Topic Highlight Open Access
Copyright ©2008 The WJG Press and Baishideng. All rights reserved.
World J Gastroenterol. Mar 21, 2008; 14(11): 1720-1733
Published online Mar 21, 2008. doi: 10.3748/wjg.14.1720
Tumor suppressor and hepatocellular carcinoma
Juliette Martin, Jean-François Dufour, Department of Clinical Pharmacology, Bern, CH-3010, Switzerland
Correspondence to: Juliette Martin, PhD, Department of Clinical Pharmacology, University of Bern, Murtenstrasse, 35, CH-3010 Bern, Switzerland. juliette.martin@ikp.unibe.ch
Telephone: +41-31-6322595
Fax: +41-31-6324997
Received: July 20, 2007
Revised: November 14, 2007
Published online: March 21, 2008

Abstract

A few signaling pathways are driving the growth of hepatocellular carcinoma. Each of these pathways possesses negative regulators. These enzymes, which normally suppress unchecked cell proliferation, are circumvented in the oncogenic process, either the over-activity of oncogenes is sufficient to annihilate the activity of tumor suppressors or tumor suppressors have been rendered ineffective. The loss of several key tumor suppressors has been described in hepatocellular carcinoma. Here, we systematically review the evidence implicating tumor suppressors in the development of hepatocellular carcinoma.

Key Words: Tumor suppressor, Hepatocellular carcinoma, Deregulation, Liver, Carcinogenesis



INTRODUCTION

Hepatocellular carcinoma (HCC) is the most common form of primary hepatic tumor and is one of the most common cancers worldwide. HCC usually develops in patients with cirrhosis. Cirrhosis may be caused by viral hepatitis (primarily hepatitis B and C), alcohol, hereditary haemochromatosis, autoimmune liver diseases and actually any disease that results in chronic inflammation of the liver. In order to understand how chronic inflammation and cirrhosis lead to initiation and progression of HCC extensive research on molecular events has been undertaken. Several intracellular signaling pathways have been closely associated with HCC: p53 pathway and DNA alteration, retinoblastoma (Rb) pathway and regulation of cell cycle, transforming growth factor-beta (TGF-β) and inhibition of cellular growth, and Wnt/beta-catenin pathway and cellular adhesion and signal transduction[1]. Deregulation in these different pathways favors the development of liver tumor. Conceptually, hepatocarcinogenesis is based on two main principles: (1) the activation of genes such as oncogenes (c-myc, β-catenin), growth factor (IGF-II, TGF-α) and telomerase enzyme inducing cellular immortalization and (2) the inactivation of genes called tumor suppressor genes (for example p53 and Rb) Their expressions can be affected by different modifications such as promoter methylation, mutations, biallelic loss and loss of heterozygosity (LOH). In liver cancer, chromosomal aberrations are frequently observed on chromosome 1, 4, 5, 6, 8, 9, 10, 11, 13, 16, 17, 20 and 22 sharing the complexity of hepatocarcinogenesis. Moreover, if some mutations are associated with initiation of carcinogenesis, other genetic alterations promote progression and clone divergence in tumors[23].

This review addresses the various tumor suppressors, which have been implicated in HCC (Table 1). Specifically, we discuss their function and involvement in the initiation or progression of HCC as well as their mechanisms of inactivation.

Table 1 Summary of tumor suppressor regulated in HCC.
SymbolNameLocationHaploinsufficiencyDownregulation in HCCMechanism of regulation
Etiology
MutationMethylationChromosome changeProtein overcrossing
p53[312142123293133180]Tumor protein p5317p13.1YesYesAFB1
HBV
HCV
Cirrhosis
p21[5153]Cyclin-dependent kinase inhibitor 1A6p21.2YesYesp53HBV
HCV
p27[60]Cyclin-dependent kinase inhibitor 1B12p13.1-12pYesYescirrhosis
p16[386668]Cyclin-dependent kinase inhibitor 2A9p21YesAFB1
p14ARF[38]Cyclin-dependent kinase inhibitor 2A9p21YesHCV
Cirrhosis
E-cadherin[7073]E-cadherin16q22.1YesHBV
HCV
Axin 1[74767879]Axis inhibitor 116p13.3YesHBV
HCV
Axin 2[757879]Axis inhibitor 217q23-q24Yes
APC[8184]Adenomatosis polyposis coli5q21-q22YesHBV
HCV
SOCS1[8990]Suppressor of cytokine signaling 116p13.13YesHCV
Cirrhosis
SOCS3[9394]Suppressor of cytokine signaling 317q25.3Yes
RASSF1A[83104105]Ras association (ralGDS/AF-6) domain family 13p21.3YesAFB1
HBV
HCV
NORE1[94108]Ras association (ralGDS/AF-6) domain family 51q32.1YesHBV
HCV
Cirrhosis
KLF6[116119]Kruppel-like factor 610p15YesHBV
HCV
Cirrhosis
PTEN[123127128130]Phosphatase and tensin homolog10q23.3YesYesmTORAFB1
HBV
Hint1Histidine triad nucleotide binding protein5q31.2Yesn.dn.dn.dn.dn.d
Hint2[144]Histidine triad nucleotide binding protein 29p13.3Yesn.dn.dn.dn.d
FHIT[149151154157]Fragile histidine triad gene3p14.2YesYesHBV
HCV
Cirrhosis
WWOX[160]WW domain containing oxidoreductase16q23.3-q24YesAFB1
PARK2[173]Parkin6q25.2-q27Yes
p53 AND ITS PATHWAY
p53

TP 53 gene, located on chromosome 17p13.1, encodes a 53 kDa DNA-binding transcription factor. The protein p53 is implicated in the control of cell cycle, apoptosis, DNA repair and angiogenesis. p53 activation has been related to various cellular and environmental changes including: (1) DNA damage (induced by UV light, gamma rays, X rays, and inhibition of topoisomerases), (2) cellular stress (hypoxia, disruption of cell adhesion) independent of DNA damage and (3) activation of growth signaling pathways[45]. p53 gene is a haploinsufficient tumor suppressor gene[67]. The loss of p53 activity has been described in many types of human tumors, particularly in 30%-60% of HCC. In many cases, these alterations contribute to progression and not to initiation of HCC[8]. In terms of prognosis, p53 alterations are generally associated with larger, less differentiated tumors and poor survival[9]. Recently, Lowe and its team observed that senescence program in correlation with the innate immune system turn off the tumor development after restoration of p53 expression in liver tumor cells[10]. Different studies described that p53 is regulated by methylation of its CpG islands in HCC[1112] but its expression is mainly regulated by genetic mutations. Mutations affecting p53 are diverse by their nature and position. The p53 gene is altered by allelic deletion and punctual mutation concentrated between exons 4 and 9 of the coding region containing the DNA binding domain. Amongst these mutations, the transversion in codon 249 (G→T), which causes an arginine to serine (R→S) substitution is present in 50% of HCCs. This genetic alteration can be a consequence of exposure to aflatoxin B1 (AFB1) which is a mycotoxin found in contaminated foods (like corn, rice, and peanuts) particularly in African and Asian countries[1314]. Kirk have proposed the use of p53 mutant DNA as a biomarker for AFB1 exposure[15]. Mutated R249S p53 protein expression may induce (1) an inhibition of apoptosis[16], (2) inhibition of p53 mediated transcription[17] and (3) stimulation of liver cell growth in vitro[1820]. Like AFB1, the hepatitis B virus (HBV) affects the activity of p53 by inducing DNA damage and mutating the p53 gene. Synergism between AFB1 and HBV has been described[2122]. The X gene of HBV (HBx) encodes a protein of 154 amino acids, which is a viral transcriptional co-activator capable of activating the expression of several proteins such as oncogenes (c-myc, c-fos), cellular growth factors and cytokines. Protein X inactivates various functions of p53[323] such as apoptosis[172427] and transcriptional activation[28]. Concerning HCV, the non-structural protein NS2-5 seems to deregulate the actions of factors controlling hepatocellular proliferation by inhibiting p21WAF and sequestering p53[29]. In transgenic mice, HCV core protein has been found in vivo to stimulate the initiation and development of tumor with the same histological characteristics of human HCC[30]. HCV core protein may interact directly with p53 and p73[3133].

In HCC, p53 level and activity are modulated by different proteins such as MDM2 (murine double minute 2) and p14ARF (Alternative Reading Frame). Oncogenic MDM2 protein contains a p53-DNA binding site and induces p53 degradation by ubiquitination and proteolysis[3435]. However, auto-regulatory feedback loop of MDM2-p53 controls the function and expression of p53 and MDM2, respectively[36]. However, the activity of MDM2 is inhibited by ARF tumor suppressor protein[37]. This protein is encoded by INK4a/ARF locus on chromosome 9p21, which is frequently affected by hypermethylation and mutations in HCC[38]. In the clinical treatment strategy of HCC, the reactivation of p53 protein is focused on stabilization of p53 activity, over-expression of ARF protein and inactivation of MDM2-p53 interaction[39]. In animals, different small molecules such as nutlins and PRIMA-1 inactivate MDM2 and increase p53 activity[4041]. Loss of cell cycle check-point control by mutation of p53 has been suggested to stimulate the metastasic potential of HCC via over-expression of L2DTL[42]. Ubiquitination and protein stability of p53 are regulated by a complex regrouping L2DLT, CDT2 and PCNA[43]. In contrast, TIP30, which is known to inhibit cell proliferation and tumorigenesis, may act as a hepatocarcinogenetic tumour suppressor[44]. This protein diminishes Bcl2/Bcl-x expression and augments p53 expression[45]. TIP30 mutants inhibited the expression of tumor suppressor genes such as p53 and E-cadherin whereas they induced positive regulation of oncogenic genes expression such as N-cadherin and c-Myc[46]. Immunohistochemical analysis of HCC and normal liver showed that a 33 kDa protein called ING1 (inhibitor of growth-1, p33 (ING1)) is expressed at a lower level in HCC. Lower ING1 protein level was associated with enhanced cyclin E kinase activity[47]. Recently, Zhu and co-workers proposed that p33 (ING1b) and p53 work in tandem to enhance apoptosis, cell cycle arrest and to inhibit cell growth in HCC. Combined expression of p33 and p53 augments p21 (WAF1/CIP1) protein causing an arrest of cell cycle at stage G0/G1 and enhancing apoptosis[48]. The p53 pathway is intercrossing with other tumor suppressors as p21, p27 and p16, which are described below.

p21

p53 directly controls a gene encoding for a protein named p21WAF/CIP1 (p21). This protein, whose gene is located on chromosome 6p21.2, acts as haploinsufficient tumor suppressor gene[49]. It functions as an inhibitor of cyclin-dependent kinases (cyclin-CDK2, CDK4) and interacts with proliferating cell nuclear antigen (PCNA), a DNA polymerase accessory factor. This protein has a regulatory role in the cell cycle, S phase DNA replication and DNA damage repair. p21 has also been described to activate the caspase 3 protein and thus induce apoptosis. A reduction of p21 expression was observed in HCC[50]. Moreover, tumor progression and poor outcome of HCC were associated at disruption of p53-p21/WAF1 cell cycle pathways[51]. In a study addressing p53 expression and apoptosis, high p53 expression was associated with cell cycle arrest and apoptosis whilst a lower level of p53 induced only cell cycle arrest. However, p21 expression could activate only cycle arrest but not apoptosis in HCC as well as in the presence of high p53 as low p53 expression[52]. The transcription of p21 gene was found to be repressed by HBV X protein (HBx) and HCV core protein[53]. In addition, the stress due to the inflammation and fibrosis of HCV-associated chronic liver diseases induced up-regulation of p21[54]. Huether and co-workers analyzed the effect of cetuximab (Erbitux), a chimeric human/mouse antibody directed against the Epidermal Growth Factor Receptor (EGFR), with or not in combination with other drugs on human hepatocellular carcinoma cell lines. The expression of the cyclin-dependent kinase inhibitors p21 and p27 (Kip1) was increased whereas the expression of cyclin D1 was decreased by cetuximab treatment. A synergistic antiproliferative effect was observed following a treatment with cetuximab combined with doxorubicin, tyrosine kinase inhibitors (erlotinib or AG1024) or the HMG-CoA-reductase inhibitor fluvastatin[55]. The expression of different proteins such as p53, p21 cyclin D implicated in enhancement of the apoptosis pathway and cell proliferation have been analyzed after treatment with antiangiogenic agent TNP-470 in a rat model of hepatocellular carcinoma. The augmentation in these angiogenic factors induced by HCC was prevented whereas a cell-cycle inhibition was generated by activation of p21 and reduction of the cyclin D-Cdk4 and cyclin E-Cdk 2 expression following animals’ treatment with TNP-470. These results suggest that TNP-470 may be efficient for anti-angiogenic therapy and treatment of human HCC[56].

