Family of Ricinus communis Monosaccharide ... - ACS Publications

Jul 10, 2017 - and Han-Hong Xu*. State Key Laboratory for Conservation and Utilization of Subtropical Agro-bioresources and Key Laboratory of Natural ...
0 downloads 0 Views 2MB Size
Subscriber access provided by Caltech Library

Article

The family of Ricinus communis monosaccharide transporters and the RcSTP1 in promoting the uptake of a glucose-fipronil conjugate Gen-Lin Mao, Ying Yan, Yan Chen, Bing-Feng Wang, Fei-Fei Xu, Zhi-xiang Zhang, Fei Lin, and Han-Hong Xu J. Agric. Food Chem., Just Accepted Manuscript • Publication Date (Web): 10 Jul 2017 Downloaded from http://pubs.acs.org on July 10, 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 40

Journal of Agricultural and Food Chemistry

1

The family of Ricinus communis monosaccharide transporters and the RcSTP1

2

in promoting the uptake of a glucose-fipronil conjugate

3

Gen-Lin Mao ,Yin Yan , Yan Chen, Bing-Feng Wang, Fei-Fei Xu, Zhi-Xiang Zhang,

4

Fei Lin*, Han-Hong Xu*

§

§

5 6 7 8 9

State

Key

Laboratory

for

Conservation

and

Utilization

of

Subtropical

10

Agro-bioresources, and Key Laboratory of Natural Pesticide and Chemical Biology,

11

Ministry of Education, South China Agricultural University, Guangzhou, Guangdong

12

510642, Guangdong, People’s Republic of China.

13 14

§

These authors contributed equally to this work

15

*To whom correspondence should be addressed. Tel: +86-20-85285127. Fax:

16

+86-20-38604926. Email: [email protected] and [email protected]

17 18 19 20 21 22

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 40

23

ABSTRACT: Enhancing the systemic distribution of a bioactive compound by

24

exploiting the plant's vascular transport system presents a means of reducing both the

25

volume and frequency of pesticide/fungicide application. The foliar uptake of the

26

glucose-fipronil

27

N-[3-cyano-1-[2,6-dichloro-4-(trifluoromethyl)phenyl]-4-[(trifluoromethyl)sulfinyl]-1

28

H-pyrazol-5-yl]-1-(β-D-glucopyranosyl)-1H-1,2,3-triazole-4-methanamine

29

achieved in castor bean (Ricinus communis), and its transport via the phloem is known

30

to be mediated by monosaccharide transporter(s) (MSTs), although neither the

31

identity of the key MST(s) involved, nor the mechanistic basis of its movement have

32

yet to be described. Based on homology with Arabidopsis thaliana sugar transporters,

33

the castor bean genome was concluded to harbour 53 genes encoding a sugar

34

transporter, falling into the eight previously defined sub-families INT, PMT, VGT,

35

STP, ERD6, pGlucT, TMT and SUT. Transcriptional profiling identified the product

36

of RcSTP1 as a candidate for mediating GTF uptake, since this gene was induced by

37

exposure of the plant to GTF. By transiently expressing RcSTP1 in onion epidermis

38

cells, the site of RcSTP1 deposition was shown to be the plasma membrane. A

39

functional analysis based on RcSTP1 expression in Xenopus laevis oocytes

40

demonstrated that its product has a high affinity for GTF. The long distance

41

root-to-shoot transport of GTF was enhanced in a transgenic soybean chimera

42

constitutively expressing RcSTP1.

43

KEYWORDS: monosaccharide transporter, Ricinus communis, glucose-fipronil,

44

transgenic soybean composite plants, Xenopus oocyte

conjugate

ACS Paragon Plus Environment

(GTF)

Page 3 of 40

Journal of Agricultural and Food Chemistry

45



46

Certain pesticidal compounds are taken up by plants and then distributed systemically

47

via both phloem and xylem. This characteristic is highly desirable, as it allows full

48

protection to be achieved following either a foliar application or a simple irrigation of

49

the roots1,2. As yet only a small number of insecticides are known to have this

50

property: one of these is spirotetramat, a relatively new product derived from tetramic

51

acid, which has been shown to be transported to both the shoot and the roots via the

52

phloem and xylem, thereby providing a good level of control against aphids,

53

whiteflies and psyllids3.

