(SCD1) deficiency affects fatty acid metabolism in ... - ACS Publications

The server is currently under maintenance and some features are disabled. Share Article. ACS Network; Twitter; Facebook; Google+; CiteULike; Email...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of South Dakota

Biotechnology and Biological Transformations

CRISPR/Cas9-mediated Stearoyl-CoA desaturase 1 (SCD1) deficiency affects fatty acid metabolism in goat mammary epithelial cells Huibin Tian, Jun Luo, Zhifei Zhang, Jiao Wu, Tianying Zhang, Sebastiano Busato, Lian Huang, Ning Song, and Massimo Bionaz J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b03545 • Publication Date (Web): 05 Sep 2018 Downloaded from http://pubs.acs.org on September 5, 2018

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

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 47

Journal of Agricultural and Food Chemistry

1

CRISPR/Cas9-mediated Stearoyl-CoA desaturase 1 (SCD1) deficiency affects

2

fatty acid metabolism in goat mammary epithelial cells

3

Huibin Tian†, Jun Luo†*, Zhifei Zhang†, Jiao Wu†, Tianying Zhang†, Sebastiano

4

Busato#, Lian Huang†, Ning Song†, Massimo Bionaz#*

5

†Shaanxi Key Laboratory of Molecular Biology for Agriculture, College of Animal

6

Science and Technology, Northwest A&F University, Yangling, 712100, China

7

#Department of Animal and Rangeland Sciences, Oregon State University, Corvallis

8

97331, USA

9

*Corresponding authors: [email protected]. and

10

[email protected]

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

11

ABSTRACT

12

Stearoyl-CoA desaturase 1 (SCD1) is a fatty acid desaturase catalyzing the cis

13

double bond formation in ∆9 position to produce monounsaturated fatty acids

14

essential for the synthesis of milk fat. Previous studies using RNAi methods have

15

provided support for a role of SCD1 in goat mammary epithelial cells (GMEC);

16

however, RNAi present several limitations that might preclude a truthful

17

understanding of the biological function of SCD1. To explore the function of SCD1

18

on fatty acid metabolism in GMEC, we used CRISPR/Cas9-mediated SCD1 knockout

19

through NHEJ (Non-Homologous End-Joining) and HDR (Homology-Directed

20

Repair) pathways in GMEC. We successfully introduced nucleotides deletion and

21

mutation in the SCD1 gene locus through NHEJ pathway, and disrupted its second

22

exon via insertion of an EGFP-PuroR segment using HDR pathway. In clones derived

23

from the latter, gene and protein expression data indicated that we obtained a

24

monoallelic SCD1 knockout. A T7EN1-mediated assay revealed absence of no

25

off-targets in the surveyed sites. The content of triacylglycerol, cholesterol and

26

desaturase index were significantly decreased as consequence of SCD1 knockout. The

27

deletion of SCD1 decreased the expression of other genes involved in de novo fatty

28

acid synthesis, including SREBF1 and FASN, as well the fatty acid transporters FABP3

29

and FABP4. The downregulation of these genes partly explains the decrease of

30

intracellular triacylglycerols. Our results indicate a successful SCD1 knockout in goat

31

mammary cells using CRISPR/Cas9. The demonstration of the successful use of

32

CRISPR/Cas9 in GMEC is an important step to produce transgenic goats to study 2

ACS Paragon Plus Environment

Page 2 of 47

Page 3 of 47

Journal of Agricultural and Food Chemistry

33

mammary biology in vivo.

34

KEYWORDS: CRISPR/Cas9, SCD1, NHEJ, HDR, goat mammary epithelial cells

35

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

36

INTRODUCTION

37

Dairy goats are of increasing global importance among grazing livestock, with

38

steadily growing populations, especially among developing countries1, 2. Furthermore,

39

goat milk has been under the spotlight in recent years3, 4, for it is a significant source

40

of short-chain fatty acids and monounsaturated fatty acids (MUFA) with reported

41

beneficial roles in human health5. The formation of the MUFA palmitoleate (16:1 n-7)

42

and oleate (18:1 n-9) relies on the activity of stearoyl-CoA desaturase 1 (SCD1), a

43

∆9-fatty acyl CoA desaturase that catalyzes the synthesis of double bonds in

44

cis-delta-9 position of long-chain fatty acids6. In lactating ruminants, the expression

45

of mammary SCD1 increases dramatically concomitantly with a large increase in fatty

46

acid synthesis during lactation7, 8. The important role of SCD1 in triacylglycerols

47

synthesis and fatty acid composition was recently confirmed in goat mammary

48

epithelial cells (GMEC) using a combination of gene overexpression and gene

49

silencing techniques9. The same study also identified SREBP1 and PPARγ as

50

important transcriptional regulators of SCD1. A partial loss-of-function approach,

51

such as the use of RNA interference (RNAi) can offer insight on gene function;

52

however, this technique presents several limitations, including low and inconsistent

53

efficiency of gene silencing10, and short-term inhibition of gene expression11.

54

Furthermore, RNAi does not allow to conduct studies when specific gene and protein

55

dosages are crucial12. Studies of the transcriptional and biological role of SCD1 would

56

undoubtedly benefit from more consistent, long term knockout models.

57

A permanent loss-of-function model might be achieved through the novel 4

ACS Paragon Plus Environment

Page 4 of 47

Page 5 of 47

Journal of Agricultural and Food Chemistry

58

Clustered Regularly Interspaced Short Palindromic Repeats (CRISPR)-mediated

59

gene editing technique. CRISPR has been used successfully in a variety of species,

60

such as bacteria, yeasts, fruit flies, mice and monkeys13-16. Because of its low cost,

61

great flexibility, and high efficiency in genome editing, it has been applied for

62

generating gene modified animal models of human diseases17, 18, to obtain transgenic

63

pigs19 and to perform in vivo study of long non-coding RNA20. CRISPR models

64

utilize an endogenous endonuclease (such as the CRISPR-associated protein 9, or

65

Cas9) to cause a DNA double strand break (DSB) at a specific site, to which the

66

nuclease is directed by a single-guide RNA (sgRNA). The cells have the ability to

67

repair their DNA21 by two main pathways: the error-prone Non-Homologous

68

End-Joining (NHEJ), which can cause random nucleotide insertion, deletion and

69

mutation22, inducing a frame shift and a subsequent early stop codon; or the more

70

accurate Homology-Directed Repair (HDR)23, which integrates a homologous DNA

71

sequence (i.e. a sister chromatin or a plasmid)24, 25.

72

While global26, 27 or tissue specific knockout28-30 have been carried out to study

73

the physiological function of SCD1 in mice, no evidence of such endeavors exists in

74

goats. Compared to mice, the identification of the role of SCD1 in lactating dairy

75

goats is of greater importance: MUFA in goat milk are mostly obtained via

76

endogenous desaturation, considering that fatty acid biohydrogenation in ruminant

77

animals markedly decreases diet-sourced unsaturated FA31.

78

This study aims at providing crucial insight on the transcriptional regulation of

79

SCD1 and its role in milk fatty acid synthesis using a CRISPR/Cas9-mediated 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

80

knockout model, through both the NHEJ and HDR repair pathways in GMEC.

81

MATERIALS AND METHODS

82

Ethics Statement. All the experimental procedures were carried out in accordance

83

with the Institutional Animal Care and Use Committee in the College of Animal

84

Science and Technology, Northwest A&F University, Yang Ling, China (permit

85

number: 15-516, date: 2015-9-13).

86

Cell culture and suitable concentration of puromycin. GMEC were isolated from

87

five Xinong Saanen dairy goats at peak lactation (60 d after parturition) as described

88

previously32, and were purified and cultured according to our previously described

89

protocol33. Briefly, GMEC were isolated by tissue block preparation technique. The

90

mammary gland tissue was cut into about 1mm3. The tissue block was cultured in 5%

91

CO2 at 37℃ with basal medium. The culture medium was changed every two or three

92

days until the epithelial cells separated from the tissue block. Then the cells were

93

digested from tissue block. The basal culture medium contained DMEM/F12 medium

94

(11320-033, Invitrogen Corp., Waltham, MA, USA), 10% fetal bovine serum

95

(10099-141, Invitrogen), 5 µg/mL bovine insulin (16634, Sigma, St. Louis, MO), 5

96

mg/L hydrocortisone (H0888, Sigma), 100 U/mL penicillin/streptomycin (080092569,

97

Harbin Pharmaceutical Group, Harbin, P. R. China), and 10 ng/mL epidermal growth

98

factor (PHG0311, Invitrogen). GMEC were incubated in 5% CO2 at 37℃ and

99

medium was changed every 24 hours. To promote lactogenesis, the cells were

100

cultured in the basal medium with prolactin (L6520, 2 µg/mL, Sigma) for 48 h before

101

performing the following experiments. 6

ACS Paragon Plus Environment

Page 6 of 47

Page 7 of 47

Journal of Agricultural and Food Chemistry

102

GMEC were seeded in six-well plate, and when the cells approached to 80-90%

103

confluence, puromycin (P8833, Sigma) was added to the culture medium at 0, 0.5, 1.0,

104

or 1.5 µg/mL concentration. After cultured for 72 hours, the cell death condition was

105

examined by microscope. The lowest lethal dose of puromycin to the GMEC was 1.0

106

µg/mL.