p27

The p27 (p27 Kip1), whose gene, a member of haploinsufficient tumor suppressor gene family[57], locates on chromosome 12p13.1-p12, is a cyclin-dependent kinase inhibitor. Cell cycle progression at G1 and cyclin E-CDK2 and cyclin D-CDK4 complexes are regulated by p27. Different studies have been executed to understand the role of p27 in development of tumor and particularly in HCC. The comparison of hepatocellular HCC with adjacent non-tumoral and normal liver tissues found that the weak expression of p27 was strongly associated with infiltration, metastasis and poor prognosis in patients affected by HCC. Moreover, cytoplasmic sequestration of p27 was observed more in HCC leading a diminution of p27 expression and was particularly characterized in early steps of hepatocarcinoma development[58]. Philipp-Staheli and co-workers have already suggested that the p27 protein might be a check point for tumor suppression and an important prognostic marker because its loss favors tumor growth[59]. Expression of p27 was lower in patients having liver cirrhosis and HCC in comparison to those without HCC level[60]. This group explained the loss of p27 expression by high level of methylation of its promoter. The high and low levels of p27 expression have been associated with prolonged survival and poor prognosis, respectively[6162]. A study proposed that functional inactivation of p27 is strongly associated with methylation of p16 and loss of its expression[63].

p16

The p16 (P16INK4) gene, located on chromosome 9p21, encodes a protein that inhibits proliferation of normal cells by binding strongly with cdk4 and cdk6. This binding prevents cdk4 and cdk6 from interacting with cyclin D and inactivation by phosphorylation of the retinoblastoma protein (Rb). In 1998, post-transcriptional regulation was found to inactivate the p16 activity in HCC and this inactivation appeared to take place during the early-stage of hepatocarcinogenesis[50]. The loss of expression of p16 and inactivation of Rb represent major events in hepatocarcinogenesis[64]. Analysis of different hepatocyte cell lines revealed that increased p16 expression is associated with decreased phosphorylation of pRB. Reciprocally, phosphorylation of Rb and increase of cell growth were associated with silencing of p16 by promoter methylation[65]. The p16 gene is silenced by hypermethylation of 5’CpG islands in its promoter[6667]. In addition to hypermethylation-induced transcriptional repression of the p16 gene, expression is also affected by mutation and deletion, although these modifications are less common[38]. As described above, aflatoxin inactivates p53. The concentration of aflatoxin B1 albumin adducts has been correlated not only with p53 mutation but also with p16 methylation stressing the importance of environmental factors in the development of HCC[68].

WNT PATHWAY

Many different proteins have been described in the Wnt pathway which function either as oncogenes (e.g beta-catenin) or as tumor suppressors (e.g E-cadherin, APC, Axin 1 and 2 proteins). About 50% of the HCC exhibited alteration of the Wnt pathway[69].

E-cadherin

E-cadherin protein, encoded by a gene located on chromosome 16q22.1, belongs to the cadherin superfamily and is a calcium dependent cell-cell adhesion glycoprotein. Loss of function induced by mutations was associated with proliferation, invasion and metastasis. Many investigations with animal models and human HCC tissues have been performed and show that E-cadherin gene expression is regulated by promoter methylation. Hypermethylation was associated with decreased E-cadherin expression but also with microvascular invasion and recurrence of HCC[70,71]. HBV and HCV affect this pathway. The presence of the HBx protein was associated with hypermethylation of the E-cadherin promoter, loss of its expression and beta-catenin accumulation[72]. In HCV-associated HCC the probability of recurrence could be linked to depressed E-cadherin expression[73].

Axin1/ axin2

Axin1 and axin2 proteins, encoded by genes located on chromosome 16p13.3 and 17q23-q24, respectively, act as negative regulators of the Wnt signaling pathway and can induce apoptosis. Axins interact with different proteins such as beta-catenin, adenomatosis polyposis coli (APC) and glycogen synthase kinase 3-beta. Mutations in the Axin1 gene have been associated with different human cancers including HCC and hepatoblastomas. Satoh and co-workers identified mutations in the axin1 gene as well as in cell lines as in HCCs and they reported that wild type protein stimulates apoptosis leading to suppression of tumor growth. The gene coding for Axin1 protein is affected by loss of heterozygosity and small deletions, mutations and by missense mutations (1 bp deletion and 12 bp insertions)[74] which most of the time target the gene in a biallelic manner[75]. Combination of loss of heterozygosity and mutation induced the inactivation of Axin1 in HCC[76]. The loss of heterozygosity also frequently affects the Axin2 gene due to it chromosomal localization and has been associated with different tumors like breast cancer and neuroblastoma[77]. The percentage of Axin1 and Axin2 mutations in HCC remains controversial. In presence of Axin mutations, the Wnt-signaling pathway is the altered leading to the accumulation of beta-catenin[7880]. Beta-catenin mutations were associated with over-expression of G-protein-coupled receptor (GPR) 49, glutamate transporter (GLT)-1 and glutamine synthetase (GS) while this correlation was not found with Axin1 mutations. These results suggest that Axin1 function might affect other signaling pathway than Wnt pathway[76].

APC

A gene located on chromosome 5q21-q22 encodes a tumor suppressor protein named adenomatosis polyposis coli (ACP). This protein has many intracellular functions including nuclear export and degradation of beta-catenin. A small region designated the mutation cluster region regroup mutations associated to diseases. To determine the role of APC protein in hepatocyte carcinogenesis, Colnot and co-workers generated a knock-out mouse model targeting exon 14 of the APC gene. They observed that 67% of analyzed mice developed HCC while Wnt/beta-catenin pathway activation was demonstrated by accumulation of different genes regulated by beta-catenin (leukocyte cell-derived chemotaxin 2, ornithine aminotransferase and glutamine synthetase, glutamate transporter 1)[81]. The disruption of APC gene in liver induces hepatocyte hyperplasia, marked hepatomegaly and rapid mortality. Like other tumor suppressor genes, APC is hypermethylated in HCC relative to non-tumor liver. Hepatitis C virus/hepatitis B virus-negative HCC showed less methylation on APC gene than in hepatitis C virus-positive HCC[82]. Methylation status of APC and other tumor suppressor genes was associated with the epigenetic instability dependent HCCs[83]. Recently, bi-allelic inactivation and nonsense mutation at codon 682 of APC gene in sporadic nodule-in-nodule-type HCC were observed using high-density array-based comparative genomic hybridization (aCGH) and direct sequencing, respectively. Both alterations lead to inactivation of APC binding to beta-catenin which may enhance the evolution of sporadic HCC[84].

RAS/JAK/STAT PATHWAY
SOCS

The suppressor of cytokine signaling (SOCS), also known as STAT-induced STAT inhibitor (SSI) protein family comprises several members including SOCS1, SOCS2 and SOCS3 which are encodes by genes located in 16p13.13, 12q, 17q25.3, respectively[8586]. These proteins function as negative regulators of cytokine signaling. SOCS1 as well as SOCS2 and SOCS3 are stimulated by cytokines and act in negative feedback loop to regulate cytokine signaling. SOCSs inhibit by direct binding the kinase activity of Janus Kinases (JAKs) proteins and so block the JAK/STAT pathway[87]. The role of SOCS1 in negative regulation of interferon-gamma and in T-cell differentiation was determined by generating a knock-out mouse model lacking the expression of SOCS1 gene. In the same way, SCOS2 and SOCS3 were demonstrated to be involved in regulation of postnatal growth and regulation of fetal liver hematopoiesis, respectively[87]. The SOCS1 gene promoter is enriched with CpG dinucleotides[88]. The decrease of SOCS1 expression was associated with aberrant methylation in CpG islands and 5’non-coding region of SOCS1 promoter and loss of heterozygosity[8990]. A significant relationship between SOCS1 methylation level and HCC transformation of cirrhotic nodules was established confirming that SOCS might act as a tumor suppressor[91]. The suppression of cell growth and activation of JAK2 was observed after recovering of SOCS1 expression by gene therapy in cells having SOCS1 silenced by hypermethylation[92]. SOCS3 promoter can also be methylated resulting in diminution of SOCS3 expression[9394]. Restoration of its expression blocks the phosphorylation of STAT3 and inhibits the proliferation. Cell growth and migration were negatively regulated by inhibition of JAK/STAT signaling pathway by SOCS3 protein in HCC[93]. The enhancement of proliferation and development of hepatic tumor, activation of STAT3 and inhibition of apoptosis were observed in absence of SOCS3 expression obtained using a conditional knockout mice approach under carcinogenic condition. These results confirmed that SOCS3 acts as tumor suppressor gene[95]. Leong and co-workers found that SOCS2 and SOCS3 were both up-regulated in hepatic cells following estrogen treatment via estrogen receptor (ER) alpha providing a mechanistic explanation for the rare cases of HCC responding to this treatment[96].

RASSF1A/NORE1

Members of the RAS superfamily are plasma membrane GTP binding proteins that modulate intracellular signal transduction pathways. A subgroup in this family contains a Ras-association domain and takes part in RAS signaling pathway[97101]. A gene located on chromosome 3p21.3 encodes a protein identified as human RAS effector homologue (RASSF1). Alternative splicing and promoter usage of this gene generates three different transcripts: RASFF1A, RASSF1B and RASSF1C. RASSF1A contains a Ras Association Domain (RA) and binds Ras in a GTP-dependent manner to affect apoptosis. The presence of CpG-islands in the promoter of the RASSF1 gene was associated with high methylation level and the loss of expression of RASSF1A. The reduction of RASSF1A expression was found in lung carcinoma and these results indicate that RASSF1A might act as a tumor suppressor[102]. Additionally, RASSF1A role is associated with the cell cycle given that it blocks the accumulation of cyclin D1 and G (1)/S-phase cell cycle progression[103]. Hypermethylation of RASSF1A induced the inactivation of RASSF1A but also correlated with environmental carcinogens such as AFB (1) and with inactivation of p16INK4a protein in hepatocellular carcinoma[104]. The high frequency of hypermethylation of RASSF1A promoter could be detected not only in tumor[83105] but also in matched plasma of patients affected by HCC[105]. Methylation level of RASSF1A was proposed as a potential marker to diagnose HCC[100106107]. Recently, a strong association was identified between CpG island methylation phenotype (CIMP) and serum concentration of alpha-fetoprotein (AFP) and inactivation of many genes involved in process of tumor suppression such as RASSF1A. Thus CIMP might be use as molecular marker of late-stage HCC development[12].

Among the subfamily of RAS effectors, the NORE1 proteins encoded by a gene located on chromosome 1q32.1 were described as a regulator of Ras dependent apoptosis. In the same manner as the RASSF1 gene, three different isoforms were identified (NORE1Aalpha, NORE1Abeta and NORE1B). NORE1A and NORE1B, which are separated by CpG islands spanning their first exons, share the Ras-association (RA) domain but the diacylglycerol (DAG) binding domain was only present on NORE1A[99]. Analysis of methylation found that NORE1A promoter was methylated whereas NORE1B was unmethylated in breast, colorectal and kidney tumor cell lines[9799]. Comparative analysis showed that NORE1 gene was not altered by methylation whereas RASSF1A gene promoter was increasingly methylated from regenerating liver to hepatocellular carcinoma nodules[107]. In another study, expression of NORE1A and SOCS 3 was inhibited by high methylation level of their promoters and their inactivation associated with a subclass of poor prognosis HCC[94]. It appears that NORE1 could be considered as tumor suppressor gene in hepatocarcinogenesis[108].

