Induction of Apoptosis in Human Papillary-Thyroid-Carcinoma BCPAP

May 22, 2018 - School of Food Science and Technology, Jiangnan University, 1800 Lihu Avenue, Wuxi , Jiangsu 214122 , China. ‡ Key Laboratory of Nucl...
4 downloads 0 Views 2MB Size
Subscriber access provided by UNIVERSITY OF TOLEDO LIBRARIES

Bioactive Constituents, Metabolites, and Functions

Diallyl trisulfide induces apoptosis of human papillary thyroid carcinoma BCPAP cells through activating MAPK signaling pathway Jie Pan, Li Zhang, Shichen Xu, Xian Cheng, Huixin Yu, Jiandong Bao, and Rong-Rong Lu J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b02243 • Publication Date (Web): 22 May 2018 Downloaded from http://pubs.acs.org on May 22, 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 31

Journal of Agricultural and Food Chemistry

1

Diallyl trisulfide induces apoptosis of human papillary thyroid

2

carcinoma BCPAP cells through activating MAPK signaling

3

pathway

4

Jie Pana,b,1, Li Zhangb,1, Shichen Xub, Xian Chengb, Huixin Yub, Jiandong Baob,

5

Rongrong Lua, *

6

a

7

Wuxi, Jiangsu 214122, China

8

b

9

Ministry of Health, 20 Qian Rong Road, Wuxi, Jiangsu 214063, China

School of Food Science and Technology, Jiangnan University, 1800 Lihu Avenue,

Jiangsu Institute of Nuclear Medicine, Key Laboratory of Nuclear Medicine,

10

1

11

*

12

University, Wuxi 214122, China

13

E-mail address: [email protected] (Rongrong Lu)

14

Phone: 0510-85329061(O)

15

Fax: + 86 510 85912155

These authors equally contributed in this work.

Correspondence author: School of Food Science and Technology, Jiangnan

16

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 31

17

Abstract

18

This study aimed to elucidate the potential effects of diallyl trisulfide (DATS) on

19

human papillary thyroid carcinoma BCPAP cells and its underling mechanisms. DATS

20

is one of the organosulfur compounds which derived from garlic. In this study, we

21

demonstrated that compared to the solvent control, DATS treatment at concentrations

22

of 5, 10 and 20 µΜ decreased cell survival rate of BCPAP cells to 84.51 ± 2.67%,

23

57.16 ± 1.18%, 41.22 ± 1.19% respectively. DATS also caused cell cycle arrest at

24

G0/G1 phase and the proportion of cells arrested in G0/G1 phase rose from 68.8 ±

25

8.38% to 80.4 ± 8.38%, which eventually resulted in cell apoptosis through a

26

mitochondrial apoptotic pathway in BCPAP cells. Further evidences showed that

27

DATS activated ERK, JNK and p38, the members of MAPK family. Moreover, ERK

28

inhibitor and JNK inhibitor, partially reversed apoptosis in BCPAP cells induced by

29

DATS treatment. Taken together, our results demonstrated that DATS exerted the

30

apoptosis-inducing effect on papillary thyroid cancer cells via activating MAPK

31

signaling pathway, which shed a light on a prospective therapeutic target for thyroid

32

cancer treatment.

33

Keywords: Diallyl trisulfide; Papillary thyroid carcinoma; Apoptosis; Cell cycle arrest;

34

MAPK signaling pathway

35

2

ACS Paragon Plus Environment

Page 3 of 31

Journal of Agricultural and Food Chemistry

36

Introduction

37

Diallyl trisulfide (DATS) is one of the organosulfur compounds (OSCs) with a strong

38

pungent smell (Fig. 1). It mainly derives from alliaceae family such as onion and

39

garlic. Garlic, a well-known functional food and its derivates have attracted more and

40

more attentions recently due to their significant biological and pharmacological

41

activities. It has been reported that garlic was used as a potent phytochemical

42

medicine in traditional China, Egypt, Greece, Rome and India for a long history1.

43

Accumulating reports have proved that DATS is the most efficient bioactive

44

component among the three diallyl polysulfides: diallyl sulfide (DAS), diallyl

45

disulfide (DADS) and diallyl trisulfide (DATS)2. The three sulfur atoms in the

46

structure of DATS may substantially contribute to its stronger bioactivity compared

47

with one sulfur atom in DAS and two sulfur atoms in DADS2, 3.

48

The bioactivities of DATS contain two aspects. Firstly, DATS has a potent

49

antioxidant activity in protecting normal cells from oxidative damage4. Secondly,

50

previous studies showed that DATS could fight against many human diseases,

51

including cardiac disease5, diabetes6, inflammation7 and cancers. It has been reported

52

that DATS exerted anti-cancer effects against pancreatic cancer8, melanoma9 and

53

breast cancer10. Moreover, the epidemiologic study in China has observed a close

54

association between intake of raw garlic and lung cancer prevention11. In terms of

55

thyroid cancers, previous study showed that DAS, another kind of OSCs, could

56

induce apoptosis of anaplastic thyroid cancer cells12. In the present study, whether

57

DATS has an anti-cancer effect on papillary thyroid carcinoma and its detailed

58

mechanisms were further investigated.

59

Thyroid cancer is one of the most common endocrine malignancy with a

60

worldwide rising incidence rate13. According to its pathological histology, thyroid 3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 31

61

carcinoma can be classified into four subtypes, namely papillary thyroid carcinoma

62

(PTC), follicular thyroid carcinoma (FTC), medullary thyroid carcinoma (MTC), and

63

anaplastic thyroid carcinoma (ATC)14. PTC is the most common subtype accounting

64

for about 80% of the thyroid cancer cases14. Currently, the major therapeutic

65

modalities for thyroid cancer include surgical removal, radioiodine ablation and

66

thyroid-stimulating hormone (TSH) suppression15. However, in some advanced

67

papillary thyroid cancers harbouring BRAF and/or TERT mutations, the tumors will

68

lose the radioiodine absorbing ability and subsequently become refractory to

69

radioiodine treatment16. Thus, alternative systemic options are needed for these

70

patients. Sorafenib, an oral multikinase inhibitor, was approved by the US Food and

71

Drug Administration (FDA) for the treatment of 131I-refractory and metastatic thyroid

72

cancer. Though sorafenib was currently considered as a first-line therapy for

73

131

74

after the treatment. Therefore, it is urgent to develop promising drugs or novel

75

strategies to improve the therapeutic efficiency.

I-refractory metastatic PTC19, a proportion of patients were still in progress even

76

Targeting apoptosis is a promising strategy for cancer therapy through regulating

77

many intracellular signals including Nrf2/Akt17, NF-κB18 and MAPK signaling

78

pathways17. Mitogen-active protein kinases (MAPKs) are serine/threonine kinases. In

79

mammalian cells, this superfamily consists of three main members: the extracellular

80

regulated kinase (ERK), c-Jun N-terminal kinase (JNK) and p3819. MAPK signaling

81

pathway participates in various cell behaviors, including cell proliferation,

82

differentiation and apoptosis20. The activation of JNK/p38 was a kind of apoptotic

83

signal while the activation of ERK was generally a survival signal17. Moreover, many

84

studies reported that ERK signaling pathway was activated in nearly 80% of PTC

85

specimen21. Constitutive ERK activation is believed to be essential for the survival 4

ACS Paragon Plus Environment

Page 5 of 31

Journal of Agricultural and Food Chemistry

86

and development of thyroid cancer cells21. However, excessive activation of MAPK

87

signaling induced by chemicals could also induce apoptosis in thyroid cancer cells22.

