Skip to main content
Erschienen in: Virology Journal 1/2016

Open Access 01.12.2016 | Research

A robust model of natural hepatitis C infection using hepatocyte-like cells derived from human induced pluripotent stem cells as a long-term host

verfasst von: Khanit Sa-ngiamsuntorn, Adisak Wongkajornsilp, Phetcharat Phanthong, Suparerk Borwornpinyo, Narisorn Kitiyanant, Wasun Chantratita, Suradej Hongeng

Erschienen in: Virology Journal | Ausgabe 1/2016

insite
INHALT
download
DOWNLOAD
print
DRUCKEN
insite
SUCHEN

Abstract

Background

Hepatitis C virus (HCV) could induce chronic liver diseases and hepatocellular carcinoma in human. The use of primary human hepatocyte as a viral host is restrained with the scarcity of tissue supply. A culture model restricted to HCV genotype 2a (JFH-1) has been established using Huh7-derived hepatocyte. Other genotypes including the wild-type virus could not propagate in Huh7, Huh7.5 and Huh7.5.1 cells.

Methods

Functional hepatocyte-like cells (HLCs) were developed from normal human iPS cells as a host for HCV infection. Mature HLCs were identified for selective hepatocyte markers, CYP450s, HCV associated receptors and HCV essential host factors. HLCs were either transfected with JFH-1 HCV RNA or infected with HCV particles derived from patient serum. The enhancing effect of α-tocopherol and the inhibitory effects of INF-α, ribavirin and sofosbuvir to HCV infection were studied. The HCV viral load and HCV RNA were assayed for the infection efficiency.

Results

The fully-developed HLCs expressed phase I, II, and III drug-metabolizing enzymes, HCV associated receptors (claudin-1, occludin, CD81, ApoE, ApoB, LDL-R) and HCV essential host factors (miR-122 and SEC14L2) comparable to the primary human hepatocyte. SEC14L2, an α-tocopherol transfer protein, was expressed in HLCs, but not in Huh7 cell, had been implicated in effective HCVser infection. The HLCs permitted not only the replication of HCV RNA, but also the production of HCV particles (HCVcc) released to the culture media. HLCs drove higher propagation of HCVcc derived from JFH-1 than did the classical host Huh7 cells. HLCs infected with either JFH-1 or wild-type HCV expressed HCV core antigen, NS5A, NS5B, NS3 and HCV negative-stand RNA. HLCs allowed entire HCV life cycle derived from either JFH-1, HCVcc or wild-type HCV (genotype 1a, 1b, 3a, 3b, 6f and 6n). Further increasing the HCVser infection in HLCs was achieved by incubating cell with α-tocopherol. The supernatant from infected HLCs could infect both naïve HLC and Huh7 cell. Treating infected HLC with INF-α and ribavirin decreased HCV RNA in both the cellular fraction and the culture medium. The HLCs reacted to HCVcc or wild-type HCV infection by upregulating TNF-α, IL-28B and IL-29.

Conclusions

This robust cell culture model for serum-derived HCV using HLCs as host cells provides a remarkable system for investigating HCV life cycle, HCV-associated hepatocellular carcinoma development and the screening for new anti HCV drugs.
Hinweise

Electronic supplementary material

The online version of this article (doi:10.​1186/​s12985-016-0519-1) contains supplementary material, which is available to authorized users.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

KS and PP performed most of the experiments, designed the protocol, performed the statistical analysis and drafted the manuscript. AW, SB, and SH designed the protocol. NK participated in the teratoma assay in nude mice. PP prepared the lentiviral vectors and iPS cell reprogramming. WC collected HCV from patient serum and provided HCV viral load detection using Abbott RealTime HCV assay. AW and SH coordinated the study. All authors have read and approved the final manuscript.
Abkürzungen
HCV
Hepatitis C virus
JFH1
Japanese Fulminant Hepatitis 1
HCVcc
Hepatitis C Virus derived from cell culture
HCVser
Hepatitis C Virus derived from patient serum
HLCs
Hepatocyte-Like Cells
hMSC
Human Mesenchymal Stem Cell
iPSC
Induced Pluripotent Stem Cell
PCR
Polymerase Chain Reaction
CYP450
Cytochrome P450
INF
Interferon
RLU
Relative Light Unit

Background

Human primary hepatocytes should have served as ideal host cells for liver-targeting pathogens, e.g., malarial parasite, hepatitis B, hepatitis C and dengue viruses. The shortcoming in the procurement of these cells and their limited life span have limited the study of host-pathogen relationships and their sequential developments, e.g., xenobiotic screening, and biotransformation. A cellular substitute that closely mimics primary hepatocyte but circumvents these shortcoming would greatly extend our understanding of host-pathogen interactions and the corresponding treatment strategies.
Hepatitis C virus (HCV) [1] is a single positive-stranded RNA virus in the genus of Hepacivirus and Flaviviridae family [2]. Chronic HCV infection led to cirrhosis and hepatocellular carcinoma [3]. The self-renewal capability of liver cell was disrupted that eventually required liver transplantation or bio-artificial liver device [4]. Liver transplantation were faced with serious and faster HCV reinfection to the graft [5, 6]. An alternative for liver transplant was the hepatocyte transplant that might alleviate the demand of donor organs [7]. Direct-acting antivirals (DAA) targeting HCV enzymes was hampered with eventual drug resistance [810]. The development of suitable culture models for HCV is critical for designing efficacious anti-HCV strategies.
The studies on HCV life cycle relied heavily on human hepatocellular carcinoma cells (Huh7 and their derivatives) [11]. HCV genotype 2a (JFH-1), but not others, could be generated from Huh7 derived cells [12, 13]. The use of hepatocellular carcinoma as cellular host could not completely mimic primary human hepatocyte. The cancer cells actively entered cellular division while the primary hepatocytes were mostly in quiescent stage [14]. Most hepatoma cell lines usually lack various functional enzymes such as CYP450s and other phase I, II and III drug metabolizing enzymes that make them not suitable for the assessment of anti-HCV drug interaction and metabolism [15, 16]. Huh7.5.1 was derived from Huh7.5 [17], which in turn was originated from Huh7 [18]. These cells carried a mutation in the retinoic-inducible gene I (RIG-I) [19]. RIG-I played a central role in viral genome recognition and host immune response. Primary human hepatocytes have been indorsed by several groups as the major host cells for HCV [2022]. However, the handling primary human hepatocytes faced several limitations: 1) The mature hepatocytes could not be readily proliferated in culture condition; 2) The donor supply was limited; and 3) The batch to batch variation was substantial [23].
Human induced pluripotent stem (iPS) cells could be generated from somatic cell through exogenous expression of Oct4, Sox2, KLF4 and c-MYC [24, 25]. Human iPS cells actively entered cellular division and could be differentiated into hepatocyte-like cells (HLCs) [26] and others. The use of HLCs derived from either iPS or embryonic stem cells as cellular hosts for HCV were recently reported [2730]. These differentiated cells displayed essential liver functions and achieved nearly mature hepatocytes [31], including α-fetoprotein, albumin, phase I and phase II drug metabolizing enzymes. HLCs also expressed known HCV host receptors involved in HCV entry (Claudin-1, Occludin, SR-BI, CD81) and supported complete life cycle of HCV genotype 2a up to 21 days [30]. Nevertheless, the maturity, homogeneity and long-term stability of these HLCs have not been revealed. The HLCs not only carried less CYP450 expression than that of primary hepatocytes, but retained several fetal hepatocyte markers.
In this studies, human iPS cells generated from hMSCs using polycistronic OSKM reprograming factors [32] were differentiated into HLCs. The CYP expressions were readily inducible upon the exposure to the prototypic inducers in a similar fashion to the primary hepatocytes. Since the HCV cell receptors and cell host factors were highly expressed, these HLCs were promptly taken as HCV hosts. The infected cells were studied for viral life cycle after the transfection/infection with HCVcc and HCVser. HLCs could sustain the replication of not only JFH-1 HCV but several wild-type HCV’s from patients’ sera. High HCV titers were detected in culture medium for at least 6 months. The released HCV virions could infect naïve HLCs and were susceptible to the treatment with interferon, ribavirin or sofosbuvir.

Methods

Generation of human induced pluripotent stem cell (hiPSC) from hMSC

The mononuclear cells were separated from bone marrow using IsoPrep (Robbins Scientific, Canada) density gradient centrifugation. Isolated cells were seeded at 2 × 106 cells on T-75 cell culture flask in Minimum Essential Medium (MEM) α (Gibco Invitrogen, NY), 10 % FBS (HyClone, GE Healthcare Life Sciences, Singapore), 100 units/mL penicillin, 100 μg/mL streptomycin at 37 °C in 5 % CO2. Lentiviral particles were produced by co-transfection of packaging psPAX2 packaging plasmid, pMD2.G vesicular stomatitis virus G envelope, and the polycistronic plasmid encoding human OCT4, SOX2, KLF4, c-MYC and dTomato [32] using X-tremeGENE HP DNA Transfection Reagent (Roche Diagnostics, USA). hMSCs were transduced with concentrated lentivirus (MOI = 0.5). Production of dTomato was observed 48 h post-transduction. The transduced cells were seeded on MEF in hES medium to promote reprogramming process and colony formation. Reprogrammed colonies were identified with StainAlive TRA-1-60 Antibody (Stemgent, MA). Positive colonies were manually picked and passaged until stable iPS cells established.

Maintaining and characterization of hiPSC

Human iPS cells were maintained in a feeder free condition on Geltrex-coated dish in E8 medium at 37 °C in 5 % CO2. The E8 medium consisted of DMEM/F12 supplemented with 64 mg/L L-ascorbic acid-2-phosphate, 14 μg/L sodium selenite, 100 μg/L FGF2, 19.4 mg/L insulin, 543 mg/L NaHCO3, 10.7 mg/L transferrin and 2 μg/L TGF-β1. The medium was changed daily. Human iPS cell colonies were split at a ratio of 1:8 every 5 - 7 d and passaged onto a new Geltrex-coated dish. The iPSC colonies after the tenth passage were characterized for pluripotent markers, e.g., alkaline phosphatase, OCT4, SOX2, NANOG, TRA-1-60, TRA-1-81, and SSEA4 using alkaline phosphatase ES characterization kit (MERCK Millipore, USA), fluorescent antibody staining and quantitative RT-PCR analysis.

