Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Mar 7.
Published in final edited form as: Clin Genet. 2011 Dec 8;81(4):303–311. doi: 10.1111/j.1399-0004.2011.01809.x

Epigenetic modifications in cancer

R Kanwal a, S Gupta a,b,c
PMCID: PMC3590802  NIHMSID: NIHMS441472  PMID: 22082348

Abstract

Cancer initiation and progression is controlled by both genetic and epigenetic events. The complexity of carcinogenesis cannot be accounted for by genetic alterations alone but also involves epigenetic changes. Epigenetics refers to the study of mechanisms that alter gene expression without altering the primary DNA sequence. Epigenetic mechanisms are heritable and reversible, and include changes in DNA methylation, histone modifications and small noncoding microRNAs (miRNA). Disruption of epigenetic processes can lead to altered gene function and malignant cellular transformation. Aberrant epigenetic modifications probably occur at a very early stage in neoplastic development, and they are widely described as essential players in cancer progression. Recent advances in epigenetics offer a better understanding of the underlying mechanism(s) of carcinogenesis and provide insight into the discovery of putative cancer biomarkers for early detection, disease monitoring, prognosis, and risk assessment. In this review, we summarize the current literature on epigenetic changes causing genetic alterations that are thought to contribute to cancer, and discuss the potential impact of epigenetics future research.

Keywords: cancer, DNA methylation, epigenetics, histone modification, microRNA


Cancer, initially recognized as a genetic disease, is now known to involve epigenetic abnormalities along with genetic alterations. It is becoming clear that microenvironment-mediated epigenetic perturbations play an important roles in the development of neoplasia (1). Epigenetics refers to the study of heritable changes in gene expression that occur without a change in DNA sequence which are sufficiently powerful to regulate the dynamics of gene expression (2). The key processes responsible for epigenetic regulation are DNA methylation, histone modifications and posttran-scriptional gene regulation by noncoding RNA commonly referred as microRNAs (3). These mechanisms are critical components in the normal development and growth of cells and their modifications contribute to neoplastic phenotypes (4).

DNA methylation

In eukaryotes, methylation is characterized by epigenetic DNA modifications which play an important role in maintenance of genome integrity, genomic imprinting, transcriptional regulation, and developmental processes (5). DNA methylation usually takes place at the 5′ position of the cytosine ring within CpG dinucleotides, and its consequence is the silencing of genes and noncoding genomic regions. The modification at 5-methyl cytosine is catalyzed by enzyme DNA methyltransferases (DNMTs) (6). There are three main DNMTs: DNMT1, which maintains the existing methylation patterns following DNA replication, and DNMT3A and DNMT3B, de novo enzymes that target unmethylated CpGs to initiate methylation and are highly expressed during embryogenesis and minimally expressed in adult tissues (7). Another family member is DNMT-3L that lacks intrinsic methyltransferase activity; it interacts with DNMT3a and 3b to facilitate methylation of retrotransposons (8). Both the establishment and maintenance of DNA methylation patterns are critical for development as mice deficient in DNMT3B or DNMT1 are embryonic lethal and DNMT3A-null mice die by 4 weeks of age (9). In normal cells, DNA methylation occurs predominantly in repetitive genomic regions, including satellite DNA and parasitic elements such as long interspersed transposable elements (LINEs) and short interspersed transposable elements (SINEs), maintaining genomic integrity (10). Methylated cytosines account for approximately 1% of total nucleotides and about 75% of all CpG dinucleotides in the human genome. The CpG dinucleotides are unevenly distributed across the human genome, but are concentrated in dense pockets called CpG islands (CGIs). About 50–60% of gene promoters lie within CpG islands, and it is estimated that the human genome contains approximately 29,000 CpG sequences (11). CpG islands particularly those associated with promoters are generally unmethylated in normal cells, providing access to transcription factors and chromatin-associated proteins for the expression of most housekeeping genes and several other regulated genes although some of them (~6%) become methylated in a tissue-specific manner during early development or in differentiated tissues (4, 12). DNA methylation can inhibit gene expression directly, by inhibiting the binding of specific transcription factors, and indirectly, by recruiting methyl-CpG-binding domain (MBD) proteins. The associated MBD family members in turn recruit histone-modifying and chromatin-remodeling complexes to methylated sites (13). To date, six methyl-CpG-binding proteins, including methylcytosine binding protein 2 (MECP2), MBD1, MBD2, MBD3, MBD4 and Kaiso, have been identified in mammals. MECP2 binds methylated DNA in vitro and in vivo. It contains a MBD at its amino terminus and a transcription repression domain (TRD) in the middle (13). Furthermore, it has been shown that nucleosome remodeling complex (NuRD) can methylate DNA by interacting with DNA methylation binding protein MBD2, which directs the NuRD complex to methylate DNA (14, 15). These and other recent findings have established that DNA cytosine methylation is a critical component of epigenetic gene regulation. A list of DNA methylation genes disrupted in various human cancers is shown in Table 1.

Table 1.

DNA methylation gene changes in various human cancers

DNA methyltransferase Function Alterations Cancer type
DNMT 1 Maintenance of methylation, repression of transcription Upregulation, mutation Colorectal cancer, ovarian cancer
DNMT3a De novo methylation during embryogenesis, imprint establishment, repression Upregulation Colorectal cancer, breast cancer, ovarian cancer, esophageal squamous cell carcinoma
DNMT3b De novo methylation during embryogenesis, repeat methylation repression Upregulation Breast cancer, hepatocellular carcinoma, colorectal cancer,
DNMT3L Interacts with DNMT3a &b and facilitate methylation Upregulation
Methyl-CpG- binding proteins
MeCP2 Transcription repression Upregulation, Mutation Prostate cancer, Rett syndrome,
MBD1 Transcription repression Upregulation, Mutation Prostate cancer, colon cancer, lung cancer
MBD2 Transcription repression Upregulation, Mutation Prostate cancer, colon cancer, lung cancer
MBD3 Transcription repression Upregulation, Mutation Colon cancer, lung cancer
MBD4 DNA repair, glycosylase domain, repair of deaminated 5-methylC Upregulation, Mutation Colon cancer, gastric cancer, endometrial cancer
Kaiso Transcription repression Upregulation Colon, intestinal, lung cancer

DNMT, DNA methyltransferase; MBD, methyl-CpG-binding domain.

