Mechanistic Insights into Stereospecific Bioactivity and Dissipation of

Jun 26, 2018 - State Key Laboratory of Pollution Control & Resource Reuse, School of ... instead of the racemate in agricultural management would redu...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIVERSITY OF THE SUNSHINE COAST

Agricultural and Environmental Chemistry

Mechanistic insights into stereospecific bioactivity and dissipation of chiral fungicide triticonazole in agricultural management Qing Zhang, Zhaoxian Zhang, Bowen Tang, Beibei Gao, Mingming Tian, Edmond Sanganyado, Hai-yan Shi, and MingHua Wang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b01771 • Publication Date (Web): 26 Jun 2018 Downloaded from http://pubs.acs.org on June 27, 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 35

Journal of Agricultural and Food Chemistry

1

Mechanistic insights into stereospecific bioactivity and

2

dissipation of chiral fungicide triticonazole in agricultural

3

management

4

Qing Zhang †, ‡, Zhaoxian Zhang †, ‡, Bowen Tang §, Beibei Gao

5

Edmond Sanganyado ┴, Haiyan Shi †, ‡, Minghua Wang †, ‡, *,

6



7

University, Nanjing 210095, P. R. China

8



9

Application, Nanjing 210095, P. R. China

†, ‡

, Mingming Tian #,

Department of Pesticide Science, College of Plant Protection, Nanjing Agricultural

State & Local Joint Engineering Research Center of Green Pesticide Invention and

10

§

11

China

12

#

13

Environment, Nanjing University, Nanjing 210023, P. R. China

14



15

*Corresponding Author: E-mail: [email protected]. Phone: +86 25 84395479.

16

Fax: +86 25 84395479.

College of Pharmaceutical Sciences, Xiamen University, Xiamen 361102, P. R.

State Key Laboratory of Pollution Control & Resource Reuse, School of the

Marine Biology Institute, Shantou University, Shantou 515063, P. R. China

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

17

ABSTRACT:

18

Research interest on chiral pesticides has increased probably because

19

enantiomers often exhibit different environmental fate and toxicity. An investigation

20

into the enantiomer-specific bioactivity of chiral triticonazole enantiomers in

21

agricultural systems revealed intriguing experimental and theoretical evidence. For

22

nine of the phytopathogens studied (Rhizoctonia solani, Fusarium verticillioide,

23

Botrytis cinerea (strawberry and tomato), Rhizoctonia cereali, Alternaria solani,

24

Gibberella zeae, Sclerotinia sclerotiorum, Pyricularia grisea), the fungicidal activity

25

data showed (R)-triticonazole was 3.11-82.89 times more potent than the

26

(S)-enantiomer. Furthermore, (R)-triticonazole inhibited ergosterol biosynthesis and

27

cell membrane synthesis more 1.80-7.34 times higher than its antipode. Homology

28

modeling and molecular docking studies suggested the distinct bioactivities of the

29

enantiomers of triticonazole were probably due to their different binding modes and

30

affinities to CYP51b. However, field studies demonstrated that (S)-triticonazole was

31

more persistent than (R)-triticonazole in fruits and vegetables. The results showed that

32

application of pure (R)-triticonazole, with its high bioactivity and relatively low

33

resistance risk, instead of the racemate in agricultural management would reduce the

34

application dosage required to eliminate carcinogenic mycotoxins and any

35

environmental risks associated with this fungicide, yielding benefits in food safety

36

and environmental protection.

37

KEYWORDS: chiral triticonazole; enantioselective bioactivities; stereoselective

38

dissipation; agricultural management; environmental protection

39 2

ACS Paragon Plus Environment

Page 2 of 35

Page 3 of 35

Journal of Agricultural and Food Chemistry

40

INTRODUCTION

41

Chirality is a ubiquitous concept in many scientific fields, including biology,

42

medicine, environmental science, chemistry, and agriculture 1. Numerous chiral

43

molecules generally interact with enzymes and biological receptors in an

44

enantioselective fashion 2. The tertiary stereocenters of chiral enantiomers are

45

important core structures which affect the molecule’s biological and pharmacological

46

properties

47

binding to one of the enantiomers in biological systems, which might lead to distinct

48

therapeutic or adverse effects

49

variety of plant pathogens for the purpose of crop protection. This class of fungicides

50

contains one or two chiral centers, but they are primarily applied as mixtures of

51

racemates for chemosynthetic and economic reasons. These compounds are known to

52

inhibit a cytochrome P450-dependent lanosterol 14α-demethylase (CYP51) via a

53

mechanism in which the heterocyclic nitrogen atom binds to the heme iron atom in

54

the binding site of the enzyme 8. It has been shown that the enantiomers of triazole

55

fungicides display dramatically different toxicities and biological activities in chiral

56

environments

57

systematic biological evaluation of all enantiomers for pharmaceutical candidates

58

Therefore, it is necessary to understand the bioactivity of each pure enantiomer for

59

chiral triazole fungicides and to study the molecular mechanisms of their

60

stereospecific bioactivity to gain a better understanding of plant diseases 11, 12.

61

3-6

. Chiral macromolecules might exhibit preferential metabolism of or

9, 10

2, 7

. Triazole fungicides are widely used to combat a

. Additionally, the US Food and Drug Administration requires a 2, 4

.

Plant pathogens cause extensive annual yield losses of staple crops worldwide 13. 3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 35

62

In addition, some of these pathogens not only reduce the yield and quality of infected

63

grain, but also produce a number of mycotoxins, such as deoxynivalenol and

64

zearalenone

65

mammals and negatively impact human health. Triazole fungicides are well known

66

for their excellent antifungal activity and relatively low resistance risk compared to

67

other commonly used fungicides, and they are thus considered the mainstay of

68

modern fungicides in agrochemical applications 15. Studies so far have shown that the

69

degradation of chiral triazole fungicides is commonly enantioselective in soil

70

For example, the microbial transformation of triadimefon to triadimenol was

71

enantioselective, with the (S)-enantiomer being preferentially transformed in soils 18.

72

In addition, (R)-flutriafol exhibited a longer elimination half-life than its antipode,

73

resulting in the accumulation of the (S)-isomer in rabbits

74

enantioselective plant uptake and accumulation of chiral triazole fungicides in crops 9,

75

10

76

chiral triazole fungicides in agricultural systems is lacking.

14

. Mycotoxins may cause adverse toxicological effects in exposed

19

9, 16-18

.

