Skip to main content
Access keys NCBI Homepage MyNCBI Homepage Main Content Main Navigation
Curr Protoc Toxicol. Author manuscript; available in PMC 2014 Nov 21.
Published in final edited form as:
PMCID: PMC4240622
NIHMSID: NIHMS161022
PMID: 25419261

An Overview of Arsenic Metabolism and Toxicity

Abstract

It is likely that at least some of the toxic and carcinogenic effects associated with exposure to inorganic arsenic are, in fact, due to actions of its methylated metabolites. Here, we provide an overview of current models for the biological methylation of arsenicals. This information provides a context for understanding the chemical, biochemical, and genetic approaches to elucidation of the formation and function of metahylated arsenicals which are presented in the following manuscripts.

Introduction

The toxic properties of arsenic (As) were recognized long before Albertus Magnus in the 13th century prepared its elemental form (Buchanan, 1962). Its use as a poison has played lethal and decisive roles in domestic and dynastic intrigues throughout history (Cullen, 2008). Inorganic As (iAs) remained a mainstay of the poisoner’s art until methods for its detection were developed in the 19th century (Jolliffe, 1993). The potency of As as a carcinogen in humans has also been apparent since the beginning of its use in industrial processes. Studies in workers exposed to iAs as arsenite (iAsIII) or arsenate (iAsV) have found increased prevalences of various internal cancers (Enterline et al., 1995; Lubin et al., 2000). In addition, arsenicals have a long history of medicinal use. Fowler’s solution (an alcoholic solution of potassium arsenite) was widely used for several centuries as a tonic for dermatological, hematological, and respiratory ailments. Synthesis of Salvarsan (arsenphenamine, compound 606) by Paul Ehrlich and its use as a treatment for syphilis marked the beginning of modern chemotherapeutics (Schwarz, 2004). Organoarsenicals were widely used as antibiotics until the 1940s when they were displaced by sulfa drugs or penicillin and other microbial antibiotics. In recent years, iAs as arsenic trioxide has returned to the therapeutic armamentarium as a treatment for acute promyelocytic leukemia (Douer and Tallman, 2005).

In the past half century, iAs has emerged as a significant environmental contaminant. Exposure to iAs as a contaminant of drinking water was first recognized as a major public health problem in Taiwan. In certain regions of that island, individuals who chronically consumed drinking water containing iAs displayed a range of skin and vascular lesions that were attributed to As exposure (Tseng et al., 1968; Tseng, 1977). Studies in Taiwan and elsewhere have demonstrated that chronic exposure is associated with increased risk of vascular disease, diabetes, and internal cancers (Smith et al., 1992; Navas-Acien et al., 2005, 2008; Chen et al., 2007; Wang et al., 2007). Through the years, other populations elsewhere in the world have also been identified as chronic users of drinking water supplies that are contaminated with iAs. There is abundant evidence of adverse health effects in all of these populations. Indeed, the plight of tens of millions of individuals exposed to iAs in Bangladesh and West Bengal has been described as a “public health emergency” (Smith et al., 2000) and as “the largest poisoning of a population in history” (Bhattacharjee, 2007), assessments that are difficult to discount given the number of affected individuals.

Given the strong epidemiological evidence of myriad adverse health effects that are attributable to chronic exposure to iAs, there has been a growing interest in understanding of the distribution and metabolism of this metalloid in humans. Coupled with a renewed interest in potential modes of actions of arsenicals as toxicants or carcinogens, this has lead to a rapid expansion of our knowledge in these fields. In this chapter, attention is devoted to the study of metabolism of As in in vitro assays and in cultured cells. In particular, the enzymatic basis of As methylation is considered. In the following paragraphs, we provide a brief summary of current models for the biological methylation of arsenicals.

