eISSN: 2449-8238
ISSN: 2392-1099
Clinical and Experimental Hepatology
Current issue Archive Manuscripts accepted About the journal Editorial board Abstracting and indexing Subscription Contact Instructions for authors Ethical standards and procedures
Editorial System
Submit your Manuscript
SCImago Journal & Country Rank
1/2015
vol. 1
 
Share:
Share:

Review article
Insulin resistance and its consequences in chronic hepatitis C

Michał Kukla
,
Damian Piotrowski
,
Marek Waluga
,
Marek Hartleb

Clinical and Experimental Hepatology 2015; 1: 17–29
Online publish date: 2015/04/30
Article file
- Insulin resistance.pdf  [0.15 MB]
Get citation
 
PlumX metrics:
 

Introduction

Chronic hepatitis C (CHC) and obesity are two major rising epidemics which present tough challenges both for clinicians and healthcare systems in terms of diagnostic, therapeutic and economic implications [1]. Obesity, especially central or visceral obesity, expressed as increased waist circumference, is definitely the feature most commonly associated with insulin resistance (IR) and metabolic syndrome (MS). Recent data seem to underestimate a prominent role for visceral and hepatic fat [nonalcoholic fatty liver disease (NAFLD) or nonalcoholic steatohepatitis (NASH)] as a pathogenic driver in the etiology of IR. However, when exploring the relationship between MS, central obesity and CHC, the relative contribution of metabolic IR and NAFLD on one hand, and hepatitis C virus (HCV)-induced IR and hepatic steatosis on the other hand, should be appreciated. HCV can directly or indirectly cause both IR and steatosis, but it is still not resolved whether this viral impact bears the same prognostic value as the metabolic counterparts [2]. As the population exposed to hepatitis C ages, the morbidity due to this disease is increasing. The rising epidemic of obesity contributes to higher prevalence of IR and type 2 diabetes mellitus (T2DM). Our understanding of the mutual association between both disease states continues to grow, but is still far from complete. This review briefly discusses the most probable mechanisms involved in IR development in the course of CHC. Molecular mechanisms for the direct and indirect HCV influence on intracellular insulin signaling are described. Subsequently, the consequences of IR/T2DM for disease progression are summarized [3].
IR is typically manifested as both decreased insulinmediated glucose uptake at the level of adipose tissue and skeletal muscle tissue (peripheral IR), and as insufficient inhibition of hepatic glucose release (hepatic IR). Both types of IR may occur simultaneously, but their intensity can be different [4]. IR is defined as a condition when normal insulin levels are not sufficient to achieve a normal metabolic response, or a condition when higher-than-normal insulin concentrations are necessary to achieve a suitable metabolic response. This definition does not allow one to indicate the type of tissue where insulin activity is measured. IR involves multiple sites: (i) the muscle, where it reduces glucose uptake and utilization, (ii) adipose tissue, where lipolysis is not adequately suppressed by insulin, with subsequent release of glycerol and free fatty acids into the bloodstream, and (iii) the liver, where IR is characterized by the overproduction of glucose despite fasting hyperinsulinaemia. Irrespective of which might have been the primary site of IR (muscle, fat or liver), all these metabolic abnormalities contribute finally to the onset of overt diabetes [5].
Evidence that hepatitis C virus causes insulin resistance
CHC should be considered not only as viral but also as a special type of metabolic disease. CHC is associated with lipid metabolism abnormalities leading to hepatic steatosis, impaired glucose metabolism leading to IR and T2DM and is related to an increased risk of carotid atherosclerosis [6]. Several experimental, epidemiological and clinical studies have confirmed the unquestioned participation of HCV in glucose metabolism perturbation [7]. The relationship between CHC and IR was first described by Allison et al. [8], who revealed that the prevalence of T2DM in cirrhotic CHC patients was significantly higher than in those with cirrhosis resulting from conditions other than CHC. Cross-sectional studies comparing the prevalence of T2DM in CHC patients with those with chronic liver disease of other origin (including chronic hepatitis B [CHB]) or with human immunodeficiency virus infection, indicated higher prevalence of T2DM in CHC [7]. Conversely, the prevalence of HCV infection in diabetic patients is markedly higher than in the general population, ranging from 5% up to 12% (34-45). An analysis which included 9841 subjects, aged 20 years or more, revealed three-fold higher risk of T2DM development in CHC patients [9]. These observations have been confirmed both in general population-based [10] and longitudinal studies [11]. HCV potentiates the risk of developing abnormalities in glucose metabolism in patients with other risk factors [11], suggesting that HCV infection accelerates T2DM onset in susceptible patients. The patients with high risk for T2DM and CHC were 11 times more likely to develop T2DM than those without HCV infection [11]. HCV increases the incidence of T2DM after liver [12] and kidney transplantation, impacting on liver fibrosis progression [13] and cardiovascular events [14]. In a recent meta-analysis of 10 studies, the pooled risk ratio for post-kidney transplantation T2DM reached 2 [15].
A relationship between CHC and glucose abnormalities at the early prediabetes stages such as impaired glucose tolerance (IGT) or IR was also found (Fig. 1) [16-18]. Hui et al. measured fasting serum insulin, C-peptide and HOMA-IR levels in 121 CHC patients and 137 healthy volunteers. All three variables were significantly up-regulated in CHC patients [17]. This finding was confirmed in a study which included 600 patients (500 with CHC and 100 with CHB) where the prevalence of T2DM was 7.6%. IR was observed in 32.4% out of 462 CHC patients without diabetes. The study comparing CHB and CHC confirmed significantly more frequent IR prevalence in the latter (35% vs. 5%) [19]. IR is observed in CHC patients in very early stages of the infection before there is a marked extent of fibrosis [17-19]. Downregulation of insulin sensitivity rather than impaired pancreatic islet cell function is suggested to be an essential pathophysiological factor contributing to the appearance of abnormal glucose homeostasis in CHC [17]. HCV appears to induce IR via mechanisms operating in the liver and at the periphery. In CHC patients without any features of MS and without advanced fibrosis, endogenous glucose production was high-normal in the basal state [20]. During a hyperinsulinemic clamp, the production of endogenous glucose was 3.5 times higher in comparison with the control group, while muscle glucose uptake was unsettled, ranging from the lower end of normal to severely impaired. The action of insulin for the process of lipolysis was not impaired [20]. However, another study found peripheral IR to be exacerbated but adipose and hepatic IR to be unchanged in CHC patients [21]. Muscle IR was positively related to viral load and subcutaneous fat and was again independent from viral genotype and hepatic steatosis grade [21]. The association between viral load and IR was also observed in other studies. The study by Moucari et al. showed IR to be related to genotype 1 or 4, but again it was associated with high serum viral load [19]. However, the study by Vanni et al. showed both peripheral and hepatic IR to be independent of viral genotype and hepatic steatosis [20]. The discrepancy in the obtained results may result from the fact that blood viral load is not stable but fluctuates during the course of the disease.
The association between T2DM and HCV genotypes also is still controversial [22-24]. The most recent, large-scale study addressing the relationship between IR and viral clearance in genotype-1, -2 and -3 patients revealed that IR was more common in patients with genotype 1 in comparison with those infected with genotype 2/3 [25].
The fact of high importance is that viral clearance after successful antiviral therapy leads to improvement of insulin sensitivity and reduces occurrence in glucose metabolism abnormalities across follow-up [26-28]. These findings represent irrefutable evidence for a cau­sal relationship between HCV and impaired glucose metabolism. The study carried out on 89 Japanese CHC patients showed that efficient antiviral therapy improved HOMA-IR results and intrahepatic expression of IRS-1 and IRS-2 [28]. The results were also confirmed in a group of 181 CHC patients infected with genotype 4 [29]. Romero-Gómez et al. assessed the influence of SVR on the incidence of IGT and T2DM in 1059 CHC patients treated with pegylated IFN-α2b and RBV [26]. SVR reduced the incidence of T2DM or IGT by half during a post-treatment follow-up of 27 months (range 9.3-67 months). Also the retrospective cohort study, which encompassed 2842 CHC patients, carried out by Arase et al. showed a two-thirds reduction in the risk of T2DM development after a mean follow-up of 6.4 years [27]. However, the study by Giordanino et al. in a group of 202 CHC patients with median follow-up of 8 years (range 5-16 years) showed contradictory results and did not reveal any improvement in insulin sensitivity in CHC patients with SVR [30]. Additionally, the above-mentioned study by Thompson et al. showed that viral clearance resulted in IR down-regulation in patients infected with genotype 1, but not in those with genotype 2/3 [25]. These results suggest a causal relationship between IR and HCV genotype 1.
To summarize, HCV evokes IR in the early stage of infection and therefore increases the risk of onset of T2DM in predisposed individuals. Some studies have indicated that IR is associated with viral load and observed more often in genotype 1 or 4 infection. Moreover, successful antiviral therapy ameliorates insulin sensitivity. All these results point to a direct viral influence on IR independent of BMI and visceral adi­posity and suggest that HCV itself may promote and exacerbate IR [3]. The assessment of IR allows one to evaluate the impact of HCV on glucose metabolism before overt T2DM occurs [17]. Nevertheless, the fine details of these mechanisms are not fully resolved and are sometimes speculative.

