Skip to main content
American Journal of Translational Research logoLink to American Journal of Translational Research
. 2010 Jan 23;2(1):105–118.

Molecular genetics of hepatocellular neoplasia

Shilpa Jain 1, Shashideep Singhal 2, Peng Lee 1,3,4, Ruliang Xu 1
PMCID: PMC2826827  PMID: 20182587

Abstract

Hepatocellular carcinoma (HCC) is the sixth most common malignancy and the third leading cause of cancer deaths worldwide. Proper classification and early identification of HCC and precursor lesions is essential to the successful treatment and survival of HCC patients. Recent molecular genetic, pathologic, and clinical data have led to the stratification of hepatic adenomas into three subgroups: those with mutant TCF1/HNF1 α gene, those with mutant β-catenin, and those without mutations in either of these loci. Hepatic adenomas with α-catenin mutations have a significantly greater risk for malignant transformation in comparison with the other two subgroups. Telangiectatic focal nodular hyperplasia has now been reclassified as telangiectatic adenoma due to the presence of non-random methylation patterns, consistent with the monoclonal origin which is similar to hepatic adenoma and HCC. HCC precursor lesions demonstrate unique molecular alterations of HSP70, CAP2, glypican 3, and glutamine synthetase that have proven useful in the histologic diagnosis of early HCC. Though specific genetic alterations depend on HCC etiology, the main proteins affected include cell membrane receptors (in particular tyrosine kinase receptors) as well as proteins involved in cell signaling (specifically Wnt/beta-catenin, Ras/Raf/MEK/ERK and PI3K/Akt/mTOR pathways), cell cycle regulation (i.e. p53, p16/INK4, cyclin/cdk complex), invasiveness (EMT, TGF-β) and DNA metabolism. Advances in gene expression profiling have provided new insights into the molecular genetics of HCC. HCCs can now be stratified into two clinically relevant groups: Class A, the low survival subclass (overall survival time 30.3± 8.02 months), shows strong expression signatures of cell proliferation and antiapoptosis genes (such as PNCA and cell cycle regulators CDK4, CCNB1, CCNA2, and CKS2) as well as genes involving ubiquitination and sumoylation; Class B, the high survival subclass (overall survival time 83.7 ±10.3 months), does not have the above expression signature. In fact, insights into HCC-specific alterations of signal transduction pathways and protein expression patterns have led to the development of new therapeutic agents with molecular targets such as EGFR, VEGF, or other multi-kinase inhibitors. In the future, these specific molecular alterations in HCC can potentially serve as diagnostic tools, prognostic markers, and/or therapeutic targets with the potential to alter clinical outcomes.

Keywords: Molecular genetics, liver cancer, hepatocellular carcinoma, HSP70, CAP2, glypican 3, glutamine synthetase, β-catenin


Hepatocellular neoplasms mainly consist of hepatic adenoma, hepatocellular carcinoma (HCC) and precuror lesions. Benign tumors such as hepatic adenoma, while usually not deadly, may cause signfincant clinical challenges inlcuding maligant transformation. HCC due to vaious etiologies is one of the major leading causes of cancer death worldwide. Although knowledge about HCC is expanding exponentially in recent years, treatment and prevention of HCC is still a big challenge, and requires our thorough understanding of the molecular mechanisms of hepatocarci no-genesis. In this review article, we summarize the recent findings on molecular genetic pathology of hepatocellular neoplasms that have potential clinical implication in diagnosis, prognostication, and/or therapy.

Hepatic Adenoma

Hepatic adenoma predominantly occurs in younger women who are of child-bearing age, with or without prolonged use of oral contraception or abnormal carbohydrate metabolism (i.e. familial diabetes mellitus, glycogen storage disease, or galactosemia). It is usually a single nodule and sporadic, but can present as multiple tumors (adenomatosis) and have a familial inheritance pattern (familial liver adenomatosis). The major clinical significance includes spontaneous hemorrhage, rupture, and, in rare instances, malignant transformation [1-3].

Molecular basis of pathogenesis

The molecular mechanisms by which an adenoma arises from hepatocytes are not well understood. However, comparative genomic hybridization (CGH) data [4] suggest associations between tumorigenesis and frequent Wnt/β-catenin activation (20-34%) [5, 6], bi-allelic loss of function mutations in the genes encoding hepatocyte nuclear factor 1 α (HNF1α) or TCF1 (50%) [7], allelic imbalances in chromosomal arms lip, 13q and 17p [8], and gains of 1q. [9]

Molecular diagnosis and classification

While the diagnosis of hepatic adenoma can generally be made histologically, with or without radiological correlation, understanding the molecular genetics of hepatic adenomas may be clinically relevant. Based upon the mutation analysis of TCF1/HNF1α and β-catenin, a study proposed to sub-classify hepatic adenomas into three groups: 1) The most common and clinically important group has HNF1α mutation and is characterized by marked steatosis, lack of cytologic abnormalities, and no inflammatory infiltrate; 2) Tumors with only β-catenin activation have frequent cytologic abnormalities and pseudo-glandular formation; and 3) Tumors without TCF1/HNF1α or β-catenin activation mutation have frequent cytologic abnormalities, ductal reaction, and dystrophic vessels with/ without inflammatory infiltrates[9]. Accordingly, compared to TCF1/HNF1α mutation, β-catenin activation is more frequently associated with a high risk of malignant transformation in adenomas [9].

Although telangiectatic focal nodular hyperpla-sia (TFNH) is morphologically similar to focal nodular hyperplasia (FNH), it has been re-classified as “telangiectatic hepatic adenoma” after recent studies indicated that it shares most molecular genetic and clinical features with conventional hepatic adenoma. Its mon-clonality has been proven by non-random inacti-vation of the X chromosome, methylation analysis of the HUMARA locus, and loss of heterozy-gosity in a genome-wide allotyping [10, 11].

Hepatic adenomatosis, which characteristically presents with more than 10 liver lesions, may run in a familial pattern, and has frequent germ-line mutations [12], and biallelic inactivation of hepatocyte nuclear factor 1α (HNF1α) [13].

Hepatocellular carcinoma

Hepatocellular carcinoma is the sixth most common malignancy and the third most common cause of cancer deaths world wide. The major risk factors include chronic viral infection (HBV and HCV), alcoholic/nonalcoholic liver disease, environmental carcinogens (i.e. aflatoxin B1 (AFB), and inherited genetic disorders (Wilson's disease, hemochromatosis, α-1-antitrypsin deficiency, and tyrosinemia) [14-17]. With a few exceptions, HCC always develops in the setting of chronic hepatitis or cirrhosis, in which there is continuous inflammation and regeneration of hepatocytes. The non-random accumulated genetic alterations or chromosomal aberrations during the processes of inflammation, regeneration, and cirrhosis lead to the development of HCC [18, 19]. This multi-step process starts from hyperplastic change, dysplasia, and early HCC, eventually resulting in full-blown HCC. Qualitative and quantitative genetic alterations precede each step of carcinogenesis.

Molecular basis of pathogenesis

HCC is a heterogeneous group of carcinomas, with largely diverse molecular alterations associated with different etiologies. HBV-associated HCC may occur in the liver without cirrhosis and is distinct from HCC related to HCV and other etiologies. Although it is still under debate, HBV may have both direct and indirect oncogenic effects on hepatocytes [20, 21]. One direct effect is that viral DNA may be integrated into the hepatocyte genome, causing disruption of chromosomal stability or tumor suppressor genes and activation of proto-oncogene [22]. Another direct oncogenic effect may be attributable to the 154-amino acid (16.5-kDa) viral protein HBx, which may transactivate or up-regulate a variety of viral and cellular genes [23, 24]. The affected genes or molecules include basal transcription machinery in the nuclei (TFIIB, TBP, and RPB5) [25, 26], the Src pathway, and the Ras/Raf signaling pathway, which may, in turn, activate several oncogenes such as c-myc, c-jun, and c-fos in cytoplasm [27-30]. HBx protein amplifies TGF-β signaling through direct interaction with Smad4, and both direct and indirect interactions with DNA repair protein UVDDB, tumor suppressor proteins (p53 and APC gene product), cell cycle regulators, growth factors and receptor genes, cytokines, genes involved in apoptosis [31], proteasome subunits [32-34], and NF-kB, a modulator of immune response [29]. Indirectly, HBV infection causes liver cell injury mediated by cellular immune responses, resulting in carcinogenesis by promoting cell death, proliferation, and genetic mutations. When the HBV envelope coding region is transferred with activated T-lymphocytes to stimulate immune response in transgenic mice, they not only develop chronic hepatitis, but eventually also develop HCC [35].

The carcinogenesis of HCV infection-associated HCC differs from that related to HBV infection, and is mainly due to the indirect effects of viral infection. HCV viral DNA is never integrated into the genome of hepatocytes[36], but HCV infection may cause the accumulation of genetic abnormalities during the degeneration-regeneration process. Viral proteins, viral core protein in particular, may interfere with intracel-lular signaling pathways (activating TNF-α receptor, Raf-1 kinase, and NF-kβ pathways, resulting in inhibition of TNF-α and Fas-mediated apoptosis) and interact with the host immune system [37,38].

Exposure to food contaminated by Aflatoxin B1 (AFB1), a fungal metabolite produced by Asper-gillus flavus and related fungi, is associated with an increased incidence of HCC [39, 40]. AFB1 and its metabolite may result in a high frequency of mutations affecting 249ser in the p53 tumor suppressor gene (codon 249 G:C to T:A transversion) [41-44]. Moreover, there is a dose-dependent relationship between p53 249ser mutation load in cells and the intake of AFB1 in non-tumorous liver tissue [44]. In areas of the world with a high prevalence of AFB1 and HBV infection, synergy exists between HBV infection and high aflatoxin exposure in hepato-carcinogenesis [45].

The molecular mechanism for carcinogenesis associated with Wilson's disease, primary hemochromatosis, and other genetic diseases affecting the liver is also related to poorly controlled immune responses to copper or other metabolite accumulation. This immune response results in inflammation and generates oxidative free radicals that damage human DNA and cause genomic alterations in hepatocyte genes association with tumor suppression, cell cycle regulation, DNA repair, and apoptosis [46-48]. Mutation of p53 was frequently observed at codon 220 (A-G) in British patients with hemochromatosis-associated HCC [49]. Another highly frequent p53 mutation is at codon 249 (G:C to T:A transversion) in non-tumorous hepa-tocytes, and it is found in patients with hemochromatosis or Wilson's disease. Mutation at codon 250 (C:G to T:A transition) is also commonly seen in Wilson's disease-associated HCC [46]. The genetic abnormalities causing Wilson's disease or hemochromatosis do not increase the risk for carcinogenesis [50, 51].

