53
- - 1 Hepatitis C Virus Infection Promotes Hepatic Gluconeogenesis 1 through an NS5A-Mediated, FoxO1-Dependent Pathway 2 3 Running title: HCV promotes hepatic gluconeogenesis 4 5 Lin Deng, 1 Ikuo Shoji, 1 Wataru Ogawa, 2 Shusaku Kaneda, 1 Tomoyoshi Soga, 3 Da-peng 6 Jiang, 1 Yoshi-Hiro Ide, 1 and Hak Hotta 17 8 Division of Microbiology, Center for Infectious Diseases 1 and Division of Diabetes, Metabolism 9 and Endocrinology 2 , Kobe University Graduate School of Medicine, 7-5-1 Kusunoki-cho, 10 Chuo-ku, Kobe 650-0017; Institute for Advanced Biosciences, Keio University 3 , 246-2 11 Mizukami, Kakuganji, Tsuruoka, Yamagata 997-0052, Japan. 12 13 Word count for the abstract: 244 14 Word count for the text: 5152 15 Key words: Hepatitis C virus; NS5A; Gluconeogenesis; FoxO1; Diabetes mellitus 16 17 *Correspondence to Hak Hotta, Division of Microbiology, Center for Infectious Disease, Kobe 18 University Graduate School of Medicine, 7-5-1 Kusunoki-cho, Chuo-ku, Kobe 650-0017, Japan. 19 Phone: +81-78-382-5500. Fax: +81-78-382-5519. E-mail: [email protected] 20 Copyright © 2011, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved. J. Virol. doi:10.1128/JVI.00146-11 JVI Accepts, published online ahead of print on 22 June 2011 on May 20, 2018 by guest http://jvi.asm.org/ Downloaded from

Hepatitis C Virus Infection Promotes Hepatic Glucon …jvi.asm.org/content/early/2011/06/22/JVI.00146-11.full.pdf · INTRODUCTION Hepatitis C virus (HCV) is a small, enveloped RNA

  • Upload
    ngoque

  • View
    219

  • Download
    2

Embed Size (px)

Citation preview

- - 1

Hepatitis C Virus Infection Promotes Hepatic Gluconeogenesis 1

through an NS5A-Mediated, FoxO1-Dependent Pathway 2

3

Running title: HCV promotes hepatic gluconeogenesis 4

5

Lin Deng,1

Ikuo Shoji,1 Wataru Ogawa,

2 Shusaku Kaneda,

1 Tomoyoshi Soga,

3 Da-peng 6

Jiang,1 Yoshi-Hiro Ide,

1 and Hak Hotta

1∗∗∗∗ 7

8

Division of Microbiology, Center for Infectious Diseases1 and Division of Diabetes, Metabolism 9

and Endocrinology2, Kobe University Graduate School of Medicine, 7-5-1 Kusunoki-cho, 10

Chuo-ku, Kobe 650-0017; Institute for Advanced Biosciences, Keio University3, 246-2 11

Mizukami, Kakuganji, Tsuruoka, Yamagata 997-0052, Japan. 12

13

Word count for the abstract: 244 14

Word count for the text: 5152 15

Key words: Hepatitis C virus; NS5A; Gluconeogenesis; FoxO1; Diabetes mellitus 16

17

*Correspondence to Hak Hotta, Division of Microbiology, Center for Infectious Disease, Kobe 18

University Graduate School of Medicine, 7-5-1 Kusunoki-cho, Chuo-ku, Kobe 650-0017, Japan. 19

Phone: +81-78-382-5500. Fax: +81-78-382-5519. E-mail: [email protected] 20

Copyright © 2011, American Society for Microbiology and/or the Listed Authors/Institutions. All Rights Reserved.J. Virol. doi:10.1128/JVI.00146-11 JVI Accepts, published online ahead of print on 22 June 2011

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 2

ABSTRACT 21

22

Chronic hepatitis C virus (HCV) infection is often associated with type 2 diabetes. However, 23

the precise mechanism underlying this association is still unclear. Here, using Huh-7.5 cells 24

either harboring HCV-1b RNA replicons or infected with HCV-2a, we showed that HCV 25

transcriptionally upregulated the genes for phosphoenolpyruvate carboxykinase (PEPCK) and 26

glucose 6-phosphatase (G6Pase), the rate-limiting enzymes for hepatic gluconeogenesis. In this 27

way, HCV enhanced the cellular production of glucose 6-phosphate (G6P) and glucose. PEPCK 28

and G6Pase gene expressions are controlled by the transcription factor forkhead box O1 29

(FoxO1). We observed that, although neither the mRNA levels nor the protein levels of FoxO1 30

expression were affected by HCV, phosphorylation of FoxO1 at Ser319 was markedly 31

diminished in HCV-infected cells compared to the control cells, resulting in increased nuclear 32

accumulation of FoxO1, which is essential for sustaining its transcriptional activity. It was 33

unlikely that the decreased FoxO1 phosphorylation was mediated through Akt inactivation as 34

we observed increased phosphorylation of Akt at Ser473 in HCV-infected cells compared to the 35

control. By using specific inhibitors of c-Jun N-terminal kinase (JNK) and reactive oxygen 36

species (ROS), we demonstrated that HCV infection induced JNK activation via increased 37

mitochondrial ROS production, resulting in the decreased FoxO1 phosphorylation, FoxO1 38

nuclear accumulation and, eventually, increased glucose production. We also found that HCV 39

NS5A mediated increased ROS production and JNK activation, which is directly linked with 40

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 3

the FoxO1-dependent increased gluconeogenesis. Taken together, these observations suggest 41

that HCV promotes hepatic gluconeogenesis through an NS5A-mediated, FoxO1-dependent 42

pathway.43

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 4

INTRODUCTION 44

45

Hepatitis C virus (HCV) is a small, enveloped RNA virus that belongs to the genus Hepacivirus 46

of the family Flaviviridae, and the molecular mechanisms underlying its viral replication are 47

currently being unraveled (40). The HCV genome encodes a single polyprotein of about 3,000 48

amino acids, which is cleaved by host and viral proteases to generate at least 10 viral proteins, 49

such as core, envelope 1 (E1) and E2, p7, NS2, NS3, NS4A, NS4B, NS5A and NS5B. HCV can 50

be classified into seven genotypes, with each genotype further classified into a number of 51

subtypes, such as HCV-1a and HCV-1b (18, 24, 59). 52

53

HCV infects more than 120 million people worldwide (57). Persistent HCV infection causes not 54

only liver diseases (chronic hepatitis, liver cirrhosis, hepatocellular carcinoma) but also 55

extrahepatic manifestations, such as type 2 diabetes (2, 11, 20, 23). While it is known that liver 56

cirrhosis impairs the glucose metabolism of the liver, there are some reports showing that HCV-57

infected patients over 40 years of age have an increased risk of type 2 diabetes compared with 58

individuals without HCV infection (43). In addition, insulin receptor substrate (IRS)-1/PI3-59

kinase signaling was more impaired in HCV-infected patients than in non-HCV-infected 60

controls (3). These studies imply that HCV infection may directly predispose the host toward 61

type 2 diabetes. However, the precise mechanisms are poorly understood. 62

63

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 5

Hepatocytes play an important role in maintaining plasma glucose homeostasis by adjusting the 64

balance between hepatic glucose production and utilization via the gluconeogenic and glycolytic 65

pathways, respectively. It has been proposed that increased hepatic glucose production is a 66

major feature of type 2 diabetes (13). It is also known that hyperglycemia and the subsequent 67

development of type 2 diabetes mellitus result, at least in part, from impaired insulin signaling 68

together with elevated glucagon (5, 19). Hepatic glucose production and utilization, the 69

physiologically opposed cascades, are regulated, at least in part, at the transcriptional level of 70

the glucose 6-phosphatase (G6Pase) and glucokinase (GK) genes, which catalyze the last and 71

the first rate-limiting steps in gluconeogenesis and glycolysis, respectively. A number of studies 72

have shown that fasting/feeding (or hormones) controls the transcription of these two enzymes 73

in the opposite directions. G6Pase transcription is negatively regulated by insulin or feeding and 74

markedly increased in a fasting state (62). On the other hand, GK transcription is positively 75

regulated by insulin or feeding and markedly decreased in a fasting state (33). It has also been 76

reported that the gene expression of gluconeogenic and glycolytic enzymes, such as G6Pase, GK 77

and phosphoenolpyruvate carboxykinase (PEPCK), another rate-limiting enzyme for hepatic 78

gluconeogenesis, is regulated by certain transcription factors, including forkhead box O1 79

(FoxO1) (26, 50, 54), hepatic nuclear factor-4α (HNF-4α) (26), Krüppel-like factor 15 (KLF15) 80

(64) and cyclic AMP (cAMP) response-element binding protein (CREB) (52, 56). Deregulation 81

of the otherwise balanced control of hepatic glucose homeostasis would potentially lead to 82

hyperglycemia and eventually type 2 diabetes. 83

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 6

84

In this study, by using Huh-7.5 cells harboring HCV-1b RNA replicons, i.e., either a 85

subgenomic RNA replicon (SGR) or a full-genomic RNA replicon (FGR) (37), and cells 86

infected with HCV-2a (14, 37, 39), we investigated the possible effects of HCV on glucose 87

metabolism. We report here that HCV promotes hepatic gluconeogenesis, resulting in increased 88

cellular glucose production in hepatocytes via an NS5A-mediated, FoxO1-dependent pathway. 89

90

MATERIALS AND METHODS 91

92

Cells, HCV RNA replicons and virus. The human hepatoma-derived cell line Huh-7.5 (7) was 93

kindly provided by Dr. C. M. Rice (The Rockefeller University, New York, NY). SGR and FGR 94

were prepared using pFK5B/2884Gly (41) (a kind gift from Dr. R. Bartenschlager, University 95

of Heidelberg, Heidelberg, Germany) and pON/C-5B (31) (a kind gift from Dr. N. Kato, 96