Teratoma formation

Approximately 2 × 106 iPS cells were subcutaneously injected into nude mice (BALB/cMlac-nu). The tumor-like tissue was collected within 8–12 weeks after injection. The tissue was fixed in 10 % formalin and underwent histological processes. The tissue sections were stained with hematoxylin and eosin and examined under a light microscope for the presence of the specialized cells derived from three germ layers. All experiments performed on laboratory animals were reviewed and approved by the Animal Care and Use Committee, Institute of Molecular Biosciences, Mahidol University.

Differentiation of hiPSC to homogeneous hepatocyte-like cell

The characterized iPS cells were taken for hepatic induction. They were seeded on Geltrex-coated 6 well plate in E8 medium until 80 % confluence and differentiated using a modified hepatic lineage development protocol [31]. Briefly, the cultured cells were maintained in endoderm differentiation medium DMEM/F12 : IMDM (1:1) (Gibco Invitrogen, NY), 100 ng/mL activin A (PeproTech, US), 10 ng/mL bFGF, 10 ng/mL BMP-4, 1 μM LY294002 and 3 μM CHIR99021 for 24 h. Cells were maintained in endoderm commitment medium DMEM/F12 : IMDM (1:1), 100 ng/mL activin A, 100 ng/mL bFGF, 10 ng/mL BMP-4 and 1 μM LY294002 for 24 h. The endoderm-like cells were further maintained in anterior definitive endoderm differentiation medium RPMI/B27 with 50 ng/mL activin A for 3 d with a daily medium replacement. The differentiated cells were maintained in hepatic lineage specification medium RPMI/B27, 20 ng/mL BMP-4 and 10 ng/mL FGF10 for 4 d with daily medium replacement. The differentiated cells were cultured in hepatocyte basal medium (HBM, Lonza, UK), 30 ng/mL oncostatin M and 50 ng/mL HGF for 10 d with medium replacement every 2 d. For functional hepatocyte induction, the cells were cultured in Williams' Media E (Invitrogen, MD, 1 % DMSO for 3 d.

Periodic acid-Schiff staining of glycogen

HLCs and hMSCs were cultured on 35-mm dish for 3d. The cells were fixed in 5 % formaldehyde, 95 % ethanol for 1 min, rinsed for 1 min under running tap water, immersed in periodic acid solution for 5 min, rinsed thrice with dH2O, immersed in Schiff’s reagent for 15 min, and rinsed with running tap water for 5–10 min. Samples were counterstained with Mayer’s hematoxylin for 1 min, rinsed with water, and assessed under light microscope.

Hepatocyte-specific markers and cytochrome P450 expressions

The total RNA was extracted from iPS cells, HLCs, or HepaRG (HPRGC10, Life Technologies) by illustra RNAspin Mini RNA Isolation Kit (GE Healthcare, Asia Pacific) following the manufacturer instruction. The total RNA was immediately used for reverse transcription to construct cDNA using the ImProm-II reverse transcription system (Promega, WI) following the manufacturer instruction. The hepatocyte markers and cytochrome P450 markers included albumin, α-fetoprotein, cytokeratin18, G-6-PD, HNF-4α, tyrosine aminotransferase and major 7 CYPs isozymes. All gene specific primers (Additional file 1) were designed using vector NTI version 11.5 (Invitrogen, MD). They were amplified using KAPA SYBR® FAST qPCR Kits (Kapa Biosystems). Each sample was measured in triplicate. PCR amplicons were confirmed using size and melting curve analysis. Expression levels were calculated using the ΔΔCt method and normalized to the endogenous GAPDH. ΔΔCt was transformed into fold change using the formula: fold change = 2- ΔΔCt.

Immunofluorescent staining

HLCs on 24-well plate were fixed with 4 % paraformaldehyde for 20 min at room temperature. The fixed cells were washed thrice with PBS, blocked and permeabilized with 0.2 % Triton X-100, 3 % BSA and 2 % normal goat serum for 1 h. Primary antibodies against pluripotent markers were TRA-1-60 (1:250, SC21705, Santa Cruz Biotechnology), SSEA4 (1:250, ab16287, Abcam), NANOG (1:250, SC33759, Santa Cruz Biotechnology), OCT4 (1:250, SC5279, Santa Cruz Biotechnology). Primary antibodies against HCV cell receptors were claudin-1 (1:100, ab15098), occluding (1:100, ab31721), CD81 (1:100, ab79559), SR-BI (1:200, ab106572) and hepatitis C core antigen (1:200, ab2740) ordered from Abcam. Primary antibodies against non-structural proteins of HCV were NS3 (1:200, SC69938), NS5A (1:200, SC52417) and NS5B (1:200, SC58146) ordered from Santa Cruz Biotechnology. The cells were incubated with the primary antibody for overnight h at 4 °C in moist chamber. After washing thrice, samples were incubated with the secondary antibody; Alexa fluor 488 conjugated goat anti-mouse IgG (1:1000, A11001) or Alexa fluor 488 conjugated goat anti-rabbit IgG (1: 1000, A11003s) or Alexa fluor 488 conjugated rabbit anti-goat IgG (1: 1000, A1178) (Life Technologies) for 1 h at 37 °C. Fibroblasts or untreated hMSCs served as negative controls. After washing thrice, the cells were counterstained with DAPI and mounted with anti-fade mounting medium in coverslip and examined under a fluorescent microscope and photographed.

The induction of major CYP450 isozymes in HepaRG and HLCs using the enzyme inducer cocktail

The expression of CYP450 levels of hepatocyte-like cells responded to enzyme inducers was evaluated after incubate with cocktail of classical drugs [33]. HepaRG and hepatocyte-like cell derived from iPS cells were seeded on 6 well-plates for 3 d. These cells were maintained in Williams’ Media E, 10 % FBS before exposure to the inducers. Both of hepatocyte cells were incubated for 72 h with the cocktail agent: 20 μM rifampicin, 50 μM omeprazole, 1 mM phenobarbital and 88 μM ethanol. The treated cells were washed with 2–3 mL PBS, detached using 0.025 % trypsin-EDTA, and neutralized with 10 % FBS in DMEM/F12. The cell pellets from hepatocyte-like cell and HepaRG were stored at -80 °C until analysis for CYP450 gene expression.

The detection of CYP450 activities

CYP1A1, CYP2B6, CYP2C9 and CYP3A4 enzyme activities were evaluated directly in all cultured cells (HepaRG, HLCs and primary human hepatocytes) attaching to the collagen type IV-coated 6-cm dish at a density of 106 cell. All cultured cells were treated with a cocktail of 20 μM rifampicin, 50 μM omeprazole, 1 mM phenobarbital and 88 μM ethanol. All treated cells were incubated for 72 h with daily medium change. CYP450 activities were assessed based on luciferase activity using the P450-glo 1A1, 2B6, 2C9 and 3A4 assays (V8751, V8321, V8791, V9001; Promega, WI). After 3-d incubation period, cells were incubated with Williams’ Media E supplemented with 100 μM Luciferin-CEE, Luciferin-2B6, Luciferin-H for 3–4 h or 3 μM Luciferin-IPA for 30–60 min. An aliquot (50 μL) of the medium was transferred to 96-well opaque white luminometer plate (Nunc, Denmark) and luciferin detection reagent was added to each well. After sitting at room temperature for 20 min, luminescence was measured with a SpectraMax M5 spectrofluorometer.

The detection of cellular hepatitis C receptors on HLCs

HCV infection requires several types of cellular receptors to permit viral entry [34]. Cellular receptors essential for HCV infection such as Claudin-1, Occludin, SR-BI, CD81 were detected using quantitative RT-PCR and immunofluorescent technique. Mature HLCs were seeded on 24-well plate and maintained until 80 % confluence in Williams’ Media E, 10 % FBS. Cells were fixed and stained with fluorescent-conjugated antibodies raised against Claudin-1, Occludin, SR-BI or CD81.

Production of HCV from JFH-1 (HCVcc)

Cell culture based hepatitis C virus (HCVcc genotype 2a) was prepared from JFH-1 system [35]. The JFH-1 plasmid was propagated in E Coli and extracted using NucleoBond Xtra Midi plasmid (MN, Germany). The plasmids were purified and linearized by a restriction enzyme XbaI, and used as a template for JFH-1 RNA synthesis. The full length HCV RNA was synthesized by TranscriptAid T7 High Yield Transcription Kit (Thermo Fisher Scientific, MA) following the manufacturer’s instruction. HCVcc was produced by transient transfection of JFH-1 RNA into Huh7 cell using either lipofectamine 3000 (Invitrogen, MA) or electroporation (Gene Pulser Xcell Electroporation Systems, Bio-Rad, CA). HCV in Huh7 supernatant was passed through 0.45 μm syringe filter, concentrated by sucrose gradient ultracentrifugation and stored at −80 °C for future use.

The infection of HCVser and HCVcc to HLCs

HCV-positive sera were collected from patients with chronic HCV infection at Ramathibodi Hospital, Mahidol University. The collection of leftover blood specimen was approved by the Ethics Committee on Research Involving Human Subjects (2556/250). All subjects are adults and provided written informed consent. Serum from individual patient with HCV load > 106 IU/mL was selected to infect HLCs. These sera carried HCV genotypes 1a, 1b, 3a, 3b, 6f and 6n that were prevalent in Thailand. For HCVser and HCVcc infection, HLCs and Huh7 cells were seeded on 6-well plate in Williams' Media E, 10 % FBS until 80 % confluence. Some host cells were pretreated with 1 μM or 5 μM α-tocopherol for 24 h prior to HCVser infection. For HCVser infection, 50 μL of infected serum in 2 mL of medium was added to host cells. For HCVcc infection, concentrated JFH-1 HCV load at 107 IU/mL in 2 mL of medium was added to host cells. After 24 h incubation, the infected cells were washed 6–10 times with 0.1 % BSA in PBS and reconstituted with 2 mL complete growth medium.

The long-term maintenance of HCV infection

For long-term HCV infection assay, HLCs and Huh7 were infected with HCVcc or HCVser. After 24 h incubation, the culture medium was renewed with 10 % FBS, DMEM/F12. The HCV-containing supernatant was collected ever 4–6 days and HCV RNA was detected using RT-PCR. During long-term cultivation process, HLCs and Huh7 cell were maintained in complete medium with 1 % concentrated lipids (Life Technologies, USA) and 1 μM (α-tocopherol). The HCV production was confirmed using the Abbott real-time HCV assay (Abbott Diagnostics, Illinois, USA).

