Characterization of Stilbene Synthase Genes in mulberry (Morus

Ma. Morus atropurpurea. Nt. Nicotiana tabacum. STS stilbene synthase. CHS chalcone synthase. CHI chalcone isomerase. 4CL. 4-coumarate:CoA ligase...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Newcastle, Australia

Article

Characterization of Stilbene Synthase Genes in mulberry (Morus atropurpurea) and Metabolic Engineering for the Production of Resveratrol in Escherichia coli Chuanhong Wang, Shuang Zhi, Changying Liu, Fengxiang Xu, Ai Chun Zhao, Xiling Wang, Yanhong Ren, Zhengang Li, and Maode Yu J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b05212 • Publication Date (Web): 07 Feb 2017 Downloaded from http://pubs.acs.org on February 9, 2017

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

Journal of Agricultural and Food Chemistry 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 36

Journal of Agricultural and Food Chemistry

1

Characterization of Stilbene Synthase Genes in mulberry (Morus atropurpurea)

2

and Metabolic Engineering for the Production of Resveratrol in Escherichia coli

3

Chuanhong Wang1, Shuang Zhi1, Changying Liu1, Fengxiang Xu1, Aichun Zhao1,

4

Xiling Wang1, Yanhong Ren1, Zhengang Li2, Maode Yu1*

5

1

6

District, Chongqing 400716, China

7

2

8

Agricultural Sciences, Mengzi, Yunnan 661100, China

College of Biotechnology, Southwest University, No.2 Tiansheng Road, BeiBei

The Sericultural and Apicultural Research Institute, Yunnan Academy of

9 10

* Corresponding author: Maode Yu

11

E-mail: [email protected]

12

Tel:+8618723079257

13

Fax: +86-023-68250191

14 15 16 17 18 19 20 21 22 23 24 25 26 27

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

28

Abstract

29

Stilbenes have been recognized for their beneficial physiological effects on

30

human health. Stilbene synthase (STS) is the key enzyme of resveratrol biosynthesis

31

and has been studied in numerous plants. Here, four MaSTS genes were isolated and

32

identified in mulberry (Morus atropurpurea Roxb.). The expression levels of MaSTS

33

genes and the accumulation of trans-resveratrol,

34

trans-mulberroside A were investigated in different plant organs. A novel

35

co-expression system that harbored 4-coumarate:CoA ligase gene (Ma4CL) and

36

MaSTS was established. Stress tests suggested that MaSTS genes participates in

37

responses to salicylic acid, abscisic acid, wounding and NaCl stresses. Additionally,

38

over-expressed MaSTS in transgenic tobacco elevated the trans-resveratrol level and

39

increased tolerance to drought and salinity stresses. These results revealed the major

40

MaSTS gene and we evaluated its function in mulberry, laying the foundation for

41

future research on stilbene metabolic pathways in mulberry.

42

Keywords: mulberry; stilbene synthase; stilbenes; resveratrol; co-expression

trans-oxyresveratrol and

43 44 45 46 47 48 49 50 51 52 2

ACS Paragon Plus Environment

Page 2 of 36

Page 3 of 36

Journal of Agricultural and Food Chemistry

53

Introduction

54

Stilbenes are non-flavonoid polyphenols, derived from the phenylpropanoid

55

pathway, that function as phytoalexins in plants to resist various biotic and abiotic

56

stresses, like UV radiation, bacteria, fungi and herbivores.1-4 To date, stilbenes have

57

been identified in more than 70 unrelated plant species, including grape, apple, peanut,

58

pine, rhubarb and sorghum.1,3,5,6 In recent decades, many studies on the physiological

59

functions of stilbenes in vivo/vitro indicated that stilbenes have significant

60

health-promoting effects on the body, such as the prevention of cancer, heart disease

61

and neurodegenerative diseases, and inhibiting α-glucosidase activities and tyrosinase

62

gene expression.7−10

63

Stilbene synthase (STS) is the pivotal enzyme in the synthesis of stilbenes. It

64

occurs in a limited number of plant species and utilizes a tetraketide intermediate that

65

condenses with three malonyl-CoA and one 4-coumaroyl-CoA molecules (Figure 1).

66

Interestingly, as another member of polyketide synthase III, chalcone synthase (CHS)

67

utilizes the same starter phenylpropanoid-CoA esters as STS (Figure 1). Moreover,

68

because CHS and STS have a high degree of similarity at the amino acid level,

69

numerous plant STS sequences are annotated as CHS in different public databases

70

based on sequence homology.11 However, these genes may in fact have other

71

metabolic roles, such as stilbene-forming activities.11,12

72

Mulberry leaves are widely known for their role in the silk production in Asian

73

countries. Human utilization of the mulberry–silkworm interaction began at least

74

5,000 years ago and greatly influenced human civilization.13 Different parts of the 3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

75

mulberry have been extensively investigated in recent years, and the plant is a good

76

source of distinct natural products that are able to positively impact human health,

77

including antioxidative, antihyperglycemic, hypolipidemic and antiatherogenic effects

78

and the inhibition of α-glucosidase activities.8,14 In addition, mulberry is rich in

79

resveratrol, oxyresveratrol and mulberroside A.8,15,16 Unfortunately, compared with

80

information on the bioactivity of stilbenes in mulberry, there are limited reports

81

related to MaSTS. Thus, it is necessary to study the functions of MaSTS genes in

82

mulberry.

83

To investigate the mechanism of MaSTS in mulberry, we determined the levels of

84

the three main stilbenes of ‘Guiyou No. 62’ (Morus atropurpurea Roxb.), and we

85

evaluated four MaSTS gene responses to salicylic acid (SA), abscisic acid (ABA),

86

wounding and NaCl stresses. Moreover, stilbenes in the fruit of a new cultivated

87

variety ‘Jialing No. 40’ (Morus atropurpurea Roxb.) (tetraploid, ‘Zhongsang5801’ ×

88

‘Naxi’, hypocotyl chromosome doubling, 2n = 4x = 56) were also investigated.

89

Additionally, we established a co-expression method to produce trans-resveratrol in

90

Escherichia coli. The tolerance of transgenic tobacco harboring MaSTS to multiple

91

abiotic stresses was also evaluated.

92

Materials and methods

93

Data retrieval and cloning of MaSTS cDNAs

94

Similarity searches were performed using the coding region of the Fallopia

95

multiflora STS (AGA35552.1), Arachis hypogaea STS (BAA78617.1), and

96

Polygonum cuspidatum STS (ACC76753.1) against the Morus Genome Database 4

ACS Paragon Plus Environment

Page 4 of 36

Page 5 of 36

Journal of Agricultural and Food Chemistry

97

(http://morus.swu.edu.cn/morusdb/). The candidate genes were identified using

98

BLASTN and SMART (http://smart.embl-heidelberg.de/). The purified PCR products

99

were cloned into the pMD19-T simple vector (Takara, Otsu, Japan) and sequenced.