Molecular genetics of precursors to HCC

Histopathologic and molecular biology studies have shown that the development of human HCCs is a multi-step process, from macroregen-erative nodules (MRN)/low-grade dysplasia (LGD), to high-grade dysplasia (HGD), to early HCC [52-54]. After an initial exposure or insult by carcinogens, it may take years or decades for humans to accumulate the necessary genetic and epigenetic damages necessary for preneo-plastic diseases to develop into HCC. These genetic damages or changes include the up-regulation of growth factors, inactivation of tumor suppressor genes, aberrant methylation, and microsatellite instability. Up-regulation of TGF-α and IGF-2 are sequelae of degeneration due to chronic inflammation, viral transactiva-tion, and hepatocellular repair and regeneration [55]. Aberrant hypo- or hypermethylation observed in chronic hepatitis and cirrhosis as well as HCC is due to increases in DNA methyltrans-ferases (DNMT) associated with chronic hepatitis and cirrhosis [56-59].

Loss of heterozygosity (LOH) and microsatellite instability occur in preneoplastic lesions and HCC [60, 61]. The gain at chromosomal locus 1q is the most common finding in dysplastic nodules and small HCCs [62]. Other informative markers include allelic chromosomal arms 1q and 14q, TATA box-binding protein (TBP) and BRCA1. LOH is detected in chromosome 8p21.3-p22 in approximately 40% of dysplastic nodules and HCCs. LOH on chromosome 11p13 is found in 15.8% of dysplastic nodules and 31.6% of HCCs. In dysplastic nodules, there is more LOH of D11S995 (33.3%) but less LOH of D11S907 (7.1%), whereas in HCCs, LOH of D11S907 (44.4%) is more frequently found than that of D11S995 (8.3%) [63]. In general, the multiplicity of allelic deletions in affected cell populations is low in chronic hepatitis, rises in dysplastic lesions, and is highest in HCCs [62, 64]. Gene profiling analysis in comparison to normal or surrounding cirrhotic tissue demonstrates that among the cDNA of 1152 genes tested, MRNs and dysplastic nodules have over 50 genes that are consistently deregulated. These deregulated genes (29 up-regulated and 24 down-regulated) include oncogenes, tumor suppressor genes, DNA repair genes, genes encoding cell growth factor and cytokines, genes encoding adhesion proteins, signal trans-duction genes, transcription factors, transcription factor/DNA binding protein genes, and housekeeping genes [65].

The unique molecular alterations seen in HCC precursor lesions may be useful for early diagnosis. However, several studies showed that the genetic or genomic alterations in preneoplastic or dysplastic nodules may not necessarily be found in HCC cells. These genetic or genomic differences suggest that not every early genomic aberration in precursor lesions is necessary or sufficient for the induction of malignant transformation of hepatocytes [66, 67]. Thus, most molecular alterations seen in preneoplastic lesions may not be suitable for diagnostic purposes. Although there are several candidate molecular markers (i.e. HSP70, CAP2, glypican 3 and glutamine synthetase) that have proven useful for the histologic diagnosis of early HCC, these results have yet to be confirmed in routine pathologic diagnosis [68].

Molecular genetics of HCC

Chromosomal abnormalities: Genomic abnormalities in HCC are largely heterogeneous due to the different molecular mechanisms of car-cinogenesis related to different etiologies and the multifactoral process of oncogenesis. Gain of chromosome 10q is unique to HCV-related HCC, while loss of 4q and 16q and gain of 11q are seen preferentially in HBV positive cases [62, 69]. Conventional cytogenetic studies and CGH have shown that most HCCs are aneuploid and harbor multiple chromosomal abnormalities, including non-random, recurrent DNA copy number losses on multiple chromosomal arms (1p, 4p, 5q, 6q, 8p, 9p, 13q, 16p, 16q, 17p) and gains on others (1q, 6p, 8q and 17q) [4, 70-74]. Chromosome 1q is the most common aberration across different geographic locations [72-74. The frequently deleted chromosome regions by LOH in HCCs contain many tumor suppressor genes and some oncogenes, (p53, Rb, p16, PTEN, DLC1, and IGF2R) [78-81]. LOH at chromosome 1p is usually seen in early, small or well-differentiated HCC [82], whereas LOH at chromosomes16p and 17p is more frequently associated with HCCs in advanced stages, aggressive tumor, and poor prognosis [83, 84]. By CGH, chromosome 8p, 17p and 19p are associated with HCC metastases [85].

Deregulation of signaling pathways: Deregulation of the major signal transduction pathways is found in all HCCs but differs with associated etiology. Abnormal activities of Wnt/β-catenin, hedgehog signaling, TGFβ, MAP/ras, IGF, apoptosis, microsatellite stability, phosphatase and tensin homolog gene (PTEN), p53, and Rb1 pathways are commonly found in HCCs, irrespective of etiology, and probably reflect common pathogenic mechanisms such as chronic liver injury and cirrhosis [9, 78, 86]. However, HCCs of different etiologies may predominantly affect certain pathways. HCV-associated HCC shows significant abnormalities in both Wnt/β-catenin and MAP kinase pathways [87, 88]. Dysfunction of Wnt/β-catenin, p53, pRb, MAP kinase [89], and cytokine signaling is more commonly seen in HBV-related HCC [87, 88, 90, 91]. Tumors associated with alcoholism have more frequent alterations in the Rb1 and p53 pathways than those caused by HCV infection [92]. The “aflatoxin-associated” p53 mutation in codon 249 is identified only in samples from areas with high aflatoxin content (Asia and Africa) [93, 94].

Abnormal Wnt signaling in HCC is exemplified by β-catenin overexpression or activation. β-catenin plays an important role in both intercellular adhesion and differentiation. Mutations in the β-catenin gene are detected in 26-41% of HCCs [95-97]. It is clinically related to less aggressive tumors than those without this mutation but harboring multiple chromosomal aberrations [98].

Deregulation of the p53 pathway is the most common cause of human carcinomas, including HCC. Loss of p53 function is observed in 25-60% of tumors [99] and occurs mostly due to allelic deletions at chromosome 17p13 and missense mutations in the specific DNA-binding domain [66, 100, 101]. p53 mutation is probably a late event in oncogenesis and is associated with both progression of HCC from an early to a more advanced stage [99, 102, 103] and HCC recurrence [104, 105]. A downstream target of zinc finger transcription factor ZBP-89 can be co-localized with p53 in the nucleus and appears to help nuclear accumulation of the p53 protein in a subset of recurrent HCC. Co-localization of p53 protein with ZBP-89 may define a subgroup of recurrent HCCs that are more sensitive to radiation or chemotherapy [106].

Retinoblastoma pathway inactivation is mainly through RB1 and CDKN2A promoter methyla-tion and rare genetic mutations. LOH at the Rb locus has been found in 25-48% of cases and strong down-regulation is seen in up to 50% of cases [79]. Rb gene is an important cycle controller and can be inactivated by mutations in the gene itself, loss of TGF-β responsiveness, and inactivation of p16, p15, or CDK4 [82, 107]. Loss of p16 protein due to inactivation of p16 by promoter hypermethylation, homozygous deletions, and point mutations may be noted in both early and late stage of HCC [108].

Microsatellite instability occurs in HCC as well as chronic hepatitis and cirrhosis [60, 61]. The incidence of microsatellite instability is higher in European HCC and in liver cirrhosis associated with HBV infection. In other parts of the world, microsatellite instability is an infrequent event [109] and only 11% of HCCs have abnormal DNA repair function. The degree of this abnormality correlates significantly with poor differentiation and portal vein involvement of HCC [110].

The Met pathway is deregulated in a subset of human HCCs. MET is an oncogene that encodes the tyrosine kinase receptor for hepatocyte growth factor (HGF) located on chromosome 7q21-q31. A subset of human HCCs with deregulated Met expression shows aggressive phenotype and poor prognosis [111].

Gene expression profiling of HCC: Recently developed technology, such as DNA microarrays and other molecular profiling techniques, has provided new insights into the molecular genetics of HCC [112-122]. Data from these techniques have demonstrated that, in most cases, transcripts that either directly or indirectly promote cell proliferation/growth are upregulated and those that inhibit cell proliferation/growth are down regulated. Many different cellular pathways are affected by these deregulated genes and gene products, including the extracellular matrix, the cytoskeleton (MMP14), oncogenes (Rho, raps homolog gene), tumor suppressor genes, MHC class IC or HLA-C, apoptosis-related genes (Dynein), signal transduction/ transla-tional regulator genes (Wnt/ β-catenin pathway members), and genes related to biotransforma-tion/metabolism (GST, monoamine oxidase, cytochromes, etc.). Moreover, gene expression profiling data confirm that HCV-related HCCs have different molecular genetics from those associated with HBV-related HCCs [123, 124], supporting the theory that these disease processes are driven by different pathophysiological mechanisms of hepatocarcinogenesis. These genes or gene products may be used as potential tumor markers that can be readily detected by serological or molecular tests. In addition, some gene profiling or signature genes have been found to be associated with greater potential for metastasis and recurrence [117, 125, 126].

Aberrant expression of MicroRNAs: MicroRNAs are small, noncoding RNAs with a stem-loop structure that are initially produced by RNA poly-merase II. They usually bind to 3’ untranslated regions of mRNA transcripts to regulate gene expression. Aberrant expression of several miRNAs has been implicated in HCC carcino-genesis, and miRNA expression signatures correlated with pathological and clinical behavior of HCC. Up-regulation of mir-221 and mir-21 could reduce tumor apoptosis and lead to angiogene-sis and invasion [127]. Receptor tyrosine kinase RAS and PI3K pathways are affected not only by down-regulated miR-1, miR-199a, and Let-7, but also by upregulated miR-2. Results may lead to cell growth, survival, motility, invasion, and metastasis [128-130]. A 20-miRNA signature has been found to be associated with HCC venous invasion and could also correlate with disease free and overall survival time [131].