OTHER PATHWAYS
KLF6

Krüppel-like factors (KLFs) are highly related zinc-finger proteins that are important components of the eukaryotic cellular transcriptional machinery. Among this protein family, Krüppel-like factor 6 (KLF6) is a ubiquitously expressed zinc finger transcription factor encoded by a gene located on chromosome 10p15[109]. The KLF6 protein binds DNA on guanine-rich core promoter elements and regulates the transcriptional activation process via its zinc fingers domains and central acidic N-terminal domain. Due to its regulatory role in transcription, different studies have been conducted to define the role of KLF6 in tumor development and has been described as a tumor suppressor in different cancer such as colon and prostate[110112]. Reduction of KLF6 expression and methylation of KLF6 promoter have been associated with human cancers[111113115] but the incidence of KLF6 variation on HCC has been described for first time by Friedman’s group[116]. They determined that loss of heterozygosity induces a loss of KLF6 expression in 50% of analysed HCC samples. In same samples, they identified several missense mutation associated with presence of HBV or HCV. Moreover, in vitro analyses of KLF6 mutations showed that only KLF6 wild type inhibits the cell growth of HepG2 cell lines by p21 protein activation whereas the different mutants did not induce any changes indicating regulatory effects of mutations on KLF6 activity. The process leading to the activation of p21 and reduction of cell proliferation by KLF6 necessitates the acetylation of KLF6 by histone acetyltransferase activity of either cyclic AMP-responsive element binding protein-binding protein or p300/CBP-associated factor. This process can be abrogated by a single mutation of lysine-to-arginine (K209R), point mutation already described in prostate cancer[117]. In recent work carried out on ovarian cancer, E-cadherin level variation was found to be mediated by direct action of KLF6 on its promoter[118]. Furthermore, induction of cellular differentiation and inhibition of cell proliferation was observed in KLF6 overexpressing HepG2 cell lines and associated with augmentation of E-cadherin and albumin expression and reduced cyclinD1 and beta-catenin expression. In the same study, the authors demonstrated also inhibitory effects of HBV and HCV infection on KLF6 expression[119]. The stability of KLF6 is modified by ubiquitination after induction of apoptosis by drugs (cisplatin, adriamycin) or UVB irradiation but not by apoptotic-dependent extrinsic/death-receptor pathway. The effects of KLF6 on tumor suppression and enhancement of chemotherapeutics response might be affected by the speed of its degradation and deregulation of its stability[120]. Apoptosis induced by upregulation of p53 and diminution of Bcl-xL expression was enhanced following KLF6 knockdown. Additionally, the arrest of cell cycle in G1 phase and expression of cyclin-dependent kinase 4 and cyclin D1 were weakened or suppressed in KLF6 silenced cells. These results bring a new perspective for the link between of KLF6 and apoptosis[121].

PTEN

This haploinsufficient gene located on chromosome 10q23.3 encodes a phosphatidylinositol-3,4,5-trisphosphate 3-phosphatase[122]. This protein dephosphorylates phosphoinositide substrates and acts as a negative regulator for intracellular levels of phosphatidylinositol-3, 4, 5-trisphosphate. PTEN is inactivated in many cases of breast, endometrial, prostate cancers. Many mutations affect PTEN gene such as missense mutations in exon 5 (K144I) and exon 7 (V255A) or silent mutations in exon 5 (P96P) in HCC from Taiwan[123]. In HBV infected liver cells, PTEN expression is also deleted following the genome integration of HBV[124]. In this Asiatic subset of HCCs, both mRNA and protein levels of PTEN were weaker in tumor than in paired para-carcinoma tissues[125126]. Moreover, a point of mutation induced a weak level of PTEN which is inversely linked with FAK phosphorylation[127]. In 47% of HCC analyzed by Sieghart and co-workers, they observed a decrease or absent of PTEN which is inversely correlated with expression of phosphorylation proteins in mTOR pathway[128]. HCC development might be susceptible to inhibition of mTOR[129]. The lessening of PTEN expression was associated with loss of promoter activity[130]. Wang and collaborators confirmed that PTEN inactivation seems resulting not only of mutation and promoter methylation but also possible other epigenetic mechanisms which remain to be defined[131]. Reduction of PTEN expression in HCC was associated with poor prognosis and progression of HCC[132]. This might be due to the inverse correlation between PTEN expression and VEGF expression[133].

HINT protein

Hint1: Hint1 is encoded by gene located on chromosome 5q31.2 and is member of the The Histidine Triad (HIT) protein family characterized the His-X-His-X-His-XX motif. Hint1 forms homodimers and each subunit can bind a nucleotide. Hint1 acts as an adenosine 5’-monophosphoramidate (AMP-NH2) hydrolase[134]. Spontaneous immortalization and enhancement of cell growth were observed in cells lacking Hint1. Squamous tumors (both papillomas and carcinomas) of the forestomach developed after treatment with chemical carcinogen N-nitrosomethylbenzylamine (NMBA) showed a greater volume and a more severe degree of malignancy in Hint1-/- mice[135]. Accordingly, the hypothesis of Hint1 role as tumor suppressor was developed and work was carried out to elucidate which signaling pathway is involved. Weiske and co-workers determined an interaction between Hint1 and pontin and reptin[136]. Pontin and reptin are often found in complexes and they possessed single-stranded DNA-stimulated ATPase and ATP-dependent DNA helicase activity but with opposite action[137138]. Moreover, both proteins interact with beta-catenin by modulating its transcriptional activity in Wnt pathway[139]. Hint1 was identified as a part of the LEF-beta-catenin transcription complex and function a as negative regulator of TCF-beta-catenin transcriptional activity, repressing expression of Wnt signaling target genes such as axin2 and cyclin D1[136]. The same authors reported that p53-induced apoptosis was influenced by Hint1, which up-regulated Bax and down-regulated Bcl-2. The pro-apoptotic activity of Hint1 was not related to its enzymatic AMP-NH2 hydrolase activity[140]. Treatment of non small cell lung cancer (NSCLC) cell line with DNA demethylating agent, 5-aza-2’-deoxycytidine up-regulated of Hint1, and this correlated with growth inhibition and reduction of tumorigenicity[141]. Weinstein’s group described Hint1 is a novel haploinsufficient tumor suppressor gene and is able to repress the cell growth and tumor progression by inhibition of AP-1 transcription factor activity in mammary tumor and colon cancer cells, respectively[142143]. The role of Hint1 in hepatocarcinogenesis remains to be explored.

Hint2: Hint2 is a mitochondrial HIT protein. This protein is encoded by a gene located on chromosome 9p13.3. Tissue profile expression of Hint2 showed that this protein is predominantly expressed in the liver and pancreas. Like its cytoplasmic homologue Hint1, Hint2 acts as an adenosine monophosphate hydrolase enzyme and this enzymatic activity was lost when the second histidine of the HIT motif is mutated. The sensitivity of cells to apoptosis was increased when Hint2 was over expressed whereas Hint2 knockdown was coincident with reduced caspase 3 expression. Subcutaneous injection of HepG2 cells over-expressing int2 in SCID mice resulted in maller tumours, which displayed more apoptosis in comparison to mice, injected with control HepG2 cells. Microarray analyses carried out on human tissues found a significant reduction of HINT 2 mRNA in HCC compared with surrounding liver tissue. This diminution of Hint2 expression was associated with a poor prognosis[144].

FRAGILE CHROMOSAL SITE GENES AND HCC
FHIT

The FHIT protein is encoded by a haploinsufficient tumor suppressor gene[145], located on chromosome 3p14.2 a place known to be one of the most fragile sites in human genome. This protein, a member of the histidine triad gene family, is a diadenosine triphosphate hydrolase. In numerous tumor types such as lung, stomach, breast, colon, aberrant forms of FHIT protein were found due to rearrangements and deletions in region of FHIT locus[146148]. Aberrant FHIT transcripts with deletions of exons and fusion of remaining exons and loss of heterozygosity were observed in HCC tissues in comparison with non-tumoral tissues and lead to the lack of FHIT protein expression. So FHIT was frequently altered in liver and might be implicated in liver tumorigenesis[149151]. FHIT has also been described as a pro-apoptotic agent. Restoration of Fhit expression activated caspase 8 and induced apoptosis in Fhit-negative cell lines[152]. Sard reported a 2-fold increase in the apoptosis-related protein Bak and in the cell cycle inhibitory protein p21, but not in Bcl-2, Bcl-X, Bax and, p53[153]. Inactivation like by promoter methylation was associated with progression and poor prognosis of HCC[154155]. In 2004, a study performed on HCC cohort from the US found over-expression of modified FHIT transcripts[156]. Fusion between exon 5 and 7 and between exon 7 and 9 were frequently observed in HCV-related[157]. Apoptosis was weaker in early HCC with negative FHIT expression than in advanced HCC with positive expression of FHIT. So, the absence of FHIT protein influencing the balance between apoptosis/proliferation may play a role in formation of HCC[158].

WWOX

WWOX (WW-domain containing oxidoreductase) gene located on another fragile site on chromosome 16q23.3-q24 encodes a 46 kDa protein having 2 WW domains, which mediate the protein-protein interaction, and a short-chain dehydrogenase/reductase domain (SRD). The WW domain-proteins are expressed in all eukaryotes and act as regulator of a wide variety of cellular functions such as RNA splicing, transcription and protein degradation. Different forms of WWOX transcripts were observed following deletions or alternative splicing in frameshift leading to the total or partial loss of different domains. Diverse missense mutations and single nucleotide polymorphisms (SNP) within the WWOX have been identified in many tumor cell lines and conduct to consider this protein as a tumor suppressor[159]. No mutations leading to WWOX abnormal transcript were found in HCC cell lines. However, loss of heterozygosity on chromosome 16q at the fragile site FRA16D was found in 29% of HCC and an association between 16q lack and R249S mutation affecting the p53 gene was determined in HCC samples from patients exposed to aflatoxin B1[160]. Decreased or absent expression of WWOX was observed in cell lines derived from human HCCs[161]. WWOX protein is able to interact with other proteins, in particular with p73 protein[162163]. The association of WWOX with p73 redistributes p73 from nucleus to cytoplasm blocking p73 transcriptional activity and enhancing the pro-apoptotic potential of WWOX[164]. A recent study performed on mouse models treated with carcinogen drugs found that the same down regulation profile for FHIT and WWOX in the liver[80]. This result completes the association between FHIT and WWOX expression observed in breast and gastric cancer[164167].

Parkin (PARK2)

Like WW domain-containing oxidoreductase gene (WWOX; 16q23) and fragile histidine triad gene (FHIT; 3p14.2) parkin is also located on fragile and unstable chromosomal region (FRA6), which can be targeted by mutational rearrangement (duplication or deletion). Parkin is a member of RBR protein family implicated in ubiquitin related-proteolytic pathway[168170]. This protein was involved in neurodegenerative diseases[171]. Parkin protects neurons and autosomal recessive juvenile parkinsonism is associated with mutations affecting this gene. Parkin seems to play a role in tumor suppression[172]. Analysis of HCC samples has been performed by Wang and collaborators. They determined that the expression of parkin was lower in HCC compared with normal liver. Transfection of Hep3B cells with parkin increased their sensitivity to apoptosis and negatively regulated their cell growth. These results prompt speculation that the absence of parkin expression may favor the development of HCC[173]. Down-regulation and loss of parkin expression was correlated with methylation of its promoter in acute lymphoblastic leukemia and chronic myelogenous leukemia (CML) in lymphoid blast crisis[174].

NEW POTENTIAL PATHWAYS IN TUMOR SUPPRESSION
DNA methyltransferase

A family of protein called DNA methyltransferase catalyses the methylation process on CpG islands leading to regulation of gene expression. Two distinct gene families: DNMT1 gene and DNMT3 gene family which regroups two related genes: DNMT3a and DNMT3b are known to maintain and induce the methylation. The function of DNMT2 is still to be clarified[175177]. High mRNA levels of DNMT1 and DNMT3 were observed in HCC in comparison of normal or non-tumoral tissues but only cytoplasmic DNMT3 expression was decreased in HCC. The first studies proposed that these proteins act during early stage of hepatocarcinogenesis and might take a part in progression of HCC[178180]. Excepted for a correlation between DNMT3a immuno-reactivity and p53 methylation levels, Park and co-worked did not observed a correlation between DNMT expression and methylation levels of different tumor suppressor genes described in HCC[11]. It seems that not only DNMTs protein but also other mechanisms are engaged in methylation changes of tumor suppressor genes in HCC. In liver cell line infected by HBV and HBV-infected HCC samples, HBVx proteins induced a change of transcription of DNMTs proteins leading to regional methylation of tumor suppressor genes[181]. The lack of two other proteins: O6-Methylguanine-DNA Methyltransferase (MGMT) and human Mut L homologue (hMLH) implicated in DNA repair system have been associated with poor prognosis and advanced stages of HCC[182]. The expression of MGMT in HCC is altered by hypermethylation of its promoter and DNA integration of HBV near to FRA10F chromosome fragile site[183184].

miroRNA and HCC

Small non-coding RNAs 19 to 25-nucleotide-long RNA called microRNAs (miRNA) define a new family of regulatory molecules. They recognize and bind the complementary sequences in 3’ untranslated regions (3’-UTR) of diverse target mRNAs[185]. By their role in control of diverse cellular processes such as proliferation, differentiation and apoptosis, miRNAs appear as new actor of regulation of tumorigenesis. However, miRNAs are themselves the target of regulation and their expressions are down- or up-regulated in various human cancer types and they can act as tumor suppressors or oncogenes[186192]. Murakami and collaborators identified five mRNAs (miRNA 199a, 199a*, 200a, 125a and 195) and three (miRNA 224, 18, p18) with lower and higher expression in HCC than in adjacent non-tumoral tissues, respectively[184]. Kutay and Gramantieri observed that miR-122a, the most represented miRNA in liver, is down regulated in HCC[193194]. Moreover, Gramantieri and collaborators showed a reverse correlation between miR-122a and cyclin G1. This miRNA appears to block the tumor growth through the inhibition of cyclin G1 expression[194]. miR-122a has been also described to facilitate the replication of HCV RNA in HCC[195]. The genomic integration of HBV in region of fragile site alters the expression of miRNA[196]. In fact, the genome integration of HBV in particularly in fragile site alters the expression of different miRNA. The expression of let-7e is lower in HCC than in non tumor liver and this inhibition is a consequence of HBV integration in FRA11B, FRA11G and FRA19A fragile sites[184187196]. miRNA-195 which modulates the expression of Bcl-2 like protein , SKI oncogenes and methyl-CpG binding protein-2 in HCC, is altered by integration of HBV in FRA17A[184196].

