1,001
Views
19
CrossRef citations to date
0
Altmetric
Research Paper

Blood glutathione redox status and global methylation of peripheral blood mononuclear cell DNA in Bangladeshi adults

, , , , , , , , , , , , , & show all
Pages 730-738 | Received 11 Apr 2013, Accepted 11 May 2013, Published online: 17 May 2013

Abstract

Oxidative stress and DNA methylation are metabolically linked through the relationship between one-carbon metabolism and the transsulfuration pathway, but possible modulating effects of oxidative stress on DNA methylation have not been extensively studied in humans. Enzymes involved in DNA methylation, including DNA methyltransferases and histone deacetylases, may show altered activity under oxidized cellular conditions. Additionally, in vitro studies suggest that glutathione (GSH) depletion leads to global DNA hypomethylation, possibly through the depletion of S-adenosylmethionine (SAM). We tested the hypothesis that a more oxidized blood GSH redox status is associated with decreased global peripheral blood mononuclear cell (PBMC) DNA methylation in a sample of Bangladeshi adults. Global PBMC DNA methylation and whole blood GSH, glutathione disulfide (GSSG), and SAM concentrations were measured in 320 adults. DNA methylation was measured by using the [3H]-methyl incorporation assay; values are inversely related to global DNA methylation. Whole blood GSH redox status (Eh) was calculated using the Nernst equation. We found that a more oxidized blood GSH Eh was associated with decreased global DNA methylation (B ± SE, 271 ± 103, p = 0.009). Blood SAM and blood GSH were associated with global DNA methylation, but these relationships did not achieve statistical significance. Our findings support the hypothesis that a more oxidized blood GSH redox status is associated with decreased global methylation of PBMC DNA. Furthermore, blood SAM does not appear to mediate this association. Future research should explore mechanisms through which cellular redox might influence global DNA methylation.

Introduction

Methylation of cytosines in CpG dinucleotides is an epigenetic mechanism involved in the regulation of gene expression and cellular differentiation.Citation1,Citation2 The methyl donor for this reaction is S-adenosylmethionine (SAM); SAM synthesis is regulated by folate-dependent one-carbon metabolism, as described in . Decreased levels of global DNA methylation are associated with increased chromatin accessibility and gene transcriptional activity,Citation3 as well as genomic instability.Citation4 Consequently, global DNA hypomethylation, along with site-specific DNA hypermethylation of tumor suppressor genes, is commonly found in tumor tissue and transformed cellsCitation5 and is believed to play a role in carcinogenesis.Citation6

Figure 1. Overview of one-carbon metabolism, transsulfuration pathway, and glutathione influx/efflux. Folate is delivered into the cell via a receptor-mediated [folate receptor alpha (FRα), beta (FRβ) or gamma (FRγ)] or carrier-mediated [solute carrier family 19, member 1 (SLC19A1), also known as RFC1] transport mechanism and is subsequently reduced to dihydrofolate (DHF) and tetrahydrofolate (THF), which then enters the one-carbon metabolic pathway. THF picks up a methyl group from serine and is converted to 5-methyl THF (5mTHF). The methyl group of 5mTHF can be transferred to Hcys via methionine synthetase (MTR), generating methionine (Met) and THF. Met is activated to form S-adenosylmethionine (SAM), which serves as the methyl donor for the methylation of CpG dinucleotides and many other substrates, yielding the methylated product and S-adenosylhomocysteine (SAH), which is hydrolyzed to regenerate Hcys. Hcys can be used to regenerate Met or be directed through the transsulfuration pathway via cystathionine-β-synthase (CBS). Glutathione (GSH), a product of the transsulfuration pathway, is a tripeptide comprised of glutamate (Glu), cysteine (Cys) and glycine (Gly). As the primary endogenous antioxidant, GSH donates an electron to reactive oxygen species (e.g., H2O2) with the enzyme glutathione peroxidase (GPx) and quickly reacts with another free radical GSH to form glutathione disulfide (GSSG). GSH can be regenerated from GSSG in a reaction catalyzed by glutathione reductase (GR). Under conditions of oxidative stress, CBS activity is upregulated to direct Hcys flux through the transsulfuration pathway for GSH production. Excess intracellular Hcys can also be exported extracellularly. GSH can be exported out of the cell and catabolized via the cell membrane enzyme γ-glutamyltransferase (GGT). GGT transfers the γ-glutamyl group to an amino acid, producing cysteinylglycine (Cys-Gly), which can be broken down to Cys and Gly via dipeptidase (DP). Cys is unstable extracellularly and rapidly oxidizes to cystine (CySS). The xC antiporter can import CySS using a transmembrane Glu gradient, and the b0+ system can directly import CySS, which can be converted back to Cys to maintain the intracellular Cys pool.

Figure 1. Overview of one-carbon metabolism, transsulfuration pathway, and glutathione influx/efflux. Folate is delivered into the cell via a receptor-mediated [folate receptor alpha (FRα), beta (FRβ) or gamma (FRγ)] or carrier-mediated [solute carrier family 19, member 1 (SLC19A1), also known as RFC1] transport mechanism and is subsequently reduced to dihydrofolate (DHF) and tetrahydrofolate (THF), which then enters the one-carbon metabolic pathway. THF picks up a methyl group from serine and is converted to 5-methyl THF (5mTHF). The methyl group of 5mTHF can be transferred to Hcys via methionine synthetase (MTR), generating methionine (Met) and THF. Met is activated to form S-adenosylmethionine (SAM), which serves as the methyl donor for the methylation of CpG dinucleotides and many other substrates, yielding the methylated product and S-adenosylhomocysteine (SAH), which is hydrolyzed to regenerate Hcys. Hcys can be used to regenerate Met or be directed through the transsulfuration pathway via cystathionine-β-synthase (CBS). Glutathione (GSH), a product of the transsulfuration pathway, is a tripeptide comprised of glutamate (Glu), cysteine (Cys) and glycine (Gly). As the primary endogenous antioxidant, GSH donates an electron to reactive oxygen species (e.g., H2O2) with the enzyme glutathione peroxidase (GPx) and quickly reacts with another free radical GSH to form glutathione disulfide (GSSG). GSH can be regenerated from GSSG in a reaction catalyzed by glutathione reductase (GR). Under conditions of oxidative stress, CBS activity is upregulated to direct Hcys flux through the transsulfuration pathway for GSH production. Excess intracellular Hcys can also be exported extracellularly. GSH can be exported out of the cell and catabolized via the cell membrane enzyme γ-glutamyltransferase (GGT). GGT transfers the γ-glutamyl group to an amino acid, producing cysteinylglycine (Cys-Gly), which can be broken down to Cys and Gly via dipeptidase (DP). Cys is unstable extracellularly and rapidly oxidizes to cystine (CySS). The xC− antiporter can import CySS using a transmembrane Glu gradient, and the b0+ system can directly import CySS, which can be converted back to Cys to maintain the intracellular Cys pool.

