A Polyketide Synthase Encoded by the Gene ... - ACS Publications

Nov 14, 2016 - Key Laboratory of Industrial Fermentation Microbiology, Ministry of Education, Tianjin University of Science and Technology,. Tianjin 3...
2 downloads 8 Views 1MB Size
Subscriber access provided by UNIV TORONTO

Article

A polyketide synthase encoded by the gene An15g07920 is involved in the biosynthesis of ochratoxin A in Aspergillus niger Jian Zhang, Liuyang Zhu, Haoyu Chen, Min Li, Xiaojuan Zhu, Qiang Gao, Depei Wang, and Ying Zhang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b03907 • Publication Date (Web): 14 Nov 2016 Downloaded from http://pubs.acs.org on November 19, 2016

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 31

Journal of Agricultural and Food Chemistry

1

A polyketide synthase encoded by the gene An15g07920 is involved in the

2

biosynthesis of ochratoxin A in Aspergillus niger

3 4

Jian Zhang,† Liuyang Zhu, †

5

and Ying Zhang *,‡

Haoyu Chen,† Min Li,† Xiaojuan Zhu,† Qiang Gao,†

Depei Wang,†

6 7



8

Science and Technology, Tianjin 300457, China

9



10

Key Laboratory of Industrial Fermentation Microbiology, Ministry of Education, Tianjin University of

Key Laboratory of Food Nutrition and Safety, Ministry of Education, Tianjin University of Science and

Technology, Tianjin 300457, China

11 12

* Corresponding author:

13

Ying Zhang (E-mail: [email protected]; Tel: +86-022-60912431)

14

15

16

17

18

19

20

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 31

21 22

ABSTRACT: The polyketide synthase gene An15g07920 was known in Aspergillus niger CBS 513.88

23

as putatively involved in the production of ochratoxin A (OTA). Genome re-sequencing analysis

24

revealed that the gene An15g07920 is also present in the ochratoxin-producing A. niger strain 1062.

25

Disruption of An15g07920 in A. niger 1062 removed its capacity to biosynthesize ochratoxin β (OTβ),

26

ochratoxin α (OTα) and OTA. These results indicate that the polyketide synthase encoded by

27

An15g07920 is a crucial player in the biosynthesis of OTA, in the pathway prior to the phenylalanine

28

ligation step. The gene An15g07920 reached its maximum transcription level before OTA accumulation

29

reached its highest level, confirming that gene transcription precedes OTA production. These findings

30

will not only help explain the mechanism of OTA production in A. niger, but also provide necessary

31

information for the development of effective diagnostic, preventive and control strategies to reduce the

32

risk of OTA contamination in foods.

33 34

KEYWORDS: Ochratoxin A, Aspergillus niger, Kinetics, Gene expression, Biosynthesis

35 36 37 38 39 40 41 42 43 2

ACS Paragon Plus Environment

Page 3 of 31

Journal of Agricultural and Food Chemistry

44 45

INTRODUCTION

46 47

Ochratoxin A (OTA) is a mycotoxin produced by Aspergillus and Penicillium spp. fungi. It has been

48

demonstrated to have nephrotoxic, immunotoxic, genotoxic, neurotoxic and teratogenic properties; it

49

was classified as a renal carcinogen of animals and a possible (Group 2B) carcinogen of humans by the

50

International Agency for Research on Cancer (IARC).1 This compound is found widely in cereals and

51

their derivatives, wine, coffee, beer, nuts, dried fruits, and even meat products.2 Due to the

52

well-established routes of human exposure and abundance of toxicological data from animal studies, the

53

European Union has established stringent regulatory limits for OTA in a wide range of food products

54

such as raw cereal grains (5 µg/kg), products derived from cereals (3 µg/kg), dried fruits (10 µg/kg),

55

roasted and instant coffee (5 and 10 µg/kg, respectively), beverages containing wine and grape juice (2

56

µg/kg), baby food (0.5µg/kg) and infant formula (0.5 µg/kg).3

57

Chemically, OTA is composed of a chlorinated type I polyketide dihydroisocoumarin moiety linked to l-

58

phenylalanine by an amide bond. Early feeding and chemical degradation experiments showed that the

59

isocoumarin moiety is produced from five acetate units and a single methionine-derived carbon added at

60

the C-7 position, while the l- phenylalanine part stems from the normal shikimic acid pathway.4

61

According to OTA structure and its proposed biosynthetic pathway, the synthesis can be expected to

62

require several proteins, including a polyketide synthase (PKS) for the biosynthesis of the polyketide

63

dihydroisocoumarin, a nonribosomal peptide synthetase (NRPS) for the ligation of the amino acid

64

phenylalanine to the polyketide moiety, and a halogenase for the chlorination step. There are three types

65

of known PKSs—modular type I PKSs, iterative type I PKSs, and type II PKSs. Fungi mainly utilize the

66

iterative type I PKSs, which have a modular organization similar to the modular type I PKSs, whereby

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 31

67

the catalytic reaction of each domain can be engaged repeatedly in an iterative fashion. Typical

68

representatives of iterative type I PKS structures include ketosynthase (KS), acyl transferase (AT),

69

ketoreductase (KR), dehydratase (DH), enoyl reductase (ER), methyltransferase (MT), thioesterase (TE),

70

and acyl carrier protein (ACP) domains.5

71

Most of the molecular aspects of OTA biosynthesis have been elucidated in Penicillium species. In P.

72

nordicum, a putative OTA biosynthetic cluster has been identified which contains biosynthetic genes

73

encoding PKS, NRPS, as well as sequences with homology to chlorinating enzymes and homology to a

74

transporter protein, respectively.6 The orthologues of these genes have also been identified in the closely

75

related species P. verrucosum.7 Within Aspergillus spp., the pks gene, the nrps gene and halogenase gene,

76

have been fully characterized and demonstrated to be required for ochratoxin production in A.

77

carbonarius.

78

characterized in A. ochraceus.

