Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2013 Nov 20.
Published in final edited form as: Nutr Cancer. 2011 Sep 14;63(7):10.1080/01635581.2011.605982. doi: 10.1080/01635581.2011.605982

Promoter methylation of E- cadherin, p16, and RAR-β2 genes in breast tumors and dietary intake of nutrients important in one-carbon metabolism

Meng-Hua Tao 1, Joel B Mason 2, Catalin Marian 3, Susan E McCann 4, Mary E Platek 1,4, Amy Millen 1, Christine Ambrosone 4, Stephen B Edge 4, Shiva S Krishnan 3, Maurizio Trevisan 1,5, Peter G Shields 3, Jo L Freudenheim 1
PMCID: PMC3834353  NIHMSID: NIHMS524755  PMID: 21916701

Abstract

Aberrant DNA methylation plays a critical role in carcinogenesis, and the availability of dietary factors involved in one carbon metabolism may to contribute to aberrant DNA methylation. We investigated the association of intake of folate, vitamins B2, B6, B12, and methionine with promoter methylation of E- cadherin, p16, and RAR-β2 genes in archived tumor tissues from incident, primary breast cancer cases in a population-based case-control study. Real time methylation-specific PCR was performed on 803 paraffin-embedded samples; usual dietary intake was queried from a food frequency questionnaire. Unconditional logistic regression was used to derive adjusted odds ratios (OR) and 95% confidence intervals (CI) for likelihood of promoter methylation for high compared to low intake of those one-carbon nutrients. Overall, in case-case comparisons, dietary intakes of folate, vitamins B2, B6, B12, and methionine were not associated with likelihood of promoter methylation of E- cadherin, p16, and RAR-β2 for all cases combined or within strata defined by menopausal status and ER status in this study. This finding, however, does not exclude the possibility that intake of such nutrients might have the ability to modulate promoter methylation in normal, or pre-malignant (dysplastic) breast tissue.

Keywords: one-carbon nutrient, breast cancer, epidemiology, epigenetics, promoter methylation

Introduction

One carbon metabolism provides methyl groups for a variety of biological process including methylation of DNA, RNA and protein, as well as for synthesis of purines and pyrimidines for DNA synthesis (13). Collectively, nutrients involved in this metabolism are sometimes called “one-carbon nutrients”. Folate, the principle methyl donor, plays a central role in the conversion of methionine to S-adenosylmethione (SAM) (1, 2). Other nutrients including vitamins B2, B6, B12, and the amino acid methionine are also involved in the one carbon metabolism pathway, potentially affecting genomic DNA methylation and synthesis and thereby causing dysregulation of gene expression (2, 48). In addition, alcohol intake has been reported to interfere with one carbon metabolism in a number of ways including negative effects on folate absorption, utilization and excretion, and on activity of metabolic enzymes (1, 9, 10). Findings regarding the association of folate, vitamins B2, B6, B12 and methionine with breast cancer risk has been inconsistent (1113).

Aberrant DNA methylation, including global hypomethylation and gene-specific hypermethylation of promoter regions, is a common epigenetic alteration in breast carcinogenesis. (1417). Little is known about the etiology of these alterations in methylation. Diets deficient in folate and other one carbon nutrients may explain the observed changes, at least in part. There is evidence from pre-clinical models that these one carbon nutrients can affect both genomic DNA methylation as well as the methylation of the promoter region of some important tumor suppressor genes (2, 1820). Moreover, in clinical trials restriction of dietary folate intake has been shown to alter genomic methylation of circulating white blood cells (21, 22). To our knowledge, associations between dietary one-carbon nutrients and DNA methylation in breast cancer have been investigated in only two other studies, both indicating an association of folate intake with DNA hypermethylation of one or more genes (23, 24).

The p16 gene, a tumor suppressor gene, is important in cell cycle regulation (16). There is evidence that p16 methylation may be critical in a cascade of events leading to increased proliferation and genetic instability in human breast cells (25). Dietary folate deficiency has been shown to increase p16 promoter methylation in an animal model (20) as well as in head and neck squamous cell carcinoma (26). In addition, other genes involved in cell adhesion (E-cadherin) (27) and hormone and receptor-mediated cell signaling (i.e. RAR-β2 (retinoic acid-binding receptor-β2)) (17) seem to be responsive to folate status. The expression of the E-cadherin gene was lower with decreasing folate in a human colonic epithelial cell line (28). Loss of expression of E - cadherin caused by promoter methylation occurs frequently in breast cancer (17), suggesting that folate status might affect the Wnt pathway through epigenetic alterations. Promoter methylation of E - cadherin, p16, and RAR-β2 genes has also been shown to occur more frequently in breast tumors than benign or adjacent nonmalignant breast tissue (29, 30) and is associated with poorly differentiated breast tumors, distant metastasis of breast tumors, and estrogen receptor (ER) status of breast tumors (3135).

Based on this apparent relationship of folate and other one carbon nutrients with DNA methylation, we hypothesized that tumors from women with breast cancer with lower dietary intake of one-carbon nutrients would have higher prevalence of methylation than those from women with higher intake. We evaluated associations of reported dietary intakes of folate, methionine, and vitamins B2, B6, B12 with promoter methylation of E - cadherin, p16, and RAR-β2 genes in breast tumors in a population-based study.