CONCLUSION

Numerous proteins involved on the control of different cellular processes such as cell proliferation, apoptosis and DNA replication are described as tumor suppressor. The deregulation of expression of these proteins by mutations and/or methylation of their promoter and viral-dependent-action contributes to the progression of hepatic cancers. Comprehensive understanding of the functions of these tumor suppressors is a prerequisite to devise innovative treatments of HCC.

DEFINITIONS
Tumor suppressor

Cell fate is controlled by division, differentiation and death. The balance between these commitments is determined by negative and positive regulations mainly through two classes of genes: the tumor suppressor genes and the proto-oncogene genes, respectively. In normal condition, tumor suppressor genes repress the formation and development of tumor but damage in their expression or function conduct to uncontrolled cell growth or cancer. Their function is impaired by mutations, loss of chromosome region or silencing by promoter methylation. Tumor suppressor genes are important targets in the quest to develop clinical therapies based on the restoration of gene function to reverse the carcinogenesis process.

Haploinsufficiency

The majority of tumor suppressor genes follow the “two-hit hypothesis”. The loss of function is associated with a damage of both alleles. In case of haploinsufficiency, the missing of only one allele is sufficient to inactivate the synthesis of gene product and to confer in a dose-dependent manner tumor sensitivity. Haploinsuffiency is associated with a few tumor suppressor genes such as p53 and PTEN.

Methylation state

Expression of genes can be regulated by methylation of their promoter. DNA methylation is the conversion of cytosine to 5-methylcytosine, which is catalyzed by DNA methyltransferase. It occurs on CpG sites mainly located in promoter region of genes. In case of cancer, aberrant methylation affects several tumor suppressor genes such as E-cadherin and SOCS1 leading to gene silencing and loss of protein function. The simultaneous methylation of CpG islands of multiple genes defines a new biomarker named CpG island methylator phenotype (CIMP). CIMP has been recognized as an important mechanism of gene regulation.

Loss of heterozygosity (LOH)

Loss of heterozygosity (LOH) refers to the loss of a single allele of a gene due to mutation, deletion of large chromosome segment and epigenetic regulating events such as methylation. LOH frequently affects tumor suppressor genes because most of them are located in chromosome fragile site. Losses in regions 1p, 4q, 6q, 8p, 13q, 16q, and 17p have been related to HCC and induce the lack of gene such as E-cadherin, WWOX, FHIT or p53.

Footnotes

Supported by The Stiftung für die Leberkranheiten, the EASL fellowship to JM and the Swiss National Foundation grant # 3100-063696 to JFD