79

there are other reports which inferred that some genes (such as those encoding PKS, NRPS, P450,

80

phenylalanine tRNA synthetase and methylase) were related to OTA synthesis based on the correlation

81

of gene expression profiles and data obtained using genome sequencing and bioinformatics methods.13-17

82

Aspergillus niger is an important industrial “workhorse” organism, with extensive applications in the

83

production of industrial enzymes, heterologous proteins and organic acids, among others.18 Whereas this

84

species has gained the generally recognized as safe (GRAS) status from the US Food and Drug

85

Administration, there have been increasingly frequent reports that some isolates of A. niger are able to

86

produce OTA, which represents a matter of great concern for its industrial application.19 Genome

87

sequencing of A. niger strain CBS 513.88 revealed the presence of the pks gene An15g07920, that has a

88

strong similarity to the pks gene of A. ochraceus involved in OTA biosynthesis.15 What’s more, Ferracin

89

et al. have found a positive association between the presence of strain-specific pks genes such as

8-10

The pks gene has also been partially characterized in A. westerdijkiae,11 and fully 12

In addition to characterizing the genes involved in OTA biosynthesis,

4

ACS Paragon Plus Environment

Page 5 of 31

Journal of Agricultural and Food Chemistry

90

An15g07920 and the capability of the respective A. niger strains to produce OTA.20 Castella and

91

Cabanes showed that a fragment of the gene An15g07920 was specific for ochratoxin-producing strains

92

of A. niger and developed a real-time PCR protocol for the detection of OTA-producing strains of the A.

93

niger aggregate.21 Recently, the correlation between pks gene expression and OTA production in A. niger

94

has also been reported.22 However, this gene has to our best knowledge not been disrupted and its role in

95

OTA biosynthesis has not been confirmed in A. niger.

96

In this work, we re-sequenced the genome of A. niger 1062 in order to find the corresponding gene

97

homologous to An15g07920, after which we have characterized and disrupted this gene by targeted gene

98

replacement using Agrobacterium tumefaciens-mediated transformation (ATMT). We further analyzed

99

the phenotypic characteristics of the wild-type and the mutant strains to explore the function of

100

An15g07920. Finally, we utilized a qRT-PCR assay to analyze An15g07920 expression. Our results

101

reveal for the first time that the gene An15g07920 is involved in the biosynthetic pathway of OTA in A.

102

niger and that it plays a crucial role in the biosynthesis pathway that precedes phenylalanine ligation.

103 104

MATERIALS AND METHODS

105 106

Strains and preservation conditions

107 108

The ochratoxin-producing wild-type strain A. niger 1062 used in this study, was previously isolated from

109

Chinese vineyard and deposited in China General Microbiological Culture Collection as No.9668. The

110

mutant strain used in this study was derived from strain 1062 and is indicated as ∆pks. A. tumefaciens

111

strain AGL-1 was kindly provided by Prof. Depei Wang (Tianjin University of Science and Technology,

112

Tianjin, China). Escherichia coli DH5α (Takara, Shiga, Japan) was used for the construction and

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 31

113

propagation of vectors. Both A. niger strains (1062 and ∆pks) were grown at 25°C on potato dextrose

114

agar plates (PDA) (Fisher Labosi) for 5 days, after which spores were collected by washing with a sterile

115

solution of 0.9% (v/v) normal saline (Fisher Labosi) and stored at -80°C in 25% (v/v) of glycerol (Fisher

116

Labosi) for further use.

117 118

DNA and RNA Extraction, cDNA Synthesis, qPCR and qRT-PCR

119 120

A. niger conidia (106/mL) were used to inoculate 250 mL Erlenmeyer flasks containing 100 mL of

121

Czapek yeast autolysate (CYA) liquid medium, and incubated at 25 °C without shaking for 7 days. The

122

mycelia were harvested by filtration through a 0.45 µm filter (Millipore), frozen in liquid nitrogen and

123

stored at -80 °C for further analysis. Genomic DNA of A. niger 1062 was extracted using the E.Z.N.A.

124

Fungal DNA kit (Omega Bio-Tek, Doraville, GA, USA) according to the manufacturer’s protocol.

125

Total RNA was extracted using the Trizol Reagent (Invitrogen,CA, USA) according to the

126

manufacturer's protocol. RNA concentrations were quantified using a NanoDrop ND-1000

127

spectrophotometer (Thermo Fisher, MA, USA), and adjusted to 0.5 mg/mL. The cDNA was synthesized

128

with the RNA samples as templates using the ReverTra Ace qPCR RT Master Mix with gDNA Remover

129

(Toyobo, Osaka, Japan).

130

Real-time genomic PCR (qPCR) was carried out in order to determine the copy numbers of integrated

131

T-DNA cassettes in the genomes of transformants. The primer pair HPHq-F/R (Table 1) was designed to

132

amplify the selection marker. The single-copy A. niger actin gene served as a reference and was

133

amplified using the ACTq-F/R primers (Table 1)23. Reactions were performed in a final volume of 20 µL,

134

comprising 10 µL of SYBR PremixEx TaqII (Takara, Tokyo, Japan), 0.5 µL of each primer (10 µM), and

135

1 µL of genomic DNA, made up to volume with PCR-grade ultrapure water. The amplification program

6

ACS Paragon Plus Environment

Page 7 of 31

Journal of Agricultural and Food Chemistry

136

was conducted with an initial step of 10 min at 95°C, followed by 40 cycles of 10 s at 95°C and 34 s at

137

60°C.

138

Real-time quantitative reverse transcription PCR (qRT-PCR) was performed to examine the expression

139

of An15g07920 at different times during growth of A. niger in CYA liquid medium, using the primers

140

KSqrt-F/R designed based on the KS domain of the An15g07920 gene (Table 1). The constitutively

141

expressed β-tubulin gene (primers Tubqrt-F/R) served as an internal reference to normalize gene

142

expression. The amplification reactions were performed in reaction volumes of 20 µL, comprising 10 µL

143

of SYBR PremixEx TaqII (Takara, Tokyo, Japan), 0.5 µL of each primer (10 µM) and 2 µL of cDNA

144

template, made up to volume with PCR-grade ultrapure water. Amplification was conducted using an

145

initial denaturation step at 95 °C for 3 min, followed by 40 cycles of 10 s at 95 °C, 30 s at 60 °C and 30 s

146

at 72 °C. The specificity of amplification was confirmed by dissociation curve analysis.

147

qPCR and qRT-PCR assays were performed and monitored using an ABI PRISM 7900HT system

148

(Applied Biosystems, Spain). The relative quantitative gene expression data were established using the

149

comparative 2−∆∆CT method. PCR efficiency of each oligonucleotide pair was calculated using linear

150

regression based on standard curves. All experiments were performed in three biological replicates, and

151

each sample was tested in triplicate, in addition to a no-template control included for each primer pair.

152 153

Genome re-sequencing, assembly and sequence analysis

154 155

Genome re-sequencing of A. niger 1062 was performed using an IRoche GS FLX system by Shanghai

156

Personalbio Biotechnology (Shanghai, China). The reads were mapped against the reference genome (A.