Molecular mechanisms of insulin resistance

Insulin is the most potent anabolic agent, which enhances the storage and synthesis of lipids, carbohydrates and proteins, and blocks their breakdown and release into the blood stream [31]. The biological action of insulin depends on a cascade of events following the interaction of insulin with its receptor on the cell membrane. The insulin receptor is a heterodimeric complex consisting of two extracellular α-subunits that bind insulin and two transmembrane β-subunits with tyrosine kinase activity. Insulin binding promotes the receptor autophosphorylation and the subsequent tyrosine phosphorylation of several insulin receptor substrates (IRS) (namely IRS-1 and IRS-2), which initiate a cascade of multifaceted events. Key transducers of insulin-mediated glucose regulation are phosphatidylinositol 3-kinase (PI3K) and the protein kinase Akt [32]. These actions are manifested via insulin’s action on a complex network of intracellular pathways in adipocytes, muscles and hepatocytes. The final result of this activation is translocation of glucose transporter 4 (GLUT-4) from the intracellular pool to the cell membrane, facilitating glucose transport into the cytoplasm [3]. In IR, there is impairment of insulin receptor binding and phosphorylation of IRS-1 and -2 in the muscle and the liver, and a dramatic decrease in PI3K activity and glucose uptake [32,33]. The most likely mechanism of IR within the muscle is serine rather than tyrosine phosphorylation of IRS-1, and similar events occurring in hepatocytes are likely to mediate IR within the liver. Some additional factors influence insulin signaling, and their activation may lead to IR. Protein tyrosine phosphatases (PTP) (especially PTP1B) [34] may dephosphorylate tyrosine residues on IRS, and phosphatidylinositol phosphate (PIP) phospha­ta­ses may dephosphorylate PIPs at position 5’ (SH2-con­taining PIP 5’-phosphatase 2, or SHIP2) or 3’ (phosphatase and tensin homolog, or PTEN) [35], attenuating insulin signaling. The suppressors of cytokine signaling (SOCS) potentiate ubiquitin-mediated IRS degradation [36]. IRS may also be inactivated due to phosphorylation at serine residues by kinases, such as mTOR substrate p70 ribosomal S6 kinase (p70S6K), protein kinase C and c-Jun N-terminal kinase (JNK) [37]. Additionally, the above-mentioned mechanisms interrupting insulin signaling may be provoked by some acquired factors encompassing hyperinsulinemia, hyperglycemia, circu­lating free fatty acids, high levels of tumor necrosis factor-α (TNF-α), IL-6, nuclear factor kappa B (NF-κB) and adipokines [38,39].