100

Plant materials and abiotic stress test

101

The mulberry fruit materials were collected from the mulberry cultivar ‘Jialing

102

No.40’ at seven different developmental stages (S1–S7), and the leaves, stems bark

103

and roots bark were excised from mature plants of ‘Guiyou No. 62’ in the mulberry

104

garden of Southwest University. Abiotic stress tests were performed as previously

105

described.17 Briefly, 1 week-old mulberry seedlings were used for ABA (50 µM), SA

106

(5 mM) and NaCl (50 mg/L) treatments, and 10-week-old mulberry plants were used

107

for mechanical wounding treatments. The wounding treatment consisted of creating

108

eight wounds along leaf veins using a sterile toothpick. All of the materials were

109

immediately frozen in liquid nitrogen and stored at −80°C, and/or freeze-dried, for

110

RNA isolation and/or high performance liquid chromatography (HPLC) analysis,

111

respectively.

112

RNA extraction, cDNA synthesis and quantitative real-time polymerase chain

113

reaction (qRT-PCR)

114

The total RNA of mulberry fruit was extracted using an RNA Extraction TransZol

115

Plant Kit (TransGen Biotech, Beijing, China), and the other tissues’ total RNA were

116

extracted using an RNA Extraction Kit (TaKaRa, Dalian, China), The RNA samples

117

were treated with DNase I (TaKaRa) to digest genomic DNA, and 2 µg of purified

118

RNA was used to synthesize cDNA with Moloney murine leukemia virus reverse 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 36

119

transcriptase (Promega, Madison, WI, USA). Six-fold-diluted cDNA was used in

120

RT-PCR and qRT-PCR. The primers were designed using the online tool of the

121

GeneScript Company (Nanjing, China) (http://www.genscript.com.cn/index.html)

122

(Table S1 and Table S2). The qRT-PCR was performed according to the

123

manufacturer’s instructions for SYBR® Premix Ex TaqTM II (TaKaRa) and conducted

124

in the StepOne Real-Time PCR System (Applied Biosystems, Foster City, CA, USA).

125

To each reaction, 2 µL of diluted cDNA was added, and the MaActin3 gene was used

126

as an internal control to normalize the relative expression of target genes. All of the

127

data was analyzed using the 2−∆∆Ct method.17,22

128

Determination of stilbenes in different tissues

129

Different parts of the mulberry were collected and dried in a vacuum freeze drier

130

(Thermo Fisher Scientific, Waltham, MA, USA). Each sample was ground in liquid

131

nitrogen, 1 mL of 75% methanol was added per 0.1 g of powder, and extracted for 30

132

min three times by ultrasonication. The filtrate was evaporated to dryness and

133

dissolved in 5 mL of methyl alcohol. Each sample was centrifuged and filtered

134

through

135

trans-oxyresveratrol and trans-mulberroside A (MUST bio-technology Co., Ltd.

136

Chengdu, China) were determined using a Waters 2487 HPLC system (Waters,

137

Milford, MA, USA) equipped with a reverse-phase 5 µm C18 column (4.6 × 250 mm)

138

(InertSustain, Tokyo, Japan). The separation temperature was 30°C, and the detection

139

occurred at 330 nm. The flow rate was 0.8 mL/min, and 10 µL samples were injected.

140

The mobile phases were (A) acetonitrile and (B) 0.1% aqueous phosphoric acid. The

a

0.22

µm

filter.

Trans-resveratrol

(Sangon,

6

ACS Paragon Plus Environment

Shanghai,

China),

Page 7 of 36

Journal of Agricultural and Food Chemistry

141

solvent gradient elution program was as follows: 0–10 min, 11% A, 89% B; 10–15

142

min, 11–24% A, 89–76% B; 15–55 min, 24% A, 76% B; and 55–60 min, 24–11% A,

143

76–89% B. In addition, an ultra-performance liquid chromatography–mass

144

spectrometer (UPLC–MS) was also used. A Waters ACQuity UPLC I class system

145

was equipped with a BEH C18 1.7 µm column and TUV/QDa detector (Waters).

146

Negative scanning mass spectra were acquired over the range from 100 to 600 m/z.

147

Establishment of a co-expression system and fermentation tests

148

We used the co-expression methods described previously, with a few

149

modifications.17 Primers for vector construction are shown in Table S3. Ma4CL2

150

carried a BamHI site at its 5'-end, and PstI and XhoI sites at its 3'-end. It was first

151

cloned into the pET28a(+) (Novagen) at BamHI/XhoI sites, which resulted in the

152

pET4CL2 cassette. The pET28a(+) fragment from AGATCTCGATCCCGCGAA to

153

GGATCC (the BamHI site) was cloned and mutated from AGATCT to CTGCAG (a

154

PstI site) and mutated from GGATCC (the BamHI site) to GGTACC (the KpnI site). It

155

was then linked to the 5'-end of MaSTS3 by fusion PCR. The digested fusion fragment

156

was inserted into the PstI/XhoI sites of pET4CL2, resulting in pET4CL-T-STS. Thus,

157

MaSTS3 in pET4CL-T-STS could be replaced by other genes that carried KpnI/XhoI

158

sites. Then, we introduced a gene of uncertain function (accession: ALS20361), which

159

was identified as a MaSTS by bioinformatics methods, and we named it as MaSTS′ in

160

this study. Thus, Ma4CL2 and MaSTS3/MaSTS′ were preceded by a T7 promoter/lac

161

operator and a ribosome-binding site (Figure 2a, b).

162

The co-expression plasmid was transformed into E. coli BL21(DE3) pLysS 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

163

(Novagen). The cells were incubated in shake flasks on a rotary shaker at 37°C until

164

the OD600 reached ~0.6, induced with 0.1 mM isopropyl-β-D-thiogalactopyranoside

165

(IPTG), and cultivated at 25°C for 5 h to produce protein. Then, the cells were

166

harvested at 5,000 ×g for 5 min and resuspended in M9 medium containing 1 mM

167

p-coumaric acid, 0.1 mM malonyl-CoA lithium salt, 0.1 mM IPTG and 50 mg/L

168

kanamycin (Kan). Fermentation was continued for 60 h at 25°C. The cells in the

169

fermentation liquor were lysed by sonication and concentrated by a rotary evaporator.

170

Subsequently, it was extracted three times with equal volumes of ethyl acetate and

171

dried by a rotary evaporator. It was then dissolved in methanol and analyzed by HPLC

172

and UPLC-MS.

173

Construction and transformation of the recombinant plasmid for tobacco

174

transformations

175

Full-length MaSTS3 cDNA fragments were cloned into the BamHI and SpeI sites

176

of pLGNL (conserved in our laboratory), which contained the cauliflower mosaic

177

virus 35S promoter, to generate a pLGNL-MaSTS3 over-expression cassette (Figure

178

2c). For tobacco transformations, Agrobacterium tumefaciens containing the plasmid

179

was transformed into tobacco (Nicotiana tabacum L.) plants using the leaf disc

180

method.18 Transgenic tobacco was selected on 1/2 Murashige and Skoog medium

181

containing 50 mg/L Kan. The positive plants were confirmed by β-glucuronidase

182

staining, and genomic PCR and qRT-PCR analyses.

