ArticlePDF Available

Antiviral Effect and Virus-Host Interactions in Response to Alpha Interferon, Gamma Interferon, Poly(I)-Poly(C), Tumor Necrosis Factor Alpha, and Ribavirin in Hepatitis C Virus Subgenomic Replicons

American Society for Microbiology
Journal of Virology
Authors:

Abstract and Figures

The recently developed hepatitis C virus (HCV) subgenomic replicon system was utilized to evaluate the efficacy of several known antiviral agents. Cell lines that persistently maintained a genotype 1b replicon were selected. The replicon resident in each cell line had acquired adaptive mutations in the NS5A region that increased colony-forming efficiency, and some replicons had acquired NS3 mutations that alone did not enhance colony-forming efficiency but were synergistic with NS5A mutations. A replicon constructed from the infectious clone of the HCV-1 strain (genotype 1a) was not capable of inducing colony formation even after the introduction of adaptive mutations identified in the genotype 1b replicon. Alpha interferon (IFN-alpha), IFN-gamma, and ribavirin exhibited antiviral activity, while double-stranded RNA (dsRNA) and tumor necrosis factor alpha did not. Analysis of transcript levels for a series of genes stimulated by IFN (ISGs) or dsRNA following treatment with IFN-alpha, IFN-gamma, and dsRNA revealed that both IFNs increased ISG transcript levels, but that some aspect of the dsRNA response pathway was defective in Huh7 cells and replicon cell lines in comparison to primary chimpanzee and tamarin hepatocytes. The colony-forming efficiency of the replicon was reduced or eliminated following replication in the presence of ribavirin, implicating the induction of error-prone replication. The potential role of error-prone replication in the synergy observed between IFN-alpha and ribavirin in attaining sustained viral clearance is discussed. These studies reveal characteristics of Huh7 cells that may contribute to their unique capacity to support HCV RNA synthesis and demonstrate the utility of the replicon system for mechanistic studies on antiviral agents.
Content may be subject to copyright.
JOURNAL OF VIROLOGY, Jan. 2003, p. 1092–1104 Vol. 77, No. 2
0022-538X/03/$08.000 DOI: 10.1128/JVI.77.2.1092–1104.2003
Copyright © 2003, American Society for Microbiology. All Rights Reserved.
Antiviral Effect and Virus-Host Interactions in Response to Alpha
Interferon, Gamma Interferon, Poly(I)-Poly(C), Tumor Necrosis
Factor Alpha, and Ribavirin in Hepatitis C Virus
Subgenomic Replicons
Robert E. Lanford,
1
* Bernadette Guerra,
1
Helen Lee,
1
Devron R. Averett,
2
Brad Pfeiffer,
1
Deborah Chavez,
1
Lena Notvall,
1
and Catherine Bigger
1
Department of Virology and Immunology, Southwest National Primate Research Center, Southwest Foundation
for Biomedical Research, San Antonio, Texas 78227,
1
and Anadys Pharmaceuticals,
San Diego, California 92121
2
Received 12 August 2002/Accepted 22 October 2002
The recently developed hepatitis C virus (HCV) subgenomic replicon system was utilized to evaluate the
efficacy of several known antiviral agents. Cell lines that persistently maintained a genotype 1b replicon were
selected. The replicon resident in each cell line had acquired adaptive mutations in the NS5A region that
increased colony-forming efficiency, and some replicons had acquired NS3 mutations that alone did not
enhance colony-forming efficiency but were synergistic with NS5A mutations. A replicon constructed from the
infectious clone of the HCV-1 strain (genotype 1a) was not capable of inducing colony formation even after the
introduction of adaptive mutations identified in the genotype 1b replicon. Alpha interferon (IFN-), IFN-,
and ribavirin exhibited antiviral activity, while double-stranded RNA (dsRNA) and tumor necrosis factor alpha
did not. Analysis of transcript levels for a series of genes stimulated by IFN (ISGs) or dsRNA following
treatment with IFN-, IFN-, and dsRNA revealed that both IFNs increased ISG transcript levels, but that
some aspect of the dsRNA response pathway was defective in Huh7 cells and replicon cell lines in comparison
to primary chimpanzee and tamarin hepatocytes. The colony-forming efficiency of the replicon was reduced or
eliminated following replication in the presence of ribavirin, implicating the induction of error-prone repli-
cation. The potential role of error-prone replication in the synergy observed between IFN- and ribavirin in
attaining sustained viral clearance is discussed. These studies reveal characteristics of Huh7 cells that may
contribute to their unique capacity to support HCV RNA synthesis and demonstrate the utility of the replicon
system for mechanistic studies on antiviral agents.
Chronic hepatitis C virus (HCV) infections are one of the
leading causes of liver disease worldwide (2). The prevalence
of HCV infections is 1 to 2%, although certain geographical
regions, age groups, and ethnic groups have much higher rates
of infection (3). Although symptoms may be mild for decades,
20% of persistently infected individuals may eventually de-
velop serious liver disease including cirrhosis and liver cancer
(2). HCV infection is the leading cause for liver transplanta-
tion in the United States (12). Although the initial use of
interferon (IFN) for treatment of chronic infections yielded
marginal results, the current therapeutic regimen of pegylated
alpha 2b IFN (IFN-2b) and ribavirin provides substantially
improved rates of sustained viral clearance of 42 and 82% for
genotype 1 and genotype 2 and 3, respectively (45). Treatment
of acute infections with standard IFN therapy without ribavirin
is highly efficacious and approaches 100% sustained viral clear-
ance (33). Nonetheless, a great need exists for improved anti-
viral agents, since many patients still do not benefit from IFN
therapy, and IFN therapy is associated with undesirable side
effects. The lack of a suitable tissue culture system has previ-
ously hampered the development of antiviral agents, but the
recent development of a replicon system for HCV (43) has
partially fulfilled this need.
HCV is a member of the Flaviviridae family. The genome is
single-stranded, positive-sense RNA (Fig. 1). Since the cloning
of the viral genome (1, 11), rapid advances have been attained
in defining viral functions (reviewed in references 5 and 56).
The 5 noncoding region (NCR) contains an internal ribosome
entry site (IRES). The amino terminus of the viral polyprotein
contains the structural proteins, the capsid and two envelope
proteins, E1 and E2. The function of p7 is not known. NS2/NS3
is a metalloprotease that cleaves NS2 from NS3. NS3 is a serine
protease and the viral helicase. NS4A is a cofactor for the
serine protease. The function of NS4B is unknown. NS5B is
the viral RNA polymerase. NS5A is a phosphoprotein that
contains a sequence known as the IFN sensitivity-determining
region (ISDR). Enomoto et al. (17) first demonstrated a rela-
tionship to sequence variation in this region and resistance to
IFN therapy. Gale and colleagues have shown that this region
interacts with PKR, providing a plausible mechanism for the
modulation of the host response to IFN (22, 23). However, the
precise function of NS5A is still unknown, and whether PKR
binding accounts for viral resistance to IFN is controversial
(54, 61). NS5A induces interleukin 8 synthesis that may con-
tribute to IFN resistance (25, 55), and NS5A has been shown
* Corresponding author. Mailing address: Department of Virology
and Immunology, Southwest Foundation for Biomedical Research,
7620 NW Loop 410, San Antonio, TX 78227. Phone: (210) 258-9445.
Fax: (210) 670-3329. E-mail: rlanford@icarus.sfbr.org.
1092
to interact with grb2 (62) and a SNARE-like protein (66). E2
has been shown to interact with PKR and may be involved in
IFN resistance as well (51, 65). Following the polyprotein open
reading frame is a 3 NCR that has a variable region, a long
poly(U)-polypyrimidine stretch, and a highly conserved 98-
nucleotide terminus.
Although HCV does not replicate in conventional tissue
culture systems, a surrogate system has been created on the
basis of a bicistronic replicon constructed by Lohmann and
coworkers (43). The 5 end of the HCV genome, including the
IRES and 12 codons of the core protein, is fused in frame with
the neomycin phosphotransferase gene, the encephalomyocar-
ditis virus (EMCV) IRES drives translation of the HCV non-
structural proteins, and the construct terminates with the 3
NCR of HCV. When synthetic RNA from this construct is
transfected into the human liver cell line Huh7, G418-resistant
colonies which persistently maintain the replicon RNA can be
isolated. The utility of this system was markedly enhanced
when Blight and coworkers (8) demonstrated that adaptive
mutations arise that permit highly efcient colony formation.
This has been reproduced in a number of studies including this
one (8, 27, 35, 42). A replicon has been developed from a
FIG. 1. Schematic of HCV genome and replicon design. The HCV genome is depicted with the 5 NCR containing an IRES and the 3 NCR
including a variable region (Var), a polyuridine or polypyrimidine stretch (U/PP), and a 98-nucleotide (nt) conserved region (CR). The depiction
of the HCV IRES was adapted from the structural study of Honda et al. (28). The open reading frame of the polyprotein is depicted as a rectangle
with demarcation of the individual viral protein domains, and the positions of some of the viral functions are depicted above. c, capsid protein;
e1 and e2, envelope proteins E1 and E2. The structure of the bicistronic replicon is illustrated below the HCV genome, with the HCV IRES and
EMCV IRES regulating translation of the neomycin phosphotransferase (NEO) gene and the nonstructural proteins of HCV, respectively. A T7
promoter is fused to the HCV 5 terminus for the production of synthetic RNA. A domain of the NS5A protein is expanded to illustrate the ISDR,
PKR binding domain, and the region of the most common adaptive replicon mutations, including the deletion from amino acids 2207 to 2254
observed by Blight et al. (8). The amino acid changes for the adaptive mutations detected in the resident replicons isolated from 10 independent
G418-resistant colonies are indicated at the bottom of the gure, and the sites of NS5A hyperphosphorylation (amino acids 2197, 2201, and 2204)
are indicated by arrows. WT, wild type.
V
OL. 77, 2003 ANTIVIRAL AGENTS AND HCV REPLICONS 1093
second genotype 1b strain (27, 32) that did not require adap-
tive mutations for high colony-forming efciency (32), and
recently replicons containing the full-length genome have been
developed (32, 53).
In this study, we have utilized the replicon system for anal-
ysis of several compounds for potential antiviral effect, includ-
ing IFN-, IFN-, tumor necrosis factor alpha (TNF-),
poly(I)-poly(C), and ribavirin. Although no antiviral effect was
observed with TNF- or poly(I)-poly(C), a synthetic double-
stranded RNA (dsRNA) and known inducer of IFN and IFN-
stimulated genes (ISGs), both IFN- and IFN- exhibited an-
tiviral effects. The antiviral effect of ribavirin in this system
could be ascribed to the induction of error-prone replication
similar to recent ndings with GB virus B (GBV-B) (38), a
surrogate model for HCV. The expression levels for a number
of ISGs were monitored before and after antiviral treatments
to begin a characterization of the virus-host interactions in-
volved in this system.
MATERIALS AND METHODS
Replicon constructs. Rep1bNeo was constructed from synthetic oligonucleo-
tides to be an exact copy of the replicon described by Lohmann et al. (43)
(GenBank accession no. AJ242652). The Rep1aNeo construct is identical to the
Rep1bNeo construct except that all HCV components were derived from the
HCV-1 infectious clone (40) (GenBank accession no. AF271632) including the 5
NCR, the region encoding NS3-NS5B, and the 3 NCR. The Rep1b/aNeo con-
struct was derived from Rep1aNeo by substituting 228 nucleotides (positions
1927 to 2155) from Rep1bNeo into Rep1aNeo, which altered 12 of the rst 73
amino acids of NS3. Adaptive mutations were detected by direct sequencing of
PCR products from NS3, NS5A, and NS5B which spanned nucleotides 3414 to
5312, 6828 to 7118, and 8855 to 9029, respectively. The specic adaptive muta-
tions NS5A-S2204I and NS3-D1431Y were introduced into replicons using PCR-
directed mutagenesis. Synthetic replicon RNA was prepared from DNA linear-
ized with ScaI (Rep1bNeo) and XbaI (Rep1aNeo and Rep1b/aNeo) using the T7
Megascript kit (Ambion, Austin, Tex.) and was puried by DNase treatment,
RNazol (Leedo, Houston, Tex.) extraction, and ethanol precipitation. RNA was
quantied by optical density, and the concentration and quality were conrmed
by agarose gel electrophoresis.
Cells and transfections. Huh7 cells were cultivated in a 1:1 mixture of Dul-
beccos modied Eagles medium and Hams F12 medium supplemented with
10% fetal bovine serum (FBS) and 50 g of gentamicin sulfate per ml. RNA
transfections were performed using DMRIE-C transfection reagent (GIBCO/
BRL, Rockville, Md.) at a ratio of 5 g of lipid to 1 g of RNA. Replicons with
low or no colony-forming efciency were transfected using 10 g of replicon
RNA per 100-mm-diameter culture dish; however, replicons with adaptive mu-
tations were transfected at 1 to 0.01 g of replicon RNA diluted into Huh7 RNA.
The RNA and lipid were diluted individually into 4.5 ml of medium without FBS,
combined, incubated for 15 min at room temperature, and added to the cultures.
Cultures were washed three times with medium without FBS prior to transfec-
tion and two times with medium containing FBS after transfection. Cultures were
exposed to the RNA-lipid mixture for6hat37°C. Culture medium was supple-
mented with 250 g of G418 per ml beginning 1 day after transfection. This
transfection protocol routinely yielded approximately 45% transfection efciency
for RNA, as evaluated by transfection of a Sindbis virus replicon expressing
-galactosidase (8). Measurement of replicon copy number and antiviral studies
were conducted with cultures 48 h after plating (cultures were approximately
60% conuent). None of the antiviral treatments employed induced noticeable
toxicity by microscopic inspection of the cultures or by measurement of glycer-
aldehyde-3-phosphate dehydrogenase (GAPDH) transcript levels. This was true
even when treatment was extended for 7 days. Primary chimpanzee and tamarin
hepatocytes were cultivated in a hormonally dened, serum-free medium as
previously described (39).
