Halobenzoquinone-Induced Developmental Toxicity, Oxidative Stress

Aug 20, 2018 - HBQ exposure increased 8-hydroxydeoxyguanosine levels, DNA fragmentation, and apoptosis in larvae, with apoptosis induction related to ...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIVERSITY OF KENTUCKY

Environmental Measurements Methods

Halobenzoquinone-induced developmental toxicity, oxidative stress, and apoptosis in zebrafish embryos Chang Wang, Xue Yang, Qi Zheng, Birget Moe, and Xing-Fang Li Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.8b02831 • Publication Date (Web): 20 Aug 2018 Downloaded from http://pubs.acs.org on August 20, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

Environmental Science & Technology

1

Halobenzoquinone-induced developmental toxicity, oxidative stress, and apoptosis

2

in zebrafish embryos

3

Chang Wang,1# Xue Yang,1, 2# Qi Zheng,3* Birget Moe4,5 and Xing-Fang Li4*

4 5

1. Institute of Environment and Health, Jianghan University, Wuhan 430056, China

6

2. School of Environmental Ecology and Biological Engineering, Wuhan Institute of Technology,

7 8

Wuhan 430025, China 3. Key Laboratory of Optoelectronic Chemical Materials and Devices, Ministry of Education, Institute

9

of Environment and Health, Jianghan University, Wuhan 430056, China

10

4. Division of Analytical and Environmental Toxicology, Department of Laboratory Medicine and

11

Pathology, Faculty of Medicine and Dentistry, University of Alberta, Edmonton, Alberta, Canada

12

T6G 2G3

13

5. Alberta Centre for Toxicology, Department of Physiology and Pharmacology, Faculty of Medicine,

14

University of Calgary, Calgary, Alberta, Canada T2N 4N1

15 16

#

17

* To whom correspondence should be addressed:

18

Qi Zheng: [email protected].

19

Xing-Fang Li: [email protected], 1-780-492-5094.

Co-first authors

20

ACS Paragon Plus Environment

Environmental Science & Technology

21

Page 2 of 29

Abstract

22

The developmental toxicity of water disinfection byproducts remains unclear. Here we report the

23

study of halobenzoquinone (HBQ)-induced in vivo developmental toxicity and oxidative stress using

24

zebrafish embryos as a model. Embryos were exposed to 0.5 to 10 µM of individual HBQs and 0.5 to 5

25

mM haloacetic acids for up to 120 hours post-fertilization (hpf). LC50 values of the HBQs at 24 hpf were

26

4.6 to 9.8 µM, while those of three haloacetic acids were up to 200 times higher at 1900 to 2600 µM.

27

HBQ exposure resulted in significant developmental malformations in larvae, including failed inflation

28

of the gas bladder, heart malformations, and curved spines. An increase in reactive oxygen species was

29

observed along with a decrease in superoxide dismutase activity and glutathione content. Additionally,

30

the antioxidant N-acetyl-L-cysteine significantly mitigated all HBQ-induced effects, supporting that

31

oxidative stress contributes to HBQ toxicity. Further experiments examined HBQ-induced effects on

32

DNA and genes. HBQ exposure increased 8-hydroxydeoxyguanosine levels, DNA fragmentation, and

33

apoptosis in larvae, with apoptosis induction related to changes in the gene expression of p53 and mdm2.

34

These results suggest that HBQs are acutely toxic, causing oxidative damage and developmental toxicity

35

to zebrafish larvae.

36

ACS Paragon Plus Environment

Page 3 of 29

37

Environmental Science & Technology

Table of Contents (TOC) Graphic

38

ACS Paragon Plus Environment

Environmental Science & Technology

39

Page 4 of 29

Introduction

40

Drinking water disinfection is an effective public health measure necessary to kill pathogenic

41

microorganisms in source water. Disinfection of reclaimed water dramatically decreases the morbidity

42

and mortality of infectious diseases.1 At present, the widely used liquid chlorine, chloramine, ozone, and

43

chlorine dioxide are known to react with natural organic matter or environmental pollutants present in

44

source water to generate disinfection byproducts (DBPs).2 Many identified DBPs are cytotoxic,

45

genotoxic, and mutagenic in vitro or in vivo,3 and epidemiological studies suggest that the consumption

46

of chlorinated drinking water is potentially associated with adverse reproductive health effects and an

47

increased risk of bladder cancer.4 As such, several DBPs and classes of DBPs are regulated in North

48

America, including chlorites, bromate, trihalomethanes, and haloacetic acids (HAAs).5, 6

49

An emerging class of DBPs are the halobenzoquinones (HBQs).7, 8 HBQ-DBPs have been detected

50

in treated drinking water and swimming pool water across North America at ng/L levels, with

51

2,6-DCBQ, the most prevalent HBQ-DBP, detected at an occurrence frequency of 100%.7,9-11

52

Quantitative structure toxicity relationship (QSTR) analysis predicted HBQs to be 1000-fold more toxic

53

in comparison to regulated DBPs, and are likely to be carcinogenic and mutagenic.8 In vitro toxicity

54

studies have reported that HBQs, including 2,6-DCBQ and 2,6-DBBQ, are cytotoxic to T24 human

55

bladder cancer cells. The supplementation of glutathione (GSH) mediated the detoxification of HBQs in

56

T24 cells,12 suggesting that HBQ cytotoxicity was related to the generation of ROS.13,

57

Quinone-induced oxidative stress has also resulted in oxidative damage to proteins and DNA, forming

58

protein and DNA adducts.15-18 The 2,5-HBQ isomers, 2,5-DCBQ and 2,5-DBBQ, have been shown to be

59

more cytotoxic than their corresponding 2,6-HBQ analogues, although the compounds share the same

60

mechanism of toxicity.19,20 Additionally, 2,6-DCBQ and 2,6-DBBQ have been shown to induce cell

61

cycle arrest of human neural stem cells in vitro, indicating the potential developmental neurotoxicity of

62

this class of DBPs.21 Although inconsistent correlations between maternal exposure to treated water and

63

developmental effects in fetuses have been shown in epidemiological studies,22 an increased risk of

64

various adverse developmental outcomes have been reported.23-25 Thus, further experimental studies

65

using in vivo models are required 1) to determine if oxidative stress contributes to HBQ-induced toxicity

ACS Paragon Plus Environment

14

Page 5 of 29

66

Environmental Science & Technology

in vivo and 2) to assess the potential developmental toxicity of this emerging class of DBPs.