The nomenclature for arsenicals used here follows common usage in the toxicological literature. Because the oxidation state of arsenic in various compounds is of interest, this distinction is indicated in the naming of compounds. Systematic names and abbreviations used for arsenicals: arsenite (arsenous acid) AsIII(OH)3, iAsIII; arsenate (arsenic acid) AsV(O)(OH)3, iAsV; (mono)methylarsonous acid CH3AsIII(OH)2, MAsIII; (mono)methylarsonic acid CH3AsV(O)(OH)2, MAsV; dimethylarsinous acid (CH3)2 AsIIIOH, DMAsIII; dimethylarsinic acid (CH3)2 AsV(O)OH, DMAsV; trimethylarsine oxide, (CH3)3AsV(O), TMAO. In cases where the oxidation state of arsenic cannot be determined or is not germane, the generic abbreviations iAs, MAs, and DMAs are used

General Aspects of Arsenic Methylation

Elucidation of the pathway for As methylation has its roots in studies in the 19th century that demonstrated that microorganisms could convert iAs to a volatile methylated product (Gosio gas, trimethylarsine). Studies of metabolism of As in microbes culminated in two comprehensive reviews of biological methylation by Frederick Challenger (Challenger, 1945, 1951) in which a pathway with steps leading from iAs to methylated products was proposed (Figure 1a). In this scheme, As in the pentavalent oxidation state is reduced to trivalency and the resulting trivalent arsenical then undergoes oxidative methylation. This model has proven sufficient to explain much of the available data on the formation of methylated arsenicals in biological systems. In subsequent years, much effort has been expended understanding the mechanistic basis for reduction of pentavalent As and for oxidative methylation of trivalent As (see Cullen et al., 1984a, b). An alternative model for As methylation (Figure 1b) has been proposed (Hayakawa et al., 2005, Naramandura et al., 2006). Here, As bound to a cellular thiol is the substrate for methylation. This thiol group could be a component of a cellular protein or the reactive moiety in glutathione (GSH), the most abundant low molecular weight thiol compound in cells. In this model, methylation of trivalent As is not oxidative; formation of pentavalent arsenicals is attributed to oxidation of As after hydrolysis of As-thiol complexes. Although this alternative model is consistent with many experimental details of the methylation process, there are data suggesting that the presence of thiol species is not an absolute requirement for biological methylation of As (see Thomas et al., 2007).

An external file that holds a picture, illustration, etc.
Object name is nihms161022f1.jpg

Two proposed pathways for the conversion of inorganic arsenic into mono-, di-, and trimethylated products. a) The Challenger scheme with alternating steps of oxidative methylation and reduction. b) An alternative pathway in which trivalent arsenic bound to glutathione (GSH) or another cellular thiol is methylated without a change in the oxidation state of arsenic. Abbreviations; iAsIII, arsenite; MAsV, methylarsonic acid; MAsIII, methylarsonous acid; DMAsV, dimethylarsinic acid; DMAsIII, dimethylarsinous acid; TMAsV, Trimethylarsine oxide; AdoMet, S-adenosylmethionine; AdoHcy, S- adenosylhomocysteine; GSH, glutathione.

Recognition of the importance of methylation in the metabolism and toxicity of As in humans has it roots in the report by Crecelius (1977) that humans convert ingested iAsV into methylated (MAs) and dimethylated arsenicals (DMAs) that are excreted in urine. In the past three decades, data have accumulated showing that the methylation of iAs is a nearly universal phenomenon in vertebrates. In most species examined, DMAs is the predominant urinary metabolite after exposure to iAs, accounting for 60 to 80% of the As in urine. Both iAs and MAs each typically account for 5 to 20% of the As in urine after exposure to iAs.