The detection of negative and positive stands of HCV RNA

HCV RNA in the supernatants was extracted with NucleoSpin RNA Virus isolation kit (MN, USA); while the intracellular RNA was extracted with illustra RNAspin Mini RNA Isolation Kit (GE Healthcare, Asia Pacific). The stand specific primers were designed using Vector NTI version 11.5 (Invitrogen, USA). The primers for positive strand are: 5′-CCCTGTGAGGAACTACTGTCTTCACGCA-3′ and 5′-CTCGCAAGCACCCTATCAGG-CAGTAC-3′. The primers for the negative strand are: 5′-GATGTACCCCATGAGGTCGG-3′ and 5′-GCGCGACAAG-GAAGACTTCG-3′ [36]. The viral RNA and total RNA from infected host cells were converted to cDNA with the ImProm-II reverse transcription system (Promega, WI). For HCV (-) stand, one microgram of isolated RNA was incubated with 12.5 μM HCV ‘reverse’ primers in a total volume of 5 μL at 70 °C for 5 min and chilled on ice-water immediately for at least 5 min. The reverse transcription mix (15 μL of 5× reaction buffer, 25 mM MgCl2, 2 mM dNTP Mix, 40 U/μL RNasin ribonuclease inhibitor, and 200 U/μL Improm-II™ RT) was added to the RNA-primer mix to make a total volume of 20 μL. The mixture was incubated at 25 °C for 5 min, and 42 °C for further 1 h. The RT reaction was terminated by heating at 70 °C for 15 min followed by chilling on ice. The HCV cDNA was amplified by KAPA SYBR® FAST qPCR Kits (Kapa Biosystems, UK) and 2.5 μM HCV ‘forward’ primer at 95 °C for 30 s, 66 °C for 45 s and 70 °C for 90 s. Hepatitis C virus (+) stand was quantified by real-time RT-PCR using the m2000sp and m2000rt instruments and RealTime HCV kit (Abbott Diagnostics, Illinois, USA).

Statistical analysis

All results of experiments were performed in triplicated. Data were expressed as means ± SD. Data from quantitative real-time RT-PCR were evaluated for statistical significance using either students’ t test or ANOVA for difference followed by post-hoc tests (Tukey’s HSD) between the treatments and their respective controls. * and ** represented statistically different data with p < 0.05 and p < 0.01 respectively.

Results

Reprogramming human bone marrow mesenchymal stem cell to iPS cells

hMSC isolated from aspirated bone marrow exhibited spindle-shaped morphology at 90–95 % confluence (Fig. 1a). After infecting hMSCs with polycistronic OSKM-dTomato lentivirus particles for 48 h, the transduced hMSCs expressing OCT4, SOX2, KLF4 and C-MYC were determined by positive dTomato (Fig. 1b and c). The dTomato-positive cells were detached and seeded on mitomycin c-treated MEF feeder at 2 × 105 cells/10-cm dish. Cell morphology of dTomato positive population transformed from spindle-shape to epithelial morphology around day 5 post-transduction (Fig. 1d). These cell populations divided and formed a small cluster of dTomato after 2-3 d (Fig. 1e). The transduced colonies expanded and transformed to a tightly packed colony with a high nucleocytoplasmic ratio (Fig. 1f). Upon reaching full reprogramming, dTomato expression was silenced and reprogramming colonies were negative for dTomato on day 13 as monitored under a fluorescence microscope (Fig. 1g). The transduced cells reached iPS as determined by their positive pluripotent surface marker TRA-1-60 using a live staining technique (Fig. 1h and i). The TRA-1-60 positive colonies were selected and picked into a new MEF-seeded plates for expansion and characterizations.

Identification of pluripotent markers in reprogramming cells and iPS cell characterizations

TRA-1-60-positive colonies were passaged and maintained on MEF in ES medium until reprogramming cells achieved a typical human ESC morphology. The cell morphology consisted of a sharp border, flat, dense colony, high nucleocytoplasmic ratio and prominent nucleoli (Fig. 1h). The karyotype analysis confirmed no chromosomal alteration (Fig. 1j). The establishment of human iPS cell was confirmed using alkaline phosphatase (AP) activity, immunofluorescent staining of pluripotent markers and the expression of pluripotent genes (Fig. 1k). The reprogramming cells were maintained in E8 medium with feeder-free condition on Geltrex-coated dishes (Invitrogen, USA) prior to the analysis (Fig. 1l). The colony of reprogramming cells highly expressed essential pluripotent marker proteins, OCT4 and SOX2, required for maintaining pluripotent status. The expressions of OCT4 and SOX2 targeted proteins such as the self-renewal transcription factor NANOG, transmembrane proteins TRA-1-60 and TRA-1-81, and a cell surface glycoprotein SSEA4 were also expressed in the reprogramming cells. Human fibroblasts were used as negative controls for non-iPS cells (Fig. 2a). Exogenous OCT4 and SOX2 produced from the lentivirus were active at the early stage of cell reprogramming. To confirm the maintenance of endogenous OCT4 and SOX2 expression, specific OCT4 and SOX2 primers were designed to detect only the mRNA from endogenous source. The RT-PCR analysis confirmed the transduced cells carried endogenous OCT4 and SOX2 (Fig. 2b). Other pluripotent markers such as REX1, NANOG, GDF3, DNM3TB and UFT1 were expressed in the transduced cells and compared with those in human ES cell (BG01V) and non-transduced hMSCs (Fig. 2b). GAPDH and NTC served as positive and negative controls for PCR products.

Spontaneous differentiation and teratoma formation

To determine whether these newly established iPS cells could differentiate into all three germ layers, the iPS cells were evaluated for ex vivo and in vivo differentiation. For spontaneous ex vivo differentiation, iPS cells were maintained in suspension culture with embryoid bodies formation within a week. After 2 weeks, embryoid bodies were seeded on gelatin-coated culture plates to allow cell adhesion and migration. The adherent cells displayed various morphologies (Fig. 3a). Immunofluorescent staining revealed differentiated cell positive for βIII-tubulin (ectoderm marker), vimentin (mesoderm marker) and α-fetoprotein (AFP, endoderm marker) (Fig. 3a). These indicated the differentiation to several lineages. For teratoma formation, one million iPS cells were resuspended in DMEM 10 % FBS and injected subcutaneously into nude mice. Four weeks after injection, teratoma was formed at the injection site. Histological examinations revealed that the teratoma contained several tissues (Fig. 3b), e.g., neural rosette (ectoderm) (Fig. 3b, left panel), mesoderm derived smooth muscle, adipose tissue, cartilage (mesoderm) (Fig. 3b, middle) and endoderm-derived gut-like epithelial (endoderm) (Fig. 3b, right). Taken together, iPS cells derived from hMSCs not only support the spontaneous ex vivo differentiation but they can form teratoma in vivo.

The hepatocyte-like cell derived from iPS cell expressed hepatic phenotypes

Human iPS cells were cultured in feeder free condition until reaching 70 % confluence before hepatic induction (Fig. 4a). During hepatic differentiation, iPS cells were monitored daily for morphological changes. Initially, iPS cells were differentiated into definitive endoderm (DE) cells (Fig. 4b) that shared the same precursor with all endodermal organs (pancreas, liver, lung, thyroid and gut) [37]. The differentiated cell transformed into round-shaped morphology (Fig. 4b) with positive staining for SOX17, an endoderm differentiation marker (Fig. 4c). The second stage, definitive endoderm cells were developed to anterior definitive endoderm (ADE) with cell morphology underwent epithelial to mesenchymal transition (EMT) [38]. Cells displayed the mesenchymal-like morphology (Fig. 4d). At the third stage of differentiation, ADE cells became hepatic progenitors (Fig. 4e). The differentiated cell at this stage was confirmed with the expression of alpha-fetoprotein (AFP) (Fig. 4f), a characteristic of hepatoblast [31]. At maturation, hepatoblasts developed into functional hepatocyte-like cells (HLCs). HLCs shared both fetal and adult hepatocyte markers, e.g., albumin and minimal level of alpha-fetoprotein (Fig. 4g). The expression of HNF4-α, found only at the end of maturation, revealed that the cells fully developed into mature hepatocytes (Fig. 4h). After 4-week differentiation, human iPS cell developed into a near homogenous population with more than 95 % pure hepatocyte-like cell. HLCs exhibited typical hepatocyte morphology, e.g., a polygonal shape, a dark cytoplasm with large nuclei containing nucleoli and bi-nucleated (Fig. 4i). Mature HLCs maintained in Williams' Media E exhibited homogeneous population (Fig. 4j), hepatic triad-like structure (Fig. 4k) and some populations produced bile pigment (Fig. 4l). HLCs exhibited polygonal morphology during passages (Fig. 4m). HLCs showed the capability to synthesize and store glycogen greater than did hMSCs as detected by Periodic acid Schiff staining (Fig. 4n and o). The HLCs were also positive for gene expressions of α-fetoprotein, albumin and HNF4-α using quantitative RT-PCR (Fig. 4p). For major metabolic functions in hepatocyte involving carbohydrate, lipid and amino acid metabolism, HLCs expressed both glucose-6-phosphate dehydrogenase (G-6-PD) and tyrosine aminotransferase (TAT). The expression levels of basic hepatocyte markers (i.e., ALB, AFP, HNF4-α, and TAT) were increased between 10–20 folds over those from undifferentiated iPS cells.

The elevated expression levels of most basal CYP450 isozymes in mature HLCs were inducible with prototypic inducers

To achieve fully functional hepatocyte, mature HLCs were maintained in 10 % FBS Williams’ Media E for 3 d and the cells were screened for CYP expression. HLCs expressed a number of CYP isozymes as illustrated by immunostaining and quantitative real-time PCR. The basal mRNA levels of CYP2B6, CYP2D6, and CYP2C9 in HLCs were 22-, 25-, and 20-fold that of HepaRG cells respectively; while CYP3A4, CYP2C19, CYP1A2 were around 10 folds. CYP2E1 in HLCs was up-regulated to 40-fold that of the undifferentiated iPS cells (Fig. 5a). Other isozymes achieved comparable expression levels to those of HepaRG and the primary human hepatocytes. To evaluate the inducibility of these CYP isozymes, the HLCs were incubated with the cocktail of chemical inducers (20 μM rifampicin, 50 μM omeprazole, 1 mM phenobarbital and 88 μM ethanol). The expressions of CYP2B6, CYP2C9, CYP2C19 and CYP3A4 in HLCs were increased to 2, 2, 4, 3.5 folds respectively after the induction with the cocktail. In comparison, the expressions of CYP2B6, CYP2C9 and CYP3A4 in HepaRG were increased to 2, 2, 3.5 folds and comparable to those in HLCs (Fig. 5b). In particular, some CYP isozymes, e.g. CYP2B6, CYP2C9, CYP2C19 and CYP2E1 were more inducible in HLCs than in HepaRG cells. The expression levels of CYPs in primary human hepatocytes were dramatically decreased post-induction by 72 h. In addition to CYP expressions, HLCs contained high levels of organic anion transporting polypeptide 2 (OATP2). The elevated OATP2 expression could be extensively up regulated to 3 folds. For protein determination, more than 50 % of HLC populations were positive for intracellular CYP1A1, CYP3A4, CYP2B6 and CYP2E1 proteins (Fig. 5cg). The immunofluorescent straining demonstrated the increasing expression of major CYP proteins (Fig. 5cg), e.g., CYP3A4, CYP2B6, CYP1A1, CYP2C9 and CYP2E1. HLCs expressed multidrug resistance-associated protein 2 (MRP2) that represented the canalicular (apical) membrane of the hepatocyte as a biliary transport (Fig. 5h). The hMSC served as a control of non-hepatic cell (Fig. 5i). The increasing of CYP450 activities was investigated in hepatocytes (Fig. 5j). Mature HLCs population could proliferate, but still maintained hepatic morphology without losing basal levels of hepatic related gene expression.