107

Construction of Cas9/sgRNA expression vector and HDR donor vector. The

108

sgRNAs which targeted to exon 2 of Capra hircus SCD1 were designed using the

109

online

110

(http://chopchop.cbu.uib.no/)34. Three sgRNAs (Figure 1A) were selected based on

111

their predicted score and lower off-target effects for Cas9/sgRNA expression vector

112

construction. For the NHEJ-mediated pathway, we chose the all-in-one vector

113

pSpCas9 (BB)-2A-Puro (plasmid#62988, PX459 V2.0, Addgene, MA, USA), a gift

114

from Feng Zhang35. The sgRNAs were synthesized as single-strand DNA

115

oligonucleotides by Invitrogen (Shanghai, China), and annealed oligonucleotides were

116

inserted into PX459 vector containing two BbsⅠ (R3539S, NEB, Ipswich, MA, USA)

117

enzyme sites according to a previous protocol35. The sgRNA with higher efficiency

118

was used for the single clone selection in both NHEJ- and HDR-mediated genome

119

editing.

CRISPR

design

tools

(http://crispr.mit.edu/)

and

CHOPCHOP

120

For HDR-mediated genome editing, pX330-U6-Chimeric_BB-CBh-hSpCas9

121

(plasmid#42230, PX330, Addgene, MA, USA, a gift from Feng Zhang) was used as

122

the Cas9/sgRNA co-expression plasmid36, inserting a sgRNA using the two BbsⅠ

123

enzyme sites. The donor vector (a gift from Key Laboratory of Animal Biotechnology, 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

124

Ministry of Agriculture, Northwest A&F University), used to deliver the homology

125

arm, contained the coding sequence of puromycin resistance and enhanced green

126

fluorescent protein (eGFP) as selection markers. These two genes were fused through

127

the porcine teschovirus-1 2A (P2A) peptide sequence37, which allows co-expression

128

of the two constructs.

129

Genomic DNA was extracted from blood samples of five goats using a Universal

130

Genomic DNA kit (CW2298S; CW Biotech, Beijing, China). A 1193 bp 5’

131

homologous arm fragment and 1063 bp 3’ homologous arm fragment were amplified

132

by PCR using PrimeSTAR® Max DNA Polymerase (R045A, Takara Bio Inc., Otsu,

133

Japan) according to the manufacturer’s protocol. The reaction was performed at 98℃

134

for 10 s, followed by 55℃ for 5 s, and 72℃ for 10 s. 5’ arm and 3’ arm clone primers

135

are shown in Table 1. The PCR products were sequenced by Invitrogen (Shanghai,

136

China). Then 5’ arm and 3’ arm fragments were inserted into the donor vector. Two

137

LoxPs in the same orientation were inserted around the selection markers, just in case

138

they were required to be removed38. A schematic representation of the donor vector

139

can be observed in Figure 2A.

140

Plasmid transfection, DNA extraction and T7EN1 assay. GMEC were cultured in

141

six-well plates to 70-80% confluence. For the NHEJ pathway, 2µg PX459 was

142

transfected using Lipofectamine™ 2000 (11668019, Invitrogen, Waltham, MA, USA)

143

according to the manufacturer’s protocol. Forty-eight hours after transfection, the

144

cells were cultured in basal medium with puromycin for four to five days. Viable cells

145

were collected for genomic DNA extraction using a Universal Genomic DNA Kit 8

ACS Paragon Plus Environment

Page 8 of 47

Page 9 of 47

Journal of Agricultural and Food Chemistry

146

(CWBIO, China). The genomic region flanking the target site was PCR amplified

147

using the test primers for NHEJ (Table 1). The PCR products were purified by PCR

148

Clean-Up Kit (AP-PCR-50, Axygen, CA, USA) according to the manufacturers’

149

instructions. Purified DNA was annealed for T7EN1 cleavage assay39 (M0302L, NEB,

150

Ipswich, MA, USA) and the enzyme digestion product was analyzed by agarose gel

151

electrophoresis. Cleaved bands intensity were measured by ImageJ software

152

(ImageLab, http://imagej.net). The frequency of PCR product enzyme digestion fcut

153

was determined by the formula (a+b)/(a+b+c), where a is the intensity of the

154

undigested PCR product and b and c are the intensities of each cleavage bands. The

155

indel occurrence in the DNA was estimated based on the binomial probability

156

distribution of duplex formation as following: Indel % = (1 − (1 −  )) ∗ 100

157 158

Then PCR products were cloned into a pMD19-T vector and sequenced by Invitrogen (Shanghai, China) to assess sequence modification in the cell pools.

159

For the HDR pathway, 1 µg of donor vector together with 1 µg of PX330 were

160

transfected into GMEC using Lipofectamine™ 2000 (Invitrogen, USA). Puromycin

161

selection was performed 48 h after transfection for four to five days until co-incubated

162

wild-type cells, treated with the same antibiotic concentration, had 100% mortality. A

163

separate batch of cells, transfected with the donor vector and a PX330 that did not

164

contain any sgRNA, were used as a control.

165

Detection of individual GMEC clones by PCR and off-target analysis. After

166

selection, cells were counted using a hemocytometer and diluted to a final 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

167

concentration of 1 cell/100 µL. Individual cells were then plated in 96-well plates and

168

cultured for 10-14 days to obtain single clone colonies. The medium for the cell

169

culture was the basal culture medium containing DMEM/F12 medium, 10% fetal

170

bovine serum, 5 µg/mL bovine insulin, 5 mg/L hydrocortisone, 100 U/mL

171

penicillin/streptomycin, and 10 ng/mL epidermal growth factor. Cells from each

172

colony were collected by trypsinization. Half of the cells were plated in the 48-well

173

plates, and the remaining cells were collected in a 1.5 mL tube, and 1 mL PBS buffer

174

was used to resuspend the cells and then the cells were centrifuged for 5 minutes at

175

400×g to discard the PBS buffer. This process was performed for three times to wash

176

the cells and cells were resuspended in lysis buffer (10 mM Tris-HCl; 50 mM KCl;

177

1.5 mM MgCl2; 0.5% Tween-20; 100 ng/µL proteinase K) for PCR analysis. The

178

lysate was incubated at 56℃ for 45 min and then at 95℃ for 10 min. For PCR

179

analysis, 2 µL of the cell lysate was added to the PCR reaction. For the

180

NHEJ-mediated knockout, test primers for NHEJ (Table 1) were used to amplify the

181

region including sgRNA target sites, while for the HDR-mediated knockout, 5’

182

junction primer and 3’ junction PCR primer (Table 1) were employed, and the PCR

183

strategy is shown in Figure 2A. The reaction conditions were as follows: 98℃ for 10 s,

184

55℃ for 5 s, and 72℃ for 5 s using PrimeSTAR® Max DNA Polymerase (Takara).

185

Mutations in the HDR single clone was also confirmed by measuring eGFP

186

fluorescence using a Leica fluorescent microscope (DMI4000B, Wetzlar, Germany).

187

Wild type cells were used as control under the same exposure and filter settings.

188

Off-target (OT) sites were predicted using the online website tool Cas-OFFinder 10

ACS Paragon Plus Environment

Page 10 of 47

Page 11 of 47

Journal of Agricultural and Food Chemistry

189

(http://www.rgenome.net/cas-offinder/)40. Mismatches ≤ 4 bp was used as criteria41, 42.

190

Genomic DNA extracted from both HDR single clone and NHEJ single clone were

191

used as templates for off-target sites PCR. T7EN1 assay was used and the PCR

192

product were inserted into pMD19-T vector for sequencing. The primers for off-target

193

detection are shown in Suppl. Table 2.

