INTIMP-03497; No of Pages 10 International Immunopharmacology xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

International Immunopharmacology journal homepage: www.elsevier.com/locate/intimp

1

Review

4Q4

Reshmi Chatterjee a, Abhisek Mitra b,⁎ a

7

a r t i c l e

8 9 10 11 12

Article history: Received 8 October 2014 Received in revised form 12 December 2014 Accepted 14 December 2014 Available online xxxx

13 14 15 16 17 18 19

Keywords: HBV HCV HCC Therapy Biomarker Clinical trials

i n f o

R O

E.S.I. Hospital, Belur, West Bengal, India University of Texas MD Anderson Cancer Center, Houston, TX, USA

a b s t r a c t

T

E

D

P

Chronic liver diseases (CLDs) such as hepatitis, alcoholic liver disease, nonalcoholic fatty liver, and their downstream effect cancer affect more than a billion of people around the world both symptomatically and asymptomatically. The major limitation for early detection and suitable medical management of CLDs and liver cancer is either the absent of symptoms or their similar manifestations as other diseases. This detection impediment has led to a steady increase in the number of people suffering from CLDs with an ultimate outcome of liver failure and undergoing transplantation. A better understanding of CLD pathogenesis has helped us to develop novel therapies for patients who are at greatest risk for CLD progression to the most serious disease cancer. With the discovery of aberrant molecular pathways in CLDs, it is now possible to delineate a road map for selecting targeted therapies for CLDs. Technological advances in imaging as well as the availability of several stable, sensitive, early, noninvasive biomarkers for distinguishing different stages of CLDs and cancer have greatly facilitated both drug target identification and real-time monitoring of response to therapy. Biomarkers are the most useful in clinical practice for liver diseases like hepatocellular carcinoma (HCC), which is associated with secretion of various tumor-related proteins or nucleotides in peripheral circulation. The need for the identification of CLD biomarkers remains high. This article reviews the etiologies of CLDs, the results of recent clinical trials of treatments for CLDs, and development of noninvasive methodologies for detecting CLDs and monitoring their progression toward HCC. © 2014 Elsevier B.V. All rights reserved.

E

C

b

O

5 6

F

3

An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer

2Q3

36

R

40 38 37 39

3.

4.

R

1. 2.

Introduction . . . . . . . . . . . . . . . . Types of CLDs and associated HCC . . . . . . 2.1. Viral hepatitis . . . . . . . . . . . . 2.2. Nonalcoholic fatty liver disease (NAFLD) 2.3. Alcoholic liver disease (ALD) . . . . . 2.4. Hepatocellular carcinoma (HCC) . . . Treatment of CLDs and HCC . . . . . . . . . 3.1. Treatment for viral hepatitis . . . . . 3.2. Treatment for NAFLD . . . . . . . . 3.3. Treatment for ALD . . . . . . . . . . 3.4. Treatment for HCC . . . . . . . . . Novel non-invasive biomarkers . . . . . . . 4.1. Serum markers . . . . . . . . . . . 4.2. Circulating tumor cells (CTCs) . . . . 4.3. Novel protein markers for HCC . . . . 4.4. Circulating microRNAs (miRNAs) . . .

N C O

43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58

Contents

U

42 41

. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . .

0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0

Abbreviations: CLD, Chronic liver disease; HBV, Hepatitis B virus; HCV, Hepatitis C virus; HBsAg, Hepatitis B surface antigen; HBeAg, Hepatitis B early antigen; NAFLD, Nonalcoholic liver disease; NASH, Nonalcoholic steatohepatitis; ALD, Alcoholic liver disease; HCC, Hepatocellular carcinoma; MAPK, Mitogen activated protein kinase; TGFβ, Transforming growth factor β; ALT, Alanine transaminase; AST, Aspartate aminotransferase; CTC, Circulating tumor cells; miRNA, MicroRNA; MRE, Magnetic resonance elastography; TE, Transient elastography ⁎ Corresponding author at: The University of Texas MD Anderson Cancer Center, 1515 Holcombe Blvd., Houston, TX 77030, USA. Tel.: +1 713 563 0440; fax: +1 713 534 5871. E-mail address: [email protected] (A. Mitra).

http://dx.doi.org/10.1016/j.intimp.2014.12.024 1567-5769/© 2014 Elsevier B.V. All rights reserved.

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35

2

59 60 61 62

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

4.5. Circulating microparticles . 4.6. Novel imaging technologies 5. Conclusion . . . . . . . . . . . References . . . . . . . . . . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

0 0 0 0

63

2.1. Viral hepatitis

101 102

Hepatitis viruses have a special tropism toward hepatocytes and an ability to cause both acute and chronic liver diseases. There are 5 different hepatitis viruses from A to E of which predominantly B and C are causing CLDs and HCC. According to World Health Organization data, more than 5 million people in the United States and 500 million people worldwide are chronically infected with a hepatitis virus. A recent study predicted that the percentage of cirrhosis patients who are infected with hepatitis C virus (HCV) is likely to increase from 25% (of 3.5 million patients with cirrhosis) to 45% by 2030 [1]. On the basis of significant advances in our understanding of the life cycles of hepatitis B virus (HBV) and HCV (Fig. 1A), the chronic infection stage of these viruses can be divided into three stages: high immunological tolerance, immune competence, and immune control. Immunological tolerance is characterized by lack of symptoms; normal liver function; in HBV infection, high levels of hepatitis B surface antigen (HBsAg) and hepatitis B envelope antigen (HBeAg) in serum, which reduce helper T responses; and in HCV infection, PD-1 expression,

89 90 91 92 93 94 95 96 97

103 104 105 106 107 108 109 110 111 112 113 114 115 116 117

C

87 88

E

85 86

R

83 84

R

81 82

O

79 80

C

77 78

N

75 76

U

73 74

F

100

71 72

O

2. Types of CLDs and associated HCC

69 70

R O

99

67 68

118 119 120 121 122 123 124 125 126 127

2.2. Nonalcoholic fatty liver disease (NAFLD)

128

NAFLD is a broad spectrum of diseases ranging from simple benign steatosis to nonalcoholic steatohepatitis (NASH) to progressive fibrosis, which can lead to cryptogenic cirrhosis and HCC. Sedentary lifestyles, poor food habits, and silent progression have contributed to the fatty liver disease epidemic in the Western world, affecting at least 25% of Americans [10]. The incidence of NAFLD has steadily increased from 57.5% to 74% in obese individuals [11]. Due to exponential growth of NAFLD, NASH is predicted to be the principal cause of cirrhosis and liver transplantation by 2020 [12]. The histological phenotypes of NASH are virtually indistinguishable from those of alcoholic steatohepatitis, which affected 6 million Americans according to a World Gastroenterology Organization report [13]. In 30–40% of NASH cases, the disease progresses to fibrosis, and in 10–15% of the cases, it progresses to cirrhosis [14]. The accumulation of fat in the liver can be mediated by several mechanisms, including lipolysis of adipose tissue, increased hepatic lipogenesis, and reduced secretion of low-density lipoprotein [15] (Fig. 1B). The risk factors associated with the transition from NAFLD to NASH have not been precisely identified; however, a multifactorial model has been proposed [16]. According to this model, pathological characteristics similar to those associated with NASH can be stemmed from different factors, including dietary intake of cholesterol [17], endotoxemia [15], interference with intestinal microbiota, Toll-like receptor 5 deficiency [18], hypoadiponectinemia [19], and endoplasmic reticulum stress due to exposure of hepatocytes to toxic lipids [20]. In a recent preclinical model showed that hepatic NKT cells are critical regulator of fibrogenesis in NASH by activating hedgehog signaling pathway to promote HSCs to stimulate pro-fibrogenic factor osetopontin [21] (Fig. 1B). Given these observations, progression to NASH is associated with lipid accumulation and inflammation in the liver.

129

2.3. Alcoholic liver disease (ALD)

159

ALD is a leading cause of liver disease-related mortality, causing 287,365 deaths in a 24-year-period in the United States [22]. ALDrelated mortality accounts for 40% of cases of cirrhosis and 28% of all liver disease-related deaths [22]. Among individuals who over a period of more than 20 years consumed on average 70 12-oz beers with 4% alcohol content per week or 70 1.5-oz glasses of wine with 11% alcohol content per week, 19% developed ALD and 7% of those with ALD developed cirrhosis [23]. From a histological perspective, ALD encompasses a broad spectrum of phenotypes, including steatosis, steatohepatitis, alcoholic hepatitis, cirrhosis, and HCC [24]. ALD-related liver injury involves acetaldehyde formation, glutathione depletion, increased lipogenesis resulting from increased

160

P

98

The liver is the principal detoxifying organ and is actively involved in clearing pathogens, toxic chemicals, and metabolic waste products from the body to maintain homeostasis. Continuous exposure of the liver to viruses, excessive alcohol consumption, excessive fat accumulation, biotransformed metabolites, and cholestasis can cause hepatic injury, which can lead to inflammation and liver degeneration. When the injury sustains for a long time, it can cause CLDs, which occur in multistage processes of fibrosis, cirrhosis, and HCC. As the diseases progress to chronic stage, there is infiltration of polymorphonuclear leukocytes into the liver, with focal or zonal necrosis, destruction of hepatocytes, and architectural disarray. According to American Liver Foundation data, approximately 30 million Americans have been affected by liver and biliary diseases. In 2007, according to data from the Centers for Disease Control and Prevention, CLDs, cirrhosis and HCC were among the top 15 leading causes of death in the United States, and in 2010 alone, 31,802 people died of liver-related disease excluding 18,910 deaths from liver cancer. Patients with alcohol-related cirrhosis account for half of all hospital admissions associated with liver cirrhosis. It is now recognized that the increasing incidence of CLDs is associated with not only cirrhosis or HCC but also systemic impairments such as chronic fatigue, non-encephalopathyrelated cognitive impairment, vasomotor disturbances, and sleep disturbances. The major cause of liver disease progression to chronic stage or early stage of cancer is the absence of or minimal clinical manifestation. Thus, failure of patients to recognize symptoms of liver diseases during acute stage is a substantial risk related to death from CLDs. It is important to increase awareness especially among people associated with liverrelated disorders for a periodic monitoring in clinics. In the last decade, both therapeutically and noninvasive detection methods for CLDs have significantly improved. This article provides an overview of current concepts in CLD and HCC epidemiology, molecular pathways, therapeutic strategies and noninvasive biomarkers for diagnosing and monitoring disease progression.