Molecular classification and prognostication

Current classification and staging of HCC is mainly based upon histomorphology and/ or associated etiology and clinical presentation. In general, these classification or staging systems can provide useful information for the management of patients, prognostication, and to some extent, therapy, but their clinical relevance and accuracy are debatable. It has been advocated that molecular approaches, such as gene expression microarray and SNP array, should be used to develop a new classification system for HCCs that better predicts clinical outcome and facilitates targeted molecular therapy [132-135].

Expression signatures found via global gene expression profiling can stratify HCCs into several clinically relevant groups. For example, using DNA microarrays containing 21,329 unique genes, 91 human HCCs were analyzed; these data subclassified HCC into two distinctive groups. Class A, a low survival subclass (overall survival time 30.3+- 8.02 months), shows a strong expression signature of cell proliferation and anti-apoptosis genes (such as PNCA and cell cycle regulators: CDK4, CCNB1, CCNA2, and CKS2) as well as genes involving ubiquitination and sumoylation. In comparison, Class B, a high survival subclass (overall survival time 30.3+- 8.02 months 83.7 +-10.3 months), does not have the above expression signature[136]. Furthermore, gene expression profiles of nontu-moral liver tissue from paraffin-embedded specimens can be used to subclassify HCC patients into different survival groups [133].

A genome-wide transcriptomic analysis of 60 HCC tumors found 16 gene signatures that classify HCC tumors into the six robust subgroups (G1-G6). Each subgroup has unique clinical and genetic characteristics based upon chromosome stability status: G1-G3 are chromosome unstable and G4-G5 are chromosome stable. Since each group of tumors has specific pathway activations (i.e., protein kinase B (AKT or PKB)) in G1-G2 and Wnt pathways in G5-6), this molecular classification can not only provide prognostic information, but also facilitate the development of targeted therapies for HCC [132].

Whole-genomic array CGH analysis of 87 HCCs revealed two groups of tumors (clusters A and B) with significant differences in chromosomal alteration profiles and clinical outcomes. Cluster A's progression is more malignant than cluster B, shows exclusive chromosomal amplifications on 1q, 6p, and 8q, and has chromosomal losses on 8p. Cluster B has a low frequency of chromosomal alterations and tends to harbor limited numbers of chromosomal alterations. Since HCC is composed of several genetically homogeneous subclasses with characteristic genetics, these data have illuminated the opportunity for using targeted molecular therapy according to specific genetic background [134].

Molecular therapeutic targets of HCC

Treatment options for early or small HCC include liver transplantation, resection, or local radiation therapy, which significantly improve patient survival. However, because patients with HCC are usually diagnosed at advanced stages of disease, the above treatment modalities and chemotherapy are rarely effective. Furthermore, there is significant clinical and genetic heterogeneity among HCCs of different etiologies, thus one or a few standard treatments may not work for all HCCs. Recently introduced molecular targeted therapies are specific for groups of HCCs with similar genetics. The targeted therapy aims to inactivate activated oncogenes, recover tumor suppressor genes, or repair other genes and molecules related to HCC development, thereby correcting abnormal genes or functions as well as biological behavior. Recently, many candidate genome-based drug targets have been discovered via microarray technology, whole-genome epigenetic aberration analysis using promoter arrays, ChlP-chip analysis, and high-throughput sequencing systems. Examples of target genes or molecules include VEGFR, EGFR, DDEFL, VANGL1, WDRPUH, Ephrin-A1, gypican-3 (GPC3), number gain 7q, PFTAIRE protein kinase 1 (PFTK1), paternally expressed 10 (PEG10), and miR-122a [137-147]. Moreover, some of these targeted therapies, such as monoclonal antibodies, small molecules and antisense drugs, have reached phase II and III clinical trials for therapeutic use, and many have been shown to be effective. Sorafenib, an oral multikinase inhibitor of vascular endothelial growth factor receptor (VEGFR) and Ras kinase, has been approved by the FDA as a molecularly targeted anticancer agent [138, 148]. Some other agents targeting similar genes or molecules are being tested in preclinical and clinical trials for HCC. Results are summarized below and partially listed in Table 1.

Table 1.

Current molecular targeted therapies in HCC

Drug Type of Drugs Molecular Targets Affected Signaling Pathways FDA Approval
Sorafenib Tyrosine kinase inhibitor VEGFR, PDGFR, RAF VEGFR, PDGFR, RAS/MAPK yes
Sunitinib Tyrosine kinase inhibitor VEGFR, PDGFR, c-kit VEGFR, PDGFR, c-kit No, phase II or 3 trials
Bevacizumab Monoclonal antibodies to ligand VEGFR VEGFR No, phase II or 3 trials
Cetuximab Monoclonal antibodies to ligand EGFR EGFR No, phase II or 3 trials
Erlotinib Tyrosine kinase inhibitor EGFR EGFR No, phase II or 3 trials
Gefitinib Tyrosine kinase inhibitor EGFR EGFR No, phase II or 3 trials
Lapatinib Tyrosine kinase inhibitor Her-2/neu Her-2/neu No, phase II or 3 trials
Rapamycin ST kinase inhibitor mTOR PIK3/Akt/mTOR No, phase II or 3 trials
Everolimus ST kinase inhibitor mTOR PIK3/Akt/mTOR No, phase II or 3 trials
XL-765 ST kinase inhibitor PI3K PIK3/Akt/mTOR No, phase II or 3 trials
Trastuzumab monoclonal antibodies to receptor Her-2/neu Her-2/neu No, phase II or 3 trials

1. Anti-Epidermal growth factor receptor (anti-EGFR) therapy: Epidermal growth factor receptor (EGFR) is frequently expressed in both human HCC cell cultures and tumor tissues. Monoclonal antibodies against EGFR, such as Cetuximab, and small molecule tyrosine kinase inhibitors, such as Gefitinib and Erlotinib, have shown therapeutic effects in both cell culture and in patients in a phase II study [137].

2. Anti-vascular endothelial growth factor (antiangiogenesis): Vascular endothelial growth factor (VEGF) is upregulated via 6p21 gain [135] in HCC and targeting VEGF in HCC has potential anti-angiogenic effects. It is one of the putative targets of Sorafenib, an oral inhibitor of the VEGF receptor and other kinases [135]. Ad ministration of this drug in patients with ad vanced HCC was shown to increase median overall survival in a phase III, randomized, placebo-controlled trial (SHARP trial) [149]. Bevaci zumab, a recombinant, humanized monoclonal antibody, inhibits VEGF and also decreases the permeability of tumor vessels and relieves ele vated tumor interstitial pressure, thus poten tially enhancing the effectiveness of chemother apy [150, 151].

3. Multikinase targets: Many reagents target multiple sites of pathways or multiple genes or products. Sorafenib inhibits both VEGF and ei ther K-ras or its downstream effectors in the RAF/MEK/ERK pathway, thereby inducing tumor cell apoptosis [152]. Similar to Sorafenib, Sunitinib is an oral and multi-targeted receptor tyrosine kinase that exerts an antiangiogenic effect by targeting the tyrosine kinases VEGFR and platelet derived growth factor receptor (PDGFR). It has shown anti-HCC activity in both xenograft models and in a phase II clinical trial in patients with unresectable or metastatic HCC [153-155].

In summary, targeted therapy has proven to be an effective treatment for certain groups of HCC patients who might not respond to conventional therapeutic modalities. With a better understanding of the molecular mechanisms of hepa-tocarcinogenesis, more therapeutic options will be offered to cure or alleviate the symptoms of HCC.