67

Zebrafish embryos are an established model for evaluating the developmental toxicity of

68

environmental pollutants,26 and for high throughput screening, due to the high fecundity, rapid

69

embryogenesis, and continuous reproduction of the transparent larvae.27 Although there are increasing

70

studies regarding the toxicity mechanisms of DBPs, information on their aquatic toxicity is limited. In

71

previous studies, adult zebrafish were used to assess the aquatic toxicity of dichloroacetonitrile and

72

2,2-dichloroacetamide, which exhibited bioaccumulation and neurotoxicity.28-30 Zebrafish embryos were

73

employed to evaluate the acute and developmental toxicity of different halogenated DBPs, with

74

chlorinated DBPs shown to be less toxic than their corresponding bromine- and iodine-substituted

75

analogues.31 Dr. Zhang's team has also established a marine polychaete model for testing developmental

76

toxicity of DBPs.32, 33 DBPs are commonly released and generated in the aquatic environment, although

77

few have been identified.34 Furthermore, potential human or environmental exposures to HBQs are not

78

limited to DBPs. TCBQ is a commonly used global fungicide, with an annual production in China over

79

2000 tons.35 TCBQ is also an oxidative metabolite of pentachlorophenol, another widely-used pesticide,

80

disinfectant, and preservative.35 Both chemicals have been identified as source water contaminants. Thus,

81

because of the ease at which these contaminants may enter the aquatic environment, it is important to

82

evaluate the toxicity of HBQs in aquatic organisms.

83

Based on previous studies, our hypothesis was that developing zebrafish embryos would show

84

developmental toxicity and oxidative damage upon exposure to five HBQs (2,6-DCBQ, 2,5-DCBQ,

85

2,5-DBBQ, TCBQ, and TBBQ). To test this hypothesis, we examined embryo mortality, the rate of

86

malformations, ROS production, changes in the activity and concentration of antioxidant proteins,

87

oxidative DNA damage, and apoptosis in zebrafish larvae to provide both morphological observations

88

and quantitative molecular biological endpoints. For comparison, two regulated HAA-DBPs and

89

iodoacetic acid (IAA) were also examined. The findings of this study will illustrate a useful method for

90

screening the developmental toxicity of emerging DBPs and evaluating their potential aquatic toxicity.

91

Materials and Methods

92

Chemicals

and

Reagents.

Standards

of

2,6-dichloro-1,4-benzoquinone

ACS Paragon Plus Environment

(2,6-DCBQ),

Environmental Science & Technology

Page 6 of 29

93

2,5-dichloro-1,4-benzoquinone

94

tetrachloro-1,4-benzoquinone (TCBQ), and tetrabromo-1,4-benzoquinone (TBBQ) were obtained from

95

Tokyo Chemical Industry (TCI; Toshima, Kita-ku, Tokyo, Japan) (Table S1). HBQs were dissolved in

96

dimethyl sulfoxide (DMSO), which was used as the vehicle control in each assay. Tricaine (MS-222,

97

150 mg/L), used to anesthetize the zebrafish larvae, and 2′,7′-dichlorofluorescin diacetate (DCFDA)

98

were purchased from Sigma-Aldrich (St. Louis, MO, USA). Acridine Orange (AO) was purchased from

99

the Biosharp Company (Shanghai, China). N-acetyl-L-cysteine (NAC) was purchased from TCI and

100

diluted in sterilized water. Dichloroacetic acid (DCA) was purchased from Macklin Biochemical

101

(Shanghai, China), dibromoacetic acid (DBA) from ZZ Standard (Shanghai, China), and IAA from

102

Aladdin Bio-Chem Technology (Shanghai, China) (Table S1).

103

Zebrafish Maintenance and Experimental Design. All experiments were performed on zebrafish

104

embryos hatched by the wild type AB strain (3-months old, from the Institute of Hydrobiology, Chinese

105

Academy of Sciences), maintained in 28 ± 0.5°C aerated water with 0.25-0.75‰ (w/v) salinity. Adult

106

fish were paired, two males and two females in each tank, and were separated the night before the

107

collection of fertilized eggs. Approximately thirty embryos were randomly selected, placed into 6-wells

108

plates, and exposed to HBQs (0-16 µM) from 4 hours post fertilization (hpf) to 120 hpf. For all

109

experiments, three replicate wells were prepared for each exposure group. A larger number of embryos

110

was used in high concentration groups in all assays except the mortality assay to ensure sufficient

111

numbers of surviving larvae for each assay. Both control and exposure groups contained less than 0.007%

112

(v/v) DMSO. Embryos were also treated in the presence or absence of 50 µM of the antioxidant NAC, as

113

it has been demonstrated previously that concentrations of NAC less than 50 µM are non-toxic and

114

effective in eliminating ROS in zebrafish embryos.36,37 The HAAs (DCA, DBA, and IAA; 0-5 mM)

115

were studied in parallel for comparison with the HBQs.

116

Morphological Observations. The morphological development of embryos after HBQ exposure (0, 1, 2,

117

5, 8, 10, 16 µM) was observed using a stereoscopic microscope (ZEISS SteREO Discovery. V12,

118

Germany) during three stages of development: pharyngula, hatching, and swimming larva. Mortality (24

119

hpf), hatching rate (60 hpf and 72 hpf), heart rate (72 hpf), uninflated swim bladder (120 hpf), and

(2,5-DCBQ),

2,5-dibromo-1,4-benzoquinone

ACS Paragon Plus Environment

(2,5-DBBQ),

Page 7 of 29

Environmental Science & Technology

120

various other malformations (72 hpf; tail injury, pericardial edema, shortened body length, shortened

121

yolk sac extension, developmental delay) were recorded.

122

Measurement of ROS and Antioxidant Molecules. ROS generation in embryos exposed to HBQs (0.5,

123

1, 5, 10 µM), HAAs (0.5, 1 mM), or H2O2 (0.5%, 1%) was measured using DCFDA. Briefly, 40 live

124

larvae (120 hpf) for each treatment group were washed with cold PBS (pH 7.4) and homogenized in cold

125

buffer (0.32 mM sucrose, 20 mM HEPES, 1 mM MgCl2, and 0.5 mM phenylmethyl sulfonylfluoride;

126

pH 7.4). The homogenate was centrifuged at 12,000 rpm at 4°C for 10 min, and the supernatants were

127

transferred to clean tubes for analysis. Fluorescence intensity was measured using a microplate reader

128

(Synergy™ 2 Multi-Mode Microplate Reader, Biotek Instruments, VT, USA), with excitation and

129

emission at 485 and 530 nm, respectively. Results are expressed as a percentage (%) relative to the

130

control. In addition, an in vivo zebrafish ROS assay was performed. Briefly, after exposure of the larvae

131

to 2,5-DCBQ (120 hpf; 0, 0.5, 2.5, 5 µM), the larvae were incubated with DCFDA in the dark for 1 h at

132

28 °C. For microscopic examination, the larvae were mounted onto glass slides with 3% methylcellulose,

133

and in vivo ROS generation was assessed using a fluorescence microscope (ZEISS Axio Vert.A1,

134

Germany). The fluorescence intensity of individual larva was quantified by area of integrated optical

135

density using Image J software (National Institutes of Health, Bethesda, MD, USA). In both the in vitro

136

and in vivo ROS assays, 2,5-DCBQ- and 2,5-DBBQ-treated larvae (120 hpf) were co-exposed to the

137

antioxidant NAC to confirm that the generated ROS induce oxidative stress. More details regarding

138

these methods are available in the Supporting Information.