Materials and Methods

Study population

Briefly, the Western New York Exposures and Breast Cancer Study (WEB Study) was a population-based case-control study, including female residents of Erie and Niagara counties of New York State, who were diagnosed with primary, histologically confirmed, incident breast cancer between 1996 and 2001 and were between 35 and 79 years of age. This report includes data from cases only. Among 1638 eligible cases, 1,170 (72%) participated. All participants provided informed consent. The protocol was approved by the Institutional Review Boards of all of the participating institutions.

Data on demographics, and other breast cancer risk factors were collected during detailed in-person interview conducted by trained interviewers. The Health Habits and History food frequency questionnaire (FFQ) was used to query habitual dietary consumption 12–24 months before diagnosis (36, 37) with some small adaptations specific to dietary practices in Western New York. Nutrient intake from the FFQ was calculated using the DIETSYS (version 3.7) nutrient analysis software developed specifically for the food frequency questionnaire. Additionally, information was collected on vitamin and mineral supplement use. Total nutrient intake, intake from supplements and diet, was calculated by adding dietary intake with supplementary intake from either multiple supplement use or singular supplement use. The bioavailability of naturally occurring folate in food varies greatly as these forms of folate are labile and easily oxidized; and synthetic folic acid obtained from supplements is much more bioavailable (38). To account for this difference in the absorption of folate from foods and folic acid from supplements, dietary folate equivalents were examined. Folic acid intake from supplements was converted to dietary folate equivalents by multiplying the supplemental amount by 1.7 and this amount was added to food folate intake to determine total folate from diet and supplements (39).

The diagnosis of breast cancer from the medical record was reconfirmed by a single pathologist from Georgetown University based on review of archived tumor blocks. Information on tumor size, histological grade, and cancer stage (as measured by tumor-node-metastasis (TNM) stage) was abstracted from medical records by trained research nurses using a standardized protocol. ER status was determined by the study pathologist, by immunohistochemical analysis as described in detail previously (35).

Tumor block promoter methylation determination

Among the 1170 breast cancer cases in the WEB study, we were able to obtain archived tumor blocks for 920 (78.6%). Tumor samples were microdissected from fixed microscope slides. Bisulfite modification was performed on tumor DNA isolated from tissue blocks in accordance with methods described previously (35, 40). Promoter methylation of E- cadherin, p16, and RAR-β2 was determined by the fluorescence based version of methylation-specific PCR (MSP) using real time PCR amplification of bisulfate converted DNA in an ABI 7900HT real time PCR system as previously described (35, 41). Briefly, each reaction contained 5 μl of Taqman Universal Master Mix (29), 4.5 μl of bisulfite treated DNA and 0.5 μl of a 60X assay by design premix containing the primers and probes that were designed for each respective gene (Applied Biosystems, Carlsbad CA); the primers and probes used in this study have been described previously (35). As a control to check for modified viable DNA, we used an assay for the ACTB gene with primers and probes specially designed for CpG free sites within the gene sequence, thus amplifying the modified DNA regardless of the methylation status. If the ACTB result was negative (i.e. no amplification signal was detected), the DNA could not be used in subsequent assays, and re-extraction and modification was attempted; the other 3 genes being assayed only if ACTB was positive. Each individual DNA sample was assayed in triplicate for each gene for quality control purposes. Additionally, as a positive control, universally methylated DNA (CpGenome; Norcross, GA) was used along with water blanks as a negative control. We had successful promoter methylation results for 803 cases.

Statistical analysis

Characteristics including reported dietary intake of participating cases with and without promoter methylation of each gene were compared with the Student’s t-test for continuous variables and the χ2 test for categorical variables. In addition, the likelihood of promoter methylation according to dietary methyl donor intake was examined with a case-case comparison of those with promoter methylation of a gene compared to those without. For these latter analyses, unconditional logistic regression models were used to estimate multivariate adjusted odds ratios (ORs) and 95% confidence intervals (95% CIs) for quartiles of folate, vitamins B2, B6, B12, and methionine intake, using the highest intake category as the referent. Tests for dose-response relationship over the categories of intake were estimated by fitting the models with exposure variables included as continuous variables. All models were adjusted for age at diagnosis, education level, race, and total energy intake. Furthermore, potential confounding effects of other demographic factors, known breast cancer risk factors, and dietary intakes of other nutrients were examined, and the covariates included in the model were those that influenced the OR by more than 10%. Because alcohol consumption can affect one-carbon metabolism through its negative impact on folate absorption and concentration of vitamin B12 (9), we further examined potential interaction between dietary one-carbon nutrients intake and alcohol consumption for each of the specific genes by evaluation of a multiplicative term in the regression model. Intakes of one-carbon nutrients and alcohol in these logistic regression analyses with the interactions were stratified into low or high intake using the median intakes as cut-off points. All statistical tests were based on two-sided probability. Analyses were performed using SAS version 9.2 (SAS Institute, Cary, NC).

Results

Demographic characteristics of cases with and without promoter methylation of E - cadherin, p16, and RAR-β2 gene are shown in Table 1. The frequency of promoter methylation was 20% (161) for E - cadherin, 25.9% (208) for p16, and 27.5% (221) for RAR-β2. Mean levels of dietary folate, vitamins B2, B6 and B12 intakes were similar among cases with methylated of each of the three genes compared to those without methylation; intakes of dietary methionine among cases with methylated p16 gene were lower than among those with unmethylated p16.

Table 1.