157

niger CBS513.88) using the program bwa

158

developed by 454 Life Sciences (Branford, Connecticut, USA), to yield a number of contigs. The

24

and were assembled de novo using Newbler (version 2.6)

7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 31

159

resulting contigs were extended to obtain longer scaffolds using SSPACE (Version 2.0). The software

160

GapFiller was used to close gaps.25 The deduced amino acid sequence was determined using the translate

161

tool from Expasy (http://web.expasy.org/translate/) while protein–protein Blast (Blastp) searches were

162

conducted in GenBank (http://www.ncbi.nlm.nih.gov). Mulitple sequence alignments were calculated

163

using MultAlin (http://prodes.toulouse.inra.fr/multalin/multalin.html). Conserved functional domains

164

were analyzed using InterProScan (http://www.ebi.ac.uk/interpro/) with default parameter settings.

165 166

Construction of gene-disruption vector

167 168

The plasmid p44 which was used for the construction of the gene-disruption vector was maintained in E.

169

coli DH5α. Restriction endonucleases and DNA-modifying enzymes were purchased from Takara (Shiga,

170

Japan). The upstream (646 bp) and downstream (507 bp) fragments flanking the hygromycin-resistance

171

gene in the p44 vector were amplified from the genomic DNA of A. niger 1062 using PCR with GXL

172

high-fidelity DNA polymerase (Takara, Shiga, Japan), with the primer sets KS5F/R and KS3F/R (Table

173

1), respectively, and the corresponding fragments purified using the MiniBEST agarose gel DNA

174

extraction kit (Takara, Shiga, Japan). The purified upstream fragment was inserted into the PstI and

175

HindIII digested p44 vector using the In-Fusion cloning kit (Takara, Shiga, Japan) to produce p44-HR1.

176

Similarly, the purified downstream fragment was inserted between the BamHI and KpnI sites of

177

p44-HR1 to generate the completed disruption vector p44-HR1-HR2 (Figure 1A). An aliquot of vector

178

DNA was used to transform E. coli DH5α chemically competent cells without prior ligation.

179

Kanamycin-resistant transformants were further screened using PCR and cloning junctions were

180

confirmed by DNA sequencing (Solarbio, China). Finally, the resulting vector p44-HR1-HR2 was

181

introduced into electrocompetent A. tumefaciens AGL-1 cells as described previously.26

8

ACS Paragon Plus Environment

Page 9 of 31

Journal of Agricultural and Food Chemistry

182

Table 1

183 184 185

A. tumefaciens-mediated transformation of A. niger

186 187

A. tumefaciens AGL-1 cells carrying the vector p44-HR1-HR2 were cultured in 3 mL of Luria Bertani

188

(LB, Miller, Solarbio, China) medium on a rotatory shaker at 180 rpm and 28 °C for 10 h. A. tumefaciens

189

cells were collected by centrifugation (2,200×g,10 min) and resuspended in fresh induction medium24

190

(IM) containing 200 µmol/L acetosyringone (AS, Solarbio, China) at a cell density corresponding to an

191

optical density at 600nm (OD600) of 0.15. After pre-incubation for 6 h under constant orbital shaking at

192

28 °C to an OD600 of 0.7 to 0.8, a 5mL aliquot of the cell suspension was mixed with an equal volume of

193

5-day-old A. niger spores (2×107/mL), and the mixture was subsequently spread onto a nitrocellulose

194

filter membrane (0.45 µm pore and 45 mm diameter, Whatman, GE Life Sciences, USA) and placed on

195

an IM agar plate containing 200 µmol/L AS. The plates were incubated at 25 °C in the dark for 48h.

196

After co-cultivation, the membrane was transferred onto a complete medium27 (CM) agar plate

197

containing hygromycin B (Solarbio, China; 100 µg/mL) as the selection agent for fungal transformants,

198

and cefotaxime (Solarbio,China; 200 µg/mL) to inhibit growth of A. tumefaciens cells.

199 200

Characterization of transformants

201 202

The transformants were screened for homologous recombination by PCR using primers KS5-F and

203

KS3-R. The parental strain had an amplicon of about 1.8 kb, and successful transformants had amplicons

204

larger than 2.5 kb. Transformants with a correct amplicon size, as predicted based on the insertion

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 31

205

cassette, were collected for further qPCR analysis (see below). A melting curve was determined for each

206

sample at the end of each run to ensure the purity of the amplified product. qPCR was performed in five

207

Replicates. The copy number was calculated according to a previously published 28 modified formula:

208

X 0 R0 = 10

209

where I X and I R represent the intercepts of the relative standard curves of target and reference genes,

210

SX

C and S R represent the slopes of the standard curves of target and reference genes, and T X and

211

CT R

are the detected threshold cycles for the amplification of the target and reference genes of a tested

212

sample, respectively.

( CTX − I X ) S X  − ( CTR − I R ) SR 

213 214

Analysis of fungal growth and conidiogenesis

215 216

For growth assessment, CYA agar plates were inoculated centrally with 5 µL of a 106 /mL suspension of

217

conidia of A. niger 1062 and the ∆pks mutant strains, respectively. Cultures were incubated at 25 °C for

218

7 days. Colony morphology was recorded and the number of conidia was determined at 24 h intervals

219

for 7 days. The conidia were washed in 10 mL of sterile water, followed by filtration through four layers

220

of lens tissues (Solarbio, China) to remove mycelial debris. The concentrations of the conidial

221

suspensions were determined using a Thoma cell counting chamber (Qiujing, Shanghai, China).

222

Six-day-old samples of A. niger conidia were prepared for scanning electron microscopy (SEM) as

223

described previously by Darah et al.29 and were observed under a Leo-Supra 50VP field emission

224

scanning electron microscope (FESEM, Carl Zeiss, Germany).

225 226

Chemical analysis of culture supernatants by HPLC-FLD and HPLC-MS 10

ACS Paragon Plus Environment

Page 11 of 31

Journal of Agricultural and Food Chemistry

227 228

For the determination of the production of OTA and its metabolites, conidia samples comprising 105

229

spores in 100 µL of water were used to inoculate Erlenmeyer flasks containing 20 mL of CYA liquid

230

medium. Incubation was carried out at 25 °C in the dark under stationary conditions for 9 days. The

231

supernatants of the liquid cultures of the two tested strains were filtered through a 0.22 µm pore filter

232

(Sartorius AG, Goettingen, Germany) and analyzed for their content of OTA and its metabolites by

233

high-performance liquid chromatography with fluorescence detection (HPLC-FLD). Liquid cultures of

234

1062 and ∆pks were also analyzed by high-performance liquid chromatography coupled to a mass

235

spectrometry system (HPLC-MS) to confirm the results and to identify other metabolites belonging to

236

the OTA biosynthetic pathway.