139

To maintain intracellular redox homeostasis, ROS and other free radicals are eliminated by

140

antioxidant systems, including antioxidant enzymes (e.g. superoxide dismutase; SOD) and non-enzyme

141

inhibitors (GSH). To assess the antioxidant response to HBQ exposure, SOD activity and GSH

142

production were measured according to the protocols of commercial kits from Nanjing Jiancheng

143

Bioengineering Institute (NJBI; Nanjing, China). Briefly, 40 live larvae were homogenized and prepared

144

as described above for the in vitro ROS assay with the supernatants collected for detection of SOD

145

activity and GSH level. The results of the SOD assay are expressed in units of SOD activity per

146

milligram of protein (U/mg), wherein 1 U of SOD is defined as the amount of sample that causes a 50%

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 29

147

inhibition of cytochrome C reduction. GSH content is expressed as a percentage (%) relative to the

148

control.

149

Determination of DNA Damage. For analysis of DNA damage, larvae were exposed to 2,5-DCBQ,

150

2,5-DBBQ, and 2,6-DCBQ (120 hpf; 1, 5, 10 µM). Approximately 40 live larvae were homogenized and

151

prepared as described above for the in vitro ROS assay, with the supernatants collected for the

152

determination of DNA damage. 8-OHdG, a biomarker of oxidative DNA damage, was measured using

153

an enzyme-linked immunosorbent assay (ELISA) according to the manufacturer’s instruction (8-OHdG

154

ELISA Kit, NJBI). In addition, a DNA ladder assay was used to study DNA fragmentation and the

155

degree of DNA damage after exposure to 2,5-DCBQ and 2,5-DBBQ (120 hpf; 1, 5, 10 µM), following

156

the manufacturer’s instruction (Apoptosis DNA Ladder Extraction Kit, KeyGEN BioTECH, Nanjing,

157

China). After DNA extraction, gel electrophoresis and imaging were performed with a gel imaging

158

system (G-BOX, Syngene, Cambridge, UK).

159

Acridine Orange (AO) Staining and Apoptosis Imaging. Larval cell apoptosis was identified by AO

160

staining, a nucleic acid-selective metachromatic stain useful for studying apoptosis patterns.38 After 96 h

161

exposure to 2,5-DCBQ or 2,5-DBBQ (0.5, 2.5, 5 µM), 15 larvae from each group (n = 3) were washed

162

twice in 30% Danieau's solution (58 mM NaCl, 0.7 mM KCl, 0.4 mM MgSO4, 0.6 mM Ca(NO3)2, and 5

163

mM HEPES, pH 7.4), transferred to 30% Danieau's solution containing 5 µg/mL AO, and incubated for

164

20 min. The larvae were anesthetized with tricaine (MS-222) prior to microscopic observation of

165

apoptosis (Carl Zeiss, SteREO Discovery.V12, Germany). Image J software was used to quantify the

166

fluorescence intensity of individual larva by area of integrated optical density.

167

Quantitative Real-time PCR. To examine HBQ-induced apoptosis, changes in expression of two genes

168

associated with apoptosis, p53 and mdm2, were measured. About 100 embryos were cultured in 10-cm

169

culture dishes for each exposure group. After treatment with 2,5-DCBQ or 2,5-DBBQ (0.5, 2.5, 5 µM)

170

for 120 hpf, 30 surviving larvae were randomly selected for qRT-PCR. Details on sample preparation

171

and analysis are described in the Supporting Information. The GenBank accession numbers and forward

172

and reverse primer sequences are listed in Table S2, while the fold changes in expression level were

173

calculated using the 2-∆∆Ct method.39

ACS Paragon Plus Environment

Page 9 of 29

Environmental Science & Technology

174

Statistical Analysis. Statistical analysis was performed using GraphPad Prism 7 (Graphpad Software, La

175

Jolla, CA, USA) and SPSS 22.0 (IBM Corp., Armonk, NY, USA). Experimental results were expressed

176

as the mean ± standard deviation (SD) or as the mean ± the standard error of the mean (SEM). Statistical

177

differences between treatment groups in the presence or absence of NAC were determined by Student’s

178

t-test. Statistical comparisons of analyses in which multiple treatment groups were tested (including

179

ROS production, SOD activity, GSH production, 8-OHdG, AO, mdm2 expression, p53 expression,

180

hatching/heart rate) were assessed using a one-way analysis of variance (ANOVA) with a Dunnett's

181

post-test.40 Differences were considered statistically significant at p < 0.05. LC50 values, defined as the

182

concentration that induces half of the maximum mortality, were calculated from a nonlinear regression

183

model. The four-parameter logistic curve regression analysis was performed according to the following

184

formula:

185

y = minimum + (maximum – minimum) ÷ [1 + (x/LC50)Hill slope]

186

where y is mortality and x is the log concentration of the test compound.

187

Results and Discussion

188

Acute Toxicity (LC50) of HBQs and HAAs. Zebrafish embryo mortality significantly increased with

189

HBQ and HAA concentration (Fig. 1A; Fig. S1); however, no significant effect on embryo hatching rate

190

at 60 hpf or 72 hpf was observed (Table S3). Mortality rates were used to calculate LC50 values for the

191

five HBQs and three HAAs at 24, 48, 72, 96, and 120 hpf (Table 1; Table S4), with the toxic potency of

192

the compounds ranked as 2,5-DCBQ ˃ 2,5-DBBQ ˃ 2,6-DCBQ ˃ TBBQ ˃ TCBQ > IAA > DBA ˃

193

DCA. The calculated LC50 values show that the acute toxicity of HBQs in zebrafish embryos is two to

194

three orders of magnitude greater than the HAAs. These findings are consistent with a previous study

195

that evaluated the developmental toxicity of 20 halogenated DBPs in the marine polychaete, Platynereis

196

dumerilii. Of the 20 DBPs, 2,5-dibromohydroquinone had the highest developmental toxicity, with a

197

reported EC50 value 100 to 1000 times lower than those of the THMs or HAAs.32, 33 To clarify that the

198

observed HBQ toxicity is linked to oxidative stress, we examined the effects of HBQs in the presence of

199

NAC, a thiol antioxidant and ROS scavenger.41 When NAC (50 µM) was added to the high

200

concentration treatment group (16 µM) of each HBQ, a significant reduction in mortality was observed

ACS Paragon Plus Environment

(Equation 1)

Environmental Science & Technology

Page 10 of 29

201

(t-test, p < 0.05; Fig. 1B), implicating oxidative stress as a mechanism of HBQ-induced embryotoxicity.