INTRODUCTION

54

It has been suggested that systemic pesticides could be engineered by various

55

relatively simple chemical modifications of the active molecule4-6. One possibility is

56

to confer phloem mobility by exploiting the activity of native transporters. An

57

example is the modified form of E-2,4-dichlorophenoxyacetyl (2,4-D), which when

58

conjugated with an α-amino acid, is readily translocated via the phloem in Vicia faba

59

via an carrier system5. Similar results have also been achieved for the herbicide

60

glyphosate and 2,4-D, which are also mediated by phloem mobility and may be

61

loaded by a phosphate transporter and an Aromatic and Neutral Amino acid

62

transporter 1 (ANT1), respectively7,8.

63

The

glucose-fipronil

conjugate

64

N-[3-cyano-1-[2,6-dichloro-4-(trifluoromethyl)phenyl]-4-[(trifluoromethyl)sulfinyl]-1

65

H-pyrazol-5-yl]-1-(β-D-glucopyranosyl)-1H-1,2,3-triazole-4-methanamine (GTF) has

66

been shown to enjoy a higher level of phloem mobility than that of fipronil itself4. The

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 40

67

fluorescence-enabled visualization of the uptake and transport of GTF in the castor

68

bean (Ricinus communis) seedling has shown that, after transiting the epidermis and

69

mesophyll, the GTF is loaded into sieve tubes and then translocated from the

70

cotyledon to the hypocotyl via the phloem9. Both phloridzin and carbonyl cyanide

71

m-chlorophenylhydrazone markedly inhibit GTF uptake, while D-glucose is only a

72

moderate competitor for GTF and sucrose is non-competitive. The implication has

73

been drawn that GTF uptake, at least in the castor bean seedling, is partially mediated

74

by

75

monosaccharides across a hydrophobic membrane11.

monosaccharide

transporters

(MSTs)10,

proteins

which

shuttle

soluble

76

The MST polypeptide family has been organized into a number of sub-groups,

77

referred to as sugar transporter proteins (STPs), polyol/monosaccharide transporters

78

(PMTs), inositol transporters (INTs), plastidial glucose transporters/suppressors of G

79

protein (pGlucTs/SBGs), tonoplastic monosaccharide transporters (TMTs), early

80

response to dehydration transporters (EDRs) and vacuolar glucose transporters

81

(VGTs)12. The STPs, PMTs and INTs are all associated with the plasma membrane

82

and facilitate the import of glucose, fructose, galactose, mannose, xylose, sorbitol,

83

mannitol, xylitol and myo-inositol; the pGlucTs export glucose from the plastid, while

84

the SGBs enable glucose to be imported into the Golgi apparatus; the TMTs and

85

ERD-like transporters assist in the import/export of monosaccharides across the

86

vacuolar membrane; finally, the VGTs are important during both germination and

87

flowering13. The MSTs as a whole are able to transport a broad spectrum of

88

monosaccharides, but individual members of the family show distinct substrate

ACS Paragon Plus Environment

Page 5 of 40

Journal of Agricultural and Food Chemistry

89

specificity11,14. The Arabidopsis thaliana genome harbors 53 distinct MST-encoding

90

genes, but to date only two have been described in the castor bean15. The focus of the

91

present study was to identify which MST(s) in castor bean is/are involved in the

92

uptake and translocation of GTF.

93 94



MATERIALS AND METHODS

95

Plant Materials, Chemicals and Plasmid Vectors. R. communis seeds were

96

obtained from the Agricultural Science Academy (Zibo, Shandong, China) and those

97

of soybean cv. Yuechun 04-5 were obtained from the Root Biology Centre (South

98

China Agricultural University, Guangzhou, China). GTF was prepared according to

99

our previously described4. GTF and fipronil were dissolved in DMSO. And in all

100

treatments, the final concentration of DMSO was below 1% (v/v). Agrobacterium

101

rhizogenes strain K599 was also provided by the Root Biology Centre (South China

102

Agricultural University). The vectors pCam-35S and pEGFP were obtained from the

103

South China Agricultural University College of Life Sciences. The vector pT7TSHA

104

was kindly provided by the Institute of Plant Protection (Chinese Academy of

105

Agricultural Sciences, Beijing, China). Six days after sowing, the cotyledons of

106

uniformly-sized seedling, with residual endosperm being removed, were incubated in