Oxidative stress is a risk factor also commonly implicated in carcinogenesis.Citation7 The body’s primary antioxidant is glutathione (GSH), a thiol-containing tripeptide (γ-glutamyl-cysteinyl-glycine) that readily donates an electron to reactive oxygen species (ROS) via glutathione peroxidase (GPx) and quickly reacts with another free radical GSH molecule to form glutathione disulfide (GSSG).Citation8 Reduced levels of GSH, increased levels of GSSG and a decrease of the ratio of GSH to GSSG are indicative of a more oxidized intracellular environment.Citation9 Since the GSH-GSSG couple is the most abundant intracellular redox pair, their absolute concentrations can be used in the Nernst equation to estimate the intracellular redox state, Eh (in mV).Citation10 Approximately 50% of the cysteine (Cys) used in the production of GSH is derived from the conversion of homocysteine (Hcys) to cystathionine in the first step of the transsulfuration pathwayCitation11 ().

Increased oxidative stress and aberrant DNA methylation often co-occur in carcinogenesis.Citation12 Although DNA methylation and oxidative stress are metabolically linked through the relationship between one-carbon metabolism and the transsulfuration pathway, possible modulating effects of oxidative stress on DNA methylation have not been extensively studied in humans. Two mechanisms have been proposed in the literature.

First, a more oxidized cellular redox state may lead to a decrease in genomic DNA methylation through redox regulation of related enzymes. A BLAST analysis of gene sequences containing possible redox-sensitive cysteine residues identified the SAM-dependent methyltransferases as potentially redox-sensitive.Citation13 Additionally, the activity of methionine adenosyltransferase (MAT), which catalyzes the enzymatic addition of methionine to adenosine for the synthesis of SAM, is decreased in environments with a more oxidized GSH/GSSG ratio.Citation14

Second, it has been hypothesized by others that GSH depletion under conditions of chronic oxidative stress may lead to decreased global DNA methylation through the depletion of SAM.Citation15 Under oxidizing conditions, cystathionine-β-synthase (CBS) activity increases to direct Hcys flux through the transsulfuration pathway for the generation of GSH.Citation16 Consequently, less Hcys is directed toward the regeneration of methionine, which may result in decreased SAM levels. In support of this hypothesis, treatment with the GSH-depleting hepatotoxin bromobenzene led to extensive depletion of liver methionine pools and hypomethylation of genomic liver DNA in Syrian hamsters.Citation17,Citation18

We tested the hypothesis that increased oxidative stress is associated with decreased global DNA methylation in a cohort of Bangladeshi adults by examining the associations of blood GSH concentrations and blood GSH redox state (Eh) with global methylation of peripheral blood mononuclear cell (PBMC) DNA. To evaluate whether the association of blood GSH redox state with global PBMC DNA methylation might be due to depletion of SAM, we also examined the associations of blood SAM with the blood GSH redox variables and DNA methylation to test for possible mediation.

Results

Demographic and clinical characteristics of the study participants are shown in . The average age was 43 y, and there were roughly equal numbers of males and females. Approximately 35% of the population was classified as underweight (BMI < 18.5 kg/m2), and 31.3% of the participants were folate deficient (plasma folate < 9 nmol/L). Due to the Folate and Oxidative Stress (FOX) study sampling design, nearly 71% of the participants used wells with water arsenic (As) > 50 ug/L (the Bangladeshi standard) as their primary drinking source. Mean blood GSH and GSSG concentrations were 491.0 and 37.3 μmol/L, respectively, and the mean blood GSH Eh was −198 mV.

Table 1. Descriptive characteristics for study sample (n = 320)

In our evaluation of potential covariates for inclusion in the regression models, we observed that water As was negatively correlated with blood GSH (Spearman r = −0.15, p = 0.008), as reported in an earlier study from our group.Citation19 Males had higher blood GSH concentrations, less oxidized blood GSH Eh values, and higher blood SAM levels than females (p < 0.0001). Ever-smokers also had higher GSH concentrations, less oxidized blood GSH Eh values, and higher SAM levels than never-smokers. Since most ever-smokers were male (113/123, 91.9%), we stratified the ever-smoking analyses by gender: No significant associations were found between GSH and smoking status after stratification, indicating that the observed differences in GSH, GSH Eh and SAM by smoking status were due to the preponderance of males in the ever-smoking group (data not shown).

To support the validity of whole blood GSH Eh as a marker of intracellular redox in blood cells, we examined relationships of blood GSH redox variables with plasma Hcys and blood S-adenosylhomocysteine (SAH), shown in . Cellular oxidative stress has been shown experimentally to decrease plasma Hcys in vitroCitation20 and in vivo.Citation21 As expected, a more oxidized blood GSH Eh was negatively correlated with both plasma Hcys (Spearman r = −0.27, p < 0.0001) and blood SAH (Spearman r = −0.42, p < 0.0001). We did not find that plasma GSH Eh was correlated with plasma Hcys (Spearman r = 0.07, p = 0.22) or blood SAH (Spearman r = 0.02, p = 0.74).

Table 2. Spearman correlation coefficients of blood and plasma glutathione redox variables with plasma Hcys and blood SAH (n = 320)

A more oxidized blood GSH Eh was significantly correlated with increased [3H]-methyl incorporation in bivariate analyses (Spearman r = 0.17, p = 0.002). Since [3H]-methyl incorporation is inversely related to genomic DNA methylation, this indicated that a more oxidized blood GSH Eh is associated with reduced global DNA methylation. In covariate-adjusted linear regression models shown in 1 mV increase in blood GSH Eh was associated with an increase in [3H]-methyl incorporation of 271 disintegrations per minute (DPM) per μg DNA (p = 0.009). Similarly, a 1 μM decrease in blood GSH was associated with an increase in [3H]-methyl incorporation of 12.7 DPM per μg DNA, but this association was not statistically significant (p = 0.11).

