Preparation of quercetin esters and their antioxidant activity

and Q-3-O-monoester,) and four diesters (Q-7,3'-O-diester, Q-3',4'-O-diester, Q-3,3'-O-. 26 diester and Q-3,4'-O-diester) were the major products as w...
2 downloads 0 Views 443KB Size
Subscriber access provided by Macquarie University

Bioactive Constituents, Metabolites, and Functions

Preparation of quercetin esters and their antioxidant activity Won Young Oh, Priyatharini Ambigaipalan, and Fereidoon Shahidi J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.9b04154 • Publication Date (Web): 29 Aug 2019 Downloaded from pubs.acs.org on August 30, 2019

Just Accepted

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.

Subscriber access provided by Macquarie University

“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

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.

Subscriber access provided by Macquarie University

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 29

Journal of Agricultural and Food Chemistry

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

Preparation of quercetin esters and their antioxidant activity

2 3 4 5

Won Young Oh, Priyatharini Ambigaipalan, and Fereidoon Shahidi*

6

Department of Biochemistry, Memorial University of Newfoundland, St. John’s, NL,

7

Canada A1B 3X9

8 9 10



Corresponding Author (*)

11



Tel.: 709 864 8552*

12



E-mail addresses: [email protected]; [email protected]; [email protected]*

13 14 15 16 17 18 19

1

ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

Journal of Agricultural and Food Chemistry

20

Abstract

21

Quercetin, a polyphenolic compound, is widely distributed in plants and has numerous

22

health benefits. However, its hydrophilicity can compromise its use in lipophilic systems.

23

For this reason, quercetin was esterified with 12 different fatty acids, as their acyl chlorides,

24

with varying chain length and degree of unsaturation. Two monoesters (Q-3'-O-monoester

25

and Q-3-O-monoester,) and four diesters (Q-7,3'-O-diester, Q-3',4'-O-diester, Q-3,3'-O-

26

diester and Q-3,4'-O-diester) were the major products as was shown by HPLC-MS and 1H-

27

NMR data. The lipophilicity of quercetin derivatives was calculated; this was found to

28

increase with fatty acid chain length. The antioxidant potential of quercetin and its

29

derivatives was evaluated by using ABTS radical cation scavenging activity; quercetin

30

showed the highest radical scavenging activity among all tested samples. Despite the

31

decrease of antioxidant activity in this study, the derivatives may show better antioxidant

32

activity in lipophilic media and display improved absorption and bioavailability in the body

33

once consumed.

34 35

Keywords: Quercetin, lipophilicity, HPLC-MS, NMR, antioxidant activity

36 37 38 39 40

2

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

41

Introduction

42

Quercetin, a flavonoid that is widely distributed in the plant kingdom, serves as a robust

43

antioxidant with numerous reported health benefits such as cardioprotection, anti-

44

inflammatory, and anticancer activities.1-3 Despite its potential beneficial effects, its low

45

absorption does not allow the body to take full advantage of it. Hence, many attempts have

46

been made to maximize its efficiency by different means, including structure modification.

47

Glycoside form and esterified phenolic compounds have been extensively studied. Graefe

48

et al.4 reported that quercetin glycosides (quercetin-3-O-rutinoside and quercetin-4'-O-

49

glucoside) have better bioavailability in the human body. However, quercetin itself was

50

not detected in human plasma. Zhong and Shahidi5,6 studied the antioxidant activity of

51

esterified epigallocatechin gallate (EGCG); EGCG derivatives showed better antioxidant

52

activity than EGCG itself. Moreover, Zhong et al.7 reported that EGCG derivatives

53

exhibited anticarcinogenic activity against colon cancer. Zhou et al.8 studied the

54

antioxidant activity of tyrosol derivatives and observed an unsaturation-dependent

55

antioxidant trend in a DNA scission assay. Pokorski et al.9 found that the commercially

56

available ascorbyl palmitate might serve as an ascorbate carrier to brain tissues.

57

Esterification with fatty acids would increase the lipophilicity of phenolic compounds.

58

Hence, quercetin derivatives may also show an improved application in food, cosmetics,

59

and possibly in biological environments. Besides the potential health benefit of quercetin,

60

fatty acids have their own advantages. For example, short-chain fatty acids such as

61

propionic acid and butyric acid are generally produced via fermentation by gut microbes.10

62

Propionic acid, which can be used for hepatic gluconeogenesis, has been shown to inhibit

63

cholesterol synthesis.10,11 On the other hand, butyric acid, an energy source for colonic

3

ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

Journal of Agricultural and Food Chemistry

64

mucosa, can prevent colon cancer.10,11 In addition, the beneficial health effects of

65

polyunsaturated fatty acids (PUFAs) have been well documented. PUFAs such as

66

eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA) have been known to

67

possess anti-inflammatory, cardioprotective, and anticancer potential.12 Moreover, one

