Metabolism of a phenylarsenical in human hepatic ... - ACS Publications

litter as a fertilizer increases the arsenic concentration in the soil, crops, and water in the vicinity. 40. 6-8 . 41. Small amounts of arsenic speci...
0 downloads 9 Views 598KB Size
Subscriber access provided by RUTGERS UNIVERSITY

Article

Metabolism of a phenylarsenical in human hepatic cells and identification of a new arsenic metabolite Qingqing Liu, Elaine M. Leslie, Birget Moe, Hongquan Zhang, Donna N. Douglas, Norman M. Kneteman, and X. Chris Le Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.7b05081 • Publication Date (Web): 27 Dec 2017 Downloaded from http://pubs.acs.org on December 28, 2017

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 free 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 accessible to all readers and 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.

Environmental Science & Technology 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

TOC art 84x47mm (300 x 300 DPI)

ACS Paragon Plus Environment

Environmental Science & Technology

Page 2 of 29

1

Metabolism of a phenylarsenical in human hepatic cells and identification of a new arsenic

2

metabolite

3 4

Qingqing Liu †, Elaine M. Leslie ‡,†, Birget Moe †, Hongquan Zhang †, Donna N. Douglas §,

5

Norman M. Kneteman §, X. Chris Le *,†

6



7

Clinical Sciences Building, University of Alberta, Edmonton, Alberta, Canada T6G 2G3

8



9

Building, University of Alberta, Edmonton, Alberta, Canada T6G 2H7

Department of Laboratory Medicine and Pathology, Faculty of Medicine and Dentistry, 10-102

Department of Physiology, Faculty of Medicine and Dentistry, 7-08A Medical Sciences

10

§

11

Sciences Centre, University of Alberta, Edmonton, Alberta, Canada T6G 2B7

12

* Corresponding author. Tel.: (780) 492-6416; Fax: 1 (780) 492-7800; E-mail: [email protected]

Department of Surgery, Faculty of Medicine and Dentistry, Walter C Mackenzie Health

13

1 ACS Paragon Plus Environment

Page 3 of 29

Environmental Science & Technology

14

ABSTRACT

15

Environmental contamination and human consumption of chickens could result in potential

16

exposure to Roxarsone (3-nitro-4-hydroxyphenylarsonic acid), an organic arsenical that has been

17

used as a chicken feed additive in many countries. However, little is known about the

18

metabolism of Roxarsone in humans. The objective of this research was to investigate the

19

metabolism of Roxarsone in human liver cells and to identify new arsenic metabolites of

20

toxicological significance. Human primary hepatocytes and hepatocellular carcinoma HepG2

21

cells were treated with 20 or 100 µM Roxarsone. Arsenic species were characterized using a

22

strategy of complementary chromatography and mass spectrometry. Results showed that

23

Roxarsone was metabolized to more than 10 arsenic species in human hepatic cells. A new

24

metabolite was identified as a thiolated Roxarsone. The 24-h IC50 values of thiolated Roxarsone

25

for A549 lung cancer cells and T24 bladder cancer cells were 380 ± 80 µM and 42 ± 10 µM,

26

respectively, which was more toxic than Roxarsone whose 24-h IC50 of A549 and T24 were 9300

27

± 1600 µM and 6800 ± 740 µM, respectively. The identification and toxicological studies of the

28

new arsenic metabolite are useful for understanding the fate of arsenic species and assessing the

29

potential impact of human exposure to Roxarsone.

30

2 ACS Paragon Plus Environment

Environmental Science & Technology

Page 4 of 29

31

INTRODUCTION

32

Inorganic arsenic is known to be a significant environment hazard. Chronic exposure to

33

inorganic arsenite (AsIII) through the ingestion of contaminated water and food results in lung,

34

skin, and bladder cancers, and has a strong association with ischemic heart disease, diabetes, and

35

respiratory disease 1. 3-nitro-4-hydroxyphenylarsonic acid (Roxarsone, Rox) is an organic

36

arsenic species which has been widely used as a chicken feed additive for over 70 years 2, 3, for

37

the purpose of infection control and growth promotion. Most of Rox taken up by chickens is

38

excreted; and chicken waste contains as much as 30-60 mg/kg of Rox 4. The chicken litter, which

39

contains Rox and other arsenic species 5, is often used as a farm fertilizer. The use of chicken

40

litter as a fertilizer increases the arsenic concentration in the soil, crops, and water in the vicinity

41

6-8

42

. Small amounts of arsenic species, including inorganic arsenicals and Rox, have been

43

detected in chicken meat and livers after feeding chickens with Rox 9, 10, 11. Residual arsenicals,

44

albeit at trace concentrations, were detected in chicken meat seven days after the feeding of Rox

45

was stopped 12. Arsenic species in chickens contribute to human dietary exposure to arsenic 2, 3.

46

Therefore, there is considerable concern regarding human exposure and health risks posed by the

47

poultry use of Rox 13. Although the European Union and the United States have stopped

48

application of Rox in poultry industry 14, 15, many countries are still using Rox as an animal feed

49

supplement 16.

50

To understand the fate of Rox and the toxicological relevance of its metabolites in

51

humans, we aimed to investigate the metabolism of Rox in human cells. Because liver is the

52

primary site for the metabolism of inorganic arsenic 17, we chose to study metabolism of Rox in

53

human liver cells. In the present work, we examined the hepatic biotransformation of Rox by

3 ACS Paragon Plus Environment

Page 5 of 29

Environmental Science & Technology

54

using human hepatocellular carcinoma HepG2 epithelial cells and primary hepatocytes from

55

human donors. We describe here results from analyses of arsenic species, identification of a new

56

metabolite, and testing of its cytotoxicity. We demonstrate the identification of thiolated Rox as a

57

new metabolite and show its higher toxicity than the original compound Rox.