Table 3. Unadjusted and adjusted regression coefficients for associations between predictors and [3H]-methyl incorporation of PBMC DNA (n = 320)

For statistical evidence to support blood SAM as a mediator of the association between blood GSH redox and global DNA methylation, we examined associations of blood SAM with blood GSH, blood GSH Eh, and [3H]-methyl incorporation. Blood SAM was negatively associated with [3H]-methyl incorporation at borderline significance (B ± SE = −4803 ± 2747, p = 0.08, ). Blood GSH concentration was a significant predictor of log-transformed blood SAM in a covariate-adjusted model (B ± SE = 0.026 ± 0.012 for a 100 μM increase in blood GSH, p = 0.04), but blood GSH Eh was not associated with blood SAM (p = 0.23). Since the association of blood GSH and [3H]-methyl incorporation did not reach statistical significance, and since blood GSH Eh and blood SAM were not associated, the result did not support the role of blood SAM as a mediator.

In exploratory analyses, we also examined associations of other variables with [3H]-methyl incorporation. Neither blood SAH nor the blood SAM/SAH ratio were significantly associated with [3H]-methyl incorporation in adjusted models (p = 0.65 and p = 0.79, respectively). An oxidized blood GSH Eh correlated positively with urinary 8-oxodG levels, per g urinary creatinine (Cr) (Spearman r = 0.17, p = 0.004); since oxidative damage of guanines in CpG dinucleotides has been hypothesized to decrease global DNA methylation,Citation22,Citation23 we also examined the association between urinary 8-oxodG/Cr levels and [3H]-methyl incorporation. However, the covariate-adjusted association was not significant (B ± SE = 293 ± 207, p = 0.16).

Discussion

The objective of this study was to examine whether oxidative stress is associated with decreased global methylation of PBMC DNA in a cross-sectional study of Bangladeshi adults. We tested this hypothesis by examining the associations of blood GSH and blood GSH Eh with PBMC DNA methylation, and we also examined whether these associations were mediated by blood SAM levels. We observed that a more oxidized blood GSH Eh was associated with decreased DNA methylation. Decreased blood GSH was not significantly associated with decreased DNA methylation, indicating that the association of blood GSH Eh with global DNA methylation depended on the concentrations and balance of blood GSH and GSSG and was not simply explained by the concentration of blood GSH. Finally, although blood SAM was positively associated with PBMC DNA methylation at marginal significance, blood GSH Eh was not associated with blood SAM. Taken together, our observations are consistent with the hypothesis that an oxidized intracellular redox state, as measured by blood GSH Eh, is associated with decreased global DNA methylation, but blood SAM does not appear to mediate this association.

Only a few other epidemiologic studies have examined the association between oxidative stress and DNA methylation. A study of 45 infertile men found that seminal ROS production was negatively correlated with sperm global DNA methylation, and three-month supplementation with antioxidants produced a significant decrease in seminal ROS and significant increase in sperm global DNA methylation.Citation24 Additionally, in a study of 61 bladder cancer patients and 45 healthy controls, total antioxidant status in urine was positively correlated with methylation of long interspersed nuclear element-1 (LINE-1) in peripheral blood leukocyte (PBL) DNA from all subjects.Citation25 To our knowledge, this is the first study to find an association between blood glutathione redox status and global DNA methylation in a generally healthy population.

While methylation of DNA is known to be a dynamic process, a mechanism for demethylation has not been clearly established. Recent work has shown that the ten-eleven translocation (Tet) family of proteins can convert 5-methylcytosine (5-mC) into 5-hydroxymethylcytosine (5-hmC),Citation26,Citation27 which is believed to be an intermediary in passive and/or active DNA demethylation.Citation28,Citation29 Interestingly, Chen et al. found that the de novo methyltransferases DNMT3a and DNMT3b act as both DNA methyltransferases and DNA dehydroxymethylases, depending upon the redox environment: treatment with dithiothreitol (DTT) or hydrogen peroxide (H2O2) inhibited and enhanced the dehydroxymethylation activities of the enzymes, respectively.Citation30 Chen et al. speculated that the dual methylation and dehydroxymethylation activities of the DNMT3 enzymes might explain the paradoxical observation that DMNT3 expression is upregulated and gene-specific DNA methylation is increased in cancer cells, yet global DNA methylation is decreased.Citation30

Histone deacetylase (HDAC) inhibition has been shown experimentally to induce DNA hypomethylation,Citation31 and it has been hypothesized that HDAC activities might also be influenced directly by cellular redox.Citation32 In human neuroblastoma cells, exposure to H2O2 induces global DNA hypomethylation, along with downregulation of HDAC3, DNMT1 and DNMT3a expression.Citation33 One potential mechanism of redox inhibition of HDACs might occur through alteration of cysteine residues: Reactive carbonyl species inactivate HDAC1 through covalent modification of conserved cysteine residues,Citation34 and exposure to ROS-producing compounds induces intramolecular disulfide bond formation in conserved cysteine residues in HDAC4.Citation35

We did not find that blood SAM was significantly associated with global DNA methylation or blood GSH Eh. Mathematical modeling of one-carbon metabolism predicts that the DNMT methylation rate is remarkably stable over wide fluctuations of SAM concentrations, primarily due to the low Km for the DNMT reaction and long-range allosteric regulation of CBS and methylenetetrahydrofolate reductase (MTHFR) by SAM.Citation36 Furthermore, the hypothesis that increased demand for GSH production depletes SAM levels is complicated by the discovery that CBS is allosterically activated by SAM but destabilized under SAM-deficient conditions.Citation37 These findings, along with the observations in this study, do not lend strong support to the hypothesis that oxidative stress leads to DNA hypomethylation through SAM depletion.

It is not clear whether whole blood concentrations of SAM and GSH reflect intracellular concentrations in other tissues. Although in vitro evidence suggests that SAM is not easily transported across cell membranes,Citation38,Citation39 in a cohort of healthy adults, the plasma SAM/SAH ratio was strongly correlated with the intracellular lymphocyte SAM/SAH ratio (r = 0.73).Citation40 A study in rats showed that an increase in ROS in liver initiated changes in the erythrocytic GSH/GSSG ratio; the authors concluded that the erythrocytic GSH/GSSG ratio reflects oxidative stress in liver and other tissues.Citation41 Additionally, an intact transsulfuration pathway has been identified in blood cells, including macrophagesCitation42 and T cells,Citation43 and transsulfuration pathway flux is upregulated in naïve and activated T cells after exposure to peroxide.Citation43 Even if SAM and GSH measurements in blood do not directly reflect the absolute concentrations in other tissues, the presence of intact one-carbon metabolism and transsulfuration pathways in blood cells suggests that blood biomarkers might be appropriate to study perturbations of these pathways by oxidative stress.