194

RNA extraction and real-time quantitative PCR (RT-qPCR). Total RNA was

195

extracted from GMEC with RNAiso Plus (9109, Takara) according to the

196

manufacturer’s protocol and 0.5 µg of total RNA was used to synthesize cDNA with

197

PrimeScript® RT Reagent Kit with gDNA Eraser (RR047A, Perfect Real Time,

198

Takara). cDNA was diluted to 400 ng/uL for each sample. RT-qPCR were performed

199

using SYBR Premix Ex Taq Ⅱ (RR820A, Perfect Real Time, Takara). All the primers

200

used for RT-qPCR are listed in the Suppl. Table 1. The real-time PCR reactions were

201

performed in a CFX-96 Real-Time PCR Detection system (Bio-Rad Laboratories Inc.,

202

Hercules, CA). The reactions were performed as 95℃ for 30 s, followed by 40 cycles

203

of 95℃ for 5 s and 60℃ for 30 s; a dissociation curve was performed at 95℃ for 10 s

204

and then from 65℃ to 95℃ with a 0.5℃ increase. The RDML files were used to

205

calculate the RT-qPCR values for each gene and sample using LinRegPCR43. Four

206

transcripts were tested as potential internal controls or reference genes using

207

geNorm44, these were ribosomal protein S9 (RPS9), ribosomal protein S15 (RPS15),

208

mitochondrial ribosomal protein L39 (MRPL39), and ubiquitously expressed

209

transcript (UXT)45. The normalization factor was calculated using the expression of

210

RPS9, MRPL39, and UXT for both HDR and NHEJ experiment. The V-value using 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

211

these three genes was < 0.10 for both experiments45.

212

Protein extraction and western blot analysis. Cells were collected and lysed in

213

ice-cold RIPA buffer (R0010, Solarbio, Beijing, China) with protease inhibitor

214

(04693132001, Roche Diagnostics Ltd, Mannheim, Germany). Protein concentration

215

was measured using BCA protein assay kit (23227, Thermo Fisher Scientific,

216

Rockford, IL). The protein was separated with 10% SDS/PAGE, transferred onto

217

nitrocellulose membrane (HATF00010, Millipore, Massachusetts, USA) by a Bio-Rad

218

Trans-Blot SD semi-dry transfer cell (Bio-Rad Laboratories Inc., Hercules, CA, USA),

219

and the membranes were blocked for 1.5 h using 5% skim milk (232100, BD,

220

Franklin Lakes, New Jersey, USA). The membranes were incubated overnight at 4℃

221

with primary antibodies for SCD1 (Cat#ab39969, Abcam, Cambridge, MA, 1:300)9,

222

SREBP1 (Cat#ab3259, Abcam, Cambridge, MA, 1:1000)46and β-actin (CW0096,

223

1:1000; CW Biotech, Beijing, China). After the membranes were washed with TBST

224

for three times, horseradish peroxidase (HRP)-conjugated goat anti-rabbit (CW0103,

225

1:2000; CW Biotech) and goat anti-mouse IgG (CW0102, 1:2000; CW Biotech) were

226

used as secondary antibodies. Signals were measured using an enhanced

227

chemiluminescent (ECL) Western blot system (1705061, Bio-rad). The density of the

228

bands was analyzed by Image J and the relative expression of protein was normalized

229

to β-actin.

230

Measurement of total cellular triacylglycerol (TAG) and cholesterol. Cells were

231

plated in 60 mm culture dishes and, upon reaching 90% confluence, cellular TAG and

232

cholesterol were detected with GPO-Trinder triacylglycerol assay kit (E1013, 12

ACS Paragon Plus Environment

Page 12 of 47

Page 13 of 47

Journal of Agricultural and Food Chemistry

233

Applypen Technologies Inc., Beijing, China) and cholesterol assay kit (E1015,

234

Applygen Technologies Inc.) as described previously47. In brief, culture medium was

235

discarded, and cells were washed for three times using PBS buffer. Then GMEC were

236

treated with lysis buffer for 10 minutes and cells were scraped. The supernatant was

237

collected by centrifugation for TAG and cholesterol assay at 550 nm using a Biotek

238

microplate reader (Winooski, VT, USA). The amount of total protein was detected

239

using BCA protein assay kit (Thermo Fisher Scientific). Cellular TAG and cholesterol

240

concentrations were normalized by total protein and displayed as µg/mg protein.

241

Measurement of fatty acids in SCD1 knockout GMEC. The SCD1 knockout

242

GMEC and control group were cultured in 60mm culture dishes until 90% confluence

243

and the cells were washed with 1 mL PBS buffer for three times. Total fatty acid was

244

extracted using 2 mL aliquot of 2.5% (vol/vol) sulfuric acid :methanol and then

245

transferred to a 10 mL glass tube for methyl esterification as previously described46.

246

Methylated lipid samples were analyzed by gas chromatography (Agilent 7890A;

247

Agilent Technologies Inc., Santa Clara, CA) using a 100 meters HP-5 column

248

(Agilent Technologies Inc.) and the hydrogen flame ionization detector as previously

249

described48. Relative proportions of C16:0, C16:1, C18:0 and C18:1 in GMEC were

250

determined as percentage of the total peak area that could be identified48. The

251

desaturation index was calculated as the ratio of unsaturated fatty acid to the sum of

252

unsaturated and saturated fatty acids, as previously reported49.

253

Statistical analysis. Statistical analysis for gene expression was performed using

254

GLM procedure of SAS (v9.4) with cell type (knockout, wild type) as main effect and 13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

255

replicates as random. For the other parameters the statistical analysis was performed

256

using SPSS 19.0 statistics software (SPSS, Inc., Chicago, IL) using Student’s t test.

257

Significant differences between the groups were considered significant at *P < 0.05.

258

RESULTS

259

Cleavage efficiency detection and construction of HDR donor vector. After treated

260

with different concentrations of puromycin, the minimal lethal dose was 1.0 µg/ml for

261

GMEC (Suppl. Figure 1). This concentration was used for cell selection. Among three

262

sgRNAs, sgRNA2 had no significant cleavage, whereas sgRNA1 and sgRNA3 had

263

8.3% and 19.4% cleavage efficiency, respectively (Figure 1B). There were different

264

types of nucleotide indels in these cell pools (Figure 1D). Based on above results, we

265

selected the sgRNA3 for the subsequent experiment.

266

CRISPR/Cas9 induced sequence modification at SCD1 gene locus by NHEJ and

267

HDR repair pathways. For the NHEJ pathway, we obtained 42 single clones and

268

only one clone was identified with a gene modification representing 24 nucleotides

269

deletion and 1 nucleotide mutation in the SCD1 exon (Figure 1C and 1E). These

270

results indicated the Cas9/sgRNA system induced double strand break in GMEC,

271

causing DNA fragments indels.

272

For the HDR pathway, among 50 single clones obtained, we detected two clones

273

that had both the 5’ homology arm and 3’ homology arm insertion, and one single

274

clone only had the 5’ homology arm insertion (Figure 2B). Among the two clones

275

with both arms insertion, only for clone 10, that had a smaller insertion in the 5’

276

compared to clone 6 (Figure 2B), cells survived and proliferated. The green 14

ACS Paragon Plus Environment

Page 14 of 47

Page 15 of 47

Journal of Agricultural and Food Chemistry

277

fluorescence could be detected in all the cells under the fluorescent microscope with

278

the clone 10 (Figure 2C) indicating successful insertion of GFP and puromycin coding

279

section in the SCD1 gene locus. The sequence in the genome at this site was shown in

280

supplementary material.

281

The test primer for NHEJ pathway was used to examine if there was

282

NHEJ-mediated genome editing in the allele in HDR pathway mediated knockout

283

cells. PCR was performed by using primers for NHEJ and using the genome DNA

284

extracted from HDR group cells as templates. Small DNA fragment were amplified.

285

This indicated only one allele contained GFP and puromycin insertion. Through

286

T7EN1 assay, no nucleotides mutation was detected in this allele that did not contain

287

GFP and puromycin insertion (Suppl. Figure 2). These results suggested that the

288

SCD1 gene locus could be repaired by HDR pathway and did not generate the indels

289

caused by NHEJ repair pathway. Taken together, these two single clones constructed

290

by NHEJ and HDR pathways had genome sequence modifications of SCD1 gene

291

locus and they were selected for further analysis.

292

The expression of SCD1 decreased in genome modified GMEC without off-target

293

effect. For off-targets detection, ten off-target sites (Figure 3A) were chosen for

294

examination by T7EN1 assay (Figure 3B and 3C). The cleavage bands were visible in

295

OT3 and OT6 both in wild type and knockout cells. These results indicated that no

296

off-targets were detected in knockout cells and there were nucleotide mutations in the

297

genome of wild type cells at off-target site 3 and 6 (Suppl. Figure 3).

298

The relative mRNA and protein expression were measured by RT-qPCR and 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

299

Western blot. The mRNA level of SCD1 decreased by about 80% in both HDR-and

300

NHEJ-mediated genome editing, and the protein level of SCD1 decreased about 50%

301

(Figure 4A and 4B). These data indicated that the use of either HDR and NHEJ and

302

selection of the viable clone resulted in monoallelic knockout of SCD1.