107

a solution containing 20 mM MES (pH 5.5), 0.25 mM MgCl2, and 0.5 mM CaCl2

108

with 100 µM GTF or without 100 µM GTF. As solvent control, a solution of 1% (v/v)

109

DMSO replaced the 100 µM GTF was also included in the treatment. The cotyledons

110

were harvested at 3 h and 6 h post treatments, snap-frozen in liquid nitrogen and

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

111

Page 6 of 40

stored at −80°C until ready for RNA extraction.

112

In silico Identification of R. communis Sugar Transporter Genes and Their

113

Phylogeny. The R. communis genome sequence, downloaded from the Phytozome

114

v9.1 database (www.phytozome.net/ricinus.php), was scanned using the HMM search

115

program for the presence of sequences harboring motifs associated with “sugar (and

116

other) transporters” (Pfam HMM profile PF00083) (hmmer.janelia.org/). The selected

117

sequences were first validated by subjecting them to a BlastP search against the set of

118

known A. thaliana sugar transporters, and finally, all available EST data related to

119

these genes were assembled following a BlastN search against the R. communis NCBI

120

dbEST dataset (http://www.ncbi.nlm.nih.gov). Predicted products of below 300

121

residues in length were excluded, since these are thought unlikely to be functional16,17.

122

The set of possible R. communis sugar transporters was aligned using the Clustal X2

123

algorithm and a phylogenetic tree was constructed based on the neighbor-joining

124

method, using routines implemented in MEGA6 software18. Sub-family identity was

125

assigned on the basis of homology with A. thaliana sequences (www.arabidopsis.org).

126

RT-PCR and Quantitative Real Time PCR (qRT-PCR) Analysis. The set of

127

RcMSTs,

obtained

from

the

128

(www.phytozome.net/ricinus.php), was targeted for amplification using primer pairs

129

listed in Table S1. RNA was extracted from frozen cotyledons using a Plant RNA kit

130

(OMEGA, Guangzhou, China). An 1 µg aliquot used as the template for synthesizing

131

the cDNA first strand, using the iScript™ Reverse Transcription system (Bio-Rad,

132

Philadelphia, PA, USA) after treating with DNase I (NEB, Ipswich, MA, England)

R.

communis

gene

ACS Paragon Plus Environment

annotation

database

Page 7 of 40

Journal of Agricultural and Food Chemistry

133

was used to remove the contaminating genomic DNA. An RT-PCR assay, based on a

134

bulked sample of cDNA template, was performed to confirm the presence of MST

135

transcript, with R. communis genomic DNA included as a positive control. The

136

resulting amplicons were visualized by electrophoretic separation through a 2% (w/v)

137

agarose gel. Genes displaying a detectable level of transcription were monitored via a

138

quantitative real time PCR (qRT-PCR) assay of the response of cotyledon explants to

139

a 6 h exposure to GTF. The qRT-PCRs were performed as described elsewhere19,

140

Relative transcript abundances were calculated from the abundance of RcActin

141

transcript (GenBank accession: NM_001323740.1, for primer sequence please see

142

Table S1), using the ∆∆Ct method20. Difference between treatment means value were

143

tested

144

(https://spss.en.softonic.com/)

for

significance

using

routines

implemented

in

SPSS

software

145

Subcellular Localization of Gene Expression. The open reading frame of

146

RcSTP1, previously known as RcHEX315, was inserted in frame in front of the

147

enhanced green fluorescent protein (EGFP) coding sequence within a pEGFP vector

148

using a two-step Ω-PCR procedure21. Specifically, in the first reaction, the chimeric

149

primer pair pGFP-RcSTP1F and pGFP-RcSTP1R (see Table S1) was used to amplify

150

the RcSTP1 sequence, and a 2-3 µL aliquot of this reaction was then used as the

151

primer for the second reaction with pEGPF as template. The final product was

152

incubated at 37℃ for 30 min in the presence 5-10 U DpnI to degrade the plasmid

153

sequence, and the product was inserted into E. coli (DH5α) competent cells. Positive

