Abstract
Gastric cancer is one of the most common malignancies and a leading cause of cancer mortality worldwide. The pathogenesis mechanisms of gastric cancer are still not fully clear. Inactivation of tumor suppressor genes and activation of oncogenes caused by genetic and epigenetic alterations are known to play significant roles in Carcinogenesis. Accumulating evidence has shown that epigenetic silencing of the tumor suppressor genes, particularly caused by hypermethylation of CpG islands in promoters, is critical to Carcinogenesis and metastasis. Here, we review the recent progress in the study of methylations of tumor suppressor genes involved in the pathogenesis of gastric cancer. We also briefly describe the mechanisms that induce tumor suppressor gene methylation and the status of translating these molecular mechanisms into clinical applications.
Keywords: Gastric cancer, methylation, tumor suppressor genes, epigenetics
Tumorigenesis is a well-known multi-step process involving inactivation of tumor suppressor genes and activation of oncogenes. Tumor suppressor genes can be inactivated by gene mutation, gene deletion, or gene silencing[1]. Evidence has shown that cancer is caused by both genetic and epigenetic alterations[2],[3]. Epigenetic information is defined as heritable information other than the DNA sequence[4]. Epigenetic changes, particularly DNA methylation, have been demonstrated to be an important factor in cancer initiation and progression. Methylation of CpG islands in a promoter region inhibits gene transcription by interfering with transcription initiation and serves as an alternative mechanism of inactivating tumor suppressor genes without gene mutations[1]. Some methylated genes identified in human cancers are classic tumor suppressor genes, which are inherited with one mutant defective allele. According to Knudson's two-hit model, complete inactivation of a tumor suppressor gene requires loss of function of both gene copies[5]. Epigenetic silencing of the remaining wild-type allele can be considered the second hit in this model. Thus, in addition to gene mutations, epigenetic gene silencing is another mechanism that fosters malignant transformation by reducing tumor suppressor gene activity and promoting aberrant activation of oncogenic signaling pathways. Epigenetic alterations may occur at different stages of tumorigenesis and malignant progression[6],[7], but some researchers believe that aberrant methylation takes place before genetic alterations. Identification of tumor suppressor genes silenced by CpG methylation is expected to reveal the molecular mechanism of tumorigenesis and potential tumor biomarkers.
Gastric cancer is one of the most common malignancies and a leading cause of cancer mortality worldwide[8]. In gastric cancers, tumor suppressor genes are inactivated more frequently by promoter methylation than by mutations[9]. A number of tumor suppressor genes, including hMLM1, p14, p15, p16, GSTP1, RASSF1, COX-2, APC, CDH1, CDH4, DAP-K, THBS1, TIMP-3, RARβ, MGMT, CHFR, DCC, RUNX3, TSLC1 and 14-3-3 sigma, are known to be silenced by hypermethylation in gastric cancer and have been reviewed before[10],[11]. Accumulation of aberrant methylations is thought to promote Carcinogenesis through activation of common cancer pathways. For instance, RAS, which regulates a signal transduction pathway linking plasma membrane receptors to many signals for growth, differentiation, and other functions, is frequently dysregulated in human neoplasms. RAS mutations have been found in only a small proportion of human gastric cancers, suggesting that other mechanisms may be involved in the activation of RAS signaling in gastric tumors. Indeed, methylation of the Ras-association domain family (RASSF) genes[12],[13] and the RAS protein activator-like 1 (RASAL1) gene[14] has been frequently detected and has led to activation of the RAS signaling pathway. Identification of methylated tumor suppressor genes and critical genetic pathways may impact the prognosis and treatment strategy selection for patients with gastric cancer.
We review here the recent progress in identifying methylated tumor suppressor genes involved in gastric cancer pathogenesis and the current status of translating that knowledge to clinical applications. Many methylated tumor suppressor genes have been identified in gastric cancer (Table 1). Because of their functional involvement in various cellular pathways that prevent cancer formation, inactivation of these genes through hypermethylation likely contributes to gastric cancer tumorigenesis. Here, we describe the genes whose functions in gastric Carcinogenesis have been proposed.
Table 1. Recently identified methylated tumor suppressor genes in gastric cancer.
Gene | Chromosomal location | Frequency of methylation (%) |
Basic function | Reference | |
Tumor tissue | Cell line | ||||
RASAL1 | 12q23-q24 | 9.5% (2/21) | 60% (6/10) | GAP1 family of GTPase-activating protein, acting as a suppressor of RAS function | [14] |
EDNRB | 13q22.3 | amean 50.42% | / | G protein coupled receptor superfamily | [15] |
LMX1a | 1q22-q23 | 82% (41/50) | 100% (5/5) | Nucleotide, transcription, regulation | [16] |
XRCC1 | 19q13.31 | 76.4% (26/34) | / | Nucleotide, repair, base excision repair | [17] |
TFPI2 | 7q | 83% (15/18) | 100% (9/9) | Playing a major role in cell migration and tumor invasion | [18] |
TFF2 | 21q22.3 | 83% (15/18) | / | / | [19] |
WWOX | 16q23.1 | 33% (24/73) | 40% (2/5) | Putatively involved in regulation of apoptosis | [20] |
TSP1 | 6p12.3 | 35.4% (34/96) | / | Cysteine-rich secretory protein (CRISP) family, motor/Contractile, structural protein | [21] |
HOPX | 4q12 | b84% (67/80) | / | Transcription factor | [22] |
GATA4/5 | 8p23.1/20q13.33 | 53.8% (43/80); 61.3% (49/80) | / | GATA zinc finger transcription factor family | [23] |
RECK | 9p13.3 | 47.5% (19 /40) | / | Acting as a negative regulator for matrix metalloproteinase 9 and 2 | [24] |
GADD45G | 9q22.1-q22.2 | c63% (63/100) | / | Ribosomal protein L7AE/gadd45 family, signaling growth factor, DNA damage response and cell growth arrest | [25] |
LRP1B | 2q21.2 | 61% (45/74) | 100% (4/4) | Low density lipoprotein receptor gene family (LDLR), transducer of extracellular signals, and may be involved in signal transduction | [26] |
HAI2/SPINT2 | 19q13.2 | 75% ( 30/40) | 100% (4/4) | Kunitz family of serine protease inhibitor, playing an important regulatory role in pericellular activation of hepatocyte growth factor/scatter factor (HGF/SF) | [27] |
UNC5C | 4q22.3 | 25% ( 9/36) | / | Dependence receptor family, unc-5 family | [28] |
RELN | 7q22.1 | 100% (15 /15) | 100% (9/9) | Reelin family, signal transduction | [29] |
Claudin-11 | 3q26.2-q26.3 | 100% (18/18) | 100% (5/5) | Claudin family, adhesion, major structural components of tight junction (TJ) strands | [30] |
CDH3 | 16q22.1 | 69% (25/36) | / | Cadherin superfamily of calcium dependent cell-cell adhesion glycoproteins | [31] |
CDH5 | 16q22.1 | 73% (11/15) | 100% (7/7) | Cadherin superfamily of calcium dependent cell-cell adhesion glycoproteins | [32] |
GRIK2 | 6q16.3 | 70% (19/27) | 75% (3/4) | Glutamate-gated ion channel (TC 1.A.10) family, receptor membrane, transport channel | [33] |
SLC19A3 | 2q36.3 | 51% (52/101) | 57% (4/7) | Reduced folate family of micronutrient transporter genes, may contribute to resistance to apoptosis in the tumors | [34] |
PTCMa | 9q22.32 | 32% (55/170) | / | Patched family, signaling, receptor | [35] |
S0CS6 | 18q22.2 | 46.8% (22/47) | 44.4% (4/9) | SOCS protein family, regulatory, signaling | [36] |
BTG4 | 11q23.1 | 73.7% (28/38) | 100% (5/5) | Tob/TBG1 family of growth inhibitory gene, antiproliferative activity, being able to induce G1 arrest | [37] |
Vimentin | 10p13 | 38% (14/37) | / | Intermediate filament family, structural protein, | [38] |
DLEC1 | 3p22.2 | 34% (30/89) | 100% (17/17) | Tumor suppressor gene(s), putatively involved in regulation of the expression of the telomere | [39] |