One-carbon nutrients intake and other characteristics of breast cancer cases, cases with (M) and without (UM) promoter methylation, WEB Study 1996–2001

E-cadherin p16 RAR-β2

M (n = 161)§ UM (n = 742)§ M (n = 208)§ UM (n = 595)§ M (n = 221)§ UM (n = 582)§
Age, y 58.0 ± 11.8 57.4 ± 11.2 58.0 ± 11.2 57.4 ± 11.3 57.4 ± 11.3 57.6 ± 11.3
Race/ ethnicity
 White 148 (91.9%) 594 (92.5%) 192 (92.3%) 550 (92.4%) 210 (95.0%) 532 (91.4%)
 Non-white 13 (8.1%) 48 (7.5%) 16 (7.7%) 45 (7.6%) 11 (5.0%) 50 (8.6%)
ER status
 + 114 (71.3%) 446 (70.0%) 136 (65.7%) 424 (71.9%) 163 (74.8%) 397 (68.6%)
 − 46 (28.7%) 191 (30%) 71 (34.3%) 166 (28.1%) 55 (25.2%) 182 (31.4%)
Alcohol consumption
 Never 20 (12.6%) 110 (17.3%) 46 (22.3%) 84 (14.2%) 41 (18.7%) 89 (15.4%)
 Ever 139 (87.4%) 527 (82.7%) 160 (77.7%) 506 (85.8%) a 178 (81.3%) 488 (84.6%)
Dietary intake
 Folate intake (μg/day) 281.7 ± 178.4 280.7 ± 178.4 275.2 ± 143.2 282.9 ± 158.9 285.5 ± 175.0 279.2 ± 146.6
 Vitamin B2 (mg/day) 1.7 ± 1.2 1.6 ± 1.2 1.5 ± 1.2 1.7 ± 1.2 1.7 ± 1.2 1.6 ± 1.2
 Vitamin B6 (mg/day) 1.4 ± 1.0 1.4 ± 1.1 1.4 ± 1.0 1.5 ± 1.1 1.4 ± 1.0 1.4 ± 1.1
 Vitamin B12 (μg/day) 3.3 ± 2.2 3.3 ± 2.1 3.1 ± 2.1 3.3 ± 2.2 3.3 ± 2.1 3.3 ± 2.2
Methionine (g/day) 1.2 ± 0.9 1.1 ± 0.8 1.1 ± 0.4 a 1.2 ± 0.6 1.2 ± 0.7 1.2 ± 0.5
§

M – methylated; UM – unmethylated;

Mean ± SD;

a

Comparison of methylated to unmethylated cases p < 0.05.

Likelihood of promoter methylation in breast tumors by intake of dietary folate, vitamins B2, B6, B12, and methionine is shown in Table 2. In these case-case comparisons, there was no association of dietary intake of folate, vitamins B2, B6, or B12 with promoter methylation of E – cadherin, p16, or RAR-β2 gene. For cases with lower intake of dietary methionine, compared to those with the highest, there was decreased likelihood of methylated E - cadherin gene (OR = 0.60; 95% CI, 0.31–1.16; p trend 0.05). There were no differences in these associations by menopausal or ER status (data not shown). We also evaluated the associations of methylation with intake of folate, vitamins B2, B6 and B12 from supplements; there were no associations with promoter methylation (data not shown). Additionally, results for supplemental and total intake (diet plus supplements) of folate, vitamins B2, B6, B12, and methionine were similar to those for dietary intake (data not shown). When we analyzed associations between intake of dietary one-carbon nutrients and likelihood of tumors with promoter methylation in at least one gene, no association was observed. We further conducted analysis stratified on menopausal status and ER status.

Table 2.

Dietary intake of one-carbon nutrients and E-cadherin, p16, and RAR-β2 promoter methylation in breast tumors: case-case comparison

Quartile of nutrients E-cadherin p16, RAR-β2

M U OR (95% CI) 1 M U OR (95% CI) 1 M U OR (95% CI) 1
Folate a
 Q4 (highest) 36 141 1.0 40 137 1.0 52 125 1.0
 Q3 35 163 0.95 (0.54–1.58) 57 141 1.18 (0.73–1.91) 53 145 0.96 (0.60–1.52)
 Q2 42 158 1.11 (0.64–1.92) 53 147 0.93 (0.56–1.54) 51 149 0.98 (0.59–1.60)
 Q1 (lowest) 48 180 1.17 (0.66–2.08) 58 170 0.76 (0.44–1.31) 65 163 1.15 (0.69–1.92)
p trend 0.48 0.19 0.61
Vitamin B2 b
 Q4 (highest) 44 143 1.0 42 145 1.0 59 128 1.0
 Q3 38 159 0.74 (0.44–1.24) 42 155 0.83 (0.50–1.39) 57 140 0.92 (0.58–1.46)
 Q2 37 161 0.69 (0.39–1.21) 70 128 1.43 (0.86–2.38) 45 153 0.68 (0.41–1.14)
 Q1 (lowest) 42 179 0.69 (0.37–1.28) 54 167 0.70 (0.39–1.27) 60 161 0.87 (0.50–1.51)
p trend 0.25 0.56 0.36
Vitamin B6 c
 Q4 (highest) 38 136 1.0 38 136 1.0 55 119 1.0
 Q3 41 166 0.93 (0.55–1.56) 53 154 1.06 (0.64–1.73) 51 156 0.78 (0.49–1.25)
 Q2 38 164 0.82 (0.47–1.44) 56 146 1.04 (0.62–1.77) 50 152 0.85 (0.51–1.41)
 Q1 (lowest) 44 176 0.91 (0.49–1.68) 61 159 0.88 (0.50–1.58) 65 155 1.10 (0.64–1.89)
p trend 0.70 0.65 0.62
Vitamin B12 d
 Q4 (highest) 39 145 1.0 45 139 1.0 56 128 1.0
 Q3 41 173 0.95 (0.57–1.59) 54 160 0.90 (0.56–1.45) 52 162 0.76 (0.48–1.20)
 Q2 43 160 1.05 (0.62–1.78) 55 148 0.96 (0.59–1.56) 56 147 0.91 (0.56–1.46)
 Q1 (lowest) 38 164 0.94 (0.53–1.68) 54 148 0.82 (0.48–1.39) 57 145 1.00 (0.60–1.66)
p trend 0.94 0.54 0.78
Methionine e
 Q4 (highest) 41 151 1.0 38 154 1.0 57 135 1.0
 Q3 50 158 1.11 (0.67–1.86) 61 147 1.51 (0.91–2.50) 58 150 0.97 (0.60–1.55)
 Q2 36 158 0.74 (0.41–1.33) 42 152 0.91 (0.52–1.62) 45 149 0.81 (0.48–1.38)
 Q1 (lowest) 34 175 0.60 (0.31–1.16) 67 142 1.41 (0.76–2.60) 61 148 1.45 (0.65–2.04)
p trend 0.05 0.68 0.74
1