154

clones were validated by sequencing. The fusion vector pEGFP-RcSTP1 was

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 40

155

transiently transformed by bombardment (PDS/1000 device, Bio-Rad) into onion

156

epidermal strips supported on an agar plate. The bombardment parameters were:

157

bombardment pressure 1,100 psi, gold particle diameter 1.0 mm, separation distance 9

158

cm, and decompression vacuum 1,000 psi. After a 16 h incubation, fluorescence was

159

captured by a laser scanning confocal microscope (Zeiss LSM780, Germany), using a

160

wavelength of 488 nm for excitation and 515-545 nm and 610 nm filters for detection.

161

Functional Analysis in Xenopus laevis Oocytes. Transporter activity was

162

determined using a two electrode voltage clamp technique22-24. The 1.6 kb opening

163

reading frame of RcSTP1 gene was amplified by primer pair pT7-RcSTP1F and

164

pT7-RcSTP1R (Table S1). The plasmid pT7TSHA was linearized by cutting at its Spe

165

℃ and Sph ℃ sites and then recombined with the RcSTP1 amplicon using an In-Fusion

166

Cloning Kit (Clontech, Mountain view, CA, USA). The resulting plasmid was

167

linearized by restriction with Sma I, then a transcribed the Capped mRNA in vitro

168

using

169

http://www.thermofisher.com/cn/zh/home/brands/ ambion.html).

170

laevis oocytes were isolated and injected with 27.6 nL (1 ng/nL) RcSTP1 cRNA and

171

incubated for 2-4 days at 18℃ in Barth’s medium supplemented with 10 µg/mL

172

gentamycin. The oocytes were then bathed in modified sodium Ringer solution (96

173

mM NaCl, 2 mM KCl, 1.8 mM CaCl2, 5 mM HEPES, 5 mM MgCl2, pH 5.5 adjusted

174

with NaOH) with continuous perfusion at a rate of 3 mL/min. The recording pipettes

175

were filled with 3M KCl, delivering an electrical resistance of 1-5 MΩ. Currents were

176

measured using a oocyte clamp amplifier (Model OC-725C, Warner Instruments,

mMessage

Mmachine

kits

ACS Paragon Plus Environment

(Thermo

Fisher,

Stage ℃ and ℃ X.

Page 9 of 40

Journal of Agricultural and Food Chemistry

177

Hamden, CT, USA), filtered at 200 Hz and digitized at 2,000 Hz. Holding potential

178

was -90 mV, and voltage pulses from -150 mV to +50 mV were applied for 100 ms.

179

Steady-state currents are presented as the mean current between 50 and 100 ms

180

following the onset of voltage pulses and were obtained by subtracting an average of

181

the background current recorded before and after substrate application. The data were

182

acquired and analyzed with the help of Digidata 1440A and pClamp10.0 software

183

(Axon Instruments Inc., Union City, CA, USA). Dose-response data were analyzed

184

using GraphPad Prism 5 (http://www.graphpad.com/scientific-software/prism/).

185

Root Uptake and Root-shoot Translocation of GTF and Fipronil in Soybean

186

Seedlings. Due to the current lack of efficient transformation technology for R.

187

communis, a soybean model was used to explore the function of RcSTP1 with respect

188

to the uptake of GTF and its translocation to the root and shoot. Four day-old soybean

189

seedlings were cultured hydroponically in half-strength Hoagland’s solution for one

190

week under controlled conditions (16 h photoperiod; light period temperature 28 ±

191

2°C, dark period temperature 22 ± 2°C; relative humidity 70%). The roots were

192

bathed for either 6 h or 12 h in half strength Hoagland’s solution containing 0.1%

193

(v/v) Tween-80 and 50 µM GTF or 50 µM fipronil, after which the plants were

194

separated into their aerial part and their roots, macerated and extracted with acetone.

