Topic Highlight Open Access
Copyright ©2013 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Gastroenterol. Nov 28, 2013; 19(44): 7910-7921
Published online Nov 28, 2013. doi: 10.3748/wjg.v19.i44.7910
Scotomas in molecular virology and epidemiology of hepatitis C virus
Yue Wang, National Institute for Viral Disease Control and Prevention, China CDC, Beijing 100052, China
Yue Wang, Collaborative Innovation Center for Diagnosis and Treatment of Infectious Diseases, Zhejiang University, Hangzhou 310058, Zhejiang Province, China
Author contributions: Wang Y solely contributed to this manuscript.
Correspondence to: Yue Wang, Professor, National Institute for Viral Disease Control and Prevention, China CDC, Xicheng District, Yingxin Rd, Beijing 100052, China. euy-tokyo@umin.ac.jp
Telephone: +86-10-63555751  Fax: +86-10-63510565
Received: September 25, 2013
Revised: October 22, 2013
Accepted: November 3, 2013
Published online: November 28, 2013

Abstract

In the 1970s, scientists learned of a new pathogen causing non-A, non-B hepatitis. Classical approaches were used to isolate and characterize this new pathogen, but it could be transmitted experimentally only to chimpanzees and progress was slow until the pathogen was identified as hepatitis C virus (HCV) in 1989. Since then, research and treatment of HCV have expanded with the development of modern biological medicine: HCV genome organization and polyprotein processing were delineated in 1993; the first three-dimensional structure of HCV nonstructural protein (NS3 serine protease) was revealed in 1996; an infectious clone of HCV complementary DNA was first constructed in 1997; interferon and ribavirin combination therapy was established in 1998 and the therapeutic strategy gradually optimized; the HCV replicon system was produced in 1999; functional HCV pseudotyped viral particles were described in 2003; and recombinant infectious HCV in tissue culture was produced successfully in 2005. Recently, tremendous advances in HCV receptor discovery, understanding the HCV lifecycle, decryption of the HCV genome and proteins, as well as new anti-HCV compounds have been reported. Because HCV is difficult to isolate and culture, researchers have had to avail themselves to the best of modern biomedical technology; some of the major achievements in HCV research have not only advanced the understanding of HCV but also promoted knowledge of virology and cellular physiology. In this review, we summarize the advancements and remaining scotomas in the molecular virology and epidemiology of HCV.

Key Words: Hepatitis C virus, Hepatitis C virus lifecycle, Molecular virology, Hepatitis C virus models, Epidemiology

Core tip: The review summarizes the advancements, as well as remaining scotomas, in the molecular virology of hepatitis C virus (HCV). We emphasize the contributions of HuH-7 hepatocellular carcinoma cell line to development of the HCV replicon, cell culture-derived HCV, and HCV pseudoparticles. In addition, we reiterate the importance of epidemiological issues because accurate assessment of HCV-related disease burden has been overlooked. This review provides a history of the fight against HCV, which has required scientists to avail themselves to the best of modern biomedical technology, which in turn has enriched our knowledge of virology and cellular physiology.



INTRODUCTION

The hepatitis C virus (HCV) is an enveloped, single-stranded, positive-sense RNA virus, classified as a Hepacivirus within the Flaviviridae family[1-3]. The 9.6-kb RNA genome contains one long open reading frame (ORF) flanked by 5’ and 3’ untranslated regions (UTR)[4-6]. The single ORF encodes an approximately 3000 amino acid (aa) polyprotein that undergoes co- and post-translational cleavage by host and viral proteases to yield 10 viral proteins, not including the F protein[7,8]. The structural proteins, nucleic acid-binding nucleocapsid core protein and envelope proteins (E1 and E2/P7) are encoded by 25% of the N-terminal portion of the genome[9]. The remaining 75% of the genome encodes the non-structural proteins, NS2, NS3, NS4A, NS4B, NS5A and NS5B[9].

Humans are the primary reservoir of HCV[10]. HCV transmission occurs primarily through exposure to infected blood and the majority of individuals with persistent infection develop chronic hepatitis, which can progress to cirrhosis or hepatocellular carcinoma[11-13]. Different from other viruses, such as influenza A viruses and human immunodeficiency viruses, HCV is difficult to isolate and culture[14-16]. Since HCV was identified in 1989[1], basic research on HCV has been being hindered by the absence of reliable, reproducible, and efficient culture systems[11]. Recently, tremendous advances in understanding the HCV replicon[15,16], the pseudo-typed HCV viral particle[17], cell based culture systems[18,19], receptors[20-24], life cycle[25,26], structural biology and HCV therapy strategy[27-29] have been gained. However, several scotomas in the molecular virology and epidemiology of HCV remain to be elucidated. This review summarizes the advancements and remaining scotomas in the molecular virology and epidemiology of HCV.

MAJOR PROGRESS IN FIGHTING HCV

Since HCV was identified in 1989[1], virological research has led to a great deal of progress in the pathogenesis, diagnosis, treatment, control and prevention of the disease[11,30]. Since virus elimination is the ultimate goal of viral disease therapy, here we emphasize two major recent achievements in hepatitis C treatment. The first achievement was the development of direct-acting antiviral (DAA) agents, which are inhibitors of the HCV protease[31-37]. Although peginterferon and ribavirin remain vital components of therapy, the emergence of DAA agents has led to an unprecedented improvement in sustained virologic response rates to approximately 94%[30,38]. This is indicative of two milestones in virology: a therapy with the highest documented antiviral effects and optimism that the virus could be eliminated by medications. The second achievement is the identification of several single-nucleotide polymorphisms associated with spontaneous and treatment-induced clearance of HCV infection[39,40]. This discovery is also a milestone because only the rare single nucleotide polymorphisms of rs12980275 and rs8099917, near the interleukin28B gene, have any reported biological effect[39,40].

CURRENT HCV MODELS

The in vitro and in vivo models for HCV have evolved significantly since the discovery of the virus. With any virus, cell-based culture systems and animal models are the essential tools for virological study, vaccine development, and antiviral drug discovery. Many viruses, such as the influenza virus, are easy to isolate and culture in cell lines[41,42]; however, HCV is difficult to isolate and propagate[15,16,18,19]. Before HCV was identified, many virologists had attempted to isolate and culture the pathogen of non-A, non-B hepatitis using traditional cell-based approaches[43]. After struggling for decades, it was determined that HCV could only survive in human or chimpanzee fetal liver cells and hepatocytes or human peripheral blood mononuclear cells[44-49]. These cells are inconvenient to obtain and have a finite lifespan in culture, so even though these early studies showed that HCV was selective with a narrow host range[50,51], these methods made little contribution to HCV research (Table 1).

Table 1 Summary of in vitro and in vivo models for hepatitis C virus.
In vitro and in vivo modelsEstablished yearAdvantagesDeficiencies
In vitro
Cultivation of HCV1993-1999Achieved cultivation of HCV in human foetal liver cells, human hepatocytes or PBMC. Illustrated HCV is quite species selective and has a narrow range of hostsRequires specific cellular factors to support viral lifecycle. Primary human and chimpanzee hepatocytes or highly differentiated cells dependent. Most of them have yielded limited success. Poor reproducibility and low levels of HCV replication
HCV replicon1995-2000Provided a cell-based model for the study on HCV genome replication
HCV VLP1998-1999Rare evidence to support that HCV structural proteins core, E1, and E2 could form VLP
HCVpp2003Provided a convenient and feasible tool for studies on viral entry, HCV receptor, neutralizing antibody, etc.
HCVcc2005A break through in production of infectious hepatitis C virus in tissue culture
In vivo
Chimpanzee1979The only recognized animal model for HCV study, played a critical role in HCV discovery and play an essential role in defining the natural history of HCVChimpanzees differ from humans in their course of infection, that chronic carriers do not develop cirrhosis or fibrosis, limited availability, cost performance, and public resistance
Tree shrew1998Might be a succedaneum for chimpanzeesPersistent HCV infection could not be established and only 25% of infected animals developed transient or intermittent viremia. Germ line was not available to a small animal model
Chimeric human liver mouse2001Exhibited prolonged infection with high viral titers following inoculation with HCV isolated from human serum. HCV can be transmitted horizontally. Drug evaluationSince the mice were immunodeficient, they were not appropriate models to study HCV pathogenesis
Genetically humanized mouse2011Represents the first immunocompetent mice model for HCV study. Allows for the studies of HCV coreceptor biology in vivoOperation is difficult

Defeated by classical approaches to isolate HCV, virologists were forced to reproduce the HCV lifecycle using split models, which included the HCV genome RNA replication model (HCV replicon)[15,16], HCV structural proteins model (virus-like particle, VLP)[52] and HCV pseudotyped viral particles model[17] (Table 1). In 1999, Bartenschlager’s group in Germany established a HCV replicon system[15], followed soon after in 2000 by Rice’s group in the United States[16]. These models simulated the structure of the subgenomic selectable HCV replicons composed of the HCV 5’-UTR, the gene encoding the neomycin phosphotransferase or firefly luciferase, the encephalomyocarditis virus internal ribosome entry site, the region encoding HCV NS2-5B or NS3-5B, the authentic 3’-UTR, and the 12-16 5’-terminal codons of the core[15,16]. The replicon could replicate autonomously in hepatic cell cultures (e.g., HuH7 cell line)[15,16], leading to a series of experiments that examined the function of 5’- and 3’-UTR and NS3 to NS5B in HCV genome replication, described the viral life cycle, and led to the development of antiviral drugs[27]. The HCV replicon was able to replicate itself within the cell; however, it was not capable of producing infectious viruses[27]. Furthermore, this replicon was not able to reproduce in HuH-7 cells with high efficiency and for an extended period of time[50]. Virologists attempted to improve the replication efficiency and modify the robust HCV replicon using several methods, including adaptive mutation hunting, to reduce the non-HCV genome and increase the HCV genome composition, by attempting to replace various wild HCV strains of different genotypes[27,50,53].