58 59 60

MATERIALS AND METHODS Materials. Stock solutions (1 mg/L) of arsenobetaine (AsB), AsIII, arsenate (AsV),

61

monomethylarsonic acid (MMA), dimethylarsinic acid (DMA), 3-amino-4-

62

hydroxyphenylarsonic acid (3AHPAA), N-acetyl-4-hydroxy-m-arsanilic acid (NAHAA), and

63

Rox were prepared from arsenobetaine (98% purity, Tri Chemical Laboratories Inc., Japan),

64

sodium m-arsenite (97%, Sigma-Aldrich, St. Louis, MO), sodium arsenate (99.4%, Sigma-

65

Aldrich), monosodium acid methane arsonate (99%, Chem Service, West Chester, PA),

66

cacodylic acid (98%, Sigma-Aldrich), 3-amino-4-hydroxyphenylarsonic acid (Pfaltz and Bauer

67

Inc., Waterbury, CT), N-acetyl-4-hydroxy-m-arsanilic acid (Pfaltz and Bauer Inc., Waterbury,

68

CT), and 3-nitro-4-hydroxyphenylarsonic acid (98.1%, Sigma-Aldrich), respectively. The

69

concentrations of the arsenic species were calibrated against a primary arsenic standard (Agilent

70

Technologies, U.S.) and were determined using inductively coupled plasma mass spectrometry

71

(ICPMS) Agilent 7500cs system (Agilent Technologies, Germany). Milli-Q18.2 MΩ·cm

72

deionized water (Millipore Corporation, Billerica, MA) was used to prepare arsenic solutions.

73

Phenol red-free Dulbecco’s modified Eagle’s medium (DMEM) were purchased from Corning

74

(Corning, NY). Eagle’s Minimum Essential Medium (EMEM) was purchased from American

75

Type Culture Collection (ATCC, Manassas, VA). Fetal bovine serum (FBS), ammonium

76

bicarbonate (NH4HCO3), and Hank’s balanced salt solution (HBSS) were purchased from

4 ACS Paragon Plus Environment

Environmental Science & Technology

77

Sigma-Aldrich (St. Louis, MO). Human recombinant insulin was purchased from Life

78

Technologies (Carlsbad, CA). Trypsin-ethylenediaminetetraacetic acid (EDTA), phosphate

79

buffered saline (PBS), nonessential amino acids (NEAA), penicillin-streptomycin solution

80

(Penstrep), L-glutamine, Dextran, RIPA buffer [50 mM Tris-HCl, pH 8.0, with 150 mM sodium

81

chloride, 1.0% Igepal CA-630 (NP-40), 0.5% sodium deoxycholate, and 0.1% sodium dodecyl

82

sulfate] were purchased from Gibco (Burlington, ON, Canada). ITS+ (6.25 µg/mL insulin, 6.25

83

µg/mL transferrin, and 6.25 ng/mL selenium) and Biocoat culture plates were purchased from

84

BD Biosciences (Franklin Lakes, NJ). Methanol was from Thermo Fisher Scientific (Waltham,

85

MA).

86

Human primary hepatocytes. Ethical approval was obtained from the University of

87

Alberta Faculty of Medicine Research Ethics Board and informed consent from all hepatocyte

88

donors. Human hepatocytes were from two adult patients undergoing hepatic resection at the

89

University of Alberta Hospital (Edmonton, Alberta, Canada) by qualified medical staff.

90

Segments of human liver tissue (15–20 cm3) were obtained from the margin of normal tissue

91

surrounding resected tumor specimens using a modified 2-step collagenase perfusion method 18

92

and cultured on collagen coated plates. Hepatocytes were seeded on type I rat tail collagen

93

Biocoat 6-well culture plates following the procedure of Roggenbeck et al. 19, except without

94

collagen overlay.

95

Page 6 of 29

For hepatocytes, ~18 h after seeding, arsenicals (1 µM AsIII, 20 µM Rox, or 100 µM Rox)

96

were added into the culture medium and then incubated for 24 hours. After 24 h, cells were

97

washed twice with 1×PBS and lysed with 200 µL RIPA buffer.

5 ACS Paragon Plus Environment

Page 7 of 29

98 99 100 101

Environmental Science & Technology

HepG2 cells. HepG2 cells (ATCC, Manassas, VA) were grown in EMEM containing 10% FBS and 1% Penstrep. HepG2 cells were used between passage 7 and 17, and incubated at 37 oC with 5% CO2. After 3 days seeding, 20 µM or 100 µM Rox was added to the culture medium and

102

incubated for 24 h. The cells were washed twice with 1×PBS, trypsinized, and lysed with

103

repeated freezing in liquid nitrogen and thawing.

104

Preparation of cell samples for analysis. For both types of cells, the cell lysates were

105

centrifuged at 8000×g for 5 min to remove the cell debris. Supernatant (300 µL) was filtered

106

with a 3000 Dalton cut-off membrane ultracentrifugation unit (Millipore, Canada) at 14000×g

107

for 30 minutes. The filtrates were used for arsenic speciation analyses. Another 10 µL of the

108

supernatant was used for the determination of protein concentration by Bradford assay. The

109

samples were stored in a -20 oC freezer.

110

Synthesis of a thiolated Rox standard. A thiolated Rox standard was synthesized

111

according to the previously reported method 20, 21. Briefly, Rox (10 mM) was mixed with Na2S

112

(16 mM) in 10 mL deionized water. H2SO4 (16 mM) was added dropwise into the incubation

113

mixture. The reaction mixture was left to stand for 1 h and the solution was diluted 104 times and

114

filtered through a 0.45-µm membrane filter. The purity of the synthesized thiolated Rox was

115

90%, which was determined on the basis of chromatographic peaks. The synthesis reaction is

116

shown in Equation 1.

117

6 ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 29

118 119

(eq.1)

120

Arsenic speciation analysis using high performance liquid chromatography (HPLC)

121

coupled with ICPMS. An Agilent 1290 series HPLC system, an anion exchange column (PRP

122

X110s, 150 mm × 4.6 mm, 7-µm particle size; Hamilton, Reno, NV) and a reversed phase

123

column (ODS-3, 100 mm × 2 mm, 3-µm particle size; Phenomenex, Torrance, CA) were used

124

for the separation of arsenicals. In the anion exchange separation, the mobile phase A contained

125

5% methanol in water, and mobile phase B contained 60 mM NH4HCO3 and 5% methanol in

126

water (pH 8.75). The flow rate was 2 mL/min. A 10 µL aliquot of a sample was injected and the

127

separation was performed in a gradient. The mobile phase B was increased from 0% to 2% in the

128

first 2 minutes, to 30% in the next 13 min, to 100% in the next 7 min, kept at 100% for 2 min,

129

and then decreased to 0% in the next 1 min. The column was equilibrated in mobile phase A for

130

4 min before the next analysis. A typical chromatogram obtained from the anion exchange

131

HPLC-ICPMS analysis of arsenic species in water is shown in Figure 1a.