88

Hence, MAPK pathway has a dual potential in the regulation of cell survival and

89

apoptosis. Previous study demonstrated that DATS induced apoptosis in prostate

90

cancer cells via activating MAPK/ERK/JNK signaling pathway23. Therefore, it

91

prompted us to study the anti-cancer efficacy and the potential effects of DATS on

92

MAPK signaling pathway in thyroid cancer cells.

93

In the present study, we demonstrated that DATS inhibited cell proliferation,

94

caused G0/G1 phase cell cycle arrest and subsequently elicited cell apoptosis in

95

papillary thyroid cancer BCPAP cells. Our results provided evidences that excessive

96

activation of MAPK signaling pathway as a central mechanism of DATS-induced

97

growth inhibition and apoptosis in thyroid cancer cells. These results proposed a

98

potential therapeutic target for PTC treatment.

99

Materials and methods

100

Chemicals, reagents and antibodies

101

Diallyl trisulfide (DATS, HPLC≥98%) and Sulforhodamine B (SRB) were purchased

102

from Sigma Aldrich (Saint Louis, Missouri, USA). Hoechst 33342 (C1022), PI

103

(ST511), JC-1 (C2005), U0126 (S1901), SP600125 (S1876), SB203580 (S1863),

104

reactive oxygen species (ROS) detection kit (S0033) and Dimethyl Sulfoxide (DMSO)

105

were purchased from Beyotime Institute of Biotechnology (Nantong, China). The

106

antibodies used in the present study were as follows: Anti-p-Rb (ab173289) was

107

purchased from Abcam (Cambridge, UK). Anti-p-JNK (cst-4668), anti-Bax (cst-2772),

108

anti-caspase-8 (cst-9746), anti-caspase-3 (cst-9665) were purchased from Cell

109

Signaling Technology (Boston, Massachusetts, USA). Anti-PARP-1 (sc-7150),

110

anti-β-actin (sc-47778), goat anti-mouse (sc-2005) or rabbit (sc-2004) IgG-HRP were 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 31

111

purchased from Santa Cruz Biotechnology (Santa Cruz, California, USA).

112

Anti-tubulin (AT819), anti-Cyclin D1 (AC853), anti-p27 (AP027), anti-Bcl-2

113

(AB112), anti-Bcl-XL (AB126-1), anti-ERK (AM076), anti-p-ERK (AM071),

114

anti-JNK (AJ518), anti-p-JNK (AJ516), anti-p38 (AM065), anti-p-p38 (AM063) were

115

purchased from Beyotime Institute of Biotechnology (Nantong, China). Other

116

chemicals were analytical grade and purchased from common source.

117

Cell culture and drug treatment

118

The human papillary thyroid carcinoma line BCPAP cells were obtained from the

119

German Collection of Micro-organisms and Cell Cultures (Braunschweig, Germany).

120

BCPAP cells were cultured in RPMI 1640 medium supplemented with 10% new born

121

bovine serum (NBS), antibiotics (penicillin 100 U/ml and streptomycin 100 U/ml).

122

Nthy-ori-3.1 cells, obtained from the German Collection of Micro-organisms and Cell

123

Cultures (Braunschweig, Germany), were cultured in RPMI 1640 medium

124

supplemented with 10% fetal bovine serum (FBS), antibiotics (penicillin 100 U/ml

125

and streptomycin 100 U/ml). BCPAP cells and Nthy-ori-3.1 cells were both

126

maintained at 37 °C in a standard humidified incubator containing 5% CO2. DATS

127

was dissolved in DMSO at 20 mM as a stock solution and stored at -20 °C until

128

diluting to indicated concentration before use. The solvent control contained an

129

equivalent amount of DMSO corresponding to the highest used concentration of

130

DATS and the final concentration of DMSO was less than or equal to 0.1%.

131

Sulforhodamine B (SRB) assay

132

The sulforhodamine B (SRB) assay was used to detect the cell number. Briefly,

133

BCPAP or Nthy-ori-3.1 cells were seeded into 96-well plates at 7500 cells per well

134

and cultured overnight. Cells were treated with various concentrations of DATS for

135

indicated time and then fixed with 10% TCA (w/v) for 1 h at 4 °C, and stained with 6

ACS Paragon Plus Environment

Page 7 of 31

Journal of Agricultural and Food Chemistry

136

4% SRB for another 30 min at 37 °C. Then, tris base (10 mM) of 100 µl was added

137

into each well and the plates were vibrated for 5 min to dissolve SRB completely. The

138

optical density of the cell suspension was measured at 565 nm by a microplate reader

139

(Epoch, Biotek). Cell survival was expressed as a percentage of SRB reduction,

140

assuming that the absorbance of solvent control group was 100%.

141

Colony formation assay

142

BCPAP cells were seeded into 24-well plates at a density of 500 cells per well and

143

cultured overnight. BCPAP cells were pretreated with different dosages of DATS

144

(5-20 µΜ) for 24 h, then the medium was replaced by complete growth medium. After

145

cultured for another 10 days, the cells were fixed with methanol and stained with

146

crystal violet for 15 min. The visible colonies were photographed by a Bio-Rad

147

Imager (SYSTEM GelDoc XR+).

148

Cell cycle analysis

149

Cell cycle distribution was determined by PI staining. Briefly, BCPAP cells were

150

seeded into 6-well plates and cultured overnight. The cells were then treated with

151

indicated concentrations of DATS for 24 h. After that, cells were harvested and fixed

152

with 4% paraformaldehyde (PFA) at 4 °C for 1 h. Then the cells were washed by PBS

153

and incubated with DNA staining reagent which contained 50 µg/ml RNase, 0.1%

154

Triton X-100, 0.1 mM EDTA (pH 7.4), and 50 µg/ml PI in the dark for 30 min. The

155

stained cells were subjected to a flow cytometer (FACSCalibur, Becton Dickinson,

156

USA) for cell cycle distribution analysis.

157

Hoechst 33342/PI staining

158

Hoechst 33342/PI staining was conducted to detect cell apoptosis. Briefly, BCPAP

159

cells were seeded into 12-well plates and cultured overnight. Then BCPAP cells were

160

incubated with different concentrations of DATS (5-20 µΜ) for 24 h. Subsequently,

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 31

161

the cells were stained with a mixture solution of Hoechst 33342 (10 µg/ml) and PI (10

162

µg/ml) at 37 °C for 15 min. Cells were observed under a fluorescence microscope

163

(Olympus, X51, Japan). Three independent fields per well were picked and cell

164

Images were photographed.

165

Mitochondrial membrane potential measurement

166

JC-1, a kind of lipophilic cationic fluorescent dye, was used to measure the

167

depolarization of mitochondria. BCPAP cells were treated with indicated

168

concentrations of DATS for 24 h and subsequently stained with JC-1 (10 µg/ml) at

169

37 °C for 30 min in the dark. After staining, cells were washed by PBS and

170

immediately analyzed by flow cytometry (FACSCalibur, Becton Dickinson, USA) at

171

FL1 and FL2 channels.

172

Reactive oxygen species (ROS) detection

173

ROS detection was performed according to the manufacturer’s protocol. After

174

different concentrations of DATS treatment, the cells were washed with ice-cold PBS

175

once, and then incubated with 20 µM of DCFH-DA at 37 °C for 20 min in the dark.

176

The cells were then harvested and resuspended in ice-cold PBS. Cells were

177

immediately analyzed by flow cytometry (FACSCalibur, Becton Dickinson, USA).