237

HPLC-FLD analysis was carried out on an Agilent 1100 series system equipped with a G1312A binary

238

pump, a G1313A autosampler, a G1321A fluorescence detector set to 333 nm excitation and 460 nm

239

emission wavelength, and an Agilent Chemstation G2170AA Windows 2000 operating system (Agilent,

240

Waldbronn, Germany). The column was a KromstarTM C18 (250 × 4.6 mm, 5 µm particles), and was

241

preceded by a Rheodyne guard filter (3 mm, 0.45 µm pore size). The mobile phase was an isocratic

242

mixture of acetonitrile/water/acetic acid (99:99:2, v/v/v) and the analytes were eluted at a flow rate of

243

1.0 mL/min.

244

HPLC-MS analysis was performed on an Agilent 1200 series system connected to an Agilent 6110 single

245

quadrupole mass spectrometer (Agilent, MA, USA) via an ESI source. The ESI was operated in negative

246

mode to detect OTα and OTβ, in positive mode to detect OTA. The parameters used for the mass

247

spectrometer in all experiments were as follows: 350 °C gas temperature, 13.0 L/min drying gas (nitrogen)

248

flow, 3.5 kg/cm2 nebulizer gas pressure and 3500 V capillary voltage. The molecular ions [M+H]+ were

11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 31

249

monitored at m/z 404 for OTA. The molecular ions [M-H]- were monitored at m/z 255 for OTα, and m/z

250

221 for OTβ.

251

The compounds were identified by comparing the retention times and mass spectra obtained from

252

samples with those obtained from pure authentic standards injected under the same conditions. OTA and

253

OTB were purchased from a commercial source (Sigma-Aldrich, St. Louis, MO, USA). OTα was

254

synthesized by hydrolysis of OTA as described by Xiao et al.30 OTβ was prepared by acid hydrolysis of

255

OTB.31 Quantitation of the compounds was performed based on measurements of peak areas.

256 257

Statistical analysis

258 259

All statistical analyses were performed using the SPSS software package, version 17.0 for Windows

260

(IBM, Chicago, IL). The OTA contents and the An15g07920 gene expression analyses were evaluated

261

using one-way analysis of variance (ANOVA). Mean differences were determined using Tukey’s

262

post-hoc test (p = 0.01). All figures were plotted using Origin software (Version 8.5, OriginLab Corp.,

263

Northampton, MA, USA).

264 265

RESULTS

266 267

Sequence Analysis

268 269

In a previous study, we have screened the A. niger strain 1062, and found that it can produce OTA in

270

significant amounts.32 In order to understand the mechanism of OTA production in A. niger 1062,

271

genome re-sequencing of this strain was performed. According to the genome annotation of A. niger

12

ACS Paragon Plus Environment

Page 13 of 31

Journal of Agricultural and Food Chemistry

272

CBS 513.88, gene An15g07920, also termed ANI_1_1836134, encodes a hypothetical protein with

273

strong similarity to the PKS fragment involved in OTA biosynthesis in A. ochraceus.15 Thus, we

274

attempted to find the A. niger 1062 orthologue of An15g07920. Comparative analysis revealed a

275

nucleotide fragment in A. niger 1062 showing 100% sequence similarity with the gene An15g07920

276

from A. niger CBS 513.88, indicating that An15g07920 is also present in A. niger 1062. An15g07920 is

277

8173 nt in length and encodes a predicted protein of 2554 amino acids, with a predicted molecular mass

278

of 277.73 kDa and a predicted pI of 5.86. Conserved domain analysis demonstrated that the gene

279

An15g07920 encoded seven highly symptomatic catalytic domains, including KS, AT, DH, C-MT, ER,

280

KR and ACP (Figure 1B).

281 282

Disruption of An15g07920 in A. niger 1062

283 284

To confirm that the product of An15g07920 is indeed involved in OTA synthesis, the sequence fragment

285

of An15g07920 encoding the KS domain was disrupted in A. niger 1062 by replacing it with a

286

hygromycin resistance cassette. The gene replacement vector was constructed based on the upstream and

287

downstream fragments of the target region, which were cloned into the vector p44 flanking the

288

hygromycin resistance marker. The resulting disruption vector (p44-HR1-HR2) was used to obtain the

289

partial KS domain coding region knockout (∆pks) mutant through ATMT (Figure 1 A and B). The

290

hygromycin B-resistant colonies appeared after approximately 4 days of incubation. Genomic DNA was

291

isolated from the transformants, and the site-specific insertions were confirmed by PCR analysis using

292

the primers KS5-F and KS3-R (Table 1) (Figure 1C). In most of the transformants, there were two

293

amplicons with different sizes - one corresponding to the wild-type locus (1776 bp) and the other to the

294

T-DNA construct (2550 bp). However, in the ∆pks strains, there was only one amplicon, which

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 31

295

corresponded to the T-DNA, because the target gene was no longer present. Thus, 6 out of the 50

296

analyzed transformants had integrated the T-DNA cassette at the partial KS domain locus via

297

homologous recombination, which means that the frequency of homologous recombination among the

298

hygromycin-resistant transformants was approximately 12%. We subsequently used qPCR to determine

299

the copy number of the integrated T-DNA cassettes that had been integrated into the genome of A. niger

300

1062 (Table 2). If a ∆pks mutant has one or several extra copies of the T-DNA cassette integrated

301

elsewhere in the genome, the phenotype of the strain cannot be unambiguously attributed solely to the

302

inactivation of An15g07920. Therefore, we selected a single knock-out mutant clone containing a single

303

T-DNA integration for further analysis.

304

Figure 1

305 306

Table 2

307 308 309

Phenotypical analysis

310 311

No statistically significant differences in growth and colony morphology were observed in the ∆pks

312

mutant when compared to the wild-type strain on non-selective media (CYA agar plates) (Figure S1 A).

313

Thus, by day 7, the average colony diameter of the ∆pks mutant was 39.2 ± 0.6 mm and the wild-type

314

reached 38.9 ± 0.8 mm. Similarly, both the ∆pks mutant and wild-type strains exhibited similar

315

sporulation behavior, with 9.5×107 ± 2.6×105 conidia per mL in the wild-type and 9.0×107 ± 4.3×105

316

conidia per mL in the ∆pks mutant at day 7 (Figure S1 C). Additionally, no significant differences of

317

conidial morphology were visible upon SEM observation (Figure S1 B).