Odds ratios and 95% confidence intervals adjusted for age, race, education, energy, lifetime alcohol intake.

a

Quartile cutoff point for dietary folate among cases (μg/day): 178.49, 252.32, 347.65.

b

Quartile cutoff point for dietary vitamin B2 among cases (mg/day): 1.16, 1.63, 2.18.

c

Quartile cutoff point for dietary vitamin B6 among cases (mg/day): 1.06, 1.46, 1.91.

d

Quartile cutoff point for dietary vitamin B12 among cases (μg/day): 2.15, 3.32, 5.10.

e

Quartile cutoff point for dietary methionine among cases (g/day): 0.86, 1.12, 1.45.

We evaluated joint effects of alcohol consumption and intake of one-carbon nutrients on promoter methylation in breast tumors (Table 3). E – cadherin promoter methylation was approximately two times more prevalent among cases with low intake of dietary folate and high alcohol than for those with high folate and low alcohol (OR, 2.05, 95% CI, 1.16–3.62). Similarly, high alcohol and low dietary vitamin B12 was associated with increased prevalence of E – cadherin methylation (OR, 1.85, 95% CI, 1.04–3.27). The difference in likelihood of methylation was associated with alcohol intake, not of the other nutrients. No multiplicative interaction for any of the combinations of alcohol and dietary one-carbon nutrients was observed.

Table 3.

Interactions between dietary intake of one-carbon nutrients and alcohol consumption and likelihood of promoter methylation of E-cadherin, p16 and RAR- β2 in breast tumors: case-case comparison

E-cadherin p16, RAR-β2

M U OR (95% CI) 1 M U OR (95% CI) 1 M U OR (95% CI) 1
Folate/ alcohol intake a
 High/ Low 28 149 1.0 48 129 1.0 51 126 1.0
 High/ High 43 155 1.74 (1.00–3.01) 49 149 0.91 (0.56–1.47) 54 144 0.89 (0.56–1.43)
 Low/ Low 39 174 1.31 (0.73–2.34) 59 154 0.77 (0.47–1.25) 61 152 1.05 (0.65–1.70)
 Low/ High 51 164 2.05 (1.16–3.62) 52 163 0.69 (0.42–1.14) 55 160 0.90 (0.55–1.47)
p interaction 0.77 0.97 0.89
Vitamin B2/ alcohol intake b
 High/ Low 38 147 1.0 42 138 1.0 57 123 1.0
 High/ High 49 155 1.60 (0.96–2.69) 42 162 0.87 (0.53–1.43) 59 145 0.86 (0.55–1.35)
 Low/ Low 34 176 0.86 (0.48–1.56) 65 145 1.21 (0.73–2.00) 55 155 0.77 (0.47–1.27)
 Low/ High 45 164 1.43 (0.81–2.54) 59 150 1.12 (0.67–1.89) 50 159 0.68 (0.41–1.13)
p interaction 0.93 0.84 0.95
Vitamin B6/ alcohol intake c
 High/ Low 30 145 1.0 45 130 1.0 52 123 1.0
 High/ High 49 157 1.74 (1.03–2.95) 46 160 0.84 (0.51–1.36) 54 152 0.83 (0.52–1.61)
 Low/ Low 37 178 1.01 (0.56–1.82) 62 153 0.88 (0.53–1.45) 60 155 1.00 (0.61–1.63)
 Low/ High 45 162 1.56 (0.87–2.78) 55 152 0.85 (0.51–1.41) 55 152 0.92 (0.56–1.52)
p interaction 0.73 0.67 0.74
Vitamin B12/ alcohol intake d
 High/ Low 29 155 1.0 49 135 1.0 53 131 1.0
 High/ High 51 163 2.00 (1.18–3.42) 50 164 0.89 (0.55–1.43) 55 159 0.82 (0.52–1.31)
 Low/ Low 38 168 1.35 (0.77–2.37) 58 148 0.92 (0.58–1.47) 59 147 1.03 (0.65–1.64)
 Low/ High 43 156 1.85 (1.04–3.27) 51 148 0.84 (0.51–1.38) 54 145 0.96 (0.59–1.56)
p interaction 0.29 0.93 0.71
Methionine/ alcohol intake e
 High/ Low 35 146 1.0 52 129 1.0 55 126 1.0
 High/ High 57 167 1.65 (1.00–2.71) 49 175 0.71 (0.45–1.14) 61 163 0.84 (0.54–1.31)
 Low/ Low 32 177 0.69 (0.39–1.24) 55 154 0.63 (0.38–1.03) 57 152 0.92 (0.57–1.50)
 Low/ High 37 152 1.07 (0.60–1.92) 52 137 0.70 (0.42–1.18) 48 141 0.84 (0.50–1.40)
p interaction 0.87 0.17 0.81
1

Odds ratios and 95% confidence intervals adjusted for age, race, education, energy, menopausal status, age at menopause, and smoking status.