178

Western blot

179

Total cell lysates were prepared by RIPA lysis buffer containing protease inhibitor

180

cocktail (1% v/v). The protein concentration was quantified by Bradford protein assay.

181

Collected proteins were separated by 10% or 15% SDS polyacrylamide gel

182

electrophoresis. The proteins were transferred to a nitrocellulose (NC) membrane and

183

blocked with 5% skim milk dissolved in Tris-buffered saline containing 0.1% Tween

184

20. Targeted proteins were detected by corresponding primary antibodies and

185

subsequently incubated with horseradish peroxidase-conjugated secondary antibodies.

8

ACS Paragon Plus Environment

Page 9 of 31

Journal of Agricultural and Food Chemistry

186

The protein bands were visualized using an ECL Western blot kit (ABXBio). The

187

β-actin or α-tubulin was used as an internal control for equal loading of total proteins.

188

Statistical analysis

189

Data were presented as mean ± S.D. and analyzed by GraphPad Prism 6 software.

190

One-way ANOVA analysis followed by Dunnett's test for multiple comparisons were

191

used to analyze the statistical significance and P < 0.05 was considered to be

192

statistically significant.

193

Results

194

DATS inhibits clonogenic survival of human papillary thyroid carcinoma BCPAP

195

cells

196

Firstly, in order to investigate the potential cytotoxic effect of DATS on BCPAP cells,

197

cellular morphological changes were observed after various concentrations of DATS

198

treatment (5-20 µΜ). As illustrated in Fig. 2A, cells exposed to DATS became smaller

199

in size and less adherent to culture plate compared to the solvent control group. Then,

200

SRB assay was used to evaluate the effect of DATS on the growth of BCPAP cells.

201

The results showed that DATS at a low concentration of 0.625 µM failed to inhibit the

202

cell growth. However, when the concentration of DATS exceeded 1.25 µΜ, the cell

203

number of each group decreased significantly (P < 0.05). The cell number of BCPAP

204

cells reduced over 50% by the treatment of DATS at 20 µΜ (Fig. 2B). Next, in vitro

205

colony formation assay was performed to further determine the growth-inhibition

206

efficiency of DATS on BCPAP cells. As shown in Fig. 2C and 2D, DATS pretreatment

207

dose-dependently decreased the clone numbers. SRB assay also demonstrated that

208

DATS inhibited the cell growth of BCPAP cells in a time-dependent manner (Fig. 2E).

209

Furthermore, the cytotoxic effect of DATS on the normal thyroid follicular epithelial

210

cells was examined. As shown in Fig. 2F, SRB assay indicated that papillary thyroid 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 31

211

carcinoma BCPAP cells were more sensitive to DATS treatment than normal thyroid

212

follicular epithelial cell line Nthy-ori-3.1 cells did. This result suggested DATS

213

selectively killed thyroid cancer cells rather than normal thyroid follicular epithelial

214

cells. Taken together, these data indicated that DATS inhibited clonogenic survival of

215

human papillary thyroid carcinoma BCPAP cells in dose- and time dependent

216

manners.

217

DATS arrests cell cycle at G0/G1 phase in BCPAP cells

218

Cell cycle is abbrently fast in the malignant cells. It has been widely accepted that

219

arresting cancer cells in a certain phase can inhibit tumor growth24. Therefore, we

220

hypothesized that the growth-inhibitory effect of DATS was due to its cell cycle arrest

221

ability. Next, the effects of DATS on cell cycle distribution were evaluated by

222

propidium iodide (PI) staining. Compared to solvent control, the percentage of cell

223

population distributing at G0/G1 phase was obviously increased after DATS treatment,

224

while cells in S and G2/M phases showed a concomitant reduction (Fig. 3A). As

225

shown in Fig. 3B, 68.8 ± 8.38% of the DATS-treated cells were arrested at G0/G1

226

after 5 µΜ of DATS treatment, and this proportion was up to 77.3 ± 11.84% and

227

80.4 ± 8.38% at the concentration of 10 µΜ and 20 µΜ, respectively. Moreover, the

228

percentage of cell population at sub-G1, which served as a feature of apoptotic cells,

229

also increased significantly in a dose-dependent manner. Moreover, the alternation of

230

the cell cycle regulating proteins involved in G0/G1 phase such as cyclin D1, p-Rb

231

and p2725 after DATS treatment were investigated. The results showed that DATS

232

treatment at 5 to 20 µM dose-dependently decreased the expression of cyclin D1 and

233

p-Rb. Accordingly, DATS treatment increased the protein level of p27 which is a CKI

234

(cyclin kinase inhibitor) halting the cell cycle progression (Fig. 3C and 3D). Taken 10

ACS Paragon Plus Environment

Page 11 of 31

Journal of Agricultural and Food Chemistry

235

together, these data indicated that DATS inhibited cell growth and caused cell cycle

236

arrest at G0/G1 phase in BCPAP cells.

237

DATS induces BCPAP cells apoptosis through a mitochondrial apoptotic

238

pathway

239

The increasedly accumulating cells in the sub-G1 peak prompted us to investigate

240

whether the growth inhibition effect of DATS on BCPAP cells was attributed to its

241

apoptosis-inducing effect. Hence, Hoechst 33342/PI staining was performed to detecte

242

the apoptotic phenotype of DATS-treated cells. As shown in Fig. 4A, DATS-treated

243

BCPAP cells exihibited evident apoptotic features with increased cell membrane

244

permeability (PI positive cells) and nuclear chromatin shrinkage (bright hoechst

245

33342 staining cells). The declined mitochondrial membrane potential was believed to

246

be an early event of apoptosis26. Thus, cell mitochondrial membrane potential was

247

detected. The proportion of JC-1 momomer cells elevated remarkly with the reduction

248

of JC-1 aggregates cells, which suggested that DATS certainly induced the

249

depolarization of mitochondrial membrane potential in BCPAP cells (Fig. 4B and Fig.

250

4C). In addition, the expressions of apoptosis-related proteins were detected by

251

western blot. As illustrated in Fig. 5A and B, the expression of anti-apoptotic protein

252

Bcl-2 and Bcl-XL was downregulated, while the expression of pro-apoptotic protein

253

Bax was upregulated. It has been reported that the increased Bax bound to the outer

254

mitochondrial membrane and Bcl-2 released from the mitochondria to initiate cell

255

apoptosis27. Moreover, the cleaved caspase-3, -8 and the cleavage of downstream

256

PARP were all increased (Fig. 5C). Besides, the decreased cell numbers induced by

257

DATS treatment could be reversed by z-VAD-fmk, a pan-caspase inhibitor (Fig. 5D).

258

Thus, we concluded that DATS treatment elicited caspase-mediated apoptosis through

259

a mitochondrial apoptotic pathway in BCPAP cells.

11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 31

260

DATS-induced apoptosis is independent of ROS generation

261

One of the well-known triggers of apoptosis is cellular ROS. Excess ROS, which is

262

mainly produced in cell mitochondria, is reported as a typical characreristic of

263

oxidative damage to mitochondrial membrane and permeability28. Indeed, the opening

264

of mitochondrial permeability transition pore (MPTP) has been demonstrated to

265

decrease the mitochondrial membrane potential29. Therefore, we next investigated

266

whether the deareased cell mitochondrial membrane potential induced by DATS was

267

triggered by ROS attack. Unexpectedly, DATS treatment did not change the

268

intracellular ROS level, even at a higher concentration of 20 µΜ (Fig. 6). These data

269

indicated that DATS-induced apoptosis was not dependent on the increased

270

production of intracellular ROS.