68

study13 found that quercetin esters (pentaacetylquercetin, 3′-t-butyl-oxycarbonyl-D-alanine

69

tetraacetylquercetin

70

tetraacetylquercetin

71

tetraacetylquercetin ester) underwent partial deacylation when they passed through three

72

different epithelial cell monolayers, namely MDCK-1, MDCK-2, and Caco-2 cells. They

73

suggested that quercetin ester derivatives might provide a useful means to increase the level

74

of systemic aglycone. In addition, this may also increase the bioavailability of the

75

deacylated acid moiety. Hence, it is of interest to examine the effect of joining two

76

biologically active components such as EPA and DHA with quercetin. For this reason,

77

quercetin was esterified with 12 different acyl chlorides varying in chain length and degree

78

of unsaturation. To the best of our knowledge, this is the first attempt to prepare quercetin

79

with a variety of fatty acids including health beneficial PUFAs as well as reporting their

80

antioxidant potential.

ester, ester,

3′-[4-(t-butyl-oxycarbonyl-amino)butanoic and

3′-[6-(t-butyl-oxycarbonyl-amino)hexanoic

acid] acid]

81 82

Materials and methods

83

Materials

84

Quercetin from Xian Yuensun Biological Technology Company Limited (Shaanxi, China)

85

and DHA single cell oil (DHASCO) from DSM (Columbia, MD, USA) were procured.

86

Professor Kazuo Miyashita kindly provided EPA ethyl ester from Mochida Pharmaceutical 4

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

87

CO., LTD. (Tokyo, Japan). Flexible thin layer chromatography (TLC) plates with silica gel

88

60A (2.5 × 7.5 cm and 20 × 20 cm) were purchased from Selecto Scientific (Suwanee, GA,

89

USA) and Sigma-Aldrich (Oakville, ON, Canada). Potassium hydroxide and perdeuterated

90

dimethyl sulphoxide (DMSO-d6) containing tetramethylsilane (TMS) were bought from

91

BDH Inc. (Toronto, ON, Canada) and Cambridge Isotope Laboratories, Inc. (Andover, MA,

92

USA), respectively. Hydrogen chloride, sodium phosphate monobasic, sodium phosphate

93

dibasic, sodium chloride, and other solvents were purchased from Fisher Scientific Ltd

94

(Ottawa, ON, Canada). All other chemicals including acyl chlorides (propionyl chloride,

95

butyryl chloride, caproyl chloride, capryloyl chloride, capryl chloride, lauroyl chloride,

96

myristoyl chloride, palmitoyl chloride, stearoyl chloride, and oleoyl chloride) were bought

97

from Sigma-Aldrich (Oakville, ON, Canada).

98

Preparation of EPA and DHA

99

Saponification of EPA ethyl ester was carried out to prepare EPA. DHA was prepared from

100

DHASCO via urea complexation as described by Wanasundara and Shahidi.14 For

101

saponification, DHASCO was stirred with potassium hydroxide, water, and 95% aqueous

102

ethanol at 62 ± 2 °C under reflux condition and a nitrogen blanket. In order to remove

103

unsaponified material, the mixture was stirred with water and hexane and the aqueous

104

portion was then collected upon full separation. Hydrochloric acid (HCl, 3 M) was

105

subsequently added to the aqueous layer to reach a pH of 1 monitored using a pH meter

106

(Mettler-Toledo; Zürich, Switzerland) and then hexane was added to extract the liberated

107

free fatty acid. The final solution was passed through a layer of anhydrous sodium sulfate

108

and the solvent was then removed by using a rotary evaporator.

5

ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29

Journal of Agricultural and Food Chemistry

109

The free fatty acids were mixed with 20% urea in 95% aqueous ethanol (w/v) and this was

110

heated at 60 °C until a clear solution was obtained. The clear solution was then in kept in

111

a cold room (4 °C) for 24 h to form urea complexed crystals which were subsequently

112

suction filtered to remove them. To the resultant solution was added an equal volume of

113

water and the pH was adjusted to 4-5 with 6 M HCl. An equal volume of hexane was added

114

and stirred. The solvent was then passed through anhydrous sodium sulphate and

115

subsequently removed by rotary evaporation.

116

Preparation of quercetin derivatives

117

Preparation of quercetin derivatives was conducted according to Zhong and Shahidi.5

118

Eicosapentaenoyl chloride and docosahexaenoyl chloride were prepared with thionyl

119

chloride and the fatty acid under investigation. Quercetin (9.06 g in ethyl acetate) was

120

stirred with pyridine at a molar ratio of 1:1. Then acyl chlorides (propionyl chloride,

121

butyryl chloride, caproyl chloride, capryloyl chloride, capryl chloride, lauroyl chloride,

122

myristoyl chloride, palmitoyl chloride, stearoyl chloride, oleoyl chloride, eicosapentaenoyl

123

chloride, and docosahexaenoyl chloride) were added dropwise under a blanket of nitrogen

124

at 50 °C while stirring continuously. Three hours later, the mixture was cooled down to

125

room temperature and washed 3 times with distilled water (60 °C). The ethyl acetate layer

126

was subsequently passed through a layer of anhydrous sodium sulphate. The product

127

mixture was obtained following the removal of the solvent.

