Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Regulatory Toxicology and Pharmacology 99 (2018) 78–88

Contents lists available at ScienceDirect

Regulatory Toxicology and Pharmacology


journal homepage: www.elsevier.com/locate/yrtph

A review on arsenic carcinogenesis: Epidemiology, metabolism, genotoxicity T


and epigenetic changes
Qing Zhou, Shuhua Xi∗
Department of Environmental and Occupational Health, Liaoning Provincial Key Laboratory of Arsenic Biological Effect and Poisoning, School of Public Health, China
Medical University, No.77 Puhe Road, Shenyang North New Area, Shenyang 110122, Liaoning Province, People's Republic of China

A R T I C LE I N FO A B S T R A C T

Keywords: Long-term exposure to arsenic (inorganic arsenic) is a world-wide environmental health concern. Arsenic is
Arsenic classified as the Group 1 human carcinogen by the International Agency for Research on Cancer (IARC).
Inorganic arsenic Epidemiological studies have established a strong association between inorganic arsenic (iAs) exposure in
Carcinogenesis drinking water and an increased incidence of cancer including bladder, liver, lung, prostate, and skin cancer. iAs
Epidemiology
also increases the risk of other diseases such as cardiovascular disease, hypertension and diabetes. The molecular
Methylation
Signal transduction
mechanisms of carcinogenesis of iAs remain poorly defined, several mechanisms have been proposed, including
genotoxicity, altered cell proliferation, oxidative stress, changes to the epigenome, disturbances of signal
transduction pathways, cytotoxicity and regenerative proliferation. In this article, we will summarize current
knowledge on the mechanisms of arsenic carcinogenesis and focus on integrating all these issues to garner a
broader perspective.

1. Introduction increased risks manifesting 40 years after arsenic exposure reduction


(Cohen et al., 2013).
Arsenic is an environmental contaminant found in soil, air, food, The mechanisms by which arsenic induces carcinogenesis are un-
and water. Chronic exposure to high levels of iAs is associated with a clear, in part due to lack of animal carcinogenicity studies for many
wide range of adverse effects, such as skin lesions, peripheral vascular years. At presents, several mechanisms of tumorigenesis induced by
diseases, reproductive toxicity, and neurological effects. In addition, arsenic compounds have been proposed, including oxidative stress,
several epidemiological studies have linked arsenic exposure to a genotoxic damage, chromosomal abnormalities, modification of gene
variety of human malignancies in skin, lungs, bladder, liver (Sankpal expression, epigenetic mechanisms, cytotoxicity and regenerative pro-
et al., 2012). Exposure occurs primarily via drinking water, but dietary liferation (Zhang et al., 2007a; b; Kitchin and Wallace, 2008; Cohen
exposure can also be substantial. Inhalation is a main route of arsenic et al., 2013). However, the main shortcomings in oxidative stress,
exposure in occupational environment (Aylward et al., 2014). It is es- genotoxicity, and chromosomal abnormalities are that most the studies
timated that more than 100 million people worldwide are exposed to supporting these modes of action being in vitro at lethal concentrations.
carcinogenic levels of iAs, the majority of which are caused by drinking Furthermore, some of the genotoxicity data in vivo have several defi-
arsenic contaminated groundwater. In some areas of the world, iAs in ciencies. Further understanding of the mechanisms of arsenic carcino-
drinking water is still at a very high level, such as Taiwan, Bangladesh, genesis is needed. Therefore, in this paper, we provide and discuss
India, China, Chile, Argentina and Mexico (Humans, 2012; Sankpal important information about the mechanisms of arsenic carcinogenesis.
et al., 2012; Cohen et al., 2013). Lung, bladder, and kidney cancer
mortality due to arsenic exposure have very long latencies, with

Abbreviations: As3MT, arsenic-3-methyl transferase enzyme; BER, base excision repair; CA, chromosomal aberration; DAPK, death-associated protein kinase; DMA,
dimethylarsinous acid; DNMT, DNA methyltransferase; DSBs, DNA double strand breaks; EGF, epidermal growth factor; ERCC2, excision repair; ERK1/2, extra-
cellular signal-regulated kinase1/; GSH, glutathione; GSTO, glutathione-S-transferase omega; JNK, c-Jun N-terminal protein Kinase; MAPK, mitogen-activated
protein kinase; MMA, monomethylarsonous acid; MN, micronuclei; Nrf2, nuclear factor erythroid 2-related factor 2; NF-кB, nuclear factor-kappa B; PI3K, phos-
phoinositide 3-kinases; ROS, reactive oxygen species; RPA1, replication protein A1; SAM, s-adenosyl-l-methionine; SCE, sister chromatid exchange; VEGF, vascular
endothelial growth factor

Corresponding author. Tel.: +86 13390587580.
E-mail address: Shxi@cmu.edu.cn (S. Xi).

https://doi.org/10.1016/j.yrtph.2018.09.010
Received 15 May 2018; Received in revised form 8 September 2018; Accepted 12 September 2018
Available online 15 September 2018
0273-2300/ © 2018 Elsevier Inc. All rights reserved.
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

2. Epidemiology of arsenic and cancer an increased risk of lung cancer was associated with median drinking
water arsenic levels in the range of 3–59 μg/L. A recent meta-regression
To date, the International Agency for Research on Cancer (IARC) by Lamm et al. (2015) indicates negative slope at low exposure levels
has confirmed the association of arsenic exposure with cancer of the (100–150 μg/L) with lung cancer. There is an increased risk of both
skin, lung, and bladder, while reports on the relationship with the liver, bladder cancer and lung cancer only at relatively high concentrations in
kidney, and prostate cancer are still limited. It is noteworthy that ar- the drinking water (> 150 μg/L). In addition, tobacco smoking com-
senic has been only related to kidney pelvis tumors rather than the bined with arsenic exposure increases the risk of lung cancer and de-
usual renal cell tumors, which are urothelial and an extension of the creasing arsenic exposure can also decrease the risk of lung cancer in
urinary bladder (Ferreccio et al., 2013). Key epidemiological evidence smokers (Chen et al., 2004a; b). Chen et al. (Chen et al., 2010a; b) found
regarding the carcinogenicity of arsenic comes from studies of Taiwan, no significant association between ingestion of well water containing
Bangladesh, Chile, and Argentina, where drinking water contains high iAs at levels of 300 μg/L or less and lung cancer among never-smokers.
concentrations of this metalloid (150 μg/L) (IARC, 2012). However, The OR was 2.25 for iAs concentrations in drinking water that are
there are few epidemiological studies to assess the carcinogenic effects greater than 300 μg/L (95% CI: 1.43–3.55). The threshold was found to
of arsenic at low to moderate levels of arsenic in drinking water and be modified by smoking. The relative contribution of iAs vs smoking
food. could not be assessed from the available data.
Until now, researches have focused on the association between lung
2.1. Skin cancer cancer and exposure to high levels of arsenic, but these studies are not
sufficient to evaluate the potential of iAs to induce lung cancer at low
According to the IARC, there is sufficient human evidence of skin exposures. In view of the potential relevance of low arsenic con-
carcinogenicity of iAs (IARC, 2004). Skin cancer was the first docu- centration to arsenic exposure, it needs further investigation.
mented carcinogenic effect from chronic exposure to iAs and has been
studied extensively (Hughes et al., 2011). Skin lesions, including skin 2.3. Bladder cancer
cancer, are characteristic of exposure to arsenic in drinking water
(Cohen et al., 2013). In skin, arsenic relates to increased risk for The IARC considers there to be sufficient evidence for arsenic to
squamous cell carcinoma (SCC) in situ, invasive SCC and basal cell increase risk of urinary bladder cancer in humans, yet risks related to
carcinoma (BCC) (Naujokas et al., 2013). Several studies have reported exposure in the 10–100 μg/L range of arsenic remain uncertain
the relationship between the chronic exposure of arsenic in drinking (Humans, 2004a; b; Meliker et al., 2010). An increased cancer mortality
water and the development of skin cancer. Arsenic induced skin cancer among cohorts exposed to extremely high arsenic levels has also been
due to high contamination of arsenic in drinking water has also been reported in Japan (Tsuda et al., 1995) and Chile (Marshall et al., 2007).
observed in West Bengal, India (Haque et al., 2003). Most studies on Studies confirm that the association between exposure to iAs in
skin cancer have found an increased risk only at concentrations above drinking water and bladder cancer is found only after exposure to iAs
100 μg/L, but a case–control study by Leonardi et al. (2012) found a concentrations greater than 100 μg/L. A prospective cohort study in
positive association between BCC and exposure to iAs through drinking northeastern Taiwan reported that the multivariate-adjusted relative
water with concentrations < 100 μg/L. Overall, the risk of skin cancer risk of urinary tract cancer was statistically significant for residents who
is increased in populations with high (> 100 μg/L) exposure to iAs in drank well water containing arsenic at levels > 100 μg/L (Chiou et al.,
drinking water. 2001). Schoen et al. (2004) concluded that an increase in bladder
While the toxicity of arsenic, in general, is well established, studies cancer is observed only in humans who were exposed to iAs in drinking
on the association between exposure to low arsenic concentration and water at high concentrations, usually several hundred μg/L. A cohort
the development of skin cancer are not enough. The association be- study by Chen et al. (Chen et al., 2010a; b) found statistically significant
tween the risk of skin cancer and concentrations of iAs in drinking increase in urinary tract carcinomas in a Taiwanese population who
water needs further exploration. ingested iAs at concentrations greater than 100 μg/L. Wang et al. (Wang
et al., 2013a; b) found an increased risk of bladder or upper urinary
2.2. Lung cancer tract urothelial carcinoma only in non-smoking Taiwanese exposed to
iAs in drinking water at concentrations of ≥350 μg/L for 10 years or
Lung is one of the major targets for tumorigenesis due to arsenic more. Results from different geographical areas indicate a threshold for
exposure (Hubaux et al., 2013). Arsenic in drinking water has been association between iAs exposure and bladder cancer at concentrations
classified as a known cause for the lung cancer by IARC. Exposure to of 100 μg/L or higher (Mink et al., 2008; Chen et al., 2010a; b; Tsuji
inorganic arsenic in various mining occupations led to lung cancer also et al., 2014).
could be discovered. Various epidemiological studies have shown the These studies have only verified the correlation between bladder
link between ingestion of iAs through drinking water and development cancer and high concentrations of arsenic, the studies at lower ex-
of human lung cancer. In a case-control study, a clear trend in lung posures have shown inconsistent results. In a lower exposure study
cancer odds ratios (OR) and 95% confidence intervals with increasing conducted in the US, no increase in bladder cancer incidence was found
concentration of arsenic in drinking water was observed (Ferreccio in population consuming drinking water containing iAs at concentra-
et al., 2000). Additionally, the associations between arsenic levels in tions of 10–100 μg/L (Meliker et al., 2010). This finding was consistent
drinking water and mortality of lung cancer using village as unit were with results of an earlier ecological study in Michigan (Meliker et al.,
assessed in ten townships in Taiwan, a significant increase in the 2007). In addition, an ecologic analysis in the United States found no
mortality of lung cancer in both genders with arsenic levels above arsenic-related increase in bladder cancer mortality was found over the
0.64 mg/L was found, but no significant effect was observed at lower exposure range of 3–60 μg/L (Lamm et al., 2004). However, an ecolo-
levels (Guo, 2004). Another study in northern Chile showed that lung gical study conducted by Mendez et al. supports an association between
and bladder cancer incidence in adults was markedly increased fol- low arsenic concentrations in drinking water (1.5–15.4 μg/L) and
lowing exposure to arsenic in early life, even up to 40 years after high bladder cancer incidence in the United States (Mendez et al., 2017).
exposures ceased (Steinmaus et al., 2014). Smith et al. (2012) found Mink et al. (2008) conducted a meta-analysis of data from eight studies
that a significant increase in lung cancer mortality in a region of Chile of low levels of iAs. The meta-analysis did not find significant asso-
where iAs concentrations in water ranged from 90 μg/L to nearly ciations between exposures to low levels of iAs in drinking water (ty-
1000 μg/L. pically < 100–200 μg/L) and bladder cancer for never smokers. The
However, Ferdosi H (Ferdosi et al., 2016) did not demonstrate that multiple epidemiologic studies support the conclusion that low