HLCs expressed all major hepatitis C virus cell associated receptors

A potential application for HLCs is for the study of pathogen-host cell interactions. The HLCs derived from iPS cells were evaluated for the capability to host HCV replication. The expression of HCV receptors is essential for virus hepatitis C viral entry. Mature HLCs expressed all major receptors for HCV entry (Fig. 6ac), e.g., specific tight junctions: claudin-1, occludin; cell surface receptor: TAPA-1 (CD81); scavenger receptor B1 (SR-BI), ApoB and ApoE. The expression of specific HCV receptors was increased together with the maturation stage of HLCs. On the cell surface of cultured HLCs, the protein products of these receptors were also detected (Fig. 6di), e.g., claudin-1, occludin, SR-BI, CD81, LDL-R and ApoE. Human fibroblast served as a negative control (Fig. 6j).

Hepatocyte-like cells support HCVcc virus production through JFH-1 system

To determine if HLCs could replace the Huh7 cells to host HCV, HLCs were transfected with JFH-1 RNA using liposomal transfection [14]. Seven days after transfection, we found cytopathic effect with the aberration of cytoplasm in HLCs, but not with mock transfection (Fig. 7a and b). These cells had lower growth rate, aggregated and formed clusters of syncytial cells (Fig 7c). These findings suggested that HCV might re-infect the HLCs in high titer. We could not observe the CPE nor the aberration of cytoplasm in mock transfection or JFH-1 transfected Huh7 cell on day 7 and day 14 (Fig. 7df). Conditioned medium from HLCs and Huh7 was harvested every 3 d and filtered through a 0.45 μM syringe filter. The suitable multiplicity of infection (MOI) was optimized using 10-fold HCV dilution. Cell viability was determined with MTT assay (Fig. 7g and h). The HCV viral load was determined and compared between different cell lines (Fig. 7i). The kinetics of viral production varied depending on the hosting cells. The maximal HCV production was achieved between 12-15 d post-transfection. HCV viral load from Huh7 was about 105 IU/mL but HLCs could provide up to 107 IU/mL (Fig. 7i), comparable to those offered (108 IU/mL) by the transfected Huh7.5 [17]. The detection of HCV in HLCs supernatant was maintained beyond 1 month. In transfected Huh7 cells, HCV viral load was decreased to 104 IU/mL within 2 weeks and completely vanished in two months (Fig. 7i and j).

Completion of the wild-type HCV life cycle in HLC and the production of HCV progeny

The naïve HLCs and naïve Huh7 cells were infected with either HCVcc (JFH-1 genotype 2a) or with HCVser of various genotypes including 1a, 1b, 3a, 3b, 6f and 6n. HCVcc could be obtained from the earlier transfection of HLCs or Huh7 with JFH-1 RNA. HCVser was collected from chronic hepatitis C patient. The cytopathic effect (CPE) in the infected cells was observed on both day 7 and 14 after infection (Fig. 8ac). The cytoplasmic localization of non-structural proteins: NS3A, NS5A, NS5B and structural protein: HCV core antigen was clearly detected. Therefore, HLCs allowed both HCV entry and the transcription of viral RNA (Fig 8dh). To ensure that the HLCs support entire HCV life cycle, the total RNA prepared from infected HLCs and Huh7 cell was investigated for the presence of HCV negative-strand RNA using specific RT-PCR primers. The viral RNA was maintained in infected cells up to 4 weeks in HLCs 2 weeks in Huh7 cells. The presence of positive-strand of viral RNA was measured in the supernatant from the infected cells using RT-PCR analysis (Fig. 8i and j). The HCVcc viral loads were 107 IU/mL and 104 IU/mL from the supernatants of HLCs and Huh7, respectively. The levels of intracellular HCV negative–strand RNA and positive-strand RNA were determined by reverse transcription PCR every week after transfection until 4 weeks. The negative-strand RNA was merely detectable on day 7 and increased on day 14 (Fig. 8i and j). For HCVser, HLCs allowed viral propagation as determined by the increasing HCV particles in culture medium (Fig. 8k) and the detection intracellular HCV RNA (Fig. 8l). Huh7 cell, in contrast, could barely hosted HCV derived from patient serum. The HCVser viral loads were 104 IU/mL from the supernatants of HLCs. The kinetics of viral replication post-infection with either HCVcc (JFH-1) or HCVser (genotype 1a) in HLCs were determined using HCV titer and intracellular HCV RNA. The kinetics of HCV production suggested that HLCs supported complete HCV replication of both HCVcc and HCVser (Fig. 8m and n). However, HLCs did allow HCVser replication beyond 4 weeks (Fig. 8n). The immunostaining and RT-PCR analysis confirmed that the entire replication cycle and formation of HCV progeny took place in HLC and the supernatants from the infected cultures could efficiently infect naïve HLCs and Huh7 cells.

The α-tocopherol-pretreated HLCs improved HCV viral load

The effect of α-tocopherol to HCVser infection was investigated using HLCs and Huh7 cell line. After the pretreatment with 1 μM or 5 μM α-tocopherol for 24 h and infection with either HCVcc (JFH-1) or HCVser genotype 1a, the infected HLCs and Huh7 cells were further maintained for 14 days. The HCVser viral load was raised to 106 IU/mL in HLCs but the HCVcc viral load was not affected by α-tocopherol. In Huh7 cell, α-tocopherol did not improve the HCV viral load in HCVser infection and HCVcc (JFH-1) infection (Fig. 9a and b).

The efficacy study of anti-HCV drugs in infected HLCs

To screen for the sensitivity of infected HLCs to anti-HCV agents, naïve HLCs were infected with 106 IU/mL of either HCVcc (from conditioned medium of HLCs) or HCVser. Different anti-HCV agents or their combinations, e.g., ribavirin (20 μM), IFN-α (10 IU/mL), and sofosbuvir (PSI-7977, 200 nM) were incubated with the 24-h infected cells for 7 d. The HCV RNA expression in HCVcc and HCVser served as drug response. The intracellular HCV RNA was determined by real-time qPCR at 4, 8, 12, and 24 d after the infection initiation. Ribavirin, IFN-α and sofosbuvir were not cytotoxic up to 100 μM, 10 IU/mL and 100 μM respectively. Either ribavirin or INF-α decreased HCV RNA to 50 % in Huh7 and HLCs. The combination of INF-α and ribavirin decreased viral RNA to 20 %. Sofosbuvir decreased viral RNA below 20 %. (Fig. 9c and d). The IC50 of INF-α toward HCV RNA level in HLCs and Huh7 cells were 5.3 and 2.4 IU/mL respectively (p < 0.01). The IC50 of ribavirin in HLCs and Huh7 were 6.9 and 3.2 μg/mL respectively (p < 0.01). The IC50 of sofosbuvir in HLCs and Huh7 were 96 and 92 nM respectively (p < 0.01) (data not shown).

Host inflammatory response and the maintenance of long-term HCVcc propagation in HLCs

Relative real-time qPCR on days 3 and 14 revealed that the expression of inflammatory markers, e.g. TNF-α, IL-28B and IL-29 were inducible during the infection. Particularly, IL-28B expression in response to HCV infection was observed earlier only in primary human hepatocyte, but also in our HLCs model (Fig. 9e). Long-term HCVcc production was evaluated by infecting naïve HLCs with viral load 107 IU/mL for 72 h. The cell culture medium that had been replaced every week still carried HCV particles up to 6 months. The kinetics of HCV represented the new viral particles releasing from infected cell to the medium (Fig. 9f). At 4-week post-infection, the HCVcc viral load varied from 104 to 105 IU/mL and increased at 2-3 months later to 105 IU/mL. The HCV viral production was still maintained at 104-105 IU/mL up to 5 months. Until 6 months, most of the HLCs died in accordance with the viral CPE.