271

DATS induces apoptosis through activating MAPK signaling pathway in BCPAP

272

cells

273

Most of thyroid cancers harbour MAPK pathway activation, especially ERK1/2

274

hyper-phosphorylation because of BRAF or RAS mutations,

275

activation is considered to be a cell survival signaling31. Moreover, supression of ERK

276

is able to attenuate the prolifitation and migration properties of thyroid cancer cells31.

277

In contrast, it has been reported that excessive activation of ERK could also induce

278

thyroid cancer cell apoptosis22. Hence, targeting this pathway maybe an effective

279

strategy in thyroid cancer treatment. In order to investigate the potential role of

280

MAPK signaling in the anti-cancer effect of DATS, we next determined both the total

281

and phosphorylated levels of MAPKs in DATS-treated BCPAP cells. Interestingly, the

282

ERK, which was constitutively actived in BCPAP cells, was further activated by

283

DATS treatment. As illistrated in Fig. 7A, after BCPAP cells were exposed to DATS

284

treatment, the phosphoryation of ERK increased in a dose-dependent manner, while

30

. In general, ERK

12

ACS Paragon Plus Environment

Page 13 of 31

Journal of Agricultural and Food Chemistry

285

the protein levle of the total ERK decreased. JNK and p38, the other two members of

286

MAPK superfamily, were also significantly activated by DATS treatment. In order to

287

further confirm the critial role of MAPK signaling pathway in DATS-induced cell

288

apoptosis , BCPAP cells were pretreated with 20 µΜ of U0126 (ERK inhibitor) for 2

289

h. As shown in Fig. 7B, U0126 pretreatment completely abolished the

290

phosphorylation of ERK1/2 induced by DATS. Meanwhile, the protein levels of Bcl-2

291

and Bax were partially reversed by U0126 in DATS-treated BCPAP cells (Fig. 7B and

292

7C). Besides, JNK and p38 inhibiton were also performed to determine the role of

293

JNK and p38 in cell apoptosis induced by DATS treatment. We found that pretreating

294

with 20 µΜ of JNK inhibitor, SP600125, could rescue cell apoptosis induced by

295

DATS treatment while it could not be restored by p38 inhibitor, SB203580 (Fig. 7D

296

and 7E). Hence, the activation of ERK and JNK but not p38 were involved in the

297

apoptosis induced by DATS. All these data confirmed that DATS elicited apoptosis in

298

BCPAP cells through activating MAPK signaling pathway.

299

Discussion

300

The laboratorial and epidemiological studies demostrated that galic and its derivates

301

had an effective bioactivity against many diseases32. Recently, the functions of DATS

302

has been tested in various disease models. Owing to its antioxidant property, DATS

303

has a cytoprotective effect on hepatocyte cells4. In addition, DATS was proved to play

304

a role in cancer prevention through different machanisms. Previous study

305

demonstrated that DATS could inhibit the proliferation as well as metastasis of breast

306

cancer cell33 and DATS also exhibited its anti-cancer character by sensitizing human

307

melanoma cancer cells to chemotherapy34. Most importantly, DATS could elicit

308

apoptosis in many cancer cells such as lymphoma cells2 and gastric carcinoma cells35.

309

In the present study, our results revealed that DATS induced apoptosis of thyroid 13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

310

Page 14 of 31

cancer cells.

311

The most common follicular cell-derived cancer is papillary thyroid carcinoma

312

(PTC), which usually has a favorable prognosis after surgery, adjuvant radioactive

313

iodine and thyroid-stimulating hormone suppression15. Radioactive iodine treatment is

314

the first-line systemic treatment in patients with advanced thyroid cancer36.

315

selectively causes radiation damages, such as DNA damage, membrane injury and

316

protein oxidation to thyroid follicular epithelial cells including thyroid cancer cells,

317

which eventually induces cell death of thyroid cancer cells37. However, a few patients

318

will become refractory to radioactive iodine treatment. Therefore, it is essential to find

319

an alternative strategy to inducing apoptosis of thyroid cancer cells as a

320

supplementary option to radioactive iodine treatment. Our present results

321

demonstrated that DATS could significantly induce the apoptosis of thyroid cancer

322

BCPAP cells.

131

I

323

In most cancer cells, unlimited cell proliferation is a consequence of the

324

dysregulated cell cycle24. Hence, the disturbance of cell cycle is a promising strategy

325

for cancer therapy24. Previous reports showed that DATS induced cell cycle arrest in

326

different kinds of cancer cells. For example, DATS caused cell cycle arrest at G2/M

327

phase in DU145 human prostate cancer cell line38. Another research also showed that

328

DATS induced cell cycle arrest at G2/M phase in human gastric carcinoma cells via

329

upregulating cyclin B1 and cyclin-dependent kinase p2139. In line with previous

330

findings, we also demonstrated that DATS treatment induced cell cycle arrest at

331

G0/G1 phase of BCPAP cells (Fig. 3), which may be the underlying mechanism of the

332

growth-inhibitory effects of DATS on BCPAP cells.

333

Cell cycle arrest spares time to repair impaired DNA and if fails, it will lead to

334

cell apoptosis40. Apoptosis is termed as type I programmed cell death, and it can be

14

ACS Paragon Plus Environment

Page 15 of 31

Journal of Agricultural and Food Chemistry

335

mediated through two different pathways: intracellular and extracellular apoptotic

336

pathways41. The two pathways differ from the stimuli of cell apoptosis. Extrinsic

337

apoptotic pathway, also called the death receptor pathway, is triggered by

338

ligand-induced activation of death receptors such as Fas and TNF-related apoptosis

339

inducing ligand (TRAIL)34. The intracellular apoptotic pathway is also known as

340

mitochondrial apoptotic pathway, which is mainly induced by intracellular signals

341

such as DNA damage, growth factor starvation and oxidative stress27. The

342

mitochondrial apoptosis pathway is regulated by Bcl-2 family members41. It has been

343

proved that the expression of Bcl-2 and Bcl-XL, the anti-apoptotic protein, will

344

downregulate, accompanied with the upregulation of the expression of Bax, the

345

pro-apoptotic protein. Eventually, Bax directly induces mitochondrial outer membrane

346

permeabilization (MOMP) followed by cytochrome c releasing to cytoplasm, which

347

activates cascade of caspases41. Once activated, caspase family will trigger a chain of

348

intracellular events that elicit cell death41. Therefore, inducing cancer cell apoptosis is

349

an effective method for cancer chemoprevention and therapy. Consistent with these

350

typical apoptotic features, our results showed that DATS-induced BCPAP cell

351

apoptosis depended on caspase cascade activation through a mitochondrial apoptotic

352

pathway (Fig. 4 and Fig. 5).

353

Mitochondria is the factory of ROS, and it is reported that excessive ROS causes

354

severe oxidative damage to intracellular molecules and eventually induces cell

355

apoptosis42. Indeed, DATS-induced apoptosis has been reported to be mediated by

356

excessive ROS production in human breast cancer cells43. Contrary to our expectation,

357

our data showed that DATS failed to altered the intracellular level of ROS in BCPAP

358

cells, indicating a non-involvement of ROS in DATS-induced apoptosis in thyroid

359

cancer cells (Fig. 6).