Okayama University, Okayama, Japan), respectively. The SGR and FGR cells are of polyclonal 97

origin to avoid clonal variation. The pFL-J6/JFH1 plasmid that encodes the entire viral genome 98

of a chimeric strain of HCV-2a (J6/JFH1) (39) was kindly provided by Dr. Rice. The HCV RNA 99

genome was transcribed in vitro from pFL-J6/JFH1 and transfected into Huh-7.5 cells to yield 100

infectious HCV particles, as described previously (14). A cell culture-adapted P-47 strain (9, 101

14) was used throughout the experiments. Virus infection was performed at a multiplicity of 102

infection (moi) of 2.0. Virus infectivity was measured by indirect immunofluorescence analysis, 103

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 7

as described below, and expressed as cell-infecting units/ml. In some experiments, SGR and 104

FGR cells, as well as HCV-infected cells at 5 days after virus infection, were treated with 1,000 105

IU/ml of interferon-α (IFN) (Sigma Chemical, St. Louis, MO) for 10 days to eliminate HCV 106

replication. 107

108

Plasmid construction. Expression plasmids for core, p7, NS2, NS3, NS3/4A, NS4A, NS4B, 109

NS5A and NS5B were reported elsewhere (15, 32). 110

111

Real-time quantitative RT-PCR. Total cellular RNA was isolated using the RNAiso reagent 112

(Takara, Kyoto, Japan), and cDNA was generated using a QuantiTect Reverse Transcription 113

system (Qiagen, Valencia, CA). Real-time quantitative PCR was performed using SYBR 114

Premix Ex Taq (Takara) with SYBR Green chemistry on an ABI PRISM 7000 system (Applied 115

Biosystems, Foster City, CA), as reported previously (37). β-glucuronidase and GAPDH were 116

used as internal controls. The primers used are shown in Table 1. 117

118

G6P production assay. Huh-7.5 cells seeded in a 10-cm dish at a density of 1.0 × 106 cells/dish 119

were infected with HCV or left uninfected. At different time points after infection, the cells 120

were washed twice with 5% mannitol solution and covered with methanol (1 ml) containing 25 121

µM each of four internal standards (3-aminopyrolidine, L-methionine sulfone, trimesate, and 2-122

morpholinoethanesulfonic acid) for enzyme inactivation. The methanol and cell mixtures were 123

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 8

collected and mixed with Milli-Q water and chloroform at ratios of 2:1:2. Both the medium and 124

cell sample solutions were then centrifuged at 20,000 × g for 15 min, and the aqueous layers 125

were collected for centrifugal filtration through a 5-kDa-cutoff filter at 9,000 × g for 2 h. The 126

extracted metabolites were concentrated with a centrifugal concentrator and stored at −80°C 127

until analysis. G6P concentrations were measured by capillary electrophoresis time-of-flight 128

mass spectrometry (CE-TOFMS), and the results were normalized to cell number as described 129

previously (60, 61). 130

131

Glucose production assay. Culture medium was replaced with 1 ml of glucose production 132

buffer consisting of glucose-free DMEM (Sigma Chemical), without phenol red, supplemented 133

with gluconeogenic substrate (2 mM sodium pyruvate and 20 mM sodium lactate). After 24 h of 134

incubation, the medium was collected and the total glucose concentration was measured by 135

using a commercial kit (Glucose CII Test Wako; Wako Pure Chemical Industries, Osaka, Japan) 136

and normalized to the cellular protein content. As the base-line of glucose production, glucose-137

free DMEM with neither sodium pyruvate nor sodium lactate was used. Glucose production via 138

gluconeogenesis = total glucose production – base-line glucose production. 139

140

Luciferase reporter assay. PEPCK gene promoter (−1263/+225) and a deletion mutant 141

(−998/+225) were inserted into the pGL3 Luciferase reporter plasmid (Promega, Madison, WI). 142

They were designated rPEPCK-P5(−1263)-pGL3basic and rPEPCK-P4(−998)-pGL3basic. 143

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 9

pRL-CMV-Renilla (Promega), which expresses Renilla luciferase, was used as an internal 144

control. Huh-7.5 cells prepared in a 12-well tissue culture plate at a density of 1.0 x 105 145

cells/well were transiently transfected with pRL-CMV-Renilla and rPEPCK-P5(−1263)-146

pGL3basic or rPEPCK-P4(−998)-pGL3basic in the presence of pEF1/NS4A, pEF1/NS5A or a 147

control vector (32). After 48 h, a luciferase assay was performed by using the Dual-Luciferase 148

Reporter assay system (Promega). Firefly and Renilla luciferase activities were measured on a 149

Lumat LB 9501 luminometer (Berthold, Bad Wildbad, Germany). Firefly luciferase activity 150

was normalized to Renilla luciferase activity for each sample. 151

152

Detection of mitochondrial reactive oxygen species (ROS). Mitochondrial ROS production 153

was analyzed as described previously (14). Briefly, cells seeded on glass coverslips in a 24-well 154

plate were incubated with 5 µM MitoSOXTM

Red (Molecular Probes, Eugene, OR) at 37ºC for 155

10 min, then fixed with 3.7% paraformaldehyde and observed under a confocal laser scanning 156

microscope (Carl Zeiss, Oberkochen, Germany). When needed, the fixed cells were subjected 157

to indirect immunofluorescence analysis to confirm HCV infection or NS5A expression, as 158

described below. 159

160

Indirect immunofluorescence. Huh7.5 cells seeded on glass coverslips in a 24-well plate were 161

infected with HCV or transfected with NS5A expression plasmid. At 5 days post-infection or 3 162

days post-transfection, the cells were fixed with 3.7% paraformaldehyde in PBS for 15 min at 163

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 10

room temperature, and permeabilized in 0.1% Triton X-100 in PBS for 15 min at room 164

temperature. Mock-infected or empty vector-transfected cells were similarly treated as controls 165

for comparison. After being washed with PBS twice, cells were consecutively stained with 166

primary and secondary antibodies. Primary antibodies used were anti-FoxO1 rabbit monoclonal 167

antibody (Cell Signaling Technology, Danvers, MA), anti-NS5A mouse monoclonal antibody 168

(Chemicon International, Temecula, CA) and HCV-infected patient’s serum. Secondary 169

antibodies used were Alexa Fluor 488-conjugated goat anti-rabbit IgG, Alexa-Fluor 594-170

conjugated goat anti-mouse IgG or anti-human IgG (Molecular Probes) and FITC-conjugated 171

goat anti-mouse IgG or anti-human IgG (MBL, Nagoya, Japan). The stained cells were 172

observed under a confocal laser scanning microscope (Carl Zeiss). 173

174

Cell fractionation and immunoblotting. Nuclear and cytoplasmic extracts from cells were 175

prepared using an NE-PER Nuclear and Cytoplasmic Extraction Reagents kit (Pierce Chemical, 176

Rockford, IL). For immunoblotting, cells were lysed with SDS sample buffer, and equal 177

amounts of protein were subjected to SDS-polyacrylamide gel electrophoresis and transferred 178

onto a polyvinylidene difluoride membrane (Millipore, Bedford, MA), which was then 179

incubated with the respective primary antibody. The primary antibodies used were mouse 180

monoclonal antibodies against HCV core (clone 2H9; a kind gift from Dr. T. Wakita, 181

Department of Virology II, National Institute of Infectious Diseases, Tokyo, Japan), NS3, NS4A, 182

NS5A, GAPDH (Chemicon), FoxO1 (Sigma Chemical), phospho-Akt (Ser473) (Cell Signaling 183

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 11

Technology), c-Myc (9E10; Santa Cruz Biotechnology, Santa Cruz, CA); rabbit polyclonal 184

antibodies against phospho-FoxO1 (Ser139), Oct-1 (Santa Cruz Biotechnology), c-Jun N-185

terminal kinase (JNK), phospho-JNK (Thr183/Tyr185), c-Jun, phospho-c-Jun (Ser63) and Akt 186

(Cell Signaling Technology); and goat polyclonal antibody against HSP60 (Santa Cruz 187

Biotechnology). Horseradish peroxidase-conjugated goat anti-mouse immunoglobulin (IgG), 188

goat anti-rabbit IgG (Molecular Probes) and donkey anti-goat IgG (Santa Cruz Biotechnology) 189

were used to visualize the respective proteins by means of an enhanced chemiluminescence 190

detection system (ECL; GE Healthcare, Buckinghamshire, UK). 191

192

Statistical analysis. Results were expressed as means ± SEM. Statistical significance was 193

evaluated by ANOVA, and defined as P < 0.05. 194

195

RESULTS 196

197

HCV upregulates gene expression of PEPCK and G6Pase and downregulates gene 198

expression of GK. 199

We first examined the expression levels of the genes for the rate-limiting enzymes in hepatic 200

gluconeogenesis, PEPCK and G6Pase, and of those for GK, which catalyzes the first step of 201

glycolysis, by means of real-time quantitative RT-PCR analysis. We observed that the PEPCK 202

and G6Pase genes were transcriptionally activated in SGR- and FGR-harboring cells (Fig. 1A, 203

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 12

1B, left panels). Similarly, the PEPCK and G6Pase genes were upregulated in HCV-infected 204

cells in a time-dependent manner, starting from 3 or 5 days post-infection (dpi) up to 14 dpi 205

(Fig. 1A, 1B, middle panels). On the other hand, the GK gene was transcriptionally 206

downregulated in SGR- and FGR-harboring cells and HCV-infected cells in a time-dependent 207

manner (Fig. 1C). It is noteworthy that gene expression of six glycolytic enzymes (not 208

including GK) was observed to be upregulated in HCV-infected cells at 1 dpi (16). 209