128

Identification of quercetin derivatives

129

The chemical structures of quercetin derivatives were determined by high-performance

130

liquid chromatography-electrospray ionization-time of flight-mass spectrometry (HPLC-

6

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

131

ESI-TOF-MS) using an Agilent 1260 HPLC unit (Agilent Technologies, Palo Alto, CA,

132

USA). A C18 column (4.6 mm × 250 mm × 5 μm; Sigma-Aldrich, Oakville, ON, Canada)

133

was used. The mobile phase was methanol/ 5% acetonitrile in water at 80:20-90:10 (v/v)

134

and the runtime varied from 0 to 100 min at 1 mL/min. The compounds were detected at

135

370 nm. After HPLC-MS analysis, the compounds were separated using TLC (hexane /

136

ethyl acetate / formic acid, 3:3:0.12, v/v/v). The compounds so obtained were then

137

dissolved in perdeuterated dimethyl sulfoxide (DMSO-d6) containing tetramethylsilane

138

(TMS) in order to record their proton nuclear magnetic resonance (1H-NMR, Bruker

139

Avance 300 MHz and 500 MHz; Bruker Biospin Co., Billerica, MA, USA) for structure

140

identification. The results were compared with the chemical shifts of quercetin and

141

literature values.

142

Separation of quercetin and its derivatives

143

Intact quercetin was removed from different mixtures of quercetin and derivatives using

144

column chromatography on silica gel. The gradients of hexane / ethyl acetate / formic acid

145

(90: 10: 2, 80: 20: 2, 70: 30: 2, 60: 40: 2, v/v/v) were employed and each collected fraction

146

was monitored by TLC (hexane / ethyl acetate / formic acid, 3: 3: 0.12, v/v/v) using UV

147

(254 nm).

148

Determination of lipophilicity

149

The lipophilicity of quercetin derivatives was theoretically calculated by using ALOGPS

150

2.1 established by Tetko and Bruneau.15 The simplified molecular input line entry (SMILE)

151

system of quercetin derivatives was obtained from ChemDraw Standard 16.0.

152

ABTS radical cation scavenging assay

7

ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29

Journal of Agricultural and Food Chemistry

153

The ABTS radical cation scavenging assay was developed by Miller et al.16 as described

154

by Ambigaipalan et al.17 with slight modifications. Solutions of ABTS (2.5 mM) and

155

AAPH (2.5 mM) in a 0.1 M phosphate buffer containing 0.15 M sodium chloride (pH 7.4)

156

were mixed at a ratio of 1:1 (v/v). The mixture (ABTS radical cation) was then heated for

157

20 min at 60 °C and subsequently stored in the dark at room temperature. Quercetin and

158

its derivatives (0.5 mg/mL) were dissolved in ethanol. The compounds and a blank (40 μL)

159

were vortexed with the ABTS radical cation solution (1.96 mL). Six minutes later, the

160

samples were measured at 734 nm. The ABTS radical cation scavenging ability of

161

quercetin and its derivatives was expressed as μmoles of trolox equivalents per mg of

162

sample.

163

DPPH radical scavenging assay

164

The DPPH scavenging capacity of quercetin and its derivatives was determined according

165

to Ambigaipalan et al.17 with slight modifications. The samples (40 μL, 0.5 mg/mL, in

166

ethanol) were mixed with 60 μL of ethanol and 0.9 mL of DPPH (0.2 mM, in ethanol), kept

167

in the dark for 30 min and monitored using electron paramagnetic resonance (EPR). The

168

DPPH scavenging capacities of quercetin and its derivatives were expressed as μmoles of

169

trolox equivalents (TE) per mg of sample.

170

Statistical analysis

171

One-way ANOVA with Tukey’s HSD was carried out at p < 0.05 level using IBM statistics

172

version 21 (IBM Corp., Armonk, NY, USA).

173

Results and discussion

174

Preparation of quercetin derivatives 8

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

175

The derivatives of quercetin were prepared using quercetin and fatty acyl chlorides. The

176

esterification reaction was monitored by thin-layer chromatography (TLC). After

177

completion of the reaction, selected crude products were subjected to HPLC-MS with

178

methanol/5% acetonitrile in water (90:10, v/v) as the mobile phase. The HPLC-MS

179

analysis of quercetin with caproic acid (Q-C6:0) showed that there were intact quercetin

180

(9.4%), monoesters (35.7%), diesters (38.7%), triesters (14.2%), and tetraester (2%). The

181

yield of crude product of Q-C6:0 was 90.6%. Oh and Shahidi18 reported that the yields of

182

resveratryl propionate and resveratryl docosahexaenoate were 74 and 37.7%, respectively.