195

The extractant was purified using an AccuBond C18 SPE solid-phase extraction kit

196

(Agilent Technologies, Santa Clara, CA, USA), following a previously published

197

protocol4. The content of the resulting solutions was analyzed using an Agilent

198

Technologies 1100 HPLC system equipped with a vacuum degasser, a quaternary

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

199

pump, an autosampler, a UV-visible detector and an Agilent C18 reverse-phase

200

column (5 µm, 250 mm × 4.6 mm i.d.) held at 30°C. The flow rate was 1 mL/min, and

201

the injection volume was 10 µL. The mobile phase consisted of a 1:1 mixture of

202

acetonitrile and water and the absorbance wavelength was 210 nm. A series of

203

standard solutions of GTF or fipronil (0.5, 1, 5, 10, 25, and 50 µM) was prepared in

204

methanol for calibration purposes. In order to measure the content of GTF and

205

fipronil, standard curves were made. The linear equation of GTF was y = 20.31x -

206

0.9439 (r = 0.9993) and that for fipronil was y = 20.26x + 0.3991 (r = 0.9999).

207

Recovery studies were conducted at three spiking levels: 0.1 mg/Kg, 0.5 mg/Kg and 1

208

mg/Kg.

209

were all >80%.

The individual mean recovery rates for GTF and fipronil in the plant tissue

210

The Heterologous Expression of RcSTP1 in Soybean Hairy Roots. The

211

RcSTP1 cDNA sequence was PCR-amplified using the primer pair p35S-STP1F and

212

p35S-STP1R (Table S1). The amplified fragment was then transferred into the

213

pCam-35S vector using the appropriate restriction enzyme sites. The resulting

214

construct, comprising the RcSTP1 opening reading frame driven by a CaMV 35S

215

promoter, was transformed into Agrobacterium rhizogenes strain K599 by

216

electroporation. Transgenic chimeric soybean plants, consisting of a wild type shoot

217

and transgenic hairy roots, were treated to induce the expression of the transgene

218

using the hypocotyl injection method described by25,26. Briefly, seeds were

219

surface-sterilized by immersion for 1 min in 3% (v/v) H2O2, rinsed with sterile water

220

and germinated in sand medium. The hypocotyls of five day-old seedlings (cotyledons

ACS Paragon Plus Environment

Page 10 of 40

Page 11 of 40

Journal of Agricultural and Food Chemistry

221

unfolded) were infected with A. rhizogenes strain K599 carrying the transgene and

222

empty vector, and the plants were kept under high humidity. When newly developed

223

hairy roots had reached a length of ~10 cm, the wild type roots were excised and the

224

plants were cultured in half-strength Hoagland's solution for 1~2 weeks to allow for

225

further growth of the hairy roots. Hairy roots harbouring an empty vector were stained

226

for GUS activity following a published method27. DNA was extracted from hairy root

227

segments, and subjected to a PCR assay targeting RcSTP1 to identify those which had

228

been successfully transformed. The roots of confirmed transformants were incubated

229

in half strength Hoagland’s solution containing 0.1% (v/v) Tween-80 and 50 µM GTF,

230

and harvested after 2 h, 6 h or 12 h. The plants were separated into upper and lower

231

parts and roots (Fig. S1). GTF extraction and sample analysis were performed as

232

described above.

233 234



RESULTS

235

The Members of the R. communis Sugar Transporter Family. A total of 99

236

protein sequences were identified as harboring a potential “sugar (and other)

237

transporters” domain was recovered. Of these, 63 shared at least 45% (and up to 84%)

238

similarity to an A. thaliana sugar transporter; all but ten of the 63 specified a product

239

longer than the 300 residues required to act as a functional transporter. Their encoding

240

genes were distributed across 33 scaffolds of the R. communis genome (Table 1). The

241

final set of 53 sugar transporters contained three sucrose transporters (SUTs) and 50

242

MSTs. MSTs were classified into seven sub-families: RcINTs (six proteins), RcPMTs

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

243

(seven), RcVGTs (two), RcSTPs (ninteen), RcERD6s (nine), RcpGlucTs (four),

244

RcTMTs (three) (Table 1). Three of the genes (RcSTP1, RcSTP9 and RcSUC2) are

245

synonyms of the genes RcHEX3, RcHEX6 and RcScr1, respectively15,28.A

246

phylogenetic analysis, based on the proteins' full length sequences with the

247

neighbor-joining method and 1000 bootstrap replicates, showed the 53 genes belong

248

to eight established sub-groups (Fig. 1).