15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 31

360

Many genetic alterations have been identified in thyroid cancer which derived

361

from follicular cells. Among them, BRAF and RAS mutations are the most common

362

mutations in papillary thyroid cancer15. BRAF or RAS mutations always drive

363

constitutive activation of MAPK signaling pathway, which is generally believed to be

364

a survival signal in cancer cells44. Hence, targeting this signaling pathway is essential

365

for the treatment of thyroid cancer. In clinical practice, selumetinib, a kind of tyrosine

366

kinase inhibitor targeting MAPK signaling pathway, is used as a promising medicine

367

in thyroid cancer therapy45. Thomas C. Beadnell reported that inhibition of MEK or

368

ERK drove synergistic inhibition of growth and induction of apoptosis in thyroid

369

cancer46. In another study, local anesthetics was also found to be able to induce

370

apoptosis in thyroid cancer cells through attenuating extracellular signal-regulated

371

kinase 1/2 (ERK1/2) activity47. Yet this conclusion remained controversial, as some

372

other studies have highlighted the fact that transient activation but not inhibition of

373

ERK could also lead to the apoptosis in thyroid cancer cell lines22. This discrepancy

374

not only exists in thyroid cells, but also in other cancer type such as hepatocytes. M.

375

D. Abid et al. found that MAPK activation promoted hepatic cellular apoptosis48. On

376

the contrary, other reports showed that suppression of ERK could also induce

377

hepatocytes apoptosis49. These paradoxical conclusions imply that the survival or

378

death fate as a consequence of modulation of MAPK signaling pathway appears to be

379

highly cell type and context dependent. Of note, how long and to what extent does

380

MAPK signaling activate may affect different downstream targets, leading to a

381

different response to various stimuli. In our study, we confirmed that DATS induced

382

an excessive activation of MAPK signaling including ERK and JNK rather than p38,

383

which played an essential role in promoting apoptosis of thyroid cancer BCPAP cells

384

(Fig. 7). 16

ACS Paragon Plus Environment

Page 17 of 31

Journal of Agricultural and Food Chemistry

385

Concerning the absorption, distribution, metabolism and excretion (ADME)

386

characteristics of DATS, it has been proven that the absorption of DATS was excellent

387

and fast50. Additionally, DATS could rapidly distribute to blood, lung, liver, and it

388

metabolized instantly in vivo50, 51. Also, no noticeable toxicity of aged garlic extract

389

on human was found in preclinical trial52. Moreover, DATS with three sulfur atoms in

390

its structure is a donor of hydrogen sulfide, which is believed as the third gaseous

391

signal molecule in vivo53. It has been reported that exposure to exogenous hydrogen

392

sulfide increased the apoptotic percentage of oral cancer cells54. Whether hydrogen

393

sulfide is also attributed to DATS-induced cell apoptosis in thyroid cancer needs

394

further investigation.

395

In conclusion, to the best of our knowledge, our study demonstrated for the first

396

time that DATS significantly inhibited cell proliferation through causing G0/G1 phase

397

cell cycle arrest in thyroid cancer BCPAP cells. Additionally, we found that DATS

398

induced ROS-independent apoptosis via activating MAPK signaling pathway.

399

ABBREVIATIONS

400

DATS, diallyl trisulfide;

401

DAS, diallyl sulfide;

402

DADS, diallyl disulfide;

403

OSCs, organosulfur compounds;

404

ROS, reactive oxygen species;

405

∆Ψm, mitochondrial membrane potential;

406

Bax, Bcl-2-associated X protein;

407

Bcl-2, B-cell lymphoma 2;

408

PARP, poly(ADP-ribose) polymerase;

409

PI, propidium iodide; 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 31

410

FBS, fetal bovine serum;

411

DMSO, dimethyl sulfoxide;

412

PBS, phosphate-buffered saline;

413

MAPK, mitogen-activated protein kinase;

414

JNK, c-Jun N-terminal kinase;

415

ERK, extracellular signal-regulated kinase;

416

PTC, papillary thyroid carcinoma;

417

Funding

418

This study was funded by the grants from the Science and Research Foundation of

419

Health Bureau of Jiangsu Province (No. H2017032), the National Natural Science

420

Foundation of China (Nos. 81602352 and 81602353), the Science and Research

421

Foundation of Health Bureau of Jiangsu Province (No. H2017032), the Natural

422

Science Foundation of Jiangsu Province (BK20171145 and BK20151119), and Wuxi

423

Municipal Commission of Health and Family Planning (Q201608).

424

Conflict of interest

425

The authors declare no competing financial interest.

426

18

ACS Paragon Plus Environment

Page 19 of 31

Journal of Agricultural and Food Chemistry

427

References

428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443 444 445 446 447 448 449 450 451 452 453 454 455 456 457 458 459 460 461 462 463 464 465 466 467 468 469

1. Rivlin, R. S., Historical perspective on the use of garlic. J. Nutr. 2001, 131, 951S-954S. 2. Shigemi, Z.; Furukawa, Y.; Hosokawa, K.; Minami, S.; Matsuhiro, J.; Nakata, S.; Watanabe, T.; Kagawa, H.; Nakagawa, K.; Takeda, H.; Fujimuro, M., Diallyl trisulfide induces apoptosis by suppressing NF-κB signaling through destabilization of TRAF6 in primary effusion lymphoma. Int. J. Oncol. 2015, 48, 293-304. 3. Chen, C.; Pung, D.; Leong, V.; Hebbar, V.; Shen, G.; Nair, S.; Li, W.; Kong, A. N., Induction of detoxifying enzymes by garlic organosulfur compounds through transcription factor Nrf2: effect of chemical structure and stress signals. Free Radic. Biol. Med. 2004, 37, 1578-1590. 4. Zhang, F.; Zhang, Y.; Wang, K.; Liu, G.; Yang, M.; Zhao, Z.; Li, S.; Cai, J.; Cao, J., Protective effect of diallyl trisulfide against naphthalene-induced oxidative stress and inflammatory damage in mice. Int. J. Immunopathol. Pharmacol. 2016, 29, 205-16. 5. Yu, L. M.; Li, S.; Tang, X. L.; Li, Z.; Zhang, J.; Xue, X. D.; Han, J. S.; Liu, Y.; Zhang, Y. J.; Zhang, Y.; Xu, Y. L.; Yang, Y.; Wang, H. S., Diallyl trisulfide ameliorates myocardial ischemia-reperfusion injury by reducing oxidative stress and endoplasmic reticulum stress-mediated apoptosis in type 1 diabetic rats: role of SIRT1 activation. Apoptosis 2017, 22, 942-954. 6. Padiya, R.; Banerjee, S. K., Garlic as an anti-diabetic agent: recent progress and patent reviews. Recent Pat. Food Nutr. Agric. 2013, 5, 105-27. 7. Miltonprabu, S.; Sumedha, N. C.; Senthilraja, P., Diallyl trisulfide, a garlic polysulfide protects against As-induced renal oxidative nephrotoxicity, apoptosis and inflammation in rats by activating the Nrf2/ARE signaling pathway. Int. Immunopharmacol. 2017, 50, 107-120. 8. Ma, H. B.; Huang, S.; Yin, X. R.; Zhang, Y.; Di, Z. L., Apoptotic pathway induced by diallyl trisulfide in pancreatic cancer cells. World J. Gastroenterol. 2014, 20, 193-203. 9. Wang, H. C.; Chu, Y. L.; Hsieh, S. C.; Sheen, L. Y., Diallyl trisulfide inhibits cell migration and invasion of human melanoma a375 cells via inhibiting integrin/facal adhesion kinase pathway. Environ. Toxicol. 2017, 32, 2352-2359. 10. Wei, Z. H.; Shan, Y. L.; Tao, L.; Liu, Y. P.; Zhu, Z. J.; Liu, Z. G.; Wu, Y. Y.; Chen, W. X.; Wang, A. Y.; Lu, Y., Diallyl trisulfides, a natural histone deacetylase inhibitor, attenuate HIF-1 synthesis, and decreases breast cancer metastasis. Mol. Carcinog. 2017, 56, 2317-2331. 11. Jin, Z. Y.; Wu, M.; Han, R. Q.; Zhang, X. F.; Wang, X. S.; Liu, A. M.; Zhou, J. Y.; Lu, Q. Y.; Zhang, Z. F.; Zhao, J. K., Raw garlic consumption as a protective factor for lung cancer, a population-based case-control study in a Chinese population. Cancer Prev. Res. (Phila.) 2013, 6, 711-8. 12. Shin, H. A.; Cha, Y. Y.; Park, M. S.; Kim, J. M.; Lim, Y. C., Diallyl sulfide induces growth inhibition and apoptosis of anaplastic thyroid cancer cells by mitochondrial signaling pathway. Oral Oncol. 2010, 46, e15-8. 13. Cabanillas, M. E.; McFadden, D. G.; Durante, C., Thyroid cancer. The Lancet 2016, 388, 2783-2795. 14. Cancer Genome Atlas Research, N., Integrated genomic characterization of papillary 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