Molecular mechanisms of HCV-induced insulin resistance

IR development in CHC results from the interaction between various host, viral and environmental factors, contributing to enhanced endogenous glucose production. The first study, carried out on fresh liver samples obtained from 42 nonobese, nondiabetic CHC patients and 10 uninfected controls, matched for age and BMI incubated ex vivo with insulin [40], revealed marked inhibition of the ability of IRS-1 to associate with the insulin receptor and thus reduced tyrosine phosphory­lation (and hence decreased activation) of IRS-1. Impaired action of IRS-1 contributes to subsequent defective PI3K and Akt phosphorylation. Surprisingly, the remaining pathway of insulin signaling (Ras/MAPK pathway) was not impaired. These data point to a direct, post-receptoral interaction between HCV and the insulin signaling pathway. Since the PI3K/Akt pathway is critical for the insulin-mediated inhibition of gluconeogenesis in the liver, the authors concluded that the observed defect may lead to increased glucose synthesis in CHC [2]. The direct action of the virus on the insulin signaling pathway has been suggested in experimental models, although various mechanisms have been implicated [40,41]. At the molecular level, upregulation of SOCS, downregulation of peroxisome proliferator activated receptor (PPAR), increased proteasome activator 28γ (PA28γ) level, activation of the mTOR pathway and intense oxidative stress are suggested mechanisms implicated in disturbances in the insulin signaling pathway. All of them may be present in a genotype-specific manner [42].
HCV core protein alone stimulates increased proteasome-mediated degradation of IRS-1, mediated by activation of a member of the SOCS family [42-45]. A HCV genotype specific mechanism of impairment of insulin signaling was observed, since expression of the HCV genotype 3 core protein led to downregulation of PPARγ and upregulation of SOCS-7, while the core protein of genotype 1 activated mTOR and induced phosphorylation of IRS-1 at inhibitory serine residues [42]. Subsequent work suggested that PPARγ may directly control the SOCS-7 level in cells expressing the HCV genotype 3 core [46]. It was later shown that the activation of SOCS family members may be a mechanism common to all major HCV genotypes [47], including genotype 1, since the variant originally associated with mTOR activation was shown to be infrequent among known isolates. Intrahepatic SOCS activation (at both the mRNA and protein levels) has been reported in several human studies, with the level of activation being correlated with obesity [48] or hepatic IR [20].
PPARs, which belong to the nuclear receptor super­family, play a pivotal role in HCV-induced IR. Their action is determined by heterodimerization with receptor X for retinoids (RXR) [49,50]. The PPAR-RXR complex initiates gene transcription after being bound with a ligand such as eicosanoids, unsaturated free fatty acids, and very low-density lipoprotein, which change its confirmation and facilitate binding to DNA at PPAR response elements [51] PPARα and PPARγ alongside RXR belong to the main nuclear receptors expressed in the liver, controlling glucose and lipid metabolism, influencing cellular differentiation and proliferation as well as regulating the inflammatory process. In the course of CHC, PPARα gene expression in the liver was decreased by 86% [52]. The study by Gottardi et al. showed that PPARγ liver expression was significantly lower in patients with genotype 3 infection compared to those with genotype 1 [52]. PPARα-mediated regulation of lipid metabolism is closely related to adiponectin [53]. Adiponectin levels seem to be lower in CHC. Accumulation of triglycerides in hepatocytes contributes to the loss of adiponectin receptors in the liver, and together with low serum adiponectin concentration leads to systemic IR [54]. Additionally, PPARγ enhances adiponectin production. As PPARγ is markedly reduced in CHC, there is another obvious pathway for IR and hepatic steatosis [54].
HCV core protein may also affect the insulin signaling via activation of the proteasome activator 28γ (PA28γ) [45]. Transgenic mice expressing the HCV core develop IR [43], which is reversible through the targeted deletion of PA28γ, suggesting that HCV may induce IR through a PA28γ-dependent pathway [45]. Activation of the PA28γ-dependent pathway by HCV core protein leads to suppression of IRS-1 tyrosine phosphorylation and IRS-2 expression and TNF-α pro­moter activation [46]. Moreover, PA28γ is involved in the development and progression of steatosis and occurrence of HCC [55]. HCV core protein inhibits insulin-mediated FoxO1 translocation from the nucleus to the cytoplasm and subsequently reduces accumulation of FoxA2 in the nucleus [56]. FoxO1 and FoxA2 belong to specific Forkhead box transcriptional regulators which modulate hepatic glucose metabolism via the PI3K/Akt signaling pathway [57].
Protein phosphatase 2A (PP2A) and NF-κB are other agents engaged in HCV-induced IR. HCV NS5A may induce IR via overexpression of PP2A [58]. PP2A, which dephosphorylate and inactivate Akt [58,59], is significantly upregulated in CHC patients [60]. In vitro, PP2A overexpression results from enhanced HCV-related endoplasmic reticulum (ER) stress [61]. However, the intrahepatic level of PP2A did not correlate with HOMA-IR [58]. In another study, the HCV core protein alone or in the presence of other viral proteins increased the serine phosphorylation of IRS-1, an effect that was abolished by inhibiting the JNK signaling pathway [62]. JNK inhibitors also restored the hepatocyte glucose uptake reduced by the HCV core expression. Thus, JNK may contribute to HCV-induced IR, as suggested also by recent data in chronic hepatitis C patients [63]. HCV NS5A potentiates the NF-κB-mediated increase in proinflammatory cytokines (i.e. IL-6, TNF-α), contributing to mitochondrial reactive oxygen species (ROS) production [64-66]. Also, NS3-induced oxidative stress may activate NF-κB, potentiating the inflammatory process and IR [67].
All the above-mentioned mechanisms imply a direct effect of viral products on the intracellular insulin signaling pathway. However, CHC is a chronic inflammatory state, in which synthesis of pro-inflammatory agents is enhanced. The close interaction between hepatocytes and various immune cells such as Kupffer cells, macrophages, lymphocytes and dendritic cells may be of importance when inflammatory circumstances appear. In the study by Vanni et al. [20], hepa­tic IR was associated with enhanced hepatic expression of IL-18. IL-18-mediated IR may result from inhibition of adiponectin expression in adipocytes [68] and activation of SOCS-3 in the adipose tissue [69]. Also the activation of JNK by the HCV core [62] may occur via pro-inflammatory cytokines, such as TNF-α The role of TNF-α in inducing IR in HCV-infected persons is debated. The IR in transgenic mice expressing the HCV core, associated with high serum concentrations of TNF-α, can be reverted by anti-TNF-α antibodies [43], implying that HCV core-expressing hepatocytes secrete TNF-α, which may then induce IR via serine phosphorylation of IRS-1. This provokes reduction of GLUT2/GLUT4 gene expression [70], resulting in decreased glucose uptake into hepatocytes and adipocytes [71]. In CHC patients, circulating TNF-α levels are increased [72-75], and may be related to IR independently of the fibrosis stage [76]. However, in a controlled study [77], serum levels of TNF-α and IL-6 in 154 non-diabetic CHC patients were found to be significantly higher compared to 75 matched uninfected controls. Unfortunately, there was no cor­relation between TNF-α serum concentration and IR.
In contrast, another study carried out in a group of 161 CHC patients found serum TNF-α levels to be re­lated to hepatic steatosis grade and HOMA-IR values, whereas serum adiponectin concentration was inversely associated with HOMA-IR values, serum TNF-α levels and steatosis grade, independently of HCV genotype and gender [78]. Summarized information on HCV proteins affecting insulin sensitivity pathways are shown in Table 1.
Another study revealed that serum levels of adiponectin and leptin levels were significantly associated with IR, but not with HCV infection itself. This observation suggests virus-specific IR in CHC to be a cytokine-independent effect [77]. The role of new adipokines in regulation of insulin sensitivity in CHC also cannot be excluded. However, the data are very limited. Vaspin was found to improve insulin sensitivity and glucose tolerance and down-regulate TNF-α synthesis [79-81]. In CHC serum vaspin levels significantly decreased in patients without or with no advanced fibrosis and increased in cases of advanced fibrosis [16,82]. This suggests vaspin to be a compensatory mechanism switching in CHC-associated insulin resistance. However, there was no relationship between vaspin levels and viral load and HOMA-IR values [82]. The next new adipokine is visfatin, which is significantly increased in CHC, and inversely associated with inflammatory activity [16]. On one hand visfatin exerts an insulin-like effect, enhancing phosphorylation of IRS-1 and expression of PPARγ [83], but on the other hand it potentiates expression of some adhesion molecules, IL-6 and TNF-α [75,84]. Chemerin is another new adipokine, the concentration of which is markedly altered in CHC, but is not associated with HOMA-IR values and viral load [16]. However, considering chemerin’s ability to inhibit TNF-α and IL-6 synthesis and to ameliorate IRS phosphorylation, to improve adipocytes’ insulin sensitivity and to stimulate adiponectin synthesis [81,85,86], it may be suggested to be another player implicated in regulation of IR in CHC. Retinol binding protein (RBP)-4 is a member of the lipocalin family, and is synthesized predominantly in hepatocytes (80%) and adipocytes (20%) [87]. It is implicated in steatogenesis, and the link has been made between insulin resistance, T2DM and elevated serum and adipose RBP-4 concentration [88-90]. The study by Petta et al. [91] showed serum RBP-4 to be independently linked to steatosis in linear regression in CHC patients. On logistic regression, RBP-4 was independently associated with moderate to severe steatosis. However, another study by Kukla et al. did not confirm these observations in CHC [92]. Further investigations involving a larger number of patients are required to better determine the exact role of novel adi­pokines in CHC.
It is well known that oxidative stress and overproduction of ROS are closely associated with metabolic abnormalities and progression of chronic liver diseases. CHC as a chronic inflammatory process and metabolic disease is characterized by exacerbated oxidative stress. HCV core protein can evoke oxidative stress in hepatocytes by increased ROS production and levels of lipid peroxidation products and reduced mitochondrial glutathione reserve [93-96]. The data regarding the contribution of oxidative stress to HCV-induced IR is contradictory. The study by Oliveira et al. carried out in 52 genotype non-3 infected CHC patients found oxidative stress to be related to IR but not steatosis [97]. In contrast, in the Vidali et al.’s study oxidative stress which was observed in 61% out of 107 genotype non-3 CHC patients, was associated with steatosis, but not with IR. Moreover, steatosis was also related to IR and fibrosis. Based on these results, the authors concluded that in genotype non-3 CHC oxidative stress and IR induce and potentiate hepatic steatosis, which in turn exacerbates both IR and oxidative stress and accelerates fibrosis progression [98].

Mechanisms of HCV-induced steatosis

Since lipids are essential in the HCV life cycle, they must be accumulated in a sufficient amount in infected hepatocytes. Hepatic steatosis apart from inflammation and fibrosis is another characteristic histological feature of CHC. It is observed in 42-73% of patients. Its prevalence is two-fold higher compared to CHB [99-101], and it is up-regulated by concomitant metabolic factors.
The suggestion of a direct HCV influence on lipid accumulation in the liver may be supported by three lines of evidence: 1) steatosis is more frequent and severe in patients with genotype 3 [102,103], pointing to the presence of specific sequences across the genotype 3 genome contributing to the storage of lipid droplets in hepatocytes; 2) viral load is associated with steatosis grade [102,103]; 3) efficient antiviral therapy leads to a smaller steatosis extent [102,104,105]. These observations are more obvious in patients with genotype 3 infection. In patients infected with the genotype non-3 virus, steatosis is more likely to be associated with metabolic abnormalities [103,105]. However, in the case of recurrence of HCV after the end of antiviral therapy, reappearance of steatosis occurs in patients in whom it had been resolved during treatment.
Although in some circumstances such as IR the uptake of fatty acids by hepatocytes is enhanced, there are well-evidenced experimental studies showing the HCV core protein to be sufficient in evoking hepatic steatosis by triglycerides accumulation [16]. The most creative seems to be genotype 3. A study of primary infection in chimpanzees revealed that HCV activates genes involved in lipid metabolism via the sterol regulatory element binding protein (SREBP), which is responsible for the transactivation of enzymes involved in the synthesis of fatty acids and cholesterol. The predominant activating effect was exerted by core and NS2 and NS4B viral proteins. Additionally, the HCV core protein activates the DNA-binding domain of retinoid receptor α, which regulates lipid synthesis [16]. Alongside activation of transcription factors involved in de novo lipogenesis, such as retinoid X receptor alpha [106] and SREBP-1c [107-110], inadequate lipoprotein secretion seems to be an essential step to trigger neutral fat accumulation, in keeping with the evidence that serum levels of apolipoprotein B and cholesterol are reduced in CHC patients in whom steatosis later responds to antiviral therapy [111]. Furthermore, the disappearance of steatosis in patients who respond to therapy is paralleled by the normalization of cholesterol and apolipoprotein B levels [111,112]. HCV core protein inhibits microsomal triglyceride transfer protein (MTP) activity, which in turn decreases assembly and secretion of apolipoprotein B-containing VLDL [113]. Interestingly, intrahepatic levels of MTP mRNA are reduced in CHC patients, especially those with steatosis and/or genotype 3 [114]. Indeed, the severity of hepatic steatosis in patients with CHC is similar to that observed in hypo-betalipoproteinemia [113]. Impairment of fatty acid oxidation resulting from HCV infection may also add to the fatty liver. Transfection of hepatoma cells with the HCV core protein results in reduced expression of PPARα, which is responsible for fatty acid degradation [108]. PPARα mRNA is markedly decreased in the liver of CHC patients [52,115], especially in those with genotype 3 infection [52]. Moreover, carnitine palmitoyl acyl-CoA transferase 1A, a target gene of PPARα responsible for mediating long-chain fatty acid transport across the mitochondrial membrane, is reduced by HCV both in vitro [116] and in CHC patients’ livers [117]. The HCV sequence responsible for the fatty accumulation is not definitively known. Probably, phenylalanine at position 164 of the core sequence of genotype 3a, which is replaced by a tyrosine in all other genotypes, is responsible for enhancement of fatty acid synthesis [108] and accumulation of large lipid droplets in hepatocytes [118]. The recent data suggest a pivotal role of micro-RNA (miRNA) in the pathogenesis of steatosis and HCV replication. Especially mir122, the most abundant liver miRNA, was found to accelerate steatosis development by up-regulation of lipogenesis and by enhancing HCV replication [119].