Linkage of Arsenic Methylation and Toxicity

Because methylated arsenicals containing pentavalent As (MAsV and DMAsV) were less acutely cytotoxic than iAsIII, it was initially assumed that biological methylation was a process that detoxified As as a prelude to its excretion (Yamauchi and Fowler, 1994). However, early work showed DMAsV, a methylated metabolite of iAs, to be a teratogen, a nephrotoxin, a tumor promoter, and complete carcinogen in the rat (Rogers et al., 1981; Murai et al., 1993; Yamamoto et al., 1995; Wei et al., 1999). These findings raised the possibility that formation of methylated arsenicals produced compounds with unique and potentially higher toxicity. Subsequent studies identified methylated arsenicals containing trivalent As, MAsIII and DMAsIII, to be urinary metabolites after exposure to iAs (Aposhian et al., 2000; Le et al., 2000; Del Razo et al., 2001) and that these compounds were more potent cytotoxins, genotoxins, and enzyme inhibitors than was iAsIII (for a review see Thomas et al., 2001). Taken together, these findings argued that the methylation of iAs should be regarded as an activation process that forms more reactive species which exert unique toxic effects. Therefore, understanding the molecular basis of As methylation may be the key to understanding its actions as a toxicant and a carcinogen.

Enzymatic Basis of Arsenic Methylation in the Context of Cellular Metabolism

Based on evidence that iAs is extensively methylated in humans and many other species, understanding the biological basis of As methylation became a prime research area. As described in the following sections of this chapter, it has been possible to demonstrate that methylation of As in many species is an enzymatically catalyzed process that is consistent with Challenger’s original scheme for this process. A single protein, arsenic (+3 oxidation state) methyltransferase (AS3MT), catalyzes both reduction and methylation reactions for As shown in Figure 2. As is typical for methyltransferases, AdoMet is the methyl donor for these reactions. Hence, the activity of this As methylating enzyme can be linked to a multiple-step process in cells that controls the availability of this prime methyl group donor. Furthermore, the activity of AS3MT is linked with that of thioredoxin reductase (TR) which catalyzes reduction of thioredoxin (Tx), a small dithiol containing protein that provides reducing equivalents to AS3MT (Waters et al., 2004a). This linkage is strengthened because MAsIII, an intermediate in the AS3MT-catalyzed pathway for As methylation, is a potent inhibitor of the activity of thioredoxin reductase (Lin et al., 1999, 2001). Integrating these different metabolic processes into a network of interconnected reactions will provide a more comprehensive understanding of the link between the metabolism of As in the cell and the effect of arsenicals on the molecular processes that underlie its actions as a toxin or carcinogen.

An external file that holds a picture, illustration, etc.
Object name is nihms161022f2.jpg

Interconnections between the cycle for arsenic methylation catalyzed by arsenic (+3 oxidation state) methyltransferase, the enzymatic cycle for regeneration of S-adenosylmethionine (AdoMet) from S-adenosylhomocysteine (AdoHcy) and the cycle for regeneration of reduced thioredoxin (Txred) from oxidized thioredoxin (Txox) in a NADPH-consuming reaction catalyzed by thioredoxin reductase (TR).

The nomenclature for genes encoding arsenic (+3 oxidation state) methyltransferase follows recommendations of the Human Genome Organization Gene Nomenclature Committee (HGNC) for the naming of genes and for the use of symbols to identify genes and proteins in the genome of humans and other species (Wright and Bruford, 2006). Symbols for genes are always italicized. Hence, the symbol for the gene encoding this protein in humans is AS3MT; in all other species, it is As3mt. Symbols for protein products of genes are not italicized. Therefore, the protein product of the human gene is AS3MT; for all other species, it is As3mt.

