Chlorpyrifos Induced Testicular Cell Apoptosis through Generation of

2 days ago - Chlorpyrifos (CPF) is the most frequently applied insecticide. Aside from effects on the neuronal cholinergic system, previous studies su...
0 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Agricultural and Environmental Chemistry

Chlorpyrifos Induced Testicular Cell Apoptosis through Generation of Reactive Oxygen Species and Phosphorylation of AMPK Rui Chen, Yang Cui, Xuelian Zhang, Yanghai Zhang, Mingyue Chen, Tong Zhou, Xianyong Lan, Wuzi Dong, and Chuanying Pan J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b03407 • Publication Date (Web): 31 Oct 2018 Downloaded from http://pubs.acs.org on November 1, 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 47

Journal of Agricultural and Food Chemistry

1

Chlorpyrifos Induced Testicular Cell Apoptosis through

2

Generation of Reactive Oxygen Species and Phosphorylation

3

of AMPK

4

Rui Chen †, Yang Cui†, Xuelian Zhang†, Yanghai Zhang†, Mingyue Chen†, Tong Zhou†,

5

Xianyong Lan†, Wuzi Dong†, Chuanying Pan†*

6



7

Shaanxi, 712100, China.

8

*The corresponding author.

College of Animal Science and Technology, Northwest A&F University, Yangling,

9 10

The first author:

Rui Chen

E-mail: [email protected]

11

The second author:

Yang Cui

E-mail: [email protected]

12

The third author:

Xuelian Zhang

E-mail: [email protected]

13

The fourth author:

Yanghai Zhang

E-mail: [email protected]

14

The fifth author:

Mingyue Chen

E-mail: [email protected]

15

The sixth author:

16

The seventh author:

Xianyong Lan

17

The eighth author:

Wuzi Dong

E-mail: [email protected]

18

The ninth author:

Chuanying Pan

E-mail: [email protected]

Tong Zhou

E-mail: [email protected] E-mail: [email protected]

19 20

The corresponding author Chuanying Pan at College of Animal Science and Technology,

21

Northwest A&F University, No. 22 Xinong Road, Yangling, Shaanxi 712100, P.R.

22

China.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

23

ABSTRACT

24

Chlorpyrifos (CPF) is the most frequently applied insecticide. Aside from effects on the

25

neuronal cholinergic system, previous studies suggested a potential relationship

26

between CPF exposure and male infertility; however, the molecular mechanism

27

remains elusive. The aim of this study was to investigate the toxic effect of CPF on

28

testicular cells and the potential mechanism via in vitro and in vivo experiments. The

29

cytotoxic effects of CPF on mouse-derived spermatogonial cell lines (GC-1), Sertoli

30

cell lines (TM4) and Leydig cell lines (TM3) were assessed by CCK-8 assay, flow

31

cytometry, TUNEL assay, quantitative RT-PCR, and western blot. The exposure to CPF

32

(10 to 50 μM) for 12 or 24 h resulted in significant death in all the three testicular cell

33

lines. The number of TUNEL-positive apoptotic cells were dose-dependent and

34

increased with raised CPF concentration. Further investigation indicated that CPF

35

induced cell cycle arrest and then promoted cell apoptosis. Additionally, CPF increased

36

reactive oxygen species (ROS) and lipid peroxidation (MDA) production, and reduced

37

mitochondrial membrane potential. The mechanism involved in cell apoptosis induced

38

by CPF was an increment of phosphorylated AMP-activated protein kinase (p-AMPK)

39

levels in the tested cells. In vivo, the expression of steroid hormone bio-synthesis related

40

genes in testis, spleen, and lung in F0 and F1 mice were downregulated when there was

41

an intraperitoneal injection or dietary supplementation of CPF. This study provides a

42

potential molecular mechanism of CPF-induced toxicity in testicular cells and a

43

theoretical basis for future treatment of male infertility.

ACS Paragon Plus Environment

Page 2 of 47

Page 3 of 47

Journal of Agricultural and Food Chemistry

44

KEYWORDS: chlorpyrifos, male reproduction toxicity, testicular cell lines,

45

oxidative stress, AMPK, F0 and F1 mice

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

46

INTRODUCTION

47

According to Phillips McDougall of Agribusiness informa, the global pesticide market

48

has increased from $25.1 billion to $56.6 billion, with an overall growth rate of 125%

49

over the past 14 years (https://phillipsmcdougall.agribusinessintelligence.informa.com).

50

Although the wide use of pesticides brings in more economic value, pesticide residue

51

can cause irreparable damage to the environment, and long-term health problems for

52

humans and animals.1 Currently, much effort has been given to study the effects of the

53

environmental factors on health, which promotes a healthy environment for the

54

reproduction of healthy offspring. Regrettably, Levine et al (2017) found that the sperm

55

count of western men dropped by 50% over the last half century.2 A major cause for the

56

observed defects in male reproductive function is exposure to industrial and agricultural

57

toxins, and the by-products of other technological advancements.3 Investigators have

58

hypothesized that the wide use of chemicals, pesticides, heavy metals, and

59

hyperthermia, are key factors that result in male infertility.4

60

Chlorpyrifos [O,O-diethyl-O-(3,5,6-trichloro-2-pyridyl)-phos-phorothioate] (CPF)

61

is a broad spectrum organophosphorus pesticide (OP) that is commonly used in

62

agricultural, industrial, and domestic applications.5,6 In 2016, the global sales of CPF

63

were $686 million, and it was predicted that the compound annual growth rate (CAGR)

64

of CPF will reach 6.1% between 2014 and 2020 (http://cn.agropages.com/). Initially,

65

CPF was used for pest control, because the peripheral cholinergic nervous systems were

66

the target of CPF in insects.7,8 CPF is able to inhibit acetylcholinesterase (AChE)

67

activity, which results in an accumulation of acetylcholine and subsequent hyperactivity

68

in the cholinergic system.6 Additionally, it has been reported to cause hepatotoxicity,9

69

developmental toxicity,10,11 genotoxicity,11 immunological abnormalities12 and cell

70

signaling transduction.13

71

In multicellular organisms, the development of gonads and germ cells is essential

72

for the transmission of genetic information to the next generation and ultimately for the

73

survival of species.14 As the most important reproductive organ in male individuals,

74

testis attract a great deal of attention. CPF was administered orally to male mice at

ACS Paragon Plus Environment

Page 4 of 47

Page 5 of 47

Journal of Agricultural and Food Chemistry

75

different doses for 4 weeks, the number of live fetuses were decreased at high dose

76

group, compared with that of control group, which also accompanied by an increased

77

number of dead fetuses in CPF treated mice. Additionally, the sperm counts and sperm

78

motility were markedly reduced, the sperm malformation rate in exposed males was

79

also went up significantly.15 This phenomenon was further confirmed by Sai et al (2014)

80

using rats as a model.16 Moreover, Sai et al (2014) also demonstrated that testosterone

81

(T) levels decreased, and there was a statistical difference between the treatment group

82

and the control group.16 The association between metabolite of CPF and serum

83

reproductive hormone levels were also explored in adult men, which showed an inverse

84

association between CPF metabolite and T concentration.17 In addition, the effect of

85

CPF on testicular oxidative damage was also studied. The expression levels of

86

glutathione (GSH) and antioxidant enzymes presented a downward trend in testis of

87

CPF-treated rats.18 Recently, research indicated that extended exposure of CPF given

88

rise to damage in the process of spermatogenesis, which probably through interference

89

with sex hormones and AchE enzyme levels, thus, resulting in reduction of fertility.19

90

Adedara et al (2017) also proved that CPF mediated toxicity along the hypothalamic-

91

pituitary-testicular axis in rats via activating of lipid peroxidation, decreasing the

92

antioxidant enzymes activities and leading to changes in testicular histology.20

93

Although the effects of CPF on male reproduction toxicity have been detected, the

94

mechanisms are unclear. Additionally, only the effects of CPF on the brain between the

95

F0 and F1 generation of mice have been studied,6,21 other organs have not been reported.

96

This study was designed to explore the potential mechanisms of CPF-induced

97

toxic effects in the three most important cell types in mouse testis, which are the mouse-

98

derived spermatogonial (GC-1), Sertoli (TM4) and Leydig (TM3) cell lines.

99

Additionally, for a deeper understanding of CPF toxicity in vivo, intraperitoneal

100

injections and the dietary intake of CPF were performed on mice to detect alterations

101

in mice testis. These results can further extend the knowledge of CPF induced toxicity

102

and provide a theoretical basis for the treatment of male infertility.

103

MATERIALS AND METHODS

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

104

Cell Culture

105

The GC-1, TM4 and TM3 cells (ATCC, VA, USA) were cultured in high-glucose

106

medium (Hyclone, MA, USA) supplemented with 10% Fetal Bovine Serum (FBS;

107

Gibco, CA, USA), and 1% Penicillin/Streptomycin (Hyclone, MA, USA), under

108

controlled conditions (37 oC, 5% CO2). At a 90% confluence, the cells were sub-

109

cultured with 0.25% trypsin (Gibco, CA, USA) for further experiments. The density of

110

the cells in the 96-well plate and the 6-well plate were 0.8×104 and 20×104, respectively.