249

Transcriptional Response of the RcMSTs to GTF Treatment. Based on

250

RT-PCR analysis of cotyledons subjected to a 6 h exposure to GTF, nine of the

251

RcMSTs showed no trace of transcription, four showed spliceosome transcripts and

252

nine were weakly transcribed (Fig. 2A, table 1). A more detailed analysis of 35 of the

253

RcMSTs (Table 1) was conducted using qRT-PCR: this showed that after a 6 h

254

exposure to GTF, 26 of them were up-regulated (seven by more than two fold) (Fig.

255

2B) and eleven were down-regulated (the only gene by more than two fold was

256

RcTMT3) (Fig. 2C). The behavior of the RcSTPs was also monitored following a 3 h

257

exposure to GTF; all of these, with the sole exception of RcSTP1, were

258

down-regulated by the treatment (data not shown), with RcSTP1 transcript abundance

259

being reduced by around five fold when compared with the negative control (Fig. 3).

260

At the 6 h time point, the abundance of RcSTP1 transcript induced by GTF was

261

significantly (p 0.05). 223x167mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 4 The sub-cellular localization of RcSTP1 in transiently transformed onion epidermal cells. The transgenes were (A, C) p35S::GFP and (B, D) p35S::RcSTP1cDNA-GFP. (A, B) Z-projections of optical sections. (C, D) individual optical sections from the center of the cells show in (A) and (B), respectively. RcSTP1 localizes to the plasma membrane and lacks cytoplasmic strands (arrowed in (C)) as introduced by the p35S::GFP transgene. 190x190mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 34 of 40

Page 35 of 40

Journal of Agricultural and Food Chemistry

Figure 5 Inward currents in X. laevis oocytes in response to the treatments with glucose and GTF. The oocytes were tested using voltage clamped at -90 mV. Non-injected oocytes are shown in the upper trace and those injected with RcSTP1 cRNA are shown in the lower trace. Oocytes were perfused with 200 µM glucose or GTF at the indicated times points. 77x38mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 6 Kinetic analysis of glucose- and GTF-induced currents. Oocytes harboring RcSTP1 were perfused at a holding potential of -50mV. The data have been fitted to the Michaelis-Menten equation. The error bars show the SE (n = 5 oocytes). (A) K0.5 value for glucose (0.059 ± 0.023 mM ). (B) K0.5 value for GTF (0.042 ± 0.003 mM). 146x50mm (220 x 220 DPI)

ACS Paragon Plus Environment

Page 36 of 40

Page 37 of 40

Journal of Agricultural and Food Chemistry

Figure 7 The distribution of fipronil and GTF in wild type soybean seedlings. The detection is carried out following a 6 h or a 12 h incubation of the roots in (A) fipronil, or (B) GTF. Data were shown in the form mean ± SE (n = 3) with a column topped by different letters are significantly different from each other as determined by Duncan’s multiple-range test (P > 0.05). 191x74mm (150 x 150 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 8 PCR-based validation of transgenicity of hairy root cultures transformed with pCaMV35S::RcSTP1. M: DL5000 size marker; lanes 1-4: template extracted from the root of putative transgenic lines. Con: hairy roots harboring an empty vector, N: non-transgenic roots. 813x437mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 38 of 40

Page 39 of 40

Journal of Agricultural and Food Chemistry

Figure 9 The distribution of GTF in transgenic soybean seedings following the incubation of hairy roots in 100 µM GTF. Plants harboring either an empty vector or pCaMV35S::RcSTP1. Samples were taken after (A) 1 h, (B) 2 h, (C) 6 h, (D) 12 h. Data shown in the form mean ± SE (n = 3). Differences being statistical significant (p < 0.05) were determined by the Students' t test, and indicated by asterisks. 201x141mm (150 x 150 DPI)

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

537x218mm (150 x 150 DPI)

ACS Paragon Plus Environment

Page 40 of 40