303

Knockout of SCD1 affected TAG, cholesterol, and the desaturation index.

304

Knockout of SCD1 in GMEC decreased significantly the content of cellular TAG and

305

cholesterol (P < 0.05) (Figure 5). The 16:1 and C18:1 desaturation index was

306

decreased significantly (P < 0.05) as consequence of the knockout of SCD1 in GMEC

307

(Figure 6). The percentage of C16:0 was increased (P < 0.05) while the percentage of

308

C16:1 cis7 (P < 0.05) and C18:1 cis9 (P < 0.01) were decreased in SCD1 knockout

309

cells compared to wild type. However, there was no significant difference for C18:0

310

between SCD1 knockout and wild type GMEC (Table 2).

311

Knockout of SCD1 affected the expression of genes related to de novo fatty acid

312

synthesis. To determine whether the knockout of SCD1 was the sole cause of the

313

decrease in TAG, cholesterol and unsaturated fatty acids and if it affected the

314

expression of other genes related to milk fat synthesis, we assessed the expression of

315

several genes related to fatty acid metabolism. Compared with the wild type GMEC,

316

knocking out SCD1 gene both using HDR or NHEJ pathway decreased the expression

317

of fatty acid transporters FABP3 and FABP4, the key de novo fatty acid gene FASN,

318

ACACA (with a tendency of P = 0.08 when HDR was used), and SREBF1 (Figure 7, 8

319

and 9). The use of NHEJ pathway also decreased the expression of ELOVL5 but

320

increased the expression of ELOVL6. The use of HDR pathway tended (P = 0.07) to 16

ACS Paragon Plus Environment

Page 16 of 47

Page 17 of 47

Journal of Agricultural and Food Chemistry

321

decrease the transcription of FADS1. The transcription of DGAT2 was significantly

322

increased when the HDR pathway was used but was only numerically increased when

323

NEHJ pathway was used. No other genes were affected by the knockout of SCD1

324

(Figure 7 and 8).

325

DISCUSSION

326

CRISPR/Cas9-mediated genome editing provides an easier and more efficient

327

method to obtain gene modification in any type of cell11, 50 compared with other

328

genome editing systems, like zinc-finger nucleases51 and transcription activator-like

329

effector nucleases52. To our knowledge, this is the first study where a genome editing

330

was performed using CRISPR/Cas9 technology by both NHEJ and HDR pathways in

331

goat mammary cells. Using the NHEJ pathway we obtained monoallelic SCD1

332

knockout by the deletion of eight amino acids and one amino acid mutation, whereas

333

in the HDR pathway the selection markers were inserted into the SCD1 genomic

334

sequence and the presence of a stop codon in the puromycin-coding sequence induced

335

a transcriptional termination, shortening the mRNA of SCD1. Taken together, our

336

results demonstrated that both the NHEJ- and HDR-mediated gene editing can be

337

applied successfully in primary goat mammary gland epithelial cells.

338

RNAi has significant off-target effects and unpredictable knockdown efficiencies

339

53

340

CRISPR/Cas9 technology has the advantage of complete knockout for a certain gene.

341

RNAi requires endogenous factors to form RISC (RNA-induced silencing complex) 54

342

while CRISPR system uses the exogenous crRNA and tracrRNA together with Cas9

. When a residual expression is enough for the protein activity, compared to RNAi,

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

343

endonuclease. As a result, the efficiency of RNAi may be more dependent from the

344

status of the cell55. A previous report indicated that compared with shRNA-based

345

system and CRISPR-interference (CRISPRi), the CRISPR technology performed best

346

in identifying essential genes with less off-target effect and consistency in different

347

cell lines56. In ruminant mammary gland, CRISPR/Cas9 genome editing tool also

348

provides a more reliable method to study fatty acid metabolism. In our experiment the

349

expression of DGAT2 was increased significantly in SCD1 knockout GMEC. This is

350

contrary to what previously observed using siRNA method9. The differences on the

351

effect of DGAT2 between the two studies might be due to off-target effects usually

352

obtained using RNAi but not CRISPR/Cas9.

353

In the HDR-directed knockout method, three single clones were selected with

354

three different status. Bands corresponding to the 3’ and 5’ junctions displayed higher

355

intensity in clone 6 when compared to clone 10 likely due to a more complete biallelic

356

deletion of SCD1; the colony, however, could not be cultured for an extended period

357

of time and underwent apoptosis in the first few days. The likely biallelic deletion of

358

SCD1 could have affected the viability of the clone, considering that SCD1 is

359

essential for the de novo formation of monounsaturated fatty acids. Furthermore, an

360

increase of saturated fatty acids can be toxic when excessive. A possible explanation

361

for this phenomenon is the presence of a biallelic SCD1 knockout in clone 6. As

362

consequence of the complete SCD1 knockout the cells could have been deprived of

363

adequate amount of MUFA, an essential substrate for phospholipid synthesis, an

364

important component of the cell membrane57. Further, SFA accumulation can trigger 18

ACS Paragon Plus Environment

Page 18 of 47

Page 19 of 47

Journal of Agricultural and Food Chemistry

365

endoplasmic reticulum stress, which would ultimately lead to cell death58.

366

Interestingly, biallelic SCD1 knockout mice are viable59,

367

monogastrics, ruminants absorb from the diet little amount of unsaturated fatty acids,

368

owned to the presence of the large biohydrogenation of unsaturated fatty acids in the

369

rumen61. Thus, it is possible that cells in ruminants are more dependent from the

370

intracellular desaturation, mainly driven by the activity of SCD1, to obtain

371

unsaturated fatty acids to be inserted into the cellular membrane62.

60

. Different than

372

It is interesting to highlight the fact that knockdown of SCD1 using siRNA in

373

GMEC did not affect cell viability9. siRNA-mediated gene knockdown is a temporary

374

gene deficiency; thus, detrimental effect of decrease or lack of SCD1 might be not

375

visible due to the brief duration of the deficiency. In CRISPR/Cas9-mediated gene

376

knockout the gene deficiency is permanent and the detrimental effect has more likely

377

to be expressed phenotypically. This might explain the difference in cell viability

378

observed between likely biallelic deletion of SCD1 using CRISPR/Cas9 and SCD1

379

knockdown using siRNA in GMEC.

380

The efficiency of any CRISPR-mediated gene editing endeavor is limited by the

381

presence of unspecific mutagenesis, or off-target effects. Although complete

382

homology is required to achieve significant rates of mutation, the literature shows that

383

mutagenesis can be detected on sequences with up to five mismatches, particularly

384

when sequence homology is maintained 10-12 nucleotides upstream of the PAM63. In

385

our study, no off-targets were detected among the ten predicted by Cas-OFFinder.

386

Although the predicted off-target site 3 and 6 have cleavage bands, they are both the 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 47

387

natural mutations of one allele as also observed in the wild type cells, thus they could

388

be recognized by T7 endonuclease.

389

Our results indicated that we successfully got SCD1 monoallelic knockout in

390

GMEC by either causing indels via the NHEJ pathway or inserting the selection

391

markers to disrupt the coding frame by HDR pathway. The disruption of SCD1 in its

392

second exon caused a 50% decrease of SCD1 protein level; mRNA expression,

393

however, was affected more dramatically. The 80% decreased in SCD1 transcript

394

observed

395

expression/activation of the transcription factor SREBP1 known to bind to the SCD1

396

promoter in goats9.

may

be

caused

by

the

significant

or

numerical

decrease

in

397

The decrease of TAG, cholesterol and fatty acid desaturation index in knockout

398

cells provided further support for a successfully knockout of SCD1; however, the

399

decrease of TAG and desaturase indexes was only approx. 25% in knockout vs. wild

400

type cells, which was similar between the clone obtained by NHEJ and the one by

401

HDR. The decrease of TAG was less than proportional compared to the decrease in

402

SCD1 protein expression, indicating that the effect on TAG synthesis was mostly

403

retained and compensated by the wild-type allele. The decrease in cholesterol was

404

more pronounced (≥40%) and with larger decrease in the clone obtained using NHEJ

405

compared to HDR. The former had a larger decrease in expression and activity of

406

SREBP1 compared to the latter, indicating that the decrease in cholesterol was partly

407

due to a decrease expression/activity of SREBP1.

408

The reduction of TAG in SCD1 knockout cells appeared to be a consequence of a 20

ACS Paragon Plus Environment

Page 21 of 47

Journal of Agricultural and Food Chemistry

409

reduced de novo fatty acid synthesis and intracellular LCFA transport. This is

410

indicated by an overall decrease in expression of FASN, SREBF1, FABP3, and FABP4.