111

Cells were allowed to attach overnight.

112

Drug Treatment

113

The CPF was purchased from Shanghai Aladdin Bio-Chem Technology Co., LTD

114

(Shanghai, China). As a lipophilic molecule, the combination of serum proteins can

115

neutralize CPF activity, therefore, the cells were transferred to a serum-free medium

116

when CPF treatment.22 The CPF was dissolved in dimethyl sulfoxide (DMSO) and the

117

final concentration of DMSO did not exceed 0.1%. The incubation time when CPF (0-

118

100 µM) treatment ranged from 0 to 24 hours and is indicated in the figures. The DMSO

119

0.1% (vehicle) was added to the control group.

120

Cell Proliferation Assay

121

The GC-1, TM4 and TM3 cells were seeded into 6-well plate overnight. Cells were then

122

treated with CPF (10, 25 or 50 µM) or the vehicle for 12 or 24 h. Subsequently, the cells

123

were harvested with 0.25% trypsin (Gibco, CA, USA) for total cell count. The number

124

of cells (N) were counted using a hemocytometer. The cell numbers were normalized

125

and graphed as a ratio of Nt/N0, the DMSO was defined as N0, and the treatment group

126

defined as Nt.23 The analyses were performed in triplicate.

127

Cell Viability Assay

128

Cell viability of the three cell lines were detected by Cell counting kit-8 (CCK-8, C0037,

129

Beyotime Institute of Biotechnology, Shanghai, China) after CPF treated for 12 or 24

130

h. Viability of DMSO group was measured as control. Cells in each well were cultured

131

in serum-free culture medium containing 10 µL of CCK-8 reaction solution for 2 h at

132

37 oC, then the absorbance at 450 nm was measured by a microplate reader. Each

133

measurement was repeated three times. The data were calculated according to the

ACS Paragon Plus Environment

Page 6 of 47

Page 7 of 47

Journal of Agricultural and Food Chemistry

134

following equation: Cell viability (%) = [(ODtreatment-ODblank)/(ODcontrol-

135

ODblank)] × 100%.24

136

Annexin-V-FLUOS and Propidium Iodide (PI) Double Staining Assay

137

According to manufacturers’ instructions, cells were harvested, washed with PBS and

138

then re-suspended in a 500 µL 1×binding buffer (Annexin-V-FLUOS 5 µL and PI 5 µL;

139

Biobox, Nanjing, China). Subsequently, cells were incubated for 30 minutes at room

140

temperature in the dark, and analyzed by flow cytometry (BD FACSAria™ III, BD

141

Biosciences, USA). The data were analyzed using FCS express 5.0 Software (De Novo

142

Software, Glendale, CA, USA).

143

TUNEL Staining

144

According to prospectus, the apoptosis cells were measured using In Situ Cell Death

145

Detection Kit (Vazyme, Jiangsu, China). The 4’6’-diamidino-2-phenylindole (DAPI;

146

CWBIO, Beijing, China) was used to visualize nucleus. Digital images were captured

147

using a Nikon Eclipse 80i fluorescence microscope camera (Tokyo, Japan).25

148

Quantitative RT-PCR (qRT-PCR)

149

The TRIzol (TaKaRa, Dalian, China) was used to collect the RNA samples of cell after

150

CPF treatment. The qRT-PCR was carried out as previously described using the primers

151

presented in Table 1.26

152

Cell Cycle Assay

153

Cultures were analyzed for cell cycle after 12 or 24 h. Cells were trypsinized, washed

154

with precooled PBS and treated with a cell cycle staining kit. Finally, cells analyzed by

155

flow cytometry (Becton Dickinson, FACSCalibur). Data were analyzed using ModFit

156

Software (Verity Software House).

157

Measurement of Reactive Oxygen Species (ROS) Production

158

The intracellular ROS level was measured by ROS Assay Kit (S0033, Beyotime

159

Institute of Biotechnology, Shanghai, China) following the manufacturer's protocol.

160

When 2′,7′-dichlorodihydrofluorescein diacetate (DCFH-DA) was oxidized by ROS to

161

2′,7′-dichlorofluorescein (DCF), the higher fluorescence intensity could be observed at

162

530 nm. The cells were suspended with high-glucose medium containing 1 µL of

163

DCFH-DA (final concentration was 10 μM/L), and were incubated for 30 min at 37 oC

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

164

in the dark. Finally, the multi-detection microplate reader (Synergy HT, BioTek,

165

Vermont, USA) was used to quantify the relative levels of fluorescence (485 nm

166

excitation and 535 nm emission). 27

167

Measurement of Malonaldehyde (MDA) Level

168

Malonaldehyde was determined by a Lipid Peroxidation MDA Assay Kit (S0131,

169

Beyotime Institute of Biotechnology, Shanghai, China). The cells were prepared as

170

described in kit instructions. The MDA concentration of each sample was evaluated by

171

a multi-detection microplate reader (SpectramMax M5) at 532 nm, using 490 nm as the

172

control. 28

173

Determination of Mitochondrial Membrane Potential (MMP)

174

The MMP (ΔΨm) was detected using fluorescent probe: JC-1 (5,5’,6,6’-Tetrachloro-

175

1,1’,3,3’-tetraethyl benzimidazolyl carbocyanine iodide) (C2006, Beyotime Institute of

176

Biotechnology, Shanghai, China). The increase green fluorescence intensity (FL1) was

177

always accompanied by mitochondrial depolarization, which can be detected by multi-

178

detection microplate reader. 29

179

Western Blot

180

Antibodies against BAX (2772), BCL2 (3498) and p-AMPK (50081) were purchased

181

from Cell Signaling Technology (Beverly, MA, USA), and p-LKB1 (sc-271924; Santa

182

Cruz Biotechnology, USA) and GAPDH (Cell Signaling Technology, USA) were also

183

employed in the experiments.

184

The cells were washed with PBS, and then lysed with RIPA (P0013B, Beyotime

185

Institute of Biotechnology, Shanghai, China), which contained 1 mmol/L of PMSF

186

(ST506, Beyotime Institute of Biotechnology, Shanghai, China). The 10% SDS-PAGE

187

was used to separate the lysates of cells, which followed by transferred to PVDF

188

membranes (Millipore, USA).25 The membranes stained with the reagents in Western

189

Bright ECL Kit were then visualized using Bio-Rad Chemidoc (Bio-Rad, USA).

190

Animal Feeding

191

The Wild-type C57BL/6 mice (6 weeks) were maintained under controlled

192

environment (22±2 oC, 55±10% humidity, 12 h reversed light-dark cycle), and fed

193

with ad libitum food and water at a pathogen-free facility. The experimental animals

ACS Paragon Plus Environment

Page 8 of 47

Page 9 of 47

Journal of Agricultural and Food Chemistry

194

and procedures used in this study were approved by the Faculty Animal Policy and

195

Welfare Committee of Northwest A&F University. The care and use of experimental

196

animals fully complied with local animal welfare laws, guidelines, and policies.

197

Procedures and Experimental Groups for Intraperitoneal Injection of CPF in

198

Mice

199

Previous findings have shown that CPF at a dose of 17.5 mg/kg given orally to male

200

rats for 30 days could induce severe testicular damage.30 In addition, Sai et al.

201

demonstrated that CPF administered orally to male rats at different dose for 90 days

202

had adverse effects on the reproductive system.31 The daily exposure doses of CPF in

203

developing countries like Sri Lanka is 94,000 ng/kg ∙ day-1,32 indicating that exposure

204

levels are lower than 17.5 mg/kg. However, farmers in Sri Lanka used higher levels of

205

CPF than what is recommended for crop protection.19 Therefore, in this study, 3-, 6-,

206

and 12-mg/kg doses were selected. The CPF could be absorbed by body via different

207

approaches, including injection, ingestion, inhalation, and dermal absorption.33

208

Injection and oral ingestion were selected as the exposure mode in this study.

209

For the CPF intraperitoneal injection experiments, animals were randomly divided

210

into four groups that contained 6 mice each (a control group and three exposure groups).

211

In the exposure groups, different doses of CPF (3-, 6-, and 12-mg/kg) were injected

212

intraperitoneally into eight-week-old mice once a day, and consecutive for 35 days

213

(lasted for one spermatogenic cycle). The control mice were administered an equal

214

volume of redistilled water and DMSO.

215

Procedures and Experimental Groups for Dietary Supplementation of CPF in

216

Mice

217

For the CPF dietary supplementation experiments, the amount of CPF added to the diet

218

corresponded to a dose of 3-, 6-, and 12 mg ∙ kg-1 / bw ∙ day-1 according to a food

219

consumption of approximately 5 g/day for each mouse.34 The CPF diet was

220

administered to eight-week-old male and female mice of the F0 generation for 80

221

consecutive days. The offspring (F1 generation) of F0 mice were sacrificed at the time

222

of weaning.

223

At the end of the treatment period, mice were subjected to a 3 h fast before being

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

224

deeply anesthetized with carbon dioxide prior to being euthanized. The tissue samples

225

of mice were fixed in Bouin’s solution or transferred into TRIzol (TaKaRa, Dalian,

226

China) for molecular analysis.