Cutoff point for lifetime total alcohol (oz): 697.18.

a

Cutoff point for diet folate (mcg/day):252.32.

b

Cutoff point for diet vitamin B2 (mg/day):1.63.

c

Cutoff point for diet vitamin B6 (mg/day):1.46.

d

Cutoff point for diet vitamin B12 (mcg/day):3.32.

e

Cutoff point for diet methionine (g/day):1.12.

Discussion

In this study, we investigated the potential association of intake of folate, methionine, and vitamins B2, B6, B12 with promoter methylation of E-cadherin, p16 and RAR-β2 genes in breast cancer. We did not observe any difference in the likelihood of promoter methylation based on diet, supplemental or total intake of the selected nutrients for any of these genes. We did see increased likelihood of E – cadherin promoter methylation associated with high alcohol and low dietary folate intakes but that the interaction was not greater than multiplicative and the increase in risk was primarily associated with alcohol consumption.

Frequencies of promoter methylation for p16 and RAR-β2 genes in our sample were similar to previous reports (31); E-cadherin promoter methylation frequency was somewhat lower than has been reported previously (29, 31, 34). This difference may be related to the particular characteristics of our sample, sample size in this or other studies or it may depend on the sensitivity of the MSP assay and differences in MSP assay design. In our study, we used the same assay conditions for each tumor DNA sample and positive and negative internal controls; our MSP analysis was reliable.

To our knowledge, there are only two previous studies of dietary intake of methyl donors in relation to breast cancer. In a case control study of 304 African-American breast cancer cases, there was increased likelihood of methylated ER α gene associated with low dietary folate intake (≤443.9 μg/day) compared to higher intake (23). In a recent study of 162 breast tumors, Christensen et al (24) measured 1,413 autosomal CpG loci associated with 773 cancer-related genes, and applied unsupervised clustering, a recursively partitioned mixture modeling (RPMM), of all CpG loci to reveal eight distinct methylation classes. They reported that alcohol intake and total dietary folate were significantly associated with methylation class membership; however, they also did not find associations between dietary folate and methylation of CpG loci in E-cadherin, p16 and RAR-β2 genes in breast tumors (24). These data suggested that the effect of folate status on DNA methylation may be gene specific (2). It is also possible that putative protective effects of intake of one-carbon nutrients on breast cancer may also be from mechanisms other than the DNA methylation pathway, such as effects on histone methylation (42). Further human studies are needed to evaluate the effects of one-carbon nutrients on other epigenetic gene regulatory mechanisms as well as promoter methylation.

Although our observations suggest that dietary intake of one-carbon vitamins have limited impact on promoter methylation of these critical tumor suppressor genes in existing breast cancers, one should not conclude from these data that availability of one-carbon nutrients cannot impact gene methylation in the human breast. Intervention trials in which human subjects have undergone dietary folate deprivation have clearly shown effects on genomic methylation in circulating white blood cells (21, 22). Moreover, data from a limited number of observational studies have implied that diminished dietary intake of one-carbon nutrients is associated with hypermethylation of tumor suppressor genes in colonic and head & neck cancers (26, 43, 44), although this has not been an entirely consistent finding (45). The endogenous factors that drive gene hypermethylation in established cancers are poorly understood, but nevertheless appear to be quite potent (46). It is entirely feasible, therefore, that alterations in a methylation phenotype produced by environmental factors of modest potency, such as the diet, affect only gene methylation in pre-neoplastic tissue, which appears to be considerably more plastic in this regard. Such a concept is also consistent with the observations from animals studies, where the aberrant gene methylation induced by methyl-deficient has been most readily demonstrated in normal, rather than neoplastic, tissue (2, 19). Such an impact could still affect the carcinogenic process.