210

When IFN treatment eliminated HCV from the cells, the observed upregulation of PEPCK and 211

G6Pase gene expression, as well as the downregulation of GK gene expression in SGR- and 212

FGR-harboring cells and HCV-infected cells, were cancelled (Fig. 1A, 1B, 1C, left and right 213

panels). Thus, our results suggest that there was a trend towards an increase in gluconeogenesis 214

in SGR- and FGR-harboring cells and HCV-infected cells. In subsequent studies we further 215

examined whether or not HCV replication was correlated with gluconeogenesis. 216

217

HCV promotes cellular production of glucose and G6P. 218

We then examined the effect of HCV on cellular glucose production. The results showed that 219

SGR- and FGR-harboring cells and HCV-infected cells produced greater amounts of glucose 220

than did the control cells (Fig. 2A, upper and middle panels). IFN treatment cancelled the 221

enhanced glucose production in SGR- and FGR-harboring cells and in HCV-infected cells (Fig. 222

2A, upper and bottom panels). We also investigated the production of G6P, which is an 223

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 13

important precursor molecule that is converted to glucose in the gluconeogenesis pathway, by 224

means of metabolome analysis. As shown in Fig. 2B, a significantly higher level of G6P was 225

accumulated in HCV-infected cells than in the controls. Taken together, these results indicate 226

that HCV indeed promotes hepatic gluconeogenesis to cause hyperglycemia. In the following 227

analyses, we examined the possible mechanisms of HCV-induced increased gluconeogenesis. 228

229

HCV suppresses FoxO1 phosphorylation at Ser319, leading to the nuclear accumulation of 230

FoxO1. 231

It has been demonstrated that FoxO1 in hepatocytes enhances gluconeogenesis through the 232

transcriptional activation of various genes, including G6Pase and PEPCK (25). To investigate 233

the possible effects of FoxO1 on HCV-induced gluconeogenesis, we examined the gene 234

expression levels of FoxO1 by real-time quantitative RT-PCR analysis. As shown in Fig. 3A, 235

there was neither upregulation nor downregulation of FoxO1 gene expression in SGR- or FGR-236

harboring cells or HCV-infected cells. The FoxO1 transcription factor is controlled by various 237

post-translational modifications, which include phosphorylation, ubiquitylation and acetylation. 238

The phosphorylated form of FoxO1 is exported from the nucleus and thereby loses its 239

transcriptional function (30). We therefore examined the phosphorylation status of FoxO1 at 240

Ser319, which is critical for FoxO1 nuclear exclusion (72). The results showed that FoxO1 241

phosphorylation at Ser319 was markedly suppressed in HCV-infected cells from 4 dpi up to 8 242

dpi, compared to that in the HCV-negative control cells (Fig. 3B, first panel), in a time-243

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 14

dependent manner that was roughly the inverse of the pattern observed for PEPCK and G6Pase 244

mRNA upregulation (Fig. 1A and 1B) and glucose production (Fig. 2A), while the total protein 245

expression levels of FoxO1 were unchanged (Fig. 3B, second panel). In this connection, 246

Banerjee et al. reported that FoxO1 phosphorylation at Ser256 was also inhibited in HCV-247

infected cells (4). Since FoxO1 is known to be phosphorylated by Akt so as to be exported from 248

the nucleus and transcriptionally inactivated (38), we examined whether Akt function was 249

suppressed through its impaired phosphorylation in HCV-infected cells. The result obtained 250

revealed that this was not the case; Akt phosphorylation was enhanced in HCV-infected cells 251

from 4 dpi up to 6 dpi compared with the control cells (Fig. 3B, third panel) while the total 252

protein expression levels of Akt were comparable (Fig. 3B, fourth panel). This result is 253

consistent with a recent observation by Burdette et al. (10) that Akt phosphorylation was 254

elevated in HCV-infected cells. These data suggest that the observed decrease in FoxO1 255

phosphorylation in HCV-infected cells is caused by a mechanism independent of Akt. 256

257

Next, we tested whether HCV indeed promoted FoxO1 nuclear accumulation. The majority of 258

FoxO1 was accumulated in the nuclear fraction in HCV-infected cells (Fig. 3C, second panel, 259

lanes 2 and 4) whereas in control cells FoxO1 was distributed in both the nuclear and 260

cytoplasmic fractions (lanes 1 and 3). Taken together, these results suggest that HCV 261

suppressed FoxO1 phosphorylation, leading to nuclear accumulation of FoxO1. 262

263

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 15

HCV-induced JNK activation is involved in the suppression of FoxO1 phosphorylation. 264

Recent studies have demonstrated that a signaling pathway that involves the stress-sensitive 265

serine/threonine kinase JNK regulates FoxO at multiple levels (36, 66). We therefore 266

investigated whether HCV induced JNK activation in Huh-7.5 cells. As shown in Fig. 4A, the 267

amount of phosphorylated (activated) JNK markedly increased in HCV-infected cells in a time-268

dependent manner similar to that observed for the suppression of FoxO1 phosphorylation, while 269

the total expression levels of JNK were unchanged. As a result, c-Jun, a key substrate for JNK, 270

was phosphorylated (activated) in HCV-infected cells, but not in the mock-infected control 271

cells. It should also be noted that the total expression levels of c-Jun in HCV-infected cells were 272

significantly higher than those in the mock-infected control cells, suggesting that c-Jun 273

activation through its phosphorylation stabilizes c-Jun protein expression in HCV-infected cells, 274

as has been proposed by Zhang et al. (71). 275

276

We next sought to determine whether JNK activation was involved in the HCV-induced 277

suppression of FoxO1 phosphorylation. HCV-infected cells at 5 days after virus infection were 278

treated with the specific JNK inhibitor SP600125 (20 µM) (6) for 24 h. The catalytic JNK 279

activity was assayed by following the phosphorylation of c-Jun. As shown in Fig. 4B, 280

SP600125 clearly prevented the phosphorylation of c-Jun, and concomitantly recovered the 281

suppression of FoxO1 phosphorylation in HCV-infected cells. These results suggest that HCV 282

activates the JNK/c-Jun signaling pathway, which induces the nuclear accumulation of FoxO1 283

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 16

by reducing its phosphorylation status. 284

285

HCV-induced mitochondrial ROS production is involved in FoxO1 phosphorylation 286

suppression, FoxO1 nuclear accumulation and increased glucose production through JNK 287

activation. 288

We previously reported that HCV infection increases mitochondrial ROS production (14). JNK 289

is known to be activated by ROS (35). We therefore sought to determine whether the HCV-290

induced increase in ROS production is an event occurring upstream of JNK activation by HCV. 291

Pretreatment of HCV-infected cells (at 6 dpi) with 5 mM N-acetyl cysteine (NAC) (a general 292

antioxidant) for 2 h significantly reduced the HCV-induced increase in ROS (Fig. 5A and 5B), 293

as revealed by using MitoSOX, a fluorescent probe specific for superoxide that selectively 294

accumulates in the mitochondrial compartment. As shown in Fig. 5C, NAC clearly prevented 295

the phosphorylation of JNK, and concomitantly recovered the suppression of FoxO1 296

phosphorylation in HCV-infected cells. These results suggest that HCV-induced ROS 297

production is involved in JNK activation, which results in the inhibition of FoxO1 298

phosphorylation. 299

300

We next investigated the effects of JNK activation and ROS production on subcellular 301

localization of FoxO1 in HCV-infected cells by indirect immunofluorescence staining. As 302

shown in Fig. 6A and 6B, FoxO1 was localized predominantly in the cytoplasm in mock-303

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 17

infected control cells. On the other hand, nuclear accumulation of FoxO1 was clearly observed 304

in approximately 35% of HCV-infected cells at 5 dpi. Treatment of HCV-infected cells with a 305

JNK inhibitor (SP600125; 20 µM, 24 h) or an antioxidant (NAC; 5 mM, 2 h) significantly 306

inhibited HCV-induced FoxO1 nuclear accumulation. 307

308

To further verify the role played by JNK activation and ROS production in HCV-induced 309

hepatic gluconeogenesis, the glucose production in SP600125- or NAC-treated HCV-infected 310

cells was assessed. Treatment with SP600125 or NAC significantly impaired HCV-induced 311

increased glucose production at 7 dpi (Fig. 6C), but did not affect the overall abundance of 312

HCV NS3 protein (Fig. 6D). We also examined the possible effects of SP600125 or NAC on 313

HCV RNA replication and infectious virus production. The results obtained revealed that 314

treatment with SP600125 (20 µM, 24 h) or NAC (5 mM, 2 h) barely affected HCV RNA 315

replication (Fig. 6E). On the other hand, a tendency was noted that infectious virus production 316

was only slightly suppressed by SP600125, but not by NAC (Fig. 6F). A short-term inhibition 317

of glucose production might not sufficiently affect HCV RNA replication or virus production. 318

319

Taken together, these results indicate that ROS-mediated JNK activation plays a key role in the 320

suppression of FoxO1 phosphorylation, nuclear accumulation of FoxO1, and enhancement of 321

glucose production in HCV-infected cells. 322

323

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 18

HCV NS5A is involved in the enhancement of glucose production. 324

To examine which HCV protein(s) is involved in the enhancement of gluconeogenesis, 325

expression constructs of each of the HCV viral proteins were transfected into Huh-7.5 cells, and 326

the gene expression levels of PEPCK and G6Pase were examined by real-time quantitative RT-327

PCR analysis. We observed that NS5A significantly promoted G6Pase gene expression (Fig. 328