411

The acetyl-CoA carboxylase was also affected by SCD1 knockout. Its expression was

412

decreased, although significant only with NHEJ, but also its activity was likely

413

inhibited by an increased proportion of C16:064 with the consequent reduction of

414

malonyl-CoA. The latter is used by FASN to synthesize long chain fatty acid65, 66.

415

It is unclear how a deficiency of SCD1 would affect the expression of other

416

lipogenic genes in our study. However, SCD1 appears to regulate SREBP1. In Scd1

417

knockout mice, the expression of Srebf1 and Fasn were both downregulated67, similar

418

to our findings. A very similar results was obtained in mice with liver specific Scd1

419

knockout where a decreased expression of Srebf1c and reduced lipogenesis were

420

detected28. SREBP1 has a central role in the regulation of milk fat synthesis-related

421

genes in all species, including goat46. As for other species, FASN is among the genes

422

regulated by SREBP1 also in goat68; thus, the decreased expression of FASN in SCD1

423

knockout cells is likely consequence of the decreased expression/activity of SREBP1.

424

An important role of SCD1 in the activity of SREBP1 was also revealed in Scd1

425

knockout mice fed fructose where the up-regulation of Srebf1 was observed only

426

when supplemented with oleic acid, the main product of SCD1, but not with palmitate

427

or stearate26. No data in this regard are available for goats or other ruminants;

428

however, oleic acid does not seem to have any transcriptional effect on SCD1 or

429

SREBP1 down-stream genes in bovine69. Furthermore, activation of SREBP1 in

430

Drosophila was inhibited by palmitic acid, which is the substrate of SCD170. In cancer 21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

431

cells, high SCD1 activity promotes an increase in MUFA-to-SFA ratios in cell

432

membrane activating tyrosine kinase receptors and their membrane-interacting

433

signaling mediators such as PI3K, Akt and Ras. These signaling mediators affect the

434

activity of mTOR that is known to regulate SREBP158. All the above studies about the

435

effect of SCD1 on SREBP1 where performed in non-ruminant species. Despite

436

important differences with ruminants, for instance the absence or minor effect on

437

transcription of genes by oleate69, it appears likely that the same effect of SCD1 on

438

SREBP1 was present in our cells.

439

ELOVL5, ELOVL6 and SCD1 jointly participate in fatty acid de novo synthesis.

440

Previous studies in GMEC indicated both ELOVL5 and ELOVL6 do not influence

441

SCD1 expression72, 73. Knockout of SCD1 in NHEJ pathway decreased the content of

442

C16:1, the substrate of ELOVL5, and increased C16:0, the substrate of ELOVL6. The

443

change of gene expression may due to their substrates alteration. The different

444

expression of ELOVL5 and ELOVL6 in NHEJ and HDR pathways knockout GMEC

445

may be caused by the cell models obtained from different ways. The different DNA

446

repair pathways would have influence on some genes expression, but the mechanism

447

is still not clear.

448

In bovine mammary gland, FABP3 likely plays an essential role in providing

449

fatty acids to SCD1 while FABP4 takes the unsaturated fatty acid coming from SCD1

450

to transport it to the enzymes involved in TAG synthesis8. It is interesting in our study

451

that the transcription of both fatty acid transporters was down-regulated by knocking

452

out SCD1. It is unclear what the physical connections between the downregulation of 22

ACS Paragon Plus Environment

Page 22 of 47

Page 23 of 47

Journal of Agricultural and Food Chemistry

453

the fatty acid binding proteins and SCD1, however it makes biological sense

454

considering the concerted function of the two FABP with SCD1 for the synthesis of

455

TAG.

456

In conclusion, we successfully obtained SCD1 monoallelic knockout GMEC by

457

CRISPR/Cas9-mediated gene editing both through NHEJ and HDR pathways. Our

458

data indicated that full SCD1 knockout is likely deleterious for GMEC. The

459

monoallelic knockout of SCD1 decreased the amount of TAG, cholesterol, and

460

desaturase index and negatively affected the expression of several key genes related to

461

de novo fatty acid synthesis. Using a more effective knockout system compared to

462

RNAi, our data confirm SCD1 being an important enzyme in synthesis of TAG and

463

unsaturated fatty acids in goat milk. It remains to be determined the physical

464

relationship between SCD1 and several of the genes which expression was decreased

465

by deleting SCD1, chiefly SREBF1 being a key transcription factor in milk fat

466

synthesis.

467 468 469

ACKNOWLEDGEMENTS This research was supported by National Natural Science Foundation of China (Beijing, China; 31772575).

470

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

471

Page 24 of 47

REFERENCES

472

1.

473

changes in world ruminant production systems. Agr. Syst. 2005, 84, 121-153.

474

2.

475

Meza-Herrera, C., Dairy goat production systems: status quo, perspectives and

476

challenges. Trop Anim Health Prod 2013, 45, 17-34.

477

3.

478

155-163.

479

4.

480

Pyzel, B.; Horbańczuk, J. O., Chemical composition, physical traits and fatty acid

481

profile of goat milk as related to the stage of lactation. Anim Sci P. 2009, 27, 311-320.

482

5.

483

Lopez-Miranda, J., The influence of olive oil on human health: not a question of fat

484

alone. Mol Nutr Food Res 2007, 51, 1199-1208.

485

6.

486

1969, 2, 193-202.

487

7.

488

Genes regulating lipid and protein metabolism are highly expressed in mammary

489

gland of lactating dairy goats. Funct Integr Genomic. 2015, 15, 309-321.

490

8.

491

the lactation cycle. BMC Genomics 2008, 9, 366.

492

9.

Bouwman, A. F.; Van der Hoek, K. W.; Eickhout, B.; Soenario, I., Exploring

Escareno, L.; Salinas-Gonzalez, H.; Wurzinger, M.; Iniguez, L.; Solkner, J.;

Haenlein, G. F. W., Goat milk in human nutrition. Small Rumin Res. 2004, 51,

Strzałkowska, N.; Jóźwik, A.; Bagnicka, E.; Krzyżewski, J.; Horbańczuk, K.;

Pérez-Jiménez,

F.;

Ruano,

J.;

Perez-Martinez,

P.;

Lopez-Segura,

F.;

Bloch, K., Enzymic synthesis of monounsaturated fatty acids. Acc Chem Res.

Shi, H.; Zhu, J.; Luo, J.; Cao, W.; Shi, H.; Yao, D.; Li, J.; Sun, Y.; Xu, H.; Yu, K.,

Bionaz, M.; Loor, J. J., Gene networks driving bovine milk fat synthesis during

Yao, D.; Luo, J.; He, Q.; Shi, H.; Li, J.; Wang, H.; Xu, H.; Chen, Z.; Yi, Y.; Loor, 24

ACS Paragon Plus Environment

Page 25 of 47

Journal of Agricultural and Food Chemistry

493

J. J., SCD1 Alters Long-Chain Fatty Acid (LCFA) Composition and Its Expression Is

494

Directly Regulated by SREBP-1 and PPARgamma 1 in Dairy Goat Mammary Cells. J

495

Cell Physiol. 2017, 232, 635-649.

496

10. Taxman, D. J.; Livingstone, L. R.; Zhang, J.; Conti, B. J.; Iocca, H. A.; Williams,

497

K. L.; Lich, J. D.; Ting, J. P.; Reed, W., Criteria for effective design, construction, and

498

gene knockdown by shRNA vectors. BMC Biotechnol 2006, 6, 7.

499

11. Gaj, T.; Gersbach, C. A.; Barbas, C. F., 3rd, ZFN, TALEN, and

500

CRISPR/Cas-based methods for genome engineering. Trends Biotechnol 2013, 31,

501

397-405.

502

12. Barrangou, R.; Birmingham, A.; Wiemann, S.; Beijersbergen, R. L.; Hornung, V.;

503

Smith, A., Advances in CRISPR-Cas9 genome engineering: lessons learned from

504

RNA interference. Nucleic Acids Res. 2015, 43, 3407-19.

505

13. Dicarlo, J. E.; Norville, J. E.; Mali, P.; Rios, X.; Aach, J.; Church, G. M., Genome

506

engineering in Saccharomyces cerevisiae using CRISPR-Cas systems. Nucleic Acids

507

Res. 2013, 41, 4336-43.

508

14. Jiang, W.; Bikard, D.; Cox, D.; Zhang, F.; Marraffini, L. A., RNA-guided editing

509

of bacterial genomes using CRISPR-Cas systems. Nat Biotechnol 2013, 31, 233-9.

510

15. Gratz, S. J.; Cummings, A. M.; Nguyen, J. N.; Hamm, D. C.; Donohue, L. K.;

511

Harrison, M. M.; Wildonger, J.; O’Connorgiles, K. M., Genome Engineering of

512

Drosophila with the CRISPR RNA-Guided Cas9 Nuclease. Genetics 2013, 194,

513

1029-35.

514

16. Yao, X.; Wang, X.; Hu, X.; Liu, Z.; Liu, J.; Zhou, H.; Shen, X.; Wei, Y.; Huang, 25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

515

Z.; Ying, W.; Wang, Y.; Nie, Y. H.; Zhang, C. C.; Li, S.; Cheng, L.; Wang, Q.; Wu, Y.;

516

Huang, P.; Sun, Q.; Shi, L.; Yang, H., Homology-mediated end joining-based targeted

517

integration using CRISPR/Cas9. Cell Res 2017, 27, 801-814.

518

17. Long, C.; Amoasii, L.; Mireault, A. A.; Mcanally, J. R.; Li, H.; Sanchezortiz, E.;

519

Bhattacharyya, S.; Shelton, J. M.; Basselduby, R.; Olson, E. N., Postnatal genome

520

editing partially restores dystrophin expression in a mouse model of muscular

521

dystrophy. Science 2016, 351, 400.

522

18. Yin, H.; Xue, W.; Chen, S.; Bogorad, R. L.; Benedetti, E.; Grompe, M.;

523

Koteliansky, V.; Sharp, P. A.; Jacks, T.; Anderson, D. G., Genome editing with Cas9 in

524

adult mice corrects a disease mutation and phenotype. Nat Biotechnol. 2014, 32, 551.

525

19. Wang, Y.; Du, Y.; Shen, B.; Zhou, X.; Li, J.; Liu, Y.; Wang, J.; Zhou, J.; Hu, B.;

526

Kang, N., Efficient generation of gene-modified pigs via injection of zygote with

527

Cas9/sgRNA. Sci Rep 2015, 5, 8256.

528

20. Han, J.; Zhang, J.; Chen, L.; Shen, B.; Zhou, J.; Hu, B.; Du, Y.; Tate, P. H.; Huang,

529

X.; Zhang, W., Efficient in vivo deletion of a large imprinted lncRNA by

530

CRISPR/Cas9. RNA Biol 2014, 11, 829-35.

531

21. Strauss, B. S., DNA Repair and Mutagenesis. Science 1995, 270, 1511

532

22. Pfeiffer, P.; Thode, S.; Hancke, J.; Vielmetter, W., Mechanisms of overlap

533

formation in nonhomologous DNA end joining. Mol Cell Biol 1994, 14, 888-895.

534

23. Aparicio, T.; Baer, R.; Gautier, J., DNA double-strand break repair pathway

535

choice and cancer. DNA Repair 2014, 19, 169-175.

536

24. Genovese, P.; Schiroli, G.; Escobar, G.; Tomaso, T. D.; Firrito, C.; Calabria, A.; 26

ACS Paragon Plus Environment

Page 26 of 47

Page 27 of 47

Journal of Agricultural and Food Chemistry

537

Moi, D.; Mazzieri, R.; Bonini, C.; Holmes, M. C.; Gregory, P. D.; van der Burg, M.;

538

Gentner, B.; Montini, E.; Lombardo, A.; Naldini, L., Targeted genome editing in

539

human repopulating haematopoietic stem cells. Nature 2014, 510, 235-240.

540

25. Liang, F.; Han, M.; Romanienko, P. J.; Jasin, M., Homology-directed repair is a

541

major double-strand break repair pathway in mammalian cells. Proc Natl Acad Sci U

542

S A 1998, 95, 5172-7.

543

26. Miyazaki, M.; Dobrzyn, A.; Man, W. C.; Chu, K.; Sampath, H.; Kim, H. J.;

544

Ntambi, J. M., Stearoyl-CoA desaturase 1 gene expression is necessary for

545

fructose-mediated induction of lipogenic gene expression by sterol regulatory

546

element-binding protein-1c-dependent and -independent mechanisms. J Biol Chem

547

2004, 279, 25164-71.

548

27. Ntambi, J. M.; Miyazaki, M.; Stoehr, J. P.; Lan, H.; Kendziorski, C. M.; Yandell,

549

B. S.; Song, Y.; Cohen, P.; Friedman, J. M.; Attie, A. D., Loss of stearoyl-CoA

550

desaturase-1 function protects mice against adiposity. Proc Natl Acad Sci U S A 2002,

551

99, 11482-6.

552

28. Miyazaki, M.; Flowers, M. T.; Sampath, H.; Chu, K.; Otzelberger, C.; Liu, X.;

553

Ntambi, J. M., Hepatic stearoyl-CoA desaturase-1 deficiency protects mice from

554

carbohydrate-induced adiposity and hepatic steatosis. Cell Metab 2007, 6, 484-96.

555

29. Sampath, H.; Flowers, M. T.; Liu, X.; Paton, C. M.; Sullivan, R.; Chu, K.; Zhao,

556

M.; Ntambi, J. M., Skin-specific deletion of stearoyl-CoA desaturase-1 alters skin

557

lipid composition and protects mice from high fat diet-induced obesity. J Biol Chem

558

2009, 284, 19961-19973. 27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

559

30. Hyun, C. K.; Kim, E. D.; Flowers, M. T.; Liu, X.; Kim, E.; Strable, M.; Ntambi, J.

560

M., Adipose-specific deletion of stearoyl-CoA desaturase 1 up-regulates the glucose

561

transporter GLUT1 in adipose tissue. Biochem Biophys Res Commun 2010, 399,

562

480-6.

563

31. Chilliard, Y.; Glasser, F.; Ferlay, A.; Bernard, L.; Rouel, J.; Doreau, M., Diet,

564

rumen biohydrogenation and nutritional quality of cow and goat milk fat. Eur J Lipid

565

Sci Tech 2010, 109, 828-855.

566

32. Zhu, J.; Sun, Y.; Luo, J.; Wu, M.; Li, J.; Cao, Y., Specificity protein 1 regulates

567

gene expression related to fatty acid metabolism in goat mammary epithelial cells. Int

568

J Mol Sci 2015, 16, 1806-20.

569

33. Wang, Z., Characterization and culture of isolated primary dairy goat mammary

570

gland epithelial cells. Chin J Biotechnol 2010, 26, 1123.

571

34. Montague, T. G.; Cruz, J. M.; Gagnon, J. A.; Church, G. M.; Valen, E.,

572

CHOPCHOP: a CRISPR/Cas9 and TALEN web tool for genome editing. Nucleic

573

Acids Res 2014, 42, W401-7.

574

35. Ran, F. A.; Hsu, P. D.; Wright, J.; Agarwala, V.; Scott, D. A.; Zhang, F., Genome

575

engineering using the CRISPR-Cas9 system. Nat Protoc 2013, 8, 2281-2308.

576

36. Cong, L.; Ran, F. A.; Cox, D.; Lin, S.; Barretto, R.; Habib, N.; Hsu, P. D.; Wu, X.;

577

Jiang, W.; Marraffini, L. A.; Zhang, F., Multiplex genome engineering using

578

CRISPR/Cas systems. Science 2013, 339, 819-23.

579

37. Kim, J. H.; Lee, S. R.; Li, L. H.; Park, H. J.; Park, J. H.; Lee, K. Y.; Kim, M. K.;

580

Shin, B. A.; Choi, S. Y., High cleavage efficiency of a 2A peptide derived from 28

ACS Paragon Plus Environment

Page 28 of 47

Page 29 of 47

Journal of Agricultural and Food Chemistry

581

porcine teschovirus-1 in human cell lines, zebrafish and mice. PLoS One 2011, 6,

582

e18556.

583

38. Yu, Y.; Tong, Q.; Li, Z.; Tian, J.; Wang, Y.; Su, F.; Wang, Y.; Liu, J.; Zhang, Y.,

584

Improved site-specific recombinase-based method to produce selectable marker- and

585

vector-backbone-free transgenic cells. Sci Rep 2014, 4, 4240.

586

39. Vouillot, L.; Thelie, A.; Pollet, N., Comparison of T7E1 and surveyor mismatch

587

cleavage assays to detect mutations triggered by engineered nucleases. G3 (Bethesda)

588

2015, 5, 407-15.

589

40. Bae, S.; Park, J.; Kim, J. S., Cas-OFFinder: a fast and versatile algorithm that

590

searches for potential off-target sites of Cas9 RNA-guided endonucleases.

591

Bioinformatics 2014, 30, 1473.

592

41. Kim, Y.; Cheong, S. A.; Lee, J. G.; Lee, S. W.; Lee, M. S.; Baek, I. J.; Sung, Y. H.,

593

Generation of knockout mice by Cpf1-mediated gene targeting. Nat Biotechnol 2016,

594

34, 808-10.