TaqMan quantitative RT-PCR for replicon RNA and host transcripts. Total
cell RNA was isolated from cell cultures using RNazol (Leedo). Replicon RNA
was quantied by a real-time, 5 exonuclease reverse transcription-PCR (RT-
PCR) (TaqMan) assay (40) using the ABI 7700 sequence detector (Perkin-Elmer
Biosystems, Foster City, Calif.). The primers and probe were derived from the 5
NCR and were selected using the Primer Express software designed for this
purpose (Perkin-Elmer Biosystems). The forward primer contains nucleotides
149 to 167 (5-TGCGGAACCGGTGAGTACA-3), the reverse primer contains
nucleotides 210 to 191 (5-CGGGTTTATCCAAGAAAGGA-3), and the probe
contains nucleotides 189 to 169 (5-CCGGTCGTCCTGGCAATTCCG-3). The
uorogenic probe was labeled with FAM (6-carboxyuorescein) and TAMRA
(6-carboxytetramethylrhodamine) and was obtained from Synthegen (Houston,
Tex.). The primers and probe were used at 10 pmol per 50-l reaction mixture.
The reactions were performed using the Brilliant Plus Single Step RT-PCR kit
(Stratagene, La Jolla, Calif.) and included a 30-min 48°C RT step, followed by 10
min at 95°C, and then 40 cycles of amplication using the universal TaqMan
RT-PCR standardized conditions: 15 s at 95°C for denaturation and 1 min at
60°C for annealing and extension. The standards used to establish genome
equivalents (ge) were synthetic RNAs transcribed from a clone of the 5 NCR of
the HCV-1 strain (40). Synthetic RNA was prepared using the T7 Megascript kit
and was puried by DNase treatment, RNazol extraction, and ethanol precipi-
tation. RNA was quantied by optical density, and 10-fold serial dilutions were
prepared from 1 million to 10 copies using tRNA as a carrier. These standards
were run in all TaqMan RT-PCR assays in order to calculate the number of ge
in the experimental samples. The conditions for quantication of transcripts
from ISGs were identical to those described above for replicon RNA. Most
assays were multiplexed using GAPDH. The primer and probe sets used in these
assays will be presented elsewhere (C. Bigger, B. Guerra, K. M. Brasky, G. B.
Hubbard, M. Beard, and R. E. Lanford, submitted for publication).
Antiviral treatments. Poly(I)-poly(C) was obtained from ICN (Costa Mesa,
Calif.) and Sigma (St. Louis, Mo.). A highly puried preparation of ribavirin
(1--
D-ribofuranosyl-1H-1,2,4-triazole-3-carboximide) was a gift from Schering
Plough Research Institute (Kenilworth, N.J.). Some commercially available
preparations of ribavirin contain trace contaminants that are often toxic in tissue
culture studies. This preparation was specically puried and selected for low
toxicity in vitro. Human IFN--2b (intron A) was obtained from Schering Plough
Research Institute. IFN- was obtained from R&D Systems (Minneapolis,
Minn.). Antiviral treatments with IFN-, IFN-, and poly(I)-poly(C) were initi-
ated using subconuent cultures, but after 24 h, the cultures reached conuency.
Treatment with ribavirin used conuent cultures to reduce the adverse effects of
ribavirin seen on rapidly dividing cells.
In vitro translation. In vitro translations were preformed with 1 g of replicon
RNA prepared as described above, heated to 70°C for 10 min, and cooled on ice.
Translation reactions were performed in a total volume of 25 l using the
Promega (Madison, Wis.) nuclease-treated rabbit reticulocyte lysate systems
supplemented to contain 90 mM KCl to enhance for IRES-driven translation.
Reaction mixtures contained 20 M amino acids lacking methionine, 10 Ci of
[
35
S]methionine (Express
35
S; Perkin-Elmer, Boston, Mass.), and RNasin (Pro
-
mega). Reactions were performed for1hat30°C and the products were exam-
ined directly by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-
PAGE).
PKR assays. For PKR assays, Huh7 cells were pretreated with 100 gof
poly(I)-poly(C) per ml for2horwith 1,000 U of IFN-2b per ml for 16 h or left
untreated. Cultures were washed twice in phosphate-buffered saline (PBS), and
cell lysates were prepared using PEB (PBS containing 1% Nonidet P-40 [NP-40],
10% glycerol, 1 mM dithiothreitol, and protease inhibitors) and were claried in
a microcentrifuge for 20 min at 4°C. PKR was immunoprecipitated for2hat4°C
using protein G beads containing 5 g of monoclonal antibody clone 13 to PKR
(BD Biosciences, San Diego, Calif.). The beads were washed ve times with PEB
and once with PKR buffer (20 mM Tris [pH 7.5], 0.1 mM EDTA, 80 mM KCl,
5% glycerol, 2 mM MgCl
2
, 2 mM MnCl
2
, and 0.2 mg of bovine serum albumin
per ml). Kinase reactions were conducted for 15 min at 30°C with PKR bound to
the beads in 50 l of PKR buffer supplemented with 5 Ci of [-
32
P]ATP
(Perkin-Elmer). PKR was eluted from the beads with SDS gel sample buffer and
analyzed by SDS-PAGE and phosphorimage analysis.
Northern blot hybridization. For analysis of replicon RNA by Northern blot
hybridization, total cellular RNA was isolated from replicon lines using RNazol.
RNA was analyzed by electrophoresis on a 1% agarose-formaldehyde gel at 30
Vfor16hat16°C and was transferred to Gene Screen Plus hybridization transfer
membranes (Perkin-Elmer) using downward capillary transfer. Membranes were
prehybridized in SDS hybridization buffer {6 SSPE [1 SSPE is 0.18 M NaCl,
10 mM NaH
2
PO
4
, and 1 mM EDTA (pH 7.7)], 10% SDS, 200 g of salmon
sperm DNA per ml, and 50% formamide} for5hat59°C and were hybridized
in the same buffer containing 10
6
counts of
32
P-labeled riboprobe per ml for 16 h
at 59°C. Riboprobes were prepared from a linearized vector containing the
neomycin phosphotransferase gene downstream of a T7 RNA polymerase pro-
moter using the Promega Riboprobe System as described by the manufacturer.
Membranes were washed twice with 1 SSC (1 SSC is 0.15 M NaCl plus 0.015
M sodium citrate) with 0.1% SDS at 23°C for 30 min and twice with 0.1 SSC
1094 LANFORD ET AL. J. VIROL.
with 0.1% SDS at 65°C for 30 min. Membranes were analyzed by phosphorimage
analysis and autoradiography.
RESULTS
Isolation of cell lines with genotype 1b replicons with adap-
tive mutations. Our studies were initiated in a manner previ-
ously described using an exact copy of the original genotype 1b
replicon described by Lohmann et al. (43) (designated
Rep1bNeo in these studies). Huh7 cells were transfected with
synthetic RNA produced from the construct using DMRIE-C,
a RNA transfection reagent. The colony-forming efciency of
this replicon in the presence of G418 selection was frequently
zero and never exceeded ve colonies per g of transfected
RNA. Ten colonies from a single experiment were expanded
for further analysis. Replicon RNA was quantied using a
TaqMan, real-time RT-PCR assay with a primer and probe set
directed at the HCV 5 NCR. The level of replicon RNA
varied from 1.7 10
6
to 1.3 10
7
ge/g of cellular RNA
(Table 1) or approximately 17 to 130 ge/cell, assuming that
copy number is uniform in cells of the same line. Clone 45 cells
were plated in identical plates at 60% conuency and analyzed
at multiple times over a 10-day period. Although a 50% decline
in replicon RNA (number of ge per microgram of cell RNA)
occurred between days 1 and 2, no further decline was ob-
served over 10 days, suggesting that the differences in replicon
number in different cell lines was not due to minor variations
in culture conuency (data not shown). Three regions (a por-
tion of NS3, NS5A, and NS5B) of the replicon present in each
cell line were PCR amplied and sequenced to determine the
consensus sequence of the resident replicon. Previously, most
adaptive mutations have been detected in a region adjacent to
the ISDR of NS5A (Fig. 1) (8, 27, 35, 42); however, adaptive
mutations have also been detected in NS3, NS4B, and NS5B
(27, 35, 42). In this study, the replicon present in each cell line
contained a single mutation within NS5A, and 2 of the 10
replicons contained mutations in NS3, while no mutations were
noted within NS5B. The NS5A mutations spanned nucleotides
6861 to 6951 and amino acids 2177 to 2204 (nucleotide and
amino acid numbers are based on HCV genomic RNA se-
quence) (Fig. 1 and Table 1). A total of six unique mutations
were observed; two were present in two different replicons
each (S2197F and S2204I), and one was present in three rep-
licons (S2202L). One aspect that was noteworthy was the mu-
tation of two of the serine residues associated with hyperphos-
phorylation of NS5A (amino acids 2197 and 2204) (64); 4 of
the 10 mutations detected involved these two serine residues
(Fig. 1 and Table 1). Clones 35 (NS5A S2204I) and 38 (NS5A
A2199T) also carried mutations in NS3 at I1278L and Q1335E,
respectively. In addition, a mutation at NS3 D1431Y was ob-
served in combination with NS5A S2204I in a cell line not
shown in Table 1. Transfection of total cellular RNA derived
from each cell line resulted in a much higher colony-forming
efciency than the parental replicon, suggesting that the
changes were adaptive mutations. Extrapolation of the Taq-
Man RT-PCR results for replicon copy number (number of ge
per microgram of cellular RNA) to the number of colonies
formed per microgram of transfected cell RNA revealed that
colony-forming efciencies ranged from 52,000 to 500,000 col-
onies/g of replicon RNA (Table 1). Since limited regions of
each replicon were sequenced, the presence of additional
adaptive mutations cannot be excluded.
Synergy of NS3 mutations with NS5A mutations and failure
of adaptive mutations to enhance the colony-forming efciency
of genotype 1a replicons. To construct a Rep1bNeo replicon
with high colony-forming efciency, one of the common NS5A
mutations at amino acid position 2204 (serine to isoleucine;
S2204I) was introduced into Rep1bNeo by PCR-directed mu-
tagenesis. While the parental replicon often yielded no colo-
nies in individual assays, Rep1bNeo-2204 produced up to 800
colonies/g of replicon RNA. Although a signicant improve-
ment over the parental replicon, the transfection of total cell
RNA (Table 1) from selected colonies suggested that addi-
tional mutations must account for the very high efciency ob-
served in the adapted replicons. Next, we tested the NS3 mu-
tation at amino acid D1431Y by itself and in concert with the
S2204I mutation. Rep1bNeo-1431 exhibited no signicant im-
provement in colony-forming efciency over the parental rep-
licon, while the NS3 mutation was highly synergistic with the
NS5A mutation (Rep1bNeo-1431/2204 [Table 2]), producing
up to 80,000 colonies/g of replicon RNA. This colony-form-
ing efciency is comparable to that determined for the repli-
cons present in selected colonies and suggests that NS3 muta-
tions complement NS5A mutations in a nonadditive
mechanism.
Previous studies have suggested that genotype 1a replicons
are not capable of forming colonies in Huh7 cells at detectable
levels (8, 27). Since these studies were conducted with a single
strain of genotype 1a (H77), it is not clear whether this is a
general property of genotype1a isolates or unique to this
strain. We constructed genotype 1a replicons using our infec-
tious clone of the HCV-1 prototype sequence (40), the only
other genotype 1a infectious clone. The construct design was
identical to that used for Rep1bNeo, except that all HCV
TABLE 1. NS5A adaptive mutations of Rep1bNeo
a
Clone
Replicon RNA
(10
6
ge/gof
cell RNA)
b
Amino acid
c
Colony-forming
efciency
(10
3
colonies/gof
replicon RNA)
d
Position Change
2 1.7 2177 D3H 500
42 10.1 2197 S3F79
45 11.9 2197 S3F87
38 9.5 2199 A3T 167
40 9.2 2200 S3R53
13 7.4 2202 S3L94
39 1.8 2202 S3L 154
43 5.3 2202 S3L52
35 13.1 2204 S3I72
37 2.6 2204 S3I 147
a
Huh7 cells were transfected with Rep1bNeo, and G418-resistant colonies
were established as cell lines.
b
The replicon copy number was determined by TaqMan RT-PCR and ex
-
pressed as ge per microgram of cellular RNA. To determine the number of ge
per microgram of RNA, cells were plated at a low density and harvested 48 h
later, when the cultures were still subconuent.
c
Regions of NS3, NS5A, and NS5B were amplied and sequenced to deter
-
mine adaptive mutations. The amino acids are given in the single-letter code.
d
To determine the colony-forming efciency of the adapted replicons, Huh7
cells were transfected with 10 g of total cell RNA isolated from each cell line,
and the colony-forming efciency in the presence of G418 was determined. The
colony-forming efciency is presented as the number of colonies per microgram
of transfected replication RNA that was extrapolated from the estimated number
of replicon ge per microgram of cell RNA.
VOL. 77, 2003 ANTIVIRAL AGENTS AND HCV REPLICONS 1095
sequences were derived from the HCV-1 clone. The resulting
replicon RNA, Rep1aNeo, was not capable of inducing colony
formation following transfection into Huh7 cells in multiple
attempts (Table 2). Analysis of the in vitro translation pattern
from this RNA in rabbit reticulolysates revealed that expres-
sion of the HCV nonstructural proteins from the EMCV IRES
was very low in comparison to Rep1bNeo, despite the fact that
translation of the neo gene from the HCV IRES was equivalent
for the two constructs (Fig. 2A). A series of chimeric constructs
revealed that a short sequence at the amino terminus of NS3
adjacent to the EMCV IRES was responsible for the differ-
ences in translation efciencies of the two constructs. To pro-
vide more-efcient translation of the HCV nonstructural pro-
teins from the Rep1aNeo replicon, a portion (228 nucleotides)
of the genotype 1a sequence was replaced with the genotype 1b
sequence to create Rep1b/aNeo, which differed from
Rep1aNeo by 12 amino acids (see Materials and Methods for
details). Presumably, a RNA structure within the genotype 1a
sequence inhibited translation from the EMCV IRES. Al-
though this chimeric construct exhibited in vitro translation
properties equivalent to Rep1bNeo (Fig. 2A), it was not capa-
ble of inducing colony formation in Huh7 cells (Table 2). Since
the Rep1bNeo replicon itself exhibited marginal colony-form-
ing ability prior to acquisition of adaptive mutations, the
NS5A-S2204I mutation was introduced alone or in combina-
tion with D1431Y into Rep1b/aNeo. These mutations did not
increase the colony-forming efciency of Rep1b/aNeo to de-
tectable levels following transfection of Huh7 cells. These re-
sults suggest that replicons based on the HCV-1 genotype 1a
sequence do not possess the capacity to induce detectable
levels of colony formation in Huh7 cells even with mutations
known to enhance genotype 1b strains. Part of our analysis of
these constructs included in vitro translation, as shown in Fig.