Identifying the Products of Arsenic Metabolism

A continuing challenge in research on metabolism of As has been development and application of analytical techniques for determination of arsenicals in biological samples. As understanding of the complexities of As metabolism has evolved, researchers have responded to the challenges of identifying the full range of metabolites and accounting quantitatively for As as the administered compound or as a metabolite by adapting and refining methods. For example, from old spot tests for the detection of As (e.g., Marsh’s or Gutzeit’s test) to current analytical techniques (e.g., hydride generation-cryotrapping-atomic absorption spectrometry), researchers have commonly exploited the formation of volatile arsines from arsenicals as a means to separate and quantify arsenicals. Indeed, there has been a continuing dialogue between analytical chemists and biologists in which results of studies of As metabolism spur development and application of new methods to determine whether previously undetected metabolites are present. For example, recognition that recombinant rat As3mt could catalyze formation of trimethylarsine oxide (TMAO) as the penultimate metabolite of iAsIII was spurred by the observation that the sum of the concentrations of MAs and DMAs in reaction mixtures did not quantitatively account for As added to reaction mixtures (Waters et al., 2004b). Application of new methods for collection and analysis permitted identification and quantitation of these metabolites. New pathways for As metabolism are presently under investigation and stimulate the development of new analytical approaches. The conversion of oxoarsenical species to thioarsenicals (for example, conversion of DMAsV to dimethylthioarsenic) occurs via a reaction with hydrogen sulfide generated in tissues (Naranmandura et al., 2007). Dimethylthioarsenic has been detected in urine of individuals chronically exposed to iAs in drinking water (Raml et al., 2007) and in urine of rats exposed to iAsV or DMAsV in drinking water (Adair et al., 2007). Mass spectrometric-based methods for quantitation of members of this class of metabolites are under development in many laboratories. These analytical methods will greatly assist in elucidating the pathway for formation of thioarsenicals and in understanding the linkage between the pathway for formation of methylated oxoarsenicals and the pathway for formation of thioarsenicals.

Footnotes

Publisher's Disclaimer: This manuscript has been reviewed in accordance with the policy of the National Health and Environmental Effects Research Laboratory, U.S. Environmental Protection Agency, and approved for publication. Approval does not signify that the contents necessarily reflect the views and policies of the Agency, nor does mention of trade names or commercial products constitute endorsement or recommendation for use.