183

Zhong and Shahidi5 prepared EGCG derivatives using stearic acid (C18:0), EPA, and DHA

184

and reported that the yields of EGCG-C18:0, EGCG-EPA, and EGCG-DHA were 56.9,

185

42.7, 30.7%, respectively. They also reported that esterification by chemical method had

186

less selectivity and specificity. Both studies suggested that structural properties might

187

affect the esterification reaction. Indeed, the resultant crude product of quercetin with

188

propionic acid (Q-C3:0) did not show any remaining intact quercetin as evidenced by

189

HPLC-MS analysis. On the other hand, the crude product of the reaction of quercetin and

190

DHA (Q-DHA) showed around 56.7% of intact quercetin (triesters and tetraester were not

191

considered). The elongation of the carbon chain and kink characteristics might cause steric

192

hindrance in the case of DHA that affected the reaction progress.

193

Identification of quercetin derivatives

194

The HPLC-MS analysis of Q-C6:0 was carried out, because the isomers of quercetin with

195

shorter chain fatty acids could not be separated in the column due to their similar

196

lipophilicity. Quercetin with longer chain fatty acids, on the other hand, could be separated

197

in the column. However, substitution of more than 3 hydroxyl groups could not be detected

9

ACS Paragon Plus Environment

Page 10 of 29

Page 11 of 29

Journal of Agricultural and Food Chemistry

198

due to its extremely increased lipophilicity. Zhong and Shahidi5 reported that although the

199

yield of EGCG-C18:0, EGCG-EPA, and EGCG-DHA might be different due to their

200

structural characteristics, the main product of resultant in all case remained the same

201

(tetraester). The crude product of Q-C6:0 was subjected to HPLC-MS with methanol/ 5%

202

acetonitrile in water 90:10 (v/v) as the mobile phase (Figure 1A). Intact quercetin (2.907-

203

3.32 min), monoesters (3.32-4.44 min), diesters (4.44-6.293 min), triesters (8.133-12.113

204

min), and tetraester (19.26-20.68 min) were detected at their expected m/z of 303.05,

205

401.12, 499.20, 597.27, and 695.35, respectively (Figure 2). After that, the crude product

206

of Q-C6:0 was subjected to HPLC-MS with methanol/ 5% acetonitrile in water 80:20 (v/v)

207

as the mobile phase to achieve better separation (Figure 1B). Since quercetin has five

208

hydroxyl groups, the number of different esterification products or those with the same

209

number of esterification in different positions could vary. In this study, four monoesters

210

(4.427, 5.067-5.2, and 6.787 min), four diesters (10.873, 13.307, 15.18, and 18.947 min),

211

and three triesters (43.42, 51.5, and 75.527 min) were detected (Table 1) by HPLC-MS.

212

The monoesters from 5.067 to 5.2 min were not fully resolved and the peak had a small

213

shoulder, which could be expected as two compounds existed (Figure 1B). In this analysis

214

(methanol/ 5% acetonitrile in water, 80:20, v/v), tetraester was not detected due to

215

increased lipophilicity. Two monoesters (15.8 and 15.5%) and three diesters (10.8, 11.2,

216

and 15.4%) were predominant products.

217

1H-NMR

218

The predominant quercetin derivatives identified by HPLC-MS were two monoesters and

219

three diesters. Since there were several possibilities to form the different degrees of

220

substitution and isomers, information on the specific location of substitution was conducted

assignment of quercetin derivatives

10

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

221

using 1H-NMR and by comparison with the parent molecule as well as literature values.

222

Although the yields of esterification with different fatty acids were different, the

223

esterification pattern for all fatty acid esters remained the same. Therefore, one compound

224

was selected for further investigation. For this study, Q-DHA was selected due to its

225

complex structure and its high lipophilicity. The crude product of Q-DHA was separated

226

by TLC (hexane/ ethyl acetate/ formic acid, 3:3:0.12, v/v/v). Each compound was

227

subsequently subjected to 1H-NMR. The chemical shifts of the parent molecule were very

228

similar to those reported in the literature.19,20 Since the separations of two monoesters and

229

four diesters were unsuccessful due to their extremely similar lipophilicity, two spectral

230

assignments of two monoesters and two diesters were carried out as a mixture. The

231

chemical shifts of two unknown monoesters (QM1 and QM2) and four unknown diesters

232

(QD1, QD2, QD3, and QD4) are summarized in Table 2. For QM1, H-6 showed a similar