2A. When the replicons containing the NS3 mutations were
compared to the parental replicons and those containing the
NS5A mutation alone, NS3 exhibited a signicantly increased
mobility by SDS-PAGE. This was true whether the D1431Y
mutation was present in the Rep1bNeo or Rep1b/aNeo back-
ground and whether the S2204I mutation was present or not
(Fig. 2B). At this time, the reason for the increased mobility of
NS3 is not understood, but it may be indicative of an altered
posttranslational modication, which in turn may play a role in
the synergy of this mutation with the NS5A mutation.
Antiviral effects of IFN-, IFN-, poly(I)-poly(C), and
TNF-. The potential antiviral activities of IFN-, IFN-,
poly(I)-poly(C), and TNF- were examined using the replicon
system. In these studies, a replicon cell line was treated with
various concentrations of the compound, and the level of rep-
licon RNA in cell lysates was quantied using TaqMan RT-
PCR assays. All assays were multiplexed for GAPDH mRNA
to ensure that the treatment had no overt adverse effect on
cellular mRNA levels. Initially, a time course study was per-
formed with each antiviral agent. Clone 45 cultures (Table 1)
were treated for 24, 48, and 72 h with 100 to 1,000 U of
IFN-2b or IFN- per ml, 100 to 1,000 g of poly(I)-poly(C)
per ml, or 200 to 1,000 U of TNF- per ml. Dramatic reduc-
tions in replicon RNA levels were observed at both concentra-
TABLE 2. Colony-forming efciency of replicon constructs
a
Replicon RNA
Colony-forming efciency
(colonies/g of replicon RNA)
Rep1bNeo ....................................................................... 5
Rep1bNeo-2204.............................................................. 800
Rep1bNeo-1431.............................................................. 0
Rep1bNeo-1431/2204..................................................... 80,000
Rep1aNeo ....................................................................... 0
Rep1b/aNeo .................................................................... 0
Rep1b/aNeo-2204........................................................... 0
Rep1b/aNeo-1431/2204.................................................. 0
a
Huh7 cells were transfected with various replicons, and the colony-forming
efciency in the presence of G418 was determined. Replicons were transfected at
10 g/100-mm-diameter dish, except for Rep1bNeo-1431/2204, which was trans-
fected at 10 ng of replicon RNA mixed with 10 g of Huh7 RNA as carrier. Each
replicon was tested in multiple experiments. Rep1bNeo frequently failed to yield
colonies, but in a single experiment yielded 50 colonies/10 g of transfected
RNA, and these colonies were the source of the cell lines in Table 1.
FIG. 2. In vitro translation of bicistronic HCV replicons.
(A) Rep1aNeo (1a), Rep1b/a (1b/a), and Rep1b (1b) synthetic RNAs
were produced by in vitro transcription reactions utilizing vectors lin-
earized at EcoRI, a site downstream of the neomycin phosphotrans-
ferase gene (Fig. 1), or at XbaIorScaI to produce full-length replicon
RNA. RNAs were in vitro translated using rabbit reticulocyte lysates in
the presence of [
35
S]methionine, and the products were analyzed by
SDS-PAGE. Rep1a produced reduced levels of NS3 compared to both
Rep1b and Rep1b/a, while translation of the Neo protein was similar
for all constructs. (B) The impact of the NS3 D1431Y mutation on the
migration of NS3 by SDS-PAGE was examined following in vitro
translation. All replicon RNAs containing the D1431Y mutation ex-
hibited an increased mobility for NS3 regardless of whether the mu-
tation was in the presence of the NS5A S2204I mutation or in the
Rep1b (1b) or Rep1a (1a) background. wt, wild type.
1096 LANFORD ET AL. J. V
IROL.
tions of IFN-2b at 72 h, i.e., 40- and 172-fold reduction at 100
and 1,000 U/ml, respectively (Fig. 3). Although substantial
antiviral activity was also observed with IFN-, the reduction in
replicon RNA was approximately three- to fourfold less (14.4-
and 44.9-fold reduction at 100 and 1,000 U/ml, respectively, at
72 h). No signicant reduction was observed with poly(I)-
poly(C), even at 1,000 g/ml (Fig. 3), or with TNF- at 1,000
U/ml (data not shown).
In part to conrm these ndings and to determine the max-
imum reduction observed with each antiviral following a more
prolonged treatment, cultures were treated for 7 to 9 days with
each antiviral. Replicon levels were reduced by 2,300- and
23,000-fold at 100 and 1,000 U/ml of IFN-2b, respectively
(Table 3). This was not an unanticipated nding, since other
studies have observed a reduction in replicon RNA following
IFN- treatment (8, 20, 27). However, extended treatment
with poly(I)-poly(C), a known inducer of dsRNA response
genes including type 1 IFN genes and ISGs, still had no anti-
viral effect. This was an unanticipated nding, since poly(I)-
poly(C) has a high level of antiviral activity for many viruses,
including GBV-B replication in primary tamarin hepatocytes
(38), which is a closely related surrogate model for HCV.
Poly(I)-poly(C) treatment was performed by direct addition of
the compound to the medium at concentrations as high at
1,000 g/ml, as well as transfection with lipid reagents without
detectable antiviral effect. Although IFN- again exhibited an-
tiviral activity, the prolonged treatment did not result in sig-
nicantly greater reduction of replicon RNA than treatment
for 72 h (100 or 1,000 U/ml of IFN- for 7 days resulted in a
2.7- and 31.9-fold reduction in replicon RNA, respectively).
Host cell response to antiviral treatments. To further ex-
plore the differences in response to IFN- and poly(I)-poly(C),
induction of dsRNA-IFN-stimulated gene was examined using
TaqMan RT-PCR assays for the transcripts of STAT1,
STAT1, PKR, ISG12, IRF-1, and IP10. Huh7 cells were com-
pared to replicon cells following 24 to 48 h of treatment with
IFN-, IFN-, or poly(I)-poly(C) (Table 4). In this experiment,
IFN- resulted in 11.8- and 104-fold reduction in replicon
RNA at 24 and 48 h, respectively, while IFN- resulted in 6.1-
and 31.2-fold reduction in replicon RNA at 24 and 48 h.
Poly(I)-poly(C) again failed to induce a signicant decline in
replicon RNA. All of the ISGs were elevated in IFN--treated
Huh7 and replicon cells; however, IRF-1 was increased by only
2.3- to 3.3-fold, and it is not clear that these values represent
signicant increases over the baseline levels of the untreated
cultures.
Although the basal levels of the ISGs were similar in both
FIG. 3. Kinetics of HCV replicon RNA decline following treat-
ment with IFN-, IFN-, and poly(I)-poly(C). Clone 45 cells were
treated with IFN- or IFN- at 100 or 1,000 U/ml and with poly(I)-
poly(C) at 100 or 1,000 g/ml and harvested at 24, 48, and 72 h. Results
from 100 (A) and 1,000 (B) U/ml or g/ml. Replicon RNA levels were
quantied by TaqMan RT-PCR and expressed as the fold change from
the levels of untreated cultures harvested at the same time points. All
values are averages from duplicate cultures.
TABLE 3. Antiviral effects of IFN-, poly(I)-poly(C), and IFN-
a
Treatment Fold reduction
b
GAPDH Ct
c
IFN- (U/ml)
0 0 18.0
100 2,300 17.6
1,000 23,000 17.6
Poly(I)-poly(C) (g/ml)
0 0 16.8
50 1.4 16.5
IFN- (U/ml)
0 0 15.9
100 2.7 15.6
1,000 31.9 15.2
a
Replicon lines were treated with IFN-, poly(I)-poly(C), or IFN- for7to9
days. In this experiment, clone 8 cells were treated with IFN-, while clone 45
cells were treated with IFN- and poly(I)-poly(C). Treatments were initiated on
conuent cultures, and the medium was changed every 2 days. No overt toxicity
was observed.
b
Replicon RNA was quantied by TaqMan RT-PCR, and values are ex
-
pressed as fold reduction compared to the values for untreated cells.
c
The TaqMan RT-PCR assay for replicon RNA was multiplexed for the
GAPDH mRNA to demonstrate a lack of effect of treatments on cellular
mRNA, and all values were normalized for GAPDH. GAPDH values are ex-
pressed as Ct, the amplication cycle at which the values exceeded the back-
ground threshold.
VOL. 77, 2003 ANTIVIRAL AGENTS AND HCV REPLICONS 1097
Huh7 cells and clone 45 cells, some variation was noted in the
degree of induction of ISGs in the two cell types following
treatment with IFN- and IFN-. While PKR was induced to
similar levels in the presence and absence of the replicon and
STAT1 and STAT1 were increased approximately twofold
more in the absence of the replicon, ISG12 and IP10 induction
by IFN- was increased by approximately 8- to 10-fold in the
presence of the replicon (Table 4). The greater induction of
ISG12 and IP10 by IFN- in the presence of the replicon was
not observed following treatment with IFN-. As expected,
some of the IFN response genes were induced differentially by
IFN- and IFN-. While IRF-1 was minimally induced by
IFN- in comparison to IFN-, the opposite was true for PKR.
The increased induction of ISG12 and IP10 by IFN- in
clone 45 cells in comparison to Huh7 cells could be due to
clonal variation of individual cells selected from the Huh7
population; alternatively, it could represent an important dif-
ference induced by either viral dsRNA or an HCV protein. To
determine the degree of variation in the IFN induction of ISGs
in Huh7 cells and different replicon lines, clones 2, 40, and 45
were compared to Huh7 cells following 24 h of treatment with
IFN- and IFN-. Stat1, ISG12, and IP10 transcripts were
examined by TaqMan RT-PCR assays. Although in all cases
the transcripts were induced in both Huh7 and replicon cell
lines in the presence of IFN- or IFN-, the differential induc-
tion of IP10 and ISG12 in cells containing the replicon in
comparison to Huh7 cells was less apparent. The induction of
IP10 transcripts in IFN--treated cells was greater in all rep-
licon lines than in Huh7 cells. In two lines (clones 40 and 45),
it was again increased by 10-fold more than in Huh7 cells, but
in a third line (clone 2), IP10 transcripts were induced only
about 2.5-fold more than in Huh7 cells (Table 5). No pattern
was observed for either Stat1 or ISG12, but in general, rep-
licon lines had higher levels of ISG induction in the presence
of IFN-. Interestingly, the pattern for IP10 in the two exper-
iments remained constant with regard to Huh7 cells and clone
45 cells treated with IFN- and IFN-. IP10 transcripts were
induced to 10-fold-greater levels by IFN- in clone 45 cells in
comparison to Huh7 cells, while IFN- induced IP10 to a
greater extent in Huh7 cells than in clone 45 cells (Table 5).
These data indicate that (i) ISG induction occurs in the pres-
ence of both IFN- and IFN-, (ii) some differential regulation
of ISGs is observed between the two IFNs, and (iii) although
some differences were observed in the presence and absence of
the replicon, some of the variation may be due to the clonal
selection of cell lines.
In contrast to IFN- and IFN-, poly(I)-poly(C) treatment
did not induce the expression of ISGs in either Huh7 cells or
clone 45 cells (Table 4). These data suggest that the parental
Huh7 cells are altered in regulation of some component of the
dsRNA response pathway, which may explain the lack of an-
tiviral effect by poly(I)-poly(C) on the replicon in these cells.
This contrasts sharply with primary chimpanzee hepatocytes,
which were highly responsive to poly(I)-poly(C). Primary chim-
panzee hepatocytes were treated with either IFN- or poly(I)-
poly(C) and analyzed for the induction of ISGs. In chimpanzee
hepatocytes, 50 g of poly(I)-poly(C) per ml was superior to
100 U of IFN- per ml in the induction of ISGs (Table 4). We
had previously demonstrated the potent antiviral properties of
poly(I)-poly(C) in the GBV-B/tamarin hepatocyte model.
Since this system permitted the direct comparison of two cell
types in which poly(I)-poly(C) had either no antiviral activity
against HCV replicons or potent antiviral activity against an
HCV surrogate, we examined poly(I)-poly(C) treatment in
Huh7 cells and primary tamarin hepatocytes. TaqMan RT-
PCR assays were used to quantify the level of induction of
IFN- transcripts at 2, 7, and 24 h after poly(I)-poly(C) addi-
tion. We chose IFN-, since it is known to be responsive to
dsRNA and has potent antiviral activity. No induction of
IFN- transcripts was noted in Huh7 cells, while 12- to 18-fold
induction occurred in tamarin hepatocytes at all three time
TABLE 5. Induction of ISGs by IFN- and IFN- in multiple
replicon lines
Treatment
and cell
a
Fold change
b
Stat-1 ISG12 IP10 HCV
IFN-
Huh7 4.5 104.5 3.0
Clone 2 6.9 403.6 7.6 9.5
Clone 40 7.0 275.0 52.9 8.1
Clone 45 4.8 130.6 43.0 4.8
IFN-
Huh7 8.4 25.0 227.3
Clone 2 8.0 47.1 33.1 7.5
Clone 40 8.2 21.9 109.7 8.7
Clone 45 6.5 15.3 75.2 4.7
a
Huh7 and clone 2, 40, and 45 cells were treated for 24 h with IFN- or IFN-
at 1,000 U/ml.
b
Levels of HCV replicon RNA and transcripts for STAT-1, IP10, and ISG12
were quantied by TaqMan RT-PCR assays. The data are averages from dupli-
cate cultures and are expressed as the fold change compared to the values for
untreated cultures.
TABLE 4. Induction of ISG transcripts by IFN-, IFN-, and
poly(I)-poly(C)
a
Cell
b
Treatment
Fold change
c
STAT1 STAT1 PKR ISG12 IRF-1 IP10
Huh7 IFN- 41.1 38.8 20.3 312 3.1 10.0
Rep IFN- 18.6 24.1 18.7 2,374 3.3 111.5
Huh7 IFN- 80.8 70.6 3.4 80.0 47.9 198.8
Rep IFN- 31.3 33.5 4.5 74.3 45.5 123.1
Huh7 Poly(I)-poly(C) 1.3 1.3 1.0 1.6 1.2 1.6
Rep Poly(I)-poly(C) 1.1 1.1 1.5 2.2 1.1 1.3
CH IFN 2.0 3.6 3.8 9.3 1.4 12.6
CH Poly(I)-poly(C) 3.7 8.9 5.3 40.4 2.2 549.0
a
Huh7 and clone 45 cell lines were treated with IFN- (1,000 U/ml), polyIC
(100 g/ml), or IFN- (1,000 U/ml). Primary chimpanzee hepatocytes were
treated with IFN- (100 U/ml) or polyIC (50 g/ml).
b
Rep, replicon clone 45 cell line; CH, chimpanzee hepatocytes.
c
Transcripts for various ISGs were quantied using TaqMan RT-PCR assays.