Relationship between insulin resistance and steatosis in HCV infection

The relationship between IR and HCV infection is complex and bidirectional. HCV induces steatosis, and the latter could also induce and exacerbate IR. Many mechanisms accounting for HCV-related steatosis can also induce IR. In addition to the inflammatory process, HCV proteins may evoke and potentiate IR and oxidative stress [120].
Steatosis has been related to up-regulated ROS production, which promotes lipid peroxidation and enhances hepatic stellate cell activation. In the case of steatosis, increased production of some proinflammatory and profibrotic cytokines is observed. Noteworthy is that steatosis makes the liver more susceptible to TNF-α-mediated inflammation, liver injury, and apoptosis [121], accelerating fibrosis progression in con­sequence [122].
On the other hand, patients with the highest grades of steatosis do not necessarily present with the highest levels of IR, and vice versa. In HCV genotype 3 infection, IR levels are comparable in patients with versus those without steatosis [123]. Studies have shown that HOMA-IR levels are higher in patients with geno­types 1 and 4 [19], and that patients with genotype 3 are those in whom HOMA-IR levels are the lowest [17]. In a study from Greece, HOMA-IR levels were comparable across viral genotypes [124], and these results suggest that the severe steatosis observed in genotype 3 may not result in increased IR. It is noteworthy that SREBP-1c not only influences production of enzymes regulating lipogenesis and reducing fatty acid β-oxidation, but is a protein interfering in intracellular insulin signaling. In mice overexpressing SREBP-1c, enhanced systemic IR and hepatic steatosis were observed. Inhibition of SOCS-1 and -3 in this model normalized the levels of SREBP-1c and ameliorated both hepatic steatosis and insulin sensitivity [125]. Oxidative damage induced by the HCV core protein may simultaneously induce steatosis and disturb insulin signaling in the hepatocyte. These observations explain why inflammation in the liver, which may contribute to IR via the mechanisms mentioned above, was positively associated with steatosis grade in some studies [104,126].

Influence on cardiovascular morbidity and mortality

The complex interaction between HCV and the MS, oxidative stress, the inflammatory process, steatosis and disturbances in adipokine profile provokes the question whether this interference may translate into an increase of cardiovascular morbidity and mortality in CHC. Convincing evidence that HCV infection promotes IR and T2DM development amplifies this suspicion. Since HCV infection is a well-known factor stimulating the inflammatory response, it is believed to lead to cardiovascular disease development by several effector mechanisms, including up-regulation of intracellular adhesion molecules, expression of anti-endothelium antibodies, and induction of oxidative stress and IR [6,127,128]. Recently, the relationships between HCV infection, atherosclerosis, and the inflammatory response have been precisely investigated. However, the data concerning the relationship between CHC and atherosclerosis is highly equivocal [6,129-131]. There are also some data pointing to a possible association between HCV infection and coronary artery disease [131-133] and stroke [134].
In the Heart and Soul Study [135], CHC patients were found to have higher TNF-α serum concentration and an increased risk of cardiac failure and death compared to those without HCV infection. Mostafa et al. [136] found a greater intima-media thickness (IMT) in CHC patients than controls. Oliveira et al. [137] compared non-obese, non-diabetic CHC patients with healthy controls and reported an intermediate cardiovascular risk, as measured by the Framingham score, a raised concentration of IL-6 and TNF-α, and a higher ratio of pro-inflammatory/anti-inflammatory cytokines (TNF-α/IL-10 and IL-6/IL-10) in CHC patients [137]. Another study found the cytokine imbalance – TNF-α/adiponectin ratio – to be associated with development of IR in CHC patients [78]. Moreover, higher concentrations of intercellular adhesion molecule-1, vascular cell adhesion molecule-1 and oxidized low-density lipoproteins have been observed in CHC patients [127,128,138]. The detection of HCV RNA sequences and intermediate replicative forms within carotid plaques suggests local pro-atherogenic action of HCV [138,139]. However, the presence of HCV RNA within carotid plaques may be due to the known interaction between HCV and LDL, and an active role of HCV in atherogenesis should be further investigated [133].
As for NAFLD [140-142], HCV-related steatosis may be suggested as a cardiometabolic risk factor [128]. CHC patients demonstrated a higher prevalence of carotid atherosclerosis than both healthy controls and NAFLD patients [131]. It is noteworthy that it occurred at a younger age, particularly in the presence of hepatic steatosis, and viral load was found to be independently associated with intima media thickness [131]. HCV viral load and HCV-related steatosis have been hypothesized to favor the development of atherosclerosis (OR 5.21 and 12.98, respectively) independently of known risk factors such as hypercholesterolemia, smoking, hypertension and inflammatory cytokines [131]. Somewhat the study by Targher et al. showed different results. Atherosclerosis assessed by carotid IMT was higher in CHC patients than in healthy controls, but less significantly than in NAFLD patients [143].
Another interesting study, on CHC Egyptian patients, did not find age- and sex-adjusted mean carotid IMT or the proportion of individuals with carotid pla-que to be different between CHC patients and healthy controls. IMT was independently associated with classical atherogenic risk factors, namely LDL cholesterol and systolic blood pressure [136]. However, when these factors appeared, CHC patients had higher IMT, suggesting a direct effect of viral infection on plaque development. HCV infection is a risk factor for earlier and facilitated occurrence of atherosclerosis via viral load and steatosis, which modulate atherogenic factors such as inflammation and dysmetabolic milieu.

Conclusions and implications for management

HCV influences various metabolic aspects (Fig. 2) [16]. Several data suggest that HCV is able to alter intrahepatic insulin signaling through various mechanisms, including direct interference of the virus with the intracellular insulin cascade or functional impairment, e.g. via increased levels of proinflammatory cytokines or through oxidative stress. In chronic hepatitis C, IR is mainly impaired in the liver, but a variable degree of peripheral IR can coexist in the same individual, possibly mediated by superimposed factors such as hepatic stea­tosis. Furthermore, the metabolic disturbances caused by HCV per se can interact with the degree of liver inflammation and fibrosis and with the classical risk factors for T2DM, further aggravating IR. In fact, the increased prevalence and incidence of T2DM carried by HCV is consistently linked to predisposing conditions. This suggests that HCV infection has the potential to trigger the phenotypic expression of metabolic derangements on a genetically determined [144,145], environmentally induced, susceptible basis. IR should be actively sought in patients with HCV also for the implications in management. If IR is present, a potential role of pharmacological therapy can be envisaged.
However, initial results with pioglitazone [146-150] or metformin [151] are contradictory. Our main efforts should rather be directed at promoting healthful dietary practices and physical activity as a cultural norm.
Aforementioned data indicate that HCV may disturb intrahepatic insulin signaling due to different mechanisms, including a direct influence of the virus on the intracellular insulin pathway or indirectly through induction of proinflammatory cytokines and enhancement of oxidative stress. Oxidative stress and metabolic disturbances caused by HCV facilitate hepatic steatosis progression, which accelerates fibrosis progression, potentiates inflammatory activity and, with the traditional risk factors for T2DM, further aggravates IR. Higher prevalence of IR and T2DM in CHC is consistently linked to predisposing circumstances.
The metabolic abnormalities observed in CHC exert a pivotal impact on the morbidity and mortality of infected patients due to higher frequency of hepatocellular carcinoma, accelerated progression of liver fibrosis, and a reduced virological response to standard of care (SOC) treatment with ribavirin and pegylated interferon. HCV infection is a risk factor for earlier and facilitated occurrence of atherosclerosis. Increased risk and prevalence of cardiovascular events in CHC point to additional studies which should focus on the exact role of HCV in atherogenesis.