233

chemical shift (δ 6.19 ppm) compared to that of quercetin (δ 6.20 ppm), which indicated

234

esterification did not occur on A ring. While H-2' and H-6' of QM1 showed downfield

235

shifts (Δδ= 0.07 and 0.14 ppm, respectively), H-5' displayed similar chemical shifts (δ 6.92

236

ppm) compared to quercetin (δ 6.89 ppm). There were two possible substitution positions

237

which were either 3'-OH or 4'-OH. Since chemical shift did not occur on H-5', the

238

substitution position could be on 3'-OH. This was similar to that given by Moalin et al.21

239

who reported chemical shifts of 3'-methylated quercetin as 6.21 (H-6), 6.49 (H-8), 6.95 (H-

240

5'), 7.70 (H-6'), and 7.77 (H-2') ppm. For QM2, H-6 was slightly shifted to downfield (Δδ=

241

0.05). Interestingly, H-2' and H-6' of QM2 showed large upfield shifts (Δδ= -0.37 and -

242

0.30 ppm, respectively), whereas H-5' showed similar chemical shifts (δ 6.92 ppm)

243

compared to quercetin (δ 6.89 ppm); this might be explained by anisotropy.22 Therefore,

11

ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29

Journal of Agricultural and Food Chemistry

244

the possible substitution position of QM2 was expected to be on 3-OH. This was also

245

similar to the finding of Moalin et al. 21 Hydrogens in positions 6 and 8 of QD1 showed

246

large downfield chemical shifts (Δδ= 0.38 and 0.56 ppm, respectively). Thus, one of the

247

substitutions could be on 7-OH. The other substitution of QD1 occurred on the B ring. The

248

chemical shift of H-2', H-5', and H-6' was similar to QM1, the other substitution could be

249

expected on 3'-OH. For these reasons, QD1 could be Q-7,3'-O-diester. For QD2, the

250

chemical shifts of H-6 (δ 6.20 ppm) and H-8 (δ 6.44 ppm) were similar to those of H-6 (δ

251

6.20 ppm) and H-8 (δ 6.42 ppm) of quercetin, which meant substitution did not occur on

252

A ring. However, large downfield chemical shifts (Δδ= 0.15 and 0.41 ppm) were observed

253

on H-2' and H-6' of QD2 compared to quercetin. In addition, H-5' of QD2 also showed

254

large downfield shift (Δδ= 0.19 ppm) compared to quercetin. According to Moalin et al.21,

255

H-6' peak placed same or higher chemical shift (downfield) than H-2' peak, when

256

methylation occurred on 4'-OH. In this work, H-6' (δ 7.96 ppm) of QD2 also showed

257

higher ppm than H-2' (δ 7.84 ppm) of QD2. Hence, the substitution positions of QD2

258

could be on 3'-OH and 4'-OH. For QD3, the chemical shift of H-6 (δ 6.21 ppm) and H-8

259

(δ 6.43 ppm) showed similar to that of quercetin (δ 6.20 and δ 6.42 ppm, respectively),

260

which means the esterification did not occur on H-5 and H-7. In addition, H-2' (δ 7.80 ppm)

261

showed a higher chemical shift than H-6' (δ 7.61 ppm), which means the esterification did

262

not occur on 4'-OH. Therefore, QD3 could be Q-3,3'-O-diester. For QD4, the H-6 and H-8

263

of QD4 showed downfield shifts (Δδ 0.06 and 0.08, respectively) compared to quercetin.

264

While H-2' of QD4 displayed upfield shift (Δδ -0.06), H-5' and H-6' showed downfield

265

shifts (Δδ= 0.21 and 0.11 ppm, respectively). Therefore, QD4 was concluded to be Q-

266

3,4'-O-diester. 12

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

267

In summary, QM1, QM2, QD1, QD2, QD3 and QD4 were tentatively identified as being

268

Q-3'-O-monoester, Q-3-O-monoester, Q-7,3'-O-diester, Q-3',4'-O-diester, Q-3,3'-O-diester,

269

and Q-3,4'-O-diester, respectively (Figure 3).

270

Lipophilicity of quercetin derivatives

271

The esterification of quercetin with fatty acyl chlorides was carried out in order to increase

272

its lipophilicity. The lipophilicity was calculated using the lipophilicity calculator program,

273

ALOGPS 2.1. The shake-flask method which is commonly used for measuring

274

lipophilicity, was not employed here due to certain technical difficulties.18,23-25 The

275

calculated lipophilicity of quercetin was 1.81. As shown in Table 3, the lipophilicity of the

276

quercetin derivatives was increased as the chain length and the number of substitution

277

increased. The increased lipophilicity was also observed in the HPLC analysis. However,

278

unsaturation disturbed the lipophilicity pattern. While the monoester and diester of Q-

279

C18:0 had lipophilicity values of 7.89 and 10.26, respectively, that of Q-C18:1 had