D

65 Q6 66

which exhausts CD8 T cells [2–4]. In addition, HCV-specific interleukin 10-producing CD8 T cells greatly facilitate progression to chronic phase in minimal necroinflammatory liver diseases [5, 6]. During the immune competence phase, hepatocytes are infected via low-density lipoprotein receptor [7], and immune activity commences. The immune active phase of chronic HCV infection is associated with HCV-specific T cells and intrahepatic lymphocytes at a higher frequency in the liver than in peripheral circulation [8, 9]. Infiltration by immune cells and subsequent eradication of HCV-infected hepatocytes constitute the major mechanism of protecting the liver from viral pathogens.

E

1. Introduction

T

64

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

130 131 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158

161 162 163 164 165 166 167 168 169 170 171

3

172

Fig. 1.

U

Q1

N C O

R

R

E

C

T

E

D

P

R O

O

F

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

177 178

permeabilization of intestine, which increases fatty acid uptake by the liver from adipose tissue, and inhibition of β-oxidation of fatty acids as a result of an increased NADH/NAD ratio in hepatocytes [24] (Fig. 1C). Chronic alcohol consumption leads to activation of the innate immune system, especially Kupffer cells, which secrete proinflammatory cytokines like tumor necrosis factor-α, and it can lead to liver injury by recruiting lymphocytes.

179

2.4. Hepatocellular carcinoma (HCC)

180 181

HCC is the fifth leading cause of cancer and second leading cause of cancer-related death in the world [25, 26]. Approximately 700,000 new HCC cases are detected and more than 600,000 deaths are

173 174 175 176

182

associated with HCC every year worldwide [27]. Approximately 75–80% of HCC frequently occurs in HBV- or HCV-infected individuals whose disease has progressed to chronic stage such as cirrhosis [28]. Extensive epidemiological studies revealed that liver carcinogenesis has very complex etiologies and, in addition to being associated with viral infection, it is linked with several other risk factors: foodstuffs containing aflatoxin B1, obesity, diabetes, alcohol consumption, and tobacco use. Furthermore, to understand the pathophysiology of HCC, it is important to study the cellular signaling pathways that play a critical role in the development and progression of liver cancer. Multiple signaling pathways undergo dysregulation in HCC: Wnt/βcatenin, p53, mitogen-activated protein kinase (MAPK), PI3K/AKT/ mTOR, and transforming growth factor β (TGFβ) (Fig. 1D) [29–34].

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

183 184 185 186 187 188 189 190 191 192 193 194 195

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

3.1. Treatment for viral hepatitis

236 237

Treatment for chronic viral hepatitis continues to improve as new data on the viral life cycle inside the host emerge. In the case of HBV, many studies indicated that the ideal therapeutic approach would be aimed at completely suppressing HBV DNA rather than achieving HBeAg seroconversion, as HBV DNA is found in many patients [41, 42]. Based on success rates in clinical trials, seven drugs have been approved by the US Food and Drug Administration for treatment of viral hepatitis: peginterferon α-2a, interferon α-2b, lamivudine, adefovir, tenofovir, entecavir, telbivudine and clevudine. Interferon-α therapy is considered as the first approved treatment for chronic HBV infected patients with minimal chance to develop drug resistant strain. Compared with interferon α-2b, peginterferon α-2a was reported to have higher efficacy with a better pharmacokinetic profile, to elicit less drug resistance, and to require a less stringent dosing regimen [43]. Combination therapy with peginterferon α-2a and lamivudine, a nucleoside analogue, suppressed HBV DNA more effectively that peginterferon α-2a alone in a randomized clinical trial [44]. In a recent in vitro study showed that high dose of IFN-α or lymphotoxin β receptor activation upregulates APOBEC3A and APOBEC3B cytidine deaminases that causes degradation of covalently closed circular DNA of HBV to inhibit HBV reactivation.

223 224

228 229 230 231 232

238 239 240 241 242 243 244 245 246 247 248 249 250 251 252 253 254 255 256

C

221 222

E

219 220

R

217 218

R

215 216

O

213 214

C

211 212

N

209 210

U

207 208

3.2. Treatment for NAFLD

295

Fatty liver disease has a direct association with obesity, insulin resistance, type 2 diabetes, and dyslipidemia. Presently, lifestyle modifications like a low-carbohydrate diet and daily exercise are considered standard recommendations for reducing steatosis. However, many patients have difficulty adhering to food and exercise regimens. Alternatively, bariatric surgery has shown promising results in reducing steatosis in obese patients; however, postoperative complications such as bleeding, wound infection, anastomotic leak, thromboembolism, bowel obstruction, gallstones, and death limit the use of this approach [49]. Pharmacological agents can also be used to alleviate steatosis and NASH. In randomized clinical trials, vitamin E and the peroxisome proliferator-activated receptor gamma agonist pioglitazone, a thiazolidinedione, consistently demonstrated histological efficacy [50]. In one recent randomized clinical trial, patients receiving vitamin E showed significant improvement in a primary histological endpoint measure with reductions of at least 1 point in hepatocyte ballooning and 1 point in lobular inflammation or steatosis score in 43% of patients [51]. In another randomized clinical trial, pioglitazone treatment improved insulin sensitivity in hepatocytes, reduced inflammation, and resuscitate liver histology in 73% of patients [50]. Vitamin E is generally preferred over pioglitazone as it is safer, more cost-effective, and has fewer side effects. Metformin, pentoxifylline, statins, and incretin mimetics are also available for treatment of NAFLD; however, histological improvement, efficacy, and long-term safety need to be validated for these agents in large clinical studies.

296 297

F

235

205 206

O

233 234

Treatment of patients with CLDs has been a big challenge for physicians because CLD is usually diagnosed at the irreversible-cirrhosis, cancer or liver failure stage. The considerable progress that has been achieved in well-performed translational studies suggests that significant improvement in treatment of CLDs may occur in the near future. Table 1 lists the results of the most recent clinical trials, and the various therapies currently used to treat CLDs and HCC are discussed in depth below.

203 204

R O

227

202

P

3. Treatment of CLDs and HCC

200 201

257 258

D

226

198 199

Despite its clinical success, peginterferon α-2a is not widely prescribed for HBV treatment due to HBV genotypic limitation (A and B). In contrast, numerous nucleoside analogues, such as lamivudine, adefovir, tenofovir, entecavir, and telbivudine, are widely used for long-term treatment of HBV infection, as these agents can inhibit virus replication and have excellent tolerance and safety profiles. Tenofovir, a reverse transcriptase inhibitor, is the latest drug approved by the US Food and Drug Administration for HBV treatment. Forty-eight weeks of tenofovir therapy showed promising results: viral suppression in 76% of patients, normalization of alanine transaminase in 68%, and HBsAg loss in 3.2% [45]. The efficacy of tenofovir is especially high in HBeAg-negative patients, with viral suppression observed in 93% of such patients in a phase III study [45]. Tenofovir can be considered the most potent medication against chronic HBV infection with no reported drug resistance. Treatment of chronic HCV infection improved significantly after the crystal structures of several HCV proteins were characterized. Triple therapy combining the protease inhibitor boceprevir or telaprevir with peginterferon-α and ribavirin achieved a sustained viral response in 59–69% of patients with chronic HCV infection [46]. Presently, a number of direct-acting antiviral agents targeting HCV RNA polymerase or nonstructural protein 5A and host-targeted compounds are under study. In 2014, there is a significant advancement in interferon or ribavirin free therapy for chronic HCV type I genotype which is affecting 70% of cases in the USA [47, 48]. FDA approved Harvoni (fixed combination of ledipasvir and sofosbuvir) and Olysio (simeprevir) for chronic HCV type I patients. An open-label study with 100 HCV infected patients were treated with fixed combination of ledipasvir and sofosbuvir alone for 8 or 12 weeks showed great potential to cure in 95% of the treated patients [47]. Similar promising result was obtained with a combination therapy of simeprevir and sofosbuvir in multi-center trials with 168 patients who had previously not responded to pegylated interferon and ribavirin. Based on 8 and 12 weeks of therapy, it showed high rates of SVR (95–100%) in patients even in those with compensated cirrhosis and non-responder to protease-inhibitor regimens. With fewer tolerable side effects and improved treatments, these emerging therapies have the potential to reduce HCV occurrence in high risk population.

T

225

Several studies showed that inactivation of the β-catenin suppressor adenomatous polyposis coli, gain-of-function mutations in β-catenin, and mutations in the negative regulators of the Wnt pathway Axin1 and Axin2 were found in 50% of HCC patients [29]. As in other cancers, p53 protein undergoes alteration due to a spectrum of mutations in different positions between exons 4 and 9 [35]. The two most frequently occurring p53 mutations in HCC are a transversion caused by aflatoxin B1 at the third base in codon 249 and a transversion in codon 250 due to the effect of oxidative stress [29]. Mutations in the p53 gene cause genomic instability and abrogate the gene's regulatory role for proteins associated with the cell cycle, thereby increasing the chance of tumor development. Another important pathway, MAPK, which plays a critical role in cellular growth and differentiation, is constitutively activated in many HCC patients. The PI3K/AKT/mTOR signaling pathway can be considered another promising target in liver carcinogenesis, given that it has been reported to undergo aberrant activation in 48% of HCC patients [36]. The mutation at Akt S473, the most frequent HCC-associated mutation in this pathway (found in 71% of HCC patients), promotes invasion, metastasis, and vascularization in HCC tumors [37]. Alteration of the TGFβ signaling pathway has also been commonly observed in HCC progression, with overexpression of TGFβ and downregulation of TGFβ receptor II reported in 40% and 37–70% of HCC patient, respectively [38]. The tumorigenic role of TGFβ signaling is thought to occur through activation of the mitogenic Ras/Raf/MAPK pathway, which promotes cell proliferation independently of SMAD [39]. Furthermore, the TGFβ pathway plays a critical role in tumor invasion and metastasis by upregulating extracellular matrix proteins and decreasing the production of enzymes that degrade those proteins [40]. The TGFβ signaling pathway thus plays important roles in both the early and the late phases of liver carcinogenesis.