We observed that an oxidized blood GSH Eh was correlated with decreased blood SAH and plasma Hcys, which is consistent with an increase in Hcys flux through the transsulfuration pathway under oxidized intracellular conditions.Citation16 Although some studies have used plasma GSH Eh as a marker of intracellular redox,Citation44,Citation45 we did not observe significant correlations of Hcys or SAH with plasma GSH Eh, which suggests that whole blood GSH Eh might be a better indicator of the intracellular redox environment, at least within blood cells. However, an oxidized plasma GSH Eh is thought to reflect oxidative stress in other tissuesCitation46 or systemic oxidative stressCitation47 and is observed in aging,Citation48 obesityCitation49 and disease states such as asthmaCitation50 and heart disease.Citation51 Indeed, we found that plasma GSH Eh, but not blood GSH Eh, was positively correlated with age (Spearman r = 0.14, p = 0.01) and BMI (Spearman r = 0.11, p = 0.05). As such, the measurements of GSH redox pairs in both plasma and whole blood might be informative to fully understand the influence of “oxidative stress” in an organism.

Redox regulation is known to influence white blood cell cycle progression and proliferation,Citation52 and thus, it is possible that our observations are explained by redox-induced shifts in PBMC cell type distributions or counts. However, the overall level of global DNA methylation across cell types would have to vary dramatically for cell distribution shifts to explain our observations. While site-specific CpG methylation levels at certain loci have been found to be associated with blood cell type,Citation53 global DNA methylation levels were not significantly different among DNA samples from granulocytes, mononuclear cells, and white blood cells in adult women.Citation54 Furthermore, since PBMCs are a subset of total WBCs, the influence of cell type variability is reduced to some extent in our study.

Our study has several limitations. First, the cross-sectional design precludes us from establishing the directionality of the relationship between blood GSH redox state and global DNA methylation. Epigenetic alterations might induce changes in the cellular redox status: for example, mice exposed to d-β-hydroxybutyrate, an inhibitor of Class I HDACs, exhibit an increase in global H3 acetylation, increase in expression of oxidative stress resistance genes and a reduction in oxidative damage from ROS.Citation55 We also cannot establish whether the observed association with global DNA methylation is explained by an upstream factor (e.g., H2O2), the GSH redox state itself or another redox-associated event, e.g., alterations of the citric acid cycle and/or the NAD+/NADH ratio.Citation32 Finally, we were unable to adjust for PBMC cell type counts or distributions in our regression models.

In conclusion, we observed that a more oxidized intracellular redox state, as measured by blood GSH Eh, was associated with a decrease in global methylation of PBMC DNA. Furthermore, blood SAM was marginally associated with an increase in global DNA methylation, but it did not mediate the association between blood GSH Eh and global DNA methylation. Future research should examine whether alteration of intracellular redox is a mechanism through which environmental exposures and other factors influence DNA methylation. It might also be of interest to examine whether redox therapies might help to prevent progressive loss of global DNA methylation in cancer and other diseases.

Materials and Methods

Eligibility criteria and study design

The FOX study was initially designed to examine the dose-response relationship between arsenic exposure and markers of oxidative stress. In the FOX study, we recruited 379 men and women aged 35–65 y between April 2007 and April 2008 in Araihazar, Bangladesh. Participants were selected based on well water As (wAs) exposure such that the final study sample represented the full range of wAs concentrations in the region, as described previously.Citation19 Participants were excluded if they were pregnant and/or planned to become pregnant within 3 mo, were currently taking nutritional supplements (within the past 3 mo) or had known diabetes, cardiovascular or renal disease or other diseases known to be associated with oxidative stress.

Oral informed consent was obtained by our Bangladeshi field staff physicians, who read an approved assent form to the study participants. This study was approved by the institutional review boards of the Bangladesh Medical Research Council and Columbia University Medical Center.

Analytic techniques

Sample collection and handling

Blood and urine samples were collected and immediately processed at our field clinic in Araihazar. Blood samples were centrifuged at 3,000 × g for 10 min at 4 °C, and red cells were separated from buffy coat and plasma. Blood and plasma aliquots were stored in freezers at −80 °C. Processing for GSH and GSSG assays was performed on freshly drawn blood and plasma samples prior to storage in −80 °C freezers. Urine samples were collected in 50 mL acid-washed polypropylene tubes and frozen at −20 °C. All blood, plasma and urine samples were transported to Dhaka on dry ice and stored at −80 °C (blood and plasma) and −20 °C (urine). Samples were then packed on dry ice and flown to Columbia University for analysis.

Whole blood and plasma glutathione and glutathione disulfide

Whole blood and plasma GSH and GSSG were assayed by the method of Jones et al.,Citation56 as previously described.Citation19 Briefly, blood samples were collected in the field laboratory with a butterfly syringe, immediately processed for derivatization, and stored at − 80 °C until delivered to Columbia University for analysis. Twenty μl of each sample was injected onto the HPLC, and metabolites were detected using a Waters 474 scanning fluorescence detector with 335 nm excitation and 515 nm emission (Waters Corp.). Intra-assay CVs were between 5 and 10%, and inter-assay CVs were between 11 and 18%.

Plasma folate and vitamin B12

Plasma folate and vitamin B12 were analyzed using a radioproteinbinding assay according to the manufacturer’s protocol (SimulTRAC-S, MP Biomedicals). This method requires heating the samples to 100 °C to denature endogenous binding substances. To determine folate concentrations, folic acid as pteroylglutamic acid was used for calibration, and its 125I-labeled analog was used as the tracer; for vitamin B12, cyanocobalamin was used for calibration, and its 57Co-labeled analog was used as the tracer. The intra- and inter-assay coefficients of variation for folate were 6% and 14%, respectively, and the intra- and inter-assay coefficients of variation for vitamin B12 were 9% and 4%, respectively.

Plasma total homocysteine

Plasma total Hcys was assayed using the method of Pfieffer et al.,Citation57 as previously described.Citation58 The intra- and inter-assay CVs for plasma total Hcys were 4 and 9%, respectively.

Blood S-adenosylmethionine and S-adenosylhomocysteine

SAM and SAH were measured in whole blood, as previously described.Citation59 Briefly, blood was thawed and vortexed, and 400 μl were added to an equal volume of 0.1 M sodium acetate, pH 6.0, plus 40% TCA. SAM and SAH were detected at 254 nm using a 996 Photodiode Array UV absorbance detector (Waters Inc.) and quantified relative to standard curves generated using purified compounds (Sigma). The inter-assay coefficient of variation was 9.6% for SAM and 16.1% for SAH.