132

In the reversed phase chromatographic separation, 0.2% formic acid in water was used as

133

mobile phase A and methanol was used as mobile phase B. The flow rate was 0.15 mL/min. A

134

10 µL aliquot of a sample was injected and the separation was performed in a gradient. The

135

mobile phase B was linearly increased from 5% to 30% in the first 10 min, then increased from

136

30% to 50% from 10 to 30 min and decreased to 5% in the next 1 min. The column was kept in

137

95% mobile phase A and 5% mobile phase B for 9 min before the next analysis.

7 ACS Paragon Plus Environment

Page 9 of 29

138

Environmental Science & Technology

The effluent from HPLC system was directly introduced into the nebulizer of a 7500cs

139

ICPMS. Arsenic was monitored at m/z 75. When separation on HPLC was run in anion exchange

140

mode, instrument parameters were shown in Supporting Information (SI) Table S1. When

141

separation was run in reversed phase mode, 20% oxygen in argon gas was used as optional gas.

142

Spray chamber temperature was decreased to -5 oC. Other parameters were kept the same as

143

shown in SI Table S1.

144

Characterization of arsenic species using HPLC coupled with electrospray

145

ionization (ESI) hybrid quadrupole time-of-flight mass spectrometry (QqToFMS). The

146

overall scheme for the characterization of arsenic species, including an unknown arsenic species,

147

is shown in SI Figure S1. Reversed phase HPLC was hyphenated with ESI QqToFMS TripleToF

148

5600 system (AB Sciex, Ontario, Canada). The separation condition was described above. The

149

instrument parameters of QqToFMS are shown in SI Table S1. Two scans, MS and MS/MS,

150

using the ToF mass spectrometer, were looped in one cycle 22, 23. These two ToF scans had the

151

same source parameter, except that the first ToF scan had collision energy (CE) of -5 eV and the

152

second one had CE of -45 eV. For the detection of arsenic species, the retention time of the

153

arsenic fragments in the second ToF scan must be the same as that of the precursor ions in the

154

first scan.

155

The first step in extracting the data from the above MS and MS/MS scans was to search

156

for the specific arsenic fragments in the second scan. Possible arsenic fragments were proposed

157

on the basis of previous reports which have shown arsenic fragment ions 9, 24-28. Their retention

158

time at which these fragment ions were detected was used to find the candidate precursors in the

159

first scan. The “Non-Targeted Peak Finding” function in the “XIC Manager” which was an add-

160

on for the PeakView™ 2.0 software (AB Sciex, Ontario, Canada) was used to find all peaks in

8 ACS Paragon Plus Environment

Environmental Science & Technology

161

the first scan. The peak finding criteria in this function were set as: ‘Approximate LC peak

162

width’ 1 min, ‘Minimum intensity in counts’ 5 counts, and ‘Chemical noise intensity multiplier’

163

1.5. Application of this function resulted in a table showing the detected ions at their respective

164

retention time. The precursor ions detected at the retention times when the intensity of arsenic

165

fragment ions was the highest were chosen. After candidate precursor ions were selected, a new

166

product ion scan in QqToFMS was used to confirm that these ions could indeed generate the

167

expected fragment ions.

Page 10 of 29

168

Arsenic speciation analysis using HPLC separation and simultaneous detection with

169

ICPMS and ESIMS. ESI triple quadrupole mass spectrometry (QqQMS) and ICPMS were used

170

simultaneously for HPLC detection 9, 29. A QTRAP 5500 system (AB Sciex, Ontario, Canada)

171

was operated in the multiple reactions monitoring (MRM) mode. After HPLC separation, the

172

effluent was split into the two mass spectrometers. The instrument parameters of the QqQMS are

173

shown in SI Table S1.

174

Cytotoxicity of thiolated Rox. Cytotoxicity of thiolated Rox, Rox, and AsIII was

175

evaluated on human bladder epithelial cancer cell line T24 (ATCC, Manassas, VA) and human

176

lung epithelial cancer cell line A549 (ATCC, Manassas, VA) using the real-time cell electronic

177

sensing technique 30.

178 179

RESULTS AND DISCUSSIONS

180

Metabolism of Rox by HepG2 cells and human primary hepatocytes. We first

181

investigated the Rox metabolism in HepG2 cells. The chromatogram obtained from anion

182

exchange HPLC-ICPMS analysis of HepG2 exposed to 20 µM Rox is shown in Figure 1b. In

183

addition to Rox, DMA, MMA, AsV, 3AHPAA, and NAHAA, several unknown arsenic species

9 ACS Paragon Plus Environment

Page 11 of 29

Environmental Science & Technology

184

were detectable in HepG2 cells after exposure to 20 µM Rox for 24 h. HepG2 cells exposed to

185

100 µM Rox for 24 h also showed the presence of these arsenic species (SI Figure S2).

186

We then examined the metabolism of Rox by human primary hepatocytes (liver cells)

187

isolated from two human liver samples. A typical chromatogram of arsenic species in the

188

hepatocytes exposed to 20 µM Rox is shown in Figure 1c. These hepatocytes were from human

189

donor 1. Rox, MMA, AsV, and some unidentified arsenic species were detectable in these

190

hepatocytes after 24 hours incubation with 20 µM Rox. Chromatogram from HPLC-ICPMS

191

analysis of hepatocytes exposed to 100 µM Rox also showed the presence of these arsenic

192

species (SI Figure S2). These results demonstrate that the human primary hepatocytes were able

193

to metabolize Rox.

194

To test that the hepatocytes used in this study was metabolically active, we measured the

195

methylation of AsIII as a control. Methylation of arsenic after treatment with AsIII has been

196

previously characterized using plated human primary hepatocytes. According to Styblo et al. 31,

197

when human hepatocytes were exposed to 1 µM AsIII, the apparent methylation rate (AMR),

198

calculated as the amount of AsIII converted to MMA and DMA per hour per 106 cells, was 3.1-

199

35.7 pmol/h/106 cells; and the ratio of DMA/MMA was 0.03-2.9 over 4 hepatocyte preparations,

200

indicating considerable inter-individual variability in methylation. In our study, the plated

201

hepatocytes exposed to 1 µM AsIII had AMR at 5 and 6.5 pmol/h/106 cells for the hepatocytes

202

from human donor 1 and 2, respectively. The ratios of DMA/MMA were 0.8 and 1.7 for the

203

hepatocytes from human donor 1 and 2, respectively. These results are at the same magnitude or

204

within the range of the results reported by Styblo et al. 31, which showed that the hepatocytes in

205

our study were metabolically active.