470 471 472 473 474 475 476 477 478 479 480 481 482 483 484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513

15. 16.

17.

18.

19.

20.

21.

22.

23.

24.

25. 26.

27.

28.

Page 20 of 31

thyroid carcinoma. Cell 2014, 159, 676-90. Carling, T.; Udelsman, R., Thyroid cancer. Annu. Rev. Med. 2014, 65, 125-37. Yang, X.; Li, J.; Li, X. Y.; Liang, Z. Y.; Gao, W.; Liang, J.; Cheng, S. J.; Lin, Y. S., TERT Promoter Mutation Predicts Radioiodine-Refractory Character in Distant Metastatic Differentiated Thyroid Cancer. J. Nucl. Med. 2017, 58, 258-265. Jiang, X. Y.; Zhu, X. S.; Xu, H. Y.; Zhao, Z. X.; Li, S. Y.; Li, S. Z.; Cai, J. H.; Cao, J. M., Diallyl trisulfide suppresses tumor growth through the attenuation of Nrf2/Akt and activation of p38/JNK and potentiates cisplatin efficacy in gastric cancer treatment. Acta Pharmacol. Sin. 2017, 38, 1048-1058. Shigemi, Z.; Furukawa, Y.; Hosokawa, K.; Minami, S.; Matsuhiro, J.; Nakata, S.; Watanabe, T.; Kagawa, H.; Nakagawa, K.; Takeda, H.; Fujimuro, M., Diallyl trisulfide induces apoptosis by suppressing NF-kappa B signaling through destabilization of TRAF6 in primary effusion lymphoma. Int. J. Oncol. 2016, 48, 293-304. Kim, W. G.; Choi, H. J.; Kim, T. Y.; Shong, Y. K.; Kim, W. B., The effect of 5-aminoimidazole-4-carboxamide-ribonucleoside was mediated by p38 mitogen activated protein kinase signaling pathway in FRO thyroid cancer cells. Korean J. Intern. Med. 2014, 29, 474-81. Zhao, Y.; You, H.; Yang, Y.; Wei, L.; Zhang, X.; Yao, L.; Fan, D.; Yu, Q., Distinctive regulation and function of PI 3K/Akt and MAPKs in doxorubicin-induced apoptosis of human lung adenocarcinoma cells. J. Cell. Biochem. 2004, 91, 621-32. Kimura, E. T.; Nikiforova, M. N.; Zhu, Z.; Knauf, J. A.; Nikiforov, Y. E.; Fagin, J. A., High prevalence of BRAF mutations in thyroid cancer: genetic evidence for constitutive activation of the RET/PTC-RAS-BRAF signaling pathway in papillary thyroid carcinoma. Cancer Res. 2003, 63, 1454-7. Shih, A.; Davis, F. B.; Lin, H. Y.; Davis, P. J., Resveratrol induces apoptosis in thyroid cancer cell lines via a MAPK- and p53-dependent mechanism. J. Clin. Endocrinol. Metab. 2002, 87, 1223-32. Xiao, D.; Choi, S.; Johnson, D. E.; Vogel, V. G.; Johnson, C. S.; Trump, D. L.; Lee, Y. J.; Singh, S. V., Diallyl trisulfide-induced apoptosis in human prostate cancer cells involves c-Jun N-terminal kinase and extracellular-signal regulated kinase-mediated phosphorylation of Bcl-2. Oncogene 2004, 23, 5594-5606. Wang, J.; Li, X. M.; Bai, Z.; Chi, B. X.; Wei, Y.; Chen, X., Curcumol induces cell cycle arrest in colon cancer cells via reactive oxygen species and Akt/ GSK3beta/cyclin D1 pathway. J. Ethnopharmacol. 2018, 210, 1-9. Lim, S.; Kaldis, P., Cdks, cyclins and CKIs: roles beyond cell cycle regulation. Development 2013, 140, 3079-93. Zamzami, N.; Marchetti, P.; Castedo, M.; Zanin, C.; Vayssiere, J. L.; Petit, P. X.; Kroemer, G., Reduction in mitochondrial potential constitutes an early irreversible step of programmed lymphocyte death in vivo. J. Exp. Med. 1995, 181, 1661-72. Radogna, F.; Dicato, M.; Diederich, M., Cancer-type-specific crosstalk between autophagy, necroptosis and apoptosis as a pharmacological target. Biochem. Pharmacol. 2015, 94, 1-11. Suski, J., Relation Between Mitochondrial Membrane Potential and ROS Formation. In Methods Mol. Biol., Humana Press Clifton, Pasaik County, New Jersey, USA 2012; Vol. 20

ACS Paragon Plus Environment

Page 21 of 31

Journal of Agricultural and Food Chemistry

514 515 516 517 518 519 520 521 522 523 524 525 526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557

29. 30.

31.

32.

33.

34.

35.

36. 37. 38.

39.

40. 41.