14

ACS Paragon Plus Environment

Page 15 of 31

Journal of Agricultural and Food Chemistry

318 319

Figure S1

320 321

Furthermore, in order to investigate the kinetics of the production of OTA and its metabolites, the ∆pks

322

mutant and wild-type strain were grown in CYA liquid medium and the culture supernatants analyzed

323

using HPLC-FLD. The results showed that the metabolite profile of the ∆pks mutant was significantly

324

different from that of the wild-type strain. Three peaks at 4.122, 5.869, and 8.911 min were present only

325

in the wild-type A. niger 1062, and were observable at 2-5 days of incubation, whereby the peak at 8.911

326

min disappeared after day 5. On the other hand, the ∆pks mutant lacked three of those peaks, as observed

327

at days 2-9 (Figure 2). The peaks eluting at 4.122, 5.869, and 8.911 min were identified as OTβ, OTα

328

and OTA, since their retention times were identical to those of the OTB hydrolysis product (OTβ), OTA

329

hydrolysis product (OTα) and OTA standard, respectively. Furthermore, HPLC-MS confirmed the bona

330

fide identities of these compounds. The peak at 4.122 showed a molecular ion [M-H]- at m/z 222.1,

331

corresponding to OTβ, the peak at 5.869 showed the molecular ions [M-H]- at m/z 255.0, corresponding

332

to OTα, and the peak at 8.911 showed a molecular ion [M+H]+ at m/z 404.1 corresponding to OTA

333

(Figure 2). Conversely, OTA was not detected at all in the culture supernatants of the ∆pks mutant,

334

indicating that the disruption of the KS domain of An15g07920 indeed resulted in the inability to

335

produce OTA, as expected. In wild-type A. niger 1062, OTA production was initially observed on day 2

336

and increased to its maximum value on day 3, after which the productions levels decreased markedly by

337

day 5 (Figure 3A). During the first 3 days of incubation, the amount of OTβ in the culture supernatants

338

steadily increased and after the third day its concentration remained practically unchanged. Similarly, the

339

amount of OTα increased until day 6 and remained stable thereafter (Figure 3A). Additionally, the

340

relative expression of the gene An15g07920 was also investigated. The highest expression level was

15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 31

341

detected on day 2 after inoculation and it decreased afterwards, showing a 3-fold higher expression at

342

day 2 compared to day 5 (Figure 3B).

343

Figure 2.

344 345

Figure 3.

346 347 348

DISCUSSION

349 350

When A. niger was given the GRAS status in numerous industrial processes, the potential for OTA

351

production by this species had not been known. With the increasing number of reports showing that

352

some A. niger strains can produce OTA, great concern has been raised about some industrial

353

processes.32-34 It is clear that a thorough characterization of the OTA biosynthetic genes is necessary to

354

determine the molecular triggers which control OTA biosynthesis in A. niger. However, to our best

355

knowledge, a clear involvement of any of the genes in OTA biosynthesis in A. niger had yet to be

356

established at the time of this study. In general, genes encoding enzymes involved in OTA biosynthesis

357

are located in physical proximity to each other, forming gene clusters that usually harbor genes for PKS,

358

NRPS, hydrolases, oxidases, methylases, transporters, and regulatory proteins.35 The in silico analyses

359

performed by Ferracin et al. have shown that the pks-locus tag An15g07920 located in the genome of the

360

ochratoxin-producing A. niger strain CBS 513.88 is absent in the non-ochratoxin-producing strain ATCC

361

1015.20 Moreover, an in vivo analysis of several Brazilian strains has shown that there is an association

362

between the presence of this particular pks gene and the corresponding strains’ capability to synthesize

363

OTA.20 All these clues suggested that An15g07920 in A. niger was closely related to the biosynthesis of

16

ACS Paragon Plus Environment

Page 17 of 31

Journal of Agricultural and Food Chemistry

364

OTA, but prior to this study the gene had not been disrupted and its role in OTA biosynthesis had not

365

been confirmed in A. niger. Thus, we used gene disruption to verify that An15g07920 is indeed required

366

for the biosynthesis of OTA. Subsequent metabolic profiling of the corresponding ∆pks mutant showed

367

that OTA, OTα and OTβ were indeed absent from its culture supernatants.

368

The gene An15g07920 is 8173 nt in length, which means that any homologous genes should most likely

369

also be greater than 8000bp in length. However, PCR-amplification and restriction-fragment based

370

cloning techniques are inefficient, inaccurate, and not always applicable with large DNA segments.

371

Therefore, genome re-sequencing of this strain was carried out in order to characterize the homologous

372

gene An15g07920 in the A. niger 1062 genome. A sequence alignment map was generated by aligning

373

reads from the sequenced pool to the A. niger CBS 513.88 reference sequence. A nucleotide fragment in

374

A. niger 1062 showing 100% similarity with the gene An15g07920 from A. niger CBS 513.88, was

375

found, demonstrating that a gene corresponding to An15g07920 is indeed present in A. niger 1062.

376

Whole genome re-sequencing aims to sequence the genome of a specific strain of an organism for which

377

a reference genome is already available, and in this study we used the genomic sequence of A. niger

378

CBS 513.88 as a basis for the re-sequencing of A. niger 1062. This has become an easy, fast and

379

low-cost method for the identification of target genes.

380

Conidial pigments of A. niger are formed from two precursor molecules - hexahydroxyl pentacyclic

381

quinoid and melanin - whereby the latter is formed from acetate, while OTA biosynthesis begins with

382

one acetate and four malonate molecules condensing to a mullein moiety.36,37 Therefore, acetate is a

383

shared precursor of conidial pigments and OTA. However, our results suggested that the biosynthesis of

384

OTA is independent of the pigments because colony color remained unchanged when An15g07920 was

385

disrupted. One could probably speculate that A. niger does not share enzymes responsible for OTA

386

biosynthesis and pigment synthesis, despite the shared precursor. Furthermore, no statistically significant

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 31

387

differences in growth and colony morphology were observed in the ∆pks mutant, indicating that the

388

results of the gene disruption were quite well segregated from overall metabolism. However, the

389

metabolic network in question is complex and it deserves further study to determine whether or not OTA

390

synthesis gene-encoded enzymes play any additional roles in fungal growth and conidiogenesis.

391

Disruption of the gene An15g07920 completely eliminated the production of OTβ, OTα and OTA during

392

9 days of culture. This result confirmed that the gene An15g07920 is involved in the biosynthetic

393

pathway of OTA and that it plays a role before the step of phenylalanine ligation. We have detected OTB

394

only at a single time-point at day 4 in A. niger 1062 (data not shown). The possible reason is that OTB is

395

not stable and can only accumulate for a short period. We never found OTB ethyl ester or OTC in the

396

culture supernatants of both the 1062 and ∆pks strains. This result is in accordance with a previous report

397

on the intermittent detection of such metabolites in A. carbonarius.9 Esterification of phenylalanine

398

might protect its carboxyl group in the binding reaction leading to OTC.38 However, our results suggest

399

that esterification of phenylalanine is not necessary for the biosynthesis of OTA. Since only OTA was

400

obtained as a pure commercial product, OTβ and OTα were synthesized by ourselves. Consequently, we

401

did not perform a strict quantitative analysis for these compounds, but used the peak areas to quantitate

402

them relatively. Even so, relative quantitative kinetic curves were able to detect increasing OTα

403

production in parallel with a decrease of OTA. One likely explanation of this is that OTα formation is

404

happening directly at the expense of OTA degradation.