7A). Moreover, both the NS5A and NS4A proteins significantly enhanced PEPCK gene 329

expression at 3 and 5 days post-transfection, respectively (Fig. 7B). The expression of each of 330

the HCV proteins except NS2 was verified by immunoblot analysis (Fig. 7C). NS2 was 331

reported to be unstable and rapidly degraded by the proteasome (22). 332

333

Next, we performed a luciferase reporter assay to examine the possible effects of NS5A and 334

NS4A on PEPCK promoter activities. The construct rPEPCK-P5(−1263)-pGL3basic carries 335

1263 bp of the PEPCK 5’-flanking region (−1263 PEPCK) and is used to monitor PEPCK 336

promoter activity. The results demonstrated that the levels of PEPCK promoter activities were 337

significantly higher in both NS5A- and NS4A-expressing cells than in the control cells (Fig. 338

6C). Interestingly, when the region of the PEPCK promoter from −1263 to −998 was deleted, 339

the activation of PEPCK promoter activity in cells expressing NS5A and NS4A was abolished. 340

These results confirmed that NS5A and NS4A activate the PEPCK promoter, leading to an 341

increase in PEPCK mRNA expression. Database searches of the deleted sequence did not find 342

any potential binding sequences for transcription factors (data not shown). 343

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 19

344

Recent data suggest that ROS production is induced in NS5A-expressing cells (17) or in 345

hepatocytes of NS5A transgenic mice (68). We therefore sough to determine whether NS5A 346

contributes to increased hepatic gluconeogenesis through the induction of ROS production. The 347

NS5A expression plasmid was transfected into Huh-7.5 cells, and ROS production was assessed 348

by MitoSOX at 3 days post-transfection. As shown in Fig. 8A and 8B, approximately 30% of 349

NS5A-expressing cells displayed a much stronger signal than that observed with the vector-350

transfected control. 351

352

We then examined whether NS5A mediated JNK/c-Jun activation and FoxO1 phosphorylation 353

inhibition. The results obtained revealed that both phosphorylation level at Ser63 and total 354

expression level of c-Jun were upregulated in NS5A-expressing cells compared to the control 355

cells transfected with the vector plasmid or in cells expressing the other HCV proteins (Fig. 8C 356

and 8D, top two panels). Concomitantly, FoxO1 phosphorylation at Ser319 was clearly 357

suppressed in NS5A- and NS4A-expressing cells compared to the control cells (Fig. 8C, 358

compare lanes 6, 5 and 1 in the third panel). NS4A, a small protein of ca. 7 kDa, forms a stable 359

complex with NS3 to function as a cofactor for NS3 serine protease and RNA helicase activities 360

(51). We previously reported that NS4A caused mitochondrial damage when expressed alone, 361

but not when coexpressed with NS3 (47). We therefore speculated that the otherwise observed 362

decrease in FoxO1 phosphorylation in NS4A-expressing cells might be canceled when 363

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 20

coexpressed with NS3. To verify this notion, we tested the FoxO1 phosphorylation in cells 364

coexpressing NS3 and NS4A. As had been expected, FoxO1 phosphorylation levels did not 365

differ between NS3/4A-coexpressing cells and the vector-transfected control (Fig. 8C, compare 366

lanes 4 and 1). 367

368

Notably, we observed that HCV core protein did not alter the phosphorylation status of c-Jun 369

and FoxO1 (Fig. 8C, compare lanes 2 and 1), with result being consistent with what was 370

observed for gene expression levels of PEPCK and G6Pase in HCV core-expressing cells (Fig. 371

7A and 7B). These results imply that core is not primarily involved in HCV-induced increased 372

gluconeogenesis under our experimental condition. Similarly, other HCV nonstructural proteins, 373

such as NS4B and NS5B, did not significantly influence the phosphorylation status of c-Jun and 374

FoxO1 (Fig. 8D). 375

376

In order to further verify the effect of NS5A on the nuclear accumulation of FoxO1, we 377

examined the subcellular localization of FoxO1 in NS5A-expressing cells by indirect 378

immunofluorescence staining. As shown in Fig. 8E and 8F, nuclear accumulation of FoxO1 was 379

clearly observed in approximately 25% of NS5A-expressing cells but not in the vector-380

transfected control. These results suggest that NS5A activates JNK/c-Jun signaling pathway via 381

increased ROS production, which results in the decreased phosphorylation and nuclear 382

accumulation of FoxO1. 383

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 21

384

Finally, we examined the effects of NS5A and NS4A on glucose production. As shown in Fig. 9, 385

the amounts of glucose were significantly increased in culture supernatants of NS5A- and 386

NS4A-expressing cells, compared with the control, at 5 days post-transfection. Again, it is 387

reasonable to assume that the observed increase in glucose production in NS4A-expressing cells 388

might be canceled when coexpressed with NS3. 389

390

These results collectively suggest that NS5A plays a role, at least to some extent, in HCV-391

induced enhancement of hepatic gluconeogenesis. 392

393

DISCUSSION 394

395

Hepatocytes play an important role in maintaining plasma glucose homeostasis by adjusting the 396

balance between hepatic glucose production and utilization via the gluconeogenic and 397

glycolytic pathways, respectively. We previously reported that HCV suppresses cellular glucose 398

uptake by downregulating the surface expression of glucose transporters GLUT1 and GLUT2 399

(37). In this study, we have demonstrated that HCV promotes FoxO1-mediated hepatic 400

gluconeogenesis, as evidenced by the increased accumulation of FoxO1 in the nucleus via the 401

reduction of its phosphorylation status (Figs. 3 and 6A, 6B), which leads to the increased 402

PEPCK and G6Pase gene expression (Fig. 1A and 1B) and subsequent upregulation of G6P and 403

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 22

glucose production (Fig. 2). Moreover, our results indicate that HCV-induced ROS production 404

causes JNK activation, which results in the decreased phosphorylation and nuclear 405

accumulation of FoxO1, leading eventually to increased glucose production (Figs. 4 to 6). Our 406

results thus suggest that FoxO1 is a prime transcription factor in the HCV-mediated progression 407

of hepatic gluconeogenesis through an ROS/JNK-dependent mechanism, as summarized in the 408

schema in Fig. 10. Our results also suggest that HCV NS5A plays a role in enhanced hepatic 409

gluconeogenesis by promoting ROS production and JNK activation (Figs. 7 to 9). In line with 410

our observations, NS5A-mediated induction of ROS production (68) and JNK activation (49) 411

was reported by other investigators. 412

413

Increasing evidence suggests that mitochondrial dysfunction is causative of insulin resistance 414

and type 2 diabetes. Mitochondrial dysfunction causes upregulation of PEPCK and G6Pase, 415

leading to increased gluconeogenesis and insulin resistance (42, 46). We previously reported 416

that HCV causes mitochondrial damage and mitochondrion-mediated apoptosis (14, 47). Our 417

current data further support the concept that altered mitochondrial function plays a role in the 418

development of increased glucose production in hepatocytes. 419

420

We and other groups have reported that HCV infection increases the production of 421

mitochondrial ROS, which plays an important role in the development and progression of 422

inflammatory liver disease mediated by HCV (12, 14). Increased mitochondrial ROS generation 423

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 23

has also been shown to be an underlying mediator of multiple forms of insulin resistance 424

including inflammation- or glucocorticoid-induced insulin resistance (27, 29). Moreover, a 425

significant correlation was observed between oxidative stress and insulin resistance in patients 426

infected with HCV genotype 1 or 2 (44). ROS has also been shown to regulate the activity of 427

the FoxO transcription factor by post-translational modifications, including phosphorylation 428

(21), deacetylation (8) and ubiquitylation (67). 429

430

Although this study showed that JNK induces the nuclear accumulation of FoxO1 by reducing 431

its phosphorylation status under oxidative stress conditions in HCV-infected cells, the precise 432

mechanism(s) of interplay between JNK and FoxO1 still remains to be addressed. It has been 433

reported that activated JNK phosphorylates IRS-1 at Ser307, which results in attenuated insulin 434

signal transduction through inhibition of tyrosine phosphorylation of IRS-1 (1). Akt is a major 435

downstream signaling protein for the insulin/IRS-1 signaling and is activated through its 436

phosphorylation on Thr308 and Ser473, the latter of which is believed to be more crucial (53). 437

Therefore, an impairment of the insulin/IRS-1 signaling pathway should involve 438

downregulation of Akt phosphorylation. However, our present data showed that Akt 439

phosphorylation on Ser473 was upregulated in HCV-infected cells at 4 and 6 dpi (Fig. 3B), 440

suggesting that an Akt-independent pathway is involved in the JNK-mediated suppression of 441

FoxO1 phosphorylation. In this connection, it should be noted that the 14-3-3 protein, a binding 442

partner for phosphorylated FoxO1 that mediates its nuclear export (72), is phosphorylated by 443

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 24

JNK and that phosphorylated 14-3-3 protein releases its binding partners, which would 444

facilitate nuclear accumulation of FoxO (63, 65, 70). Further studies are needed to elucidate this 445

issue. 446

447

Another trigger that causes excessive JNK activation and insulin resistance is endoplasmic 448

reticulum (ER) stress (28, 48). Several studies have reported that HCV infection induces ER 449

stress (34, 55). Under our experimental condition, however, we did not detect significant ER 450

stress in HCV-infected cells (14). It is thus likely that ER stress was not the primary cause of 451

increased gluconeogenesis in our experimental system using Huh-7.5 cells and the P-47 strain 452

of HCV J6/JFH-1 (9, 14). 453

454

Notably, our present data showed that cells harboring SGR or FGR and HCV-infected cells 455

produced greater amounts of glucose than did the control cells (Fig. 2A), however, the changes 456

in the phosphorylation status of FoxO1 and JNK in SGR- and FGR-harboring cells were not so 457

significant compared to the virus-infected cells (data not shown). One of the reasons for this 458

difference is that SGR- and FGR- harboring cells were obtained through a longer cultivation in 459

a selection medium for a month or more and that the balance of host gene induction may be 460

somewhat different from that in virus-infected cells. Therefore, it is possible that, in addition to 461

the JNK-FoxO1 pathway, another signaling pathway(s) is involved in increased 462

gluconeogenesis in SGR- and FGR-harboring cells. Studies on this issue are now under way in 463