References

  • 1.Edmondson HA, Henderson B, Benton B. Liver-cell adenomas associated with use of oral contraceptives. N Engl J Med. 1976;294:470–472. doi: 10.1056/NEJM197602262940904. [DOI] [PubMed] [Google Scholar]
  • 2.Flejou JF, Barge J, Menu Y, Degott C, Bismuth H, Potet F, Benhamou JP. Liver adenomatosis. An entity distinct from liver adenoma? Gastroen-terology. 1985;89:1132–1138. [PubMed] [Google Scholar]
  • 3.Foster JH, Donohue TA, Berman MM. Familial liver-cell adenomas and diabetes mellitus. N Engl J Med. 1978;299:239–241. doi: 10.1056/NEJM197808032990508. [DOI] [PubMed] [Google Scholar]
  • 4.Steinemann D, Skawran B, Becker T, Tauscher M, Weigmann A, Wingen L, Tauscher S, Hinrich-sen T, Hertz S, Flemming P, Flik J, Wiese B, Kreipe H, Lichter P, Schlegelberger B, Wilkens L. Assessment of differentiation and progression of hepatic tumors using array-based comparative genomic hybridization. Clin Gastro-enterol Hepatol. 2006;4:1283–1291. doi: 10.1016/j.cgh.2006.07.010. [DOI] [PubMed] [Google Scholar]
  • 5.Yamada Y, Yoshimi N, Sugie S, Suzui M, Matsu-naga K, Kawabata K, Hara A, Mori H. Beta-catenin (Ctnnb1) gene mutations in diethylni-trosamine (DEN)-induced liver tumors in male F344 rats. Jpn J Cancer Res. 1999;90:824–828. doi: 10.1111/j.1349-7006.1999.tb00822.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Devereux TR, Anna CH, Foley JF, White CM, Sills RC, Barrett JC. Mutation of beta-catenin is an early event in chemically induced mouse hepatocellular carcinogenesis. Oncogene. 1999;18:4726–4733. doi: 10.1038/sj.onc.1202858. [DOI] [PubMed] [Google Scholar]
  • 7.Yamagata K, Oda N, Kaisaki PJ, Menzel S, Fu-ruta H, Vaxillaire M, Southam L, Cox RD, Lathrop GM, Boriraj VV, Chen X, Cox NJ, Oda Y, Yano H, Le Beau MM, Yamada S, Nishigori H, Takeda J, Fajans SS, Hattersley AT, Iwasaki N, Hansen T, Pedersen O, Polonsky KS, Bell Gl, et al. Mutations in the hepatocyte nuclear factor-1alpha gene in maturity-onset diabetes of the young (MODY3) Nature. 1996;384:455–458. doi: 10.1038/384455a0. [DOI] [PubMed] [Google Scholar]
  • 8.Cai YR, Gong L, Teng XY, Zhang HT, Wang CF, Wei GL, Guo L, Ding F, Liu ZH, Pan QJ, Su Q. Clonality and allelotype analyses of focal nodular hyperplasia compared with hepatocellular adenoma and carcinoma. World J Gastroenterol. 2009;15:4695–4708. doi: 10.3748/wjg.15.4695. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Zucman-Rossi J, Jeannot E, Nhieu JT, Scoazec JY, Guettier C, Rebouissou S, Bacq Y, Leteurtre E, Paradis V, Michalak S, Wendum D, Chiche L, Fabre M, Mellottee L, Laurent C, Partensky C, Castaing D, Zafrani ES, Laurent-Puig P, Balabaud C, Bioulac-Sage P. Genotype-phenotype correlation in hepatocellular adenoma: new classification and relationship with HCC. Hepatology. 2006;43:515–524. doi: 10.1002/hep.21068. [DOI] [PubMed] [Google Scholar]
  • 10.Bioulac-Sage P, Rebouissou S, Sa Cunha A, Jeannot E, Lepreux S, Blanc JF, Blanche H, Le Bail B, Saric J, Laurent-Puig P, Balabaud C, Zucman-Rossi J. Clinical, morphologic, and molecular features defining so-called telangiectatic focal nodular hyperplasias of the liver. Gastroen-terology. 2005;128:1211–1218. doi: 10.1053/j.gastro.2005.02.004. [DOI] [PubMed] [Google Scholar]
  • 11.Paradis V, Benzekri A, Dargere D, Bieche I, Laurendeau I, Vilgrain V, Belghiti J, Vidaud M, Degott C, Bedossa P. Telangiectatic focal nodular hyperplasia: a variant of hepatocellular adenoma. Gastroenterology. 2004;126:1323–1329. doi: 10.1053/j.gastro.2004.02.005. [DOI] [PubMed] [Google Scholar]
  • 12.Bacq Y, Jacquemin E, Balabaud C, Jeannot E, Scotto B, Branchereau S, Laurent C, Bourlier P, Pariente D, de Muret A, Fabre M, Bioulac-Sage P, Zucman-Rossi J. Familial liver adenomatosis associated with hepatocyte nuclear factor 1alpha inactivation. Gastroenterology. 2003;125:1470–1475. doi: 10.1016/j.gastro.2003.07.012. [DOI] [PubMed] [Google Scholar]
  • 13.Bluteau O, Jeannot E, Bioulac-Sage P, Marques JM, Blanc JF, Bui H, Beaudoin JC, Franco D, Balabaud C, Laurent-Puig P, Zucman-Rossi J. Bi-allelic inactivation of TCF1 in hepatic adenomas. Nat Genet. 2002;32:312–315. doi: 10.1038/ng1001. [DOI] [PubMed] [Google Scholar]
  • 14.Parkin DM, Bray F, Ferlay J, Pisani P. Global cancer statistics, 2002. CA Cancer J Clin. 2005;55:74–108. doi: 10.3322/canjclin.55.2.74. [DOI] [PubMed] [Google Scholar]
  • 15.Clark JM. The epidemiology of nonalcoholic fatty liver disease in adults. J Clin Gastroenterol. 2006;40(Suppl 1):S5–10. doi: 10.1097/01.mcg.0000168638.84840.ff. [DOI] [PubMed] [Google Scholar]
  • 16.McGlynn KA, London WT. Epidemiology and natural history of hepatocellular carcinoma. Best Pract Res Clin Gastroenterol. 2005;19:3–23. doi: 10.1016/j.bpg.2004.10.004. [DOI] [PubMed] [Google Scholar]
  • 17.Motola-Kuba D, Zamora-Valdes D, Uribe M, Mendez-Sanchez N. Hepatocellular carcinoma. An overview. Ann Hepatol. 2006;5:16–24. [PubMed] [Google Scholar]
  • 18.Durr R, Caselmann WH. Carcinogenesis of primary liver malignancies. Langenbecks Arch Surg. 2000;385:154–161. doi: 10.1007/s004230050259. [DOI] [PubMed] [Google Scholar]
  • 19.El-Serag HB, Mason AC. Risk factors for the rising rates of primary liver cancer in the United States. Arch Intern Med. 2000;160:3227–3230. doi: 10.1001/archinte.160.21.3227. [DOI] [PubMed] [Google Scholar]
  • 20.Paterlini-Brechot P, Vona G, Brechot C. Circulating tumorous cells in patients with hepatocellular carcinoma. Clinical impact and future directions. Semin Cancer Biol. 2000;10:241–249. doi: 10.1006/scbi.2000.0323. [DOI] [PubMed] [Google Scholar]
  • 21.Brechot C, Gozuacik D, Murakami Y, Paterlini-Brechot P. Molecular bases for the development of hepatitis B virus (HBV)-related hepatocellular carcinoma (HCC) Semin Cancer Biol. 2000;10:211–231. doi: 10.1006/scbi.2000.0321. [DOI] [PubMed] [Google Scholar]
  • 22.Gozuacik D, Murakami Y, Saigo K, Chami M, Mugnier C, Lagorce D, Okanoue T, Urashima T, Brechot C, Paterlini-Brechot P. Identification of human cancer-related genes by naturally occurring Hepatitis B Virus DNA tagging. Oncogene. 2001;20:6233–6240. doi: 10.1038/sj.onc.1204835. [DOI] [PubMed] [Google Scholar]
  • 23.Murakami S. Hepatitis B virus X protea multifunctional viral regulator. J Gastroenterol. 2001;36:651–660. doi: 10.1007/s005350170027. [DOI] [PubMed] [Google Scholar]
  • 24.Caselmann WH. Trans-activation of cellular genes by hepatitis B virus proteins: a possible mechanism of hepatocarcinogenesis. Adv Virus Res. 1996;47:253–302. doi: 10.1016/s0065-3527(08)60737-x. [DOI] [PubMed] [Google Scholar]
  • 25.Qadri I, Maguire HF, Siddiqui A. Hepatitis B virus transactivator protein X interacts with the TATA-binding protein. Proc Natl Acad Sci U S A. 1995;92:1003–1007. doi: 10.1073/pnas.92.4.1003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Haviv I, Shamay M, Doitsh G, Shaul Y. Hepatitis B virus pX targets TFIIB in transcription coac-tivation. Mol Cell Biol. 1998;18:1562–1569. doi: 10.1128/mcb.18.3.1562. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Natoli G, Avantaggiati ML, Chirillo P, Costanzo A, Artini M, Balsano C, Levrero M. Induction of the DNA-binding activity of c-jun/c-fos het-erodimers by the hepatitis B virus transactivator pX. Mol Cell Biol. 1994;14:989–998. doi: 10.1128/mcb.14.2.989. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Benn J, Schneider RJ. Hepatitis B virus HBx protein activates Ras-GTP complex formation and establishes a Ras, Raf, MAP kinase signaling cascade. Proc Natl Acad Sci U S A. 1994;91:10350–10354. doi: 10.1073/pnas.91.22.10350. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Su F, Schneider RJ. Hepatitis B virus HBx protein activates transcription factor NF-kappaB by acting on multiple cytoplasmic inhibitors of rel-related proteins. J Virol. 1996;70:4558–4566. doi: 10.1128/jvi.70.7.4558-4566.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Terradillos O, Billet O, Renard CA, Levy R, Molina T, Briand P, Buendia MA. The hepatitis B virus X gene potentiates c-myc-induced liver oncogenesis in transgenic mice. Oncogene. 1997;14:395–404. doi: 10.1038/sj.onc.1200850. [DOI] [PubMed] [Google Scholar]
  • 31.Wu CG, Salvay DM, Forgues M, Valerie K, Farns-worth J, Markin RS, Wang XW. Distinctive gene expression profiles associated with Hepatitis B virus x protein. Oncogene. 2001;20:3674–3682. doi: 10.1038/sj.onc.1204481. [DOI] [PubMed] [Google Scholar]