595

42. Hsu, P. D.; Scott, D. A.; Weinstein, J. A.; Ran, F. A.; Konermann, S.; Agarwala, V.;

596

Li, Y.; Fine, E. J.; Wu, X.; Shalem, O., DNA targeting specificity of RNA-guided

597

Cas9 nucleases. Nat Biotechnol 2013, 31, 827-832.

598

43. Ramakers, C.; Ruijter, J. M.; Deprez, R. H. L.; Moorman, A. F. M.,

599

Assumption-free analysis of quantitative real-time polymerase chain reaction (PCR)

600

data. Neurosci Lett 2003, 339, 62-66.

601

44. Vandesompele, J.; De Preter, K.; Pattyn, F.; Poppe, B.; Van Roy, N.; De Paepe, A.;

602

Speleman, F., Accurate normalization of real-time quantitative RT-PCR data by 29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

603

geometric averaging of multiple internal control genes. Genome Biol 2002, 3,

604

research0034. 1.

605

45. Bionaz, M.; Loor, J. J., Identification of reference genes for quantitative real-time

606

PCR in the bovine mammary gland during the lactation cycle. Physiol Genomics 2007,

607

29, 312-9.

608

46. Xu, H. F.; Luo, J.; Zhao, W. S.; Yang, Y. C.; Tian, H. B.; Shi, H. B.; Bionaz, M.,

609

Overexpression of SREBP1 (sterol regulatory element binding protein 1) promotes de

610

novo fatty acid synthesis and triacylglycerol accumulation in goat mammary epithelial

611

cells. J Dairy Sci 2016, 99, 783-795.

612

47. Luo, J.; Zhu, J. J.; Sun, Y. T.; Shi, H. B.; Li, J., Inhibitions of FASN suppress

613

triglyceride synthesis via the control of malonyl-CoA in goat mammary epithelial

614

cells. Anim Reprod Sci 2016, 57, 1624-1630.

615

48. Wang, W.; Luo, J.; Zhong, Y.; Lin, X. Z.; Shi, H. B.; Zhu, J. J.; Li, J.; Sun, Y. T.;

616

Zhao, W. S., Goat liver X receptor alpha, molecular cloning, functional

617

characterization and regulating fatty acid synthesis in epithelial cells of goat

618

mammary glands. Gene 2012, 505, 114-20.

619

49. Morales, M. S.; Palmquist, D. L.; Weiss, W. P., Effects of Fat Source and Copper

620

on Unsaturation of Blood and Milk Triacylglycerol Fatty Acids in Holstein and Jersey

621

Cows 1. J Dairy Sci 2000, 83, 2105-2111.

622

50. Wei, C.; Liu, J.; Yu, Z.; Zhang, B.; Gao, G.; Jiao, R., TALEN or Cas9 - rapid,

623

efficient and specific choices for genome modifications. J Genet Genomics 2013, 40,

624

281-9. 30

ACS Paragon Plus Environment

Page 30 of 47

Page 31 of 47

Journal of Agricultural and Food Chemistry

625

51. Maeder, M. L.; Thibodeau-Beganny, S.; Osiak, A.; Wright, D. A.; Anthony, R. M.;

626

Eichtinger, M.; Jiang, T.; Foley, J. E.; Winfrey, R. J.; Townsend, J. A.; Unger-Wallace,

627

E.; Sander, J. D.; Muller-Lerch, F.; Fu, F.; Pearlberg, J.; Gobel, C.; Dassie, J. P.;

628

Pruett-Miller, S. M.; Porteus, M. H.; Sgroi, D. C.; Iafrate, A. J.; Dobbs, D.; McCray, P.

629

B., Jr.; Cathomen, T.; Voytas, D. F.; Joung, J. K., Rapid "open-source" engineering of

630

customized zinc-finger nucleases for highly efficient gene modification. Mol Cell

631

2008, 31, 294-301.

632

52. Bogdanove, A. J.; Voytas, D. F., TAL effectors: customizable proteins for DNA

633

targeting. Science 2011, 333, 1843-6.

634

53. Boettcher, M.; McManus, M. T., Choosing the Right Tool for the Job: RNAi,

635

TALEN, or CRISPR. Mol Cell 2015, 58, 575-85.

636

54. Hannon, G. J., RNA interference. Nature 2002, 418, 244.

637

55. Ali, N.; Manoharan, V. N., RNA folding and hydrolysis terms explain ATP

638

independence of RNA interference in human systems. Oligonucleotides 2009, 19,

639

341-6.

640

56. Evers, B.; Jastrzebski, K.; Heijmans, J. P.; Grernrum, W.; Beijersbergen, R. L.;

641

Bernards, R., CRISPR knockout screening outperforms shRNA and CRISPRi in

642

identifying essential genes. Nat Biotechnol 2016, 34, 631-3.

643

57. Zelles, L., Phospholipid fatty acid profiles in selected members of soil microbial

644

communities. Chemosphere 1997, 35, 275-294.

645

58. Igal, R. A., Stearoyl CoA desaturase-1: New insights into a central regulator of

646

cancer metabolism. Biochim Biophys Acta 2016, 1861, 1865-1880. 31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

647

59. Dobrzyn, P.; Dobrzyn, A.; Miyazaki, M.; Cohen, P.; Asilmaz, E.; Hardie, D. G.;

648

Friedman, J. M.; Ntambi, J. M., Stearoyl-CoA desaturase 1 deficiency increases fatty

649

acid oxidation by activating AMP-activated protein kinase in liver. Proc Natl Acad Sci

650

U S A 2004, 101, 6409-14.

651

60. Miyazaki, M.; Man, W. C.; Ntambi, J. M., Targeted disruption of stearoyl-CoA

652

desaturase1 gene in mice causes atrophy of sebaceous and meibomian glands and

653

depletion of wax esters in the eyelid. J Nutr 2001, 131, 2260-2268.

654

61. Jenkins, T. C.; Wallace, R. J.; Moate, P. J.; Mosley, E. E., Board-invited review:

655

Recent advances in biohydrogenation of unsaturated fatty acids within the rumen

656

microbial ecosystem. J Anim Sci 2008, 86, 397-412.

657

62. Ntambi, J. M., Stearoyl-CoA Desaturase Genes in Lipid Metabolism. 2013.

658

63. Kuscu, C.; Arslan, S.; Singh, R.; Thorpe, J.; Adli, M., Genome-wide analysis

659

reveals characteristics of off-target sites bound by the Cas9 endonuclease. Nat

660

Biotechnol 2014, 32, 677-83.

661

64. Paton, C. M.; Ntambi, J. M., Biochemical and physiological function of

662

stearoyl-CoA desaturase. Am J Physiol Endocrinol Metab 2009, 297, 28-37.

663

65. Jenni, S.; Leibundgut, M.; Boehringer, D.; Frick, C.; Mikolásek, B.; Ban, N.,

664

Structure of Fungal Fatty Acid Synthase and Implications for Iterative Substrate

665

Shuttling. Science 2007, 316, 254.

666

66. Livore, V. I.; Tripodi, K. E.; Uttaro, A. D., Elongation of polyunsaturated fatty

667

acids in trypanosomatids. FEBS J 2010, 274, 264-274.

668

67. Ntambi, J. M.; Miyazaki, M.; Stoehr, J. P.; Lan, H.; Kendziorski, C. M.; Yandell, 32

ACS Paragon Plus Environment

Page 32 of 47

Page 33 of 47

Journal of Agricultural and Food Chemistry

669

B. S.; Song, Y.; Cohen, P.; Friedman, J. M.; Attie, A. D., Loss of Stearoyl-CoA

670

Desaturase-1 Function Protects Mice against Adiposity. Proc Natl Acad Sci U S A

671

2002, 99, 11482-6.

672

68. Li, J.; Luo, J.; Xu, H.; Wang, M.; Zhu, J.; Shi, H.; Haile, A. B.; Wang, H.; Sun, Y.,

673

Fatty acid synthase promoter: characterization, and transcriptional regulation by sterol

674

regulatory element binding protein-1 in goat mammary epithelial cells. Gene 2015,

675

561, 157-64.

676

69. Bionaz, M.; Osorio, J.; Loor, J. J., TRIENNIAL LACTATION SYMPOSIUM:

677

Nutrigenomics in dairy cows: Nutrients, transcription factors, and techniques. J Anim

678

Sci 2015, 93, 5531-5553.

679

70. Seegmiller, A. C.; Dobrosotskaya, I.; Goldstein, J. L.; Ho, Y. K.; Brown, M. S.;

680

Rawson, R. B., The SREBP Pathway in : Regulation by Palmitate, Not Sterols. Dev

681

Cell 2002, 2, 229-238.