Histone modification

Histone modifications influence chromatin structure which plays an important role in gene regulation and carcinogenesis (16). Chromatin is a highly ordered structure consisting of repeats of nucleosomes connected by linker DNA. Each nucleosome encompasses ~146 bp of DNA wrapped around an octamer of histone proteins. These octamers consist of two subunits of each of the following core histone proteins: H2A, H2B, H3 and H4 (17). Chromatin consists of DNA, histones, and non-histone proteins condensed into nucleoprotein complexes and functions as the physiological template of all eukaryotic genetic information (18). Histones are small basic proteins containing a globular domain and a flexible charged NH2 terminus known as the histone tail, which protrudes from the nucleosome. Regulation of gene expression occurs through posttranslational modifications of the histone tails provided by covalent modifications including acetylation, methylation, phosphorylation, ubiquitination, sumoylation, proline isomerization, and ADP ribosylation (1921). Posttranslational modifications to histone tails govern the structural status of chromatin and the resulting transcriptional status of genes within a particular locus. These modifications are reversible and are controlled by a group of enzymes including histone acetyltransferases (HATs) and deacetylases (HDACs), methyltransferases (HMTs) and demethylases (HDMs), kinases, phosphatases, ubiquitin ligases and deubiquitinases, SUMO ligases and proteases which add and remove such modifications (20, 22). Chromatin is divided into two distinct conformation states: heterochromatin, which is densely compacted and transcriptionally inert and euchromatin, which is decondensed and transcriptionally active (17, 23). Euchromatin is characterized by high levels of acetylation and trimethylated H3K4, H3K36 and H3K79. On the other hand, heterochromatin is characterized by low levels of acetylation and high levels of H3K9, H3K27 and H4K20 methylation (12, 24, 25). Studies have shown that histone modification levels are predictive for gene expression. Actively transcribed genes are characterized by high levels of H3K4me3, H3K27ac, H2BK5-azacytidine (H2BK5ac) and H4K20me1 in the promoter and H3K79me1 and H4K20me1 along the gene body (26). Genome-wide studies have revealed that various combinations of histone modifications in a specific genomic region can lead to a more ‘open’ or ‘closed’ chromatin structure resulting in the activation or repression of gene expression (12, 24, 25). A list of the histone-modifying genes altered during carcinogenesis is presented in Table 2.

Table 2.

Histone modification genes altered in various human cancers

Histone deacetylases Alterations Cancer type
HDAC1 Upregulation/Downregulation Colorectal cancer, cervical dysplasias, endometrial stromal sarcomas, gastric carcinomas, colon cancer
HDAC2 Upregulation/Mutation Multiple gastric carcinomas, colon cancer
HDAC3 Upregulation Colon cancer
HDAC4 Upregulation/Downregulation /Mutation Prostate, breast and colon cancer
HDAC5 Under expression Colon cancer, AML
HDAC6 Upregulation Breast, AML
HDAC7 Upregulation Colon cancer
HDAC8 Upregulation Colon cancer
SIRT1 Upregulation/Downregulation Colon cancer
SIRT2 Downregulation, Deletion Glioma
SIRT3 Upregulation Breast cancer
SIRT4 Downregulation AML
SIRT7 Upregulation Breast, thyroid carcinoma
Histone acetyl transferases
P300 Mutation, Translocation, Deletions Colorectal, breast, ovarian, hepatocellular and oral carcinomas,
CBP Mutation, Translocation, Deletions Colon, breast, ovarian and AML
PCAF Mutation Colon
MOZ Translocation Hematologic malignancy
MORF Translocation Hematologic, uterine leiomyomata
Tip60 Downregulation, translocation Colorectal, prostate cancer
GCN5 Downregulation, mutation Prostate, breast ovary
PCAF Rare mutation Colon cancer
HBO1 Upregulation Testis, breast, ovary and bladder cancer
Histone methyltransferase
MLL1 Translocation, Amplification Hematologic malignancies
MLL2 Amplification Glioma, pancreatic cancer
MLL3 Mutation, Deletion Hematologic malignancies, colon cancer
MLL4 Amplification Solid tumor
SUV39H1-2 Mutation overexpression Ovarian, colon
G9a Gene Repression Colon, gastric, breast cancer
RIZ1/ PRDM2 Underexpression, Mutations Colorectal, gastric cancer
EVI1 Chromosomal rearrangement Myeloid leukemia
EZH2 Amplification, Upregulation Prostate, breast ovarian cancer
SUZ12 Upregulation Prostate, breast ovarian cancer
BMI1 Upregulation Prostate, breast ovarian cancer
NSD1 Translocation, Upregulation AML
NSD2 Translocation Multiple myeloma
NSD2 Translocation Multiple myeloma
NSD3 Translocation, Amplification AML, breast cancer
SYMD2 Upregulation Breast, colon cancer
DOT1 Upregulation AML
Histone demethylases
LSD1/BHC110 Downregulation Breast cancer
JARID1A-D Upregulation /Inactivation mutation Leukemia, prostate, breast renal carcinoma
JHDM2a, 2b Upregulation T-cell lymphoma
JMJD2A/JHDM3A Upregulation Squamous cell carcinomas, lung cancer
JMJD3 Downregulation Lung, liver cancer
UTX Inactivation mutation Squamous cell carcinomas, renal cell carcinomas

CBP, CREB binding protein; Dot1, disruptor of telomeric silencing EZH2, enhancer of zest homolog2; Gcn5, general control nonderepressible; HBO1, histone acetyltransferase binding to ORC1; HDAC, histone deacetylase; JARID, Jumonji, AT-rich interactive-domain JHDM, JmjC domain-containing histone demethylase 1; JMJD, Jumonji domain containing 2; LSD1, lysine specific demethylase 1; MLL, myeloid/lymphoid or mixedlineage leukemia-associated protein; Morf, MOZ-related factors; MOZ, monocytic leukemia zinc finger protein; NSD1, nuclear receptor-binding SET-domain protein 1; p300, E1A binding protein p300; PCAF, p300/CBP-associated factor; PRMT, protein arginine methyltransferase 1; RIZ1, retinoblastoma protein-interacting zinc finger 1; SIRT, Sir2 histone deacetylase gene family; SMYD2, split SET/MYND domain-containing histone H3 lysine 36-specific methyltransferase; SUV39H, suppressor of variation 3–9 homolog; TIP60, human HIV-1 Tat interactive protein 60; UTX, ubiquitously transcribed tetratricopeptide repeat, X chromosome.