References
1.  Saffroy R, Pham P, Lemoine A, Debuire B. Molecular biology and hepatocellular carcinoma: current status and future prospects. Ann Biol Clin (Paris). 2004;62:649-656.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Thorgeirsson SS, Grisham JW. Molecular pathogenesis of human hepatocellular carcinoma. Nat Genet. 2002;31:339-346.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Cha C, Dematteo RP. Molecular mechanisms in hepatocellular carcinoma development. Best Pract Res Clin Gastroenterol. 2005;19:25-37.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Appella E, Anderson CW. Signaling to p53: breaking the posttranslational modification code. Pathol Biol (Paris). 2000;48:227-245.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Pluquet O, Hainaut P. Genotoxic and non-genotoxic pathways of p53 induction. Cancer Lett. 2001;174:1-15.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Venkatachalam S, Shi YP, Jones SN, Vogel H, Bradley A, Pinkel D, Donehower LA. Retention of wild-type p53 in tumors from p53 heterozygous mice: reduction of p53 dosage can promote cancer formation. EMBO J. 1998;17:4657-4667.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Lynch CJ, Milner J. Loss of one p53 allele results in four-fold reduction of p53 mRNA and protein: a basis for p53 haplo-insufficiency. Oncogene. 2006;25:3463-3470.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Teramoto T, Satonaka K, Kitazawa S, Fujimori T, Hayashi K, Maeda S. p53 gene abnormalities are closely related to hepatoviral infections and occur at a late stage of hepatocarcinogenesis. Cancer Res. 1994;54:231-235.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Qin LX, Tang ZY. The prognostic molecular markers in hepatocellular carcinoma. World J Gastroenterol. 2002;8:385-392.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  Xue W, Zender L, Miething C, Dickins RA, Hernando E, Krizhanovsky V, Cordon-Cardo C, Lowe SW. Senescence and tumour clearance is triggered by p53 restoration in murine liver carcinomas. Nature. 2007;445:656-660.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Park HJ, Yu E, Shim YH. DNA methyltransferase expression and DNA hypermethylation in human hepatocellular carcinoma. Cancer Lett. 2006;233:271-278.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Zhang C, Li Z, Cheng Y, Jia F, Li R, Wu M, Li K, Wei L. CpG island methylator phenotype association with elevated serum alpha-fetoprotein level in hepatocellular carcinoma. Clin Cancer Res. 2007;13:944-952.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Bressac B, Kew M, Wands J, Ozturk M. Selective G to T mutations of p53 gene in hepatocellular carcinoma from southern Africa. Nature. 1991;350:429-431.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Montesano R, Hainaut P, Wild CP. Hepatocellular carcinoma: from gene to public health. J Natl Cancer Inst. 1997;89:1844-1851.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Kirk GD, Camus-Randon AM, Mendy M, Goedert JJ, Merle P, Trepo C, Brechot C, Hainaut P, Montesano R. Ser-249 p53 mutations in plasma DNA of patients with hepatocellular carcinoma from The Gambia. J Natl Cancer Inst. 2000;92:148-153.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Forrester K, Lupold SE, Ott VL, Chay CH, Band V, Wang XW, Harris CC. Effects of p53 mutants on wild-type p53-mediated transactivation are cell type dependent. Oncogene. 1995;10:2103-2111.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Wang XW, Gibson MK, Vermeulen W, Yeh H, Forrester K, Sturzbecher HW, Hoeijmakers JH, Harris CC. Abrogation of p53-induced apoptosis by the hepatitis B virus X gene. Cancer Res. 1995;55:6012-6016.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  Ponchel F, Puisieux A, Tabone E, Michot JP, Froschl G, Morel AP, Frobourg T, Fontaniere B, Oberhammer F, Ozturk M. Hepatocarcinoma-specific mutant p53-249ser induces mitotic activity but has no effect on transforming growth factor beta 1-mediated apoptosis. Cancer Res. 1994;54:2064-2068.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Puisieux A, Ji J, Guillot C, Legros Y, Soussi T, Isselbacher K, Ozturk M. p53-mediated cellular response to DNA damage in cells with replicative hepatitis B virus. Proc Natl Acad Sci USA. 1995;92:1342-1346.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Dumenco L, Oguey D, Wu J, Messier N, Fausto N. Introduction of a murine p53 mutation corresponding to human codon 249 into a murine hepatocyte cell line results in growth advantage, but not in transformation. Hepatology. 1995;22:1279-1288.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Scorsone KA, Zhou YZ, Butel JS, Slagle BL. p53 mutations cluster at codon 249 in hepatitis B virus-positive hepatocellular carcinomas from China. Cancer Res. 1992;52:1635-1638.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Aguilar F, Harris CC, Sun T, Hollstein M, Cerutti P. Geographic variation of p53 mutational profile in nonmalignant human liver. Science. 1994;264:1317-1319.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Staib F, Hussain SP, Hofseth LJ, Wang XW, Harris CC. TP53 and liver carcinogenesis. Hum Mutat. 2003;21:201-216.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Lucito R, Schneider RJ. Hepatitis B virus X protein activates transcription factor NF-kappa B without a requirement for protein kinase C. J Virol. 1992;66:983-991.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Feitelson MA, Zhu M, Duan LX, London WT. Hepatitis B x antigen and p53 are associated in vitro and in liver tissues from patients with primary hepatocellular carcinoma. Oncogene. 1993;8:1109-1117.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Bennett MR, Evan GI, Schwartz SM. Apoptosis of rat vascular smooth muscle cells is regulated by p53-dependent and -independent pathways. Circ Res. 1995;77:266-273.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  Miura N, Horikawa I, Nishimoto A, Ohmura H, Ito H, Hirohashi S, Shay JW, Oshimura M. Progressive telomere shortening and telomerase reactivation during hepatocellular carcinogenesis. Cancer Genet Cytogenet. 1997;93:56-62.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Wang XW, Forrester K, Yeh H, Feitelson MA, Gu JR, Harris CC. Hepatitis B virus X protein inhibits p53 sequence-specific DNA binding, transcriptional activity, and association with transcription factor ERCC3. Proc Natl Acad Sci USA. 1994;91:2230-2234.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Tan SL, Katze MG. How hepatitis C virus counteracts the interferon response: the jury is still out on NS5A. Virology. 2001;284:1-12.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Moriya K, Fujie H, Shintani Y, Yotsuyanagi H, Tsutsumi T, Ishibashi K, Matsuura Y, Kimura S, Miyamura T, Koike K. The core protein of hepatitis C virus induces hepatocellular carcinoma in transgenic mice. Nat Med. 1998;4:1065-1067.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Lee MN, Jung EY, Kwun HJ, Jun HK, Yu DY, Choi YH, Jang KL. Hepatitis C virus core protein represses the p21 promoter through inhibition of a TGF-beta pathway. J Gen Virol. 2002;83:2145-2151.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Yamanaka T, Uchida M, Doi T. Innate form of HCV core protein plays an important role in the localization and the function of HCV core protein. Biochem Biophys Res Commun. 2002;294:521-527.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  Alisi A, Giambartolomei S, Cupelli F, Merlo P, Fontemaggi G, Spaziani A, Balsano C. Physical and functional interaction between HCV core protein and the different p73 isoforms. Oncogene. 2003;22:2573-2580.  [PubMed]  [DOI]  [Cited in This Article: ]
34.  Bode AM, Dong Z. Post-translational modification of p53 in tumorigenesis. Nat Rev Cancer. 2004;4:793-805.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Appella E. Modulation of p53 function in cellular regulation. Eur J Biochem. 2001;268:2763.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Wu X, Bayle JH, Olson D, Levine AJ. The p53-mdm-2 autoregulatory feedback loop. Genes Dev. 1993;7:1126-1132.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Lin J, Zhu MH. [Interactive pathway of ARF-mdm2-p53]. Ai Zheng. 2003;22:328-330.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Anzola M, Cuevas N, Lopez-Martinez M, Saiz A, Burgos JJ, Martinez de Pancorboa M. P14ARF gene alterations in human hepatocellular carcinoma. Eur J Gastroenterol Hepatol. 2004;16:19-26.  [PubMed]  [DOI]  [Cited in This Article: ]
39.  Guan YS, La Z, Yang L, He Q, Li P. p53 gene in treatment of hepatic carcinoma: status quo. World J Gastroenterol. 2007;13:985-992.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Klein C, Vassilev LT. Targeting the p53-MDM2 interaction to treat cancer. Br J Cancer. 2004;91:1415-1419.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Bykov VJ, Selivanova G, Wiman KG. Small molecules that reactivate mutant p53. Eur J Cancer. 2003;39:1828-1834.  [PubMed]  [DOI]  [Cited in This Article: ]
42.  Pan HW, Chou HY, Liu SH, Peng SY, Liu CL, Hsu HC. Role of L2DTL, cell cycle-regulated nuclear and centrosome protein, in aggressive hepatocellular carcinoma. Cell Cycle. 2006;5:2676-2687.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Banks D, Wu M, Higa LA, Gavrilova N, Quan J, Ye T, Kobayashi R, Sun H, Zhang H. L2DTL/CDT2 and PCNA interact with p53 and regulate p53 polyubiquitination and protein stability through MDM2 and CUL4A/DDB1 complexes. Cell Cycle. 2006;5:1719-1729.  [PubMed]  [DOI]  [Cited in This Article: ]
44.  Ito M, Jiang C, Krumm K, Zhang X, Pecha J, Zhao J, Guo Y, Roeder RG, Xiao H. TIP30 deficiency increases susceptibility to tumorigenesis. Cancer Res. 2003;63:8763-8767.  [PubMed]  [DOI]  [Cited in This Article: ]
45.  Zhao J, Zhang X, Shi M, Xu H, Jin J, Ni H, Yang S, Dai J, Wu M, Guo Y. TIP30 inhibits growth of HCC cell lines and inhibits HCC xenografts in mice in combination with 5-FU. Hepatology. 2006;44:205-215.  [PubMed]  [DOI]  [Cited in This Article: ]
46.  Jiang C, Pecha J, Hoshino I, Ankrapp D, Xiao H. TIP30 mutant derived from hepatocellular carcinoma specimens promotes growth of HepG2 cells through up-regulation of N-cadherin. Cancer Res. 2007;67:3574-3582.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Ohgi T, Masaki T, Nakai S, Morishita A, Yukimasa S, Nagai M, Miyauchi Y, Funaki T, Kurokohchi K, Watanabe S. Expression of p33 (ING1) in hepatocellular carcinoma: relationships to tumour differentiation and cyclin E kinase activity. Scand J Gastroenterol. 2002;37:1440-1448.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  Zhu Z, Lin J, Qu JH, Feitelson MA, Ni CR, Li FM, Zhu MH. Inhibitory effect of tumor suppressor p33 (ING1b) and its synergy with p53 gene in hepatocellular carcinoma. World J Gastroenterol. 2005;11:1903-1909.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Jackson RJ, Engelman RW, Coppola D, Cantor AB, Wharton W, Pledger WJ. p21Cip1 nullizygosity increases tumor metastasis in irradiated mice. Cancer Res. 2003;63:3021-3025.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Hui AM, Makuuchi M, Li X. Cell cycle regulators and human hepatocarcinogenesis. Hepatogastroenterology. 1998;45:1635-1642.  [PubMed]  [DOI]  [Cited in This Article: ]
51.  Lee TK, Man K, Poon RT, Lo CM, Ng IO, Fan ST. Disruption of p53-p21/WAF1 cell cycle pathway contributes to progression and worse clinical outcome of hepatocellular carcinoma. Oncol Rep. 2004;12:25-31.  [PubMed]  [DOI]  [Cited in This Article: ]
52.  Lai PB, Chi TY, Chen GG. Different levels of p53 induced either apoptosis or cell cycle arrest in a doxycycline-regulated hepatocellular carcinoma cell line in vitro. Apoptosis. 2007;12:387-393.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Han HJ, Jung EY, Lee WJ, Jang KL. Cooperative repression of cyclin-dependent kinase inhibitor p21 gene expression by hepatitis B virus X protein and hepatitis C virus core protein. FEBS Lett. 2002;518:169-172.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Wagayama H, Shiraki K, Sugimoto K, Ito T, Fujikawa K, Yamanaka T, Takase K, Nakano T. High expression of p21WAF1/CIP1 is correlated with human hepatocellular carcinoma in patients with hepatitis C virus-associated chronic liver diseases. Hum Pathol. 2002;33:429-434.  [PubMed]  [DOI]  [Cited in This Article: ]