Discussion

Our study demonstrated that HLCs could host HCV for at least 6 months while retaining the capability to allow a complete viral life cycle. In these regards, HLCs are superior to the classical HCV host, the Huh7 cells. Despite HLCs had been studied in association with HCV, only a brief infection with JFH-1 had been addressed. These earlier studies employed immature and heterogeneous HLCs derived from iPS/ES cells as host cells for HCVcc infection [27, 28, 30]. These cells did not express all of major functional CYP450s and were functionally immature that constrained the study of HCV-host interactions [3942]. We expanded this observation by elucidating the mature HLCs not only as a viral target, but as a producer of new viral particles to complete a viral life cycle.
We modified the previous hepatic differentiation protocol [31, 43, 44] to achieve homogenous population of fully functional HLCs by extending the maturation step. The differentiated cells were challenged with a cocktail of CYP inducers. During hepatic differentiation, the differentiated cells expressed hepatic lineage markers (Sox17, AFP and albumin). In final step, we observed the homogeneous of hepatocyte-like cell with bile canaliculi, hepatic triad-like structure and bile pigments. HLCs were positive for hepatocyte selective markers, e.g., HNF-4α and tyrosine aminotransferase (TAT). The expression level of hepatic markers was higher than those of HepaRG cell, but were comparable to those of primary hepatocyte.
The xenobiotic biotransformation is a crucial and specialized function of hepatocyte. Previous studies could drive the differentiation of human iPS cells to only immature hepatocytes [45, 46]. Our study demonstrated that HLCs express phase I (CYP2B6, CYP2D6, CYP2C9 CYP2C19, CYP1A2 and CYP2E1), phase II drug-metabolizing enzyme (OATP2) and phase III drug transporters (MRP2) at levels comparable to those of primary human hepatocytes, but higher than HepaRG, the hepatocellular carcinoma cell line frequently used in pharmaceutical industry. The expression of all major CYP450s was inducible with the standard cocktail of inducers [33].
Cellular polarity is crucial for maintaining hepatic phenotype and the susceptibility to HCV [47]. At the end of hepatic induction, four types of HCV receptors on the basolateral membrane were detected, i.e., claudin-1, occludin, SR-BI and CD81. HCV particles triggered the increase in CD81 and claudin-1 endocytosis via clathrin [48]. Clathrin-dependent endocytosis and LDL metabolism involved in the processes of HCV entry [4951]. ApoB and ApoE were involved in the production of HCVcc in Huh7 cells [52]. Both apolipoproteins and LDL-receptors on apical membrane correlated with HCV entry and emission from host cells [53]. The production of infectious particles depended on VLDL secretion pathway in hepatocytes [54]. We observed the increasing expressions of ApoE, ApoB and LDL receptors after maturation.
Since the discovery of JFH-1 replication in Huh7 cells [11], subclones of Huh7 (Huh7.5 and Huh7.5.1) had been developed as a replacement for primary hepatocytes. HCV has been classified into seven major genotypes and numerous of subtypes [5557]. Nonetheless, JFH-1 is a single unusual HCV strain of genotype 2a. Other genotypes could not be maintained in Huh7 cells due to cell tropism and the restrict host range. The robust propagation systems of full-length HCV clones of the TN strain (1a) [58], H77 strain (genotype 1a) [59], JFH-2 (2a) and s310 strain (3a) [60] were developed but the construction of other HCV genotypes has not been success. HepG2, Hep3B and HEK293T cells allowed HCVcc replication [6163]. However, these cells require exogenous host factors (e.g., miR-122, claudin-1 and ApoE) for complete propagation of HCVcc [63]. Novel culture models that can host various HCV genotypes are required. Our study showed that HCV production was sustained for six months in HLCs that had been either transfected with full-length JFH-1 RNA or infected with HCV genotype 1a, 1b, 3a, 3b ,6f and 6n. No requirement for exogenous expression of cellular host factors was required for the production of HCVcc. HCVcc and HCVser produced from HLCs could infect naïve hepatocytes.
The essential HCV host factors, miR-122 and SEC14L2, facilitated the propagation of HCV in cultured hepatocyte. SEC14-like protein 2 (SEC14L2) promoted HCVser infection by upregulating vitamin E-mediated protection against lipid peroxidation. SEC14L2 together with α-tocopherol transfer protein and cellular retinol-binding protein involved in cholesterol biosynthetic pathway [64]. Both miR-122 and SEC14L2 were highly expressed in HLCs and primary human hepatocytes, but not Huh7 cells, correlated with the capability to host HCVser propagation.
After transfection with HCV RNA or infection with patient serum, we found the CPE in both transfected and infected HLCs within 7-d post-transfection. HCV viral load in culture was 104 IU/mL on day 3 and 108 IU/mL on day 15; while the intracellular HCV RNA was detectable starting on day 3. Taken together, HLCs supported complete HCV life cycle from JFH-1 system and from patient serum.
Host cell immune response play a major role in antiviral activity which is naturally occurred in animal models but defective in current cell culture model of HCV [65]. HLCs were investigated for cellular immune responses by infecting with HCVcc for 14 d and inflammatory cytokines were detected on day 3 and day 14. The cellular response to HCV infection was clearly observed in HLCs. In HLCs, the expressions of inflammatory cytokines, e.g., TNF-α, IL-28B and IL-29 were up-regulated during the infection that mimicked the natural infection in the primary hepatocytes. IL-28B is associated with the sensitivity of IFN treatment of chronic hepatitis C patients [66, 67]. These results are typical of an ongoing inflammatory response in cells that could serve as indicators for subsequent antiviral strategies.
Our study implicated that our HLCs could substitute the primary hepatocytes to serve as a generalized host for HCV. A complete life cycle of the wild-type HCV could be driven in our culture system. With this encouraging observation, an array of applications of HLCs for the study of hepatocyte-pathogen interactions would be the next direction. The list of potential pathogens included hepatitis B and dengue viruses that have been conventionally limited to human biopsies, surgical resection, organ donation and hepatocellular carcinoma cells from clinical specimens. The obtained host-pathogen interactions would provide a platform for the screening of efficacious antiviral regimens in accordance with personalized medicine.

Conclusions

Here, we have developed the functional homogenous HLCs starting from hMSCs through iPS cell reprogramming. These HLCs were confirmed for eliciting major hepatic functions including CYP450 activities. They exhibited the capability for hosting long-term HCVser infection in cell culture condition, bypassing the earlier limitation to the genotype 2a. Together with the expression of SEC14L2 and the addition of α-tocopherol, the HCVser infection in HLCs were further intensified. We demonstrated that the production of HCV genotype 1a, 1b, 3a, 3b ,6f and 6n were sustained for at least six months in these HLCs. The HCV-infection was controlled by the upregulation of TNF-α, IL-28B and IL-29. Taken together, this novel HCV culture could serve as a model for HCV biology, host immune response, and antiviral strategy.
Human mesenchymal stem cells (hMSC) were isolated from aspirated bone marrow of consenting healthy donors (n = 3). This procedure received an approval from the Ethics Committee on Research Involving Human Subjects at Ramathibodi Hospital, Mahidol University (2554/404). All subjects are adults and provided written informed consent.

Acknowledgements

The authors would like to thank Prof. Takaji Wakita (National Institute of Infectious Diseases, Japan) for his generous gift of the full-length JFH1 HCV plasmid.

Funding

This study was funded by the research grants of Mahidol University and the Thailand Research Fund (TRF) separately awarded to A. Wongkajornsilp and K. Sa-ngiamsuntorn. K. Sa-ngiamsuntorn is a recipient of Mahidol Postdoctoral Fellowship. A. Wongkajornsilp is a recipient of Chalermprakiat grant of the Faculty of Medicine Siriraj Hospital, Mahidol University.
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://​creativecommons.​org/​licenses/​by/​4.​0/​), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The Creative Commons Public Domain Dedication waiver (http://​creativecommons.​org/​publicdomain/​zero/​1.​0/​) applies to the data made available in this article, unless otherwise stated.

Competing interests

The authors declare that they have no competing interests.