79
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

exposure to iAs does not increase the risk of bladder cancer (Tsuji et al., arsenite undergoes a series of sequential oxidative-methylation and
2014). reduction steps (Vahter, 2002). In the human arsenic metabolic
pathway, iAsV is converted to iAsIII, with subsequent methylation to
2.4. Liver cancer monomethylated and dimethylated arsenicals, respectively. Methyla-
tion is an important step in conversion of arsenic from inorganic to the
The epidemiological data for the skin, lung and urinary bladder is organic form, and the process consumes both S-adenosyl-l-methionine
widely accepted, whereas the association of long-term iAs exposure (SAM) and glutathione (GSH) (Drobna et al., 2005). The arsenic-3-
through drinking water with risk of liver cancer mortality remains methyl transferase enzyme (As3MT) metabolizes inorganic forms to
controversial (Liu and Waalkes, 2008). IARC listes the liver as a po- methylated forms (MMA and DMA), including both trivalent and pen-
tential organ for arsenic carcinogenesis (IARC, 2004). The increase of tavalent forms, with SAM as the methyl-donating cofactor. (Suzuki
liver cancer mortality in population drinking water with elevated ar- et al., 2002; Naranmandura et al., 2006). The dimethyl form also can be
senic was first reported in the southwestern population of Taiwan (Chen further methylated to trimethyl arsenic oxide (TMAO) in rodents and
et al., 1985). Moreover, increased liver cancer mortality associated with does not occur in humans except at extremely high exposures (Cohen
drinking water arsenic levels have been reported in populations from et al., 2013). In the biotransformation of inorganic arsenic, glutathione-
Inner Mongolia, China, Bangladesh, and Japan (Liu and Waalkes, S-transferase omega (GSTO) catalyzes the reduction of arsenate, MMAV,
2008). A meta-analysis indicated that long-term iAs exposure through and DMAV to the +3 arsenic species which are more toxic. Chowdhury
drinking water increased the risk of liver cancer mortality (Wang et al., et al. (2006). reported on the complete methylation of arsenate to
2014). In addition, an investigation in Chile revealed that exposure to DMAIII in GSTO knockout mice. There appears to be other enzyme(s)
arsenic in drinking water during early childhood may result in an in- other than GSTO able to reduce arsenic(V) species but to a lesser extent.
creased in childhood liver cancer mortality (Liaw et al., 2008). Al- While there is evidence that GSTO can reduce pentavalent arsenicals in
though studies conducted in Argentina, Chile, and other regions have vitro, its in vivo relevance is lacking (Nemeti et al., 2015). The general
reported a relationship of liver cancer with arsenic levels in drinking scheme of reactions is AsV + 2e→ AsIII + Me+→ MMAV + 2e →
water, no consensus has been made on this relationship mostly because MMAIII + Me+→DMAV + 2e→ DMAIII. Since arsenic metabolism in-
of a limited availability to representative data (Morales et al., 2000; volves oxidative methylation with S-adenosyl methionine as the methyl
Baastrup et al., 2008). donor, some nutrients such as folate levels affecting one carbon meta-
bolism may influence the arsenic methylation patterns, leading diffi-
2.5. Prostate cancer culties in exposure analysis (Gamble et al., 2005).
In the environment, the main forms of arsenic are iAsV and iAsIII
Multiple studies in humans reveal an association between environ- (National Research Council, 2001). The trivalent forms of arsenic,
mental inorganic arsenic exposure and prostate cancer mortality or especially organic trivalent arsenicals, are highly reactive and cyto-
incidence, prostate is considered to be a potential target for iAs carci- toxic, with activity at low micromolar concentrations (Cohen et al.,
nogenesis (Benbrahim-Tallaa and Waalkes, 2008). A US prospective 2013). It has been demonstrated that the trivalent forms of arsenic are
cohort study suggested that low to moderate exposure to inorganic the toxic forms in a variety of cell systems. The trivalent forms of or-
arsenic was prospectively associated with increased mortality of pros- ganic arsenic, MMAIII and DMAIII, have been also detected in the urine
tate cancer (Garcia-Esquinas et al., 2013). Roh et al. conducted an of rodents and humans that were exposed to high concentrations of iAs
ecologic study in Iowa, which showed a significant dose-dependent (Valenzuela et al., 2005). But the organic pentavalent forms of arsenic
association between low-level (1.08–2.06 μg/L) arsenic exposure and are mostly much less reactive, with activity generally at millimolar
prostate cancer (Roh et al., 2017). Moreover, in a US based ecologic concentrations (Cohen et al., 2013).
study, counties with higher mean arsenic levels in community water Notably, arsenic methylation is considered to be a bioactivation and
systems had significantly higher prostate cancer incidence. However, an detoxification pathways. The overall toxicity of arsenic is dependent on
individual-level study of prostate cancer incidence and low-level ar- the rate of methylation of the iAsIII. In vitro study has shown that
senic exposure is needed (Bulka et al., 2016). MMAIII and DMAIII are much more toxic than the pentavalent forms,
Overall, a reliable relationship between ingestion of iAs and cancers and trivalent methylated arsenicals are more toxic than trivalent in-
including skin, lung, and bladder cancer was only found in drinking organic arsenic (Mass et al., 2001). Chen et al. suggested that a higher
water concentrations exceeding 100 μg/L (Mink et al., 2008; Tsuji et al., ratio of MMA to DMA in urine is associated with an increased risk of
2014). Lack of significant associations may result from methodological bladder cancer development (Chen et al., 2003; Chen et al., 2004a; b).
issues (e.g., inadequate statistical power and by potential mis- Recent studies showed that the higher percentage of methylated arsenic
classification of exposure) (National Research Council, 2001). In addi- in urine, the more evidence of increased risk of skin and bladder cancer
tion, exposure to inorganic and organic arsenicals via food supply, (Moe et al., 2016; Khairul et al., 2017). Moreover, it has been found
which can confound epidemiology studies considerably, especially at that MMAIII and DMAIII can inhibit the activities of many enzymes and
low exposures to arsenic in water (Aylward et al., 2014). are more genotoxic than iAsIII (Dopp et al., 2010).

3. Metabolism of arsenic 4. Genotoxicity

The metabolism of arsenic is related to its toxicity. Inorganic arsenic Genotoxicity can be produced either by direct interaction with DNA
is absorbed in the gastrointestinal tract when ingested through food or or by indirect effects that produce errors in DNA. These indirect effects
drinking water. Arsenic uptake is strongly dependent on its oxidation can involve micronucleus (MN) formation, Sister chromatid exchange
state. iAsV enters cells through membrane transporters, such as the (SCE) or chromosomal aberrations (CAs). In addition, inhibition of DNA
sodium/phosphate co-transporters. The methylated pentavalent or- repair enzymes could also lead to an indirect genotoxic process.
ganic forms of arsenic are not transported, whereas the trivalent forms
(iAsIII) are readily transported (Rosen, 2002; Cohen et al., 2013). iAsIII 4.1. Direct damage to DNA
enters the cell via aquaglyceroporin channels and transporters (Yang
et al., 2012). The major transporters for the trivalent arsenicals include The chromosomal effects and DNA damage were postulated to be
P-glycoproteins, ATP-binding cassette transporters (multidrug-resistant critical events in the initiation and progression of human cancer.
proteins), and glucose transporters (Cohen et al., 2013). The most of Nesnow et al. (2002) demonstrated that DNA reaction did not occur
circulating arsenic undergoes bio-transformation in hepatocytes where with arsenicals. A direct interaction of arsenicals with DNA would lead

80
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

to formation of DNA adduct, and no arsenic containing DNA adduct has Arsenite interferes with the DNA damage response at multiple levels,
been reported. these effects of arsenic exposure likely contribute to its roles as a car-
cinogen.
4.2. Indirect damage to DNA
5. Epigenetic changes
The iAs-mediated chromosomal instability occurs frequently at
centromeres, leading to the formation of acentric chromosomes or the Recent a large body of studies on the mechanism of arsenic carci-
fusion of centromeres between two chromosomes (Kesari et al., 2012). nogenesis have shown the potential involvement of altered epigenetic
Fusion of two chromosomes at their centromeres can cause improper regulation in gene expression changes induced by arsenic exposure
chromosome segregation, which results in aneuploidy or micronuclei (Bjorklund et al., 2017). Epigenetics refers to the reversible regulation
formation (Xie et al., 2014). However, a fusion occurring at chromo- of gene expression, independent of DNA sequences. Epigenetic me-
somal ends may result in the formation of ring-like structures and/or chanisms allow cells to rapidly alter long-term transcriptional activity,
participate in abnormal SCE (Kesari et al., 2012). At low doses, iAs does thereby permitting coordinated changes in gene expression without
not cause DNA base pair mutations; instead, generate double-stranded permanently altering the sequence of DNA. Epigenetic alteration, which
breaks (Xie et al., 2014), which results in large-scale CAs (Klein et al., is not a genotoxic effect, leads to heritable phenomena that regulate
2007). gene expression and thus could be considered a form of potentially
Increased incidences of CAs, SCE and MN have been reported from reversible DNA modification. Epigenetic alterations are also believed to
the human populations exposed to arsenic through drinking water play an important role in the complex mode of action of iAs. The main
(Ostrosky-Wegman et al., 1991; Lerda, 1994; Dulout et al., 1996; epigenetic mechanisms are DNA methylation in gene promoter regions
Gonsebatt et al., 1997; Maki-Paakkanen et al., 1998; Mahata et al., that regulate gene expression, covalent posttranslational histone mod-
2003). MN are detected in exfoliated bladder cells, buccal cells, sputum ifications, and microRNAs (miRNAs) expression changes associated
cells, and lymphocytes from arsenic-exposed humans (Tian et al., 2001; with iAs exposure.
Basu et al., 2002). In vitro studies, concentrations of arsenite are
usually extremely high, frequently greater than 10 μmol/L, which 5.1. DNA methylation
would be lethal to most cell lines (Kligerman et al., 2003; Gentry et al.,
2010). Such high concentrations of trivalent arsenicals usually used in DNA methylation is one of the several epigenetic mechanisms which
vitro could not be reached in vivo at environmentally relevant con- cells use to control gene expression. DNA methylation provides suffi-
centrations of iAs. In animal studies, the positive results appear at ex- cient control over genes regulating cell differentiation, proliferation
tremely high doses of arsenic, which would be expected to produce and organism development, and these changes can be passed to
cytotoxicity and cell death. daughter cells during replication (Herceg, 2007). Global genomic DNA
However, there are numerous difficulties in interpretation of these methylation is a sign of many types of cancer. DNA methylation is the
studies. To begin with, most of these studies have not shown a dose covalent addition of a methyl group to the 5th carbon (C5) position of
response in relation to arsenic exposure as assessing either drinking cytosine-rich region, known as CpG islands (Bjorklund et al., 2017). In
water levels or urinary arsenic. Exposure to tobacco and tobacco pro- normal pathologically cells, promoter CpG islands regions are typically
ducts affect MN formation (Reali et al., 1987; Humans, 2004a; b; Roy unmethylated.
et al., 2016). This is particularly likely as a confounding factor given the Methylation is a major and important step in arsenic bio-
extensive use of betel quid in some populations and the relatively high transformation. These methylation steps facilitate arsenic excretion but
smoking rates (Humans, 2004a; b). Another possible variety among consume SAM at the same time. SAM, as the methyl donor for the
populations could be folate levels, which vary between populations and majority of methyltransferases, modifies DNA, RNA, histones and other
affect arsenic metabolism and toxicity (Gamble et al., 2006, 2007; proteins, dictating replicational, transcriptional and translational fide-
Peters et al., 2015). Folate is also associated with variability in micro- lity, mismatch repair, chromatin modeling, epigenetic modifications
nucleus formation (Lindberg et al., 2007; Bull et al., 2012). and imprinting, and thus plays an important role in cancer research
(Loenen, 2006). In humans, DNA methylation is initiated by the denovo
4.3. Inhibition of DNA repair DNA methyltransferase 3 (DNMT3A and B) and maintained by DNA
methyltransferase 1 (DNMT1) (Bestor, 2000). Following low-dose iAs
Interaction with a variety of DNA repair proteins could potentially exposure, the expression of the DNMTs is reduced, which results in less
lead to an increased risk of genotoxicity (Banerjee et al., 2008; Chen methylation at target sites (Reichard et al., 2007; Cheng et al., 2012).
et al., 2010a; b; Cohen et al., 2013). Inhibition of DNA repair has been When cells metabolize arsenic, the arsenic methyltransferase, AS3MT,
implicated as a mechanism of arsenic contaminated drinking water transfers a methyl group from SAM to the arsenite, depleting the
induced carcinogenesis. Gentry et al. (2010) highlighted inhibition of available methyl groups needed to the DNMTs for DNA methylation
DNA repair is a mode of action of arsenic carcinogenic effect. Arsenic (Reichard et al., 2007).
can inhibit nucleotide excision repair (NER), base excision repair (BER), Hypo- and hypermethylation of DNA in promoter regions are known
and mismatch repair (Hossain et al., 2012). Arsenic interfering with to mediate gene expression and to suppress gene expression, respec-
DNA repair is expected to promote mutation of key tumor suppressor tively. Both global DNA hypomethylation and hypermethylation in the
genes such as tumor protein P53 (TP53) in patients exposed to arsenic, promoter region of tumor suppressor genes are often associated with
leading to increased risk of bladder cancer (Kelsey et al., 2005). Ar- carcinogenesis and cancer outcomes. Population studies have shown
senic-mediated oxidative damage in enzymes is also reported to inter- that individuals having relatively lower capacity to methylate arsenic to
fere with the DNA repair mechanisms by either inhibiting ligation or DMAV are at greater risk for skin cancers, bladder cancer, and periph-
down-regulating the gene expression of DNA repair enzymes such as eral vascular disease (Chen et al., 2003; Tseng et al., 2005). A hospital-
DNA polymerase β (Sinha and Roy, 2011). Additionally, arsenic induces based case-control study performed in northern Mexico found that
formation of DNA adducts and leaves unreplicated stretches in re- women with higher %MMA and/or primary methylation index (PMI)
plicating DNA followed by DNA double strand breaks (DSBs) when the had an increased bladder cancer risk (Wei et al., 2017). Additionally,
adduct is excised. As arsenic impairs DNA repair mechanisms, the un- human skin exposure to iAs also led to global hypomethylation of tumor
repaired DSBs can lead to chromosome or chromatid type aberrations tissues (Zhang et al., 2007a; b). Arsenic exposure has been reported to
(Wang et al., 2001). induce DNA hypomethylation in vitro and in animal studies. For ex-
In summary, arsenic negatively impacts DNA repair capability. ample, rats (Uthus and Davis, 2005) and mice (Okoji et al., 2002; Chen