202

Exposure-induced Oxidative Stress. All five HBQs significantly induced ROS in zebrafish embryos in a

203

dose-dependent manner at 120 hpf (ANOVA with Dunnett’s post-test, p < 0.05; Fig. 2A1). In

204

comparison to the HAA-DBPs (Fig. 2A2), the HBQs generated similar levels of ROS, but at much lower

205

concentrations, as 10 µM of each HBQ produced as much ROS as 1000 µM of each HAA (Fig. 2). Thus,

206

the more potent embryotoxicity of the HBQs in relation to the HAAs is likely influenced by their ability

207

to induce ROS. Again, it was found that NAC co-exposure mitigated the HBQ-induced effects; 50 µM

208

of the antioxidant was sufficient to eliminate excess ROS in the 2,5-DCBQ and 2,5-DBBQ treatment

209

groups (Fig. 2A1). For the localization of ROS production in tissue, in vivo imaging of DCFH-DA in

210

whole zebrafish larvae was performed after HBQ exposure. Background fluorescence from normal

211

physiological processes was observed in both the control and DCFDA treatment groups (Fig. 3A, B).

212

However, increased fluorescence over the background was clearly visible in each of the HBQ treatment

213

groups (Fig.3 D-F). The fluorescence of the 5 µM 2,5-DCBQ treatment group (Fig. 3F) was greatly

214

reduced in the presence of exogenous NAC (Fig. 3C), but the adverse effects were not fully attenuated.

215

Fluorescence intensity was greatest in the eyes, head, digestive tract, and yolk sac of the larvae, but also

216

visible in the muscle tissue of the tail. In the higher concentration 2,5-DCBQ treatment groups (Fig. 3E,

217

F), pericardial edema is clearly distinguished under fluorescence (Fig. 3C, E, F). The fluorescence

218

intensity of each of the treatment groups was quantified in Figure S2A.

219

SOD activity and GSH depletion were also investigated as indicators of oxidative stress, as both

220

antioxidants are critically important for the elimination of excess ROS to prevent oxidative damage.42

221

With increasing concentration of each HBQ, SOD activity was concurrently inhibited (ANOVA with

222

Dunnett’s post-test, p < 0.05; Fig. 2B). Likewise, GSH depletion was dose-dependent in the 2,5-DCBQ

223

and 2,5-DBBQ treatment groups (ANOVA with Dunnett’s post-test; Fig. 2C), consistent with in vitro

224

studies that found HBQs reduced GSH levels in quinone-treated PC12, T24, and HepG2 cells.12,43, 44

225

Both SOD activity and GSH concentrations were significantly recovered with NAC co-exposure (t-test,

226

p < 0.05; Fig. 2B, C). As a free radical scavenger, GSH responds rapidly to ROS produced by toxic

227

chemicals, 45 while SOD effectively maintains the oxidant balance of the highly reactive superoxide

ACS Paragon Plus Environment

Page 11 of 29

Environmental Science & Technology

228

anion. The resulting reduction in free GSH and SOD activity increases the vulnerability of zebrafish

229

larvae to HBQ-induced oxidative damage. The production of ROS is dependent on the semi-quinone and

230

hydroxyl radical structural features of HBQs.5, 16 Because of their electron deficient structure, HBQs are

231

reactive electrophiles which can readily react with various bio-nucleophiles, such as proteins and nucleic

232

acids. The nucleophilic attack of GSH by quinones and HBQs has been demonstrated.16, 46 Thus, the

233

reduction in GSH is likely a result of both GSH oxidation and conjugation.46

234

DNA Damage and Apoptosis. Oxidative DNA damage in zebrafish larvae was assessed via the

235

measurement of the DNA lesion, 8-OHdG, at 120 hpf. The content of 8-OHdG significantly increased at

236

120 hpf in the high dose treatment groups of 2,6-DCBQ, 2,5-DCBQ, and 2,5-DBBQ (1, 5, 10 µM)

237

(ANOVA with Dunnett’s post-test, p < 0.05; Fig. 4A), indicating DNA oxidation. This is consistent with

238

in vitro studies that found HBQs significantly induced 8-OHdG formation in T24 cells and Chinese

239

hamster lung cells.15, 47 Simultaneously, measurements (Fig. 4B) of DNA fragmentation clearly showed

240

HBQ-induced DNA damage, as the DNA smear extended throughout the gel lane in the high dose

241

treatment groups of 2,5-DCBQ (Fig. 4B1) and 2,5-DBBQ (Fig. 4B2). No significant fragmentation was

242

visible in the control group (Fig. 4B1, second lane from left). DNA fragmentation is a hallmark of

243

apoptosis, which could be induced as a result of irreparable HBQ-induced oxidative DNA damage. Thus,

244

we examined apoptotic cells in vivo in whole zebrafish larvae using AO staining. Treatment

245

concentrations of 2,5-DCBQ (Fig. 5A2-4) and 2,5-DBBQ (Fig. 5B2-4) exhibited higher fluorescence

246

intensity in the heart, brain, yolk sac, and mouth of the larvae in comparison to the control group (Fig.

247

5A1). Furthermore, co-exposure of 2,5-DCBQ (5 µM) with 50 µM NAC resulted in fewer apoptotic

248

cells, demonstrated by the lower fluorescence intensity, and prevented pericardial edema (Fig. 5B1). The

249

quantified fluorescence intensity of each treatment group is available in the Supporting Information

250

(Figure S2B).

251

For further examination of HBQ-induced apoptosis, two genes (p53, mdm2) associated with

252

apoptosis were examined in larvae at 120 hpf after exposure to 0.5, 2.5, or 5 µM 2,5-DCBQ or

253

2,5-DBBQ. The pattern of p53 expression was similar in larvae exposed to the HBQs investigated, as

254

p53 was significantly up-regulated (1.73-fold, 1.42-fold) in the 0.5 µM 2,5-DCBQ or 2,5-DBBQ

ACS Paragon Plus Environment

Environmental Science & Technology

Page 12 of 29

255

treatment groups, but was unchanged at higher concentrations (2.5, 5 µM) (ANOVA with Dunnett’s

256

post-test, p < 0.05; Fig. S3). Alternatively, mdm2 expression differed between the 2,5-DCBQ and

257

2,5-DBBQ treatment groups. Compared to the control, mdm2 was significantly down-regulated in larvae

258

exposed to 2.5 or 5 µM 2,5-DCBQ (2.71-fold, 2.76-fold) or 5 µM 2,5-DBBQ (2.12-fold), but was

259

significantly up-regulated (1.46-fold) in the 2.5 µM 2,5-DBBQ treatment group (ANOVA with Dunnett’s

260

post-test, p < 0.05; Fig. S3).