10 ACS Paragon Plus Environment

Environmental Science & Technology

206

Page 12 of 29

The concentrations of arsenic species in HepG2 and human primary hepatocytes after 20

207

µM and 100 µM Rox exposure are summarized in Table 1. When cells were exposed to 20 µM

208

Rox, HepG2 cells produced more DMA, 3AHPAA, and NAHAA (not detected in hepatocytes)

209

but less MMA and AsV (3-fold lower for HepG2) than hepatocytes from human donor 1. When

210

cells were exposed to 100 µM Rox, the comparison between hepatocytes from human donor 1

211

and HepG2 cells suggests that the primary hepatocytes have the potential to produce more MMA

212

(200-fold), AsV (2-fold), and NAHAA (8-fold) than HepG2.

213

Consistent with the previous study showing considerable inter-individual variability in

214

the methylations of AsIII by hepatocytes from four donors 31, we have also observed inter-

215

individual variability in the metabolic activity of our hepatocyte preparations for Rox.

216

Hepatocytes from human donor 2 produced more AsV but less DMA, MMA, 3AHPAA, and

217

NAHAA than HepG2 cells. The AsV concentration generated in the hepatocytes from human

218

donor 2 was at least 15- and 5-times more than the ones from HepG2 cells when the cells were

219

exposed to 20 µM and 100 µM Rox, respectively. DMA, MMA, 3AHPAA and NAHAA were

220

non-detectable in hepatocytes from human donor 2 at both Rox concentration levels.

221

Rox has relatively low bioavailability compared to other arsenic species, such as AsIII and

222

AsB 32. Our study has demonstrated that the low percentage of the administrated Rox taken up

223

by HepG2 cells or human primary hepatocytes undergoes metabolism. The sum concentration of

224

metabolites only accounted for 9.7-10.9 % of the sum concentration of all arsenic species

225

detected in HepG2, and the principal product of Rox biotransformation in HepG2 cells was

226

3AHPAA. These results agree with that of Moody and Williams 33 who found that hens

227

metabolized 18% Rox to 3AHPAA as the major metabolite.

11 ACS Paragon Plus Environment

Page 13 of 29

Environmental Science & Technology

228

The metabolites that we have measured in HepG2 cells and human primary hepatocytes

229

were in general agreement with the results of US FDA 34, showing the presence of AsV, DMA,

230

MMA, 3AHPAA, and NAHAA in the livers of chickens fed Rox. In our study, several

231

unidentified arsenic species were detected in HepG2 cells and hepatocytes after Rox exposure.

232

Identifying these arsenic species can provide further information on illustrating the cellular

233

metabolism of Rox in humans and understanding the toxicological significance of Rox

234

metabolism.

235

Although liver is the primary site for the metabolism of inorganic arsenic, it is not clear if

236

liver is also the primary site for the metabolism of Rox. This is a limitation of using cultured

237

hepatocytes as compared to whole-body models.

238

Hepatocytes grown in a monolayer rather than in other configurations (e.g., sandwich

239

cultured human hepatocytes or 3-D cultures) can have reduced biotransformation abilities. Thus,

240

the metabolic conversion of Rox in the current study is potentially underestimated. Our results of

241

the comparison between suspension culture and plate monolayer culture of hepatocytes provided

242

supportive information. In the same batch of experiment where we tested the AsIII methylation in

243

plate culture of the hepatocytes, we also tested the methylation of AsIII in suspension culture of

244

the hepatocytes (SI Section 3). The plated hepatocytes exposed to 1 µM AsIII had AMR of 5 and

245

6.5 pmol/h/106 cells for human 1 and 2, respectively. The AMR of suspended cells exposed to 1

246

µM AsIII was 22.9 pmol/h/106 cells for human 1 and 34.3 pmol/h/106 cells for human 2. These

247

results indicated that the AsIII methylation ability of hepatocytes in suspension culture was higher

248

than that in plate culture.

249

Hepatocytes have reductive intracellular environment with high concentrations of

250

glutathione; this condition would favor the production of trivalent arsenicals. The HPLC-ICPMS

12 ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 29

251

method used here could separate MMAIII from other arsenicals but could not separate DMAIII

252

from DMAV. We did not observe any MMAIII in our cell samples. MMAIII and DMAIII could

253

easily be oxidized to the pentavalent MMAV and DMAV.35 The present study did not take any

254

special measure to preserve the pentavalent methylarsenicals. Thus, the measured concentrations

255

of monomethyl and dimethyl arsenic species represented both the trivalent and pentavalent forms

256

of methylated arsenicals.

257 258

Identification of arsenic species and a new arsenic metabolite. Figure 1c shows the

259

presence of an arsenic species (Un) at a retention time of 25.7 min. The concentration of Un in

260

the hepatocytes, estimated against the calibration of Rox, was at low µg/L levels. We first

261

confirmed that the new arsenic species (Un) was present only in the hepatocyte samples (cell

262

lysate and medium), but not in the culture medium control without incubation with the

263

hepatocytes (SI Figure S3a). On a strong anion exchange column, the Un species eluted after

264

Rox, and the retention time of Un did not match any of the common arsenic species that have

265

standards available (SI Figure S3b). Separation of arsenic species on a reversed phase column

266

(SI Figure S3c), collection of the Un fraction, and reanalysis of the collected fraction using anion

267

exchange chromatography further confirmed the chromatographic behavior of the new arsenic

268

species (Un) (SI Figure S3d).

269

To identify the Un arsenic species, we developed a strategy that took advantages of the

270

complementary information, such as high-resolution MS and MS/MS scans from quadrupole

271

time-of-flight measurements, MRM from triple quadrupole linear ion trap analysis, and the

272

HPLC retention time information, along with simultaneous detection of ICPMS and ESIMS. The

273

details of the methodology are described in SI and the work flow is summarized in SI Figure S1.