Noncoding RNAs

MicroRNAs (miRNAs) are small ncRNAs of ~22 nucleotides and are involved in posttranslational gene silencing by controlling mRNA translation into proteins (27). miRNAs induce heritable changes in gene expression without altering DNA sequence and thus contribute to the epigenetic landscape. In addition, miR-NAs can both regulate and be regulated by other epigenetic mechanisms (28). Although miRNA are vital to normal cell physiology their mis-expression has been linked to several diseases, including cancer (4). Cancer development and miRNA profiles are now being used to classify human cancers (23, 28, 27). Approximately, 1000 miRNA genes have been computationally predicted in the human genome with each miRNA targeting multiple protein-coding transcripts. It has been predicted that miRNAs regulate the translation rate of more than 60% of protein-coding genes (7), and participate in the regulation of cellular processes. Like mRNAs, miRNAs are mainly transcribed by RNA polymerase-II although miRNA synthesis is known to occur by RNA polymerase-III in those miRNAs that reside near tRNA, Alu and mammalian-wide interspersed sequences (28). The first identified miRNAs, the products of the Caenorhabditis elegans genes lin-4 and let-7, have important roles in controlling developmental timing and probably act by regulating mRNA translation. When lin-4 or let-7 is inactivated, specific epithelial cells undergo additional cell divisions instead of their normal differentiation (27). A list of miRNAs dysregulated in cancer is presented in Table 3.

Table 3.

MicroRNAs alteration in various human cancers

MicroRNAs Target gene(s) Alterations Cancer type
miR-127 Bcl-6 Upregulation Bladder cancer
miR-124 CDK6 Upregulation Colon cancer
miR-223 NFI-A, MEF2C Upregulation Acute myeloid leukemia
miR-34b/34c p53 network, CDK6, E2F3 Upregulation Colon cancer
miR-17, miR-92 c-MYC Upregulation Lung cancer
miR-372, miR-373 RAS, p53,CD44 Upregulation Testicular germ cell tumor and breast cancer
miR-21 PDCD4, PTEN, TPM1, REC, TIMP3, BCL2 Upregulation Glioblastoma, breast, lung, prostate,
miR-155 RHOA Upregulation Burkitt’s lymphoma, breast, colon, and lung cancers
miR-146 NF-κB Upregulation Breast, pancreas and prostate cancers
miR-92b PRMT5 Upregulation Brain primary tumors
miR-520 CD44 Upregulation Breast cancer
miR-10b HOXD10 Upregulation Metastatic breast cancer
miR-9 CDH1 Upregulation Breast cancer
miR-127, miR-199a BCL6, E2F1 Upregulation Cervical cancer
miR-421 CBX7, RBMXL1 Upregulation Gastric cancer
miR-1228, miR-195, miR30b, miR-32, miR345 CDKN2A, NF2, and JUN Upregulation Malignant mesothelioma
miR-190, miR-196 HGF Upregulation Pancreatic cancer
miR-125 AKT, ERBB2-4, FGF, FGFR, IGF, MAPKs, MMP11, SP1, TNF, VEGF Upregulation Breast cancer
miR-29 DNMT3a&b Upregulation Lung cancer
miR-1 FoxP1 Upregulation Hepatocarcinoma
miR-9-3 E-cadherin Upregulation Colorectal, melanoma, head and neck cancer
miR-34a CD44, Notch1 Upregulation Hematological, prostate, breast cancer
miR-181c Notch4, K-Ras Upregulation Gastric, colorectal cancer
miR-200c, miR-141, miR-429 ZEB1/ZEB2 Upregulation Colorectal, breast, lung cancer
miR-126 CRK1, PIK3R2, SPRED1, VCAM1 Downregulation Breast and lung cancer
miR-146a, miR-146b ROCK1, IRAK1, TRAF6 Downregulation Prostate cancer and papillary thyroid carcinomas
miR-340, miR-421, miR-658 MYC, RB, PTEN Downregulation Lymph node metastasis and gastric cancer
let-7a-3 RAS, IGF-II Downregulation Lung and ovarian cancer
miR-221, miR-222 CDKN1C/P57 and CDKN1B/P27 Downregulation Hepatocellular carcinoma
miR-9 NF-κB Downregulation Ovarian and lung cancer
miR-218, miR-145 PXN Downregulation Breast, lung and prostate cancer
miR-25, miR-32, miR-142 ITGAα1 Downregulation Lung cancer and solid tumor
miR-124, miR-183 ITGBβ1 Downregulation Lung cancer
miR-143 ERK5 Downregulation Cervical cancer
miR-372, miR-373 LATS2 Downregulation Testicular germ cell cancer
miR-181 VGFR Downregulation Lung cancer
miR-370 MAP3K8 Downregulation Cholangiocarcinoma
miR-342 ER, PR and HER2 Downregulation Breast and colon cancer
miR-145 ER Downregulation Colon and breast cancer
miR-124, miR-183 ITGB1 β Downregulation Lung cancer
miR-101 EZH2 Downregulation Prostate, breast , Lung
miR-143 DNMT3a Downregulation Colorectal cancer
miR-9-1 FGF family Downregulation Breast, ovarian, pancreas cancer
miR-137 CDK6, E2F6, LSD Downregulation Glioblastoma, oral, colorectal cancer
miR-129-2 SOX4 Downregulation Gastric, endometrial, colorectal cancer
miR-145 OCT/SOX2/KLF Downregulation Prostate cancer
miR-148 TGIF2 Downregulation Colorectal, melanoma, head and neck cancer
miR-199a IKKB Downregulation Hepatocarcinoma, testicular ovarian cancer
miR-203 ABL1 Downregulation Hematological, liver cancer
miR-205 ZEB1/ZEB2 Downregulation Bladder, breast, prostate cancer
miR-335 SOX4/TNC Downregulation Breast cancer
miR-342 PDGFRA Downregulation Ovarian, breast cancer