405

qRT-PCR was carried out to analyze the expression profile of An15g07920 during the production of

406

OTA by A. niger grown in CYA liquid medium, and the results of this analysis demonstrated a clear

407

correlation between An15g07920 expression and OTA production. The maximum expression of the gene

408

An15g07920 was observed at day 2, while OTA accumulation reached its highest level a day later (day

409

3), confirming that gene transcription preceded OTA detection. A similar timing relationship between

18

ACS Paragon Plus Environment

Page 19 of 31

Journal of Agricultural and Food Chemistry

410

gene transcription and OTA production has also been observed in a number of other OTA-producing

411

Aspergilli. 8,11,22.39

412

In conclusion, inactivation of the gene An15g07920 allowed us to determine its functional role in OTA

413

biosynthesis of A. niger. Future analysis of the OTA gene cluster in this fungus will proceed with the

414

characterization of additional genes found in close proximity to the nrps and pks genes that may be

415

involved in mycotoxin biosynthesis. Complete elucidation of this biosynthetic pathway will not only

416

help explain OTA production, but will also provide necessary information for the development of

417

effective diagnostic, preventive and control strategies to reduce the risk of OTA contamination in foods

418

and other products that utilize this nominally GRAS-status organism in their production processes.

419 420

ASSOCIATED CONTENT

421

Supporting Information description

422

Figure S1. Phenotypic analysis of wild-type (WT) A. niger strain 1062 and the ∆pks mutant strains (A)

423

Colony view of the two strains inoculated in CYA agar plates at different time intervals after inoculation

424

(1 -7 days). (B) SEM images (14,000 × magnification) of conidia derived from CYA agar plates after 4

425

days inoculation. (C) Amount of conidia formation in CYA agar plates at different time intervals after

426

inoculation (1-7 days). The data are the mean ± SD of three determinations from three separate

427

experiments.

428

AUTHOR INFORMATION

429

Corresponding Author

430

* Tel: +86-022-60912431. E-mail: (Y.Z.) [email protected]

431

Notes

432

The authors declare no competing financial interest.

433

Funding Sources 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 31

434

This work was supported by the National Natural Science Foundation of China (31471725, 31370075,

435

and 31201354 ), Tianjin enterprise postdoctoral innovation project (2015); Tianjin funded training

436

selected outstanding postdoctoral program of internationalization (2014).

437 438

REFERENCES

439

(1) International Agency for Research Cancer (IARC). Some naturally occurring substances, food items

440

and constituents, heterocyclic aromatic amines and mycotoxins. IARC monographs on the evaluation of

441

carcinogenic risks to humans. Lyon, France, 1993; Vol.56.

442

(2) Duarte, S.C.; Lino, C.M.; Pena, A. Mycotoxin food and feed regulation and the specific case of

443

ochratoxin A: a review of the worldwide status. Food Addit Contam, Part A. 2010, 27, 1440-1450.

444

(3) European Commission. Commission Regulation (EC) No 1881/2006 of 19 December 2006 setting

445

maximum levels for certain contaminants in foodstuffs. Off. J. Eur. Union 2006, L364, 5-24.

446

(4) Harris, J.P.; Mantle, P.G. Biosynthesis of ochratoxins by Aspergillus ochraceus. Phytochemistry.

447

2001, 58, 709-716.

448

(5) Cox, R.J.; Simpson, T.J. Fungal type I polyketide synthases. Method Enzymol. 2009, 459, 49-78.

449

(6) Geisen, R.; Schmidt-Heydt, M.; Karolewiez, A. A gene cluster of the ochratoxin A biosynthetic genes

450

in Penicillium. Mycotoxin Res. 2006, 22, 134-141.

451

(7) Abbas, A.; Coghlan, A.; O'Callaghan, J.; García-Estrada, C.; Martín, J.F.; Dobson, A.D. Functional

452

characterization of the polyketide synthase gene required for ochratoxin A biosynthesis in Penicillium

453

verrucosum. Int. J. Food Microbiol. 2013, 161, 172-181.

454

(8) Gallo, A.; Knox, B.P.; Bruno, K.S.; Solfrizzo, M.; Baker, S.E.; Perrone, G. Identification and

455

characterization of the polyketide synthase involved in ochratoxin A biosynthesis in Aspergillus

456

carbonarius. Int J Food Microbiol. 2014, 179, 10-7.

20

ACS Paragon Plus Environment

Page 21 of 31

Journal of Agricultural and Food Chemistry

457

(9) Gallo, A.; Bruno, K.S.; Solfrizzo, M.;Perrone, G.; Mulè, G.; Visconti, A.; Baker, S.E. New insight

458

into the ochratoxin A biosynthetic pathway through deletion of a nonribosomal peptide synthetase gene

459

in Aspergillus carbonarius.Appl Environ Microbiol. 2012, 78, 8208-8218

460

(10) Ferrara, M.; Perrone, G.; Gambacorta, L.; Epifani, F.; Solfrizzo, M.; Gallo, A.

461

Halogenase Involved in the Biosynthesis of Ochratoxin A in Aspergillus carbonarius.Appl Environ

462

Microbiol. 2016, 82, 5631-5641

463

(11) Bacha, N.; Atoui, A.; Mathieu, F.; Liboz, T.; Lebrihi, A. Aspergillus westerdijkiae polyketide

464

synthase gene “aoks1” is involved in the biosynthesis of ochratoxin A. Fungal Genet. Biol. 2009, 46,

465

77-84.

466

(12) Wang, L.; Wang, Y.; Wang, Q.; Liu, F.; Selvaraj, J. N.; Liu, L.; Xing, F.; Zhao, Y.; Zhou, L.; Liu,

467

Functional Characterization of New Polyketide Synthase Genes Involved in Ochratoxin A Biosynthesis

468

in Aspergillus Ochraceus fc-1. Toxins 2015, 7, 2723-2738

469

(13) Färber, P.; Geisen, R. Analysis of differentially-expressed Ochratoxin A biosynthesis genes of

470

Penicillium nordicum. Eur J Plant Pathol. 2004, 110(5), 661-669

471

(14) O'Callaghan, J.; Stapleton, P.C.; Dobson, A.D.W. Ochratoxin A biosynthetic genes in Aspergillus

472

ochraceus are differentially regulated by pH and nutritional stimuli. Fungal Genet. Biol. 2006, 43,

473

213-221.