55.  Huether A, Höpfner M, Baradari V, Schuppan D, Scherübl H. EGFR blockade by cetuximab alone or as combination therapy for growth control of hepatocellular cancer. Biochem Pharmacol70(11):1568-1578.  [PubMed]  [DOI]  [Cited in This Article: ]
56.  Mauriz JL, Gonzalez P, Duran MC, Molpeceres V, Culebras JM, Gonzalez-Gallego J. Cell-cycle inhibition by TNP-470 in an in vivo model of hepatocarcinoma is mediated by a p53 and p21WAF1/CIP1 mechanism. Transl Res. 2007;149:46-53.  [PubMed]  [DOI]  [Cited in This Article: ]
57.  Fero ML, Randel E, Gurley KE, Roberts JM, Kemp CJ. The murine gene p27Kip1 is haplo-insufficient for tumour suppression. Nature. 1998;396:177-180.  [PubMed]  [DOI]  [Cited in This Article: ]
58.  Nan KJ, Jing Z, Gong L. Expression and altered subcellular localization of the cyclin-dependent kinase inhibitor p27Kip1 in hepatocellular carcinoma. World J Gastroenterol. 2004;10:1425-1430.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Philipp-Staheli J, Payne SR, Kemp CJ. p27 (Kip1): regulation and function of a haploinsufficient tumor suppressor and its misregulation in cancer. Exp Cell Res. 2001;264:148-168.  [PubMed]  [DOI]  [Cited in This Article: ]
60.  Lei PP, Zhang ZJ, Shen LJ, Li JY, Zou Q, Zhang HX. Expression and hypermethylation of p27 kip1 in hepatocarcinogenesis. World J Gastroenterol. 2005;11:4587-4591.  [PubMed]  [DOI]  [Cited in This Article: ]
61.  Ito Y, Matsuura N, Sakon M, Miyoshi E, Noda K, Takeda T, Umeshita K, Nagano H, Nakamori S, Dono K. Expression and prognostic roles of the G1-S modulators in hepatocellular carcinoma: p27 independently predicts the recurrence. Hepatology. 1999;30:90-99.  [PubMed]  [DOI]  [Cited in This Article: ]
62.  Fiorentino M, Altimari A, D’Errico A, Cukor B, Barozzi C, Loda M, Grigioni WF. Acquired expression of p27 is a favorable prognostic indicator in patients with hepatocellular carcinoma. Clin Cancer Res. 2000;6:3966-3972.  [PubMed]  [DOI]  [Cited in This Article: ]
63.  Matsuda Y, Ichida T. p16 and p27 are functionally correlated during the progress of hepatocarcinogenesis. Med Mol Morphol. 2006;39:169-175.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Azechi H, Nishida N, Fukuda Y, Nishimura T, Minata M, Katsuma H, Kuno M, Ito T, Komeda T, Kita R. Disruption of the p16/cyclin D1/retinoblastoma protein pathway in the majority of human hepatocellular carcinomas. Oncology. 2001;60:346-354.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Maeta Y, Shiota G, Okano J, Murawaki Y. Effect of promoter methylation of the p16 gene on phosphorylation of retinoblastoma gene product and growth of hepatocellular carcinoma cells. Tumour Biol. 2005;26:300-305.  [PubMed]  [DOI]  [Cited in This Article: ]
66.  Cho JW, Jeong YW, Han SW, Park JB, Jang BC, Baek WK, Kwon TK, Park JW, Kim SP, Suh MH. Aberrant p16INK4A RNA transcripts expressed in hepatocellular carcinoma cell lines regulate pRb phosphorylation by binding with CDK4, resulting in delayed cell cycle progression. Liver Int. 2003;23:194-200.  [PubMed]  [DOI]  [Cited in This Article: ]
67.  Qin Y, Liu JY, Li B, Sun ZL, Sun ZF. Association of low p16INK4a and p15INK4b mRNAs expression with their CpG islands methylation with human hepatocellular carcinogenesis. World J Gastroenterol. 2004;10:1276-1280.  [PubMed]  [DOI]  [Cited in This Article: ]
68.  Zhang YJ, Rossner P Jr, Chen Y, Agrawal M, Wang Q, Wang L, Ahsan H, Yu MW, Lee PH, Santella RM. Aflatoxin B1 and polycyclic aromatic hydrocarbon adducts, p53 mutations and p16 methylation in liver tissue and plasma of hepatocellular carcinoma patients. Int J Cancer. 2006;119:985-991.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Boyault S, Rickman DS, de Reynies A, Balabaud C, Rebouissou S, Jeannot E, Herault A, Saric J, Belghiti J, Franco D. Transcriptome classification of HCC is related to gene alterations and to new therapeutic targets. Hepatology. 2007;45:42-52.  [PubMed]  [DOI]  [Cited in This Article: ]
70.  Calvisi DF, Ladu S, Conner EA, Factor VM, Thorgeirsson SS. Disregulation of E-cadherin in transgenic mouse models of liver cancer. Lab Invest. 2004;84:1137-1147.  [PubMed]  [DOI]  [Cited in This Article: ]
71.  Kwon GY, Yoo BC, Koh KC, Cho JW, Park WS, Park CK. Promoter methylation of E-cadherin in hepatocellular carcinomas and dysplastic nodules. J Korean Med Sci. 2005;20:242-247.  [PubMed]  [DOI]  [Cited in This Article: ]
72.  Liu J, Lian Z, Han S, Waye MM, Wang H, Wu MC, Wu K, Ding J, Arbuthnot P, Kew M. Downregulation of E-cadherin by hepatitis B virus X antigen in hepatocellullar carcinoma. Oncogene. 2006;25:1008-1017.  [PubMed]  [DOI]  [Cited in This Article: ]
73.  Iso Y, Sawada T, Okada T, Kubota K. Loss of E-cadherin mRNA and gain of osteopontin mRNA are useful markers for detecting early recurrence of HCV-related hepatocellular carcinoma. J Surg Oncol. 2005;92:304-311.  [PubMed]  [DOI]  [Cited in This Article: ]
74.  Satoh S, Daigo Y, Furukawa Y, Kato T, Miwa N, Nishiwaki T, Kawasoe T, Ishiguro H, Fujita M, Tokino T. AXIN1 mutations in hepatocellular carcinomas, and growth suppression in cancer cells by virus-mediated transfer of AXIN1. Nat Genet. 2000;24:245-250.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Salahshor S, Woodgett JR. The links between axin and carcinogenesis. J Clin Pathol. 2005;58:225-236.  [PubMed]  [DOI]  [Cited in This Article: ]
76.  Zucman-Rossi J, Benhamouche S, Godard C, Boyault S, Grimber G, Balabaud C, Cunha AS, Bioulac-Sage P, Perret C. Differential effects of inactivated Axin1 and activated beta-catenin mutations in human hepatocellular carcinomas. Oncogene. 2007;26:774-780.  [PubMed]  [DOI]  [Cited in This Article: ]
77.  Mai M, Qian C, Yokomizo A, Smith DI, Liu W. Cloning of the human homolog of conductin (AXIN2), a gene mapping to chromosome 17q23-q24. Genomics. 1999;55:341-344.  [PubMed]  [DOI]  [Cited in This Article: ]
78.  Taniguchi K, Roberts LR, Aderca IN, Dong X, Qian C, Murphy LM, Nagorney DM, Burgart LJ, Roche PC, Smith DI. Mutational spectrum of beta-catenin, AXIN1, and AXIN2 in hepatocellular carcinomas and hepatoblastomas. Oncogene. 2002;21:4863-4871.  [PubMed]  [DOI]  [Cited in This Article: ]
79.  Ishizaki Y, Ikeda S, Fujimori M, Shimizu Y, Kurihara T, Itamoto T, Kikuchi A, Okajima M, Asahara T. Immunohistochemical analysis and mutational analyses of beta-catenin, Axin family and APC genes in hepatocellular carcinomas. Int J Oncol. 2004;24:1077-1083.  [PubMed]  [DOI]  [Cited in This Article: ]
80.  Iida M, Anna CH, Holliday WM, Collins JB, Cunningham ML, Sills RC, Devereux TR. Unique patterns of gene expression changes in liver after treatment of mice for 2 weeks with different known carcinogens and non-carcinogens. Carcinogenesis. 2005;26:689-699.  [PubMed]  [DOI]  [Cited in This Article: ]
81.  Colnot S, Decaens T, Niwa-Kawakita M, Godard C, Hamard G, Kahn A, Giovannini M, Perret C. Liver-targeted disruption of Apc in mice activates beta-catenin signaling and leads to hepatocellular carcinomas. Proc Natl Acad Sci USA. 2004;101:17216-17221.  [PubMed]  [DOI]  [Cited in This Article: ]
82.  Yang B, Guo M, Herman JG, Clark DP. Aberrant promoter methylation profiles of tumor suppressor genes in hepatocellular carcinoma. Am J Pathol. 2003;163:1101-1107.  [PubMed]  [DOI]  [Cited in This Article: ]
83.  Katoh H, Shibata T, Kokubu A, Ojima H, Fukayama M, Kanai Y, Hirohashi S. Epigenetic instability and chromosomal instability in hepatocellular carcinoma. Am J Pathol. 2006;168:1375-1384.  [PubMed]  [DOI]  [Cited in This Article: ]
84.  Katoh H, Shibata T, Kokubu A, Ojima H, Kosuge T, Kanai Y, Hirohashi S. Genetic inactivation of the APC gene contributes to the malignant progression of sporadic hepatocellular carcinoma: a case report. Genes Chromosomes Cancer. 2006;45:1050-1057.  [PubMed]  [DOI]  [Cited in This Article: ]
85.  Yoshimura A, Ohkubo T, Kiguchi T, Jenkins NA, Gilbert DJ, Copeland NG, Hara T, Miyajima A. A novel cytokine-inducible gene CIS encodes an SH2-containing protein that binds to tyrosine-phosphorylated interleukin 3 and erythropoietin receptors. EMBO J. 1995;14:2816-2826.  [PubMed]  [DOI]  [Cited in This Article: ]
86.  Hilton DJ, Richardson RT, Alexander WS, Viney EM, Willson TA, Sprigg NS, Starr R, Nicholson SE, Metcalf D, Nicola NA. Twenty proteins containing a C-terminal SOCS box form five structural classes. Proc Natl Acad Sci USA. 1998;95:114-119.  [PubMed]  [DOI]  [Cited in This Article: ]
87.  Krebs DL, Hilton DJ. SOCS proteins: negative regulators of cytokine signaling. Stem Cells. 2001;19:378-387.  [PubMed]  [DOI]  [Cited in This Article: ]
88.  Nagai H, Kim YS, Lee KT, Chu MY, Konishi N, Fujimoto J, Baba M, Matsubara K, Emi M. Inactivation of SSI-1, a JAK/STAT inhibitor, in human hepatocellular carcinomas, as revealed by two-dimensional electrophoresis. J Hepatol. 2001;34:416-421.  [PubMed]  [DOI]  [Cited in This Article: ]
89.  Nagai H, Kim YS, Konishi N, Baba M, Kubota T, Yoshimura A, Emi M. Combined hypermethylation and chromosome loss associated with inactivation of SSI-1/SOCS-1/JAB gene in human hepatocellular carcinomas. Cancer Lett. 2002;186:59-65.  [PubMed]  [DOI]  [Cited in This Article: ]
90.  Miyoshi H, Fujie H, Moriya K, Shintani Y, Tsutsumi T, Makuuchi M, Kimura S, Koike K. Methylation status of suppressor of cytokine signaling-1 gene in hepatocellular carcinoma. J Gastroenterol. 2004;39:563-569.  [PubMed]  [DOI]  [Cited in This Article: ]
91.  Okochi O, Hibi K, Sakai M, Inoue S, Takeda S, Kaneko T, Nakao A. Methylation-mediated silencing of SOCS-1 gene in hepatocellular carcinoma derived from cirrhosis. Clin Cancer Res. 2003;9:5295-5298.  [PubMed]  [DOI]  [Cited in This Article: ]
92.  Yoshikawa H, Matsubara K, Qian GS, Jackson P, Groopman JD, Manning JE, Harris CC, Herman JG. SOCS-1, a negative regulator of the JAK/STAT pathway, is silenced by methylation in human hepatocellular carcinoma and shows growth-suppression activity. Nat Genet. 2001;28:29-35.  [PubMed]  [DOI]  [Cited in This Article: ]
93.  Niwa Y, Kanda H, Shikauchi Y, Saiura A, Matsubara K, Kitagawa T, Yamamoto J, Kubo T, Yoshikawa H. Methylation silencing of SOCS-3 promotes cell growth and migration by enhancing JAK/STAT and FAK signalings in human hepatocellular carcinoma. Oncogene. 2005;24:6406-6417.  [PubMed]  [DOI]  [Cited in This Article: ]
94.  Calvisi DF, Ladu S, Gorden A, Farina M, Conner EA, Lee JS, Factor VM, Thorgeirsson SS. Ubiquitous activation of Ras and Jak/Stat pathways in human HCC. Gastroenterology. 2006;130:1117-1128.  [PubMed]  [DOI]  [Cited in This Article: ]
95.  Ogata H, Kobayashi T, Chinen T, Takaki H, Sanada T, Minoda Y, Koga K, Takaesu G, Maehara Y, Iida M. Deletion of the SOCS3 gene in liver parenchymal cells promotes hepatitis-induced hepatocarcinogenesis. Gastroenterology. 2006;131:179-193.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Leong GM, Moverare S, Brce J, Doyle N, Sjogren K, Dahlman-Wright K, Gustafsson JA, Ho KK, Ohlsson C, Leung KC. Estrogen up-regulates hepatic expression of suppressors of cytokine signaling-2 and -3 in vivo and in vitro. Endocrinology. 2004;145:5525-5531.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  Tommasi S, Dammann R, Jin SG, Zhang Xf XF, Avruch J, Pfeifer GP. RASSF3 and NORE1: identification and cloning of two human homologues of the putative tumor suppressor gene RASSF1. Oncogene. 2002;21:2713-2720.  [PubMed]  [DOI]  [Cited in This Article: ]
98.  Zabarovsky ER, Lerman MI, Minna JD. Tumor suppressor genes on chromosome 3p involved in the pathogenesis of lung and other cancers. Oncogene. 2002;21:6915-6935.  [PubMed]  [DOI]  [Cited in This Article: ]