  • 32.Chirillo P, Pagano S, Natoli G, Puri PL, Burgio VL, Balsano C, Levrero M. The hepatitis B virus X gene induces p53-mediated programmed cell death. Proc Natl Acad Sci U S A. 1997;94:8162–8167. doi: 10.1073/pnas.94.15.8162. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Elmore LW, Hancock AR, Chang SF, Wang XW, Chang S, Callahan CP, Geller DA, Will H, Harris CC. Hepatitis B virus X protein and p53 tumor suppressor interactions in the modulation of apoptosis. Proc Natl Acad Sci U S A. 1997;94:14707–14712. doi: 10.1073/pnas.94.26.14707. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Hu Z, Zhang Z, Doo E, Coux O, Goldberg AL, Liang TJ. Hepatitis B virus X protein is both a substrate and a potential inhibitor of the protea-some complex. J Virol. 1999;73:7231–7240. doi: 10.1128/jvi.73.9.7231-7240.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Nakamoto Y, Guidotti LG, Kuhlen CV, Fowler P, Chisari FV. Immune pathogenesis of hepatocellular carcinoma. J Exp Med. 1998;188:341–350. doi: 10.1084/jem.188.2.341. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Kasai Y, Takeda S, Takagi H. Pathogenesis of hepatocellular carcinoma: a review from the viewpoint of molecular analysis. Semin Surg Oncol. 1996;12:155–159. doi: 10.1002/(SICI)1098-2388(199605/06)12:3<155::AID-SSU2>3.0.CO;2-6. [DOI] [PubMed] [Google Scholar]
  • 37.Chen CM, You LR, Hwang LH, Lee YH. Direct interaction of hepatitis C virus core protein with the cellular lymphotoxin-beta receptor modulates the signal pathway of the lymphotoxin-beta receptor. J Virol. 1997;71:9417–9426. doi: 10.1128/jvi.71.12.9417-9426.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Block TM, Mehta AS, Fimmel CJ, Jordan R. Molecular viral oncology of hepatocellular carcinoma. Oncogene. 2003;22:5093–5107. doi: 10.1038/sj.onc.1206557. [DOI] [PubMed] [Google Scholar]
  • 39.Newberne PM. Chemical carcinogenesis: my-cotoxins and other chemicals to which humans are exposed. Semin Liver Dis. 1984;4:122–135. doi: 10.1055/s-2008-1040652. [DOI] [PubMed] [Google Scholar]
  • 40.Henry SH, Bosch FX, Bowers JC. Aflatoxin, hepatitis and worldwide liver cancer risks. Adv Exp Med Biol. 2002;504:229–233. doi: 10.1007/978-1-4615-0629-4_24. [DOI] [PubMed] [Google Scholar]
  • 41.Smela ME, Hamm ML, Henderson PT, Harris CM, Harris TM, Essigmann JM. The aflatoxin B(1) formamidopyrimidine adduct plays a major role in causing the types of mutations observed in human hepatocellular carcinoma. Proc Natl Acad Sci U S A. 2002;99:6655–6660. doi: 10.1073/pnas.102167699. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Smela ME, Currier SS, Bailey EA, Essigmann JM. The chemistry and biology of aflatoxin B(1): from mutational spectrometry to carcinogenesis. Carcinogenesis. 2001;22:535–545. doi: 10.1093/carcin/22.4.535. [DOI] [PubMed] [Google Scholar]
  • 43.Ozturk M. p53 mutation in hepatocellular carcinoma after aflatoxin exposure. Lancet. 1991;338:1356–1359. doi: 10.1016/0140-6736(91)92236-u. [DOI] [PubMed] [Google Scholar]
  • 44.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. doi: 10.1126/science.8191284. [DOI] [PubMed] [Google Scholar]
  • 45.Li Y, Su JJ, Qin LL, Yang C, Ban KC, Yan RQ. Synergistic effect of hepatitis B virus and aflatoxin B1 in hepatocarcinogenesis in tree shrews. Ann Acad Med Singapore. 1999;28:67–71. [PubMed] [Google Scholar]
  • 46.Hussain SP, Harris CC. Molecular epidemiology and carcinogenesis: endogenous and exogenous carcinogens. Mutat Res. 2000;462:311–322. doi: 10.1016/s1383-5742(00)00015-6. [DOI] [PubMed] [Google Scholar]
  • 47.Narayanan VS, Fitch CA, Levenson CW. Tumor suppressor protein p53 mRNA and subcellu-lar localization are altered by changes in cellular copper in human Hep G2 cells. J Nutr. 2001;131:1427–1432. doi: 10.1093/jn/131.5.1427. [DOI] [PubMed] [Google Scholar]
  • 48.Wang XW, Hussain SP, Huo TI, Wu CG, Forgues M, Hofseth U, Brechot C, Harris CC. Molecular pathogenesis of human hepatocellular carcinoma. Toxicology. 2002:181–182. 43–47. doi: 10.1016/s0300-483x(02)00253-6. [DOI] [PubMed] [Google Scholar]
  • 49.Vautier G, Bomford AB, Portmann BC, Metivier E, Williams R, Ryder SD. p53 mutations in British patients with hepatocellular carcinoma: clustering in genetic hemochromatosis. Gastroen-terology. 1999;117:154–160. doi: 10.1016/s0016-5085(99)70562-7. [DOI] [PubMed] [Google Scholar]
  • 50.Boige V, Castera L, de Roux N, Ganne-Carrie N, Ducot B, Pelletier G, Beaugrand M, Buffet C. Lack of association between HFE gene mutations and hepatocellular carcinoma in patients with cirrhosis. Gut. 2003;52:1178–1181. doi: 10.1136/gut.52.8.1178. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Cauza E, Peck-Radosavljevic M, Ulrich-Pur H, Datz C, Gschwantler M, Schoniger-Hekele M, Hackl F, Polli C, Rasoul-Rockenschaub S, Muller C, Wrba F, Gangl A, Ferenci P. Mutations of the HFE gene in patients with hepatocellular carcinoma. Am J Gastroenterol. 2003;98:442–447. doi: 10.1111/j.1572-0241.2003.07222.x. [DOI] [PubMed] [Google Scholar]
  • 52.Theise ND, Park YN, Kojiro M. Dysplastic nodules and hepatocarcinogenesis. Clin Liver Dis. 2002;6:497–512. doi: 10.1016/s1089-3261(02)00006-5. [DOI] [PubMed] [Google Scholar]
  • 53.Orsatti G, Theise ND, Thung SN, Paronetto F. DNA image cytometric analysis of macroregen-erative nodules (adenomatous hyperplasia) of the liver: evidence in support of their preneo-plastic nature. Hepatology. 1993;17:621–627. doi: 10.1002/hep.1840170416. [DOI] [PubMed] [Google Scholar]
  • 54.Takayama T, Makuuchi M, Hirohashi S, Sakamoto M, Okazaki N, Takayasu K, Kosuge T, Mo-too Y, Yamazaki S, Hasegawa H. Malignant transformation of adenomatous hyperplasia to hepatocellular carcinoma. Lancet. 1990;336:1150–1153. doi: 10.1016/0140-6736(90)92768-d. [DOI] [PubMed] [Google Scholar]
  • 55.Grisham JW. Interspecies comparison of liver carcinogenesis: implications for cancer risk assessment. Carcinogenesis. 1997;18:59–81. doi: 10.1093/carcin/18.1.59. [DOI] [PubMed] [Google Scholar]
  • 56.Kanai Y, Ushijima S, Tsuda H, Sakamoto M, Hirohashi S. Aberrant DNA methylation precedes loss of heterozygosity on chromosome 16 in chronic hepatitis and liver cirrhosis. Cancer Lett. 2000;148:73–80. doi: 10.1016/s0304-3835(99)00316-x. [DOI] [PubMed] [Google Scholar]
  • 57.Lin CH, Hsieh SY, Sheen IS, Lee WC, Chen TC, Shyu WC, Liaw YF. Genome-wide hypomethy-lation in hepatocellular carcinogenesis. Cancer Res. 2001;61:4238–4243. [PubMed] [Google Scholar]
  • 58.Shen L, Fang J, Qiu D, Zhang T, Yang J, Chen S, Xiao S. Correlation between DNA methylation and pathological changes in human hepatocellular carcinoma. Hepatogastroenterology. 1998;45:1753–1759. [PubMed] [Google Scholar]
  • 59.Shim YH, Yoon GS, Choi HJ, Chung YH, Yu E. p16 Hypermethylation in the early stage of hepatitis B virus-associated hepatocarcinogenesis. Cancer Lett. 2003;190:213–219. doi: 10.1016/s0304-3835(02)00613-4. [DOI] [PubMed] [Google Scholar]
  • 60.Karachristos A, Liloglou T, Field JK, Deligiorgi E, Kouskouni E, Spandidos DA. Microsatellite instability and p53 mutations in hepatocellular carcinoma. Mol Cell Biol Res Commun. 1999;2:155–161. doi: 10.1006/mcbr.1999.0170. [DOI] [PubMed] [Google Scholar]
  • 61.Piao Z, Kim H, Malkhosyan S, Park C. Frequent chromosomal instability but no microsatellite instability in hepatocellular carcinomas. Int J Oncol. 2000;17:507–512. doi: 10.3892/ijo.17.3.507. [DOI] [PubMed] [Google Scholar]
  • 62.Nishida N, Nishimura T, Ito T, Komeda T, Fukuda Y, Nakao K. Chromosomal instability and human hepatocarcinogenesis. Histol Histopathol. 2003;18:897–909. doi: 10.14670/HH-18.897. [DOI] [PubMed] [Google Scholar]
  • 63.Kahng YS, Lee YS, Kim BK, Park WS, Lee JY, Kang CS. Loss of heterozygosity of chromosome 8p and lip in the dysplastic nodule and hepatocellular carcinoma. J Gastroenterol Hepatol. 2003;18:430–436. doi: 10.1046/j.1440-1746.2003.02997.x. [DOI] [PubMed] [Google Scholar]
  • 64.Maggioni M, Coggi G, Cassani B, Bianchi P, Ro-magnoli S, Mandelli A, Borzio M, Colombo P, Roncalli M. Molecular changes in hepatocellular dysplastic nodules on microdissected liver biopsies. Hepatology. 2000;32:942–946. doi: 10.1053/jhep.2000.18425. [DOI] [PubMed] [Google Scholar]
  • 65.Anders RA, Yerian LM, Tretiakova M, Davison JM, Quigg RJ, Domer PH, Hoberg J, Hart J. cDNA microarray analysis of macroregenerative and dysplastic nodules in end-stage hepatitis C virus-induced cirrhosis. Am J Pathol. 2003;162:991–1000. doi: 10.1016/S0002-9440(10)63893-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Staib F, Hussain SP, Hofseth U, Wang XW, Harris CC. TP53 and liver carcinogenesis. Hum Mutat. 2003;21:201–216. doi: 10.1002/humu.10176. [DOI] [PubMed] [Google Scholar]
  • 67.Roncalli M, Bianchi P, Grimaldi GC, Ricci D, Laghi L, Maggioni M, Opocher E, Borzio M, Coggi G. Fractional allelic loss in non-end-stage cirrhosis: correlations with hepatocellular carcinoma development during follow-up. Hepatology. 2000;31:846–850. doi: 10.1053/he.2000.5790. [DOI] [PubMed] [Google Scholar]