Isolation of PBMC DNA

PBMC DNA was isolated from 4 mL PBMC lysate using 1ml Protein Precipitation Solution (5-Prime), and standard isopropanol extraction was conducted using the manufacturer’s protocol. DNA samples were then frozen at −20 °C until further analysis.

Genomic DNA methylation

Genomic DNA methylation was measured using the [3H]-methyl incorporation assay of Balaghi and Wagner,Citation60 as previously described.Citation61 The assay employs 3H-labeled SAM and SssI methylase to add 3H-labeled methyl groups to unmethylated CpG sequences. Thus, the DP) values are inversely related to global DNA methylation. PicoGreen dsDNA Quantitation Reagent (Molecular Probes) was used to quantify the amount of double-stranded DNA in each reaction. DPM values were expressed per μg DNA. The intra- and inter-assay coefficients of variation were 3.4% and 10.4%, respectively.

Urinary 8-oxodG

Urinary 8-oxodG was measured using the “New 8-OHdG Check” ELISA kit at the Genox Corporation’s laboratory. Samples were measured in triplicate. The detection level for this assay was 0.64 ng/mL urine. A significant proportion of the study sample (n = 56 individuals, or 18%) had urinary 8-oxodG levels below the detection limit; levels for these individuals were set to one-half of the detection limit, or 0.32 ng/mL. The intra- and inter-assay coefficients of variation were 7.5% and 7.7%, respectively. Urinary Cr was analyzed with a colorimetric assay based on the Jaffe reactionCitation62 to adjust for urine concentration.

Water As

Procedures for field sample collection and laboratory analyses were described previously.Citation63,Citation64 Briefly, water samples were collected in 20-mL polyethylene scintillation vials and acidified to 1% with high-purity Optima HCl (Fisher Scientific) at least 48 h before analysis.Citation65 Water samples were analyzed by high-resolution inductively coupled plasma mass spectrometry after 1:10 dilution and addition of a Ge spike to correct fluctuations in instrument sensitivity. The detection limit of the method is <0.2 μg/L. A standard with an As concentration of 51 μg/L was run multiple times in each batch. The intra- and inter-assay coefficients of variation (CVs) for this standard were 6.0% and 3.8%, respectively.

Calculation of the reduction potential (Eh)

The reduction potentials of the blood and plasma GSH/GSSG redox pairs were calculated using the Nernst equation, Eh = Eo + (RT/nF)ln[(acceptor)/(donor)], where Eo is the standard potential for the redox couple at the defined pH, R is the gas constant, T is the absolute temperature, F is Faraday's constant, and n is the number of electrons transferred.Citation46 For GSH and GSSG, the equation simplifies to Eh (mV) = −264 – 30 log[(GSH)Citation2/(GSSG)], where (GSH) and (GSSG) are molar concentrations and the Eo value assumes a physiologic pH of 7.4.Citation46 A more positive Eh value reflects a more oxidized state.

Statistical methods

Descriptive statistics (means and standard deviations) were calculated for the characteristics of the sample. Spearman correlations were used to examine bivariate associations between blood GSH variables, [3H]-methyl incorporation, and other continuous covariates. Wilcoxon rank-sum test was used to detect differences in continuous measures between categories of binary variables.

Linear regression analyses were constructed with blood GSH redox variables and blood SAM as predictors and [3H]-methyl incorporation as the outcome, with and without controlling for confounding factors. Proper transformations were applied to the continuous independent variables with skewed distribution to reduce the impact of extreme values. Certain confounders (gender, age, cigarette smoking and water As) were selected a priori based on biologic plausibility and previous studies in the literature.Citation61 Other potential confounders (plasma folate, plasma Hcys, plasma vitamin B12, betel nut chewing, BMI, television ownership, education) were considered by examining their bivariate associations with blood GSH redox variables, blood SAM and [3H]-methyl incorporation. The control variables in the final models were those that were related to the outcome and main predictors and resulted in an appreciable (>5%) change in the regression coefficient for the association between a predictor and the outcome. Since cigarette smoking did not predict [3H]-methyl incorporation independently of gender, it was excluded from the final model. The final regression models contained gender, age, and water As as covariates.

Blood SAM was considered to be a partial mediator of the association between blood GSH and [3H]-methyl incorporation if it met the following criteria: (1) Blood GSH (predictor) was significantly associated with blood SAM (mediator); (2) Blood GSH was significantly associated with [3H]-methyl incorporation (outcome); (3) Blood SAM was significantly associated with [3H]-methyl incorporation; (4) inclusion of blood SAM in a model of blood GSH predicting [3H]-methyl incorporation resulted in attenuation of the regression coefficient.

Values for [3H]-methyl incorporation were excluded from the analysis if duplicates had coefficients of variation >15% (n = 48), DNA assay inputs below 10 μg/ml (n = 8), or values that were extreme outliers, defined as values that exceeded the boxplot of the DPM values by more than three interquartile ranges (n = 2). Additionally, we were unable to isolate DNA from n = 1 lysate, resulting in a final sample size of 320. All statistical analyses were conducted using SAS (version 9.3; SAS Institute Inc.).

Abbreviations:
8-oxodG=

urinary 8-oxo-2’deoxyguanosine

As=

arsenic

BMI=

body mass index

CBS=

cystathionine-β-synthase

uCr=

urinary creatinine

CV=

coefficient of variation

Cys=

cysteine

DPM=

disintegrations per minute

DTT=

dithiothreitol

Eh=

reduction potential of redox couple, in millivolts

FOX=

folate and oxidative stress study

GFAA=

graphite furnace atomic absorption

GSH=

glutathione

GSSG=

glutathione disulfide

H2O2=

hydrogen peroxide

Hcys=

homocysteine

ICP-MS=

inductively coupled mass spectrometry

LINE-1=

long interspersed nuclear element-1

MAT=

methionine adenosyltransferase

MTHFR=

methylenetetrahydrofolate reductase

PBL=

peripheral blood leukocyte

PBMC=

peripheral blood mononuclear cell

SAH=

S-adenosylhomocysteine

SAM=

S-adenosylmethionine

wAs=

water As

Acknowledgments

This work was supported by grants RO1 CA133595, RO1 ES017875, P42 ES10349, P30 ES09089, 5-R25-CA 094061 and T32 CA009529-24 from the National Institutes of Health. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of Environmental Health Sciences, the National Cancer Institute or the National Institutes of Health. We wish to thank Dr. Frederick Domann for his help with the preparation of this manuscript.