261

Both DNA damage and oxidative stress elicit the activation of p53,48 which in turn activates

262

apoptotic gene expression, leading to apoptosis.49,

50

263

significantly up-regulated in the lowest exposure group of 2,5-DCBQ and 2,5-DBBQ (0.5 µM), but was

264

not statistically changed in the higher exposure groups (2.5 and 5 µM). This is most likely related to the

265

expression levels of mdm2. The p53 protein is negatively regulated by mdm2, which suppresses p53

266

transcription.51 Hence, in the higher 2,5-DCBQ exposure groups, where mdm2 mRNA expression was

267

significantly down-regulated, upregulation of p53 was not necessary to maintain levels of p53 due to the

268

lack of its inhibitor. However, in the 2,5-DBBQ treatment groups, the relationship between p53 and

269

mdm2 is not as clear. These findings illustrate the delicate balance between these two proteins and the

270

control of apoptosis, and may indicate differing mechanisms of action between 2,5-DCBQ and

271

2,5-DBBQ. Nevertheless, the perturbation of these apoptosis genes indicate that HBQ exposure can

272

affect apoptosis signaling pathways in developing zebrafish.

273

HBQ-Induced Malformations in Zebrafish Larvae. 2,6-DCBQ, 2,5-DCBQ, and 2,5-DBBQ induced the

274

highest levels of mortality in exposed zebrafish embryos. Because 2,6-DCBQ is the most prevalent

275

HBQ-DBP detected in treated tap water,8 larvae exposed to these three HBQs were examined

276

morphologically to assess abnormal development. At 72 hpf, various physical malformations were

277

observed in the zebrafish larvae. As shown in Figure 6, the observed abnormalities included heart

278

malformations, curved spine and caudal fin, pericardial edema, and shortened body length and yolk sac

279

extension. Malformations were observed for exposure doses as low as 2 µM (Fig. 6B), where 2,5-DCBQ

280

clearly affected yolk sac extension and body length. Higher doses of 2,5-DCBQ (Fig. 6C-E) resulted in

281

more lethal damages, including severe tail injuries and pericardial edema. This was also observed in

Interestingly, p53 mRNA expression was

ACS Paragon Plus Environment

Page 13 of 29

Environmental Science & Technology

282

higher doses of 2,6-DCBQ (Fig. 6H-J), where caudal scoliosis and pericardial edema were identified

283

(Fig. S4). Lower doses of 2,6-DCBQ significantly inhibited the rate of larval development

284

(developmental delay; Table S5). Uninflated swim bladders were observed in up to 20% of the larvae

285

exposed to 2,6-DCBQ, 2,5-DCBQ, or 2,5-DBBQ (Table S5; Fig. S5). Consistent with our other findings,

286

the number of deformities was significantly reduced with NAC co-exposure (Fig. 6F), with the

287

malformation rate of the 10 µM treatment groups decreasing from nearly 80% to less than 15% (Table

288

S5).

289

Oxidative stress and apoptosis have both been implicated in developmental toxicity during zebrafish

290

embryogenesis.36, 52 Environmental stresses induce malformations and mortality in fish in both natural

291

and aquaculture environments, while stress-induced apoptosis is thought to contribute to abnormal

292

development during embryogenesis.53 Upon HBQ exposure, several developmental malformations were

293

observed in zebrafish embryos. Low doses of HBQs also prevented the expansion of the gas bladder (Fig.

294

6), which further contributed to limited motility and ingestion, and eventually increased mortality.54

295

Heart edema and the separation of the atrium-ventricle were observed (Fig. 6E, S4), suggesting that the

296

developing heart may be a sensitive target organ in zebrafish embryos. Malformations of the heart and

297

pericardium may impact cardiac function and lead to an abnormal heartbeat and circulation failure,

298

resulting in stunted development (Fig. 6I). However, no adverse effects of HBQs or HAAs on zebrafish

299

heart rate were observed at 72 hpf (Table S6). The high percentage of apoptotic cells around the heart

300

(Fig. 5) may partially account for the resulting heart malformations.

301

This study presents an efficient approach to evaluate the toxicity and potential mechanisms of DBPs

302

using zebrafish embryos as a model of developmental toxicity. The results of this study show that HBQs

303

induce ROS generation and inhibit the cells’ anti-oxidative response in developing zebrafish, resulting in

304

death, physical malformations, oxidative DNA damage, and apoptosis. These findings were consistent

305

with in vitro studies, which linked the cytotoxicity of HBQs to ROS generation, antioxidant enzyme

306

inhibition, and oxidative DNA damage in mammalian cells.55-57 The acute toxicity and ROS induction of

307

HBQs was up to 200 times more potent than the two regulated HAAs (DCA and DBA) and one of the

308

most toxic DBPs, IAA. Hence, ours and others results support that aromatic halogenated DBPs,

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 29

309

including halogenated phenolic and halopyrrole DBPs, induce significantly greater developmental

310

toxicity than aliphatic DBPs.58-60 More research into the pharmacodynamics of aliphatic and aromatic

311

halogenated DBPs are required to understand the toxicity difference of these DBP classes in vivo.

312

Furthermore, long-term exposures with environmentally-relevant concentrations of HBQs are required

313

to fully understand their toxicological significance.

314

ACS Paragon Plus Environment

Page 15 of 29

Environmental Science & Technology

315

ASSOCIATED CONTENT

316

Supporting Information

317

The Supporting Information is available free of charge on the ACS Publications website, including

318

additional details of the methods and results (six tables and five figures).

319

AUTHOR INFORMATION

320

Corresponding Authors

321

*E-mail: [email protected].

322

[email protected].

323

ORCID

324

Xing-Fang Li: 0000-0003-1844-7700 641

325

Funding

326

This work was financially supported by grants from the National Natural Science Foundation of China

327

(21677062, 21507155), and grants from the Natural Sciences and Engineering Research Council of

328

Canada (NSERC), Alberta Health, and Alberta Innovates-Energy and Environmental Solutions.

329

Acknowledgements

330

We thank Dr. Mengxi Cao from Jianghan University for her kind assistance with zebrafish feeding and

331

experiments.

332

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 29

333

References

334

1. Calderon, R. L. The epidemiology of chemical contaminants of drinking water. Food Chem. Toxicol.

335

2000, 38 (Suppl 1), S13-S20.

336

2. L Sedlak, D.; von Gunten, U. The Chlorine Dilemma. Science. 2011, 331 (6013), 42-43.

337

3. Richardson, S.; Plewa, M.; Wagner, E.; Schoeny, R.; Demarini, D. Occurrence, genotoxicity, and

338

carcinogenicity of regulated and emerging disinfection by-products in drinking water: a review and

339

roadmap for research. Mutat. Res. 2007, 636 (1-3), 178-242.