Pseudotyped viral particles are commonly known as lentiviral vectors. These vectors are composed primarily of three viral elements; the gag-pol, which forms the viral structure, recognizes the viral genome and is responsible for the genome lifecycle; the viral mimic genome, which provides the genome elements that will be recognized by gag-pol and ensures complete viral RNA metabolism; and the envelope proteins, which are presented onto the artificial viral particle[54-58]. Additionally, a lentiviral vector contains a reporter gene inserted into the artificial viral genome. Although the lentiviral vector was used widely in gene transduction, presenting an HCV envelope protein functionally in this viral particle was not considered. Virologists tried to generate the HCV VLP[52], because classic virological experience told us that VLP of a certain virus could be produced by cloning and expressing virus structural proteins, and Liang’s group at the National Institutes of Health (Bethesda, MD) was successful in establishing the HCV-like particles using a baculovirus expression vector system[52]. In 2003, French virologist Bartosch et al[17] produced HCV pseudoparticles (HCVpp) using viral elements derived from murine leukemia virus. The HCVpp system led to advanced studies that identified a neutralizing antibody against HCV[59], explored HCV receptors and described the structure and function of the HCV envelope proteins[60-63].

In 2005, the Japanese virologist Wakita obtained a genotype 2a HCV strain (JFH-1) from a Japanese patient with a rare case of fulminant hepatitis C[18]. Based on the experience and methods accumulated in studying the HCV replicon, Wakita and his group rescued HCV in the HuH7 cell line, which was designated as HCVcc, for cell-culture-derived HCV[18]. HuH-7 cells infected with cloned and in vitro transcribed JFH-1 genomes produced viruses that were capable of infecting naïve HuH-7 cells[18]. In addition, the virus particles could be neutralized with a monoclonal antibody against the viral glycoprotein E2[18]. The study was the first in vitro experiment that showed the complete lifecycle of HCV. More importantly, virus obtained from the cell cultures was highly infectious in chimpanzees and immunodeficient mice with partial human livers[64].

As early as the 1970s, it was known that the etiological agent responsible for non-A and non-B hepatitis could be transmitted to chimpanzees[65], and chimpanzees were subsequently recognized as the only animal model of HCV[66] (Table 1). Chimpanzees played a critical role in defining the natural history of HCV[66] and since they are closely related to humans, any study of chimpanzees could reflect more closely what happens in humans than other animal models. However, chimpanzees that are chronic carriers of HCV do not develop cirrhosis or fibrosis[66,67], which are the most important consequences of HCV infection in humans.

Because chimpanzee studies are expensive and restricted by ethical responsibilities[67], scientists diverted their attention to other small animal models, such as the tree shrew and a chimeric human liver mouse. Xie et al[68] demonstrated that Tupaia could be infected by HCV when severely immunosuppressed; however, persistent HCV infection could not be established and only 25% of infected animals developed transient or intermittent viremia[51]. By genetically manipulating the urokinase-type plasminogen activator transgenic mouse, Mercer et al[69] transplanted normal human hepatocytes into severe combined immunodeficient mice carrying a plasminogen activator transgene. The chimeric mice exhibited prolonged infection with high viral titers following inoculation with HCV isolated from human serum[69]. Since the mice were immunodeficient, they were not appropriate models for investigation of HCV pathogenesis, although they were useful in assessing the activity of antiviral drugs, as well as neutralizing antibodies[51] (Table 1).

Mouse models of HCV provided little information about the human hepatocellular factors required for HCV entry. Thus, Ploss et al[24] introduced human CD81, scavenger receptor type B class 1, claudin 1, and OCLN genes into mice using a recombinant adenovirus expression system. They found that mice expressing these human factors were sufficient for HCV infection[24]. This system allowed for the investigation of HCV co-receptor biology in vivo and evaluation of passive immunization strategies and, therefore, represented the first immunocompetent small animal model for HCV[70] (Table 1).

SCOTOMAS IN MOLECULAR VIROLOGY

Although much progress has been made in all aspects of HCV research in the last few decades, we are still far from achieving the ultimate goal of complete HCV control and prevention. Thus, a better understanding of the HCV life cycle is essential to optimize the antiviral strategy. As mentioned above, the major challenges to HCV research are that HCV is difficult to isolate and culture, and no vaccine is available[14-16,71]. However, rather than summarizing the many achievements in HCV research, we have chosen to enumerate the scotomas in HCV molecular virology.

Structural biology of the HCV particle

Since HCV was first proposed to be a distinct infectious pathogen, virologists of that era attempted to visualize this enigmatic microbe using electron microscopy[43]. Different from other hepatitis viruses, including hepatitis A virus, hepatitis B virus and hepatitis E virus and the other viruses within the Flaviviridae family[43], no clear electron microscope image of HCV was reported until Chisari’s and Rice’s groups provided high-resolution images of highly enriched cell culture-derived HCV (HCVcc) particles in 2010 and 2013, respectively[72,73]. The reason for the difficulty in observing the crude HCV particle in HCV-harboring tissue remains unclear. The viral titer should not be an issue since viral copies in blood samples produced by HCV RNA are > 106/mL[74]. Serum-derived HCV particles are associated with the lipoprotein components apolipoprotein A-I (apoA-I), apoB-48, apoB-100, apoC-I and apoE[75]. The interaction between virus particles and serum lipoproteins suggests that HCV may form hybrid lipoviral particles[75] that facilitate virus entry into hepatocytes and protect the virus from the host immune response. In addition, the lipoprotein components might affect the morphological observation of crude HCV particles by electron microscopy. Our current knowledge of HCV morphology indicates that HCV particles are 40-80 nm in diameter, pleiomorphic, lack obvious symmetry or surface features and contain electron-dense cores[72,73]. The lack of details describing the overall architecture of HCV limits the ability of molecular biologists to study HCV structural biology and topology.

Molecular virology of HCV structural proteins

A quarter of the N-terminal region of the HCV polyprotein encodes the core structural protein and glycoproteins E1 and E2, which are believed to be incorporated into the HCV particle[17]. Based on general virological knowledge, the core protein should be a major component of the viral capsid, responsible for viral genome RNA recognition, binding and packaging[27]. Glycoproteins E1 and E2 located in the viral surface are also called envelope proteins; E1 and E2 are responsible for receptor recognition, receptor binding, endocytosis and membrane fusion[27]. The mature core protein contains a positively charged N-terminal RNA binding domain and a C-terminal domain that consists of two amphipathic helices and a palmitoylated cysteine residue to facilitate peripheral membrane binding[27]. Antibodies against core proteins are important for HCV serological detection[43]. Previous studies demonstrated that the core protein is involved in many pathogenic processes[43]. Furthermore, the core protein induces hepatocellular carcinoma in transgenic mice[76] and is a potent inhibitor of RNA silencing-based antiviral response[77]. However, the basic function of the core protein in capsid formation remains unknown. Although the region between amino acids 82 and 102 contains a tryptophan-rich sequence involved in homotypic core proteins interaction[78], the HCV core particle had not been successfully produced. In vitro nucleocapsid reconstitution experiments using the 1-124 or 1-179 core segments and structured RNA molecules have yielded irregular particles larger than those reported by the limited electron microscopy observations[79]. Scientists in China at Xiamen University successfully generated human papillomavirus VLP[80], hepatitis E virus VLP[81] and hepatitis B virus core capsids but failed to produce the HCV capsid (personal communication). In addition, the crystal structure of the HCV core protein is not yet available (Table 2).

Table 2 Summary of the properties of hepatitis C virus structural proteins.
CoreE1E2p7
Genome location342-914915-14901491-25792580-2769
Translation processing siteRough ER
Amino acid composition19119236363
Molecular weight (kDa)21-2333-3570-727
GlycosylationNoYesYesNo
CleavageER signal peptidase and SPP
Crystal structureNot availableRevealed
Functional unitDimerHeterodimer?Hexamer
Common functionViral particle formation. Core, E1 and E2, together with p7 and NS2, are required for virus assembly (assembly module)
Unique functionCapsid protein, viral particle formation, viral genome recognizing and packaging. Interacts with cLDs in early viral particle formation process. Counters host antiviral factors and involves pathogenesisEnvelope glycoproteins, interact with SRB1, CD81, CLDN1, OCLN, etc. to trigger viral entry. Promote fusion with the endosomal membrane. Counter host immune response via hypervariable regionsViroporin. Has key roles in organizing the virus assembly complex. p7-NS2 complex interacts with the NS3-4A enzyme to retrieve core protein from cLDs to form viral particle
Major scotomasHow do the core form the viral capsid? The signals and processes that mediate RNA packaging are largely unknown. What impeded us to resolve the structure of the viral glycoproteins? What is the real process in HCV entry? How are these receptors and co-receptors temporally and spatially used to ensure the early infection processes?

E1 and E2 are type I transmembrane glycoproteins assumed to be class II fusion proteins, with N-terminal ectodomains of 160 and 334 amino acids, respectively, and a short C-terminal transmembrane domain of approximately 30 amino acids[27,43]. Studies of HCVpp have indicated that 14 amino acids from the HCV core and 12 amino acids from the E1 C-terminus are required for E1 and E2 function[62]. The hemagglutinin and neuraminidase of influenza A viruses matches each other in a relative slack manner regardless of gene homology[54-58], while the matching pattern of HCV E1 and E2 is relatively strict. We separated E1 and E2 of HCV genotypes 1a, 1b, and 2a into two individual expression plasmids and replaced the transmembrane domains of 1b and 2a E1 and E2 with that of genotype 1a. The complementation features of E1 and E2, as well as the contributions of both the ecto- and transmembrane domains to the formation of the E1E2 complex, were evaluated using the HCVpp system[63]. We found that 1aE2 could not only complement its native 1aE1 but also 1bE1; in genotype 1b, glycoprotein complex formation is dependent primarily on the overall biological characteristics of the intact native E1 and E2; in genotype 2a, although the interaction of intact native E1 and E2 is critical for the formation of the glycoprotein complex, the ectodomain made a greater contribution than did the transmembrane domain[63]. This study suggested that E1 and E2 formed a functional envelope protein complex dependent on E1 and E2 expression[63]. E1 and E2 are assembled as non-covalent heterodimers[82,83], although the number of E1 molecules necessary to aggregate with E2 for biological function has not been elucidated. We highlight this scotoma because the envelope proteins of many viruses do not function simply in a 1:1 ratio[54-58] and viral proteins are not translated in equal numbers[43]. A lack of understanding of this point will impede receptor discovery, regardless of how many receptors and co-receptors are identified[75]. The crystal structure of the dengue virus glycoprotein, which is another member of the Flaviviridae family, was revealed in 2004[84]. By contrast, virologists failed to produce the HCV E1 or E2 crystal, which limits our understanding of the biological characteristics of HCV envelope proteins. Since one of the most important biological functions of HCV envelope proteins is membrane fusion, E1 and E2 are assumed to be class II fusion proteins[83]. This assumption has been challenged by recent studies suggesting that HCV and pestiviruses share an uncharacterized mechanism of membrane fusion[85]. These contradictory issues are common in HCV research and await further advances in our understanding of HCV virology (Table 2).

Molecular virology of non-structural HCV proteins

HCV proteins can be categorized into an assembly module (from core to NS2) and a replication module (from NS3 to NS5B) on the basis of viral essential functions[27,75]. Details on the function of structural and non-structural HCV proteins are lacking due to the limitations of in vitro models. There is also some controversy on topics such as whether the assembly module is necessary for viral particle formation or whether p7 is a structural or non-structural protein. Although the HCV replicon system provided solid evidence that non-structural proteins activate HCV RNA replication in vitro[15,16], some unresolved issues remain. These include why it is not possible to turn this system into a fully competent HCV cell culture model or why all replicons, except for the genotype 2a JFH-1 clone, contain cell-culture-adaptive mutations that when introduced back into viral genome, render it non-infectious in chimpanzees[67].

The p7 polypeptide is a small, 63-aa intrinsic membrane protein with a double-membrane-spanning topology in which its N- and C-terminal ends face the ER lumen[27]. Recent data indicate that p7 can mediate membrane ion permeability and form hexamers[86,87]. The three-dimensional structure of a hexameric p7 channel revealed a highly tilted, flower-shaped protein architecture with six protruding petals oriented toward the ER lumen[86,87]. These structural and membrane-permeability properties suggest that p7 belongs to the viroporin family and could play an important role in viral particle release and maturation[86,87]. However, the role of p7 in calcium and ion metabolism is unknown. Furthermore, HCVpp with or without p7 showed no changes in viral particle formation and pp infectivity[62,63]. A study of the closely related GB virus B, which infects tamarins and has an analogous but larger protein, p13, showed that p13 is processed into two components p6 and p7, and that p6 was dispensable while p7 was essential for infectivity[88].

NS2 is a metal-dependent proteinase, whose functions are dependent on the interaction with p7 and NS3[27]. Although the NS2 protease is dispensable for RNA replication, NS2 participates in proteolytic cleavage at the NS2-NS3 junction of the polyprotein[27]. The transmembrane and protease domains of NS2 are required for infectious virus assembly[89]. Why NS2 is critical for viral particle formation remains unknown and the interactions between NS2 and other structural and non-structural viral proteins to form an unknown viral particle formation network should be explored (Table 3).

Table 3 Summary of the properties of hepatitis C virus non-structural proteins.
NS2NS3NS4ANS4BNS5ANS5B
Genome location2769-34193420-53125313-54745475-62576258-76017602-9378
Translation processing siteRough ER
Amino acid composition810-10261027-16571658-17111712-19721973-24202421-3012
Molecular weight (kDa)21-237082756-5865-68
CleavageViral cysteine protease NS2-3 and the serine-type protease activity of the viral NS3-NS4A complex
Crystal structureC-terminal (aa904-1026) was solvedRevealedRevealedNot availableRevealedRevealed
Functional unitHomodimerMonomer or oligomerMonomerOligomerHomodimerMonomer
Common functionReplication module
Unique functionA metal-dependent proteinase, many functions dependent on the interaction with P7 and NS3. Participation in proteolytic cleavage at the NS2-NS3 junction of the polyprotein. Both the TMDs and protease domain of NS2 are required for the production of virus particlesThe DAA targeting protein. NS3 was anchored in ER membrane by cofactor NS4A. NS3-4A complex has serine-type protease activity and NTPase/RNA helicase activities. Nonspecific cleavage of two critical interferon induction proteins: MAVS and TRIFThe central portion of NS4A, residues 21-32, intercalates into NS3 and activates the protease activity by stabilizing this protease subdomain and contributing to the substrate recognition site. The C-terminal acidic portion of NS4A interacts with the NS3 helicase and other HCV proteins and contributes to RNA replication as well as assemblyA master organizer of replication complex formation. NTPase activity? RNA binding?Produced as multiple phospho-variants. RNA-binding phosphoprotein involved in RNA replication. Phosphorylation of a specific serine residue within the C-terminus by CKIIα is essential for virus assembly. The interaction of NS5A with the cLD-bound core protein is the key steps in HCV assemblyRNA-dependent RNA polymerase
Major scotomasHow HCV particles are organized? What is the accurate duty of each nonstructural protein in viral lifecycle? How do the nonstructural proteins utilize host cellular factors for its own survival? Why HCV lifecycle is tightly associated with components of LDLs and VLDLs?

NS3 is a 70-kDa multifunctional protein anchored by the cofactor NS4A[27,89]. NS2/NS3 junction cleavage is essential to liberate fully functional NS3 protein[27,89]. NS3-4A is a non-covalent complex, with a serine protease located in the N-terminus (aa 1-180) and an NTPase/RNA helicase in the C-terminus (aa 181-631)[27,89]. The substrate specificity of NS3-4A is low and causes non-specific cleavage of host proteins; e.g., mitochondrial antiviral-signaling (MAVS) and TIR domain-containing adaptor inducing interferon β (TRIF), and thus might impact host IFN response[90]. The central portion of NS4A, residues 21-32, intercalates into NS3 and activates the protease activity by stabilizing this protease subdomain and contributing to the substrate recognition site[27]. The C-terminal acidic portion of NS4A interacts with the NS3 helicase and other HCV proteins and contributes to RNA replication, as well as assembly[27]. The DAAs telaprevir and boceprevir[31-37] are inhibitors targeting the NS3-4A protease that displayed promising effects in clinical trials, indicating that the NS3-4A protease is critical for viral life cycle (Table 3).

NS4B is a poorly characterized hydrophobic 27-kDa protein[27] comprised of a 66-aa N-terminal portion, a 120-aa central portion, and a 70-aa C-terminal portion[91]. Four transmembrane-spanning regions were predicted in the central portion, while the N-terminal portion plays an important role in assembly of a functional replication complex[27,91]. Einav et al[92,93] and Thompson et al[94] demonstrated that NS4B harbors NTPase activity and has a role in viral assembly.

NS5A is a 447-aa membrane-associated protein that plays an important role in modulating HCV RNA replication and particle formation[91]. NS5A can be detected in basally phosphorylated and hyper-phosphorylated forms with molecular weights of 56- and 58-kDa, respectively[95,96]. NS5A is comprised of four domains: a N-terminal membrane anchor and three domains separated by two low complexity sequences[27,91]. The three separated domains are domain 1, aa 36-213; domain 2, aa 250-342 and domain 3, aa 356-447. Domains 1 and 2 are involved in RNA replication and domain 1 is involved in cellular lipid drop binding, domain 3 is essential for viral assembly and is involved in interaction with the core protein accumulated in cellular lipid drops[27,91]. Although studies showed that NS5A is critical for HCV RNA replication, deletions in D2 and D3 are tolerated in RNA replication[97], and viable replicons and viruses harboring GFP insertions displayed no change on HCV RNA replication[98]. Phosphorylation of a specific serine residue within C-terminal by casein kinase IIα is essential for virus assembly[99]. The interaction of NS5A with the cytosolic lipid droplets-bound core protein is a key step in HCV assembly[97,100,101].

NS5B is an RNA-dependent RNA polymerase (RdRp). Its crystal structure was revealed in 1999[102]; the active site is highly conserved and located in the palm subdomain[91]. The low substrate specificity allows for the incorporation of ribavirin into nascent RNA. Thus, ribavirin remains a perfect RNA analog in HCV therapy[103]. Although recombinant NS5B is available and its crystal structure is known[104,105], its role in HCV RNA replication remains unclear (Table 3).

5’-non-translated regions and 3’ non-translated regions

Viral non-translated regions (NTRs) and non-coding regions, harbor important biological functions, involving viral genome reorganization, replication, translation initiation, and viral assembly[106]. The HCV 5’NTR contains 341 bp (H77 strain, NCBI Reference Sequence: NC_004102.1)[107]. The predicted secondary structure of the HCV 5’NTR consists of four domains (domains I-IV, numbered from 5’ to 3’), and the largest domain III was further categorized into sub-domains a-f[107]. The major functional unit in the HCV 5’NTR is an IRES, which includes three domains (II-IV)[106-109]. Initiation of protein synthesis in host cells utilized by HCV is different from mRNA translation in eukaryotes because HCV initiates viral protein synthesis via its IRES, which is known as internal translation initiation. This process is a cap-independent mechanism of recruiting, positioning and activating the host cellular protein synthesis machinery driven by the HCV IRES[106-109], which is relatively weak in directing protein translation compared to the IRESs of other viruses, and may contribute to an insufficient host immune response[110]. Although structural and biochemical studies of the IRES found in HCV have provided the most detailed information thus far regarding the mechanism of IRES driven translation, unresolved issues remain. For example, it is unknown whether the HCV IRES acts as one determining factor for hepatotropism or how the HCV 5’NTR interacts with the 3’NTR to support HCV RNA replication and polyprotein translation. Additionally, the biological impact of the NTR to each hepatitis virus remains unclear since the HCV NTRs have a different structure compared to those of hepatitis A and E viruses. Finally, the interaction of miR-122 with the HCV 5’NTR to facilitate replication of viral RNA remains to be fully elucidated[111].

The 3’ terminal of any genome is technically difficult to identify and the available complete sequence of the HCV 3’NTR is unusual. The 3’ UTR is divided into three structurally distinct domains from 5’ to 3’, an upstream variable region of about 40 nucleotides, a long poly (U)-poly (U/UC) tract and a 98-nucleotide (3’ X) sequence that forms three stem-loop structures[43,106]. The long poly (U)-poly (U/UC) tract was a major obstacle to obtaining an HCV genomic clone because no known DNA polymerase could amplify this region and the fidelity of a reverse transcriptase in this region was suspect. The function of the 3’ UTR remains to be determined[43,106]. It may play an important role in minus intermediate RNA and genome RNA synthesis during HCV RNA replication[43,106] since variable region deletions of RNA replicons could replicate, albeit at a much lower level[112]. However, deletion of either the poly (U/UC) or the 3’X was not viable, which suggested that the poly (U/UC) and 3’X regions are critical for HCV RNA replication[107].

HuH7 cell line

HCV researchers should be familiar with the human hepatocellular carcinoma cell line HuH-7, also known as Huh7 or HuH7. This cell line is critical because the HCV replicon, HCVcc, and HCVpp are all dependent on this cell line or its derivatives, indicating that it harbors all critical factors for HCV replication, assembly, budding and entry[43,113-115]. HuH-7 is a well-differentiated hepatocyte-derived hepatocellular carcinoma cell line that originated from a liver tumor in a 57-year-old Japanese male in 1982 (http://huh7.com/). It was established by scientists at Okayama University of Japan in the 1980s (http://cellbank.nibio.go.jp/legacy/celldata/jcrb0403.htm). HuH-7 remains the only hepatocellular carcinoma cell line that can fully support the HCV life cycle. Improvements in hepatocellular carcinoma cell line isolation could provide more effective HCV-supporting cell lines; alternatively the advances in induced pluripotent stem cells could result in a breakthrough in HCV culture and isolation.

SCOTOMAS IN EPIDEMIOLOGY OF HCV

HCV carries a large disease burden in some countries and is the second most studied virus. Most HCV infections are subclinical with a long and insidious disease course. However, epidemiological surveillance of HCV is relatively weak compared to some acute respiratory transmission diseases (http://www.who.int/influenza/surveillance_monitoring/en/).

Chronic HCV carriers are the only reservoir of HCV since chimpanzees could be infected with HCV only experimentally[43,106]. HCV transmission occurs primarily through exposure to HCV-infected blood[43,106]. Blood transfusion, solid organ transplantation from an infected donor, and unsafe medical practices were the major transmission routes before HCV was identified in 1989[43,106]. Beginning in the early 1990s, strict screening of blood donors and precise control over the blood supply were implemented by national governments[43,106,116]. The majority of HCV infections are now limited to specific subpopulations, such as intravenous drug users and patients with certain hemopathies[117]. Although unsafe medical practices, occupational exposure to infected blood, maternal-fetal transmission, sex with an infected person and high-risk sexual practices are believed to be HCV transmission routes, the rate of acquisition of infection by these routes is low[118]. The average incubation period is 6-10 weeks and most virologists and hepatologists consider that up to 80% of HCV infected persons do not eliminate HCV spontaneously[43,106]. Cirrhosis and liver failure develop in 10%-20% of chronic HCV carriers; 1%-5% of chronic HCV carriers develop hepatocellular carcinoma[43,106] (Table 4).

Table 4 Epidemiological features of hepatitis C virus infection.
Epidemiological indexCurrent consensus
Source of infectionChronic HCV carriers
Route of transmissionHCV transmission occurs primarily through exposure to infected blood. Past: Receiving infected blood or organ transplantation, from accidental exposure to infected blood, and sexual transmission in persons with high risk behaviours. Present: HCV is usually spread by sharing infected needles with a chronic HCV carrier, and some people acquire the infection through nonparenteral means that have not been fully defined.
Susceptible populationGeneral population
Incubation periodAverage 6-10 wk
Prevalence and incidence3% of the world’s population have HCV
Rate of chronic infectionUp to more than 80%
Outcome of chronic infection10%-20% of chronic HCV carriers may develop into cirrhosis and liver failure. 1%-5% of chronic HCV carriers are associated with the development of hepatocellular carcinoma
Molecular epidemiologyHCV is classified into eleven major genotypes (designated as 1-11), many subtypes (designated a, b, c, etc.), and about 100 different strains (numbered 1, 2, 3, etc.) based on the genomic sequence heterogeneity. Genotypes 1-3 have a worldwide distribution. Types 1a and 1b are the most common, accounting for about 60% of global infections. Type 2 is less frequently represented than type 1. Type 3 is endemic in southeast Asia and is variably distributed in different countries. Genotype 4 is principally found in the Middle East, Egypt, and central Africa. Type 5 is almost exclusively found in South Africa, and genotypes 6-11 are distributed in Asia.
StabilityHCV is inactivated by exposure to lipid solvents or detergents, heating at 60 °C for 10 h or 100 °C for 2 min in aqueous solution, formaldehyde (1:2000) at 37 °C for 72 h, β-propriolactone and UV irradiation.
VaccineNot available

The World Health Organization estimates that up to 3% of the global population is infected with HCV (http://www.who.int/csr/disease/hepatitis/Hepc.pdf), and the peak disease burden is expected around 2020[119]. However, these estimations lack sufficient evidence. Firstly, it is hard to track the origin of the data since most authors citing this statistic used inaccurate citations. Secondly, the HCV serological detection kit has undergone at least three iterations[117]. The first test developed in 1990 detected antibody to a single epitope within the core protein by enzyme linked immunosorbent assay and provided data on 170 million HCV carriers, even though it was plagued by poor sensitivity[43]. Third-generation enzyme immunoassays included antibodies against multiple antigens, which increased the sensitivity significantly[43], although no large-scale serological investigations have been performed. In China, a nationwide HCV serological survey performed in 2006 showed the prevalence of anti-HCV antibodies to be < 0.5% among more than 80000 Chinese subjects[116]. Furthermore, the rates of HCV were much lower than those of hepatitis B among clinical inpatient and outpatient populations, which was significantly different from a Japanese population[119,120]. Epidemiology is important because it will provide basic knowledge of disease and inaccurate epidemiological data will lead to inaccuracies in our knowledge of the disease burden, natural history and therapeutic efficacy.

The number of people that will become chronic carriers after HCV infection remains unknown. Scientists believe that as many as 40%-80% of HCV infections will develop into chronic infections[121-123] (Table 3). While these estimates are also likely inaccurate, how and when people are infected must be determined to ascertain a more precise figure. One recent cross-sectional study performed in intravenous drug users challenged the current assumptions regarding the rate of chronic infection; in that study as many as 77.8% of individuals cleared HCV infection without the need for anti-viral therapy[117].

Footnotes

P- Reviewers: Teschke R, Wong GLH S- Editor: Gou SX L- Editor: Wang TQ 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: ]  [Cited by in Crossref: 4996]  [Cited by in F6Publishing: 4592]  [Article Influence: 131.2]  [Reference Citation Analysis (0)]
2.  Alter MJ, Hadler SC, Judson FN, Mares A, Alexander WJ, Hu PY, Miller JK, Moyer LA, Fields HA, Bradley DW. Risk factors for acute non-A, non-B hepatitis in the United States and association with hepatitis C virus infection. JAMA. 1990;264:2231-2235.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 337]  [Cited by in F6Publishing: 331]  [Article Influence: 9.7]  [Reference Citation Analysis (0)]
3.  Weiner AJ, Kuo G, Bradley DW, Bonino F, Saracco G, Lee C, Rosenblatt J, Choo QL, Houghton M. Detection of hepatitis C viral sequences in non-A, non-B hepatitis. Lancet. 1990;335:1-3.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 543]  [Cited by in F6Publishing: 548]  [Article Influence: 16.1]  [Reference Citation Analysis (0)]
4.  Miller RH, Purcell RH. Hepatitis C virus shares amino acid sequence similarity with pestiviruses and flaviviruses as well as members of two plant virus supergroups. Proc Natl Acad Sci USA. 1990;87:2057-2061.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Purcell R. The hepatitis C virus: overview. Hepatology. 1997;26:11S-14S.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 80]  [Article Influence: 3.0]  [Reference Citation Analysis (0)]
6.  Pawlotsky JM. Genetic heterogeneity and properties of hepatitis C virus. Acta Gastroenterol Belg. 1998;61:189-191.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Niepmann M. Hepatitis C virus RNA translation. Curr Top Microbiol Immunol. 2013;369:143-166.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 57]  [Article Influence: 5.2]  [Reference Citation Analysis (0)]
8.  Hellen CU, Pestova TV. Translation of hepatitis C virus RNA. J Viral Hepat. 1999;6:79-87.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 69]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
9.  Lohmann V. Hepatitis C virus RNA replication. Curr Top Microbiol Immunol. 2013;369:167-198.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 36]  [Cited by in F6Publishing: 71]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
10.  Alter MJ. Epidemiology of hepatitis C virus infection. World J Gastroenterol. 2007;13:2436-2441.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  NIH Consensus Statement on Management of Hepatitis C: 2002 NIH Consens State Sci Statements. 2002;19:1-46.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Di Bisceglie AM, Simpson LH, Lotze MT, Hoofnagle JH. Development of hepatocellular carcinoma among patients with chronic liver disease due to hepatitis C viral infection. J Clin Gastroenterol. 1994;19:222-226.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Purcell RH. Does non-A, non-B hepatitis cause hepatocellular carcinoma? Cancer Detect Prev. 1989;14:203-207.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Takikawa S, Matsuura Y, Miyamura T. [The present studies of the development of HCV vaccine]. Nihon Rinsho. 2001;59:1379-1383.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Lohmann V, Körner F, Koch J, Herian U, Theilmann L, Bartenschlager R. Replication of subgenomic hepatitis C virus RNAs in a hepatoma cell line. Science. 1999;285:110-113.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2294]  [Cited by in F6Publishing: 2224]  [Article Influence: 89.0]  [Reference Citation Analysis (0)]
16.  Blight KJ, Kolykhalov AA, Rice CM. Efficient initiation of HCV RNA replication in cell culture. Science. 2000;290:1972-1974.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1154]  [Cited by in F6Publishing: 1139]  [Article Influence: 47.5]  [Reference Citation Analysis (0)]
17.  Bartosch B, Dubuisson J, Cosset FL. Infectious hepatitis C virus pseudo-particles containing functional E1-E2 envelope protein complexes. J Exp Med. 2003;197:633-642.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 885]  [Cited by in F6Publishing: 868]  [Article Influence: 41.3]  [Reference Citation Analysis (0)]
18.  Wakita T, Pietschmann T, Kato T, Date T, Miyamoto M, Zhao Z, Murthy K, Habermann A, Kräusslich HG, Mizokami M. Production of infectious hepatitis C virus in tissue culture from a cloned viral genome. Nat Med. 2005;11:791-796.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2241]  [Cited by in F6Publishing: 2239]  [Article Influence: 117.8]  [Reference Citation Analysis (0)]
19.  Lindenbach BD, Evans MJ, Syder AJ, Wölk B, Tellinghuisen TL, Liu CC, Maruyama T, Hynes RO, Burton DR, McKeating JA. Complete replication of hepatitis C virus in cell culture. Science. 2005;309:623-626.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1843]  [Cited by in F6Publishing: 1817]  [Article Influence: 95.6]  [Reference Citation Analysis (0)]
20.  Agnello V, Abel G, Elfahal M, Knight GB, Zhang QX. Hepatitis C virus and other flaviviridae viruses enter cells via low density lipoprotein receptor. Proc Natl Acad Sci USA. 1999;96:12766-12771.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 702]  [Cited by in F6Publishing: 688]  [Article Influence: 27.5]  [Reference Citation Analysis (0)]
21.  Scarselli E, Ansuini H, Cerino R, Roccasecca RM, Acali S, Filocamo G, Traboni C, Nicosia A, Cortese R, Vitelli A. The human scavenger receptor class B type I is a novel candidate receptor for the hepatitis C virus. EMBO J. 2002;21:5017-5025.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 888]  [Cited by in F6Publishing: 868]  [Article Influence: 39.5]  [Reference Citation Analysis (0)]
22.  Evans MJ, von Hahn T, Tscherne DM, Syder AJ, Panis M, Wölk B, Hatziioannou T, McKeating JA, Bieniasz PD, Rice CM. Claudin-1 is a hepatitis C virus co-receptor required for a late step in entry. Nature. 2007;446:801-805.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 930]  [Cited by in F6Publishing: 915]  [Article Influence: 53.8]  [Reference Citation Analysis (0)]
23.  Pileri P, Uematsu Y, Campagnoli S, Galli G, Falugi F, Petracca R, Weiner AJ, Houghton M, Rosa D, Grandi G. Binding of hepatitis C virus to CD81. Science. 1998;282:938-941.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1572]  [Cited by in F6Publishing: 1521]  [Article Influence: 58.5]  [Reference Citation Analysis (0)]
24.  Ploss A, Evans MJ, Gaysinskaya VA, Panis M, You H, de Jong YP, Rice CM. Human occludin is a hepatitis C virus entry factor required for infection of mouse cells. Nature. 2009;457:882-886.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 715]  [Cited by in F6Publishing: 705]  [Article Influence: 47.0]  [Reference Citation Analysis (0)]
25.  Lindenbach BD, Rice CM. Unravelling hepatitis C virus replication from genome to function. Nature. 2005;436:933-938.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 602]  [Cited by in F6Publishing: 583]  [Article Influence: 30.7]  [Reference Citation Analysis (0)]
26.  Moradpour D, Penin F, Rice CM. Replication of hepatitis C virus. Nat Rev Microbiol. 2007;5:453-463.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Bartenschlager R, Lohmann V, Penin F. The molecular and structural basis of advanced antiviral therapy for hepatitis C virus infection. Nat Rev Microbiol. 2013;11:482-496.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 287]  [Cited by in F6Publishing: 276]  [Article Influence: 25.1]  [Reference Citation Analysis (0)]
28.  Jacobson IM, Pawlotsky JM, Afdhal NH, Dusheiko GM, Forns X, Jensen DM, Poordad F, Schulz J. A practical guide for the use of boceprevir and telaprevir for the treatment of hepatitis C. J Viral Hepat. 2012;19 Suppl 2:1-26.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 55]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]
29.  Ghany MG, Nelson DR, Strader DB, Thomas DL, Seeff LB. An update on treatment of genotype 1 chronic hepatitis C virus infection: 2011 practice guideline by the American Association for the Study of Liver Diseases. Hepatology. 2011;54:1433-1444.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 803]  [Cited by in F6Publishing: 834]  [Article Influence: 64.2]  [Reference Citation Analysis (0)]
30.  Ghany MG, Strader DB, Thomas DL, Seeff LB. Diagnosis, management, and treatment of hepatitis C: an update. Hepatology. 2009;49:1335-1374.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2252]  [Cited by in F6Publishing: 2192]  [Article Influence: 146.1]  [Reference Citation Analysis (1)]
31.  Kwo PY, Lawitz EJ, McCone J, Schiff ER, Vierling JM, Pound D, Davis MN, Galati JS, Gordon SC, Ravendhran N. Efficacy of boceprevir, an NS3 protease inhibitor, in combination with peginterferon alfa-2b and ribavirin in treatment-naive patients with genotype 1 hepatitis C infection (SPRINT-1): an open-label, randomised, multicentre phase 2 trial. Lancet. 2010;376:705-716.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 519]  [Cited by in F6Publishing: 567]  [Article Influence: 40.5]  [Reference Citation Analysis (0)]
32.  Poordad F, McCone J, Bacon BR, Bruno S, Manns MP, Sulkowski MS, Jacobson IM, Reddy KR, Goodman ZD, Boparai N. Boceprevir for untreated chronic HCV genotype 1 infection. N Engl J Med. 2011;364:1195-1206.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1948]  [Cited by in F6Publishing: 1951]  [Article Influence: 150.1]  [Reference Citation Analysis (0)]
33.  Bacon BR, Gordon SC, Lawitz E, Marcellin P, Vierling JM, Zeuzem S, Poordad F, Goodman ZD, Sings HL, Boparai N. Boceprevir for previously treated chronic HCV genotype 1 infection. N Engl J Med. 2011;364:1207-1217.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1287]  [Cited by in F6Publishing: 1288]  [Article Influence: 99.1]  [Reference Citation Analysis (0)]
34.  McHutchison JG, Everson GT, Gordon SC, Jacobson IM, Sulkowski M, Kauffman R, McNair L, Alam J, Muir AJ. Telaprevir with peginterferon and ribavirin for chronic HCV genotype 1 infection. N Engl J Med. 2009;360:1827-1838.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 851]  [Cited by in F6Publishing: 916]  [Article Influence: 61.1]  [Reference Citation Analysis (0)]
35.  Hézode C, Forestier N, Dusheiko G, Ferenci P, Pol S, Goeser T, Bronowicki JP, Bourlière M, Gharakhanian S, Bengtsson L. Telaprevir and peginterferon with or without ribavirin for chronic HCV infection. N Engl J Med. 2009;360:1839-1850.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 837]  [Cited by in F6Publishing: 888]  [Article Influence: 59.2]  [Reference Citation Analysis (0)]
36.  Jacobson IM, McHutchison JG, Dusheiko G, Di Bisceglie AM, Reddy KR, Bzowej NH, Marcellin P, Muir AJ, Ferenci P, Flisiak R. Telaprevir for previously untreated chronic hepatitis C virus infection. N Engl J Med. 2011;364:2405-2416.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1866]  [Cited by in F6Publishing: 1835]  [Article Influence: 141.2]  [Reference Citation Analysis (0)]
37.  Zeuzem S, Andreone P, Pol S, Lawitz E, Diago M, Roberts S, Focaccia R, Younossi Z, Foster GR, Horban A. Telaprevir for retreatment of HCV infection. N Engl J Med. 2011;364:2417-2428.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1205]  [Cited by in F6Publishing: 1194]  [Article Influence: 91.8]  [Reference Citation Analysis (0)]
38.  Cartwright EJ, Miller L. Novel drugs in the management of difficult-to-treat hepatitis C genotypes. Hepatic Medicine: Evidence and Research. 2013;5:53-61.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 7]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
39.  Thomas DL, Thio CL, Martin MP, Qi Y, Ge D, O’Huigin C, Kidd J, Kidd K, Khakoo SI, Alexander G. Genetic variation in IL28B and spontaneous clearance of hepatitis C virus. Nature. 2009;461:798-801.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1667]  [Cited by in F6Publishing: 1635]  [Article Influence: 109.0]  [Reference Citation Analysis (0)]
40.  Ge D, Fellay J, Thompson AJ, Simon JS, Shianna KV, Urban TJ, Heinzen EL, Qiu P, Bertelsen AH, Muir AJ. Genetic variation in IL28B predicts hepatitis C treatment-induced viral clearance. Nature. 2009;461:399-401.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 2776]  [Cited by in F6Publishing: 2666]  [Article Influence: 177.7]  [Reference Citation Analysis (0)]
41.  Krauss S, Walker D, Webster RG. Influenza virus isolation. Methods Mol Biol. 2012;865:11-24.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 18]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
42.  Diane S, Leland , Christine C. Ginocchio. Role of Cell Culture for Virus Detection in the Age of Technology. Clin Microbiol Rev. 2007;20:49-78.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 369]  [Cited by in F6Publishing: 303]  [Article Influence: 17.8]  [Reference Citation Analysis (0)]
43.  Thomas HC, Lemon SM, Zuckerman AJ.  Viral Hepatitis. 3rd ed. Hoboken: Blackwell Publishing 2005; 824-840.  [PubMed]  [DOI]  [Cited in This Article: ]
44.  Ito T, Mukaigawa J, Zuo J, Hirabayashi Y, Mitamura K, Yasui K. Cultivation of hepatitis C virus in primary hepatocyte culture from patients with chronic hepatitis C results in release of high titre infectious virus. J Gen Virol. 1996;77:1043-1054.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 67]  [Article Influence: 2.4]  [Reference Citation Analysis (0)]
45.  Shimizu YK, Iwamoto A, Hijikata M, Purcell RH, Yoshikura H. Evidence for in vitro replication of hepatitis C virus genome in a human T-cell line. Proc Natl Acad Sci USA. 1992;89:5477-5481.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 226]  [Cited by in F6Publishing: 233]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
46.  Kato N, Nakazawa T, Mizutani T, Shimotohno K. Susceptibility of human T-lymphotropic virus type I infected cell line MT-2 to hepatitis C virus infection. Biochem Biophys Res Commun. 1995;206:863-869.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Lanford RE, Sureau C, Jacob JR, White R, Fuerst TR. Demonstration of in vitro infection of chimpanzee hepatocytes with hepatitis C virus using strand-specific RT/PCR. Virology. 1994;202:606-614.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  Iacovacci S, Manzin A, Barca S, Sargiacomo M, Serafino A, Valli MB, Macioce G, Hassan HJ, Ponzetto A, Clementi M. Molecular characterization and dynamics of hepatitis C virus replication in human fetal hepatocytes infected in vitro. Hepatology. 1997;26:1328-1337.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Fournier C, Sureau C, Coste J, Ducos J, Pageaux G, Larrey D, Domergue J, Maurel P. In vitro infection of adult normal human hepatocytes in primary culture by hepatitis C virus. J Gen Virol. 1998;79:2367-2374.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Duverlie G, Wychowski C. Cell culture systems for the hepatitis C virus. World J Gastroenterol. 2007;13:2442-2445.  [PubMed]  [DOI]  [Cited in This Article: ]
51.  Carcamo WC, Nguyen CQ. Advancement in the development of models for hepatitis C research. J Biomed Biotechnol. 2012;2012:346761.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 7]  [Cited by in F6Publishing: 8]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
52.  Baumert TF, Vergalla J, Satoi J, Thomson M, Lechmann M, Herion D, Greenberg HB, Ito S, Liang TJ. Hepatitis C virus-like particles synthesized in insect cells as a potential vaccine candidate. Gastroenterology. 1999;117:1397-1407.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 90]  [Cited by in F6Publishing: 94]  [Article Influence: 3.8]  [Reference Citation Analysis (0)]
53.  Scheel TK, Prentoe J, Carlsen TH, Mikkelsen LS, Gottwein JM, Bukh J. Analysis of functional differences between hepatitis C virus NS5A of genotypes 1-7 in infectious cell culture systems. PLoS Pathog. 2012;8:e1002696.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 31]  [Cited by in F6Publishing: 32]  [Article Influence: 2.7]  [Reference Citation Analysis (0)]
54.  Wu J, Zhang F, Wang M, Xu C, Song J, Zhou J, Lin X, Zhang Y, Wu X, Tan W. Characterization of neuraminidases from the highly pathogenic avian H5N1 and 2009 pandemic H1N1 influenza A viruses. PLoS One. 2010;5:e15825.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 13]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
55.  Liu Y, Liu X, Fang J, Shen X, Chen W, Lin X, Li H, Tan W, Wang Y, Zhao P. Characterization of antibodies specific for hemagglutinin and neuraminidase proteins of the 1918 and 2009 pandemic H1N1 viruses. Vaccine. 2010;29:183-190.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 6]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
56.  Du N, Zhou J, Lin X, Zhang Y, Yang X, Wang Y, Shu Y. Differential activation of NK cells by influenza A pseudotype H5N1 and 1918 and 2009 pandemic H1N1 viruses. J Virol. 2010;84:7822-7831.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 41]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
57.  Zhang Y, Lin X, Wang G, Zhou J, Lu J, Zhao H, Zhang F, Wu J, Xu C, Du N. Neuraminidase and hemagglutinin matching patterns of a highly pathogenic avian and two pandemic H1N1 influenza A viruses. PLoS One. 2010;5:e9167.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 1.7]  [Reference Citation Analysis (0)]
58.  Zhang Y, Lin X, Zhang F, Wu J, Tan W, Bi S, Zhou J, Shu Y, Wang Y. Hemagglutinin and neuraminidase matching patterns of two influenza A virus strains related to the 1918 and 2009 global pandemics. Biochem Biophys Res Commun. 2009;387:405-408.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 19]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
59.  Bartosch B, Bukh J, Meunier JC, Granier C, Engle RE, Blackwelder WC, Emerson SU, Cosset FL, Purcell RH. In vitro assay for neutralizing antibody to hepatitis C virus: evidence for broadly conserved neutralization epitopes. Proc Natl Acad Sci USA. 2003;100:14199-14204.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 253]  [Cited by in F6Publishing: 241]  [Article Influence: 11.5]  [Reference Citation Analysis (0)]
60.  Hsu M, Zhang J, Flint M, Logvinoff C, Cheng-Mayer C, Rice CM, McKeating JA. Hepatitis C virus glycoproteins mediate pH-dependent cell entry of pseudotyped retroviral particles. Proc Natl Acad Sci USA. 2003;100:7271-7276.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 655]  [Cited by in F6Publishing: 636]  [Article Influence: 30.3]  [Reference Citation Analysis (0)]
61.  Cocquerel L, Voisset C, Dubuisson J. Hepatitis C virus entry: potential receptors and their biological functions. J Gen Virol. 2006;87:1075-1084.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 132]  [Cited by in F6Publishing: 141]  [Article Influence: 7.8]  [Reference Citation Analysis (0)]
62.  Bian T, Zhou Y, Bi S, Tan W, Wang Y. HCV envelope protein function is dependent on the peptides preceding the glycoproteins. Biochem Biophys Res Commun. 2009;378:118-122.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 14]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
63.  Lin X, Zhang Y, Bi S, Lu J, Zhao H, Tan W, Li D, Wang Y. Hepatitis C virus envelope glycoproteins complementation patterns and the role of the ecto- and transmembrane domains. Biochem Biophys Res Commun. 2009;385:257-262.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 6]  [Cited by in F6Publishing: 8]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
64.  Lindenbach BD, Meuleman P, Ploss A, Vanwolleghem T, Syder AJ, McKeating JA, Lanford RE, Feinstone SM, Major ME, Leroux-Roels G. Cell culture-grown hepatitis C virus is infectious in vivo and can be recultured in vitro. Proc Natl Acad Sci USA. 2006;103:3805-3809.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 353]  [Cited by in F6Publishing: 338]  [Article Influence: 18.8]  [Reference Citation Analysis (0)]
65.  Prince AM. Non-A, non-B hepatitis viruses. Annu Rev Microbiol. 1983;37:217-232.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 11]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
66.  Bukh J. Animal models for the study of hepatitis C virus infection and related liver disease. Gastroenterology. 2012;142:1279-1287.e3.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 104]  [Article Influence: 8.7]  [Reference Citation Analysis (0)]
67.  Bukh J. A critical role for the chimpanzee model in the study of hepatitis C. Hepatology. 2004;39:1469-1475.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 136]  [Cited by in F6Publishing: 130]  [Article Influence: 6.5]  [Reference Citation Analysis (0)]
68.  Xie ZC, Riezu-Boj JI, Lasarte JJ, Guillen J, Su JH, Civeira MP, Prieto J. Transmission of hepatitis C virus infection to tree shrews. Virology. 1998;244:513-520.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Mercer DF, Schiller DE, Elliott JF, Douglas DN, Hao C, Rinfret A, Addison WR, Fischer KP, Churchill TA, Lakey JR. Hepatitis C virus replication in mice with chimeric human livers. Nat Med. 2001;7:927-933.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 693]  [Cited by in F6Publishing: 665]  [Article Influence: 28.9]  [Reference Citation Analysis (0)]
70.  MacArthur KL, Wu CH, Wu GY. Animal models for the study of hepatitis C virus infection and replication. World J Gastroenterol. 2012;18:2909-2913.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 10]  [Cited by in F6Publishing: 11]  [Article Influence: 0.9]  [Reference Citation Analysis (1)]
71.  Manns MP, Foster GR, Rockstroh JK, Zeuzem S, Zoulim F, Houghton M. The way forward in HCV treatment--finding the right path. Nat Rev Drug Discov. 2007;6:991-1000.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 236]  [Cited by in F6Publishing: 245]  [Article Influence: 14.4]  [Reference Citation Analysis (0)]
72.  Gastaminza P, Dryden KA, Boyd B, Wood MR, Law M, Yeager M, Chisari FV. Ultrastructural and biophysical characterization of hepatitis C virus particles produced in cell culture. J Virol. 2010;84:10999-11009.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 166]  [Cited by in F6Publishing: 166]  [Article Influence: 11.9]  [Reference Citation Analysis (0)]
73.  Catanese MT, Uryu K, Kopp M, Edwards TJ, Andrus L, Rice WJ, Silvestry M, Kuhn RJ, Rice CM. Ultrastructural analysis of hepatitis C virus particles. Proc Natl Acad Sci USA. 2013;110:9505-9510.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 208]  [Cited by in F6Publishing: 201]  [Article Influence: 18.3]  [Reference Citation Analysis (0)]
74.  Wang Y, Kato N, Hoshida Y, Yoshida H, Taniguchi H, Goto T, Moriyama M, Otsuka M, Shiina S, Shiratori Y. Interleukin-1beta gene polymorphisms associated with hepatocellular carcinoma in hepatitis C virus infection. Hepatology. 2003;37:65-71.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Lindenbach BD, Rice CM. The ins and outs of hepatitis C virus entry and assembly. Nat Rev Microbiol. 2013;11:688-700.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 275]  [Cited by in F6Publishing: 274]  [Article Influence: 24.9]  [Reference Citation Analysis (0)]
76.  Moriya K, Fujie H, Shintani Y, Yotsuyanagi H, Tsutsumi T, Ishibashi K, Matsuura Y, Kimura S, Miyamura T, Koike K. The core protein of hepatitis C virus induces hepatocellular carcinoma in transgenic mice. Nat Med. 1998;4:1065-1067.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 931]  [Cited by in F6Publishing: 895]  [Article Influence: 34.4]  [Reference Citation Analysis (0)]
77.  Wang Y, Kato N, Jazag A, Dharel N, Otsuka M, Taniguchi H, Kawabe T, Omata M. Hepatitis C virus core protein is a potent inhibitor of RNA silencing-based antiviral response. Gastroenterology. 2006;130:883-892.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 88]  [Cited by in F6Publishing: 88]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
78.  Nolandt O, Kern V, Müller H, Pfaff E, Theilmann L, Welker R, Kräusslich HG. Analysis of hepatitis C virus core protein interaction domains. J Gen Virol. 1997;78:1331-1340.  [PubMed]  [DOI]  [Cited in This Article: ]
79.  Boulant S, Vanbelle C, Ebel C, Penin F, Lavergne JP. Hepatitis C virus core protein is a dimeric alpha-helical protein exhibiting membrane protein features. J Virol. 2005;79:11353-11365.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 112]  [Cited by in F6Publishing: 121]  [Article Influence: 6.4]  [Reference Citation Analysis (0)]
80.  Lu WX, Cheng T, Li SW, Pan HR, Shen WT, Chen YX, Zhang T, Zheng Z, Zhang J, Xia NS. [Establishment and application of human papillomavirus type 16 pseudovirions neutralization assay]. Sheng Wu Gong Cheng Xue Bao. 2006;22:990-995.  [PubMed]  [DOI]  [Cited in This Article: ]
81.  Wu T, Li SW, Zhang J, Ng MH, Xia NS, Zhao Q. Hepatitis E vaccine development: a 14 year odyssey. Hum Vaccin Immunother. 2012;8:823-827.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 73]  [Cited by in F6Publishing: 76]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
82.  Op De Beeck A, Voisset C, Bartosch B, Ciczora Y, Cocquerel L, Keck Z, Foung S, Cosset FL, Dubuisson J. Characterization of functional hepatitis C virus envelope glycoproteins. J Virol. 2004;78:2994-3002.  [PubMed]  [DOI]  [Cited in This Article: ]
83.  Voisset C, Dubuisson J. Functional hepatitis C virus envelope glycoproteins. Biol Cell. 2004;96:413-420.  [PubMed]  [DOI]  [Cited in This Article: ]
84.  Modis Y, Ogata S, Clements D, Harrison SC. A ligand-binding pocket in the dengue virus envelope glycoprotein. Proc Natl Acad Sci USA. 2003;100:6986-6991.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 823]  [Cited by in F6Publishing: 768]  [Article Influence: 36.6]  [Reference Citation Analysis (0)]
85.  Li Y, Wang J, Kanai R, Modis Y. Crystal structure of glycoprotein E2 from bovine viral diarrhea virus. Proc Natl Acad Sci USA. 2013;110:6805-6810.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 90]  [Cited by in F6Publishing: 99]  [Article Influence: 9.0]  [Reference Citation Analysis (0)]
86.  OuYang B, Xie S, Berardi MJ, Zhao X, Dev J, Yu W, Sun B, Chou JJ. Unusual architecture of the p7 channel from hepatitis C virus. Nature. 2013;498:521-525.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 197]  [Cited by in F6Publishing: 207]  [Article Influence: 18.8]  [Reference Citation Analysis (0)]
87.  Luik P, Chew C, Aittoniemi J, Chang J, Wentworth P, Dwek RA, Biggin PC, Vénien-Bryan C, Zitzmann N. The 3-dimensional structure of a hepatitis C virus p7 ion channel by electron microscopy. Proc Natl Acad Sci USA. 2009;106:12712-12716.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 122]  [Cited by in F6Publishing: 128]  [Article Influence: 8.5]  [Reference Citation Analysis (0)]
88.  Takikawa S, Engle RE, Emerson SU, Purcell RH, St Claire M, Bukh J. Functional analyses of GB virus B p13 protein: development of a recombinant GB virus B hepatitis virus with a p7 protein. Proc Natl Acad Sci USA. 2006;103:3345-3350.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 24]  [Cited by in F6Publishing: 25]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
89.  Penin F, Dubuisson J, Rey FA, Moradpour D, Pawlotsky JM. Structural biology of hepatitis C virus. Hepatology. 2004;39:5-19.  [PubMed]  [DOI]  [Cited in This Article: ]
90.  Meylan E, Curran J, Hofmann K, Moradpour D, Binder M, Bartenschlager R, Tschopp J. Cardif is an adaptor protein in the RIG-I antiviral pathway and is targeted by hepatitis C virus. Nature. 2005;437:1167-1172.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1933]  [Cited by in F6Publishing: 1854]  [Article Influence: 97.6]  [Reference Citation Analysis (0)]
91.  Moradpour D, Penin F. Hepatitis C virus proteins: from structure to function. Curr Top Microbiol Immunol. 2013;369:113-142.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 66]  [Cited by in F6Publishing: 149]  [Article Influence: 13.5]  [Reference Citation Analysis (0)]
92.  Einav S, Sklan EH, Moon HM, Gehrig E, Liu P, Hao Y, Lowe AW, Glenn JS. The nucleotide binding motif of hepatitis C virus NS4B can mediate cellular transformation and tumor formation without Ha-ras co-transfection. Hepatology. 2008;47:827-835.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 39]  [Cited by in F6Publishing: 44]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
93.  Einav S, Gerber D, Bryson PD, Sklan EH, Elazar M, Maerkl SJ, Glenn JS, Quake SR. Discovery of a hepatitis C target and its pharmacological inhibitors by microfluidic affinity analysis. Nat Biotechnol. 2008;26:1019-1027.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 183]  [Cited by in F6Publishing: 192]  [Article Influence: 12.0]  [Reference Citation Analysis (0)]
94.  Thompson AA, Zou A, Yan J, Duggal R, Hao W, Molina D, Cronin CN, Wells PA. Biochemical characterization of recombinant hepatitis C virus nonstructural protein 4B: evidence for ATP/GTP hydrolysis and adenylate kinase activity. Biochemistry. 2009;48:906-916.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 41]  [Cited by in F6Publishing: 44]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
95.  Tanji Y, Kaneko T, Satoh S, Shimotohno K. Phosphorylation of hepatitis C virus-encoded nonstructural protein NS5A. J Virol. 1995;69:3980-3986.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Hirota M, Satoh S, Asabe S, Kohara M, Tsukiyama-Kohara K, Kato N, Hijikata M, Shimotohno K. Phosphorylation of nonstructural 5A protein of hepatitis C virus: HCV group-specific hyperphosphorylation. Virology. 1999;257:130-137.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  Appel N, Zayas M, Miller S, Krijnse-Locker J, Schaller T, Friebe P, Kallis S, Engel U, Bartenschlager R. Essential role of domain III of nonstructural protein 5A for hepatitis C virus infectious particle assembly. PLoS Pathog. 2008;4:e1000035.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 363]  [Cited by in F6Publishing: 366]  [Article Influence: 22.9]  [Reference Citation Analysis (0)]
98.  Moradpour D, Evans MJ, Gosert R, Yuan Z, Blum HE, Goff SP, Lindenbach BD, Rice CM. Insertion of green fluorescent protein into nonstructural protein 5A allows direct visualization of functional hepatitis C virus replication complexes. J Virol. 2004;78:7400-7409.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 202]  [Cited by in F6Publishing: 216]  [Article Influence: 10.8]  [Reference Citation Analysis (0)]
99.  Tellinghuisen TL, Foss KL, Treadaway J. Regulation of hepatitis C virion production via phosphorylation of the NS5A protein. PLoS Pathog. 2008;4:e1000032.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 313]  [Cited by in F6Publishing: 317]  [Article Influence: 19.8]  [Reference Citation Analysis (0)]
100.  Miyanari Y, Atsuzawa K, Usuda N, Watashi K, Hishiki T, Zayas M, Bartenschlager R, Wakita T, Hijikata M, Shimotohno K. The lipid droplet is an important organelle for hepatitis C virus production. Nat Cell Biol. 2007;9:1089-1097.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 937]  [Cited by in F6Publishing: 948]  [Article Influence: 55.8]  [Reference Citation Analysis (0)]
101.  Masaki T, Suzuki R, Murakami K, Aizaki H, Ishii K, Murayama A, Date T, Matsuura Y, Miyamura T, Wakita T. Interaction of hepatitis C virus nonstructural protein 5A with core protein is critical for the production of infectious virus particles. J Virol. 2008;82:7964-7976.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 280]  [Cited by in F6Publishing: 285]  [Article Influence: 17.8]  [Reference Citation Analysis (0)]
102.  Lesburg CA, Cable MB, Ferrari E, Hong Z, Mannarino AF, Weber PC. Crystal structure of the RNA-dependent RNA polymerase from hepatitis C virus reveals a fully encircled active site. Nat Struct Biol. 1999;6:937-943.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 597]  [Cited by in F6Publishing: 621]  [Article Influence: 24.8]  [Reference Citation Analysis (0)]
103.  Deore RR, Chern JW. NS5B RNA dependent RNA polymerase inhibitors: the promising approach to treat hepatitis C virus infections. Curr Med Chem. 2010;17:3806-3826.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 30]  [Cited by in F6Publishing: 34]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
104.  Behrens SE, Tomei L, De Francesco R. Identification and properties of the RNA-dependent RNA polymerase of hepatitis C virus. EMBO J. 1996;15:12-22.  [PubMed]  [DOI]  [Cited in This Article: ]
105.  Lohmann V, Körner F, Herian U, Bartenschlager R. Biochemical properties of hepatitis C virus NS5B RNA-dependent RNA polymerase and identification of amino acid sequence motifs essential for enzymatic activity. J Virol. 1997;71:8416-8428.  [PubMed]  [DOI]  [Cited in This Article: ]
106.  Knipe DM, Howley PM.  Field Virology. 5th ed. Philadelphia: Lippincott Williams and Wilkins Immunology 2007; .  [PubMed]  [DOI]  [Cited in This Article: ]
107.  Fraser CS, Doudna JA. Structural and mechanistic insights into hepatitis C viral translation initiation. Nat Rev Microbiol. 2007;5:29-38.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 166]  [Cited by in F6Publishing: 159]  [Article Influence: 8.8]  [Reference Citation Analysis (0)]
108.  Tsukiyama-Kohara K, Iizuka N, Kohara M, Nomoto A. Internal ribosome entry site within hepatitis C virus RNA. J Virol. 1992;66:1476-1483.  [PubMed]  [DOI]  [Cited in This Article: ]
109.  Wang C, Sarnow P, Siddiqui A. A conserved helical element is essential for internal initiation of translation of hepatitis C virus RNA. J Virol. 1994;68:7301-7307.  [PubMed]  [DOI]  [Cited in This Article: ]
110.  Kato J, Kato N, Moriyama M, Goto T, Taniguchi H, Shiratori Y, Omata M. Interferons specifically suppress the translation from the internal ribosome entry site of hepatitis C virus through a double-stranded RNA-activated protein kinase-independent pathway. J Infect Dis. 2002;186:155-163.  [PubMed]  [DOI]  [Cited in This Article: ]
111.  Jopling CL, Yi M, Lancaster AM, Lemon SM, Sarnow P. Modulation of hepatitis C virus RNA abundance by a liver-specific MicroRNA. Science. 2005;309:1577-1581.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 1993]  [Cited by in F6Publishing: 1929]  [Article Influence: 101.5]  [Reference Citation Analysis (0)]
112.  Song Y, Friebe P, Tzima E, Jünemann C, Bartenschlager R, Niepmann M. The hepatitis C virus RNA 3’-untranslated region strongly enhances translation directed by the internal ribosome entry site. J Virol. 2006;80:11579-11588.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 110]  [Cited by in F6Publishing: 113]  [Article Influence: 6.3]  [Reference Citation Analysis (0)]
113.  Sainz B, Chisari FV. Production of infectious hepatitis C virus by well-differentiated, growth-arrested human hepatoma-derived cells. J Virol. 2006;80:10253-10257.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 80]  [Cited by in F6Publishing: 89]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
114.  Cai Z, Zhang C, Chang KS, Jiang J, Ahn BC, Wakita T, Liang TJ, Luo G. Robust production of infectious hepatitis C virus (HCV) from stably HCV cDNA-transfected human hepatoma cells. J Virol. 2005;79:13963-13973.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 125]  [Cited by in F6Publishing: 132]  [Article Influence: 7.3]  [Reference Citation Analysis (0)]
115.  Blight KJ, McKeating JA, Rice CM. Highly permissive cell lines for subgenomic and genomic hepatitis C virus RNA replication. J Virol. 2002;76:13001-13014.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 980]  [Cited by in F6Publishing: 995]  [Article Influence: 45.2]  [Reference Citation Analysis (0)]
116.  Lu J, Zhou Y, Lin X, Jiang Y, Tian R, Zhang Y, Wu J, Zhang F, Zhang Y, Wang Y. General epidemiological parameters of viral hepatitis A, B, C, and E in six regions of China: a cross-sectional study in 2007. PLoS One. 2009;4:e8467.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 89]  [Article Influence: 5.9]  [Reference Citation Analysis (0)]
117.  Tao YL, Tang YF, Qiu JP, Cai XF, Shen XT, Wang YX, Zhao XT. Prevalence of hepatitis C infection among intravenous drug users in Shanghai. World J Gastroenterol. 2013;19:5320-5325.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in CrossRef: 7]  [Cited by in F6Publishing: 8]  [Article Influence: 0.7]  [Reference Citation Analysis (0)]
118.  Cohen DE, Russell CJ, Golub SA, Mayer KH. Prevalence of hepatitis C virus infection among men who have sex with men at a Boston community health center and its association with markers of high-risk behavior. AIDS Patient Care STDS. 2006;20:557-564.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 26]  [Article Influence: 1.4]  [Reference Citation Analysis (0)]
119.  Hajarizadeh B, Grebely J, Dore GJ. Epidemiology and natural history of HCV infection. Nat Rev Gastroenterol Hepatol. 2013;10:553-562.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 704]  [Cited by in F6Publishing: 707]  [Article Influence: 64.3]  [Reference Citation Analysis (0)]
120.  Imamura M, Chayama K. [HCV-HBV infection]. Nihon Rinsho. 2006;64:1310-1313.  [PubMed]  [DOI]  [Cited in This Article: ]
121.  Leone N, Rizzetto M. Natural history of hepatitis C virus infection: from chronic hepatitis to cirrhosis, to hepatocellular carcinoma. Minerva Gastroenterol Dietol. 2005;51:31-46.  [PubMed]  [DOI]  [Cited in This Article: ]
122.  Morgan RL, Baack B, Smith BD, Yartel A, Pitasi M, Falck-Ytter Y. Eradication of hepatitis C virus infection and the development of hepatocellular carcinoma: a meta-analysis of observational studies. Ann Intern Med. 2013;158:329-337.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 631]  [Cited by in F6Publishing: 610]  [Article Influence: 55.5]  [Reference Citation Analysis (0)]
123.  El Khoury AC, Klimack WK, Wallace C, Razavi H. Economic burden of hepatitis C-associated diseases in the United States. J Viral Hepat. 2012;19:153-160.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 57]  [Cited by in F6Publishing: 55]  [Article Influence: 4.6]  [Reference Citation Analysis (0)]