. Little is known about

. Although these compounds are widely used, vital stereospecific knowledge of the

77

To better understand the enantioselective bioactivity and degradation of chiral

78

triazole fungicides in agriculture, one of the broad-spectrum, systemic chiral triazole

79

fungicides triticonazole was investigated at length. In the present study,

80

stereoselectivity was evaluated during biodegradation in four kinds of fruits and

81

vegetables by applying racemic triticonazole in field conditions. Individual

82

enantiomers were used for characterizing their stereospecific bioactivity to nine

83

common plant pathogens in lab conditions. Furthermore, structural analysis and 4

ACS Paragon Plus Environment

Page 5 of 35

Journal of Agricultural and Food Chemistry

84

computer modeling technology was used to investigate how the selective interactions

85

of triticonazole enantiomers with cytochrome P450 14α-sterol demethylases (CYP51)

86

in plant pathogens lead to their enantioselective bioactivity. The results of this

87

research will be helpful in elucidating the molecular mechanisms of the

88

enantioselective inhibition of fungal cell-wall synthesis in plant pathogens targeted by

89

chiral triazole fungicides possessing multiple chiral centers. This knowledge will also

90

establish a more complete understanding of the different therapeutic or adverse effects

91

which characterize each enantiomer. The insight gained from this study will be helpful

92

in promoting the application of bioactive enantiomers while reducing the unwanted

93

effects of invalid enantiomers as high-risk pollutants in agricultural management.

94

MATERIALS AND METHODS

95

Chemicals, phytopathogens, and culture conditions. Racemic triticonazole

96

(98.2% purity), rac-(E)-5-[ (4-chlorophenyl) methylidene]-2,2-dimethyl-1-(1H-1,2,4-

97

triazol-1-ylmethyl) cyclopentanol (Fig. S1), and ergosterol (98.0% purity) were

98

purchased from the China Standard Material Center (Beijing, China). Two pure

99

enantiomers of triticonazole (purity ≥99.0%) were prepared from the Shanghai

100

Chiralway Biotech Co., Ltd (Shanghai, China). The ee value of (R)-triticonazole and

101

(S)-triticonazole were 99.72% and 99.52%, respectively. The cellulose tris

102

(3-chloro-4-methylphenylcarbamate) chromatographic column (Lux Cellulose-2,

103

Phenomenex, Torrance, USA) packed with 3-μm particles (250 mm×4.6 mm i.d.) was

104

used for the chiral separation of triticonazole. For achiral analysis, a C18 column

105

(150×4.6 mm, 4.6 µm) was used. Sample extraction was conducted using a Cleanert 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

106

Florisil-SPE cartridges (500 mg, 6 mL) obtained from Agela Technologies (Tianjin,

107

China). All chromatographic solutions (acetonitrile, methanol, and n-hexane) were

108

purchased from Tedia (Fairfield, USA). Analytical grade reagents (sodium chloride,

109

potassium hydroxide, anhydrous sodium sulfate, and anhydrous magnesium sulfate)

110

were purchased from commercial sources. Purified water was prepared using a

111

MUL-9000 water purification system (Nanjing Zongxin Water Equipment Co. Ltd,

112

China). The stock solution of (R)-, (S), and rac-triticonazole were each stored at

113

-20 °C, and the two enantiopure standards were found to be stable over six months.

114

Nine of the test phytopathogens (Rhizoctonia solani, Fusarium verticillioide,

115

Botrytis cinerea (strawberry and tomato), Rhizoctonia cereali, Alternaria solani,

116

Gibberella zeae, Sclerotinia sclerotiorum, Pyricularia grisea) were obtained from the

117

Key Laboratory of Pesticides, Nanjing Agricultural University (Nanjing, China).The

118

Potato Dextrose Agar (PDA) consisted of 200 g of potato, 20 g of agar, and 20 g of

119

dextrose for every 1 L of distilled water, and the Potato Dextrose (PD) was based on

120

PDA without the agar.

121

Bioassay of fungicidal activities. Rac-triticonazole and its enantiomers were

122

prepared in acetone solvent. A high concentration of (R)-, (S)- and rac-triticonazole

123

were added into separate unsolidified PDA after the autoclaved media had reached a

124

temperature of approximately 45-50 °C. Both of enantiomers are not interconversion

125

and racemization in PDA solution. To avoid the influence of acetone solvent on the

126

growth of phytopathogens, the concentration of acetone in the PDA medium was kept

127

below 0.1 ml L-1. Mycelial plugs (5 mm in diameter) from the margins of actively 6

ACS Paragon Plus Environment

Page 6 of 35

Page 7 of 35

Journal of Agricultural and Food Chemistry

128

growing colonies were transferred to solidified PDA plates which contained different

129

concentrations of (R), (S), and rac-triticonazole in aseptic conditions. The same

130

concentration of acetone was added into plates as a positive control. Each isolate was

131

incubated in the dark at 25 °C with three replicates. Since, the growth rate of each

132

phytopathogen was different, all colony diameters were determined when the control

133

colony diameter had reached approximately 7 cm. The percent growth inhibition was

134

calculated according to the following equation 9, 10:

135

Inhibition (%) = (colony diameter in the control - colony diameter in the treatment) ×

136

100 / (colony diameter in the control-5 mm)

(1)

137

The median effective concentrations (EC50) of (R), (S), and rac-triticonazole for

138

the nine phytopathogens were calculated by a linear regression of the colony diameter

139

to a log of the transformed chemical concentration using DPS data-processing

140

software.

141

Extraction and analysis of ergosterol in phytopathogens. Mycelia of nine

142

phytopathogens were cultured in 1000-mL flasks (15 mycelial plugs per flask)

143

containing 400 mL of liquid potato dextrose (PD) in a rotary shaker (160 rpm) at 25°C,

144

whereas the mycelial plugs (5 mm diameter) were harvested from actively growing

145

colonies. When the mycelial plugs grew to the appropriate size after several days, the

146

different triticonazole enantiomers were separately added into liquid potato dextrose

147

(PD). The spiking concentrations of (R), (S), and rac-triticonazole were close to the

148

EC90 values for each pathogen and described in the Table S1. The liquid potato

149

dextrose without triticonazole served as a control and all treatments were in triplicate. 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 35

150

After a 72 h incubation, ergosterol was extracted from each phytopathogen according

151

to previous research with some modifications, and then ergosterol concentrations were

152

determined using high performance liquid chromatography (HPLC)

153

mycelia were collected by filtration and washed with sterile water several times, and