Disclosure of Potential Conflicts of Interest

No potential conflicts of interest were disclosed.

References

  • Jones PA, Taylor SM. Cellular differentiation, cytidine analogs and DNA methylation. Cell 1980; 20:85 - 93; http://dx.doi.org/10.1016/0092-8674(80)90237-8; PMID: 6156004
  • Razin A, Riggs AD. DNA methylation and gene function. Science 1980; 210:604 - 10; http://dx.doi.org/10.1126/science.6254144; PMID: 6254144
  • Thurman RE, Rynes E, Humbert R, Vierstra J, Maurano MT, Haugen E, et al. The accessible chromatin landscape of the human genome. Nature 2012; 489:75 - 82; http://dx.doi.org/10.1038/nature11232; PMID: 22955617
  • Lengauer C, Kinzler KW, Vogelstein B. DNA methylation and genetic instability in colorectal cancer cells. Proc Natl Acad Sci U S A 1997; 94:2545 - 50; http://dx.doi.org/10.1073/pnas.94.6.2545; PMID: 9122232
  • Das PM, Singal R. DNA methylation and cancer. J Clin Oncol 2004; 22:4632 - 42; http://dx.doi.org/10.1200/JCO.2004.07.151; PMID: 15542813
  • Robertson KD, Jones PA. DNA methylation: past, present and future directions. Carcinogenesis 2000; 21:461 - 7; http://dx.doi.org/10.1093/carcin/21.3.461; PMID: 10688866
  • Klaunig JE, Kamendulis LM. The role of oxidative stress in carcinogenesis. Annu Rev Pharmacol Toxicol 2004; 44:239 - 67; http://dx.doi.org/10.1146/annurev.pharmtox.44.101802.121851; PMID: 14744246
  • Forman HJ, Zhang H, Rinna A. Glutathione: overview of its protective roles, measurement, and biosynthesis. Mol Aspects Med 2009; 30:1 - 12; http://dx.doi.org/10.1016/j.mam.2008.08.006; PMID: 18796312
  • Jones DP. Radical-free biology of oxidative stress. Am J Physiol Cell Physiol 2008; 295:C849 - 68; http://dx.doi.org/10.1152/ajpcell.00283.2008; PMID: 18684987
  • Schafer FQ, Buettner GR. Redox environment of the cell as viewed through the redox state of the glutathione disulfide/glutathione couple. Free Radic Biol Med 2001; 30:1191 - 212; http://dx.doi.org/10.1016/S0891-5849(01)00480-4; PMID: 11368918
  • Beatty PW, Reed DJ. Involvement of the cystathionine pathway in the biosynthesis of glutathione by isolated rat hepatocytes. Arch Biochem Biophys 1980; 204:80 - 7; http://dx.doi.org/10.1016/0003-9861(80)90009-0; PMID: 7425648
  • Franco R, Schoneveld O, Georgakilas AG, Panayiotidis MI. Oxidative stress, DNA methylation and carcinogenesis. Cancer Lett 2008; 266:6 - 11; http://dx.doi.org/10.1016/j.canlet.2008.02.026; PMID: 18372104
  • Fomenko DE, Xing W, Adair BM, Thomas DJ, Gladyshev VN. High-throughput identification of catalytic redox-active cysteine residues. Science 2007; 315:387 - 9; http://dx.doi.org/10.1126/science.1133114; PMID: 17234949
  • Pajares MA, Durán C, Corrales F, Pliego MM, Mato JM. Modulation of rat liver S-adenosylmethionine synthetase activity by glutathione. J Biol Chem 1992; 267:17598 - 605; PMID: 1517209
  • Hitchler MJ, Domann FE. An epigenetic perspective on the free radical theory of development. Free Radic Biol Med 2007; 43:1023 - 36; http://dx.doi.org/10.1016/j.freeradbiomed.2007.06.027; PMID: 17761298
  • Mosharov E, Cranford MR, Banerjee R. The quantitatively important relationship between homocysteine metabolism and glutathione synthesis by the transsulfuration pathway and its regulation by redox changes. Biochemistry 2000; 39:13005 - 11; http://dx.doi.org/10.1021/bi001088w; PMID: 11041866
  • Lertratanangkoon K, Wu CJ, Savaraj N, Thomas ML. Alterations of DNA methylation by glutathione depletion. Cancer Lett 1997; 120:149 - 56; http://dx.doi.org/10.1016/S0304-3835(97)00300-5; PMID: 9461031
  • Lertratanangkoon K, Orkiszewski RS, Scimeca JM. Methyl-donor deficiency due to chemically induced glutathione depletion. Cancer Res 1996; 56:995 - 1005; PMID: 8640792
  • Hall MN, Niedzwiecki M, Liu X, Harper KN, Alam S, Slavkovich V, et al. Chronic arsenic exposure and blood glutathione and glutathione disulfide concentrations in Bangladeshi adults. Environ Health Perspect 2013; In press
  • Hultberg M, Hultberg B. Oxidative stress decreases extracellular homocysteine concentration in human hepatoma (HepG2) cell cultures. Chem Biol Interact 2007; 165:54 - 8; http://dx.doi.org/10.1016/j.cbi.2006.10.009; PMID: 17141750
  • Rogers EJ, Chen S, Chan A. Folate deficiency and plasma homocysteine during increased oxidative stress. N Engl J Med 2007; 357:421 - 2; http://dx.doi.org/10.1056/NEJMc066569; PMID: 17652662
  • Weitzman SA, Turk PW, Milkowski DH, Kozlowski K. Free radical adducts induce alterations in DNA cytosine methylation. Proc Natl Acad Sci U S A 1994; 91:1261 - 4; http://dx.doi.org/10.1073/pnas.91.4.1261; PMID: 8108398
  • Valinluck V, Tsai HH, Rogstad DK, Burdzy A, Bird A, Sowers LC. Oxidative damage to methyl-CpG sequences inhibits the binding of the methyl-CpG binding domain (MBD) of methyl-CpG binding protein 2 (MeCP2). Nucleic Acids Res 2004; 32:4100 - 8; http://dx.doi.org/10.1093/nar/gkh739; PMID: 15302911
  • Tunc O, Tremellen K. Oxidative DNA damage impairs global sperm DNA methylation in infertile men. J Assist Reprod Genet 2009; 26:537 - 44; http://dx.doi.org/10.1007/s10815-009-9346-2; PMID: 19876730