  • 68.Sakamoto M, Mori T, Masugi Y, Effendi K, Rie I, Du W. Candidate molecular markers for histological diagnosis of early hepatocellular carcinoma. Intervirology. 2008;51(Suppl 1):42–45. doi: 10.1159/000122603. [DOI] [PubMed] [Google Scholar]
  • 69.Kusano N, Shiraishi K, Kubo K, Oga A, Okita K, Sasaki K. Genetic aberrations detected by comparative genomic hybridization in hepatocellular carcinomas: their relationship to clinicopa-thological features. Hepatology. 1999;29:1858–1862. doi: 10.1002/hep.510290636. [DOI] [PubMed] [Google Scholar]
  • 70.Buendia MA. Genetics of hepatocellular carcinoma. Semin Cancer Biol. 2000;10:185–200. doi: 10.1006/scbi.2000.0319. [DOI] [PubMed] [Google Scholar]
  • 71.Nagai H, Pineau P, Tiollais P, Buendia MA, Dejean A. Comprehensive allelotyping of human hepatocellular carcinoma. Oncogene. 1997;14:2927–2933. doi: 10.1038/sj.onc.1201136. [DOI] [PubMed] [Google Scholar]
  • 72.Terracciano L, Tornillo L. Cytogenetic alterations in liver cell tumors as detected by comparative genomic hybridization. Pathologica. 2003;95:71–82. [PubMed] [Google Scholar]
  • 73.Guan XY, Fang Y, Sham J, Kwong D, Zhang Y, Liang Q, Li H, Zhou H, Trent J. Recurrent chromosome alterations in hepatocellular carcinoma detected by comparative genomic hybridization. Genes Chromosomes Cancer. 2001;30:110. [PubMed] [Google Scholar]
  • 74.Chang J, Kim NG, Piao Z, Park C, Park KS, Paik YK, Lee WJ, Kim BR, Kim H. Assessment of chromosomal losses and gains in hepatocellular carcinoma. Cancer Lett. 2002;182:193–202. doi: 10.1016/s0304-3835(02)00083-6. [DOI] [PubMed] [Google Scholar]
  • 75.Wilkens L, Bredt M, Flemming P, Mengel M, Becker T, Klempnauer J, Kreipe H. Comparative genomic hybridization (CGH) and fluorescence in situ hybridization (FISH) in the diagnosis of hepatocellular carcinoma. J Hepatobiliary Pancreat Surg. 2002;9:304–311. doi: 10.1007/s005340200034. [DOI] [PubMed] [Google Scholar]
  • 76.Wilkens L, Flemming P, Bredt M, Kreipe H. Detection of chromosomal imbalances in hepatocellular carcinoma. Expert Rev Mol Diagn. 2002;2:120–128. doi: 10.1586/14737159.2.2.120. [DOI] [PubMed] [Google Scholar]
  • 77.Simon D, Knowles BB, Weith A. Abnormalities of chromosome 1 and loss of heterozygosity on 1p in primary hepatomas. Oncogene. 1991;6:765–770. [PubMed] [Google Scholar]
  • 78.Edamoto Y, Hara A, Biernat W, Terracciano L, Cathomas G, Riehle HM, Matsuda M, Fujii H, Scoazec JY, Ohgaki H. Alterations of RBI, p53 and Wnt pathways in hepatocellular carcinomas associated with hepatitis C, hepatitis B and alcoholic liver cirrhosis. Int J Cancer. 2003;106:334–341. doi: 10.1002/ijc.11254. [DOI] [PubMed] [Google Scholar]
  • 79.Hsia CC, Di Bisceglie AM, Kleiner DE, Jr, Farshid M, Tabor E. RB tumor suppressor gene expression in hepatocellular carcinomas from patients infected with the hepatitis B virus. J Med Virol. 1994;44:67–73. doi: 10.1002/jmv.1890440113. [DOI] [PubMed] [Google Scholar]
  • 80.Matsuda Y, Ichida T, Genda T, Yamagiwa S, Ao-yagi Y, Asakura H. Loss of p16 contributes to p27 sequestration by cyclin D(1)-cyclin-dependent kinase 4 complexes and poor prognosis in hepatocellular carcinoma. Clin Cancer Res. 2003;9:3389–3396. [PubMed] [Google Scholar]
  • 81.Ng IO, Liang ZD, Cao L, Lee TK. DLC-1 is deleted in primary hepatocellular carcinoma and exerts inhibitory effects on the proliferation of hepatoma cell lines with deleted DLC-1. Cancer Res. 2000;60:6581–6584. [PubMed] [Google Scholar]
  • 82.Kuroki T, Fujiwara Y, Tsuchiya E, Nakamori S, Imaoka S, Kanematsu T, Nakamura Y. Accumulation of genetic changes during development and progression of hepatocellular carcinoma: loss of heterozygosity of chromosome arm 1p occurs at an early stage of hepatocarcinogene-sis. Genes Chromosomes Cancer. 1995;13:163–167. doi: 10.1002/gcc.2870130305. [DOI] [PubMed] [Google Scholar]
  • 83.Lin YW, Sheu JC, Huang GT, Lee HS, Chen CH, Wang JT, Lee PH, Lu FJ. Chromosomal abnormality in hepatocellular carcinoma by comparative genomic hybridisation in Taiwan. Eur J Cancer. 1999;35:652–658. doi: 10.1016/s0959-8049(98)00430-4. [DOI] [PubMed] [Google Scholar]
  • 84.Tamura S, Nakamori S, Kuroki T, Sasaki Y, Furu-kawa H, Ishikawa O, Imaoka S, Nakamura Y. Association of cumulative allelic losses with tumor aggressiveness in hepatocellular carcinoma. J Hepatol. 1997;27:669–676. doi: 10.1016/s0168-8278(97)80084-0. [DOI] [PubMed] [Google Scholar]
  • 85.Zhang LH, Qin LX, Ma ZC, Ye SL, Liu YK, Ye QH, Wu X, Huang W, Tang ZY. Allelic imbalance regions on chromosomes 8p, 17p and 19p related to metastasis of hepatocellular carcinoma: comparison between matched primary and me-tastatic lesions in 22 patients by genome-wide microsatellite analysis. J Cancer Res Clin Oncol. 2003;129:279–286. doi: 10.1007/s00432-002-0407-5. [DOI] [PubMed] [Google Scholar]
  • 86.Aravalli RN, Steer CJ, Cressman EN. Molecular mechanisms of hepatocellular carcinoma. Hepatology. 2008;48:2047–2063. doi: 10.1002/hep.22580. [DOI] [PubMed] [Google Scholar]
  • 87.Colnot S, Decaens T, Niwa-Kawakita M, Godard C, Hamard G, Kahn A, Giovannini M, Perret C. Liver-targeted disruption of Ape in mice activates beta-catenin signaling and leads to hepatocellular carcinomas. Proc Natl Acad Sci U S A. 2004;101:17216–17221. doi: 10.1073/pnas.0404761101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Bai F, Nakanishi Y, Takayama K, Pei XH, Inoue K, Harada T, Izumi M, Hara N. Codon 64 of K-ras gene mutation pattern in hepatocellular carcinomas induced by bleomycin and 1-nitropyrene in A/J mice. Teratog Carcinog Mutagen. 2003;(Suppl 1):161–170. doi: 10.1002/tcm.10071. [DOI] [PubMed] [Google Scholar]
  • 89.Budhu A, Forgues M, Ye QH, Jia HL, He P, Zanetti KA, Kammula US, Chen Y, Qin LX, Tang ZY, Wang XW. Prediction of venous metastases, recurrence, and prognosis in hepatocellular carcinoma based on a unique immune response signature of the liver microenvironment. Cancer Cell. 2006;10:99–111. doi: 10.1016/j.ccr.2006.06.016. [DOI] [PubMed] [Google Scholar]
  • 90.Azechi H, Nishida N, Fukuda Y, Nishimura T, Minata M, Katsuma H, Kuno M, Ito T, Komeda T, Kita R, Takahashi R, Nakao K. Disruption of the p16/cyclin D1/retinoblastoma protein pathway in the majority of human hepatocellular carcinomas. Oncology. 2001;60:346–354. doi: 10.1159/000058531. [DOI] [PubMed] [Google Scholar]
  • 91.Yoshida T, Hisamoto T, Akiba J, Koga H, Nakamura K, Tokunaga Y, Hanada S, Kumemura H, Maeyama M, Harada M, Ogata H, Yano H, Kojiro M, Ueno T, Yoshimura A, Sata M. Spreds, inhibitors of the Ras/ERK signal transduction, are dysregulated in human hepatocellular carcinoma and linked to the malignant phenotype of tumors. Oncogene. 2006;25:6056–6066. doi: 10.1038/sj.onc.1209635. [DOI] [PubMed] [Google Scholar]
  • 92.Marchio A, Pineau P, Meddeb M, Terris B, Tiol-lais P, Bernheim A, Dejean A. Distinct chromosomal abnormality pattern in primary liver cancer of non-B, non-C patients. Oncogene. 2000;19:3733–3738. doi: 10.1038/sj.onc.1203713. [DOI] [PubMed] [Google Scholar]
  • 93.Huang XH, Sun LH, Lu DD, Sun Y, Ma U, Zhang XR, Huang J, Yu L. Codon 249 mutation in exon 7 of p53 gene in plasma DNA: maybe a new early diagnostic marker of hepatocellular carcinoma in Qidong risk area, China. World J Gastroenterol. 2003;9:692–695. doi: 10.3748/wjg.v9.i4.692. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.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. doi: 10.1093/jnci/92.2.148. [DOI] [PubMed] [Google Scholar]
  • 95.Hsu HC, Jeng YM, Mao TL, Chu JS, Lai PL, Peng SY. Beta-catenin mutations are associated with a subset of low-stage hepatocellular carcinoma negative for hepatitis B virus and with favorable prognosis. Am J Pathol. 2000;157:763–770. doi: 10.1016/s0002-9440(10)64590-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Mao TL, Chu JS, Jeng YM, Lai PL, Hsu HC. Expression of mutant nuclear beta-catenin correlates with non-invasive hepatocellular carcinoma, absence of portal vein spread, and good prognosis. J Pathol. 2001;193:95–101. doi: 10.1002/1096-9896(2000)9999:9999<::AID-PATH720>3.0.CO;2-3. [DOI] [PubMed] [Google Scholar]
  • 97.Terris B, Pineau P, Bregeaud L, Valla D, Belghiti J, Tiollais P, Degott C, Dejean A. Close correlation between beta-catenin gene alterations and nuclear accumulation of the protein in human hepatocellular carcinomas. Oncogene. 1999;18:6583–6588. doi: 10.1038/sj.onc.1203051. [DOI] [PubMed] [Google Scholar]
  • 98.Wong CM, Fan ST, Ng IO. beta-Catenin mutation and overexpression in hepatocellular carcinoma: clinicopathologic and prognostic significance. Cancer. 2001;92:136–145. doi: 10.1002/1097-0142(20010701)92:1<136::aid-cncr1301>3.0.co;2-r. [DOI] [PubMed] [Google Scholar]
  • 99.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 hepatocar-cinogenesis. Cancer Res. 1994;54:231–235. [PubMed] [Google Scholar]
  • 100.Debuire B, Paterlini P, Pontisso P, Basso G, May E. Analysis of the p53 gene in European hepatocellular carcinomas and hepatoblas-tomas. Oncogene. 1993;8:2303–2306. [PubMed] [Google Scholar]
  • 101.Lin Y, Shi CY, Li B, Soo BH, Mohammed-Ali S, Wee A, Oon CJ, Mack PO, Chan SH. Tumour suppressor p53 and Rb genes in human hepatocellular carcinoma. Ann Acad Med Singapore. 1996;25:22–30. [PubMed] [Google Scholar]
  • 102.Oda T, Tsuda H, Sakamoto M, Hirohashi S. Different mutations of the p53 gene in nodule-in-nodule hepatocellular carcinoma as a evidence for multistage progression. Cancer Lett. 1994;83:197–200. doi: 10.1016/0304-3835(94)90319-0. [DOI] [PubMed] [Google Scholar]
  • 103.Tanaka S, Toh Y, Adachi E, Matsumata T, Mori R, Sugimachi K. Tumor progression in hepatocellular carcinoma may be mediated by p53 mutation. Cancer Res. 1993;53:2884–2887. [PubMed] [Google Scholar]
  • 104.Sheen IS, Jeng KS, Wu JY. Is p53 gene mutation an indicatior of the biological behaviors of recurrence of hepatocellular carcinoma? World J Gastroenterol. 2003;9:1202–1207. doi: 10.3748/wjg.v9.i6.1202. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Jeng KS, Sheen IS, Chen BF, Wu JY. Is the p53 gene mutation of prognostic value in hepatocellular carcinoma after resection? Arch Surg. 2000;135:1329–1333. doi: 10.1001/archsurg.135.11.1329. [DOI] [PubMed] [Google Scholar]
  • 106.Chen GG, Merchant JL, Lai PB, Ho RL, Hu X, Okada M, Huang SF, Chui AK, Law DJ, Li YG, Lau WY, Li AK. Mutation of p53 in recurrent hepatocellular carcinoma and its association with the expression of ZBP-89. Am J Pathol. 2003;162:1823–1829. doi: 10.1016/S0002-9440(10)64317-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Zhang X, Xu HJ, Murakami Y, Sachse R, Ya-shima K, Hirohashi S, Hu SX, Benedict WF, Sekiya T. Deletions of chromosome 13q, mutations in Retinoblastoma 1, and retinoblastoma protein state in human hepatocellular carcinoma. Cancer Res. 1994;54:4177–4182. [PubMed] [Google Scholar]
  • 108.Hui AM, Sakamoto M, Kanai Y, Ino Y, Gotoh M, Yokota J, Hirohashi S. Inactivation of p16INK4 in hepatocellular carcinoma. Hepatology. 1996;24:575–579. doi: 10.1002/hep.510240319. [DOI] [PubMed] [Google Scholar]
  • 109.Yamamoto H, Itoh F, Fukushima H, Kaneto H, Sasaki S, Ohmura T, Satoh T, Karino Y, Endo T, Toyota J, Imai K. Infrequent widespread microsatellite instability in hepatocellular carcinomas. Int J Oncol. 2000;16:543–547. doi: 10.3892/ijo.16.3.543. [DOI] [PubMed] [Google Scholar]
  • 110.Kondo Y, Kanai Y, Sakamoto M, Mizokami M, Ueda R, Hirohashi S. Genetic instability and aberrant DNA methylation in chronic hepatitis and cirrhosis-A comprehensive study of loss of heterozygosity and microsatellite instability at 39 loci and DNA hypermethylation on 8 CpG islands in microdissected specimens from patients with hepatocellular carcinoma. Hepatology. 2000;32:970–979. doi: 10.1053/jhep.2000.19797. [DOI] [PubMed] [Google Scholar]
  • 111.Kaposi-Novak P, Lee JS, Gomez-Quiroz L, Cou-louarn C, Factor VM, Thorgeirsson SS. Met-regulated expression signature defines a subset of human hepatocellular carcinomas with poor prognosis and aggressive phenotype. J Clin Invest. 2006;116:1582–1595. doi: 10.1172/JCI27236. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Chung EJ, Sung YK, Farooq M, Kim Y, Im S, Tak WY, Hwang YJ, Kim Yl, Han HS, Kim JC, Kim MK. Gene expression profile analysis in human hepatocellular carcinoma by cDNA microarray. Mol Cells. 2002;14:382–387. [PubMed] [Google Scholar]
  • 113.Lau WY, Lai PB, Leung MF, Leung BC, Wong N, Chen G, Leung TW, Liew CT. Differential gene expression of hepatocellular carcinoma using cDNA microarray analysis. Oncol Res. 2000;12:59–69. doi: 10.3727/096504001108747530. [DOI] [PubMed] [Google Scholar]
  • 114.Li Y, Tang R, Xu H, Qiu M, Chen Q, Chen J, Fu Z, Ying K, Xie Y, Mao Y. Discovery and analysis of hepatocellular carcinoma genes using cDNA microarrays. J Cancer Res Clin Oncol. 2002;128:369–379. doi: 10.1007/s00432-002-0347-0. [DOI] [PubMed] [Google Scholar]
  • 115.Xu XR, Huang J, Xu ZG, Qian BZ, Zhu ZD, Yan Q, Cai T, Zhang X, Xiao HS, Qu J, Liu F, Huang QH, Cheng ZH, Li NG, Du JJ, Hu W, Shen KT, Lu G, Fu G, Zhong M, Xu SH, Gu WY, Huang W, Zhao XT, Hu GX, Gu JR, Chen Z, Han ZG. Insight into hepatocellular carcinogenesis at transcrip-tome level by comparing gene expression profiles of hepatocellular carcinoma with those of corresponding noncancerous liver. Proc Natl Acad Sci U S A. 2001;98:15089–15094. doi: 10.1073/pnas.241522398. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Smith MW, Yue ZN, Geiss GK, Sadovnikova NY, Carter VS, Boix L, Lazaro CA, Rosenberg GB, Bumgarner RE, Fausto N, Bruix J, Katze MG. Identification of novel tumor markers in hepatitis C virus-associated hepatocellular carcinoma. Cancer Res. 2003;63:859–864. [PubMed] [Google Scholar]
  • 117.Cheung ST, Chen X, Guan XY, Wong SY, Tai LS, Ng IO, So S, Fan ST. Identify metastasis-associated genes in hepatocellular carcinoma through clonality delineation for multinodular tumor. Cancer Res. 2002;62:4711–4721. [PubMed] [Google Scholar]
  • 118.Shirota Y, Kaneko S, Honda M, Kawai HF, Kobayashi K. Identification of differentially expressed genes in hepatocellular carcinoma with cDNA microarrays. Hepatology. 2001;33:832–840. doi: 10.1053/jhep.2001.23003. [DOI] [PubMed] [Google Scholar]
  • 119.Okabe H, Satoh S, Kato T, Kitahara O, Yana-gawa R, Yamaoka Y, Tsunoda T, Furukawa Y, Nakamura Y. Genome-wide analysis of gene expression in human hepatocellular carcinomas using cDNA microarray: identification of genes involved in viral carcinogenesis and tumor progression. Cancer Res. 2001;61:2129–2137. [PubMed] [Google Scholar]
  • 120.Kondoh N, Shuda M, Tanaka K, Wakatsuki T, Hada A, Yamamoto M. Enhanced expression of S8, L12, L23a, L27 and L30 ribosomal protein mRNAs in human hepatocellular carcinoma. Anticancer Res. 2001;21:2429–2433. [PubMed] [Google Scholar]
  • 121.Yamashita T, Kaneko S, Hashimoto S, Sato T, Nagai S, Toyoda N, Suzuki T, Kobayashi K, Matsushima K. Serial analysis of gene expression in chronic hepatitis C and hepatocellular carcinoma. Biochem Biophys Res Commun. 2001;282:647–654. doi: 10.1006/bbrc.2001.4610. [DOI] [PubMed] [Google Scholar]
  • 122.Lee JS, Thorgeirsson SS. Genome-scale profiling of gene expression in hepatocellular carcinoma: classification, survival prediction, and identification of therapeutic targets. Gas-troenterology. 2004;127:S51–55. doi: 10.1053/j.gastro.2004.09.015. [DOI] [PubMed] [Google Scholar]
  • 123.lizuka N, Oka M, Yamada-Okabe H, Mori N, Tamesa T, Okada T, Takemoto N, Tangoku A, Hamada K, Nakayama H, Miyamoto T, Uchi-mura S, Hamamoto Y. Comparison of gene expression profiles between hepatitis B virus-and hepatitis C virus-infected hepatocellular carcinoma by oligonucleotide microarray data on the basis of a supervised learning method. Cancer Res. 2002;62:3939–3944. [PubMed] [Google Scholar]
  • 124.Delpuech O, Trabut JB, Carnot F, Feuillard J, Brechot C, Kremsdorf D. Identification, using cDNA macroarray analysis, of distinct gene expression profiles associated with pathological and virological features of hepatocellular carcinoma. Oncogene. 2002;21:2926–2937. doi: 10.1038/sj.onc.1205392. [DOI] [PubMed] [Google Scholar]
  • 125.Li Y, Tang Y, Ye L, Liu B, Liu K, Chen J, Xue Q. Establishment of a hepatocellular carcinoma cell line with unique metastatic characteristics through in vivo selection and screening for metastasis-related genes through cDNA microar-ray. J Cancer Res Clin Oncol. 2003;129:43–51. doi: 10.1007/s00432-002-0396-4. [DOI] [PubMed] [Google Scholar]
  • 126.lizuka N, Oka M, Yamada-Okabe H, Nishida M, Maeda Y, Mori N, Takao T, Tamesa T, Tangoku A, Tabuchi H, Hamada K, Nakayama H, Ishitsuka H, Miyamoto T, Hirabayashi A, Uchi-mura S, Hamamoto Y. Oligonucleotide mi-croarray for prediction of early intrahepatic recurrence of hepatocellular carcinoma after curative resection. Lancet. 2003;361:923–929. doi: 10.1016/S0140-6736(03)12775-4. [DOI] [PubMed] [Google Scholar]
  • 127.Meng F, Henson R, Wehbe-Janek H, Ghoshal K, Jacob ST, Patel T. MicroRNA-21 regulates expression of the PTEN tumor suppressor gene in human hepatocellular cancer. Gastroenterology. 2007;133:647–658. doi: 10.1053/j.gastro.2007.05.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Fornari F, Gramantieri L, Ferracin M, Veronese A, Sabbioni S, Calin GA, Grazi GL, Giovannini C, Croce CM, Bolondi L, Negrini M. MiR-221 controls CDKN1C/p57 and CDKN1B/p27 expression in human hepatocellular carcinoma. Oncogene. 2008;27:5651–5661. doi: 10.1038/onc.2008.178. [DOI] [PubMed] [Google Scholar]
  • 129.Gramantieri L, Fornari F, Callegari E, Sabbioni S, Lanza G, Croce CM, Bolondi L, Negrini M. MicroRNA involvement in hepatocellular carcinoma. J Cell Mol Med. 2008;12:2189–2204. doi: 10.1111/j.1582-4934.2008.00533.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Gramantieri L, Fornari F, Ferracin M, Veronese A, Sabbioni S, Calin GA, Grazi GL, Croce CM, Bolondi L, Negrini M. MicroRNA-221 targets Bmf in hepatocellular carcinoma and correlates with tumor multifocality. Clin Cancer Res. 2009;15:5073–5081. doi: 10.1158/1078-0432.CCR-09-0092. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Budhu A, Jia HL, Forgues M, Liu CG, Goldstein D, Lam A, Zanetti KA, Ye QH, Qin LX, Croce CM, Tang ZY, Wang XW. Identification of metastasis-related microRNAs in hepatocellular carcinoma. Hepatology. 2008;47:897–907. doi: 10.1002/hep.22160. [DOI] [PubMed] [Google Scholar]
  • 132.Boyault S, Rickman DS, de Reynies A, Balabaud C, Rebouissou S, Jeannot E, Herault A, Saric J, Belghiti J, Franco D, Bioulac-Sage P, Laurent-Puig P, Zucman-Rossi J. Transcriptome classification of HCC is related to gene alterations and to new therapeutic targets. Hepatology. 2007;45:42–52. doi: 10.1002/hep.21467. [DOI] [PubMed] [Google Scholar]
  • 133.Hoshida Y, Villanueva A, Kobayashi M, Peix J, Chiang DY, Camargo A, Gupta S, Moore J, Wrobel MJ, Lerner J, Reich M, Chan JA, Glick-man JN, Ikeda K, Hashimoto M, Watanabe G, Daidone MG, Roayaie S, Schwartz M, Thung S, Salvesen HB, Gabriel S, Mazzaferro V, Bruix J, Friedman SL, Kumada H, Llovet JM, Golub TR. Gene expression in fixed tissues and outcome in hepatocellular carcinoma. N Engl J Med. 2008;359:1995–2004. doi: 10.1056/NEJMoa0804525. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Katoh H, Ojima H, Kokubu A, Saito S, Kondo T, Kosuge T, Hosoda F, Imoto I, Inazawa J, Hiroha-shi S, Shibata T. Genetically distinct and clinically relevant classification of hepatocellular carcinoma: putative therapeutic targets. Gastroenterology. 2007;133:1475–1486. doi: 10.1053/j.gastro.2007.08.038. [DOI] [PubMed] [Google Scholar]
  • 135.Chiang DY, Villanueva A, Hoshida Y, Peix J, Newell P, Minguez B, LeBlanc AC, Donovan DJ, Thung SN, Sole M, Tovar V, Alsinet C, Ramos AH, Barretina J, Roayaie S, Schwartz M, Wax-man S, Bruix J, Mazzaferro V, Ligon AH, Najfeld V, Friedman SL, Sellers WR, Meyerson M, Llovet JM. Focal gains of VEGFA and molecular classification of hepatocellular carcinoma. Cancer Res. 2008;68:6779–6788. doi: 10.1158/0008-5472.CAN-08-0742. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Lee JS, Chu IS, Heo J, Calvisi DF, Sun Z, Ros-kams T, Durnez A, Demetris AJ, Thorgeirsson SS. Classification and prediction of survival in hepatocellular carcinoma by gene expression profiling. Hepatology. 2004;40:667–676. doi: 10.1002/hep.20375. [DOI] [PubMed] [Google Scholar]
  • 137.Zender L, Kubicka S. Molecular pathogene-sis and targeted therapy of hepatocellular carcinoma. Onkologie. 2008;31:550–555. doi: 10.1159/000151586. [DOI] [PubMed] [Google Scholar]
  • 138.Midorikawa Y, Sugiyama Y, Aburatani H. Molecular targets for liver cancer therapy: From screening of target genes to clinical trials. Hepatol Res. 2009 doi: 10.1111/j.1872-034X.2009.00583.x. [DOI] [PubMed] [Google Scholar]
  • 139.Okabe H, Furukawa Y, Kato T, Hasegawa S, Yamaoka Y, Nakamura Y. Isolation of development and differentiation enhancing factor-like 1 (DDEFL1) as a drug target for hepatocellular carcinomas. Int J Oncol. 2004;24:43–48. [PubMed] [Google Scholar]
  • 140.Okabe H, Satoh S, Furukawa Y, Kato T, Hasegawa S, Nakajima Y, Yamaoka Y, Nakamura Y. Involvement of PEG10 in human hepatocellular carcinogenesis through interaction with SIAH1. Cancer Res. 2003;63:3043–3048. [PubMed] [Google Scholar]
  • 141.Yagyu R, Hamamoto R, Furukawa Y, Okabe H, Yamamura T, Nakamura Y. Isolation and characterization of a novel human gene, VANGL1, as a therapeutic target for hepatocellular carcinoma. Int J Oncol. 2002;20:1173–1178. [PubMed] [Google Scholar]
  • 142.Silva FP, Hamamoto R, Nakamura Y, Furukawa Y. WDRPUH, a novel WD-repeat-containing protein, is highly expressed in human hepatocellular carcinoma and involved in cell proliferation. Neoplasia. 2005;7:348–355. doi: 10.1593/neo.04544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Midorikawa Y, Ishikawa S, Iwanari H, Imamura T, Sakamoto H, Miyazono K, Kodama T, Makuu-chi M, Aburatani H. Glypican-3, overex-pressed in hepatocellular carcinoma, modulates FGF2 and BMP-7 signaling. Int J Cancer. 2003;103:455–465. doi: 10.1002/ijc.10856. [DOI] [PubMed] [Google Scholar]
  • 144.Pang EY, Bai AH, To KF, Sy SM, Wong NL, Lai PB, Squire JA, Wong N. Identification of PFTAIRE protein kinase 1, a novel cell division cycle-2 related gene, in the motile phenotype of hepatocellular carcinoma cells. Hepatology. 2007;46:436–445. doi: 10.1002/hep.21691. [DOI] [PubMed] [Google Scholar]
  • 145.Sy SM, Lai PB, Pang E, Wong NL, To KF, Johnson PJ, Wong N. Novel identification of zyxin upregulations in the motile phenotype of hepatocellular carcinoma. Mod Pathol. 2006;19:1108–1116. doi: 10.1038/modpathol.3800626. [DOI] [PubMed] [Google Scholar]
  • 146.Ip WK, Lai PB, Wong NL, Sy SM, Beheshti B, Squire JA, Wong N. Identification of PEG10 as a progression related biomarker for hepatocellular carcinoma. Cancer Lett. 2007;250:284–291. doi: 10.1016/j.canlet.2006.10.012. [DOI] [PubMed] [Google Scholar]
  • 147.Gramantieri L, Ferracin M, Fornari F, Veronese A, Sabbioni S, Liu CG, Calin GA, Giovannini C, Ferrazzi E, Grazi GL, Croce CM, Bolondi L, Negrini M. Cyclin Gl is a target of miR-122a, a microRNA frequently down-regulated in human hepatocellular carcinoma. Cancer Res. 2007;67:6092–6099. doi: 10.1158/0008-5472.CAN-06-4607. [DOI] [PubMed] [Google Scholar]
  • 148.Thomas M. Molecular targeted therapy for hepatocellular carcinoma. J Gastroenterol. 2009;44(Suppl 19):136–141. doi: 10.1007/s00535-008-2252-z. [DOI] [PubMed] [Google Scholar]
  • 149.Llovet JM, Ricci S, Mazzaferro V, Hilgard P, Gane E, Blanc JF, de Oliveira AC, Santoro A, Raoul JL, Forner A, Schwartz M, Porta C, Zeuzem S, Bolondi L, Greten TF, Galle PR, Seitz JF, Borbath I, Haussinger D, Giannaris T, Shan M, Moscovici M, Voliotis D, Bruix J. Soraf-enib in advanced hepatocellular carcinoma. N Engl J Med. 2008;359:378–390. doi: 10.1056/NEJMoa0708857. [DOI] [PubMed] [Google Scholar]
  • 150.Thomas MB, Chadha R, Glover K, Wang X, Morris J, Brown T, Rashid A, Dancey J, Ab-bruzzese JL. Phase 2 study of erlotinib in patients with unresectable hepatocellular carcinoma. Cancer. 2007;110:1059–1067. doi: 10.1002/cncr.22886. [DOI] [PubMed] [Google Scholar]
  • 151.Zhu AX, Stuart K, Blaszkowsky LS, Muzikansky A, Reitberg DP, Clark JW, Enzinger PC, Bhar-gava P, Meyerhardt JA, Horgan K, Fuchs CS, Ryan DP. Phase 2 study of cetuximab in patients with advanced hepatocellular carcinoma. Cancer. 2007;110:581–589. doi: 10.1002/cncr.22829. [DOI] [PubMed] [Google Scholar]
  • 152.Liu L, Cao Y, Chen C, Zhang X, McNabola A, Wilkie D, Wilhelm S, Lynch M, Carter C. Sorafenib blocks the RAF/MEK/ERK pathway, inhibits tumor angiogenesis, and induces tumor cell apoptosis in hepatocellular carcinoma model PLC/PRF/5. Cancer Res. 2006;66:11851–11858. doi: 10.1158/0008-5472.CAN-06-1377. [DOI] [PubMed] [Google Scholar]
  • 153.Faivre S, Raymond E, Boucher E, Douillard J, Lim HY, Kim JS, Zappa M, Lanzalone S, Lin X, Deprimo S, Harmon C, Ruiz-Garcia A, Lechuga MJ, Cheng AL. Safety and efficacy of sunit-inib in patients with advanced hepatocellular carcinoma: an open-label, multicentre, phase II study. Lancet Oncol. 2009;10:794–800. doi: 10.1016/S1470-2045(09)70171-8. [DOI] [PubMed] [Google Scholar]
  • 154.Papaetis GS, Syrigos KN. Sunitinib: a multi-targeted receptor tyrosine kinase inhibitor in the era of molecular cancer therapies. BioDrugs. 2009;23:377–389. doi: 10.2165/11318860-000000000-00000. [DOI] [PubMed] [Google Scholar]
  • 155.Huynh H, Ngo VC, Choo SP, Poon D, Koong HN, Thng CH, Toh HC, Zheng L, Ong LC, Jin Y, Song IC, Chang AP, Ong HS, Chung AY, Chow PK, Soo KC. Sunitinib (SUTENT, SU11248) suppresses tumor growth and induces apoptosis in xenograft models of human hepatocellular car- [DOI] [PubMed]

Articles from American Journal of Translational Research are provided here courtesy of e-Century Publishing Corporation

RESOURCES