154

the mycelia were dried in a vacuum freeze drier. Approximately 0.5 g of dry mycelia

155

were broken with liquid nitrogen and subsequently suspended in 20 mL of freshly

156

prepared methanolic KOH (10 %, w/v) solvent. The mixtures was homogenized for 1

157

min, sonicated for 10 min, and then saponified at 80 °C for 90 minutes. To extract

158

ergosterol, a mixture of 5 mL distilled water and 10 mL n-hexane was added followed

159

by vigorous vortexing for 1 min. The samples were then centrifuged for 5 min at 3000

160

rpm. The ergosterol extraction procedure was repeated three times, then the n-hexane

161

supernatants from each extraction stage were pooled, evaporated to dryness, and then

162

re-dissolved in 10 mL of methanol-water (98:2, v/v). Before analysis, the samples

163

were filtered through a 0.22-µm filter membrane (Tengda, Tianjin, China). HPLC

164

analyses were conducted on an Agilent 1200 HPLC system (Agilent, USA) with a

165

ultraviolet detector. The separation of ergosterol was accomplished on a C18 column

166

(150×4.6 mm, 4.6 µm). Isocratic elution was conducted using a mobile phase

167

comprising of methanol and water (98:2, v/v) at a flow rate of 1.0 mL min-1. The

168

injection volume was 20 μL and the detection wavelength was set at 282 nm.

20-23

. Briefly,

169

Cell membrane permeability of phytopathogens. Prior to collection of plant

170

pathogen mycelia, the same concentration of triticonazole enantiomers as used in the

171

last experiment was separately added into liquid PD and incubated for 36 h (Table S1). 8

ACS Paragon Plus Environment

Page 9 of 35

Journal of Agricultural and Food Chemistry

172

After incubation, the mycelia were filtered on a double gauze, and thoroughly washed

173

using distilled water. The mycelia (0.500 g) were suspended in separate centrifuge

174

tubes containing 20 mL of double-distilled water 24. All treatments were performed in

175

triplicate at each time point. Then, the electrical conductivity in this system was

176

assessed by determining the cell membrane permeability at different time intervals

177

using a conductivity meter (CON510 Eutech/Oakton, Singapore). Finally, the

178

centrifuge tubes were put into boiling water for 5 min to allow for thorough

179

breakdown of the cell membrane, and final conductivity was determined. The relative

180

conductivity of the various mycelia was calculated according to the following

181

equation:

182 183

Relative conductivity (%) = Conductivity at different time points / Final conductivity × 100

(2)

184

Field experiments. Four kinds of vegetables (tomatoes, cucumbers, spinach,

185

Chinese cabbage) were prepared under field conditions in Nanjing, China. An

186

experimental field with no application history of triazole fungicide was divided into

187

several plots, each with an area of 30 m2 and insulated with a buffer zone. When the

188

vegetables are grown in appropriate stages according to the BBCH-scale, triticonazole

189

(2.5%, FS) was used as a foliar treatment with an application rate of approximately

190

400 g a.i. ha-1 (two times the recommended dose). Besides triticonazole, no other

191

triazole fungicides were applied to the crops during the field trials. Each field

192

experimental treatment was set up in triplicate for the dissipation rate study.

193

Representative vegetable samples were collected at different periods of time after 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

194

spraying. All samples were separately ground and stored in glass containers at -20 °C.

195

Extraction and analysis for vegetables. A 20 g aliquot of each vegetable

196

sample was extracted using 50 mL of ethyl acetate in a 100 mL Teflon centrifuge tube.

197

The extraction and cleanup procedures for chiral triticonazole were based on a

198

previous method with modifications

199

min followed by sonication for 15 min. After that, triticonazole was salted out from

200

the aqueous phase using 2.0 g of sodium chloride. The extraction mixture was

201

vigorously vortexed for 1 min and then centrifuged at 4,000 rpm for 5 min. The ethyl

202

acetate supernatant (~25-mL) was collected and evaporated prior to sample

203

purification using Florisil-SPE. Following purification, the sample was filtered

204

through a 0.22 μm filter before HPLC analysis. The chiral separation and

205

determination of enantiomers were according the previous established methods25.

206 207 208

25

. Briefly, the samples were homogenized for 3

Enantiomeric fraction (EF) described the enantiomeric composition of chiral compounds via the following equation 9: EF = R/(R + S)

(3)

209

where (S)-(+)-triticonazole was eluted first, followed by (R)-(-)-triticonazole on the

210

chiral chromatographic column. The R and S in this equation was the concentration of

211

(R) and (S) enantiomers, respectively.

212

Homology modeling and molecular docking. Cytochrome P450 14α-sterol

213

demethylases (CYP51) are essential enzymes in the process of sterol biosynthesis in

214

phytopathogens [26]. Triazole fungicides used in treatment of topical and systemic

215

mycoses function as inhibitors of CYP51, thus it is important to further explore the 10

ACS Paragon Plus Environment

Page 10 of 35

Page 11 of 35

Journal of Agricultural and Food Chemistry

216

mechanisms of triticonazole’s enantioselective bioactivity. The primary sequences of

217

nine phytopathogens are very similar to those of Aspergillus fumigatus, with the

218

sequences of six plant pathogens being over 58.5% similar, and the remaining three

219

being over 22.2% similar (Table S2). Therefore, the crystal structure (PDB: 4UYM)

220

of template proteins (CYP51B) for Aspergillus fumigatus was used for homology

221

modeling. Three-dimensional models of CYP51B in six plant pathogens with high

222

sequence similarity (Botrytis cinerea (strawberry and tomato), Gibberella zeae,

223

Fusarium verticillioide, Sclerotinia sclerotiorum, Pyricularia grisea) were built based

224

on the aligned sequences by homology modeling using the Prime application in

225

Schrodinger Maestro Suite 2015 (Prime, version 3.9, Schrodinger, LLC, New York,

226

NY, 2015). The initial three-dimensional structural models were selected and

227

energy-minimized for further refinement using the OPLS 2005 force field in

228

Schrodinger Prime. The validated models were optimized using the Protein

229

Preparation Wizard in Schrodinger Maestro Suite 2015 prior to simulated docking

230

with chiral triticonazole.