There are several strengths and weakness which should be taken into account in assessment of these results regarding diet and methylation in tumors. Strengths of our study include the population-based study design, large sample size with archived tissues obtained from breast cancers, and detailed information on a wide range of potential confounders. Additionally, the FFQ used in the study has been studied and is consistent with other instruments of this kind; it includes most foods usually consumed by the population in Western New York State. However, several limitations, common to other observational studies on diet and cancer, need to be considered. There are well known limitations of the FFQ for assessment of usual intake and this non-differential error would result in bias toward the null in estimation of associations. Another concern of this study is recall bias; cases may differentially recall exposures of dietary because of their cancer diagnosis. However, this bias would not affect these case-case comparisons of tumor characteristics, particularly in that the participants would not be aware of their methylation status. Although the mean level of dietary folate intake for cases in our study was lower than the previous study examining this question (23), it was similar to that reported in the National Health and Nutrition Examination Surveys (NHANES) study during the same period (47). The lower intakes in our population may partly account for the lack of an association with DNA methylation in breast cancer in our study or it may be that there is truly no association. In spite of the relatively large size of the study, the statistical power in subgroups of our study remained limited due to the low frequencies of the promoter methylation, reducing our ability to identify weak associations. Furthermore, we were unable to obtain archived tumor tissue for all breast cancer cases. Compared to those for whom we were unable to obtain tissue, cases with breast tumor tissue were slightly younger at diagnosis and tended to have tumors of more advanced stage; they were similar in terms of tumor size, histological grade, nuclear grade, ER and PR status. It is not likely that there would be a difference in selection of breast cancer cases by promoter methylation for specific gene. Both age and TNM stage are unrelated to methylation of these genes in this population (35); therefore selection bias is not likely, particularly for the case-case comparisons. Another concern with interpretation of these findings is that our study was limited to an examination of methylation in three genes. While these genes were chosen because they are frequently methylated in breast tumors and because there was some evidence that they would be responsive to folate status in breast carcinogenesis, we cannot rule out that methyl donor intake may affect promoter methylation of genes other than those included in our study. Examination of intake of one-carbon nutrients in relation to other genes remains an important area of inquiry. In addition, the methodology used in this study was limited to examination of a single CpG island in the promoter regions. It is assumed that these single regions are sentinels for gene silencing and methylation of other CpG islands, but it is possible that in some women, aberrant methylation of these genes in sequences that we did not study also affects their expression.

Understanding the etiology of aberrant DNA methylation will likely be informative in understanding breast carcinogenesis and in targeting prevention efforts. There are indications that dietary one-carbon nutrients intake may contribute to aberrant DNA methylation. In our study, we found no association between one-carbon nutrients intake and likelihood of gene promoter methylation for E- cadherin, p16 or RAR-β2 in breast tumors or any indication of more than additive interaction for folate and alcohol for E- cadherin promoter methylation. The fact that we examined this question only in tumor tissue, and the fact that several potent forces appear to drive aberrant DNA methylation in neoplasms that may override the effects of altered nutrient availability, may be a responsible for the lack of associations..

Acknowledgments

This study would not have been possible without the support of all the study participants and the research staff of the Western New York Exposures and Breast Cancer (WEB) Study. This work was supported in part by the National Institute on Alcohol Abuse and Alcoholism (P50-AA09802), the Department of Defense (DAMD 179616202, DAMD 17030446), and United States Public Health Service (USPHS) from the National Cancer Institute (R01CA 092040).