183

Physiological and abiotic stress tolerance analysis of transgenic tobaccos

184

The T1 transgenic lines and wild type (WT) tobacco samples were cultivated in a 8

ACS Paragon Plus Environment

Page 8 of 36

Page 9 of 36

Journal of Agricultural and Food Chemistry

185

climate chamber (27°C, 14 h day/10 h night, 8,000 Lx). After growing ~3 weeks, they

186

were subjected to 40°C, salt (400 mM NaCl) and drought [20% polyethylene glycol

187

(PEG) 6000] stresses, independently, for ~2 weeks. Leaves from the same positions

188

were collected for free proline and malonaldehyde (MDA) content measurements. The

189

proline content analysis was performed according to a previously published method19

190

in which 0.5 g of fresh leaves were cut into small pieces, homogenized in 5 mL of 3%

191

sulfosalicylic acid, boiled 10 min and centrifuged at 4,000 ×g for 10 min. A 2 mL

192

extract was incubated with 2 mL of ninhydrin reagent, which contained 2.5% (w/v)

193

ninhydrin, 40% 6 M phosphoric acid and 60% (v/v) glacial acetic acid, and 2 mL of

194

glacial acetic acid, and then it was continuously boiled for 30 min. After the reaction

195

was terminated in an ice bath, 4 mL of toluene was added and vortexed. The reaction

196

mixture was centrifuged at 3,000 ×g for 5 min, and the absorbance of the supernatant

197

at 520 nm was determined using a spectrometer (TECHCOMP, Shanghai, China).

198

The MDA content was determined as described by previous research.20 Briefly,

199

0.5 g of fresh tobacco leaves was homogenized in 5 mL 10% trichloroacetic acid and

200

centrifuged at 12,000 ×g for 10 min. Then, 2 mL 0.6% thiobarbituric acid dissolved in

201

10% trichloroacetic acid was added to 2 mL of the supernatant. The mixture was

202

boiled for 20 min and then terminated in an ice bath. Then, it was centrifuged at

203

12,000 ×g for 3 min, and the absorbance of the supernatant was determined at 532,

204

450 and 600 nm. The MDA content was calculated as described previously.20

205

To analyze the effects of MaSTS3 on other genes in transgenic tobacco (N.

206

tabacum), the transcription of Nt4CL1 (U50845), Nt4CL2 (U50846) and NtCHS 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

207

(XM_016638898 and XM_016634418) was determined by qRT-PCR (primers are

208

listed in Table S4). In addition, to analyze the metabolites of transgenic tobacco,

209

extracts were prepared from each line and extracted using the method described

210

above.

211

Statistical analyses

212

All experiments were performed in triplicate, and the results were expressed as

213

means ± standard deviation (SD). The statistical analysis was performed using SPSS

214

Statistics 18.0 (SPSS Inc., Chicago, IL, USA) with Duncan's multiple range tests.

215

Figures were drawn using OriginPro 7.5 (OriginLab, Northampton, MA, USA). Mean

216

values that were significantly different within treatments were designated with an

217

asterisk.

218

Results

219

Expression of MaSTS genes in different tissues and the contents of three stilbenes

220

in different tissues

221

Based on a multiple sequence alignment against the Morus genome and NCBI

222

databases, four candidate STS genes were selected, and they were identified by

223

cloning their corresponding cDNAs, and named as MaSTS1, MaSTS2, MaSTS3 and

224

MaSTS4, respectively (File S1). To investigate the relationships between the

225

expression levels of MaSTS genes and the accumulations of stilbenes, transcription

226

levels of MaSTS genes were quantitatively measured by qRT-PCR, and the contents of

227

the three stilbenes were determined by HPLC (Figure 3). An expression analysis

228

revealed that MaSTS genes were expressed in various mulberry tissues, but exhibited 10

ACS Paragon Plus Environment

Page 10 of 36

Page 11 of 36

Journal of Agricultural and Food Chemistry

229

significant differences in their magnitudes of expression (Figure 4, a1). MaSTS1 was

230

mainly expressed in root bark, stem bark and old leaves, while MaSTS3 was highly

231

expressed in root bark, stem bark and branch bark. MaSTS2 and MaSTS4 were highly

232

expressed in root bark. Additionally, all of the MaSTS genes were expressed at

233

relatively low levels in young leaves. Using the method above, the contents of the

234

three stilbenes were determined. The mulberroside A content was greater than the

235

oxyresveratrol and resveratrol contents in all of the selected tissues (Figure 4, a2–a4).

236

The resveratrol content showed a similar pattern as mulberroside A, although

237

resveratrol was not detected in young leaves (Figure 4, a2). Oxyresveratrol had a high

238

content in root bark and branch bark (Figure 4, a3). Moreover, mulberroside A was the

239

most abundant in root bark, reaching 9.09 mg/g of dry weight (mg/g DW), followed

240

by stem bark and branch bark (Figure 4, a4).

241

The dynamics of the three stilbene’s contents during mulberry fruit development

242

from S1 to S7 (Figure 4, b1) were determined. As in other tissues, mulberroside A was

243

the main stilbene form present. However, resveratrol existed throughout fruit

244

development and reached its maximum of 0.06 (mg/g DW) at S7 (Figure 4, b2). The

245

mulberroside A content’s trend was identical to that of oxyresveratrol, and they

246

reached their maximum levels of 0.802 (mg/g DW) and 0.084 (mg/g DW),

247

respectively, at S5 (Figure 4, b3 and b4). The four MaSTS showed expression

248

differences during fruit development (Figure 4, c1–c4). MaSTS1 increased with the

249

fruit developmental stage and reached its maximum at S7, although it was transiently

250

down-regulated at S5 (Figure 4, c1). MaSTS2 and MaSTS3 were lowly expressed at 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

251

the early stages, but their expression increased sharply and reached maximum levels

252

at S6 and S4, respectively (Figure 4, c2 and c3). MaSTS4 performed irregularly and

253

peaked at S3 (Figure 4, c4).

254

Fermentation test

255

The recombined plasmids of pET4CL2, pET4CL-T-STS3 and pET4CL-T-STS′

256

were fermented in E. coli BL21 (Figure 5a). The color of the fermentation broth

257

showed obvious differences after 60 h at 25°C. The broth of pET4CL2 was milky

258

white, while pET4CL-T-STS3 was canary yellow and pET4CL-T-STS′ was yellow

259

(Figure 5b), indicating that the function of MaSTS3 was different from that of MaSTS′.

260

In addition, about 0.187 mg/L trans-resveratrol was produced in the pET4CL-T-STS3

261

fermentation broth, and naringenin was detected in the pET4CL-T-STS′ ethyl acetate

262

extract (Figure 5c).

263

Dynamic expression of MaSTS genes under abiotic stresses

264

MaSTS genes were sensitive to ABA treatments. MaSTS1 and MaSTS2 showed a

265

“W-shaped” expression pattern, reached their minimum values at 1 h and 12 h,

266

respectively, and peaked at 3 h and 24 h, respectively (Figure 6 a, b). MaSTS3 and

267

MaSTS4 showed a completely different expression pattern. MaSTS3 decreased

268

immediately and maintained a low expression level compared with the control (Figure

269

6 c). However, MaSTS4 increased immediately after ABA treatment and maintained a

270

high level (Figure 6 d).