231

The docking between the CYP51b models from five types of plant pathogens and

232

chiral triticonazole enantiomers was performed using the Glide approach integrated in

233

Maestro 10.1 (Schrodinger, LLC, 2015), with standard precision mode. The binding

234

site was centered on the co-factor HEM (heme) and the metal-ligand interactions were

235

set as docking constraints in the Glide grid generation process, in which the remaining

236

settings used default parameters. Two enantiomers of chiral triticonazole were

237

separately docked into the binding site after preparation using LigPrep with the 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 35

238

docking protocol. Glide Score (GScore) as a scoring function is used to describe each

239

docking complex conformation, and the values generated are negative values. The

240

best-ranked complex conformation of each enantiomer by GScore was selected to

241

analyze the different binding modes with CYP51b for chiral triticonazole. A higher

242

absolute docking value of GScore means that the docking between ligand and receptor

243

is more stable 27, 28.

244

RESULTS AND DISCUSSION

245

Enantioselective

bioactivity

in

phytopathogens.

The

stereochemical

246

bioactivity of chiral triticonazole for nine phytopathogens was determined using a

247

biological assay based on the inhibition of bacteriostatic circle method. The racemic

248

mixture, rac-triticonazole, exhibited antifungal activity against all phytopathogens

249

with EC50 values of less than 1.00 mg kg-1, exhibited antifungal activity against all

250

phytopathogens with EC50 values less than 1.00mg kg-1, except for Pyriculariagrisea

251

where an EC50 value of 2.16 mg kg-1 was obtained (Table 1). Furthermore, the

252

antifungal activity of rac-triticonazole against Botrytis cinereal varied between

253

different host plant species. For example, based on EC50 values, rac-triticonazole was

254

2.33 times more effective in Botrytis cinereal infected strawberry plants than tomato.

255

As shown in Table 1, the two enantiomers of triticonazole possessed different

256

fungicidal activities and the order of potency was (R)-triticonazole > rac-triticonazole >

257

(S)-triticonazole in all bioassays. It was observed (R)-triticonazole was 3.11-82.89

258

times more active than the (S)-enantiomer. However, enantioselective bioactivity was

259

low against Pyriculariagrisea with the (R)-enantiomer only 3.11 times more potent 12

ACS Paragon Plus Environment

Page 13 of 35

Journal of Agricultural and Food Chemistry

260

than its antipode. In addition, the (R)-enantiomer was significantly more bioactivity

261

against Fusarium verticillioide and Rhizoctoniasolani with EC50 values 43.65-82.89

262

times higher than the (S)-enantiomer. Thus, the antifungal activity of triticonazole

263

may be attributed primarily to the (R)-enantiomer. The EC50 values of triticonazole

264

from the experiment showed in eight phytopathogens (Rhizoctoniasolani, Fusarium

265

verticillioide, Botrytis cinerea (strawberry and tomato), Rhizoctoniacereali,

266

Alternaria

267

contributed approximately 92.07-98.81% of the activity against plant pathogens in the

268

racemate. In previous studies on triazole fungicides, a 24.2-fold and 1000-fold

269

difference was observed between the chiral enantiomers of difenoconazole and

270

paclobutrazol, respectively 9, 29. Therefore, only one of the enantiomers of current-use

271

chiral fungicides may exhibit pathogenic control properties in agricultural

272

management.

solani,

Gibberellazeae,

Sclerotiniasclerotiorum),

(R)-enantiomer

273

Enantioselectivity in ergosterol biosynthesis. To further explore the underlying

274

mechanism of triticonazole’s enantioselective bioactivity on phytopathogens, the

275

inhibition of ergosterol biosynthesis by chiral enantiomers was investigated using the

276

chromatographic analysis. The concentrations of ergosterol for all treatments are

277

shown in Fig. 1. All triticonazole treatments significantly affected the ability of

278

pathogenic

279

Enantioselectivity was observed in the inhibitory effects on ergosterol biosynthesis in

280

nine phytopathogens for the two enantiomers (Table S3). Compared to the control,

281

around 25.63-77.58%, 18.23-64.27%, and 3.63-36.52% of ergosterol biosynthesis was

cells

to

synthesize

ergosterol

following

13

ACS Paragon Plus Environment

exposure

for

72

h.

Journal of Agricultural and Food Chemistry

282

inhibited by (R), (rac), and (S)-triticonazole, respectively. The inhibition of ergosterol

283

production was generally higher (1.80-7.34 times) following (R)-triticonazole

284

exposure than (S)-triticonazole against the nine pathogens. The enantioselective

285

inhibition of ergosterol production observed was in agreement with the

286

enantioselective bioactivity found in exposed phytopathogens. A positive correlation

287

(r=0.86) between the enantioselective bioactivity of the nine phytopathogens and the

288

inhibition of ergosterol biosynthesis for the triticonazole enantiomers was obtained by

289

fitting the bioactive ratio value (R/S) to the corresponding inhibition of ergosterol

290

ratio value in Fig. 2. The results suggest that the distinct bioactivities observed for

291

chiral triazole fungicides was mainly caused by the loss of the ability to synthesize

292

ergosterol in phytopathogenic cells.

293

Enantioselectivity in cell membrane permeability. Many commercial chiral

294

triazole fungicides are capable of affecting gene expression and may thus further

295

block the process of ergosterol biosynthesis

296

process for some commercial chiral triazole fungicides which are designed to

297

eliminate plant pathogens

298

substances in fungal cell membranes. Chiral triazole fungicides inhibit the formation

299

of cell membranes and ultimately the exchange of substances across the membrane,

300

resulting in cell death

301

concentrations of racemic and enantiopure triticonazole (Table S1), and the electrical

302

conductivity was detected after different exposure times. Following triticonazole

303

treatment, electrical conductivity gradually increased with increasing treatment time

20

21

. Ergosterol biosynthesis is the target

26

. This is significant since ergosterols are essential

. Nine plant pathogenic cells were treated with appropriate

14

ACS Paragon Plus Environment

Page 14 of 35

Page 15 of 35

Journal of Agricultural and Food Chemistry

304

and finally reached equilibrium after incubating for 180 h (Fig. S2). The results show

305

that (R)-, (S)-and, rac-triticonazole differentially enhanced the electrical conductivity

306

of the pathogenic cells. However, the electrical conductivity of nine plant pathogenic

307

cells amended with (S)-triticonazole were not significantly different from the control

308

for some of the plant pathogens (i.e. Botrytis cinerea, Pyricularia grisea, Sclerotinia

309

sclerotiorum). Furthermore, (R)-triticonazole substantially increased electrical

310

conductivity by 21.72 to 76.42% under the same treatment conditions, indicating that

311

the (R)-isomer is more potent in this regard (Table 2). Therefore, the enantioselective

312

cell membrane permeability for plant pathogenic cells may offer a more

313

comprehensive insight on the mechanism of enantioselective bioactivity caused by

314

chiral triazole fungicides. Cytochrome P450 enzyme are essential catalytic enzymes in

315

fungal ergosterol biosynthesis

316

against plant pathogenic cell requires further research to be understood, especially in

317

regard to the enantioselective relationship between chiral triazole fungicides and

318

phytopathogenic CYP51b.