340

4. Michaud, D.; Kogevinas, M.; Cantor, K.; Villanueva, C.; Garcia-Closas, M.; Rothman, N.; Malats, N.;

341

Real, F.; Serra, C.; Garcia-Closas, R.; Tardon, A.; Carrato, A.; Dosemeci, M.; Silverman, D. Total fluid

342

and water consumption and the joint effect of exposure to disinfection by-products on risk of bladder

343

cancer. Environ. Health Perspect. 2007, 115 (11), 1569-1572.

344

5. Yang, S.; Wagner, B. A.; Lehmler, H. J.; Buettner, G. R. Semiquinone Radicals from Oxygenated

345

Polychlorinated Biphenyls: Electron Paramagnetic Resonance Studies. Chem. Res. Toxicol. 2008, 21 (7),

346

1359-1367.

347

6. Hua, G.; Reckhow, D., Comparison of disinfection byproduct formation from chlorine and alternative

348

disinfectants. Water Res. 2007, 41 (8), 1667-78.

349

7. Zhao, Y.; Anichina, J.; Lu, X.; Bull, R.; Krasner, S.; Hrudey, S.; Li, X. Occurrence and formation of

350

chloro- and bromo-benzoquinones during drinking water disinfection. Water Res. 2012, 46 (14),

351

4351-4360.

352

8. Zhai, H.; Zhang, X. Formation and decomposition of new and unknown polar brominated disinfection

353

byproducts during chlorination. Environ. Sci. Tech. 2011, 45 (22), 9750-9757.

354

9. Qin, F.; Zhao, Y.; Zhao, Y.; Boyd, J.; Zhou, W.; Li, X. A toxic disinfection by-product,

355

2,6-dichloro-1,4-benzoquinone, identified in drinking water. Angew. Chem. Int. Ed. Engl. 2010, 49 (4),

356

790-792.

357

10. Zhao, Y.; Qin, F.; Boyd, J.; Anichina, J.; Li, X., Characterization and determination of chloro- and

358

bromo-benzoquinones as new chlorination disinfection byproducts in drinking water. Anal. Chem. 2010,

359

82 (11), 4599-4605.

ACS Paragon Plus Environment

Page 17 of 29

Environmental Science & Technology

360

11. Wang, W.; Qian, Y.; Boyd, J.; Wu, M.; Hrudey, S.; Li, X. Halobenzoquinones in swimming pool

361

waters and their formation from personal care products. Environ. Sci. Technol. 2013, 47 (7), 3275-3282.

362

12. Li, J.; Wang, W.; Zhang, H.; Le, X.; Li, X. Glutathione-mediated detoxification of

363

halobenzoquinone drinking water disinfection byproducts in T24 cells. Toxicol. Sci. 2014, 141 (2),

364

335-343.

365

13. Bolton, J.; A. Trush, M.; M. Penning, T.; Dryhurst, G.; J. Monks, T. Role of Quinones in Toxicology.

366

Chem. Res. Toxicol. 2000, 13 (3), 135-160.

367

14. Monks, T. J.; Jones, D. J. The Metabolism and Toxicity of Quinones, Quinonimines, Quinone

368

Methides, and Quinone-Thioethers. Curr. Drug Metabolism. 2002, 3 (4), 425-438.

369

15. Du, H.; Li, J.; Moe, B.; McGuigan, C. F.; Shen, S.; Li, X. F. Cytotoxicity and oxidative damage

370

induced by halobenzoquinones to T24 bladder cancer cells. Environ. Sci. Tech. 2013, 47 (6), 2823-2830.

371

16. Song, Y.; Wagner, B.; Witmer, J.; Lehmler, H.; Buettner, G. Nonenzymatic displacement of chlorine

372

and formation of free radicals upon the reaction of glutathione with PCB quinones. Proc. Natl. Acad. Sci.

373

U.S.A. 2009, 106 (24), 9725-9730.

374

17. Nguyen, M. L.; Westerhoff, P.; Baker, L.; Hu, Q.; Esparza-Soto, M.; Sommerfeld, M. Characteristics

375

and Reactivity of Algae-Produced Dissolved Organic Carbon. J. Environ. Eng. 2005, 131 (11),

376

1574-1582.

377

18. Waidyanatha, S.; McDonald, T. A.; Lin, P. H.; Rappaport, S. M. Measurement of Hemoglobin and

378

Albumin Adducts of Tetrachlorobenzoquinone. Chem. Res. Toxicol. 1994, 7 (3), 463-468.

379

19. Li, J.; Moe, B.; Vemula, S.; Wang, W.; Li, X-F. Emerging Disinfection Byproducts,

380

Halobenzoquinones: Effects of Isomeric Structure and Halogen Substitution on Cytotoxicity, Formation

381

of Reactive Oxygen Species, and Genotoxicity. Environ. Sci. Tech. 2016, 50, 6744−6752.

382

20. Li, J.; Moe, B.; Liu, Y.; Li, X-F. Halobenzoquinone-Induced Alteration of Gene Expression

383

Associated with Oxidative Stress Signaling Pathways. Environ. Sci. Tech. 2018, 52, 6576−6584.

384

21. Fu, K.; Li, J.; Vemula, S.; Moe, B.; Li, X. F. Effects of halobenzoquinone and haloacetic acid water

385

disinfection byproducts on human neural stem cells. J. Environ. Sci. 2017, 58 (5), 239-249.

386

22. Nieuwenhuijsen, M.; Grellier, J.; Smith, R.; Iszatt, N.; Bennett, J.; Best, N.; Toledano, M. The

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 29

387

epidemiology and possible mechanisms of disinfection by-products in drinking water. Philos Trans.

388

2009, 367 (1904), 4043-4076.

389

23. Colman, J.; Rice, G. E.; Wright, J. M.; Hunter, E. S.; Teuschler, L. K.; Lipscomb, J. C.; Hertzberg, R.

390

C.; Simmons, J. E.; Fransen, M.; Osier, M.; Narotsky, M. G. Identification of developmentally toxic

391

drinking water disinfection byproducts and evaluation of data relevant to mode of action. Toxicol.

392

Applied Pharm. 2011, 254 (2), 100-126.

393

24. Levallois, P.; Gingras, S.; Marcoux, S.; Legay, C.; Catto, C.; Rodriguez, M.; Tardif, R. Maternal

394

exposure to drinking-water chlorination by-products and small-for-gestational-age neonates.

395

Epidemiology. 2012, 23 (2), 267-276.