271

Under the SA treatment, MaSTS1, MaSTS2 and MaSTS3 showed similar

272

expression patterns, peaking at 6 h and then being down-regulated (Figure 6 a–c). 12

ACS Paragon Plus Environment

Page 12 of 36

Page 13 of 36

Journal of Agricultural and Food Chemistry

273

However, MaSTS4 showed a different pattern, reaching its maximum level at 3 h

274

(Figure 6, d).

275

For the NaCl treatment, the four MaSTS genes showed different reactions.

276

MaSTS1 and MaSTS2 increased immediately and peaked at 3 h and 1 h, respectively

277

(Figure 6 a, b). Then, MaSTS1 decreased gently, while MaSTS2 increased once again

278

after 12 h and reached its maximum at 24 h. MaSTS3 showed a “U-shaped”

279

expression pattern, which decreased sharply before 12 h, then increased continuously

280

and finally reached a maximum at 24 h (Figure 6 c). However, MaSTS4 was

281

expressed at a low level for 12 h but then greatly increased, reaching a maximum that

282

was 14.03-fold greater than that of the control at 24 h (Figure 6 d).

283

The MaSTS genes exhibited various responses to the wounding treatment.

284

MaSTS1 and MaSTS2 showed similar expression patterns and peaked at 6 h. Their

285

peak values were 52.93- and 50.91-fold greater, respectively, than that of the control

286

(Figure 6 e). Interestingly, MaSTS3 and MaSTS4 also showed similar expression

287

patterns but reached maximum levels at 12 h. However, MaSTS4’s level was more

288

than 61.09-fold greater than that of the control, while MaSTS3’s level was only

289

9.46-fold greater (Figure 6 e).

290

Analysis of other genes regulated by MaSTS3 and the enhanced tolerance of

291

transgenic tobacco to multiple abiotic stresses

292

STS and CHS in plants used the same substrates (Figure 1); therefore, we

293

selected upstream and downstream genes of CHS in N. tabacum to determine the

294

genes that were regulated by MaSTS3. The expression level of MaSTS3 in six 13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

295

transgenic lines was quantified using qRT-PCR. Three differently expressing

296

transgenic lines were selected for the next test. The qRT-PCR showed the expression

297

levels of MaSTS3 were significantly higher than WT (Figure 7, a1). Nt4CL1 in

298

over-expression line 4 (OE4) and OE5 was up-regulated, but there was no obvious

299

change in OE6 (Figure 7, a2). Interestingly, Nt4CL2 in WT was significantly

300

up-regulated, and its expression trend was similar to that of MaSTS3 (Figure 7, a3).

301

However, the expression levels of NtCHS in OE4 and OE6 were not different from

302

that of WT, while the expression of NtCHS in OE5 increased (Figure 7, a4).

303

Additionally, the total flavonoid contents in transgenic tobacco lines were also

304

determined. The flower color was not significantly different between the transgenic

305

tobacco and the WT (Figure 7, b1). The total flavonoid contents in transgenic tobacco

306

lines showed less obvious reductions compared with the control group (Figure 7, b2).

307

Additionally, HPLC chromatograms revealed trans-resveratrol was accumulated, with

308

45.167 (µg/g FW), which was not presented in the WT (Figure 7, b3).

309

Three transgenic tobacco lines and WT seedlings were subjected to various

310

abiotic stresses to characterize the functions of MaSTS3 under NaCl, heat and PEG

311

stress. After ~2-week treatments, transgenic lines and WT showed some differences

312

compared with the control (Figure 7, b4 and c1). WT showed more sensitivity than

313

transgenic lines to salt and heat treatments, and the leaves started to turn yellow.

314

However, there were no evident morphological differences between the transgenic

315

line and the WT seedlings under drought stress. The proline contents of OE4 and OE5

316

were higher than that of WT under the PEG treatment, although OE6 showed no 14

ACS Paragon Plus Environment

Page 14 of 36

Page 15 of 36

Journal of Agricultural and Food Chemistry

317

significant change compared with WT (Figure 7, c2). Additionally, the proline

318

contents of all of the transgenic lines were higher than that of WT under the NaCl and

319

heat treatments (Figure 7, c3 and c4). The MDA content was also determined. All of

320

the transgenic lines had lower MDA contents than the WT under PEG and NaCl

321

treatments (Figure 7, d1–d3). However, the MDA contents of OE5 and OE6 were

322

higher than WT under the heat treatment (Figure 7, d4).

323

Discussion

324

Mulberry is widely distributed in China and contains more phenolic and flavonoid

325

compounds than some other fruits and vegetables.21 The root, branch bark and leaves

326

of mulberry are used as traditional medicinal materials, and the fruit is used as

327

nutritional foodstuff.22 In addition, mulberry is well known for its capacity to resist

328

harsh environments. It can grow in areas severely affected by desertification,

329

including sand damage, drought and saline stress.

330

Although all of the MaSTS genes were detected in selected tissues, they showed

331

obvious differences. MaSTS3 was highly expressed in root bark, stem bark and branch

332

bark. In addition, the expression levels of MaSTS3 were in accordance with the

333

stilbene contents of select tissues. The expression trend of MaSTS3 was also similar to

334

the trend of the stilbene contents during fruit developmental stages. Thus, MaSTS3

335

was the major STS gene in mulberry.

336

Mulberroside A was the major stilbene in mulberry, followed by oxyresveratrol.

337

Nevertheless, the parent nucleus had the lowest level of trihydroxystilbene resveratrol

338

in the different tissues. This was similar to the results of other studies.23-25 The major 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

339

stilbene, 2,3,5,4'-tetra-hydroxy-stilbene-2-O-β-D-glucoside, was abundant in the

340

rhizomes and old stems of F. multiflora.24 A similar situation occurred in the genus

341

Picea, in which the stilbene glucosides astringin and isorhapontin were at high

342

concentrations in the roots and bark.23,25 In vitro and in vivo assays revealed that STS

343

catalyzed the condensation of one 4-coumaroyl-CoA and three molecules of

344

malonyl-CoA to form the trihydroxystilbene resveratrol.11,23 However, over-expressed

345

PaSTS1 in transgenic Norway spruce showed that significantly higher amounts of the

346

tetrahydroxystilbene glycosides isorhapontin and astringin were produced.23 Similarly,

347

over-expressed SbSTS in Arabidopsis tt4 mutants could lead to the accumulation of

348

cis-resveratrol glucoside (piceid), which is the major stilbene in the transgenic lines.26

349

Those results suggested that the first step in the biosynthesis of stilbenes in plants was

350

the formation of resveratrol, which was then further modified by hydroxylation,

351

O-methylation and O-glucosylation.23,25,26 Additionally, previous studies showed that

352

trans-resveratrol accumulates in the early stages of seedling roots. Additionally, the

353

contents of trans-oxyresveratrol and trans-mulberroside A increased while resveratrol

354

decreased during mulberry early development.27 Therefore, we speculate that

355

oxyresveratrol and mulberroside A are probably transformed from resveratrol through

356

oxidation, O-methylation or other derivatizations in mulberry.