227

Hematoxylin and Eosin (H&E) Staining

228

Testis samples collected from mice were fixed in Bouin’s solution overnight before

229

being dehydrated and embedded in paraffin. The 5 μm cross-sections were adhered to

230

precoated glass slides. The H&E staining of the paraffinembedded sections were

231

conducted to observe histology.35

232

For the CPF intraperitoneal injection experiments, the samples of control group,

233

3- and 6-mg/kg CPF injection group were collected at 2 weeks and 35 days, respectively,

234

and three samples were taken for each group. While for 12-mg/kg CPF injection group,

235

except for sampling at 2 weeks, the mice were not in good condition at 18 days and

236

were not reared for 35 days, thus, the samples at 18 days were taken (n=3).

237

For the CPF dietary supplementation experiments, the testis samples of control

238

group, 3-, 6-, and 12 mg ∙ kg-1 / bw ∙ day-1 CPF diet feeding F0 mice were collected

239

after 80 days of feeding. Three samples in each group were taken. The offspring (F1

240

generation) of F0 mice were sacrificed at the time of weaning, and three samples were

241

collected in each group.

242

Statistical Analysis

243

Statistical treatment of the data was performed with SPSS 19.0 software (SPSS,

244

Chicago, IL, USA) using one-way ANOVA and Student t-test. The Student t-test was

245

used for the two-group comparisons, while the ANOVA with a Tukey HSD post-hoc

246

test was applied to multi-group comparisons.36 All data were expressed as the mean ±

247

standard error (SE) of the three independent experiments and were considered

248

statistically significant when the P value was less than 0.05 (*) and 0.01 (**).

249

RESULTS

250

CPF Induced Morphological Changes and Cytotoxicity

251

A dose responsive effect, with a CPF concentration that ranged from 0-100 µM, was

252

tested on GC-1, TM4, and TM3 cells at different treatment times to evaluate the toxicity

ACS Paragon Plus Environment

Page 10 of 47

Page 11 of 47

Journal of Agricultural and Food Chemistry

253

of CPF. As CPF is a lipophilic molecule, the combination of CPF and serum proteins

254

may compromise CPF activity, therefore, the cells were transferred to a serum-free

255

medium when CPF treatment.22 In order to avoid the negative effects of long-term

256

serum starvation during CPF treatment, the GC-1 cells were exposed for 24 h, while

257

TM4 and TM3 cells were exposed for 12 h. Cell viability decreased with raised

258

concentrations and prolonged treatment of CPF. Since 100 µM CPF resulted in a

259

decrease in cell viability greater than 50% for GC-1 and TM4 cells (data not shown),

260

50 µM and lower concentrations were chosen for further assayed in the subsequent

261

experiments.

262

As shown in Figure 1, a low CPF concentration (10 µM) had no obvious effect on

263

cell morphology. Nevertheless, the shape and density of the cells changed with raised

264

CPF concentration: lower cell density, cell shrinkage, round cells and a shedding

265

morphology.

266

Cell viability was also assessed in this study using a CCK-8 assay. The results

267

from the CCK-8 assay showed that the 10 and 25 µM concentration of CPF significantly

268

decreased the viability of the GC-1 and TM4 cells, and this response was still present

269

at higher concentrations (Figure 2A,B). The viability of TM3 cells were slightly

270

affected by CPF (Figure 2C).

271

To further assess the CPF-induced cytotoxicity in the three cell types, the total cell

272

number was counted. Consistent with the cell viability results, the CPF (0-50 µM)

273

concentration dependently decreased the total cell number of the GC-1 and TM4 cells

274

(Figure 2A´,B´). Unlike the moderate decrease in cell viability, the total number of TM3

275

cells was statistically significant decreased with the raised concentration of CPF (Figure

276

2C´).

277

CPF Promoted Cell Apoptosis

278

First of all, the TUNEL staining was used to assess whether cell apoptosis caused

279

morphological changes of CPF treated cells. The principle of Terminal

280

deoxynucleotidyl transferase (TdT) dUTP nick end labeling (TUNEL) is to attach dUTP

281

to the 3’ ends of double- and single-stranded DNA breaks using TdT enzyme in cells.37

282

The 25 µM CPF concentration dramatically increased the number of TUNEL positive

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

283

cells in GC-1 cell line, and this response was still present at higher concentrations (50

284

µM) (Figure 3A). For TM4 cells, the number of TUNEL positive cells were increased

285

obviously when exposed to 50 µM CPF, compared with control group, 10 and 25 µM

286

CPF group (Figure 3B). Nevertheless, no significant changes were observed in TM3

287

cells when exposed to different concentrations of CPF (Figure 3C). The above

288

experimental results hinted that the sensitivity of these three cell lines to CPF was

289

different. GC-1 cells were the most sensitive cell line, followed by TM4 cells, while

290

TM3 cells were the least sensitive.

291

Subsequently, cell apoptosis was further monitored with an Annexin V/PI stain

292

followed by flow cytometry analysis in the tested cells. The translocation of

293

phospholipid phosphatidylserine to the outer plasma membrane was regarded as the

294

early feature of apoptosis.25 In GC-1 cells, the percentage of early apoptosis cells

295

(Annexin V+/PI−) that was induced by the different concentrations (0, 10, 25, and 50

296

µM) of CPF were 2.635%, 2.770%, 2.255%, and 2.880%, respectively (Figure 4A,B;

297

Table 2). There was an upward trend in the percentage of late apoptosis (Annexin

298

V+/PI+) when cells were exposed to CPF (0 µM: 10.200%, 10 µM: 13.450%, 25 µM:

299

22.400% and 50 µM: 32.400%) (Figure 4A,B; Table 2). Unlike GC-1 cells, no obvious

300

changes in the percentage of late apoptosis (Annexin V+/PI+) were observed in TM4

301

cells, while a very significant increase in the percentage of early apoptosis (Annexin

302

V+/PI−) was detected when TM4 cells exposed to CPF (0 µM: 18.867%, 10 µM:

303

18.800%, 25 µM: 29.200% and 50 µM: 29.750%) (Figure 4C,D; Table 2). For the TM3

304

cells, the percentage of early apoptosis (Annexin V+/PI−) that was induced by the

305

different concentrations (0, 10, 25, and 50 µM) of CPF were 3.715%, 3.410%, 7.005%,

306

and 10.900%, respectively (Figure 4E,F; Table 2). A very significant increasing trend

307

in the percentage of late apoptosis (Annexin V+/PI+) was detected when TM3 cells

308

exposed to CPF (0 µM: 1.625%, 10 µM: 1.075%, 25 µM: 2.950% and 50 µM: 8.620%)

309

(Figure 4E,F; Table 2).

310

Next, the expressions of apoptosis-related genes were also detected in GC-1 and

311

TM4 cells. The results demonstrated that the expression levels of the cell apoptosis

312

genes (p53, Puma, Caspase 3 and Caspase 9) had at least a 2‑fold increase in both cell

ACS Paragon Plus Environment

Page 12 of 47

Page 13 of 47

Journal of Agricultural and Food Chemistry

313

lines when exposed to high concentrations of CPF, compared to that of control groups

314

(Figure 5A,6A).

315

As shown in Figure 5 and Figure 6, verification at the protein level further proved

316

that cell apoptosis was significantly induced by CPF in GC-1 and TM4 cells. The 10

317

and 25 µM concentration of CPF led to a decrease in BCL2 expression, while there was

318

noticeable change in BAX in GC-1 cells (Figure 5C). Importantly, the ratio of

319

BAX/BCL2 was up-regulated, and the expression of Caspase 9 increased when GC-1

320

cells were exposed to CPF (Figure 5B,C). Similarly, the ratio of BAX/BCL2 increased

321

in TM4 cells. In contrast to GC-1 cells, the expression of BCL2 was not influenced, but

322

there was up-regulation of BAX in the TM4 cells (Figure 6B).

323

CPF Induced Cell Cycle Arrest

324

An additional study was carried out to explore the effects of CPF on cell cycle in the

325

three cell lines. To evaluate this action, cell cycle was assessed by PI staining and

326

measured by flow cytometry. For GC-1 cells, the percentage of cells in the G1 phase

327

was dose-dependent and increased with an increase in the CPF treatment concentration

328

(Figure 7A,B). The proportion of GC-1 cells in the G1 phase ranged from 43.447% to

329

71.053% in the 10 µM to 50 µM CPF concentration treatments, respectively, and the

330

control group was 33.247% (Table 3). Additionally, the increased percentage of cells in

331

the G1 phase was accompanied by a decreased percentage of cells in both S and G2

332

phases (Figure 7A,B). For TM4 cells, CPF induced S cell cycle arrest with a

333

concentration-dependency (Figure 7C,D). Exposure to CPF dramatically shuffled the

334

cells in the cell cycle compartments as evidenced by a considerable accumulation of

335

cells in S phase (0 µM: 19.807%, 10 µM: 19.233%, 25 µM: 44.880% and 50 µM:

336

52.237%), and a remarkable reduction of cells in G1 phase (0 µM: 54.013%, 10 µM:

337

56.723%, 25 µM: 44.343% and 50 µM: 40.217%) and G2 phase (0 µM: 26.180%, 10

338

µM: 24.047%, 25 µM: 10.773% and 50 µM: 7.550%) (Table 3). For TM3 cells, consist

339

with the results of TM4 cells, a significant increment of cells in S phase was observed

340

at the 50 µM CPF treatment (35.595%) (Table 3), compared with control group

341

(11.460%). In addition, CPF resulted in a decreasing proportion of G1 phase cells,

342

though no statistically significant differences were detected at G2 phase (Figure 7E,F).