Disclosure

Authors report no conflict of interest.

References

1. Huang JF, Yu ML, Dai, CY, et al. Glucose abnormalities in hepa­titis C virus infection. Kaohsiung J Med Sci 2013; 29: 61-68.
2. Bugianesi E, Salamone F, Negro F. The interaction of metabolic factors with HCV infection: Does it matter? J Hepatol 2012; 56 Suppl 1: S56-65.
3. Eslam M, Khattab MA, Harrison SA. Insulin resistance and hepatitis C: an evolving story. Gut 2011; 60: 1139-1151.
4. Sykiotis GP, Papavassiliou AG. Serine phosphorylation of insulin receptor substrate-1: a novel target for the eversal of insulin resistance. Mol Endocrinol 2001; 15: 1864-1869.
5. Matthews DR, Hosker JP, Rudenski AS, et al. Homeostasis model assessment: insulin resistance and beta-cell function from plasma fasting glucose and insulin concentrations in man. Diabetologia 1985; 28: 412-419.
6. Ishizaka N, Ishizaka Y, Takahashi E, et al. Association between hepatitis C virus seropositivity, carotid-artery plaque, and intimamedia thickening. Lancet 2002; 359: 133-135.
7. Negro F. Abnormalities of lipid metabolism in hepatitis C virus infection. Gut 2010; 59: 1279-1287.
8. Allison ME, Wreghitt T, Palmer CR, et al. Evidence for a link between hepatitis C virus infection and diabetes mellitus in a cirrhotic population. J Hepatol 1994; 21: 1135-1139.
9. White DL, Ratziu V, El-Serag HB. Hepatitis C infection and risk of diabetes: a systematic review and metaanalysis. J Hepatol 2008; 49: 831-844.
10. Mehta SH, Brancati FL, Sulkowski MS, et al. Prevalence of type 2 diabetes mellitus among persons with hepatitis C virus infection in the United States. Ann Intern Med 2000; 133: 592-599.
11. Mehta SH, Brancati FL, Strathdee SA, et al. Hepatitis C virus infection and incident type 2 diabetes. Hepatology 2003; 38: 50-56.
12. Chen T, Jia H, Li J, et al. New onset diabetes mellitus after liver transplantation and hepatitis C virus infection: meta-analysis of clinical studies. Transpl Int 2009; 22: 408-415.
13. Foxton MR, Quaglia A, Muiesan P, et al. The impact of diabetes mellitus on fibrosis progression in patients transplanted for hepatitis C. Am J Transplant 2006; 6: 1922-1929.
14. Demirci MS, Toz H, Yilmaz F, et al. Risk factors and consequences of post-transplant diabetes mellitus. Clin Transplant 2010; 24: E170-E177.
15. Fabrizi F, Messa P, Martin P, et al. Hepatitis C virus infection and post-transplant diabetes mellitus among renal transplant patients: a meta-analysis. Int J Artif Organs 2008; 31: 675-682.
16. Kukla M, Mazur W, Bułdak RJ, et al. Potential role of leptin, adi­ponectin and three novel adipokines – visfatin, chemerin and vaspin – in chronic hepatitis. Mol Med 2011; 17: 1397-1410.
17. Hui JM, Sud A, Farrell GC, et al. Insulin resistance is associated with chronic hepatitis C virus infection and fibrosis progression [corrected]. Gastroenterology 2003; 125: 1695-1704.
18. Yoneda M, Saito S, Ikeda T, et al. Hepatitis C virus directly associates with insulin resistance independent of the visceral fat area in nonobese and nondiabetic patients. J Viral Hepat 2007; 14: 600-607.
19. Moucari R, Asselah T, Cazals-Hatem D, et al. Insulin resistance in chronic hepatitis C: association with genotypes 1 and 4, serum HCV RNA level, and liver fibrosis. Gastroenterology 2008; 134: 416-423.
20. Vanni E, Abate ML, Gentilcore E, et al. Sites and mechanisms of insulin resistance in nonobese, nondiabetic patients with chronic hepatitis C. Hepatology 2009; 50: 697-706.
21. Milner KL, van der Poorten D, Trenell M, et al. Chronic hepatitis C is associated with peripheral rather than hepatic insulin resistance. Gastroenterology 2010; 138: 932-941.e931-933.
22. Knobler H, Schihmanter R, Zifroni A, et al. Increased risk of type 2 diabetes in noncirrhotic patients with chronic hepatitis C virus infection. Mayo Clin Proc 2000; 75: 355-359.
23. Lecube A, Hernandez C, Genesca J, et al. Glucose abnormalities in patients with hepatitis C virus infection: epidemiology and pathogenesis. Diabetes Care 2006; 29: 1140-1149.
24. Petit JM, Bour JB, Galland-Jos C, et al. Risk factors for diabetes mellitus and early insulin resistance in chronic hepatitis C. J Hepatol 2001; 35: 279-283.
25. Thompson AJ, Patel K, Chuang WL, et al. Viral clearance is associated with improved insulin resistance in genotype 1 chronic hepatitis C but not genotype 2/3. Gut 2012; 61: 128-134.
26. Romero-Gómez M, Fernández-Rodríguez CM, Andrade RJ, et al. Effect of sustained virological response to treatment on the incidence of abnormal glucose values in chronic hepatitis C. J Hepatol 2008; 48: 721-727.
27. Arase Y, Suzuki F, Suzuki Y, et al. Sustained virological response reduces incidence of onset of type 2 diabetes in chronic hepatitis C. Hepatology 2009; 49: 739-744.
28. Kawaguchi T, Ide T, Taniguchi E, et al. Clearance of HCV improves insulin resistance, beta-cell function, and hepatic expression of insulin receptor substrate 1 and 2. Am J Gastroenterol 2007; 102: 570-576.
29. Chehadeh W, Abdella N, Ben-Nakhi A, et al. Risk factors for the development of diabetes mellitus in chronic hepatitis C virus genotype 4 infection. J Gastroenterol Hepatol 2009; 24: 42-48.
30. Giordanino C, Bugianesi E, Smedile A, et al. Incidence of type 2 diabetes mellitus and glucose abnormalities in patients with chronic hepatitis C infection by response to treatment: results of a cohort study. Am J Gastroenterol 2008; 103: 2481-2487.
31. Saltiel AR, Kahn CR. Insulin signaling and the regulation of glucose and lipid metabolism. Nature 2001; 414: 799-806.
32. Biddinger SB, Kahn CR. From mice to men: insights into the insulin resistance syndromes. Annu Rev Physiol 2006; 68: 123-158.
33. Bugianesi E, McCullough AJ, Marchesini G. Insulin resistance: a metabolic pathway to chronic liver disease. Hepatology 2005; 42: 987-1000.
34. Koren S, Fantus IG. Inhibition of the protein tyrosine phosphatase PTP1B: potential therapy for obesity, insulin resistance and type-2 diabetes mellitus. Best Pract Res Clin Endocrinol Metab 2007; 21: 621-640.
35. Vinciguerra M, Foti M. PTEN and SHIP2 phosphoinositide phosphatases as negative regulators of insulin signalling. Arch Physiol Biochem 2006; 112: 89-104.
36. Rui L, Yuan M, Frantz D, et al. SOCS-1 and SOCS-3 block insulin signaling by ubiquitin-mediated degradation of IRS1 and IRS2. J Biol Chem 2002; 277: 42394-42398.
37. Tanti JF, Jager J. Cellular mechanisms of insulin resistance: role of stress-regulated serine kinases and insulin receptor substrates (IRS) serine phosphorylation. Curr Opin Pharmacol 2009; 9: 753-762.
38. Virkamaki A, Ueki K, Kahn CR. Protein-protein interaction in insulin signaling and the molecular mechanisms of insulin resistance. J Clin Invest 1999; 103: 931-943.
39. Kukla M, Berdowska A, Gabriel A, et al. Association between hepatic angiogenesis and serum adipokine profile in non-obese chronic hepatitis C patients. Pol J Pathol 2011; 4: 218-228.
40. Aytug S, Reich D, Sapiro LE, et al. Impaired IRS-1/PI3-kinase signaling in patients with HCV: a mechanism for increased prevalence of type 2 diabetes. Hepatology 2003; 38: 1384-1392.
41. Kawaguchi T, Yoshida T, Harada M, et al. Hepatitis C virus down- regulates insulin receptor substrates 1 and 2 through up-regulation of suppressor of cytokine signaling 3. Am J Pathol 2004; 165: 1499-1508.