The data are averages from duplicate cultures and are expressed as the fold
change compared to the values for untreated cultures. The data for IFN- and
IFN- treatment are from the 24-h harvest, since this time point had the highest
values. For polyIC, similar values were obtained for the 24- and 48-h time points.
The 48-h values are shown to demonstrate that a delayed response was not
missed. The chimpanzee hepatocytes were harvested at 72 h for both treatments.
Additional cells were not available to repeat the chimpanzee hepatocyte study at
multiple time points. The level of the replicon RNA was decreased by 104- and
31.2-fold in IFN-- and IFN--treated cells at the 48-h time point. No decrease
in replicon RNA was observed following polyIC treatment.
1098 LANFORD ET AL. J. VIROL.
points (Fig. 4). Since numerous different approaches to induc-
tion of ISG transcription with poly(I)-poly(C) were attempted
in Huh7 cells without success, we questioned whether poly(I)-
poly(C) was entering the cells and inducing latent dsRNA
binding proteins. To test this, Huh7 cells were treated with
poly(I)-poly(C) for2horwith IFN- for 16 h, and PKR was
immunoprecipitated and used in kinase assays. In the presence
of dsRNA, PKR dimerizes and undergoes autophosphoryla-
tion, and PKR kinase activity can also be induced by IFN-.In
comparison to untreated Huh7 cells, both poly(I)-poly(C) and
IFN- dramatically induced PKR kinase activity (Fig. 5).
These data suggest that Huh7 cells are responsive to poly(I)-
poly(C) but that they do not respond in a typical fashion to
dsRNA, since no antiviral activity or induction of ISG tran-
scription was apparent using this compound.
Ribavirin induces error-prone replication of HCV replicon
RNA. Ribavirin has an antiviral effect for a number of viruses
in tissue culture (14, 47, 58). The monophosphate form of
ribavirin is an IMP dehydrogenase (IMPDH) inhibitor, and at
least some of the in vitro antiviral effect has been ascribed to
the reduction of GTP pools by inhibition of IMPDH. Although
this has deleterious effects on cellular metabolism, these effects
are tolerated in short-term studies in nondividing cell cultures.
Ribavirin is used in combination with IFN- for the treatment
of HCV infections and provides a signicant improvement in
the rate of sustained viral clearance in comparison to the rate
for IFN- alone. However, ribavirin monotherapy does not
result in a signicant reduction in the level of viremia despite
a marked improvement in liver disease (9, 16). One mechanism
suggested to explain the improvement with ribavirin therapy
involves an immunomodulatory activity possessed by ribavirin
that promotes a Th1-biased immune response (19, 30, 50, 60).
However, the mechanism of the synergistic effect that ribavirin
has with IFN- for increasing the percentage of patients with
sustained viral clearance remains uncertain. We recently dem-
onstrated a direct antiviral effect of ribavirin on GBV-B virus
that involves induction of error-prone replication due to riba-
virin triphosphate incorporation (38). The following experi-
ments were conducted to determine whether ribavirin has an
antiviral effect on the HCV replicon and whether this effect
could be ascribed to induction of error-prone replication.
Experiments with ribavirin were conducted with conuent
nondividing cultures to minimize the adverse effects from GTP
pool reduction. In addition, a highly puried preparation of
ribavirin was used; the preparation was devoid of trace con-
taminants that often cause toxicity in tissue culture studies
unrelated to the IMPDH inhibition by ribavirin (see Materials
and Methods). Initially, the level of ribavirin was titrated to
determine the optimal dose for antiviral effect with minimal
cellular toxicity. A replicon cell line was treated with 50 to 400
M ribavirin for 9 days, and total cell RNA was examined by
TaqMan RT-PCR assay to determine the effects on replicon
RNA and GAPDH mRNA levels. Little effect was observed at
50 or 100 M, while treatment with 200 and 400 M ribavirin
resulted in 19.3- and 2,900-fold reduction in replicon RNA,
respectively (Table 6). No overt toxicity or decrease in the level
of GAPDH mRNA was observed at any treatment level, but
some decline in the secretion of apolipoprotein B was observed
at the higher ribavirin concentrations (data not shown). A time
course study revealed that treatment of cultures with 400 M
ribavirin reduced replicon RNA from 2.2-fold at 24 h to 24.3-
fold at 72 h when analyzed by TaqMan RT-PCR (Fig. 6C).
Analysis of the same cellular RNAs for replicon RNA levels by
Northern blot hybridization and quantication by phosphorim-
ager analysis of the blot provided essentially identical results
for the percent decrease in replicon RNA in ribavirin-treated
cultures (Fig. 6B). Since only the area of the gel representing
intact replicon RNA is used in phosphorimage analysis, these
results suggest that the replicon RNA was not degraded in
ribavirin-treated cultures. The autoradiogram of the Northern
FIG. 5. PKR activation by poly(I)-poly(C) in Huh7 cells. Huh7
cells were treated with 100 g of poly(I)-poly(C) per ml for2horwith
1,000 U of IFN- per ml for 16 h prior to harvest. PKR was immuno-
precipitated from cell lysates, and kinase reactions were conducted
with [-
32
P]ATP with PKR still bound to the antibody beads. Phos
-
phorylated PKR was analyzed by SDS-PAGE and autoradiography
(top) or by phosphorimage analysis (bottom). Phosphorimage values
are expressed as arbitrary units and represent the volume in individual
PKR bands shown in the blot.
FIG. 4. IFN- transcription in response to poly(I)-poly(C). Huh7,
clone 45, and primary tamarin hepatocyte cultures were harvested
after 2, 7, and 24 h of treatment with 100 g of poly(I)-poly(C) per ml.
The levels of IFN- transcripts were quantied in total cell RNA by
TaqMan RT-PCR and were expressed as fold change in comparison to
the levels of untreated cells. All values are averages of duplicate
cultures.
VOL. 77, 2003 ANTIVIRAL AGENTS AND HCV REPLICONS 1099
also demonstrated that the replicon RNA from treated cul-
tures was not detectably degraded (Fig. 6A); however, due to
the 24.3-fold decrease in replicon RNA at 72 h, overexposure
of the autoradiogram was required to conrm that the RNA at
this time point was not degraded (data not shown).
The cellular RNAs from the same time course study were
used to evaluate whether the antiviral effect was due to error-
prone replication. To test this hypothesis, Huh7 cells were
transfected with total cellular RNA derived from treated and
untreated cultures to determine whether a decrease occurred
in the colony-forming capacity of the replicon in treated cul-
tures. This test for error-prone replication avoids the issue of
ribavirin toxicity and is similar to the approach used for
GBV-B (38) where we demonstrated that virus from treated
cultures had a reduced specic infectivity when used to infect
new cultures. In the current studies, cells were transfected with
equivalent numbers of ge of replicon RNA derived from
treated and untreated cultures. For each time point, cells were
transfected with 10 g of total cell RNA from the treated
cultures (the maximum amount of cellular RNA that can be
efciently transfected), and the RNA from untreated cultures
was adjusted with Huh7 cell RNA such that cells were trans-
fected with equal amounts of cell RNA containing equal rep-
licon copy numbers from treated and untreated cultures (Table
7).
Colony-forming activity for the RNA from treated cultures
was reduced by 3.1- and 6.4-fold at 24 and 48 h of treatment,
respectively, while RNA from cultures treated for 72 h was
devoid of colony-forming activity (Table 7). These data dem-
onstrate a progressive loss in colony-forming activity for the
replicon RNA with increased exposure to ribavirin and support
the conclusion that ribavirin has the capacity to affect HCV
RNA replication through a mechanism of error-prone replica-
tion.
DISCUSSION
The development of a replicon system for HCV has pro-
vided a means to approach a number of biological questions
previously not possible. One of the primary uses of this system
will be the evaluation of antiviral agents and virus-host inter-
actions. Persistence of the replicon is very sensitive to IFN-,
conrming that IFN- has a direct antiviral effect on the virus
(8, 20, 27). In this study, we have extended these observations
to include IFN- and ribavirin, and we have demonstrated the
absence of antiviral effect for poly(I)-poly(C) and TNF-.
The successful application of the replicon system to HCV
research is dependent in part upon the realization that specic
adaptive mutations dramatically increase the ability of the rep-
licon to persist and induce colony formation (8). Adaptive
mutations have been described by several laboratories (8, 27,
35, 42), and the scope of these mutations was increased in this
study. As in previous studies, the primary adaptive mutations
TABLE 6. Titration of antiviral effect of ribavirin
a
Ribavirin
concn (M)
Replicon
(10
6
ge/gof
cell RNA)
b
Fold
reduction
c
GAPDH
Ct
d
0 5.8 0 17.4
50 3.4 1.7 17.6
100 2.5 2.3 17.6
200 0.3 19.3 17.6
400 0.002 2,900 17.6
a
Clone 8 cells were plated and allowed to reach conuency for 3 days and then
were treated with ribavirin for 9 days.
b
Replicon RNA was quantied by TaqMan RT-PCR and expressed as copy
number (ge) per microgram of cell RNA.
c
The data are expressed as the fold change compared to the values for
untreated cultures.
d
The TaqMan RT-PCR assay for replicon RNA was multiplexed for the
GAPDH mRNA to demonstrate a lack of overt toxicity or decrease in cellular
mRNA in ribavirin-treated cells. GAPDH values are expressed as Ct, the am-
plication cycle at which the values exceeded the background threshold.
FIG. 6. Antiviral effect of ribavirin on HCV replicon RNA. Clone
24 cells were cultivated with () or without () 400 M ribavirin for
24, 48, or 72 h. Replicon RNA was analyzed by agarose gel electro-
phoresis, Northern blot hybridization, and autoradiography (A). RNA
in the same blot was quantied by phosphorimage analysis (B), or the
replicon RNA was quantied by TaqMan RT-PCR (C). Values in
panels B and C are expressed as percentages of untreated, control
cultures at each time point.
1100 LANFORD ET AL. J. V
IROL.
were localized to a small domain adjacent to the ISDR and
PKR binding sites of NS5A (Fig. 1), although adaptive muta-
tions have been observed in NS3, NS4B, and NS5B (27, 35, 42).
We found that mutations in NS3 frequently arise in conjunc-
tion with the NS5A mutations. Although the NS3 mutation at
D1431Y did not increase colony-forming efciency of the pa-
rental replicon, it was highly synergistic with the S2204I muta-
tion in NS5A. The synergy between NS3 and NS5A mutations
has been observed previously (35), but this specic combina-
tion of mutations has not been previously observed. Previously
described mutations in NS3 span a region of 407 amino acids
(positions 1202 to 1609) (27, 42) within the helicase domain of
NS3. In our studies, a single NS5A S2204I mutation was asso-
ciated with NS3 mutations separated by 153 amino acids (po-
sitions 1278 to 1431). The mechanisms of adaptive mutations
and the synergy between NS3 and NS5A mutations are not
currently understood but may involve changes in the interac-
tions among viral proteins and in the interactions of viral
proteins with specic host factors. The NS3 D1431Y mutation
increased the mobility of NS3 by SDS-PAGE, whether it was in
the genotype 1b or 1a background, which is suggestive of an
altered posttranslational modication. The adaptive mutations
for genotype 1b did not confer an increased colony-forming
ability to genotype 1a replicons in Huh7 cells, and they did not
extend the host range of the genotype 1b replicon to cell lines
other than Huh7 (data not shown). Previous studies that ex-
amined genotype 1a replicons were restricted to the use of the
H77 strain (8, 27). In this study, we extended the apparent
inactivity of genotype 1a replicons to include the infectious
clone derived from the HCV-1 sequence (40), suggesting that
this may be a general attribute of genotype 1a strains in Huh7
cells, rather than an isolated observation with a single clone.
Currently, the replicon model is limited to the original geno-
type 1b strain employed by Lohmann et al. (43) and one other
genotype 1b isolate (27, 32), the HCV-N infectious clone (6).
Replicons based on the HCV-N sequence do not require adap-
tive mutations for high colony-forming efciency. A four-ami-
no-acid insertion in the ISDR appears to be responsible for the
inherent colony-forming ability of this clone (32). In addition
to being restricted to two genotype 1b isolates, colony forma-
tion with HCV replicons is limited to a single human liver cell
line, Huh7. Undoubtedly, these limitations will not persist for
long. Recently, full-length bicistronic and monocistronic con-
structs that replicate in Huh7 cells have been developed (32,
53), but the production of infectious particles has not been
demonstrated.
In this study, we have extended our observations on the
antiviral activity of ribavirin to include the subgenomic repli-
cons of HCV. We previously demonstrated a reduction in
specic infectivity for GBV-B produced in the presence of
ribavirin that was attributed to the incorporation of ribavirin
triphosphate by the GBV-B polymerase and an accompanying
increase in error-prone replication (38). This effect has also
been observed with poliovirus (13). HCV replicon RNA ob-
tained from cultures exposed to ribavirin exhibited a decrease
in colony-forming efciency when transfected into untreated
cells, and the decrease was proportional to the duration of
ribavirin exposure. Recent in vitro observations with HCV
NS5B protein suggest that the HCV RNA polymerase can
incorporate ribavirin triphosphate utilizing a synthetic tem-
plate (44).