357

The resveratrol STS genes were originally described in peanuts and grapes,11,28

358

which accumulate elevated levels of the stilbene following biotic and abiotic stresses,

359

including protecting plants against fungal invasions.23,29 SA is an important signal

360

molecule in the plant host’s defense response process, and the endogenous SA levels 16

ACS Paragon Plus Environment

Page 16 of 36

Page 17 of 36

Journal of Agricultural and Food Chemistry

361

were involved in the activation of pathogenesis-related gene expression.30 However,

362

all of the MaSTS genes participated in the up/down-regulatory processes, indicating

363

that MaSTS genes are involved in feedback regulation when stimulated by SA. This is

364

similar to previous research on the expression pattern of Vitis’ STS under powdery

365

mildew infection.28,29 SA signaling in plant defense is part of a complex network,

366

including feedback loops.31-33 Therefore, we speculated that exogenous SA influences

367

the endogenous SA levels at the first step, then triggers MaSTS feedback loops in the

368

next process.

369

As another important signaling molecule involved in the plant immune response,

370

ABA improves plant tolerances to salt and dehydration.34-36 In plant cells, salt and

371

drought lead to increased reactive oxygen species levels, which then stimulate the

372

synthesis and accumulation of ABA in roots.37,38 Here, MaSTS genes were sensitive to

373

ABA stress. In particular, MaSTS3 showed a complicated expression level that was

374

always down-regulated while MaSTS4 was always up-regulated. The recent

375

immunolocalization of STS revealed that stilbene biosynthesis takes place within the

376

cell wall.39,40 Interestingly, the upstream Ma4CLs were up-regulated significantly to

377

synthesize lignin when mulberry were stimulated by ABA, even though Ma4CL was

378

related to flavonoid synthesis in normal times.17 The process reduced the substrate

379

level of MaSTS, which may explain the genes’ down-regulation. In addition, the

380

expression of MaSTS genes under NaCl stress differed from that under ABA stress.

381

This implies that MaSTS genes respond to salt stress through an ABA-independent

382

signal transduction pathway. 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

383

When plants suffer from mechanical damage, the cell walls of the damaged sites

384

are first strengthened by crosslinking proteins to prevent dehydration and possible

385

pathogen infections, and then phenylpropanoid derivatives are required to be

386

synthesized in the subsequent step.17,41 As an outstanding phytoalexin, stilbene is a

387

potential fungicidal agent.4,25 This may explain why the wounding treatment led to a

388

high induction of MaSTS genes.

389

STS was frequently used to modify plant secondary metabolism to elevate the

390

self-defense capacity or the nutritional quality of crops.6,42,43 However, most reports

391

focused on increased tolerance against microbial pathogens.43,44 Transgenic tobacco

392

improved the tolerance capability, at different magnitudes, to PEG and NaCl

393

treatments, as evidenced by the higher proline and the lower MDA contents compared

394

with in WT. Additionally, over-expressed MaSTS3 genes modified the transcription of

395

endogenous genes in transgenic tobacco. Interestingly, the expression pattern of

396

Nt4CL2 was in accordance with that of MaSTS3. Thus, Nt4CL2 was associated with

397

stilbene synthesis, and the result was consistent with previous research showing that

398

Nt4CL2 was more likely to biosynthesize non-lignins.45 Additionally, NtCHS

399

expression was distinctly different among the transgenic lines, with only OE5 being

400

up-regulated, while OE4 and OE6 remained unchanged, compared with WT. As

401

described previously, CHS utilizes the same substrates as STS;11 therefore, the

402

over-expression of MaSTS3 may result in foreign MaSTS3 competing with

403

endogenous NtCHS for a limited substrate. However, the expression levels of

404

MaSTS3 in OE4 and OE6 were higher than in WT but lower than in OE5, which 18

ACS Paragon Plus Environment

Page 18 of 36

Page 19 of 36

Journal of Agricultural and Food Chemistry

405

meant that the substrate content was enough for both MaSTS3 and NtCHS. This

406

would eliminate the competition between MaSTS3 and NtCHS. Previously, plants

407

were transformed with STS genes from different plants, leading to an accumulation of

408

resveratrol or its derivatives. For instance, transformations of Arachis hypogea44 and

409

grapevine46 STS genes led to the accumulation of resveratrol in tobacco. Other

410

STS-coding genes have also been transformed to Arabidopsis thaliana, such as

411

SbSTS1 (Sorghum bicolor)11 and PcRS (Polygonum cuspidatum),43 which resulted in

412

piceid accumulation. As expected, HPLC chromatograms of the extracts from

413

transgenic tobaccos had obvious peaks that were not detected in the extracts of the

414

WT plants, and UPLC–MS detected a resveratrol (227 m/z) signal. This indicated that

415

trans-resveratrol was accumulated after MaSTS3 over-expression.

416

Resveratrol has been recognized for its benefits to human health. Currently, an

417

increasing demand for resveratrol for cosmetic, nutraceutical, and putative

418

pharmaceutical uses makes its production from a sustainable source a necessity.47

419

However, because of the low content in natural products, the environmental pollution

420

and disruption caused by the excessive use of raw materials has become increasingly

421

serious.17,48 It is necessary to establish a reliable alternative method of resveratrol

422

production under controlled conditions.47 As indicated by the stilbene contents

423

determined above, there is a highly efficient stilbene production system in mulberry,

424

which can be simulated in vitro to produce resveratrol. We built a recombinant

425

plasmid harboring two promoter/lac operators to independently regulate Ma4CL2 and

426

MaSTS3. Although the use of biotechnology through recombinant bacteria has been 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

427

reported,49,50 the methods did not suit this study because a soluble Ma4CL protein

428

could not be obtained.17 Additionally, CHS is highly similar to STS, leading to

429

numerous plant STS being annotated as CHS genes in different public databases.5 The

430

most used method for identifying CHS/STS is the determination of the product by

431

expressing it in transgenic plants or by enzymatic reactions in vitro.11,26 However, the

432

substrates are not easy to produce. Thus, we established this system to conveniently

433

distinguish CHS/STS. We improved the recombinant plasmid, adding a KpnI cloning

434

site at the 5′ of MaSTS3 that easily allows the insertion of other STS/CHSs. In our

435

previous study, we identified MaSTS′ genes by bioinformatics method. However, our

436

data demonstrate that ALS20361, in fact, encodes a CHS enzyme.

437

‘Jialing No. 40’, which is a new polyploidy variety of fruit-producing mulberry,

438

was bred in our laboratory in recent years,22 and it possesses a high product yield of

439

good quality fruit and leaves. In conclusion, the fruit of ‘Jialing No. 40’ not only have

440

a high yield but are also rich in stilbenes. Four MaSTS genes were identified in this

441

study, and the stress tests suggested that MaSTS3 genes participate in a series of

442

abiotic and biotic stresses. Genetic transformation tests indicated that MaSTS3 could

443

accumulate trans-resveratrol and improve tobacco tolerance to abiotic stresses.