13 ACS Paragon Plus Environment

Page 15 of 29

274

Environmental Science & Technology

Briefly, from each HPLC analysis, we detected both precursor ions and fragment ions

275

using a high-resolution quadrupole time-of-flight mass spectrometer, by acquiring both MS and

276

MS/MS scans. Building on the previously reported mass spectral information (SI Table S2)

277

about the common arsenic species, our analyses of the precursor ions, the fragment ions, and

278

their inter-relations (SI Figures S4-S6) enabled us to narrow down the candidate list (SI Tables

279

S2 and S3) and achieve a tentative identification. We supported our identification by performing

280

repeated analyses in both the negative ionization mode (SI Figures S4-S6) and the positive

281

ionization mode (SI Figures S7 and S8).

282

We further confirmed the identification of Un as a thiolated Rox, using the authentic

283

arsenic compound that we synthesized (SI Figures S9 and S10). Using the mass spectral

284

information we obtained from the thiolated Rox (SI Figure S10) and Rox (SI Figure S11), we

285

constructed an MRM method for the detection of Rox and its newly identified hepatic metabolite,

286

thiolated Rox. Reversed phase HPLC separation with simultaneous detection by ICPMS (Figure

287

2, top panel) and ESIMS detection (Figure 2, bottom panel) shows consistent results of Rox and

288

the thiolated Rox in the human primary hepatocyte sample. We have also detected the thiolated

289

Rox as a new metabolite of Rox in HepG2 cells.

290 291

The cytotoxicity of the thiolated Rox. The cytotoxicity of the thiolated Rox was then

292

studied on two cell lines, urinary bladder cancer cell T24 and lung cancer cell A549. In the same

293

batch of experiments, we evaluated the cytotoxicity of AsIII as a control. The 24-h IC50 values for

294

AsIII were 42 ± 1 µM in A549 cells and 5.2 ± 0.2 µM in T24 cells, which were consistent with

295

previous studies 30. The results of the toxicological experiments showed that the 24-h IC50 of

296

thiolated Rox for A549 and T24 were 380 ± 80 µM and 42 ± 10 µM, respectively, whereas the

14 ACS Paragon Plus Environment

Environmental Science & Technology

297

24-h IC50 of Rox for A549 and T24 were 9300 ± 1600 µM and 6800 ± 740 µM, respectively.

298

Therefore, the toxicity of the thiolated Rox was 23-161 fold higher than that of Rox. These

299

results are consistent with previous cytotoxicity studies which suggested that the thiolated

300

arsenic species are generally more toxic than their oxygenated analogues 36-42.

301

Page 16 of 29

We have considered whether the impurities Rox and in the synthesized thiolated Rox

302

could contribute to the observed IC50 values. From the HPLC-ICPMS analysis of the synthesized

303

thiolated Rox, we determined that the impurities were AsV and Rox, accounting for 0.9% and

304

9.3% of total arsenic, respectively. The 24-h IC50 values of AsV was 1400 ± 130 µM for A549

305

cells and 85 ± 6 µM for T24 cells (SI Table S4). When the apparent thiolated Rox was at its IC50

306

value, e.g. about 400 µM for A549, the concentration of impurity as AsV (< 0.9%) was 4 µM

307

which was much less than the IC50 of AsV. Thus the low concentration of AsV as an impurity

308

present in the thiolated Rox had little contribution to the observed cytotoxicity of thiolated Rox.

309 310

IMPLICATIONS AND OUTLOOK.

311

Trace amounts of arsenic species, including Rox, have been previously detected in chicken fed

312

Rox-containing food 12, 43. Rox is widely used as a poultry feed additive in many countries 44-47

313

although it is no longer approved for this use in the United States 14, 48 and the European Union 15.

314

A recent market-basket survey conducted in the United States showed that 46% of raw chicken

315

meat samples purchased in the local food markets in the U.S. contained detectable Rox 3. The

316

exact source(s) of the detected Rox in these chicken meat samples are not known. However,

317

consumers of chicken meat are at risk of exposure to the residual Rox. Prior to this work, there is

318

no study on the human metabolism of Rox. Our study, using both normal cells (human primary

15 ACS Paragon Plus Environment

Page 17 of 29

Environmental Science & Technology

319

hepatocytes) and cancer cells (HepG2), demonstrates that human liver cells can metabolize Rox,

320

forming several arsenic species.

321

We determined the concentrations of arsenic species in the cell lysate after the filtration

322

through a 3kD cut-off membrane filter. This fraction accounted for (72 ± 10)% (mean ± SD) as

323

compared to the total concentration of arsenic in the whole cell digest (SI Section 1). These

324

results suggest that there was still arsenic remaining in the cell debris and/or in the high

325

molecular weight fraction (>3kD) in the filtrate. The remaining arsenic might be present in a

326

protein-bound form. Further studies could be carried out using size-exclusion chromatography to

327

determine both the protein-bound and free arsenicals in the cells.

328

One of the metabolites detected in the human liver cell samples is identified as a thiolated

329

Rox. This new arsenic species is more toxic than Rox. The thiolated Rox has IC50 values of 42

330

µM for T24 cells and 380 µM for A549 cells, representing 1-2 orders of magnitude lower than

331

the IC50 values for Rox. How the thiolated Rox exerts higher cytotoxicity than Rox to these cells

332

remains unclear. Higher cellular absorption could contribute to increased cytotoxicity 38.

333

Determination of cellular arsenic levels in cells exposed to Rox or thiolated Rox would be useful

334

in future studies.

335

There is no report on the quantitative amounts or estimations of human dietary intake of

336

Rox. Our findings of Rox metabolism in human hepatic cells and the discovery of the more toxic

337

arsenic metabolites warrant further research on human exposure, metabolism, and potential

338

health effects of these arsenic species.

339 340

341

16 ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 29

342

ACKNOWLEDGEMENTS

343

This work was supported by the Natural Sciences and Engineering Research Council of Canada,

344

Canadian Institutes of Health Research, the Canada Research Chairs Program, Alberta Innovates,

345

and Alberta Health.

346 347

SUPPORTING INFORMATION. Determination of recovery of arsenic from cell lysate

348

filtrates and recovery from the chromatographic column. Methylation of AsIII in suspension

349

culture of the hepatocytes. Determination of arsenic species using HPLC-ICPMS.