99.  Hesson L, Dallol A, Minna JD, Maher ER, Latif F. NORE1A, a homologue of RASSF1A tumour suppressor gene is inactivated in human cancers. Oncogene. 2003;22:947-954.  [PubMed]  [DOI]  [Cited in This Article: ]
100.  Schagdarsurengin U, Wilkens L, Steinemann D, Flemming P, Kreipe HH, Pfeifer GP, Schlegelberger B, Dammann R. Frequent epigenetic inactivation of the RASSF1A gene in hepatocellular carcinoma. Oncogene. 2003;22:1866-1871.  [PubMed]  [DOI]  [Cited in This Article: ]
101.  Aoyama Y, Avruch J, Zhang XF. Nore1 inhibits tumor cell growth independent of Ras or the MST1/2 kinases. Oncogene. 2004;23:3426-3433.  [PubMed]  [DOI]  [Cited in This Article: ]
102.  Dammann R, Li C, Yoon JH, Chin PL, Bates S, Pfeifer GP. Epigenetic inactivation of a RAS association domain family protein from the lung tumour suppressor locus 3p21.3. Nat Genet. 2000;25:315-319.  [PubMed]  [DOI]  [Cited in This Article: ]
103.  Shivakumar L, Minna J, Sakamaki T, Pestell R, White MA. The RASSF1A tumor suppressor blocks cell cycle progression and inhibits cyclin D1 accumulation. Mol Cell Biol. 2002;22:4309-4318.  [PubMed]  [DOI]  [Cited in This Article: ]
104.  Zhang YJ, Ahsan H, Chen Y, Lunn RM, Wang LY, Chen SY, Lee PH, Chen CJ, Santella RM. High frequency of promoter hypermethylation of RASSF1A and p16 and its relationship to aflatoxin B1-DNA adduct levels in human hepatocellular carcinoma. Mol Carcinog. 2002;35:85-92.  [PubMed]  [DOI]  [Cited in This Article: ]
105.  Yeo W, Wong N, Wong WL, Lai PB, Zhong S, Johnson PJ. High frequency of promoter hypermethylation of RASSF1A in tumor and plasma of patients with hepatocellular carcinoma. Liver Int. 2005;25:266-272.  [PubMed]  [DOI]  [Cited in This Article: ]
106.  Zhong S, Yeo W, Tang MW, Wong N, Lai PB, Johnson PJ. Intensive hypermethylation of the CpG island of Ras association domain family 1A in hepatitis B virus-associated hepatocellular carcinomas. Clin Cancer Res. 2003;9:3376-3382.  [PubMed]  [DOI]  [Cited in This Article: ]
107.  Di Gioia S, Bianchi P, Destro A, Grizzi F, Malesci A, Laghi L, Levrero M, Morabito A, Roncalli M. Quantitative evaluation of RASSF1A methylation in the non-lesional, regenerative and neoplastic liver. BMC Cancer. 2006;6:89.  [PubMed]  [DOI]  [Cited in This Article: ]
108.  Macheiner D, Heller G, Kappel S, Bichler C, Stattner S, Ziegler B, Kandioler D, Wrba F, Schulte-Hermann R, Zochbauer-Muller S. NORE1B, a candidate tumor suppressor, is epigenetically silenced in human hepatocellular carcinoma. J Hepatol. 2006;45:81-89.  [PubMed]  [DOI]  [Cited in This Article: ]
109.  Ratziu V, Lalazar A, Wong L, Dang Q, Collins C, Shaulian E, Jensen S, Friedman SL. Zf9, a Kruppel-like transcription factor up-regulated in vivo during early hepatic fibrosis. Proc Natl Acad Sci USA. 1998;95:9500-9505.  [PubMed]  [DOI]  [Cited in This Article: ]
110.  Narla G, Heath KE, Reeves HL, Li D, Giono LE, Kimmelman AC, Glucksman MJ, Narla J, Eng FJ, Chan AM. KLF6, a candidate tumor suppressor gene mutated in prostate cancer. Science. 2001;294:2563-2566.  [PubMed]  [DOI]  [Cited in This Article: ]
111.  Chen C, Hyytinen ER, Sun X, Helin HJ, Koivisto PA, Frierson HF Jr, Vessella RL, Dong JT. Deletion, mutation, and loss of expression of KLF6 in human prostate cancer. Am J Pathol. 2003;162:1349-1354.  [PubMed]  [DOI]  [Cited in This Article: ]
112.  Reeves HL, Narla G, Ogunbiyi O, Haq AI, Katz A, Benzeno S, Hod E, Harpaz N, Goldberg S, Tal-Kremer S. Kruppel-like factor 6 (KLF6) is a tumor-suppressor gene frequently inactivated in colorectal cancer. Gastroenterology. 2004;126:1090-1103.  [PubMed]  [DOI]  [Cited in This Article: ]
113.  Yamashita K, Upadhyay S, Osada M, Hoque MO, Xiao Y, Mori M, Sato F, Meltzer SJ, Sidransky D. Pharmacologic unmasking of epigenetically silenced tumor suppressor genes in esophageal squamous cell carcinoma. Cancer Cell. 2002;2:485-495.  [PubMed]  [DOI]  [Cited in This Article: ]
114.  Ito G, Uchiyama M, Kondo M, Mori S, Usami N, Maeda O, Kawabe T, Hasegawa Y, Shimokata K, Sekido Y. Kruppel-like factor 6 is frequently down-regulated and induces apoptosis in non-small cell lung cancer cells. Cancer Res. 2004;64:3838-3843.  [PubMed]  [DOI]  [Cited in This Article: ]
115.  Glinsky GV, Glinskii AB, Stephenson AJ, Hoffman RM, Gerald WL. Gene expression profiling predicts clinical outcome of prostate cancer. J Clin Invest. 2004;113:913-923.  [PubMed]  [DOI]  [Cited in This Article: ]
116.  Kremer-Tal S, Reeves HL, Narla G, Thung SN, Schwartz M, Difeo A, Katz A, Bruix J, Bioulac-Sage P, Martignetti JA. Frequent inactivation of the tumor suppressor Kruppel-like factor 6 (KLF6) in hepatocellular carcinoma. Hepatology. 2004;40:1047-1052.  [PubMed]  [DOI]  [Cited in This Article: ]
117.  Li D, Yea S, Dolios G, Martignetti JA, Narla G, Wang R, Walsh MJ, Friedman SL. Regulation of Kruppel-like factor 6 tumor suppressor activity by acetylation. Cancer Res. 2005;65:9216-9225.  [PubMed]  [DOI]  [Cited in This Article: ]
118.  DiFeo A, Narla G, Camacho-Vanegas O, Nishio H, Rose SL, Buller RE, Friedman SL, Walsh MJ, Martignetti JA. E-cadherin is a novel transcriptional target of the KLF6 tumor suppressor. Oncogene. 2006;25:6026-6031.  [PubMed]  [DOI]  [Cited in This Article: ]
119.  Kremer-Tal S, Narla G, Chen Y, Hod E, DiFeo A, Yea S, Lee JS, Schwartz M, Thung SN, Fiel IM. Downregulation of KLF6 is an early event in hepatocarcinogenesis, and stimulates proliferation while reducing differentiation. J Hepatol. 2007;46:645-654.  [PubMed]  [DOI]  [Cited in This Article: ]
120.  Banck MS, Beaven SW, Narla G, Walsh MJ, Friedman SL, Beutler AS. KLF6 degradation after apoptotic DNA damage. FEBS Lett. 2006;580:6981-6986.  [PubMed]  [DOI]  [Cited in This Article: ]
121.  Sirach E, Bureau C, Peron JM, Pradayrol L, Vinel JP, Buscail L, Cordelier P. KLF6 transcription factor protects hepatocellular carcinoma-derived cells from apoptosis. Cell Death Differ. 2007;14:1202-1210.  [PubMed]  [DOI]  [Cited in This Article: ]
122.  Sulis ML, Parsons R. PTEN: from pathology to biology. Trends Cell Biol. 2003;13:478-483.  [PubMed]  [DOI]  [Cited in This Article: ]
123.  Yao YJ, Ping XL, Zhang H, Chen FF, Lee PK, Ahsan H, Chen CJ, Lee PH, Peacocke M, Santella RM. PTEN/MMAC1 mutations in hepatocellular carcinomas. Oncogene. 1999;18:3181-3185.  [PubMed]  [DOI]  [Cited in This Article: ]
124.  Chung TW, Lee YC, Ko JH, Kim CH. Hepatitis B Virus X protein modulates the expression of PTEN by inhibiting the function of p53, a transcriptional activator in liver cells. Cancer Res. 2003;63:3453-3458.  [PubMed]  [DOI]  [Cited in This Article: ]
125.  Wan XW, Jiang M, Cao HF, He YQ, Liu SQ, Qiu XH, Wu MC, Wang HY. The alteration of PTEN tumor suppressor expression and its association with the histopathological features of human primary hepatocellular carcinoma. J Cancer Res Clin Oncol. 2003;129:100-106.  [PubMed]  [DOI]  [Cited in This Article: ]
126.  Dong-Dong L, Xi-Ran Z, Xiang-Rong C. Expression and significance of new tumor suppressor gene PTEN in primary liver cancer. J Cell Mol Med. 2003;7:67-71.  [PubMed]  [DOI]  [Cited in This Article: ]
127.  Zhang L, Yu Q, He J, Zha X. Study of the PTEN gene expression and FAK phosphorylation in human hepatocarcinoma tissues and cell lines. Mol Cell Biochem. 2004;262:25-33.  [PubMed]  [DOI]  [Cited in This Article: ]
128.  Sieghart W, Fuereder T, Schmid K, Cejka D, Werzowa J, Wrba F, Wang X, Gruber D, Rasoul-Rockenschaub S, Peck-Radosavljevic M. Mammalian target of rapamycin pathway activity in hepatocellular carcinomas of patients undergoing liver transplantation. Transplantation. 2007;83:425-432.  [PubMed]  [DOI]  [Cited in This Article: ]
129.  Semela D, Piguet AC, Kolev M, Schmitter K, Hlushchuk R, Djonov V, Stoupis C, Dufour JF. Vascular remodeling and antitumoral effects of mTOR inhibition in a rat model of hepatocellular carcinoma. J Hepatol. 2007;46:840-848.  [PubMed]  [DOI]  [Cited in This Article: ]
130.  Ma DZ, Xu Z, Liang YL, Su JM, Li ZX, Zhang W, Wang LY, Zha XL. Down-regulation of PTEN expression due to loss of promoter activity in human hepatocellular carcinoma cell lines. World J Gastroenterol. 2005;11:4472-4477.  [PubMed]  [DOI]  [Cited in This Article: ]
131.  Wang L, Wang WL, Zhang Y, Guo SP, Zhang J, Li QL. Epigenetic and genetic alterations of PTEN in hepatocellular carcinoma. Hepatol Res. 2007;37:389-396.  [PubMed]  [DOI]  [Cited in This Article: ]
132.  Hu TH, Huang CC, Lin PR, Chang HW, Ger LP, Lin YW, Changchien CS, Lee CM, Tai MH. Expression and prognostic role of tumor suppressor gene PTEN/MMAC1/TEP1 in hepatocellular carcinoma. Cancer. 2003;97:1929-1940.  [PubMed]  [DOI]  [Cited in This Article: ]
133.  Mi D, Yi J, Liu E, Li X. Relationsip between PTEN and VEGF expression and clinicopathological characteristics in HCC. J Huazhong Univ Sci Technolog Med Sci. 2006;26:682-685.  [PubMed]  [DOI]  [Cited in This Article: ]
134.  Bieganowski P, Garrison PN, Hodawadekar SC, Faye G, Barnes LD, Brenner C. Adenosine monophosphoramidase activity of Hint and Hnt1 supports function of Kin28, Ccl1, and Tfb3. J Biol Chem. 2002;277:10852-10860.  [PubMed]  [DOI]  [Cited in This Article: ]
135.  Su T, Suzui M, Wang L, Lin CS, Xing WQ, Weinstein IB. Deletion of histidine triad nucleotide-binding protein 1/PKC-interacting protein in mice enhances cell growth and carcinogenesis. Proc Natl Acad Sci USA. 2003;100:7824-7829.  [PubMed]  [DOI]  [Cited in This Article: ]
136.  Weiske J, Huber O. The histidine triad protein Hint1 interacts with Pontin and Reptin and inhibits TCF-beta-catenin-mediated transcription. J Cell Sci. 2005;118:3117-3129.  [PubMed]  [DOI]  [Cited in This Article: ]
137.  Kanemaki M, Kurokawa Y, Matsu-ura T, Makino Y, Masani A, Okazaki K, Morishita T, Tamura TA. TIP49b, a new RuvB-like DNA helicase, is included in a complex together with another RuvB-like DNA helicase, TIP49a. J Biol Chem. 1999;274:22437-22444.  [PubMed]  [DOI]  [Cited in This Article: ]
138.  Makino Y, Kanemaki M, Kurokawa Y, Koji T, Tamura T. A rat RuvB-like protein, TIP49a, is a germ cell-enriched novel DNA helicase. J Biol Chem. 1999;274:15329-15335.  [PubMed]  [DOI]  [Cited in This Article: ]
139.  Bauer A, Chauvet S, Huber O, Usseglio F, Rothbacher U, Aragnol D, Kemler R, Pradel J. Pontin52 and reptin52 function as antagonistic regulators of beta-catenin signalling activity. EMBO J. 2000;19:6121-6130.  [PubMed]  [DOI]  [Cited in This Article: ]
140.  Weiske J, Huber O. The histidine triad protein Hint1 triggers apoptosis independent of its enzymatic activity. J Biol Chem. 2006;281:27356-27366.  [PubMed]  [DOI]  [Cited in This Article: ]
141.  Yuan BZ, Jefferson AM, Popescu NC, Reynolds SH. Aberrant gene expression in human non small cell lung carcinoma cells exposed to demethylating agent 5-aza-2’-deoxycytidine. Neoplasia. 2004;6:412-419.  [PubMed]  [DOI]  [Cited in This Article: ]
142.  Li H, Zhang Y, Su T, Santella RM, Weinstein IB. Hint1 is a haplo-insufficient tumor suppressor in mice. Oncogene. 2006;25:713-721.  [PubMed]  [DOI]  [Cited in This Article: ]
143.  Wang L, Zhang Y, Li H, Xu Z, Santella RM, Weinstein IB. Hint1 inhibits growth and activator protein-1 activity in human colon cancer cells. Cancer Res. 2007;67:4700-4708.  [PubMed]  [DOI]  [Cited in This Article: ]
144.  Martin J, Magnino F, Schmidt K, Piguet AC, Lee JS, Semela D, St-Pierre MV, Ziemiecki A, Cassio D, Brenner C. Hint2, a mitochondrial apoptotic sensitizer down-regulated in hepatocellular carcinoma. Gastroenterology. 2006;130:2179-2188.  [PubMed]  [DOI]  [Cited in This Article: ]
145.  Fong LY, Fidanza V, Zanesi N, Lock LF, Siracusa LD, Mancini R, Siprashvili Z, Ottey M, Martin SE, Druck T. Muir-Torre-like syndrome in Fhit-deficient mice. Proc Natl Acad Sci USA. 2000;97:4742-4747.  [PubMed]  [DOI]  [Cited in This Article: ]
146.  Inoue H, Ishii H, Alder H, Snyder E, Druck T, Huebner K, Croce CM. Sequence of the FRA3B common fragile region: implications for the mechanism of FHIT deletion. Proc Natl Acad Sci USA. 1997;94:14584-14589.  [PubMed]  [DOI]  [Cited in This Article: ]