BCL2, B-cell lymphoma 2 protein; CBX7, chromobox 7; CD44, cluster differentiation 44; CDH1, cadherin-1; CDK6, cyclin D kinase 6; CDNK2A, cyclin-dependent kinase inhibitor 2A; CRK1, Cdc2-related kinase1; ER, estrogen receptor; ERBB2-4 or (HER4), human epidermal growth factor receptor 4; ERK5, extracellular signal-regulated kinase 5; FGFR, fibroblast growth factor receptor; FOXP1, forkhead box protein O1; HGF, hepatocyte growth factor; HOXD10, homeobox D10; IGF-II, insulin-like growth factor 2; IRAK1, interleukin-1 receptor associated kinase-1; KLF, Kruppel-like factors; LATS2, large tumor-suppressor, homolog 2; MAPKs, mitogen-activated protein kinase; MEF2C, myocyte enhancer factor 2C; MMP11, matrix metalloproteinase11; NF2, neurofibromatosis, type 2; NFIA, nuclear factor 1 A-type; NF-kB, nuclear factor-κB; OCT4, octamer-binding transcription factor 4; P53, tumor protein 53; PDcD4, programmed cell death 4; PDGFRA, alpha-type platelet-derived growth factor receptor; PIK3R2, Phosphatidylinositol 3-kinase regulatory subunit beta; PR, progesterone receptor; PRNT5, protein arginine N-methyltransferase 5; PTEN, phosphatase and tensin homolog; PXN, paxilin; ITGβ1, integrin beta-1; RAS, rat sarcoma; Rb, retinoblastoma; RBMX L1, RNA binding motif protein X-linked; RECK, reversion inducing cysteine rich protein kazal Motif; ROCK1, rho-associated, coiled-coil containing protein kinase 1; ROHA, Ras homolog gene family member A; SOX4, SRY (sex determining region Y)-box 4; SPRED1, sprouty-related, EVH1 domain containing 1; TNFα, tumor necrosis factor-alpha; TPM1, tropomyosin 1; TRAF6, TNF receptor associated factor 6; VCAM, vascular cell adhesion molecule; VEGF, vascular endothelial growth factor; ZEB1, zinc finger E-box-binding homeobox 1.

Epigenetic modifications in cancer

In human cancers, aberrant epigenomics are known to contribute to various phases of neoplastic development including initiation, promotion, invasion, metastases and chemotherapy resistance. It has been proposed recently that more than 300 genes and gene products are epigenetically altered in various human cancers (4). A link between DNA methylation and cancer was first shown in 1983, when it was shown that the genomes of cancer cells are hypomethylated relative to their normal counterparts (29). In terms of DNA methylation, cancer cells show genome-wide hypomethylation and site-specific CpG island promoter hypermethylation (30, 31). Hypermethylation at specific genes typically affects promoter CpG-islands inactivating transcription. Hypermethylation is observed at specific CpG islands. The transcriptional inactivation caused by promoter hypermethylation affects genes involved in the main cellular pathways: DNA repair [hMLH1 (mismatch repair gene 1), MGMT (O6-methylguanine–DNA methyltransferase), WRN (Werner syndrome, RecQ helikase like), BRCA1 (breast cancer 1)], cell cycle control (p16 INK4a, p15 I NK4b, RB), Ras signaling {RASSF1A [Ras association (RalGDS/AF-6) domain family member 1], NOREIA}, apoptosis [TMS1 ( target of methylation-induced silencing 1), DAPK1 (death-associated protein kinase), WIF-1, SFRP1 ], metastasis [cadherin 1 (CDH1), CDH13, PCDH10 ], detoxification [GSTP1 (glutathione S-transferase pi 1)], hormone response (ESR1, ESR2 ), vitamin response [RARB2 (retinoic acid receptor b2), CRBP1] and p53 network [p14 ARF , p73 (also known as TP73 ), HIC-1 ] among others. This provides tumor cells with a growth advantage and increases their genetic instability and aggressiveness (22, 30). Hypomethylation in tumor cells is primarily caused by the loss of methylation from repetitive regions of the genome and the resulting genomic instability is a hallmark of tumor cells. In addition, imprinting patterns are often disturbed (32). In cancers, hypomethylation is often associated with oncogenes. c-Myc, a transcription factor that acts as an oncogene, is one of the widely reported hypomethylated genes in cancers. Hypomethylation at specific promoters can activate the aberrant expression of oncogenes and induce loss of imprinting (LOI). The most common LOI event due to hypomethy-lation is insulin-like growth factor 2 (IGF2 ), which has been reported in a wide range of tumor types, including breast, liver, lung and colon cancer (33, 34). S100P in pancreatic cancer, SNCG in breast and ovarian cancers and melanoma-associated gene (MAGE ) and dipeptidyl peptidase 6 (DPP6 ) in melanomas are well-studied examples of hypomethylated genes in cancer (12, 35).

In addition to changes in DNA methylation, histone modification patterns are also altered in human tumors. Recent studies have shown that histone modification levels are predictive for gene expression. Actively transcribed genes are characterized by high levels of H3K4me3, H3K27ac, H2BK5ac and H4K20me1 in the promoter and H3K79me1 and H4K20me1 along the gene body (26). Loss of acetylation is mediated by HDACs that have been found to be over-expressed or mutated in different tumor types. Aberrant expression of both HMTs and HDMs are observed in various cancer types (36). A recent study has described inactivating mutations in the histone methyltransferase SETD2 and in the histone demethylase UTX and JARID1C in renal carcinomas (36, 37). H3 acetylation and H3K9 dimethylation can discriminate between cancerous and nonmalignant prostate tissue and H3K4 trimethylation can predict occurrence of prostate-specific antigen serum level elevation after prostatectomy for cancer (28). EZH2 (enhancer of zeste homolog 2) expression is an independent prognostic marker that is correlated with the aggressiveness of prostate, breast and endometrial cancers (38).