30

. The precise impact of chiral triazole fungicides

319

Molecular interactions between chiral triticonazole enantiomers and

320

CYP51b. The homology modeling and molecular docking techniques are performed

321

to further explain the stereoselective molecular interactions for chiral enantiomers in

322

the phytopathogens. The composition of the three dimensional active group (i.e.

323

hydroxy, benzene, and pyrazolone ring) in triticonazole enantiomers strongly affects

324

their binding affinities to CYP51b enzymes. Both enantiomers can be tightly bound to

325

the active cavities of CYP51b, and the enantiomers are shown at different distances 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

326

from the key amino acids present (Fig. 3). For five of the phytopathogens studied (i.e.

327

Botrytis cinerea, Gibberella zeae, Fusarium verticillioide, Sclerotinia sclerotiorum,

328

Pyricularia grisea), several residues are in the active site of CYP51b within a distance

329

of 4 Å and 5 Å for three-dimensional and two-dimensional representations,

330

respectively (Fig. 3). All active site residues are summarized in Table S4. These

331

residues are closely related to the observed binding affinities for triticonazole

332

enantiomers since they determine the three-dimensional conformation of the

333

enzyme-ligand interaction. The binding affinities depend on the structure of chiral

334

compounds and of the receptor proteins themselves

335

groups (i.e. hydroxy and benzene) present in chiral triticonazole easily generate

336

hydrogen bonds and Pi-Pi stacking interactions with amino acid residues in the

337

enzyme active site. In addition, the ferric ions of heme are present in the active pocket

338

of CYP51b, allowing for Pi-cation and metal coordination interactions in the presence

339

of triticonazole. Furthermore, hydrophobic effects play an important role in

340

combining the hydrophobic functional groups of triticonazole with amino acid

341

residues 33, 34.

25, 31, 32

. Some of the functional

342

For five phytopathogens with high sequence similarity, the binding activity of

343

the association between the two enantiomers and CYP51b was investigated in terms

344

of their docking energy 31, 33. The top scored protonation formed between triticonazole

345

and CYP51b was selected, since docking score includes both binding affinity and

346

tautomeric ratio 35, 36. The two enantiomers showed distinct binding modes with amino

347

acid residues in the enzyme active site (Fig. 3). The binding affinity values of the 16

ACS Paragon Plus Environment

Page 16 of 35

Page 17 of 35

Journal of Agricultural and Food Chemistry

348

(R)-enantiomer were consistently lower than those of the (S)-isomer, suggesting that

349

(R) bound strongly to CYP51b, while (S) had poor binding activity to CYP51b in

350

phytopathogens (Table S4). The different binding activity with chiral triticonazole

351

could be synthetically caused by H-bonds, Pi-Pi stacking, Pi-cation interactions, metal

352

coordination, and hydrophobic effects, which are critical for stereoselective molecular

353

recognition

354

(R)-enantiomer exhibited greater bioactivity against common plant pathogens.

355

Therefore, the distinct bioactivities of chiral enantiomers in triazole fungicides may

356

derive from variable binding affinities and binding modes to CYP51b.

37, 38

. The results were consistent with the experimental results that the

357

Enantioselective dissipation in fruits and vegetables. The change of the value

358

of EF is a powerful metric to indicate enantioselective behavior of chiral pollution in

359

biological processes, and it was used to evaluate the enantioselective dissipation of

360

chiral triticonazole in vegetables under field conditions

361

relatively high concentrations of triticonazole after the foliar spray treatment,

362

followed by a gradual decreased with time. The dissipation of triticonazole

363

enantiomers followed first order kinetics (R2 ≥ 0.927), and the half-lives for both

364

enantiomers are shown in Table S5. The half-lives of the more bioactive

365

(R)-enantiomer were consistently shorter than those of (S)-triticonazole in the fruits

366

and vegetables studied. The average EF value for triticonazole remained constant (EF

367

= 0.5) at the beginning of the foliar spray treatment (0-5d for cucumber, 0-7d for

368

Chinese cabbage and tomato, and 0-14 d for spinach) in four plants (Fig. 4), but the

369

EF value instantly increased to 0.576-0.675 in the late stages (0.576 for tomatoes, 17

ACS Paragon Plus Environment

21

. The vegetables contained

Journal of Agricultural and Food Chemistry

Page 18 of 35

370

0.609 for spinach, 0.624 for Chinese cabbage and 0.675 for cucumbers). Preferential

371

dissipation of the (R)-enantiomer resulted in relative enrichment of the (S)-enantiomer.

372

Hence, the less bioactive (S)-enantiomer is more persistent in fruits and vegetables,

373

thus humans may be exposed to it through dietary intake. Chiral switching to

374

(R)-triticonazole from the commercial racemate will probably help in crop protection

375

against pathogens while reducing potential public health risks. The enzymatic systems

376

in plants have been shown in previous studies to play an important role in the

377

stereoselective degradation, metabolism, and accumulation of chiral pollutants

378

The absence of enantioselectivity at the beginning of the foliar spray treatment may be

379

due to the fact that the system of stereoselective enzymes had limited contact with

380

chiral triticonazole during this initial period. After this initial lag period following

381

pesticide application, the transportation and distribution of triticonazole throughout

382

the plant tissues increased in fruits and vegetables. Therefore, the system of

383

stereoselective enzymes was exposed to the chiral fungicides, resulting in a significant

384

enantioselective dissipation of the enantiomers in the vegetables.

9, 39

.

385

In this present work, Cell membrane permeability, ergosterol biosynthesis, and

386

molecular interaction investigations revealed significant differences between (R)-, (S)-,

387

and rac-triticonazole. Thus, the current study demonstrated the bioactivity and

388

dissipation of chiral triticonazole in agricultural ecosystems was enantiomer-specific.