Because ribavirin monotherapy does not result in a signi-
cant reduction in viral load, the improvement in liver disease
realized during ribavirin monotherapy cannot be attributed to
the antiviral properties of ribavirin demonstrated in these stud-
ies. The immunomodulatory activity of ribavirin may be in-
volved in both the improvement of liver disease during mono-
therapy and the increased rate of sustained viral clearance
during combination therapy (for a recent review on ribavirin,
see reference 41). However, the possibility that the synergism
with IFN- is due in part to ribavirin-induced error-prone
replication cannot be dismissed. The small increase in error
rate that may occur at the levels of ribavirin used in the clinic
may result in too few lethal mutations to impact a large pool of
replicating viral RNA. In contrast, this error rate may have a
signicant impact on HCV survival once IFN has reduced the
viral load to the extent where the viral RNA is undetectable by
RT-PCR and thus facilitate sustained viral clearance (37). This
would suggest that three phases are involved in sustained viral
clearance in IFN-ribavirin therapy. The initial rapid decline
of virus (phase I) is most likely due to the direct antiviral effect
of IFN-. A second more gradual and variable decline in viral
load (phase II) may be due to the death of infected hepato-
cytes. Ribavirin does not appear to play a signicant role in
either of the rst two phases. A third phase in which ribavirin
induces biologically signicant mutagenesis may occur after
replicating viral RNA has been reduced to very low levels. If
this hypothesis is correct, then treatment with higher levels of
ribavirin may lead to an increase in sustained viral clearance;
however, at higher levels, the toxic side effects of ribavirin
necessitate short-term therapy. Indeed, induction of error-
prone replication following short-term, high-dose, intravenous
ribavirin therapy may be the mechanism responsible for the
successful treatment of hemorrhagic fevers induced by arena-
viruses (Lassa, Junin, and Machupo viruses) and bunyaviruses
(Hantaan virus) (18, 29, 34, 48). For HCV, improved ribavirin
therapy may be possible by specic targeting of ribavirin to the
liver to decrease extrahepatic toxicity and potentially increase
efcacy if error-prone replication is indeed a mechanism for
synergy with IFN-. The results of future clinical studies with
TABLE 7. Effect of ribavirin on replicon colony-forming efciency
a
Harvest
time (h)
Ribavirin
b
Transfected
replicon (ge)
Colony-forming efciency
c
(colonies/g of replicon)
24 2.6 10
8
140,000
2.6 10
8
45,333
48 8.3 10
7
93,750
8.3 10
7
14,583
72 3.1 10
7
138,888
3.1 10
7
0
a
Clone 24 cells were treated with or without 400 M ribavirin for 24 to 72 h.
Cell RNA was harvested at each time point, and the replicon copy number was
determined (Fig. 6). RNA from untreated cultures was adjusted with Huh7 RNA
such that transfections of treated and untreated RNAs were performed with
identical replicon copy number (ge) and such that each transfection was per-
formed with 10 g of total cell RNA per 100-mm-diameter dish.
b
Symbols: , no ribavirin; , 400 M ribavirin.
c
Colony-forming efciency was expressed as the number of colonies per mi
-
crogram of transfected replicon RNA which was extrapolated from the ge of
replicon transfected at each time point as shown in the table.
VOL. 77, 2003 ANTIVIRAL AGENTS AND HCV REPLICONS 1101
drugs that separate the antiviral and immunomodulatory ef-
fects of ribavirin likely will resolve this mechanistic uncertainty.
In this study, we also observed antiviral activity with IFN-
and IFN- but not with poly(I)-poly(C) and TNF-. A number
of laboratories have observed the direct antiviral effect of
IFN- in the replicon system (8, 20, 27), and while this paper
was under review, a publication by Frese et al. (21) demon-
strated the antiviral effect with IFN-, as well. This antiviral
effect was independent of the production of nitric oxide or the
depletion of tryptophan, pathways known to be involved in the
antiviral activity of IFN- in some systems. We examined
IFN- because of its known antiviral effect in many systems
and our observations that IFN- transcripts are increased in
the livers of chimpanzees undergoing HCV viral clearance
(R. E. Lanford, unpublished data). Following binding to dis-
tinct receptors, both types of IFNs utilize the JAK-STAT signal
transduction pathway to activate ISG transcription (for re-
views, see references 10, 57, and 59). The primary IFN-/
response is mediated through the binding of ISRE promoter
elements by the transcription factor ISGF3, a complex of
STAT-1, STAT-2, and IRF-9, while the primary type II IFN
response is accomplished by the binding of GAS promoter
elements by GAF, a homodimer of STAT-1. Many similarities
exist in the signal transduction pathways induced by type I and
II IFNs, including the induction of transcription of large, par-
tially overlapping sets of ISGs (15), most of which have poorly
understood or no known function. Thus, although the best-
characterized antiviral mechanisms of type I and II IFNs differ,
in some systems they may actually involve similar ISG re-
sponses. IFN- can activate ISGF3, and this response is de-
pendent upon activation of the type I IFN pathway (63). We
initiated studies into the common induction of genes in Huh7
cells by IFN- and IFN-. The basal level of ISG transcripts
did not differ signicantly between Huh7 cells and replicon
lines, suggesting a lack of signicant response to the presence
of viral dsRNA. Most ISGs were induced by both IFN- and
IFN-, although some differences were observed. For example,
IRF-1 transcription is known to be induced to a greater extent
by IFN- than by IFN- (15, 49), and this difference was
observed in this study in the presence or absence of the repli-
con. IFN- induced IP10 to a much greater extent in replicon
lines than in Huh7 cells; however, some of the differences in
the level of induction of specic genes may be due to the clonal
nature of the replicon lines. At this time, the exact mechanisms
of antiviral activity of IFN- and IFN- in the replicon system
remain unresolved.
We had previously demonstrated a potent antiviral activity
for poly(I)-poly(C) in GBV-B-infected primary tamarin hepa-
tocytes (38). The nding that the replicon system was sensitive
to IFN-, but not to poly(I)-poly(C), prompted an exploration
of the dsRNA response in these cells. ISG transcript levels did
not increase in Huh7 cells or replicon lines following poly(I)-
poly(C) treatment. In contrast, the induction of ISGs in pri-
mary chimpanzee hepatocytes was greater for poly(I)-poly(C)
than for IFN-. Since poly(I)-poly(C) has a high level of anti-
viral activity against GBV-B in primary tamarin hepatocytes,
we compared this model system directly with Huh7 cells for the
induction of IFN- transcripts following poly(I)-poly(C) treat-
ment. Although tamarin hepatocytes exhibited 12- to 18-fold
increases in IFN- transcripts 2 to 24 h after poly(I)-poly(C)
treatment, no signicant increase in IFN- transcripts was
noted in Huh7 cells. Some response to dsRNA was detected in
Huh7 cells and replicon lines, since PKR became phosphory-
lated following poly(I)-poly(C) treatment, so the defect in
dsRNA-induced transcription of ISGs in Huh7 cells is inde-
pendent of, or downstream of, PKR activation. After the initial
review of this manuscript, Pugheber et al. (52) published
similar observations with regard to poly(I)-poly(C)-induced
phosphorylation of PKR in Huh7 cells, as well as activation of
IRF-1. Although some induction of ISG transcription was ob-
served in their study, the antiviral impact of poly(I)-poly(C) on
replicon RNA was not examined. It is of interest that the
poly(I)-poly(C)-induced degradation of HBV transcripts in
transgenic mice replicating HBV was not altered in PKR and
IRF-1 knockout mice, indicating that each of these antiviral
pathways is dispensable in this system (26).
Our ndings indicate that Huh7 cells are defective in some
portion of the signaling pathway for dsRNA but that this defect
does not prevent response to IFN- or IFN-. Whether this
defect is related to the permissiveness of Huh7 cells for HCV
replicons is not known, but the defect in dsRNA response
suggests that information gained from the replicon system with
regard to the antiviral mechanism of IFN- may not be entirely
applicable to HCV-infected individuals. Microarray analysis of
serial liver samples from an acute resolving HCV infection in
the chimpanzee model demonstrated high levels of ISG tran-
script induction within 2 days of infection (7). We interpreted
this response to be indicative of a robust dsRNA response in
infected cells that resulted in the secretion of type I IFNs,
which in turn would create zones of cells with ISG induction
adjacent to each infected cell. Since cells within each zone
would be resistant to HCV infection, a reduction in available
replication space would occur as the infection spread in the
liver (for greater discussion of these observations, see refer-
ences 7 and 37). In this manner, the innate antiviral response
may control HCV infections until a T-cell response can elim-
inate infected cells.
The cellular response to dsRNA has been extensively char-
acterized but is still only partially understood. Microarray anal-
ysis of the response to dsRNA revealed that 175 genes exhib-
ited increased expression and 95 genes exhibited decreased
expression (24). This study was conducted in a cell type defec-
tive for the production of type 1 IFNs; therefore, an autocrine
loop from secreted IFN did not complicate interpretation of
the results. Several latent proteins are activated by dsRNA.
The antiviral pathway of 2,5 oligoadenylate synthetase is ac-
tivated by dsRNA, and it in turn activates RNase L, which
possesses potent antiviral activity. MxA possesses antiviral ac-
tivity and is activated by dsRNA, but MxA does not appear to
be involved in the antiviral activity of IFN- in the replicon
system (20). PKR is activated by dsRNA, which leads to the
phosphorylation and inactivation of IB, the inhibitor of NF-
B. Activation of NF-B is required for the dsRNA induction
of IFN- and some ISGs (36, 68). Under some conditions, the
response to dsRNA requires both IRF-1 and STAT1 (4) but
is not dependent on PKR (68) or ISGF3 (4). At least some of
the activity of dsRNA is conveyed by the latent dsRNA-acti-
vated factor transcription factor, or DRAF1, which is com-
prised, in part, of IRF-3 and CBP/p300 (67). Recent studies
have also demonstrated a requirement for p53 in the dsRNA
1102 LANFORD ET AL. J. VIROL.
response (31). At this time, the defect in the dsRNA response
pathway in Huh7 cells is not known, and it is not known
whether this inuences the permissiveness for HCV replicons.
In some replicon lines, IP10 and ISG12 were induced to a
greater extent by IFN- in comparison to Huh7 cells. These
data imply that IFN- and the presence of the replicon may act
together to increase the level of expression of these genes. The
role of the replicon in this induction could be the presence of
dsRNA, since IFN- and dsRNA are known to behave syner-
gistically (46). As with the PKR activation by poly(I)-poly(C),
this would imply that at least portions of the dsRNA response
pathway are intact in Huh7 cells, at least under the conditions
of IFN- stimulation.
Our studies did not detect suppression of ISG induction by
IFN- in replicon cells in comparison to Huh7 cells. Although
in some studies sequence variation in NS5A has been impli-
cated in the degree of IFN- responsiveness in patients, the
exact mechanism by which NS5A impacts IFN- responsive-
ness has not been determined. In vitro studies have suggested
that NS5A suppresses the IFN response by virtue of its inter-
action with PKR (23) or by induction of interleukin 8 expres-
sion (25, 55). The lack of detectable suppression of the IFN-
response in the replicon cells may have been due to a number
of factors, including the high level of IFN exposure, decien-
cies in the dsRNA response in Huh7 cells, or the effects of
adaptive mutations present in the replicons. However, the re-
sults of these studies do suggest that further investigation of
the interaction of subgenomic replicons with the cellular fac-
tors involved in viral resistance may improve our understand-
ing of the mechanism(s) of IFN resistance among different
genotypes of HCV. Examination of the newly developed full-
length replicons (32, 53) will be important as well, since mul-
tiple HCV proteins may modulate the host response to dsRNA
and IFNs.
ACKNOWLEDGMENTS
This work was supported in part by NIH grants U19 AI40035, RO1
AI49574, and P51 RR13986.
We thank Stuart Ray and David Thomas for insightful discussions
on the potential antiviral role of ribavirin once the replicating viral
population has been reduced by IFN treatment.
REFERENCES
1. Alter, H. J., and M. Houghton. 2000. Hepatitis C virus and eliminating
post-transfusion hepatitis. Nat. Med. 6:10821086.
2. Alter, H. J., and L. B. Seeff. 2000. Recovery, persistence, and sequelae in
hepatitis C virus infection: a perspective on long-term outcome. Semin. Liver
Dis. 20:1735.
3. Alter, M. J., D. Kruszon-Moran, O. V. Nainan, G. M. McQuillan, F. X. Gao,
L. A. Moyer, R. A. Kaslow, and H. S. Margolis. 1999. The prevalence of
hepatitis C virus infection in the United States, 1988 through 1994. N. Engl.
J. Med. 341:556562.
4. Bandyopadhyay, S. K., G. T. Leonard, Jr., T. Bandyopadhyay, G. R. Stark,
and G. C. Sen. 1995. Transcriptional induction by double-stranded RNA is
mediated by interferon-stimulated response elements without activation of
interferon-stimulated gene factor 3. J. Biol. Chem. 270:1962419629.
5. Bartenschlager, R., and V. Lohmann. 2000. Replication of hepatitis C virus.
J. Gen. Virol. 81:16311648.
6. Beard, M. R., G. Abell, M. Honda, A. Carroll, M. Gartland, B. Clarke, K.
Suzuki, R. Lanford, D. V. Sangar, and S. M. Lemon. 1999. An infectious
molecular clone of a Japanese genotype 1b hepatitis C virus. Hepatology
30:316324.
7. Bigger, C. B., K. M. Brasky, and R. E. Lanford. 2001. DNA microarray
analysis of chimpanzee liver during acute resolving hepatitis C virus infec-
tion. J. Virol. 75:70597066.
8. Blight, K. J., A. A. Kolykhalov, and C. M. Rice. 2000. Efcient initiation of
HCV RNA replication in cell culture. Science 290:19721974.
9. Bodenheimer, H. C., Jr., K. L. Lindsay, G. L. Davis, J. H. Lewis, S. N. Thung,
and L. B. Seeff. 1997. Tolerance and efcacy of oral ribavirin treatment of
chronic hepatitis C: a multicenter trial. Hepatology 26:473477.
10. Boehm, U., T. Klamp, M. Groot, and J. Howard. 1997. Cellular responses to
interferon-. Annu. Rev. Immunol. 15:749795.
11. Choo, Q.-L., G. Kuo, A. J. Weiner, L. J. Overby, D. W. Bradley, and M.
Houghton. 1989. Isolation of a cDNA clone derived from blood-borne
non-A, non-B viral hepatitis genome. Science 244:359362.
12. Conry-Cantilena, C., M. VanRaden, J. Gibble, J. Melpolder, A. O. Shakil, L.
Viladomiu, L. Cheung, A. DiBisceglie, J. Hoofnagle, J. W. Shih, R. Kaslow,
P. Ness, and H. J. Alter. 1996. Routes of infection, viremia, and liver disease
in blood donors found to have hepatitis C virus infection. N. Engl. J. Med.
334:16911696.
13. Crotty, S., C. E. Cameron, and R. Andino. 2001. RNA virus error catastro-
phe: direct molecular test by using ribavirin. Proc. Natl. Acad. Sci. USA
98:68956900.
14. De Clercq, E. 1993. Antiviral agents: characteristic activity spectrum depend-
ing on the molecular target with which they interact. Adv. Virus Res. 42:1
55.