Because abnormal cell proliferation is a hallmark of human cancers, it seems possible that miRNA expression patterns might denote the malignant state. Indeed, altered expression of a few miRNAs has been found in some tumor types (39, 40). The first association between miRNA and cancer development was described in chronic lymphocytic leukemia with chromosome 13q14 deletion. This deletion deregulates miRNA-15 and miRNA-16 (41). Most of the targets of these two miRNAs are involved in cell growth and cell cycle. The let-7 is one of the most widely studied miRNA families in cancer. Alterations of let-7 function have been described in several human cancer types, including carcinomas of the head and neck region, lung, colon, rectum and ovary. It acts mainly as a tumor-suppressor miRNA (28). miRNA-145 is a well-known tumor-suppressor miRNA downregulated in many human cancers owing to aberrant DNA methylation of its promoter and/or p53 mutations (42). This miRNA is a pluripotency repressor which regulates silencing of OCT, SOX2 and KLF4 in human embryonic stem cells; these genes are required for cell self-renewal and pluripotency maintenance (43). Interestingly, it is becoming apparent that the expression of epigenetic regulatory enzymes such as DNMT, HATs, and HMTs can be controlled by miRNAs (44). In particular, the miRNA-29 family can directly regulate the expression of DNMTs such that downregulation of this family of miRNAs in small-cell lung cancer results in increased expression of DNMT3A and 3B causing a global genomic hypermethylation and specific methylation-induced silencing of tumor-suppressor genes such as FHIT and WWOX (45, 46).

Influence of epigenetics on cancer genetics

It is evident that discrete genetic alterations in neoplastic cells alone cannot explain multistep carcinogenesis whereby tumor cells are able to express diverse phenotypes during the complex phases of tumor development and progression. In fact, cancer cells have an altered epigenome compared to the tissues from which they arise. Deregulated epigenetic mechanisms may initiate genetic instability, resulting in the acquisition of genetic mutations in tumor-suppressor genes and activating genetic mutations in oncogenes. Moreover, epigenetic disruptions in tumors are generally of a clonal nature, indicating occurrence in early generations of cells (47). It is well known that 5-methylcytosine (m5C) residues are ‘hot spots’ for mutations, which can destabilize gene structure and function. One-third of germ-line point mutations leading to human genetic diseases occur at CpGs and most of these mutations are C→T transitions (48). This is because m5C is highly mutable by deamination, resulting in transitional mutations (i.e. C→T) at CpGs. In view of the symmetry of these CpG motifs, the methylcytosine on the opposite strand may also be affected, leading to (G→A) changes. As a consequence, CpGs are hot spots for mutations, in a variety of genes. (48, 49) G→A transitions are found in 44.8% cases of leukemia and myelodysplasia, and in 60% of colon cancer cases. C→T and tandem CC-TT mutations are found in basal cell and squamous cell carcinomas (50). Methylation increases the rate of hydrolytic deamination and also increases the reactivity of neighboring guanines to electrophiles (50, 51). The oxidation of m5C may contribute to the high frequency of C→T transitions at CpG sequences. Oxygen radicals can react with m5C to oxidize the 5, 6-double bond; the intermediate product, m5C glycol, then deaminates to form thymine glycol (52, 53). Oxidative stress can contribute to tumor development not only through genetic but also through epigenetic mechanisms. As noted earlier, the presence of hydroxyl radicals can cause a wide range of DNA lesions including base modifications, deletions, strand breakage and chromosomal rearrangements. Such DNA lesions have been shown to interfere with the ability of DNA to function as a substrate for the DNMTs, resulting in global hypomethylation (54). The presence of 8-OHdG in CpG dinucleotide sequences has been shown to strongly inhibit methylation of adjacent cytosine residues (55, 56). In addition, 8-OHdG may not be recognized by proof-reading enzymes and thus may persist as a mutation resulting in G→T transversions (57, 58). These studies suggest that oxidative DNA damage can affect patterns of DNA methylation leading to aberrant gene expression and possibly contributing to the development of malignancy.

Hereditary cancer genes and cancer predisposition

Studies of familial cancer have identified a group of genes whose mutational inactivation results in predisposition to a characteristic spectrum of cancers. The tumor-suppressor gene, RB which is mutated in retinoblastoma, was the first hereditary cancer gene to be identified. Subsequently, other tumor-suppressor genes operating through a diverse range of mechanisms were identified in other familial cancers, e.g. adenomatosis polyposis coli (APC) mutations were identified in cases of familial adenomatous polyposis coli and p16 INK4a mutations were identified in cases of familial melanoma (59). DNA repair genes such as the BRCA1, MLH1, and MSH2 are also often involved in predisposition to familial cancer (60, 61). Specific allele sequence variants such as single nucleotide polymorphisms (SNPs) affect the probability of CpG islands methylation in the cis region. Cis-acting DNA mutations have been shown to cause constitutional epi-mutations early in cancer development (62). One example of epigenetic silencing caused by sequence alterations in cis is the methylation of expanded CGG repeats within the FMR1 promoter in the fragile X mental retardation syndrome (63). Little is known about what causes variation in methylation levels and how this variation relates to future pathogenesis. Genetic factors may play key roles in maintaining a normal methylation pattern, but environmental factors also play an important role. Polymorphisms in MTHFR, DNMT3b, and MTR genes, involved in methyl metabolism or DNA methylation, are some of the obvious genetic candidates for variation in methylation in both normal and tumor tissues (64).

Epigenetic biomarkers

Methylated DNA sequences represent potential bio-markers for diagnosis, staging, prediction of prognosis and monitoring of response to therapy. Epigenetic markers can be evaluated in resected tumors or in body fluids (65). For example, the frequency of occurrence of hypermethylated CDH13, MYOD1, MGMT, p16 INK4a and RASSF1A genes varies significantly among cancer types. These changes can be detected in plasma DNA and urine. Recently, combined hypermethylation assays for small number of genes such as RASSF1A, RARβ2, APC and GSTP1 have been used to discriminate between benign and cancerous changes in the prostate (66, 67). Furthermore, repetitive DNA elements such as SINEs and LINEs and other repetitive sequences are often hypomethylated. Their presence has been used to characterize some human cancers, but their clinical utility remains questionable (68).

Histone modification patterns also provide prognostic and diagnostic information in cancer. Repressive chromatin structures characterized by particular histone modifications such as H3K9, H3K27 and H4K20 methylation may precipitate DNA methylation. Generalized changes in chromatin structure and histone modification, such as increased H3K4 dimethylation and H3K18 acetylation activation, are associated with poor prognosis (69). Whether the extent of these changes correlates with alteration in gene activity is a major limitation.

More recently, miRNA has been proposed as potential epigenetic biomarkers in the diagnosis of cancer. Some of the miRNA such as miR-199a, miR-200a, miR-146, miR-214, miR-221 and miR-222 have been found to be upregulated, whereas miR-100 is down-regulated in human cancers (44). The miRNA let-71 has been recently designated as a tumor-suppressor and miR-429, miR-200a and miR-200b were found to be clustered on a single primary transcript regulated by the epithelial-to-mesenchymal transition (70). Studies have shown that two other miRNAs, miR-21 and miR-181a can be used to identify the presence or absence of a malignant phenotype. A group of 27 miRNA has been shown to be significantly associated with chemotherapy response and proposed as possible prognostic and diagnostic biomarker (71, 72).

Conclusions

The epigenetic modification patterns associated with the development and progression of cancer are potentially clinically useful. The development of DNA methylation markers may prove useful for early cancer detection, establishing a diagnosis of cancer, or predicting the prognosis in cancer cases. Recent advances in epigenomic approaches allow mapping of the methylation/acetylation state and miRNA levels in the genome with high accuracy, which may help in the identification of biomarkers for various diseases. Understanding the molecular events that initiate and maintain epigenetic gene silencing could lead to the development of clinical strategies for the prevention and therapy of cancer.

Acknowledgments

The reference from author’s laboratory listed in this review was supported by United States Public Health Service Grants RO1 CA115491 and R21 CA109424. We apologize to those investigators whose original work could not be cited owing to the space limitations.

Footnotes

Conflict of interest

Nothing to declare.

References

  • 1.Berger SL, Kouzarides T, Shiekhattar R, et al. An operational definition of epigenetics. Genes Dev. 2009;23:781–783. doi: 10.1101/gad.1787609. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Rodenhiser D, Mann M. Epigenetics and human disease: translating basic biology into clinical applications. CMAJ. 2006;174:341–348. doi: 10.1503/cmaj.050774. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Ducasse M, Brown MA. Epigenetic aberrations and cancer. Mol Cancer. 2006;5:60. doi: 10.1186/1476-4598-5-60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Kanwal R, Gupta S. Epigenetics and cancer. J Appl Physiol. 2010;109:598–605. doi: 10.1152/japplphysiol.00066.2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Wu SC, Zhang Y. Active DNA demethylation: many roads lead to Rome. Nat Rev Mol Cell Biol. 2010;11:607–620. doi: 10.1038/nrm2950. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Rodriguez-Paredes M, Esteller M. Cancer epigenetics reaches mainstream oncology. Nat Med. 2011;17:330–339. doi: 10.1038/nm.2305. [DOI] [PubMed] [Google Scholar]
  • 7.Burge CB, Friedman RC, Farh KKH, et al. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009;19:92–105. doi: 10.1101/gr.082701.108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Fuks F, Deplus R, Brenner C, et al. Dnmt3L is a transcriptional repressor that recruits histone deacetylase. Nucleic Acids Res. 2002;30:3831–3838. doi: 10.1093/nar/gkf509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Li E, Okano M, Bell DW, et al. DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell. 1999;99:247–257. doi: 10.1016/s0092-8674(00)81656-6. [DOI] [PubMed] [Google Scholar]
  • 10.Robertson KD. DNA methylation and human disease. Nat Rev Genet. 2005;6:597–610. doi: 10.1038/nrg1655. [DOI] [PubMed] [Google Scholar]
  • 11.Pradhan S, Kim JK, Samaranayake M. Epigenetic mechanisms in mammals. Cell Mol Life Sci. 2009;66:596–612. doi: 10.1007/s00018-008-8432-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Esteller M, Portela A. Epigenetic modifications and human disease. Nat Biotechnol. 2010;28:1057–1068. doi: 10.1038/nbt.1685. [DOI] [PubMed] [Google Scholar]
  • 13.Bogdanovic O, Veenstra GJ. DNA methylation and methyl-CpG binding proteins: developmental requirements and function. Chromosoma. 2009;118:549–565. doi: 10.1007/s00412-009-0221-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Dehan P, Kustermans G, Guenin S, et al. DNA methylation and cancer diagnosis: new methods and applications. Expert Rev Mol Diagn. 2009;9:651–657. doi: 10.1586/erm.09.53. [DOI] [PubMed] [Google Scholar]
  • 15.Lai AY, Wade PA. Cancer biology and NuRD: a multifaceted chromatin remodelling complex. Nat Rev Cancer. 2011;11:588–596. doi: 10.1038/nrc3091. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Cedar H, Bergman Y. Linking DNA methylation and histone modification: patterns and paradigms. Nat Rev Genet. 2009;10:295–304. doi: 10.1038/nrg2540. [DOI] [PubMed] [Google Scholar]
  • 17.Johnstone RW, Bots M. Rational combinations using HDAC inhibitors. Clin Cancer Res. 2009;15:3970–3977. doi: 10.1158/1078-0432.CCR-08-2786. [DOI] [PubMed] [Google Scholar]
  • 18.Suganuma T, Workman JL. Signals and combinatorial functions of histone modifications. Annu Rev Biochem. 2011;80:473–499. doi: 10.1146/annurev-biochem-061809-175347. [DOI] [PubMed] [Google Scholar]
  • 19.Cohen I, Poreba E, Kamieniarz K, et al. Histone modifiers in cancer: friends or foes? Genes Cancer. 2011;2:631–647. doi: 10.1177/1947601911417176. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Kouzarides T. Chromatin modifications and their function. Cell. 2007;128:693–705. doi: 10.1016/j.cell.2007.02.005. [DOI] [PubMed] [Google Scholar]
  • 21.Bannister AJ, Kouzarides T. Regulation of chromatin by histone modifications. Cell Res. 2011;21:381–395. doi: 10.1038/cr.2011.22. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Esteller M. Epigenetic gene silencing in cancer: the DNA hypermethylome. Hum Mol Genet. 2007;16 (Spec1):R50–R59. doi: 10.1093/hmg/ddm018. [DOI] [PubMed] [Google Scholar]
  • 23.Iorio MV, Piovan C, Croce CM. Interplay between microRNAs and the epigenetic machinery: an intricate network. Biochem Biophys Acta. 2010;1799:694–701. doi: 10.1016/j.bbagrm.2010.05.005. [DOI] [PubMed] [Google Scholar]
  • 24.Li B, Carey M, Workman JL. The role of chromatin during transcription. Cell. 2007;128:707–719. doi: 10.1016/j.cell.2007.01.015. [DOI] [PubMed] [Google Scholar]
  • 25.Izzo A, Schneider R. Chatting histone modifications in mammals. Brief Funct Genomics. 2010;9:429–443. doi: 10.1093/bfgp/elq024. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Karlic R, Chung HR, Lasserre J, et al. Histone modification levels are predictive for gene expression. Proc Natl Acad Sci U S A. 2010;107:2926–2931. doi: 10.1073/pnas.0909344107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Golub TR, Lu J, Getz G, et al. MicroRNA expression profiles classify human cancers. Nature. 2005;435:834–838. doi: 10.1038/nature03702. [DOI] [PubMed] [Google Scholar]
  • 28.Lopez-Serra P, Esteller M. DNA methylation-associated silencing of tumor-suppressor microRNAs in cancer. Oncogene. 2011;16:1–14. doi: 10.1038/onc.2011.354. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Bruserud O, Stapnes C, Ersvaer E, et al. Histone deacetylase inhibitors in cancer treatment: a review of the clinical toxicity and the modulation of gene expression in cancer cell. Curr Pharm Biotechnol. 2007;8:388–400. doi: 10.2174/138920107783018417. [DOI] [PubMed] [Google Scholar]
  • 30.Shen L, Kondo Y, Guo Y, et al. Genome-wide profiling of DNA methylation reveals a class of normally methylated CpG island promoters. PLoS Genet. 2007;3:2023–2036. doi: 10.1371/journal.pgen.0030181. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Sincic N, Herceg Z. DNA methylation and cancer: ghosts and angels above the genes. Curr Opin Oncol. 2011;23:69–76. doi: 10.1097/CCO.0b013e3283412eb4. [DOI] [PubMed] [Google Scholar]
  • 32.Markowitz SD, Li M, Chen WD, et al. Sensitive digital quantification of DNA methylation in clinical samples. Nat Biotechnol. 2009;27:858–U118. doi: 10.1038/nbt.1559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Chan WY, Cheung HH, Lee TL, et al. DNA methylation of cancer genome. Birth Defects Res C. 2009;87:335–50. doi: 10.1002/bdrc.20163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Cui HM. Loss of imprinting of IGF2 as an epigenetic marker for the risk of human cancer. Dis Markers. 2007;23:105–112. doi: 10.1155/2007/363464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Sadikovic B, Al-Romaih K, Squire JA, et al. Cause and consequences of genetic and epigenetic alterations in human cancer. Curr Genomics. 2008;9:394–408. doi: 10.2174/138920208785699580. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Allis CD, Chi P, Wang GG. Covalent histone modifications - miswrit-ten, misinterpreted and mis-erased in human cancers. Nat Rev Cancer. 2010;10:457–469. doi: 10.1038/nrc2876. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Teh B, Dalgliesh GL, Furge K, et al. Systematic sequencing of renal carcinoma reveals inactivation of histone modifying genes. Nature. 2010;463:360–363. doi: 10.1038/nature08672. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Akslen LA, Bachmann IM, Halvorsen OJ, et al. EZH2 expression is associated with high proliferation rate and aggressive tumor subgroups in cutaneous melanoma and cancers of the endometrium, prostate, and breast. J Clin Oncol. 2006;24:268–273. doi: 10.1200/JCO.2005.01.5180. [DOI] [PubMed] [Google Scholar]
  • 39.Michael MZ, O’Connor SM, Pellekaan NGV, et al. Reduced accumulation of specific microRNAs in colorectal neoplasia. Mol Cancer Res. 2003;1:882–891. [PubMed] [Google Scholar]
  • 40.Olson P, Lu J, Zhang H, et al. MicroRNA dynamics in the stages of tumorigenesis correlate with hallmark capabilities of cancer. Genes Dev. 2009;23:2152–2165. doi: 10.1101/gad.1820109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Calin GA, Dumitru CD, Shimizu M, et al. Frequent deletions and down-regulation of micro-RNA genes miR15 and miR16 at 13q14 in chronic lymphocytic leukemia. Proc Natl Acad Sci U S A. 2002;99:15524–1529. doi: 10.1073/pnas.242606799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Dahiya R, Suh SO, Chen Y, et al. MicroRNA-145 is regulated by DNA methylation and p53 gene mutation in prostate cancer. Carcinogenesis. 2011;32:772–778. doi: 10.1093/carcin/bgr036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Kosik KS, Xu N, Papagiannakopoulos T, et al. MicroRNA-145 Regulates OCT4, SOX2, and KLF4 and Represses Pluripotency in Human Embryonic Stem Cells. Cell. 2009;137:647–658. doi: 10.1016/j.cell.2009.02.038. [DOI] [PubMed] [Google Scholar]
  • 44.Esteller M, Guil S. DNA methylomes, histone codes and miRNAs: tying it all together. Int J Biochem Cell. 2009;B41:87–95. doi: 10.1016/j.biocel.2008.09.005. [DOI] [PubMed] [Google Scholar]
  • 45.Croce CM, Fabbri M, Garzon R, et al. MicroRNA-29 family reverts aberrant methylation in lung cancer by targeting DNA methyl-transferases 3A and 3B. Proc Natl Acad Sci U S A. 2007;104:15805–15810. doi: 10.1073/pnas.0707628104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Denis H, Ndlovu MN, Fuks F. Regulation of mammalian DNA methyl-transferases: a route to new mechanisms. EMBO Rep. 2011;12:647–656. doi: 10.1038/embor.2011.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Ellis L, Atadja PW, Johnstone RW. Epigenetics in cancer: targeting chromatin modifications. Mol Cancer Ther. 2009;8:1409–1420. doi: 10.1158/1535-7163.MCT-08-0860. [DOI] [PubMed] [Google Scholar]
  • 48.Lee DH, O’Connor TR, Pfeifer GP. Oxidative DNA damage induced by copper and hydrogen peroxide promotes CG -> TT tandem mutations at methylated CpG dinucleotides in nucleotide excision repair-deficient cells. Nucleic Acids Res. 2002;30:3566–3573. doi: 10.1093/nar/gkf478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Soussi T, Beroud C. Significance of TP53 mutations in human cancer: a critical analysis of mutations at CpG dinucleotides. Hum Mutat. 2003;21:192–200. doi: 10.1002/humu.10189. [DOI] [PubMed] [Google Scholar]
  • 50.Parry TE. Mutagenic mechanisms in leukemia and cancer: a new concept - cytosine lack could be as mutagenic as cytosine deamination. Leuk Res. 2006;30:1079–1083. doi: 10.1016/j.leukres.2005.12.019. [DOI] [PubMed] [Google Scholar]
  • 51.Zuo SJ, Boorstein RJ, Teebor GW. Oxidative Damage to 5-Methyl-cytosine in DNA. Nucleic Acids Res. 1995;23:3239–3243. doi: 10.1093/nar/23.16.3239. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Chen ZQ, Zhang CH, Kim C, et al. Quantum mechanics study and Monte Carlo simulation on the hydrolytic deamination of 5-methylcytosine glycol. Phys Chem Chem Phys. 2011;13:6471–6483. doi: 10.1039/c0cp02783a. [DOI] [PubMed] [Google Scholar]
  • 53.Ziegel R, Shallop A, Upadhyaya P, et al. Endogenous 5-methylcytosine protects neighboring guanines from N7 and O6-methylation and O6-pyridyloxobutylation by the tobacco carcinogen 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone. Biochemistry. 2004;43:540–549. doi: 10.1021/bi035259j. [DOI] [PubMed] [Google Scholar]
  • 54.Donkena KV, Young CY, Tindall DJ. Oxidative stress and DNA methylation in prostate cancer. Obstet Gynecol Int. 2010:302051-1–14. doi: 10.1155/2010/302051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Sowers LC, Valinluck V, Tsai HH, et al. Oxidative damage to methyl-CpG sequences inhibits the binding of the methyl-CpG binding domain (MBD) of methyl-CpG binding protein 2 (MeCP2) Nucleic Acids Res. 2004;32:4100–4108. doi: 10.1093/nar/gkh739. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Maltseva DV, Baykov AA, Jeltsch A, et al. Impact of 7,8-dihydro-8-oxoguanine on methylation of the CpG site by Dnmt3a. Biochemistry. 2009;48:1361–1368. doi: 10.1021/bi801947f. [DOI] [PubMed] [Google Scholar]
  • 57.Tudek B, Winczura A, Janik J. Involvement of oxidatively damaged DNA and repair in cancer development and aging. Am J Transl Res. 2010;2:254–284. [PMC free article] [PubMed] [Google Scholar]
  • 58.Franco R, Schoneveld O, Georgakilas AG, et al. Oxidative stress, DNA methylation and carcinogenesis. Cancer Lett. 2008;266:6–11. doi: 10.1016/j.canlet.2008.02.026. [DOI] [PubMed] [Google Scholar]
  • 59.Rahman N, Scott RH. Cancer genes associated with phenotypes in monoallelic and biallelic mutation carriers: new lessons from old players. Hum Mol Genet. 2007;16 (Spec1):R60–R66. doi: 10.1093/hmg/ddm026. [DOI] [PubMed] [Google Scholar]
  • 60.Stoger R, Kajimura TM, Brown WT, et al. Epigenetic variation illustrated by DNA methylation patterns of the fragile-X gene FMR1. Hum Mol Genet. 1997;6:1791–1801. doi: 10.1093/hmg/6.11.1791. [DOI] [PubMed] [Google Scholar]
  • 61.Bacolod MD, Schemmann GS, Giardina SF, et al. Emerging paradigms in cancer genetics: some important findings from high-density single nucleotide polymorphism array studies. Cancer Res. 2009;69:723–727. doi: 10.1158/0008-5472.CAN-08-3543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Chen H, Taylor NP, Sotamaa KM, et al. Evidence for heritable predisposition to epigenetic silencing of MLH1. Int J Cancer. 2007;120:1684–1688. doi: 10.1002/ijc.22406. [DOI] [PubMed] [Google Scholar]
  • 63.Dobrovic A, Kristensen LS. Direct genotyping of single nucleotide polymorphisms in methyl metabolism genes using probe-free high-resolution melting analysis. Cancer Epidem Biomar. 2008;17:1240–1247. doi: 10.1158/1055-9965.EPI-07-2531. [DOI] [PubMed] [Google Scholar]
  • 64.Herman JG, Esteller M, Garcia-Foncillas J, et al. Inactivation of the DNA-repair gene MGMT and the clinical response of gliomas to alkylating agents. New Engl J Med. 2000;343:1350–1354. doi: 10.1056/NEJM200011093431901. [DOI] [PubMed] [Google Scholar]
  • 65.Lin HJL, Zuo T, Chao JR, et al. Seed in soil, with an epigenetic view. Biochem Biophys Acta. 2009;1790:920–924. doi: 10.1016/j.bbagen.2008.12.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Lyko F, Brown R. DNA methyltransferase inhibitors and the development of epigenetic cancer therapies. J Natl Cancer Inst. 2005;97:1498–1506. doi: 10.1093/jnci/dji311. [DOI] [PubMed] [Google Scholar]
  • 67.McCabe MT, Brandes JC, Vertino PM. Cancer DNA methylation: molecular mechanisms and clinical implications. Clin Cancer Res. 2009;15:3927–3937. doi: 10.1158/1078-0432.CCR-08-2784. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Metias SM, Lianidou E, Yousef GM. MicroRNAs in clinical oncology: at the crossroads between promises and problems. J Clin Pathol. 2009;62:771–776. doi: 10.1136/jcp.2009.064717. [DOI] [PubMed] [Google Scholar]
  • 69.Myzak MC, Dashwood RH. Histone deacetylases as targets for dietary cancer preventive agents: lessons learned with butyrate, diallyl disulfide, and sulforaphane. Curr Drug Targets. 2006;7:443–452. doi: 10.2174/138945006776359467. [DOI] [PubMed] [Google Scholar]
  • 70.Pandey M, Shukla S, Gupta S. Promoter demethylation and chromatin remodeling by green tea polyphenols leads to re-expression of GSTP1 in human prostate cancer cells. Int J Cancer. 2010;126:2520–2533. doi: 10.1002/ijc.24988. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Paranjape T, Slack FJ, Weidhaas JB. MicroRNAs: tools for cancer diagnostics. Gut. 2009;58:1546–5154. doi: 10.1136/gut.2009.179531. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Peter ME. Let-7 and miR-200 microRNAs Guardians against pluripotency and cancer progression. Cell Cycle. 2009;8:843–852. doi: 10.4161/cc.8.6.7907. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES