Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2016 Jan 1.
Published in final edited form as: Transl Res. 2014 Mar 12;165(1):74–90. doi: 10.1016/j.trsl.2014.03.001

Epigenetics of Lung Cancer

Scott M Langevin 1, Robert A Kratzke 2, Karl T Kelsey 3,4,
PMCID: PMC4162853  NIHMSID: NIHMS576894  PMID: 24686037

Abstract

Lung cancer is the leading cause of cancer-related mortality in the United States. Epigenetic alterations, including DNA methylation, histone modifications, and non-coding RNA expression, have widely been reported in the literature to play a major role in the genesis of lung cancer. The goal of this review is to summarize the common epigenetic changes associated with lung cancer to give some clarity to its etiology, and provide an overview of the potential translational applications of these changes, including applications for early detection, diagnosis, prognostication, and therapeutics.

Keywords: DNA methylation, histone modifications, microRNA, lnc-RNA, non-coding RNA, biomarkers, NSCLC, small cell carcinoma, pulmonary carcinoma

INTRODUCTION

Despite large-scale reductions in cigarette consumption over the past several decades, lung cancer remains the leading cause of cancer-related mortality in the United States and is the second leading cause of death overall, after heart disease [1]. While lung cancer rates have steadily declined among men since the 1980s and seem to have plateaued among women, there still remain an estimated 228,190 new cases and 159,480 deaths each year [1]. This high mortality rate is driven by the high incidence of this disease coupled with its dismal 5-year survival rate of only 17% [1].

The vast majority of lung cancer is can be characterized as small cell (neuroendocrine) carcinoma (SCLC) or non-small cell carcinoma (NSCLC), which broadly includes squamous cell carcinoma, adenocarcinoma, and large cell carcinoma subtypes [2]. NSCLC is by far the more common of the types, accounting for approximately 85% of all lung cancer cases [3]. While these histologies share a common organ of origin and some molecular attributes, they also exhibit unique molecular traits and represent distinct diseases, and it has become clear that it is important to make the distinction between the histologies for the purpose of treating the disease.

While smoking remains the major risk factor for all histologies (especially small cell and squamous cell carcinoma), it is important to note that only some 10% of smokers will ultimately develop lung cancer [4]. Moreover, not all lung cancer patients have a smoking history, as there are other risk factors for the disease. Adenocarcinoma, which accounts for approximately 37% of all lung cancers in the United States, is the most common form among non-smokers [5]. Globally, an estimated 15% of men and 53% of women with lung cancer are never-smokers [6]. Other risk factors for the disease include radon, asbestos, and environmental/occupational exposure to polycyclic aromatic hydrocarbons and other pollutants [2]. However, as with smoking, not everyone who is exposed to these environmental factors will go on to develop lung cancer.

The carcinogenic process is driven by the accumulation of genetic and epigenetic alterations that result in dysregulation of key oncogenes, tumor suppressor genes, and DNA repair/housekeeping genes. The probability of incurring these pathologically important events is largely dependent upon the individual exposome in conjunction with interpersonal phenotypic variability. While genetic heterogeneity accounts for some of the variable risk, it does not explain this phenomenon in toto [7]. Epigenetic variability, including DNA methylation, histone modifications, and non-coding RNA expression, also contribute the phenotypic makeup of an individual (e.g. xenobiotic metabolism, DNA repair capacity, immunity, etc.) and, accordingly, risk of malignancy.

This underscores the importance of enhancing our knowledge of lung cancer epigenetics (in addition to genetics) to fully comprehend the pathogenesis of this disease. Moreover, continued expansion our understanding of various epigenetic events involved with different types of lung cancer expands the potential battery of diagnostic and prognostic biomarkers available to clinicians, as well as introducing new avenues for the discovery of novel therapeutic targets. The primary objective of this review is to summarize the common epigenetic events in lung cancer and provide an overview of the potential translational applications of these events for the management of this disease.

EPIGENETICS & ETIOLOGY

Tumorigenesis

Lung cancer involves an accumulation of genetic and epigenetic events in the respiratory epithelium [8]. While somatic genetic aberrations, such as mutations and copy number alterations, play a well-known role in oncogenesis, epigenetic alterations are in fact more frequent than somatic mutations in lung cancer [9].

Tumor suppressor gene inactivation through promoter methylation, often referred to as hypermethylation, is a hallmark of lung cancer and is an early event in the carcinogenic process [10, 11]. Promoter methylation of specific tumor suppressor genes, along with the overall number of hypermethylated genes, increases with neoplastic progression from hyperplasia to adenocarcinoma [12, 13]. Promoter methylation can couple with mutation or deletion events to inactivate a tumor suppressor gene (i.e. a different inactivation event for each allele). This is because inactivation of one allele for dominantly acting suppressor gene loci is generally insufficient to lead to clonal selection, since the protein can still be produced from one normal allele. However, there is also evidence that, at some gene loci, both copies do not necessarily need to be inactivated to adversely impact the cell, but rather that partial inactivation of one allele, giving rise to haploinsufficiency, (i.e. one wild-type allele is insufficient to provide full functionality) can contribute to carcinogenesis [14]. In these cases, inactivation of one allele by promoter methylation would be sufficient for clonal selection.

Many of the tumor suppressor genes that are hypermethylated in lung cancer are also frequently hypermethylated in other types of solid tumors [6]. Some are specific, although many are not, similar to what is observed for somatic mutations. In premalignant and malignant states, promoter methylation is frequently observed in genes involved with crucial functions, including cell cycle control, proliferation, apoptosis, cellular adhesion, motility, and DNA repair.

Some of the most oft-studied genes in the context of promoter methylation in lung cancer include p16INK4a, RASSF1A, APC, RARβ, CDH1, CDH13, DAPK, FHIT, and MGMT. While p16INK4a is frequently methylated, mutated, or deleted in NSCLC, with estimates for the prevalence of alteration of this gene are around 60%, p14arf, which is also encoded on the CDKN2A gene, is much less commonly inactivated (~8–30% of NSCLC) [15, 16]. Moreover, although a common event in NSCLC, p16INK4a is disrupted in less than 10% of SCLC [15]. Additionally, RASSF1A is deleted or methylated in 30–40% of NSCLC and 70–100% of SCLC [15], FHIT is deleted or methylated in 40–70% of NSCLC and 50–80% of SCLC [15], and TSLC1 is methylated in an estimated 85% of NSCLC [15]. A more extensive list of known, commonly hypermethylated genes in lung cancer is provided in Table 1.

Table 1.

Select genes frequently reported to undergo promoter methylation in lung cancer.

Gene Locus (Ensembl) Frequency in NSCLC Gene function(s) References
CDKN2A/p16INK4a 9.21.3 22–47% a Cyclin Dependent Kinase Inhibitor 2A/p16: key TSG involved in cell cycle arrest at the G1/S-phase checkpoint through inhibition of CDK4/6 [5, 1631]
FHIT 3p14.2 34–47% a Fragile Histidine Triad: member of the histidine triad family involved in purine metabolism; TSG that regulates genes essential for cell proliferation and induction of apoptosis [16, 23, 27, 3437]
APC 5q22.2 30–96% a Adenomatous Polyposis Coli: TSG that acts as a negative regulator of Wnt and also is involved in cell migration and adhesion, transcriptional activation, and apoptosis [16, 17, 27, 28, 3841]
RASSF1A 3p21.31 25–45% a Ras Association (RalGDS/AF-6) Domain Family Member 1: putative TSG involved in apoptosis and cell cycle control [5, 16, 17, 27, 28, 31, 3638, 4245]
DAPK 9q21.33 16–45% a Death-associated protein kinase: putative TSG involved in apoptotic signaling [16, 28, 29, 44, 46]
RAR β 9p24.2 26–45% b Retinoic acid receptor beta: involved in cell growth and differentiation [16, 17, 37, 40, 47]
MGMT 10q26.3 11–38% a O-6-methylguanine DNA methyltransferase: DNA repair enzyme that removes alkyl lesions from the O6 position of guanine [17, 22, 4850]
SHOX2 3q25.32 91–95% c Homeobox family gene involved in gene transcription with putative involvement in cell growth and differentiation [51, 52]
RUNX3 1p36.11 25% d Runt-domain containing transcription factor that functions as a TSG [36, 5357]
CDH13 16q23.3 28–30% b H-cadherin: TSG involved in regulation of cell growth, proliferation, and apoptosis [5, 28, 31, 40, 58]
CDH1 16q22.1 12–58% a E-cadherin; TSG that negatively regulates cell-cell adhesion, growth, motility, and proliferation; loss of function in cancer is thought to increase proliferation, invasion, and metastasis [17, 28, 38, 42, 58]
TSLC1 11q23.3 37–44% d Cell adhesion molecule 1: TSG involved in cell-cell adhesion [5961]
ASC/TMS1 16p11.2 35–47% a PYD And CARD Domain-Containing Protein: crucial mediator of apoptosis and inflammation [62, 63]
DAL1 18p11.31 55–57% d Erythrocyte Membrane Protein Bank 4.1-Like 3; TSG involved in cell cycle arrest and apoptosis [59, 64]
PTEN 10q23.31 26% b Phosphatase And Tensin Homolog: TSG involved in cell cycle control and survival by acting as a negative regulator of the AKT/PKB pathway [22, 65, 66]
GSTP1 11q13.2 15% b Glutathione S-Transferase P1: functions as a phase II xenobiotic detoxification enzyme [17, 27, 67, 68]

Abbreviations: NSCLC = non-small cell lung cancer; SCLC = small cell lung cancer; TSG = tumor suppressor gene

a

Estimate based on [32, 33]

b

Estimate based on [32]

c

Estimate includes NSCLC and SCLC [51, 52]

d

Estimate based on [33]

Hypermethylation of CDKN2A may occur early in the genesis of some lung cancers, having been identified in premalignant lesions [69]. Promoter methylation of RASSF1A, APC, ESR1, ABCB1, MT1G, and HOXC9 have been associated with Stage I NSCLC [70], suggesting that they too may occur relatively early on in the development of the cancer. CpG island methylation of homeobox-associated genes is also common Stage I lung cancer, appearing in nearly all early-stage tumors [71]. Conversely, other commonly hypermethylated genes, such as hDAB2IP, H-Cadherin, DAL-1, and FBN2, have been associated with advanced stage NSCLC [64, 72, 73], suggesting that these changes may occur at a later point in the carcinogenic process. It is important to note, however, that later involvement does not preclude the importance of the modification in the development of the disease, as they may play key roles in the ability of the cancer to continue to flourish in its advanced state, evade host immunity or exogenous cancer treatments, or to metastasize locally and/or systemically. Further, it is critical that these generalized “temporal” observations are kept in perspective, as lung cancer is a very heterogeneous disease and each tumor is unique; an early event in one tumor may not occur until later on in another. Moreover, it must be considered that most of these observations are based on cross-sectional measurement of malignant or premalignant tissue, and that linear consideration of the sequential timing of molecular aberrations likely represents a gross oversimplification of cancer development and progression.

There is some evidence for CpG island methylator phenotype (CIMP), a tumor phenotype characterized by widespread hypermethylation, in lung cancer [7477]. This is not wholly surprising, given that DNA methyltransferases (DNMT), the group of enzymes that catalyze the covalent attachment of the methyl group to the cytosine base, are upregulated in NSCLC [68, 78, 79].

In contrast to gene-specific hypermethylation, which can occur early on in cancer development, genomic hypomethylation, a genome-wide loss of methylation, may be a late event in the genesis of lung cancer, at least for adenocarcinoma [5]. However, there is not presently a clear consensus on the timing, as Anisowicz found that hypomethylation was associated with NSCLC progression from normal to lung cancer [80]. Regardless, widespread hypomethylation has been associated with genomic instability in NSCLC [81] and can result in oncogene activation [82, 83] and loss of imprinting [84]. In lung cancer, hypomethylation tends to occur at nuclear elements, LTR elements, segmental duplicates, and subtelomeric regions (loss of methylation is much less common at non-repetitive sequences) [85]. In addition to the genomic loss of methyl content, gene-specific hypomethylation has been reported for several loci, including MAGEA [86, 87], TKTL1 [86, 88], BORIS [89, 90], DDR1 [31] 14-3-3σ [91, 92], and TMSB10 [83, 93]. MAGE overexpression with an associated loss of methylation has been observed in 75–80% of NSCLC [94].

CpG methylation can also induce point mutations through deamination of 5-methylcytosine (5-meC) or enhancement of exongenous carcinogens. Methylated cytosine can undergo hydrolytic deamination causing a C to T transition [95]. More than 30% of disease-related germline point mutations occur at CpG dinucleotides [95]. Furthermore, nearly half of all somatic and one-third of all germline p53 mutations take place at methylated CpGs, and many common p53 mutations that manifest in somatic cells are caused by C to T transitions, including “hot spot” mutations at codons 248, 273, and 282 [96]. The risk of p53 mutation at 5-meC is 10-fold that of unmethylated cytosine, and CpG dinucleotides in these regions have been observed to be methylated in normal tissue [96]. Secondly, DNA methylation can enhance the mutagenic effect of exogenous carcinogens [95]. An example of this is the affinity of benzo(α)pyrene diol epoxide (BPDE) for adduct formation on guanine bases adjacent to 5-meC, resulting in G to T transversions in aerodigestive tract cancers in smokers [9799]. Similarly, acrolein has an affinity for binding 5-meC, which can induce C to T transitions [100].

Histone deacetylases (HDAC) are overexpressed in lung cancer [101103]. HDACs catalyze the removal of acetyl groups on the histone tail resulting in a transcriptionally inactive heterochromatic state [104]. Likewise, SIN3A (part of an HDAC repressor complex) is downregulated in NSCLC [105]. Cross talk occurs, mediated by proteins with methyl-binding domains that can recruit histone modifying enzymes (MBDs, KAISO, MeCP2), such as HDAC, that closely couples DNA methylation and histone modifications [9]. Relative to normal lung, lung cancer undergoes H4K5/H4K8 hyperacetylation, H4K12/H4K16 hypoacetylation, and H4K20me3 [106]. Lower global levels of H4K20me3 can be detected in precursor lesions and is particularly common in squamous cancers [106].

Other post-translational modifiers that have been reported as overexpressed in lung cancer relative to normal lung tissue include polycomb group genes (PcG), which are crucial epigenetic regulators of stem cell (and cancer stem cell) survival and pluripotency. Specifically, upregulation has been observed for PcGs that form the Polycomb repressive complexes (PRC1 and PRC2) [107118], particularly for the respective catalytic subunits, B-cell-specific Moloney Murine Leukemia Virus Integration Site 1 (BMI-1), which is involved in gene silencing through ubiquitylation of histone 2A (H2A) [119], and Enhancer of Zeste Homologue 2 (EZH2) [111114, 116118], which controls gene expression through histone H3 Lysine 27 trimethylation or by interacting with DNMTs to signal transcriptional repression [119].[109] BMI-1 is overexpressed at higher levels in SCLC relative to NSCLC [109], although, unlike EZH2, it is constitutively expressed in normal lung tissue [120, 121]. Upregulation of Polycomb repressive complex genes have been associated with lung cancer proliferation, survival, and epithelial-mesenchymal transition (EMT) [108, 109, 112, 122, 123].

Non-coding RNA are a class of RNA sequences that are transcribed but do not encode proteins. MicroRNA are small non-coding RNA molecules (~18–22 nt) that can negatively regulate the expression of hundreds of mRNA targets and are frequently dysregulated in lung cancer. Two recent meta-analyses have reported on microRNAs for which expression is commonly reported to be dysregulated in lung cancer [124, 125], which are in summarized in Table 2. In particular, miR-196a and miR-200b are highly overexpressed in lung cancer, with estimated fold-changes exceeding 23- and 37-times, respectively [124]. Much less is presently known about the precise functions of long non-coding RNA (lncRNA), although there is evidence that they function as protein regulators and structural organizers, in addition to regulation of gene expression [126]. Two lncRNA in particular that have been implicated in lung cancer are metastasis-associated lung adenocarcinoma transcript 1 (MALAT1) and Hox transcript antisense intergenic RNA (HOTAIR). MALAT1 is a highly conserved lncRNA sequence of ~8000 nt in length [127] that is overexpressed in NSCLC (including adenocarcinoma, squamous cell carcinoma, and large cell carcinoma) [128, 129] and is thought to play a role in cell motility, invasion, and metastasis, at least in part through posttranslational inhibition of expression of genes regulating these functions [130132]. HOTAIR is also a highly conserved lncRNA (at least in mammals) that is also reported to be overexpressed in lung cancer [133135] and interacts with PRC2 and histone regulatory complexes, and regulates HOXD expression [126].

Table 2.

MicroRNAs commonly dysregulated in lung cancer.

microRNAa Locus (Ensembl) Top 10 putative targets (TargetScanHuman v.6.2)
Upregulated
miR-21 17q23.1 ZNF367; GPR64; YOD1; PHF14; PLEKHA1; PIKFYVE; PBRM1; GATAD2B; SCML2; VCL
miR-31 9p21.3 RSBN1; ARHGEF2; IDE; NR5A2; SH2D1A; ZNF512; PRKCE; PIK3C2A; PEX; SATB2
miR-92b 1q22 CD69; FNIP1; SLC12A5; MAN2A1; ACTC1; BFXW7; ASPH; DCAF6; SYN2; EFR3A
miR-182 7q32.2 RGS17; MITF; MFAP3; CTTN; TMEM20; EDNRB; ACTR2; CAMSAP1L1; EPAS1; NPTX1
miR-183 7q32.2 ABAT; AKAP12; PIGX; PFN2; PTPN4; REV1; ITGB1; KCNK10; C20orf177; FRMD6
miR-193b 16p13.12 ABI2; IL17RD; SLC10A6; DCAF7; FLI1; JUB; SON; ERBB4; FHDC1; IDE
miR-196a 17q21.32 HOXC8; HOXA7; SLC9A6; HOXA9; KLHL23; PHOSPHO2-KLHL23; ZMYND11; PACRGL; HOXB8; MAP3K1
miR-200b 1p36.33 ZEB1; FAM122C; ZEB2; LRP1B; WIPF1; ELL2; MCFD2; RECK; SEC23A; C7orf58
miR-203 14q32.33 ZNF281; CAMTA1; B3GNT5; LIFR; ABCE1; NUDT21; AFF4; PRPS2; SEMA5A; SMAD9
miR-205 1q32.2 ZNF606; CMTM4; DMXL2; BTBD3; LPCAT1; SECISBP2L; YES1; C16orf52; CHN1; DLG2
miR-210 11p15.5 THSD7A; ISCU; ZNF462; NR1D2; DIMT1L; FAM116A; ARMC1; NEUROD2; ELFN2; SYNGAP1
miR-708 11q14.1 GON4L; SEMA4C; KIAA0355; HOXB3; ALG9; FOXJ3; EFR3B; DCUN1D5; MPL; RNF150

Downregulated
miR-30a 6q13 PPARGC1B; ANKRA2; MKRN3; LMBR1; EED; LHX8; KLHL20; SNX16; SCN2A; CELSR3
miR-30b 8q24.22 PPARGC1B; ANKRA2; MKRN3; LMBR1; EED; LHX8; KLHL20; SNX16; SCN2A; CELSR3
miR-30d 8q24.22 PPARGC1B; ANKRA2; MKRN3; LMBR1; EED; LHX8; KLHL20; SNX16; SCN2A; CELSR3
miR-101 1p31.3 FAM108C1; GLTSCR1; ZNF654; EYA1; FLRT3; CYDL; TNPO1; LCOR; MYCN; TET2
miR-126-3p 9q34.3 PTPN9; PLXNB2; RGS3; KANK2; EFHD2; CAMSAP1; ZNF219; SPRED1; LRP6; RNF165
miR-126-5p 9q34.3 [not available in the TargetScanHuman database]
miR-138 3p21.32 GPR124; CREB3L2; RMND5A; NFIX; WWC1; RARA; PPIP5K1; SLC35F1; SYT13; CLEC1A
miR-139-5p 11q13.4 TMF1; USP6NL; TBX1; SCAPER; NDRG2; DCBLD2; SLC9A2; PRDM16; EBF1; MORN4
miR-140-3p 16q22.1 GOCP; LOC221710; STAG2; ACVR2B; VGLL2; CBL; UBAP2L; FAIM; PITPNM3; MARCKS
miR-143 5q32 DENND1B; VASH1; ASAP3; SLC30A8; ABL2; TTPA; SLC25A15; MSI2; EPM2AIP1; AKAP6
miR-145 5q32 TPM3; FSCN1; SRGAP2; FAM108C1; NAV3; ABCE1; KCNA4; PLDN; FLI1; SEMA3A
miR-451 17q11.2 OSR1; PSMB8; TTN; TSC1; C11orf30; AEBP2; S1PR2; MEX3C; CAB39; SAMD4B
miR-486-3p 8p11.21 NAT15; RGAG4; SF3A1; TMEM132E; GDI1; FLOT2; LPPR2; DMBX1; FYCO1; CNP
miR-486-5p 8p11.21 RAB8B; EFHC1; ST3GAL1; SSH3; RBM4; TTBK1; UNC45A; NUCB1; TAOK1; SDC3
let-7ab 22q13.31 C14orf28; FIGNL2; MHGA2; LIN28B; TRIM71; IGDCC3; ARID3B; STARD9; PTAFR; THRSP
a

Based on two recent meta-analyses of microRNA expression in human lung cancer [124, 125]

b

Although not considered in either meta-analysis, this is an important and historic tumor suppressor microRNA that has been oft-studied in the context of lung cancer and thus the authors have decided to include it in this table

There have also been several well-documented examples of interaction between DNA methylation and microRNAs in lung cancer. For example, miR-124a [136], miR-34b/c [137, 138], miR-886-3p (in SCLC) [139], miR-9-3 [140] and miR-193a [140] have all been reported as frequently hypermethylated in lung tumors, while let-7a-3 [141] is often hypomethylated. Moreover, miR-29 targets the transcripts of de novo DNMT (DNMT3a/b). Additionally, expression of miR-29 is inversely associated with DNMT3a/b expression in lung tumors, and has been shown to restore normal DNA methylation patterns in experimental models (in vitro and in vivo) [142].

Some epigenetic alterations reported for lung cancer may be smoking-specific (i.e. only found in lung tumors of smokers). Genes reported to undergo smoking-specific promoter methylation include APC, FHIT, RASSF1A, and CCND2 [21, 35, 40, 143, 144]. Also, the frequency of promoter methylation of p16INK4a, MGMT, RASSF1A, MTHFR, and FHIT is higher in the NSCLC tumors of smokers relative to non-smokers [21, 22, 145147]. Moreover, RARβ, p16INK4a, FHIT, and RASSF1A promoter methylation increases with increasing smoking intensity [21, 148150]. Interestingly, genes whose silencing is associated with duration or amount of tobacco smoking are likely later stage contributors to this disease, as long-term exposure to carcinogenic stimuli would imply a later selection of existing clones. Experimentally, genomic hypomethylation and promoter methylation of RASSF1A and RARβ were observed when normal small airway epithelial cells and immortalized bronchial epithelial were exposed to cigarette smoke condensate [151]. There is also experimental evidence indicating that cigarette condensate decreases nuclear levels of H4K16ac and H4K2me3 in respiratory epithelial cells [152]. Conversely, RASSF2, TNFRSF10, BHLHB5 and BOLL have been reported to be more frequently hypermethylated in NSCLC from never-smokers [153, 154].

DNMT1 expression is particularly high in lung cancer tumors of smokers [155]. The tobacco-specific nitrosamine, 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone (NNK), has been associated with increased levels of DNMT1. Experimental evidence indicates that NNK does not affect the actual DNMT1 mRNA level but rather inhibits DNMT1 degradation via Akt signaling, which stabilizes the DNMT1 protein by decreasing ubiquitylation, resulting in elevated accumulation and, accordingly, aberrant tumor suppressor gene methylation [151, 155, 156]. More specifically, NNK stimulates Akt, which inactivates GSK3β ser/thr kinase, which would otherwise phosphorylate DNMT1 and recruit ubiquitylases. This is similarly observed for laryngeal [157] and esophageal squamous cancers [158]. Corroborating this point, NNK has been shown to induce promoter hypermethylation of tumor suppressor genes in lung tumors in mice and rats [159161].

Moreover, chronic inflammation, which occurs in response to cigarette smoking, also plays an important role in lung cancer development, stimulating cellular turnover and proliferation. Inflammation has long been associated with DNA methylation in lung cancer [162164]. There is evidence that reactive oxygen species (ROS), such as are generated during chronic inflammation, target transcriptional repressors, giving rise to increased levels of DNA methylation [165]. Cigarette smoke also inhibits the metabolism and storage of folate [166]. Nitrates, nitrous oxide, cyanates and isocyanates found in tobacco smoke have been shown to transform folate, a major source of methyl groups for one-carbon metabolism, into a biologically inactive compound in experimental models [167, 168]. In further support of this, reduced serum folate levels have been observed in smokers relative to non-smokers [169, 170]. One-carbon metabolism is a critical pathway in the DNA methylation process, and depletion of folate can negatively impact the availability of s-adenosylmethionine (SAM), the primary methyl donor in the cytosine methylation reaction. Consequently, folate deficiency can result in chromosomal damage through impaired nucleotide synthesis and aberrant DNA methylation [171, 172].

To this point, any discussion in this review relating epigenetics to cancer risk factors has primarily focused on tobacco-smoke associated events. However, as discussed earlier, not all lung cancer patients, particularly adenocarcinoma patients, have a history of smoking. Exposure to radon gas is another major risk factor for lung cancer. Experimentally, radon-exposed transformed bronchial epithelial cells exhibit altered microRNA expression profiles [173]. Further, a study of Chinese uranium miners found a dose-response between radon exposure and promoter methylation of p16INK4a and MGMT in sputum samples [174], although this was not observed in a similar study of uranium miners by Pulling et al [175].

Tissue specificity

Historically, adenocarcinoma and squamous cell carcinoma have been considered together, along with large cell carcinoma, in a heterogeneous class of lung cancers referred to as “non-small cell carcinoma’. However, this broad categorization may no longer be appropriate given our understanding of the biological and clinical differences between the histologic subtypes. Making the distinction between squamous cell carcinoma and adenocarcinoma based on morphology alone can be a challenge, necessitating additional markers for accurate discrimination.

The importance of discriminating NSCLC by histology was highlighted by data from a clinical trial of the VEGF antibody bevacizumab, in which elevated toxicity (severe pulmonary hemorrhage) was observed among patients with pulmonary squamous cell carcinoma but not adenocarcinoma [176]. Bevacizumab, in combination with platinum-based chemotherapy, has shown to be efficacious in non-squamous NSCLC in two Phase III trials (E4599 and AVAiL) [177, 178]. Further, Pemetrexed, an antimetabolite, is more efficacious in terms of overall and progression-free survival among patients with non-squamous NSCLC relative to other histologies [179].

Distinct DNA methylation patterns have been identified for adenocarcinoma and squamous cell carcinoma, respectively [180183]. Kwon and colleagues identified a methylation panel that could distinguish histologic class of NSCLC [180]. In another regard, Christensen and colleagues were able to accurately discriminate lung adenocarcinoma from mesothelioma, which can also be challenging but is very important to differentiate diagnostically, based on the DNA methylation profile [184]. There are no significant large-scale differences in overall DNA methylation levels by lung cancer histology [32], although there are locus-specific differences. Additionally, there are also reports of histology-specific microRNA expression patterns. For example, miR-205 is reported to be SCC-specific [185, 186], while miR-124a has been reported as specific for adenocarcinoma [136].

Early detection and diagnosis

Lung cancer mortality could be significantly reduced with earlier detection of the disease. However, only about 15% of lung tumors are localized at diagnosis, with the majority presenting at an advanced stage [2]. Five-year survival for lung cancer is markedly better for early-stage patients, with a less than 10% 5-year survival for advanced stage patients versus better than 70% for early-patients [187]. Spiral computed tomography (CT) has shown promise for the early detection of lung cancer but has a high false-positive rate [188], with up to up to 30% of indeterminate nodules identified by CT ultimately found to be benign [189], indicating a need for development of additional markers to enhance the specificity.

As previously discussed, promoter hypermethylation can be an early event in lung carcinogenesis, and as such, may have utility in early detection of the disease. For instance, promoter hypermethylation of p16INK4a has been observed in NSCLC precursor lesions [13, 18], and PTPRN2 promoter methylation is reported to be an early event in pulmonary adenocarcinoma, with detectable changes in the premalignant atypical adenomatous hyperplasia [5].

Importantly, some of these early epigenetic events can be detected by non- or minimally-invasive sample collection techniques, an important characteristic for cancer screening applications. For example, aberrant DNA methylation can be detected in sputum [190, 191], bronchoalveolar aspirate/lavage [39, 145, 192194], and saliva [195, 196] in lung cancer patients. Furthermore, CDKN2A and MGMT promoter methylation was detected in sputum up to 3-years prior to lung cancer diagnosis [20], and promoter methylation of p16INK4a, MGMT, PAX5b, DAPK, GATA5, and RASSF1A was detected in sputum 18 months prior to lung cancer diagnosis in another study [197]. However, specificity can be an issue for some of these early markers, as they can also be detected in individuals who will not ever go on to develop disease. To this point, promoter methylation of p16INK4a has been detected in sputum from former and current smokers [10], which underscores the importance of multimarker panels. However, not all single markers are non-specific, as exemplified by SHOX2 promoter methylation, which has demonstrated good sensitivity (68–78%) and specificity (95–96%) for NSCLC in bronchial aspirates (AUC = 86–94%) [51, 193].

Additionally, cancer-specific DNA methylation [38, 198], microRNA [199, 200], and lncRNA [201] profiles have been identified in the blood of lung cancer patients. Meta-analysis on the diagnostic value of circulating microRNAs for lung cancer reports a meta-ROC of 0.92 [202], highlighting the potential diagnostic capacity.

PROGNOSTIC BIOMARKERS

Identification of novel biomarkers to aid in prediction of outcome and tumor response is of paramount importance to optimize therapeutic efficacy and prevent over- or under-treatment of lung cancer patients. Epigenetic biomarkers, in particular DNA methylation and microRNA expression, have unique properties that make them well-suited as potential prognostic markers, and accordingly have been widely studied and reported in the literature.

Not surprisingly, many of the hypermethylated tumor suppressor genes that are implicated in lung cancer oncogenesis have also been reported to be associated with prognosis. Promoter methylation of RASSF1A [16, 36, 203], PTEN, DAPK [29], p16INK4a [21, 24, 203206], Wif-1, CXCL12 [207], DLEC1 [208], MLH1 [208], CDH1 [58, 204], CDH13 [58], APC [41, 209], RUNX3 [36], SPARC, and DAL1 have all been associated with NSCLC outcome [22, 36, 64, 207, 208, 210212]. Additionally, DNMT1 overexpression in NSCLC is associated with decreased survival [78, 79, 213]; and DNMT3b only in patients <65 years old [213]. Along similar lines, CIMP has also been correlated with prognosis in NSCLC [74].

Relative to advanced stage lung cancers, chemotherapeutic recommendations are not as clear for early stage disease, with no true consensus as to the optimal approach [214]. Early stage lung cancer can be locally controlled but exhibits a high recurrence rate. Completely resected Stage IB and II tumors have a nearly 50% recurrence rate with a median time to recurrence of 1-year [215]. Fewer recur in Stage IA tumors, although certain IA subsets have high recurrence rates [216]. Methylation of p16INK4a, RASSF1A, CDH13, and APC has associated with early recurrence in surgically treated Stage I NSCLCs [210]. The combination of FHIT and p16INK4a promoter methylation has also been associated with recurrence in Stage I NSCLC [205].

Several global histone modifications have been associated with lung cancer survival; in particular decreased levels of H3K4diMe have been associated with poorer outcome [217, 218]. Additionally, the combination of several histone modifications have been reported to predict survival (H3K4me2, H3K9ac, H2AK5ac) [217], and H4K20me3 down-regulation has been associated with poorer survival in Stage I lung adenocarcinoma patients [106].

Higher expression of the PRC2 protein, EZH2, in NSCLC tumor tissue has been associated with worse overall [112, 118, 219] and disease-free survival [112] in several studies[117]. Likewise, BMI-1 expression has also been correlated with poorer patient outcomes [107, 121, 220], and amplification and concordant overexpression of the PcG, PHC3, has recently been associated with poorer outcomes in NSCLC [221].

Noncoding RNA have also been associated with lung cancer outcomes. A number of miRNA profiles of varying size and combination have been reported in tumor tissue [222]. In particular, mir-21and mir-155 have been the most oft studied as prognostic markers. In a recent meta-analysis, high miR-21 expression was associated with poorer overall survival in NSCLC (metaHR = 2.32, 95% CI: 1.17–4.62; 6 studies) and recurrence-free survival in adenocarcinoma (metaHR = 2.43, 95% CI: 1.67–3.54; 3 studies) [222]. Likewise, in the same meta-analysis, miR-155 was reported to be associated with poorer recurrence-free survival in NSCLC (metaHR = 1.42, 95% CI: 1.10–1.83; 5 studies). Additionally, expression of the lncRNA MALAT1 has been associated with metastasis and poorer survival in NSCLC patients [128, 129], and HOTAIR expression has been associated with aggressive behavior, poorer outcomes in NSCLC [133135, 223].

MicroRNA are often released from tissue during apoptosis or in exosomes or microvesicles and can be detected in the blood stream [224]. Circulating microRNA have been associated with prognosis. Several studies have reported an association between microRNA expression in blood or plasma and lung cancer outcome [225232]. In particular, high-levels of circulating miR-21 have been consistently reported to be associated with outcome [225, 229, 231, 232].

Methylation of circulating DNA may also be of use in outcome prediction, although it is important to distinguish free-circulating DNA from leukocytic DNA, since DNA methylation marks are tightly coupled to cellular differentiation and vary by cell type [233]. Thus, without careful study design, blood-based methylation profiles can be confounded by variation in relative circulating proportions of leukocyte types associated with outcome, such as immune response [7]. Ramirez and colleagues found that 14-3-3 sigma methylation in pretreatment serum may be an important predictor of NSCLC outcome in patients treated with platinum-based chemotherapy [234]. Similarly, Ponomaryova and colleagues found that increased methylation of RARβ 2 in circulating DNA was associated with lung cancer progression [235].

Recent evidence suggests that epigenetic mechanisms may be, at least in some instances, major players in treatment resistance. A number of studies have reported an association between altered expression of a variety of miRNA and sensitivity to gefitinib [236241], erlotinib [242, 243], cisplatin [244259], and radiotherapy [253, 260265]. Elevated expression of HOTAIR [223] and EZH2 [117] have also each been associated with cisplatin-resistance.

In a very elegant study, Sharma et al. demonstrated the capacity of a small subset of stem-like cells in two NSCLC cell-lines to undergo chromatin remodeling after treatment with erlotinib and cisplatin treatment to acquire treatment-resistance [266]. The authors found that the treatment-resistant tumor cells overexpressed the histone demethylase, KDM5A, and exhibited reduced H3K4 di- and trimethylation, and that these resistant cells were sensitive to HDAC inhibitors. These findings are extremely important, as they (1) demonstrate a major role for epigenetic plasticity in acquired drug-resistance, and (2) suggest the potential for histone post-translational modifications as therapeutic targets in drug-resistant NSCLC cells.

Intratumoral heterogeneity and clonal selection of treatment-resistant subpopulations of tumor cells has long been considered as another possible mechanism that can drive acquired treatment resistance in tumors. This notion of intratumoral genetic heterogeneity has recently been confirmed in renal cell carcinoma [267], and very likely is the case in other solid tumor types as well. Moreover, although not assessed in the study, cancer cells are presumably also epigenetically heterogenous within the tumor, as epigenetic alterations are common activating/inactivating events in carcinogenesis, and thus would be susceptible to stochastic selection as well. However, at this point, more research on this topic is needed to determine the extent and generalizability of these findings.

PHARMACOLOGIC TARGETS

Since DNA methylation and transcriptionally inactive heterochromatin conformation are frequent causes of tumor suppressor gene inactivation, therapeutic strategies that globally target epigenetic inactivation of genes have been pioneered for cancer treatment. Experimentally, HDAC [268273] and DNMT [273, 274] inhibitors can restore tumor suppressor gene expression and, consequently, critical pathway function for small-cell and NSCLC patients. Other inhibitors of post-translational modifying proteins have also demonstrated promising antitumor properties in studies of lung cancer in vitro, including the Bromodomain and extra terminal domain (BET) protein inhibitor, JQ1 [275], and EZH2 inhibitors, 3-Deazaneplanocin A (DZNeP) [116, 276] and GSK126 [116].

Clinically, two classes of epigenetic regimens have been clinically tested for NSCLC [277]: (1) DNMT inhibitors and (2) HDAC inhibitors. The HDAC inhibitors vorinostat and romidepsin are each currently approved by the FDA for treatment of cutaneous T-cell lymphoma [278]. However, none are currently approved for treatment of lung cancer, although a recently completed clinical trial demonstrated promising results from combining 5-azacytidine, a DNA demethylating agent, with entinostat, an HDAC inhibitor, for the treatment of advanced chemorefractory NSCLC [279]. Additionally, other epigenetic treatment strategies are currently in clinical trials. The majority of these strategies combine HDAC inhibitors, with or without the DNMT inhibitor 5-azacytidine, with cytotoxic drugs, although one phase II trial is presently underway to test the efficacy of fluoro-2-deoxycytidine (FdCyd), a DNMT inhibitor, in conjunction with tetrahydrouridine, a cytidine deaminase inhibitor that prevents FdCyd from being metabolized, in the treatment of NSCLC. A complete list of ongoing trials (as of December 31, 2013) is presented in Table 3 [140].

Table 3.

Ongoing clinical trials involving epigenetic therapeutic agents for treatment of lung cancer patients, as of December 31, 2013.

Clinical Trial Phase Protocol IDs Intervention(s)
HDAC Inhibitors
Chidamide in Combination With Carboplatin and Paclitaxel in Advanced Non-small Cell Lung Cancer II CDM204, NCT01836679 Evaluate the efficacy and safety of chidamide (CS055/HBI8000; benzamide class HDI) combined with paclitaxel and carboplatin in patients with advanced NSCLC
Chemotherapy With or Without Epigenetic Priming in NSCLC Patients II J1309, NCT01846897 To evaluate the efficacy of epigenetic priming by 5-azacytidine (U-18496; DNMT inhibitor) and entinostat (MS 275/SNDX-275; benzamide class HDI) prior to treatment with standard chemotherapy for NSCLC patients
Phase II Anti-PD1 Epigenetic Priming Study in NSCLC II J1353, NA_00084192, NCT01928576 Determine the response rate to Nivolumab following epigenetic priming with 5-azacitidine (U-18496; DNMT inhibitor) with or without entinostat (MS 275/SNDX-275; benzamide class HDI) for NSCLC patients
Treatment of Locally Advanced Non-Small Cell Lung Cancer (NSCLC) I FER-TH-031, NCI-2010-01913, NCT01059552 Determine the maximum tolerated dose of the combination of vorinostat (L-001079038; suberoylanilide hydroxamic acid HDI), cisplatin, pemetrexed, and radiation therapy in patients with unresectable stage IIIA/IIIB NSCLC
Study of Cisplatin and Pemetrexed in Combination With Panobinostat in Solid Tumors I UCDCC#220, NCT01336842 Determine if panobinostat (LBH589; cinnamic hydroxamic acid analogue HDI) can be safely combined with cisplatin and pemetrexed without increasing side effects and that the combination will be more efficacious than platinum-based doublet chemotherapy alone in solid tumor patients (including NSCLC)
A Trial of Oral 5-azacitidine in Combination With Romidepsin in Advanced Solid Tumors, With an Expansion Cohort in Non-small Cell Lung Cancer I J11102, NA_00052054, NCT01537744 Determine if 5-azacitidine (U-18496; DNMT inhibitor) combined with intravenous romidepsin (FK228/FR901228/NSC 630176; depsipeptide HDI) is effective in the treatment of advanced solid tumors (including NSCLC)

DNMT Inhibitors
A Multi-Histology Phase II Study of 5-Fluoro-2-Deoxycytidine With Tetrahydrouridine (FdCyd + THU) II 090214, 09-C-0214, NCT00978250 Evaluate the efficacy of fluoro-2-deoxycytidine (FdCyd; DNMT inhibitor) combined with tetrahydrouridine for treatment of solid tumor patients (including NSCLC)

Abbreviations: HDAC = histone deacetylase; DNMT = DNA methyltransferase; NSCLC = non-small cell lung cancer; HDI = histone deacetylase inhibitor

CONCLUSIONS

Epigenetics play a powerful role in the etiology of lung cancer and have potential utility as diagnostic and prognostic biomarkers. This may be particularly true for DNA methylation marks, due to their relative stability and amenability to PCR-based measurement, and microRNAs, which tend to be much more stable than mRNA and can each regulate up to hundreds of different gene transcripts. Expanding our understanding of how epigenetic events contribute to the genesis of lung cancer, and how these can be translated into clinically-relevant biomarkers and therapeutic targets, will enhance our ability to properly manage lung cancer and ultimately reduce the heavy global burden of this devastating disease.

Acknowledgments

This work was supported by the National Institutes of Health (R01CA100679 and R01CA126939 to KTK; and K22CA172358 to SML).

Footnotes

All authors have all read the journal’s authorship agreement and policy on conflicts of interest and have no conflicts to declare.

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Siegel R, Naishadham D, Jemal A. Cancer statistics, 2013. CA: a cancer journal for clinicians. 2013;63:11–30. doi: 10.3322/caac.21166. [DOI] [PubMed] [Google Scholar]
  • 2.Dela Cruz CS, Tanoue LT, Matthay RA. Lung cancer: epidemiology, etiology, and prevention. Clinics in chest medicine. 2011;32:605–44. doi: 10.1016/j.ccm.2011.09.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Lu F, Zhang HT. DNA methylation and nonsmall cell lung cancer. Anatomical record. 2011;294:1787–95. doi: 10.1002/ar.21471. [DOI] [PubMed] [Google Scholar]
  • 4.Massion PP, Carbone DP. The molecular basis of lung cancer: molecular abnormalities and therapeutic implications. Respiratory research. 2003;4:12. doi: 10.1186/1465-9921-4-12. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Selamat SA, Galler JS, Joshi AD, Fyfe MN, Campan M, Siegmund KD, et al. DNA methylation changes in atypical adenomatous hyperplasia, adenocarcinoma in situ, and lung adenocarcinoma. PloS one. 2011;6:e21443. doi: 10.1371/journal.pone.0021443. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Herceg Z, Vaissiere T. Epigenetic mechanisms and cancer: an interface between the environment and the genome. Epigenetics : official journal of the DNA Methylation Society. 2011;6:804–19. doi: 10.4161/epi.6.7.16262. [DOI] [PubMed] [Google Scholar]
  • 7.Langevin SM, Kelsey KT. The fate is not always written in the genes: epigenomics in epidemiologic studies. Environmental and molecular mutagenesis. 2013;54:533–41. doi: 10.1002/em.21762. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Dumitrescu RG. Epigenetic markers of early tumor development. Methods in molecular biology. 2012;863:3–14. doi: 10.1007/978-1-61779-612-8_1. [DOI] [PubMed] [Google Scholar]
  • 9.Brzezianska E, Dutkowska A, Antczak A. The significance of epigenetic alterations in lung carcinogenesis. Molecular biology reports. 2013;40:309–25. doi: 10.1007/s11033-012-2063-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Belinsky SA, Klinge DM, Dekker JD, Smith MW, Bocklage TJ, Gilliland FD, et al. Gene promoter methylation in plasma and sputum increases with lung cancer risk. Clinical cancer research : an official journal of the American Association for Cancer Research. 2005;11:6505–11. doi: 10.1158/1078-0432.CCR-05-0625. [DOI] [PubMed] [Google Scholar]
  • 11.Zochbauer-Muller S, Minna JD, Gazdar AF. Aberrant DNA methylation in lung cancer: biological and clinical implications. The oncologist. 2002;7:451–7. doi: 10.1634/theoncologist.7-5-451. [DOI] [PubMed] [Google Scholar]
  • 12.Chung JH, Lee HJ, Kim BH, Cho NY, Kang GH. DNA methylation profile during multistage progression of pulmonary adenocarcinomas. Virchows Archiv : an international journal of pathology. 2011;459:201–11. doi: 10.1007/s00428-011-1079-9. [DOI] [PubMed] [Google Scholar]
  • 13.Licchesi JD, Westra WH, Hooker CM, Herman JG. Promoter hypermethylation of hallmark cancer genes in atypical adenomatous hyperplasia of the lung. Clinical cancer research : an official journal of the American Association for Cancer Research. 2008;14:2570–8. doi: 10.1158/1078-0432.CCR-07-2033. [DOI] [PubMed] [Google Scholar]
  • 14.Berger AH, Knudson AG, Pandolfi PP. A continuum model for tumour suppression. Nature. 2011;476:163–9. doi: 10.1038/nature10275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Toyooka S, Mitsudomi T, Soh J, Aokage K, Yamane M, Oto T, et al. Molecular oncology of lung cancer. General thoracic and cardiovascular surgery. 2011;59:527–37. doi: 10.1007/s11748-010-0743-3. [DOI] [PubMed] [Google Scholar]
  • 16.Fischer JR, Ohnmacht U, Rieger N, Zemaitis M, Stoffregen C, Manegold C, et al. Prognostic significance of RASSF1A promoter methylation on survival of non-small cell lung cancer patients treated with gemcitabine. Lung cancer. 2007;56:115–23. doi: 10.1016/j.lungcan.2006.11.016. [DOI] [PubMed] [Google Scholar]
  • 17.Topaloglu O, Hoque MO, Tokumaru Y, Lee J, Ratovitski E, Sidransky D, et al. Detection of promoter hypermethylation of multiple genes in the tumor and bronchoalveolar lavage of patients with lung cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2004;10:2284–8. doi: 10.1158/1078-0432.ccr-1111-3. [DOI] [PubMed] [Google Scholar]
  • 18.Belinsky SA, Nikula KJ, Palmisano WA, Michels R, Saccomanno G, Gabrielson E, et al. Aberrant methylation of p16(INK4a) is an early event in lung cancer and a potential biomarker for early diagnosis. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:11891–6. doi: 10.1073/pnas.95.20.11891. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Nuovo GJ, Plaia TW, Belinsky SA, Baylin SB, Herman JG. In situ detection of the hypermethylation-induced inactivation of the p16 gene as an early event in oncogenesis. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:12754–9. doi: 10.1073/pnas.96.22.12754. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Palmisano WA, Divine KK, Saccomanno G, Gilliland FD, Baylin SB, Herman JG, et al. Predicting lung cancer by detecting aberrant promoter methylation in sputum. Cancer research. 2000;60:5954–8. [PubMed] [Google Scholar]
  • 21.Kim DH, Nelson HH, Wiencke JK, Zheng S, Christiani DC, Wain JC, et al. p16(INK4a) and histology-specific methylation of CpG islands by exposure to tobacco smoke in non-small cell lung cancer. Cancer research. 2001;61:3419–24. [PubMed] [Google Scholar]
  • 22.Buckingham L, Penfield Faber L, Kim A, Liptay M, Barger C, Basu S, et al. PTEN, RASSF1 and DAPK site-specific hypermethylation and outcome in surgically treated stage I and II nonsmall cell lung cancer patients. International journal of cancer Journal international du cancer. 2010;126:1630–9. doi: 10.1002/ijc.24896. [DOI] [PubMed] [Google Scholar]
  • 23.Nakata S, Sugio K, Uramoto H, Oyama T, Hanagiri T, Morita M, et al. The methylation status and protein expression of CDH1, p16(INK4A), and fragile histidine triad in nonsmall cell lung carcinoma: epigenetic silencing, clinical features, and prognostic significance. Cancer. 2006;106:2190–9. doi: 10.1002/cncr.21870. [DOI] [PubMed] [Google Scholar]
  • 24.Sterlacci W, Tzankov A, Veits L, Zelger B, Bihl MP, Foerster A, et al. A comprehensive analysis of p16 expression, gene status, and promoter hypermethylation in surgically resected non-small cell lung carcinomas. Journal of thoracic oncology : official publication of the International Association for the Study of Lung Cancer. 2011;6:1649–57. doi: 10.1097/JTO.0b013e3182295745. [DOI] [PubMed] [Google Scholar]
  • 25.Tsou JA, Galler JS, Siegmund KD, Laird PW, Turla S, Cozen W, et al. Identification of a panel of sensitive and specific DNA methylation markers for lung adenocarcinoma. Molecular cancer. 2007;6:70. doi: 10.1186/1476-4598-6-70. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Gasco M, Bell AK, Heath V, Sullivan A, Smith P, Hiller L, et al. Epigenetic inactivation of 14-3-3 sigma in oral carcinoma: association with p16(INK4a) silencing and human papillomavirus negativity. Cancer research. 2002;62:2072–6. [PubMed] [Google Scholar]
  • 27.Kim DS, Cha SI, Lee JH, Lee YM, Choi JE, Kim MJ, et al. Aberrant DNA methylation profiles of non-small cell lung cancers in a Korean population. Lung cancer. 2007;58:1–6. doi: 10.1016/j.lungcan.2007.04.008. [DOI] [PubMed] [Google Scholar]
  • 28.Shivapurkar N, Stastny V, Suzuki M, Wistuba, Li L, Zheng Y, et al. Application of a methylation gene panel by quantitative PCR for lung cancers. Cancer letters. 2007;247:56–71. doi: 10.1016/j.canlet.2006.03.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Tang X, Khuri FR, Lee JJ, Kemp BL, Liu D, Hong WK, et al. Hypermethylation of the death-associated protein (DAP) kinase promoter and aggressiveness in stage I non-small-cell lung cancer. Journal of the National Cancer Institute. 2000;92:1511–6. doi: 10.1093/jnci/92.18.1511. [DOI] [PubMed] [Google Scholar]
  • 30.Dammann R, Strunnikova M, Schagdarsurengin U, Rastetter M, Papritz M, Hattenhorst UE, et al. CpG island methylation and expression of tumour-associated genes in lung carcinoma. European journal of cancer. 2005;41:1223–36. doi: 10.1016/j.ejca.2005.02.020. [DOI] [PubMed] [Google Scholar]
  • 31.Nelson HH, Marsit CJ, Christensen BC, Houseman EA, Kontic M, Wiemels JL, et al. Key epigenetic changes associated with lung cancer development: results from dense methylation array profiling. Epigenetics : official journal of the DNA Methylation Society. 2012;7:559–66. doi: 10.4161/epi.20219. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Chen C, Yin N, Yin B, Lu Q. DNA methylation in thoracic neoplasms. Cancer letters. 2011;301:7–16. doi: 10.1016/j.canlet.2010.10.017. [DOI] [PubMed] [Google Scholar]
  • 33.Heller G, Zielinski CC, Zochbauer-Muller S. Lung cancer: from single-gene methylation to methylome profiling. Cancer metastasis reviews. 2010;29:95–107. doi: 10.1007/s10555-010-9203-x. [DOI] [PubMed] [Google Scholar]
  • 34.Li W, Deng J, Jiang P, Tang J. Association of 5′-CpG island hypermethylation of the FHIT gene with lung cancer in southern-central Chinese population. Cancer biology & therapy. 2010;10:997–1000. doi: 10.4161/cbt.10.10.13231. [DOI] [PubMed] [Google Scholar]
  • 35.Verri C, Roz L, Conte D, Liloglou T, Livio A, Vesin A, et al. Fragile histidine triad gene inactivation in lung cancer: the European Early Lung Cancer project. American journal of respiratory and critical care medicine. 2009;179:396–401. doi: 10.1164/rccm.200807-1153OC. [DOI] [PubMed] [Google Scholar]
  • 36.Yanagawa N, Tamura G, Oizumi H, Kanauchi N, Endoh M, Sadahiro M, et al. Promoter hypermethylation of RASSF1A and RUNX3 genes as an independent prognostic prediction marker in surgically resected non-small cell lung cancers. Lung cancer. 2007;58:131–8. doi: 10.1016/j.lungcan.2007.05.011. [DOI] [PubMed] [Google Scholar]
  • 37.Tomizawa Y, Iijima H, Nomoto T, Iwasaki Y, Otani Y, Tsuchiya S, et al. Clinicopathological significance of aberrant methylation of RARbeta2 at 3p24, RASSF1A at 3p21.3, and FHIT at 3p14.2 in patients with non-small cell lung cancer. Lung cancer. 2004;46:305–12. doi: 10.1016/j.lungcan.2004.05.003. [DOI] [PubMed] [Google Scholar]
  • 38.Begum S, Brait M, Dasgupta S, Ostrow KL, Zahurak M, Carvalho AL, et al. An epigenetic marker panel for detection of lung cancer using cell-free serum DNA. Clinical cancer research : an official journal of the American Association for Cancer Research. 2011;17:4494–503. doi: 10.1158/1078-0432.CCR-10-3436. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Grote HJ, Schmiemann V, Kiel S, Bocking A, Kappes R, Gabbert HE, et al. Aberrant methylation of the adenomatous polyposis coli promoter 1A in bronchial aspirates from patients with suspected lung cancer. International journal of cancer Journal international du cancer. 2004;110:751–5. doi: 10.1002/ijc.20196. [DOI] [PubMed] [Google Scholar]
  • 40.Toyooka S, Maruyama R, Toyooka KO, McLerran D, Feng Z, Fukuyama Y, et al. Smoke exposure, histologic type and geography-related differences in the methylation profiles of non-small cell lung cancer. International journal of cancer Journal international du cancer. 2003;103:153–60. doi: 10.1002/ijc.10787. [DOI] [PubMed] [Google Scholar]
  • 41.Brabender J, Usadel H, Danenberg KD, Metzger R, Schneider PM, Lord RV, et al. Adenomatous polyposis coli gene promoter hypermethylation in non-small cell lung cancer is associated with survival. Oncogene. 2001;20:3528–32. doi: 10.1038/sj.onc.1204455. [DOI] [PubMed] [Google Scholar]
  • 42.Ji M, Guan H, Gao C, Shi B, Hou P. Highly frequent promoter methylation and PIK3CA amplification in non-small cell lung cancer (NSCLC) BMC cancer. 2011;11:147. doi: 10.1186/1471-2407-11-147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Lee SM, Lee WK, Kim DS, Park JY. Quantitative promoter hypermethylation analysis of RASSF1A in lung cancer: comparison with methylation-specific PCR technique and clinical significance. Molecular medicine reports. 2012;5:239–44. doi: 10.3892/mmr.2011.608. [DOI] [PubMed] [Google Scholar]
  • 44.Niklinska W, Naumnik W, Sulewska A, Kozlowski M, Pankiewicz W, Milewski R. Prognostic significance of DAPK and RASSF1A promoter hypermethylation in non-small cell lung cancer (NSCLC) Folia histochemica et cytobiologica / Polish Academy of Sciences, Polish Histochemical and Cytochemical Society. 2009;47:275–80. doi: 10.2478/v10042-009-0091-2. [DOI] [PubMed] [Google Scholar]
  • 45.Dammann R, Takahashi T, Pfeifer GP. The CpG island of the novel tumor suppressor gene RASSF1A is intensely methylated in primary small cell lung carcinomas. Oncogene. 2001;20:3563–7. doi: 10.1038/sj.onc.1204469. [DOI] [PubMed] [Google Scholar]
  • 46.Kim DH, Nelson HH, Wiencke JK, Christiani DC, Wain JC, Mark EJ, et al. Promoter methylation of DAP-kinase: association with advanced stage in non-small cell lung cancer. Oncogene. 2001;20:1765–70. doi: 10.1038/sj.onc.1204302. [DOI] [PubMed] [Google Scholar]
  • 47.Virmani AK, Rathi A, Zochbauer-Muller S, Sacchi N, Fukuyama Y, Bryant D, et al. Promoter methylation and silencing of the retinoic acid receptor-beta gene in lung carcinomas. Journal of the National Cancer Institute. 2000;92:1303–7. doi: 10.1093/jnci/92.16.1303. [DOI] [PubMed] [Google Scholar]
  • 48.Ekim M, Caner V, Buyukpinarbasili N, Tepeli E, Elmas L, Bagci G. Determination of O(6)-methylguanine DNA methyltransferase promoter methylation in non-small cell lung cancer. Genetic testing and molecular biomarkers. 2011;15:357–60. doi: 10.1089/gtmb.2010.0211. [DOI] [PubMed] [Google Scholar]
  • 49.Lai JC, Cheng YW, Goan YG, Chang JT, Wu TC, Chen CY, et al. Promoter methylation of O(6)-methylguanine-DNA-methyltransferase in lung cancer is regulated by p53. DNA repair. 2008;7:1352–63. doi: 10.1016/j.dnarep.2008.04.016. [DOI] [PubMed] [Google Scholar]
  • 50.Brabender J, Usadel H, Metzger R, Schneider PM, Park J, Salonga D, et al. Quantitative O(6)-methylguanine DNA methyltransferase methylation analysis in curatively resected non-small cell lung cancer: associations with clinical outcome. Clinical cancer research : an official journal of the American Association for Cancer Research. 2003;9:223–7. [PubMed] [Google Scholar]
  • 51.Schmidt B, Liebenberg V, Dietrich D, Schlegel T, Kneip C, Seegebarth A, et al. SHOX2 DNA methylation is a biomarker for the diagnosis of lung cancer based on bronchial aspirates. BMC cancer. 2010;10:600. doi: 10.1186/1471-2407-10-600. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Schneider KU, Dietrich D, Fleischhacker M, Leschber G, Merk J, Schaper F, et al. Correlation of SHOX2 gene amplification and DNA methylation in lung cancer tumors. BMC cancer. 2011;11:102. doi: 10.1186/1471-2407-11-102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Yu GP, Ji Y, Chen GQ, Huang B, Shen K, Wu S, et al. Application of RUNX3 gene promoter methylation in the diagnosis of non-small cell lung cancer. Oncology letters. 2012;3:159–62. doi: 10.3892/ol.2011.450. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Omar MF, Ito K, Nga ME, Soo R, Peh BK, Ismail TM, et al. RUNX3 downregulation in human lung adenocarcinoma is independent of p53, EGFR or KRAS status. Pathology oncology research : POR. 2012;18:783–92. doi: 10.1007/s12253-011-9485-5. [DOI] [PubMed] [Google Scholar]
  • 55.Sato K, Tomizawa Y, Iijima H, Saito R, Ishizuka T, Nakajima T, et al. Epigenetic inactivation of the RUNX3 gene in lung cancer. Oncology reports. 2006;15:129–35. [PubMed] [Google Scholar]
  • 56.Kim TY, Lee HJ, Hwang KS, Lee M, Kim JW, Bang YJ, et al. Methylation of RUNX3 in various types of human cancers and premalignant stages of gastric carcinoma. Laboratory investigation; a journal of technical methods and pathology. 2004;84:479–84. doi: 10.1038/labinvest.3700060. [DOI] [PubMed] [Google Scholar]
  • 57.Li QL, Kim HR, Kim WJ, Choi JK, Lee YH, Kim HM, et al. Transcriptional silencing of the RUNX3 gene by CpG hypermethylation is associated with lung cancer. Biochemical and biophysical research communications. 2004;314:223–8. doi: 10.1016/j.bbrc.2003.12.079. [DOI] [PubMed] [Google Scholar]
  • 58.Kim DS, Kim MJ, Lee JY, Kim YZ, Kim EJ, Park JY. Aberrant methylation of E-cadherin and H-cadherin genes in nonsmall cell lung cancer and its relation to clinicopathologic features. Cancer. 2007;110:2785–92. doi: 10.1002/cncr.23113. [DOI] [PubMed] [Google Scholar]
  • 59.Heller G, Fong KM, Girard L, Seidl S, End-Pfutzenreuter A, Lang G, et al. Expression and methylation pattern of TSLC1 cascade genes in lung carcinomas. Oncogene. 2006;25:959–68. doi: 10.1038/sj.onc.1209115. [DOI] [PubMed] [Google Scholar]
  • 60.Kikuchi S, Yamada D, Fukami T, Maruyama T, Ito A, Asamura H, et al. Hypermethylation of the TSLC1/IGSF4 promoter is associated with tobacco smoking and a poor prognosis in primary nonsmall cell lung carcinoma. Cancer. 2006;106:1751–8. doi: 10.1002/cncr.21800. [DOI] [PubMed] [Google Scholar]
  • 61.Fukami T, Fukuhara H, Kuramochi M, Maruyama T, Isogai K, Sakamoto M, et al. Promoter methylation of the TSLC1 gene in advanced lung tumors and various cancer cell lines. International journal of cancer Journal international du cancer. 2003;107:53–9. doi: 10.1002/ijc.11348. [DOI] [PubMed] [Google Scholar]
  • 62.Ramirez JL, Sarries C, de Castro PL, Roig B, Queralt C, Escuin D, et al. Methylation patterns and K-ras mutations in tumor and paired serum of resected non-small-cell lung cancer patients. Cancer letters. 2003;193:207–16. doi: 10.1016/s0304-3835(02)00740-1. [DOI] [PubMed] [Google Scholar]
  • 63.Virmani A, Rathi A, Sugio K, Sathyanarayana UG, Toyooka S, Kischel FC, et al. Aberrant methylation of TMS1 in small cell, non small cell lung cancer and breast cancer. International journal of cancer Journal international du cancer. 2003;106:198–204. doi: 10.1002/ijc.11206. [DOI] [PubMed] [Google Scholar]
  • 64.Kikuchi S, Yamada D, Fukami T, Masuda M, Sakurai-Yageta M, Williams YN, et al. Promoter methylation of DAL-1/4.1B predicts poor prognosis in non-small cell lung cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2005;11:2954–61. doi: 10.1158/1078-0432.CCR-04-2206. [DOI] [PubMed] [Google Scholar]
  • 65.Marsit CJ, Zheng S, Aldape K, Hinds PW, Nelson HH, Wiencke JK, et al. PTEN expression in non-small-cell lung cancer: evaluating its relation to tumor characteristics, allelic loss, and epigenetic alteration. Human pathology. 2005;36:768–76. doi: 10.1016/j.humpath.2005.05.006. [DOI] [PubMed] [Google Scholar]
  • 66.Soria JC, Lee HY, Lee JI, Wang L, Issa JP, Kemp BL, et al. Lack of PTEN expression in non-small cell lung cancer could be related to promoter methylation. Clinical cancer research : an official journal of the American Association for Cancer Research. 2002;8:1178–84. [PubMed] [Google Scholar]
  • 67.Grimminger PP, Maus MK, Schneider PM, Metzger R, Holscher AH, Sugita H, et al. Glutathione S-transferase PI (GST-PI) mRNA expression and DNA methylation is involved in the pathogenesis and prognosis of NSCLC. Lung cancer. 2012;78:87–91. doi: 10.1016/j.lungcan.2012.07.008. [DOI] [PubMed] [Google Scholar]
  • 68.Vallbohmer D, Brabender J, Yang D, Schneider PM, Metzger R, Danenberg KD, et al. DNA methyltransferases messenger RNA expression and aberrant methylation of CpG islands in non-small-cell lung cancer: association and prognostic value. Clinical lung cancer. 2006;8:39–44. doi: 10.3816/CLC.2006.n.031. [DOI] [PubMed] [Google Scholar]
  • 69.Belinsky SA. Silencing of genes by promoter hypermethylation: key event in rodent and human lung cancer. Carcinogenesis. 2005;26:1481–7. doi: 10.1093/carcin/bgi020. [DOI] [PubMed] [Google Scholar]
  • 70.Lin Q, Geng J, Ma K, Yu J, Sun J, Shen Z, et al. RASSF1A, APC, ESR1, ABCB1 and HOXC9, but not p16INK4A, DAPK1, PTEN and MT1G genes were frequently methylated in the stage I non-small cell lung cancer in China. Journal of cancer research and clinical oncology. 2009;135:1675–84. doi: 10.1007/s00432-009-0614-4. [DOI] [PubMed] [Google Scholar]
  • 71.Rauch T, Wang Z, Zhang X, Zhong X, Wu X, Lau SK, et al. Homeobox gene methylation in lung cancer studied by genome-wide analysis with a microarray-based methylated CpG island recovery assay. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:5527–32. doi: 10.1073/pnas.0701059104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Yano M, Toyooka S, Tsukuda K, Dote H, Ouchida M, Hanabata T, et al. Aberrant promoter methylation of human DAB2 interactive protein (hDAB2IP) gene in lung cancers. International journal of cancer Journal international du cancer. 2005;113:59–66. doi: 10.1002/ijc.20531. [DOI] [PubMed] [Google Scholar]
  • 73.Chen H, Suzuki M, Nakamura Y, Ohira M, Ando S, Iida T, et al. Aberrant methylation of FBN2 in human non-small cell lung cancer. Lung cancer. 2005;50:43–9. doi: 10.1016/j.lungcan.2005.04.013. [DOI] [PubMed] [Google Scholar]
  • 74.Liu Z, Li W, Lei Z, Zhao J, Chen XF, Liu R, et al. CpG island methylator phenotype involving chromosome 3p confers an increased risk of non-small cell lung cancer. Journal of thoracic oncology : official publication of the International Association for the Study of Lung Cancer. 2010;5:790–7. doi: 10.1097/jto.0b013e3181d862f5. [DOI] [PubMed] [Google Scholar]
  • 75.Liu Z, Zhao J, Chen XF, Li W, Liu R, Lei Z, et al. CpG island methylator phenotype involving tumor suppressor genes located on chromosome 3p in non-small cell lung cancer. Lung cancer. 2008;62:15–22. doi: 10.1016/j.lungcan.2008.02.005. [DOI] [PubMed] [Google Scholar]
  • 76.Marsit CJ, Houseman EA, Christensen BC, Eddy K, Bueno R, Sugarbaker DJ, et al. Examination of a CpG island methylator phenotype and implications of methylation profiles in solid tumors. Cancer research. 2006;66:10621–9. doi: 10.1158/0008-5472.CAN-06-1687. [DOI] [PubMed] [Google Scholar]
  • 77.Shinjo K, Okamoto Y, An B, Yokoyama T, Takeuchi I, Fujii M, et al. Integrated analysis of genetic and epigenetic alterations reveals CpG island methylator phenotype associated with distinct clinical characters of lung adenocarcinoma. Carcinogenesis. 2012;33:1277–85. doi: 10.1093/carcin/bgs154. [DOI] [PubMed] [Google Scholar]
  • 78.Kim H, Kwon YM, Kim JS, Han J, Shim YM, Park J, et al. Elevated mRNA levels of DNA methyltransferase-1 as an independent prognostic factor in primary nonsmall cell lung cancer. Cancer. 2006;107:1042–9. doi: 10.1002/cncr.22087. [DOI] [PubMed] [Google Scholar]
  • 79.Lin RK, Hsu HS, Chang JW, Chen CY, Chen JT, Wang YC. Alteration of DNA methyltransferases contributes to 5′CpG methylation and poor prognosis in lung cancer. Lung cancer. 2007;55:205–13. doi: 10.1016/j.lungcan.2006.10.022. [DOI] [PubMed] [Google Scholar]
  • 80.Anisowicz A, Huang H, Braunschweiger KI, Liu Z, Giese H, Wang H, et al. A high-throughput and sensitive method to measure global DNA methylation: application in lung cancer. BMC cancer. 2008;8:222. doi: 10.1186/1471-2407-8-222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Daskalos A, Logotheti S, Markopoulou S, Xinarianos G, Gosney JR, Kastania AN, et al. Global DNA hypomethylation-induced DeltaNp73 transcriptional activation in non-small cell lung cancer. Cancer letters. 2011;300:79–86. doi: 10.1016/j.canlet.2010.09.009. [DOI] [PubMed] [Google Scholar]
  • 82.Feinberg AP, Vogelstein B. Hypomethylation of ras oncogenes in primary human cancers. Biochemical and biophysical research communications. 1983;111:47–54. doi: 10.1016/s0006-291x(83)80115-6. [DOI] [PubMed] [Google Scholar]
  • 83.Feinberg AP, Vogelstein B. Hypomethylation distinguishes genes of some human cancers from their normal counterparts. Nature. 1983;301:89–92. doi: 10.1038/301089a0. [DOI] [PubMed] [Google Scholar]
  • 84.Kondo M, Suzuki H, Ueda R, Osada H, Takagi K, Takahashi T, et al. Frequent loss of imprinting of the H19 gene is often associated with its overexpression in human lung cancers. Oncogene. 1995;10:1193–8. [PubMed] [Google Scholar]
  • 85.Rauch TA, Zhong X, Wu X, Wang M, Kernstine KH, Wang Z, et al. High-resolution mapping of DNA hypermethylation and hypomethylation in lung cancer. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:252–7. doi: 10.1073/pnas.0710735105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Glazer CA, Smith IM, Ochs MF, Begum S, Westra W, Chang SS, et al. Integrative discovery of epigenetically derepressed cancer testis antigens in NSCLC. PloS one. 2009;4:e8189. doi: 10.1371/journal.pone.0008189. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Kim SH, Lee S, Lee CH, Lee MK, Kim YD, Shin DH, et al. Expression of cancer-testis antigens MAGE-A3/6 and NY-ESO-1 in non-small-cell lung carcinomas and their relationship with immune cell infiltration. Lung. 2009;187:401–11. doi: 10.1007/s00408-009-9181-3. [DOI] [PubMed] [Google Scholar]
  • 88.Kayser G, Sienel W, Kubitz B, Mattern D, Stickeler E, Passlick B, et al. Poor outcome in primary non-small cell lung cancers is predicted by transketolase TKTL1 expression. Pathology. 2011;43:719–24. doi: 10.1097/PAT.0b013e32834c352b. [DOI] [PubMed] [Google Scholar]
  • 89.Renaud S, Pugacheva EM, Delgado MD, Braunschweig R, Abdullaev Z, Loukinov D, et al. Expression of the CTCF-paralogous cancer-testis gene, brother of the regulator of imprinted sites (BORIS), is regulated by three alternative promoters modulated by CpG methylation and by CTCF and p53 transcription factors. Nucleic acids research. 2007;35:7372–88. doi: 10.1093/nar/gkm896. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Hong JA, Kang Y, Abdullaev Z, Flanagan PT, Pack SD, Fischette MR, et al. Reciprocal binding of CTCF and BORIS to the NY-ESO-1 promoter coincides with derepression of this cancer-testis gene in lung cancer cells. Cancer research. 2005;65:7763–74. doi: 10.1158/0008-5472.CAN-05-0823. [DOI] [PubMed] [Google Scholar]
  • 91.Radhakrishnan VM, Jensen TJ, Cui H, Futscher BW, Martinez JD. Hypomethylation of the 14-3-3sigma promoter leads to increased expression in non-small cell lung cancer. Genes, chromosomes & cancer. 2011;50:830–6. doi: 10.1002/gcc.20904. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Shiba-Ishii A, Noguchi M. Aberrant stratifin overexpression is regulated by tumor-associated CpG demethylation in lung adenocarcinoma. The American journal of pathology. 2012;180:1653–62. doi: 10.1016/j.ajpath.2011.12.014. [DOI] [PubMed] [Google Scholar]
  • 93.Gu Y, Wang C, Wang Y, Qiu X, Wang E. Expression of thymosin beta10 and its role in non-small cell lung cancer. Human pathology. 2009;40:117–24. doi: 10.1016/j.humpath.2008.06.023. [DOI] [PubMed] [Google Scholar]
  • 94.Jang SJ, Soria JC, Wang L, Hassan KA, Morice RC, Walsh GL, et al. Activation of melanoma antigen tumor antigens occurs early in lung carcinogenesis. Cancer research. 2001;61:7959–63. [PubMed] [Google Scholar]
  • 95.Gronbaek K, Hother C, Jones PA. Epigenetic changes in cancer. Apmis. 2007;115:1039–59. doi: 10.1111/j.1600-0463.2007.apm_636.xml.x. [DOI] [PubMed] [Google Scholar]
  • 96.Rideout WM, 3rd, Coetzee GA, Olumi AF, Jones PA. 5-Methylcytosine as an endogenous mutagen in the human LDL receptor and p53 genes. Science (New York, NY. 1990;249:1288–90. doi: 10.1126/science.1697983. [DOI] [PubMed] [Google Scholar]
  • 97.Chen JX, Zheng Y, West M, Tang MS. Carcinogens preferentially bind at methylated CpG in the p53 mutational hot spots. Cancer research. 1998;58:2070–5. [PubMed] [Google Scholar]
  • 98.Denissenko MF, Pao A, Tang M, Pfeifer GP. Preferential formation of benzo [a]pyrene adducts at lung cancer mutational hotspots in P53. Science (New York, NY. 1996;274:430–2. doi: 10.1126/science.274.5286.430. [DOI] [PubMed] [Google Scholar]
  • 99.Yoon JH, Smith LE, Feng Z, Tang M, Lee CS, Pfeifer GP. Methylated CpG dinucleotides are the preferential targets for G-to-T transversion mutations induced by benzo [a]pyrene diol epoxide in mammalian cells: similarities with the p53 mutation spectrum in smoking-associated lung cancers. Cancer research. 2001;61:7110–7. [PubMed] [Google Scholar]
  • 100.Feng Z, Hu W, Hu Y, Tang MS. Acrolein is a major cigarette-related lung cancer agent: Preferential binding at p53 mutational hotspots and inhibition of DNA repair. Proceedings of the National Academy of Sciences of the United States of America. 2006;103:15404–9. doi: 10.1073/pnas.0607031103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 101.Sasaki H, Moriyama S, Nakashima Y, Kobayashi Y, Kiriyama M, Fukai I, et al. Histone deacetylase 1 mRNA expression in lung cancer. Lung cancer. 2004;46:171–8. doi: 10.1016/j.lungcan.2004.03.021. [DOI] [PubMed] [Google Scholar]
  • 102.Nakagawa M, Oda Y, Eguchi T, Aishima S, Yao T, Hosoi F, et al. Expression profile of class I histone deacetylases in human cancer tissues. Oncology reports. 2007;18:769–74. [PubMed] [Google Scholar]
  • 103.Bartling B, Hofmann HS, Boettger T, Hansen G, Burdach S, Silber RE, et al. Comparative application of antibody and gene array for expression profiling in human squamous cell lung carcinoma. Lung cancer. 2005;49:145–54. doi: 10.1016/j.lungcan.2005.02.006. [DOI] [PubMed] [Google Scholar]
  • 104.Esteller M. Epigenetics in cancer. The New England journal of medicine. 2008;358:1148–59. doi: 10.1056/NEJMra072067. [DOI] [PubMed] [Google Scholar]
  • 105.Suzuki H, Ouchida M, Yamamoto H, Yano M, Toyooka S, Aoe M, et al. Decreased expression of the SIN3A gene, a candidate tumor suppressor located at the prevalent allelic loss region 15q23 in non-small cell lung cancer. Lung cancer. 2008;59:24–31. doi: 10.1016/j.lungcan.2007.08.002. [DOI] [PubMed] [Google Scholar]
  • 106.Van Den Broeck A, Brambilla E, Moro-Sibilot D, Lantuejoul S, Brambilla C, Eymin B, et al. Loss of histone H4K20 trimethylation occurs in preneoplasia and influences prognosis of non-small cell lung cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2008;14:7237–45. doi: 10.1158/1078-0432.CCR-08-0869. [DOI] [PubMed] [Google Scholar]
  • 107.Hu J, Liu YL, Piao SL, Yang DD, Yang YM, Cai L. Expression patterns of USP22 and potential targets BMI-1, PTEN, p-AKT in non-small-cell lung cancer. Lung cancer. 2012;77:593–9. doi: 10.1016/j.lungcan.2012.05.112. [DOI] [PubMed] [Google Scholar]
  • 108.Huang J, Qiu Y, Chen G, Huang L, He J. The relationship between Bmi-1 and the epithelial-mesenchymal transition in lung squamous cell carcinoma. Medical oncology. 2012;29:1606–13. doi: 10.1007/s12032-011-9998-5. [DOI] [PubMed] [Google Scholar]
  • 109.Kimura M, Takenobu H, Akita N, Nakazawa A, Ochiai H, Shimozato O, et al. Bmi1 regulates cell fate via tumor suppressor WWOX repression in small-cell lung cancer cells. Cancer science. 2011;102:983–90. doi: 10.1111/j.1349-7006.2011.01891.x. [DOI] [PubMed] [Google Scholar]
  • 110.Vonlanthen S, Heighway J, Altermatt HJ, Gugger M, Kappeler A, Borner MM, et al. The bmi-1 oncoprotein is differentially expressed in non-small cell lung cancer and correlates with INK4A-ARF locus expression. British journal of cancer. 2001;84:1372–6. doi: 10.1054/bjoc.2001.1791. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Breuer RH, Snijders PJ, Smit EF, Sutedja TG, Sewalt RG, Otte AP, et al. Increased expression of the EZH2 polycomb group gene in BMI-1-positive neoplastic cells during bronchial carcinogenesis. Neoplasia. 2004;6:736–43. doi: 10.1593/neo.04160. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Cao W, de Ribeiro RO, Liu D, Saintigny P, Xia R, Xue Y, et al. EZH2 promotes malignant behaviors via cell cycle dysregulation and its mRNA level associates with prognosis of patient with non-small cell lung cancer. PloS one. 2012;7:e52984. doi: 10.1371/journal.pone.0052984. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Takawa M, Masuda K, Kunizaki M, Daigo Y, Takagi K, Iwai Y, et al. Validation of the histone methyltransferase EZH2 as a therapeutic target for various types of human cancer and as a prognostic marker. Cancer science. 2011;102:1298–305. doi: 10.1111/j.1349-7006.2011.01958.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 114.Wan L, Li X, Shen H, Bai X. Quantitative analysis of EZH2 expression and its correlations with lung cancer patients’ clinical pathological characteristics. Clinical & translational oncology : official publication of the Federation of Spanish Oncology Societies and of the National Cancer Institute of Mexico. 2013;15:132–8. doi: 10.1007/s12094-012-0897-9. [DOI] [PubMed] [Google Scholar]
  • 115.Zhou Y, Wan C, Liu Y, Lv L, Chen B, Ni R, et al. Polycomb Group Oncogene RING1 is Over-expressed in Non-Small Cell Lung Cancer. Pathology oncology research : POR. 2014 doi: 10.1007/s12253-013-9727-9. [DOI] [PubMed] [Google Scholar]
  • 116.Sato T, Kaneda A, Tsuji S, Isagawa T, Yamamoto S, Fujita T, et al. PRC2 overexpression and PRC2-target gene repression relating to poorer prognosis in small cell lung cancer. Scientific reports. 2013;3:1911. doi: 10.1038/srep01911. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Lv Y, Yuan C, Xiao X, Wang X, Ji X, Yu H, et al. The expression and significance of the enhancer of zeste homolog 2 in lung adenocarcinoma. Oncology reports. 2012;28:147–54. doi: 10.3892/or.2012.1787. [DOI] [PubMed] [Google Scholar]
  • 118.Huqun, Ishikawa R, Zhang J, Miyazawa H, Goto Y, Shimizu Y, et al. Enhancer of zeste homolog 2 is a novel prognostic biomarker in nonsmall cell lung cancer. Cancer. 2012;118:1599–606. doi: 10.1002/cncr.26441. [DOI] [PubMed] [Google Scholar]
  • 119.Crea F, Paolicchi E, Marquez VE, Danesi R. Polycomb genes and cancer: time for clinical application? Critical reviews in oncology/hematology. 2012;83:184–93. doi: 10.1016/j.critrevonc.2011.10.007. [DOI] [PubMed] [Google Scholar]
  • 120.Koch LK, Zhou H, Ellinger J, Biermann K, Holler T, von Rucker A, et al. Stem cell marker expression in small cell lung carcinoma and developing lung tissue. Human pathology. 2008;39:1597–605. doi: 10.1016/j.humpath.2008.03.008. [DOI] [PubMed] [Google Scholar]
  • 121.Vrzalikova K, Skarda J, Ehrmann J, Murray PG, Fridman E, Kopolovic J, et al. Prognostic value of Bmi-1 oncoprotein expression in NSCLC patients: a tissue microarray study. Journal of cancer research and clinical oncology. 2008;134:1037–42. doi: 10.1007/s00432-008-0361-y. [DOI] [PubMed] [Google Scholar]
  • 122.Dovey JS, Zacharek SJ, Kim CF, Lees JA. Bmi1 is critical for lung tumorigenesis and bronchioalveolar stem cell expansion. Proceedings of the National Academy of Sciences of the United States of America. 2008;105:11857–62. doi: 10.1073/pnas.0803574105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Liu L, Andrews LG, Tollefsbol TO. Loss of the human polycomb group protein BMI1 promotes cancer-specific cell death. Oncogene. 2006;25:4370–5. doi: 10.1038/sj.onc.1209454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Guan P, Yin Z, Li X, Wu W, Zhou B. Meta-analysis of human lung cancer microRNA expression profiling studies comparing cancer tissues with normal tissues. Journal of experimental & clinical cancer research : CR. 2012;31:54. doi: 10.1186/1756-9966-31-54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Vosa U, Vooder T, Kolde R, Vilo J, Metspalu A, Annilo T. Meta-analysis of microRNA expression in lung cancer. International journal of cancer Journal international du cancer. 2013;132:2884–93. doi: 10.1002/ijc.27981. [DOI] [PubMed] [Google Scholar]
  • 126.Shi X, Sun M, Liu H, Yao Y, Song Y. Long non-coding RNAs: a new frontier in the study of human diseases. Cancer letters. 2013;339:159–66. doi: 10.1016/j.canlet.2013.06.013. [DOI] [PubMed] [Google Scholar]
  • 127.Gutschner T, Hammerle M, Diederichs S. MALAT1 -- a paradigm for long noncoding RNA function in cancer. Journal of molecular medicine. 2013;91:791–801. doi: 10.1007/s00109-013-1028-y. [DOI] [PubMed] [Google Scholar]
  • 128.Ji P, Diederichs S, Wang W, Boing S, Metzger R, Schneider PM, et al. MALAT-1, a novel noncoding RNA, and thymosin beta4 predict metastasis and survival in early-stage non-small cell lung cancer. Oncogene. 2003;22:8031–41. doi: 10.1038/sj.onc.1206928. [DOI] [PubMed] [Google Scholar]
  • 129.Schmidt LH, Spieker T, Koschmieder S, Schaffers S, Humberg J, Jungen D, et al. The long noncoding MALAT-1 RNA indicates a poor prognosis in non-small cell lung cancer and induces migration and tumor growth. Journal of thoracic oncology : official publication of the International Association for the Study of Lung Cancer. 2011;6:1984–92. doi: 10.1097/JTO.0b013e3182307eac. [DOI] [PubMed] [Google Scholar]
  • 130.Gutschner T, Hammerle M, Eissmann M, Hsu J, Kim Y, Hung G, et al. The noncoding RNA MALAT1 is a critical regulator of the metastasis phenotype of lung cancer cells. Cancer research. 2013;73:1180–9. doi: 10.1158/0008-5472.CAN-12-2850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Tano K, Mizuno R, Okada T, Rakwal R, Shibato J, Masuo Y, et al. MALAT-1 enhances cell motility of lung adenocarcinoma cells by influencing the expression of motility-related genes. FEBS letters. 2010;584:4575–80. doi: 10.1016/j.febslet.2010.10.008. [DOI] [PubMed] [Google Scholar]
  • 132.Gutschner T, Baas M, Diederichs S. Noncoding RNA gene silencing through genomic integration of RNA destabilizing elements using zinc finger nucleases. Genome research. 2011;21:1944–54. doi: 10.1101/gr.122358.111. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Liu XH, Liu ZL, Sun M, Liu J, Wang ZX, De W. The long non-coding RNA HOTAIR indicates a poor prognosis and promotes metastasis in non-small cell lung cancer. BMC cancer. 2013;13:464. doi: 10.1186/1471-2407-13-464. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Nakagawa T, Endo H, Yokoyama M, Abe J, Tamai K, Tanaka N, et al. Large noncoding RNA HOTAIR enhances aggressive biological behavior and is associated with short disease-free survival in human non-small cell lung cancer. Biochemical and biophysical research communications. 2013;436:319–24. doi: 10.1016/j.bbrc.2013.05.101. [DOI] [PubMed] [Google Scholar]
  • 135.Zhuang Y, Wang X, Nguyen HT, Zhuo Y, Cui X, Fewell C, et al. Induction of long intergenic non-coding RNA HOTAIR in lung cancer cells by type I collagen. Journal of hematology & oncology. 2013;6:35. doi: 10.1186/1756-8722-6-35. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Lujambio A, Ropero S, Ballestar E, Fraga MF, Cerrato C, Setien F, et al. Genetic unmasking of an epigenetically silenced microRNA in human cancer cells. Cancer research. 2007;67:1424–9. doi: 10.1158/0008-5472.CAN-06-4218. [DOI] [PubMed] [Google Scholar]
  • 137.Wang Z, Chen Z, Gao Y, Li N, Li B, Tan F, et al. DNA hypermethylation of microRNA-34b/c has prognostic value for stage non-small cell lung cancer. Cancer biology & therapy. 2011;11:490–6. doi: 10.4161/cbt.11.5.14550. [DOI] [PubMed] [Google Scholar]
  • 138.Nadal E, Chen G, Gallegos M, Lin L, Ferrer-Torres D, Truini A, et al. Epigenetic Inactivation of microRNA-34b/c Predicts Poor Disease-Free Survival in Early-Stage Lung Adenocarcinoma. Clinical cancer research : an official journal of the American Association for Cancer Research. 2013 doi: 10.1158/1078-0432.CCR-13-0736. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 139.Cao J, Song Y, Bi N, Shen J, Liu W, Fan J, et al. DNA methylation-mediated repression of miR-886-3p predicts poor outcome of human small cell lung cancer. Cancer research. 2013;73:3326–35. doi: 10.1158/0008-5472.CAN-12-3055. [DOI] [PubMed] [Google Scholar]
  • 140.Heller G, Weinzierl M, Noll C, Babinsky V, Ziegler B, Altenberger C, et al. Genome-wide miRNA expression profiling identifies miR-9-3 and miR-193a as targets for DNA methylation in non-small cell lung cancers. Clinical cancer research : an official journal of the American Association for Cancer Research. 2012;18:1619–29. doi: 10.1158/1078-0432.CCR-11-2450. [DOI] [PubMed] [Google Scholar]
  • 141.Brueckner B, Stresemann C, Kuner R, Mund C, Musch T, Meister M, et al. The human let-7a-3 locus contains an epigenetically regulated microRNA gene with oncogenic function. Cancer research. 2007;67:1419–23. doi: 10.1158/0008-5472.CAN-06-4074. [DOI] [PubMed] [Google Scholar]
  • 142.Fabbri M, Garzon R, Cimmino A, Liu Z, Zanesi N, Callegari E, et al. MicroRNA-29 family reverts aberrant methylation in lung cancer by targeting DNA methyltransferases 3A and 3B. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:15805–10. doi: 10.1073/pnas.0707628104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 143.Feng Q, Hawes SE, Stern JE, Wiens L, Lu H, Dong ZM, et al. DNA methylation in tumor and matched normal tissues from non-small cell lung cancer patients. Cancer epidemiology, biomarkers & prevention : a publication of the American Association for Cancer Research, cosponsored by the American Society of Preventive Oncology. 2008;17:645–54. doi: 10.1158/1055-9965.EPI-07-2518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Toyooka S, Suzuki M, Tsuda T, Toyooka KO, Maruyama R, Tsukuda K, et al. Dose effect of smoking on aberrant methylation in non-small cell lung cancers. International journal of cancer Journal international du cancer. 2004;110:462–4. doi: 10.1002/ijc.20125. [DOI] [PubMed] [Google Scholar]
  • 145.Kim H, Kwon YM, Kim JS, Lee H, Park JH, Shim YM, et al. Tumor-specific methylation in bronchial lavage for the early detection of non-small-cell lung cancer. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2004;22:2363–70. doi: 10.1200/JCO.2004.10.077. [DOI] [PubMed] [Google Scholar]
  • 146.Liu Y, Lan Q, Siegfried JM, Luketich JD, Keohavong P. Aberrant promoter methylation of p16 and MGMT genes in lung tumors from smoking and never-smoking lung cancer patients. Neoplasia. 2006;8:46–51. doi: 10.1593/neo.05586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Vaissiere T, Hung RJ, Zaridze D, Moukeria A, Cuenin C, Fasolo V, et al. Quantitative analysis of DNA methylation profiles in lung cancer identifies aberrant DNA methylation of specific genes and its association with gender and cancer risk factors. Cancer research. 2009;69:243–52. doi: 10.1158/0008-5472.CAN-08-2489. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 148.Yanagawa N, Tamura G, Oizumi H, Endoh M, Sadahiro M, Motoyama T. Inverse correlation between EGFR mutation and FHIT, RASSF1A and RUNX3 methylation in lung adenocarcinoma: relation with smoking status. Anticancer research. 2011;31:1211–4. [PubMed] [Google Scholar]
  • 149.Hong YS, Roh MS, Kim NY, Lee HJ, Kim HK, Lee KE, et al. Hypermethylation of p16INK4a in Korean non-small cell lung cancer patients. Journal of Korean medical science. 2007;22 (Suppl):S32–7. doi: 10.3346/jkms.2007.22.S.S32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Andujar P, Wang J, Descatha A, Galateau-Salle F, Abd-Alsamad I, Billon-Galland MA, et al. p16INK4A inactivation mechanisms in non-small-cell lung cancer patients occupationally exposed to asbestos. Lung cancer. 2010;67:23–30. doi: 10.1016/j.lungcan.2009.03.018. [DOI] [PubMed] [Google Scholar]
  • 151.Liu F, Killian JK, Yang M, Walker RL, Hong JA, Zhang M, et al. Epigenomic alterations and gene expression profiles in respiratory epithelia exposed to cigarette smoke condensate. Oncogene. 2010;29:3650–64. doi: 10.1038/onc.2010.129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Marwick JA, Kirkham PA, Stevenson CS, Danahay H, Giddings J, Butler K, et al. Cigarette smoke alters chromatin remodeling and induces proinflammatory genes in rat lungs. American journal of respiratory cell and molecular biology. 2004;31:633–42. doi: 10.1165/rcmb.2004-0006OC. [DOI] [PubMed] [Google Scholar]
  • 153.Kaira K, Sunaga N, Tomizawa Y, Yanagitani N, Ishizuka T, Saito R, et al. Epigenetic inactivation of the RAS-effector gene RASSF2 in lung cancers. International journal of oncology. 2007;31:169–73. [PubMed] [Google Scholar]
  • 154.Tessema M, Yu YY, Stidley CA, Machida EO, Schuebel KE, Baylin SB, et al. Concomitant promoter methylation of multiple genes in lung adenocarcinomas from current, former and never smokers. Carcinogenesis. 2009;30:1132–8. doi: 10.1093/carcin/bgp114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 155.Lin RK, Hsieh YS, Lin P, Hsu HS, Chen CY, Tang YA, et al. The tobacco-specific carcinogen NNK induces DNA methyltransferase 1 accumulation and tumor suppressor gene hypermethylation in mice and lung cancer patients. The Journal of clinical investigation. 2010;120:521–32. doi: 10.1172/JCI40706. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Damiani LA, Yingling CM, Leng S, Romo PE, Nakamura J, Belinsky SA. Carcinogen-induced gene promoter hypermethylation is mediated by DNMT1 and causal for transformation of immortalized bronchial epithelial cells. Cancer research. 2008;68:9005–14. doi: 10.1158/0008-5472.CAN-08-1276. [DOI] [PubMed] [Google Scholar]
  • 157.Wang J, Xu Y, Li J, Sun X, Wang LP, Ji WY. The tobacco-specific carcinogen NNK induces DNA methyltransferase 1 accumulation in laryngeal carcinoma. Oral oncology. 2012;48:541–6. doi: 10.1016/j.oraloncology.2012.01.007. [DOI] [PubMed] [Google Scholar]
  • 158.Wang J, Zhao SL, Li Y, Meng M, Qin CY. 4-(Methylnitrosamino)-1-(3-pyridyl)-1-butanone induces retinoic acid receptor beta hypermethylation through DNA methyltransferase 1 accumulation in esophageal squamous epithelial cells. Asian Pacific journal of cancer prevention : APJCP. 2012;13:2207–12. doi: 10.7314/apjcp.2012.13.5.2207. [DOI] [PubMed] [Google Scholar]
  • 159.Hutt JA, Vuillemenot BR, Barr EB, Grimes MJ, Hahn FF, Hobbs CH, et al. Life-span inhalation exposure to mainstream cigarette smoke induces lung cancer in B6C3F1 mice through genetic and epigenetic pathways. Carcinogenesis. 2005;26:1999–2009. doi: 10.1093/carcin/bgi150. [DOI] [PubMed] [Google Scholar]
  • 160.Pulling LC, Vuillemenot BR, Hutt JA, Devereux TR, Belinsky SA. Aberrant promoter hypermethylation of the death-associated protein kinase gene is early and frequent in murine lung tumors induced by cigarette smoke and tobacco carcinogens. Cancer research. 2004;64:3844–8. doi: 10.1158/0008-5472.CAN-03-2119. [DOI] [PubMed] [Google Scholar]
  • 161.Vuillemenot BR, Hutt JA, Belinsky SA. Gene promoter hypermethylation in mouse lung tumors. Molecular cancer research : MCR. 2006;4:267–73. doi: 10.1158/1541-7786.MCR-05-0218. [DOI] [PubMed] [Google Scholar]
  • 162.Balkwill F, Coussens LM. Cancer: an inflammatory link. Nature. 2004;431:405–6. doi: 10.1038/431405a. [DOI] [PubMed] [Google Scholar]
  • 163.Coussens LM, Werb Z. Inflammation and cancer. Nature. 2002;420:860–7. doi: 10.1038/nature01322. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 164.Meng X, Riordan NH. Cancer is a functional repair tissue. Medical hypotheses. 2006;66:486–90. doi: 10.1016/j.mehy.2005.09.041. [DOI] [PubMed] [Google Scholar]
  • 165.O’Hagan HM, Wang W, Sen S, Destefano Shields C, Lee SS, Zhang YW, et al. Oxidative damage targets complexes containing DNA methyltransferases, SIRT1, and polycomb members to promoter CpG Islands. Cancer cell. 2011;20:606–19. doi: 10.1016/j.ccr.2011.09.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Northrop-Clewes CA, Thurnham DI. Monitoring micronutrients in cigarette smokers. Clinica chimica acta; international journal of clinical chemistry. 2007;377:14–38. doi: 10.1016/j.cca.2006.08.028. [DOI] [PubMed] [Google Scholar]
  • 167.Khaled MA, Krumdieck CL. Association of folate molecules as determined by proton NMR: implications on enzyme binding. Biochemical and biophysical research communications. 1985;130:1273–80. doi: 10.1016/0006-291x(85)91752-8. [DOI] [PubMed] [Google Scholar]
  • 168.Abu Khaled M, Watkins CL, Krumdieck CL. Inactivation of B12 and folate coenzymes by butyl nitrite as observed by NMR: implications on one-carbon transfer mechanism. Biochemical and biophysical research communications. 1986;135:201–7. doi: 10.1016/0006-291x(86)90963-0. [DOI] [PubMed] [Google Scholar]
  • 169.Ortega RM, Lopez-Sobaler AM, Gonzalez-Gross MM, Redondo RM, Marzana I, Zamora MJ, et al. Influence of smoking on folate intake and blood folate concentrations in a group of elderly Spanish men. Journal of the American College of Nutrition. 1994;13:68–72. doi: 10.1080/07315724.1994.10718374. [DOI] [PubMed] [Google Scholar]
  • 170.Piyathilake CJ, Macaluso M, Hine RJ, Richards EW, Krumdieck CL. Local and systemic effects of cigarette smoking on folate and vitamin B-12. The American journal of clinical nutrition. 1994;60:559–66. doi: 10.1093/ajcn/60.4.559. [DOI] [PubMed] [Google Scholar]
  • 171.Blount BC, Mack MM, Wehr CM, MacGregor JT, Hiatt RA, Wang G, et al. Folate deficiency causes uracil misincorporation into human DNA and chromosome breakage: implications for cancer and neuronal damage. Proceedings of the National Academy of Sciences of the United States of America. 1997;94:3290–5. doi: 10.1073/pnas.94.7.3290. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Fang JY, Xiao SD. Folic acid, polymorphism of methyl-group metabolism genes, and DNA methylation in relation to GI carcinogenesis. Journal of gastroenterology. 2003;38:821–9. doi: 10.1007/s00535-003-1207-7. [DOI] [PubMed] [Google Scholar]
  • 173.Cui FM, Li JX, Chen Q, Du HB, Zhang SY, Nie JH, et al. Radon-induced alterations in micro-RNA expression profiles in transformed BEAS2B cells. Journal of toxicology and environmental health Part A. 2013;76:107–19. doi: 10.1080/15287394.2013.738176. [DOI] [PubMed] [Google Scholar]
  • 174.Su S, Jin Y, Zhang W, Yang L, Shen Y, Cao Y, et al. Aberrant promoter methylation of p16(INK4a) and O(6)-methylguanine-DNA methyltransferase genes in workers at a Chinese uranium mine. Journal of occupational health. 2006;48:261–6. doi: 10.1539/joh.48.261. [DOI] [PubMed] [Google Scholar]
  • 175.Pulling LC, Divine KK, Klinge DM, Gilliland FD, Kang T, Schwartz AG, et al. Promoter hypermethylation of the O6-methylguanine-DNA methyltransferase gene: more common in lung adenocarcinomas from never-smokers than smokers and associated with tumor progression. Cancer research. 2003;63:4842–8. [PubMed] [Google Scholar]
  • 176.Johnson DH, Fehrenbacher L, Novotny WF, Herbst RS, Nemunaitis JJ, Jablons DM, et al. Randomized phase II trial comparing bevacizumab plus carboplatin and paclitaxel with carboplatin and paclitaxel alone in previously untreated locally advanced or metastatic non-small-cell lung cancer. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2004;22:2184–91. doi: 10.1200/JCO.2004.11.022. [DOI] [PubMed] [Google Scholar]
  • 177.Sandler A, Gray R, Perry MC, Brahmer J, Schiller JH, Dowlati A, et al. Paclitaxel-carboplatin alone or with bevacizumab for non-small-cell lung cancer. The New England journal of medicine. 2006;355:2542–50. doi: 10.1056/NEJMoa061884. [DOI] [PubMed] [Google Scholar]
  • 178.Reck M, von Pawel J, Zatloukal P, Ramlau R, Gorbounova V, Hirsh V, et al. Phase III trial of cisplatin plus gemcitabine with either placebo or bevacizumab as first-line therapy for nonsquamous non-small-cell lung cancer: AVAil. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2009;27:1227–34. doi: 10.1200/JCO.2007.14.5466. [DOI] [PubMed] [Google Scholar]
  • 179.Scagliotti G, Brodowicz T, Shepherd FA, Zielinski C, Vansteenkiste J, Manegold C, et al. Treatment-by-histology interaction analyses in three phase III trials show superiority of pemetrexed in nonsquamous non-small cell lung cancer. Journal of thoracic oncology : official publication of the International Association for the Study of Lung Cancer. 2011;6:64–70. doi: 10.1097/JTO.0b013e3181f7c6d4. [DOI] [PubMed] [Google Scholar]
  • 180.Kwon YJ, Lee SJ, Koh JS, Kim SH, Lee HW, Kang MC, et al. Genome-wide analysis of DNA methylation and the gene expression change in lung cancer. Journal of thoracic oncology : official publication of the International Association for the Study of Lung Cancer. 2012;7:20–33. doi: 10.1097/JTO.0b013e3182307f62. [DOI] [PubMed] [Google Scholar]
  • 181.Castro M, Grau L, Puerta P, Gimenez L, Venditti J, Quadrelli S, et al. Multiplexed methylation profiles of tumor suppressor genes and clinical outcome in lung cancer. Journal of translational medicine. 2010;8:86. doi: 10.1186/1479-5876-8-86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Hawes SE, Stern JE, Feng Q, Wiens LW, Rasey JS, Lu H, et al. DNA hypermethylation of tumors from non-small cell lung cancer (NSCLC) patients is associated with gender and histologic type. Lung cancer. 2010;69:172–9. doi: 10.1016/j.lungcan.2009.11.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Lockwood WW, Wilson IM, Coe BP, Chari R, Pikor LA, Thu KL, et al. Divergent genomic and epigenomic landscapes of lung cancer subtypes underscore the selection of different oncogenic pathways during tumor development. PloS one. 2012;7:e37775. doi: 10.1371/journal.pone.0037775. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 184.Christensen BC, Marsit CJ, Houseman EA, Godleski JJ, Longacker JL, Zheng S, et al. Differentiation of lung adenocarcinoma, pleural mesothelioma, and nonmalignant pulmonary tissues using DNA methylation profiles. Cancer research. 2009;69:6315–21. doi: 10.1158/0008-5472.CAN-09-1073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 185.Bishop JA, Benjamin H, Cholakh H, Chajut A, Clark DP, Westra WH. Accurate classification of non-small cell lung carcinoma using a novel microRNA-based approach. Clinical cancer research : an official journal of the American Association for Cancer Research. 2010;16:610–9. doi: 10.1158/1078-0432.CCR-09-2638. [DOI] [PubMed] [Google Scholar]
  • 186.Lebanony D, Benjamin H, Gilad S, Ezagouri M, Dov A, Ashkenazi K, et al. Diagnostic assay based on hsa-miR-205 expression distinguishes squamous from nonsquamous non-small-cell lung carcinoma. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2009;27:2030–7. doi: 10.1200/JCO.2008.19.4134. [DOI] [PubMed] [Google Scholar]
  • 187.Hoffman PC, Mauer AM, Vokes EE. Lung cancer. Lancet. 2000;355:479–85. doi: 10.1016/S0140-6736(00)82038-3. [DOI] [PubMed] [Google Scholar]
  • 188.Field JK, Baldwin D, Brain K, Devaraj A, Eisen T, Duffy SW, et al. CT screening for lung cancer in the UK: position statement by UKLS investigators following the NLST report. Thorax. 2011;66:736–7. doi: 10.1136/thoraxjnl-2011-200351. [DOI] [PubMed] [Google Scholar]
  • 189.Isbell JM, Deppen S, Putnam JB, Jr, Nesbitt JC, Lambright ES, Dawes A, et al. Existing general population models inaccurately predict lung cancer risk in patients referred for surgical evaluation. The Annals of thoracic surgery. 2011;91:227–33. doi: 10.1016/j.athoracsur.2010.08.054. discussion 33. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Machida EO, Brock MV, Hooker CM, Nakayama J, Ishida A, Amano J, et al. Hypermethylation of ASC/TMS1 is a sputum marker for late-stage lung cancer. Cancer research. 2006;66:6210–8. doi: 10.1158/0008-5472.CAN-05-4447. [DOI] [PubMed] [Google Scholar]
  • 191.Leng S, Do K, Yingling CM, Picchi MA, Wolf HJ, Kennedy TC, et al. Defining a gene promoter methylation signature in sputum for lung cancer risk assessment. Clinical cancer research : an official journal of the American Association for Cancer Research. 2012;18:3387–95. doi: 10.1158/1078-0432.CCR-11-3049. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 192.de Fraipont F, Moro-Sibilot D, Michelland S, Brambilla E, Brambilla C, Favrot MC. Promoter methylation of genes in bronchial lavages: a marker for early diagnosis of primary and relapsing non-small cell lung cancer? Lung cancer. 2005;50:199–209. doi: 10.1016/j.lungcan.2005.05.019. [DOI] [PubMed] [Google Scholar]
  • 193.Dietrich D, Kneip C, Raji O, Liloglou T, Seegebarth A, Schlegel T, et al. Performance evaluation of the DNA methylation biomarker SHOX2 for the aid in diagnosis of lung cancer based on the analysis of bronchial aspirates. International journal of oncology. 2012;40:825–32. doi: 10.3892/ijo.2011.1264. [DOI] [PubMed] [Google Scholar]
  • 194.Schmiemann V, Bocking A, Kazimirek M, Onofre AS, Gabbert HE, Kappes R, et al. Methylation assay for the diagnosis of lung cancer on bronchial aspirates: a cohort study. Clinical cancer research : an official journal of the American Association for Cancer Research. 2005;11:7728–34. doi: 10.1158/1078-0432.CCR-05-0999. [DOI] [PubMed] [Google Scholar]
  • 195.Hu YC, Sidransky D, Ahrendt SA. Molecular detection approaches for smoking associated tumors. Oncogene. 2002;21:7289–97. doi: 10.1038/sj.onc.1205805. [DOI] [PubMed] [Google Scholar]
  • 196.Simkin M, Abdalla M, El-Mogy M, Haj-Ahmad Y. Differences in the quantity of DNA found in the urine and saliva of smokers versus nonsmokers: implications for the timing of epigenetic events. Epigenomics. 2012;4:343–52. doi: 10.2217/epi.12.24. [DOI] [PubMed] [Google Scholar]
  • 197.Belinsky SA, Liechty KC, Gentry FD, Wolf HJ, Rogers J, Vu K, et al. Promoter hypermethylation of multiple genes in sputum precedes lung cancer incidence in a high-risk cohort. Cancer research. 2006;66:3338–44. doi: 10.1158/0008-5472.CAN-05-3408. [DOI] [PubMed] [Google Scholar]
  • 198.Greenberg AK, Rimal B, Felner K, Zafar S, Hung J, Eylers E, et al. S-adenosylmethionine as a biomarker for the early detection of lung cancer. Chest. 2007;132:1247–52. doi: 10.1378/chest.07-0622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Shen J, Todd NW, Zhang H, Yu L, Lingxiao X, Mei Y, et al. Plasma microRNAs as potential biomarkers for non-small-cell lung cancer. Laboratory investigation; a journal of technical methods and pathology. 2011;91:579–87. doi: 10.1038/labinvest.2010.194. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Ma J, Li N, Guarnera M, Jiang F. Quantification of Plasma miRNAs by Digital PCR for Cancer Diagnosis. Biomarker insights. 2013;8:127–36. doi: 10.4137/BMI.S13154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Weber DG, Johnen G, Casjens S, Bryk O, Pesch B, Jockel KH, et al. Evaluation of long noncoding RNA MALAT1 as a candidate blood-based biomarker for the diagnosis of non-small cell lung cancer. BMC research notes. 2013;6:518. doi: 10.1186/1756-0500-6-518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 202.Shen Y, Wang T, Yang T, Hu Q, Wan C, Chen L, et al. Diagnostic value of circulating microRNAs for lung cancer: a meta-analysis. Genetic testing and molecular biomarkers. 2013;17:359–66. doi: 10.1089/gtmb.2012.0370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Wang J, Lee JJ, Wang L, Liu DD, Lu C, Fan YH, et al. Value of p16INK4a and RASSF1A promoter hypermethylation in prognosis of patients with resectable non-small cell lung cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2004;10:6119–25. doi: 10.1158/1078-0432.CCR-04-0652. [DOI] [PubMed] [Google Scholar]
  • 204.Gu J, Berman D, Lu C, Wistuba, Roth JA, Frazier M, et al. Aberrant promoter methylation profile and association with survival in patients with non-small cell lung cancer. Clinical cancer research : an official journal of the American Association for Cancer Research. 2006;12:7329–38. doi: 10.1158/1078-0432.CCR-06-0894. [DOI] [PubMed] [Google Scholar]
  • 205.Kim JS, Kim JW, Han J, Shim YM, Park J, Kim DH. Cohypermethylation of p16 and FHIT promoters as a prognostic factor of recurrence in surgically resected stage I non-small cell lung cancer. Cancer research. 2006;66:4049–54. doi: 10.1158/0008-5472.CAN-05-3813. [DOI] [PubMed] [Google Scholar]
  • 206.Ota N, Kawakami K, Okuda T, Takehara A, Hiranuma C, Oyama K, et al. Prognostic significance of p16(INK4a) hypermethylation in non-small cell lung cancer is evident by quantitative DNA methylation analysis. Anticancer research. 2006;26:3729–32. [PubMed] [Google Scholar]
  • 207.Suzuki M, Mohamed S, Nakajima T, Kubo R, Tian L, Fujiwara T, et al. Aberrant methylation of CXCL12 in non-small cell lung cancer is associated with an unfavorable prognosis. International journal of oncology. 2008;33:113–9. [PubMed] [Google Scholar]
  • 208.Seng TJ, Currey N, Cooper WA, Lee CS, Chan C, Horvath L, et al. DLEC1 and MLH1 promoter methylation are associated with poor prognosis in non-small cell lung carcinoma. British journal of cancer. 2008;99:375–82. doi: 10.1038/sj.bjc.6604452. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 209.Usadel H, Brabender J, Danenberg KD, Jeronimo C, Harden S, Engles J, et al. Quantitative adenomatous polyposis coli promoter methylation analysis in tumor tissue, serum, and plasma DNA of patients with lung cancer. Cancer research. 2002;62:371–5. [PubMed] [Google Scholar]
  • 210.Brock MV, Hooker CM, Ota-Machida E, Han Y, Guo M, Ames S, et al. DNA methylation markers and early recurrence in stage I lung cancer. The New England journal of medicine. 2008;358:1118–28. doi: 10.1056/NEJMoa0706550. [DOI] [PubMed] [Google Scholar]
  • 211.Suzuki M, Hao C, Takahashi T, Shigematsu H, Shivapurkar N, Sathyanarayana UG, et al. Aberrant methylation of SPARC in human lung cancers. British journal of cancer. 2005;92:942–8. doi: 10.1038/sj.bjc.6602376. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 212.Yoshino M, Suzuki M, Tian L, Moriya Y, Hoshino H, Okamoto T, et al. Promoter hypermethylation of the p16 and Wif-1 genes as an independent prognostic marker in stage IA non-small cell lung cancers. International journal of oncology. 2009;35:1201–9. doi: 10.3892/ijo_00000437. [DOI] [PubMed] [Google Scholar]
  • 213.Xing J, Stewart DJ, Gu J, Lu C, Spitz MR, Wu X. Expression of methylation-related genes is associated with overall survival in patients with non-small cell lung cancer. British journal of cancer. 2008;98:1716–22. doi: 10.1038/sj.bjc.6604343. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 214.Strauss GM, Herndon JE, 2nd, Maddaus MA, Johnstone DW, Johnson EA, Harpole DH, et al. Adjuvant paclitaxel plus carboplatin compared with observation in stage IB non-small-cell lung cancer: CALGB 9633 with the Cancer and Leukemia Group B, Radiation Therapy Oncology Group, and North Central Cancer Treatment Group Study Groups. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2008;26:5043–51. doi: 10.1200/JCO.2008.16.4855. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 215.Kelsey CR, Marks LB, Hollis D, Hubbs JL, Ready NE, D’Amico TA, et al. Local recurrence after surgery for early stage lung cancer: an 11-year experience with 975 patients. Cancer. 2009;115:5218–27. doi: 10.1002/cncr.24625. [DOI] [PubMed] [Google Scholar]
  • 216.Shoji F, Haro A, Yoshida T, Ito K, Morodomi Y, Yano T, et al. Prognostic significance of intratumoral blood vessel invasion in pathologic stage IA non-small cell lung cancer. The Annals of thoracic surgery. 2010;89:864–9. doi: 10.1016/j.athoracsur.2009.09.047. [DOI] [PubMed] [Google Scholar]
  • 217.Barlesi F, Giaccone G, Gallegos-Ruiz MI, Loundou A, Span SW, Lefesvre P, et al. Global histone modifications predict prognosis of resected non small-cell lung cancer. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2007;25:4358–64. doi: 10.1200/JCO.2007.11.2599. [DOI] [PubMed] [Google Scholar]
  • 218.Seligson DB, Horvath S, McBrian MA, Mah V, Yu H, Tze S, et al. Global levels of histone modifications predict prognosis in different cancers. The American journal of pathology. 2009;174:1619–28. doi: 10.2353/ajpath.2009.080874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Kikuchi J, Kinoshita I, Shimizu Y, Kikuchi E, Konishi J, Oizumi S, et al. Distinctive expression of the polycomb group proteins Bmi1 polycomb ring finger oncogene and enhancer of zeste homolog 2 in nonsmall cell lung cancers and their clinical and clinicopathologic significance. Cancer. 2010;116:3015–24. doi: 10.1002/cncr.25128. [DOI] [PubMed] [Google Scholar]
  • 220.Zhang XY, Dong QG, Huang JS, Huang AM, Shi CL, Jin B, et al. The expression of stem cell-related indicators as a prognostic factor in human lung adenocarcinoma. Journal of surgical oncology. 2010;102:856–62. doi: 10.1002/jso.21718. [DOI] [PubMed] [Google Scholar]
  • 221.Crea F, Sun L, Pikor L, Frumento P, Lam WL, Helgason CD. Mutational analysis of Polycomb genes in solid tumours identifies PHC3 amplification as a possible cancer-driving genetic alteration. British journal of cancer. 2013;109:1699–702. doi: 10.1038/bjc.2013.454. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Yang M, Shen H, Qiu C, Ni Y, Wang L, Dong W, et al. High expression of miR-21 and miR-155 predicts recurrence and unfavourable survival in non-small cell lung cancer. European journal of cancer. 2013;49:604–15. doi: 10.1016/j.ejca.2012.09.031. [DOI] [PubMed] [Google Scholar]
  • 223.Liu Z, Sun M, Lu K, Liu J, Zhang M, Wu W, et al. The long noncoding RNA HOTAIR contributes to cisplatin resistance of human lung adenocarcinoma cells via downregualtion of p21(WAF1/CIP1) expression. PloS one. 2013;8:e77293. doi: 10.1371/journal.pone.0077293. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 224.Zandberga E, Kozirovskis V, Abols A, Andrejeva D, Purkalne G, Line A. Cell-free microRNAs as diagnostic, prognostic, and predictive biomarkers for lung cancer. Genes, chromosomes & cancer. 2013;52:356–69. doi: 10.1002/gcc.22032. [DOI] [PubMed] [Google Scholar]