Authors’ contributions

KS and PP performed most of the experiments, designed the protocol, performed the statistical analysis and drafted the manuscript. AW, SB, and SH designed the protocol. NK participated in the teratoma assay in nude mice. PP prepared the lentiviral vectors and iPS cell reprogramming. WC collected HCV from patient serum and provided HCV viral load detection using Abbott RealTime HCV assay. AW and SH coordinated the study. All authors have read and approved the final manuscript.
insite
INHALT
download
DOWNLOAD
print
DRUCKEN
Literatur
1.
Zurück zum Zitat Pawlotsky JM, Chevaliez S, McHutchison JG. The hepatitis C virus life cycle as a target for new antiviral therapies. Gastroenterology. 2007;132:1979–98.CrossRefPubMed Pawlotsky JM, Chevaliez S, McHutchison JG. The hepatitis C virus life cycle as a target for new antiviral therapies. Gastroenterology. 2007;132:1979–98.CrossRefPubMed
2.
Zurück zum Zitat Simmonds P, Bukh J, Combet C, Deleage G, Enomoto N, Feinstone S, Halfon P, Inchauspe G, Kuiken C, Maertens G, et al. Consensus proposals for a unified system of nomenclature of hepatitis C virus genotypes. Hepatology. 2005;42:962–73. Simmonds P, Bukh J, Combet C, Deleage G, Enomoto N, Feinstone S, Halfon P, Inchauspe G, Kuiken C, Maertens G, et al. Consensus proposals for a unified system of nomenclature of hepatitis C virus genotypes. Hepatology. 2005;42:962–73.
3.
5.
Zurück zum Zitat Joshi D, Pinzani M, Carey I, Agarwal K. Recurrent HCV after liver transplantation-mechanisms, assessment and therapy. Nat Rev Gastroenterol Hepatol. 2014;11:710–21.CrossRefPubMed Joshi D, Pinzani M, Carey I, Agarwal K. Recurrent HCV after liver transplantation-mechanisms, assessment and therapy. Nat Rev Gastroenterol Hepatol. 2014;11:710–21.CrossRefPubMed
6.
Zurück zum Zitat Xiao F, Fofana I, Heydmann L, Barth H, Soulier E, Habersetzer F, Doffoel M, Bukh J, Patel AH, Zeisel MB, Baumert TF. Hepatitis C virus cell-cell transmission and resistance to direct-acting antiviral agents. PLoS Pathog. 2014;10:e1004128. Xiao F, Fofana I, Heydmann L, Barth H, Soulier E, Habersetzer F, Doffoel M, Bukh J, Patel AH, Zeisel MB, Baumert TF. Hepatitis C virus cell-cell transmission and resistance to direct-acting antiviral agents. PLoS Pathog. 2014;10:e1004128.
7.
Zurück zum Zitat Soltys KA, Soto-Gutierrez A, Nagaya M, Baskin KM, Deutsch M, Ito R, Shneider BL, Squires R, Vockley J, Guha C. Barriers to the successful treatment of liver disease by hepatocyte transplantation. J Hepatol. 2010;53:769–74.CrossRefPubMedPubMedCentral Soltys KA, Soto-Gutierrez A, Nagaya M, Baskin KM, Deutsch M, Ito R, Shneider BL, Squires R, Vockley J, Guha C. Barriers to the successful treatment of liver disease by hepatocyte transplantation. J Hepatol. 2010;53:769–74.CrossRefPubMedPubMedCentral
8.
Zurück zum Zitat Jazwinski AB, Muir AJ. Direct-acting antiviral medications for chronic hepatitis C virus infection. Gastroenterol Hepatol (N Y). 2011;7:154–62. Jazwinski AB, Muir AJ. Direct-acting antiviral medications for chronic hepatitis C virus infection. Gastroenterol Hepatol (N Y). 2011;7:154–62.
9.
Zurück zum Zitat Halfon P, Locarnini S. Hepatitis C virus resistance to protease inhibitors. J Hepatol. 2011;55:192–206.CrossRefPubMed Halfon P, Locarnini S. Hepatitis C virus resistance to protease inhibitors. J Hepatol. 2011;55:192–206.CrossRefPubMed
10.
Zurück zum Zitat Vinaixa C, Rubin A, Aguilera V, Berenguer M. Recurrence of hepatitis C after liver transplantation. Ann Gastroenterol. 2013;26:304–13.PubMedPubMedCentral Vinaixa C, Rubin A, Aguilera V, Berenguer M. Recurrence of hepatitis C after liver transplantation. Ann Gastroenterol. 2013;26:304–13.PubMedPubMedCentral
11.
Zurück zum Zitat Wakita T, Kato T. Development of an Infectious HCV Cell Culture System. In Hepatitis C Viruses: Genomes and Molecular Biology. Edited by Tan SL. Norfolk (UK): Horizon Bioscience; 2006. Wakita T, Kato T. Development of an Infectious HCV Cell Culture System. In Hepatitis C Viruses: Genomes and Molecular Biology. Edited by Tan SL. Norfolk (UK): Horizon Bioscience; 2006.
12.
Zurück zum Zitat Wakita T, Pietschmann T, Kato T, Date T, Miyamoto M, Zhao Z, Murthy K, Habermann A, Krausslich HG, Mizokami M, et al. Production of infectious hepatitis C virus in tissue culture from a cloned viral genome. Nat Med. 2005;11:791–6. Wakita T, Pietschmann T, Kato T, Date T, Miyamoto M, Zhao Z, Murthy K, Habermann A, Krausslich HG, Mizokami M, et al. Production of infectious hepatitis C virus in tissue culture from a cloned viral genome. Nat Med. 2005;11:791–6.
13.
Zurück zum Zitat Catanese MT, Dorner M. Advances in experimental systems to study hepatitis C virus in vitro and in vivo. Virology. 2015;479–480:221–33.CrossRefPubMed Catanese MT, Dorner M. Advances in experimental systems to study hepatitis C virus in vitro and in vivo. Virology. 2015;479–480:221–33.CrossRefPubMed
14.
Zurück zum Zitat Bartenschlager R, Pietschmann T. Efficient hepatitis C virus cell culture system: what a difference the host cell makes. Proc Natl Acad Sci U S A. 2005;102:9739–40.CrossRefPubMedPubMedCentral Bartenschlager R, Pietschmann T. Efficient hepatitis C virus cell culture system: what a difference the host cell makes. Proc Natl Acad Sci U S A. 2005;102:9739–40.CrossRefPubMedPubMedCentral
15.
Zurück zum Zitat Blight KJ, McKeating JA, Marcotrigiano J, Rice CM. Efficient replication of hepatitis C virus genotype 1a RNAs in cell culture. J Virol. 2003;77:3181–90.CrossRefPubMedPubMedCentral Blight KJ, McKeating JA, Marcotrigiano J, Rice CM. Efficient replication of hepatitis C virus genotype 1a RNAs in cell culture. J Virol. 2003;77:3181–90.CrossRefPubMedPubMedCentral
16.
Zurück zum Zitat Sa-ngiamsuntorn K, Wongkajornsilp A, Kasetsinsombat K, Duangsa-ard S, Nuntakarn L, Borwornpinyo S, Akarasereenont P, Limsrichamrern S, Hongeng S. Upregulation of CYP 450 s expression of immortalized hepatocyte-like cells derived from mesenchymal stem cells by enzyme inducers. BMC Biotechnol. 2011;11:89. Sa-ngiamsuntorn K, Wongkajornsilp A, Kasetsinsombat K, Duangsa-ard S, Nuntakarn L, Borwornpinyo S, Akarasereenont P, Limsrichamrern S, Hongeng S. Upregulation of CYP 450 s expression of immortalized hepatocyte-like cells derived from mesenchymal stem cells by enzyme inducers. BMC Biotechnol. 2011;11:89.
17.
Zurück zum Zitat Zhong J, Gastaminza P, Cheng G, Kapadia S, Kato T, Burton DR, Wieland SF, Uprichard SL, Wakita T, Chisari FV. Robust hepatitis C virus infection in vitro. Proc Natl Acad Sci U S A. 2005;102:9294–9. Zhong J, Gastaminza P, Cheng G, Kapadia S, Kato T, Burton DR, Wieland SF, Uprichard SL, Wakita T, Chisari FV. Robust hepatitis C virus infection in vitro. Proc Natl Acad Sci U S A. 2005;102:9294–9.
18.
Zurück zum Zitat Blight KJ, McKeating JA, Rice CM. Highly permissive cell lines for subgenomic and genomic hepatitis C virus RNA replication. J Virol. 2002;76:13001–14.CrossRefPubMedPubMedCentral Blight KJ, McKeating JA, Rice CM. Highly permissive cell lines for subgenomic and genomic hepatitis C virus RNA replication. J Virol. 2002;76:13001–14.CrossRefPubMedPubMedCentral
19.
Zurück zum Zitat Breiman A, Grandvaux N, Lin R, Ottone C, Akira S, Yoneyama M, Fujita T, Hiscott J, Meurs EF. Inhibition of RIG-I-dependent signaling to the interferon pathway during hepatitis C virus expression and restoration of signaling by IKKepsilon. J Virol. 2005;79:3969–78. Breiman A, Grandvaux N, Lin R, Ottone C, Akira S, Yoneyama M, Fujita T, Hiscott J, Meurs EF. Inhibition of RIG-I-dependent signaling to the interferon pathway during hepatitis C virus expression and restoration of signaling by IKKepsilon. J Virol. 2005;79:3969–78.
20.
Zurück zum Zitat Helle F, Brochot E, Fournier C, Descamps V, Izquierdo L, Hoffmann TW, Morel V, Herpe YE, Bengrine A, Belouzard S, et al. Permissivity of primary human hepatocytes and different hepatoma cell lines to cell culture adapted hepatitis C virus. PLoS One. 2013;8:e70809. Helle F, Brochot E, Fournier C, Descamps V, Izquierdo L, Hoffmann TW, Morel V, Herpe YE, Bengrine A, Belouzard S, et al. Permissivity of primary human hepatocytes and different hepatoma cell lines to cell culture adapted hepatitis C virus. PLoS One. 2013;8:e70809.
21.
Zurück zum Zitat Podevin P, Carpentier A, Pene V, Aoudjehane L, Carriere M, Zaidi S, Hernandez C, Calle V, Meritet JF, Scatton O, et al. Production of infectious hepatitis C virus in primary cultures of human adult hepatocytes. Gastroenterology. 2010;139:1355–64. Podevin P, Carpentier A, Pene V, Aoudjehane L, Carriere M, Zaidi S, Hernandez C, Calle V, Meritet JF, Scatton O, et al. Production of infectious hepatitis C virus in primary cultures of human adult hepatocytes. Gastroenterology. 2010;139:1355–64.
22.
Zurück zum Zitat Gondeau C, Briolotti P, Razafy F, Duret C, Rubbo PA, Helle F, Reme T, Ripault MP, Ducos J, Fabre JM, et al. In vitro infection of primary human hepatocytes by HCV-positive sera: insights on a highly relevant model. Gut. 2014;63:1490–500. Gondeau C, Briolotti P, Razafy F, Duret C, Rubbo PA, Helle F, Reme T, Ripault MP, Ducos J, Fabre JM, et al. In vitro infection of primary human hepatocytes by HCV-positive sera: insights on a highly relevant model. Gut. 2014;63:1490–500.
23.
24.
Zurück zum Zitat Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006;126:663–76.CrossRefPubMed Takahashi K, Yamanaka S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell. 2006;126:663–76.CrossRefPubMed
25.
Zurück zum Zitat Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131:861–72. Takahashi K, Tanabe K, Ohnuki M, Narita M, Ichisaka T, Tomoda K, Yamanaka S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell. 2007;131:861–72.