References

  • 1.Choi SW, Mason JB. Folate status: effects on pathways of colorectal carcinogenesis. J Nutr. 2002;132:2413S–2418S. doi: 10.1093/jn/132.8.2413S. [DOI] [PubMed] [Google Scholar]
  • 2.Kim YI. Nutritional epigenetics: impact of folate deficiency on DNA methylation and colon cancer susceptibility. J Nutr. 2005;135:2703–2709. doi: 10.1093/jn/135.11.2703. [DOI] [PubMed] [Google Scholar]
  • 3.Mason JB. Biomarkers of nutrient exposure and status in one-carbon (methyl) metabolism. J Nutr. 2003;133:941S–947S. doi: 10.1093/jn/133.3.941S. [DOI] [PubMed] [Google Scholar]
  • 4.Ames BN. DNA damage from micronutrient deficiencies is likely to be a major cause of cancer. Mutation Res. 2001;475:7–20. doi: 10.1016/s0027-5107(01)00070-7. [DOI] [PubMed] [Google Scholar]
  • 5.Davis CD, Uthus EO. DNA methylation, cancer susceptibility, and nutrient interactions. Exp Biol Med (Maywood) 2004;229:988–995. doi: 10.1177/153537020422901002. [DOI] [PubMed] [Google Scholar]
  • 6.Hustad S, Ueland PM, Vollset SE, Zhang Y, Bjørke-Monsen AL, et al. Riboflavin as a determinant of plasma total homocysteine: effect modification by the methylenetetrahydrofolate reductase C677T polymorphism. Clin Chem. 2000;46:1065–1071. [PubMed] [Google Scholar]
  • 7.Jacques PF, Kalmbach R, Bagley PJ, Russo GT, Rogers G, et al. The relationship between riboflavin and plasma total homocysteine in the Framingham Offspring cohort is influenced by folate status and the C677T transition in the methylenetetrahydrofolate reductase gene. J Nutr. 2002;132:283–288. doi: 10.1093/jn/132.2.283. [DOI] [PubMed] [Google Scholar]
  • 8.Moat SJ, Ashfield-Watt PA, Powers HJ, Newcombe RG, McDowell IF. Effect of riboflavin status on the homocysteine-lowering effect of folate in relation to the MTHFR (C677T) genotype. Clin Chem. 2003;49:295–302. doi: 10.1373/49.2.295. [DOI] [PubMed] [Google Scholar]
  • 9.Dumitrescu RG, Shields PG. The etiology of alcohol-induced breast cancer. Alcohol. 2005;35:213–225. doi: 10.1016/j.alcohol.2005.04.005. [DOI] [PubMed] [Google Scholar]
  • 10.Mason JB, Choi SW. Effects of alcohol on folate metabolism: implications for carcinogenesis. Alcohol. 2005;35:235–241. doi: 10.1016/j.alcohol.2005.03.012. [DOI] [PubMed] [Google Scholar]
  • 11.Larsson SC, Giovannucci E, Wolk A. Folate and risk of breast cancer: a meta-analysis. J Natl Cancer Inst. 2007;99:64–76. doi: 10.1093/jnci/djk006. [DOI] [PubMed] [Google Scholar]
  • 12.Lewis SJ, Harbord RM, Harris R, Smith GD. Meta-analyses of observational and genetic association studies of folate intakes or levels and breast cancer risk. J Natl Cancer Inst. 2006;98:1607–1622. doi: 10.1093/jnci/djj440. [DOI] [PubMed] [Google Scholar]
  • 13.Kabat GC, Miller AB, Jain M, Rohan TE. Dietary intake of selected B vitamins in relation to risk of major cancers in women. Br J Cancer. 2008;99:816–821. doi: 10.1038/sj.bjc.6604540. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Wajed SA, Laird PW, DeMeester TR. DNA methylation: an alternative pathway to cancer. Ann Surg. 2001;234:10–20. doi: 10.1097/00000658-200107000-00003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Szyf M, Pakneshan P, Rabbani SA. DNA methylation and breast cancer. Biochem Parmacol. 2004;68:1187–1197. doi: 10.1016/j.bcp.2004.04.030. [DOI] [PubMed] [Google Scholar]
  • 16.Esteller M, Corn PG, Baylin SB, Herman JG. A gene hypermethylation profile of human cancer. Cancer Res. 2001;61:3225–3229. [PubMed] [Google Scholar]
  • 17.Widschwendter M, Jones PA. DNA methylation and breast carcinogenesis. Oncogene. 2002;21:5462–5482. doi: 10.1038/sj.onc.1205606. [DOI] [PubMed] [Google Scholar]
  • 18.Keyes MK, Jang H, Mason JB, Liu Z, Crott JW, et al. Older age and dietary folate are determinants of genomic and p16-specific DNA methylation in mouse colon. J Nutr. 2007;137:1713–1717. doi: 10.1093/jn/137.7.1713. [DOI] [PubMed] [Google Scholar]
  • 19.Ross SA. Diet and DNA methylation interactions in cancer prevention. Ann N Y Acad Sci. 2003;983:197–207. doi: 10.1111/j.1749-6632.2003.tb05974.x. [DOI] [PubMed] [Google Scholar]
  • 20.Pogribny IP, James SJ. De novo methylation of the p16INK4A gene in early preneoplastic liver and tumors induced by folate/methyl deficiency in rats. Cancer Lett. 2002;187:69–75. doi: 10.1016/s0304-3835(02)00408-1. [DOI] [PubMed] [Google Scholar]
  • 21.Jacob RA, Gretz DM, Taylor PC, James SJ, Pogribny IP, et al. Moderate folate depletion increases plasma homocysteine and decreases lymphocyte DNA methylation in postmenopausal women. J Nutr. 1998;128:1204–1212. doi: 10.1093/jn/128.7.1204. [DOI] [PubMed] [Google Scholar]
  • 22.Rampersaud G, Kauwell G, Hutson A, Cerda J, Bailey L. Genomic DNA methylation decreases in response to moderate folate depletion in elderly women. Am J Clin Nutr. 2000;72:998–1003. doi: 10.1093/ajcn/72.4.998. [DOI] [PubMed] [Google Scholar]
  • 23.Zhu K, Davidson NE, Hunter S, Yang X, Payne-Wilks K, et al. Methyl-group dietary intake and risk of breast cancer among African-American women: a case-control study by methylation status of the estrogen receptor alpha genes. Cancer Causes Control. 2003;14:827–36. doi: 10.1023/b:caco.0000003823.97506.be. [DOI] [PubMed] [Google Scholar]
  • 24.Christensen BC, Kelsey KT, Zheng S, Houseman EA, Marsit CJ, et al. Breast cancer DNA methylation profiles are associated with tumor size and alcohol and folate intake. PLos Genet. 2010;6:e1001043. doi: 10.1371/journal.pgen.1001043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Novak P, Jensen TJ, Garbe JC, Stampfer MR, Futscher BW. Stepwise DNA methylation changes are linked to escape from defined proliferation barriers and mammary epithelial cell immortalization. Cancer Res. 2009;69:5251–5258. doi: 10.1158/0008-5472.CAN-08-4977. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Kraunz KS, Hsiung D, McClean MD, Liu M, Osanyingbemi JO, et al. Dietary folate is associated with p16INK4A methylation in head and neck squamous cell carcinoma. Int J Cancer. 2006;119:1553–1557. doi: 10.1002/ijc.22013. [DOI] [PubMed] [Google Scholar]