Literature Cited

  • Adair BM, Moore T, Conklin SD, Creed JT, Wolf DC, Thomas DJ. Tissue distribution and urinary excretion of dimethylated arsenic and its metabolites in dimethylarsinic acid- or arsenate-treated rats. Toxicol Appl Pharmacol. 2007;222:235–242. [PubMed] [Google Scholar]
  • Aposhian HV, Gurzau ES, Le XC, Gurzau A, Healy SM, Lu X, Ma M, Yip L, Zakharyan RA, Maiorino RM, Dart RC, Tircus MG, Gonzalez-Ramirez D, Morgan DL, Avram D, Aposhian MM. Occurrence of monomethylarsonous acid in urine of humans exposed to inorganic arsenic. Chem Res Toxicol. 2000;13:693–697. [PubMed] [Google Scholar]
  • Bhattacharjee Y. A sluggish response to humanity’s biggest mass poisoning. Science. 2007;315:1659–1661. [PubMed] [Google Scholar]
  • Buchanan WD. Toxicity of arsenic compounds. Amsterdam: Elsevier; 1962. pp. 1–13. [Google Scholar]
  • Challenger F. Biological methylation. Chem Rev. 1945;36:315–361. [Google Scholar]
  • Challenger F. Biological methylation. Advan Enzymol. 1951;12:432–491. [PubMed] [Google Scholar]
  • Chen CJ, Wang SL, Chiou JM, Tseng CH, Chiou HY, Hsueh YM, Chen SY, Wu M, Lai MS. Arsenic and diabetes and hypertension in human populations: a review. Toxicol Appl Pharmacol. 2007;222:298–304. [PubMed] [Google Scholar]
  • Crecelius EA. Changes in the chemical speciation of arsenic following ingestion by man. Environ Health Perspect. 1977;19:147–150. [PMC free article] [PubMed] [Google Scholar]
  • Cullen WR. The sociochemistry of an element. Cambridge: RSC Publishing; 2008. Is arsenic an aphrodisiac? pp. 130–214. [Google Scholar]
  • Cullen WR, McBride BC, Reglinski J. The reduction of trimethylarsine oxide to trimethylarsine by thiols: a mechanistic model for the biological reduction of arsenicals. J Inorg Biochem. 1984a;21:45–60. [Google Scholar]
  • Cullen WR, McBride BC, Reglinski J. The reaction of methylarsenicals with thiols: Some biological implications. J Inorg Biochem. 1984b;21:179–194. [Google Scholar]
  • Del Razo LM, Styblo M, Cullen WR, Thomas DJ. Determination of trivalent methylated arsenicals in biological matrices. Toxicol Appl Pharmacol. 2001;174:282–293. [PubMed] [Google Scholar]
  • Douer D, Tallman MS. Arsenic trioxide: new clinical experience with an old medication in hematologic malignancies. J Clin Oncol. 2005;23:2396–2410. [PubMed] [Google Scholar]
  • Enterline PE, Day R, Marsh GM. Cancers related to exposure to arsenic at a copper smelter. Occup Environ Med. 1995;52:28–32. [PMC free article] [PubMed] [Google Scholar]
  • Hayakawa T, Kobayashi Y, Cui X, Hirano S. A new metabolic pathway of arsenite: arsenic-glutathione complexes are substrates for human arsenic methyltransferase Cyt19. Arch Toxicol. 2005;79:183–191. [PubMed] [Google Scholar]
  • Jolliffe DM. A history of the use of arsenicals in man. J Royal Soc Med. 1993;86:287–289. [PMC free article] [PubMed] [Google Scholar]
  • Le XC, Ma M, Cullen WR, Aposhian HV, Lu X, Zheng B. Determination of monomethylarsonous acid, a key arsenic methylation intermediate, in human urine. Environ Health Perspect. 2000;108:1015–1018. [PMC free article] [PubMed] [Google Scholar]
  • Lin S, Cullen WR, Thomas DJ. Methylarsenicals and arsinothiols are potent inhibitors of mouse liver thioredoxin reductase. Chem Res Toxicol. 1999;12:924–930. [PubMed] [Google Scholar]
  • Lin S, Del Razo LM, Styblo M, Wang C, Cullen WR, Thomas DJ. Arsenicals inhibit thioredoxin reductase in cultured rat hepatocytes. Chem Res Toxicol. 2001;14:305–311. [PubMed] [Google Scholar]
  • Lubin JH, Pottern LM, Stone BJ, Fraumeni JF., Jr Respiratory cancer in a cohort of copper smelter workers: results from more than 50 years of follow-up. Am J Epidemiol. 2000;151:554–565. [PubMed] [Google Scholar]
  • Murai T, Iwata H, Otoshi T, Endo G, Horiguchi S, Fukushima S. Renal lesions induced in F344/DuCrj rats by 4 weeks oral administration of dimethylarsenic acid. Toxicol Lett. 1993;66:53–61. [PubMed] [Google Scholar]
  • Naranmandura H, Suzuki N, Iwata K, Hirano S, Suzuki KT. Arsenic metabolism and thioarsenicals in hamsters and rats. Chem Res Toxicol. 2007;20:616–624. [PubMed] [Google Scholar]