196 197

E

4

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284 285 286 287 288 289 290 291 292 293 294

298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx t1:1 t1:2

5

Table 1 An overview of recently completed clinical trials for CLDs.

t1:3

Drug name

Type of CLD

Phase Outcome of trial of trial

Mechanism of action

References

t1:4 t1:5 t1:6 t1:7

HBsAg–HBIG immunogenic complex therapeutic vaccine Besifovir HEPLISAV

HBV

III

[108]

HBV HBV

IIb III

Modulation viral antigen processing and presentation Inhibitor of HBV polymerase HBV vaccine targets TLR9

t1:8

NASVAC

HBV

III

t1:9

Daclatasvir plus sofosbuvir

HCV

III

t1:10

HCV

III

HCV

II

HCV protease inhibition and immune stimulation Vaccine targeting NS3, NS4 and NS5B of HCV Farnesoid X receptor agonist

[112]

t1:12

Simeprevir plus peg-interferon and ribavirin TG4040 plus peg-interferon and ribavirin FLINT

t1:13

Spironolactone and vitamin E

NAFLD II

[115]

t1:14

Phlebotomy

NAFLD II

t1:15

Obeticholic acid

NAFLD II

t1:16 t1:17

Ezetimibe Cysteamine bitartrate

NAFLD II NAFLD IIb

t1:18

Pioglitazone

NASH

IV

t1:19

Liraglutide

NASH

II

Iron reduction and significant improvement in liver histology Decrease insulin resistance, liver inflammation and fibrosis. Ameliorate hepatic fibrosis Reduces ALT, AST values but no significant improvement in body mass index. Significant improvement from lobular inflammation, steatosis and necroinflammation. Well tolerated and improves liver histology

t1:20

Glucocorticoids and N-acetylcysteine

ALD

III

Improves 1 month survival of patients.

t1:21 t1:22 t1:23 t1:24 t1:25 Q2 t1:26 t1:27

X-linked inhibitor of apoptosis (XIAP) antisense AEG3515 and sorafenib Tivantinib Sorafenib and gemcitabine Sorafenib and tegafur/uracil MGN-3 with interventional therapy Pegylated arginine deaminase

HCC

II

More promising than sorafenib alone.

Improve insulin resistance in NAFLD patients Targeting iron overload to increase oxidative stress Farnesoid X receptor ligand to regulate glucose and lipid metabolism. Inhibition of cholesterol absorption Potent antioxidant to increase GSH production Decrease lipid accumulation and increase glucose uptake Glucagon-like-peptide-1 agonist to increase insulin secretion Inhibits AP-1 and NF-κB and anti-oxidant Multiple cellular pathway blocker

HCC HCC HCC HCC HCC

II II II II II

Potential alternative option for advanced HCC patients. Not better than sorafenib alone. More efficacious than sorafenib alone. Useful for the treatment of HCC. Promising and well tolerated.

t1:28

TRC105 anti-angiogenic monoclonal antibody 5-Fluorouracil/leucovorin (FOLFOX4)

HCC

II

HCC

III

Well tolerated in patients. Suitable as a combination therapy. Increased response rate and median overall survival.

C

T

E

D

P

F

R O

NAFLD IIb

O

Superior response compared to peg-interferon or ribavarin Decreases body weight, γ-glutamyl transpeptidase and NAS score Decrease in serum insulin and insulin-resistance

[109] Dynavax Technologies Vaccine comprises of HBsAg and HBcAg [110], to cause immunogenicity NCT01374308 NS5A inhibitor [111]

Oral inhibitor of MET Multiple cellular pathway blocker Multiple cellular pathway blocker Potent immune system booster Systemic deprivation of arginine in cancer cells Targeting CD105 to inhibit tumor angiogenesis. Cytotoxicity

[113] [114]

[116] [117] [118] [119] [120] [121] [122] [123] [124] [125] [126] [127] [128] [129] [130]

R

R

t1:29

Complete seroconversion as early as in 30 days and seroprotection with highest α-HBs titer at day 90. High rate of SVR response among patients infected with HCV 1, 2, or 3 Well tolerated and high rate of SVR response

E

t1:11

HBeAg seroconversion is decreased. Overdosing caused immune fatigue. Similar effect as entecavir. Complete seroprotection after three doses.

3.3. Treatment for ALD

322

Successful strategies for management of ALD include medication, liver transplantation, behavioral therapy, and psychosocial assistance to facilitate complete abstinence from drinking. Primary medications approved by the US Food and Drug Administration for ALD are disulfiram (an inhibitor of alcohol dehydrogenase), acamprosate (an anticraving drug), baclofen (a γ-aminobutyric acid B receptor agonist), and naltrexone (an opioid antagonist). For alcoholic patients with hepatitis, corticosteroids, pentoxifylline, nutritional therapy, and liver transplantation are advised. Because of heterogeneity in clinical outcomes, especially in hepatotoxicity, for the various drug treatments for ALD, the search continues for new compounds and molecular targets. Chemokines, the extracellular matrix protein osteopontin, gut microbiota, endocannabinoid receptors, and the inflammasome are considered key targets for designing next-generation treatments for ALD. Besides medication, cognitive–behavioral therapies can provide selfmotivation for coping with the urge of binge drinking, analysis of situation, and skills for abstaining from alcohol. These nonmedicinal approaches offer highly structured instructions and assignments and include support of family members in directing patients toward healthy activities. Such modification of behaviors related to alcohol consumption can be a safer mode of treatment than medication.

325 326 327 328 329 330 331 332 333 334 335 336 337 338 339 340 341 342

U

323 324

N C O

321

3.4. Treatment for HCC

343

Selection of appropriate therapy for HCC should take into account tumor size and location, underlying liver function, the presence or absence of cirrhosis, and the physical status of the patient. Sorafenib, surgical resection, transcatheter arterial chemoembolization (TACE), radiofrequency ablation (RFA), ethanol ablation, combination therapy, and liver transplantation are the most common treatments for HCC. Sorafenib, a receptor tyrosine kinase inhibitor, showed promising results in patients with advanced HCC: the survival rate in those patients was 10.7 months, compared to 7.9 months for the placebo group [52]. In that clinical trial, drug-induced grade 3 toxic effects were reported in very few patients: diarrhea in 8% of the patients, hand-foot syndrome in 8%, hypertension in 2%, and abdominal pain in 1%. The authors of that study concluded that sorafenib alone or in combination can be effective against HCC patients [52]. Liver transplantation is considered the second-best option for treatment for HCC, as it has been reported to increase the patient survival rate by 67–91% [53]. The only drawbacks of liver transplantation are the need to administer immunosuppressive drugs, which might cause tumors to relapse, and the fact that only 5% of HCC patients are eligible for this procedure [54]. Another approach, surgical resection, can be used to remove a tumor along with surrounding hepatic tissue if the non-cirrhotic region is greater than 25% of the total

344

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361 362 363 364

395

4.1. Serum markers

396 397

Serum markers that reflect hepatic dysfunction are referred to as indirect markers, whereas serum markers that predict changes in extracellular matrix turnover are referred to as direct markers. Potential indirect markers include alanine transaminase, aspartate aminotransferase, γ-glutamyl transpeptidase, total bilirubin, alkaline phosphatase, and albumin. As these substances leak from injured hepatocytes into the bloodstream, their levels in serum increase above the normal range. Hematological variables such as prothrombin index and platelet count are also considered indirect biomarkers of CLD. In contrast to indirect markers, direct biomarkers reflect modifications in molecular pathogenesis of fibrosis or metabolite levels. Direct markers proposed for detecting fibrosis include glycoproteins such as serum hyaluronate [59], laminin [60], and YKL-40 [61]; collagens such as procollagen III N-peptide [62] and type IV collagen [63]; and collagenases such as tissue inhibitor of metalloproteinase-1 [64] and matrix metalloproteinases, along with collagenase inhibitors [57]. Some of the recent studies provided a better understanding to state whether NAFLD can be considered as a potential risk factor for coronary artery diseases (CAD). Based on the serum concentrations of heat shock protein-70 (HSP-70) and gamma-glutamyl transferase (γ-GT), it can be stated NAFLD has insignificant role in determining CAD [65, 66]. However, circulating levels of HSP-70, γ-GT and Bcl-2 were elevated from lower to higher grades of steatosis. These markers could be utilized as potential tools to understand the progression of NAFLD [65–67]. In ALD, conventional biomarkers are not suitable for distinguishing alcohol-based liver toxic effects from coexisting causes of liver dysfunction. Researchers are also looking for serum markers that reveal information about recent drinking activity, termed state markers, and markers of genetic proclivity for alcohol dependence, termed trait markers [68]. With improvements in measuring techniques, carbohydrate-deficient transferrin, mean corpuscular volume, and N-acetyl-β-hexosaminidase have

387 388 389 390 391 392

398 399 400 401 402 403 404 405 406 407 408 409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426

C

385 386

E

384

R

382 383

R

380 381

O

378 379

C

376 377

N

374 375

U

372 373

4.2. Circulating tumor cells (CTCs)

464

Identification and quantitation of CTCs undoubtedly have significant clinical correlation with tumor aggressiveness, prediction of recurrence and monitoring response of patients toward treatment. The discovery of novel cell surface markers to detect these rare CTCs from peripheral blood of HCC patients is growing. In the past year, in 2013, epithelial cell adhesion molecule, asialoglycoprotein receptor and intracellular adhesion molecule 1 were proposed as novel markers of CTCs in HCC patients [26, 30, 77–80]. These novel CTC markers can potentially complement traditional methods of monitoring disease progression and predicting overall survival for HCC patients.

465 466

4.3. Novel protein markers for HCC

475

Alpha-fetoprotein has also been used as a serological biomarker of HCC. However, because of alpha-fetoprotein's poor detection sensitivity at the early stage of HCC and high serum levels in patients with nonmalignant CLD, alternative protein markers of HCC have been proposed. Sall4, Glypican-3, dickkopf-1 and talin-1 are promising candidates, having demonstrated greater sensitivity and specificity for CLD than has alpha-fetoprotein [32, 81–84]. These proteins are completely undetectable in sera of healthy individuals.