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

343

In light of the above observations, the expression of the cell cycle regulators was

344

examined in the GC-1, TM4 and TM3 cells with or without the CPF treatment. Cyclin

345

is a family of protein that controls the progression of cells through cell cycle.38 In the

346

three cells, expression levels of p53 and cyclin-dependent kinase inhibitor p21CIP, a

347

product of a p53-activated gene, increased when cells were exposed to CPF (Figure 8).

348

39,40

349

and CCNE1 (encoding Cyclin E1) was observed (Figure 8A). Nevertheless, the

350

expression of CCNA2 (encoding Cyclin A2) and CCNB1 (encoding Cyclin B1) was

351

decreased in CPF treated TM4 and TM3 cells (Figure 8B,C). These data suggest that

352

CPF suppressed cell proliferation via cell cycle arrest.

353

CPF Up-regulated ROS Level and Promoted AMPK Phosphorylation

354

To elucidate the underlying mechanism of cell apoptosis induced by CPF, the

355

phosphorylation of AMP-activated protein kinase (AMPK) was analyzed. The presence

356

of p-AMPK was evaluated by a western blot analysis of specific antibodies in GC-1

357

and TM4 cells exposed to CPF at different concentrations for 12 or 24 h. As shown in

358

Figure 9A, the 25 and 50 µM concentrations of CPF induced a significant increment of

359

AMPK phosphorylation in GC-1 cells compared to that of the control group, whereas

360

there was no change in the 10 µM CPF concentration group.

For GC-1 cells, a decrease in expression levels of CCND1 (encoding Cyclin D1)

361

Many environmental compounds could modify the oxidative balance, which led

362

to the inhibition of cell proliferation and induce apoptosis,41 therefore, we also detected

363

whether CPF could affect the redox balance in GC-1 cells. The ROS generation during

364

the GC-1 cells were treated with different concentrations of CPF were studied, and a

365

significant increase in cells treat with CPF was also noticed, compared to the control

366

group (Figure 9D,E,F). As the reduction of MMP is often associated with apoptosis,42

367

MMP during the CPF treatment was evaluated. The results revealed that the addition of

368

CPF could reduce MMP in GC-1 cells (Figure 9B). Moreover, elevated MDA level was

369

associated with an increase in the concentration of CPF (Figure 9C). These data

370

suggested that ROS production was involved in the regulation of CPF-induced cell

371

apoptosis.

372

To prove whether AMPK and ROS regulated the survival of TM4 cells after the

ACS Paragon Plus Environment

Page 14 of 47

Page 15 of 47

Journal of Agricultural and Food Chemistry

373

CPF treatment, p-AMPK and ROS levels in TM4 cells were detected. Consistent with

374

the results of GC-1 cells, the addition of 25 µM CPF concentration resulted in up-

375

regulation of p-AMPK (Figure 10A,B), and ROS levels also increased in TM4 cells

376

compared to that of the control (Figure 10C).

377

Intraperitoneal Injection of CPF reduced the Cell Number of Testicular

378

Seminiferous Tubules

379

There were no significant differences in the organ coefficients, and morphologic

380

changes were observed in the testis for the 3 and 6 mg/kg CPF concentration injection

381

group and control group at 2 weeks and 35 days (Figure S1; Table S1,S2). Interestingly,

382

one mouse died after the intraperitoneal injection at 18 days in the 12 mg/kg group, the

383

neural reflex of the other mice in that group was slow, and the number of cells in the

384

testicular seminiferous were also decreased compared to that of the control, the

385

percentage of damaged seminiferous tubules was 28.54% (Figure 11A,B). The qRT-

386

PCR results certified that the expression of the germ, Sertoli, Leydig and apoptosis

387

related genes were altered in the 12 mg/kg CPF concentration injection group (Figure

388

11C). These data suggested that acute toxicity accumulation of CPF may damage male

389

reproduction in mice.

390

Dietary Supplementation of CPF Resulted in Expression Changes in Steroid

391

Hormone Bio-synthesis related genes in F0 and F1 Mice

392

In 2006, Meeker et al. analyzed the association of male reproductive hormones with

393

CPF and its metabolic product.17 In this study, the expression of steroid hormone bio-

394

synthesis related genes, such as StAR, HSD3B1 and HSD17B3, was detected. The

395

results showed that the expression of these three genes was markedly reduced in the F0

396

generation after fed CPF for 80 days (Figure 12A). Additionally, the expression of genes

397

that is associated with steroid hormone synthesis, the development of germ cells, Sertoli

398

cells and Leydig cells, cell proliferation and apoptosis, showed a downward trend in

399

the 12 mg/kg CPF concentration treatment group compared to that of the control group

400

(Figure 12B).

401

Figure 12C provides an overview of the effects of dietary CPF on the other organs

402

in mice. Overall, the expression of StAR, HSD3B1 and HSD17B3 decreased in the testis,

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

403

spleen, and lung in the F0 generation after the 12 mg/kg CPF concentration group when

404

fed for 80 days compared to that of the control group. In the kidney, the expression of

405

HSD3B1 and HSD17B3 showed a downward trend in the CPF treatment group

406

compared to that of the control group, but, interestingly, StAR expression levels were

407

significantly elevated in the CPF-treated kidney samples compared to that of the control

408

group.

409

F1 generation samples were collected at the time of weaning (3 and 4 w). There

410

were no significant differences in the body and tissue weights following the 1 and 6

411

mg/kg CPF exposure for any of the mice (Table S3). Nevertheless, in three-week-old

412

F1 mice, with an increase in CPF concentration, the expression of HSD3B1 and

413

HSD17B3 showed a downward trend (Figure 13A). However, no significant differences

414

were observed in four-week-old F1 mice (Figure 13B).

415

Additionally, no significant differences in morphologic changes were observed

416

using HE staining in neither F0 mice nor F1 mice testis tissue (data not shown).

417

DISCUSSION

418

Chlorpyrifos (CPF) is extensively used for various purposes. The widespread use of

419

CPF has stimulated research on the possible existence of effects related to reproductive

420

toxic activity.4,43,44 However, male reproductive toxicity mechanisms induced by CPF

421

has not been thoroughly studied. In this study, the toxic effects of CPF in vitro and in

422

vivo were detected.

423

Various stressors, including pesticides, are existed in the environment and are able

424

to cause DNA damage.45 Researchers demonstrated that the percentages of sperms with

425

DNA damage in CPF-exposed animals were significantly higher than control group.1

426

This phenomenon was also proved by using CPF and cypermethrin combined rat as

427

model. Significant increments in sperm DNA fragmentation index were manifested in

428

exposure group.46 In addition, CPF-induced DNA damage and apoptosis were observed

429

in larval Drosophila midgut tissues, which was proved by a markedly increase in the

430

Comet parameters, viz, tail length (mm), TM (arbitrary units) and tail DNA (%) of the

431

exposed individuals.43 In 2015, Li et al. indicated that the addition of CPF in the culture

ACS Paragon Plus Environment

Page 16 of 47

Page 17 of 47

Journal of Agricultural and Food Chemistry

432

medium induced a dramatically concentration- and time-dependent augment in HeLa

433

and HEK293 cell apoptosis, which was also accompanied by single-strand DNA breaks

434

in CPF treated cells compared to that of the control.47 Normally, p53 plays a pivotal

435

role in cell cycle regulation and the induction of apoptosis when mammalian cells are

436

subjected to stress conditions, such as hypoxia, radiation, chemotherapeutic drugs, or

437

DNA damage.48,49 In the process of the cell cycle, the Cyclin-dependent kinase inhibitor

438

p21CIP (a downstream gene of p53) could be combined with a series of Cyclin-cdk

439

complexes to inhibit the activity of protein kinase, thereby arresting the cell cycle.39,40

440

Additionally, the p53 tumor suppressor could also mediate apoptosis through Bax

441

transactivation, the release of mitochondrial cytochrome c, and caspase-9 activation,

442

which is usually followed by the activation of caspase-3, -6, and -7.50

443

In this study, the addition of CPF to the culture medium led to a decrease in cell

444

number and viability in GC-1, TM4 and TM3 cells. The expression of p53 and p21CIP

445

increased when all testicular cells were exposed to CPF. Moreover, the expression

446

levels of cell apoptosis genes (p53, Puma, Caspase 3 and Caspase 9) increased when

447

GC-1 and TM4 cells exposed to high concentrations of CPF, compared to that of with

448

the control groups. Afterwards, the effect of the CPF on cell cycle was studied. The

449

results showed that CPF could induce G1 phase, S phase and S phase arrest in GC-1,

450

TM4 and TM3 cells, respectively. In addition, consistent with the results of flow

451

cytometry, a decrease in the expression levels of CCND1 (encoding Cyclin D1) and

452

CCNE1 (encoding Cyclin E1) in GC-1 cells, as well as a decrease in the expression

453

levels of CCNA2 (encoding Cyclin A2) and CCNB1 (encoding Cyclin B1) in both TM4

454

and TM3 cells were also observed. The differences of the results were related to the

455

different cell types. These data indicated that CPF may induce DNA damage first by

456

arresting the cell cycle, suppressing cell proliferation, and promoting cell apoptosis,

457

until, finally, there is a reduction in the number of GC-1, TM4, and TM3 cells. Future

458

studies should investigate the DNA damage induced by CPF in these cells.