ZIC1 | 3q24 | 94.6% (35/37) | 100% (7/7) | Transcription factor | [40] |
Hsulf-1 | 8q13.3 | 81.3% (13/16) | 100% (3/3) | Sulfatase family, enzyme | [41] |
MAL | 2q11.1 | 65.8% (133/202) | / | MAL proteolipid family, cellular trafficking transport | [42] |
FBLN1 | 22q13.31 | 84% (86/102) | 71% (5/7) | Structural protein | [43] |
TCF4 | 18q21.2 | dmean 3.3% | / | Basic helix-loop-helix (BHLH) family of transcription factors, Wnt signaling pathway | [44] |
CACNA2D3 | 3p21.1 | 30% (24/80) | / | Alpha-2/delta subunit family, having voltage-gated ion channel activity | [45] |
DCBLD2 | 3q12.1 | amean 12.2% | 36.4% (4/11) | Neuropilin family, adhesion | [46] |
PKD1 | 16p13.3 | amean 19.5% | 72.7% (8/11) | Polycystin family, involved in cell-cell/matrix interaction and the regulation of several signalling pathways linked to cell proliferation | [47] |
TSPYL5 | 8q22.1 | 63.9% (23/3) | 70% (7/10) | Nucleosome assembly protein (NAP) family, involved in nucleosome assembly | [48] |
IQGAP2 | 5q13.3 | 47% (28/59) | 33.3% (3/9) | Ras GTPase-activating protein family, Wnt/beta-catenin signaling pathway | [49] |
TMS1 | 16p12-p11.2 | 32.1% (26/81) | / | CARD containing adaptor protein family, adaptor, signal transduction | [50] |
DAPK1 | 9q21.33 | 22.2% (18/81) | / | Protein kinase superfamily, enzyme, signaling, adhesion inhibitory effect, inducing death coupling the control of apoptosis to metastasis | [51] |
NMDAR2B | 12p13.1 | 61% (17/28) | 60% (6/10) | Glutamate-regulated family of ion channels, receptor membrane, transport channel, glutamate signaling pathway | [52] |
CD99 | Xp22.32 | 16.9% (15/89) | / | CD99 family, adhesion, antigen, signal transduction mediated death signaling (novel death pathway) | [53] |
ADRA1B | 5q33.1 | 70.6% (24/34) | / | G protein coupled receptor superfamily, signal transduction, mediating its action by association with G proteins that activate a phosphatidylinositol-calcium second messenger system | [54] |
VEGFC | 4q34.3 | 29% (9/31) | 36.4% (4/11) | PDGF/VEGF growth factor family, signal transduction | [55] |
SMAD4 | 18q21.2 | 5% (4/73) | 0% (0/9) | Dwarfin (DWA/B)/Smad family, DNA associated, transcription factor, critical mediator of TGFB and BMP signaling pathways | [56] |
NES1/hk10 | 19q13.33 | 90.9% (10/11) | 71.4% (5/7) | Kallikrein family, peptidase S1 family, enzyme | [57] |
IGFBP3 | 7p12.3 | 67% (16/24) | 46% (6/13) | Sterol desaturase family, signaling cytokine growth factor | [58] |
DFNA5 | 7p15.3 | 52% (46/89) | 100% (12/12) | Gasdermin family, plays a role in the TP53-regulated cellular response to genotoxic stress probably by cooperating with TP53 | [59] |
LIMS2 | 2q14.3 | 53% (51/96) | 80% (8/10) | PINCH protein family, novel LIM domain-containing gene, adhesion, regulatory | [60] |
HLTF | 3q24 | 38% (98/256) | / | SWI/SNF family of chromatin remodeling complex, actin-dependent regulator of chromatin structure CHFR-mediated downregulation of HLTF may help protect against cancer | [61] |
BMP2 | 20p12.3 | 42.9% (24/56) | 50% (2/4) | Transforming growth factor-beta (TGFB) superfamily, signaling growth factor, activating PI-3 kinase/Akt pathway | [62] |
EBF3 | 10q26.3 | 40.4% (42/104) | 60% (6/10) | Cell cycle progression and apoptosis | [63] |
SST | 3q28 | 93% (30/32) | 100% (7/7) | A primary inhibitor of gastrin-stimulated gastric acid secretion | [64] |
HIN1 | 5q35.3 | 57.8% (26/45) | 80% (4/5) | Uteroglobin/Clara cell secretory protein family, secretory, signaling cytokine, tumor suppressor may be mediated through the AKT signaling pathway | [65] |
IRX1 | 5p15.33 | 51.9% (8/15) | 100% (7/7) | TALE/IRO homeobox family, transcription factor | [67] |
CMTM3 | 16q22.1 | 44% (28/63) | 68.8% (11/16) | Chaperone/stress, signaling | [68] |
S0X2 | 3q26.33 | 16.2% (12/74) | 20% (2/10) | Transcription factor, can inhibit beta-catenin-driven reporter gene expression | [69] |
ZNF382 | 19q13.12 | 63.6%(7/11) | 100% (15/15) | Transcription factor | [70] |
UCHL1 | 4p13 | 77% (53/69) | 88.2% (15/17) | Ubiquitin carboxyl-terminal hydrolase family 1, enzyme, transcription factor | [71] |
hSRBC | 11p15.4 | 41% (46/111) | 73% (11/15) | STICK (substrates that interact with C-kinase) superfamily of PKC-binding protein, | [72] |
OPCML | 11q25 | 64% (7/11) | 100% (17/17) | Immunoglobulin superfamily, adhesion | [73] |
PLCD1 | 3p22.2 | 62% (61/98) | 84% (16/19) | Enzyme, signal transduction | [77] |
DLC1-i4 | 8p22 | 82% (9/11) | 12.5% (2/16) | RhoA pathway | [80] |
Sox17 | 8q11.23 | / | 100% (2/2) | Transcription factor, | [92] |
POPDC3 | 6q21 | 64.5% (15/18) | 73% (8/11) | Unknown/unspecified | [93] |
BVES | 6q21 | 69% (53/76) | 73% (8/11) | Interacting with GEFT, a GEF for Rho-family GTPases, and colocalizing in adult skeletal muscle | [93] |
PCDH10 | 4q28.3 | 82% (85/104) | 94% (16/17) | Protocadherin subfamily, cadherin superfamily of calcium dependent cell-cell adhesion glycoproteins | [95] |
FBP1 | 9q22.32 | 33% (33/101) | 57% (4/7) | FBPase class 1 family, enzyme, catalyzing the hydrolysis of fructose 1,6-bisphosphate to fructose 6-phosphate and inorganic phosphate | [97] |
LRRC3B | 3p24.1 | / | 90.9% (10/11) | IFN Signaling gene | [98] |
a, pyrosequencing; b, Q-MSP; c, high-resolution melting (HRM) analysis; d, methylight; e, bisulfit sequencing.
Methylated Tumor Suppressor Genes that Regulate Cell Cycle and Apoptosis
Dysregulated proliferation is a hallmark of tumor cells. Defects in many of the molecules that regulate the cell cycle have been implicated in cancer initiation and progression. These include p53, Rb, p107 and pRb2/p130, and CDK inhibitors (p15, p16, p18, p19, p21, p27), all of which act to prevent cell cycle progression until all repairs to damaged DNA have been completed. In this section, we will review tumor suppressor genes that regulate cell cycle and apoptosis that have been found to be methylated in gastric cancers.
Tumor suppressor genes which are most recently found to be methylated in gastric cancers regulate cell cycle progression and apoptosis. Kim et al.[63] reported that promoter methylation of early B-cell factor-3 (EBF3) was detected in 40.4% (42/104) of gastric cancer tissues but not in normal gastric tissues. The EBFs are a group of four highly conserved, DNA-binding transcription factors with an atypical zinc-finger domain and a helix-loop-helix domain. Functional analysis demonstrates that EBF3 represses gastric cancer cell growth and migration but activates cell cycle arrest and apoptosis. EBF3 up-regulates p21 and p27 in gastric cancer cells but represses CDK2, cyclin D1, cyclin D2, and Rb as previously reported in different types of cancers[64],[65].
Bcl-2-like 10 (BCL2L10) protein is a member of the Bcl-2 family. Interestingly, in addition to their central roles in apoptosis regulation, the Bcl-2 family of proteins influences the cell cycle, specifically the transition between quiescence and proliferation. BCL2L10 methylation was detected in 44.5% of gastric cancers and 21.34% of normal gastric mucosae (P < 0.05) by bisulphite sequencing[66]. The pro-apoptotic effect of BCL2L10 and growth promotion by siRNA targeting BCL2L10 in gastric cancer cells suggests that BCL2L10 may be a tumor suppressor by inducing apoptosis through mitochondrial pathways.
The iroquois homeobox protein 1 (IRX1) gene is located on chromosome 5p15.33, which is a cancer susceptibility locus. IRX1 transcription was suppressed by hypermethylation in gastric cancer, and restoring IRX1 expression in SGC-7901 and NCI-N87 gastric cancer cells inhibited tumor growth, invasion, and tumorigenesis in vitro and in vivo. IRX1 directly targets bradykinin receptor B2 (BDKRB2), an angiogenesis-related gene, as well as histone H2B type 2-E (HIST2H2BE) and fibroblast growth factor 70 (FGF7), cell proliferation and invasion related genes. Hypermethylation of IRX1 was detected not only in primary gastric cancer tissues but also in the peripheral blood cells of gastric cancer patients, suggesting that IRX1 could potentially serve as a biomarker for gastric cancer[67].