26.
Zurück zum Zitat Chun YS, Byun K, Lee B. Induced pluripotent stem cells and personalized medicine: current progress and future perspectives. Anat Cell Biol. 2011;44:245–55.CrossRefPubMedPubMedCentral Chun YS, Byun K, Lee B. Induced pluripotent stem cells and personalized medicine: current progress and future perspectives. Anat Cell Biol. 2011;44:245–55.CrossRefPubMedPubMedCentral
27.
Zurück zum Zitat Schwartz RE, Trehan K, Andrus L, Sheahan TP, Ploss A, Duncan SA, Rice CM, Bhatia SN. Modeling hepatitis C virus infection using human induced pluripotent stem cells. Proc Natl Acad Sci U S A. 2012;109:2544–8. Schwartz RE, Trehan K, Andrus L, Sheahan TP, Ploss A, Duncan SA, Rice CM, Bhatia SN. Modeling hepatitis C virus infection using human induced pluripotent stem cells. Proc Natl Acad Sci U S A. 2012;109:2544–8.
28.
Zurück zum Zitat Roelandt P, Obeid S, Paeshuyse J, Vanhove J, Van Lommel A, Nahmias Y, Nevens F, Neyts J, Verfaillie CM. Human pluripotent stem cell-derived hepatocytes support complete replication of hepatitis C virus. J Hepatol. 2012;57:246–51. Roelandt P, Obeid S, Paeshuyse J, Vanhove J, Van Lommel A, Nahmias Y, Nevens F, Neyts J, Verfaillie CM. Human pluripotent stem cell-derived hepatocytes support complete replication of hepatitis C virus. J Hepatol. 2012;57:246–51.
29.
Zurück zum Zitat Si-Tayeb K, Duclos-Vallee JC, Petit MA. Hepatocyte-like cells differentiated from human induced pluripotent stem cells (iHLCs) are permissive to hepatitis C virus (HCV) infection: HCV study gets personal. J Hepatol. 2012;57:689–91.CrossRefPubMed Si-Tayeb K, Duclos-Vallee JC, Petit MA. Hepatocyte-like cells differentiated from human induced pluripotent stem cells (iHLCs) are permissive to hepatitis C virus (HCV) infection: HCV study gets personal. J Hepatol. 2012;57:689–91.CrossRefPubMed
30.
Zurück zum Zitat Wu X, Robotham JM, Lee E, Dalton S, Kneteman NM, Gilbert DM, Tang H. Productive hepatitis C virus infection of stem cell-derived hepatocytes reveals a critical transition to viral permissiveness during differentiation. PLoS Pathog. 2012;8:e1002617. Wu X, Robotham JM, Lee E, Dalton S, Kneteman NM, Gilbert DM, Tang H. Productive hepatitis C virus infection of stem cell-derived hepatocytes reveals a critical transition to viral permissiveness during differentiation. PLoS Pathog. 2012;8:e1002617.
31.
32.
Zurück zum Zitat Warlich E, Kuehle J, Cantz T, Brugman MH, Maetzig T, Galla M, Filipczyk AA, Halle S, Klump H, Scholer HR, et al. Lentiviral vector design and imaging approaches to visualize the early stages of cellular reprogramming. Mol Ther. 2011;19:782–9. Warlich E, Kuehle J, Cantz T, Brugman MH, Maetzig T, Galla M, Filipczyk AA, Halle S, Klump H, Scholer HR, et al. Lentiviral vector design and imaging approaches to visualize the early stages of cellular reprogramming. Mol Ther. 2011;19:782–9.
33.
Zurück zum Zitat Ulvestad M, Nordell P, Asplund A, Rehnstrom M, Jacobsson S, Holmgren G, Davidson L, Brolen G, Edsbagge J, Bjorquist P, et al. Drug metabolizing enzyme and transporter protein profiles of hepatocytes derived from human embryonic and induced pluripotent stem cells. Biochem Pharmacol. 2013;86:691–702. Ulvestad M, Nordell P, Asplund A, Rehnstrom M, Jacobsson S, Holmgren G, Davidson L, Brolen G, Edsbagge J, Bjorquist P, et al. Drug metabolizing enzyme and transporter protein profiles of hepatocytes derived from human embryonic and induced pluripotent stem cells. Biochem Pharmacol. 2013;86:691–702.
34.
Zurück zum Zitat Vieyres G, Pietschmann T. Entry and replication of recombinant hepatitis C viruses in cell culture. Methods. 2013;59:233–48.CrossRefPubMed Vieyres G, Pietschmann T. Entry and replication of recombinant hepatitis C viruses in cell culture. Methods. 2013;59:233–48.CrossRefPubMed
35.
Zurück zum Zitat Yu X, Uprichard SL: Cell-based hepatitis C virus infection fluorescence resonance energy transfer (FRET) assay for antiviral compound screening. Curr Protoc Microbiol 2010, Chapter 17:Unit 17 15. Yu X, Uprichard SL: Cell-based hepatitis C virus infection fluorescence resonance energy transfer (FRET) assay for antiviral compound screening. Curr Protoc Microbiol 2010, Chapter 17:Unit 17 15.
36.
Zurück zum Zitat Raymond VA, Selliah S, Ethier C, Houle R, Jouan L, Maniere T, Lamarre D, Willems B, Bilodeau M. Primary cultures of human hepatocytes isolated from hepatitis C virus-infected cirrhotic livers as a model to study hepatitis C infection. Liver Int. 2009;29:942–9. Raymond VA, Selliah S, Ethier C, Houle R, Jouan L, Maniere T, Lamarre D, Willems B, Bilodeau M. Primary cultures of human hepatocytes isolated from hepatitis C virus-infected cirrhotic livers as a model to study hepatitis C infection. Liver Int. 2009;29:942–9.
37.
Zurück zum Zitat Zorn AM. Liver development. In StemBook. Cambridge (MA): The Stem Cell Research Community; 2008. Zorn AM. Liver development. In StemBook. Cambridge (MA): The Stem Cell Research Community; 2008.
38.
Zurück zum Zitat Hu A, Shang C, Li Q, Sun N, Wu L, Ma Y, Jiao X, Min J, Zeng G, He X. Epithelial-mesenchymal transition delayed by E-cad to promote tissue formation in hepatic differentiation of mouse embryonic stem cells in vitro. Stem Cells Dev. 2014;23:877–87. Hu A, Shang C, Li Q, Sun N, Wu L, Ma Y, Jiao X, Min J, Zeng G, He X. Epithelial-mesenchymal transition delayed by E-cad to promote tissue formation in hepatic differentiation of mouse embryonic stem cells in vitro. Stem Cells Dev. 2014;23:877–87.
39.
Zurück zum Zitat Touboul T, Hannan NR, Corbineau S, Martinez A, Martinet C, Branchereau S, Mainot S, Strick-Marchand H, Pedersen R, Di Santo J, et al. Generation of functional hepatocytes from human embryonic stem cells under chemically defined conditions that recapitulate liver development. Hepatology. 2010;51:1754–65. Touboul T, Hannan NR, Corbineau S, Martinez A, Martinet C, Branchereau S, Mainot S, Strick-Marchand H, Pedersen R, Di Santo J, et al. Generation of functional hepatocytes from human embryonic stem cells under chemically defined conditions that recapitulate liver development. Hepatology. 2010;51:1754–65.
40.
Zurück zum Zitat Song Z, Cai J, Liu Y, Zhao D, Yong J, Duo S, Song X, Guo Y, Zhao Y, Qin H, et al. Efficient generation of hepatocyte-like cells from human induced pluripotent stem cells. Cell Res. 2009;19:1233–42. Song Z, Cai J, Liu Y, Zhao D, Yong J, Duo S, Song X, Guo Y, Zhao Y, Qin H, et al. Efficient generation of hepatocyte-like cells from human induced pluripotent stem cells. Cell Res. 2009;19:1233–42.
41.
Zurück zum Zitat Chen YF, Tseng CY, Wang HW, Kuo HC, Yang VW, Lee OK. Rapid generation of mature hepatocyte-like cells from human induced pluripotent stem cells by an efficient three-step protocol. Hepatology. 2012;55:1193–203.CrossRefPubMedPubMedCentral Chen YF, Tseng CY, Wang HW, Kuo HC, Yang VW, Lee OK. Rapid generation of mature hepatocyte-like cells from human induced pluripotent stem cells by an efficient three-step protocol. Hepatology. 2012;55:1193–203.CrossRefPubMedPubMedCentral
42.
Zurück zum Zitat Si-Tayeb K, Noto FK, Nagaoka M, Li J, Battle MA, Duris C, North PE, Dalton S, Duncan SA. Highly efficient generation of human hepatocyte-like cells from induced pluripotent stem cells. Hepatology. 2010;51:297–305. Si-Tayeb K, Noto FK, Nagaoka M, Li J, Battle MA, Duris C, North PE, Dalton S, Duncan SA. Highly efficient generation of human hepatocyte-like cells from induced pluripotent stem cells. Hepatology. 2010;51:297–305.
43.
Zurück zum Zitat Wongkajornsilp A, Sa-Ngiamsuntorn K, Hongeng S. Development of immortalized hepatocyte-like cells from hMSCs. Methods Mol Biol. 2012;826:73–87.CrossRefPubMed Wongkajornsilp A, Sa-Ngiamsuntorn K, Hongeng S. Development of immortalized hepatocyte-like cells from hMSCs. Methods Mol Biol. 2012;826:73–87.CrossRefPubMed
44.
Zurück zum Zitat Zhao D, Chen S, Duo S, Xiang C, Jia J, Guo M, Lai W, Lu S, Deng H. Promotion of the efficient metabolic maturation of human pluripotent stem cell-derived hepatocytes by correcting specification defects. Cell Res. 2013;23:157–61. Zhao D, Chen S, Duo S, Xiang C, Jia J, Guo M, Lai W, Lu S, Deng H. Promotion of the efficient metabolic maturation of human pluripotent stem cell-derived hepatocytes by correcting specification defects. Cell Res. 2013;23:157–61.
45.
Zurück zum Zitat Chistiakov DA, Chistiakov PA. Strategies to produce hepatocytes and hepatocyte-like cells from pluripotent stem cells. Hepatol Res. 2012;42:111–9.CrossRefPubMed Chistiakov DA, Chistiakov PA. Strategies to produce hepatocytes and hepatocyte-like cells from pluripotent stem cells. Hepatol Res. 2012;42:111–9.CrossRefPubMed
46.
Zurück zum Zitat Du Y, Wang J, Jia J, Song N, Xiang C, Xu J, Hou Z, Su X, Liu B, Jiang T, et al. Human hepatocytes with drug metabolic function induced from fibroblasts by lineage reprogramming. Cell Stem Cell. 2014;14:394–403. Du Y, Wang J, Jia J, Song N, Xiang C, Xu J, Hou Z, Su X, Liu B, Jiang T, et al. Human hepatocytes with drug metabolic function induced from fibroblasts by lineage reprogramming. Cell Stem Cell. 2014;14:394–403.
47.
Zurück zum Zitat Sainz Jr B, TenCate V, Uprichard SL. Three-dimensional Huh7 cell culture system for the study of Hepatitis C virus infection. Virol J. 2009;6:103.CrossRefPubMedPubMedCentral Sainz Jr B, TenCate V, Uprichard SL. Three-dimensional Huh7 cell culture system for the study of Hepatitis C virus infection. Virol J. 2009;6:103.CrossRefPubMedPubMedCentral
48.