459

Although no information about CPF induced oxidative stress on testicular cells

460

were reported, the in vivo experiments were extensive. The activation of lipid

461

peroxidation and the decrease of antioxidant enzymes activities were observed in CPF

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

462

treated rats.20 Association between reproductive toxicity and antioxidant enzymes

463

catalase activities, superoxide dismutase (SOD), glutathione peroxidase (GPx), and

464

glutathione S-transferase (GST) as well as glutathione (GSH) level exposed to CPF

465

compounds has been demonstrated in rats.51 Consistent with this, an obvious positive

466

correlation between ROS generation and CPF addition were detected in this study.52,53

467

To further examine the causes of CPF-induced oxidative stress, MMP was measured.

468

The data indicated that CPF mediated cell apoptosis was associated with a reduction in

469

MMP. Lipid peroxidation is a free radical-driven reaction that leads to membrane

470

damage by a reaction of oxygen with polyunsaturated fatty acids.54 Previous studies

471

had also indicated that the increment of MDA content was correlated to cellular

472

membrane damage in exposed cells, which could then derivate cells to death.55 An

473

augment of MDA was detected when GC-1 cells were exposed to CPF at a dose capable

474

of increasing the ROS generation.

475

The aforementioned results have shown that the addition of CPF could promote

476

cell apoptosis. The mechanism involved in CPF-induced cell apoptosis was also

477

examined in this study. Numerous studies have revealed that activation of MAPK and

478

EGFR/ERK1/2 are involved in CPF-induced apoptosis. CPF could induce human

479

neuroblastoma SH-SY5Y cell apoptosis through the MAPK pathway.55 A recent study

480

proved that CPF inhibited breast cancer cells MCF-7 and MDA-MB-231 proliferation

481

via an incremental phosphorylation of p-ERK1/2 levels that were mediated by oxidative

482

stress.5 In contrast, the function of AMP-activated protein kinase (AMPK) in CPF

483

induced toxic reactions has not been reported. AMPK is a crucial enzyme protein that

484

links energy induction with metabolic reactions, and it is also involved in regulating the

485

processes of glycometabolism, fat metabolism, protein metabolism, and maintaining

486

the homeostasis of energy in cells.56 There are increasing indications that some types of

487

cellular stress activate AMPK by ROS, such as H2O2.56 In cultured cells, AMPK is

488

activated by reactive oxygen species (ROS) such as H2O2. In 2008, research indicated

489

that Britannin (Bri) could induce apoptosis in liver cancer cell lines via AMPK

490

activation that was regulated by ROS.57 Moreover, Chen et al. proved that a hypoxia

491

treatment triggered AMPK activation in H9c2 cells in a time dependent manner and led

ACS Paragon Plus Environment

Page 18 of 47

Page 19 of 47

Journal of Agricultural and Food Chemistry

492

to cardiomyocyte injury.58 Consistent with this, AMPK was activated when testicular

493

cells were exposed to CPF in this study. Although this study confirmed the role of

494

AMPK in CPF-induced reproductive toxicity, the upstream and downstream signaling

495

pathways of AMPK are unknown and needs to be studied further.

496

CPF inhibits AchE activity and reduces monoamine levels that are needed for

497

adequate hypothalamic-pituitary-gonadal axis (HPGA) activity, which in turn directs

498

the toxicity of male hormones, induces sperm DNA damage, or causes sperm epigenetic

499

changes.4,17 To this point, the expression of steroid hormone bio-synthesis related genes

500

was detected in this study between the CPF treatment group and the control group. The

501

data from the in vivo experiments initially demonstrated that the expression of steroid

502

hormone bio-synthesis related genes in the testis, spleen, and lung of F0 mice was

503

downregulated when there was intraperitoneal injection or dietary supplementation of

504

CPF. It is worth noting that the intraperitoneal injection of CPF in mice led to a

505

significantly increasing trends in the expression of StAR in the CPF treatment group,

506

compared to that of control group. Nevertheless, the exact mechanism needs to be

507

studied further.

508

For F1 generations, the expression of HSD3B1 and HSD17B3 showed a downward

509

trend with an increase in CPF concentration of three-week-old mice, while no

510

significant differences were observed in four-week-old F1 mice. Considering the

511

expression of these two genes in testis was related to the development of Leydig cells,

512

which showed an increase tendency during the transition from progenitor Leydig cells

513

(PLCs) (around postnatal day 14) to immature Leydig cell (ILCs) (around postnatal day

514

28).59,60 It is well known that the cells with lower differentiation degree were more

515

sensitive to drugs. Back to our study, compared to ILCs, PLCs were more sensitive to

516

CPF. Thus, the number of PLCs decreased in three-week-old mice, resulted in the

517

decrease expression of these two genes.

518

In conclusion, we propose a model for the regulation of CPF for the survival of

519

testicular cells (Figure 14). The addition of CPF in the culture medium resulted in a

520

reduction in cell number and cell viability. CPF decreased the cellular membrane

521

potential of mitochondria, up-regulated the levels of ROS and MDA, activated the

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

522

phosphorylation of AMPK, and thereby, arrested the cell cycle, suppressed cell

523

proliferation, promoted cell apoptosis, and regulated the survival of testicular cells. In

524

vivo, the expression of steroid hormone bio-synthesis related genes in testis, spleen and

525

lungs of the F0 and F1 mice was downregulated when there was an intraperitoneal

526

injection or dietary supplementation of CPF. This study replenishes the research on the

527

regulation of CPF in male reproduction and provides a theoretical basis for the

528

treatment of male infertility in the future.

ACS Paragon Plus Environment

Page 20 of 47

Page 21 of 47

Journal of Agricultural and Food Chemistry

529

AUTHOR INFORMATION

530

Corresponding Authors

531

*E-mail: [email protected].

532

Funding

533

This work was supported by the National Natural Science Fund (regional project)

534

(No.31760650), The Key Research and Development Program of Shaanxi Province

535

(agricultural field) (No.2017NY-064).

536

Notes

537

The authors declare no competing financial interest.

Tel: 86-13772098751

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

538

ABBREVIATIONS USED

539

CPF, chlorpyrifos; OP, organophosphorous pesticide; AChE, acetylcholinesterase; GC-

540

1, mouse-derived spermatogonial cell lines; TM4, mouse Sertoli cell lines; TM3, mouse

541

Leydig cell lines; DMSO, dimethyl sulfoxide; qRT-PCR, quantitative RT-PCR; ROS,

542

reactive oxygen species; MDA, malonaldehyde; SOD, superoxide dismutase; GPx,

543

glutathione peroxidase; p-AMPK, phosphorylated AMP-activated protein kinase;

544

DCFH-DA, dichlorodihydrofluorescein diacetate; DCF, dichlorofluorescein; T,

545

testosterone; FSH, follicle-stimulating hormone; LH, luteinizing hormone; GSH,

546

glutathione; SHBG, sex hormone-binding globulin; ILCs, immature Leydig cell; PLCs,

547

progenitor Leydig cells; HPGA, hypothalamic-pituitary-gonadal axis; TUNEL,

548

terminal deoxynucleotidyl transferase (TdT) dUTP nick end labeling.

ACS Paragon Plus Environment

Page 22 of 47

Page 23 of 47

Journal of Agricultural and Food Chemistry

549

REFERENCES

550

(1) Babazadeh, M.; Najafi, G. Effect of chlorpyrifos on sperm characteristics and

551

testicular tissue changes in adult male rats. Vet. Res. Forum. 2017, 8, 319-26.

552

(2) Levine, H.; Jørgensen, N.; Martino-Andrade, A.; Mendiola, J.; Weksler-Derri, D.;

553

Mindlis, I.; Pinotti, R.; Swan, S. H. Temporal trends in sperm count: a systematic

554

review and meta-regression analysis. Hum. Reprod. Update 2017, 23, 646-59.

555 556 557 558

(3) Jenardhanan, P.; Panneerselvam, M.; Mathur, P. P. Effect of environmental contaminants on spermatogenesis. Semin. Cell Dev. Biol. 2016, 59, 126-40. (4) Mima, M.; Greenwald, D.; Ohlander, S. Environmental toxins and male fertility. Curr. Urol. Rep. 2018, 19, 50.

559

(5) Ventura, C.; Venturino, A.; Miret, N.; Randi, A.; Rivera, E.; Núñez, M.; Cocca, C.

560

Chlorpyrifos inhibits cell proliferation through ERK1/2 phosphorylation in breast

561

cancer cell lines. Chemosphere 2015, 120, 343-50.