The CKLF-like MARVEL transmembrane domain-containing family (CMTM) is a novel family of proteins linking chemokines and the transmembrane-4 superfamily. Wang et al.[68] reported that CMTM3 is frequently methylated in many carcinoma cell lines and some primary tumors including gastric cancer, resulting in loss of its expression at both the mRNA and protein levels. Ectopic restoration of CMTM3 expression in tumor cells leads to the suppression of cell growth and induction of apoptosis with caspase-3 activation, suggesting that CMTM3 may function as a tumor suppressor.
Hypermethylation signals were also observed in some cultured and primary gastric cancers with little or no expression of transcription factor SOX2, a SOX transcription factor[69]. Among the 52 patients with advanced gastric cancers who were tested, those having SOX2 gene methylation in their cancer tissues had a significantly shorter survival than did those without the methylation (P = 0.006). Exogenous expression of SOX2 inhibited cell growth through apoptosis and cell cycle arrest. SOX2-overexpressing cells exhibited characteristics of apoptosis, such as DNA laddering and caspase-3 activation. Cell cycle arrest was associated with decreased levels of cyclin D1 and phosphorylated Rb as well as an increased p27 level.
Cheng et al.[70] reported that zinc-finger transcription factor ZNF382 functions as a tumor suppressor in many types of carcinomas including gastric cancer. Ectopic expression of ZNF382 in tumor cells in which the gene was silenced significantly reduced clonogenicity and proliferation and induced apoptosis. Cheng et al.[70] further found that ZNF382 inhibited NF-κB and AP-1 signaling and down-regulated the expression of multiple oncogenes, including MYC, MITF, HMGA2, CDK6, STAT3, STAT5B, ID1, and IKBKE, most likely through heterochromatin silencing.
Methylation of the Ubiquitin carboxyl-terminal hydrolase L1 (UCHL1)[71] was detected in primary digestive tumors, including 77% (53/69) of gastric carcinomas, but not in or occasionally in paired adjacent non-tumor tissues. Restoring UCHL1 expression in cell lines where the gene had been silenced significantly inhibited their growth and colony formation ability by inhibiting cell proliferation, causing cell cycle arrest in G2/M phase, and inducing apoptosis through the intrinsic caspase-dependent pathway. Moreover, UCHL1 directly interacts with p53 and stabilizes p53 through the ubiquitination.
Similarly, SDR-related gene product that binds to c-kinase (hSRBC), another novel methylated tumor suppressor gene in gastric cancer, increased the stability of p53 and expression of p53 target genes, such as p21Waf1, PUMA, and NOXA. hSRBC-mediated cell cycle arrest and apoptosis were abolished by blockade of p53 function[72]. These results suggested that epigenetic inactivation of hSRBC contributes to the malignant progression of gastric tumors, in part, through attenuated p53 response to stress. Moreover, opioid binding protein/cell adhesion molecule-like gene (OPCML) is also a stress- and p53-responsive gene, but this response is epigenetically impaired when the OPCML promoter becomes methylated. Ecotopic expression of OPCML led to significant inhibition of both anchorage-dependent and -independent growth of carcinoma cells in which the gene had been silenced[73].
As reviewed above, many methylated tumor suppressor genes in gastric cancer contribute to the dysregulation of the important aspects of tumorigenesis: cell cycle and apoptosis. The understanding of these molecular mechanisms will provide accesses to novel molecular therapeutic strategies for inhibiting oncogenic activity of signaling pathways.
Methylated Tumor Suppressor Genes that Regulate Cell Adhesion and Invasion
Morbidity in most cancer patients is not due to primary cancer but to metastatic disease. The high death rate of gastric cancer strikingly correlates with the high metastatic capacity of most gastric cancers. Thus, understanding tumor progression to the metastatic state and changes in highly aggressive cells is important for the development of novel approaches to diagnose and treat aggressive malignancies.
Detachment of cells with increased motility from a primary tumor is the first step of cancer invasion and metastasis. Gastric cancer cells with fibroblastoid morphological changes show increased motility and invasiveness due to decreased cell-cell adhesion, which is reminiscent of epithelial-mesenchymal transition (EMT) during embryonic development. The dynamic regulation of cell-cell adhesion is crucial for developmental processes including tumorigenesis. With reduced cell-cell adhesiveness, cancer cells can disobey the social order and result in destruction of the histological structure[74]. A common and early requirement for cell motility is actin polymerization, which drives the formation of cell protrusions that are used to adhere to the extracellular matrix, define the direction of migration and initiate cell crawling. Animal cells respond to signaling at the plasma membrane by remodeling their actin cytoskeleton[75]. On the other hand, the movement of cells into a tightly woven extracellular matrix also require an active proteolytic system, which can cleave a path for cell migration[76]. So the genes regulating the host cell cytoskeleton rearrangement and proteolytic activity also have important functions in cell morphogenesis and motility.
Our group reported that epigenetic inactivation of phospholipase C delta 1 (PLCD1) is common and tumor-specific in gastric cancer and that PLCD1 acts as a functional tumor suppressor in gastric Carcinogenesis[77]. Located at the important tumor suppressor locus 3p22, PLCD1 encodes an enzyme that regulates energy metabolism, calcium homeostasis, and intracellular movements. Ectopic expression of PLCD1 in gastric tumor cells with silenced PLCD1 dramatically inhibited clonogenicity and migration, possibly through down-regulation of MMP7 expression and hampered cyto-skeletal reorganization via phosphorylation and inactivation of cofilin.
Deleted in liver cancer 1 (DLC1) is another tumor suppressor gene involved in the regulation of cytoskeletal organization and other functions and is frequently methylated in many cancers, including gastric cancer[78],[79]. Recently, its new isoform, DLC1 isoform 4 (DLC1-i4), was also found to be silenced epigenetically, to have tumor inhibitory properties in multiple carcinomas, and to be regulated by p53[80].
E-cadherin is essential for cell-cell adhesion of epithelial cells. The down-regulation of E-cadherin may favor dissociation of cancer cells from one another, facilitating their invasion through the basal membrane. Although mutation and allelic loss have been confirmed as major mechanisms for E-cadherin (CDH1) gene inactivation in many malignancies[81], CDH1 promoter methylation could be frequently detected in gastric carcinoma[82].
As a co-partner of E-cadherin, β-catenin not only is critical for cell adhesion in the membrane and cytoplasm of cells, but also plays a role as a transcription activating protein in nuclei[83]. Aberrant activation of the Wnt/β-catenin signaling pathway is frequently found in many cancers, including gastric cancer[84],[85]. In addition to genetic deletions and point mutations, a number of negative regulators of Wnt signaling, including secreted frizzled-related protein (SFRP), dickkopf homolog 2 (DKK2), dickkopf homolog 2 (DKK3) and WNT inhibitory factor 1 (WIF1), were frequently methylated in gastric cancer[86]–[91]. Recently, other genes involving Wnt pathway were also reported as methylated tumor suppressor genes in gastric cancer. For example, SRY-box containing gene 17 (SOX17) was reported to be indispensable for embryonic development and a candidate tumor suppressor gene that antagonizes the canonical Wnt/β-catenin signaling pathway in colorectal cancer. Treatment with a demethylating agent induced SOX17 expression in gastric cancer cells, thus indicating the down-regulation of SOX17 by methylation. Transgenic expression of SOX17 suppressed dysplastic tumor development in the stomach of K19-Wnt1/C2mE mice by suppressing Wnt activity. These results suggested that SOX17 protects benign tumors from malignant progression at an early stage of tumorigenesis, and down-regulation of SOX17 contributes to malignant progression through promotion of Wnt activity[92].
The Popeye domain-containing (POPDC) genes BVES, POPDC2, and POPDC3 encode proteins that regulate cell-cell adhesion and cell migration during development. BVES and POPDC3 were reported to be hypermethylated in 69% and 64% of the gastric cancer tissues, respectively. Knockdown of POPDC3 in SNU-216 cells caused an increase in cell migration and invasion. Promoter hypermethylation is a causal event for long-term repression of BVES and POPDC3, whereas EGF stimulation is an immediate repression mechanism for both genes in gastric cancer. BVES, POPDC3, and E-cadherin mRNAs were down-regulated and Snail mRNA expression was up-regulated in EGF-induced EMT in SNU-216 cells[93].
Interestingly, accumulating evidence suggests that a major subfamily within the cadherin superfamily, protocadherins, frequently act as tumor suppressor genes. Inactivation of these genes through promoter methylation is significantly correlated with tumor development[81]. Recent studies have shown that the main structural and functional properties of protocadherins are distinct from those of classical cadherins. They may not have the strong cell-cell adhesion activity but do have other functions, such as mediating specificity of cell-cell interactions and signal transduction[94]. Yu et al.[95] reported that protocadherin 10 (PCDH10) was silenced or down-regulated in 94% (16 of 17) of gastric cancer cell lines. Furthermore, PCDH10 methylation was detected in 82% (85 of 104) of gastric tumors compared with 37% (38 of 104) of paired non-tumor tissues (P < 0.001). Re-expression of PCDH10 reduced colony formation in vitro and tumor growth in vivo. It also inhibited cell proliferation, induced cell apoptosis, and repressed cell invasion, possibly through up-regulation of the pro-apoptosis genes Fas, caspase 8, Jun, and CDKN1A; the anti-proliferation gene FGFR; and the anti-invasion gene HTATIP2. PCDH10 methylation at early stages of gastric Carcinogenesis is an independent prognostic factor.
Methylated Tumor Suppressor Genes with Other Functions
In cancer cells, glucose is often converted into lactic acid, which is known as the “Warburg effect.” The reason that cancer cells have a higher rate of aerobic glycolysis, but not oxidative phosphorylation, remains largely unclear[96]. Recently, fructose-1,6-bisphosphatase-1 (FBP1), which functions to antagonize glycolysis, was reported to be down-regulated through the NF-κB pathway in RAS-transformed NIH3T3 cells. FBP1 was hypermethylated in 57% (4/7) of gastric cancer cell lines and 33% (33/101) of gastric carcinomas. Inhibition of NF-κB restored FBP1 expression, partially through demethylation of FBP1 promoter. Restoration of FBP1 expression suppressed anchorage-independent growth, indicating the relevance of FBP1 down-regulation in Carcinogenesis[97].
Leucine-rich repeat-containing 3B (LRRC3B) is an evolutionary highly conserved leucine-rich repeat-containing protein, but its biological significance is unknown. Recently, LRRC3B was identified as a putative tumor suppressor gene and is reportedly silenced in gastric cancers by epigenetic mechanisms[98]. Stable transfection of LRRC3B in gastric cancer cell line SNU-601 inhibited anchorage-dependent and anchorage-independent colony formation. Moreover, LRRC3B expression suppressed tumorigenesis in nude mice. Microarray analysis of LRRC3B-expressing xenograft tumors showed induction of immune response-related genes and IFN signaling genes. Hematoxylin and eosin (H&E)-stained sections of LRRC3B-expressing xenograft tumors showed lymphocyte infiltration in the region. These results suggested that LRRC3B silencing in cancer may play an important role in tumor escape from immune surveillance.
DNA Methylation and Helicobacter Pylori Infection in Gastric Cancer
As reviewed above, aberrant DNA methylation of promoter CpG islands is one of the major inactivating mechanisms of tumor suppressor genes and is deeply involved in gastric Carcinogenesis. Moreover, in gastric cancer, DNA methylation from the pre-malignant till the most aggressive stage of cancer can be defined. For instance, many genes such as THBD, LOX, HRASLS, FLNc, and HAND1 were found to be infrequently methylated in non-cancerous gastric mucosae, in addition to their frequent methylation in cancers[99]. Similar findings were reported for CDH1, DAPK, p14, THBS1, and TIMP-1[100]. The presence of trace amounts of methylation in non-cancerous gastric mucosae suggested that some gastric carcinogens could have induced the methylation. The most important gastric carcinogenic factor is Helicobacter pylori (H. pylori) infection, which increases the risk of developing gastric cancers by 2.2- to 21.0-fold[101],[102]. Recent studies to quantify DNA methylation changes induced by the H. pylori infection in light of imflammatory reactions have been performed in both animal models and in the clinical setting. The findings suggest that H. pylori infection induces aberrant methylation in gastric mucosae and that levels of accumulated methylation are associated with gastric cancer risk. Notably, although H. pylori infection is important to trigger inflammation capable of inducing aberrant DNA methylation, some inflammation processes appear to be critical in induction of aberrant DNA methylation[103]–[105]. Indeed, the inflammation induced by H. pylori infection, not H. pylori itself, was critically involved in methylation induction. H. pylori may induce methylation of promoters containing CpG islands by release of reactive oxygen species (ROS) and nitric oxide (NO) and by activation of DNA methyltransferase[106],[107]. More recently, Sepulveda et al.[108] suggested that permanent DNA methylation after H. pylori eradication may occur in stem cells, thus promoting tumorigenesis. Considering the significant involvement of aberrant DNA methylation of CpG islands in human cancers, further identification of the inducing factors and mechanisms can be expected to provide novel targets for cancer prevention.
In addition to regional hypermethylation, global hypomethylation is also purported to be a hallmark of cancer cells[109] and is known to cause chromosomal instability as an early event during gastric Carcinogenesis[110]. However, Yoshida et al.[111] reported that H. pylori infection potently induces Alu and Sata hypomethylation in gastric mucosae as an early event during gastric Carcinogenesis, whereas global hypomethylation is present only in some individuals. Thus, the use of hypomethylation as a risk marker has not been considered realistic.
Clinical Application
Aberrant hypermethylation of promoter regions of specific genes is a key event in the formation and progression of cancer. Various tumor-specific aberrant DNA methylations have been identified. An accumulating number of studies have reported that gene silencing by DNA methylation may be established at an early stage in the multistep process of Carcinogenesis. Methylations of specific genes or methylation patterns of groups of genes were found to be associated with chemotherapy response and prognosis. Thus, aberrant DNA methylation can be applied to cancer diagnostics in three ways: as a marker to detect cancer cells or cancer-derived DNA, as a marker to predict prognosis, and as a biomarker for the assessment of therapeutic response. Methylation analysis has an advantage in that it can be performed using chemically stable DNA (compared to RNA). Moreover, detecting gene methylation is easier than detecting gene mutation because the exact location of a mutation is usually unknown, making it difficult to specifically amplify DNA molecules with an embedded mutation in excess of wild-type molecules. Detection of aberrant methylation can provide confirmation of the presence of intact cancer cells or cancer-derived DNA in bodily fluids, such as blood, urine, sputum, saliva, and stools. Recently, the methylations of BNIP3, CHFR, CYP1B1, MINT25, SFRP2, and RASSF2 were analyzed in 107 specimens of peritoneal fluid by quantitative methylation-specific polymerase chain reaction (MSP)[112]. The results showed that DNA methylation in periotoneal fluid is a possible marker for detecting occult neoplastic cells on the peritoneum. Methylation analysis along with a cytological examination might, therefore, improve the positive detection of cancer cells in the peritoneal fluid of gastric cancer patients.
The development of the bisulfate conversion technique that reproducibly changes unmethylated cytosines to uracil but leaves methylated cytosines unchanged rapidly increased progress in DNA methylation detection. Bisulfite sequencing, MSP, and combined bisulfite restriction analysis (COBRA) were all developed on the basis of bisulfite conversion[113]–[115] Among these technologies, MSP is a subjective, gel-based assay and cannot provide quantitative information. To date, several real-time MSP methods, such as bisulfite treatment in combination with MethyLight[116], quantitative multiplex-MSP(QM-MSP)[117],[118], or pyrosequencing[119], have been developed and used in DNA methylation studies. These methods have facilitated quantitative detection of minimal amounts of aberrant DNA methylation. For example, a recent study showed that one pyrosequencing analysis of global and site-specific DNA methylation in peripheral blood samples from 105 gastric cancer patients provides quantitative DNA methylation values that may serve as important prognostic indicators[120]. Methylated genes are predicted to be a new generation of cancer biomarkers. However, prior to clinical application, these findings require validation in prospective clinical studies.
Epigenetic change is heritable but has plasticity. During the last few decades, an increasing number of drugs targeting DNA methylation have been improved to increase efficacy and to decrease toxicity. DNA demethylating agents have been shown to be effective therapeutics for hematological malignancies. 5-azacyti-dine, the most successful epigenetic drug to date, is currently recommended as the first-line treatment of high-risk myelodysplastic syndromes (MDS). New methods for gene-specific epigenetic modification are being developed and tested in solid tumors[121].
Gastric cancer is a biologically heterogeneous disease with various molecular tumor subsets that likely respond distinctly to therapy. Discovery of biomarkers that improve disease characterization may make optimized or personalized therapy possible. Methylation of genes involved in DNA repair and maintaining genome integrity (e.g. MGMT, hMLH1, WRN, and FANCF), as well as genes involved in cell cycle checkpoints (e.g. CHFR, 14-3-3σ, CDK10, and p73) all reportedly influence sensitivity to chemotherapeutic drugs, suggesting that DNA methylation could serve as a molecular marker for predicting tumor responsiveness to chemotherapy[122]. An investigation of DNA methylation using specialized high-throughput platforms can potentially be applied to further stratify patients to individualized therapies.
Conclusions
Methylated tumor suppressor genes are being intensively investigated in gastric cancer, but the underlying functions and mechanisms need to be carefully examined. Future investigations of the connections between H. pylori infection, gastric cancer, and epigenetic changes will greatly expand our understanding. We believe that information obtained from studying DNA methylation will have an impact on cancer prevention, diagnostics, and treatment, and will contribute to cancer elimination. However, proper selection of methylation markers is crucial for sensitive and specific detection, and further technological advances are necessary in the future.
Acknowledgments
This study was supported by grants from National Natural Science Foundation of China (No. 30770920 and 81071651), Zhejiang Provincial Natural Science Foundation of China (No. R2100213, 2009C33142, Z2090056 and WKJ2009-2-028), and 973 Project (No. 2010CB834300).
References
- 1.Jones PA, Laird PW. Cancer epigenetics comes of age. Nat Genet. 1999;21:63–67. doi: 10.1038/5947. [DOI] [PubMed] [Google Scholar]
- 2.Jones PA, Baylin SB. The fundamental role of epigenetic events in cancer. Nat Rev Genet. 2002;3:415–428. doi: 10.1038/nrg816. [DOI] [PubMed] [Google Scholar]
- 3.Belinsky SA. Gene-promoter hypermethylation as a biomarker in lung cancer. Nat Rev Cancer. 2004;4:707–717. doi: 10.1038/nrc1432. [DOI] [PubMed] [Google Scholar]
- 4.Herman JG, Baylin SB. Gene silencing in cancer in association with promoter hypermethylation. N Engl J Med. 2003;349:2042–2054. doi: 10.1056/NEJMra023075. [DOI] [PubMed] [Google Scholar]
- 5.Knudson AG. Mutation and cancer: statistical study of retinoblastoma. Proc Natl Acad Sci USA. 1971;68:820–823. doi: 10.1073/pnas.68.4.820. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Sharma S, Kelly TK, Jones PA. Epigenetics in cancer. Carcinogenesis. 2010;31:27–36. doi: 10.1093/carcin/bgp220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Taby R, Issa JP. Cancer epigenetics. CA Cancer J Clin. 2010;60:376–92. doi: 10.3322/caac.20085. [DOI] [PubMed] [Google Scholar]
- 8.Bertuccio P, Chatenoud L, Levi F, et al. Recent patterns in gastric cancer: a global overview. Int J Cancer. 2009;125:666–673. doi: 10.1002/ijc.24290. [DOI] [PubMed] [Google Scholar]
- 9.Choi IS, Wu TT. Epigenetic alterations in gastric Carcinogenesis. Cell Res. 2005;15:247–254. doi: 10.1038/sj.cr.7290293. [DOI] [PubMed] [Google Scholar]
- 10.Sato F, Meltzer SJ. CpG island hypermethylation in progression of esophageal and gastric cancer. Cancer. 2006;106:483–493. doi: 10.1002/cncr.21657. [DOI] [PubMed] [Google Scholar]
- 11.Tamura G. Alterations of tumor suppressor and tumor-related genes in the development and progression of gastric cancer. World J Gastroenterol. 2006;12:192–198. doi: 10.3748/wjg.v12.i2.192. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Byun DS, Lee MG, Chae KS, et al. Frequent epigenetic inactivation of RASSF1A by aberrant promoter hypermethylation in human gastric adenocarcinoma. Cancer Res. 2001;61:7034–7038. [PubMed] [Google Scholar]
- 13.Guo W, Dong Z, Chen Z, et al. Aberrant CpG island hypermethylation of RASSF1A in gastric cardia adenocarcinoma. Cancer Invest. 2009;27:459–465. doi: 10.1080/07357900802620828. [DOI] [PubMed] [Google Scholar]
- 14.Seto M, Ohta M, Ikenoue T, et al. Reduced expression of RAS protein activator like-1 in gastric cancer. Int J Cancer. 2011;128:1293–1302. doi: 10.1002/ijc.25459. [DOI] [PubMed] [Google Scholar]
- 15.Tao K, Wu C, Wu K, et al. Quantitative analysis of promoter methylation of the EDNRB gene in gastric cancer. Med Oncol. 2012;29:107–112. doi: 10.1007/s12032-010-9805-8. [DOI] [PubMed] [Google Scholar]
- 16.Dong W, Feng L, Xie Y, et al. Hypermethylation-mediated reduction of LMX1A expression in gastric cancer. Cancer Sci. 2011;102:361–366. doi: 10.1111/j.1349-7006.2010.01804.x. [DOI] [PubMed] [Google Scholar]
- 17.Wang P, Tang JT, Peng YS, et al. XRCC1 downregulated through promoter hypermethylation is involved in human gastric Carcinogenesis. J Dig Dis. 2010;11:343–351. doi: 10.1111/j.1751-2980.2010.00459.x. [DOI] [PubMed] [Google Scholar]
- 18.Takada H, Wakabayashi N, Dohi O, et al. Tissue factor pathway inhibitor 2 (TFPI2) is frequently silenced by aberrant promoter hypermethylation in gastric cancer. Cancer Genet Cytogenet. 2010;197:16–24. doi: 10.1016/j.cancergencyto.2009.11.004. [DOI] [PubMed] [Google Scholar]
- 19.Hong SJ, Oh JH, Jung YC, et al. DNA methylation patterns of ulcer-healing genes associated with the normal gastric mucosa of gastric cancers. J Korean Med Sci. 2010;25:405–417. doi: 10.3346/jkms.2010.25.3.405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Maeda N, Semba S, Nakayama S, et al. Loss of WW domain-containing oxidoreductase expression in the progression and development of gastric carcinoma: clinical and histopathologic correlations. Virchows Arch. 2010;457:423–432. doi: 10.1007/s00428-010-0956-y. [DOI] [PubMed] [Google Scholar]
- 21.Guo W, Dong Z, He M, et al. Aberrant methylation of thrombospondin-1 and its association with reduced expression in gastric cardia adenocarcinoma. J Biomed Biotechnol. 2010;2010:721485. doi: 10.1155/2010/721485. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Ooki A, Yamashita K, Kikuchi S, et al. Potential utility of HOP homeobox gene promoter methylation as a marker of tumor aggressiveness in gastric cancer. Oncogene. 2010;29:3263–3275. doi: 10.1038/onc.2010.76. [DOI] [PubMed] [Google Scholar]
- 23.Wen XZ, Akiyama Y, Pan KF, et al. Methylation of GATA-4 and GATA-5 and development of sporadic gastric carcinomas. World J Gastroenterol. 2010;16:1201–1208. doi: 10.3748/wjg.v16.i10.1201. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Du YY, Dai DQ, Yang Z. Role of RECK methylation in gastric cancer and its clinical significance. World J Gastroenterol. 2010;16:904–908. doi: 10.3748/wjg.v16.i7.904. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Zhang W, Li T, Shao Y, et al. Semi-quantitative detection of GADD45-gamma methylation levels in gastric, colorectal and pancreatic cancers using methylation-sensitive high-resolution melting analysis. J Cancer Res Clin Oncol. 2010;136:1267–1273. doi: 10.1007/s00432-010-0777-z. [DOI] [PubMed] [Google Scholar]
- 26.Lu YJ, Wu CS, Li HP, et al. Aberrant methylation impairs low density lipoprotein receptor-related protein 1B tumor suppressor function in gastric cancer. Genes Chromosomes Cancer. 2010;49:412–424. doi: 10.1002/gcc.20752. [DOI] [PubMed] [Google Scholar]
- 27.Dong W, Chen X, Xie J, et al. Epigenetic inactivation and tumor suppressor activity of HAI-2/SPINT2 in gastric cancer. Int J Cancer. 2010;127:1526–1534. doi: 10.1002/ijc.25161. [DOI] [PubMed] [Google Scholar]
- 28.Hibi K, Sakata M, Sakuraba K, et al. Changes in UNC5C gene methylation during human gastric Carcinogenesis. Anticancer Res. 2009;29:4397–4399. [PubMed] [Google Scholar]
- 29.Dohi O, Takada H, Wakabayashi N, et al. Epigenetic silencing of RELN in gastric cancer. Int J Oncol. 2010;36:85–92. [PubMed] [Google Scholar]
- 30.Agarwal R, Mori Y, Cheng Y, et al. Silencing of claudin-11 is associated with increased invasiveness of gastric cancer cells. PLoS One. 2009;4:e8002. doi: 10.1371/journal.pone.0008002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Hibi K, Kitamura YH, Mizukami H, et al. Frequent CDH3 demethylation in advanced gastric carcinoma. Anticancer Res. 2009;29:3945–3947. [PubMed] [Google Scholar]
- 32.Wang X, Lau KK, So LK, et al. CHD5 is down-regulated through promoter hypermethylation in gastric cancer. J Biomed Sci. 2009;16:95. doi: 10.1186/1423-0127-16-95. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Wu CS, Lu YJ, Li HP, Hsueh C, et al. Glutamate receptor, ionotropic, kainate 2 silencing by DNA hypermethylation possesses tumor suppressor function in gastric cancer. Int J Cancer. 2010;126:2542–2552. doi: 10.1002/ijc.24958. [DOI] [PubMed] [Google Scholar]
- 34.Liu X, Lam EK, Wang X, et al. Promoter hypermethylation mediates downregulation of thiamine receptor SLC19A3 in gastric cancer. Tumour Biol. 2009;30:242–248. doi: 10.1159/000243767. [DOI] [PubMed] [Google Scholar]
- 35.Du P, Ye HR, Gao J, et al. Methylation of PTCH1a gene in a subset of gastric cancers. World J Gastroenterol. 2009;15:3799–3806. doi: 10.3748/wjg.15.3799. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Lai RH, Hsiao YW, Wang MJ, et al. SOCS6, down-regulated in gastric cancer, inhibits cell proliferation and colony formation. Cancer Lett. 2010;288:75–85. doi: 10.1016/j.canlet.2009.06.025. [DOI] [PubMed] [Google Scholar]
- 37.Dong W, Tu S, Xie J, et al. Frequent promoter hypermethylation and transcriptional downregulation of BTG4 gene in gastric cancer. Biochem Biophys Res Commun. 2009;387:132–138. doi: 10.1016/j.bbrc.2009.06.140. [DOI] [PubMed] [Google Scholar]
- 38.Kitamura YH, Shirahata A, Sakata M, et al. Frequent methylation of Vimentin in well-differentiated gastric carcinoma. Anticancer Res. 2009;29:2227–2229. [PubMed] [Google Scholar]
- 39.Ying J, Poon FF, Yu J, et al. DLEC1 is a functional 3p22.3 tumour suppressor silenced by promoter CpG methylation in colon and gastric cancers. Br J Cancer. 2009;100:663–669. doi: 10.1038/sj.bjc.6604888. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Wang LJ, Jin HC, Wang X, et al. ZIC1 is downregulated through promoter hypermethylation in gastric cancer. Biochem Biophys Res Commun. 2009;379:959–963. doi: 10.1016/j.bbrc.2008.12.180. [DOI] [PubMed] [Google Scholar]
- 41.Chen Z, Fan JQ, Li J, et al. Promoter hypermethylation correlates with the Hsulf-1 silencing in human breast and gastric cancer. Int J Cancer. 2009;124:739–744. doi: 10.1002/ijc.23960. [DOI] [PubMed] [Google Scholar]
- 42.Buffart TE, Overmeer RM, Steenbergen RD, et al. MAL promoter hypermethylation as a novel prognostic marker in gastric cancer. Br J Cancer. 2008;99:1802–1807. doi: 10.1038/sj.bjc.6604777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Cheng YY, Jin H, Liu X, et al. Fibulin 1 is downregulated through promoter hypermethylation in gastric cancer. Br J Cancer. 2008;99:2083–2087. doi: 10.1038/sj.bjc.6604760. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Kim SK, Jang HR, Kim JH, et al. CpG methylation in exon 1 of transcription factor 4 increases with age in normal gastric mucosa and is associated with gene silencing in intestinal-type gastric cancers. Carcinogenesis. 2008;29:1623–1631. doi: 10.1093/carcin/bgn110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Wanajo A, Sasaki A, Nagasaki H, et al. Methylation of the calcium channel-related gene, CACNA2D3, is frequent and a poor prognostic factor in gastric cancer. Gastroenterology. 2008;135:580–590. doi: 10.1053/j.gastro.2008.05.041. [DOI] [PubMed] [Google Scholar]
- 46.Kim M, Lee KT, Jang HR, et al. Epigenetic down-regulation and suppressive role of DCBLD2 in gastric cancer cell proliferation and invasion. Mol Cancer Res. 2008;6:222–230. doi: 10.1158/1541-7786.MCR-07-0142. [DOI] [PubMed] [Google Scholar]
- 47.Kim M, Jang HR, Kim JH, et al. Epigenetic inactivation of protein kinase D1 in gastric cancer and its role in gastric cancer cell migration and invasion. Carcinogenesis. 2008;29:629–637. doi: 10.1093/carcin/bgm291. [DOI] [PubMed] [Google Scholar]
- 48.Jung Y, Park J, Bang YJ, et al. Gene silencing of TSPYL5 mediated by aberrant promoter methylation in gastric cancers. Lab Invest. 2008;88:153–160. doi: 10.1038/labinvest.3700706. [DOI] [PubMed] [Google Scholar]
- 49.Jin SH, Akiyama Y, Fukamachi H, et al. IQGAP2 inactivation through aberrant promoter methylation and promotion of invasion in gastric cancer cells. Int J Cancer. 2008;122:1040–1046. doi: 10.1002/ijc.23181. [DOI] [PubMed] [Google Scholar]
- 50.Moriai R, Tsuji N, Kobayashi D, et al. A proapoptotic caspase recruitment domain protein gene, TMS1, is hypermethylated in human breast and gastric cancers. Anticancer Res. 2002;22:4163–4168. [PubMed] [Google Scholar]
- 51.Milne AN, Sitarz R, Carvalho R, et al. Molecular analysis of primary gastric cancer, corresponding xenografts, and 2 novel gastric carcinoma cell lines reveals novel alterations in gastric Carcinogenesis. Hum Pathol. 2007;38:903–913. doi: 10.1016/j.humpath.2006.12.010. [DOI] [PubMed] [Google Scholar]
- 52.Liu JW, Kim MS, Nagpal J, et al. Quantitative hypermethylation of NMDAR2B in human gastric cancer. Int J Cancer. 2007;121:1994–2000. doi: 10.1002/ijc.22934. [DOI] [PubMed] [Google Scholar]
- 53.Lee JH, Kim SH, Wang LH, et al. Clinical significance of CD99 down-regulation in gastric adenocarcinoma. Clin Cancer Res. 2007;13:2584–2591. doi: 10.1158/1078-0432.CCR-06-1785. [DOI] [PubMed] [Google Scholar]
- 54.Noda H, Miyaji Y, Nakanishi A, et al. Frequent reduced expression of alpha-1B-adrenergic receptor caused by aberrant promoter methylation in gastric cancers. Br J Cancer. 2007;96:383–390. doi: 10.1038/sj.bjc.6603555. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Matsumura S, Oue N, Mitani Y, et al. DNA demethylation of vascular endothelial growth factor-C is associated with gene expression and its possible involvement of lymphangiogenesis in gastric cancer. Int J Cancer. 2007;120:1689–1695. doi: 10.1002/ijc.22433. [DOI] [PubMed] [Google Scholar]
- 56.Wang LH, Kim SH, Lee JH, et al. Inactivation of SMAD4 tumor suppressor gene during gastric carcinoma progression. Clin Cancer Res. 2007;13:102–110. doi: 10.1158/1078-0432.CCR-06-1467. [DOI] [PubMed] [Google Scholar]
- 57.Huang W, Zhong J, Wu LY, et al. Downregulation and CpG island hypermethylation of NES1/hK10 gene in the pathogenesis of human gastric cancer. Cancer Lett. 2007;251:78–85. doi: 10.1016/j.canlet.2006.11.006. [DOI] [PubMed] [Google Scholar]
- 58.Tomii K, Tsukuda K, Toyooka S, et al. Aberrant promoter methylation of insulin-like growth factor binding protein-3 gene in human cancers. Int J Cancer. 2007;120:566–573. doi: 10.1002/ijc.22341. [DOI] [PubMed] [Google Scholar]
- 59.Akino K, Toyota M, Suzuki H, et al. Identification of DFNA5 as a target of epigenetic inactivation in gastric cancer. Cancer Sci. 2007;98:88–95. doi: 10.1111/j.1349-7006.2006.00351.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Kim SK, Jang HR, Kim JH, et al. The epigenetic silencing of LIMS2 in gastric cancer and its inhibitory effect on cell migration. Biochem Biophys Res Commun. 2006;349:1032–1040. doi: 10.1016/j.bbrc.2006.08.128. [DOI] [PubMed] [Google Scholar]
- 61.Oue N, Mitani Y, Motoshita J, et al. Accumulation of DNA methylation is associated with tumor stage in gastric cancer. Cancer. 2006;106:1250–1259. doi: 10.1002/cncr.21754. [DOI] [PubMed] [Google Scholar]
- 62.Wen XZ, Akiya ma Y, Baylin SB, et al. Frequent epigenetic silencing of the bone morphogenetic protein 2 gene through methylation in gastric carcinomas. Oncogene. 2006;25:2666–2673. doi: 10.1038/sj.onc.1209297. [DOI] [PubMed] [Google Scholar]
- 63.Kim J, Min SY, Lee HE, et al. Aberrant DNA methylation and tumor suppressive activity of the EBF3 gene in gastric carcinoma. Int J Cancer. 2012;130:817–826. doi: 10.1002/ijc.26038. [DOI] [PubMed] [Google Scholar]
- 64.Liao D. Emerging roles of the EBF family of transcription factors in tumor suppression. Mol Cancer Res. 2009;7:1893–1901. doi: 10.1158/1541-7786.MCR-09-0229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Zhao LY, Niu Y, Santiago A, et al. An EBF3-mediated transcriptional program that induces cell cycle arrest and apoptosis. Cancer Res. 2006;66:9445–9452. doi: 10.1158/0008-5472.CAN-06-1713. [DOI] [PubMed] [Google Scholar]
- 66.Mikata R, Fukai K, Imazeki F, et al. BCL2L10 is frequently silenced by promoter hypermethylation in gastric cancer. Oncol Rep. 2010;23:1701–1708. doi: 10.3892/or_00000814. [DOI] [PubMed] [Google Scholar]
- 67.Guo X, Liu W, Pan Y, et al. Homeobox gene IRX1 is a tumor suppressor gene in gastric carcinoma. Oncogene. 2010;29:3908–3920. doi: 10.1038/onc.2010.143. [DOI] [PubMed] [Google Scholar]
- 68.Wang Y, Li J, Cui Y, et al. CMTM3, located at the critical tumor suppressor locus 16q22.1, is silenced by CpG methylation in carcinomas and inhibits tumor cell growth through inducing apoptosis. Cancer Res. 2009;69:5194–5201. doi: 10.1158/0008-5472.CAN-08-3694. [DOI] [PubMed] [Google Scholar]
- 69.Otsubo T, Akiyama Y, Yanagihara K, et al. SOX2 is frequently downregulated in gastric cancers and inhibits cell growth through cell-cycle arrest and apoptosis. Br J Cancer. 2008;98:824–831. doi: 10.1038/sj.bjc.6604193. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Cheng Y, Geng H, Cheng SH, et al. KRAB zinc finger protein ZNF382 is a proapoptotic tumor suppressor that represses multiple oncogenes and is commonly silenced in multiple carcinomas. Cancer Res. 2010;70:6516–6526. doi: 10.1158/0008-5472.CAN-09-4566. [DOI] [PubMed] [Google Scholar]
- 71.Yu J, Tao Q, Cheung KF, et al. Epigenetic identification of Ubiquitin carboxyl-terminal hydrolase L1 as a functional tumor suppressor and biomarker for hepatocellular carcinoma and other digestive tumors. Hepatology. 2008;48:508–518. doi: 10.1002/hep.22343. [DOI] [PubMed] [Google Scholar]
- 72.Lee JH, Byun DS, Lee MG, et al. Frequent epigenetic inactivation of hSRBC in gastric cancer and its implication in attenuated p53 response to stresses. Int J Cancer. 2008;122:1573–1584. doi: 10.1002/ijc.23166. [DOI] [PubMed] [Google Scholar]
- 73.Cui Y, Ying Y, van Hasselt A, et al. OPCML is a broad tumor suppressor for multiple carcinomas and lymphomas with frequently epigenetic inactivation. PLoS One. 2008;3:e2990. doi: 10.1371/journal.pone.0002990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Hirohashi S, Kanai Y. Cell adhesion system and human cancer morphogenesis. Cancer Sci. 2003;94:575–581. doi: 10.1111/j.1349-7006.2003.tb01485.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Ridley AJ, Hall A. The small GTP-binding protein Rho regulates the assembly of focal adhesions and actin stress bers in response to growth factors. Cell. 1992;70:389–399. doi: 10.1016/0092-8674(92)90163-7. [DOI] [PubMed] [Google Scholar]
- 76.Singer CF, Marbaix E, Lemoine P, et al. Local cytokines induce differential expression of matrix metallo-proteinases but not their tissue inhibitors in human endometrial broblasts. Eur J Biochem. 1999;259:40–45. doi: 10.1046/j.1432-1327.1999.00001.x. [DOI] [PubMed] [Google Scholar]
- 77.Hu XT, Zhang FB, Fan YC, et al. Phospholipase C delta 1 is a novel 3p22.3 tumor suppressor involved in cytoskeleton organization, with its epigenetic silencing correlated with high-stage gastric cancer. Oncogene. 2009;28:2466–2475. doi: 10.1038/onc.2009.92. [DOI] [PubMed] [Google Scholar]
- 78.Kim TY, Jong HS, Song SH, et al. Transcriptional silencing of the DLC-1 tumor suppressor gene by epigenetic mechanism in gastric cancer cells. Oncogene. 2003;22:3943–3951. doi: 10.1038/sj.onc.1206573. [DOI] [PubMed] [Google Scholar]
- 79.Yuan BZ, Durkin ME, Popescu NC. Promoter hypermethylation of DLC-1, a candidate tumor suppressor gene, in several common human cancers. Cancer Genet Cytogenet. 2003;140:113–117. doi: 10.1016/s0165-4608(02)00674-x. [DOI] [PubMed] [Google Scholar]
- 80.Low JS, Tao Q, Ng KM, et al. A novel isoform of the 8p22 tumor suppressor gene DLC1 suppresses tumor growth and is frequently silenced in multiple common tumors. Oncogene. 2011;30:1923–1935. doi: 10.1038/onc.2010.576. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Berx G, van Roy F. Involvement of members of the cadherin superfamily in cancer. Cold Spring Harb Perspect Biol. 2009;1:a003129. doi: 10.1101/cshperspect.a003129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Graziano F, Arduini F, Ruzzo A, et al. Combined analysis of E-cadherin gene (CDH1) promoter hypermethylation and E-cadherin protein expression in patients with gastric cancer: implications for treatment with demethylating drugs. Ann Oncol. 2004;15:489–492. doi: 10.1093/annonc/mdh108. [DOI] [PubMed] [Google Scholar]
- 83.Cadigan KM, Nusse R. Wnt signaling: a common theme in animal development. Genes Dev. 1997;11:3286–3305. doi: 10.1101/gad.11.24.3286. [DOI] [PubMed] [Google Scholar]
- 84.Wu WK, Cho CH, Lee CW, et al. Dysregulation of cellular signaling in gastric cancer. Cancer Lett. 2010;295:144–153. doi: 10.1016/j.canlet.2010.04.025. [DOI] [PubMed] [Google Scholar]
- 85.Kolligs FT, Bommer G, G?ke B. Wnt/beta-catenin/tcf signaling: a critical pathway in gastrointestinal tumorigenesis. Digestion. 2002;66:131–144. doi: 10.1159/000066755. [DOI] [PubMed] [Google Scholar]
- 86.Nojima M, Suzuki H, Toyota M, et al. Frequent epigenetic inactivation of SFRP genes and constitutive activation of Wnt signaling in gastric cancer. Oncogene. 2007;26:4699–4713. doi: 10.1038/sj.onc.1210259. [DOI] [PubMed] [Google Scholar]
- 87.Zhao CH, Bu XM, Zhang N. Hypermethylation and aberrant expression of Wnt antagonist secreted frizzled-related protein 1 in gastric cancer. World J Gastroenterol. 2007;13:2214–2217. doi: 10.3748/wjg.v13.i15.2214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Cheng YY, Yu J, Wong YP, et al. Frequent epigenetic inactivation of secreted frizzled-related protein 2 (SFRP2) by promoter methylation in human gastric cancer. Br J Cancer. 2007;97:895–901. doi: 10.1038/sj.bjc.6603968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Sato H, Suzuki H, Toyota M, et al. Frequent epigenetic inactivation of DICKKOPF family genes in human gastrointestinal tumors. Carcinogenesis. 2007;28:2459–2466. doi: 10.1093/carcin/bgm178. [DOI] [PubMed] [Google Scholar]
- 90.Yu J, Tao Q, Cheng YY, et al. Promoter methylation of the Wnt/beta-catenin signaling antagonist Dkk-3 is associated with poor survival in gastric cancer. Cancer. 2009;115:49–60. doi: 10.1002/cncr.23989. [DOI] [PubMed] [Google Scholar]
- 91.Taniguchi H, Yamamoto H, Hirata T, et al. Frequent epigenetic inactivation of Wnt inhibitory factor-1 in human gastrointestinal cancers. Oncogene. 2005;24:7946–7952. doi: 10.1038/sj.onc.1208910. [DOI] [PubMed] [Google Scholar]
- 92.Du YC, Oshima H, Oguma K, et al. Induction and down-regulation of Sox17 and its possible roles during the course of gastrointestinal tumorigenesis. Gastroenterology. 2009;137:1346–1357. doi: 10.1053/j.gastro.2009.06.041. [DOI] [PubMed] [Google Scholar]
- 93.Kim M, Jang HR, Haam K, et al. Frequent silencing of popeye domain-containing genes, BVES and POPDC3, is associated with promoter hypermethylation in gastric cancer. Carcinogenesis. 2010;31:1685–1693. doi: 10.1093/carcin/bgq144. [DOI] [PubMed] [Google Scholar]
- 94.Morishita H, Yagi T. Protocadherin family: diversity, structure, and function. Curr Opin Cell Biol. 2007;19:584–592. doi: 10.1016/j.ceb.2007.09.006. [DOI] [PubMed] [Google Scholar]
- 95.Yu J, Cheng YY, Tao Q, et al. Methylation of protocadherin 10, a novel tumor suppressor, is associated with poor prognosis in patients with gastric cancer. Gastroenterology. 2009;136:640–651. doi: 10.1053/j.gastro.2008.10.050. [DOI] [PubMed] [Google Scholar]
- 96.Cairns RA, Harris IS, Mak TW. Regulation of cancer cell metabolism. Nat Rev Cancer. 2011;11:85–95. doi: 10.1038/nrc2981. [DOI] [PubMed] [Google Scholar]
- 97.Liu X, Wang X, Zhang J, et al. Warburg effect revisited: an epigenetic link between glycolysis and gastric Carcinogenesis. Oncogene. 2010;29:442–450. doi: 10.1038/onc.2009.332. [DOI] [PubMed] [Google Scholar]
- 98.Kim M, Kim JH, Jang HR, et al. LRRC3B, encoding a leucine-rich repeat-containing protein, is a putative tumor suppressor gene in gastric cancer. Cancer Res. 2008;68:7147–7155. doi: 10.1158/0008-5472.CAN-08-0667. [DOI] [PubMed] [Google Scholar]
- 99.Kaneda A, Kaminishi M, Yanagihara K, et al. Identification of silencing of nine genes in human gastric cancers. Cancer Res. 2002;62:6645–6650. [PubMed] [Google Scholar]
- 100.Waki T, Tamura G, Tsuchiya T, et al. Promoter methylation status of E-cadherin, hMLH1, and p16 genes in nonneoplastic gastric epithelia. Am J Pathol. 2002;161:399–403. doi: 10.1016/S0002-9440(10)64195-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Chan AO, Lam SK, Wong BC, et al. Promoter methylation of E-cadherin gene in gastric mucosa associated with Helicobacter pylori infection and in gastric cancer. Gut. 2003;52:502–506. doi: 10.1136/gut.52.4.502. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Kang GH, Lee HJ, Hwang KS, et al. Aberrant CpG island hypermethylation of chronic gastritis, in relation to aging, gender, intestinal metaplasia, and chronic inflammation. Am J Pathol. 2003;163:1551–1556. doi: 10.1016/S0002-9440(10)63511-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Forman D, Webb P, Parsonnet J. H pylori and gastric cancer. Lancet. 1994;343:243–244. [PubMed] [Google Scholar]
- 104.Uemura N, Okamoto S, Yamamoto S, et al. Helicobacter pylori infection and the development of gastric cancer. N Engl J Med. 2001;345:784–789. doi: 10.1056/NEJMoa001999. [DOI] [PubMed] [Google Scholar]
- 105.Hur K, Niwa T, Toyoda T, et al. Insufficient role of cell proliferation in aberrant DNA methylation induction and involvment of specific types of inflammation. Carcinogenesis. 2011;32:35–41. doi: 10.1093/carcin/bgq219. [DOI] [PubMed] [Google Scholar]
- 106.Niwa T, Ushijima T. Induction of epigenetic alterations by chronic inflammation and its significance on Carcinogenesis. Adv Genet. 2010;71:41–56. doi: 10.1016/B978-0-12-380864-6.00002-X. [DOI] [PubMed] [Google Scholar]
- 107.Nardone G, Rocco A, Malfertheiner P. Helicobacter pylori and molecular events in precancerous gastric lesions. Aliment Pharmacol Ther. 2004;20:261–270. doi: 10.1111/j.1365-2036.2004.02075.x. [DOI] [PubMed] [Google Scholar]
- 108.Sepulveda AR, Yao Y, Yan W, et al. CpG methylation and reduced expression of O6-methylguanine DNAmethyltransferase is associatedwith H. pylori infection. Gastroenterology. 2010;138:1836–1844. doi: 10.1053/j.gastro.2009.12.042. [DOI] [PubMed] [Google Scholar]
- 109.Tomita H, Hirata A, Yamada Y, et al. Suppressive effect of global DNA hypomethylation on gastric Carcinogenesis. Carcinogenesis. 2010;31:1627–1633. doi: 10.1093/carcin/bgq129. [DOI] [PubMed] [Google Scholar]
- 110.Jones PA, Baylin SB. The fundamental role of epigenetic events in cancer. Nat Rev Genet. 2002;3:415–428. doi: 10.1038/nrg816. [DOI] [PubMed] [Google Scholar]
- 111.Yoshida T, Yamashita S, Takamura-Enya T, et al. Alu and Satalpha hypomethylation in Helicobacter pylori-infected gastric mucosae. Int J Cancer. 2011;128:33–39. doi: 10.1002/ijc.25534. [DOI] [PubMed] [Google Scholar]
- 112.Hiraki M, Kitajima Y, Koga Y, et al. Aberrant gene methylation is a biomarker for the detection of cancer cells in peritoneal wash samples from advanced gastric cancer patients. Ann Surg Oncol. 2011;18:3013–3019. doi: 10.1245/s10434-011-1636-0. [DOI] [PubMed] [Google Scholar]
- 113.Frommer M, McDonald LE, Millar DS, et al. A genomic sequencing protocol that yields a positive display of 5-methylcytosine residues in individual DNA strands. Proc Natl Acad Sci USA. 1992;89:1827–1831. doi: 10.1073/pnas.89.5.1827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Herman JG, Graff JR, Myohanen S, et al. Methylation-specific PCR: a novel PCR assay for methylation status of CpG islands. Proc Natl Acad Sci USA. 1996;93:9821–9826. doi: 10.1073/pnas.93.18.9821. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Xiong Z, Laird PW. COBRA: a sensitive and quantitative DNA methylation assay. Nucleic Acids Res. 1997;25:2532–2534. doi: 10.1093/nar/25.12.2532. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Eads CA, Danenberg KD, Kawakami K, et al. MethyLight: a high-throughput assay to measure DNA methylation. Nucleic Acids Res. 2000;28:E32. doi: 10.1093/nar/28.8.e32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Fackler MJ, McVeigh M, Mehrotra J, et al. Quantitative multiplex methylation-specific PCR assay for the detection of promoter hypermethylation in multiple genes in breast cancer. Cancer Res. 2004;64:4442–4452. doi: 10.1158/0008-5472.CAN-03-3341. [DOI] [PubMed] [Google Scholar]
- 118.Swift-Scanlan T, Blackford A, Argani P, et al. Two-color quantitative multiplex methylation-specific PCR. Biotechniques. 2006;40:210–219. doi: 10.2144/000112097. [DOI] [PubMed] [Google Scholar]
- 119.Tost J, Gut IG. DNA methylation analysis by pyrosequencing. Nat Protoc. 2007;2:2265–2275. doi: 10.1038/nprot.2007.314. [DOI] [PubMed] [Google Scholar]
- 120.Al-Moundhri MS, Al-Nabhani M, Tarantini L, et al. The prognostic significance of whole blood global and specific DNA methylation levels in gastric adenocarcinoma. PLoS One. 2010;5:e15585. doi: 10.1371/journal.pone.0015585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Yang X, Lay F, Han H, et al. Targeting DNA methylation for epigenetic therapy. Trends Pharmacol Sci. 2010;31:536–546. doi: 10.1016/j.tips.2010.08.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Toyota M, Suzuki H, Yamashita T, et al. Cancer epigenomics: implications of DNA methylation in personalized cancer therapy. Cancer Sci. 2009;100:787–791. doi: 10.1111/j.1349-7006.2009.01095.x. [DOI] [PMC free article] [PubMed] [Google Scholar]