444

Furthermore, the novel co-expression system of Ma4CL2 and MaSTS3 was

445

established, and trans-resveratrol was successful produced by fermentation. In

446

particular, the co-expression system could be used to identify the functions of

447

CHS/STS. This study provides the basis for future research on the stilbene metabolic

448

pathway in mulberry. 20

ACS Paragon Plus Environment

Page 20 of 36

Page 21 of 36

Journal of Agricultural and Food Chemistry

449

Abbreviations Used Ma Nt STS CHS CHI 4CL RT-PCR qRT-PCR ABA SA WT OE DW FW IPTG Kan

Morus atropurpurea Nicotiana tabacum stilbene synthase chalcone synthase chalcone isomerase 4-coumarate:CoA ligase reverse transcription polymerase chain reaction real-time quantitative polymerase chain reaction abscisic acid salicylic acid wild type over-expression dry weight fresh weight isopropyl-β-D-thiogalactopyranoside kanamycin

450

451 452

Acknowledgements We thank associate professor Li Xu and Haipeng Lu for providing HPLC

453

detection, and Hu Chen for providing UPLC–MS detection.

454

Funding

455

The work was funded by the China Special Fund for Agro-scientific Research in

456

the Public Interest (grant No. 201403064), Fundamental Research Funds for the

457

Central Universities (grant No. XDJK2016D024), the China Agriculture Research

458

System (grant No.CARS-22), and the National Natural Science Foundation of China

459

(grant No. 31360190)

460

Notes 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

461

462 463 464

The authors declare no competing financial interest.

Supporting Information Table S1. The primers used to isolate the MaSTS genes in Jialing No.40/ Guiyou62.

465

Table S2. qRT-PCR Primers for MaSTS genes.

466

Table S3. Oligonucleotides for vector construction in this study.

467

Table S4. qRT-PCR Primers for transgenic tobaccos.

468

File S1. The cDNA sequences of MaSTS genes isolated from Jialing No.40/

469

Guiyou No. 62.

470 471 472 473 474 475 476 477 478 479 480 481 482 22

ACS Paragon Plus Environment

Page 22 of 36

Page 23 of 36

Journal of Agricultural and Food Chemistry

483

References

484 485 486 487 488 489 490 491 492 493 494 495 496 497 498 499 500 501 502 503 504 505 506 507 508 509 510 511 512 513 514 515 516 517 518 519 520 521 522 523 524 525

1.

Chong, J.; Poutaraud, A.; Hugueney, P., Metabolism and roles of stilbenes in plants. Plant Science

2009, 177, 143-155. 2.

Dixon, R. A.; Paiva, N. L., Stress-lnduced Phenylpropanoid Metabolism. Plant Cell 1995, 7,

1085-1097. 3.

Aisyah, S.; Gruppen, H.; Slager, M.; Helmink, B.; Vincken, J. P., Modification of Prenylated

Stilbenoids in Peanut (Arachis hypogaea) Seedlings by the Same Fungi That Elicited Them: The Fungus Strikes Back. J. Agric. Food Chem. 2015, 63, 9260-8. 4.

He, D.; Jian, W.; Liu, X.; Shen, H.; Song, S., Synthesis, biological evaluation, and

structure-activity relationship study of novel stilbene derivatives as potential fungicidal agents. J. Agric. Food Chem. 2015, 63, 1370-7. 5.

Lo, C.; Coolbaugh, R. C.; Nicholson, R. L., Molecular characterization and in silico expression

analysis of a chalcone synthase gene family in Sorghum bicolor. Physiol. Mol. Plant P. 2002, 61, 179-188. 6.

Farneti, B.; Masuero, D.; Costa, F.; Magnago, P.; Malnoy, M.; Costa, G.; Vrhovsek, U.; Mattivi, F.,

Is there room for improving the nutraceutical composition of apple? J. Agric. Food Chem. 2015, 63, 2750-9. 7.

Hsieh, T. C.; Wang, Z.; Deng, H.; Wu, J. M., Identification of glutathione sulfotransferase-pi

(GSTP1) as a new resveratrol targeting protein (RTP) and studies of resveratrol-responsive protein changes by resveratrol affinity chromatography. Anticancer Res. 2008, 28, 29-36. 8.

He, H.; Lu, Y. H., Comparison of inhibitory activities and mechanisms of five mulberry plant

bioactive components against alpha-glucosidase. J. Agric. Food Chem. 2013, 61, 8110-9. 9.

Seyed, M. A.; Jantan, I.; Bukhari, S. N.; Vijayaraghavan, K., A Comprehensive Review on the

Chemotherapeutic Potential of Piceatannol for Cancer Treatment, with Mechanistic Insights. J. Agric. Food Chem. 2016, 64, 725-37. 10. Gomez-Zorita, S.; Fernandez-Quintela, A.; Lasa, A.; Aguirre, L.; Rimando, A. M.; Portillo, M. P., Pterostilbene, a dimethyl ether derivative of resveratrol, reduces fat accumulation in rats fed an obesogenic diet. J. Agric. Food Chem. 2014, 62, 8371-8. 11. Yu, C. K.; Springob, K.; Schmidt, J.; Nicholson, R. L.; Chu, I. K.; Yip, W. K.; Lo, C., A Stilbene Synthase Gene (SbSTS1) Is Involved in Host and Nonhost Defense Responses in Sorghum. Plant Physiol. 2005, 138, 393-401. 12. Springob, K.; Nakajima, J. I.; Yamazaki, M.; Saito, K., Recent Advances in the Biosynthesis and Accumulation of Anthocyanins. Nat. Prod. Rep. 2003, 20, 288-303. 13. He, N.; Zhang, C.; Qi, X.; Zhao, S.; Tao, Y.; Yang, G.; Lee, T. H.; Wang, X.; Cai, Q.; Li, D., Draft genome sequence of the mulberry tree Morus notabilis. Nat. Commun. 2013, 4, 2445-2445. 14. Harauma, A.; Murayama, T.; Ikeyama, K.; Sano, H.; Arai, H.; Takano, R.; Kita, T.; Hara, S.; Kamei, K.; Yokode, M., Mulberry leaf powder prevents atherosclerosis in apolipoprotein E-deficient mice. Biochem. Bioph. Res. Co. 2007, 358, 751-6. 15. Mei, M.; Ruan, J. Q.; Wu, W. J.; Zhou, R. N.; Lei, J. P.; Zhao, H. Y.; Yan, R.; Wang, Y. T., In vitro pharmacokinetic characterization of mulberroside A, the main polyhydroxylated stilbene in mulberry (Morus alba L.), and its bacterial metabolite oxyresveratrol in traditional oral use. J. Agric. Food Chem. 2012, 60, 2299-308. 16. Song, W.; Wang, H. J.; Bucheli, P.; Zhang, P. F.; Wei, D. Z.; Lu, Y. H., Phytochemical profiles of 23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