350

Characterization of arsenic species using ESIMS. Re-analyses of arsenic species along with

351

synthesized new arsenic species. The operating conditions for ICPMS and ESIMS. A list of

352

expected arsenic-containing fragment ions. A list of candidate precursor ions. The 24- and 48-

353

hour IC50 of arsenic species for A549 and T24 cells. Schematic of identification of new arsenic

354

species. Anion exchange HPLC-ICPMS chromatograms of HepG2, hepatocytes and cell medium.

355

Reversed-phase HPLC-ESI QqToFMS chromatograms of cell medium in negative and positive

356

ionization modes. Mass spectra of Rox and thiolated Rox on QqToFMS. Reversed-phase HPLC

357

with ICPMS and ESI-MRM chromatograms of cell samples.

358 359 360 361 362 363

REFERENCES (1) National Research Council. Critical aspects of EPA’s IRIS assessment of inorganic arsenic: Interim report. National Academies Press, Washington DC. 2014. (2) Silbergeld, E.K.; Nachman, K. The environmental and public health risks associated with arsenical use in animal feeds. Ann. N.Y. Acad. Sci. 2008, 1140, 346–357.

17 ACS Paragon Plus Environment

Page 19 of 29

364

Environmental Science & Technology

(3) Nachman, K.E,; Baron, P.A.; Raber, G.; Francesconi, K.A.; Navas-Acien, A.; Love, D.C.

365

Roxarsone, inorganic arsenic, and other arsenic species in chicken: a U.S.-based market

366

basket sample. Environ. Health Perspect. 2013, 121, 818–24.

367

(4) Garbarino, J. R.; Bednar, A. J.; Rutherford, D. W.; Beyer, R. S.; Wershaw, R. L.

368

Environmental fate of roxarsone in poultry litter. I. Degradation of roxarsone during

369

composting. Environ. Sci. Technol. 2003, 37, 1509–1514.

370

(5) Stolz, J.F.; Perera, E.; Kilonzo, B.; Kail, B.; Crable, B.; Fisher, E.; et al.

371

Biotransformation of 3-nitro-4-hydroxybenzene arsonic acid (Roxarsone) and release of

372

inorganic arsenic by clostridium species. Environ. Sci. Technol. 2007, 41, 818–823.

373

(6) Brown, B. L.; Slaughter, A. D.; Schreiber, M. E. Controls on roxarsone transport in

374

agricultural watersheds. Appl. Geochem. 2005, 20, 123-133.

375

(7) Yao, L.; Huang, L.; He, Z.; Zhou, C.; Lu, W.; Bai, C. Delivery of roxarsone via chicken

376

diet → chicken → chicken manure → soil → rice plant. Sci. Total Environ. 2016, 566–

377

567, 1152–1158.

378

(8) Oyewumi, O.; Schreiber, M.E. Using column experiments to examine transport of As and

379

other trace elements released from poultry litter: Implications for trace element mobility

380

in agricultural watersheds. Environ. Pollut. 2017, 227, 223–233.

381

(9) Peng, H.; Hu, B.; Liu, Q.; Yang, Z.; Lu, X.; Huang, R.; et al. Liquid chromatography

382

combined with atomic and molecular mass spectrometry for speciation of arsenic in

383

chicken liver. J. Chromatogr. A. 2014, 1370, 40–9.

384

(10) Liu, Q.; Peng, H.; Lu, X.; Le, X. C. Enzyme-assisted extraction and liquid

385

chromatography mass spectrometry for the determination of arsenic species in chicken

386

meat. Anal. Chim. Acta. 2015, 888, 1–9.

18 ACS Paragon Plus Environment

Environmental Science & Technology

Page 20 of 29

387

(11) Liu, Q.; Lu, X.; Peng, H.; Popowich, A.; Tao, J.; Uppal, J. S.; Yan, X.; Boe, D.; Le, X. C.

388

Speciation of arsenic – A review of phenylarsenicals and related arsenic metabolites.

389

TrAC Trends Anal. Chem. 2017. doi: 10.1016/j.trac.2017.10.006

390

(12) Liu, Q.; Peng, H.; Lu, X.; Zuidhof, M.J.; Li, X-F.; Le, X.C. Arsenic species in chicken

391

breast: Temporal variations of metabolites, elimination kinetics, and residual

392

concentrations. Environ. Health Perspect. 2016, 124,1174–1181.

393

(13) Sapkota, A. R.; Lefferts, L. Y.; McKenzie, S.; Walker, P. What so we feed to food-

394

production animals? A review of animal feed ingredients and their potential impacts on

395

human health. Environ. Health Perspect. 2007, 115, 663–670.

396

(14) Nachman, K.E.; Ginsberg, G.L.; Miller, M.D.; Murray, C.J.; Nigra, A.E.; Pendergrast,

397

C.B. Mitigating dietary arsenic exposure: Current status in the United States and

398

recommendations for an improved path forward. Sci. Total Environ. 2017, 581–582, 221–

399

236.

400

(15) European Commission. Council directive 1999/29/EC of 22 April 1999 on the

401

undesirable substances and products in animal nutrition. 1999. Available:

402

https://publications.europa.eu/en/publication-detail/-/publication/ff906a95-6766-4342-

403

9bbb-7ae6594aa8c3 (accessed on March 17, 2017)

404

(16) Yao, L.; Huang, L.; He, Z.; Zhou, C.; Li, G. Occurrence of arsenic impurities in

405

organoarsenics and animal feeds. J. Agric. Food Chem. 2013, 61, 320–324.

406 407 408 409

(17) US EPA. Toxicological review of ingested inorganic arsenic. CAS No. 7440-38-2. US EPA, Washington, DC. 2010. (18) Mercer, D. F., Schiller, D. E., Elliott, J. F., Douglas, D. N., Hao, C., Rinfret, A., Addison, W. R., Fischer, K. P., Churchill, T. A., Lakey, J. R., Tyrrell, D. L., and Kneteman, N. M.

19 ACS Paragon Plus Environment

Page 21 of 29

Environmental Science & Technology

410

Hepatitis C virus replication in mice with chimeric human livers. Nat. Med. 2001, 7, 927–

411

33.