389

The different binding affinities between the individual triticonazole enantiomers and

390

receptor proteins of CYP51b demonstrated the enantioselective bioactivities of chiral

391

triazole fungicides at molecular level. In addition, although (R)-triticonazole was 18

ACS Paragon Plus Environment

Page 19 of 35

Journal of Agricultural and Food Chemistry

392

found to be more bioactive than (S)-triticonazole against nine phytopathogens,

393

(R)-triticonazole was less persistent than (S)-triticonazole in various kinds of fruits

394

and vegetables under field conditions. Therefore, chiral switching to the more

395

bioactive and less persistent (R)-triticonazole from the currently used racemate would

396

reduce the required application dosage and hence the associated environmental risks

397

while still maintaining the fungicidal efficacy. Enantioselectivity is commonly

398

observed in several aspects of environmental science (i.e. adsorption, degradation,

399

transportation, excretion and ecotoxicology) for numerous chiral pesticides. Therefore,

400

the significance of enantioselective effects for chiral agrochemicals must be

401

considered in order to improve their risk assessment and minimize dangers to public

402

health.

403

ABBREVIATIONS USED

404

PDA, potato dextrose agar; PD, potato dextrose; EC50, effective concentrations;

405

HPLC, high performance liquid chromatography; GScore, glide score; EF,

406

enantiomeric fraction.

407

AUTHOR INFORMATION

408

Corresponding Author

409

*Phone: +86 25 84395479. Fax: +86 25 84395479. E-mail: [email protected].

410

Funding

411

This study was supported by the National Key Research and Development Program of

412

China (2016YFD0200207).

413

Notes 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

414

Conflicts of interest the authors do not have any potential conflict of interest to

415

declare.

416

ASSOCIATED CONTENT

417

Supporting Information

418

Supporting materials (Fig. S1-S2 and Table S1-S5) are available free of charge via the

419

Internet at http://pubs.acs.org.

420 421

REFERENCES

422

[1] Morris, R. E.; Bu, X. Induction of chiral porous solids containing only achiral

423 424 425

building blocks. Nat. chem. 2010, 2, 353-361. [2] Shi, S.; Wong, Z.; Buchwald, S. L. Copper-catalysed enantioselective stereodivergent synthesis of amino alcohols. Nature 2016, 532, 353-356.

426

[3] Wu, L.; Wang, F.; Wan, X.; Wang, D.; Chen, P.; Liu, G. Asymmetric cu-catalyzed

427

intermolecular trifluoromethylarylation of styrenes: enantioselective arylation of

428

benzylic radicals. J. Am. Chem. Soc. 2017, 139, 2904-2907.

429

[4] Xu, S.; Oda, A.; Kamada, H.; Negishi, E. Highly enantioselective synthesis of γ-,

430

δ-, and ε-chiral 1-alkanols via Zr-catalyzed asymmetric carboalumination of

431

alkenes (ZACA)-Cu- or Pd-catalyzed cross-coupling. P. Natl. Acad. Sci. 2014,

432

111, 8368-8373.

433

[5] Sanganyado, E.; Lu, Z.; Fu, Q.; Daniel, S.; Jay, G. Chiral pharmaceuticals: A

434

review on their environmental occurrence and fate processes. Water Res. 2017,

435

527-542. 20

ACS Paragon Plus Environment

Page 20 of 35

Page 21 of 35

436 437

Journal of Agricultural and Food Chemistry

[6] Jin, B.; Rolle, M. Joint interpretation of enantiomer and stable isotope fractionation for chiral pesticides degradation. Water Res. 2016, 178-186.

438

[7] Lawrence, L. J.; Casida,BJ.E. Stereospecific action of pyrethroid insecticides on

439

the gamma-aminobutyric acid receptor-ionophore complex. Science. 1983, 221,

440

1399-1401.

441

[8] Lepesheva, G.I.; Waterman, M.R. Sterol 14α-demethylase cytochrome P450

442

(CYP51), a P450 in all biological kingdoms. Bba-Gen. Subjects. 2007, 1770,

443

467-477.

444

[9] Dong, F.; Li, J.; Chankvetadze, B.; Cheng, Y.; Xu, J.; Liu, X.; Li, Y.; Chen, X.;

445

Bertucci, C.; Tedesco, D.; Zanasi, R.; Zheng, Y. Chiral triazole fungicide

446

difenoconazole: Absolute stereochemistry, stereoselective bioactivity, aquatic

447

toxicity, and environmental behavior in vegetables and soil. Environ. Sci.

448

Technol. 2013, 47, 3386-3394.

449

[10] Zhang, Q.; Hua, X.; Shi, H.; Liu, J.; Tian, M.; Wang, M. Enantioselective

450

bioactivity, acute toxicity and dissipation in vegetables of the chiral triazole

451

fungicide flutriafol. J. Hazard. Mater. 2015, 284, 65-72.

452 453

[11] Corić, I.; List, B. Asymmetric spiroacetalization catalysed by confined Bronsted acids. Nature. 2012, 483, 315-319.

454

[12] Lewis, D.L.; Garrison, A.W.; Wommack, K.E.; Whittemore, A.; Steudler, P.;

455

Melillo, J. Influence of environmental changes on degradation of chiral

456

pollutants in soils. Nature. 1999, 401, 898-901.

457

[13] Giraldo, M.C. Valent, B. Filamentous plant pathogen effectors in action. Nature 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

458

Reviews Microbiology. 2013, 11, 800-814.

459

[14] Proctor, R.H.; Hohn, T. M.; Mccormick, S.P. Reduced virulence of gibberella

460

zeae caused by disruption of a trichothecene toxin biosynthetic gene. Mol. Plant

461

Microbe. In. 1995, 8, 593-601.

462

[15] Mazur, C.S.; Kenneke, J. F. Cross-species comparison of conazole fungicide

463

metabolites using rat and rainbow trout (onchorhynchus mykiss) hepatic

464

microsomes and purified human CYP 3A4. Environ. Sci. Technol. 2008, 42,

465

947-954.

466

[16] Li, Y.; Dong, F.; Liu, X.; Xu, J.; Li, J.; Kong, Z.; Chen, X.; Zheng, Y.

467

Environmental behavior of the chiral triazole fungicide fenbuconazole and its

468

chiral metabolites: enantioselective transformation and degradation in soils.

469

Environ. Sci. Technol. 2012, 46, 2675-2683.

470

[17] Buerge, I.J.; Poiger, T.; Müller, D. M.; Buser, H. R. Influence of pH on the

471

stereoselective degradation of the fungicides epoxiconazole and cyproconazole

472

in soils. Environ. Sci. Technol. 2006, 40, 5443-5050.