81
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

et al., 2004a; b; Xie et al., 2004) exposed to iAsIII for several weeks chromatin structure and regulate gene expression (Kouzarides, 2007).
displayed global DNA hypomethylation and aberrant gene expression. Chromatin is structured within the cell nucleus in units called nucleo-
Studies in cell lines in vitro yielded similar results, with a reduction in somes, in which DNA is packaged within the cell. The nucleosome core
global genomic DNA methylation resulting from iAsIII exposure particle consists of stretches of DNA wrapped in left-handed super he-
(Reichard et al., 2007). Urothelial carcinoma patients with high arsenic lical turns around a histone octamer consisting of two copies each of the
exposure show death-associated protein kinase (DAPK) promoter hy- core histones H2A, H2B, H3, and H4 (Kouzarides, 2007).
permethylation with low expression of the DAPK gene (Chen et al.,
2007). Hypermethylation and lower protein expression of DAPK were 5.3. Histone methylation
also detected in iAs treated SV40-immortalized human uroepithelial
cells (Chai et al., 2007). Histone methylation is also a reversible process, which occurs in
Conversely, some research indicates that iAs exposure leads to hy- both lysine and arginine residues (Wysocka et al., 2006). In mammals,
permethylation at the promoters of specific tumor-suppressor genes histone methylation is usually found on histone H3 and H4, although it
(Miao et al., 2015). In iAs-exposed human hepatocytes, significant hy- also occurs on H2A or H2B. AsIII treatment can lead to differential ef-
permethylation of the promoters of genes involved in DNA repair such fects on the methylation of H3 lysine residues, including increased H3
as excision repair cross-complementation group 2 (ERCC2) and re- lysine 9 dimethylation (H3K9me2) and H3 lysine 4 trimethylation
plication protein A1 (RPA1), and of genes associated with the Wnt (H3K4me3) and decreased H3 lysine 27 trimethylation (H3K27me3)
pathway like c-MYC (MYC) and Wnt family member 2B (WNT2B), was (Zhou et al., 2008). Elevated H3K9me2, mediated by increased levels of
observed (Miao et al., 2015). Dose-dependent hypermethylation at the histone methyltransferase G9a protein, correlates with transcriptional
promoter region of several tumor suppressor genes was induced by repression (Peterson and Laniel, 2004; Zhou et al., 2008); and elevated
arsenic exposure in vitro and in vivo (Stabile et al., 2005; Zhang et al., H3K9me2 was involved in the silencing of tumor suppressors in the
2007a; b). Significant hypermethylation of the promoters of the tumor cancer cell lines (Esteve et al., 2007). Arsenite induces global and gene
suppressor p16, and the DNA repair gene, mutl homolog 1 (MLH1) specific changes in DNA methylation and alteration in global histone
(Hossain et al., 2012; Lu et al., 2014), was observed in whole blood methylation levels, suggesting that epigenetic mechanisms contribute
obtained from humans chronically exposed to iAs. Additionally, the to arsenite-induced aberrant gene expression. However, data on the
promoter region of the p53 tumor suppressor gene was also found hyper patterns of histone methylation induced by arsenic exposure are lim-
methylated in arsenic-induced skin cancer patients (Chanda et al., ited, and further studies are required to decipher the relationship be-
2006). In a population-based study of bladder cancer, a significant in- tween altered histone methylation and gene expression, as well as its
crease in methylation at the RASSF1A and PRSS3 promoters was found effect on arsenic-induced carcinogenesis.
(Chai et al., 2007). This outcome was recapitulated in arsenic-induced
lung cancer in A/J mice, in which the arsenic exposure reduced the 5.4. Histone acetylation
expression of RASSF1A resulting from hypermethylation of its promoter
region (Chai et al., 2007). Histone acetylation is a dynamic and reversible event (Glozak and
Methylation of ingested iAs to MMAs and DMAs relies on folate- Seto, 2007), which stimulates the nucleotide excision repair pathway in
dependent one-carbon metabolism. Most studies have not taken into vivo by increasing the chromatin accessibility (Escargueil et al., 2008).
account controlling for this confounding factor, particularly for studies More recently, changes in histone H3 acetylation have been observed in
from Asia, including India, Bangladesh, and China, where folate defi- AsIII and MMAIII - induced malignant transformation of human ur-
ciency can be quite common. This can greatly influence the result of othelial cells in vitro. These modifications are arsenic specific, sug-
studies. In a cross-sectional study, Gamble ascertained that folate con- gesting that arsenicals may participate in tumorigenesis by altering the
ditions are associated with lower arsenic methylation (Gamble et al., epigenetic terrain of select genes (Jensen et al., 2008). Genome-wide
2005). Subsequent study provides compelling evidence that folic acid analysis suggests that iAs exposure during embryonic development
supplementation to adults with low plasma concentrations of folate causes global hypo-acetylation at H3K9 (Jensen et al., 2008). Recently,
would result in increased arsenic methylation (Gamble et al., 2006, study showed that the global level of H4K16 acetylation in human
2007). bladder epithelial cells was reduced in a dose- and time-dependent
Clearly, the results of various studies demonstrate that iAs can in- manner by both AsIII and MMAIII treatment (Liu et al., 2015). A re-
duce global levels of DNA hypomethylation and promoter specific hy- duction in H4K16 acetylation caused by chronic arsenic exposure could
permethylation of tumor suppressor genes and suggest that iAs-induced contribute to bladder carcinogenesis in bladder epithelial cells. More-
disruption of DNA methylation may be important in carcinogenesis. But over, knockdown of MYST1, the gene responsible for H4K16 acetyla-
the current available studies are essentially descriptive and difficult to tion, resulted in increased cytotoxicity from arsenical exposure in
interpret because of the complexity of the study populations and limited human bladder epithelial cells, suggesting that H4K16 acetylation may
information provided in the reports. Studies systematically in- be important for resistance to arsenic-induced toxicity (Jo et al., 2009).
vestigating DNA methylation on a genome wide level in arsenic-ex- These studies clearly support an association of iAs/MMAIII exposure
posed cell lines and in target tissues are needed. Such studies would and altered histone acetylation, but further work is needed to under-
help to clarify potential effects of arsenic exposure on DNA methylation stand the underlying mechanisms and to clarify the effect of altered
and carcinogenesis. histone acetylation on arsenic-induced toxicity and carcinogenesis.

5.2. Histone modifications 5.5. Histone phosphorylation

In addition to DNA methylation, histone modifications could be Besides, phosphorylation of core histone proteins may be deeply
potential biomarkers for the assessment of arsenic exposure and pre- involved in arsenic-induced carcinogenicity (Ren et al., 2011). All four
diction of adverse health effects (Ma et al., 2016). Covalent post- core histone proteins, H2A, H2B, H3, and H4 can be post-translationally
translational modifications of histone proteins are known to play an modified by phosphorylation, and are targets of protein kinases
important role in chromatin remodeling and thereby in regulation of (Peterson and Laniel, 2004). Cyclin-dependent kinases are believed to
gene expression, causing various cellular responses to environmental be responsible for H1 phosphorylation. Several kinases can phosphor-
influences (Liu et al., 2015). To date, published studies on histone ylate H2A and H2B. Phosphorylation of H3 has been specifically im-
modifications and arsenic toxicity have focused on acetylation, me- plicated in cell cycle progression and regulation of gene expression
thylation, and phosphorylation. These modifications modulate (Houben et al., 2007). Similarly, phosphorylation of histone H4 (serine