474

(15) Pel, H.J; de Winde, J.H.; Archer, D.B.; Dyer, P. S.; Hofmann, G.; Schaap, P. J.; Turner, G.; de Vries,

475

R. P.; Albang, R.; Albermann, K.; Andersen, M. R.; Bendtsen, J. D.; Benen, J. A E.; van den Berg, M.;

476

Breestraat, S.; Caddick, M. X.;

477

A. J M.; Dekker, P.; van Dijck, P. W M.;

478

Geysens, S.; Goosen, C.; Groot, G. S. P.; de Groot, P. W. J.; Guillemette, T.; Henrissat, B.; Herweijer, M.;

479

van den Hombergh, J. P. T. W.; van den Hondel, C. A. M. J. J.;

Identification of a

Y.

Contreras, R.; Cornell, M.; Coutinho, P. M.; Danchin, E. G J.; Debets, van Dijk, A.; Dijkhuizen, L.; Driessen, A. J. M.;,d’Enfert, C.;

van der Heijden, R. T. J. M.; van der

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 31

480

Kaaij, R. M.; Klis, F. M.; Kools, H. J.; Kubicek, C. P.; van Kuyk, P. A .; Lauber, J.; Lu, X.; vander

481

Maarel, M. J. E. C.; Meulenberg, R.; Menke, H.; Mortimer, M. A.; Nielsen, J.; Oliver, S. G.; Olsthoorn,

482

M.; Pal, K.; van Peij, N. N. M. E.; Ram, A. F J.; Rinas, U.; Roubos, J. A.; Sagt, C. M. J.; Schmoll, M.;

483

Sun, J.; Ussery, D.; Varga, J.; Vervecken, W.; van de Vondervoort, P. J. J.; Wedler, H.; Wo¨sten, H. A. B.;

484

Zeng, A.P.; van Ooyen, A. J. J.; Visser, J.; Stam, H.Genome sequencing and analysis of the versatile cell

485

factory Aspergillus niger CBS 513.88. Nat Biotechnol. 2007, 25, 221-231

486

(16) Han, X.; Chakrabortti, A.; Zhu, J.; Liang, Z. X.; Li, J. Sequencing and functional annotation of the

487

whole genome of the filamentous.fungus Aspergillus westerdijkiae. BMC Genomics. 2016, 17, 633.

488

(17) Gil-Serna, J.; Vázquez, C.; González-Jaén, M. T.; Patiño, B. Clustered array of ochratoxin A

489

biosynthetic genes in Aspergillus steynii and their expression patterns in permissive conditions. Int J

490

Food Microbiol. 2015, 214:102-108.

491

(18) Frisvad, J.C; Larsen, T.O; Thrane, U.; Meijer, M.; Varga, J.; Samson, R.A.; Nielsen, K.F. Fumonisin

492

and ochratoxin production in industrial Aspergillus niger strains. PLoS One. 2011, 6, e23496.

493

(19) Blumenthal, C.Z. Production of toxic metabolites in Aspergillus niger, Aspergillus oryzae, and

494

Trichoderma reesei: Justification of mycotoxin testing in food grade enzyme preparations derived from

495

the three fungi. Regul Toxicol Pharm. 2004, 39, 214-228.

496

(20) Ferracin, L.M.; Fier, C.B.; Vieira, M.L.C.; Vitorello, C.B.M.; Varani, A.M.; Rossi, M.M.; Santos,

497

M.M.; Taniwaki, M.H.; Iamanaka, B.T.; Fungaro, M.H.P. Strain-specific polyketide synthase genes of

498

Aspergillus niger. Int. J. Food Microbiol. 2012, 155, 137-145.

499

(21) Castella, G.; Cabanes, F. J. Development of a real time PCR system for detection of ochratoxin

500

A-producing strains of the Aspergillus niger aggregate. Food Control. 2011, 22, 1367-1372.

22

ACS Paragon Plus Environment

Page 23 of 31

Journal of Agricultural and Food Chemistry

501

(22) Castella, G.; Alborch, L.; Bragulat, M.R.; Cabanes, F.J. Real time quantitative expression study of a

502

polyketide synthase gene related to ochratoxin a biosynthesis in Aspergillus niger. Food Control 2015,

503

53, 147-150.

504

(23) Dave, K.K.; Punekar, N. S. Expression of lactate dehydrogenase in Aspergillus niger for L-lactic

505

acid production.PLoS One. 2015, 10: e0145459.

506

(24) Li, H.; Durbin, R. Fast and accurate short read alignment with Burrows-Wheeler transform.

507

Bioinformatics 2009, 25, 1754-1760.

508

(25) Nadalin, F.; Vezzi, F.; Policriti, A. GapFiller: a de novo assembly approach to fill the gap within

509

paired reads. BMC Bioinformatics. 2012, 13(Suppl 14), S8

510

(26) den Dulk-Ras, A.; Hooykaas, P.J.J. Electroporation of Agrobacterium tumefaciens. Methods Mol

511

Biol.,1995, 55, 63-72.

512

(27) Michielse, C. B.; Hooykaas, P. J.; van den Hondel, C. A.; Ram, A.F. Agrobacterium-mediated

513

transformation of the filamentous Aspergillus awamori. Nat. Protoc. 2008, 3, 1671-1678.

514

(28) Weng, H.B.; Pan, A.H.; Yang, L.T.; Zhang, C.M.; Liu, Z.L.; Zhang, D. B. Estimating transgene

515

copy number by real-time PCR assay using HMG I/Y as an endogenous reference gene in transgenic

516

rapeseed. Plant Mol Biol Rep. 2004, 22, 289-300.

517

(29) Darah, I.; Sumathi, G.; Jain, K.; Lim, S.H. Influence of agitation speed on tannase production and

518

morphology of Aspergillus niger FETL FT3 in submerged fermentation. Appl Biochem Biotech. 2011,

519

165, 1682-1690

520

(30) Xiao, H.; Marquardt, R.R.; Frohlich, A.A.; Ling, Y.Z. Synthesis and structural elucidation of

521

analogs of ochratoxin A. J. Agric. Food Chem. 1995, 43, 524–530.

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 31

522

(31) Bredenkamp, M. W.; Dillen, J. L. M.; van Rooyen, P. H.; Steyn, P. S. Crystal structures and

523

onformational analysis of ochratoxin A and B: probing the chemical structure causing toxicity. J. Chem.

524

Soc. Perkin Trans. II 1989:1835-1839.