  • Patchsung M, Boonla C, Amnattrakul P, Dissayabutra T, Mutirangura A, Tosukhowong P. Long interspersed nuclear element-1 hypomethylation and oxidative stress: correlation and bladder cancer diagnostic potential. PLoS One 2012; 7:e37009; http://dx.doi.org/10.1371/journal.pone.0037009; PMID: 22615872
  • Ito S, D’Alessio AC, Taranova OV, Hong K, Sowers LC, Zhang Y. Role of Tet proteins in 5mC to 5hmC conversion, ES-cell self-renewal and inner cell mass specification. Nature 2010; 466:1129 - 33; http://dx.doi.org/10.1038/nature09303; PMID: 20639862
  • Tahiliani M, Koh KP, Shen Y, Pastor WA, Bandukwala H, Brudno Y, et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 2009; 324:930 - 5; http://dx.doi.org/10.1126/science.1170116; PMID: 19372391
  • Guo JU, Su Y, Zhong C, Ming GL, Song H. Hydroxylation of 5-methylcytosine by TET1 promotes active DNA demethylation in the adult brain. Cell 2011; 145:423 - 34; http://dx.doi.org/10.1016/j.cell.2011.03.022; PMID: 21496894
  • Wu SC, Zhang Y. Active DNA demethylation: many roads lead to Rome. Nat Rev Mol Cell Biol 2010; 11:607 - 20; http://dx.doi.org/10.1038/nrm2950; PMID: 20683471
  • Chen CC, Wang KY, Shen CK. The mammalian de novo DNA methyltransferases DNMT3A and DNMT3B are also DNA 5-hydroxymethylcytosine dehydroxymethylases. J Biol Chem 2012; 287:33116 - 21; http://dx.doi.org/10.1074/jbc.C112.406975; PMID: 22898819
  • Selker EU. Trichostatin A causes selective loss of DNA methylation in Neurospora. Proc Natl Acad Sci U S A 1998; 95:9430 - 5; http://dx.doi.org/10.1073/pnas.95.16.9430; PMID: 9689097
  • Cyr AR, Domann FE. The redox basis of epigenetic modifications: from mechanisms to functional consequences. Antioxid Redox Signal 2011; 15:551 - 89; http://dx.doi.org/10.1089/ars.2010.3492; PMID: 20919933
  • Gu X, Sun J, Li S, Wu X, Li L. Oxidative stress induces DNA demethylation and histone acetylation in SH-SY5Y cells: potential epigenetic mechanisms in gene transcription in Aβ production. Neurobiol Aging 2013; 34:1069 - 79; http://dx.doi.org/10.1016/j.neurobiolaging.2012.10.013; PMID: 23141413
  • Doyle K, Fitzpatrick FA. Redox signaling, alkylation (carbonylation) of conserved cysteines inactivates class I histone deacetylases 1, 2, and 3 and antagonizes their transcriptional repressor function. J Biol Chem 2010; 285:17417 - 24; http://dx.doi.org/10.1074/jbc.M109.089250; PMID: 20385560
  • Ago T, Liu T, Zhai P, Chen W, Li H, Molkentin JD, et al. A redox-dependent pathway for regulating class II HDACs and cardiac hypertrophy. Cell 2008; 133:978 - 93; http://dx.doi.org/10.1016/j.cell.2008.04.041; PMID: 18555775
  • Nijhout HF, Reed MC, Anderson DF, Mattingly JC, James SJ, Ulrich CM. Long-range allosteric interactions between the folate and methionine cycles stabilize DNA methylation reaction rate. Epigenetics 2006; 1:81 - 7; http://dx.doi.org/10.4161/epi.1.2.2677; PMID: 17998813
  • Prudova A, Bauman Z, Braun A, Vitvitsky V, Lu SC, Banerjee R. S-adenosylmethionine stabilizes cystathionine beta-synthase and modulates redox capacity. Proc Natl Acad Sci U S A 2006; 103:6489 - 94; http://dx.doi.org/10.1073/pnas.0509531103; PMID: 16614071
  • Van Phi L, Söling HD. Methyl group transfer from exogenous S-adenosylmethionine on to plasma-membrane phospholipids without cellular uptake in isolated hepatocytes. Biochem J 1982; 206:481 - 7; PMID: 6184050
  • McMillan JM, Walle UK, Walle T. S-adenosyl-L-methionine: transcellular transport and uptake by Caco-2 cells and hepatocytes. J Pharm Pharmacol 2005; 57:599 - 605; http://dx.doi.org/10.1211/0022357056082; PMID: 15901349
  • Melnyk S, Pogribna M, Pogribny IP, Yi P, James SJ. Measurement of plasma and intracellular S-adenosylmethionine and S-adenosylhomocysteine utilizing coulometric electrochemical detection: alterations with plasma homocysteine and pyridoxal 5′-phosphate concentrations. Clin Chem 2000; 46:265 - 72; PMID: 10657384
  • Siems W, Mueller M, Garbe S, Gerber G. Damage of erythrocytes by activated oxygen generated in hypoxic rat liver. Free Radic Res Commun 1987; 4:31 - 9; http://dx.doi.org/10.3109/10715768709088086; PMID: 3506895
  • Garg S, Vitvitsky V, Gendelman HE, Banerjee R. Monocyte differentiation, activation, and mycobacterial killing are linked to transsulfuration-dependent redox metabolism. J Biol Chem 2006; 281:38712 - 20; http://dx.doi.org/10.1074/jbc.M606235200; PMID: 17046819
  • Garg SK, Yan Z, Vitvitsky V, Banerjee R. Differential dependence on cysteine from transsulfuration versus transport during T cell activation. Antioxid Redox Signal 2011; 15:39 - 47; http://dx.doi.org/10.1089/ars.2010.3496; PMID: 20673163
  • Mannery YO, Ziegler TR, Park Y, Jones DP. Oxidation of plasma cysteine/cystine and GSH/GSSG redox potentials by acetaminophen and sulfur amino acid insufficiency in humans. J Pharmacol Exp Ther 2010; 333:939 - 47; http://dx.doi.org/10.1124/jpet.110.166421; PMID: 20207721
  • Ashfaq S, Abramson JL, Jones DP, Rhodes SD, Weintraub WS, Hooper WC, et al. The relationship between plasma levels of oxidized and reduced thiols and early atherosclerosis in healthy adults. J Am Coll Cardiol 2006; 47:1005 - 11; http://dx.doi.org/10.1016/j.jacc.2005.09.063; PMID: 16516085