412

(19) Roggenbeck, B.A.; Carew, M.W.; Charrois, G.J.; Douglas, D.N.; Kneteman, N.M.; Lu,

413

X.; et al. Characterization of arsenic hepatobiliary transport using sandwich-cultured

414

human hepatocytes. Toxicol. Sci. 2015, 145, 307–20.

415 416 417

(20) Naranmandura, H.; Suzuki, N.; Suzuki, K.T. Trivalent arsenicals are bound to proteins during reductive methylation. Chem. Res. Toxicol. 2006, 19, 1010–1018. (21) Cullen, W. R.; Liu, Q.; Lu, X.; McKnight-Whitford, A.; Peng, H.; Popowich, A.; Yan, X.;

418

Zhang, Q.; Fricke, M.; Sun, H.; et al. Methylated and thiolated arsenic species for

419

environmental and health research — A review on synthesis and characterization. J.

420

Environ. Sci. 2016, 49, 7–27.

421

(22) Zhu, X.; Chen, Y.; Subramanian, R. Comparison of information-dependent acquisition,

422

SWATH, and MS All techniques in metabolite identification study employing ultrahigh-

423

performance liquid chromatography–quadrupole time-of-flight mass spectrometry. Anal.

424

Chem. 2014, 86, 1202–1209.

425

(23) Wrona, M.; Mauriala, T.; Bateman, K.P.; Mortishire-Smith, R.J.; O’Connor, D. “All-in-

426

One” analysis for metabolite identification using liquid chromatography/hybrid

427

quadrupole time-of-flight mass spectrometry with collision energy switching. Rapid

428

Commun. Mass Spectrom. 2006, 19, 2597–2602.

429

(24) Amayo, K. O.; Petursdottir, A.; Newcombe, C.; Gunnlaugsdottir, H.; Raab, A.; Krupp, E.

430

M.; Feldmann, J. Identification and quantification of arsenolipids using reversed-phase

431

HPLC coupled simultaneously to high-resolution ICPMS and high-resolution

432

electrospray MS without species-specific standards. Anal. Chem. 2011, 83, 3589–3595.

20 ACS Paragon Plus Environment

Environmental Science & Technology

433

Page 22 of 29

(25) Amayo, K. O.; Raab, A.; Krupp, E. M.; Gunnlaugsdottir, H.; Feldmann, J. Novel

434

Identification of arsenolipids using chemical derivatizations in conjunction with RP-

435

HPLC-ICPMS/ESMS. Anal. Chem. 2013, 85, 9321–9327.

436

(26) Arroyo-Abad, U.; Lischka, S.; Piechotta, C.; Mattusch, J.; Reemtsma, T. Determination

437

and identification of hydrophilic and hydrophobic arsenic species in methanol extract of

438

fresh cod liver by RP-HPLC with simultaneous ICP-MS and ESI-Q-TOF-MS detection.

439

Food Chem. 2013, 141, 3093–3102.

440 441 442

(27) Wallschläger, D.; Stadey, C.J. Determination of (Oxy)thioarsenates in sulfidic waters. Anal. Chem. 2007, 79, 3873–80. (28) Hansen, H.R.; Raab, A.; Jaspars, M.; Milne, B.F.; Feldmann, J. Sulfur-containing

443

arsenical mistaken for dimethylarsinous acid [DMA(III)] and identified as a natural

444

metabolite in urine: major implications for studies on arsenic metabolism and toxicity.

445

Chem. Res. Toxicol. 2004, 17, 1086–91.

446

(29) Yang, Z.; Peng, H.; Lu, X.; Liu, Q.; Huang, R.; Hu, B.; Kachanoski, G.; Zuidhof, M. J.;

447

Le, X. C. Arsenic metabolites, including N-acetyl-4-hydroxy-m-arsanilic acid, in chicken

448

litter from a Roxarsone-feeding study involving 1600 chickens. Environ. Sci. Technol.

449

2016, 50, 6737–6743.

450

(30) Moe, B.; Peng, H.; Lu, X.; Chen, B.; Chen, L. W. L.; Gabos, S.; Li, X.-F.; Le, X. C.

451

Comparative cytotoxicity of fourteen trivalent and pentavalent arsenic species determined

452

using real-time cell sensing. J. Environ. Sci. 2016, 49, 113–124.

453

(31) Styblo, M.; Del Razo, L. M.; LeCluyse, E. L.; Hamilton, G. A.; Wang, C.; Cullen, W. R.;

454

Thomas, D. J. Metabolism of arsenic in primary cultures of human and rat hepatocytes.

455

Chem. Res. Toxicol. 1999, 12, 560–565.

21 ACS Paragon Plus Environment

Page 23 of 29

456

Environmental Science & Technology

(32) Liu, Q.; Leslie, E. M.; Le, X. C. Accumulation and transport of Roxarsone,

457

arsenobetaine, and inorganic arsenic using the human immortalized Caco-2 cell line. J.

458

Agric. Food Chem. 2016, 64, 8902–8908.

459 460 461

(33) Moody, J. P.; and Williams, R. T. The metabolism of 4-hydroxy-3-nitrophenylarsonic acid in hens. Food Cosmet. Toxicol. 1964, 2, 707–715. (34) US FDA. Provide data on various arsenic species present in broilers treated with

462

roxarsone: Comparison with untreated birds. 2011. Available:

463

https://www.fda.gov/downloads/AnimalVeterinary/SafetyHealth/ProductSafetyInformati

464

on/UCM257545.pdf (accessed on March 17, 2017).

465

(35) Gong, Z.; Lu, X.; Cullen, W.R.; Le, X.C. Unstable trivalent arsenic metabolites,

466

monomethylarsonous acid and dimethylarsinous acid. J. Anal. At. Spectrom. 2001, 16,

467

1409-1413.

468

(36) Wang, Q. Q.; Thomas, D. J.; Naranmandura, H. Importance of being thiomethylated:

469

Formation, fate, and effects of methylated thioarsenicals. Chem. Res. Toxicol. 2015, 28,

470

281–289.

471 472 473

(37) Sun, Y.; Liu, G.; Cai, Y. Thiolated arsenicals in arsenic metabolism: Occurrence, formation, and biological implications. J. Environ. Sci. 2016, 49, 59–73. (38) Naranmandura, H.; Ogra, Y.; Iwata, K.; Lee, J.; Suzuki, K. T.; Weinfeld, M.; Le, X. C.