476 477

4.4. Circulating microRNAs (miRNAs)

484

F

393 394

Clinical correlation of response to treatment with disease progression is critical for proper assessment and management of CLDs and HCC. Liver biopsy-based histological studies aimed at determining the extent of architectural distortions in CLDs and HCC are considered the most predictive and specific indicator for selecting a treatment regimen. However, liver biopsy has a number of well-documented limitations. This invasive procedure can lead to bleeding and/or significant pain. It can produce a sampling error due to small specimen size. It can be expensive. It does not allow for dynamic evaluation over time. And different pathologists may interpret histological findings differently [56, 57]. Predictive noninvasive biomarkers have been proposed as an adjunct tool to be used with biopsies. Ideal noninvasive biomarkers for CLDs and HCC need to have high sensitivity, specificity, and reproducibility. They need to outperform traditional liver function tests, be easily detectable in both preclinical and clinical settings, and be correlated with histopathological characteristics of liver diseases [58]. Noninvasive biomarkers include serum markers and findings obtained with imaging techniques and transient elastography (TE). With the help of “omics” technologies such as genomics, proteomics, and metabolomics, research on serum biomarkers continues to advance toward the goal of clinically managing disease progression with greater sensitivity and specificity. With the large number of studies being conducted on serum biomarkers for CLDs and HCC, novel markers that have most of the desirable characteristics described above will likely be identified in the near future.

O

371

R O

4. Novel non-invasive biomarkers

427 428

P

370

367 368

shown promising sensitivities and specificities as state markers [68]. As trait markers, adenylyl cyclase activity [69], gamma-aminobutyric acid [70], dopamine receptor [71], β-endorphin [72], and serotonin transporter [73] have been reported to differentiate alcoholics and nonalcoholics. Furthermore metabolic byproducts of alcohol in the liver are considered as direct serum markers of CLD. These byproducts include acetaldehyde [68], ethyl glucuronide [74], ethyl sulfate [75], 5-hydroxytryptophol [76], dolichols [68], and salsolinol [68]. Usually, these markers have a halflife of only a few hours in biological fluids in cases of acute alcohol consumption; however, in habitual alcoholics, these markers have been detected from 24 h to a few days after alcohol consumption. Increasingly, metabolic products are gaining importance in diagnosis of CLDs because they are not subject to dissimulation by patients. Much attention in CLD serum marker research has focused on detection of disease at the asymptomatic stage and on monitoring disease progression. All noninvasive methodologies being developed have potential limitations, such as the fact that the most commonly used enzymes and proteins are not liver specific and the fact that the level of liver dysfunction has to be high for a marker of CLD to be detectable in serum. In addition, the rate of excretion may alter the concentration of direct-marker metabolic products in serum. In light of these important limitations, comprehensive metaanalyses of these biomarkers need to be conducted, and algorithms predicting clinical outcomes need to be developed. Of the several hundred serum markers of CLD available, none has been independently validated with optimal accuracy for early or longitudinal progression of disease. The diagnostic accuracy of a biomarker is measured as area under the receiver operating characteristic curve. Typically, the values near the ends of the area under the curve indicate either high sensitivity and low specificity or low sensitivity and high specificity, and these results are clinically meaningful for patient prognosis. However, results in the intermediate range, with moderate sensitivity and specificity, do not reliably indicate disease progression. Because of these shortcomings, researchers have created panels of different combinations of biochemical and clinical markers to integrate noninvasive and histopathological methods for diagnosis and treatment of CLDs.

T

369

liver. The 5-year postoperative survival rate for HCC patients who undergo surgical resection is around 40–70%; however, the recurrence rate is almost 70% [31, 55]. Although these various treatment strategies for HCC have shown promise, large clinical trials are required to accurately ascertain recurrence and overall survival rates of patients.

D

365 366

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

E

6

429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463

467 468 469 470 471 472 473 474

478 479 480 481 482 483

Extensive research on miRNAs has linked these small noncoding 485 RNA molecules to liver biology and hepatic dysfunction [85] and 486

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

521 522

Microparticles (MPs) are heterogeneous population of spherical structures with a size range between 100 and 1000 nm [33]. MPs are budding from the plasma membrane with known to express antigen from their parental cells. These are elevated in many disease states such as inflammation, auto immune disorders, malignancies, atherosclerosis, renal failure and numerous infections [34]. Recent studies illustrated that MPs can be considered as a non-invasive biomarker in acute liver failure, chronic hepatitis C and NAFLD where these showed correlation with the extent of histological inflammation in paired biopsies [89–92]. Presently, large number of patient screening and efficient enrichment technology are required to utilize it as universal marker for CLDs.

523

4.6. Novel imaging technologies

524 525

With the significant advances made in the past few decades, imaging technologies are gradually gaining prominence among clinical methods of diagnosing liver diseases. The transition from anatomical to molecular imaging has initiated a new era in noninvasive detection of liver diseases. The imaging methods most commonly used for liver disease are ultrasound, computed tomography, magnetic resonance imaging, positron emission tomography, molecular imaging, transient elastography (TE), controlled attenuate parameter (CAP) and magnetic resonance elastography (MRE). Of the currently available imaging platforms, translational imaging technologies such as positron emission tomography, MRE, and TE are matching with the requirements of clinical needs. Molecular imaging can detect changes in expression of surface proteins on a cell and metabolic functions of inflammatory and tumor cells that might correlate with disease progression. With the availability of new molecular probes for inflammation and tumors, molecular imaging technologies have the potential to visualize, characterize, and offer greater diagnostic flexibility for clinical assessment of liver pathology [93–95]. Researchers are also investigating the plasticity of liver tissue in CLD progression. An active area of investigation is correlation of the degree of liver stiffness due to fibrosis with pathophysiological changes in liver tissue, as it is well known that the liver is stiff at the cirrhosis stage and HCC [96]. The contractile property of the liver has been utilized to measure shear elasticity by transmitting longitudinal waves. A study using a rodent model showed that liver stiffness begins much

515 516 517 518 519 520

526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548

Table 2 Most commonly deregulated miRNAs in chronic liver diseases.

t2:1 t2:2

D

513 514

T

506 507

C

504 505

E

502 503

R

500 501

R

498 499

N C O

496 497

U

494 495

F

511 512

493

O

4.5. Circulating microparticles

491 492

549 550

R O

510

489 490

earlier than collagen deposition [97]. In this method, an ultrasound transducer probe emits a low-frequency (50 Hz) vibration that generates a shear wave and disseminates it into the liver [98]. The clinical accuracy of TE is particularly high in detecting progressive evolution of cirrhosis, for which the sensitivity of TE is greater than 90% [99]. The inter- and intra-observer variability of TE are minimal according to studies of cirrhosis and HCC patients. However, TE has several limitations. This technique is not suitable for detecting early and intermediate stages of fibrosis or measuring complex advanced portal hypertension [96]. In addition, TE is not recommended for obese patients (body mass index above 28 kg/m2) or patients with ascites or narrow intercostal spacing [100] or in case of extrahepatic cholestasis [101]. Due to potential limitations of TE, the recent imaging modality CAP has been utilized efficiently to detect fibrosis and different grades of steatosis [101]. It uses Fibroscan M probe to measure the ultrasound attenuation with values range from 100 to 400 dB/M for steatosis detection [102]. A recent study linking CAP values with steatosis grades (S1 to S3) among 135 patients revealed that it can differentiate more than two grades apart precisely compared to semi-quantitative grading and the sensitivity of this method varied between 73 and 83% and the specificity was 84–96% [103]. Supporting to these observations, a prospective study with more than 5300 examinations showed that technical failure of CAP was observed in only 7.7% of cases [101]. Another major finding in this study was a direct relationship between the mean CAP values with increasing number of alcohol drinks per week. With such a high accuracy, low sampling error rate and a clear diagnostic concept, CAP technology is a promising tool to characterize major histopathological aspects of patients with steatosis. Another elasticity-based imaging technique, MRE, is a threedimensional approach that can use the phase contrast method to visualize propagation of acoustic shear waves transmitted through the liver. The underlying theoretical principles of MRE and TE are similar, but MRE has several advantages over TE: a minimal sampling error rate, the ability to scan almost the whole liver using a three-dimensional displacement vector, and the ability of compression waves to penetrate throughout the liver (making MRE suitable for obese patients and patients with ascites). Furthermore, MRE has excellent sensitivity and specificity, with values above 85% for detecting hepatic fibrosis at stages 2 to 4, based on receiver operating characteristic curve analyses [98]. Over the years, MRE has gained significant importance as a tool for CLD diagnosis and treatment, particularly in helping to match hepatitis patients with appropriate therapy. Imaging evaluation for accurate assessment of response to therapy, recurrent and new tumors are essential for proper management of HCC patients. Based on technological advancements and upgraded algorithms, contrast-enhanced ultrasound (CE-US), multidetector computer tomography, nuclear medicine and MRI (gadolinium, diffusionweighted imaging, blood oxygen level dependent, perfusion-weighted imaging) are recently using for better correlation between hemodynamics and pathophysiology [104]. Size of liver nodules (either N2 or b2 cm) is an important aspect for specificity and sensitivity of the

P

508 509

elucidated their roles in regulating lipid- and glucose-metabolizing enzymes, inflammation, fibrosis, and cancer-related signaling pathways in the liver (Table 2). Elevated levels of circulating miRNAs in serum are highly correlated with pathophysiological changes in inflammationmediated liver diseases. Usually, these extracellular miRNAs are detected in exosomes, microvesicles, or bound to proteins that facilitate extreme stability in acidic environments and RNase resistance. These properties can extend the half-lives of miRNAs by several days over the half-lives of markers in current clinical use. Numerous studies have suggested that elevated levels of miRNA-122 (miR-122) in serum are correlated with various CLDs such as HCV infection, NAFLD, NASH and ALD. Elevated serum levels of other extracellular miRNAs have also been linked to CLDs. miR-155, for example, was found to be elevated in both sera and peripheral monocytes of patients with chronic HCV infection [86]. In patients with chronic HCV infection or NAFLD, serum levels of miR-34a and miR-16 were correlated with disease severity [87]. In patients with cirrhosis, miR-513-3p and miR-571 serum levels were significantly higher than in control individuals [88]. Collectively, these findings suggest that circulating miRNAs may be more, sensitive and more specific than other serum markers for early detection of CLD and for gauging response to treatment. The levels of extracellular miRNAs in serum are much lower than in whole blood, and sophisticated instruments are required to obtain precise measurements.