42. Pazienza V, Clément S, Pugnale P, et al. The hepatitis C virus core protein of genotypes 3a and 1b downregulates insulin receptor substrate 1 through genotype-specific mechanisms. Hepatology 2007; 45: 1164-1171.
43. Shintani Y, Fujie H, Miyoshi H, et al. Regulation of glucose transporter in the development of insulin resistance. Gastroenterology 2004; 126: 840-848.
44. Banerjee S, Saito K, Ait-Goughoulte M, et al. Hepatitis C virus core protein upregulates serine phosphorylation of insulin receptor substrate-1 and impairs the downstream akt/protein kinase B signaling pathway for insulin resistance. J Virol 2008; 82: 2606-2612.
45. Miyamoto H, Moriishi K, Moriya K, et al. Involvement of the PA28gamma-dependent pathway in insulin resistance induced by hepatitis C virus core protein. J Virol 2007; 81: 1727-1735.
46. Pazienza V, Vinciguerra M, Andriulli A, et al. Hepatitis C virus core protein genotype 3a increases SOCS-7 expression through PPAR-g in Huh-7 cells. J Gen Virol 2010; 91: 1678-1686.
47. Pascarella S, Clément S, Guilloux K, et al. Effects of hepatitis C virus on suppressor of cytokine signaling mRNA levels: comparison between different genotypes and core protein sequence analysis. J Med Virol 2011; 83: 1005-1015.
48. Walsh MJ, Jonsson JR, Richardson MM, et al. Non-response to antiviral therapy is associated with obesity and increased hepatic expression of suppressor of cytokine signalling 3 (SOCS-3) in patients with chronic hepatitis C, viral genotype 1. Gut 2006; 55: 529-535.
49. Desvergne B, Wahli W. Peroxisome proliferator-activated receptors: nuclear control of metabolism. Endocr Rev 1999; 20: 649-688.
50. Bardot O, Aldridge TC, Latruffe N, et al. PPAR-RXR heterodimer activates a peroxisome proliferator response element upstream of the bifunctional enzyme gene. Biochem Biophys Res Commun 1993; 192: 37-45.
51. Gearing KL, Gottlicher M, Teboul M, et al. Interaction of the peroxisome-proliferator-activated receptor and retinoid X receptor. Proc Natl Acad Sci USA 1993; 90: 1440-1444.
52. de Gottardi A, Pazienza V, Pugnale P, et al. Peroxisome proliferator-activated receptor-alpha and-gamma mRNA levels are reduced in chronic hepatitis C with steatosis and genotype 3 infection. Aliment Pharmacol Ther 2006; 23: 107-114.
53. Tanaka N, Moriya K, Kiyosawa K, et al. PPAR alpha activation is essential for HCV core protein-induced hepatic steatosis and hepatocellular carcinoma in mice. J Clin Invest 2008; 118: 683-694.
54. Haluzik M, Parizkova J, Haluzik MM. Adiponectin and its role in the obesity-induced insulin resistance and related complications. Physiol Res 2004; 53: 123-129.
55. Moriishi K, Mochizuki R, Moriya K, et al. Critical role of PA28gamma in hepatitis C virus-associated steatogenesis and hepatocarcinogenesis. Proc Natl Acad Sci USA 2007; 104: 1661-1666.
56. Banerjee A, Meyer K, Mazumdar B, et al. Hepatitis C virus differentially modulates activation of forkhead transcription factors and insulin-induced metabolic gene expression. J Virol 2010; 84: 5936-5946.
57. Higuchi S, Kubota M, Iguchi K, et al. Transcriptional regulation of aquaporin 3 by insulin. J Cell Biochem 2007; 102: 1051-1058.
58. Bernsmeier C, Duong FH, Christen V, et al. Virus-induced over-expression of protein phosphatase 2A inhibits insulin signaling in chronic hepatitis C. J Hepatol 2008; 49: 429-440.
59. Duong FH, Christen V, Berke JM, et al. Upregulation of protein phosphatase 2Ac by hepatitis C virus modulates NS3 helicase activity through inhibition of protein arginine methyltransferase 1. J Virol 2005; 79: 15342-15350.
60. Ugi S, Imamura T, Maegawa H, et al. Protein phosphatase 2A negatively regulates insulin metabolic signaling pathway by inhibiting Akt (protein kinase B) activity in 3T3-L1 adipocytes. Mol Cell Biol 2004; 24: 8778-8789.
61. Christen V, Treves S, Duong FH, et al. Activation of endoplasmic reticulum stress response by hepatitis viruses up-regulates protein phosphatase 2A. Hepatology 2007; 46: 558-565.
62. Banerjee S, Saito K, Ait-Goughoulte M, et al. Hepatitis C virus core protein upregulates serine phosphorylation substrate-1 and impairs the downstream Akt/protein kinase B signaling pathway for insulin resistance. J Virol 2008; 82: 2606-2612.
63. García-Monzó C, Lo Iacono O, Mayoral R, et al. Hepatic insulin resistance is associated with increased apoptosis and fibrogenesis in nonalcoholic steatohepatitis and chronic hepatitis C. J Hepatol 2011; 54: 142-152.
64. Tardif KD, Mori K, Siddiqui A. Hepatitis C virus subgenomic replicons induce endoplasmic reticulum stress activating an intracellular signaling pathway. J Virol 2002; 76: 7453-7459.
65. Gong G, Waris G, Tanveer R, et al. Human hepatitis C virus NS5A protein alters intracellular calcium levels, induces oxidative stress, and activates STAT-3 and NF-kappa B. Proc Natl Acad Sci USA 2001; 98: 9599-9604.
66. Riordan SM, Skinner NA, Kurtovic J, et al. Toll-like receptor expression in chronic hepatitis C: correlation with proinflammatory cytokine levels and liver injury. Inflamm Res 2006; 55: 279-285.
67. Bureau C, Bernad J, Chaouche N, et al. Nonstructural 3 protein of hepatitis C virus triggers an oxidative burst in human monocytes via activation of NADPH oxidase. J Biol Chem 2001; 276: 23077-23083.
68. Chandrasekar B, Boylston WH, Venkatachalam K, et al. Adiponectin blocks interleukin-18-mediated endothelial cell death via APPL1-dependent AMP-activated protein kinase (AMPK) activation and IKK/NF-kappaB/PTEN suppression. J Biol Chem 2008; 283: 24889-24898.
69. Emanuelli B, Peraldi P, Filloux C, et al. SOCS-3 inhibits insulin signaling and is up-regulated in response to tumor necrosis factor-alpha in the adipose tissue of obese mice. J Biol Chem 2001; 276: 47944-47949.
70. Im SS, Kwon SK, Kim TH, et al. Regulation of glucose transporter type 4-isoform-gene expression in muscle and adipocytes. IUBMB Life 2007; 59: 134-145.
71. Ohmura E, Hosaka D, Yazawa M, et al. Association of free fatty acids (FFA) and tumor necrosis factor-alpha (TNF-αlpha) and insulin-resistant metabolic disorder. Horm Metab Res 2007; 39: 212-217.
72. Crespo J, Rivero M, Fábrega E, et al. Plasma leptin and TNF-αlpha levels in chronic hepatitis C patients and their relationship to hepatic fibrosis. Dig Dis Sci 2002; 47: 1604-1610.
73. Knobler H, Zhornicky T, Sandler A, et al. Tumor necrosis factor-alpha-induced insulin resistance may mediate the hepatitis C virus-diabetes association. Am J Gastroenterol 2003; 98: 2751-2756.
74. Zylberberg H, Rimaniol AC, Pol S, et al. Soluble tumor necrosis factor receptors in chronic hepatitis C: a correlation with histological fibrosis and activity. J Hepatol 1999; 30: 185-191.
75. Kukla M, Warakomska I, Gabriel A, et al. Serum levels of sICAM-1, TNFα, sTNF-R1 and sTNF-R2 in patients with chronic hepatitis C treated with pegylated interferon and ribavirin. Exp Clin Hepatol 2008; 4: OR12-OR20.
76. Maeno T, Okumura A, Ishikawa T, et al. Mechanisms of increased insulin resistance in non-cirrhotic patients with chronic hepatitis C virus infection. J Gastroenterol Hepatol 2003; 18: 1358-1363.