396

25. Grazuleviciene, R.; Kapustinskiene, V.; Vencloviene, J.; Buinauskiene, J.; Nieuwenhuijsen, M. Risk

397

of congenital anomalies in relation to the uptake of trihalomethane from drinking water during

398

pregnancy. Occup. Environ. Med. 2013, 70 (4), 274-282.

399

26. Scholz, S.; Fischer, S.; Gündel, U.; Küster, E.; Luckenbach, T.; Voelker, D. The zebrafish embryo

400

model in environmental risk assessment-applications beyond acute toxicity testing. Environ. Sci.

401

Pollution Res. 2008, 15 (5), 394-404.

402

27. Spitsbergen, J. M.; Kent, M. L. The state of the art of the zebrafish model for toxicology and

403

toxicologic pathology research-advantages and current limitations. Toxicol. Pathol. 2003, 31, 62-87.

404

28. Yu, S.; Lin, T.; Chen, W.; Tao, H. The toxicity of a new disinfection by-product,

405

2,2-dichloroacetamide (Dcacam), on adult zebrafish (Danio rerio) and its occurrence in the chlorinated

406

drinking water. Chemosphere. 2015, 139, 40-46.

407

29. Lin, T.; Zhou, D.; Dong, J.; Jiang, F.; Chen, W. Acute toxicity of dichloroacetonitrile (Dcan), a

408

typical nitrogenous disinfection by-product (N-DBP), on zebrafish (Danio rerio). Ecotoxicol. Environ.

409

Saf. 2016, 133, 97-104.

410

30. Lin, T.; Zhou, D.; Yu, S.; Chen, W. The removal process of 2,2-dichloroacetamide (Dcacam), a new

411

disinfection by-product, in drinking water treatment process and its toxicity on zebrafish. Chemosphere.

412

2016, 159, 403-411.

413

31. Hanigan, D.; Truong, L.; Simonich, M.; Tanguay, R.; Westerhoff, P. Zebrafish embryo toxicity of 15

ACS Paragon Plus Environment

Page 19 of 29

Environmental Science & Technology

414

chlorinated, brominated, and iodinated disinfection by-products. J. Environ. Sci. 2017, 58 (8), 302-310.

415

32. Yang, M.; Liu, X.; Zhang, X.; Richardson, S. D. Comparative toxicity of Chlorinated Saline and

416

Freshwater Wastewater Effluents to Marine Organisms. Environ. sci. Tech. 2015, 49 (24), 14475-14483.

417

33. Yang, M.; Zhang, X. Comparative developmental toxicity of new aromatic halogenated DBPs in a

418

chlorinated saline sewage effluent to the marine polychaete Platynereis dumerilii. Environ. sci. Tech.

419

2013, 47 (19), 10868-10876.

420

34. Richardson, S. D.; Postigo, C., Drinking Water Disinfection By-products. In Emerging Organic

421

Contaminants and Human Health, Barceló, D., Ed. Springer Berlin Heidelberg. 2012, 20, 93-137.

422

35. Liu, W.; Tao, F.; Zhang, W.; Li, S.; Zheng, M., Contamination and emission factors of PCDD/Fs,

423

unintentional PCBs, HxCBz, PeCBz and polychlorophenols in chloranil in China. Chemosphere. 2012,

424

86 (3), 248-251.

425

36. Usenko, C. Y.; Harper, S. L.; Tanguay, R. L. Fullerene C60 exposure elicits an oxidative stress

426

response in embryonic zebrafish. Toxicol. Applied. Pharm. 2008, 229 (1), 44-55.

427

37. Liu, H.; Sheng, N.; Zhang, W.; Dai, J. Toxic effects of perfluorononanoic acid on the development

428

of Zebrafish (Danio rerio) embryos. J. Environ. Sci. 2015, 32 (Suppl C), 26-34.

429

38. Chan, P.; Cheng, S. Cadmium-induced ectopic apoptosis in zebrafish embryos. Arch. Toxicol. 2003,

430

77 (2), 69-79.

431

39. Livak, K. J.; Schmittgen, T. D. Analysis of Relative Gene Expression Data Using Real-Time

432

Quantitative PCR and the 2−∆∆CT Method. Methods. 2001, 25 (4), 402-408.

433

40. McHugh, M.L. Multiple comparison analysis testing in ANOVA. Biochemia Medica 2011, 21(3),

434

203–209.

435

41. Aruoma, O.; Halliwell, B.; Hoey, B.; Butler, J. The antioxidant action of N-acetylcysteine: its

436

reaction with hydrogen peroxide, hydroxyl radical, superoxide, and hypochlorous acid. Free Radic. Biol.

437

Med. 1989, 6 (6), 593-597.

438

42. Li, X.; Liu, Y.; Song, L.; Liu, J. Responses of antioxidant systems in the hepatocytes of common

439

carp (Cyprinus carpio L.) to the toxicity of microcystin-LR. Toxicon. 2003, 42 (1), 85-89.

440

43. Siraki, A.; Chan, T.; O'Brien, P. Application of quantitative structure-toxicity relationships for the

ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 29

441

comparison of the cytotoxicity of 14 p-benzoquinone congeners in primary cultured rat hepatocytes

442

versus PC12 cells. Toxicol. Sci. 2004, 81 (1), 148-159.

443

44. Liu, J.; Song, E.; Liu, L.; Ma, X.; Tian, X.; Dong, H.; Song, Y., Polychlorinated biphenyl quinone

444

metabolites lead to oxidative stress in HepG2 cells and the protective role of dihydrolipoic acid. Toxicol.

445

In Vitro. 2012, 26 (6), 841-848.

446

45. Hayes, J.; McLellan, L. Glutathione and glutathione-dependent enzymes represent a co-ordinately

447

regulated defence against oxidative stress. Free Radic. Res. 1999, 31 (4), 273-300.

448

46. Wang, Wei.; Qian, Y. C.; Li, J.; Aljuhani, N.; Siraki, A. G.; Le, X. C.; Li X. F. Characterization of

449

Mechanisms of Glutathione Conjugation with Halobenzoquinones in Solution and HepG2 Cells. Environ.

450

Sci. Tech. 2018, 52 (5), 2898-2908.

451

47. Yin, R.; Zhang, D.; Song, Y.; Zhu, B.-Z.; Wang, H. Potent DNA damage by polyhalogenated

452

quinones and H2O2 via a metal-independent and intercalation-enhanced oxidation mechanism. Sci. Rep.

453

2013, 3 (7436), 1269.

454

48. Levine, A. p53, the cellular gatekeeper for growth and division. Cell. 1997, 88 (3), 323-331.

455

49. Langheinrich, U.; Hennen, E.; Stott, G.; Vacun, G. Zebrafish as a model organism for the

456

identification and characterization of drugs and genes affecting p53 signaling. Curr. Biol. 2002, 12 (23),

457

2023-2028.