E

487 488

7

551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599

Circulating miRNAs

Liver diseases

References

t2:3

miR-16 miR-21 miR-34a miR-101 miR-122 miR-125b miR-155 miR-192 miR-194 miR-199 miR-223

HCC, HCV, HBV HCC, HCV HCC, NAFLD HCC, HBV HCC, ALD, HCV, HBV and NAFLD HCC, HCV, ALD HCC, ALD, HCV HCC, HCV, NAFLD HBV, HCV, HCC HCC, HCV, HCC, HBV

[87,131] [132,133] [134,135] [131,136] [131,133,137,138] [139–141] [141–143] [144,145,140] [131,146,147] [148,149] [133]

t2:4 t2:5 t2:6 t2:7 t2:8 t2:9 t2:10 t2:11 t2:12 t2:13 t2:14

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

648

References

625 626 627 628

633 634 635 636 637 638 639 640 641 642 643 644 645 646

649 650 651 652 653 654 655 656 657 658 659 660 661 662 663

C

623 624

E

621 622

R

619 620

R

617 618

O

615 616

C

613 614

N

611 612

U

609 610

F

647

The considerable progress in our understanding of CLDs and HCC at both the molecular and the cellular levels has greatly facilitated the design of numerous drugs and combination therapies that may have clinical importance. Additional multicenter, long-term therapy with comprehensive clinical trials of novel agents and combination treatments are strongly recommended to avoid the potential relapse of CLDs and HCC. Noninvasive biomarkers for detecting and monitoring CLDs and HCC are becoming clinically valuable tools for assessing disease severity. The limitations associated with biomarkers are low sensitivity and specificity in disease staging. This improper assessment of CLDs could be a potential risk factor for disease progression to HCC. Clinical evidence suggests that combinations of biomarkers improve diagnosis and therapy of CLDs and HCC could be avoided. This review provides comprehensive information about the recent updates on therapeutic and diagnostic tools for proper management of patients affected with CLDs and its associated HCC.

607 608

O

632

606

R O

5. Conclusion

604 605

P

631

602 603

[6] Abel M, Sene D, Pol S, Bourliere M, Poynard T, Charlotte F, et al. Intrahepatic virusspecific IL-10-producing CD8 T cells prevent liver damage during chronic hepatitis C virus infection. Hepatology 2006;44:1607–16. [7] Albecka A, Belouzard S, Op de Beeck A, Descamps V, Goueslain L, Bertrand-Michel J, et al. Role of low-density lipoprotein receptor in the hepatitis C virus life cycle. Hepatology 2012;55:998–1007. [8] Rehermann B. Hepatitis C, virus versus innate and adaptive immune responses: a tale of coevolution and coexistence. J Clin Invest 2009;119:1745–54. [9] Li S, Vriend LE, Nasser IA, Popov Y, Afdhal NH, Koziel MJ, et al. Hepatitis C virusspecific T-cell-derived transforming growth factor beta is associated with slow hepatic fibrogenesis. Hepatology 2012;56:2094–105. [10] Michelotti GA, Machado MV, Diehl AM. NAFLD, NASH and liver cancer. Nat Rev Gastroenterol Hepatol 2013;10:656–65. [11] Sass DA, Chang P, Chopra KB. Nonalcoholic fatty liver disease: a clinical review. Dig Dis Sci 2005;50:171–80. [12] Wree A, Broderick L, Canbay A, Hoffman HM, Feldstein AE. From NAFLD to NASH to cirrhosis—new insights into disease mechanisms. Nat Rev Gastroenterol Hepatol 2013;10:627–36. [13] Erickson SK. Nonalcoholic fatty liver disease. J Lipid Res 2009;50:S412–6 [Suppl.]. [14] M I, Singh C, Ganie MA, Alsayari K. NASH: the hepatic injury of metabolic syndrome: a brief update. Int J Health Sci (Qassim) 2009;3:265–70. [15] Tilg H, Moschen AR. Evolution of inflammation in nonalcoholic fatty liver disease: the multiple parallel hits hypothesis. Hepatology 2010;52:1836–46. [16] Coulon S, Legry V, Heindryckx F, Van Steenkiste C, Casteleyn C, Olievier K, et al. Role of vascular endothelial growth factor in the pathophysiology of nonalcoholic steatohepatitis in two rodent models. Hepatology 2013;57:1793–805. [17] Wouters K, van Gorp PJ, Bieghs V, Gijbels MJ, Duimel H, Lutjohann D, et al. Dietary cholesterol, rather than liver steatosis, leads to hepatic inflammation in hyperlipidemic mouse models of nonalcoholic steatohepatitis. Hepatology 2008;48:474–86. [18] Vijay-Kumar M, Aitken JD, Carvalho FA, Cullender TC, Mwangi S, Srinivasan S, et al. Metabolic syndrome and altered gut microbiota in mice lacking Toll-like receptor 5. Science 2010;328:228–31. [19] Musso G, Gambino R, Biroli G, Carello M, Faga E, Pacini G, et al. Hypoadiponectinemia predicts the severity of hepatic fibrosis and pancreatic Beta-cell dysfunction in nondiabetic nonobese patients with nonalcoholic steatohepatitis. Am J Gastroenterol 2005;100:2438–46. [20] Lee JS, Zheng Z, Mendez R, Ha SW, Xie Y, Zhang K. Pharmacologic ER stress induces non-alcoholic steatohepatitis in an animal model. Toxicol Lett 2012;211:29–38. [21] Syn WK, Agboola KM, Swiderska M, Michelotti GA, Liaskou E, Pang H, et al. NKTassociated hedgehog and osteopontin drive fibrogenesis in non-alcoholic fatty liver disease. Gut 2012;61:1323–9. [22] Paula H, Asrani SK, Boetticher NC, Pedersen R, Shah VH, Kim WR. Alcoholic liver disease-related mortality in the United States: 1980–2003. Am J Gastroenterol 2010;105:1782–7. [23] Huang YW, YSS, Kao JH. Pathogenesis and management of alcoholic liver cirrhosis: a review. Hepat Med Evid Res 2010;2011:1–11. [24] Orman ES, Odena G, Bataller R. Alcoholic liver disease: pathogenesis, management, and novel targets for therapy. J Gastroenterol Hepatol 2013;28(Suppl. 1):77–84. [25] Parkin DM. Global cancer statistics in the year 2000. Lancet Oncol 2001;2:533–43. [26] Mitra A, Satelli A, Xia X, Cutrera J, Lopa M, Li S. Cell-surface Vimentin (csVim): a mislocalized protein for isolating csVimentin CD133 novel stem-like hepatocellular carcinoma cells expressing EMT markers. Int J Cancer 2014. [27] Dhanasekaran R, Limaye A, Cabrera R. Hepatocellular carcinoma: current trends in worldwide epidemiology, risk factors, diagnosis, and therapeutics. Hepat Med 2012;4:19–37. [28] Arzumanyan A, Reis HM, Feitelson MA. Pathogenic mechanisms in HBV- and HCV-associated hepatocellular carcinoma. Nat Rev Cancer 2013;13:123–35. [29] Aravalli RN, Steer CJ, Cressman EN. Molecular mechanisms of hepatocellular carcinoma. Hepatology 2008;48:2047–63. [30] Satelli A, Mitra A, Cutrera JJ, Devarie M, Xia X, Ingram DR, et al. Universal marker and detection tool for human sarcoma circulating tumor cells. Cancer Res 2014; 74:1645–50. [31] Mitra A, Satelli A, Yan J, Xueqing X, Gagea M, Hunter CA, et al. IL-30 (IL27p28) attenuates liver fibrosis through inducing NKG2D-rae1 interaction between NKT and activated hepatic stellate cells in mice. Hepatology 2014;60:2027–39. [32] Mitra A, Mishra L, Li S. Technologies for deriving primary tumor cells for use in personalized cancer therapy. Trends Biotechnol 2013;31:347–54. [33] Mitra A, Ross JA, Rodriguez G, Nagy ZS, Wilson HL, Kirken RA. Signal transducer and activator of transcription 5b (Stat5b) serine 193 is a novel cytokine-induced phospho-regulatory site that is constitutively activated in primary hematopoietic malignancies. J Biol Chem 2012;287:16596–608. [34] Nagy ZS, LeBaron MJ, Ross JA, Mitra A, Rui H, Kirken RA. STAT5 regulation of BCL10 parallels constitutive NFkappaB activation in lymphoid tumor cells. Mol Cancer 2009;8:67. [35] Qi LN, Bai T, Chen ZS, Wu FX, Chen YY, De Xiang B, et al. The p53 mutation spectrum in hepatocellular carcinoma from Guangxi, China: role of chronic hepatitis B virus infection and aflatoxin B1 exposure. Liver Int 2014. [36] Villanueva A, Chiang DY, Newell P, Peix J, Thung S, Alsinet C, et al. Pivotal role of mTOR signaling in hepatocellular carcinoma. Gastroenterology 2008;135: 1972–83 [1983 e1971-1911]. [37] Chen JS, Wang Q, Fu XH, Huang XH, Chen XL, Cao LQ, et al. Involvement of PI3K/ PTEN/AKT/mTOR pathway in invasion and metastasis in hepatocellular carcinoma: association with MMP-9. Hepatol Res 2009;39:177–86. [38] Morris SM, Baek JY, Koszarek A, Kanngurn S, Knoblaugh SE, Grady WM. Transforming growth factor-beta signaling promotes hepatocarcinogenesis induced by p53 loss. Hepatology 2012;55:121–31.

T

629 630

imaging technologies to characterize arterial-phase hypervascularity and later washout. In case of nodule size above 2 cm, only a single imaging technique (CT or MRI) is needed to establish confidently its vascular characteristic profile [104]. However smaller lesions ranging from 1 to 2 cm show diagnostic challenges on imaging due to atypical vascularization and features of cirrhotic and dysplastic nodules [105, 106]. As per the guidelines of the American Association for the Study of Liver Diseases (AASLD), two coincident dynamic imaging modalities are required to diagnose malignancy in small neoplastic lesions. Contrary to this, Sangiovanni et al suggested sequential application rather than combined approach for imaging techniques to detect 1–2 cm nodules for improving sensitivity while maintaining high specificity. During HCC surveillance to identify 1–2 cm malignant tumors, single imaging technique was able to detect with 100% specificity and 66% sensitivity compared to 66% specificity and 34% sensitivity in combined procedures [106]. In another study published by Khalili et al, developed an algorithm to detect an additional 20% nodules with a sensitivity of 62%, specificity of 79% and a 73% save of liver biopsies to detect [107]. The authors suggested to use only fine-needle biopsy examination of indeterminate nodules showing either arterial hypervascularity on CT or MRI scan or the presence of a typical synchronous HCC [107]. In conclusion, biphasic combination of arterial phase hypervascularity and delayed phase hypovascularity are essential for the best sensitivity and specificity for diagnosing 1–2 cm HCC nodules with minimal exposure of radiation dose for patients who are likely to undergo multiple follow-up imaging studies. In light of the significant progress that has been made in the use of imaging technologies as noninvasive tools for diagnosing and monitoring CLDs and HCC, these technologies are likely to play an increasingly important role in routine clinical practices, with a potential to decrease the need for liver biopsies.