82
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

1) increases during the cell cycle, which is believed to be regulated by with lung cancer (Xue et al., 2016). Moreover, inhibition of miR-155
casein kinase 2 (Barber et al., 2004). Only a few studies have demon- also decreased the cell proliferation, and triggered the apoptotic cell
strated a connection between histone phosphorylation and arsenic-in- death in lung adenocarcinoma cells (Lv et al., 2016). These findings
duced carcinogenicity. Studies have suggested that arsenic induced H3 indicate miR-155 may be served as the specific biomarker for diagnosis
phosphorylation might be responsible for the up-regulation of the on- and prognostic prediction of lung cancer. ROS generation resulting
cogenes c-fos and c-jun (Li et al., 2003). Because arsenite exposure is from arsenic exposure is thought to play a large role in arsenic-induced
known to activate JNK and p38/Mpk2 kinase by inhibition of the cor- carcinogenesis and toxicity (Flora et al., 2007; Suarez et al., 2007) and
responding protein phosphatases, phosphorylation of histone H3 via the could potentially alter these miRNAs in a similar manner (Marsit et al.,
JNK/SAPK pathway might be a common mechanism of metal-induced 2006). The induced changes in miRNA expression were not stable,
histone modification. In addition, Ray et al. reported that iAs induces a suggesting that chronic exposure may be necessary to permanently alter
coordinated regulation of Nrf2 and histone H3S10 phosphorylation, expression of miRNAs (Marsit et al., 2006).
which activates the human heme oxygenase-1gene (HMOX1) (Ray Further studies are necessary to clarify whether chronic exposure to
et al., 2015). These studies suggest that H3 phosphorylation may also arsenic can alter miRNA expression and what biological effects are re-
contribute to As-induced carcinogenesis. lated to the altered miRNA expression. All these epigenetic marks are
associated with iAs-induced carcinogenesis, additional research is re-
5.6. MicroRNAs quired to clarify the mechanisms driving each system. More studies
comparing epigenetic alterations mediated by different arsenicals are
Several studies have shown that differential miRNA expression can crucial, considering that toxicities and potential to cause cancer of ar-
be a hallmark of cancer and miRNAs can function as potential onco- senicals vary considerably. Expanded research in this area will de-
genes or tumor suppressor genes influencing tumor development, pro- lineate the role of iAs exposure in the initiation and development of
gression and response to therapy (Gonzalez et al., 2015; States, 2015). cancer.
miRNAs constitute another epigenetic mechanism of gene regulation
affecting development, growth, and the response to stress. miRNAs are 6. Signal transduction pathways
single strand small non-coding RNAs that function as negative reg-
ulators of target mRNA expression at the post-transcriptional level, Signal transduction pathways transmit extracellular signals, via an
which participate in diverse biological regulatory events and are tran- intracellular series of signaling molecules into gene. Cellular processes
scribed mainly from non-protein-coding regions of the genome (He and such as proliferation, differentiation, and apoptosis are directed and
Hannon, 2004). Each miRNA is thought to target several hundred managed by these pathways or cascades. Arsenic can alter signal
genes, and as many as 30% of mammalian genes are regulated by transduction, which leads to activation or inhibition of transcription
miRNA (Lewis et al., 2005). miRNAs can modify essential cellular factors, regulatory proteins which bind to DNA and regulates gene
functions such as proliferation, differentiation and programmed death transcription (Huang et al., 2004; Kumagai and Sumi, 2007; Platanias,
(He et al., 2005). The exact mechanisms by which expression of target 2009).
mRNA is repressed are still under investigation but may include the NF-κB is an inducible transcription factor involved in diverse sides
inhibition of protein synthesis, the degradation of target mRNAs, and of cellular processes, NF-κB activation in cancers has been linked to
the translocation of target mRNAs into cytoplasmic processing bodies tumor cell resistance to apoptosis and necrosis, increasing proliferation,
(Suarez et al., 2007). miRNAs bind to the 3′-UTR (untranslated regions) angiogenesis, and metastasis (Chen et al., 2016). NF-κB is transiently
of target mRNAs through base pairing, stimulating mRNA degradation activated by a variety of signals, including cytokines, chemokines,
or inhibiting protein translation (Hei and Filipic, 2004; Flora et al., bacterial and viral products, and free radicals. Chen et al. recently
2007). found that arsenite caused the oxidative damage by activation of Nrf2
Several in vitro and human studies demonstrate that arsenic-in- and NF-κB-mediated pathways in cultured 16 E6/E7 transformed
duced alterations in miRNA gene expression. A recently study provide normal human bronchial epithelial (16-HBE) cells (Chen et al., 2017). A
strong evidence that the differential expression of miRNAs and target recently study showed that the transformation of HBE cells induced by
genes at early stage of arsenite exposure may contribute to arsenic in- chronic exposure to sodium arsenite is mediated by decreased Nrf2
duced carcinogenesis (Al-Eryani et al., 2018). In prostate cancer, miR- level and its downstream NQO1 and HO-1 protein, which subsequently
150 is upregulated, and is additionally correlated with tumor recur- promote the malignant proliferation (Gu et al., 2016).
rence and metastasis, as well as poor overall survival; and miR-150 Epidermal growth factor (EGF) binds to EGF receptor (EGFR) and
could serve as a novel and reliable prognostic biomarker for prostate ultimately activates mitogen-activated protein kinase (MAPK) path-
cancer patients (Dezhong et al., 2015). miR-21 is a key oncogene, which ways, which plays an important role in cellular stress response, in-
is highly overexpressed in a wide range of cancers. miR-21 plays an flammation, and growth. Arsenic can stimulate c-Src activity, which can
important role in the initiation and progression of cancer and is im- then activate EGFR by physical interaction resulting in two unique
plicated in multiple malignancy-related processes such as cell pro- tyrosine phosphorylation events, leading to ligand-independent EGFR
liferation, apoptosis, invasion, and metastasis. Knockdown of miR-21 phosphorylation and constitutive activation (Simeonova and Luster,
can reduce proliferation and tumor growth in MCF- 7 cells, and reduce 2002). Arsenic can also induce activation of components of the EGFR
invasion and metastasis in MDAMB-231 cells (Pratheeshkumar et al., pathway in lung epithelial cells, such as Ras, Raf, MEK and ERK through
2016). Upregulation of miR-21 can enhance the transformation po- ROS (Zhang et al., 2007a; b; Miao et al., 2015). EGFR, an upstream
tential of cell by targeting programmed cell-death 4 (PDCD4) (Luo effector of MAPK signaling, is normally inhibited by protein tyrosine
et al., 2015). Gao founded that iAs-mediated autophagy was supported phosphatases (PTPs) such as MAPK phosphatase 1. MAPK pathways
by miR-21-induced PTEN-ERK signaling, and iAs-induced angiogenesis include the extracellular signal-regulated kinase1/2 (ERK1/2), p38
occurred through a microRNA 425-5p-regulated CCM3 (Gao et al., MAPK, and stress-activated protein kinase/Jun N-terminal protein ki-
2016). Furthermore, miR-155, an oncogenic miRNA, has been found to nase (SAPK/JNK) pathways. Erk1/2 is involved in growth and differ-
be increased in many types of cancer (Xue et al., 2016). Chen C et al. entiation while p38 MAPK and SAPK/JNK are important in stress-in-
demonstrated for the first time that inhibition of miR-155 by its specific duced inflammation and mitochondrial-dependent apoptosis. Our
inhibitor could remarkably attenuate the malignant phenotypes and previous studies found that arsenic activated p-ERK, p-P38 and p-JNK
promote apoptotic cell death in arsenite-transformed cells (Chen et al., in SV-HUC-1 cells (Wang et al., 2013a; b). Arsenic-exposed hepatocel-
2017). Results from clinical trials showed that overexpressed miR-155 lular carcinoma cells display overexpression of EGFR (Sung et al.,
in the serum results in low survival rate and poor prognosis of patients 2012), while in leukemia cell lines, iAsIII is capable of activating Rac1

83
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

GTPases resulting in downstream engagement of the JNK pathway and immune system to some extent, the tumors related to arsenic, particu-
increasing cell survival and proliferation (Herbert and Snow, 2012). larly skin, lung, and bladder, are not increased in individuals who are
Furthermore, acute exposure to arsenite can stimulate the PI3K/ immunosuppressed (National Research Council, 1999). Essentially,
AKT phosphorylation cascade, leading to cellular transformation char- immunosuppression is only related to virally associated malignancies
acterized by increased proliferation and anchorage-independent growth such as lymphoma, Kaposi's sarcoma, various squamous cell carcinomas
(Liu et al., 2012). In human bronchial epithelial cells (HBECs), iAsIII can of mucous membranes, such as the cervix and oral cavity, and to a
stimulate AKT and the consequent release of vascular endothelial limited extent, liver cancer. (Cohen et al., 1991, 2004).
growth factor (VEGF), inducing cell migration through different me-
chanisms (Wang et al., 2012). During malignant transformation of stem
8. Cytotoxicity and regeneration
cells, arsenite has also been shown to suppress expression of PTEN, an
important inhibitor of PI3K/AKT signaling (Stueckle et al., 2012).
The mechanisms of cancer induced by arsenicals appear to involve
Evidence also suggests that this metalloid works through AKT to pro-
cytotoxicity with consequent regeneration (Cohen et al., 2006, 2013;
mote epithelial to mesenchymal transition (EMT), invasion, and mi-
Dodmane et al., 2013). Most commonly hyperplasia involves not only
gration of arsenic-transformed bronchial epithelial cells (Wang et al.,
an increase in the cell number but an increase in the replication rate
2012).
(Cohen et al., 2016). The evidence is strong and accumulating that
Endocrine disruptors (EDs) can mimic or antagonize the effects of
arsenic induced carcinogenesis of urinary bladder, skin and lung in-
estrogens in target tissue (Casals-Casas and Desvergne, 2011). Evidence
volves cytotoxicity and regeneration (Cohen et al., 2013; Dodmane
has accumulated that estrogen receptors α and β (ERα and ERβ) are
et al., 2013; Efremenko et al., 2015). Trivalent arsenicals (iAsIII,
expressed in lung tumors, acting both in membrane-initiated and nu-
MMAIII, and DMAIII) are highly toxic to cells at concentrations ranging
clear signaling (Siegfried and Stabile, 2014). Arsenic are EDs that ac-
from 0.1 μmol/L to less than 10 μmol/L. Concentrations less than
tivate genomic ERα, through interaction with the zinc fingers of the
0.1 μmol/L are adaptive, and concentrations ≥10 μmol/L are lethal to
DNA binding domain and/or activation of protein kinase C and/or di-
all cell types (Cohen et al., 2002, 2013; Gentry et al., 2010). High doses
rect interaction with the ligand binding domain (Fechner et al., 2011).
of DMAV orally administered for 10 weeks to rats increase urothelial
Several studies demonstrate that environmentally relevant concentra-
cell proliferation and cytotoxic changes, including cellular necrosis,
tions of NaAsO2 induces rapid activation of MAPK and cellular pro-
suggesting that administration of DMAV results in cytotoxicity with
liferation in human lung adenocarcinoma cells. Further, this activation
necrosis, followed by regenerative hyperplasia of the bladder epithe-
is mediated through ER, Src, EGFR, and GPER, and arsenite is de-
lium. Cohen et al. (2006) suggested that the toxic effects observed in
monstrated as EDs in lung adenocarcinoma (Stabile et al., 2005;
the rat urinary bladder induced by DMAV may be caused by DMAIII
Marquez-Garban et al., 2007; Pietras and Marquez-Garban, 2007).
formation in vivo.
Moreover, intense expression of ERα is observed in liver tumors and
Cytotoxicity has also been proposed as a nonlinear mode of action
tumor-surrounding normal tissues after gestational arsenic exposure in
for the bladder carcinogenicity of arsenic (Wei et al., 2005; Cohen et al.,
mice (Waalkes et al., 2004, 2006). ERα overexpression also occurs in
2006). Epidemiologic studies also support this conclusion. For the
uterine tumors, urinary bladder carcinoma (Waalkes et al., 2006).
cancer endpoint, it is not cancer that is induced by the arsenic but a
Overexpression of ERα and ER-linked cyclin D1 was also evident in
variety of toxicities, such as arseniasis skin changes, bronchitis, or ur-
liver biopsy samples of arsenicosis patients in Guizhou, China
othelial toxicity, which lead to regenerative proliferation and ulti-
(Marquez-Garban et al., 2007). Although aberrant ER signaling path-
mately an indirect induction of cancer.
ways may be important in carcinogenesis induced by arsenic, it may not
apply to all targets of arsenic carcinogenesis.
9. Conclusion
7. Immunosuppression
Inorganic arsenic is a human carcinogen targeting mainly urinary
The multiple effects of As on immune system tend to decrease the bladder, lung and skin. Genetic changes, inhibition of DNA repair,
immune surveillance system and increase the rate of infection, auto- changes to the epigenome and signal transduction are involved in ar-
immune disease, cancer and other immune mediated problems (Haque senic carcinogenesis. The metabolism of arsenic may also play a role in
et al., 2017). Recent studies found that arsenic disrupts macrophage this process. The biotransformation of iAsV into iAsIII and its methylated
function and exacerbate lipopolysaccharide-induced inflammation that conjugates play a crucial role in arsenic carcinogenicity at both genetic
is mediated by macrophages and monocytes (Srivastava et al., 2013). and epigenetic levels. Compared with iAs, MMAIII and DMAIII exert
Inflammation is a well-known hallmark event for cancer progression, several unique biological effects and are more cytotoxic and genotoxic.
which plays a key role in protecting the damaged tissues after injury, The methylation of iAs is considered a mechanism of activation of its
but also leads to several diseases, including various cancers (Rao et al., toxicity and carcinogenicity. In addition, the difference in susceptibility
2017). Increased secretion of inflammatory biomarkers like IL-8, TNF- to arsenic is related to arsenic carcinogenesis. The individual suscept-
α, and TGF-β is related with iAs exposure in urothelial cells (Liu et al., ibility may result from differences in metabolism genes, DNA repair,
2014). Chronic low-level iAs exposure (11–50 μg/L) evoke systemic cell transport, immune response, antioxidant defenses and cell cycle
inflammation by upregulating pro-inflammatory mediators like TNF-α, control. Evidence supporting arsenic carcinogenesis comes from in vitro
IL-6, IL-8, IL-12, and CRP (Prasad and Sinha, 2017). Dutta has reported investigations on human cells and epidemiological studies. Further
induction of various pro inflammatory and inflammatory markers in studies are still necessary for arsenic carcinogenesis.
chronic low level iAs exposed population (Dutta et al., 2015). These
results show that prolonged arsenic exposure can impair immune re-
Author contributions
sponses against inflammation. However, there are few in-depth and
systematic studies to determine whether arsenic causes cancer through
Qing Zhou conceived and drafted this manuscript. Shuhua Xi re-
inflammatory diseases. It is not clear whether arsenic changes the signal
vised and approved the final manuscript.
of inflammation first, and then causes chronic disease. Extensive la-
boratory studies (both in vitro and in vivo) are warranted to elucidate
the direct effects of arsenic on inflammation and cancer (Dangleben Conflicts of interest
et al., 2013).
Although there is evidence that inorganic arsenic can suppress the The authors declare no conflict of interest.