458

50. Saeed Sheikh, M.; Fornace, A., Role of p53 family member in apoptosis. 2000, 182, 171-181.

459

51. Moll, U. M.; Petrenko, O. The MDM2-p53 Interaction. Mol. Cancer. Res. 2003, 1 (14), 1001-1008.

460

52. Yamashita, M. Apoptosis in zebrafish development. Comp. Biochem. Physiol. Part B: Biochem. Mol.

461

Biol. 2003, 136 (4), 731-742.

462

53. Takeshi, Y.; Setsuko, T.; Michiaki, Y. Stress-induced apoptosis by heat shock, UV and γ-ray

463

irradiation in zebrafish embryos detected by increased caspase activity and whole-mount TUNEL

464

staining. Fisheries Science 2001, 67 (2), 333-340.

465

54. Li, J.; Liang, Y.; Zhang, X.; Lu, J.; Zhang, J.; Ruan, T.; Zhou, Q.; Jiang, G. Impaired gas bladder

466

inflation in zebrafish exposed to a novel heterocyclic brominated flame retardant tris(2,3-dibromopropyl)

467

isocyanurate. Environ. Sci. Tech. 2011, 45 (22), 9750-9757.

ACS Paragon Plus Environment

Page 21 of 29

Environmental Science & Technology

468

55. Felix, L. M.; Vidal, A. M.; Serafim, C.; Valentim, A. M.; Antunes, L. M.; Campos, S.; Matos, M.;

469

Monteiro, S. M.; Coimbra, A. M. Ketamine-induced oxidative stress at different developmental stages of

470

zebrafish (Danio rerio) embryos. RSC Advances. 2016, 6 (66), 61254-61266.

471

56. Hileman, E.; Achanta, G.; Huang, P. Superoxide dismutase: an emerging target for cancer

472

therapeutics. Expert Opin. Ther. Targets. 2001, 5 (6), 697-710.

473

57. Xu, D.; Hu, L.; Su, C.; Xia, X.; Zhang, P.; Fu, J.; Wang, W.; Xu, D.; Du, H.; Hu, Q.; Song, E.; Song,

474

Y., Tetrachloro-p-benzoquinone induces hepatic oxidative damage and inflammatory response, but not

475

apoptosis in mouse: The prevention of curcumin. Toxicol. Applied Pharm. 2014, 280 (2), 305-313.

476

58. Yang, Mengting, and X. Zhang. Halopyrroles: A New Group of Highly Toxic Disinfection

477

Byproducts Formed in Chlorinated Saline Wastewater. Environ. Sci. Tech. 2014, 48 (20), 11846-11852.

478

59. Yang Pan; Xiangru Zhang; Yu Li. Identification, toxicity and control of iodinated disinfection

479

byproducts in cooking with simulated chlor(am)inated tap water and iodized table salt. Water Research.

480

2016, 88, 60-68.

481

60. Jiang, J.; Zhang, X.; Zhu, X.; Li, Y. Removal of intermediate aromatic halogenated DBPs by

482

activated carbon absorption: A new approach to controlling halogenated DBPs in chlorinated drinking

483

water. Environ. Sci. Tech. 2017, 51, 3435-3444.

ACS Paragon Plus Environment

Environmental Science & Technology

484

Page 22 of 29

Table 1. Calculated LC50 Values of the HBQs and HAAs at 24 hpf Compounds Halobenzoquinones (HBQs) 2,5-DCBQ 2,6-DCBQ 2,5-DBBQ TBBQ TCBQ Haloacetic acids (HAAs) IAA DBA DCA

LC50 (mean ± SD) 2,5-dichloro-1,4-benzoquinone 2,6-dichloro-1,4-benzoquinone 2,5-dibromo-1,4-benzoquinone tetrabromo-1,4-benzoquinone tetrachloro-1,4-benzoquinone

µM 4.6 ± 0.2 6.6 ± 0.2 5.6 ± 0.2 9.4 ± 0.3 9.8 ± 0.4

Iodoacetic acid Dibromoacetic acid Dichloroacetic acid

1900 ± 240 2200 ± 190 2600 ± 170

485 486 487

ACS Paragon Plus Environment

Page 23 of 29

Environmental Science & Technology

488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503

Figure 1. Effect of HBQs (1, 2, 5, 8, 10, 16 µM) on the mortality of zebrafish embryos at 24 hpf (A);

504

Effect of NAC co-exposure on HBQ-induced mortality (B). Error bars indicate mean ± SD. Three

505

replicate experiments were performed for each concentration (n = 3). t-test: HBQ group vs. respective

506

HBQ + NAC group: ***p < 0.001.

507 508

ACS Paragon Plus Environment

Environmental Science & Technology

Page 24 of 29

509

510 511

Figure 2. ROS formation in larvae treated with 2,5-DCBQ or 2,5-DBBQ (1, 5, 10 µM) with or without

512

NAC co-exposure (A1). Effects of different concentrations (0.5, 1 mM) of DCA, DBA, IAA, or H2O2 on

513

ROS production in comparison to 2,5-DCBQ (A2). Effects of different concentrations (1, 5, 10 µM) of

514

five HBQs on SOD activity with or without NAC co-exposure (B). Effect of different concentrations (1,

515

5, 10 µM) of 2,5-DCBQ and 2,5-DBBQ on GSH levels with or without NAC co-exposure (C). Error

516

bars indicate mean ± SEM. Three replicate experiments were performed for each group. ANOVA with

517

Dunnett’s post-test: *p < 0.05, **p < 0.01, ***p < 0.001, ****p < 0.0001.

518

ACS Paragon Plus Environment

Page 25 of 29

Environmental Science & Technology

519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534

Figure 3. In vivo assessment of ROS production in larvae at 120 hpf in the negative control (A),

535

DCFDA control (B), and the 2,5-DCBQ treatment groups (0.5, 2.5, 5 µM) (D-F). Co-exposure of 5 µM

536

2,5-DCBQ with 50 µM NAC (C) attenuated the detected fluorescence. FE: fluorescence enhancement;

537

PE: pericardial edema; SYSE: shortened yolk sac extension; SAV: separation of the atrium-ventricle;

538

USB: uninflated swim bladder; ISB: inflated swim bladder. Scale bar, 450 µm.

539 540

ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 29

541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558

Figure 4. Increase in 8-OHdG lesions at 120 hpf in the genomic DNA of zebrafish larvae after HBQ (1,

559

5, 10 µM) exposure (A). DNA fragmentation is also observed in the 2,5-DCBQ (B1) and 2,5-DBBQ (B2)

560

treatment groups (1, 5, 10 µM). Error bars indicate the mean ± SEM. Three replicate experiments were

561

performed for each exposure group (n = 3). ANOVA with Dunnett’s post-test: *p