D

600 601

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

E

8

[1] Davis GL, Alter MJ El-Serag H, Poynard T, Jennings LW. Aging of hepatitis C virus (HCV)-infected persons in the United States: a multiple cohort model of HCV prevalence and disease progression. Gastroenterology 2010;138:513–21 [521 e511-516]. [2] Urbani S, Amadei B, Tola D, Massari M, Schivazappa S, Missale G, et al. PD-1 expression in acute hepatitis C virus (HCV) infection is associated with HCV-specific CD8 exhaustion. J Virol 2006;80:11398–403. [3] Samal J, Kandpal M, Vivekanandan P. Molecular mechanisms underlying occult hepatitis B virus infection. Clin Microbiol Rev 2012;25:142–63. [4] Kondo Y, Ninomiya M, Kakazu E, Kimura O, Shimosegawa T. Hepatitis B surface antigen could contribute to the immunopathogenesis of hepatitis B virus infection. ISRN Gastroenterol 2013;2013:935295. [5] Flynn JK, Dore GJ, Hellard M, Yeung B, Rawlinson WD, White PA, et al. Early IL-10 predominant responses are associated with progression to chronic hepatitis C virus infection in injecting drug users. J Viral Hepat 2011;18:549–61.

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 Q7 709 710 711 712 713 714 715 Q8 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731 732 733 734 735 736 737 738 739 740 Q9 741 742 743 744 745 746 747 748 749

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

N C O

R

R

E

C

D

P

R O

O

F

[70] Krystal JH, Staley J, Mason G, Petrakis IL, Kaufman J, Harris RA, et al. Gammaaminobutyric acid type A receptors and alcoholism: intoxication, dependence, vulnerability, and treatment. Arch Gen Psychiatry 2006;63:957–68. [71] Munafo MR, Johnstone EC, Welsh KI, Walton RT. Association between the DRD2 gene Taq1A (C32806T) polymorphism and alcohol consumption in social drinkers. Pharmacogenomics J 2005;5:96–101. [72] Zalewska-Kaszubska J, Czarnecka E. Deficit in beta-endorphin peptide and tendency to alcohol abuse. Peptides 2005;26:701–5. [73] Oroszi G, Goldman D. Alcoholism: genes and mechanisms. Pharmacogenomics 2004;5:1037–48. [74] Wurst FM, Skipper GE, Weinmann W. Ethyl glucuronide—the direct ethanol metabolite on the threshold from science to routine use. Addiction 2003;98 (Suppl. 2):51–61. [75] Helander A, Beck O. Mass spectrometric identification of ethyl sulfate as an ethanol metabolite in humans. Clin Chem 2004;50:936–7. [76] Stephanson N, Helander A, Beck O. Alcohol biomarker analysis: simultaneous determination of 5-hydroxytryptophol glucuronide and 5-hydroxyindoleacetic acid by direct injection of urine using ultra-performance liquid chromatography– tandem mass spectrometry. J Mass Spectrom 2007;42:940–9. [77] Schulze K, Gasch C, Staufer K, Nashan B, Lohse AW, Pantel K, et al. Presence of EpCAM-positive circulating tumor cells as biomarker for systemic disease strongly correlates to survival in patients with hepatocellular carcinoma. Int J Cancer 2013; 133:2165–71. [78] Li YM, Xu SC, Li J, Han KQ, Pi HF, Zheng L, et al. Epithelial–mesenchymal transition markers expressed in circulating tumor cells in hepatocellular carcinoma patients with different stages of disease. Cell Death Dis 2013;4:e831. [79] Liu S, Li N, Yu X, Xiao X, Cheng K, Hu J, et al. Expression of intercellular adhesion molecule 1 by hepatocellular carcinoma stem cells and circulating tumor cells. Gastroenterology 2013;144:1031–41 [e1010]. [80] Satelli A, Brownlee Z, Mitra A, Meng QH, Li S. Circulating tumor cell enumeration with a combination of epithelial cell adhesion molecule- and cell-surface Vimentin-based methods for monitoring breast cancer therapeutic response. Clin Chem 2014. [81] Han SX, Wang JI, Guo XJ, He CC, Ying X, Ma JL, et al. Serum SALL4 is a novel prognosis biomarker with tumor recurrence and poor survival of patients in hepatocellular carcinoma. J Immunol Res 2014. [82] Capurro M, Wanless IR, Sherman M, Deboer G, Shi W, Miyoshi E, et al. Glypican-3: a novel serum and histochemical marker for hepatocellular carcinoma. Gastroenterology 2003;125:89–97. [83] Shen Q, Fan J, Yang XR, Tan Y, Zhao W, Xu Y, et al. Serum DKK1 as a protein biomarker for the diagnosis of hepatocellular carcinoma: a large-scale, multicentre study. Lancet Oncol 2012;13:817–26. [84] Zhang JL, Qian YB, Zhu LX, Xiong QR. Talin1, a valuable marker for diagnosis and prognostic assessment of human hepatocelluar carcinomas. Asian Pac J Cancer Prev 2011;12:3265–9. [85] Ura S, Honda M, Yamashita T, Ueda T, Takatori H, Nishino R, et al. Differential microRNA expression between hepatitis B and hepatitis C leading disease progression to hepatocellular carcinoma. Hepatology 2009;49:1098–112. [86] Bala S, Tilahun Y, Taha O, Alao H, Kodys K, Catalano D, et al. Increased microRNA155 expression in the serum and peripheral monocytes in chronic HCV infection. J Transl Med 2012;10:151. [87] Cermelli S, Ruggieri A, Marrero JA, Ioannou GN, Beretta L. Circulating microRNAs in patients with chronic hepatitis C and non-alcoholic fatty liver disease. PLoS One 2011;6:e23937. [88] Wang XW, Heegaard NH, Orum H. MicroRNAs in liver disease. Gastroenterology 2012;142:1431–43. [89] Kornek M, Popov Y, Libermann TA, Afdhal NH, Schuppan D. Human T cell microparticles circulate in blood of hepatitis patients and induce fibrolytic activation of hepatic stellate cells. Hepatology 2011;53:230–42. [90] Kornek M, Lynch M, Mehta SH, Lai M, Exley M, Afdhal NH, et al. Circulating microparticles as disease-specific biomarkers of severity of inflammation in patients with hepatitis C or nonalcoholic steatohepatitis. Gastroenterology 2012;143:448–58. [91] Kornek M, Schuppan D. Microparticles: modulators and biomarkers of liver disease. J Hepatol 2012;57:1144–6. [92] Stravitz RT, Bowling R, Bradford RL, Key NS, Glover S, Thacker LR, et al. Role of procoagulant microparticles in mediating complications and outcome of acute liver injury/acute liver failure. Hepatology 2013;58:304–13. [93] Wu C, Li F, Niu G, Chen X. PET imaging of inflammation biomarkers. Theranostics 2013;3:448–66. [94] Ittrich H, Peldschus K, Raabe N, Kaul M, Adam G. Superparamagnetic iron oxide nanoparticles in biomedicine: applications and developments in diagnostics and therapy. Rofo 2013;185:1149–66. [95] Sclavons C, Burtea C, Boutry S, Laurent S, Vander Elst L, Muller RN. Phage display screening for tumor necrosis factor-alpha-binding peptides: detection of inflammation in a mouse model of hepatitis. Int J Pept 2013;2013:348409. [96] Rockey DC. Noninvasive assessment of liver fibrosis and portal hypertension with transient elastography. Gastroenterology 2008;134:8–14. [97] Georges PC, Hui JJ, Gombos Z, McCormick ME, Wang AY, Uemura M, et al. Increased stiffness of the rat liver precedes matrix deposition: implications for fibrosis. Am J Physiol Gastrointest Liver Physiol 2007;293:G1147–54. [98] Talwalkar JA. Elastography for detecting hepatic fibrosis: options and considerations. Gastroenterology 2008;135:299–302. [99] Chon YE, Choi EH, Song KJ, Park JY, Kim do Y, Han KH, et al. Performance of transient elastography for the staging of liver fibrosis in patients with chronic hepatitis B: a meta-analysis. PLoS One 2012;7:e44930.