810, pp 183-205. Ly, J. D.; Grubb, D. R.; Lawen, A., The mitochondrial membrane potential (deltapsi(m)) in apoptosis; an update. Apoptosis 2003, 8, 115-28. Son, H. Y.; Hwangbo, Y.; Yoo, S. K.; Im, S. W.; Yang, S. D.; Kwak, S. J.; Park, M. S.; Kwak, S. H.; Cho, S. W.; Ryu, J. S.; Kim, J.; Jung, Y. S.; Kim, T. H.; Kim, S. J.; Lee, K. E.; Park, D. J.; Cho, N. H.; Sung, J.; Seo, J. S.; Lee, E. K.; Park, Y. J.; Kim, J. I., Genome-wide association and expression quantitative trait loci studies identify multiple susceptibility loci for thyroid cancer. Nature communications 2017, 8, 15966. Zhou, Q. Y.; Chen, J.; Feng, J. L.; Xu, Y. N.; Zheng, W. J.; Wang, J. D., SOSTDC1 inhibits follicular thyroid cancer cell proliferation, migration, and EMT via suppressing PI3K/Akt and MAPK/Erk signaling pathways. Mol. Cell. Biochem. 2017, 435, 87-95. Zhang, L.; Cheng, X.; Gao, Y.; Bao, J.; Guan, H.; Lu, R.; Yu, H.; Xu, Q.; Sun, Y., Induction of ROS-independent DNA damage by curcumin leads to G2/M cell cycle arrest and apoptosis in human papillary thyroid carcinoma BCPAP cells. Food Funct. 2016, 7, 315-25. Chandra-Kuntal, K.; Lee, J.; Singh, S. V., Critical role for reactive oxygen species in apoptosis induction and cell migration inhibition by diallyl trisulfide, a cancer chemopreventive component of garlic. Breast Cancer Res. Treat. 2013, 138, 69-79. Murai, M.; Inoue, T.; Suzuki-Karasaki, M.; Ochiai, T.; Ra, C.; Nishida, S.; Suzuki-Karasaki, Y., Diallyl trisulfide sensitizes human melanoma cells to TRAIL-induced cell death by promoting endoplasmic reticulum-mediated apoptosis. Int. J. Oncol. 2012, 41, 2029-37. Choi, Y. H., Diallyl trisulfide induces apoptosis and mitotic arrest in AGS human gastric carcinoma cells through reactive oxygen species-mediated activation of AMP-activated protein kinase. Biomedicine & pharmacotherapy = Biomedecine & pharmacotherapie 2017, 94, 63-71. Cabanillas, M. E.; McFadden, D. G.; Durante, C., Thyroid cancer. Lancet 2016, 388, 2783-2795. Ryu, J., Diagnostic and Therapeutic Approaches to Radioactive Iodine Refractory Differentiated Thyroid Cancer. Medicine 2012, 55, 403. Herman-Antosiewicz, A.; Kim, Y.-A.; Kim, S.-H.; Xiao, D.; Singh, S. V., Diallyl Trisulfide-Induced G2/M Phase Cell Cycle Arrest in DU145 Cells Is Associated with Delayed Nuclear Translocation of Cyclin-Dependent Kinase 1. Pharm. Res. 2010, 27, 1072-1079. Shin, S. S.; Song, J. H.; Hwang, B.; Park, S. L.; Kim, W. T.; Park, S. S.; Kim, W. J.; Moon, S. K., Angiopoietin-like protein 4 potentiates DATS-induced inhibition of proliferation, migration, and invasion of bladder cancer EJ cells; involvement of G(2)/M-phase cell cycle arrest, signaling pathways, and transcription factors-mediated MMP-9 expression. Food Nutr. Res. 2017, 61, 1338918. Malumbres, M.; Barbacid, M., Cell cycle, CDKs and cancer: a changing paradigm. Nat. Rev. Cancer 2009, 9, 153-66. Radogna, F.; Dicato, M.; Diederich, M., Cancer-type-specific crosstalk between autophagy, necroptosis and apoptosis as a pharmacological target. Biochem. Pharmacol. 2015, 94, 1-11. 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601

Page 22 of 31

42. Pierce, G. B.; Parchment, R. E.; Lewellyn, A. L., Hydrogen peroxide as a mediator of programmed cell death in the blastocyst. Differentiation 1991, 46, 181-6. 43. Na, H. K.; Kim, E. H.; Choi, M. A.; Park, J. M.; Kim, D. H.; Surh, Y. J., Diallyl trisulfide induces apoptosis in human breast cancer cells through ROS-mediated activation of JNK and AP-1. Biochem. Pharmacol. 2012, 84, 1241-50. 44. Xing, M., Molecular pathogenesis and mechanisms of thyroid cancer. Nat. Rev. Cancer 2013, 13, 184-99. 45. Hayes, D. N.; Lucas, A. S.; Tanvetyanon, T.; Krzyzanowska, M. K.; Chung, C. H.; Murphy, B. A.; Gilbert, J.; Mehra, R.; Moore, D. T.; Sheikh, A.; Hoskins, J.; Hayward, M. C.; Zhao, N.; O'Connor, W.; Weck, K. E.; Cohen, R. B.; Cohen, E. E., Phase II efficacy and pharmacogenomic study of Selumetinib (AZD6244; ARRY-142886) in iodine-131 refractory papillary thyroid carcinoma with or without follicular elements. Clin. Cancer Res. 2012, 18, 2056-65. 46. Beadnell, T. C.; Mishall, K. M.; Zhou, Q.; Riffert, S. M.; Wuensch, K. E.; Kessler, B. E.; Corpuz, M. L.; Jing, X.; Kim, J.; Wang, G.; Tan, A. C.; Schweppe, R. E., The Mitogen-Activated Protein Kinase Pathway Facilitates Resistance to the Src Inhibitor Dasatinib in Thyroid Cancer. Mol. Cancer Ther. 2016, 15, 1952-63. 47. Chang, Y. C.; Hsu, Y. C.; Liu, C. L.; Huang, S. Y.; Hu, M. C.; Cheng, S. P., Local anesthetics induce apoptosis in human thyroid cancer cells through the mitogen-activated protein kinase pathway. PLoS One 2014, 9, e89563. 48. Abid, M. D. N.; Chen, J.; Xiang, M.; Zhou, J.; Chen, X. P.; Gong, F. L., Khat (Catha edulis) generates reactive oxygen species and promotes hepatic cell apoptosis via MAPK activation. Int. J. Mol. Med. 2013, 32, 389-395. 49. Du, X. L.; Shi, Z.; Peng, Z. C.; Zhao, C. X.; Zhang, Y. M.; Wang, Z.; Li, X. B.; Liu, G. W.; Li, X. W., Acetoacetate induces hepatocytes apoptosis by the ROS-mediated MAPKs pathway in ketotic cows. J. Cell. Physiol. 2017, 232, 3296-3308. 50. Sun, X.; Guo, T.; He, J.; Zhao, M.; Yan, M.; Cui, F.; Deng, Y., Determination of the concentration of diallyl trisulfide in rat whole blood using gas chromatography with electron-capture detection and identification of its major metabolite with gas chromatography mass spectrometry. Yakugaku Zasshi Journal of the Pharmaceutical Society of Japan 2006, 126, 521-7. 51. Lawson, L. D.; Wang, Z. J., Allicin and allicin-derived garlic compounds increase breath acetone through allyl methyl sulfide: use in measuring allicin bioavailability. Journal of Agricultural & Food Chemistry 2005, 53, 1974-1983. 52. Tanaka, S.; Haruma, K.; Yoshihara, M.; Kajiyama, G.; Kira, K.; Amagase, H.; Chayama, K., Aged garlic extract has potential suppressive effect on colorectal adenomas in humans. J Nutr 2006, 136, 821S-826S. 53. Liang, D.; Wu, H.; Wong, M. W.; Huang, D., Diallyl Trisulfide Is a Fast H2S Donor, but Diallyl Disulfide Is a Slow One: The Reaction Pathways and Intermediates of Glutathione with Polysulfides. Org. Lett. 2015, 17, 4196-9. 54. Murata, T.; Sato, T.; Kamoda, T.; Moriyama, H.; Kumazawa, Y.; Hanada, N., Differential susceptibility to hydrogen sulfide-induced apoptosis between PHLDA1-overexpressing oral cancer cell lines and oral keratinocytes: role of PHLDA1 as an apoptosis suppressor. Exp. Cell Res. 2014, 320, 247-57. 22

ACS Paragon Plus Environment

Page 23 of 31

Journal of Agricultural and Food Chemistry

602

Figure Captions

603

Figure 1. The chemical structure of DATS.

