Assessing the genotoxicity of two commonly occurring byproducts of water disinfection: Chloral hydrate and bromal hydrate

https://doi.org/10.1016/j.mrgentox.2016.11.009Get rights and content

Highlights

  • Genotoxicity of chloral hydrate and bromal hydrate investigated via 3 in vitro assays.

  • Chloral hydrate not statistically exhibiting significant genotoxic effects.

  • Mutagenic activity of bromal hydrate in the Salmonella strain TA100.

  • Significant DNA lesions in CHO cells induced by bromal hydrate.

  • Higher genotoxicities of brominated DBPs in comparison to chlorinated species.

Abstract

Water disinfection treatments result in the formation of disinfection byproducts (DBPs) that have been linked to adverse human health outcomes including higher incidence of bladder and colorectal cancer. However, data about the genotoxicity of DBPs is limited to only a small fraction of compounds. Chloral hydrate (CH) and bromal hydrate (BH) are two trihaloacetaldehydes commonly detected in disinfected waters, but little is known about their genotoxicity, especially BH.

We investigated the genotoxicity of CH and BH using a test battery that includes three in vitro genotoxicity assays.

We conducted the Ames test using Salmonella bacterial strains TA97a, TA98, TA100 and TA102, and the alkaline comet assay and the micronucleus test both using Chinese hamster ovary cells. We carried out the tests in the absence and presence of the metabolic fraction S9 mix.

CH did not exhibit statistically significant genotoxic effects in any of the three assays. In contrast, BH exhibited mutagenic activity in the Salmonella strain TA100 and induced statistically significant DNA lesions in CHO cells as appeared in the comet assay. The genotoxic potential of BH in both assays decreased in the presence of the metabolic fraction S9 mix. BH did not induce chromosomal damage in CHO cells.

Our results show that BH exhibited genotoxic activity by causing mutations and primary DNA damage while CH did not induce genotoxic effects. Our findings highlight concerns about the higher genotoxicity of brominated DBPs in comparison to their chlorinated analogues.

Introduction

One of the most significant public health advances of the twentieth century was the adoption of drinking water disinfection in many countries [1]. This practice has sharply reduced the incidence of infectious diseases such as cholera, typhoid, and dysentery [2], [3]. After this dramatic success, disinfection practices have been introduced into swimming pools and other recreational water venues to ensure the elimination of pathogenic microorganisms and the prevention of waterborne disease outbreaks [4]. However, disinfection treatments result in the undesirable formation of chemical contaminants known as disinfection byproducts (DBPs), in consequence to reactions taking place between disinfectants and organic matter present in water [5], [6]. Exposure to DBPs in humans can take place through ingestion of drinking water or inhalation and dermal absorption during showering or swimming [7], [8], [9], [10]. Many studies have suggested associations between exposure to DBPs and adverse health effects. Increased incidence of asthma [11], bladder cancer [12], [13], and colorectal cancer [14] have been reported. Adverse pregnancy outcomes such as spontaneous abortions [15], stillbirth [16], and fetal growth restriction [17] have also been noted. To date, more than six hundred DBPs including trihalomethanes, haloacids, halonitriles, haloaldeydes, haloketones, halonitromethanes, haloamines, haloamides, haloalcohols, and halobenzoquinones have been identified in disinfected waters [9], [18], [19], [20], [21], [22], [23]. Several laboratory-controlled studies have been conducted to evaluate potential toxicities of DBPs providing evidence about cytotoxic, genotoxic, carcinogenic and teratogenic potentials [20], [24], [25], [26], [27], [28]. However, the toxicological data are limited to only a small fraction of identified DBPs. In consequence, many DBPs that have been detected in disinfected waters remain with unknown toxicological profiles. Chloral hydrate and bromal hydrate, the hydrated forms of trichloroacetaldehyde and tribromoacetaldehyde respectively, belong to the chemical class of haloacetaldehydes. This class of DBPs has been reported to be one of the most abundant DBP classes by weight [19], [25], [29], [30]. Occurrence studies have shown that the predominant trhihaloacetaldehyde in chlorinated waters is chloral hydrate, while bromal hydrate is the predominant trihaloacetaldehyde in chlorinated waters containing high levels of bromide [19], [31]. In a recent study, BH was detected as one of the degradation byproducts of benzophenone-3, a UV filter commonly used in sunscreens, in chlorinated swimming pools filled with seawater [32].

Toxicokinetic studies have shown that CH is rapidly absorbed after oral administration, and enters the liver where it undergoes extensive metabolism in rodents [33], [34] and in humans [35], [36]. Studies of the potential carcinogenicity of CH in mice have demonstrated that it is able to induce hepatocellular adenomas and carcinomas, and exposure to CH has been associated with increases in malignant lymphoma and adenoma of the pituitary gland [37], [38], [39]. However, there was still no persuasive evidence to connect chloral hydrate exposure and the development of cancers in humans [40]. CH was also found to induce significant aneugenic effects in mice [41]. Furthermore, micronuclei were produced in germ cells of male mice treated intraperitoneally with CH [42]. CH was also reported to be able to lead to chromosomal loss in mouse spermatids [43] and in human lymphocytes [44]. Nevertheless, most of the investigations incorporated only one or two in vitro assays [25], [45] and results from genotoxicity assessment of CH remain inconclusive. Concerning BH, although little is known about its toxicity, the U.S. Environmental Protection Agency (EPA) included this compound to the list of priority DBPs to be monitored in a nationwide occurrence study [46] due to anticipations of potential toxicity based on alarming structure-activity relationships [47]. To address this scarcity of data, we analyzed the genotoxicity of CH and BH using a battery of three genotoxicity assays, namely the Ames test, the comet assay, and the micronucleus assay. The use of a test battery is critical since no single genotoxicity test is capable of detecting all genotoxic mechanisms [48]. We performed the three assays in the absence and presence of the metabolic activation fraction S9 mix to assess the effects of metabolic reactions on the toxicity of the two compounds.

Section snippets

Chemicals

The identifiers and structures of CH and BH are shown in Table 1. CH (crystallized, ≥98%) was obtained from Sigma-Aldrich (China). BH was prepared by adding tribromoacetaldehyde (bromal, Sigma-Aldrich, UK, 97% purity), to ultrapure water and then recrystallizing the product from a small volume of water. Ultrapure water was produced from a Millipore water system (resistivity = 18.2 MΩ.cm). Before toxicological analyses, stock solutions were prepared in dimethyl sulfoxide (DMSO, Chromasolv plus,

The Ames test

We evaluated the capacity of CH and BH to induce mutations in DNA using the Ames test in four Salmonella tester strains: TA97a, TA98, TA100, and TA102. The assay was carried out in the absence and presence of exogenous metabolic activation (S9 mix). As shown in Table 2, CH did not exhibit mutagenic effects in the tested strains, unlike BH, which induced mutagenic activity in the strain TA100. The results of the regression analyses for the dose-response relationship of BH in the presence and

Discussion

Several epidemiologic studies have suggested that exposure to DBPs is associated with increased incidence of cancer, particularly bladder and colorectal cancers [12], [13], [14]. The analysis of genotoxicity of DBPs allows identifying the chemical species that could be responsible for carcinogenicity [48]. Genotoxicity testing detects carcinogens that are thought to act primarily via a mechanism involving direct genetic damage [48]. CH has been frequently detected as a predominantly occurring

Conclusions

Our findings show that CH did not induce any genotoxic effects using the Ames test, the comet assay, and the micronucleus test, while BH was mutagenic in the Salmonella assay and caused genomic damage in the comet assay but not in the micronucleus test. These results imply that the mechanism of genotoxicity of BH involves inducing mutations and DNA damage but not chromosomal aberrations. The genotoxicity of BH decreased in the presence of metabolic fractions suggesting the formation of less

Acknowledgment

T.M. acknowledges the French Ministry of Higher Education and Research for his doctoral scholarship.

References (76)

  • O.H. Lowry et al.

    Protein measurement with the Folin phenol reagent

    J. Biol. Chem.

    (1951)
  • D.M. Maron et al.

    Revised methods for the salmonella mutagenicity test

    Mutat. Res.

    (1983)
  • M. De Méo et al.

    Optimization of the Salmonella/mammalian microsome assay for urine mutagenesis by experimental designs

    Mutat. Res. Genet. Toxicol.

    (1996)
  • B.S. Kim et al.

    Statistical methods for the Ames Salmonella assay: a review

    Mutat. Res. Mutat. Res.

    (1999)
  • M. De Méo et al.

    Genotoxic activity of potassium permanganate in acidic solutions

    Mutat. Res. Toxicol.

    (1991)
  • P.L. Olive et al.

    Induction and rejoining of radiation-induced DNA single-strand breaks: tail moment as a function of position in the cell cycle

    Mutat. Res. Repair.

    (1993)
  • M. Kirsch-Volders et al.

    Report from the in vitro micronucleus assay working group

    Mutat. Res. Toxicol. Environ. Mutagen.

    (2003)
  • J. Lee et al.

    Production of various disinfection byproducts in indoor swimming pool waters treated with different disinfection methods

    Int. J. Hyg. Environ. Health

    (2010)
  • M. Kiffe et al.

    Characterization of cytotoxic and genotoxic effects of different compounds in CHO K5 cells with the comet assay (single-cell gel electrophoresis assay)

    Mutat. Res. Toxicol. Environ. Mutagen.

    (2003)
  • D.W. Fairbairn et al.

    The comet assay: a comprehensive review

    Mutat. Res.

    (1995)
  • A. Matsuoka et al.

    Evaluation of the micronucleus test using a Chinese hamster cell line as an alternative to the conventional in vitro chromosomal aberration test

    Mutat. Res. Mutagen. Relat. Subj.

    (1992)
  • I. Decordier et al.

    The in vitro micronucleus test: from past to future

    Mutat. Res. − Genet. Toxicol. Environ. Mutagen.

    (2006)
  • B. Miller et al.

    Evaluation of the in vitro micronucleus test as an alternative to the in vitro chromosomal aberration assay: position of the GUM working group on the in vitro micronucleus test

    Mutat. Res. − Rev. Mutat. Res.

    (1998)
  • L. Vian et al.

    Evaluation of hydroquinone and chloral hydrate on the in vitro micronucleus test on isolated lymphocytes

    Mutat. Res.

    (1995)
  • K. Harrington-Brock et al.

    Mutagenicity of three disinfection by-products: di- and trichloroacetic acid and chloral hydrate in L5178Y/TK +/− (−)3.7.2C mouse lymphoma cells

    Mutat. Res.

    (1998)
  • R.J. Bull et al.

    Evaluation of mutagenic and carcinogenic properties of brominated and chlorinated acetonitriles: by-products of chlorination

    Fundam. Appl. Toxicol.

    (1985)
  • G.A. Boorman et al.

    Drinking water disinfection byproducts: review and approach to toxicity evaluation

    Environ. Health Persp.

    (1999)
  • D. Cutler et al.

    The role of public health improvements in health advances: the twentieth-century United States

    Demography

    (2005)
  • J. Orme et al.

    Health effects of disinfectants and disinfection by-products: a regulatory perspective

    Water Chlor.

    (1990)
  • Guidelines for safe recreational water environments: volume 2

    (2006)
  • J. Rook

    Formation of haloforms during chlorination of natural waters

    J. Water Treat. Exam.

    (1974)
  • L.C. Backer et al.

    Household exposures to drinking water disinfection by-products: whole blood trihalomethane levels

    J. Expo. Anal. Environ. Epidemiol.

    (2000)
  • A.M. Miles et al.

    Comparison of trihalomethanes in tap water and blood

    Environ. Sci. Technol.

    (2002)
  • S.D. Richardson et al.

    What’s in the pool? A comprehensive identification of disinfection by-products and assessment of mutagenicity of chlorinated and brominated swimming pool water

    Environ. Health Perspect.

    (2010)
  • X. Xu et al.

    Dermal uptake of chloroform and haloketones during bathing

    J. Expo. Anal. Environ. Epidemiol.

    (2005)
  • A. Bernard et al.

    Chlorinated pool attendance atopy, and the risk of asthma during childhood

    Environ. Health Perspect.

    (2006)
  • C.M. Villanueva et al.

    Disinfection byproducts and bladder cancer: a pooled analysis

    Epidemiology

    (2004)
  • C.M. Villanueva et al.

    Bladder cancer and exposure to water disinfection by-products through ingestion bathing, showering, and swimming in pools

    Am. J. Epidemiol.

    (2007)
  • Cited by (28)

    View all citing articles on Scopus
    View full text