  • 27.Graff JR, Herman JG, Lapidus RG, Chopra H, Xu R, et al. E-cadherin expression is silence by DNA hypermethylation in human breast and prostate carcinomas. Cancer Res. 1995;55:5195–5199. [PubMed] [Google Scholar]
  • 28.Crott JW, Liu Z, Keyes MK, Choi SW, Jang H, et al. Moderate folate depletion modulates the expression of selected genes involved in cell cycle, intracellular signaling and folate uptake in human colonic epithelial cell lines. J Nutr Biochem. 2008;19:328–335. doi: 10.1016/j.jnutbio.2007.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Parrella P, Poeta ML, Gallo AP, Prencipe M, Scintu M, et al. Nonrandom distribution of aberrant promoter methylation of cancer-related genes in sporadic breast tumors. Clin Cancer Res. 2004;10:5349–5354. doi: 10.1158/1078-0432.CCR-04-0555. [DOI] [PubMed] [Google Scholar]
  • 30.Krassenstein R, Sauter E, Dulaimi E, Battagli C, Ehya H, et al. Detection of breast cancer in nipple aspirate fluid by CpG island hypermethylation. Clin Cancer Res. 2004;10:28–32. doi: 10.1158/1078-0432.ccr-0410-3. [DOI] [PubMed] [Google Scholar]
  • 31.Li SY, Rong M, Iacopetta B. DNA hypermethylation in breast cancer and its association with clinicopathological features. Cancer Lett. 2006;237:272–280. doi: 10.1016/j.canlet.2005.06.011. [DOI] [PubMed] [Google Scholar]
  • 32.Mehrotra J, Vali M, McVeigh M, Kominsky SL, Fackler MJ, et al. Very high frequency of hypermthylated genes in breast cancer metastasis to the bone, brain, and lung. Clin Cancer Res. 2004;10:3104–3109. doi: 10.1158/1078-0432.ccr-03-0118. [DOI] [PubMed] [Google Scholar]
  • 33.Yan PS, Perry MR, Laux DE, Asare AL, Caldwell CW, et al. CpG island arrays: an application toward deciphering epigenetic signatures of breast cancer. Clin Cancer Res. 2000;6:1432–1438. [PubMed] [Google Scholar]
  • 34.Shinozaki M, Hoon DS, Giuliano AE, Hansen NM, Wang HJ, et al. Distinct hypermethylation profile of primary breast cancer is associated with sentinel lymph node metastasis. Clin Cancer Res. 2005;11:2156–2162. doi: 10.1158/1078-0432.CCR-04-1810. [DOI] [PubMed] [Google Scholar]
  • 35.Tao MH, Shields PG, Nie J, Millen A, Ambrosone CB, et al. DNA hypermethylation and clinicopathological features in breast cancer: the Western New York Exposures and Breast Cancer (WEB) Study. Breast Cancer Res Treat. 2009;114:559–568. doi: 10.1007/s10549-008-0028-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.McCann SE, McCann WE, Hong CC, Marshall JR, Edge SB, et al. Dietary patterns related to glycemic index and load and risk of premenopausal and postmenopausal breast cancer in the Western New York Exposure and Breast Cancer Study. Am J Clin Nutr. 2007;86:465–471. doi: 10.1093/ajcn/86.2.465. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Block G, Hartman AM, Dresser CM, Carroll MD, Gannon J, et al. A data-based approach to diet questionnaire design and testing. Am J Epidemiol. 1986;124:453–469. doi: 10.1093/oxfordjournals.aje.a114416. [DOI] [PubMed] [Google Scholar]
  • 38.Gropper SS, Smith JL, Groff JL. Advanced Nutrition and Human Metabolism. 5. Belmont, CA: Wadsworth; 2009. [Google Scholar]
  • 39.Suitor CW, Bailey LB. Dietary folate equivalents: interpretation and application. J Am Diet Assoc. 2000;100:88–94. doi: 10.1016/S0002-8223(00)00027-4. [DOI] [PubMed] [Google Scholar]
  • 40.Jeronimo C, Usadel H, Henrique R, Oliveira J, Lopes C, et al. Quantitation of GSTP1 methylation in non-neoplastic prostatic tissue and organ-confined prostate adenocarcinoma. J Natl Cancer Inst. 2001;93:1747–1452. doi: 10.1093/jnci/93.22.1747. [DOI] [PubMed] [Google Scholar]
  • 41.Eads CA, Danenberg KD, Kawakami K, Saltz LB, Blake C, et al. MethyLight: a high-throughput assay to measure DNA methylation. Nucleic Acids Res. 2000;28:e32. doi: 10.1093/nar/28.8.e32. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Dobosy JR, Fu VX, Desotelle JA, Srinivasan R, Kenowski ML, et al. A methyl-deficient diet modifies histone methylation and alters Igf2 and H19 repression in the prostate. Prostate. 2008;68:1187–1195. doi: 10.1002/pros.20782. [DOI] [PubMed] [Google Scholar]
  • 43.van Engeland M, Weijenberg MP, Roemen GM, Brink M, de Bruïne AP, et al. Effects of dietary folate and alcohol intake on promoter methylation in sporadic colorectal cancer: the Netherlands cohort study on diet and cancer. Cancer Res. 2003;63:3133–3137. [PubMed] [Google Scholar]
  • 44.de Vogel S, Bongaerts BW, Wouters KA, Kester AD, Schouten LJ, et al. Associations of dietary methyl donor intake with MLH1 promoter hypermethylation and related molecular phenotypes in sporadic colorectal cancer. Carcinogenesis. 2008;29:1765–1773. doi: 10.1093/carcin/bgn074. [DOI] [PubMed] [Google Scholar]
  • 45.Slattery ML, Curtin K, Sweeney C, Levin TR, Potter J, et al. Diet and lifestyle factor associations with CpG island methylator phenotype and BRAF mutations in colon cancer. Int J Cancer. 2006;120:656–663. doi: 10.1002/ijc.22342. [DOI] [PubMed] [Google Scholar]
  • 46.Sharma S, Kelly T, Jones P. Epigenetics in Cancer. Carcinogenesis. 2010;31:27–36. doi: 10.1093/carcin/bgp220. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Dietrich M, Brown CJ, Block G. The effect of folate fortification of cereal-grain products on blood folate status, dietary folate intake, and dietary folate sources among adult non-supplement users in the United States. J Am Coll Nutr. 2005;24:266–274. doi: 10.1080/07315724.2005.10719474. [DOI] [PubMed] [Google Scholar]

RESOURCES