474

Evidence for toxicity differences between inorganic arsenite and thioarsenicals in human

475

bladder cancer cells. Toxicol. Appl. Pharmacol. 2009, 238, 133–140.

476

(39) Ochi, T.; Kita, K.; Suzuki, T.; Rumpler, A.; Goessler, W.; Francesconi, K. A. Cytotoxic,

477

genotoxic and cell-cycle disruptive effects of thio-dimethylarsinate in cultured human

478

cells and the role of glutathione. Toxicol. Appl. Pharmacol. 2008, 228, 59–67.

22 ACS Paragon Plus Environment

Environmental Science & Technology

479

Page 24 of 29

(40) Leffers, L.; Unterberg, M.; Bartel, M.; Hoppe, C.; Pieper, I.; Stertmann, J.; Ebert, F.;

480

Humpf, H. U.; Schwerdtle, T. In vitro toxicological characterisation of the S-containing

481

arsenic metabolites thio-dimethylarsinic acid and dimethylarsinic glutathione. Toxicology.

482

2013, 305, 109–119.

483

(41) Bartel, M.; Ebert, F.; Leffers, L.; Karst, U.; Schwerdtle, T. Toxicological characterization

484

of the inorganic and organic arsenic metabolite thio-DMA in cultured human lung cells. J.

485

Toxicol. 2011, 2011, 373141.

486

(42) Naranmandura, H.; Carew, M. W.; Xu, S.; Lee, J.; Leslie, E. M.; Weinfeld, M.; Le, X. C.

487

Comparative toxicity of arsenic metabolites in human bladder cancer EJ-1 cells. Chem.

488

Res. Toxicol. 2011, 24, 1586–1596.

489

(43) Peng, H.; Hu, B.; Liu, Q.; Li, J.; Li, X.-F.; Zhang, H.; Le, X. C. Methylated

490

phenylarsenical metabolites discovered in chicken liver. Angew. Chemie Int. Ed. 2017, 56,

491

6773–6777.

492

(44) Hu, Y.; Zhang, W.; Cheng, H.; Tao, S. Public health risk of arsenic species in chicken

493

tissues from live poultry markets of Guangdong Province, China. Environ. Sci. Technol.

494

2017, 51, 3508–3517.

495

(45) Batista, B. L.; Grotto, D.; Carneiro, M. F. H.; Barbosa, F. Evaluation of the concentration

496

of nonessential and essential elements in chicken, pork, and beef samples produced in

497

Brazil. J. Toxicol. Environ. Health. A. 2012, 75, 1269–1279.

498

(46) Mafla, S.; Moraga, R.; León, C. G.; Guzmán-Fierro, V. G.; Yañez, J.; Smith, C. T.;

499

Mondaca, M. A.; Campos, V. L. Biodegradation of roxarsone by a bacterial community

500

of underground water and its toxic impact. World J. Microbiol. Biotechnol. 2015, 31,

501

1267–1277.

23 ACS Paragon Plus Environment

Page 25 of 29

502

Environmental Science & Technology

(47) Eom, H. Y.; Yang, D.-H.; Suh, J. H.; Kim, U.; Kim, J.; Cho, H.-D.; Han, S. B.

503

Determination of residual arsenic compounds in chicken muscle by ultra-performance

504

liquid chromatography coupled with ultraviolet detection after pre-column derivatization

505

with toluene-3,4-dithiol. J. Chromatogr. B. 2015, 1006, 151–157.

506

(48) Fisher, D. J.; Yonkos, L. T.; Staver, K. W. Environmental concerns of Roxarsone in

507

broiler poultry feed and litter in Maryland, USA. Environ. Sci. Technol. 2015, 49, 1999–

508

2012.

24 ACS Paragon Plus Environment

Environmental Science & Technology

Page 26 of 29

509

Table 1. The intracellular concentrations of arsenic species (µg/g protein) present in HepG2 cells

510

(n=3) and plated human primary hepatocytes from two individuals, designated Human 1 and

511

Human 2. Cells were incubated with 20 or 100 µM Rox for 24 h prior to the HPLC-ICPMS

512

analyses.

20 µM Rox (µg/g protein) 100 µM Rox (µg/g protein) HepG2 cells HepG2 cells Human 1 Human 2 Human 1 Human 2 (mean ± SD) (mean ± SD) DMA 0.008 ± 0.002 N.D.a N.D. N.D. N.D. 0.022 ± 0.001 N.D. N.D. N.D. N.D. MMA 0.005 ± 0.001 0.8 0.8 N.D. N.D. 0.010 ± 0.002 2 2 N.D. N.D. V As 0.08 ± 0.05 0.3 0.3 2 2 0.32 ± 0.06 0.7 0.6 2 2 a 3AHPAA 1.1 ± 0.2 N.D. N.D. N.D. N.D. 2.8 ± 0.5 2 1 N.D. N.D. NAHAA 0.03 ± 0.01 N.D. N.D. N.D. N.D. 0.05 ± 0.01 0.4 0.2 N.D. N.D. a 513 N.D.: non-detectable (below detection limit). SD: standard deviation.

25 ACS Paragon Plus Environment

Page 27 of 29

Environmental Science & Technology

Figure 1. Chromatograms obtained from anion exchange HPLC-ICPMS analyses of (a) arsenic standards, (b) HepG2 cells exposed to 20 µM Rox, and (c) primary hepatocytes exposed to 20 µM Rox. The primary hepatocytes were from human donor 1. A new arsenic species to be identified is labelled as Un.

ACS Paragon Plus Environment

Environmental Science & Technology

Figure 2. Chromatograms from reversed phase separation hyphenated with simultaneous (a) ICPMS detection and (b) ESI MS/MS detection of a hepatocyte sample. In ESI MS/MS, 5 transition ion pairs (278/95, 278/107, 278/111, 278/123, and 278/139) of thiolated Rox and 2 transition ion pairs (262/107 and 262/123) of Rox were monitored. The hepatocytes were prepared from a liver slice donated by human 1. Hepatocytes were incubated with 100 µM Rox for 24 h. Incubation medium from this culture was analyzed for arsenic species.

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

Environmental Science & Technology

ACS Paragon Plus Environment