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 25

our laboratory. 464

465

We observed that HCV infection modulated, either positively or negatively, the transcription of 466

the PEPCK, G6Pase and GK genes at 3 to 5 dpi (Fig. 1). Virus infection, in general, causes 467

dynamically changing induction and suppression of a wide variety of host genes. For example, 468

expression levels of certain genes, such as the interferon genes, increase during an early phase 469

of virus infection, e.g., 1 dpi, but return to normal levels within a few days in cell culture 470

system. On the other hand, virus infection-induced expression of other genes, such as the ERK 471

gene, remains for a prolonged period of time (data not shown). Also, some of the gene products 472

induced in the acute phase may suppress expression of other genes. Under this balanced 473

condition, it is quite possible that certain genes are induced only at a later timing, e.g., 3 to 5 dpi, 474

but not immediately after virus infection. 475

476

It has been reported that HCV core protein-expressing transgenic mice exhibit marked insulin 477

resistance by inhibiting IRS-1 tyrosine phosphorylation and Akt phosphorylation (45, 58). 478

However, our present results showed that HCV NS5A, but not the core protein, was associated 479

with increased gluconeogenesis. Moreover, it was recently reported that HCV infection 480

significantly inhibited cellular glucose levels at 10 dpi (69), which is quite opposite to what we 481

observed in the present study. These results collectively suggest the possibility that multiple 482

pathways are involved in glucose metabolism in HCV-infected cells. Also, the possible effect(s) 483

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 26

of the dysregulation of hepatic gluconeogenesis on the HCV life cycle needs to be clarified. 484

485

In conclusion, our present results collectively suggest that HCV promotes hepatic 486

gluconeogenesis, resulting in increased glucose production in hepatocytes via an NS5A-487

mediated, FoxO1-dependent pathway. 488

489

ACKNOWLEDGMENTS 490

The authors are grateful to Dr. C. M. Rice (The Rockefeller University, New York, NY) for 491

providing Huh-7.5 cells and pFL-J6/JFH1, Dr. R. Bartenschlager (University of Heidelberg, 492

Heidelberg, Germany) for providing an HCV subgenomic RNA replicon (pFK5B/2884Gly), 493

and Dr. N. Kato (Okayama University, Okayama, Japan) for providing an HCV full-length RNA 494

replicon (pON/C-5B). Thanks are also due to Dr. T. Adachi (Kyoto Prefectural University of 495

Medicine, Kyoto, Japan), Ms. K. Igarashi, Ms. K. Kashikura, and Ms. A. Suzuki (Keio 496

University, Yamagata, Japan) for their technical assistance. This work was supported in part by 497

grants-in-aid for Research on Hepatitis from the Ministry of Health, Labour and Welfare, Japan, 498

and the Japan Initiative for Global Research Network on Infectious Diseases (J-GRID) program 499

of Ministry of Education, Culture, Sports, Science and Technology, Japan. This study was also 500

carried out as part of the Global Center of Excellence program of Kobe University Graduate 501

School of Medicine, and the Science and Technology Research Partnership for Sustainable 502

Development (SATREPS) program of Japan Science and Technology Agency (JST) and Japan 503

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

- - 27

International Cooperation Agency (JICA).504

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

28

FIGURE LEGENDS 505

506

FIG. 1. HCV upregulates gene expression of PEPCK and G6Pase and downregulates gene 507

expression of GK. Quantitative RT-PCR analysis was performed to quantify PEPCK (A), 508

G6Pase (B), and GK (C) mRNA expression in SGR- and FGR-harboring cells and HCV-509

infected cells (moi = 2), and the results were normalized to β-glucuronidase mRNA 510

expression levels. In parallel, SGR- and FGR-harboring cells and HCV-infected cells (at 5 511

dpi) were treated with IFN (1,000 IU/ml) for 10 days to eliminate HCV replication before 512

being subjected to quantitative RT-PCR. Data represent mean ± SEM of three independent 513

experiments, and the values for the control cells were arbitrarily expressed as 1.0. *, P<0.01 514

compared with the control. 515

516

FIG. 2. HCV promotes the production of glucose and G6P. (A) Extracellular glucose 517

production was measured in SGR- and FGR-harboring cells and HCV-infected cells (moi = 2), 518

and normalized to total cellular protein expression. In parallel, SGR- and FGR-harboring 519

cells and HCV-infected cells (at 5 dpi) were treated with IFN (1,000 IU/ml) for 10 days to 520

eliminate HCV replication before being subjected to glucose production analysis. Data 521

represent mean ± SEM of three independent experiments and the value for the control cells 522

was arbitrarily expressed as 1.0. *, P<0.01 compared with the control. (B) Cellular G6Pase 523

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

29

production was measured in HCV-infected cells (moi = 2), and the results were normalized to 524

cell number. Data represent mean ± SEM of three independent experiments. *, P<0.01 525

compared with the control. 526

527

FIG. 3. HCV suppresses FoxO1 phosphorylation, leading to nuclear accumulation of FoxO1 528

(A) Quantitative RT-PCR analysis was performed to determine FoxO1 mRNA expression in 529

SGR- and FGR-harboring cells and HCV-infected cells (moi = 2) and normalized to β-530

glucuronidase mRNA expression levels. (B) The expression levels of FoxO1, phospho-531

FoxO1 (Ser319), Akt, and phospho-Akt (Ser473) were analyzed by immunoblotting in HCV-532

infected cells and mock-infected control cells. Blots were reprobed with antibodies 533

recognizing NS3 and GAPDH. The amounts of GAPDH were measured as an internal control 534

to verify equal amounts of sample loading. (C) Cytoplasmic and nuclear fractions were 535

prepared from HCV-infected cells and the mock-infected control at 4 dpi, and were analyzed 536

by immunoblotting using antibodies against FoxO1, phospho-FoxO1 (Ser319), NS3, Hsp60, 537

and Oct-1. The amounts of Hsp60 and Oct-1 were measured to verify that they were equal to 538

the amounts of cytoplasmic and nuclear fractions, respectively. 539

540

FIG. 4. HCV-induced JNK activation is required for the suppression of FoxO1 541

phosphorylation. (A) HCV activates the JNK/c-Jun signaling pathway. Activation 542

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

30

(phosphorylation) of JNK (Thr183/Tyr185) and c-Jun (Ser63) in whole-cell lysates of HCV-543

infected cells and the mock-infected control was analyzed by immunoblotting. Blots were 544

reprobed with antibodies recognizing total JNK and c-Jun, NS3 and GAPDH. The amounts of 545

GAPDH were measured as an internal control to verify equal amounts of sample loading. (B) 546

Pretreatment with JNK inhibitor SP600125 abrogates HCV-induced c-Jun activation and 547

FoxO1 phosphorylation suppression. Phosphorylation of c-Jun (Ser63) and that of FoxO1 548

(Ser319) were analyzed by immunoblotting at 6 dpi in HCV-infected cells and the mock-549

infected control with or without SP600125 pretreatment (20 µM, 24 h). Blots were reprobed 550

with antibodies recognizing total c-Jun and FoxO1, NS3 and GAPDH. The amounts of 551

GAPDH were measured as an internal control to verify equal amounts of sample loading. 552

553

FIG. 5. HCV-induced production of mitochondrial ROS suppresses FoxO1 phosphorylation 554

through activation of JNK. (A) Pretreatment with NAC abrogates HCV-induced increased 555

production of mitochondrial ROS. HCV-infected cells and the mock-infected control were 556

pretreated with 5 mM NAC for 2 h at 6 dpi. The cell were then incubated with MitoSOX 557

(upper row) and then stained for HCV antigens by using an HCV-infected patient’s serum, 558

followed by FITC-conjugated goat anti-human IgG (bottom row). (B) Quantification of 559

MitoSOX-stained cells. The percentages of cells stained with MitoSOX were determined for 560

HCV-infected cells and the mock-infected control with or without NAC pretreatment. Data 561

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

31

represent mean ± SEM of two independent experiments. *, P<0.01. (C) NAC pretreatment 562

abrogates HCV-induced JNK activation and FoxO1 phosphorylation suppression. 563

Phosphorylation of JNK (Thr183/Tyr185) and that of FoxO1 (Ser319) were analyzed by 564

immunoblotting at 6 dpi in HCV-infected cells and the mock-infected control with or without 565

NAC pretreatment (5 mM, 2h). The blots were reprobed with antibodies recognizing total 566

JNK and FoxO1, NS3 and GAPDH. The amounts of GAPDH were measured as an internal 567

control to verify equal amounts of sample loading. 568

569

FIG. 6. HCV-induced JNK activation and ROS production are involved in FoxO1 nuclear 570

accumulation and increased glucose production. (A) Subcellular localization of FoxO1 in 571

HCV-infected cells and the mock-infected control with or without a JNK inhibitor 572

(SP600125; 20 µM, 24 h) or an antioxidant (NAC; 5 mM, 2 h) pretreatment at 5 dpi was 573

examined by confocal microscopy. After fixation and permeabilization, the cells were 574

incubated with anti-FoxO1 rabbit monoclonal antibody followed by Alexa Fluor 488-575

conjugated goat anti-rabbit IgG (upper row), and with an HCV-infected patient’s serum 576

followed by Alexa Fluor 594-conjugated goat anti-human IgG (bottom row). (B) The 577

percentages of cells with FoxO1 nuclear localization were determined for HCV-infected cells 578

and the mock-infected control with or without SP600125 or NAC pretreatment. Data 579

represent mean ± SEM of two independent experiments. *, P<0.01. (C) Extracellular glucose 580