526 527 528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569

different mulberry (Morus sp.) species from China. J. Agric. Food Chem. 2009, 57, 9133-40. 17. Wang, C.-H.; Yu, J.; Cai, Y.-X.; Zhu, P.-P.; Liu, C.-Y.; Zhao, A.-C.; Lu, R.-H.; Li, M.-J.; Xu, F.-X.; Yu, M.-D., Characterization and Functional Analysis of 4-Coumarate:CoA Ligase Genes in Mulberry (vol 11, e0155814, 2016). Plos One 2016, 11. 18. Voelker T, S. A., Chrispeels MJ., Differences in expression between two seed lectin alleles obtained from normal and lectin-deficient beans are maintained in transgenic tobacco. Embo J. 1987, 6, 3571-3578. 19. Bates, L. S.; Waldren, R. P.; Teare, I. D., Rapid determination of free proline for water-stress studies. Plant Soil 1973, 39, 205-207. 20. Hodges, D. M.; DeLong, J. M.; Forney, C. F.; Prange, R. K., Improving the thiobarbituric acid-reactive-substances assay for estimating lipid peroxidation in plant tissues containing anthocyanin and other interfering compounds. Planta 1999, 207, 604-611. 21. Lin, J. Y.; Tang, C. Y., Determination of total phenolic and flavonoid contents in selected fruits and vegetables, as well as their stimulatory effects on mouse splenocyte proliferation. Food Chem. 2007, 101, 140-147. 22. Liu, C.; Zhao, A.; Zhu, P.; Li, J.; Han, L.; Wang, X.; Fan, W.; Lu, R.; Wang, C.; Li, Z.; Lu, C.; Yu, M., Characterization and expression of genes involved in the ethylene biosynthesis and signal transduction during ripening of mulberry fruit. PLoS One 2015, 10, e0122081. 23. Hammerbacher, A.; Ralph, S. G.; Bohlmann, J.; Fenning, T. M.; Gershenzon, J.; Schmidt, A., Biosynthesis of the major tetrahydroxystilbenes in spruce, astringin and isorhapontin, proceeds via resveratrol and is enhanced by fungal infection. Plant Physiol. 2011, 157, 876-90. 24. Sheng, S. J.; Liu, Z. Y.; Wei, Z.; Li, S.; Zhao, S. J., Molecular analysis of a type III polyketide synthase gene in Fallopia multiflora. Biologia 2010, 65, 939-946. 25. Hammerbacher, A.; Schmidt, A.; Wadke, N.; Wright, L. P.; Schneider, B.; Bohlmann, J.; Brand, W. A.; Fenning, T. M.; Gershenzon, J.; Paetz, C., A common fungal associate of the spruce bark beetle metabolizes the stilbene defenses of Norway spruce. Plant Physiol. 2013, 162, 1324-36. 26. Yu, C. K.; Lam, C. N.; Springob, K.; Schmidt, J.; Chu, I. K.; Lo, C., Constitutive accumulation of cis-piceid in transgenic Arabidopsis overexpressing a sorghum stilbene synthase gene. Plant Cell Physiol. 2006, 47, 1017-21. 27. Zhou, J.; Li, S. X.; Wang, W.; Guo, X. Y.; Lu, X. Y.; Yan, X. P.; Huang, D.; Wei, B. Y.; Cao, L., Variations in the levels of mulberroside A, oxyresveratrol, and resveratrol in mulberries in different seasons and during growth. The Scientific World J. 2013, 2013, 380692-380692. 28. Wiese, W.; Vornam, B.; Krause, E.; Kindl, H., Structural organization and differential expression of three stilbene synthase genes located on a 13 kb grapevine DNA fragment. Plant Mol. Biol. 1994, 26, 667-77. 29. Xu, W.; Yu, Y.; Zhou, Q.; Ding, J.; Dai, L.; Xie, X.; Xu, Y.; Zhang, C.; Wang, Y., Expression pattern, genomic structure, and promoter analysis of the gene encoding stilbene synthase from Chinese wild Vitis pseudoreticulata. J. Exp. Bot. 2011, 62, 2745-61. 30. Shah, J., The salicylic acid loop in plant defense. Curr. Opin. Plant Biol. 2003, 6, 365-371. 31. Wildermuth, M. C.; Dewdney, J.; Wu, G.; Ausubel, F. M., Isochorismate synthase is required to synthesize salicylic acid for plant defence. Nature 2001, 414, 562-5. 32. Verberne, M. C.; Verpoorte, R.; Bol, J. F.; Mercado-Blanco, J.; Linthorst, H. J., Overproduction of salicylic acid in plants by bacterial transgenes enhances pathogen resistance. Nat. Biotechnol. 2000, 18, 779-83. 24

ACS Paragon Plus Environment

Page 24 of 36

Page 25 of 36

Journal of Agricultural and Food Chemistry

570 571 572 573 574 575 576 577 578 579 580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613

33. Shah, J.; Kachroo, P.; Klessig, D. F., The Arabidopsis ssi1 mutation restores pathogenesis-related gene expression in npr1 plants and renders defensin gene expression salicylic acid dependent. Plant Cell 1999, 11, 191-206. 34. Schroeder, J. I.; Kwak, J. M.; Allen, G. J., Guard cell abscisic acid signalling and engineering drought hardiness in plants. Nature 2001, 410, 327-30. 35. Zhu, J. K., Salt and Drought Stress Signal Transduction in Plants. Annu. Rev. Plant Biol. 2002, 53, 247-73. 36. Cutler, S. R.; Rodriguez, P. L.; Finkelstein, R. R.; Abrams, S. R., Abscisic Acid: Emergence of a Core Signaling Network. Annu. Rev. Plant Biol. 2010, 61, 651-79. 37. Jiang, M.; Zhang, J., Effect of abscisic acid on active oxygen species, antioxidative defence system and oxidative damage in leaves of maize [Zea mays] seedlings. Plant Cell Physiol. 2001, 42, 1265-1273. 38. Jiang, M.; Zhang, J., Water stress-induced abscisic acid accumulation triggers the increased generation of reactive oxygen species and up-regulates the activities of antioxidant enzymes in maize leaves. J Exp. Bot. 2002, 53, 2401-10. 39. Pan, Q. H.; Lei, W.; Li, J. M., Amounts and subcellular localization of stilbene synthase in response of grape berries to UV irradiation. Plant Science 2009, 176, 360-366. 40. Fornara, V.; Onelli, E.; Sparvoli, F.; Rossoni, M.; Aina, R.; Marino, G.; Citterio, S., Localization of stilbene synthase in Vitis vinifera L. during berry development. Protoplasma 2008, 233, 83-93. 41. Bruxelles, G. L. d.; Roberts, M. R., Signals Regulating Multiple Responses to Wounding and Herbivores. Crit. Rev. Plant Sci. 2001, 20, 487-521. 42. Schwekendiek, A.; Spring, O.; Heyerick, A.; Pickel, B.; Pitsch, N. T.; Peschke, F.; De, K. D.; Weber, G., Constitutive expression of a grapevine stilbene synthase gene in transgenic hop (Humulus lupulus L.) yields resveratrol and its derivatives in substantial quantities. J. Agric. Food Chem. 2007, 55, 7002-9. 43. Liu, Z.; Zhuang, C.; Sheng, S.; Shao, L.; Zhao, W.; Zhao, S., Overexpression of a resveratrol synthase gene (PcRS) from Polygonum cuspidatum in transgenic Arabidopsis causes the accumulation of trans-piceid with antifungal activity. Plant cell rep. 2011, 30, 2027-36. 44. Hain, R.; Bieseler, B.; Kindl, H.; Schroder, G.; Stocker, R., Expression of a stilbene synthase gene in Nicotiana tabacum results in synthesis of the phytoalexin resveratrol. Plant Mol. Biol. 1990, 15, 325-335. 45. Lee, D.; Douglas, C. J., Two Divergent Members of a Tobacco 4-CoumarateCoenzyme A Ligase (4CL) Gene Family. Plant physiol. Bioch. 1996, 112, 193-205. 46. Hain, R.; Reif, H.; Krause, E.; Langebartels, R.; Kindl, H.; Vornam, B.; Wiese, W.; Schmelzer, E.; Schreier, P.; Stöcker, R., Disease resistance results from foreign phytoalexin expression in a novel plant. Nature 1993, 361, 153-156. 47. Donnez, D.; Jeandet, P.; Clément, C.; Courot, E., Bioproduction of resveratrol and stilbene derivatives by plant cells and microorganisms. Trends Biotechnol. 2009, 27, 706-13. 48. Wang, C.; Liu, C.; Liu, J.; Xiang, W.; Huang, X.; Xu, L., Antioxidant Activity of Medicine Mulberry (Morus nigra) in Xinjiang. Scientia Silvae Sinicae 2014, 50, 53-59. 49. Lim, C. G.; Fowler, Z. L.; Hueller, T.; Schaffer, S.; Koffas, M. A., High-yield resveratrol production in engineered Escherichia coli. Appl. Environ. Microb. 2011, 77, 3451-60. 50. Katsuyama, Y.; Funa, N.; Miyahisa, I.; Horinouchi, S., Synthesis of unnatural flavonoids and stilbenes by exploiting the plant biosynthetic pathway in Escherichia coli. Chemistry & biology 2007, 25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656