473

[18] Garrison, A. W.; Avants, J. K.; Jones, W. J. Microbial transformation of

474

triadimefon to triadimenol in soils: selective production rates of triadimenol

475

stereoisomers affect exposure and risk. Environ. Sci. Technol. 2011, 45,

476

2186-2193.

477

[19] Shen, Z.; Zhang, P.; Xu, X.; Wang, X.; Zhou, Z.; Liu, D. Gender-related

478

differences in stereoselective degradation of flutriafol in rabbits. J. Agric. Food

479

Chem. 2011, 59, 10071-10077. 22

ACS Paragon Plus Environment

Page 22 of 35

Page 23 of 35

Journal of Agricultural and Food Chemistry

480

[20] Sun, X.; Wang, J.; Feng, D.; Ma, Z.; Li, H. PdCYP51B, a new putative sterol

481

14α-demethylase gene of Penicillium digitatum involved in resistance to imazalil

482

and other fungicides inhibiting ergosterol synthesis. Appl. Microbiol. Biotechnol.

483

2011, 91, 1107-1119.

484

[21] Ouyang, Q.; Tao, N.; Jing, G. Transcriptional profiling analysis of penicillium

485

digitatum, the causal agent of citrus green mold, unravels an inhibited ergosterol

486

biosynthesis pathway in response to citral. BMC Genomics. 2016, 17, 599.

487 488

[22] Nes, W. D. Biosynthesis of cholesterol and other sterols. Chem. Rev. 2011, 111, 6423-6451.

489

[23] Alcazar-Fuoli, L.; Mellado, E.; Garcia-Effron, G.; Lopez, J. F.; Grimalt, J. O.;

490

Cuenca-Estrella, J. M.; Rodriguez-Tudela, J. L. Ergosterol biosynthesis pathway

491

in Aspergillus fumigatus. Steroids. 2008, 73, 339-347.

492

[24] Duan, Y.; Ge, C.; Liu, S.; Chen, C.; Zhou, M. Effect of phenylpyrrole fungicide

493

fludioxonil on morphological and physiological characteristics of Sclerotinia

494

sclerotiorum. Pestic. Biochem. Phys. 2013, 106, 61-67.

495

[25] Zhang, Q.; Gao, B.; Tian, M.; Shi, H.; Hua, X.; Wang, M. Enantioseparation and

496

determination of triticonazole enantiomers in fruits, vegetables, and soil using

497

efficient extraction and clean-up methods. J. Chromatogr. B. 2016, 1009,

498

130-137.

499

[26] Podust, L.M.; Poulos, T. L.; Waterman, M. R. Crystal structure of cytochrome

500

P450 14α-sterol demethylase (CYP51) from mycobacterium tuberculosis in

501

complex with azole inhibitors. P. Natl. Acad. Sci. 2001, 98, 3068-3073. 23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 35

502

[27] Tian, M.; Zhang, Q.; Hua, X.; Tang, B.; Gao, B.; Wang, M. Systemic

503

stereoselectivity study of flufiprole: Stereoselective bioactivity, acute toxicity

504

and environmental fate. J. Hazard. Mater. 2016, 320, 487-494.

505

[28] Zhang, J.; Begum, A.; Brännström, K.; Grundström, C.; Iakovleva, I.; Olofsson,

506

A.; Sauer-Eriksson, A. E.; Andersson, P. L. A structure-based virtual screening

507

protocol for in silico identification of potential thyroid disrupting chemicals

508

targeting transthyretin. Environ. Sci. Technol. 2016, 50, 11984-11993.

509

[29] Burden, R.S.; Carter, G. A.; Clark, T.; Cooke, D.T.; Croker, S. J.; Deas, A. H. B.;

510

Hedden, P.; James, C. S.; Lenton, J. R. Comparative activity of the enantiomers

511

of triadimenol and paclobutrazol as inhibitors of fungal growth and plant sterol

512

and gibberellin biosynthesis. Pest Manag. Sci. 2010, 21, 253-267.

513

[30] Trzaskos, J. M.; Fischer, R. T.; Favata, M. F. Mechanistic studies of lanosterol

514

c-32

demethylation.

conditions

which

promote

oxysterol

intermediate

515

accumulation during the demethylation process. J. Biol. Chem. 1986, 26,

516

16937-16942.

517

[31] Zhai, G.; Hu, D.; Lehmler, H. J.; Schnoor, J. L. Enantioselective

518

biotransformation of chiral PCBs in whole poplar plants. Environ. Sci. Technol.

519

2011, 45, 2308-2316.

520

[32] Khalaf, H.; Larsson, A.; Berg, H.; Mccrindle, R.; Arsenault, G.; Olsson, P. E.

521

Diastereomers

of

the

brominated

flame

retardant

522

dibromoethyl)cyclohexane induce androgen receptor activation in the HepG2

523

hepatocellular carcinoma cell line and the LNCaP prostate cancer cell line. 24

ACS Paragon Plus Environment

1,2-dibromo-4-(1,2

Page 25 of 35

524

Journal of Agricultural and Food Chemistry

Environ. Health Perspect. 2009, 117, 1853-1859.

525

[33] Zhao, M.; Zhang, Y.; Zhuang, S.; Zhang, Q.; Lu, C.; Liu, W. Disruption of the

526

hormonal network and the enantioselectivity of bifenthrin in trophoblast:

527

Maternal-fetal health risk of chiral pesticides. Environ. Sci. Technol. 2014, 48,

528

8109-8116.

529

[34] He, X.; Dong, X.; Zou, D.; Yu, Y.; Fang, Q.; Zhang, Q.; Zhao, M.

530

Enantioselective effects of o,p′-DDT on cell invasion and adhesion of breast

531

cancer cells: chirality in cancer development. Environ. Sci. Technol. 2015, 49,

532

10028-10037.

533

[35] Shelley, J. C.; Cholleti, A.; Frye, L. L.; Greenwood, J. R.; Timlin, M. R.;

534

Uchimaya, M. Epik: A software program for pKa prediction and protonation state

535

generation for drug-like molecules. J. Comput. Aided. Mol. Des. 2017, 21,

536

681-691.

537

[36] Friesner, R. A.; Banks, J. L.; Murphy, R. B.; Halgren, T. A.; Klicic, J. J.; Mainz,

538

D. T.; Repasky, M. P.; Knoll, E. H.; Shelley, M.; Perry, J. K.; Shaw, D. E.;

539

Francis, P.; Shenkin, P. S. Glide: A New Approach for Rapid, Accurate Docking

540

and Scoring. 1. Method and Assessment of Docking Accuracy. J. Med. Chem.