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

32

production was measured in HCV-infected cells and the mock-infected control with or 581

without SP600125 or NAC pretreatment at 7 dpi, and normalized to total cellular protein 582

expression. Data represent mean ± SEM of two independent experiments and the value for 583

the control cells was arbitrarily expressed as 1.0. *, P<0.01. (D) Cellular expression levels of 584

NS3 in HCV-infected cells and the mock-infected control with or without sodium lactate (SL), 585

sodium pyruvate (SP), SP600125 or NAC are shown. The amounts of GAPDH were 586

measured as an internal control to verify equal amounts of sample loading. (E) Amounts of 587

HCV RNA were measured by quantitative RT-PCR analysis in HCV-infected cells treated 588

with SP600125 or NAC or left untreated at 6 dpi. The amounts were normalized to GAPDH 589

mRNA expression levels. Data represent mean ± SEM of two independent experiments, and 590

the value for the non-treated HCV-infected cells was arbitrarily expressed as 1.0. (F) Virus 591

infectivity in the culture supernatants of HCV-infected cells treated with SP600125 or NAC 592

or left untreated at 6 dpi was measured. Data represent mean ± SEM of two independent 593

experiments. CIU, cell-infecting units. 594

595

FIG. 7. HCV NS5A is involved in increased mRNA expression for G6Pase and PEPCK. 596

Huh7.5 cells were respectively transfected with the indicated HCV viral protein expression 597

plasmids. At 3 and 5 days post-transfection, quantitative RT-PCR analysis of mRNA for 598

G6Pase (A) and PEPCK (B) were conducted, and the results were normalized to β-599

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

33

glucuronidase mRNA expression levels. Data represent mean ± SEM of three independent 600

experiments, and the values for the control cells were arbitrarily expressed as 1.0. *, P<0.01 601

compared with the control. (C) At 3 days post-transfection, the expression levels of each of 602

the HCV proteins were examined by immunoblot analysis using antibodies against c-Myc, 603

core, NS4A and GAPDH. The amounts of GAPDH served as an internal control to verify 604

equal amounts of sample loading. (D) NS5A and NS4A enhance PEPCK promoter activity. 605

NS5A and NS4A expression plasmids were each co-transfected with rPEPCK-P5(-1263)-606

pGL3basic or rPEPCK-P4(-998)-pGL3basic in Huh7.5 cells. At 48 h after transfection, the 607

PEPCK promoter activities were measured using a luciferase reporter assay. Data represent 608

mean ± SEM of three independent experiments, and the values for the control cells were 609

arbitrarily expressed as 1.0. *, P<0.05 compared with the control. 610

611

FIG. 8. HCV NS5A is involved in increased ROS production, JNK activation, FoxO1 612

phosphorylation suppression and FoxO1 nuclear accumulation. (A) NS5A promotes ROS 613

production. Huh-7.5 cells transfected with NS5A expression plasmid or the empty control 614

(vector) were incubated with MitoSOX (upper row) at 3 days post-transfection, and then 615

stained for NS5A by using anti-NS5A mouse monoclonal antibody, followed by FITC-616

conjugated goat anti-mouse IgG (bottom row). (B) Quantification of MitoSOX-stained cells. 617

The percentages of cells stained with MitoSOX were determined for NS5A-expressing cells 618

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

34

and the control. Data represent mean ± SEM of two independent experiments. *, P<0.01. (C) 619

& (D) HCV NS5A activates c-Jun phosphorylation and suppresses FoxO1 phosphorylation. 620

Huh-7.5 cells transfected with the indicated HCV viral protein expression plasmids were 621

harvested at 3 days post-transfection, and the whole-cell lysates were subjected to 622

immunoblot analysis using antibodies against phospho-c-Jun (Ser63), c-Jun, phospho-FoxO1 623

(Ser319), FoxO1, GAPDH, and core, NS3, NS4A, NS5A (C) or and c-Myc (D). The amounts 624

of GAPDH were measured as an internal control to verify equal amounts of sample loading. 625

(E) NS5A facilitates FoxO1 nuclear accumulation. Huh-7.5 cells transfected with NS5A 626

expression plasmid or the empty control (vector) were fixed and permeabilized at 3 days 627

post-transfection. The cells were incubated with anti-FoxO1 rabbit monoclonal antibody 628

followed by Alexa Fluor 488-conjugated goat anti-rabbit IgG (upper row), or with anti-NS5A 629

mouse monoclonal antibody followed by Alexa Fluor 594-conjugated goat anti-mouse IgG 630

(bottom row). (F) The percentages of cells with nuclear localization of FoxO1 were 631

determined for NS5A-expressing cells and the control. Data represent mean ± SEM of two 632

independent experiments. *, P<0.01. 633

634

FIG. 9. HCV NS5A and NS4A enhance glucose production. (A) Huh7.5 cells were 635

transfected with either NS5A or NS4A expression plasmid. At 5 days post-transfection, 636

extracellular glucose production was measured and normalized to the total cellular protein 637

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

35

expression. Data represent mean ± SEM of two independent experiments, and the values for 638

the control cells were arbitrarily expressed as 1.0. *, P<0.05 compared with the control. (B) 639

Cellular expression levels of NS4A and NS5A in the absence and presence of sodium lactate 640

(SL) and sodium pyruvate (SP) are shown. 641

642

FIG. 10. Schematic presentation of the HCV-dysregulated hepatic gluconeogenesis signaling 643

pathway. HCV induces mitochondrial dysfunction (14). This results in increased ROS 644

production and JNK activation, which induces the nuclear accumulation of FoxO1 by 645

reducing its phosphorylation status. Consequently, PEPCK and G6Pase gene expressions are 646

upreglulated, leading to upregulation of G6P and glucose production. NS5A plays a role in 647

HCV-induced gluconeogenesis via induction of ROS production. IRE, insulin response 648

element.649

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

36

650

REFERENCES 651

652

1. Aguirre, V., T. Uchida, L. Yenush, R. Davis, and M. F. White. 2000. The c-Jun 653

NH(2)-terminal kinase promotes insulin resistance during association with insulin 654

receptor substrate-1 and phosphorylation of Ser(307). J. Biol. Chem. 275:9047-9054. 655

2. Alaei, M., and F. Negro. 2008. Hepatitis C virus and glucose and lipid metabolism. 656

Diabetes Metab. 34:692-700. 657

3. Aytug, S., D. Reich, L. E. Sapiro, D. Bernstein, and N. Begum. 2003. Impaired 658

IRS-1/PI3-kinase signaling in patients with HCV: a mechanism for increased 659

prevalence of type 2 diabetes. Hepatology 38:1384-1392. 660

4. Banerjee, A., K. Meyer, B. Mazumdar, R. B. Ray, and R. Ray. 2010. Hepatitis C 661

virus differentially modulates activation of forkhead transcription factors and insulin-662

induced metabolic gene expression. J. Virol. 84:5936-5946. 663

5. Baron, A. D., L. Schaeffer, P. Shragg, and O. G. Kolterman. 1987. Role of 664

hyperglucagonemia in maintenance of increased rates of hepatic glucose output in 665

type II diabetics. Diabetes 36:274-283. 666

6. Bennett, B. L., D. T. Sasaki, B. W. Murray, E. C. O'Leary, S. T. Sakata, W. Xu, J. 667

C. Leisten, A. Motiwala, S. Pierce, Y. Satoh, S. S. Bhagwat, A. M. Manning, and 668

D. W. Anderson. 2001. SP600125, an anthrapyrazolone inhibitor of Jun N-terminal 669

kinase. Proc. Natl. Acad. Sci. U. S. A. 98:13681-13686. 670

7. Blight, K. J., J. A. McKeating, and C. M. Rice. 2002. Highly permissive cell lines 671

for subgenomic and genomic hepatitis C virus RNA replication. J. Virol. 76:13001-672

13014. 673

8. Brunet, A., L. B. Sweeney, J. F. Sturgill, K. F. Chua, P. L. Greer, Y. Lin, H. Tran, 674

S. E. Ross, R. Mostoslavsky, H. Y. Cohen, L. S. Hu, H. L. Cheng, M. P. 675

Jedrychowski, S. P. Gygi, D. A. Sinclair, F. W. Alt, and M. E. Greenberg. 2004. 676

Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. 677

Science 303:2011-2015. 678

9. Bungyoku, Y., I. Shoji, T. Makine, T. Adachi, K. Hayashida, M. Nagano-Fujii, Y. 679

H. Ide, L. Deng, and H. Hotta. 2009. Efficient production of infectious hepatitis C 680

virus with adaptive mutations in cultured hepatoma cells. J. Gen. Virol. 90:1681-1691. 681

10. Burdette, D., M. Olivarez, and G. Waris. 2010. Activation of transcription factor 682

Nrf2 by hepatitis C virus induces the cell-survival pathway. J. Gen. Virol. 91:681-690. 683

11. Caronia, S., K. Taylor, L. Pagliaro, C. Carr, U. Palazzo, J. Petrik, S. O'Rahilly, S. 684

Shore, B. D. Tom, and G. J. Alexander. 1999. Further evidence for an association 685

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

37

between non-insulin-dependent diabetes mellitus and chronic hepatitis C virus 686

infection. Hepatology 30:1059-1063. 687

12. Choi, J., and J. H. Ou. 2006. Mechanisms of liver injury. III. Oxidative stress in the 688

pathogenesis of hepatitis C virus. Am. J. Physiol. Gastrointest. Liver Physiol. 689

290:G847-851. 690

13. Clore, J. N., J. Stillman, and H. Sugerman. 2000. Glucose-6-phosphatase flux in 691

vitro is increased in type 2 diabetes. Diabetes 49:969-974. 692

14. Deng, L., T. Adachi, K. Kitayama, Y. Bungyoku, S. Kitazawa, S. Ishido, I. Shoji, 693

and H. Hotta. 2008. Hepatitis C virus infection induces apoptosis through a Bax-694

triggered, mitochondrion-mediated, caspase 3-dependent pathway. J. Virol. 82:10375-695

10385. 696

15. Deng, L., M. Nagano-Fujii, M. Tanaka, Y. Nomura-Takigawa, M. Ikeda, N. Kato, 697

K. Sada, and H. Hotta. 2006. NS3 protein of Hepatitis C virus associates with the 698

tumour suppressor p53 and inhibits its function in an NS3 sequence-dependent 699

manner. J. Gen. Virol. 87:1703-1713. 700

16. Diamond, D. L., A. J. Syder, J. M. Jacobs, C. M. Sorensen, K. A. Walters, S. C. 701

Proll, J. E. McDermott, M. A. Gritsenko, Q. Zhang, R. Zhao, T. O. Metz, D. G. 702

Camp, 2nd, K. M. Waters, R. D. Smith, C. M. Rice, and M. G. Katze. 2010. 703

Temporal proteome and lipidome profiles reveal hepatitis C virus-associated 704

reprogramming of hepatocellular metabolism and bioenergetics. PLoS Pathog. 705

6:e1000719. 706

17. Dionisio, N., M. V. Garcia-Mediavilla, S. Sanchez-Campos, P. L. Majano, I. 707

Benedicto, J. A. Rosado, G. M. Salido, and J. Gonzalez-Gallego. 2009. Hepatitis C 708

virus NS5A and core proteins induce oxidative stress-mediated calcium signalling 709

alterations in hepatocytes. J. Hepatol. 50:872-882. 710

18. Doi, H., C. Apichartpiyakul, K. I. Ohba, M. Mizokami, and H. Hotta. 1996. 711

Hepatitis C virus (HCV) subtype prevalence in Chiang Mai, Thailand, and 712

identification of novel subtypes of HCV major type 6. J. Clin. Microbiol. 34:569-574. 713

19. Dunning, B. E., and J. E. Gerich. 2007. The role of alpha-cell dysregulation in 714

fasting and postprandial hyperglycemia in type 2 diabetes and therapeutic 715

implications. Endocr. Rev. 28:253-283. 716

20. Eslam, M., M. A. Khattab, and S. A. Harrison. 2011. Insulin resistance and 717

hepatitis C: an evolving story. Gut. 718

21. Essers, M. A., S. Weijzen, A. M. de Vries-Smits, I. Saarloos, N. D. de Ruiter, J. L. 719

Bos, and B. M. Burgering. 2004. FOXO transcription factor activation by oxidative 720

stress mediated by the small GTPase Ral and JNK. EMBO J. 23:4802-4812. 721

22. Franck, N., J. Le Seyec, C. Guguen-Guillouzo, and L. Erdtmann. 2005. Hepatitis 722

C virus NS2 protein is phosphorylated by the protein kinase CK2 and targeted for 723

degradation to the proteasome. J. Virol. 79:2700-2708. 724

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

38

23. Galossi, A., R. Guarisco, L. Bellis, and C. Puoti. 2007. Extrahepatic manifestations 725

of chronic HCV infection. J. Gastrointestin. Liver Dis. 16:65-73. 726

24. Gottwein, J. M., T. K. Scheel, T. B. Jensen, J. B. Lademann, J. C. Prentoe, M. L. 727

Knudsen, A. M. Hoegh, and J. Bukh. 2009. Development and characterization of 728

hepatitis C virus genotype 1-7 cell culture systems: role of CD81 and scavenger 729

receptor class B type I and effect of antiviral drugs. Hepatology 49:364-377. 730

25. Gross, D. N., A. P. van den Heuvel, and M. J. Birnbaum. 2008. The role of FoxO 731

in the regulation of metabolism. Oncogene 27:2320-2336. 732

26. Hirota, K., J. Sakamaki, J. Ishida, Y. Shimamoto, S. Nishihara, N. Kodama, K. 733

Ohta, M. Yamamoto, K. Tanimoto, and A. Fukamizu. 2008. A combination of 734

HNF-4 and Foxo1 is required for reciprocal transcriptional regulation of glucokinase 735

and glucose-6-phosphatase genes in response to fasting and feeding. J. Biol. Chem. 736

283:32432-32441. 737

27. Hoehn, K. L., A. B. Salmon, C. Hohnen-Behrens, N. Turner, A. J. Hoy, G. J. 738

Maghzal, R. Stocker, H. Van Remmen, E. W. Kraegen, G. J. Cooney, A. R. 739

Richardson, and D. E. James. 2009. Insulin resistance is a cellular antioxidant 740

defense mechanism. Proc. Natl. Acad. Sci. U. S. A. 106:17787-17792. 741

28. Hotamisligil, G. S. 2005. Role of endoplasmic reticulum stress and c-Jun NH2-742

terminal kinase pathways in inflammation and origin of obesity and diabetes. Diabetes 743

54 Suppl 2:S73-78. 744

29. Houstis, N., E. D. Rosen, and E. S. Lander. 2006. Reactive oxygen species have a 745

causal role in multiple forms of insulin resistance. Nature 440:944-948. 746

30. Huang, H., and D. J. Tindall. 2007. Dynamic FoxO transcription factors. J. Cell Sci. 747

120:2479-2487. 748

31. Ikeda, M., K. Abe, H. Dansako, T. Nakamura, K. Naka, and N. Kato. 2005. 749

Efficient replication of a full-length hepatitis C virus genome, strain O, in cell culture, 750

and development of a luciferase reporter system. Biochem. Biophys. Res. Commun. 751

329:1350-1359. 752

32. Inubushi, S., M. Nagano-Fujii, K. Kitayama, M. Tanaka, C. An, H. Yokozaki, H. 753

Yamamura, H. Nuriya, M. Kohara, K. Sada, and H. Hotta. 2008. Hepatitis C virus 754

NS5A protein interacts with and negatively regulates the non-receptor protein tyrosine 755

kinase Syk. J. Gen. Virol. 89:1231-1242. 756

33. Iynedjian, P. B., P. R. Pilot, T. Nouspikel, J. L. Milburn, C. Quaade, S. Hughes, C. 757

Ucla, and C. B. Newgard. 1989. Differential expression and regulation of the 758

glucokinase gene in liver and islets of Langerhans. Proc. Natl. Acad. Sci. U. S. A. 759

86:7838-7842. 760

34. Joyce, M. A., K. A. Walters, S. E. Lamb, M. M. Yeh, L. F. Zhu, N. Kneteman, J. S. 761

Doyle, M. G. Katze, and D. L. Tyrrell. 2009. HCV induces oxidative and ER stress, 762

and sensitizes infected cells to apoptosis in SCID/Alb-uPA mice. PLoS Pathog. 763

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

39

5:e1000291. 764

35. Kamata, H., S. Honda, S. Maeda, L. Chang, H. Hirata, and M. Karin. 2005. 765

Reactive oxygen species promote TNFalpha-induced death and sustained JNK 766

activation by inhibiting MAP kinase phosphatases. Cell 120:649-661. 767

36. Karpac, J., and H. Jasper. 2009. Insulin and JNK: optimizing metabolic homeostasis 768

and lifespan. Trends Endocrinol. Metab. 20:100-106. 769

37. Kasai, D., T. Adachi, L. Deng, M. Nagano-Fujii, K. Sada, M. Ikeda, N. Kato, Y. H. 770

Ide, I. Shoji, and H. Hotta. 2009. HCV replication suppresses cellular glucose 771

uptake through down-regulation of cell surface expression of glucose transporters. J. 772

Hepatol. 50:883-894. 773

38. Kops, G. J., and B. M. Burgering. 1999. Forkhead transcription factors: new insights 774

into protein kinase B (c-akt) signaling. J. Mol. Med. 77:656-665. 775

39. Lindenbach, B. D., M. J. Evans, A. J. Syder, B. Wolk, T. L. Tellinghuisen, C. C. 776

Liu, T. Maruyama, R. O. Hynes, D. R. Burton, J. A. McKeating, and C. M. Rice. 777

2005. Complete replication of hepatitis C virus in cell culture. Science 309:623-626. 778

40. Lindenbach, B. D., and C. M. Rice. 2005. Unravelling hepatitis C virus replication 779

from genome to function. Nature 436:933-938. 780

41. Lohmann, V., F. Korner, A. Dobierzewska, and R. Bartenschlager. 2001. 781

Mutations in hepatitis C virus RNAs conferring cell culture adaptation. J. Virol. 782

75:1437-1449. 783

42. Lowell, B. B., and G. I. Shulman. 2005. Mitochondrial dysfunction and type 2 784

diabetes. Science 307:384-387. 785

43. Mehta, S. H., F. L. Brancati, M. S. Sulkowski, S. A. Strathdee, M. Szklo, and D. L. 786

Thomas. 2000. Prevalence of type 2 diabetes mellitus among persons with hepatitis C 787

virus infection in the United States. Ann. Intern. Med. 133:592-599. 788

44. Mitsuyoshi, H., Y. Itoh, Y. Sumida, M. Minami, K. Yasui, T. Nakashima, and T. 789

Okanoue. 2008. Evidence of oxidative stress as a cofactor in the development of 790

insulin resistance in patients with chronic hepatitis C. Hepatol. Res. 38:348-353. 791

45. Miyamoto, H., K. Moriishi, K. Moriya, S. Murata, K. Tanaka, T. Suzuki, T. 792

Miyamura, K. Koike, and Y. Matsuura. 2007. Involvement of the PA28gamma-793

dependent pathway in insulin resistance induced by hepatitis C virus core protein. J. 794

Virol. 81:1727-1735. 795

46. Morino, K., K. F. Petersen, and G. I. Shulman. 2006. Molecular mechanisms of 796

insulin resistance in humans and their potential links with mitochondrial dysfunction. 797

Diabetes 55 Suppl 2:S9-S15. 798

47. Nomura-Takigawa, Y., M. Nagano-Fujii, L. Deng, S. Kitazawa, S. Ishido, K. Sada, 799

and H. Hotta. 2006. Non-structural protein 4A of Hepatitis C virus accumulates on 800

mitochondria and renders the cells prone to undergoing mitochondria-mediated 801

apoptosis. J. Gen. Virol. 87:1935-1945. 802

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

40

48. Ozcan, U., Q. Cao, E. Yilmaz, A. H. Lee, N. N. Iwakoshi, E. Ozdelen, G. Tuncman, 803

C. Gorgun, L. H. Glimcher, and G. S. Hotamisligil. 2004. Endoplasmic reticulum 804

stress links obesity, insulin action, and type 2 diabetes. Science 306:457-461. 805

49. Park, K. J., S. H. Choi, D. H. Choi, J. M. Park, S. W. Yie, S. Y. Lee, and S. B. 806

Hwang. 2003. 1Hepatitis C virus NS5A protein modulates c-Jun N-terminal kinase 807

through interaction with tumor necrosis factor receptor-associated factor 2. J. Biol. 808

Chem. 278:30711-30718. 809

50. Puigserver, P., J. Rhee, J. Donovan, C. J. Walkey, J. C. Yoon, F. Oriente, Y. 810

Kitamura, J. Altomonte, H. Dong, D. Accili, and B. M. Spiegelman. 2003. Insulin-811

regulated hepatic gluconeogenesis through FOXO1-PGC-1alpha interaction. Nature 812

423:550-555. 813

51. Reed, K. E., and C. M. Rice. 2000. Overview of hepatitis C virus genome structure, 814

polyprotein processing, and protein properties. Curr. Top Microbiol. Immunol. 815

242:55-84. 816

52. Rozance, P. J., S. W. Limesand, J. S. Barry, L. D. Brown, S. R. Thorn, D. 817

LoTurco, T. R. Regnault, J. E. Friedman, and W. W. Hay, Jr. 2008. Chronic late-818

gestation hypoglycemia upregulates hepatic PEPCK associated with increased 819

PGC1alpha mRNA and phosphorylated CREB in fetal sheep. Am. J. Physiol. 820

Endocrinol. Metab. 294:E365-370. 821

53. Sale, E. M., and G. J. Sale. 2008. Protein kinase B: signalling roles and therapeutic 822

targeting. Cell Mol. Life Sci. 65:113-127. 823

54. Schmoll, D., K. S. Walker, D. R. Alessi, R. Grempler, A. Burchell, S. Guo, R. 824

Walther, and T. G. Unterman. 2000. Regulation of glucose-6-phosphatase gene 825

expression by protein kinase Balpha and the forkhead transcription factor FKHR. 826

Evidence for insulin response unit-dependent and -independent effects of insulin on 827

promoter activity. J. Biol. Chem. 275:36324-36333. 828

55. Sekine-Osajima, Y., N. Sakamoto, K. Mishima, M. Nakagawa, Y. Itsui, M. Tasaka, 829

Y. Nishimura-Sakurai, C. H. Chen, T. Kanai, K. Tsuchiya, T. Wakita, N. 830

Enomoto, and M. Watanabe. 2008. Development of plaque assays for hepatitis C 831

virus-JFH1 strain and isolation of mutants with enhanced cytopathogenicity and 832

replication capacity. Virology 371:71-85. 833

56. Seo, H. Y., M. K. Kim, A. K. Min, H. S. Kim, S. Y. Ryu, N. K. Kim, K. M. Lee, H. 834

J. Kim, H. S. Choi, K. U. Lee, K. G. Park, and I. K. Lee. 2010. Endoplasmic 835

reticulum stress-induced activation of activating transcription factor 6 decreases 836

cAMP-stimulated hepatic gluconeogenesis via inhibition of CREB. Endocrinology 837

151:561-568. 838

57. Shepard, C. W., L. Finelli, and M. J. Alter. 2005. Global epidemiology of hepatitis 839

C virus infection. Lancet Infect. Dis. 5:558-567. 840

58. Shintani, Y., H. Fujie, H. Miyoshi, T. Tsutsumi, K. Tsukamoto, S. Kimura, K. 841

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

41

Moriya, and K. Koike. 2004. Hepatitis C virus infection and diabetes: direct 842

involvement of the virus in the development of insulin resistance. Gastroenterology 843

126:840-848. 844

59. Simmonds, P., J. Bukh, C. Combet, G. Deleage, N. Enomoto, S. Feinstone, P. 845

Halfon, G. Inchauspe, C. Kuiken, G. Maertens, M. Mizokami, D. G. Murphy, H. 846

Okamoto, J. M. Pawlotsky, F. Penin, E. Sablon, I. T. Shin, L. J. Stuyver, H. J. 847

Thiel, S. Viazov, A. J. Weiner, and A. Widell. 2005. Consensus proposals for a 848

unified system of nomenclature of hepatitis C virus genotypes. Hepatology 42:962-849

973. 850

60. Soga, T., R. Baran, M. Suematsu, Y. Ueno, S. Ikeda, T. Sakurakawa, Y. Kakazu, 851

T. Ishikawa, M. Robert, T. Nishioka, and M. Tomita. 2006. Differential 852

metabolomics reveals ophthalmic acid as an oxidative stress biomarker indicating 853

hepatic glutathione consumption. J. Biol. Chem. 281:16768-16776. 854

61. Soga, T., K. Igarashi, C. Ito, K. Mizobuchi, H. P. Zimmermann, and M. Tomita. 855

2009. Metabolomic profiling of anionic metabolites by capillary electrophoresis mass 856

spectrometry. Anal. Chem. 81:6165-6174. 857

62. Streeper, R. S., C. A. Svitek, S. Chapman, L. E. Greenbaum, R. Taub, and R. M. 858

O'Brien. 1997. A multicomponent insulin response sequence mediates a strong 859

repression of mouse glucose-6-phosphatase gene transcription by insulin. J. Biol. 860

Chem. 272:11698-11701. 861

63. Sunayama, J., F. Tsuruta, N. Masuyama, and Y. Gotoh. 2005. JNK antagonizes 862

Akt-mediated survival signals by phosphorylating 14-3-3. J. Cell Biol. 170:295-304. 863

64. Takashima, M., W. Ogawa, K. Hayashi, H. Inoue, S. Kinoshita, Y. Okamoto, H. 864

Sakaue, Y. Wataoka, A. Emi, Y. Senga, Y. Matsuki, E. Watanabe, R. Hiramatsu, 865

and M. Kasuga. 2010. Role of KLF15 in regulation of hepatic gluconeogenesis and 866

metformin action. Diabetes 59:1608-1615. 867

65. Tsuruta, F., J. Sunayama, Y. Mori, S. Hattori, S. Shimizu, Y. Tsujimoto, K. 868

Yoshioka, N. Masuyama, and Y. Gotoh. 2004. JNK promotes Bax translocation to 869

mitochondria through phosphorylation of 14-3-3 proteins. EMBO J. 23:1889-1899. 870

66. van der Horst, A., and B. M. Burgering. 2007. Stressing the role of FoxO proteins 871

in lifespan and disease. Nat. Rev. Mol. Cell Biol. 8:440-450. 872

67. van der Horst, A., A. M. de Vries-Smits, A. B. Brenkman, M. H. van Triest, N. 873

van den Broek, F. Colland, M. M. Maurice, and B. M. Burgering. 2006. FOXO4 874

transcriptional activity is regulated by monoubiquitination and USP7/HAUSP. Nat. 875

Cell Biol. 8:1064-1073. 876

68. Wang, A. G., D. S. Lee, H. B. Moon, J. M. Kim, K. H. Cho, S. H. Choi, H. L. Ha, 877

Y. H. Han, D. G. Kim, S. B. Hwang, and D. Y. Yu. 2009. Non-structural 5A protein 878

of hepatitis C virus induces a range of liver pathology in transgenic mice. J. Pathol. 879

219:253-262. 880

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

42

69. Woodhouse, S. D., R. Narayan, S. Latham, S. Lee, R. Antrobus, B. Gangadharan, 881

S. Luo, G. P. Schroth, P. Klenerman, and N. Zitzmann. 2010. Transcriptome 882

sequencing, microarray, and proteomic analyses reveal cellular and metabolic impact 883

of hepatitis C virus infection in vitro. Hepatology 52:443-453. 884

70. Yoshida, K., T. Yamaguchi, T. Natsume, D. Kufe, and Y. Miki. 2005. JNK 885

phosphorylation of 14-3-3 proteins regulates nuclear targeting of c-Abl in the 886

apoptotic response to DNA damage. Nat. Cell Biol. 7:278-285. 887

71. Zhang, S., J. Liu, G. MacGibbon, M. Dragunow, and G. J. Cooper. 2002. 888

Increased expression and activation of c-Jun contributes to human amylin-induced 889

apoptosis in pancreatic islet beta-cells. J. Mol. Biol. 324:271-285. 890

72. Zhao, X., L. Gan, H. Pan, D. Kan, M. Majeski, S. A. Adam, and T. G. Unterman. 891

2004. Multiple elements regulate nuclear/cytoplasmic shuttling of FOXO1: 892

characterization of phosphorylation- and 14-3-3-dependent and -independent 893

mechanisms. Biochem. J. 378:839-849. 894

895

896

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from

on May 20, 2018 by guest

http://jvi.asm.org/

Dow

nloaded from