562

(6) Pallotta, M. M.; Ronca, R.; Carotenuto, R.; Porreca, I.; Turano, M.; Ambrosino, C.;

563

Capriglione, T. Specific effects of chronic dietary exposure to chlorpyrifos on brain

564

gene expression-a mouse study. Int. J. Mol. Sci. 2017, 18, doi: 10.3390/ijms18112467.

565

(7) Mileson, B. E.; Chambers, J. E.; Chen, W. L.; Dettbarn, W.; Ehrich, M.; Eldefrawi,

566

A. T.; Gaylor, D. W.; Hamernik, K.; Hodgson, E.; Karczmar, A. G.; Padilla, S.; Pope,

567

C. N.; Richardson, R. J.; Saunders, D. R.; Sheets, L. P.; Sultatos, L. G.; Wallace, K.

568

B. Common mechanism of toxicity: a case study of organophosphorus pesticides.

569

Toxicol. Sci. 1998,1, 8-20.

570

(8) Parran, D. K.; Magnin, G.; Li, W.; Jortner, B. S.; Ehrich, M. Chlorpyrifos alters

571

functional integrity and structure of an in vitro BBB model: co-cultures of bovine

572

endothelial cells and neonatal rat astrocytes. Neurotoxicology 2005, 26, 77-88.

573

(9) Goel, A.; Dani, V.; Dhawan, D. K. Chlorpyrifos-induced alterations in the activities

574

of carbohydrate metabolizing enzymes in rat liver: the role of zinc. Toxicol. Lett. 2006,

575

163, 235-41.

576

(10) Howard, A. S.; Bucelli, R.; Jett, D. A.; Bruun, D.; Yang, D.; Lein, P. J. Chlorpyrifos

577

exerts opposing effects on axonal and dendritic growth in primary neuronal cultures.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

578

Toxicol. Appl. Pharmacol. 2005, 207, 112-24.

579

(11) Tian, Y.; Ishikawa, H.; Yamaguchi, T.; Yamauchi, T.; Yokoyama, K. Teratogenicity

580

and developmental toxicity of chlorpyrifos. Maternal exposure during organogenesis

581

in mice. Reprod. Toxicol. 2005, 20, 267-70.

582

(12) Rahman, M. F.; Mahboob, M.; Danadevi, K.; Saleha Banu, B.; Grover, P.

583

Assessment of genotoxic effects of chloropyriphos and acephate by the comet assay

584

in mice leucocytes. Mutat. Res. 2002, 516, 139-47.

585

(13) Schuh, R. A.; Lein, P. J.; Beckles, R. A.; Jett, D. A. Noncholinesterase mechanisms

586

of chlorpyrifos neurotoxicity: altered phosphorylation of Ca2+/cAMP response

587

element binding protein in cultured neurons. Toxicol. Appl. Pharmacol. 2002, 182,

588

176-85.

589 590

(14) Dolci, S.; Campolo, F.; De Felici, M. Gonadal development and germ cell tumors in mouse and humans. Semin. Cell Dev. Biol. 2015, 45, 114-23.

591

(15) Farag, A. T.; Radwan, A. H.; Sorour, F.; El, Okazy. A.; El-Agamy, el-S.; El-Sebae,

592

Ael-K. Chlorpyrifos induced reproductive toxicity in male mice. Reprod. Toxicol.

593

2010, 29, 80-5.

594

(16) Sai, L.; Li, X.; Liu, Y.; Guo, Q.; Xie, L.; Yu, G.; Bo, C.; Zhang, Z.; Li, L. Effects

595

of chlorpyrifos on reproductive toxicology of male rats. Environ. Toxicol. 2014, 29,

596

1083-8.

597 598

(17) Meeker, J. D.; Ryan, L.; Barr, D. B.; Hauser, R. Exposure to nonpersistent insecticides and male reproductive hormones. Epidemiology 2006, 17, 61-8.

599

(18) Mosbah, R.; Yousef, M. I.; Maranghi, F.; Mantovani, A. Protective role of Nigella

600

sativa oil against reproductive toxicity, hormonal alterations, and oxidative damage

601

induced by chlorpyrifos in male rats. Toxicol. Ind. Health. 2016, 32, 1266-77.

602

(19) Peiris, D. C.; Dhanushka, T. Low doses of chlorpyrifos interfere with

603

spermatogenesis of rats through reduction of sex hormones. Environ. Sci. Pollut. Res.

604

Int. 2017, 24, 20859-67.

605

(20) Adedara, I, A.; Owoeye, O.; Ajayi, B. O.; Awogbindin, I. O.; Rocha, J. B. T.;

606

Farombi, E. O. Diphenyl diselenide abrogates chlorpyrifos-induced hypothalamic-

607

pituitary-testicular axis impairment in rats. Biochem. Biophys. Res. Commun. 2017,

ACS Paragon Plus Environment

Page 24 of 47

Page 25 of 47

608

Journal of Agricultural and Food Chemistry

503, 171-176.

609

(21) Venerosi, A.; Tait, S.; Stecca, L.; Chiarotti, F.; De Felice, A.; Cometa, M. F6.; Volpe,

610

M. T.; Calamandrei, G.; Ricceri, L. Effects of maternal chlorpyrifos diet on social

611

investigation and brain neuroendocrine markers in the offspring - a mouse study.

612

Environ. Health. 2015, 14, 32.

613

(22) Lee, J. E.; Park, J. H.; Shin, I. C.; Koh, H. C. Reactive oxygen species regulated

614

mitochondria-mediated apoptosis in PC12 cells exposed to chlorpyrifos. Toxicol. Appl.

615

Pharmacol. 2012, 263, 148-62.

616

(23) Zakharevich, M.; Kattan, J. M.; Chen, J. L.; Lin, B. R.; Cervantes, A. E.; Chung,

617

D. D.; Frausto, R. F.; Aldave, A. J. Elucidating the molecular basis of PPCD: Effects

618

of decreased ZEB1 expression on corneal endothelial cell function. Mol. Vis. 2017,

619

23, 740-52.

620

(24) Yang, Z. R.; Liu, M.; Peng, X. L.; Lei, X. F.; Zhang, J. X.; Dong, W. G. Noscapine

621

induces mitochondria-mediated apoptosis in human colon cancer cells in vivo and in

622

vitro. Biochem. Biophys. Res. Commun. 2012, 421, 627-33.

623

(25) Liu, T.; Chen, X.; Li, T.; Li, X.; Lyu, Y.; Fan, X.; Zhang, P.; Zeng, W. Histone

624

methyltransferase

SETDB1

maintains

survival

of

mouse

spermatogonial

625

stem/progenitor cells via PTEN/AKT/FOXO1 pathway. Biochim. Biophys. Acta. 2017,

626

1860, 1094-102.

627

(26) Chen, R.; Du, J.; Ma, L.; Wang, L. Q.; Xie, S. S.; Yang, C. M.; Lan, X. Y.; Pan, C.

628

Y.; Dong, W. Z. Comparative microRNAome analysis of the testis and ovary of the

629

Chinese giant salamander. Reproduction 2017, 154, 169-79.

630

(27) Zhu, Z.; Ren, Z.; Fan, X.; Pan, Y.; Lv, S.; Pan, C.; Lei, A.; Zeng, W. Cysteine

631

protects rabbit spermatozoa against reactive oxygen species-induced damages. PLoS

632

One 2017, 12, e0181110.

633

(28) Zhan, Y.; Gong, K.; Chen, C.; Wang, H.; Li, W. P38 MAP kinase functions as a

634

switch in MS-275-induced reactive oxygen species-dependent autophagy and

635

apoptosis in human colon cancer cells. Free Radic. Biol. Med. 2012, 53, 532-43.

636 637

(29) Xin, H.; Liu, X. H.; Zhu, Y. Z. Herba leonurine attenuates doxorubicin-induced apoptosis in H9c2 cardiac muscle cells. Eur. J. Pharmacol. 2009, 612, 75-9.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

638 639

(30) Joshi, S. C, Mathur, R.; Gulati, N. Testicular toxicity of chlorpyrifos (an organophosphate pesticide) in albino rat. Toxicol. Ind. Health 2007, 23, 439-44.

640

(31) Sai, L.; Li, X.; Liu, Y.; Guo, Q.; Xie, L.; Yu, G.; Bo, C.; Zhang, Z.; Li, L. Effects

641

of chlorpyrifos on reproductive toxicology of male rats. Environ. Toxicol. 2014, 29,

642

1083-8.

643

(32) Phung, D. T.; Connell, D.; Miller, G.; Chu, C. Probabilistic assessment of

644

chlorpyrifos exposure to rice farmers in Viet Nam. J. Expo. Sci. Environ. Epidemiol.

645

2012, 22, 417-23.

646

(33) Salazar-Arredondo, E.; de Jesús Solís-Heredia, M.; Rojas-García, E.; Hernández-

647

Ochoa, I.; Quintanilla-Vega, B. Sperm chromatin alteration and DNA damage by

648

methyl-parathion, chlorpyrifos and diazinon and their oxon metabolites in human

649

spermatozoa. Reprod. Toxicol. 2008, 25, 455-60.