84
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

Acknowledgments Chen, Y.C., Guo, Y.L., Su, H.J., Hsueh, Y.M., Smith, T.J., Ryan, L.M., Lee, M.S., Chao, S.C.,
Lee, J.Y., Christiani, D.C., 2003. Arsenic methylation and skin cancer risk in south-
western Taiwan. J. Occup. Environ. Med. 45 (3), 241–248.
This work was supported by the gs1:National Natural Science Cheng, T.F., Choudhuri, S., Muldoon-Jacobs, K., 2012. Epigenetic targets of some tox-
Foundation of China (81673207 and 81373023). icologically relevant metals: a review of the literature. J. Appl. Toxicol. 32 (9),
643–653.
Chiou, H.Y., Chiou, S.T., Hsu, Y.H., Chou, Y.L., Tseng, C.H., Wei, M.L., Chen, C.J., 2001.
Transparency document Incidence of transitional cell carcinoma and arsenic in drinking water: a follow-up
study of 8,102 residents in an arseniasis-endemic area in northeastern Taiwan. Am. J.
Transparency document related to this article can be found online at Epidemiol. 153 (5), 411–418.
Chowdhury, U.K., Zakharyan, R.A., Hernandez, A., Avram, M.D., Kopplin, M.J.,
https://doi.org/10.1016/j.yrtph.2018.09.010. Aposhian, H.V., 2006. Glutathione-S-transferase-omega [MMA(V) reductase]
knockout mice: enzyme and arsenic species concentrations in tissues after arsenate
References administration. Toxicol. Appl. Pharmacol. 216 (3), 446–457.
Cohen, S.M., Arnold, L.L., Beck, B.D., Lewis, A.S., Eldan, M., 2013. Evaluation of the
carcinogenicity of inorganic arsenic. Crit. Rev. Toxicol. 43 (9), 711–752.
Al-Eryani, L., Waigel, S., Tyagi, A., Peremarti, J., Jenkins, S.F., Damodaran, C., States, Cohen, S.M., Arnold, L.L., Eldan, M., Lewis, A.S., Beck, B.D., 2006. Methylated arsenicals:
J.C., 2018. Differentially expressed mRNA targets of differentially expressed miRNAs the implications of metabolism and carcinogenicity studies in rodents to human risk
predict changes in the TP53 Axis and carcinogenesis related pathways in human assessment. Crit. Rev. Toxicol. 36 (2), 99–133.
keratinocytes chronically exposed to arsenic. Toxicol. Sci. 162 (2), 645–654. Cohen, S.M., Arnold, L.L., Uzvolgyi, E., Cano, M., St John, M., Yamamoto, S., Lu, X., Le,
Aylward, L.L., Ramasamy, S., Hays, S.M., Schoeny, R., Kirman, C.R., 2014. Evaluation of X.C., 2002. Possible role of dimethylarsinous acid in dimethylarsinic acid-induced
urinary speciated arsenic in NHANES: issues in interpretation in the context of po- urothelial toxicity and regeneration in the rat. Chem. Res. Toxicol. 15 (9),
tential inorganic arsenic exposure. Regul. Toxicol. Pharmacol. 69 (1), 49–54. 1150–1157.
Baastrup, R., Sorensen, M., Balstrom, T., Frederiksen, K., Larsen, C.L., Tjonneland, A., Cohen, S.M., Chowdhury, A., Arnold, L.L., 2016. Inorganic arsenic: a non-genotoxic
Overvad, K., Raaschou-Nielsen, O., 2008. Arsenic in drinking-water and risk for carcinogen. J. Environ. Sci. (China) 49, 28–37.
cancer in Denmark. Environ. Health Perspect. 116 (2), 231–237. Cohen, S.M., Klaunig, J., Meek, M.E., Hill, R.N., Pastoor, T., Lehman-McKeeman, L.,
Banerjee, M., Sarma, N., Biswas, R., Roy, J., Mukherjee, A., Giri, A.K., 2008. DNA repair Bucher, J., Longfellow, D.G., Seed, J., Dellarco, V., Fenner-Crisp, P., Patton, D., 2004.
deficiency leads to susceptibility to develop arsenic-induced premalignant skin le- Evaluating the human relevance of chemically induced animal tumors. Toxicol. Sci.
sions. Int. J. Canc. 123 (2), 283–287. 78 (2), 181–186.
Barber, C.M., Turner, F.B., Wang, Y., Hagstrom, K., Taverna, S.D., Mollah, S., Ueberheide, Cohen, S.M., Purtilo, D.T., Ellwein, L.B., 1991. Ideas in pathology. Pivotal role of in-
B., Meyer, B.J., Hunt, D.F., Cheung, P., Allis, C.D., 2004. The enhancement of histone creased cell proliferation in human carcinogenesis. Mod. Pathol. 4 (3), 371–382.
H4 and H2A serine 1 phosphorylation during mitosis and S-phase is evolutionarily Dangleben, N.L., Skibola, C.F., Smith, M.T., 2013. Arsenic immunotoxicity: a review.
conserved. Chromosoma 112 (7), 360–371. Environ. Health 12 (1), 73.
Basu, A., Mahata, J., Roy, A.K., Sarkar, J.N., Poddar, G., Nandy, A.K., Sarkar, P.K., Dutta, Dezhong, L., Xiaoyi, Z., Xianlian, L., Hongyan, Z., Guohua, Z., Bo, S., Shenglei, Z., Lian, Z.,
P.K., Banerjee, A., Das, M., Ray, K., Roychaudhury, S., Natarajan, A.T., Nilsson, R., 2015. miR-150 is a factor of survival in prostate cancer patients. J BUON 20 (1),
Giri, A.K., 2002. Enhanced frequency of micronuclei in individuals exposed to arsenic 173–179.
through drinking water in West Bengal, India. Mutat. Res. 516 (1–2), 29–40. Dodmane, P.R., Arnold, L.L., Kakiuchi-Kiyota, S., Qiu, F., Liu, X., Rennard, S.I., Cohen,
Benbrahim-Tallaa, L., Waalkes, M.P., 2008. Inorganic arsenic and human prostate cancer. S.M., 2013. Cytotoxicity and gene expression changes induced by inorganic and or-
Environ. Health Perspect. 116 (2), 158–164. ganic trivalent arsenicals in human cells. Toxicology 312, 18–29.
Bestor, T.H., 2000. The DNA methyltransferases of mammals. Hum. Mol. Genet. 9 (16), Dopp, E., von Recklinghausen, U., Diaz-Bone, R., Hirner, A.V., Rettenmeier, A.W., 2010.
2395–2402. Cellular uptake, subcellular distribution and toxicity of arsenic compounds in me-
Bjorklund, G., Aaseth, J., Chirumbolo, S., Urbina, M.A., Uddin, R., 2017. Effects of arsenic thylating and non-methylating cells. Environ. Res. 110 (5), 435–442.
toxicity beyond epigenetic modifications. Environ. Geochem. Health 40 (3), 955–965. Drobna, Z., Waters, S.B., Devesa, V., Harmon, A.W., Thomas, D.J., Styblo, M., 2005.
Bulka, C.M., Jones, R.M., Turyk, M.E., Stayner, L.T., Argos, M., 2016. Arsenic in drinking Metabolism and toxicity of arsenic in human urothelial cells expressing rat arsenic
water and prostate cancer in Illinois counties: an ecologic study. Environ. Res. 148, (+3 oxidation state)-methyltransferase. Toxicol. Appl. Pharmacol. 207 (2), 147–159.
450–456. Dulout, F.N., Grillo, C.A., Seoane, A.I., Maderna, C.R., Nilsson, R., Vahter, M., Darroudi,
Bull, C.F., Mayrhofer, G., Zeegers, D., Mun, G.L., Hande, M.P., Fenech, M.F., 2012. Folate F., Natarajan, A.T., 1996. Chromosomal aberrations in peripheral blood lymphocytes
deficiency is associated with the formation of complex nuclear anomalies in the cy- from native Andean women and children from northwestern Argentina exposed to
tokinesis-block micronucleus cytome assay. Environ. Mol. Mutagen. 53 (4), 311–323. arsenic in drinking water. Mutat. Res. 370 (3–4), 151–158.
Casals-Casas, C., Desvergne, B., 2011. Endocrine disruptors: from endocrine to metabolic Dutta, K., Prasad, P., Sinha, D., 2015. Chronic low level arsenic exposure evokes in-
disruption. Annu. Rev. Physiol. 73, 135–162. flammatory responses and DNA damage. Int. J. Hyg Environ. Health 218 (6),
Chai, C.Y., Huang, Y.C., Hung, W.C., Kang, W.Y., Chen, W.T., 2007. Arsenic salts induced 564–574.
autophagic cell death and hypermethylation of DAPK promoter in SV-40 im- Efremenko, A.Y., Seagrave, J., Clewell, H.J., Van Landingham, C., Gentry, P.R., Yager,
mortalized human uroepithelial cells. Toxicol. Lett. 173 (1), 48–56. J.W., 2015. Evaluation of gene expression changes in human primary lung epithelial
Chanda, S., Dasgupta, U.B., Guhamazumder, D., Gupta, M., Chaudhuri, U., Lahiri, S., Das, cells following 24-hr exposures to inorganic arsenic and its methylated metabolites
S., Ghosh, N., Chatterjee, D., 2006. DNA hypermethylation of promoter of gene p53 and to arsenic trioxide. Environ. Mol. Mutagen. 56 (5), 477–490.
and p16 in arsenic-exposed people with and without malignancy. Toxicol. Sci. 89 (2), Escargueil, A.E., Soares, D.G., Salvador, M., Larsen, A.K., Henriques, J.A., 2008. What
431–437. histone code for DNA repair? Mutat. Res. 658 (3), 259–270.
Chen, C., Jiang, X., Gu, S., Zhang, Z., 2017. MicroRNA-155 regulates arsenite-induced Esteve, P.O., Chin, H.G., Pradhan, S., 2007. Molecular mechanisms of transactivation and
malignant transformation by targeting Nrf2-mediated oxidative damage in human doxorubicin-mediated repression of survivin gene in cancer cells. J. Biol. Chem. 282
bronchial epithelial cells. Toxicol. Lett. 278, 38–47. (4), 2615–2625.
Chen, C.J., Chuang, Y.C., Lin, T.M., Wu, H.Y., 1985. Malignant neoplasms among re- Fechner, P., Damdimopoulou, P., Gauglitz, G., 2011. Biosensors paving the way to un-
sidents of a Blackfoot disease-endemic area in Taiwan: high-arsenic artesian well derstanding the interaction between cadmium and the estrogen receptor alpha. PLoS
water and cancers. Canc. Res. 45 (11 Pt 2), 5895–5899. One 6 (8), e23048.
Chen, C.L., Chiou, H.Y., Hsu, L.I., Hsueh, Y.M., Wu, M.M., Chen, C.J., 2010a. Ingested Ferdosi, H., Dissen, E.K., Afari-Dwamena, N.A., Li, J., Chen, R., Feinleib, M., Lamm, S.H.,
arsenic, characteristics of well water consumption and risk of different histological 2016. Arsenic in drinking water and lung cancer mortality in the United States: an
types of lung cancer in northeastern Taiwan. Environ. Res. 110 (5), 455–462. analysis based on US counties and 30 Years of observation (1950-1979). J Environ
Chen, C.L., Chiou, H.Y., Hsu, L.I., Hsueh, Y.M., Wu, M.M., Wang, Y.H., Chen, C.J., 2010b. Public Health 2016, 1602929.
Arsenic in drinking water and risk of urinary tract cancer: a follow-up study from Ferreccio, C., Gonzalez, C., Milosavjlevic, V., Marshall, G., Sancha, A.M., Smith, A.H.,
northeastern Taiwan. Canc. Epidemiol. Biomarkers Prev. 19 (1), 101–110. 2000. Lung cancer and arsenic concentrations in drinking water in Chile.
Chen, C.L., Hsu, L.I., Chiou, H.Y., Hsueh, Y.M., Chen, S.Y., Wu, M.M., Chen, C.J., Epidemiology 11 (6), 673–679.
Blackfoot Disease Study, G., 2004a. Ingested arsenic, cigarette smoking, and lung Ferreccio, C., Smith, A.H., Steinmaus, C., 2013. Case-control study of arsenic in drinking
cancer risk: a follow-up study in arseniasis-endemic areas in Taiwan. J. Am. Med. water and kidney cancer in uniquely exposed Northern Chile. Am. J. Epidemiol. 178
Assoc. 292 (24), 2984–2990. (5), 813–818.
Chen, G., Mao, J., Zhao, J., Zhang, Y., Li, T., Wang, C., Xu, L., Hu, Q., Wang, X., Jiang, S., Flora, S.J., Bhadauria, S., Kannan, G.M., Singh, N., 2007. Arsenic induced oxidative stress
Nie, X., Wu, Q., 2016. Arsenic trioxide mediates HAPI microglia inflammatory re- and the role of antioxidant supplementation during chelation: a review. J. Environ.
sponse and the secretion of inflammatory cytokine IL-6 via Akt/NF-kappaB signaling Biol. 28 (2 Suppl. l), 333–347.
pathway. Regul. Toxicol. Pharmacol. 81, 480–488. Gamble, M.V., Liu, X., Ahsan, H., Pilsner, J.R., Ilievski, V., Slavkovich, V., Parvez, F.,
Chen, H., Li, S., Liu, J., Diwan, B.A., Barrett, J.C., Waalkes, M.P., 2004b. Chronic in- Chen, Y., Levy, D., Factor-Litvak, P., Graziano, J.H., 2006. Folate and arsenic meta-
organic arsenic exposure induces hepatic global and individual gene hypomethyla- bolism: a double-blind, placebo-controlled folic acid-supplementation trial in
tion: implications for arsenic hepatocarcinogenesis. Carcinogenesis 25 (9), Bangladesh. Am. J. Clin. Nutr. 84 (5), 1093–1101.
1779–1786. Gamble, M.V., Liu, X., Ahsan, H., Pilsner, R., Ilievski, V., Slavkovich, V., Parvez, F., Levy,
Chen, W.T., Hung, W.C., Kang, W.Y., Huang, Y.C., Chai, C.Y., 2007. Urothelial carcinomas D., Factor-Litvak, P., Graziano, J.H., 2005. Folate, homocysteine, and arsenic meta-
arising in arsenic-contaminated areas are associated with hypermethylation of the bolism in arsenic-exposed individuals in Bangladesh. Environ. Health Perspect. 113
gene promoter of the death-associated protein kinase. Histopathology 51 (6), (12), 1683–1688.
785–792. Gamble, M.V., Liu, X., Slavkovich, V., Pilsner, J.R., Ilievski, V., Factor-Litvak, P., Levy, D.,

85
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

Alam, S., Islam, M., Parvez, F., Ahsan, H., Graziano, J.H., 2007. Folic acid supple- arsenicals–oxidative stress theory of arsenic carcinogenesis. Toxicol. Appl.
mentation lowers blood arsenic. Am. J. Clin. Nutr. 86 (4), 1202–1209. Pharmacol. 232 (2), 252–257.
Gao, Y., Yin, Y., Xing, X., Zhao, Z., Lu, Y., Sun, Y., Zhuang, Z., Wang, M., Ji, W., He, Y., Klein, C.B., Leszczynska, J., Hickey, C., Rossman, T.G., 2007. Further evidence against a
2016. Arsenic-induced anti-angiogenesis via miR-425-5p-regulated CCM3. Toxicol. direct genotoxic mode of action for arsenic-induced cancer. Toxicol. Appl.
Lett. 254, 22–31. Pharmacol. 222 (3), 289–297.
Garcia-Esquinas, E., Pollan, M., Umans, J.G., Francesconi, K.A., Goessler, W., Guallar, E., Kligerman, A.D., Doerr, C.L., Tennant, A.H., Harrington-Brock, K., Allen, J.W., Winkfield,
Howard, B., Farley, J., Best, L.G., Navas-Acien, A., 2013. Arsenic exposure and cancer E., Poorman-Allen, P., Kundu, B., Funasaka, K., Roop, B.C., Mass, M.J., DeMarini,
mortality in a US-based prospective cohort: the strong heart study. Canc. Epidemiol. D.M., 2003. Methylated trivalent arsenicals as candidate ultimate genotoxic forms of
Biomarkers Prev. 22 (11), 1944–1953. arsenic: induction of chromosomal mutations but not gene mutations. Environ. Mol.
Gentry, P.R., McDonald, T.B., Sullivan, D.E., Shipp, A.M., Yager, J.W., Clewell 3rd, H.J., Mutagen. 42 (3), 192–205.
2010. Analysis of genomic dose-response information on arsenic to inform key events Kouzarides, T., 2007. Chromatin modifications and their function. Cell (4), 693–705.
in a mode of action for carcinogenicity. Environ. Mol. Mutagen. 51 (1), 1–14. Kumagai, Y., Sumi, D., 2007. Arsenic: signal transduction, transcription factor, and bio-
Glozak, M.A., Seto, E., 2007. Histone deacetylases and cancer. Oncogene 26 (37), transformation involved in cellular response and toxicity. Annu. Rev. Pharmacol.
5420–5432. Toxicol. 47, 243–262.
Gonsebatt, M.E., Vega, L., Salazar, A.M., Montero, R., Guzman, P., Blas, J., Del Razo, L.M., Lamm, S.H., Engel, A., Kruse, M.B., Feinleib, M., Byrd, D.M., Lai, S., Wilson, R., 2004.
Garcia-Vargas, G., Albores, A., Cebrian, M.E., Kelsh, M., Ostrosky-Wegman, P., 1997. Arsenic in drinking water and bladder cancer mortality in the United States: an
Cytogenetic effects in human exposure to arsenic. Mutat. Res. 386 (3), 219–228. analysis based on 133 U.S. counties and 30 years of observation. J. Occup. Environ.
Gonzalez, H., Lema, C., Kirken, R.A., Maldonado, R.A., Varela-Ramirez, A., Aguilera, R.J., Med. 46 (3), 298–306.
2015. Arsenic-exposed keratinocytes exhibit differential microRNAs expression pro- Lamm, S.H., Ferdosi, H., Dissen, E.K., Li, J., Ahn, J., 2015. A systematic review and meta-
file; potential implication of miR-21, miR-200a and miR-141 in melanoma pathway. regression analysis of lung cancer risk and inorganic arsenic in drinking water. Int. J.
Clin Cancer Drugs 2 (2), 138–147. Environ. Res. Publ. Health 12 (12), 15498–15515.
Gu, S., Chen, C., Jiang, X., Zhang, Z., 2016. Role of nuclear factor E2 related factor 2 Leonardi, G., Vahter, M., Clemens, F., Goessler, W., Gurzau, E., Hemminki, K., Hough, R.,
signaling pathway in malignant transformation induced by sodium arsenite. Wei Koppova, K., Kumar, R., Rudnai, P., Surdu, S., Fletcher, T., 2012. Inorganic arsenic
Sheng Yan Jiu 45 (3), 356–361. and basal cell carcinoma in areas of Hungary, Romania, and Slovakia: a case-control
Guo, H.R., 2004. Arsenic level in drinking water and mortality of lung cancer (Taiwan). study. Environ. Health Perspect. 120 (5), 721–726.
Cancer Causes Control 15 (2), 171–177. Lerda, D., 1994. Sister-chromatid exchange (SCE) among individuals chronically exposed
Haque, R., Chaudhary, A., Sadaf, N., 2017. Immunomodulatory role of arsenic in reg- to arsenic in drinking water. Mutat. Res. 312 (2), 111–120.
ulatory T cells. Endocr. Metab. Immune Disord. - Drug Targets 17 (3), 176–181. Lewis, B.P., Burge, C.B., Bartel, D.P., 2005. Conserved seed pairing, often flanked by
Haque, R., Mazumder, D.N., Samanta, S., Ghosh, N., Kalman, D., Smith, M.M., Mitra, S., adenosines, indicates that thousands of human genes are microRNA targets. Cell 120
Santra, A., Lahiri, S., Das, S., De, B.K., Smith, A.H., 2003. Arsenic in drinking water (1), 15–20.
and skin lesions: dose-response data from West Bengal, India. Epidemiology 14 (2), Li, J., Gorospe, M., Barnes, J., Liu, Y., 2003. Tumor promoter arsenite stimulates histone
174–182. H3 phosphoacetylation of proto-oncogenes c-fos and c-jun chromatin in human di-
He, L., Hannon, G.J., 2004. MicroRNAs: small RNAs with a big role in gene regulation. ploid fibroblasts. J. Biol. Chem. 278 (15), 13183–13191.
Nat. Rev. Genet. 5 (7), 522–531. Liaw, J., Marshall, G., Yuan, Y., Ferreccio, C., Steinmaus, C., Smith, A.H., 2008. Increased
He, L., Thomson, J.M., Hemann, M.T., Hernando-Monge, E., Mu, D., Goodson, S., Powers, childhood liver cancer mortality and arsenic in drinking water in northern Chile.
S., Cordon-Cardo, C., Lowe, S.W., Hannon, G.J., Hammond, S.M., 2005. A microRNA Canc. Epidemiol. Biomarkers Prev. 17 (8), 1982–1987.
polycistron as a potential human oncogene. Nature 435 (7043), 828–833. Lindberg, H.K., Wang, X., Jarventaus, H., Falck, G.C., Norppa, H., Fenech, M., 2007.
Hei, T.K., Filipic, M., 2004. Role of oxidative damage in the genotoxicity of arsenic. Free Origin of nuclear buds and micronuclei in normal and folate-deprived human lym-
Radic. Biol. Med. 37 (5), 574–581. phocytes. Mutat. Res. 617 (1–2), 33–45.
Herbert, K.J., Snow, E.T., 2012. Modulation of arsenic-induced epidermal growth factor Liu, D., Wu, D., Zhao, L., Yang, Y., Ding, J., Dong, L., Hu, L., Wang, F., Zhao, X., Cai, Y.,
receptor pathway signalling by resveratrol. Chem. Biol. Interact. 198 (1–3), 38–48. Jin, J., 2015. Arsenic trioxide reduces global histone H4 acetylation at lysine 16
Herceg, Z., 2007. Epigenetics and cancer: towards an evaluation of the impact of en- through direct binding to histone acetyltransferase hMOF in human cells. PLoS One
vironmental and dietary factors. Mutagenesis 22 (2), 91–103. 10 (10), e0141014.
Hossain, M.B., Vahter, M., Concha, G., Broberg, K., 2012. Environmental arsenic exposure Liu, J., Chen, B., Lu, Y., Guan, Y., Chen, F., 2012. JNK-dependent Stat3 phosphorylation
and DNA methylation of the tumor suppressor gene p16 and the DNA repair gene contributes to Akt activation in response to arsenic exposure. Toxicol. Sci. 129 (2),
MLH1: effect of arsenic metabolism and genotype. Metallomics 4 (11), 1167–1175. 363–371.
Houben, A., Demidov, D., Caperta, A.D., Karimi, R., Agueci, F., Vlasenko, L., 2007. Liu, J., Waalkes, M.P., 2008. Liver is a target of arsenic carcinogenesis. Toxicol. Sci. 105
Phosphorylation of histone H3 in plants–a dynamic affair. Biochim. Biophys. Acta (1), 24–32.
1769 (5–6), 308–315. Liu, S., Sun, Q., Wang, F., Zhang, L., Song, Y., Xi, S., Sun, G., 2014. Arsenic induced
Huang, C., Ke, Q., Costa, M., Sh, X., 2004. Molecular mechanisms of arsenic carcino- overexpression of inflammatory cytokines based on the human urothelial cell model
genesis. Mol. Cell. Biochem. 255 (1–2), 57–66. in vitro and urinary secretion of individuals chronically exposed to arsenic. Chem.
Hubaux, R., Becker-Santos, D.D., Enfield, K.S., Rowbotham, D., Lam, S., Lam, W.L., Res. Toxicol. 27 (11), 1934–1942.
Martinez, V.D., 2013. Molecular features in arsenic-induced lung tumors. Mol. Canc. Loenen, W.A., 2006. S-adenosylmethionine: jack of all trades and master of everything?
12, 20. Biochem. Soc. Trans. 34 (Pt 2), 330–333.
Hughes, M.F., Beck, B.D., Chen, Y., Lewis, A.S., Thomas, D.J., 2011. Arsenic exposure and Lu, G., Xu, H., Chang, Wu, Z., Yao, X., Zhang, S., Li, Z., Bai, J., Cai, Q., Zhang, W., 2014.
toxicology: a historical perspective. Toxicol. Sci. 123 (2), 305–332. Arsenic exposure is associated with DNA hypermethylation of the tumor suppressor
Humans, I. W. G. o. t. E. o. C. R. t., 2004a. Betel-quid and areca-nut chewing and some gene p16. J. Occup. Med. Toxicol. 9 (1), 42.
areca-nut derived nitrosamines. IARC Monogr. Eval. Carcinog. Risks Hum. 85, 1–334. Luo, F., Ji, J., Liu, Y., Xu, Y., Zheng, G., Jing, J., Wang, B., Xu, W., Shi, L., Lu, X., Liu, Q.,
Humans, I. W. G. o. t. E. o. C. R. t., 2004b. Some drinking-water disinfectants and con- 2015. MicroRNA-21, up-regulated by arsenite, directs the epithelial-mesenchymal
taminants, including arsenic. Monographs on chloramine, chloral and chloral hy- transition and enhances the invasive potential of transformed human bronchial epi-
drate, dichloroacetic acid, trichloroacetic acid and 3-chloro-4-(dichloromethyl)-5- thelial cells by targeting PDCD4. Toxicol. Lett. 232 (1), 301–309.
hydroxy-2(5H)-furanone. IARC Monogr. Eval. Carcinog. Risks Hum. 84, 269–477. Lv, L., An, X., Li, H., Ma, L., 2016. Effect of miR-155 knockdown on the reversal of
Humans, I. W. G. o. t. E. o. C. R. t., 2012. Arsenic, metals, fibres, and dusts. IARC Monogr. doxorubicin resistance in human lung cancer A549/dox cells. Oncol Lett 11 (2),
Eval. Carcinog. Risks Hum. 100 (Pt C), 11–465. 1161–1166.
Jensen, T.J., Novak, P., Eblin, K.E., Gandolfi, A.J., Futscher, B.W., 2008. Epigenetic re- Ma, L., Li, J., Zhan, Z., Chen, L., Li, D., Bai, Q., Gao, C., Li, J., Zeng, X., He, Z., Wang, S.,
modeling during arsenical-induced malignant transformation. Carcinogenesis 29 (8), Xiao, Y., Chen, W., Zhang, A., 2016. Specific histone modification responds to ar-
1500–1508. senic-induced oxidative stress. Toxicol. Appl. Pharmacol. 302, 52–61.
International Agency for Research on Cancer, 2004. Some drinkingwater disinfectants Mahata, J., Basu, A., Ghoshal, S., Sarkar, J.N., Roy, A.K., Poddar, G., Nandy, A.K.,
and contaminants, including arsenic. IARC Monogr. Eval. Carcinog. Risks Hum. 84, Banerjee, A., Ray, K., Natarajan, A.T., Nilsson, R., Giri, A.K., 2003. Chromosomal
39–229. aberrations and sister chromatid exchanges in individuals exposed to arsenic through
International Agency for Research on Cancer, 2012. Arsenic, metals, fibres, and dusts. drinking water in West Bengal, India. Mutat. Res. 534 (1–2), 133–143.
IARC Monogr. Eval. Carcinog. Risks Hum. 100, 41–93. Maki-Paakkanen, J., Kurttio, P., Paldy, A., Pekkanen, J., 1998. Association between the
Jo, W.J., Ren, X., Chu, F., Aleshin, M., Wintz, H., Burlingame, A., Smith, M.T., Vulpe, clastogenic effect in peripheral lymphocytes and human exposure to arsenic through
C.D., Zhang, L., 2009. Acetylated H4K16 by MYST1 protects UROtsa cells from ar- drinking water. Environ. Mol. Mutagen. 32 (4), 301–313.
senic toxicity and is decreased following chronic arsenic exposure. Toxicol. Appl. Marquez-Garban, D.C., Chen, H.W., Fishbein, M.C., Goodglick, L., Pietras, R.J., 2007.
Pharmacol. 241 (3), 294–302. Estrogen receptor signaling pathways in human non-small cell lung cancer. Steroids
Kelsey, K.T., Hirao, T., Hirao, S., Devi-Ashok, T., Nelson, H.H., Andrew, A., Colt, J., Baris, 72 (2), 135–143.
D., Morris, J.S., Schned, A., Karagas, M., 2005. TP53 alterations and patterns of Marshall, G., Ferreccio, C., Yuan, Y., Bates, M.N., Steinmaus, C., Selvin, S., Liaw, J.,
carcinogen exposure in a U.S. population-based study of bladder cancer. Int. J. Canc. Smith, A.H., 2007. Fifty-year study of lung and bladder cancer mortality in Chile
117 (3), 370–375. related to arsenic in drinking water. J. Natl. Cancer Inst. 99 (12), 920–928.
Kesari, V.P., Kumar, A., Khan, P.K., 2012. Genotoxic potential of arsenic at its reference Marsit, C.J., Eddy, K., Kelsey, K.T., 2006. MicroRNA responses to cellular stress. Canc.
dose. Ecotoxicol. Environ. Saf. 80, 126–131. Res. 66 (22), 10843–10848.
Khairul, I., Wang, Q.Q., Jiang, Y.H., Wang, C., Naranmandura, H., 2017. Metabolism, Mass, M.J., Tennant, A., Roop, B.C., Cullen, W.R., Styblo, M., Thomas, D.J., Kligerman,
toxicity and anticancer activities of arsenic compounds. Oncotarget 8 (14), A.D., 2001. Methylated trivalent arsenic species are genotoxic. Chem. Res. Toxicol.
23905–23926. 14 (4), 355–361.
Kitchin, K.T., Wallace, K., 2008. Evidence against the nuclear in situ binding of Meliker, J.R., Slotnick, M.J., AvRuskin, G.A., Schottenfeld, D., Jacquez, G.M., Wilson,

86
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

M.L., Goovaerts, P., Franzblau, A., Nriagu, J.O., 2010. Lifetime exposure to arsenic in Sankpal, U.T., Pius, H., Khan, M., Shukoor, M.I., Maliakal, P., Lee, C.M., Abdelrahim, M.,
drinking water and bladder cancer: a population-based case-control study in Connelly, S.F., Basha, R., 2012. Environmental factors in causing human cancers:
Michigan, USA. Cancer Causes Control 21 (5), 745–757. emphasis on tumorigenesis. Tumour Biol 33 (5), 1265–1274.
Meliker, J.R., Wahl, R.L., Cameron, L.L., Nriagu, J.O., 2007. Arsenic in drinking water Schoen, A., Beck, B., Sharma, R., Dube, E., 2004. Arsenic toxicity at low doses: epide-
and cerebrovascular disease, diabetes mellitus, and kidney disease in Michigan: a miological and mode of action considerations. Toxicol. Appl. Pharmacol. 198 (3),
standardized mortality ratio analysis. Environ. Health 6, 4. 253–267.
Mendez Jr., W.M., Eftim, S., Cohen, J., Warren, I., Cowden, J., Lee, J.S., Sams, R., 2017. Siegfried, J.M., Stabile, L.P., 2014. Estrongenic steroid hormones in lung cancer. Semin.
Relationships between arsenic concentrations in drinking water and lung and bladder Oncol. 41 (1), 5–16.
cancer incidence in U.S. counties. J. Expo. Sci. Environ. Epidemiol. 27 (3), 235–243. Simeonova, P.P., Luster, M.I., 2002. Arsenic carcinogenicity: relevance of c-Src activation.
Miao, Z., Wu, L., Lu, M., Meng, X., Gao, B., Qiao, X., Zhang, W., Xue, D., 2015. Analysis of Mol. Cell. Biochem. 234–235 (1–2), 277–282.
the transcriptional regulation of cancer-related genes by aberrant DNA methylation Sinha, D., Roy, M., 2011. Antagonistic role of tea against sodium arsenite-induced oxi-
of the cis-regulation sites in the promoter region during hepatocyte carcinogenesis dative DNA damage and inhibition of DNA repair in Swiss albino mice. J. Environ.
caused by arsenic. Oncotarget 6 (25), 21493–21506. Pathol. Toxicol. Oncol. 30 (4), 311–322.
Mink, P.J., Alexander, D.D., Barraj, L.M., Kelsh, M.A., Tsuji, J.S., 2008. Low-level arsenic Smith, A.H., Marshall, G., Liaw, J., Yuan, Y., Ferreccio, C., Steinmaus, C., 2012. Mortality
exposure in drinking water and bladder cancer: a review and meta-analysis. Regul. in young adults following in utero and childhood exposure to arsenic in drinking
Toxicol. Pharmacol. 52 (3), 299–310. water. Environ. Health Perspect. 120 (11), 1527–1531.
Moe, B., Peng, H., Lu, X., Chen, B., Chen, L.W.L., Gabos, S., Li, X.F., Le, X.C., 2016. Srivastava, R.K., Li, C., Chaudhary, S.C., Ballestas, M.E., Elmets, C.A., Robbins, D.J.,
Comparative cytotoxicity of fourteen trivalent and pentavalent arsenic species de- Matalon, S., Deshane, J.S., Afaq, F., Bickers, D.R., Athar, M., 2013. Unfolded protein
termined using real-time cell sensing. J. Environ. Sci. (China) 49, 113–124. response (UPR) signaling regulates arsenic trioxide-mediated macrophage innate
Morales, K.H., Ryan, L., Kuo, T.L., Wu, M.M., Chen, C.J., 2000. Risk of internal cancers immune function disruption. Toxicol. Appl. Pharmacol. 272 (3), 879–887.
from arsenic in drinking water. Environ. Health Perspect. 108 (7), 655–661. Stabile, L.P., Lyker, J.S., Gubish, C.T., Zhang, W., Grandis, J.R., Siegfried, J.M., 2005.
Naranmandura, H., Suzuki, N., Suzuki, K.T., 2006. Trivalent arsenicals are bound to Combined targeting of the estrogen receptor and the epidermal growth factor re-
proteins during reductive methylation. Chem. Res. Toxicol. 19 (8), 1010–1018. ceptor in non-small cell lung cancer shows enhanced antiproliferative effects. Canc.
National Research Council (US), 1999. Arsenic in Drinking Water. Washington (DC). Res. 65 (4), 1459–1470.
National Research Council (US), 2001. Arsenic in Drinking Water: 2001 Update. States, J.C., 2015. Disruption of mitotic progression by arsenic. Biol. Trace Elem. Res. 166
Washington (DC). (1), 34–40.
Naujokas, M.F., Anderson, B., Ahsan, H., Aposhian, H.V., Graziano, J.H., Thompson, C., Steinmaus, C., Ferreccio, C., Acevedo, J., Yuan, Y., Liaw, J., Duran, V., Cuevas, S., Garcia,
Suk, W.A., 2013. The broad scope of health effects from chronic arsenic exposure: J., Meza, R., Valdes, R., Valdes, G., Benitez, H., VanderLinde, V., Villagra, V., Cantor,
update on a worldwide public health problem. Environ. Health Perspect. 121 (3), K.P., Moore, L.E., Perez, S.G., Steinmaus, S., Smith, A.H., 2014. Increased lung and
295–302. bladder cancer incidence in adults after in utero and early-life arsenic exposure. Canc.
Nemeti, B., Poor, M., Gregus, Z., 2015. Reduction of the pentavalent arsenical dimethy- Epidemiol. Biomarkers Prev. 23 (8), 1529–1538.
larsinic acid and the GSTO1 substrate S-(4-Nitrophenacyl)glutathione by rat liver Stueckle, T.A., Lu, Y., Davis, M.E., Wang, L., Jiang, B.H., Holaskova, I., Schafer, R.,
cytosol: analyzing the role of GSTO1 in arsenic reduction. Chem. Res. Toxicol. 28 Barnett, J.B., Rojanasakul, Y., 2012. Chronic occupational exposure to arsenic in-
(11), 2199–2209. duces carcinogenic gene signaling networks and neoplastic transformation in human
Nesnow, S., Roop, B.C., Lambert, G., Kadiiska, M., Mason, R.P., Cullen, W.R., Mass, M.J., lung epithelial cells. Toxicol. Appl. Pharmacol. 261 (2), 204–216.
2002. DNA damage induced by methylated trivalent arsenicals is mediated by re- Suarez, Y., Fernandez-Hernando, C., Pober, J.S., Sessa, W.C., 2007. Dicer dependent
active oxygen species. Chem. Res. Toxicol. 15 (12), 1627–1634. microRNAs regulate gene expression and functions in human endothelial cells. Circ.
Okoji, R.S., Yu, R.C., Maronpot, R.R., Froines, J.R., 2002. Sodium arsenite administration Res. 100 (8), 1164–1173.
via drinking water increases genome-wide and Ha-ras DNA hypomethylation in Sung, T.I., Wang, Y.J., Chen, C.Y., Hung, T.L., Guo, H.R., 2012. Increased serum level of
methyl-deficient C57BL/6J mice. Carcinogenesis 23 (5), 777–785. epidermal growth factor receptor in liver cancer patients and its association with
Ostrosky-Wegman, P., Gonsebatt, M.E., Montero, R., Vega, L., Barba, H., Espinosa, J., exposure to arsenic. Sci. Total Environ. 424, 74–78.
Palao, A., Cortinas, C., Garcia-Vargas, G., del Razo, L.M., et al., 1991. Lymphocyte Suzuki, K.T., Mandal, B.K., Ogra, Y., 2002. Speciation of arsenic in body fluids. Talanta 58
proliferation kinetics and genotoxic findings in a pilot study on individuals chroni- (1), 111–119.
cally exposed to arsenic in Mexico. Mutat. Res. 250 (1–2), 477–482. Tian, D., Ma, H., Feng, Z., Xia, Y., Le, X.C., Ni, Z., Allen, J., Collins, B., Schreinemachers,
Peters, B.A., Hall, M.N., Liu, X., Parvez, F., Sanchez, T.R., van Geen, A., Mey, J.L., D., Mumford, J.L., 2001. Analyses of micronuclei in exfoliated epithelial cells from
Siddique, A.B., Shahriar, H., Uddin, M.N., Islam, T., Balac, O., Ilievski, V., Factor- individuals chronically exposed to arsenic via drinking water in inner Mongolia,
Litvak, P., Graziano, J.H., Gamble, M.V., 2015. Folic acid and creatine as therapeutic China. J. Toxicol. Environ. Health 64 (6), 473–484.
approaches to lower blood arsenic: a randomized controlled trial. Environ. Health Tseng, C.H., Huang, Y.K., Huang, Y.L., Chung, C.J., Yang, M.H., Chen, C.J., Hsueh, Y.M.,
Perspect. 123 (12), 1294–1301. 2005. Arsenic exposure, urinary arsenic speciation, and peripheral vascular disease in
Peterson, C.L., Laniel, M.A., 2004. Histones and histone modifications. Curr. Biol. 14 (14), Blackfoot disease-hyperendemic villages in Taiwan. Toxicol. Appl. Pharmacol. 206
R546–R551. (3), 299–308.
Pietras, R.J., Marquez-Garban, D.C., 2007. Membrane-associated estrogen receptor sig- Tsuda, T., Babazono, A., Yamamoto, E., Kurumatani, N., Mino, Y., Ogawa, T., Kishi, Y.,
naling pathways in human cancers. Clin. Canc. Res. 13 (16), 4672–4676. Aoyama, H., 1995. Ingested arsenic and internal cancer: a historical cohort study
Platanias, L.C., 2009. Biological responses to arsenic compounds. J. Biol. Chem. 284 (28), followed for 33 years. Am. J. Epidemiol. 141 (3), 198–209.
18583–18587. Tsuji, J.S., Alexander, D.D., Perez, V., Mink, P.J., 2014. Arsenic exposure and bladder
Prasad, P., Sinha, D., 2017. Low-level arsenic causes chronic inflammation and suppresses cancer: quantitative assessment of studies in human populations to detect risks at low
expression of phagocytic receptors. Environ. Sci. Pollut. Res. Int. 24 (12), doses. Toxicology 317, 17–30.
11708–11721. Uthus, E.O., Davis, C., 2005. Dietary arsenic affects dimethylhydrazine-induced aberrant
Pratheeshkumar, P., Son, Y.O., Divya, S.P., Wang, L., Zhang, Z., Shi, X., 2016. Oncogenic crypt formation and hepatic global DNA methylation and DNA methyltransferase
transformation of human lung bronchial epithelial cells induced by arsenic involves activity in rats. Biol. Trace Elem. Res. 103 (2), 133–145.
ROS-dependent activation of STAT3-miR-21-PDCD4 mechanism. Sci. Rep. 6, 37227. Vahter, M., 2002. Mechanisms of arsenic biotransformation. Toxicology 181–182,
Rao, C.V., Pal, S., Mohammed, A., Farooqui, M., Doescher, M.P., Asch, A.S., Yamada, 211–217.
H.Y., 2017. Biological effects and epidemiological consequences of arsenic exposure, Valenzuela, O.L., Borja-Aburto, V.H., Garcia-Vargas, G.G., Cruz-Gonzalez, M.B., Garcia-
and reagents that can ameliorate arsenic damage in vivo. Oncotarget 8 (34), Montalvo, E.A., Calderon-Aranda, E.S., Del Razo, L.M., 2005. Urinary trivalent me-
57605–57621. thylated arsenic species in a population chronically exposed to inorganic arsenic.
Ray, P.D., Huang, B.W., Tsuji, Y., 2015. Coordinated regulation of Nrf2 and histone H3 Environ. Health Perspect. 113 (3), 250–254.
serine 10 phosphorylation in arsenite-activated transcription of the human heme Waalkes, M.P., Liu, J., Chen, H., Xie, Y., Achanzar, W.E., Zhou, Y.S., Cheng, M.L., Diwan,
oxygenase-1 gene. Biochim. Biophys. Acta 1849 (10), 1277–1288. B.A., 2004. Estrogen signaling in livers of male mice with hepatocellular carcinoma
Reali, D., Di Marino, F., Bahramandpour, S., Carducci, A., Barale, R., Loprieno, N., 1987. induced by exposure to arsenic in utero. J. Natl. Cancer Inst. 96 (6), 466–474.
Micronuclei in exfoliated urothelial cells and urine mutagenicity in smokers. Mutat. Waalkes, M.P., Liu, J., Ward, J.M., Powell, D.A., Diwan, B.A., 2006. Urogenital carcino-
Res. 192 (2), 145–149. genesis in female CD1 mice induced by in utero arsenic exposure is exacerbated by
Reichard, J.F., Schnekenburger, M., Puga, A., 2007. Long term low-dose arsenic exposure postnatal diethylstilbestrol treatment. Canc. Res. 66 (3), 1337–1345.
induces loss of DNA methylation. Biochem. Biophys. Res. Commun. 352 (1), Wang, F., Liu, S., Xi, S., Yan, L., Wang, H., Song, Y., Sun, G., 2013a. Arsenic induces the
188–192. expressions of angiogenesis-related factors through PI3K and MAPK pathways in SV-
Ren, X., McHale, C.M., Skibola, C.F., Smith, A.H., Smith, M.T., Zhang, L., 2011. An HUC-1 human uroepithelial cells. Toxicol. Lett. 222 (3), 303–311.
emerging role for epigenetic dysregulation in arsenic toxicity and carcinogenesis. Wang, T.S., Hsu, T.Y., Chung, C.H., Wang, A.S., Bau, D.T., Jan, K.Y., 2001. Arsenite in-
Environ. Health Perspect. 119 (1), 11–19. duces oxidative DNA adducts and DNA-protein cross-links in mammalian cells. Free
Roh, T., Lynch, C.F., Weyer, P., Wang, K., Kelly, K.M., Ludewig, G., 2017. Low-level ar- Radic. Biol. Med. 31 (3), 321–330.
senic exposure from drinking water is associated with prostate cancer in Iowa. Wang, W., Cheng, S., Zhang, D., 2014. Association of inorganic arsenic exposure with
Environ. Res. 159, 338–343. liver cancer mortality: a meta-analysis. Environ. Res. 135, 120–125.
Rosen, B.P., 2002. Transport and detoxification systems for transition metals, heavy Wang, Y.H., Yeh, S.D., Wu, M.M., Liu, C.T., Shen, C.H., Shen, K.H., Pu, Y.S., Hsu, L.I.,
metals and metalloids in eukaryotic and prokaryotic microbes. Comp. Biochem. Chiou, H.Y., Chen, C.J., 2013b. Comparing the joint effect of arsenic exposure, ci-
Physiol. Mol. Integr. Physiol. 133 (3), 689–693. garette smoking and risk genotypes of vascular endothelial growth factor on upper
Roy, P., Mukherjee, A., Giri, S., 2016. Evaluation of genetic damage in tobacco and ar- urinary tract urothelial carcinoma and bladder cancer. J. Hazard Mater. 262,
senic exposed population of Southern Assam, India using buccal cytome assay and 1139–1146.
comet assay. Ecotoxicol. Environ. Saf. 124, 169–176. Wang, Z., Yang, J., Fisher, T., Xiao, H., Jiang, Y., Yang, C., 2012. Akt activation is

87
Q. Zhou, S. Xi Regulatory Toxicology and Pharmacology 99 (2018) 78–88

responsible for enhanced migratory and invasive behavior of arsenic-transformed 112 (12), 1255–1263.
human bronchial epithelial cells. Environ. Health Perspect. 120 (1), 92–97. Xue, X., Liu, Y., Wang, Y., Meng, M., Wang, K., Zang, X., Zhao, S., Sun, X., Cui, L., Pan, L.,
Wei, B., Yu, J., Kong, C., Li, H., Yang, L., Guo, Z., Cui, N., Xia, Y., Wu, K., 2017. An Liu, S., 2016. MiR-21 and MiR-155 promote non-small cell lung cancer progression
investigation of the health effects caused by exposure to arsenic from drinking water by downregulating SOCS1, SOCS6, and PTEN. Oncotarget 7 (51), 84508–84519.
and coal combustion: arsenic exposure and metabolism. Environ. Sci. Pollut. Res. Int. Yang, H.C., Fu, H.L., Lin, Y.F., Rosen, B.P., 2012. Pathways of arsenic uptake and efflux.
24 (33), 25947–25954. Curr. Top. Membr. 69, 325–358.
Wei, M., Arnold, L., Cano, M., Cohen, S.M., 2005. Effects of co-administration of anti- Zhang, A., Feng, H., Yang, G., Pan, X., Jiang, X., Huang, X., Dong, X., Yang, D., Xie, Y.,
oxidants and arsenicals on the rat urinary bladder epithelium. Toxicol. Sci. 83 (2), Peng, L., Jun, L., Hu, C., Jian, L., Wang, X., 2007a. Unventilated indoor coal-fired
237–245. stoves in Guizhou province, China: cellular and genetic damage in villagers exposed
Wysocka, J., Allis, C.D., Coonrod, S., 2006. Histone arginine methylation and its dynamic to arsenic in food and air. Environ. Health Perspect. 115 (4), 653–658.
regulation. Front. Biosci. 11, 344–355. Zhang, A.H., Bin, H.H., Pan, X.L., Xi, X.G., 2007b. Analysis of p16 gene mutation, deletion
Xie, H., Huang, S., Martin, S., Wise Sr., J.P., 2014. Arsenic is cytotoxic and genotoxic to and methylation in patients with arseniasis produced by indoor unventilated-stove
primary human lung cells. Mutat. Res. Genet. Toxicol. Environ. Mutagen 760, 33–41. coal usage in Guizhou, China. J. Toxicol. Environ. Health 70 (11), 970–975.
Xie, Y., Trouba, K.J., Liu, J., Waalkes, M.P., Germolec, D.R., 2004. Biokinetics and sub- Zhou, X., Sun, H., Ellen, T.P., Chen, H., Costa, M., 2008. Arsenite alters global histone H3
chronic toxic effects of oral arsenite, arsenate, monomethylarsonic acid, and di- methylation. Carcinogenesis 29 (9), 1831–1836.
methylarsinic acid in v-Ha-ras transgenic (Tg.AC) mice. Environ. Health Perspect.

88

You might also like