Zurück zum Zitat Farquhar MJ, Hu K, Harris HJ, Davis C, Brimacombe CL, Fletcher SJ, Baumert TF, Rappoport JZ, Balfe P, McKeating JA. Hepatitis C virus induces CD81 and claudin-1 endocytosis. J Virol. 2012;86:4305–16. Farquhar MJ, Hu K, Harris HJ, Davis C, Brimacombe CL, Fletcher SJ, Baumert TF, Rappoport JZ, Balfe P, McKeating JA. Hepatitis C virus induces CD81 and claudin-1 endocytosis. J Virol. 2012;86:4305–16.
49.
Zurück zum Zitat Meertens L, Bertaux C, Dragic T. Hepatitis C virus entry requires a critical postinternalization step and delivery to early endosomes via clathrin-coated vesicles. J Virol. 2006;80:11571–8.CrossRefPubMedPubMedCentral Meertens L, Bertaux C, Dragic T. Hepatitis C virus entry requires a critical postinternalization step and delivery to early endosomes via clathrin-coated vesicles. J Virol. 2006;80:11571–8.CrossRefPubMedPubMedCentral
50.
Zurück zum Zitat Blanchard E, Belouzard S, Goueslain L, Wakita T, Dubuisson J, Wychowski C, Rouille Y. Hepatitis C virus entry depends on clathrin-mediated endocytosis. J Virol. 2006;80:6964–72. Blanchard E, Belouzard S, Goueslain L, Wakita T, Dubuisson J, Wychowski C, Rouille Y. Hepatitis C virus entry depends on clathrin-mediated endocytosis. J Virol. 2006;80:6964–72.
51.
Zurück zum Zitat Coller KE, Berger KL, Heaton NS, Cooper JD, Yoon R, Randall G. RNA interference and single particle tracking analysis of hepatitis C virus endocytosis. PLoS Pathog. 2009;5:e1000702.CrossRefPubMedPubMedCentral Coller KE, Berger KL, Heaton NS, Cooper JD, Yoon R, Randall G. RNA interference and single particle tracking analysis of hepatitis C virus endocytosis. PLoS Pathog. 2009;5:e1000702.CrossRefPubMedPubMedCentral
52.
Zurück zum Zitat Shiokawa M, Fukuhara T, Ono C, Yamamoto S, Okamoto T, Watanabe N, Wakita T, Matsuura Y. Novel permissive cell lines for complete propagation of hepatitis C virus. J Virol. 2014;88:5578–94. Shiokawa M, Fukuhara T, Ono C, Yamamoto S, Okamoto T, Watanabe N, Wakita T, Matsuura Y. Novel permissive cell lines for complete propagation of hepatitis C virus. J Virol. 2014;88:5578–94.
53.
Zurück zum Zitat Owen DM, Huang H, Ye J, Gale Jr M. Apolipoprotein E on hepatitis C virion facilitates infection through interaction with low-density lipoprotein receptor. Virology. 2009;394:99–108.CrossRefPubMedPubMedCentral Owen DM, Huang H, Ye J, Gale Jr M. Apolipoprotein E on hepatitis C virion facilitates infection through interaction with low-density lipoprotein receptor. Virology. 2009;394:99–108.CrossRefPubMedPubMedCentral
54.
Zurück zum Zitat Gastaminza P, Cheng G, Wieland S, Zhong J, Liao W, Chisari FV. Cellular determinants of hepatitis C virus assembly, maturation, degradation, and secretion. J Virol. 2008;82:2120–9.CrossRefPubMedPubMedCentral Gastaminza P, Cheng G, Wieland S, Zhong J, Liao W, Chisari FV. Cellular determinants of hepatitis C virus assembly, maturation, degradation, and secretion. J Virol. 2008;82:2120–9.CrossRefPubMedPubMedCentral
55.
Zurück zum Zitat Gottwein JM, Scheel TK, Jensen TB, Lademann JB, Prentoe JC, Knudsen ML, Hoegh AM, Bukh J. Development and characterization of hepatitis C virus genotype 1-7 cell culture systems: role of CD81 and scavenger receptor class B type I and effect of antiviral drugs. Hepatology. 2009;49:364–77. Gottwein JM, Scheel TK, Jensen TB, Lademann JB, Prentoe JC, Knudsen ML, Hoegh AM, Bukh J. Development and characterization of hepatitis C virus genotype 1-7 cell culture systems: role of CD81 and scavenger receptor class B type I and effect of antiviral drugs. Hepatology. 2009;49:364–77.
56.
Zurück zum Zitat Scheel TK, Gottwein JM, Mikkelsen LS, Jensen TB, Bukh J. Recombinant HCV variants with NS5A from genotypes 1-7 have different sensitivities to an NS5A inhibitor but not interferon-alpha. Gastroenterology. 2011;140:1032–42.CrossRefPubMed Scheel TK, Gottwein JM, Mikkelsen LS, Jensen TB, Bukh J. Recombinant HCV variants with NS5A from genotypes 1-7 have different sensitivities to an NS5A inhibitor but not interferon-alpha. Gastroenterology. 2011;140:1032–42.CrossRefPubMed
57.
Zurück zum Zitat Tariq H, Manzoor S, Parvaiz F, Javed F, Fatima K, Qadri I. An overview: in vitro models of HCV replication in different cell cultures. Infect Genet Evol. 2012;12:13–20.CrossRefPubMed Tariq H, Manzoor S, Parvaiz F, Javed F, Fatima K, Qadri I. An overview: in vitro models of HCV replication in different cell cultures. Infect Genet Evol. 2012;12:13–20.CrossRefPubMed
58.
Zurück zum Zitat Li YP, Ramirez S, Jensen SB, Purcell RH, Gottwein JM, Bukh J. Highly efficient full-length hepatitis C virus genotype 1 (strain TN) infectious culture system. Proc Natl Acad Sci U S A. 2012;109:19757–62.CrossRefPubMedPubMedCentral Li YP, Ramirez S, Jensen SB, Purcell RH, Gottwein JM, Bukh J. Highly efficient full-length hepatitis C virus genotype 1 (strain TN) infectious culture system. Proc Natl Acad Sci U S A. 2012;109:19757–62.CrossRefPubMedPubMedCentral
59.
Zurück zum Zitat Yi M, Villanueva RA, Thomas DL, Wakita T, Lemon SM. Production of infectious genotype 1a hepatitis C virus (Hutchinson strain) in cultured human hepatoma cells. Proc Natl Acad Sci U S A. 2006;103:2310–5.CrossRefPubMedPubMedCentral Yi M, Villanueva RA, Thomas DL, Wakita T, Lemon SM. Production of infectious genotype 1a hepatitis C virus (Hutchinson strain) in cultured human hepatoma cells. Proc Natl Acad Sci U S A. 2006;103:2310–5.CrossRefPubMedPubMedCentral
60.
Zurück zum Zitat Saeed M, Gondeau C, Hmwe S, Yokokawa H, Date T, Suzuki T, Kato T, Maurel P, Wakita T. Replication of hepatitis C virus genotype 3a in cultured cells. Gastroenterology. 2013;144:56–8. e57. Saeed M, Gondeau C, Hmwe S, Yokokawa H, Date T, Suzuki T, Kato T, Maurel P, Wakita T. Replication of hepatitis C virus genotype 3a in cultured cells. Gastroenterology. 2013;144:56–8. e57.
61.
Zurück zum Zitat Narbus CM, Israelow B, Sourisseau M, Michta ML, Hopcraft SE, Zeiner GM, Evans MJ. HepG2 cells expressing microRNA miR-122 support the entire hepatitis C virus life cycle. J Virol. 2011;85:12087–92. Narbus CM, Israelow B, Sourisseau M, Michta ML, Hopcraft SE, Zeiner GM, Evans MJ. HepG2 cells expressing microRNA miR-122 support the entire hepatitis C virus life cycle. J Virol. 2011;85:12087–92.
62.
Zurück zum Zitat Kambara H, Fukuhara T, Shiokawa M, Ono C, Ohara Y, Kamitani W, Matsuura Y. Establishment of a novel permissive cell line for the propagation of hepatitis C virus by expression of microRNA miR122. J Virol. 2012;86:1382–93. Kambara H, Fukuhara T, Shiokawa M, Ono C, Ohara Y, Kamitani W, Matsuura Y. Establishment of a novel permissive cell line for the propagation of hepatitis C virus by expression of microRNA miR122. J Virol. 2012;86:1382–93.
63.
Zurück zum Zitat Da Costa D, Turek M, Felmlee DJ, Girardi E, Pfeffer S, Long G, Bartenschlager R, Zeisel MB, Baumert TF. Reconstitution of the entire hepatitis C virus life cycle in nonhepatic cells. J Virol. 2012;86:11919–25. Da Costa D, Turek M, Felmlee DJ, Girardi E, Pfeffer S, Long G, Bartenschlager R, Zeisel MB, Baumert TF. Reconstitution of the entire hepatitis C virus life cycle in nonhepatic cells. J Virol. 2012;86:11919–25.
64.
Zurück zum Zitat Saeed M, Andreo U, Chung HY, Espiritu C, Branch AD, Silva JM, Rice CM. SEC14L2 enables pan-genotype HCV replication in cell culture. Nature. 2015;524:471–5. Saeed M, Andreo U, Chung HY, Espiritu C, Branch AD, Silva JM, Rice CM. SEC14L2 enables pan-genotype HCV replication in cell culture. Nature. 2015;524:471–5.
65.
Zurück zum Zitat Schoggins JW, Rice CM. Innate immune responses to hepatitis C virus. Curr Top Microbiol Immunol. 2013;369:219–42.PubMed Schoggins JW, Rice CM. Innate immune responses to hepatitis C virus. Curr Top Microbiol Immunol. 2013;369:219–42.PubMed
66.
Zurück zum Zitat Fukuhara T, Taketomi A, Motomura T, Okano S, Ninomiya A, Abe T, Uchiyama H, Soejima Y, Shirabe K, Matsuura Y, Maehara Y. Variants in IL28B in liver recipients and donors correlate with response to peg-interferon and ribavirin therapy for recurrent hepatitis C. Gastroenterology. 2010;139:1577–85. 1585 e1571-1573. Fukuhara T, Taketomi A, Motomura T, Okano S, Ninomiya A, Abe T, Uchiyama H, Soejima Y, Shirabe K, Matsuura Y, Maehara Y. Variants in IL28B in liver recipients and donors correlate with response to peg-interferon and ribavirin therapy for recurrent hepatitis C. Gastroenterology. 2010;139:1577–85. 1585 e1571-1573.
67.
Zurück zum Zitat Suppiah V, Moldovan M, Ahlenstiel G, Berg T, Weltman M, Abate ML, Bassendine M, Spengler U, Dore GJ, Powell E, et al. IL28B is associated with response to chronic hepatitis C interferon-alpha and ribavirin therapy. Nat Genet. 2009;41:1100–4. Suppiah V, Moldovan M, Ahlenstiel G, Berg T, Weltman M, Abate ML, Bassendine M, Spengler U, Dore GJ, Powell E, et al. IL28B is associated with response to chronic hepatitis C interferon-alpha and ribavirin therapy. Nat Genet. 2009;41:1100–4.
Metadaten
Titel
A robust model of natural hepatitis C infection using hepatocyte-like cells derived from human induced pluripotent stem cells as a long-term host
verfasst von
Khanit Sa-ngiamsuntorn
Adisak Wongkajornsilp
Phetcharat Phanthong
Suparerk Borwornpinyo
Narisorn Kitiyanant
Wasun Chantratita
Suradej Hongeng
Publikationsdatum
01.12.2016
Verlag
BioMed Central
Erschienen in
Virology Journal / Ausgabe 1/2016
Elektronische ISSN: 1743-422X
DOI
https://doi.org/10.1186/s12985-016-0519-1

Weitere Artikel der Ausgabe 1/2016

Virology Journal 1/2016 Zur Ausgabe

Leitlinien kompakt für die Innere Medizin

Mit medbee Pocketcards sicher entscheiden.

Seit 2022 gehört die medbee GmbH zum Springer Medizin Verlag

Update Innere Medizin

Bestellen Sie unseren Fach-Newsletter und bleiben Sie gut informiert.