77. Cua IH, Hui JM, Bandara P, et al. Insulin resistance and liver injury in hepatitis C is not associated with virus-specific changes in adipocytokines. Hepatology 2007; 46: 66-73.
78. Durante-Mangoni E, Zampino R, Marrone A, et al. Hepatic steatosis and insulin resistance are associated with serum imbalance of adiponectin/tumour necrosis factor-alpha in chronic hepatitis C patients. Aliment Pharmacol Ther 2006; 24: 1349-1357.
79. Klöting N, Berndt J, Kralisch S, et al. Vaspin gene expression in human adipose tissue: association with obesity and type 2 diabetes. Biochem Biophys Res Commun 2006; 339: 430-436.
80. Wada J. Vaspin: a novel serpin with insulin-sensitizing effects. Expert Opin Investig Drugs 2008; 17: 327-333.
81. Rabe K, Lehrke M, Parhofer KG, et al. Adipokines and insulin resistance. Mol Med 2008; 14: 741-751.
82. Kukla M, Waluga M, Sawczyn T, et al. Serum vaspin may be a good indicator of fibrosis in chronic hepatitis V and is not altered by antiviral therapy. Pol J Pathol 2012; 63: 213-220.
83. Aytug S, Reich D, Sapiro LE, et al. Impaired IRS-1/PI3-kinase signaling in patients with HCV: a mechanism for increased prevalence of type 2 diabetes. Hepatology 2003; 38: 1384-1392.
84. Dahl TB, Haukeland JW, Yndestad A, et al. Intracellular nicotinamide phosphoribosylotransferase protects against hepatocyte apoptosis and is down-regulated in nonalcoholic fatty liver disease. J Clin Endocrinol Metab 2010; 95: 3039-3047.
85. Yoshimura T, Oppenheim JJ. Chemerin reveals a chimeric nature. J Exp Med 2008; 205: 2187-2190.
86. Cash JL, Hart R, Russ A, et al. Synthetic chemerin-derived peptides suppress inflammation through ChemR23. J Exp Med 2008; 205: 767-775.
87. van Dam RM, Hu FB. Lipocalins and insulin resistance: etiological role of retinolbinding protein 4 and lipocalin-2? Clin Chem 2007; 53: 5-7.
88. Yang Q, Graham TE, Mody N, et al. Serum retinol binding protein contributes to insulin resistance in obesity and type 2 diabetes. Nature 2005; 436: 356-362.
89. Abahusain MA, Wright J, Dickerson JW, et al. Retinol, alpha- tocopherol and carotenoids in diabetes. Eur J Clin Nutr 1999; 53: 630-635.
90. Basualdo CG, Wein EE, Basu TK. Vitamin A (retinol) status of first nation adults with non-insulin-dependent diabetes mellitus. J Am Coll Nutr 1997; 16: 39-45.
91. Petta S, Tripodo C, Grimaudo S, et al. High liver RBP4 protein content is associated with histological features in patients with genotype 1 chronic hepatitis C and with nonalcoholic steatohepatitis. Dig Liver Dis 2011; 43: 404-410.
92. Kukla M, Berdowska A, Stygar D, et al. Serum FGF21 and RBP4 levels in patients with chronic hepatitis C. Scand J Gastroenterol 2012; 47: 1037-1047.
93. Korenaga M, Wang T, Li Y, et al. Hepatitis C virus core protein inhibits mitochondrial electron transport and increases ROS production. J Biol Chem 2005; 280: 37481-37488.
94. Okuda M, Li K, Beard MR, et al. Mitochondrial injury, oxidative stress, and antioxidant gene expression are induced by hepatitis C virus core protein. Gastroenterology 2002; 122: 366-375.
95. Moriya K, Nakagawa K, Santa T, et al. Oxidative stress in the absence of inflammation in a mouse model for hepatitis C virus associated hepatocarcinogenesis. Cancer Res 2001; 61: 4365-4370.
96. Abdalla MY, Ahmad IM, Spitz DR, et al. Hepatitis C virus-core and non structural proteins lead to different effects on cellular antioxidant defenses. J Med Virol 2005; 76: 489-487.
97. Oliveira AC, Parise ER, Catarino RM, et al. Insulin resistance and not steatosis is associated with modifications in oxidative stress markers in chronic hepatitis C, non-3 genotype. Free Radic Res 2009; 43: 1187-1194.
98. Vidali M, Tripodi MF, Ivaldi A, et al. Interplay between oxidative stress and hepatic steatosis in the progression of chronić hepatitis C. J Hepatol 2008; 48: 399-406.
99. Czaja AJ, Carpenter HA, Santrach PJ, et al. Host- and disease- specific factors affecting steatosis in chronic hepatitis C. J Hepatol 1998; 29: 198-206.
100. Thomopoulos KC, Arvaniti V, Tsamantas AC, et al. Prevalence of liver steatosis in patients with chronic hepatitis B: a study of associated factors and of relationship with fibrosis. Eur J Gastroenterol Hepatol 2006; 18: 233-237.
101. Peng D, Han Y, Ding H, et al. Hepatic steatosis in chronic hepatitis B patients is associated with metabolic factors more than viral factors. J Gastroenterol Hepatol 2008; 23: 1082-1088.
102. Rubbia-Brandt L, Quadri R, Abid K, et al. Hepatocyte steatosis is a cytopathic effect of hepatitis C virus genotype 3. J Hepatol 2000; 33: 106-115.
103. Adinolfi LE, Gambardella M, Andreana A, et al. Steatosis accelerates the progression of liver damage of chronic hepatitis C patients and correlates with specific HCV genotype and visceral obesity. Hepatology 2001; 33: 1358-1364.
104. Kumar D, Farrell GC, Fung C, et al. Hepatitis C virus genotype 3 is cytopathic to hepatocytes. Genotype-specific reversal of hepatic steatosis after sustained response to antiviral therapy. Hepatology 2002; 36: 1266-1272.
105. Poynard T, Ratziu V, McHutchison J, et al. Effect of treatment with peginterferon or interferon alfa-2b and ribavirin on steatosis in patients infected with hepatitis C. Hepatology 2003; 38: 75-85.
106. Tsutsumi T, Suzuki T, Shimoike T, et al. Interaction of hepatitis C virus core protein with retinoid X receptor alpha modulates its transcriptional activity. Hepatology 2002; 35: 937-946.
107. Su AI, Pezacki JP, Wodicka L, et al. Genomic analysis of the host response to hepatitis C virus infection. Proc Natl Acad Sci USA 2002; 99: 15669-15674.
108. Jackel-Cram C, Babiuk LA, Liu Q. Up-regulation of fatty acid synthase promoter by hepatitis C virus core protein: genotype-3a core has a stronger effect than genotype-1b core. J Hepatol 2007; 46: 999-1008.
109. Waris G, Felmlee DJ, Negro F, et al. Hepatitis C virus induces the proteolytic cleavage of sterol regulatory element binding proteins (SREBPs) and stimulates the phosphorylation of SREBPs via oxidative stress. J Virol 2007; 81: 8122-8130.
110. Lerat H, Kammoun HL, Hainault I, et al. Hepatitis C virus proteins induce lipogenesis and defective triglyceride secretion in transgenic mice. J Biol Chem 2009; 284: 33466-33474.
111. Serfaty L, Andreani T, Giral P, et al. Hepatitis C virus induced hypobetalipoproteinemia: a possible mechanism for steatosis in chronic hepatitis C. J Hepatol 2001; 34: 428-434.
112. Hofer H, Bankl HC, Wrba F, et al. Hepatocellular fat accumulation and low serum cholesterol in patients infected with HCV-3a. Am J Gastroenterol 2002; 97: 2880-2885.
113. Perlemuter G, Sabile A, Letteron P, et al. Hepatitis C virus core protein inhibits microsomal triglyceride transfer protein activity and very low density lipoprotein secretion: a model of viral- related steatosis. Faseb J 2002; 16: 185-194.
114. Mirandola S, Realdon S, Iqbal J, et al. Liver microsomal trigly­ceride transfer protein is involved in hepatitis C liver steatosis. Gastroenterology 2006; 130: 1661-1669.
115. Dharancy S, Malapel M, Perlemuter G, et al. Impaired expression of the peroxisome proliferator-activated receptor alpha during hepatitis C virus infection. Gastroenterology 2005; 128: 334-342.
116. Yamaguchi A, Tazuma S, Nishioka T, et al. Hepatitis C virus core protein modulates fatty acid metabolism and thereby causes lipid accumulation in the liver. Dig Dis Sci 2005; 50: 1361-1371.
117. Cheng Y, Dharancy S, Malapel M, et al. Hepatitis C virus infection down-regulates the expression of peroxisome proliferatoractivated receptor alpha and carnitine palmitoyl acyl-CoA transferase 1A. World J Gastroenterol 2005; 11: 7591-7596.
118. Hourioux C, Patient R, Morin A, et al. The genotype 3-specific hepatitis C virus core protein residue phenylalanine 164 increases steatosis in an in vitro cellular model. Gut 2007; 56: 1302-1308.
119. Hui JM, Sud A, Farrell GC, et al. Insulin resistance is associated with chronic hepatitis C and virus infection fibrosis progression [corrected]. Gastroenterology 2003; 125: 1695-1704.
120. Sheikh MY, Choi J, Qadri I, et al. Hepatitis C virus infection: molecular pathways to metabolic syndrome. Hepatology 2008; 47: 2127-2133.
121. Walsh MJ, Vanags DM, Clouston AD, et al. Steatosis and liver cell apoptosis in chronic hepatitis C: a mechanism for increased liver injury. Hepatology 2004; 39: 1230-1238.
122. Leandro G, Mangia A, Hui J, et al. Relationship between steatosis, inflammation, and fibrosis in chronic hepatitis C: a meta-analysis of individual patient data. Gastroenterology 2006; 130: 1636-1642.
123. Fartoux L, Poujol-Robert A, Guéchot J, et al. Insulin resistance is a cause of steatosis and fibrosis progression in chronic hepatitis C. Gut 2005; 54: 1003-1008.
124. Tsochatzis E, Manolakopoulos S, Papatheodoridis GV, et al. Serum HCV RNA levels and HCV genotype do not affect insulin resistance in nondiabetic patients with chronic hepatitis C: a multicentre study. Aliment Pharmacol Ther 2009; 30: 947-954.
125. Ueki K, Kondo T, Tseng YH, et al. Central role of suppressors of cytokine signaling proteins in hepatic steatosis, insulin resistance, and the metabolic syndrome in the mouse. Proc Natl Acad Sci USA 2004; 101: 10422-10427.
126. Asselah T, Boyer N, Guimont MC, et al. Liver fibrosis is not associated with steatosis but with necroinflammation in French patients with chronic hepatitis C. Gut 2003; 52: 1638-1643.
127. Żwirska-Korczala K, Kukla M, Ziółkowski A, et al. Association of serum sVCAM-1 concentration with fibrosis stage and inflammatory activity grade in chronic hepatitis C patients. Exp Clin Hepatology 2006; 2: 31-35.
128. Kukla M, Żwirska-Korczala K, Gabriel A, et al. sPECAM-1 and sVCAM-1: Role in Pathogenesis and Diagnosis of Chronic Hepatitis C and Association with Response to Antiviral Therapy. Therap Adv Gastroenterol 2009; 2: 79-90.
129. Ishizaka Y, Ishizaka N, Takahashi E, et al. Association between hepatitis C virus core protein and carotid atherosclerosis. Circ J 2003; 67: 26-30.
130. Völzke H, Schwahn C, Wolff B, et al. Hepatitis B and C virus infection and the risk of atherosclerosis in a general population. Atherosclerosis 2004; 174: 99-103.
131. Adinolfi LE, Restivo L, Zampino R, et al. Chronic HCV infection is a risk of atherosclerosis. Role of HCV and HCV related steatosis. Atherosclerosis 2012; 221: 496-502.
132. Vassalle C, Masini S, Bianchi F, et al. Evidence for association between hepatitis C virus seropositivity and coronary artery disease. Heart 2004; 90: 565-566.
133. Butt AA, Xiaoqiang W, Budoff M, et al. Hepatitis C virus infection and the risk of coronary disease. Clin Infect Dis 2009; 49: 225-232.
134. Liao CC, Su TC, Sung FC, et al. Does hepatitis C virus infection increase risk for stroke? A population-based cohort study. PLoS One 2012; 7: e31527.
135. Tsui JI, Whooley MA, Monto A, et al. Association of hepatitis C virus seropositivity with inflammatory markers and heart failure in persons with coronary heart disease: data from the Heart and Soul study. J Card Fail 2009; 15: 451-456.
136. Mostafa A, Mohamed MK, Saeed M, et al. Hepatitis C infection and clearance: impact on atherosclerosis and cardiometabolic risk factors. Gut 2010; 59: 1135-1140.
137. Oliveira CP, Kappel CR, Siqueira ER, et al. Effects of hepatitis C virus on cardiovascular risk in infected patients: a comparative study. Int J Cardiol 2013; 164: 221-226.
138. Boddi M, Abbate R, Chellini B, et al. HCV infection facilitates asymptomatic carotid atherosclerosis: preliminary report of HCV RNA localization in human carotid plaques. Dig Liver Dis 2007; 39 Suppl 1: S55-S60.
139. Boddi M, Abbate R, Chellini B, et al. Hepatitis C virus RNA localization in human carotid plaques. J Clin Virol 2010; 47: 72-75.
140. Volzke H, Robinson DM, Kleine V, et al. Hepatic steatosis is associated with an increased risk of carotid atherosclerosis. World J Gastroenterol 2005; 11: 1848-1853.
141. Fracanzani AL, Burdick L, Raselli S, et al. Carotid artery intima-media thickness in nonalcoholic fatty liver disease. Am J Med 2008; 121: 72-78.
142. Gastaldelli A, Basta G. Ectopic fat and cardiovascular disease: what is the link? Nutr Metab Cardiovasc Dis 2010; 20: 481-490.
143. Targher G, Bertolini L, Padovani R, et al. Differences and similarities in early atherosclerosis between patients with non-alcoholic steatohepatitis and chronic hepatitis B and C. J Hepatol 2007; 46: 1126-1132.
144. Joy T, Lahiry P, Pollex RL, et al. Genetics of metabolic syndrome. Curr Diab Rep 2008; 8: 141-148.
145. Sookoian S, Pirola CJ. Metabolic syndrome: from the genetics to the pathophysiology. Curr Hypertens Rep 2011; 13: 149-157.
146. Overbeck K, Genné D, Golay A, et al. Pioglitazone in chronic hepatitis C not responding to pegylated interferon-alpha and ribavirin. J Hepatol 2008; 49: 295-298.
147. Elgouhari HM, Cesario KB, Lopez R, et al. Pioglitazone improves early virologic kinetic response to PEG IFN/RBV combination therapy in hepatitis C genotype 1 naive pts. Hepatology 2008; 48 Suppl: 383A.
148. Conjeevaram H, Burant CF, McKenna, et al. A randomized, double-blind, placebo-controlled study of PPAR-gamma agonist pioglitazone given in combination with peginterferon and ribavirin in patients with genotype-1 chronic hepatitis C. Hepatology 2008; 48 Suppl: 384A.
149. Harrison SF, Hamzeh M, Lentz E, et al. Virologic and metabolic responses in chronic hepatitis C (CHC) patients with insulin resistance (IR) treated with pioglitazone and peginterferon alpha-2a plus ribavirin. J Hepatol 2010; 52: S129.
150. Khattab M, Emad M, Abdelaleem A, et al. Pioglitazone improves virological response to peginterferon alpha-2b/ribavirin combination therapy in hepatitis C genotype 4 patients with insulin resistance. Liver Int 2010; 30: 447-454.
151. Romero-Gómez M, Diago M, Andrade RJ, et al. Treatment of insulin resistance with metformin in naïve genotype 1 chronic hepatitis C patients receiving peginterferon alfa-2a plus ribavirin. Hepatology 2009; 50: 1702-1708.
Copyright: © 2015 Clinical and Experimental Hepatology. This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International (CC BY-NC-SA 4.0) License (http://creativecommons.org/licenses/by-nc-sa/4.0/), allowing third parties to copy and redistribute the material in any medium or format and to remix, transform, and build upon the material, provided the original work is properly cited and states its license.
Quick links
© 2024 Termedia Sp. z o.o.
Developed by Bentus.