525

(32) Zhang, J.; Wang, X. X.; Zhang, Y.; Gao, Q. Screening and identification of an ochratoxin

526

A-producing Aspergillus niger strain. Microbiol. China, 2015, 42, 1010-1016

527

(33) Taniwaki, M.H.; Pitt, J.I.; Teixeira, A.A.; Iamanaka, B.T. The source of ochratoxin A in Brazilian

528

coffee and its formation in relation to processing methods. Int J Food Microbiol. 2003, 82, 173-179

529

(34) Lucchetta, G.; Bazzo, I.; Cortivo, D.G.;Stringher, L.; Bellotto, D.; Borgo, M.; Angelini, E.

530

Occurrence of black Aspergilli and ochratoxin A on grapes in Italy. Toxins, 2010, 2, 840-855

531

(35) Turner, G. Genomics and secondary metabolism in Aspergillus. In: Machida, M.,Gomi, K. (Eds.),

532

Aspergillus: Molecular Biology and Genomics. Caister Academic Press, Norfolk, 2010, pp. 139-155.

533

(36) Jørgensen, T. R.; Park, J.; Arentshorst, M.; van Welzen, A. M.; Lamers, G.; Vankuyk, P.A.; Damveld,

534

R.A.; van den Hondel, C.A.; Nielsen, K. F.; Frisvad, J.C.; Ram, A. F. The molecular and genetic basis of

535

conidial pigmentation in Aspergillus niger. Fungal Genet. Biol. 2011, 48, 544-553

536

(37) Steyn, P. S.; Holzapfel, W. H. The biosynthesis of the ochratoxins, metabolites of Aspergillus

537

ochraceus. Phytochemistry 1970, 9, 1977-1983

538

(38) Huff, W. E.; Hamilton, P. B. Mycotoxins—their biosynthesis in fungi:ochratoxins—metabolites of

539

combined pathways. J. Food Prot. 1979, 42, 815-820.

540

(39) Gallo, A.; Perrone, G.; Solfrizzo, M.; Epifani, F.; Abbas, A.; Dobson, A. D. W.; Mulè, G.

541

Characterisation of a pks gene which is expressed during ochratoxin A production by Aspergillus

542

carbonarius. Int. J. Food Microbiol. 2009, 129, 8-15.

543

24

ACS Paragon Plus Environment

Page 25 of 31

Journal of Agricultural and Food Chemistry

Figure captions

Figure 1. Schematic diagram of gene An15g07920 disruption. (A) Map of plasmid p44-HR1-HR2. (B) Sketch map of the replacement of the targeting gene using a homologous recombination method. (C) Transformants confirmed by the PCR method: M, marker (DL5000); lanes 1 , wild-type strain; lanes 2 , ectopic transformant; lanes 3, ∆pks mutants.

Figure 2. HPLC-FLD chromatograms and mass spectra. (A) Culture medium control; (B) Culture supernatant of wild-type A. niger strain 1062; (C) Culture supernatant of the ∆pks mutant strain. Retention times: OTA: 8.911 min; OTα: 5.869 min; OTβ , 4.122 min.

Figure 3. Kinetics of OTA, OTβ and OTα production (A) and Time-profile of the expression of the gene An15g07920 (B) in A. niger 1062 during growth on CYA media

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 31

Tables

Table 1 Primers used in this study Primers

Function

Sequence (5’ to 3’)

KS5-F

upstream fragment amplification

CCCAAGCTTTGCATGCAAGTACGCCAATGG

KS5-R

upstream fragment amplification

TGCACTGCAGTGGTGGGTAAGGCTATCAGAC

KS3-F

downstream fragment amplification

GGGGTACCGATTCCTCAACGCGATGACCT

KS3-R

downstream fragment amplification

CGGAATTCGAATGCTGCAAGGACTCGC

HPHq-F

qPCR for gene copy number (target gene)

GGCTCCAACAATGTCCTGAC

HPHq-R

qPCR for gene copy number (target gene)

CGTCTGCTGCTCCATACAAG

ACTq-F

qPCR for gene copy number (reference gene)

TTGCGGTACAGCCTCCATTG

ACTq-R

qPCR for gene copy number (reference gene)

CGCTTGGACTGTGCCTCATC

KSqrt-F

qRT-PCR (target gene)

CGGATGACCTCTAAAGCAG

KSqrt-R

qRT-PCR (target gene)

GAAGAATGTCCCACCACC

Tubqrt-F

qRT-PCR (reference gene)

CTTCTGACACGCAGGATAG

Tubqrt-R

qRT-PCR (reference gene)

ACGGCACGAGGAACATAC

26

ACS Paragon Plus Environment

Page 27 of 31

Journal of Agricultural and Food Chemistry

Table 2 Estimation of the number of T-DNA copies that have been integrated in the genome of the mutants Slope

Inter

Ct HYG

Ct actin

Copy number

HYG

Actin

HYG

Actin

Wild type

△pks

Wild type

△pks

Wild type

△pks

-4.117

-3.399

38.053

30.049

30.97

21.46

16.96

16.69

0.008

1.279

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 31

Figure graphics

A

C

B

Figure 1

28

ACS Paragon Plus Environment

Page 29 of 31

Journal of Agricultural and Food Chemistry

LU 17.5 15 12.5 10 7.5 5 2.5

A

0 LU 17.5 15 12.5 10 7.5 5 2.5

2

4

B

8

OTα

OTβ

0 LU 17.5 15 12.5 10 7.5 5 2.5

6

10

min

OTA

2

4

6

8

10

min

2

4

6

8

10

min

C

221.1

0 100

OTβ

Max: 1978

80

60

368.1

293.1

278.8

255.1

216.9

198.8

179.0

187.0

169.0

158.9

125.1

20

132.9

40

0 200

250

300

OTα

m/z

350

255.0

150 100

Max: 1432

80

556.0

380.9

300.9

232.7

217.0

191.1

160.8

20

172.9

129.2

40

255.9

195.9

60

0 200

300

400

80

OTA

m/z

500

404.1

100

100

Max: 3471

429.4

414.1

365.3

377.3

328.2

346.0

295.4

314.3

270.2

230.9 226.3

243.9

194.0 205.1

170.1 181.1

150.2

136.1

120.8

103.9

40

20

406.0

60

0 100

200

300

400

500

m/z

Figure. 2.

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

B

700

OTA OTα OTβ

Peak area (LU*S)

600

1.2

Differences expressed in multiples

A

500

400

300

200

100

0

Page 30 of 31

1.0

0.8

0.6

0.4

0.2

0.0

1

2

3

4

5

6

7

8

9

10

2

Time (d)

3

4

5

6

Time (d)

Figure 3.

30

ACS Paragon Plus Environment

Page 31 of 31

Journal of Agricultural and Food Chemistry

TOC Graphic

31

ACS Paragon Plus Environment