682

71. Shi, H. B.; Du, Y.; Zhang, C. H.; Sun, C.; He, Y. L.; Wu, Y. H.; Liu, J. X.; Luo, J.;

683

Loor, J. J., Fatty acid elongase 5 (ELOVL5) alters the synthesis of long-chain

684

unsaturated fatty acids in goat mammary epithelial cells. J Dairy Sci 2018, 101,

685

4586-4594.

686

72. Shi, H. B.; Wu, M.; Zhu, J. J.; Zhang, C. H.; Yao, D. W.; Luo, J.; Loor, J. J., Fatty

687

acid elongase 6 plays a role in the synthesis of long-chain fatty acids in goat

688

mammary epithelial cells. J Dairy Sci 2017, 100, 4987-4995.

689 690 33

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 34 of 47

691

Table 1. Test primers for NHEJ and 5’ and 3’ homology arms for HDR pathway

692

induced SCD1 knockout. Primer Test primer for NHEJ

Sequence (5’ to 3’) F: GGGAGTGAAGTGGTCCCTAC

Size (bp) 583

R: CAGCCCCAACACCGAAATTA 5’ homology arm clone primer

5-F: AGCACCCCATACCCAAGACT

1193

5-R: GGATAAGGAGGGCCCAAAGC 3’ homology arm clone primer

3-F: TCATCTCTCATTTCAGGGCG

1063

3-R: TGTGTCTGCAGCATCCAGTT 5’ junction primer

5j F: AATAAGAGCCCTTCCTGGTTT

1697

5j R: GGTTCTACGTTAGTGGGAGTTT 3’ junction primer

3j F: TCATAATCAGCCATACCACA 3j R: CCTAGTGCCCATCCATTT

693 694

34

ACS Paragon Plus Environment

1698

Page 35 of 47

Journal of Agricultural and Food Chemistry

695

Table 2. Effects of SCD1 monoallelic knockout on fatty acid composition in GMEC. Fatty acid

HDR-CTR

HDR-KO

Fatty acid

NHEJ-CTR

NHEJ-KO

C16:0(%)

23.19a±0.15

26.73b±0.70

C16:0(%)

23.41A±0.49

25.38B±0.23

C16:1(%)

2.15a±0.03

1.90b±0.00

C16:1(%)

1.97a±0.09

1.69b±0.00

C18:0(%)

17.34±0.15

16.80±0.20

C18:0(%)

17.34±0.15

17.31±0.33

C18:1 cis9(%)

20.56a±0.62

18.66b±0.15

C18:1 cis9(%)

20.10a±1.27

16.98b±0.14

696

Fatty acid data are reported as proportion of the total fatty acids. Statistical

697

significance between knockout (KO) and wild type GMEC (CTR) was as follow:

698

lower case letters, P < 0.05; upper case letters, P < 0.01. Data are presented as means

699

± SEM.

700

35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

701

Figure legends

702

Figure 1. sgRNAs target sites selection of SCD1 gene in Capra hircus genome and

703

NHEJ-mediated gene modification. (A) SCD1 gene locus and sgRNAs design. PAM

704

motifs are bold and underlined. (B) Cleavage efficiency of Cas9/sgRNA at three target

705

sites. CTR denotes the DNA isolated from GMEC transfected with pSpCas9

706

(BB)-2A-Puro without sgRNA sequence. The cleavage efficiency was quantified with

707

T7EN1 cleavage assay and analyzed by ImageJ. (C) Cleavage efficiency of the

708

GMEC single clone through NHEJ-mediated SCD1 gene modification by T7EN1

709

assay. (D) The sequences of modified SCD1 alleles. sgRNA target sequences are in

710

red text; PAM motifs are in green text bold and underlined; mutations are red and

711

lowercase; deletions (-), mutations (m) or wild type (WT) are shown to the right of

712

each sequence. (E) The sequence of the single clone detected in picture (C).

713

Figure 2. Insertion and selection of the SCD1 transgenic single clone by

714

HDR-mediated genome editing. (A) Schematic representation of the HDR donor

715

vector and 5’ or 3’ junction PCR primers position. EGFP: enhanced green

716

fluorescence protein; P2A: porcine teschovirus-1 2A; PuroR: puromycin resistance

717

selection marker; 5j F: 5’ junction PCR forward primer; 5j R: 5’ junction PCR reverse

718

primer; 3j F: 3’ junction PCR forward primer; 3j R: 3’ junction PCR reverse primer. 5j

719

F and 3j R were the primers for the regions outsides the homologous arms in the

720

genome, and 5j R and 3j F were primers for the donor vector region. (B) 5’ and 3’

721

junction PCR for the GMEC single clones detection which were transfected with both

722

sgRNA/Cas9 vector and donor vector. (C) Photograph of the single clone 10 GMEC

723

(KO) and a wild type GMEC counterpart (CTR). Scale bar is 100 µm.

724

Figure 3. Off-target sites of sgRNA3 in NHEJ- and HDR-mediated SCD1 knockout

725

GMEC. The sgRNA sequences end with PAM motif. (A) 10 off-target sites predicted

726

by online website were selected. The off-target nucleotide bases matching with the

727

sgRNA and the PAM sequence are highlighted in orange and red respectively. (B)

728

Off-target sites detection in NHEJ-mediated knockout single clone and (C)

729

HDR-mediated knockout single clone. Top pictures in B and C are PCR products and 36

ACS Paragon Plus Environment

Page 36 of 47

Page 37 of 47

Journal of Agricultural and Food Chemistry

730

bottom pictures are the results of T7EN1 cleavage assay. OT: off-target, WT: wild

731

type for each off-target site in control group.

732

Figure 4. SCD1 knockout in GMEC both through HDR and NHEJ pathway.

733

Real-time quantitative PCR analysis and protein level of SCD1 in NHEJ (A) and

734

HDR (B) knockout GMEC. Values are mean ± SEM for three independent

735

experiments. *P < 0.05 v. control.

736

Figure 5. Monoallelic knockout of SCD1 decreased the content of cellular

737

triacylglycerol (TAG) and cholesterol. Quadrants A and C report the amount of

738

triacylglycerol (TAG) and quadrants B and D the amount of cholesterol in NHEJ and

739

HDR groups. Values are presented as mean ± SEM for three independent experiments.

740

*P < 0.05 v. control.

741

Figure 6. Effects of SCD1 knockout on fatty acid desaturation index. Desaturation

742

index of C16:1 (A and C) and C18:1 (B and D) in NHEJ and HDR pathway SCD1

743

monoallelic knockout compared with wild type GMEC. Values are mean ± SEM for

744

three independent experiments. *P < 0.05 v. control.

745

Figure 7. Monoallelic knockout of SCD1 through NHEJ pathway affects mRNA

746

expression level of genes related to milk fat synthesis. Values are mean ± SEM for

747

three independent experiments. *P < 0.05 v. control.

748

Figure 8. Monoallelic knockout of SCD1 through HDR pathway affected mRNA

749

expression level of genes related to milk fat synthesis. Values are mean ± SEM for

750

three independent experiments. *P < 0.05 v. control.

751

Figure 9. Relative mRNA expression and protein level of precursor SREBP1 and

752

nuclear SREBP1 (mature SREBP1) in NHEJ-KO and HDR-KO groups. Values are

753

mean ± SEM. *P < 0.05 v. control.

754

37

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

755 756

Figures Figure 1.

757 758

38

ACS Paragon Plus Environment

Page 38 of 47

Page 39 of 47

Journal of Agricultural and Food Chemistry

759

Figure 2.

760 761

39

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

762

Figure 3.

763 764

40

ACS Paragon Plus Environment

Page 40 of 47

Page 41 of 47

Journal of Agricultural and Food Chemistry

765

Figure 4.

766 767

41

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

768

Figure 5.

769 770

42

ACS Paragon Plus Environment

Page 42 of 47

Page 43 of 47

Journal of Agricultural and Food Chemistry

771

Figure 6.

772 773

43

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

774

Figure 7.

775 776

44

ACS Paragon Plus Environment

Page 44 of 47

Page 45 of 47

Journal of Agricultural and Food Chemistry

777

Figure 8.

778 779

45

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

780

Figure 9.

781 782

46

ACS Paragon Plus Environment

Page 46 of 47

Page 47 of 47

Journal of Agricultural and Food Chemistry

783

TOC graphic

784 785

Legends: This study aims at investigating the function of SCD1 gene using

786

CRISPR/Cas9 technology in goat mammary epithelial cells. Firstly, non-homology

787

end joining (NHEJ) and homology directed repair (HDR) pathway were both used to

788

get the SCD1 monoallelic knockout cells. Then the content of cellular triacylglycerol,

789

cholesterol, fatty acid and fatty acid metabolism gene expression were measured. Our

790

results confirmed the important role of SCD1 on milk fat metabolism in goat

791

mammary cells.

792

47

ACS Paragon Plus Environment