280

lipophilicity values of 7.87 and 10.14, respectively. Moreover, Q-EPA showed a drastic

281

decrease of its lipophilicity compared to Q-C18:0 and Q-C18:1, despite the elongation of

282

chain length. This might be due to the nature of unsaturation. Carbon-carbon double bonds

283

with slight polarity have higher electron density than carbon-carbon single bonds.26 Hence,

284

the large number of unsaturation of EPA (C20:5) could enhance electron density so that

285

the lipophilicity of Q-EPA might decrease, compared to Q-C18:0. The extent of decreased

286

lipophilicity of Q-EPA was increased as the number of substitution increased (diester >

287

monoester). A similar observation was also made by Zhong and Shahidi5 who reported that

288

the lipophilicity of EGCG derivatives was in the order EGCG-SA (stearic acid) > EGCG-

289

EPA > EGCG-DHA; EGCG-EPA and EGCG-DHA showed a nonsignificant difference (P

13

ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29

Journal of Agricultural and Food Chemistry

290

> 0.05) as was evaluated by shake flask method. They also reported that the HPLC retention

291

time obtained was in the order EGCG-SA > EGCG-DHA > EGCG-EPA. A similar

292

lipophilicity trend was exhibited when resveratrol was esterified.18 The lipophilicity of

293

resveratrol (R) esterified with SA (9.30) showed the highest value followed by R-DHA

294

(8.59) and R-EPA (8.47).

295

ABTS radical cation scavenging assay

296

Quercetin is known to be a very powerful antioxidant. Esterification could affect its

297

antioxidant activity either positively or negatively. To evaluate antioxidant activity of

298

quercetin and its derivatives, ABTS radical cation scavenging assay was employed. In this

299

study, intact quercetin was removed then the mixtures of quercetin esters were used in this

300

evaluation assay. As shown in Figure 4, quercetin showed the highest ABTS radical cation

301

scavenging activity among all tested compounds. Among derivatives, Q-C3:0 and Q-C4:0

302

showed better ABTS radical cation scavenging activity than other derivatives. Oh and

303

Shahidi17 reported a similar behaviour for resveratrol and its derivatives. Resveratrol

304

showed the highest ABTS radical cation scavenging activity among all compounds

305

examined. In addition, resveratrol with short-chain fatty acids showed better antioxidant

306

activity than other derivatives. According to Mainini et al.27, quercetin esters showed

307

weaker DPPH and ABTS radical cation scavenging activity compared to the parent

308

molecule. This might be due to their loss of most active hydroxyl groups upon esterification.

309

In addition, the powerful antioxidant activity of quercetin has been known to originate from

310

a combination of the 3',4'-dihydroxyl structure (catechol) in the B ring, 3- and 5-hydroxyl

311

groups in the A and B rings, and the 2,3-double bond and 4-oxo group in the C ring.28-30

312

The quercetin derivatives which were tested here might not have the 3',4'-dihydroxy

14

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

313

structure in the B ring due to their esterification. Hence this could be another reason why

314

quercetin showed the highest antioxidant activity in this study.

315

DPPH radical scavenging assay

316

The determination of DPPH radical scavenging capacity of quercetin and its esters was

317

carried out to study their electron and/or hydrogen donation.31 DPPH radical is

318

commercially available without any preparation to generate the radical, thus it is commonly

319

used as an initial step in the evaluation of antioxidant activity.32 As reflected in Figure 5,

320

quercetin showed the highest DPPH radical scavenging activity among all tested samples.

321

This might be due to the decreased number of hydroxyl groups, which can donate their

322

hydrogen and/or electron. Mainini et al.27 also observed that quercetin showed the highest

323

DPPH radical scavenging activity compared to quercetin esters tested. In addition,

324

quercetin with short-chain fatty acids exhibited a better DPPH radical scavenging activity

325

compared to quercetin with long-chain fatty acids, except for Q-C16:0. Moreover, Q-C18:0

326

and Q-C18:1 exhibited the lowest DPPH radical scavenging activity. According to Oh and

327

Shahidi18, resveratrol with C18:0 and C18:1 displayed the highest DPPH radical

328

scavenging activity. DPPH radical is lipophilic, thus it was hypothesised that quercetin

329

with long-chain fatty acids would show better radical scavenging activity due to their

330

increased affinity to lipophilic DPPH. Sun et al.33 reported that tyrosol with short- and

331

medium-chain fatty acids also showed better DPPH radical scavenging activity than tyrosol

332

with long-chain fatty acids, except for tyrosol with DHA. However, Oh and Shahidi18

333

reported that resveratrol with short- and medium-chain fatty acids exhibited lower DPPH

334

radical scavenging activity than resveratrol with long-chain fatty acids. Furthermore,

335

Zhong and Shahidi5 observed that DPPH radical scavenging activity of epigallocatechin

15

ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29

Journal of Agricultural and Food Chemistry

336

gallate (EGCG) derivatives compared to the parent EGCG was enhanced dramatically due

337

to esterification. The effect of esterification on antioxidant activity was found to be

338

compound-dependent; this requires further evaluation.