E

T

[39] Javelaud D, Mauviel A. Crosstalk mechanisms between the mitogen-activated protein kinase pathways and Smad signaling downstream of TGF-beta: implications for carcinogenesis. Oncogene 2005;24:5742–50. [40] Zhang YE. Non-Smad pathways in TGF-beta signaling. Cell Res 2009;19:128–39. [41] Fung J, Yuen MF, Yuen JC, Wong DK, Lai CL. Low serum HBV DNA levels and development of hepatocellular carcinoma in patients with chronic hepatitis B: a case– control study. Aliment Pharmacol Ther 2007;26:377–82. [42] Yuan HJ, Yuen MF, Ka-Ho Wong D, Sablon E, Lai CL. The relationship between HBV– DNA levels and cirrhosis-related complications in Chinese with chronic hepatitis B. J Viral Hepat 2005;12:373–9. [43] Sonneveld MJ, Janssen HL. Chronic hepatitis B: peginterferon or nucleos(t)ide analogues? Liver Int 2011;31(Suppl. 1):78–84. [44] Lau GK, Piratvisuth T, Luo KX, Marcellin P, Thongsawat S, Cooksley G, et al. Peginterferon Alfa-2a, lamivudine, and the combination for HBeAg-positive chronic hepatitis B. N Engl J Med 2005;352:2682–95. [45] Marcellin P, Heathcote EJ, Buti M, Gane E, de Man RA, Krastev Z, et al. Tenofovir disoproxil fumarate versus adefovir dipivoxil for chronic hepatitis B. N Engl J Med 2008;359:2442–55. [46] Sarrazin C, Hezode C, Zeuzem S, Pawlotsky JM. Antiviral strategies in hepatitis C virus infection. J Hepatol 2012;56(Suppl. 1):S88–S100. [47] Lawitz E, Poordad FF, Pang PS, Hyland RH, Ding X, Mo H, et al. Sofosbuvir and ledipasvir fixed-dose combination with and without ribavirin in treatment-naive and previously treated patients with genotype 1 hepatitis C virus infection (LONESTAR): an open-label, randomised, phase 2 trial. Lancet 2014;383:515–23. [48] Manos MM, Shvachko VA, Murphy RC, Arduino JM, Shire NJ. Distribution of hepatitis C virus genotypes in a diverse US integrated health care population. J Med Virol 2012;84:1744–50. [49] Brethauer SA, Chand B, Schauer PR. Risks and benefits of bariatric surgery: current evidence. Cleve Clin J Med 2006;73:993–1007. [50] Lomonaco R, Sunny NE, Bril F, Cusi K. Nonalcoholic fatty liver disease: current issues and novel treatment approaches. Drugs 2013;73:1–14. [51] Sanyal AJ, Chalasani N, Kowdley KV, McCullough A, Diehl AM, Bass NM, et al. Pioglitazone, vitamin E, or placebo for nonalcoholic steatohepatitis. N Engl J Med 2010;362:1675–85. [52] Llovet JM, Ricci S, Mazzaferro V, Hilgard P, Gane E, Blanc JF, et al. Sorafenib in advanced hepatocellular carcinoma. N Engl J Med 2008;359:378–90. [53] Vitale A, Gringeri E, Valmasoni M, D'Amico F, Carraro A, Pauletto A, et al. Long-term results of liver transplantation for hepatocellular carcinoma: an update of the University of Padova experience. Transplant Proc 2007;39:1892–4. [54] Rougier P, Mitry E, Barbare JC, Taieb J. Hepatocellular carcinoma (HCC): an update. Semin Oncol 2007;34:S12–20. [55] Duffy JP, Hiatt JR, Busuttil RW. Surgical resection of hepatocellular carcinoma. Cancer J 2008;14:100–10. [56] Gangadharan B, Antrobus R, Dwek RA, Zitzmann N. Novel serum biomarker candidates for liver fibrosis in hepatitis C patients. Clin Chem 2007;53:1792–9. [57] Adams LA. Biomarkers of liver fibrosis. J Gastroenterol Hepatol 2011;26:802–9. [58] Ozer J, Ratner M, Shaw M, Bailey W, Schomaker S. The current state of serum biomarkers of hepatotoxicity. Toxicology 2008;245:194–205. [59] Guechot J, Laudat A, Loria A, Serfaty L, Poupon R, Giboudeau J. Diagnostic accuracy of hyaluronan and type III procollagen amino-terminal peptide serum assays as markers of liver fibrosis in chronic viral hepatitis C evaluated by ROC curve analysis. Clin Chem 1996;42:558–63. [60] Parsian H, Rahimipour A, Nouri M, Somi MH, Qujeq D, Fard MK, et al. Serum hyaluronic acid and laminin as biomarkers in liver fibrosis. J Gastrointestin Liver Dis 2010;19:169–74. [61] Saitou Y, Shiraki K, Yamanaka Y, Yamaguchi Y, Kawakita T, Yamamoto N, et al. Noninvasive estimation of liver fibrosis and response to interferon therapy by a serum fibrogenesis marker, YKL-40, in patients with HCV-associated liver disease. World J Gastroenterol 2005;11:476–81. [62] Tanwar S, Trembling PM, Guha IN, Parkes J, Kaye P, Burt AD, et al. Validation of terminal peptide of procollagen III for the detection and assessment of nonalcoholic steatohepatitis in patients with nonalcoholic fatty liver disease. Hepatology 2013; 57:103–11. [63] Murawaki Y, Koda M, Okamoto K, Mimura K, Kawasaki H. Diagnostic value of serum type IV collagen test in comparison with platelet count for predicting the fibrotic stage in patients with chronic hepatitis C. J Gastroenterol Hepatol 2001; 16:777–81. [64] Flisiak R, Maxwell P, Prokopowicz D, Timms PM, Panasiuk A. Plasma tissue inhibitor of metalloproteinases-1 and transforming growth factor beta 1—possible noninvasive biomarkers of hepatic fibrosis in patients with chronic B and C hepatitis. Hepatogastroenterology 2002;49:1369–72. [65] Tarantino G, Finelli C, Colao A, Capone D, Tarantino M, Grimaldi E, et al. Are hepatic steatosis and carotid intima media thickness associated in obese patients with normal or slightly elevated gamma-glutamyl-transferase? J Transl Med 2012;10:50. [66] Tarantino G, Colao A, Capone D, Conca P, Tarantino M, Grimaldi E, et al. Circulating levels of cytochrome C, gamma-glutamyl transferase, triglycerides and unconjugated bilirubin in overweight/obese patients with non-alcoholic fatty liver disease. J Biol Regul Homeost Agents 2011;25:47–56. [67] Tarantino G, Scopacasa F, Colao A, Capone D, Tarantino M, Grimaldi E, et al. Serum Bcl-2 concentrations in overweight-obese subjects with nonalcoholic fatty liver disease. World J Gastroenterol 2011;17:5280–8. [68] Peterson K. Biomarkers for alcohol use and abuse. Alcohol Res Health 2005;28: 30–7. [69] Hoffman PL, Glanz J, Tabakoff B. Platelet adenylyl cyclase activity as a state or trait marker in alcohol dependence: results of the WHO/ISBRA study on state and trait markers of alcohol use and dependence. Alcohol Clin Exp Res 2002;26:1078–87.

U

750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773 774 775 776 777 778 779 780 781 782 783 784 785 786 787 788 789 790 791 792 793 794 795 796 797 798 799 800 801 802 803 804 805 806 807 808 809 810 811 812 813 814 815 816 817 818 819 820 821 822 823 824 825 826 827 828 829 830 831 832 833 834 835

9

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

836 837 838 839 840 841 842 843 844 845 846 847 848 849 850 851 852 853 854 855 856 857 858 859 860 861 862 863 864 865 866 867 868 Q10 869 870 871 Q11 872 873 874 875 876 877 878 879 880 881 882 883 884 885 886 887 888 889 890 891 892 893 894 895 896 897 898 899 900 901 902 903 904 905 906 907 908 909 910 911 912 913 914 915 916 917 918 919 920 921

[128]

[129]

[130]

[131] [132]

[133]

D

[134]

C

E

R

R

O

C

N

F

[127]

O

[126]

R O

[125]

apoptosis (XIAP) antisense AEG35156 in combination with sorafenib in patients with advanced hepatocellular carcinoma (HCC). Asco Annual Meeting; 2012. Santoro A, Rimassa L, Borbath I, Daniele B, Salvagni S, Van Laethem JL, et al. Tivantinib for second-line treatment of advanced hepatocellular carcinoma: a randomised, placebo-controlled phase 2 study. Lancet Oncol 2013;14:55–63. Naqi N, Ahmad S, Murad S, Khattak J. Efficacy and safety of sorafenib–gemcitabine combination therapy in advanced hepatocellular carcinoma: an open-label phase II feasibility study. Hematol Oncol Stem Cell Ther 2014;7:27–31. Hsu CH, Shen YC, Lin ZZ, Chen PJ, Shao YY, Ding YH, et al. Phase II study of combining sorafenib with metronomic tegafur/uracil for advanced hepatocellular carcinoma. J Hepatol 2010;53:126–31. Bang MH, Van Riep T, Thinh NT, Song le H, Dung TT, Van Truong L, et al. Arabinoxylan rice bran (MGN-3) enhances the effects of interventional therapies for the treatment of hepatocellular carcinoma: a three-year randomized clinical trial. Anticancer Res 2010;30:5145–51. Glazer ES, Piccirillo M, Albino V, Di Giacomo R, Palaia R, Mastro AA, et al. Phase II study of pegylated arginine deiminase for nonresectable and metastatic hepatocellular carcinoma. J Clin Oncol 2010;28:2220–6. Duffy OER Austin G, Varkey Ulahannan Susanna, Fioravanti Suzanne, Venkatesan Aradhana, Turkbey Ismail B, Choyke Peter L, et al. A phase II study of TR105 in patients with hepatocellular carcinoma (HCC) who have progressed on sorafenib. Asco Annual Meeting; 2013. Qin YB S, Ye S, Fan J, Lim H, Cho JY, Thongprasert S, et al. Phase III study of oxaliplatin plus 5-fluorouracil/leucovorin (FOLFOX4) versus doxorubicin as palliative systemic chemotherapy in advanced HCC in Asian patients. Asco Annual Meeting; 2010. Ji F, Yang B, Peng X, Ding H, You H, Tien P. Circulating microRNAs in hepatitis B virus-infected patients. J Viral Hepat 2011;18:e242–51. Chen Y, Chen J, Wang H, Shi J, Wu K, Liu S, et al. HCV-induced miR-21 contributes to evasion of host immune system by targeting MyD88 and IRAK1. PLoS Pathog 2013; 9:e1003248. Xu J, Wu C, Che X, Wang L, Yu D, Zhang T, et al. Circulating microRNAs, miR-21, miR-122, and miR-223, in patients with hepatocellular carcinoma or chronic hepatitis. Mol Carcinog 2011;50:136–42. Dang Y, Luo D, Rong M, Chen G. Underexpression of miR-34a in hepatocellular carcinoma and its contribution towards enhancement of proliferating inhibitory effects of agents targeting c-MET. PLoS One 2013;8:e61054. Rottiers V, Naar AM. MicroRNAs in metabolism and metabolic disorders. Nat Rev Mol Cell Biol 2012;13:239–50. Su H, Yang JR, Xu T, Huang J, Xu L, Yuan Y, et al. MicroRNA-101, down-regulated in hepatocellular carcinoma, promotes apoptosis and suppresses tumorigenicity. Cancer Res 2009;69:1135–42. Bala S, Petrasek J, Mundkur S, Catalano D, Levin I, Ward J, et al. Circulating microRNAs in exosomes indicate hepatocyte injury and inflammation in alcoholic, drug-induced, and inflammatory liver diseases. Hepatology 2012;56:1946–57. Wen J, Friedman JR. miR-122 regulates hepatic lipid metabolism and tumor suppression. J Clin Invest 2012;122:2773–6. Liang L, Wong CM, Ying Q, Fan DN, Huang S, Ding J, et al. MicroRNA-125b suppressed human liver cancer cell proliferation and metastasis by directly targeting oncogene LIN28B2. Hepatology 2010;52:1731–40. Szabo G, Bala S. MicroRNAs in liver disease. Nat Rev Gastroenterol Hepatol 2013; 10:542–52. Bala S, Szabo G. MicroRNA signature in alcoholic liver disease. Int J Hepatol 2012; 2012:498232. Wang B, Majumder S, Nuovo G, Kutay H, Volinia S, Patel T, et al. Role of microRNA155 at early stages of hepatocarcinogenesis induced by choline-deficient and amino acid-defined diet in C57BL/6 mice. Hepatology 2009;50:1152–61. Zhang Y, Wei W, Cheng N, Wang K, Li B, Jiang X, et al. Hepatitis C virus-induced upregulation of microRNA-155 promotes hepatocarcinogenesis by activating Wnt signaling. Hepatology 2012;56:1631–40. Cai G, Liu Y. Diagnosing advanced versus early-stage hepatocellular carcinoma. J Clin Oncol 2012;30:2167 [author reply 2168]. van der Meer AJ, Farid WR, Sonneveld MJ, de Ruiter PE, Boonstra A, van Vuuren AJ, et al. Sensitive detection of hepatocellular injury in chronic hepatitis C patients with circulating hepatocyte-derived microRNA-122. J Viral Hepat 2013;20:158–66. Budhu A, Jia HL, Forgues M, Liu CG, Goldstein D, Lam A, et al. Identification of metastasis-related microRNAs in hepatocellular carcinoma. Hepatology 2008;47: 897–907. Hoffmann TW, Duverlie G, Bengrine A. MicroRNAs and hepatitis C virus: toward the end of miR-122 supremacy. Virol J 2012;9:109. Murakami Y, Yasuda T, Saigo K, Urashima T, Toyoda H, Okanoue T, et al. Comprehensive analysis of microRNA expression patterns in hepatocellular carcinoma and non-tumorous tissues. Oncogene 2006;25:2537–45. Bala S, Marcos M, Szabo G. Emerging role of microRNAs in liver diseases. World J Gastroenterol 2009;15:5633–40.