  • Jones DP. Redox potential of GSH/GSSG couple: assay and biological significance. Methods Enzymol 2002; 348:93 - 112; http://dx.doi.org/10.1016/S0076-6879(02)48630-2; PMID: 11885298
  • Calabrese V, Cornelius C, Leso V, Trovato-Salinaro A, Ventimiglia B, Cavallaro M, et al. Oxidative stress, glutathione status, sirtuin and cellular stress response in type 2 diabetes. Biochim Biophys Acta 2012; 1822:729 - 36; http://dx.doi.org/10.1016/j.bbadis.2011.12.003; PMID: 22186191
  • Samiec PS, Drews-Botsch C, Flagg EW, Kurtz JC, Sternberg P Jr., Reed RL, et al. Glutathione in human plasma: decline in association with aging, age-related macular degeneration, and diabetes. Free Radic Biol Med 1998; 24:699 - 704; http://dx.doi.org/10.1016/S0891-5849(97)00286-4; PMID: 9586798
  • João Cabrera E, Valezi AC, Delfino VD, Lavado EL, Barbosa DS. Reduction in plasma levels of inflammatory and oxidative stress indicators after Roux-en-Y gastric bypass. Obes Surg 2010; 20:42 - 9; http://dx.doi.org/10.1007/s11695-009-9988-2; PMID: 19826889
  • Fitzpatrick AM, Stephenson ST, Hadley GR, Burwell L, Penugonda M, Simon DM, et al. Thiol redox disturbances in children with severe asthma are associated with posttranslational modification of the transcription factor nuclear factor (erythroid-derived 2)-like 2. J Allergy Clin Immunol 2011; 127:1604 - 11; http://dx.doi.org/10.1016/j.jaci.2011.03.031; PMID: 21514635
  • Huang YS, Zhi YF, Kong SY, Wang QL, Xu BS, Wang SR. [Plasma glutathione of patients with coronary heart disease measured by fluorospectrophotometer]. Guang Pu Xue Yu Guang Pu Fen Xi 2006; 26:936 - 40; PMID: 16883873
  • Iwata S, Hori T, Sato N, Ueda-Taniguchi Y, Yamabe T, Nakamura H, et al. Thiol-mediated redox regulation of lymphocyte proliferation. Possible involvement of adult T cell leukemia-derived factor and glutathione in transferrin receptor expression. J Immunol 1994; 152:5633 - 42; PMID: 8207197
  • Talens RP, Boomsma DI, Tobi EW, Kremer D, Jukema JW, Willemsen G, et al. Variation, patterns, and temporal stability of DNA methylation: considerations for epigenetic epidemiology. FASEB J 2010; 24:3135 - 44; http://dx.doi.org/10.1096/fj.09-150490; PMID: 20385621
  • Wu HC, Delgado-Cruzata L, Flom JD, Kappil M, Ferris JS, Liao Y, et al. Global methylation profiles in DNA from different blood cell types. Epigenetics 2011; 6:76 - 85; http://dx.doi.org/10.4161/epi.6.1.13391; PMID: 20890131
  • Shimazu T, Hirschey MD, Newman J, He W, Shirakawa K, Le Moan N, et al. Suppression of oxidative stress by β-hydroxybutyrate, an endogenous histone deacetylase inhibitor. Science 2013; 339:211 - 4; http://dx.doi.org/10.1126/science.1227166; PMID: 23223453
  • Jones DP, Carlson JL, Samiec PS, Sternberg P Jr., Mody VC Jr., Reed RL, et al. Glutathione measurement in human plasma. Evaluation of sample collection, storage and derivatization conditions for analysis of dansyl derivatives by HPLC. Clin Chim Acta 1998; 275:175 - 84; http://dx.doi.org/10.1016/S0009-8981(98)00089-8; PMID: 9721075
  • Pfeiffer CM, Huff DL, Gunter EW. Rapid and accurate HPLC assay for plasma total homocysteine and cysteine in a clinical laboratory setting. Clin Chem 1999; 45:290 - 2; PMID: 9931056
  • Gamble MV, Ahsan H, Liu X, Factor-Litvak P, Ilievski V, Slavkovich V, et al. Folate and cobalamin deficiencies and hyperhomocysteinemia in Bangladesh. Am J Clin Nutr 2005; 81:1372 - 7; PMID: 15941889
  • Poirier LA, Wise CK, Delongchamp RR, Sinha R. Blood determinations of S-adenosylmethionine, S-adenosylhomocysteine, and homocysteine: correlations with diet. Cancer Epidemiol Biomarkers Prev 2001; 10:649 - 55; PMID: 11401915
  • Balaghi M, Wagner C. DNA methylation in folate deficiency: use of CpG methylase. Biochem Biophys Res Commun 1993; 193:1184 - 90; http://dx.doi.org/10.1006/bbrc.1993.1750; PMID: 8323540
  • Pilsner JR, Liu X, Ahsan H, Ilievski V, Slavkovich V, Levy D, et al. Genomic methylation of peripheral blood leukocyte DNA: influences of arsenic and folate in Bangladeshi adults. Am J Clin Nutr 2007; 86:1179 - 86; PMID: 17921400
  • Slot C. Plasma creatinine determination. A new and specific Jaffe reaction method. Scand J Clin Lab Invest 1965; 17:381 - 7; http://dx.doi.org/10.3109/00365516509077065; PMID: 5838275
  • Cheng Z, Zheng Y, Mortlock R, Van Geen A. Rapid multi-element analysis of groundwater by high-resolution inductively coupled plasma mass spectrometry. Anal Bioanal Chem 2004; 379:512 - 8; http://dx.doi.org/10.1007/s00216-004-2618-x; PMID: 15098084
  • Van Geen A, Cheng Z, Seddique AA, Hoque MA, Gelman A, Graziano JH, et al. Reliability of a commercial kit to test groundwater for arsenic in Bangladesh. Environ Sci Technol 2005; 39:299 - 303; http://dx.doi.org/10.1021/es0491073; PMID: 15667109
  • van Geen A, Cheng Z, Jia Q, Seddique AA, Rahman MW, Rahman MM, et al. Monitoring 51 community wells in Araihazar, Bangladesh, for up to 5 years: implications for arsenic mitigation. J Environ Sci Health A Tox Hazard Subst Environ Eng 2007; 42:1729 - 40; http://dx.doi.org/10.1080/10934520701564236; PMID: 17952774

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.