  • 225.Markou A, Sourvinou I, Vorkas PA, Yousef GM, Lianidou E. Clinical evaluation of microRNA expression profiling in non small cell lung cancer. Lung cancer. 2013;81:388–96. doi: 10.1016/j.lungcan.2013.05.007. [DOI] [PubMed] [Google Scholar]
  • 226.Sanfiorenzo C, Ilie MI, Belaid A, Barlesi F, Mouroux J, Marquette CH, et al. Two panels of plasma microRNAs as non-invasive biomarkers for prediction of recurrence in resectable NSCLC. PloS one. 2013;8:e54596. doi: 10.1371/journal.pone.0054596. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Boeri M, Verri C, Conte D, Roz L, Modena P, Facchinetti F, et al. MicroRNA signatures in tissues and plasma predict development and prognosis of computed tomography detected lung cancer. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:3713–8. doi: 10.1073/pnas.1100048108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Hu Z, Chen X, Zhao Y, Tian T, Jin G, Shu Y, et al. Serum microRNA signatures identified in a genome-wide serum microRNA expression profiling predict survival of non-small-cell lung cancer. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2010;28:1721–6. doi: 10.1200/JCO.2009.24.9342. [DOI] [PubMed] [Google Scholar]
  • 229.Liu XG, Zhu WY, Huang YY, Ma LN, Zhou SQ, Wang YK, et al. High expression of serum miR-21 and tumor miR-200c associated with poor prognosis in patients with lung cancer. Medical oncology. 2012;29:618–26. doi: 10.1007/s12032-011-9923-y. [DOI] [PubMed] [Google Scholar]
  • 230.Silva J, Garcia V, Zaballos A, Provencio M, Lombardia L, Almonacid L, et al. Vesicle-related microRNAs in plasma of nonsmall cell lung cancer patients and correlation with survival. The European respiratory journal. 2011;37:617–23. doi: 10.1183/09031936.00029610. [DOI] [PubMed] [Google Scholar]
  • 231.Wang ZX, Bian HB, Wang JR, Cheng ZX, Wang KM, De W. Prognostic significance of serum miRNA-21 expression in human non-small cell lung cancer. Journal of surgical oncology. 2011;104:847–51. doi: 10.1002/jso.22008. [DOI] [PubMed] [Google Scholar]
  • 232.Wei J, Gao W, Zhu CJ, Liu YQ, Mei Z, Cheng T, et al. Identification of plasma microRNA-21 as a biomarker for early detection and chemosensitivity of non-small cell lung cancer. Chinese journal of cancer. 2011;30:407–14. doi: 10.5732/cjc.010.10522. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Houseman EA, Accomando WP, Koestler DC, Christensen BC, Marsit CJ, Nelson HH, et al. DNA methylation arrays as surrogate measures of cell mixture distribution. BMC bioinformatics. 2012;13:86. doi: 10.1186/1471-2105-13-86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Ramirez JL, Rosell R, Taron M, Sanchez-Ronco M, Alberola V, de Las Penas R, et al. 14-3-3sigma methylation in pretreatment serum circulating DNA of cisplatin-plus-gemcitabine-treated advanced non-small-cell lung cancer patients predicts survival: The Spanish Lung Cancer Group. Journal of clinical oncology : official journal of the American Society of Clinical Oncology. 2005;23:9105–12. doi: 10.1200/JCO.2005.02.2905. [DOI] [PubMed] [Google Scholar]
  • 235.Ponomaryova AA, Rykova EY, Cherdyntseva NV, Skvortsova TE, Dobrodeev AY, Zav’yalov AA, et al. RARbeta2 gene methylation level in the circulating DNA from blood of patients with lung cancer. European journal of cancer prevention : the official journal of the European Cancer Prevention Organisation. 2011;20:453–5. doi: 10.1097/CEJ.0b013e3283498eb4. [DOI] [PubMed] [Google Scholar]
  • 236.Li B, Ren S, Li X, Wang Y, Garfield D, Zhou S, et al. MiR-21 overexpression is associated with acquired resistance of EGFR-TKI in non-small cell lung cancer. Lung cancer. 2014;83:146–53. doi: 10.1016/j.lungcan.2013.11.003. [DOI] [PubMed] [Google Scholar]
  • 237.Wang YS, Wang YH, Xia HP, Zhou SW, Schmid-Bindert G, Zhou CC. MicroRNA-214 regulates the acquired resistance to gefitinib via the PTEN/AKT pathway in EGFR-mutant cell lines. Asian Pacific journal of cancer prevention : APJCP. 2012;13:255–60. doi: 10.7314/apjcp.2012.13.1.255. [DOI] [PubMed] [Google Scholar]
  • 238.Garofalo M, Romano G, Di Leva G, Nuovo G, Jeon YJ, Ngankeu A, et al. EGFR and MET receptor tyrosine kinase-altered microRNA expression induces tumorigenesis and gefitinib resistance in lung cancers. Nature medicine. 2012;18:74–82. doi: 10.1038/nm.2577. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
  • 239.Zhong M, Ma X, Sun C, Chen L. MicroRNAs reduce tumor growth and contribute to enhance cytotoxicity induced by gefitinib in non-small cell lung cancer. Chemico-biological interactions. 2010;184:431–8. doi: 10.1016/j.cbi.2010.01.025. [DOI] [PubMed] [Google Scholar]
  • 240.Kitamura K, Seike M, Okano T, Matsuda K, Miyanaga A, Mizutani H, et al. MiR-134/487b/655 Cluster Regulates TGF-beta-Induced Epithelial-Mesenchymal Transition and Drug Resistance to Gefitinib by Targeting MAGI2 in Lung Adenocarcinoma Cells. Molecular cancer therapeutics. 2014;13:444–53. doi: 10.1158/1535-7163.MCT-13-0448. [DOI] [PubMed] [Google Scholar]
  • 241.Cao M, Seike M, Soeno C, Mizutani H, Kitamura K, Minegishi Y, et al. MiR-23a regulates TGF-beta-induced epithelial-mesenchymal transition by targeting E-cadherin in lung cancer cells. International journal of oncology. 2012;41:869–75. doi: 10.3892/ijo.2012.1535. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 242.Cufi S, Bonavia R, Vazquez-Martin A, Oliveras-Ferraros C, Corominas-Faja B, Cuyas E, et al. Silibinin suppresses EMT-driven erlotinib resistance by reversing the high miR-21/low miR-200c signature in vivo. Scientific reports. 2013;3:2459. doi: 10.1038/srep02459. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Bryant JL, Britson J, Balko JM, Willian M, Timmons R, Frolov A, et al. A microRNA gene expression signature predicts response to erlotinib in epithelial cancer cell lines and targets EMT. British journal of cancer. 2012;106:148–56. doi: 10.1038/bjc.2011.465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Zhan M, Qu Q, Wang G, Zhou H. Let-7c sensitizes acquired cisplatin-resistant A549 cells by targeting ABCC2 and Bcl-XL. Die Pharmazie. 2013;68:955–61. [PubMed] [Google Scholar]
  • 245.Dong Z, Zhong Z, Yang L, Wang S, Gong Z. MicroRNA-31 inhibits cisplatin-induced apoptosis in non-small cell lung cancer cells by regulating the drug transporter ABCB9. Cancer letters. 2014;343:249–57. doi: 10.1016/j.canlet.2013.09.034. [DOI] [PubMed] [Google Scholar]
  • 246.Li Y, Li L, Guan Y, Liu X, Meng Q, Guo Q. MiR-92b regulates the cell growth, cisplatin chemosensitivity of A549 non small cell lung cancer cell line and target PTEN. Biochemical and biophysical research communications. 2013;440:604–10. doi: 10.1016/j.bbrc.2013.09.111. [DOI] [PubMed] [Google Scholar]
  • 247.Song L, Li Y, Li W, Wu S, Li Z. MiR-495 enhances the sensitivity of non-small cell lung cancer cells to platinum by modulation of copper-transporting P-type adenosine triphosphatase A (ATP7A) Journal of cellular biochemistry. 2013 doi: 10.1002/jcb.24665. [DOI] [PubMed] [Google Scholar]
  • 248.Qiu T, Zhou L, Wang T, Xu J, Wang J, Chen W, et al. miR-503 regulates the resistance of non-small cell lung cancer cells to cisplatin by targeting Bcl-2. International journal of molecular medicine. 2013;32:593–8. doi: 10.3892/ijmm.2013.1439. [DOI] [PubMed] [Google Scholar]
  • 249.Xiang Q, Tang H, Yu J, Yin J, Yang X, Lei X. MicroRNA-98 sensitizes cisplatin-resistant human lung adenocarcinoma cells by up-regulation of HMGA2. Die Pharmazie. 2013;68:274–81. [PubMed] [Google Scholar]
  • 250.Zhou L, Qiu T, Xu J, Wang T, Wang J, Zhou X, et al. miR-135a/b modulate cisplatin resistance of human lung cancer cell line by targeting MCL1. Pathology oncology research : POR. 2013;19:677–83. doi: 10.1007/s12253-013-9630-4. [DOI] [PubMed] [Google Scholar]
  • 251.Pouliot LM, Shen DW, Suzuki T, Hall MD, Gottesman MM. Contributions of microRNA dysregulation to cisplatin resistance in adenocarcinoma cells. Experimental cell research. 2013;319:566–74. doi: 10.1016/j.yexcr.2012.10.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Zang YS, Zhong YF, Fang Z, Li B, An J. MiR-155 inhibits the sensitivity of lung cancer cells to cisplatin via negative regulation of Apaf-1 expression. Cancer gene therapy. 2012;19:773–8. doi: 10.1038/cgt.2012.60. [DOI] [PubMed] [Google Scholar]
  • 253.Liu ZL, Wang H, Liu J, Wang ZX. MicroRNA-21 (miR-21) expression promotes growth, metastasis, and chemo- or radioresistance in non-small cell lung cancer cells by targeting PTEN. Molecular and cellular biochemistry. 2013;372:35–45. doi: 10.1007/s11010-012-1443-3. [DOI] [PubMed] [Google Scholar]
  • 254.Zhu W, Xu H, Zhu D, Zhi H, Wang T, Wang J, et al. miR-200bc/429 cluster modulates multidrug resistance of human cancer cell lines by targeting BCL2 and XIAP. Cancer chemotherapy and pharmacology. 2012;69:723–31. doi: 10.1007/s00280-011-1752-3. [DOI] [PubMed] [Google Scholar]
  • 255.Ceppi P, Mudduluru G, Kumarswamy R, Rapa I, Scagliotti GV, Papotti M, et al. Loss of miR-200c expression induces an aggressive, invasive, and chemoresistant phenotype in non-small cell lung cancer. Molecular cancer research : MCR. 2010;8:1207–16. doi: 10.1158/1541-7786.MCR-10-0052. [DOI] [PubMed] [Google Scholar]
  • 256.Wang Q, Zhong M, Liu W, Li J, Huang J, Zheng L. Alterations of microRNAs in cisplatin-resistant human non-small cell lung cancer cells (A549/DDP) Experimental lung research. 2011;37:427–34. doi: 10.3109/01902148.2011.584263. [DOI] [PubMed] [Google Scholar]
  • 257.Bian HB, Pan X, Yang JS, Wang ZX, De W. Upregulation of microRNA-451 increases cisplatin sensitivity of non-small cell lung cancer cell line (A549) Journal of experimental & clinical cancer research : CR. 2011;30:20. doi: 10.1186/1756-9966-30-20. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Zhu W, Zhu D, Lu S, Wang T, Wang J, Jiang B, et al. miR-497 modulates multidrug resistance of human cancer cell lines by targeting BCL2. Medical oncology. 2012;29:384–91. doi: 10.1007/s12032-010-9797-4. [DOI] [PubMed] [Google Scholar]
  • 259.Zhu W, Shan X, Wang T, Shu Y, Liu P. miR-181b modulates multidrug resistance by targeting BCL2 in human cancer cell lines. International journal of cancer Journal international du cancer. 2010;127:2520–9. doi: 10.1002/ijc.25260. [DOI] [PubMed] [Google Scholar]
  • 260.Wang XC, Wang W, Zhang ZB, Zhao J, Tan XG, Luo JC. Overexpression of miRNA-21 promotes radiation-resistance of non-small cell lung cancer. Radiation oncology. 2013;8:146. doi: 10.1186/1748-717X-8-146. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 261.Grosso S, Doyen J, Parks SK, Bertero T, Paye A, Cardinaud B, et al. MiR-210 promotes a hypoxic phenotype and increases radioresistance in human lung cancer cell lines. Cell death & disease. 2013;4:e544. doi: 10.1038/cddis.2013.71. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Salim H, Akbar NS, Zong D, Vaculova AH, Lewensohn R, Moshfegh A, et al. miRNA-214 modulates radiotherapy response of non-small cell lung cancer cells through regulation of p38MAPK, apoptosis and senescence. British journal of cancer. 2012;107:1361–73. doi: 10.1038/bjc.2012.382. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Lee KM, Choi EJ, Kim IA. microRNA-7 increases radiosensitivity of human cancer cells with activated EGFR-associated signaling. Radiotherapy and oncology : journal of the European Society for Therapeutic Radiology and Oncology. 2011;101:171–6. doi: 10.1016/j.radonc.2011.05.050. [DOI] [PubMed] [Google Scholar]
  • 264.Arora H, Qureshi R, Jin S, Park AK, Park WY. miR-9 and let-7g enhance the sensitivity to ionizing radiation by suppression of NFkappaB1. Experimental & molecular medicine. 2011;43:298–304. doi: 10.3858/emm.2011.43.5.031. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 265.Jeong SH, Wu HG, Park WY. LIN28B confers radio-resistance through the posttranscriptional control of KRAS. Experimental & molecular medicine. 2009;41:912–8. doi: 10.3858/emm.2009.41.12.097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Sharma SV, Lee DY, Li B, Quinlan MP, Takahashi F, Maheswaran S, et al. A chromatin-mediated reversible drug-tolerant state in cancer cell subpopulations. Cell. 2010;141:69–80. doi: 10.1016/j.cell.2010.02.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 267.Gerlinger M, Rowan AJ, Horswell S, Larkin J, Endesfelder D, Gronroos E, et al. Intratumor heterogeneity and branched evolution revealed by multiregion sequencing. The New England journal of medicine. 2012;366:883–92. doi: 10.1056/NEJMoa1113205. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 268.Kodani M, Igishi T, Matsumoto S, Chikumi H, Shigeoka Y, Nakanishi H, et al. Suppression of phosphatidylinositol 3-kinase/Akt signaling pathway is a determinant of the sensitivity to a novel histone deacetylase inhibitor, FK228, in lung adenocarcinoma cells. Oncology reports. 2005;13:477–83. [PubMed] [Google Scholar]
  • 269.Yu XD, Wang SY, Chen GA, Hou CM, Zhao M, Hong JA, et al. Apoptosis induced by depsipeptide FK228 coincides with inhibition of survival signaling in lung cancer cells. Cancer journal. 2007;13:105–13. doi: 10.1097/PPO.0b013e318046eedc. [DOI] [PubMed] [Google Scholar]
  • 270.Choi YH. Induction of apoptosis by trichostatin A, a histone deacetylase inhibitor, is associated with inhibition of cyclooxygenase-2 activity in human non-small cell lung cancer cells. International journal of oncology. 2005;27:473–9. [PubMed] [Google Scholar]
  • 271.Doi S, Soda H, Oka M, Tsurutani J, Kitazaki T, Nakamura Y, et al. The histone deacetylase inhibitor FR901228 induces caspase-dependent apoptosis via the mitochondrial pathway in small cell lung cancer cells. Molecular cancer therapeutics. 2004;3:1397–402. [PubMed] [Google Scholar]
  • 272.Tong M, Ding Y, Tai HH. Histone deacetylase inhibitors and transforming growth factor-beta induce 15-hydroxyprostaglandin dehydrogenase expression in human lung adenocarcinoma cells. Biochemical pharmacology. 2006;72:701–9. doi: 10.1016/j.bcp.2006.06.004. [DOI] [PubMed] [Google Scholar]
  • 273.Kaminskyy VO, Surova OV, Vaculova A, Zhivotovsky B. Combined inhibition of DNA methyltransferase and histone deacetylase restores caspase-8 expression and sensitizes SCLC cells to TRAIL. Carcinogenesis. 2011;32:1450–8. doi: 10.1093/carcin/bgr135. [DOI] [PubMed] [Google Scholar]
  • 274.Nam JS, Ino Y, Kanai Y, Sakamoto M, Hirohashi S. 5-aza-2′-deoxycytidine restores the E-cadherin system in E-cadherin-silenced cancer cells and reduces cancer metastasis. Clinical & experimental metastasis. 2004;21:49–56. doi: 10.1023/b:clin.0000017180.19881.c1. [DOI] [PubMed] [Google Scholar]
  • 275.Lockwood WW, Zejnullahu K, Bradner JE, Varmus H. Sensitivity of human lung adenocarcinoma cell lines to targeted inhibition of BET epigenetic signaling proteins. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:19408–13. doi: 10.1073/pnas.1216363109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 276.Rao M, Chinnasamy N, Hong JA, Zhang Y, Zhang M, Xi S, et al. Inhibition of histone lysine methylation enhances cancer-testis antigen expression in lung cancer cells: implications for adoptive immunotherapy of cancer. Cancer research. 2011;71:4192–204. doi: 10.1158/0008-5472.CAN-10-2442. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 277.Lawless MW, O’Byrne KJ, Gray SG. Oxidative stress induced lung cancer and COPD: opportunities for epigenetic therapy. Journal of cellular and molecular medicine. 2009;13:2800–21. doi: 10.1111/j.1582-4934.2009.00845.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 278.Rangwala S, Zhang C, Duvic M. HDAC inhibitors for the treatment of cutaneous T-cell lymphomas. Future medicinal chemistry. 2012;4:471–86. doi: 10.4155/fmc.12.6. [DOI] [PubMed] [Google Scholar]
  • 279.Juergens RA, Wrangle J, Vendetti FP, Murphy SC, Zhao M, Coleman B, et al. Combination epigenetic therapy has efficacy in patients with refractory advanced non-small cell lung cancer. Cancer discovery. 2011;1:598–607. doi: 10.1158/2159-8290.CD-11-0214. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES