SOF and the NS5A inhibitor ledipasvir (LDV) were approved as a once-daily co-formulation for the treatment of HCV genotype 1. This combination was designed to quickly stop viral replication and prevent the spread of resistant strains [57]. After the course of 12 weeks of treatment, the SOF/LDV combination achieved SVR rates of 94–99% both with and without ribavirin. For the treatment of HCV genotype 1, SOF and SMV received approval in 2014 [57]. The NS3/4A protease inhibitors paritaprevir, ombitasvir, dasabuvir, and non-nucleoside NS5B polymerase inhibitors, as well as other protease inhibitors, can all be considerably improved by the HIV-1 protease inhibitor ritonavir. Patients with severe liver disease and all HCV genotypes were included in the investigations. SVR rates of more than 95% were achieved with regimens needing only 8 weeks of treatment, which is significant [58]. Another important factor is that innovative DAAs are extremely effective in some populations, such as elderly people, IDUs, people with severe liver disease, chronic renal illness, hemoglobinopathies, and HIV-1/HCV coinfection. In order to give medications with a higher metabolic profile and a wider ability to inhibit various HCV genotypes and variants, new DAAs are being developed [59].

6.1.5 Host-HCV interaction’s effects on treatment

Pegylated interferon and ribavirin were previously the two main treatments for HCV infection, and SVR was only reached in a very small percentage of those who received treatment. Alopecia, arthralgia, sleeplessness, pyrexia, headaches, myalgia, tinnitus, and depression were among the unfavourable side effects frequently reported by patients getting interferon-based therapy [59]. DAAs have a quicker duration of treatment and a greater SVR than interferons, in contrast to being lower toxic and more effective. IL-1 induces persistent stimulation of innate immune-mediated inflammation. Innate immune activation has been found to be inhibited by DAA medication by reducing IL-1 release and NF phosphorylation. Because of this, there is fewer inflammation, which also means that liver fibrosis and damage are reduced [60]. The chemokines CXCL10 and CXCL11, which attract innate immune cells, are expressed less when DAA is treated with medication. The effectiveness of NK cells has also been linked to DAA treatment. The balance of the innate immune system is restored by reversing the impaired innate immunity caused by reduced chemokine release and normalising NK cell function. ISGs (interferon stimulated genes) were shown to be higher at baseline in HCV patients who had undergone DAA treatment by Alao et al. [61]. This finding raises the possibility that innate immunity plays a role in the elimination of HCV. It is crucial to understand that RIG-I and TLR3 signalling are obstructed by the HCV NS3/4A protease, which pierces the human proteins MAVS and TRIF. It is unclear whether the direct antiviral activity of NS3/4A protease inhibitors or their capacity to activate the innate immune system’s defence against viruses by preventing TRIF and MAVS hydrolysis is what ultimately rids the body of the virus. A viral infection is likely to spontaneously clear if a powerful early humoral immune reaction is developed via neutralising antibodies as during initial stages of an HCV infection. In HCV-infected individuals, early and substantial neutralising antibody development is connected with acquired immunity against viral persistence. It has also been demonstrated that the spontaneous clearance of acute HCV induces memory T-cell-driven preventative immunity [62]. This protective response is partially effective, but it is unable to protect reinfection with HCV strains that did not stimulate pre-existing memory T cells. Even though research on HCV vaccines is in various phases, none have achieved FDA approval. According to studies by Law et al., a singular oral HCV vaccine produced extensive attempt to cross antibodies against certain HCV genotypes [63]. Additionally, T-cell-mediated reactions were evoked. A human prophylactic T-cell-based HCV immunisation promoted the formation of both CD4+ and CD8+ T cells, according to the study by Swadling et al. The four following factors increase the likelihood of HCV infection: HCV has four main characteristics: (1) a high error-prone mutation rate with the ability to evade selective pressures by neutralising antibodies and CD8+ T cells; (2) genomic variability with seven genetic variants and more than 65 subtypes that differ in nucleotide pattern; (3) a mutational rate happening in the variable region zone 1 of E2 with the potential for HVR 1 to prevent antibody conditional to E2; and (4) cell-to-cell transmission of HCV [64]. Because circulating HCV binds to plasma lipoprotein to form an infected hybrid lipoviral particle (LVP), that promotes viral continuation and infection by limiting protective antibody access to envelop glycoprotein, the development of an effective HCV vaccine is greatly hindered. More research and development are required to create effective and safe HCV vaccines that encourage the production of pass immunoglobulins that target epitopes that are retained all over HCV genotypes and are unconnected to HCV escape. This is due to a risk of re-infection following HCV therapy. Given the notable nucleotide sequence changes between genotypes, it should be efficient against a variety of HCV strains [65]. A cell-mediated immune response as well as cross-neutralising antibodies can be generated by an HCV vaccination, while humoral immunity produced by vaccine highly immunogenic is insufficient to protect against HCV infections.

6.1.6 HCV resistance in relation to directly acting antiviral drugs

The majority of current anti HCV therapies do not use IFN. IFN-free methods frequently combine various DAA subcategories that focus on NS3/4A, NS5A, and NS5B as shown in Table 1. Alternative classifications for NS5B polymerase inhibitors include nucleotide or non-nucleotide counterparts [66]. The patient’s cirrhotic condition is one of the most important host & viral factors that are known to affect the responsiveness to anti HCV treatment. Individuals with HCV infection who have never taken the DAA may have mutation spectra with substitutions linked to resistance (RASs). Globally, there is a wide variation in the prevalence of RASs based upon that viral genotype and origin [67]. It is not improbable that these mutations could be selected and impact a patient’s reaction to medication. Any patient with HCV infection should have their medication resistance profile reviewed before beginning treatment. The American Association for the Study of Liver Diseases (AASLD) is indeed the organisation that strongly advises testing for antiviral resistance in patients who have not responded to NS5A inhibitors (and also in some patients who are dependent on DAA treatment who are treatment-naive), especially for genotypes (GT) 1a and 3. A rising number of studies have discovered that DAAs are less effective in treating HCV patients with susceptible strains. After 8 weeks and 12 weeks following diagnosis with ledipasvir/sofosbuvir, correspondingly, in diagnosis and treatment and treatment-naive persons, well before NS5A RASs with a 100-fold greater degree of sensitivity to ledipasvir than wildtype virus led to poorer SVR rates [68]. Findings from 35 clinical studies done in 22 countries suggest that the baseline presence of NS5A mutations affects the efficacy of a ledipasvir/sofosbuvir combo in GT1-infected individuals. People with GT1a infection who have had treatment are particularly impacted by this outcome. Elbasvir, a recently approved NS5A inhibitor, and grazoprevir, a recently approved NS3/4A protease inhibitor, were used in clinical studies on HCV resistance to demonstrate that amino acid replacements in NS5A at positions Met28, Gln30, Leu31, and Tyr93 significantly reduced treatment effectiveness when the variable viruses were pervasive at baseline in GT1a-infected patient populations (rates of 70% vs. 98% SVR12 for patients with and without NS5A mutations). According to the current recommendations, the identification of drug sensitivity for therapeutic measures would almost certainly be followed by the customisation of antiviral therapy and a higher probability of success [69]. Because of their limited fitness effect, NS5A RASs are an indication of RASs that, once identified, continue to develop over time at a specific rate. Treatment, which now affects between 2 and 10% of those with HCV, is frequently (though not always) connected to RAS selection. The European Association for the Study of the Liver (EASL) has released numerous reference materials and clinical guidelines that include lists of RASs that offer reduced susceptibility to DAAs in people and in vitro [70]. RAS incidence is a situation that is developing and rapidly evolving when different DAAs and DAA combinations are licenced to treat HCV infections. This scenario is anticipated by quasispecies dynamics. RASs in NS3/4A and NS5A are frequently employed in individuals whose antifungal medication with NS3/4A and NS5A inhibitors failed [71]. After therapeutic failings with NS5B inhibitor-containing regimens, RASs in NS5B are detected much less frequently. Important residues of amino acids that resist nearly all NS3/4A inhibitors include Gln80, Arg155, Ala156, and Asp168. Locations Met28, Gln30, Leu31, Pro58, and Tyr93 in NS5A commonly choose replacements. In either situation, there is not much resistance. The long-term persistence of NS5A enzyme inhibitor mutations following the failure of NS5A inhibitor treatment also affects the therapeutic significance of these polymorphisms. Despite the relatively high barrier to resistance, several NS5B mutations, especially Leu159, Ser282, Cys316, Leu320, and Val321, have been linked to sofosbuvir resistance [72]. It has been demonstrated that the region of NS5B between amino acids 314 and 565 is sensitive to the non-nucleoside counterpart dasabuvir. Regarding RAS mutations connected to DAA drug failure, we need hard data. The HCV Italian popular resistance Network described the RAS profiles obtained by population sequencing study of 200 virological failures. They found a variety of tolerance tendencies that vary depending on the virus genetic, DAA regimen, and research study population, with a greater RAS prevalence at failure than previously reported [73]. Sarrazin et al. thoroughly analysed RAS patterns in a group of 626 non-responder/breakthrough and relapser patients to 2322 DAA-naive patients with HCV GT1 to GT4 [74]. They belonged to the group studying European HCV resistance. They discovered that the medication combinations’ target regions, subtypes, and genotypes had a wide range of difficult activities. R155K in GT1a and D168E/V in GT1b RASs were usually utilised in NS3 after the failure of simeprevir- and paritaprevir-based treatments. Between GT1a and GT1b, a unique resistance profile in NS5A was found. When daclatasvir, ledipasvir, and ombitasvir therapy failed, Q30H/R were seen in GT1a patients, but Y93H were commonly picked in GT1b practiced medicine with NS5A inhibitors [75]. After treatment with daclatasvir/sofosbuvir, Y93H was frequently observed in GT3a. Patients may be moved to a nucleotide analogue and a blocker against a protein that was not targeted in the original therapy if they do not respond to the existing formulations of sofosbuvir with an NS3 or NS5A inhibitor. In reducing HCV viral load in patients, the recently licenced DAA sofosbuvir/velpatasvir/voxilaprevir seems to be more efficient than prior DAA-based drugs like NS3 and NS5A inhibitors [76]. The barrier to resistance is too low without sofosbuvir for the glecaprevir/pibrentasvir combination to be taken into consideration for the retreatment of patients who failed NS5A inhibitors. The glecaprevir/pibrentasvir combination cannot be considered for the reoperation of patient groups who lost NS5A inhibition because the barrier to resistance is too low. Sofosbuvir should be added as well [77]. Triple (or quadruple, if ribavirin is included) or quadruple medication combinations, one for each target, are widely employed. This approach is based on the notion that genetic resistance is increased when several amino acid alterations are required to produce the resistant phenotype. However, the following needs to be emphasised: A small number of RASs have been labelled as drug-class RASs, meaning they develop resistance to numerous medications in the same class. Because of their limited fitness effect, some RASs, once identified, continue to grow over time at a specified rate, as shown by NS5A RASs. As a result, the effectiveness of re-treatment alternatives may be affected, even with enhanced antiviral combinations [78]. Scientist team has identified a unique HCV antibiotic resistance mechanism that differs from the conventional resistance based on real amino acid alterations, complicating the understanding of HCV resistance. This technique was discovered after an HCV community was subjected to IFN 100 times in a clone cell culture without any treatment and shown partial sensitivity despite having not received the medicine. This discovery was made in respect to several anti-HCV medications, each of which has a unique mechanism of action for effectively combating the virus (telaprevir, daclatasvir, ribavirin, cyclosporine A) [79]. Strength exercise was also resisted when sofosbuvir levels were high. This data refutes the hypothesis that the observed decline in viral susceptibility to inhibitors is caused by changes in the receptor that mediates receptor susceptibility: No substitutions that could be categorised as providing resistance to one of the inhibitors used were found in the I UDS analysis; similar levels of opposition were seen when infections with high-fitness virus infections were carried out over a 1000-fold range of multiplicity of infection (MOI), indicating that progeny manufacturing did not rely on the presence of averse strains, which would be clearly lowered with a 1000-fold decrease in MOI; and studied into a probable in vivo parallel and the clinical implications of strength and conditioning medication resistance has shown studies that demonstrate a proportion of patients cannot endure therapy even if there is no proof that resistance substitutions have arisen in the intended target [80].

DAA classDAA (directed genotypes)
NS5B polymerase inhibitors of nucleotide and nucleosideDaclatasvir (3)
Sofosbuvir (1–4)
Pibrentasvir (1–6)
Ombitasvir (1,4)
NS3/4A protease inhibitors (Pls)Sunvepra (1,4)
Glecaprevir (1–6)
Grazoprevir (1,3,4)
Voxilaprevir (1–6)
Galexox (1)
NS5A inhibitorsVelpatasvir (1–6)
Ombitasvir (1)
Elbasvir (1,6)
Inhibitors of the non-nucleoside NS5B polymeraseDasabuvir (1)

Table 1.

The four kinds of DAAs, which are the cornerstone of anti-HCV therapy and are used in various combinations.

6.1.7 Tools for assessing drug resistance include ultra-deep sequencing (UDS) technology

A fundamental problem in the investigation of HCV resistance is the standardisation of techniques for analysing amino acid alterations. The ability to identify RASs using various antiviral resistant testing techniques varies. Population sequencing might be helpful if such virus quasispecies were forced to keep the mutation that confers antiviral resistance [81]. On the other hand, hand, the architecture of a viral group is determined by the collection of mutations that go beyond the consensus sequence. Currently, it is possible to examine the mutant spectrum makeup using molecular clones, Sanger sequencing, and highly delicate methods such UDS platforms. Now, in order to develop bioinformatic methods for the research of viral quasispecies, the following significant obstacles must be overcome. A biased representation of the mutant spectrum may result from subpopulations being amplified by insufficient oligonucleotide primers. To determine a reasonable cut-off level for mutant frequency, erroneous mutations discovered during the sequenced and amplification phases should be checked in testing phases with regular clones. Artefactual crossover can be decreased by altering the PCR settings. Previously, 10,000 reads of sequencing depth were used to determine a cut-off number of 1% mutation frequency [82]. On the question of whether a given RAS can be considered clinically significant above a specific threshold for mutant regularity, the literature is mixed. It has been suggested that if the cutoff is 15%, the existence of a RAS may be disregarded. We do not believe that, or any other snipped level has any theoretical or practical support, for a variety of reasons. The detection rate for a particular variation depends on the quantity of readings (covers) obtained, the detection technique, and the composition of the HCV mutant spectrum. The upgraded UDS can find a certain variant more frequently depending on the number of readings. Due to how quickly it diverged into many genotypes and subtypes, HCV has truly come to represent true unpredictability [83]. For this reason, it would be preferable to have deep sequenced data both during the time of therapeutic failure and prior to the introduction of a new medicine, particularly if a lot of time passed since the failure of the last treatment. This is due to the possibility of the mutant spectra changing in a matter of seconds.

6.1.8 PRRs that detect viruses innately

The network of PRRs and related signalling pathways, which cause the production of IFN and inflammatory genes, are a crucial part of the innate immune system. The germline-encoded receptors (PRRs) found in plants, worms, drosophila, and mammals have not changed during the history of evolution. PRRs are encoded inside the host organism’s germline, in contrast to the immunoglobulin receptors of the adaptive immune system, which are created through somatic gene rearrangements. These PRRs detect conserved microbial characteristics to prevent infection [84]. The classes of viral-sensing PRRs that will be discussed in this study include inflammasomes, Toll-like receptors (TLRs), RIG-I-like receptors (RLRs), C-type lectin receptors (CLRs), and DNA sensors as shown in Table 2. Although these receptors are capable of detecting a wide variety of microorganisms, microbial metabolites, and host-derived harm molecular patterns, we will focus on the role of these receptors in the detection of viruses (DAMPs). Given the significance of these defences in defending the host from viral infections, it is not surprising that viruses have developed a range of ways to circumvent these antiviral defences. This makes it possible for viruses to proliferate and spread illness. Therefore, the topic of our discussion will be the most recent advancements in viral immune escape techniques.

TLR(Ligand)Virus(genome)OutcomesReference
TLR7 (ssRNA)Rota virus (dsRNA)Favours host
West Nile virus (ss(+)RNAHarmful/protective[85]
Avian InfluenzaFavours host[85]
Chikungunya virusHarmful to host[86]
HIVDamaging to host[84, 87]
HSV1Favours host[86]
VSVFavours host[88]
Enterovirus 71Favours host[87]
TLR3 (dsRNA)PoliovirusFavours host[89]
Huntaan virusFavours host[90]
CoxasackievirusFavours host[91]
Punta Toro virusDamaging to host[84]
EMCVFavours host[89]
InfluenzaDamaging to host[89]
HCVFavours host[90]
TLR4 (virus coat protein)InfluenzaDamaging to host[92]
RSVFavours host[92]
VSVFavours host[92]
TLR6 (virus coat protein)Dengue virusHarmful to host[93]
RSVFavours host
TLR9 (CpG DNA)Murine gammaherpesvirus 68Favours host[94]
MCMVFavours host[95]
ECTVFavours host[94]
HSVFavours host[96]
Dengue virusFavours host[95]
TLR2 (virus coat protein)HSVFavours host[92]
RSVFavours host[93]
HCVHarmful to host[93]
TLR8 (ssRNA)West Nile virusHarmful to host[97]

Table 2.

TLR-virus interactions and their effects on the host and the virus.

6.2 Adaptive immunity and DAAs treatment

The emergence of HCV-specific T cells that produce the memory T cell marker CD127 and the antiapoptotic molecule Bcl-2 is proof that cytotoxic T cell reaction is restored after a person’s natural recovery from HCV infection [98]. Weiland et al. found that people with chronic HCV infection had persistent TCF1 + CD127 + PD-1+ HCV-specific T cells that displayed signs of exhaustion and memory, and that were present both before and after DAA treatment. This subgroup was elevated in one patient who survived and had HCV-specific CD8 T cells that were terminally exhausted [99]. These CD8 T cells were different from memory CD8 T cells, which surface during spontaneous resolution (CD127+ PD-1−). Aregay et al. investigated the possibility that the diminished CD8T cellular response to HCV could be restored by DAA treatment. After HCV cure, every one of the 40 patients who underwent DAA treatment still had diminished functioning in their worn-out HCV-specific CD8 T cells [100]. Following HCV treatment, they observed a consistent shift in the subset makeup of CD8 T cells across all cirrhotic patients. The researchers deduced that the seriousness of the liver fibrosis in these cases was correlated with the prevalence of CD8 T cells with hyperfunction. These findings imply that the decreased phenotype of CD8 T cell reactions specific to HCV is not restored by DAA-induced HCV therapy. CD4 helper T cells are essential for an effective HCV-specific CD8 T cell reaction and for the spontaneous clearance of infection. Intrahepatic regulatory CD4 T cells and CD4(+) CD25(+) FoxP3(+) T-reg cells are more prevalent in patients with chronic HCV infection and do not decrease after viral eradication with IFN- or DAA-based therapy. Langhans et al. looked at T-reg cells at 12 and 24 weeks before and after taking DAAs. They discovered that T-reg cells survive after viral eradication for a long time and that DAAs do not alter their activation status [101]. Activation of the B cell response is associated with B cell-related illnesses such as combined cryoglobulins in 40–60% of patients and CV in 10–15% of patients. Comarmond et al. looked at the effect of DAAs on the CV of 27 patients with HCV infection. They discovered that DAAs for CV led to the recovery of peripheral B cellular responses in around 88.9% of patients [102]. 90.2% of patients who underwent DAA treatment for vasculitis had a full clinical reaction, which was associated with a virological response, according to the Saadoun et al.’s examination of the vasculitis remission in 41 patients [103].

6.2.1 Liver disease-related proliferative signalling caused by HCV

The genesis and development of liver problems are reportedly aided by a variety of cell signalling alterations brought on by HCV infection, either indirectly or directly as shown in Figure 3. As HCV evolved, it altered these pathways to favour replication and persistence, which had a significant effect on viral pathogenicity and liver disease [104]. EGFR signalling is encouraged by the activation of the TGF and EGF pathways. STAT3 route: Both directly and indirectly, via the stimulation of NS5A and EGFR, which promotes the production of ROS, the activity of the core protein activates STAT3. Additionally, HCV silences PTPRD and SOCS3, two rival STAT3 regulators, by utilising miR-135a-5p and miR-19a. The TGF-pathway the UPR, which encourages NF-B activity, as well as the centre proteins, which directly interact with SMAD3, are two mechanisms by which HCV activates the TGF pathway. Endoglin (CD105) expression brought on by HCV encourages angiogenesis transmission and TGF-pathway stimulation. The HCV core triggers HIF-1, which boosts VEGF synthesis. Similar to this, STAT3-dependent androgen receptor stimulation allows HCV to enhance VEGF expression.

Figure 3.

Chronic HCV infection-related hepatocellular carcinoma processes include a mix of virus-mediated (direct), host-mediated (indirect), and host-related bystander effects.

6.2.2 Immune system’s role in the progression of liver damage after HCV infection that has lasted a long time

HCV lacks a dormant stage through its life span and is widely characterised as non-cytopathic, despite reports of apoptosis induction. It consequently persistently obstructs the liver’s capacity to preserve homeostasis, resulting in stress and inflammation. A proinflammatory environment is produced by non-parenchymal cells that have been triggered by innate immune responses, and this milieu plays a significant role in the progression of liver illness from fibrosis to cirrhosis and HCC [105]. HCV is mostly recognised by TLR3 and RIG-I since it is a prototypical positive-stranded RNA virus. Nevertheless, it has a number of defences to prevent innate immune identification and control the subsequent IFN response. HCV enhances TLR3 signalling in monocytes and macrophages during a persistent infection. This causes the inflammasome to activate without IFN induction and the release of proinflammatory cytokines such interleukins (IL) [106]. It appears that natural killing (NK) cells are activated, IL-18 is produced, and NLR3P inflammasomes are activated when macrophages recognise HCV-infected hepatocytes. The STAT3 signalling pathway is important during inflammation in complement to its pro-viral and proliferation effects. The signalling pathways STAT3 and NF-B are stimulated by the proinflammatory IL-6 and TNF, and when they are frequently activated, they can hasten the onset of liver problems and the emergence of HCC [107]. When JAK/STAT signalling is activated in neighbouring cells as a result of NF-B signalling, that also boosts IFN production, antiviral genes are produced. HCV proteins specifically target these actions while attenuating the innate antiviral response. Infected cells are also prevented from apoptosing by the HCV proteins core and NS5A, which do this by triggering NF-kB and AKT serine/threonine kinase (AKT).

Chronic HCV infection disturbs the equilibrium between both the ligand and the Fas receptor (FasR, CD95) (FasL, CD95L). It has been found that FasL-positive T lymphocytes interact with FasR-exposed hepatocytes to kill liver cells. FasR and FasL expression on hepatocytes and T cells, respectively, are strongly correlated with the severity of liver damage during chronic HCV infection [108]. Additionally, HCC exhibits nearly no FasR expression, which suggests a decreased susceptibility to T cell-mediated cytotoxicity and may increase the survival of tumour-genic cells. This result is also influenced by the absence of a functional T cell reaction during chronic infection. T cells, in particular CD8+ T cells, get exhausted after prolonged and continuous exposure to HCV antigens. Additionally, a lot of evidence points to the possibility of productive HCV infection of immune cells, particularly T cells, but it is unclear how much this can influence the immune response specifically directed against HCV [99]. According to the available data, myeloid-derived suppressor cells (MDSCs) in HCC patients and liver cancer animal models increase tumour growth and the incidence of metastases. HSCs are essential for both the control of the extracellular matrix of the liver and the healing of wounds.

6.2.3 Clinical implications of HCV-induced HCC risk biomarker identification

Just 30 years after discovery, HCV is now treatable thanks to the outstanding work of researchers, doctors, and the pharmaceutical industry. However, even in patients with severe liver disease, the viral cure brought on by treatment cannot entirely eliminate HCV-associated comorbid and HCC risk [109]. It is thought that epidemiologic peak of HCV-associated liver problems and HCC has not yet been reached because of the comparatively large time lag among virus infection liver damage and the emergence of HCC. This brings to light two important unmet medical needs for the clinical treatment of individuals with SVR: effective and secure chemopreventive programmes that support these individuals by tackling virus-specific pro-oncogenic mechanisms, epigenetic signatures, and liver fibrosis [110]. Viable biomarkers to estimate the proportion of SVR patients who are more likely to develop HCC. Finding a precise and comprehensive biomarker to assess the risk of HCC is difficult due to the relative range of individual aberrations. A 186-gene transcriptional profile that predicts HCC risk has been found in early-stage cirrhosis caused by HCV and non-tumour tissue close to HCC lesions [111]. Recent biomarkers, like the PES, that are based on virus-induced epigenetic modifications offer a novel perspective for determining the likelihood of HCC reappearance in HCV patients after SVR and make it possible to choose these people for clinical studies investigating HCC chemoprevention. Numerous extrahepatic side effects, including mixed cryoglobulinemia and B cell lymphomas, have been associated with chronic HCV infection. Though the mechanisms underlying these problems have not yet been fully elucidated, it has been hypothesised that TGF and IL-6 play a part in their development [112]. For the HCC treatment that has already developed, a number of medicines that target HCV-relevant signalling pathways have also been suggested. In order to prevent and maybe treat HCC caused by other related liver cirrhosis aetiologies, such as NAFLD, new drug and individualised therapy strategies may be developed with the knowledge acquired from HCV-induced liver disease.

Advertisement

7. Conclusion

Since liver cirrhosis develops in one-third of people with persistent HCV infections, HCV is a prominent cause of liver disease. The acute hepatitis C may go away on its own or develop into chronic HCV infection, depending on the virus and host factors. To provide innate cellular immunity, NK cells release type II IFN and TNF, which stop viral multiplication by noncytolytic mechanisms, in addition to perforin and granzyme, which kill invading pathogens through cytolytic-dependent pathways. The adapted cellular approach to HCV infection, which eradicates the virus equally cytolytically and noncytolytically, depends heavily on CD8+ T cells. CD4+ T cells assist APC, B cells, and CD8+ T cells. Reduced capacity to control HCV infection is associated with cellular immunity failure. HCV escape mutation, reduced antigen presentation by HCV-infected DC, T cell exhaustion brought on by persistent HCV antigens, inadequate T cell primed by DC and intrahepatic antigen-presenting cells, and the formation of a tolerogenic intrahepatic milieu are all contributing factors to the persistence of HCV infection. Understanding how the host as well as the virus interact in respect of the variables that promote the settlement of the acute stage of a hepatitis C and the immunological evasive strategies the virus uses to sustain its survival in the host is critical. However, viral characteristics include HCV genetic variation, HCV cell-to-cell transmission, a rapid mutation rate, and the production of infectious lipoviral components hinder efforts to develop an HCV vaccine.

Advertisement

Conflict of interest

Authors declare no conflict of interest.

References

  1. 1. Lapa D et al. Hepatitis C virus genetic variability, human immune response, and genome polymorphisms: Which is the interplay? Cell. 2019;8(4):305
  2. 2. Kwok AJ, Mentzer A, Knight JC. Host genetics and infectious disease: New tools, insights and translational opportunities. Nature Reviews Genetics. 2021;22(3):137-153
  3. 3. Ansari MA et al. Interferon lambda 4 impacts the genetic diversity of hepatitis C virus. eLife. 2019;8:e42463
  4. 4. An P et al. Host and viral genetic variation in HBV-related hepatocellular carcinoma. Frontiers in Genetics. 2018;9:261
  5. 5. Sanjuán R, Domingo-Calap P. Genetic diversity and evolution of viral populations. Encyclopedia of Virology. 2021;2021:53-61
  6. 6. Martin NK et al. Prioritization of HCV treatment in the direct-acting antiviral era: An economic evaluation. Journal of Hepatology. 2016;65(1):17-25
  7. 7. Nishiya AS et al. Phylogenetic analysis of the emergence of main hepatitis C virus subtypes in São Paulo, Brazil. Brazilian Journal of Infectious Diseases. 2015;19:473-478
  8. 8. Choumet V, Desprès P. Dengue and other flavivirus infections. Revue Scientifique et Technique. 2015;34(2):473-472
  9. 9. Maggi F, Focosi D, Pistello M. How current direct-acting antiviral and novel cell culture systems for HCV are shaping therapy and molecular diagnosis of chronic HCV infection? Current Drug Targets. 2017;18(7):811-825
  10. 10. Paul D, Madan V, Bartenschlager R. Hepatitis C virus RNA replication and assembly: Living on the fat of the land. Cell Host & Microbe. 2014;16(5):569-579
  11. 11. Bartenschlager R, Lohmann V, Penin F. The molecular and structural basis of advanced antiviral therapy for hepatitis C virus infection. Nature Reviews Microbiology. 2013;11(7):482-496
  12. 12. Xiao F et al. Hepatitis C virus cell-cell transmission and resistance to direct-acting antiviral agents. PLoS Pathogens. 2014;10(5):e1004128
  13. 13. Yoon JC et al. Cell-to-cell contact with hepatitis C virus-infected cells reduces functional capacity of natural killer cells. Journal of Virology. 2011;85(23):12557-12569
  14. 14. Barretto N et al. Determining the involvement and therapeutic implications of host cellular factors in hepatitis C virus cell-to-cell spread. Journal of Virology. 2014;88(9):5050-5061
  15. 15. Wakita T et al. Production of infectious hepatitis C virus in tissue culture from a cloned viral genome. Nature Medicine. 2005;11(7):791-796
  16. 16. Suzuki R et al. Trans-complemented hepatitis C virus particles as a versatile tool for study of virus assembly and infection. Virology. 2012;432(1):29-38
  17. 17. Yang D-R, Zhu H-Z. Hepatitis C virus and antiviral innate immunity: Who wins at tug-of-war? World Journal of Gastroenterology. 2015;21(13):3786
  18. 18. Jayasekera CR et al. Treating hepatitis C in lower-income countries. The New England Journal of Medicine. 2014;370(20):1869-1871
  19. 19. Westbrook R, Dusheiko G. Natural history of hepatitis C. Journal of Hepatology. 2014;61(Suppl. 1):S58-S68. DOI: 10.1016/j. jhep.2014.07. 012
  20. 20. Webster DP, Klenerman P, Dusheiko GM. Hepatitis C. Lancet. 2015;385(9973):1124-1135
  21. 21. El-Serag HB. Epidemiology of viral hepatitis and hepatocellular carcinoma. Gastroenterology. 2012;142(6):1264-1273.e1
  22. 22. Thein HH et al. Estimation of stage-specific fibrosis progression rates in chronic hepatitis C virus infection: A meta-analysis and meta-regression. Hepatology. 2008;48(2):418-431
  23. 23. Xu F et al. All-cause mortality and progression risks to hepatic decompensation and hepatocellular carcinoma in patients infected with hepatitis C virus. Clinical Infectious Diseases. 2016;62(3):289-297
  24. 24. Lamontagne RJ, Bagga S, Bouchard MJ. Hepatitis B virus molecular biology and pathogenesis. Hepatoma Research. 2016;2:163
  25. 25. Zeisel MB, Felmlee DJ, Baumert TF. Hepatitis C virus entry. In: Hepatitis C Virus: From Molecular Virology to Antiviral Therapy. HAL Open Science. 2013. pp. 87-112
  26. 26. Lohmann V. Hepatitis C virus RNA replication. In: Bartenschlager R, editor. Hepatitis C Virus: From Molecular Virology to Antiviral Therapy. Berlin, Heidelberg: Springer-Verlag; 2013. pp. 167-198
  27. 27. Moradpour D, Penin F. Hepatitis C virus proteins: From structure to function. In: Bartenschlager R, editor. Hepatitis C Virus: From Molecular Virology to Antiviral Therapy. Berlin, Heidelberg: Springer; 2013. pp. 113-142
  28. 28. Scheel TK, Rice CM. Understanding the hepatitis C virus life cycle paves the way for highly effective therapies. Nature Medicine. 2013;19(7):837-849
  29. 29. Dustin LB et al. Hepatitis C virus: Life cycle in cells, infection and host response, and analysis of molecular markers influencing the outcome of infection and response to therapy. Clinical Microbiology and Infection. 2016;22(10):826-832
  30. 30. Dustin LB. Too low to measure, infectious nonetheless. Blood, The Journal of the American Society of Hematology. 2012;119(26):6181-6182
  31. 31. Schoggins JW, Rice CM. Innate immune responses to hepatitis C virus. In: Bartenschlager R, editor. Hepatitis C Virus: From Molecular Virology to Antiviral Therapy. Berlin, Heidelberg: Springer; 2013. pp. 219-242
  32. 32. Horner SM, Gale M. Regulation of hepatic innate immunity by hepatitis C virus. Nature Medicine. 2013;19(7):879-888
  33. 33. Wei D et al. The molecular chaperone GRP78 contributes to toll-like receptor 3-mediated innate immune response to hepatitis C virus in hepatocytes. Journal of Biological Chemistry. 2016;291(23):12294-12309
  34. 34. Sheahan T et al. Interferon lambda alleles predict innate antiviral immune responses and hepatitis C virus permissiveness. Cell Host & Microbe. 2014;15(2):190-202
  35. 35. Holz L, Rehermann B. T cell responses in hepatitis C virus infection: Historical overview and goals for future research. Antiviral Research. 2015;114:96-105
  36. 36. Honegger JR, Zhou Y, Walker CM, Semin LD. Will there be a vaccine to prevent HCV infection? In: Jean-Michel P, editor. Seminars in Liver Disease. Thieme Medical Publishers; 2014;34:79-88
  37. 37. Dazert E et al. Loss of viral fitness and cross-recognition by CD8+ T cells limit HCV escape from a protective HLA-B27–restricted human immune response. The Journal of Clinical Investigation. 2009;119(2):376-386
  38. 38. Cashman SB, Marsden BD, Dustin LB. The humoral immune response to HCV: Understanding is key to vaccine development. Frontiers in Immunology. 2014;5:550
  39. 39. De Jong YP et al. Broadly neutralizing antibodies abrogate established hepatitis C virus infection. Science Translational Medicine. 2014;6(254):254ra129
  40. 40. Klenerman P, Thimme R. T cell responses in hepatitis C: The good, the bad and the unconventional. Gut. 2012;61(8):1226-1234
  41. 41. Laidlaw SM, Dustin LB. Interferon lambda: Opportunities, risks, and uncertainties in the fight against HCV. Frontiers in Immunology. 2014;5:545
  42. 42. Park S-H, Rehermann B. Immune responses to HCV and other hepatitis viruses. Immunity. 2014;40(1):13-24
  43. 43. Cornberg M, Wedemeyer H. Hepatitis C virus infection from the perspective of heterologous immunity. Current Opinion in Virology. 2016;16:41-48
  44. 44. Martinez M, Franco S. Therapy implications of hepatitis C virus genetic diversity. Viruses. 2021;13:41
  45. 45. Lemey P et al. Synonymous substitution rates predict HIV disease progression as a result of underlying replication dynamics. PLoS Computational Biology. 2007;3(2):e29
  46. 46. Alter HJ et al. Reflections on the history of HCV: A posthumous examination. Clinical Liver Disease. 2020;15(Suppl. 1):S64
  47. 47. Fafi-Kremer S et al. Viral entry and escape from antibody-mediated neutralization influence hepatitis C virus reinfection in liver transplantation. Journal of Experimental Medicine. 2010;207(9):2019-2031
  48. 48. Harouaka D et al. Diminished viral replication and compartmentalization of hepatitis C virus in hepatocellular carcinoma tissue. Proceedings of the National Academy of Sciences. 2016;113(5):1375-1380
  49. 49. Hedegaard DL et al. High resolution sequencing of hepatitis C virus reveals limited intra-hepatic compartmentalization in end-stage liver disease. Journal of Hepatology. 2017;66(1):28-38
  50. 50. Bukh J. The history of hepatitis C virus (HCV): Basic research reveals unique features in phylogeny, evolution and the viral life cycle with new perspectives for epidemic control. Journal of Hepatology. 2016;65(1):S2-S21
  51. 51. Stuart JD, Salinas E, Grakoui A. Immune system control of hepatitis C virus infection. Current Opinion in Virology. 2021;46:36-44
  52. 52. Franco S et al. Detection of a sexually transmitted hepatitis C virus protease inhibitor-resistance variant in a human immunodeficiency virus–infected homosexual man. Gastroenterology. 2014;147(3):599-601.e1
  53. 53. Wrensch F et al. Interferon-induced transmembrane proteins mediate viral evasion in acute and chronic hepatitis C virus infection. Hepatology. 2019;70(5):1506-1520
  54. 54. Ansaldi F et al. Hepatitis C virus in the new era: Perspectives in epidemiology, prevention, diagnostics and predictors of response to therapy. World Journal of Gastroenterology: WJG. 2014;20(29):9633
  55. 55. Cornberg M, Tacke F, Karlsen TH. Clinical practice guidelines of the European Association for the study of the liver–advancing methodology but preserving practicability. Journal of Hepatology. 2019;70(1):5-7
  56. 56. Lawitz E et al. Sofosbuvir for previously untreated chronic hepatitis C infection. New England Journal of Medicine. 2013;368(20):1878-1887
  57. 57. Afdhal N et al. Ledipasvir and sofosbuvir for previously treated HCV genotype 1 infection. New England Journal of Medicine. 2014;370(16):1483-1493
  58. 58. Lazarus JV, Roel E, Elsharkawy AM. Hepatitis C virus epidemiology and the impact of interferon-free hepatitis C virus therapy. Cold Spring Harbor Perspectives in Medicine. 2020;10(3):a036913
  59. 59. Wang G et al. Synthesis and anti-HCV activity of sugar-modified guanosine analogues: Discovery of AL-611 as an HCV NS5B polymerase inhibitor for the treatment of chronic hepatitis C. Journal of Medicinal Chemistry. 2020;63(18):10380-10395
  60. 60. Goutagny N et al. Evidence of viral replication in circulating dendritic cells during hepatitis C virus infection. The Journal of Infectious Diseases. 2003;187(12):1951-1958
  61. 61. Alao H et al. Baseline intrahepatic and peripheral innate immunity are associated with hepatitis C virus clearance during direct-acting antiviral therapy. Hepatology. 2018;68(6):2078-2088
  62. 62. Cormier EG et al. L-SIGN (CD209L) and DC-SIGN (CD209) mediate transinfection of liver cells by hepatitis C virus. Proceedings of the National Academy of Sciences. 2004;101(39):14067-14072
  63. 63. Law JLM et al. A hepatitis C virus (HCV) vaccine comprising envelope glycoproteins gpE1/gpE2 derived from a single isolate elicits broad cross-genotype neutralizing antibodies in humans. PLoS One. 2013;8(3):e59776
  64. 64. Kondo Y et al. Lymphotropic HCV strain can infect human primary naïve CD4+ cells and affect their proliferation and IFN-γ secretion activity. Journal of Gastroenterology. 2011;46(2):232-241
  65. 65. Wang RY et al. Preferential association of hepatitis C virus with CD19+ B cells is mediated by complement system. Hepatology. 2016;64(6):1900-1910
  66. 66. European Association for the Study of the Liver. EASL recommendations on treatment of hepatitis C 2018. Journal of Hepatology. 2018;69(2):461-511
  67. 67. Chen Z-W et al. Global prevalence of pre-existing HCV variants resistant to direct-acting antiviral agents (DAAs): Mining the GenBank HCV genome data. Scientific Reports. 2016;6(1):1-9
  68. 68. Yang S et al. Prevalence of NS5B resistance-associated variants in treatment-naive Asian patients with chronic hepatitis C. Archives of Virology. 2018;163(2):467-473
  69. 69. Bertoli A et al. Prevalence of single and multiple natural NS3, NS5A and NS5B resistance-associated substitutions in hepatitis C virus genotypes 1-4 in Italy. Scientific Reports. 2018;8(1):1-11
  70. 70. Sorbo MC et al. Hepatitis C virus drug resistance associated substitutions and their clinical relevance: Update 2018. Drug Resistance Updates. 2018;37:17-39
  71. 71. Calleja JL et al. NS5A resistance: Clinical implications and treatment possibilities. AIDS Reviews. 2016;18(1):15-22
  72. 72. Tong X et al. In vivo emergence of a novel mutant L159F/L320F in the NS5B polymerase confers low-level resistance to the HCV polymerase inhibitors mericitabine and sofosbuvir. The Journal of Infectious Diseases. 2014;209(5):668-675
  73. 73. Hedskog C et al. Evolution of the HCV viral population from a patient with S282T detected at relapse after sofosbuvir monotherapy. Journal of Viral Hepatitis. 2015;22(11):871-881
  74. 74. Sarrazin C. The importance of resistance to direct antiviral drugs in HCV infection in clinical practice. Journal of Hepatology. 2016;64(2):486-504
  75. 75. Di Maio VC et al. Multiclass HCV resistance to direct-acting antiviral failure in real-life patients advocates for tailored second-line therapies. Liver International. 2017;37(4):514-528
  76. 76. Dietz J et al. Patterns of resistance-associated substitutions in patients with chronic HCV infection following treatment with direct-acting antivirals. Gastroenterology. 2018;154(4):976-988.e4
  77. 77. Bourlière M et al. Sofosbuvir, velpatasvir, and voxilaprevir for previously treated HCV infection. New England Journal of Medicine. 2017;376(22):2134-2146
  78. 78. Perales C et al. Response of hepatitis C virus to long-term passage in the presence of alpha interferon: Multiple mutations and a common phenotype. Journal of Virology. 2013;87(13):7593-7607
  79. 79. Gallego I et al. Barrier-independent, fitness-associated differences in sofosbuvir efficacy against hepatitis C virus. Antimicrobial Agents and Chemotherapy. 2016;60(6):3786-3793
  80. 80. Gallego I et al. Resistance of high fitness hepatitis C virus to lethal mutagenesis. Virology. 2018;523:100-109
  81. 81. Fourati S, Pawlotsky J-M. Virologic tools for HCV drug resistance testing. Viruses. 2015;7(12):6346-6359
  82. 82. Quer J et al. Deep sequencing in the management of hepatitis virus infections. Virus Research. 2017;239:115-125
  83. 83. Bull RA et al. A method for near full-length amplification and sequencing for six hepatitis C virus genotypes. BMC Genomics. 2016;17(1):1-10
  84. 84. Carty M, Guy C, Bowie AG. Detection of viral infections by innate immunity. Biochemical Pharmacology. 2021;183:114316
  85. 85. Town T et al. Toll-like receptor 7 mitigates lethal West Nile encephalitis via interleukin 23-dependent immune cell infiltration and homing. Immunity. 2009;30(2):242-253
  86. 86. Luo Z et al. EV71 infection induces neurodegeneration via activating TLR7 signaling and IL-6 production. PLoS Pathogens. 2019;15(11):e1008142
  87. 87. Solmaz G et al. TLR7 controls VSV replication in CD169+ SCS macrophages and associated viral neuroinvasion. Frontiers in Immunology. 2019;10:466
  88. 88. Meier A et al. Sex differences in the Toll-like receptor–mediated response of plasmacytoid dendritic cells to HIV-1. Nature Medicine. 2009;15(8):955-959
  89. 89. Tsai YT et al. Human TLR3 recognizes dengue virus and modulates viral replication in vitro. Cellular Microbiology. 2009;11(4):604-615
  90. 90. Handke W et al. Hantaan virus triggers TLR3-dependent innate immune responses. The Journal of Immunology. 2009;182(5):2849-2858
  91. 91. Gowen BB et al. TLR3 deletion limits mortality and disease severity due to Phlebovirus infection. The Journal of Immunology. 2006;177(9):6301-6307
  92. 92. Murawski MR et al. Respiratory syncytial virus activates innate immunity through Toll-like receptor 2. Journal of Virology. 2009;83(3):1492-1500
  93. 93. Chen J, Ng MM-L, Chu JJH. Activation of TLR2 and TLR6 by dengue NS1 protein and its implications in the immunopathogenesis of dengue virus infection. PLoS Pathogens. 2015;11(7):e1005053
  94. 94. Puttur F et al. Conventional dendritic cells confer protection against mouse cytomegalovirus infection via TLR9 and MyD88 signaling. Cell Reports. 2016;17(4):1113-1127
  95. 95. Lai JH et al. Infection with the dengue RNA virus activates TLR9 signaling in human dendritic cells. EMBO Reports. 2018;19(8):e46182
  96. 96. Rogers GL et al. Plasmacytoid and conventional dendritic cells cooperate in crosspriming AAV capsid-specific CD8+ T cells. Blood, The Journal of the American Society of Hematology. 2017;129(24):3184-3195
  97. 97. Paul AM et al. TLR8 couples SOCS-1 and restrains TLR7-mediated antiviral immunity, exacerbating West Nile virus infection in mice. The Journal of Immunology. 2016;197(11):4425-4435
  98. 98. Walker MR et al. Clearance of hepatitis C virus is associated with early and potent but narrowly-directed, Envelope-specific antibodies. Scientific Reports. 2019;9(1):1-14
  99. 99. Wieland D et al. TCF1+ hepatitis C virus-specific CD8+ T cells are maintained after cessation of chronic antigen stimulation. Nature Communications. 2017;8(1):1-13
  100. 100. Aregay A et al. Elimination of hepatitis C virus has limited impact on the functional and mitochondrial impairment of HCV-specific CD8+ T cell responses. Journal of Hepatology. 2019;71(5):889-899
  101. 101. Langhans B et al. Increased peripheral CD4+ regulatory T cells persist after successful direct-acting antiviral treatment of chronic hepatitis C. Journal of Hepatology. 2017;66(5):888-896
  102. 102. Comarmond C et al. Direct-acting antiviral therapy restores immune tolerance to patients with hepatitis C virus–induced cryoglobulinemia vasculitis. Gastroenterology. 2017;152(8):2052-2062.e2
  103. 103. Saadoun D et al. Efficacy and safety of sofosbuvir plus daclatasvir for treatment of HCV-associated cryoglobulinemia vasculitis. Gastroenterology. 2017;153(1):49-52.e5
  104. 104. Hoshida Y et al. Pathogenesis and prevention of hepatitis C virus-induced hepatocellular carcinoma. Journal of Hepatology. 2014;61(1):S79-S90
  105. 105. Ferreira AR et al. Hepatitis C virus: Evading the intracellular innate immunity. Journal of Clinical Medicine. 2020;9(3):790
  106. 106. Chattergoon MA et al. HIV and HCV activate the inflammasome in monocytes and macrophages via endosomal Toll-like receptors without induction of type 1 interferon. PLoS Pathogens. 2014;10(5):e1004082
  107. 107. Shoukry NH. Hepatitis C vaccines, antibodies, and T cells. Frontiers in Immunology. 2018;9:1480
  108. 108. Casey JL, Feld JJ, MacParland SA. Restoration of HCV-specific immune responses with antiviral therapy: A case for DAA treatment in acute HCV infection. Cell. 2019;8(4):317
  109. 109. Kulik L, El-Serag HB. Epidemiology and management of hepatocellular carcinoma. Gastroenterology. 2019;156(2):477-491.e1
  110. 110. Guichard C et al. Integrated analysis of somatic mutations and focal copy-number changes identifies key genes and pathways in hepatocellular carcinoma. Nature Genetics. 2012;44(6):694-698
  111. 111. Schulze K et al. Exome sequencing of hepatocellular carcinomas identifies new mutational signatures and potential therapeutic targets. Nature Genetics. 2015;47(5):505-511
  112. 112. Pol S, Vallet-Pichard A, Hermine O. Extrahepatic cancers and chronic HCV infection. Nature Reviews. Gastroenterology & Hepatology. 2018;15(5):283-290

Written By

Quratulain Maqsood, Maria Hussain and Aleena Sumrin

Submitted: 22 December 2022 Reviewed: 23 December 2022 Published: 17 February 2023