15. Der, S. D., A. Zhou, B. R. Williams, and R. H. Silverman. 1998. Identication
of genes differentially regulated by interferon alpha, beta, or gamma using
oligonucleotide arrays. Proc. Natl. Acad. Sci. USA 95:1562315628.
16. Dusheiko, G., J. Main, H. Thomas, O. Reichard, C. Lee, A. Dhillon, S.
Rassam, A. Fryden, H. Reesink, M. Bassendine, G. Norkrans, T. Cuypers, N.
Lelie, P. Telfer, J. Watson, C. Weegink, P. Sillikens, and O. Weiland. 1996.
Ribavirin treatment for patients with chronic hepatitis C: results of a place-
bo-controlled study. J. Hepatol. 25:591598.
17. Enomoto, N., I. Sakuma, Y. Asahina, M. Kurosaki, T. Murakami, C.
Yamamoto, and N. Izumi. 1996. Interferon sensitivity determining sequence
of the hepatitis C virus genome. Hepatology 24:460461.
18. Enria, D. A., and J. I. Maiztegui. 1994. Antiviral treatment of argentine
hemorrhagic fever. Antiviral Res. 23:2331.
19. Fang, S. H., L. H. Hwang, D. S. Chen, and B. L. Chiang. 2000. Ribavirin
enhancement of hepatitis C virus core antigen-specic type 1 T helper cell
response correlates with the increased IL-12 level. J. Hepatol. 33:791798.
20. Frese, M., T. Pietschmann, D. Moradpour, O. Haller, and R. Barten-
schlager. 2001. Interferon- inhibits hepatitis C virus subgenomic RNA
replication by an MxA-independent pathway. J. Gen. Virol. 82:723733.
21. Frese, M., V. Schwarzle, K. Barth, N. Krieger, V. Lohmann, S. Mihm, O.
Haller, and R. Bartenschlager. 2002. Interferon-gamma inhibits replication
of subgenomic and genomic hepatitis C virus RNAs. Hepatology 35:694703.
22. Gale, M., C. M. Blakely, B. Kwieciszewski, S. L. Tan, M. Dossett, N. M.
Tang, M. J. Korth, S. J. Polyak, D. R. Gretch, and M. G. Katze. 1998.
Control of PKR protein kinase by hepatitis C virus nonstructural 5A protein:
molecular mechanisms of kinase regulation. Mol. Cell. Biol. 18:52085218.
23. Gale, M. J., M. J. Korth, N. M. Tang, S. L. Tan, D. A. Hopkins, T. E. Dever,
S. J. Polyak, D. R. Gretch, and M. G. Katze. 1997. Evidence that hepatitis C
virus resistance to interferon is mediated through repression of the PKR
protein kinase by the nonstructural 5A protein. Virology 230:217227.
24. Geiss, G., G. Jin, J. Guo, R. Bumgarner, M. G. Katze, and G. Sen. 2001. A
comprehensive view of regulation of gene expression by double-stranded
RNA-mediated cell signaling. J. Biol. Chem. 276:3017830182.
25. Girard, S., P. Shalhoub, P. Lescure, A. Sabile, D. E. Misek, S. Hanash, C.
Brechot, and L. Beretta. 2002. An altered cellular response to interferon and
up-regulation of interleukin-8 induced by the hepatitis C viral protein NS5A
uncovered by microarray analysis. Virology 295:272283.
26. Guidotti, L. G., A. Morris, H. Mendez, R. Koch, R. H. Silverman, B. R. G.
Williams, and F. V. Chisari. 2002. Interferon-regulated pathways that con-
trol hepatitis B virus replication in transgenic mice. J. Virol. 76:26172621.
27. Guo, J. T., V. Bichko, and C. Seeger. 2001. Effect of alpha interferon on the
hepatitis C virus replicon. J. Virol. 75:85168523.
28. Honda, M., M. R. Beard, L. H. Ping, and S. M. Lemon. 1999. A phyloge-
netically conserved stem-loop structure at the 5 border of the internal
ribosome entry site of hepatitis C virus is required for cap-independent viral
translation. J. Virol. 73:11651174.
29. Huggins, J. W., C. M. Hsiang, T. M. Cosgriff, J. Y. Guang, J. I. Smith, Z. O.
Wu, J. W. DeDuc, Z. M. Zheng, J. M. Meegan, Q. N. Wang, D. D. Oland,
X. E. Gui, P. H. Gibbs, G. H. Yuan, and T. M. Zhang. 1991. Prospective,
double-blind, concurrent, placebo-controlled clinical trial of intravenous
ribavirin therapy of hemorrhagic fever with renal syndrome. J. Infect. Dis.
164:11191127.
30. Hultgren, C., D. R. Milich, O. Weiland, and M. Sa¨llberg. 1998. The antiviral
compound ribavirin modulates the T helper (Th)1/Th2 subset balance in
hepatitis B and C virus-specic immune responses. J. Gen. Virol. 79:2381
2391.
31. Hummer, B. T., X. L. Li, and B. A. Hassel. 2001. Role for p53 in gene
induction by double-stranded RNA. J. Virol. 75:77747777.
32. Ikeda, M., M. K. Yi, K. Li, and S. A. Lemon. 2002. Selectable subgenomic
and genome-length dicistronic RNAs derived from an infectious molecular
clone of the HCV-N strain of hepatitis C virus replicate efciently in cultured
Huh7 cells. J. Virol. 76:29973006.
33. Jaeckel, E., M. Cornberg, H. Wedemeyer, T. Santantonio, J. Mayer, M.
VOL. 77, 2003 ANTIVIRAL AGENTS AND HCV REPLICONS 1103
Zankel, G. Pastore, M. Dietrich, C. Trautwein, and M. Manns. 2001. Treat-
ment of acute hepatitis C with interferon alpha-2b. N. Engl. J. Med. 345:16.
34. Kilgore, P. E., T. G. Ksiazek, P. E. Rollin, J. N. Mills, M. R. Villagra, M. J.
Montenegro, M. A. Costales, L. C. Paredes, and C. J. Peters. 1997. Treat-
ment of bolivian hemorrhagic fever with intravenous ribavirin. Clin. Infect.
Dis. 24:718722.
35. Krieger, N., V. Lohmann, and R. Bartenschlager. 2001. Enhancement of
hepatitis C virus RNA replication by cell culture-adaptive mutations. J. Vi-
rol. 75:46144624.
36. Kumar, A., Y. L. Yang, V. Flati, S. Der, S. Kadereit, A. Deb, J. Haque, L.
Reis, C. Weissmann, and B. R. Williams. 1997. Decient cytokine signaling
in mouse embryo broblasts with a targeted deletion in the PKR gene: role
of IRF-1 and NF-B. EMBO J. 16:406416.
37. Lanford, R. E., and C. Bigger. 2002. Advances in model systems for hepatitis
C virus research. Virology 293:19.
38. Lanford, R. E., D. Chavez, B. Guerra, J. Y. N. Lau, Z. Hong, K. M. Brasky,
and B. Beames. 2001. Ribavirin induces error-prone replication of GB virus
B in primary tamarin hepatocytes. J. Virol. 75:80748081.
39. Lanford, R. E., and L. E. Estlack. 1998. A cultivation method for highly
differentiated primary chimpanzee hepatocytes permissive for hepatitis C
virus replication. Methods Mol. Med. 19:501516.
40. Lanford, R. E., H. Lee, D. Chavez, B. Guerra, and K. M. Brasky. 2001.
Infectious cDNA clone of the hepatitis C virus genotype 1 prototype se-
quence. J. Gen. Virol. 82:12911297.
41. Lau, J. Y., R. C. Tam, T. J. Liang, and Z. Hong. 2002. Mechanism of action
of ribavirin in the combination treatment of chronic HCV infection. Hepa-
tology 35:10021009.
42. Lohmann, V., F. Korner, A. Dobierzewska, and R. Bartenschlager. 2001.
Mutations in hepatitis C virus RNAs conferring cell culture adaptation.
J. Virol. 75:14371449.
43. Lohmann, V., F. Ko¨rner, J. O. Koch, U. Herian, L. Theilmann, and R.
Bartenschlager. 1999. Replication of subgenomic hepatitis C virus RNAs in
a hepatoma cell line. Science 285:110113.
44. Maag, D., C. Castro, Z. Hong, and C. E. Cameron. 2001. Hepatitis C virus
RNA-dependent RNA polymerase (NS5B) as a mediator of the antiviral
activity of ribavirin. J. Biol. Chem. 276:4609446098.
45. Manns, M. P., J. G. McHutchison, S. C. Gordon, V. K. Rustgi, M. Shiffman,
R. Reindollar, Z. D. Goodman, M. H. Ling, and J. K. Albrecht. 2001.
Peginterferon alfa-2b plus ribavirin compared with interferon alfa-2b plus
ribavirin for initial treatment of chronic hepatitis C: a randomised trial.
Lancet 358:958965.
46. Marcus, P. I., and M. J. Sekellick. 2001. Combined sequential treatment with
interferon and dsRNA abrogates virus resistance to interferon action. J. In-
terferon Cytokine Res. 21:423429.
47. Markland, W., T. J. McQuaid, J. Jain, and A. D. Kwong. 2000. Broad-
spectrum antiviral activity of the IMP dehydrogenase inhibitor VX-497: a
comparison with ribavirin and demonstration of antiviral additivity with
alpha interferon. Antimicrob. Agents Chemother. 44:859866.
48. McCormick, J. B., I. J. King, P. A. Webb, C. L. Scribner, R. B. Craven, K. M.
Johnson, L. H. Elliott, and R. Belmont-Williams. 1986. Lassa fever/Effective
therapy with ribavirin. N. Engl. J. Med. 314:2026.
49. Melen, K., P. Keskinen, A. Lehtonen, and I. Julkunen. 2000. Interferon-
induced gene expression and signaling in human hepatoma cell lines.
J. Hepatol. 33:764772.
50. Ning, Q., D. Brown, J. Parodo, M. Cattral, R. Gorczynski, E. Cole, L. Fung,
J. W. Ding, M. F. Liu, O. Rotstein, M. J. Phillips, and G. Levy. 1998.
Ribavirin inhibits viral-induced macrophage production of TNF, IL-1, the
procoagulant fgl2 prothrombinase and preserves Th1 cytokine production
but inhibits Th2 cytokine response. J. Immunol. 160:34873493.
51. Pavio, N., D. R. Taylor, and M. M. C. Lai. 2002. Detection of a novel
unglycosylated form of hepatitis C virus E2 envelope protein that is located
in the cytosol and interacts with PKR. J. Virol. 76:12651272.
52. Pugheber, J., B. Fredericksen, R. Sumpter, Jr., C. Wang, F. Ware, D. L.
Sodora, and M. Gale, Jr. 2002. Regulation of PKR and IRF-1 during hep-
atitis C virus RNA replication. Proc. Natl. Acad. Sci. USA 99:46504655.
53. Pietschmann, T., V. Lohmann, A. Kaul, N. Krieger, G. Rinck, G. Rutter, D.
Strand, and R. Bartenschlager. 2002. Persistent and transient replication of
full-length hepatitis C virus genomes in cell culture. J. Virol. 76:40084021.
54. Podevin, P., A. Sabile, R. Gajardo, N. Delhem, A. Abadie, P. Y. Lozach, L.
Beretta, and C. Brechot. 2001. Expression of hepatitis C virus NS5A natural
mutants in a hepatocytic cell line inhibits the antiviral effect of interferon in
a PKR-independent manner. Hepatology 33:15031511.
55. Polyak, S. J., K. S. A. Khabar, D. M. Paschal, H. J. Ezelle, G. Duverlie, G. N.
Barber, D. E. Levy, N. Mukaida, and D. R. Gretch. 2001. Hepatitis C virus
nonstructural 5A protein induces interleukin-8, leading to partial inhibition
of the interferon-induced antiviral response. J. Virol. 75:60956106.
56. Reed, K. E., and C. M. Rice. 1998. Molecular characterization of hepatitis C
virus. Curr. Stud. Hematol. Blood Transfus. 62:137.
57. Samuel, C. E. 2001. Antiviral actions of interferons. Clin. Microbiol. Rev.
14:778809.
58. Sidwell, R. W., J. H. Huffman, G. P. Khare, L. B. Allen, J. T. Witkowski, and
R. K. Robins. 1972. Broad-spectrum antiviral activity of virazole: 1--
D-
ribofuranosyl-1,2,4-triazole-3-carboxamide. Science 177:705706.
59. Stark, G. R., I. M. Kerr, B. R. Williams, R. H. Silverman, and R. D.
Schreiber. 1998. How cells respond to interferons. Annu. Rev. Biochem.
67:227264.
60. Tam, R. C., B. Pai, J. Bard, C. Lim, D. R. Averett, U. T. Phan, and T.
Milovanovic. 1999. Ribavirin polarizes human T cell responses towards a
type 1 cytokine prole. J. Hepatol. 30:376382.
61. Tan, S. L., and M. G. Katze. 2001. How hepatitis C virus counteracts the
interferon response: the jury is still out on NS5A. Virology 284:112.
62. Tan, S. L., H. Nakao, Y. He, S. Vijaysri, P. Neddermann, B. L. Jacobs, B. J.
Mayer, and M. G. Katze. 1999. NS5A, a nonstructural protein of hepatitis C
virus, binds growth factor receptor-bound protein 2 adaptor protein in a Src
homology 3 domain/ligand-dependent manner and perturbs mitogenic sig-
naling. Proc. Natl. Acad. Sci. USA 96:55335538.
63. Taniguchi, T., and A. Takaoka. 2001. A weak signal for strong responses:
interferon-alpha/beta revisited. Nat. Rev. Mol. Cell Biol. 2:378386.
64. Tanji, Y., T. Kaneko, S. Satoh, and K. Shimotohno. 1995. Phosphorylation of
hepatitis C virus-encoded nonstructural protein NS5A. J. Virol. 69:3980
3986.
65. Taylor, D. R., S. T. Shi, P. R. Romano, G. N. Barber, and M. M. Lai. 1999.
Inhibition of the interferon-inducible protein kinase PKR by HCV E2 pro-
tein. Science 285:107110.
66. Tu, H., L. Gao, S. T. Shi, D. R. Taylor, T. Yang, A. K. Mircheff, Y. M. Wen,
A. E. Gorbalenya, S. B. Hwang, and M. M. Lai. 1999. Hepatitis C virus RNA
polymerase and NS5A complex with a SNARE-like protein. Virology 263:
3041.