339

In conclusion, esterification of quercetin with the fatty acyl chlorides of different chain

340

length and degree of unsaturation afforded derivatives whose structures were confirmed by

341

HPLC-MS and 1H-NMR. These were mainly the Q-3'-O-monoester, Q-3-O-monoester, Q-

342

7,3'-O-diester, Q-3',4'-O-diester, Q-3,3'-O-diester and Q-3,4'-O-diester. As expected, the

343

lipophilicity of quercetin derivatives was increased. The antioxidant activity of quercetin

344

and its derivatives for scavenging the ABTS radical cation and DPPH radical showed that

345

quercetin had the highest activity among all tested sample. Quercetin with short-chain fatty

346

acids (Q-C3:0 and Q-C4:0) possessed a better antioxidant activity than the others in ABTS

347

radical cation scavenging assay, whereas quercetin with C3:0, C4:0, C6:0, and C8:0

348

showed better antioxidant activity in DPPH radical scavenging assy. Although ABTS

349

radical cation and DPPH radical scavenging assay provided important information about

350

antioxidant activity, this needs to be further extended to include both in vitro and in vivo

351

studies.

352 353

Acknowledgement

354

This work was financially supported through a Discovery Grant from the Natural Science

355

and Engineering Research Council (NSERC) of Canada to one of us (FS).

356 357

16

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 18 of 29

358

Table 1. Composition (peak area, %) of quercetin-caproic acid ester (Q-C6:0) derivatives

359

based on HPLC analysis.a

360

Peak number

Monoesters

Diesters

Triesters

1

6.0

10.8

4.5

2

15.8

5.2

7.1

3

15.5

11.2

4.9

4

3.6

15.4

N/A*

a Intact

quercetin and tetraester were not considered. * Not available.

361 362 363 364 365 366 367 368 369 370

17

ACS Paragon Plus Environment

Page 19 of 29

371

Journal of Agricultural and Food Chemistry

Table 2. 1H chemical shift (δ in ppm) of quercetin and Q-C6:0 derivatives. 1H

Quercetin (Q)a

QM1b

QM2 b

QD1 a

QD2 a

QD3 a

QD4 a

6

6.20

6.19

6.24

6.58

6.20

6.21

6.26

8

6.42

6.47

6.46

6.99

6.44

6.43

6.50

2'

7.69

7.75

7.32

7.76

7.84

7.80

7.54

5'

6.89

6.94

6.89

6.92

7.08

7.14

7.10

6'

7.55

7.68

7.25

7.60

7.96

7.61

7.66

position

372

The compounds were tentatively identified as Q-3'-O-monoester (QM1), Q-3-O-monoester

373

(QM2), Q-7,3'-O-diester (QD1), Q-3',4'-O-diester (QD2), Q-3,3'-O-diester (QD3), and Q-

374

3,4'-O-diester (QD4).

375

a 1H-NMR

(300 MHz) was used.

376

b 1H-NMR

(500 MHz) was used.

377 378 379 380 381 382

Table 3. Lipophilicity of quercetin (Q) and its derivatives. Q (1.81)

Monoester

18

ACS Paragon Plus Environment

Diester

Journal of Agricultural and Food Chemistry

Page 20 of 29

Q-C3:0

2.88

3.70

Q-C4:0

3.13

4.27

Q-C6:0

3.81

5.51

Q-C8:0

4.65

6.65

Q-C10:0

5.38

7.92

Q-C12:0

6.08

8.97

Q-C14:0

6.78

9.66

Q-C16:0

7.38

10.00

Q-C18:0

7.89

10.26

Q-C18:1

7.87

10.14

Q-EPA

7.23

8.87

Q-DHA

7.48

8.89

383

The lipophilicity of quercetin (1.81) and its derivatives was

384

calculated using ALOGPS 2.1. The SMILE structures of the

385

compounds were drawn using ChemDraw Standard 16.0.