650

(34) Venerosi, A.; Tait, S.; Stecca, L.; Chiarotti, F.; De Felice, A.; Cometa, M. F.; Volpe,

651

M. T.; Calamandrei, G.; Ricceri, L. Effects of maternal chlorpyrifos diet on social

652

investigation and brain neuroendocrine markers in the offspring - a mouse study.

653

Environ. Health 2015, 14, 32.

654

(35) Heidari, B.; Rahmati-Ahmadabadi, M.; Akhondi, M. M.; Zarnani, A. H.; Jeddi-

655

Tehrani, M.; Shirazi, A.; Naderi, M. M.; Behzadi, B. Isolation, identification, and

656

culture of goat spermatogonial stem cells using c-kit and PGP9.5 markers. J. Assist.

657

Reprod. Genet. 2012, 29, 1029-38.

658

(36) Cui, Y.; Yan, H.; Wang, K.; Xu, H.; Zhang, X.; Zhu, H.; Liu, J.; Qu, L.; Lan, X.;

659

Pan, C. Insertion/deletion within the KDM6A gene is significantly associated with

660

litter size in goat. Front. Genet. 2018, 9, 91.

661 662 663 664

(37) Fayzullina, S.; Martin, L. J. Detection and analysis of DNA damage in mouse skeletal muscle in situ using the TUNEL method. J. Vis. Exp. 2014, 16, 94. (38) Galderisi, U.; Jori, F. P.; Giordano, A. Cell cycle regulation and neural differentiation. Oncogene 2003, 22, 5208-19.

665

(39) Imamura, K.; Ogura, T.; Kishimoto, A.; Kaminishi, M.; Esumi, H. Cell cycle

666

regulation via p53 phosphorylation by a 5'-AMP activated protein kinase activator, 5-

667

aminoimidazole-4-carboxamide-1-beta-D-ribofuranoside, in a human hepatocellular

ACS Paragon Plus Environment

Page 26 of 47

Page 27 of 47

668

Journal of Agricultural and Food Chemistry

carcinoma cell line. Biochem. Biophys. Res. Commun. 2001, 287, 562-7.

669

(40) Jones, R. G.; Plas, D. R.; Kubek, S.; Buzzai, M.; Mu, J.; Xu, Y.; Birnbaum, M. J.;

670

Thompson, C. B. AMP-activated protein kinase induces a p53-dependent metabolic

671

checkpoint. Mol. Cell 2005, 18, 283-93.

672

(41) Itziou, A.; Kaloyianni, M.; Dimitriadis, V. K. Effects of organic contaminants in

673

reactive oxygen species, protein carbonylation and DNA damage on digestive gland

674

and haemolymph of land snails. Chemosphere 2011, 85, 1101-7.

675

(42) Hung, J. H.; Chen, C. Y.; Omar, H. A.; Huang, K. Y.; Tsao, C. C.; Chiu, C. C.; Chen,

676

Y. L.; Chen, P. H.; Teng, Y. N. Reactive oxygen species mediate Terbufos-induced

677

apoptosis in mouse testicular cell lines via the modulation of cell cycle and pro-

678

apoptotic proteins. Environ. Toxicol. 2016, 31, 1888-98.

679

(43) Gupta, S. C.; Mishra, M.; Sharma, A.; Deepak Balaji, T. G.; Kumar, R.; Mishra, R.

680

K.; Chowdhuri, D. K. Chlorpyrifos induces apoptosis and DNA damage in Drosophila

681

through generation of reactive oxygen species. Ecotoxicol. Environ. Saf. 2010, 73,

682

1415-23.

683

(44) Eaton, D. L.; Daroff, R. B.; Autrup, H.; Bridges, J.; Buffler, P.; Costa, L. G.; Coyle,

684

J.; McKhann, G.; Mobley, W. C.; Nadel, L.; Neubert, D.; Schulte-Hermann, R.;

685

Spencer, P. S. Review of the toxicology of chlorpyrifos with an emphasis on human

686

exposure and neurodevelopment. Crit. Rev. Toxicol. 2008, 38, Suppl 2:1-125.

687 688

(45) Wilson, D. M. 3rd.; Sofinowski, T. M.; McNeill, D. R. Repair mechanisms for oxidative DNA damage. Front. Biosci. 2003, 8, d963-81.

689

(46) Alaa-Eldin, E. A.; El-Shafei, D. A.; Abouhashem, N. S. Individual and combined

690

effect of chlorpyrifos and cypermethrin on reproductive system of adult male albino

691

rats. Environ. Sci. Pollut. Res. Int. 2017, 24, 1532-1543.

692

(47) Li, D.; Huang, Q.; Lu, M.; Zhang, L.; Yang, Z.; Zong, M.; Tao, L. The

693

organophosphate insecticide chlorpyrifos confers its genotoxic effects by inducing

694

DNA damage and cell apoptosis. Chemosphere 2015, 135, 387-93.

695 696 697

(48) Wang, X. W.; Harris, C. C. p53 tumor-suppressor gene: clues to molecular carcinogenesis. J. Cell Physiol. 1997, 173, 247-55. (49) Horn, H. F.; Vousden, K. H. Coping with stress: multiple ways to activate p53.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

698

Oncogene 2007, 26, 1306-16.

699

(50) Cheng, W. L.; Lin, T. Y.; Tseng, Y. H.; Chu, F. H.; Chueh, P. J.; Kuo, Y. H.; Wang,

700

S. Y. Inhibitory effect of human breast cancer cell proliferation via p21-mediated G1

701

cell cycle arrest by araliadiol isolated from Aralia cordata Thunb. Planta Med. 2011,

702

77, 164-8.

703

(51) Abolaji, A. O.; Ojo, M.; Afolabi, T. T.; Arowoogun, M. D.; Nwawolor, D.; Farombi,

704

E. O. Protective properties of 6-gingerol-rich fraction from Zingiber officinale

705

(Ginger) on chlorpyrifos-induced oxidative damage and inflammation in the brain,

706

ovary and uterus of rats. Chem. Biol. Interact. 2017, 270, 15-23.

707 708 709 710 711 712

(52) Fleury, C.; Mignotte, B.; Vayssière, J. L. Mitochondrial reactive oxygen species in cell death signaling. Biochimie 2002, 84, 131-41. (53) Loh, K. P.; Huang, S. H.; De Silva, R.; Tan, B. K.; Zhu, Y. Z. Oxidative stress: apoptosis in neuronal injury. Curr. Alzheimer Res. 2006, 3, 327-37. (54) Barrera, G. Oxidative stress and lipid peroxidation products in cancer progression and therapy. ISRN Oncol. 2012, 137289.

713

(55) Ziech, D.; Franco, R.; Georgakilas, A. G.; Georgakila, S.; Malamou-Mitsi, V.;

714

Schoneveld, O.; Pappa, A.; Panayiotidis, M. I. The role of reactive oxygen species

715

and oxidative stress in environmental carcinogenesis and biomarker development.

716

Chem. Biol. Interact. 2010, 188, 334-9.

717 718

(56) Hardie, D. G.; Ross, F. A.; Hawley, S. A. AMPK: a nutrient and energy sensor that maintains energy homeostasis. Nat. Rev. Mol. Cell Biol. 2012, 13, 251-62.

719

(57) Cui, Y. Q.; Liu, Y. J.; Zhang, F. The suppressive effects of Britannin (Bri) on human

720

liver cancer through inducing apoptosis and autophagy via AMPK activation

721

regulated by ROS. Biochem. Biophys. Res. Commun. 2018, 497, 916-23.

722

(58) Chen, X.; Li, X.; Zhang, W.; He, J.; Xu, B.; Lei, B.; Wang, Z.; Cates, C.; Rousselle,

723

T.; Li, J. Activation of AMPK inhibits inflammatory response during hypoxia and

724

reoxygenation through modulating JNK-mediated NF-κB pathway. Metabolism 2018,

725

83, 256-70.

726 727

(59) Oh, Y. S.; Seo, J. T.; Ahn, H. S.; Gye, M. C. Expression of cubilin in mouse testes and Leydig cells. Andrologia 2016, 48, 325-32.

ACS Paragon Plus Environment

Page 28 of 47

Page 29 of 47

728 729

Journal of Agricultural and Food Chemistry

(60) Ye, L.; Li, X.; Li, L.; Chen, H.; Ge, R. S. Insights into the development of the adult Leydig cell lineage from stem Leydig cells. Front Physiol. 2017, 8, 430.