67. Weaver, B. K., K. P. Kumar, and N. C. Reich. 1998. Interferon regulatory
factor 3 and CREB-binding protein/p300 are subunits of double-stranded
RNA-activated transcription factor DRAF1. Mol. Cell. Biol. 18:13591368.
68. Yang, Y. L., L. F. Reis, J. Pavlovic, A. Aguzzi, R. Schafer, A. Kumar, B. R.
Williams, M. Aguet, and C. Weissmann. 1995. Decient signaling in mice
devoid of double-stranded RNA-dependent protein kinase. EMBO J. 14:
60956106.
1104 LANFORD ET AL. J. VIROL.
... 13,14 The only therapeutic agent currently recommended by the WHO for the treatment of CCHF is ribavirin, administered either orally or intravenously. 1 However, its effectiveness in vitro and in vivo remains controversial, and a case-controlled clinical trial has never been conducted. 12,[15][16][17][18][19][20] Ribavirin is a broad-spectrum antiviral guanosine analog, which has been routinely used in combination with interferon-α to treat hepatitis C. It has also been indicated for the treatment of Lassa fever and respiratory-syncytial virus infections 10,15,[21][22][23] and has efficacy against herpes, vesicular stomatitis, and influenza viruses. 24 Several mechanisms of action have been proposed for ribavirin, including inhibition of mRNA capping, impact on host cell gene expression, inflammation and immunomodulation, through its ability to decrease intracellular guanosine triphosphate (GTP). ...
... 21 The latter is suggested to work by erroneous substitutions of ribavirin triphosphates (RTPs) for guanosine triphosphates (GTPs), which have ambiguous base-pairing properties with uridine triphosphate or cytidine triphosphate, leading to an increase in G-to-A and C-to-T (equivalent to C-to-U in RNA sequence) mutations. [27][28][29][30] While this mechanism of action has been found to occur in vitro for hepatitis C virus, 22 other studies have undermined this hypothesis in vivo. 31 A recent clinical study inferred that ribavirin treatment increased the rate of CCHFV genome mutagenesis, however, this was based only on a single case report and lacked appropriate controls. ...
Article
Crimean-Congo haemorrhagic fever (CCHF) is the most widespread tick-borne viral haemorrhagic fever (VHF) affecting humans, and yet a licensed drug against the virus (CCHFV) is still not available. While several studies have suggested the efficacy of ribavirin against CCHFV, current literature remains inconclusive. In this study, we have utilized next-generation sequencing to investigate the mutagenic effect of ribavirin on the CCHFV genome during clinical disease. Samples collected from CCHF patients receiving ribavirin treatment or supportive care only at Sivas Cumhuriyet University Hospital, Turkey, were analysed. By comparing the frequency of mutations in each group, we found little evidence of an overall mutagenic effect. This suggests that ribavirin, administered at the acute stages of CCHFV infection (at the WHO recommended dose) is unable to induce lethal mutagenesis that would cause an extinction event in the CCHFV population and reduce viremia. This article is protected by copyright. All rights reserved.
... This emphasizes the need to address the role of IFN-III in human listeriosis. As for Huh7 cells, they do not produce any IFN in response to Listeria, as previously observed when these cells are challenged with other IFN-triggering stimuli, such as dsRNA (Lanford et al., 2003), poly(I-C) (Li et al., 2005) or hepatitis C virus (Israelow et al., 2014). A possible explanation for this lack of response is that Huh7 cells have a defective TLR3 signaling pathway (Li et al., 2005). ...
... These differences in IFN type and expression kinetics may have important physiological impacts and highlight the need to study the role of IFN-III in human listeriosis. As for Huh7 cells, they do not produce any IFN in response to Lm, as previously observed when these cells are confronted with other IFN-triggering stimuli, such as double stranded RNA (dsRNA) (Lanford et al., 2003), poly I-C (a synthetic dsRNA) (Li et al., 2005b) or HCV (Israelow et al., 2014). A possible explanation for this unresponsiveness is that Huh7 cells have a defective TLR3 signalling pathway (Li et al., 2005b). ...
Thesis
Listeria monocytogenes (Lm) is a facultative intracellular pathogen that causes severe foodborne illness in pregnant women and immunocompromised individuals. After the intestinal phase of infection, the liver plays a central role in the clearance of bacteria. It is also a primary target organ, in which Lm replicates extensively in hepatocytes, the parenchymal liver cells. Recent data suggest that during long-term infection of hepatocytes, a bacterial subpopulation can escape eradication by entering a persistence phase in intracellular vacuoles called LisCVs. The first axis of my thesis was to examine the host response to this long-term infection in hepatocytes with the objective of identifying a common transcriptomic signature in several hepatocyte models. Cellular models of persistent infection were established in HepG2 and Huh7 human hepatocyte cell lines and primary mouse hepatocytes. Lm consistently entered the persistence phase after three days of infection in these cells, while inducing a potent interferon response, of type I in primary mouse hepatocytes and type III in HepG2, while Huh7 cells remained unresponsive. RNA-sequencing analysis identified a profoundly altered transcriptional landscape from which a common signature of long-term Lm infection emerged, characterized by (i) the upregulation of a set of genes involved in antiviral immunity and (ii) the downregulation of many genes encoding acute phase proteins, particularly those involved in the complement and coagulation systems. This transcriptional block on acute phase protein coding gene expression was maintained in the presence of pro-inflammatory cytokine stimulation. Quantitative proteomics analysis of the hepatocyte secretome revealed reduced protein abundance that correlated with transcriptomic downregulation. Infection also altered the expression of cholesterol metabolism-associated genes in human hepatocytes that was independent of the interferon response. The second axis of my thesis involved investigating the role of the epigenetic factors BAHD1 and MIER in long term Lm infection in hepatocytes. The BAHD1-MIER association forms the scaffold of a recently described chromatin-repressive complex, belonging to the histone deacetylase (HDAC) family. BAHD1 was previously shown to repress the interferon response upon Lm infection of colon epithelial cells. This work expands on these results, by providing novel data on the inhibition of interferon-stimulated genes by BAHD1-MIER in hepatocytes. The results also suggest that, in a mixed population of infected and uninfected hepatocytes, uninfected bystander cells are the major producers of interferon and interferon stimulated gene products in response to interferon after three days of infection. This work strongly suggests that long-term infection with Lm profoundly deregulates the secretory and metabolic functions of hepatocytes, which could generate an environment favourable to the establishment of persistent infection through the reduced abundance of crucial hepatocyte specific innate immune proteins. At the same time, it opens up multiple avenues to explore the mechanisms of host transcriptional regulation during persistent infection and the role of uninfected bystander cells in subverting infection-mediated transcriptional repression.
... Notably, impairing autophagy in other cell types, including Huh7 cells and HUVECs, showed no effect on RVFV replication. Unlike macrophages, Huh7 cells and HUVECs have compromised immune responses (41)(42)(43)(44). This further indicates that immune suppression is the major mechanism underlying the autophagy-dependent promotion of RVFV replication, at least in macrophages. ...
Article
Full-text available
Rift Valley fever virus (RVFV) is a mosquito-borne bunyavirus that causes severe and potentially fatal hemorrhagic fever in humans. Autophagy is a self-degradative process that can restrict viral replication at multiple infection steps. In this study, we evaluated the effects of RVFV-triggered autophagy on viral replication and immune responses. Our results showed that RVFV infection triggered autophagosome formation and induced complete autophagy. Impairing autophagy flux by depleting autophagy-related gene 5 (ATG5), ATG7, or sequestosome 1 (SQSTM1) or treatment with autophagy inhibitors markedly reduced viral RNA synthesis and progeny virus production. Mechanistically, our findings demonstrated that the RVFV nucleoprotein (NP) C-terminal domain interacts with the autophagy receptor SQSTM1 and promotes the SQSTM1-microtubule-associated protein 1 light chain 3 B (LC3B) interaction and autophagy. Deletion of the NP C-terminal domain impaired the interaction between NP and SQSTM1 and its ability to trigger autophagy. Notably, RVFV-triggered autophagy promoted viral infection in macrophages but not in other tested cell types, including Huh7 hepatocytes and human umbilical vein endothelial cells, suggesting cell type specificity of this mechanism. It was further revealed that RVFV NP-triggered autophagy dampens antiviral innate immune responses in infected macrophages to promote viral replication. These results provide novel insights into the mechanisms of RVFV-triggered autophagy and indicate the potential of targeting the autophagy pathway to develop antivirals against RVFV. IMPORTANCE We showed that RVFV infection induced the complete autophagy process. Depletion of the core autophagy genes ATG5, ATG7, or SQSTM1 or pharmacologic inhibition of autophagy in macrophages strongly suppressed RVFV replication. We further revealed that the RVFV NP C-terminal domain interacted with SQSTM1 and enhanced the SQSTM1/LC3B interaction to promote autophagy. RVFV NP-triggered autophagy strongly inhibited virus-induced expression of interferon-stimulated genes in infected macrophages but not in other tested cell types. Our study provides novel insights into the mechanisms of RVFV-triggered autophagy and highlights the potential of targeting autophagy flux to develop antivirals against this virus.
... Successful treatment of FIP with new antiviral compounds, e.g., nucleoside analog GS-441524 and 3C-like protease inhibitor, is looking very promising [24,27]; however, they are not universally effective. It has been suggested that antiviral drugs can act synergistically with immunomodulatory treatments to improve patient survival in a number of viral diseases, such as human influenza virus, hepatitis C virus and human immunodeficiency virus (HIV) [28][29][30] and combinations of these antiviral drugs with immune-modulators, such as PI, may potentially improve the long-term treatment outcomes for cats with FIP [31]. ...
Article
Full-text available
Feline infectious peritonitis (FIP) remains a major diagnostic and treatment challenge in feline medicine. An ineffective immune response is an important component of FIP pathophysiology; hence treatment with an immune stimulant such as Polyprenyl Immunostimulant™ (PI), which enhances cell-mediated immunity by upregulating the innate immune response via Toll-like receptors, is a rational approach. Records of cats with FIP treated with PI orally for over 365 days were retrospectively studied. Of these cats (n = 174), records were obtained for n = 103 cats with appropriate clinical signs and clinical pathology. Of these, n = 29 had FIP confirmed by immunohistochemistry (IHC) or reverse transcription polymerase-chain-reaction (RT-PCR). Most of the cats (25/29; 86%) had non-effusive FIP, and only 4/29 cats (14%) had effusive FIP. The mean survival time (MST) was 2927 days (eight years); with 55% of the cats (16/29) still being alive at the time data collection, and 45% (13/29) having died. A persistently low hematocrit plus low albumin:globulin (A:G) ratio, despite treatment, was a negative prognostic indicator. It took a mean of ~182 days and ~375 days, respectively, for anemia and low A:G ratio to resolve in the cats that presented with these laboratory changes. This study shows that PI is beneficial in the treatment of FIP, and more studies are needed to establish the best protocols of use.
... Moreover, ribavirin is a mutagenic nucleoside analog in rapidly replicating viruses, inducing error catastrophes and competing with physiological nucleotides for the HCV RdRp active site [29]. Ribavirin also induces lethal mutagenesis and error-prone HCV replication [30][31][32]. We have reported previously that A→G and G→A mutations increased in comparison with the total number of transition mutations [32]. ...
Article
Full-text available
Hepatitis A virus (HAV) is a causative agent of acute hepatitis and can occasionally induce acute liver failure. However, specific potent anti-HAV drug is not available on the market currently. Thus, we investigated several novel therapeutic drugs through a drug repositioning approach, targeting ribonucleic acid (RNA)-dependent RNA polymerase and RNA-dependent deoxyribonucleic acid polymerase. In the present study, we examined the anti-HAV activity of 18 drugs by measuring the HAV subgenomic replicon and HAV HA11-1299 genotype IIIA replication in human hepatoma cell lines, using a reporter assay and real-time reverse transcription polymerase chain reaction, respectively. Mutagenesis of the HAV 5’ untranslated region was also examined by next-generation sequencing. These specific parameters were explored because lethal mutagenesis has emerged as a novel potential therapeutic approach to treat RNA virus infections. Favipiravir inhibited HAV replication in both Huh7 and PLC/PRF/5 cells, although ribavirin inhibited HAV replication in only Huh7 cells. Next-generation sequencing demonstrated that favipiravir could introduce nucleotide mutations into the HAV genome more than ribavirin. In conclusion, favipiravir could introduce nucleotide mutations into the HAV genome and work as an antiviral against HAV infection. Provided that further in vivo experiments confirm its efficacy, favipiravir would be useful for the treatment of severe HAV infection.
... Ribavirin stops particularly the viral replication by inhibiting the mechanism of duplication of viral nucleic acid as after the penetration in a host cell, the virus gets the control of cellular metabolism and uses the cellular machinery for its own replication and virus assembly. These facts are supported by the fi ndings of Fenton and Potter, [30] John, [31] Connor et al., [32] Hultgren et al., [33] Lucia et al., [34] Robert et al., [35] Fernandez et al., [18] Marina et al., [36] Ramos et al., [37] and Wang et al. [38] Herbal extracts contain a large number of therapeutically active compounds, and these phytochemicals are a potential source of molecular drug design and development against a large number of infectious viral diseases. It may conclude here that the aqueous extract of Glycyrrhiza contains the anti-viral activity due to glycyrrhizic acid; however, further studies are needed to determine the exact role of the potentially active ingredient in the viral inhibition in Newcastle disease. ...
Article
Full-text available
Background: The Newcastle disease represents as one of the most infectious viral diseases, which afflicts almost every species of bird. The causative agent of the disease is a single-stranded RNA virus with rapid replication capability. Objective: This study was performed to evaluate the comparative anti-viral efficacy and toxicity of Glycyrrhiza glabra aqueous extract and ribavirin against the Newcastle disease virus. Materials and Methods: The embryonated eggs were divided into six groups (A, B, C, D, E and F). Groups A, B, C, and D were further subdivided into three subgroups. The virus was identified by a hemagglutination inhibition test. Spot hemagglutination test and viability of embryos were also evaluated. Three different concentrations i-e., 30 mg/100 ml, 60 mg/100 ml, and 120 mg/100 ml of the Glycyrrhiza aqueous extract and 10 μg/ml, 20 μg/ml, and 40 μg/ml ribavirin in deionized water was evaluated for its toxicity and anti-viral activity in the embryonated eggs. Results: 60 mg/100 ml concentration of Glycyrrhiza extract did not produce any toxicity in the embryonated eggs and showed anti-viral activity against the virus. Similarly, 20 μg/ml ribavirin was non-toxic in the embryonated eggs and contained anti-viral activity. Conclusion: It may conclude from the presented study that 60 mg/100 ml Glycyrrhiza extract inhibits replication of Newcastle disease virus and is non-toxic in the embryonated eggs. So, Glycyrrhiza glabra extract may be further evaluated in future to determine the potentially active compounds for their anti-viral activity against the Newcastle disease virus. Furthermore, the mechanism of action of these active phytochemicals as an antiviral agent would be helpful to elucidate the pathogenesis of the disease.