14, 613-21.

26

ACS Paragon Plus Environment

Page 26 of 36

Page 27 of 36

Journal of Agricultural and Food Chemistry

657

Figure captions

658

Fig. 1. Biosynthetic pathways of resveratrol and chalcone. The enzymes of general

659

phenylpropanoid metabolism, which are connected by black arrows, consist of

660

phenylalanine ammonia-lyase (PAL), cinnamic 4-hydroxylase (C4H), and 4CL.

661

STS and CHS used the same substrates, and naringenin chalcone is usually

662

spontaneously converted to naringenin in vitro.

663

Fig. 2. Schematic representation of the strategy used for constructing the

664

co-expression vector pET4CL-T-STS (a and b), 28 indicates the fragment cloned

665

from pET28a(+); Construction the vector for tobacco transformations (c).

666

Fig. 3. Chromatograms of three standard stilbenes and mulberry extractions. (a)

667

standard chromatogram of the three stilbenes. (b) and (c) chromatograms of root

668

bark and fruit extractions, respectively.

669

Fig. 4. Expression of MaSTS genes (a1) and the content of stilbenes in root bark (RB),

670

stem bark (SB), branch bark (BB), old leaf (OL) and young leaf (YL) (a2-a4);

671

b1*, images [previously published (17)] indicate fruit development from S1 to

672

S7. In addition, the contents of stilbenes (b2–b4) and the expression levels of

673

MaSTS genes (c1–c4) in fruits are shown, ND means that no resveratrol was

674

detected.

675

Fig. 5. Biosynthesis routes of resveratrol constructed in recombinant E. coli (a) and

676

fermentation broth (b); (c1–c3) chromatograms of pET4CL, pET4CL-T-STS3 and

677

pET4CL-T-STS′, respectively.

678

Fig. 6. The relative expression levels of MaSTS genes under a series of treatments. a, 27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

679

b, c and d indicate the expression levels of four MaSTS genes under ABA, SA,

680

NaCl and wounding treatments, respectively. “*” indicates p < 0.05, “**”

681

indicates p < 0.01.

682

Fig. 7. Analysis of the expression levels of MaSTS3, Nt4CL1, Nt4CL2 and NtCHS in

683

transgenic tobaccos (a1–a4); The total flavonoid contents of transgenic tobacco

684

flowers (b1 and b2), chromatograms of tobacco extracts (b3) and phenotypes of

685

WT and transgenic lines after two week treatments (b4). Proline (c1–c4) and

686

MDA contents (d1–d4) for control, PEG, NaCl and heat treatments, respectively.

687

“*” indicates p < 0.05, “**” indicates p < 0.01.

28

ACS Paragon Plus Environment

Page 28 of 36

Page 29 of 36

Journal of Agricultural and Food Chemistry

Figure 1. Biosynthetic pathways of resveratrol and chalcone. The enzymes of general phenylpropanoid metabolism, which are connected by black arrows, consist of phenylalanine ammonia-lyase (PAL), cinnamic 4-hydroxylase (C4H), and 4CL. STS and CHS used the same substrates, and naringenin chalcone is usually spontaneously converted to naringenin in vitro. 190x155mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 2. Schematic representation of the strategy used for constructing the co-expression vector pET4CL-TSTS (a and b), 28 indicates the fragment cloned from pET28a(+); Construction the vector for tobacco transformations (c). 156x93mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 30 of 36

Page 31 of 36

Journal of Agricultural and Food Chemistry

Figure 3. Chromatograms of three standard stilbenes and mulberry extractions. (a) standard chromatogram of the three stilbenes. (b) and (c) chromatograms of root bark and fruit extractions, respectively. 142x155mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 4. Expression of MaSTS genes (a1) and the content of stilbenes in root bark (RB), stem bark (SB), branch bark (BB), old leaf (OL) and young leaf (YL) (a2-a4); b1*, images [previously published (17)] indicate fruit development from S1 to S7. In addition, the contents of stilbenes (b2–b4) and the expression levels of MaSTS genes (c1–c4) in fruits are shown, ND means that no resveratrol was detected. 167x93mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 32 of 36

Page 33 of 36

Journal of Agricultural and Food Chemistry

Figure 5. Biosynthesis routes of resveratrol constructed in recombinant E. coli (a) and fermentation broth (b); (c1–c3) chromatograms of pET4CL, pET4CL-T-STS3 and pET4CL-T-STS′, respectively. 179x87mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 6. The relative expression levels of MaSTS genes under a series of treatments. a, b, c and d indicate the expression levels of four MaSTS genes under ABA, SA, NaCl and wounding treatments, respectively. “*” indicates p < 0.05, “**” indicates p < 0.01. 141x155mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 34 of 36

Page 35 of 36

Journal of Agricultural and Food Chemistry

Figure 7. Analysis of the expression levels of MaSTS3, Nt4CL1, Nt4CL2 and NtCHS in transgenic tobaccos (a1–a4); The total flavonoid contents of transgenic tobacco flowers (b1 and b2), chromatograms of tobacco extracts (b3) and phenotypes of WT and transgenic lines after two week treatments (b4). Proline (c1–c4) and MDA contents (d1–d4) for control, PEG, NaCl and heat treatments, respectively. “*” indicates p < 0.05, “**” indicates p < 0.01. 196x170mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

table of contents (TOC) 56x47mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 36 of 36