147.  Croce CM, Sozzi G, Huebner K. Role of FHIT in human cancer. J Clin Oncol. 1999;17:1618-1624.  [PubMed]  [DOI]  [Cited in This Article: ]
148.  Huebner K, Garrison PN, Barnes LD, Croce CM. The role of the FHIT/FRA3B locus in cancer. Annu Rev Genet. 1998;32:7-31.  [PubMed]  [DOI]  [Cited in This Article: ]
149.  Chen YJ, Chen PH, Chang JG. Aberrant FHIT transcripts in hepatocellular carcinomas. Br J Cancer. 1998;77:417-420.  [PubMed]  [DOI]  [Cited in This Article: ]
150.  Gramantieri L, Chieco P, Di Tomaso M, Masi L, Piscaglia F, Brillanti S, Gaiani S, Valgimigli M, Mazziotti A, Bolondi L. Aberrant fragile histidine triad gene transcripts in primary hepatocellular carcinoma and liver cirrhosis. Clin Cancer Res. 1999;5:3468-3475.  [PubMed]  [DOI]  [Cited in This Article: ]
151.  Yuan BZ, Keck-Waggoner C, Zimonjic DB, Thorgeirsson SS, Popescu NC. Alterations of the FHIT gene in human hepatocellular carcinoma. Cancer Res. 2000;60:1049-1053.  [PubMed]  [DOI]  [Cited in This Article: ]
152.  Roz L, Gramegna M, Ishii H, Croce CM, Sozzi G. Restoration of fragile histidine triad (FHIT) expression induces apoptosis and suppresses tumorigenicity in lung and cervical cancer cell lines. Proc Natl Acad Sci USA. 2002;99:3615-3620.  [PubMed]  [DOI]  [Cited in This Article: ]
153.  Sard L, Accornero P, Tornielli S, Delia D, Bunone G, Campiglio M, Colombo MP, Gramegna M, Croce CM, Pierotti MA. The tumor-suppressor gene FHIT is involved in the regulation of apoptosis and in cell cycle control. Proc Natl Acad Sci USA. 1999;96:8489-8492.  [PubMed]  [DOI]  [Cited in This Article: ]
154.  Zhao P, Song X, Nin YY, Lu YL, Li XH. Loss of fragile histidine triad protein in human hepatocellular carcinoma. World J Gastroenterol. 2003;9:1216-1219.  [PubMed]  [DOI]  [Cited in This Article: ]
155.  Sun Y, Geng XP, Zhu LX, Xiong QR, Qian YB, Dong GY, Li XM. Clinicopathological significance of aberrant methylation of the fragile histidine triad gene in patients with hepatocellular carcinoma. Zhonghua Waike Zazhi. 2006;44:609-612.  [PubMed]  [DOI]  [Cited in This Article: ]
156.  Kannangai R, Sahin F, Adegbola O, Ashfaq R, Su GH, Torbenson M. FHIT mRNA and protein expression in hepatocellular carcinoma. Mod Pathol. 2004;17:653-659.  [PubMed]  [DOI]  [Cited in This Article: ]
157.  Zekri AR, Bahnassy AA, Hafez M, El-Shehaby AM, Sherif GM, Khaled HM, Zakhary N. Alterations of the fragile histidine triad gene in hepatitis C virus-associated hepatocellular carcinoma. J Gastroenterol Hepatol. 2005;20:87-94.  [PubMed]  [DOI]  [Cited in This Article: ]
158.  Nan KJ, Ruan ZP, Jing Z, Qin HX, Wang HY, Guo H, Xu R. Expression of fragile histidine triad in primary hepatocellular carcinoma and its relation with cell proliferation and apoptosis. World J Gastroenterol. 2005;11:228-231.  [PubMed]  [DOI]  [Cited in This Article: ]
159.  Paige AJ, Taylor KJ, Taylor C, Hillier SG, Farrington S, Scott D, Porteous DJ, Smyth JF, Gabra H, Watson JE. WWOX: a candidate tumor suppressor gene involved in multiple tumor types. Proc Natl Acad Sci USA. 2001;98:11417-11422.  [PubMed]  [DOI]  [Cited in This Article: ]
160.  Yakicier MC, Legoix P, Vaury C, Gressin L, Tubacher E, Capron F, Bayer J, Degott C, Balabaud C, Zucman-Rossi J. Identification of homozygous deletions at chromosome 16q23 in aflatoxin B1 exposed hepatocellular carcinoma. Oncogene. 2001;20:5232-5238.  [PubMed]  [DOI]  [Cited in This Article: ]
161.  Park SW, Ludes-Meyers J, Zimonjic DB, Durkin ME, Popescu NC, Aldaz CM. Frequent downregulation and loss of WWOX gene expression in human hepatocellular carcinoma. Br J Cancer. 2004;91:753-759.  [PubMed]  [DOI]  [Cited in This Article: ]
162.  Herath NI, Kew MC, Whitehall VL, Walsh MD, Jass JR, Khanna KK, Young J, Powell LW, Leggett BA, Macdonald GA. p73 is up-regulated in a subset of hepatocellular carcinomas. Hepatology. 2000;31:601-605.  [PubMed]  [DOI]  [Cited in This Article: ]
163.  Zemel R, Koren C, Bachmatove L, Avigad S, Kaganovsky E, Okon E, Ben-Ari Z, Grief F, Ben-Yehoyada M, Shaul Y. p73 overexpression and nuclear accumulation in hepatitis C virus-associated hepatocellular carcinoma. Dig Dis Sci. 2002;47:716-722.  [PubMed]  [DOI]  [Cited in This Article: ]
164.  Aqeilan RI, Pekarsky Y, Herrero JJ, Palamarchuk A, Letofsky J, Druck T, Trapasso F, Han SY, Melino G, Huebner K. Functional association between Wwox tumor suppressor protein and p73, a p53 homolog. Proc Natl Acad Sci USA. 2004;101:4401-4406.  [PubMed]  [DOI]  [Cited in This Article: ]
165.  Guler G, Uner A, Guler N, Han SY, Iliopoulos D, Hauck WW, McCue P, Huebner K. The fragile genes FHIT and WWOX are inactivated coordinately in invasive breast carcinoma. Cancer. 2004;100:1605-1614.  [PubMed]  [DOI]  [Cited in This Article: ]
166.  Iliopoulos D, Guler G, Han SY, Druck T, Ottey M, McCorkell KA, Huebner K. Roles of FHIT and WWOX fragile genes in cancer. Cancer Lett. 2006;232:27-36.  [PubMed]  [DOI]  [Cited in This Article: ]
167.  Iliopoulos D, Fabbri M, Druck T, Qin HR, Han SY, Huebner K. Inhibition of breast cancer cell growth in vitro and in vivo: effect of restoration of Wwox expression. Clin Cancer Res. 2007;13:268-274.  [PubMed]  [DOI]  [Cited in This Article: ]
168.  Imai Y, Soda M, Takahashi R. Parkin suppresses unfolded protein stress-induced cell death through its E3 ubiquitin-protein ligase activity. J Biol Chem. 2000;275:35661-35664.  [PubMed]  [DOI]  [Cited in This Article: ]
169.  Shimura H, Hattori N, Kubo S, Mizuno Y, Asakawa S, Minoshima S, Shimizu N, Iwai K, Chiba T, Tanaka K. Familial Parkinson disease gene product, parkin, is a ubiquitin-protein ligase. Nat Genet. 2000;25:302-305.  [PubMed]  [DOI]  [Cited in This Article: ]
170.  Ren Y, Zhao J, Feng J. Parkin binds to alpha/beta tubulin and increases their ubiquitination and degradation. J Neurosci. 2003;23:3316-3324.  [PubMed]  [DOI]  [Cited in This Article: ]
171.  Marin I, Lucas JI, Gradilla AC, Ferrus A. Parkin and relatives: the RBR family of ubiquitin ligases. Physiol Genomics. 2004;17:253-263.  [PubMed]  [DOI]  [Cited in This Article: ]
172.  Denison SR, Wang F, Becker NA, Schule B, Kock N, Phillips LA, Klein C, Smith DI. Alterations in the common fragile site gene Parkin in ovarian and other cancers. Oncogene. 2003;22:8370-8378.  [PubMed]  [DOI]  [Cited in This Article: ]
173.  Wang F, Denison S, Lai JP, Philips LA, Montoya D, Kock N, Schule B, Klein C, Shridhar V, Roberts LR. Parkin gene alterations in hepatocellular carcinoma. Genes Chromosomes Cancer. 2004;40:85-96.  [PubMed]  [DOI]  [Cited in This Article: ]
174.  Agirre X, Roman-Gomez J, Vazquez I, Jimenez-Velasco A, Garate L, Montiel-Duarte C, Artieda P, Cordeu L, Lahortiga I, Calasanz MJ. Abnormal methylation of the common PARK2 and PACRG promoter is associated with downregulation of gene expression in acute lymphoblastic leukemia and chronic myeloid leukemia. Int J Cancer. 2006;118:1945-1953.  [PubMed]  [DOI]  [Cited in This Article: ]
175.  Bestor TH. Activation of mammalian DNA methyltransferase by cleavage of a Zn binding regulatory domain. EMBO J. 1992;11:2611-2617.  [PubMed]  [DOI]  [Cited in This Article: ]
176.  Okano M, Bell DW, Haber DA, Li E. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell. 1999;99:247-257.  [PubMed]  [DOI]  [Cited in This Article: ]
177.  Okano M, Xie S, Li E. Cloning and characterization of a family of novel mammalian DNA (cytosine-5) methyltransferases. Nat Genet. 1998;19:219-220.  [PubMed]  [DOI]  [Cited in This Article: ]
178.  Nagai M, Nakamura A, Makino R, Mitamura K. Expression of DNA (5-cytosin)-methyltransferases (DNMTs) in hepatocellular carcinomas. Hepatol Res. 2003;26:186-191.  [PubMed]  [DOI]  [Cited in This Article: ]
179.  Saito Y, Kanai Y, Nakagawa T, Sakamoto M, Saito H, Ishii H, Hirohashi S. Increased protein expression of DNA methyltransferase (DNMT) 1 is significantly correlated with the malignant potential and poor prognosis of human hepatocellular carcinomas. Int J Cancer. 2003;105:527-532.  [PubMed]  [DOI]  [Cited in This Article: ]
180.  Choi MS, Shim YH, Hwa JY, Lee SK, Ro JY, Kim JS, Yu E. Expression of DNA methyltransferases in multistep hepatocarcinogenesis. Hum Pathol. 2003;34:11-17.  [PubMed]  [DOI]  [Cited in This Article: ]
181.  Park IY, Sohn BH, Yu E, Suh DJ, Chung YH, Lee JH, Surzycki SJ, Lee YI. Aberrant epigenetic modifications in hepatocarcinogenesis induced by hepatitis B virus X protein. Gastroenterology. 2007;132:1476-1494.  [PubMed]  [DOI]  [Cited in This Article: ]
182.  Matsukura S, Miyazaki K, Yakushiji H, Ogawa A, Chen Y, Sekiguchi M. Combined loss of expression of O6-methylguanine-DNA methyltransferase and hMLH1 accelerates progression of hepatocellular carcinoma. J Surg Oncol. 2003;82:194-200.  [PubMed]  [DOI]  [Cited in This Article: ]
183.  Matsukura S, Soejima H, Nakagawachi T, Yakushiji H, Ogawa A, Fukuhara M, Miyazaki K, Nakabeppu Y, Sekiguchi M, Mukai T. CpG methylation of MGMT and hMLH1 promoter in hepatocellular carcinoma associated with hepatitis viral infection. Br J Cancer. 2003;88:521-529.  [PubMed]  [DOI]  [Cited in This Article: ]
184.  Murakami Y, Yasuda T, Saigo K, Urashima T, Toyoda H, Okanoue T, Shimotohno K. Comprehensive analysis of microRNA expression patterns in hepatocellular carcinoma and non-tumorous tissues. Oncogene. 2006;25:2537-2545.  [PubMed]  [DOI]  [Cited in This Article: ]
185.  Miska EA. How microRNAs control cell division, differentiation and death. Curr Opin Genet Dev. 2005;15:563-568.  [PubMed]  [DOI]  [Cited in This Article: ]
186.  Michael MZ, O’ Connor SM, van Holst Pellekaan NG, Young GP, James RJ. Reduced accumulation of specific microRNAs in colorectal neoplasia. Mol Cancer Res. 2003;1:882-891.  [PubMed]  [DOI]  [Cited in This Article: ]
187.  Calin GA, Sevignani C, Dumitru CD, Hyslop T, Noch E, Yendamuri S, Shimizu M, Rattan S, Bullrich F, Negrini M. Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers. Proc Natl Acad Sci USA. 2004;101:2999-3004.  [PubMed]  [DOI]  [Cited in This Article: ]
188.  Iorio MV, Ferracin M, Liu CG, Veronese A, Spizzo R, Sabbioni S, Magri E, Pedriali M, Fabbri M, Campiglio M. MicroRNA gene expression deregulation in human breast cancer. Cancer Res. 2005;65:7065-7070.  [PubMed]  [DOI]  [Cited in This Article: ]
189.  Lu J, Getz G, Miska EA, Alvarez-Saavedra E, Lamb J, Peck D, Sweet-Cordero A, Ebert BL, Mak RH, Ferrando AA. MicroRNA expression profiles classify human cancers. Nature. 2005;435:834-838.  [PubMed]  [DOI]  [Cited in This Article: ]
190.  He L, Thomson JM, Hemann MT, Hernando-Monge E, Mu D, Goodson S, Powers S, Cordon-Cardo C, Lowe SW, Hannon GJ. A microRNA polycistron as a potential human oncogene. Nature. 2005;435:828-833.  [PubMed]  [DOI]  [Cited in This Article: ]
191.  Volinia S, Calin GA, Liu CG, Ambs S, Cimmino A, Petrocca F, Visone R, Iorio M, Roldo C, Ferracin M. A microRNA expression signature of human solid tumors defines cancer gene targets. Proc Natl Acad Sci USA. 2006;103:2257-2261.  [PubMed]  [DOI]  [Cited in This Article: ]
192.  Kent OA, Mendell JT. A small piece in the cancer puzzle: microRNAs as tumor suppressors and oncogenes. Oncogene. 2006;25:6188-6196.  [PubMed]  [DOI]  [Cited in This Article: ]
193.  Kutay H, Bai S, Datta J, Motiwala T, Pogribny I, Frankel W, Jacob ST, Ghoshal K. Downregulation of miR-122 in the rodent and human hepatocellular carcinomas. J Cell Biochem. 2006;99:671-678.  [PubMed]  [DOI]  [Cited in This Article: ]
194.  Gramantieri L, Ferracin M, Fornari F, Veronese A, Sabbioni S, Liu CG, Calin GA, Giovannini C, Ferrazzi E, Grazi GL. Cyclin G1 is a target of miR-122a, a microRNA frequently down-regulated in human hepatocellular carcinoma. Cancer Res. 2007;67:6092-6099.  [PubMed]  [DOI]  [Cited in This Article: ]
195.  Jopling CL, Yi M, Lancaster AM, Lemon SM, Sarnow P. Modulation of hepatitis C virus RNA abundance by a liver-specific MicroRNA. Science. 2005;309:1577-1581.  [PubMed]  [DOI]  [Cited in This Article: ]
196.  Feitelson MA, Lee J. Hepatitis B virus integration, fragile sites, and hepatocarcinogenesis. Cancer Lett. 2007;252:157-170.  [PubMed]  [DOI]  [Cited in This Article: ]