541

2004, 47, 1739-1749.

542

[37] Iakovleva, I.; Brännström, K.; Nilsson, L.; Gharibyan, A. L.; Begum, A.; Anan, I.;

543

Walfridsson, M.; mSauer-Eriksson, A. E.; Olofsson, A. Enthalpic forces correlate

544

with the selectivity of transthyretin-stabilizing ligands in human plasma. J. Med.

545

Chem. 2015, 58, 6507-6515. 25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

546

[38] Alhamadsheh, M. M.; Connelly, S.; Cho, A.; Reixach, N.; Powers, E. T.; Pan,

547

D.W.; Wilson, I. A.; Kelly, J. W.; Graef, I. A. Potent kinetic stabilizers that

548

prevent transthyretin-mediated cardiomyocyte proteotoxicity. Sci. Transl. Med.

549

2011, 3, 81.

550

[39] Huang, H.; Zhang, S.; Lv, J.; Wen, B.; Wang, S.; Wu, T. Experimental and

551

theoretical evidence for diastereomer- and enantiomer-specific accumulation and

552

biotransformation of HBCD in maize roots. Environ. Sci. Technol. 2016, 50,

553

12205-12213.

554

26

ACS Paragon Plus Environment

Page 26 of 35

Page 27 of 35

Journal of Agricultural and Food Chemistry

555

Figure Legends

556

Figure 1. The enantioselective inhibition of ergosterol biosynthesis by chiral

557

triticonazole enantiomers.

558

Figure 2. The relationship between the R/S ratio value of enantioselective bioactivity

559

and the enantioselective inhibition of ergosterol.

560

Figure 3. The two-dimensional and three-dimensional enantiomer-specific binding

561

modes for triticonazole enantiomers bound to CYP51b.

562

Figure 4. Time development of EF values for the triticonazole enantiomers in

563

vegetables.

564

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 35

Table 1. EC50 values of chiral triticonazole enantiomers for nine plant pathogens. R-triticonazole

Species

Rac-triticonazole

S-triticonazole

R/Sd

EC50a (mg kg-1)

R2b

Pc

EC50 (mg kg-1)

R2

P

EC50 (mg kg-1)

R2

P

Rhizoctonia solani

0.0060±0.0005

0.9497

0.0173

0.0091±0.0007

0.9516

0.0019

0.2619±0.0057

0.9873

0.0027

43.65

Fusarium verticillioide

0.0335±0.0019

0.9574

0.0105

0.1153±0.0039

0.9830

0.0026

2.7769±0.0431

0.9870

0.0018

82.89

Botrytis cinerea (strawberry) Rhizoctonia cereali

0.1190±0.0036

0.9456

0.0048

0.4307±0.0121

0.9421

0.0058

1.3825±0.0578

0.9917

0.0057

11.62

0.1632±0.0043

0.9710

0.0231

0.3485±0.0193

0.9755

0.0134

1.9580±0.0713

0.9967

0.0151

12.00

Botrytis cinerea (tomato) Alternaria solani

0.2771±0.0203

0.9916

0.0071

0.7431±0.0363

0.9847

0.0047

4.2013±0.2151

0.9448

0.0071

15.16

0.3084±0.0176

0.9847

0.0238

0.7934±0.0219

0.9887

0.0313

4.8694±0.2637

0.9921

0.0054

15.79

Gibberella zeae

0.4066±0.0183

0.9896

0.0137

0.8034±0.0605

0.9757

0.0171

6.0840±0.4179

0.9938

0.0079

14.96

Sclerotinia sclerotiorum

0.5236±0.0471

0.9900

0.0091

0.9070±0.0786

0.9910

0.0121

11.2996±1.0021

0.9704

0.0191

21.58

Pyricularia grisea

1.7661±0.0893

0.9798

0.0275

2.1607±0.1563

0.9679

0.0029

5.4854±0.4873

0.9966

0.0023

3.11

a the effective concentration that results in a 50% reduction in population growth compared to the control. b represents the correlation coefficient. c represents the probability (associated with the t-test). d the EC50 value of R-triticonazole to S-triticonazole ratio.

28

ACS Paragon Plus Environment

Page 29 of 35

Journal of Agricultural and Food Chemistry

Table 2. Inhibited cell membrane permeability (%) in plant pathogens by chiral triticonazole enantiomers.

Gibberella

Rhizoctonia

Botrytis cinerea

Rhizoctonia

Alternaria

Botrytis cinerea

Fusarium

Sclerotinia

Pyricularia

zeae

cereali

(strawberry)

solani

solani

(tomato)

verticillioide

sclerotiorum

grisea

(R)-triticonazole

30.30

24.98

45.40

37.60

52.16

76.42

21.72

73.20

69.85

(Rac)-triticonazole

26.95

21.42

42.19

33.25

48.78

73.41

17.68

69.51

63.61

(S)-triticonazole

24.15

20.07

40.38

32.25

43.89

70.32

16.95

65.54

59.73

Untreated control

21.59

15.85

37.73

31.45

41.81

67.82

15.97

63.34

57.90

Chiral enanatiomers

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 1. The enantioselective inhibition of ergosterol biosynthesis by chiral triticonazole enantiomers in nine plant pathogens.

30

ACS Paragon Plus Environment

Page 30 of 35

Page 31 of 35

Journal of Agricultural and Food Chemistry

Figure 2. The relationship between the R/S ratio value of enantioselective bioactivity and the enantioselective inhibition of ergosterol.

31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3. The two-dimensional dimensional and three-dimensional enantiomer-specific specific binding modes for pure triticonazole enantiomers bound to CYP51b in five plant pathogens verticilllioide C Botrytis cinerea,, D Sclerotinia (A Gibberella zeae, B Fusarium verticilllioide, sclerotiorum, E Pyricularia grisea) grisea with high sequence similarity.

32

ACS Paragon Plus Environment

Page 32 of 35

Page 33 of 35

Journal of Agricultural and Food Chemistry

Figure 4. Time development of EF values for the triticonazole enantiomers in spinach (A) , cucumber (B), (B) Chinese cabbage (C) and tomato (D).

33

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

TOC Graphic:

34

ACS Paragon Plus Environment

Page 34 of 35

Page 35 of 35

Journal of Agricultural and Food Chemistry

509x285mm (150 x 150 DPI)

ACS Paragon Plus Environment