730

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

731

Table 1. Primers Used in qRT-PCR Gene names

Forward primers (from 5’ to 3’)

Reverse primers (from 5’ to 3’)

Caspase 9

CCACTGCCTCATCATCAAC

TGTGCCATCTCCATCAAA

Caspase 3

AGTTCCCGGGTGCTGTCTAT

GCCATGGTCTTTCTGCTCAC

p53

ATGCGGTTCGGGTCCAAAAT

CTAAATGGCAGTCGTTCTCTCC

Puma

AGCAGCACTTAGAGTCGCC

CCTGGGTAAGGGGAGGAGT

CCND1

TGCTGCAAATGGAACTGCTT

CCACAAAGGTCTGTGCATGCT

CCNE1

GTGGCTCCGACCTTTCAGTC

CACAGTCTTGTCAATCTTGGCA

P21CIP

CCTGGTGATGTCCGACCTG

CCATGAGCGCATCGCAATC

CCNB1

AAGGTGCCTGTGTGTGAACC

GTCAGCCCCATCATCTGCG

CCNA2

GCCTTCACCATTCATGTGGAT

TTGCTGCGGGTAAAGAGACAG

Akt

ATGAACGACGTAGCCATTGTG

TTGTAGCCAATAAAGGTGCCAT

PLZF

CTGCGGAAAACGGTTCCTG

GTGCCAGTATGGGTCTGTCT

GDNF

TCCAACTGGGGGTCTACGG

GCCACGACATCCCATAACTTCAT

Stra8

ACAACCTAAGGAAGGCAGTTTAC

GACCTCCTCTAAGCTGTTGGG

HSD3B1

TGGACAAAGTATTCCGACCAGA

GGCACACTTGCTTGAACACAG

HSD17B3

AGGTTCTCGCAGCACCTTTTT

CATCGCCTGCTCCGGTAATC

StAR

GGTTCTCAGCTGGAAGACACT

ACCTCGTCCCCATTCTCCTG

LHR

GCCTCAGCCGACTATCACTC

GGAGGTTGTCAAAGGCATTAGC

ꞵ-Actin

TTGCTGACAGGATGCAGAAG

ACTCCTGCTTGCTGATCCACAT

732

ACS Paragon Plus Environment

Page 30 of 47

Page 31 of 47

Journal of Agricultural and Food Chemistry

733

Table 2. Percent of Cell Apoptosis at Different Periods When GC-1, TM4 and TM3 Cells

734

Exposure to Different Concentrations of CPF Group

Early apoptosis (%)

Late apoptosis (%)

Total apoptosis (%)

DMSO

2.635±0.085b

10.200±0.100d

12.835±0.185c

CPF 10 µM

2.770±0.050b

13.450±0.450c

16.220±0.400c

CPF 25 µM

2.255±0.155ab

22.400±0.600b

24.655±0.755b

CPF 50 µM

2.880±1.010a

32.400±1.400a

35.280±2.410a

DMSO

18.867±1.425b

2.480±0.384b

21.947±1.808b

CPF 10 µM

18.800±0.900b

2.185±0.145b

20.985±1.045b

CPF 25 µM

29.200±1.500a

1.610±0.210b

30.610±1.490a

CPF 50 µM

29.750±0.350a

4.995±0.495a

34.745±0.845a

DMSO

3.715±0.745c

1.625±0.885b

5.340±1.630c

CPF 10 µM

3.410±0.580c

1.075±0.085b

4.485±0.495c

CPF 25 µM

7.005±0.395b

2.950±0.680b

9.955±1.075b

CPF 50 µM

10.900±0.400a

8.620±0.200a

19.520±0.600a

GC-1 cells

TM4 cells

TM3 cells

735

Note: The values with different letters (a, b, c and d) within the same column differ significantly at

736

P < 0.05 or P < 0.01.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

737

Page 32 of 47

Table 3. Cell Cycle Distributions of GC-1, TM4 and TM3 Cells Using CPF Group

G1 (%)

S (%)

G2 (%)

DMSO

33.247±0.964d

46.587±2.222a

20.167±1.380a

CPF 10 µM

43.447±1.191c

41.730±1.168a

14.827±0.815b

CPF 25 µM

55.543±0.987b

35.987±1.698b

8.473±1.042c

CPF 50 µM

71.053±0.919a

23.707±1.669c

5.240±0.780c

DMSO

54.013±0.238a

19.807±0.855c

26.180±0.635a

CPF 10 µM

56.723±0.644a

19.233±0.456c

24.047±0.259a

CPF 25 µM

44.343±0.367b

44.880±0.921b

10.773±0.703b

CPF 50 µM

40.217±0.519c

52.237±0.524a

7.550±0.448c

DMSO

63.500±0.970a

11.460±0.550c

25.040±1.520

CPF 10 µM

64.047±0.158a

12.043±0.358c

23.910±0.219

CPF 25 µM

52.610±0.771b

24.140±0.573b

23.253±0.327

CPF 50 µM

41.265±0.005c

35.595±0.145a

23.140±0.140

GC-1 cells

TM4 cells

TM3 cells

738

Note: The values with different letters (a, b, c and d) within the same column differ significantly

739

at P < 0.05 or P < 0.01.

ACS Paragon Plus Environment

Page 33 of 47

Journal of Agricultural and Food Chemistry

740

741 742

Figure 1. The phenotype of GC-1, TM4 and TM3 cells after CPF exposed for 12 or 24 h. (A) GC-

743

1 cells treated with CPF of different concentrations for 24 h. (B) TM4 and TM3 cells treated with

744

CPF of different concentrations for 12 h. Bar=100 μm.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

745 746

Figure 2. The cell viability and total cell number of GC-1, TM4 and TM3 cells after CPF treated

747

for 12 or 24 h. A and A´, B and B´, C and C´ represent cell viability and total cell number of GC-1,

748

TM4 and TM3 cells, respectively. Note: the values with different letters (a, b, c and d) differ

749

significantly at P < 0.05 or P < 0.01 level. NS means no significant differences.

ACS Paragon Plus Environment

Page 34 of 47

Page 35 of 47

Journal of Agricultural and Food Chemistry

750 751

Figure 3. TUNEL staining of GC-1, TM4 and TM3 cell lines treated with different concentrations

752

of CPF. A, B and C were the results of GC-1, TM4 and TM3 cells, respectively. Red: TUNEL

753

positive cells; Blue: nuclear (counterstaining of DNA).

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

754 755

Figure 4. Flow cytometry was used to detect apoptosis of GC-1, TM4 and TM3 cells after CPF

756

treatment. A and B, C and D, E and F were the results of GC-1, TM4 and TM3 cells, respectively.

757

The data were presented as the percentage of apoptotic cells. Early apoptotic cells were in green

758

square and late apoptotic cells were in red square.

ACS Paragon Plus Environment

Page 36 of 47

Page 37 of 47

Journal of Agricultural and Food Chemistry

759 760

Figure 5. The qPCR and western blot were used to detect the expression of apoptosis-related genes

761

in CPF treated GC-1 cells. (A) The expression of p53, puma, Caspase 3 and Caspase 9 of GC-1

762

after CPF treatment. (B) Expression of cleaved-Caspase 9 protein were detected in GC-1 cells.

763

Intensity analysis of cleaved-Caspase 9 ratio was calculated by Image J. ** P < 0.01 (C) Expression

764

of BAX and BCL2 protein were detected in GC-1 cells after CPF treated for 24 h. Intensity analysis

765

of BAX/BCL2 ratio was calculated by Image J. Note: the values with different letters (a, b, c and d)

766

differ significantly at P < 0.05 or P < 0.01 level.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

767 768

Figure 6. The qPCR and western blot were used to detect the expression of apoptosis-related genes

769

in CPF treated TM4 cells. (A) The expression of p53, puma, Caspase 3 and Caspase 9 of TM4 after

770

CPF treatment. (B) Expression of BAX and BCL2 protein were detected in TM4 cells after CPF

771

treated for 12 h. Note: the values with different letters (a, b, c and d) differ significantly at P < 0.05

772

or P < 0.01 level.

ACS Paragon Plus Environment

Page 38 of 47

Page 39 of 47

Journal of Agricultural and Food Chemistry

773 774

Figure 7. CPF modifications on cell cycle distribution of GC-1, TM4 and TM3 cells. Cells were

775

exposed to CPF (10, 25 and 50 µM) or vehicle for 12 or 24 h. A and B, C and D, E and F were the

776

results of GC-1, TM4 and TM3 cells, respectively. Cells were stained with propidium iodide (PI),

777

and analyzed for DNA content by flow cytometry. Note: the values with different letters (a, b, c and

778

d) differ significantly at P < 0.05 or P < 0.01 level. NS means no significant differences.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

779 780

Figure 8. The qRT-PCR was used to detect the expression of cell cycle regulators in CPF treated

781

GC-1, TM4 and TM3 cells. The expression of p21CIP, CCND1, CCNE1, CCNA2 and CCNB1 were

782

detected. A, B and C were the results of GC-1, TM4 and TM3 cells, respectively. Note: the values

783

with different letters (a, b, c and d) differ significantly at P < 0.05 or P < 0.01 level.

ACS Paragon Plus Environment

Page 40 of 47

Page 41 of 47

Journal of Agricultural and Food Chemistry

784 785 786

Figure 9. The expression of p-AMPK, mitochondrial membrane potentials, ROS level and MDA

787

level were detected in GC-1 cells after CPF treatment. (A) presented the expression of p-AMPK

788

protein in CPF treated GC-1 cells. (B) Detection of mitochondrial membrane potentials. (C)

789

Detection of MDA level. (D) Flow cytometry was applied to assess oxidative stress after CPF treated.

790

(E) and (F) indicated ROS levels in CPF treated cells compared to control group.

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

791 792

Figure 10. The expression of p-AMPK and ROS level were detected in TM4 cells after CPF

793

treatment. (A) Western blot analysis of p-AMPK expression in CPF treated TM4 cells. (B) p-AMPK

794

intensity analysis was calculated by Image J.**P