... This emphasizes the need to address the role of IFN-III in human listeriosis. As for Huh7 cells, they do not produce any IFN in response to Listeria, as previously observed when these cells are challenged with other IFN-triggering stimuli, such as dsRNA (Lanford et al., 2003), poly(I-C) (Li et al., 2005) or hepatitis C virus (Israelow et al., 2014). A possible explanation for this lack of response is that Huh7 cells have a defective TLR3 signaling pathway (Li et al., 2005). ...
Article
Full-text available
Listeria monocytogenes causes severe foodborne illness in pregnant women and immunocompromised individuals. After the intestinal phase of infection, the liver plays a central role in the clearance of this pathogen through its important functions in immunity. However, recent evidence suggests that during long-term infection of hepatocytes, a subpopulation of Listeria may escape eradication by entering a persistence phase in intracellular vacuoles. Here, we examine whether this long-term infection alters hepatocyte defense pathways, which may be instrumental for bacterial persistence. We first optimized cell models of persistent infection in human hepatocyte cell lines HepG2 and Huh7 and primary mouse hepatocytes (PMH). In these cells, Listeria efficiently entered the persistence phase after three days of infection, while inducing a potent interferon response, of type I in PMH and type III in HepG2, while Huh7 remained unresponsive. RNA-sequencing analysis identified a common signature of long-term Listeria infection characterized by the overexpression of a set of genes involved in antiviral immunity and the under-expression of many acute phase protein (APP) genes, particularly involved in the complement and coagulation systems. Infection also altered the expression of cholesterol metabolism-associated genes in HepG2 and Huh7 cells. The decrease in APP transcripts was correlated with lower protein abundance in the secretome of infected cells, as shown by proteomics, and also occurred in the presence of APP inducers (IL-6 or IL-1β). Collectively, these results reveal that long-term infection with Listeria profoundly deregulates the innate immune functions of hepatocytes, which could generate an environment favorable to the establishment of persistent infection.
... Type I IFN plays an essential role in the immune defense against viruses. During the HCV infection and replicon replication, the reduction of ISG expression is observed, hence the antiviral response is suppressed [108]. IFN-α administration to HCVinfected patients activates ISG expression [109], causing a decreased, almost undetectable, HCV level. ...
Article
Full-text available
Interferons type I (IFN-I), activated following a bacterial or viral infection, play a major role in the induction and regulation of the immune system. The immune response results in viral RNA and binds to receptors such as RIG-I-like receptors (RLRs) or Toll-like receptors, leading to the IFN-I signaling cascade. Thanks to its cellular function, IFN-I is widely used in therapies for such diseases as multiple sclerosis (MS) and hepatitis C disease (HCD). MS is a neurological, autoimmune, chronic inflammatory disease of the central nervous system (CNS). During MS, nerve cell demyelination is observed due to the myelin heaths and oligodendrocyte damage. As a result, neuronal signal and neuron communication are attenuated. The mechanism of MS is still unknown. MS therapy applies interferon-β (IFN-β). IFN-β therapy has been used since the last century, but the therapeutic mechanism of IFN-β has not been completely understood. MS can lead to four syndromes: clinically isolated syndrome (CIS), relapsing-remitting MS (RRMS), primary progressive MS (PPMS), and secondary progressive MS (SPMS). HCD occurs as a result of infection with the hepatitis C virus (HCV), belonging to the Flaviviridae family. HCV is a blood-borne virus with a positive single-stranded RNA. A vaccine for HCV is not available yet. HCD can lead to liver damage or cancer. In HCD interferon-α therapy (IFN-α) is applied. As with MS, the mechanism of IFN-α therapy is not completely known.
... IFNs are the direct antiviral cytokines as the first line of defense against virus [20,21]. For instant, IFN-γ has potent antiviral activity against HCV in the subgenomic replicon system [22]. Besides, IFN-γ has the most potent immunomodulatory activity of all the interferons. ...
Article
Background: Ligustrum purpurascens has been used as a traditional herb for over 2,000 years in China. This study was design to investigate the modulation of antiviral cytokines and reduction in lung inflammation of virus-infected mice by the glycosides isolated from Ligustrum purpurascens. Methods: Ligustrum purpurascens glycosides (LPG) were isolated from the leaves of Ligustrum purpurascens. Proliferation of spleen lymphocytes were investigated after LPG treatment. The in vitro and in vivo cytokine modulation of LPG was studied. Furthermore, the anti-inflammatory and antiviral activities of LPG, with the potential to reduce inflammatory lung disorders, were investigated by influenza A virus infected mice. Results: LPG could significantly promote the proliferation, and also could stimulate the production of IFN-γ by spleen lymphocytes in a dose-dependent manner. IFN-γ expression level was increased significantly compared to the control and presented a dose-dependent manner in vitro. Furthermore, LPG inhibit the expression of TNF-α and IL-10, which return to normal level in the cyclophosphamide-induced mice model in vivo. Besides, the histopathological analysis indicated LPG reduced acute lung injury in mice infected with influenza virus. Conclusion: This study suggested that LPG could increase the exression of IFN-γ, immunoregulation and decrease lung inflammation of virus-infected mice., et al. Modulation of antiviral cytokine production and lung protection against influenza virus by the glycosides from Ligustrum purpurascens. Asian Toxicol Res.
Article
Despite the excellent antiviral potency of direct-acting antivirals (DAAs) against hepatitis C virus (HCV), emergence of drug-resistant viral mutations remains a potential challenge. Sofobuvir (SOF), a nucleotide analog targeting HCV NS5B - RNA-dependent RNA polymerase (RdRp), constitutes a key component of many anti-HCV cocktail regimens and confers a high barrier for developing drug resistance. The serine to threonine mutation at the amino acid position 282 of NS5B (S282T) is the mostly documented SOF resistance-associated substitution (RAS), but severely hampers the virus fitness. In this study, we first developed new genotype 1b (GT1b) subgenomic replicon cells, denoted PR52D4 and PR52D9, directly from a GT1b clinical isolate. Next, we obtained SOF-resistant and replication-competent PR52D4 replicon by culturing the replicon cells in the presence of SOF. Sequencing analysis showed that the selected replicon harbored two mutations K74R and S282T in NS5B. Reverse genetics analysis showed that while PR52D4 consisting of either single mutation K74R or S282T could not replicate efficiently, the engineering of the both mutations led to a replication-competent and SOF-resistant PR52D4 replicon. Furthermore, we showed that the K74R mutation could also rescue the replication deficiency of the S282T mutation in Con1, another GT1b replicon as well as in JFH1, a GT2a replicon. Structural modeling analysis suggested that K74R might help maintain an active catalytic conformation of S282T by engaging with Y296. In conclusion, we identified the combination of two NS5B mutations S282T and K74R as a novel RAS that confers a substantial resistance to SOF while retains the HCV replication capacity.
Article
Full-text available
Ribavirin, a synthetic guanosine analogue, possesses a broad spectrum of activity against DNA and RNA viruses. It has been previously shown to attenuate the course of fulminant hepatitis in mice produced by murine hepatitis virus strain 3. We therefore studied the effects of ribavirin on murine hepatitis virus strain 3 replication, macrophage production of proinflammatory mediators including TNF, IL-1, and the procoagulant activity (PCA), fgl2 prothrombinase; and Th1/Th2 cytokine production. Although ribavirin had inhibitory effects on viral replication (<1 log), even at high concentrations complete eradication of the virus was not seen. In contrast, at physiologic concentrations (up to 500 μg/ml), ribavirin markedly reduced viral-induced parameters of macrophage activation. With ribavirin treatment, the concentrations of PCA, TNF-α and IL-1β all decreased to basal concentrations: PCA from 941 ± 80 to 34 ± 11 mU/106 cells; TNF-α from 10.73 ± 2.15 to 2.74 ± 0.93 ng/ml; and IL-1β from 155.91 ± 22.62 to 5.74 ± 0.70 pg/ml. The inhibitory effects of ribavirin were at the level of gene transcription as evidenced by Northern analysis. Both in vitro and in vivo, ribavirin inhibited the production of IL-4 by Th2 cells, whereas it did not diminish the production of IFN-γ in Th1 cells. In contrast, ribavirin had no inhibitory effect on TNF-α and IL-1β production in LPS-stimulated macrophages. These results suggest that the beneficial effects of ribavirin are mediated by inhibition of induction of macrophage proinflammatory cytokines and Th2 cytokines while preserving Th1 cytokines.
Article
Full-text available
The interferon (IFN)-induced double-stranded RNA (dsRNA)-activated Ser/Thr protein kinase (PKR) plays a role in the antiviral and antiproliferative effects of IFN. PKR phosphorylates initiation factor eIF2, thereby inhibiting protein synthesis, and also activates the transcription factor, nuclear factor-B (NF-B), by phosphorylating the inhibitor of NF-B, IB. Mice devoid of functional PKR (Pkr°/°) derived by targeted gene disruption exhibit a diminished response to IFN- and poly(rI:rC) (pIC). In embryo fibroblasts derived from Pkr°/° mice, interferon regulatory factor 1 (IRF-1) or guanylate binding protein (Gbp) promoter–reporter constructs were unresponsive to IFN- or pIC but response could be restored by co-transfection with PKR. The lack of responsiveness could be attributed to a diminished activation of IRF-1 and/or NF-B in response to IFN- or pIC. Thus, PKR acts as a signal transducer for IFN-stimulated genes dependent on the transcription factors IRF-1 and NF-B.
Article
Full-text available
Although hepatitis C virus (HCV) infection is an emerging global epidemic causing severe liver disorders, the molecular mechanisms of HCV pathogenesis remain elusive. The NS5A nonstructural protein of HCV contains several proline-rich sequences consistent with Src homology (SH) 3-binding sites found in cellular signaling molecules. Here, we demonstrate that NS5A specifically bound to growth factor receptor-bound protein 2 (Grb2) adaptor protein. Immunoblot analysis of anti-Grb2 immune complexes derived from HeLa S3 cells infected with a recombinant vaccinia virus (VV) expressing NS5A revealed an interaction between NS5A and Grb2 in vivo. An inactivating point mutation in the N-terminal SH3 domain, but not in the C-terminal SH3 domain, of Grb2 displayed significant diminished binding to NS5A. However, the same mutation in both SH3 regions completely abrogated Grb2 binding to NS5A, implying that the two SH3 domains bind in cooperative fashion to NS5A. Further, mutational analysis of NS5A assigned the SH3-binding region to a proline-rich motif that is highly conserved among HCV genotypes. Importantly, phosphorylation of extracellular signal-regulated kinases 1 and 2 (ERK1/2) was inhibited in HeLa S3 cells infected with NS5A-expressing recombinant VV but not recombinant VV control. Additionally, HeLa cells stably expressing NS5A were refractory to ERK1/2 phosphorylation induced by exogenous epidermal growth factor. Moreover, the coupling of NS5A to Grb2 in these cells was induced by epidermal growth factor stimulation. Therefore, NS5A may function to perturb Grb2-mediated signaling pathways by selectively targeting the adaptor. These findings highlight a viral interceptor of cellular signaling with potential implications for HCV pathogenesis.
Article
Hepatitis C virus (HCV) infection is a global health problem affecting an estimated 170 million individuals worldwide. We report the identification of multiple independent adaptive mutations that cluster in the HCV nonstructural protein NS5A and confer increased replicative ability in vitro. Among these adaptive mutations were a single amino acid substitution that allowed HCV RNA replication in 10% of transfected hepatoma cells and a deletion of 47 amino acids encompassing the interferon (IFN) sensitivity determining region (ISDR). Independent of the ISDR, IFN-α rapidly inhibited HCV RNA replication in vitro. This work establishes a robust, cell-based system for genetic and functional analyses of HCV replication.
Article
Hepatitis C virus (HCV) NS5A is a phosphoprotein that possesses a cryptic trans-activation activity. To investigate its potential role in viral replication, we searched for the cellular proteins interacting with NS5A protein by yeast two-hybrid screening of a human hepatocyte cDNA library. We identified a newly discovered soluble N-ethylmaleimide-sensitive factor attachment protein receptor-like protein termed human vesicle-associated membrane protein-associated protein of 33 kDa (hVAP-33). In vitro binding assay and in vivo coimmunoprecipitation studies confirmed the interaction between hVAP-33 and NS5A. Interestingly, hVAP-33 was also shown to interact with NS5B, the viral RNA-dependent RNA polymerase. NS5A and NS5B bind to different domains of hVAP-33: NS5A binds to the C-terminus, whereas NS5B binds to the N-terminus of hVAP-33. Immunofluorescent staining showed a significant colocalization of hVAP-33 with both NS5A and NS5B proteins. hVAP-33 contains a coiled-coil domain followed by a membrane-spanning domain at its C-terminus. Cell fractionation analysis revealed that hVAP-33 is predominantly associated with the ER, the Golgi complex, and the prelysosomal membrane, consistent with its potential role in intracellular membrane trafficking. These interactions provide a mechanism for membrane association of the HCV RNA replication complex and further suggest that NS5A is a part of the viral RNA replication complex.
Article
The liver performs a wide array of functions, a few of which include the synthesis and secretion of most of the plasma proteins, including the lipoproteins, cholesterol, and bile acid metabolism, and detoxification of the blood. In vitro analysis of most liver functions has been hampered by the difficulties encountered in isolating and maintaining functional cultures of primary hepatocytes. Although in vivo the liver has an amazing capacity for regeneration, hepatocytes in culture have limited proliferation capacity and are normally short-lived. We have developed methods for the isolation and cultivation of highly differentiated primate hepatocyte cultures that can be maintained for over 100 d without significant loss of differentiated function. This system has been used in our lab for the analysis of lipoprotein synthesis and hepatotropic virus replication. This chapter is designed to provide a detailed methodology of our approach. Numerous alternative hepatocyte cultivation systems have been described, but owing to space limitations, these systems will not be described here. A number of excellent reviews are dedicated to this subject, one of which is in a previous volume of this series (1), and another that is an entire book dedicated to the subject (2).