  • Naranmandura H, Suzuki N, Suzuki KT. Trivalent arsenicals are bound to proteins during reductive methylation. Chem Res Toxicol. 2006;19:1010–1018. [PubMed] [Google Scholar]
  • Navas-Acien A, Sharrett AR, Silbergeld EK, Schwartz BS, Nachman KE, Burke TA, Guallar E. Arsenic exposure and cardiovascular disease: a systematic review of the epidemiologic evidence. Am J Epidemiol. 2005;162:1037–1049. [PubMed] [Google Scholar]
  • Navas-Acien A, Silbergeld EK, Pastor-Barriuso R, Guallar E. Arsenic exposure and prevalence of type 2 diabetes in US adults. JAMA. 2008;300:814–822. [PubMed] [Google Scholar]
  • Raml R, Rumpler A, Goessler W, Vahter M, Li L, Ochi T, Francesconi KA. Thio-dimethylarsinate is a common metabolite in urine samples from arsenic-exposed women in Bangladesh. Toxicol Appl Pharmacol. 2007;222:374–380. [PubMed] [Google Scholar]
  • Rogers EH, Chernoff N, Kavlock RJ. The teratogenic potential of cacodylic acid in the rat and mouse. Drug Chem Toxicol. 1981;4:49–61. [PubMed] [Google Scholar]
  • Schwarz RS. Paul Ehrlich’s magic bullets. New Engl J Med. 2004;350:1079–1080. [PubMed] [Google Scholar]
  • Smith AH, Hopenhayn-Rich C, Bates MN, Goeden HM, Hertz-Picciotto I, Duggan HM, Wood R, Kosnett MJ, Smith MT. Cancer risks from arsenic in drinking water. Environ Health Perspect. 1992;97:259–267. [PMC free article] [PubMed] [Google Scholar]
  • Smith AH, Lingas EO, Rahman M. Contamination of drinking-water by arsenic in Bangladesh: a public health emergency. Bull World Health Organ. 2000;78:1093–1103. [PMC free article] [PubMed] [Google Scholar]
  • Thomas DJ, Li J, Waters SB, Xing W, Adair BM, Drobna Z, Devesa V, Styblo M. Arsenic (+3 oxidation state) methyltransferase and the methylation of arsenicals. Exp Biol Med. 2007;232:3–13. [PMC free article] [PubMed] [Google Scholar]
  • Thomas DJ, Styblo M, Lin S. The cellular metabolism and systemic toxicity of arsenic. Toxicol Appl Pharmacol. 2001;176:127–144. [PubMed] [Google Scholar]
  • Tseng WP. Effects and dose--response relationships of skin cancer and blackfoot disease with arsenic. Environ Health Perspect. 1977;19:109–119. [PMC free article] [PubMed] [Google Scholar]
  • Tseng WP, Chu HM, How SW, Fong JM, Lin CS, Yeh S. Prevalence of skin cancer in an endemic area of chronic arsenicism in Taiwan. J Natl Cancer Inst. 1968;40:453–463. [PubMed] [Google Scholar]
  • Wang CH, Hsiao CK, Chen CL, Hsu LI, Chiou HY, Chen SY, Hsueh YM, Wu MM, Chen CJ. A review of the epidemiologic literature on the role of environmental arsenic exposure and cardiovascular diseases. Toxicol Appl Pharmacol. 2007;222:315–326. [PubMed] [Google Scholar]
  • Waters SB, Devesa V, Del Razo LM, Styblo M, Thomas DJ. Endogenous reductants support the catalytic function of recombinant rat cyt19, an arsenic methyltransferase. Chem Res Toxicol. 2004a;17:404–409. [PubMed] [Google Scholar]
  • Waters SB, Devesa V, Fricke M, Creed J, Styblo M, Thomas DJ. Glutathione modulates recombinant rat arsenic (+3 oxidation state) methyltransferase-catalyzed formation of trimethylarsine oxide and trimethylarsine. Chem Res Toxicol. 2004b;17:1621–1629. [PubMed] [Google Scholar]
  • Wei M, Wanibuchi H, Yamamoto S, Li W, Fukushima S. Urinary bladder carcinogenicity of dimethylarsinic acid in male F344 rats. Carcinogenesis. 1999;20:1873–1876. [PubMed] [Google Scholar]
  • Wright MW, Bruford EA. Human and orthologous gene nomenclature. Gene. 2006;369:1–6. [PubMed] [Google Scholar]
  • Yamamoto S, Konishi Y, Matsuda T, Murai T, Shibata MA, Matsui-Yuasa I, Otani S, Kuroda K, Endo G, Fukushima S. Cancer induction by an organic arsenic compound, dimethylarsinic acid (cacodylic acid), in F344/DuCrj rats after pretreatment with five carcinogens. Cancer Res. 1995;55:1271–1276. [PubMed] [Google Scholar]
  • Yamauchi H, Fowler BA. Toxicity and metabolism of inorganic and methylated arsenicals. In: Nriagu JO, editor. Arsenic in the Environment, Part II: Human Health and Ecosystem Effects. New York: Wiley; 1994. pp. 35–43. [Google Scholar]
-