P

[124]

[135] [136]

T

[100] Sagir A, Erhardt A, Schmitt M, Haussinger D. Transient elastography is unreliable for detection of cirrhosis in patients with acute liver damage. Hepatology 2008; 47:592–5. [101] de Ledinghen V, Vergniol J, Capdepont M, Chermak F, Hiriart JB, Cassinotto C, et al. Controlled attenuation parameter (CAP) for the diagnosis of steatosis: a prospective study of 5323 examinations. J Hepatol 2014;60:1026–31. [102] Sasso M, Beaugrand M, de Ledinghen V, Douvin C, Marcellin P, Poupon R, et al. Controlled attenuation parameter (CAP): a novel VCTE guided ultrasonic attenuation measurement for the evaluation of hepatic steatosis: preliminary study and validation in a cohort of patients with chronic liver disease from various causes. Ultrasound Med Biol 2010;36:1825–35. [103] Chon YE, Jung KS, Kim SU, Park JY, Park YN, Kim do Y. Controlled attenuation parameter (CAP) for detection of hepatic steatosis in patients with chronic liver diseases: a prospective study of a native Korean population. Liver Int 2014;34: 102–9. [104] Lencioni R. Evolving strategies in the diagnosis of hepatocellular carcinoma. J Hepatol 2011;54:184–6. [105] Bremner KE, Bayoumi AM, Sherman M, Krahn MD. Management of solitary 1 cm to 2 cm liver nodules in patients with compensated cirrhosis: a decision analysis. Can J Gastroenterol 2007;21:491–500. [106] Sangiovanni A, Manini MA, Iavarone M, Romeo R, Forzenigo LV, Fraquelli M, et al. The diagnostic and economic impact of contrast imaging techniques in the diagnosis of small hepatocellular carcinoma in cirrhosis. Gut 2010;59:638–44. [107] Khalili K, Kim TK, Jang HJ, Yazdi LK, Guindi M, Sherman M. Indeterminate 1–2-cm nodules found on hepatocellular carcinoma surveillance: biopsy for all, some, or none? Hepatology 2011;54:2048–54. [108] Xu DZ, Wang XY, Shen XL, Gong GZ, Ren H, Guo LM, et al. Results of a phase III clinical trial with an HBsAg–HBIG immunogenic complex therapeutic vaccine for chronic hepatitis B patients: experiences and findings. J Hepatol 2013;59:450–6. [109] Lai CL, Ahn SH, Lee KS, Um SH, Cho M, Yoon SK, et al. Phase IIb multicentred randomised trial of besifovir (LB80380) versus entecavir in Asian patients with chronic hepatitis B. Gut 2013. [110] Betancourt AA, Delgado CA, Estevez ZC, Martinez JC, Rios GV, Aureoles-Rosello SR, et al. Phase I clinical trial in healthy adults of a nasal vaccine candidate containing recombinant hepatitis B surface and core antigens. Int J Infect Dis 2007;11: 394–401. [111] Sulkowski MS, Gardiner DF, Rodriguez-Torres M, Reddy KR, Hassanein T, Jacobson I, et al. Daclatasvir plus sofosbuvir for previously treated or untreated chronic HCV infection. N Engl J Med 2014;370:211–21. [112] Forns X, Lawitz E, Zeuzem S, Gane E, Bronowicki JP, Andreone P, et al. Simeprevir with peginterferon and ribavirin leads to high rates of SVR in patients with HCV genotype 1 who relapsed after previous therapy: a phase 3 trial. Gastroenterology 2014. [113] Di Bisceglie AM, Janczweska-Kazek E, Habersetzer F, Mazur W, Stanciu C, Carreno V, et al. Efficacy of immunotherapy with TG4040, peg-interferon, and ribavirin in a phase 2 study of patients with chronic HCV infection. Gastroenterology 2014. [114] Sanyal Arun J, Sunder M, Henry Robert, Marschall Hanns-Ulrich, Morrow Linda, Sciacca Cathi I, et al. A new therapy for nonalcoholic fatty liver disease and diabetes? INT‐747 — the first FXR hepatic therapeutic study. AASLD 2009;50(S4):A183. [115] Polyzos SA, Kountouras J, Zafeiriadou E, Patsiaoura K, Katsiki E, Deretzi G, et al. Effect of spironolactone and vitamin E on serum metabolic parameters and insulin resistance in patients with nonalcoholic fatty liver disease. J Renin Angiotensin Aldosterone Syst 2011;12:498–503. [116] Beaton MD, Chakrabarti S, Levstik M, Speechley M, Marotta P, Adams P. Phase II clinical trial of phlebotomy for non-alcoholic fatty liver disease. Aliment Pharmacol Ther 2013;37:720–9. [117] Mudaliar S, Henry RR, Sanyal AJ, Morrow L, Marschall HU, Kipnes M, et al. Efficacy and safety of the farnesoid X receptor agonist obeticholic acid in patients with type 2 diabetes and nonalcoholic fatty liver disease. Gastroenterology 2013;145:574–82 [e571]. [118] Takeshita Y, Takamura T, Honda M, Kita Y, Zen Y, Kato KI, et al. The effects of ezetimibe on non-alcoholic fatty liver disease and glucose metabolism: a randomised controlled trial. Diabetologia 2014. [119] Dohil R, Schmeltzer S, Cabrera BL, Wang T, Durelle J, Duke KB, et al. Enteric-coated cysteamine for the treatment of paediatric non-alcoholic fatty liver disease. Aliment Pharmacol Ther 2011;33:1036–44. [120] Boettcher E, Csako G, Pucino F, Wesley R, Loomba R. Meta-analysis: pioglitazone improves liver histology and fibrosis in patients with non-alcoholic steatohepatitis. Aliment Pharmacol Ther 2012;35:66–75. [121] Armstrong MJ, Barton D, Gaunt P, Hull D, Guo K, Stocken D, et al. Liraglutide efficacy and action in non-alcoholic steatohepatitis (LEAN): study protocol for a phase II multicentre, double-blinded, randomised, controlled trial. BMJ Open 2013;3: e003995. [122] Nguyen-Khac E, Thevenot T, Piquet MA, Benferhat S, Goria O, Chatelain D, et al. Glucocorticoids plus N-acetylcysteine in severe alcoholic hepatitis. N Engl J Med 2011;365:1781–9. [123] Ann Sing Lee BCYZ, Yiu Cheung Foon, Kwong Philip, Chi Kin Cheng Ashley, Lai Maria, Kwok Chloe, et al. Randomized phase II study of the x-linked inhibitor of

U

922 923 924 925 926 927 928 929 930 931 932 933 Q12 934 935 936 937 938 939 940 941 942 943 944 945 946 947 948 949 950 951 952 953 Q13 954 955 956 957 958 959 960 961 962 963 964 Q14 965 966 967 Q15 968 969 970 971 972 973 974 975 976 977 978 979 980 981 982 983 984 Q16 985 986 987 988 989 990 991 992 993 994 995 996 997 998 999

R. Chatterjee, A. Mitra / International Immunopharmacology xxx (2014) xxx–xxx

E

10

[137]

[138] [139]

[140] [141] [142]

[143]

[144] [145]

[146]

[147] [148]

[149]

Please cite this article as: Chatterjee R, Mitra A, An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer, Int Immunopharmacol (2014), http://dx.doi.org/10.1016/j.intimp.2014.12.024

1000 1001 1002 1003 1004 1005 1006 1007 1008 1009 1010 1011 1012 1013 1014 1015 1016 1017 1018 1019 1020 1021 1022 1023 1024 1025 1026 1027 1028 1029 1030 1031 1032 1033 1034 1035 1036 1037 1038 1039 1040 1041 1042 1043 1044 1045 1046 1047 1048 1049 1050 1051 1052 1053 1054 1055 1056 1057 1058 1059 1060 1061 1062 1063 1064 1065 1066 1067 1068 1069 1070 1071 1072 1073 1074

An overview of effective therapies and recent advances in biomarkers for chronic liver diseases and associated liver cancer.

Chronic liver diseases (CLDs) such as hepatitis, alcoholic liver disease, nonalcoholic fatty liver, and their downstream effect cancer affect more tha...
1MB Sizes 0 Downloads 8 Views