604 605

Figure 2. DATS inhibits the colony formation ability of BCPAP cells. (A) DATS

606

treatment resulted in the morphological change of BCPAP cells. BCPAP cells were

607

exposed to different concentrations of DATS (5-20 µΜ) for 24 h and then

608

photographed with inverted microscope. Scale bar, 20 µm. (B) DATS inhibited cell

609

growth of BCPAP cells in a dose-dependent manner. Cells were treated with indicated

610

concentrations of DATS for 24 h and then the cell numbers were measured by SRB

611

assay. (C) DATS inhibited the colony formation ability of BCPAP cells. The cells

612

were incubated with various concentrations of DATS for 24 h and then stained with

613

crystal violet. Colonies consisting of more than 50 cells were counted for each group.

614

(D) Quantitative analysis of clone numbers in (C). (E) DATS inhibited the cell growth

615

of BCPAP cells in a time-dependent manner. Cells were treated with DATS at 20 µΜ

616

for indicated periods and then the cell numbers were measured by SRB assay. (F)

617

DATS exhibited selective cytotoxicity on thyroid cancer cells. BCPAP and

618

Nthy-ori-3.1 cells were treated with 5-20 µΜ of DATS for 24 h and the cell numbers

619

were determined by SRB assay. All data were represented as mean ± S.D. *P < 0.05,

620

**

621

Figure 3. DATS causes G0/G1 cell cycle arrest in BCPAP cells. (A) BCPAP cells

622

were treated with different concentrations of DATS for 24 h. Cell cycle distributions

623

were detected by flow cytometry and analyzed by FlowJo software. M1, M2, M3 and

624

M4 gates represented the sub-G1 phase, G0/G1 phase, S phase and G2/M phase,

625

respectively. (B) Cell cycle distributions after DATS treatment in BCPAP cells were

626

analyzed and the data were plotted in the bar chart. (C) After BCPAP cells were

P < 0.01, ***P < 0.001 vs. SC group. SC, solvent control.

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 31

627

treated with various concentrations of DATS (5-20 µΜ), the protein levels of Cyclin

628

D1 and p-Rb and p27 were determined by western blot. Tubulin was used as an

629

internal control. (D) The bands quantification of (C) were determined from three

630

independent experiments (n = 3). *P < 0.05 vs. SC group of cyclin D1, #P < 0.05 and

631

##

632

Figure 4. DATS induces apoptosis of BCPAP cells. (A) BCPAP cells were exposed

633

to different concentrations of DATS for 24 h and then stained with Hoechst 33342 and

634

propidium iodide (PI). White arrows indicated the PI positive cells and green arrows

635

indicated the bright Hoechst 33342 staining cells. Images were taken under a

636

fluorescence microscope and representive image were shown. Scale bar, 20 µm. (B)

637

Cells were incubated with DATS for 24 h and subjected to mitochondrial membrane

638

potential analysis by flow cytometry. (C) The quatification of JC-1 monomer cells in

639

(B). The symbol ''*'' is compared with SC group in statistical analysis of JC-1

640

monomer cells. One-way ANOVA analysis followed by Dunnett's test for multiple

641

comparisons was used to analyze the statistical significance. *P < 0.05 vs. SC group.

642

SC, solvent control.

643

Fig. 5. DATS-induced apoptosis is mediated by caspase activation. (A) DATS

644

induced apoptosis of BCPAP cells in a mitochondria-mediated pathway. After BCPAP

645

cells were treated with various concentrations of DATS, the protein levels of Bcl-2,

646

Bax, Bcl-XL. Actin and tubulin were used as internal controls. (B) The bands

647

quantification of Bcl-XL, Bcl-2 and Bax were analyzed by Gel Image System of

648

Tanon. *P < 0.05 vs. SC group of Bcl-2. #P < 0.05 vs. SC group of Bcl-XL. &P < 0.05

649

vs. SC group of Bax. SC, solvent control. (C) BCPAP cells were treated with various

650

concentrations of DATS and then the protein levels of cleaved capase-3, -8 and PARP

651

were determined by western blot. C-PARP, C-Caspase-8 and C-Caspase-3 represented

P < 0.01 vs. SC group of p-Rb, &P < 0.05 vs. SC group of p27. SC, solvent control.

24

ACS Paragon Plus Environment

Page 25 of 31

Journal of Agricultural and Food Chemistry

652

their cleaved form respectively. Actin and tubulin were used as internal controls. The

653

symbol ''*'' represented the quantification of their cleaved form of caspase-8,caspase-3

654

and PARP. (D) DATS-induced decrease in cell survival were resuced by z-VAD-fmk.

655

BCPAP cells were co-treated with indicated concentrations of DATS (5-20 µΜ) and

656

z-VAD-fmk (10 µΜ or 20 µΜ) and then the cell survival was measured by SRB assay.

657

The symbol ''*'' is compared with the individual DATS treatment group represented as

658

a white column. *P < 0.05, ***P < 0.001 vs. SC group. SC, solvent control.

659

Figure 6. DATS-induced apoptosis is independent of ROS production. (A) BCPAP

660

cells were collected after different dosages of DATS treatment for 24 h. Then the cells

661

were stained with 10 µM of DCFH-DA at 37 °C for 20 min and analyzed by flow

662

cytometry. (B) Histogram indicated the quantification of mean fluorescent intensity

663

(MFI) in three independent experiments. N.S., no significant difference.

664

Figure 7. DATS activates MAPK signaling pathway in BCPAP cells. (A) BCPAP

665

cells were treated with indicated concentrations of DATS for 24 h. The protein levels

666

of the total and phosphorylated ERK, JNK and p38 were determined by western blot.

667

Actin and tubulin were used as internal controls. (B) BCPAP cells were pretreated

668

with U0126 (20 µΜ) for 2 h and then incubated with DATS (20 µΜ) for another 24 h.

669

The protein levels of Bax, Bcl-2, total and phosphorylated ERK were determined by

670

western blot. Actin was used as an internal control. (C) The quantification of Bcl-2

671

and Bax in (B). (D) BCPAP cell were pretreated with 20 µΜ of SP600125 for 2 h and

672

then the cells were co-treated with 20 µΜ of DATS and SP600125 for another 24 h.

673

Then the level of p-JNK and cleaved PARP were determined by western blot. The

674

symbol ''*'' represented the quantification of cleaved PARP. (E) BCPAP cells were

675

pretreated with 20 µΜ of SB203580 for 2 h and then the cells were co-treated with 20

676

µΜ of DATS and SB203580 for additional 24 h. Then the level of p-p38 and cleaved

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 31

677

PARP were determined by western blot. Tubulin was used as an internal control. The

678

symbol ''*'' represented the quantification of cleaved form of PARP.

679

26

ACS Paragon Plus Environment

Page 27 of 31

Journal of Agricultural and Food Chemistry

680

Figure 1. The chemical structure of DATS.

681 682

Figure 2. DATS inhibits the colony formation ability of BCPAP cells.

683 27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

684

Page 28 of 31

Figure 3. DATS causes G0/G1 cell cycle arrest in BCPAP cells.

685 686

Fig. 4 DATS induces apoptosis of BCPAP cells.

687 688

Fig. 5. DATS-induced apoptosis is mediated by caspase activation.

28

ACS Paragon Plus Environment

Page 29 of 31

Journal of Agricultural and Food Chemistry

689 690

Figure 6. DATS-induced apoptosis is independent of ROS production.

691 692

Figure 7. DATS activates MAPK signaling pathway in BCPAP cells.

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 31

693 694

30

ACS Paragon Plus Environment

Page 31 of 31

Journal of Agricultural and Food Chemistry

695

Graphic for table of Contents

696

31

ACS Paragon Plus Environment