386 387 388

19

ACS Paragon Plus Environment

Page 21 of 29

Journal of Agricultural and Food Chemistry

389

x10 3 DAD1 - B:Sig=370.0,16.0 Ref=off WO6129a.d 3.560 1.2

a b

1 0.8

c

d

4.613

0.6

A e

5.800

0.4 0.2

8.453

11.187

* 20.133

0 1

2

3

4

5

6

7

8

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 Response Units vs. Acquisition Time (min)

390 x10 3 1

DAD1 - B:Sig=370.0,16.0

a

0.8

x10 3 1 0.8 0.6

DAD1 - B:Sig=370.0,16.0 * 5.067

10.873 6.787

1

2

3

18.947 43.420 51.500

0 10

15

B

c 15.180

9.240

18.947

0

10.873

5

b

3.347

0.2

c

0.2

391

Ref=off WO6129d.d

0.4

a b

0.4

Ref=off WO6129d.d 0.6

* 5.067

20

25

4

5

d

6

7

8

9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 2 Response Units vs. Acquisition Time (min)

30 35 40 45 50 55 60 65 70 Response Units vs. Acquisition Time (min)

75.527 75

80

85

90

95

392

Figure 1. HPLC chromatogram of Q-C6:0 with methanol/ 5% acetonitrile in water (A)

393

90:10 and (B) 80:20 as mobile phase (a, quercetin (Q); b, Q-C6:0 monoesters; c, Q-C6:0

394

diesters; d, Q-C6:0 triesters; e, Q-C6:0 teterester).

395 396 397 398

20

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 29

x10 6 + Scan (3.255-3.570 min, 20 Scans) WO6129d.d

1.2

A

303.0514

1.6 1.4

400

399

600

800

1221.9889

922.0096

832.2404

0

731.2166

0.4 0.2

622.0297

439.1014

0.6

521.1793

1 0.8

1000 1200 1400 1600 1800 Counts vs. Mass-to-Charge (m/z)

2000

2200

2400

x10 6 + Scan (4.946-5.012, 5.626-6.605 min, 65 Scans) WO6129d.d 1.6

401.1242

B

1.4 1.2 1

400

400

600

800

1221.9882

1019.3652

922.0089

0

823.2212

0.2

683.2308

0.4

457.1502 521.1793

0.6

284.3322

0.8

1000 1200 1400 1600 1800 Counts vs. Mass-to-Charge (m/z)

2000

2200

2400

x10 6 + Scan (18.280-18.611, 20.021-21.464 min, 109 Scans) WO6129d.d

400

401

600

800

1221.9885

931.5388

555.2232

0

401.1241

0.5 0.25

284.3323

1

1019.3666

1.5 1.25 0.75

C

499.1974

1.75

1000 1200 1400 1600 1800 Counts vs. Mass-to-Charge (m/z)

2000

2200

2400

x10 6 + Scan (41.680-42.144, 44.366-45.229 min, 82 Scans) WO6129d.d

D

597.2704

1.5 1.25

402

300

600

1221.9880

500

922.0091

400

665.3070 704.2289

499.1976 539.4468

0

391.2853

0.25

343.3394

0.5

284.3323

1 0.75

700 800 900 1000 1100 1200 Counts vs. Mass-to-Charge (m/z)

1300

1400

1500

403

Figure 2. Mass spectra of quercetin (Q) and Q-C6:0 derivatives (A, quercetin (Q); B, Q-

404

C6:0 monoesters; C, Q-C6:0 diesters; D, Q-C6:0 triesters).

21

ACS Paragon Plus Environment

Page 23 of 29

Journal of Agricultural and Food Chemistry

405 406 407

408 409 410

Figure 3. Structures of quercetin. Red circles present substitution occurring position.

411 412

22

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

413

a

10 8 6

b

c

d

efg

ef

gh

e

d

h

fg

ef

ef

4 2 0

Q QC3 :0 QC4 :0 QC6 :0 QC8 :0 QC1 0: 0 QC1 2: 0 QC1 4: 0 QC1 6: 0 QC1 8: 0 QC1 8: 1 QEP A QDH A

ABTS (μmol Trolox eq./mg sample)

12

Page 24 of 29

414

Figure 4. ABTS radical cation scavenging activity of quercetin (Q) and its derivatives

415

expressed as trolox equivalents. All determinations were repeated three times using the

416

same combined starting materials and mean values and standard deviations reported. The

417

different letters are significantly different (P < 0.05). Columns with same letters such as

418

“a” and “ab” are not significantly different.

419 420 421 422 423

23

ACS Paragon Plus Environment

Page 25 of 29

Journal of Agricultural and Food Chemistry

a

3.50 3.00 2.50 2.00 1.50

bc bc

1.00

bc

b

cd

de

cd

b

f

ef

cd

de

0.50

C8 Q- :0 C1 0 Q - :0 C1 2 Q - :0 C1 4 Q - :0 C1 6 Q - :0 C1 8 Q - :0 C1 8: 1 QEP A QDH A

:0

Q-

C6

:0 C4

Q-

424

Q-

QC3

:0

0.00

Q

DPPH (μmol Trolox eq./mg sample)

4.00

425

Figure 5. DPPH radical scavenging activity of quercetin (Q) and its derivatives expressed

426

as trolox equivalents. All determinations were repeated three times using the same

427

combined starting materials and mean values and standard deviations reported. The

428

different letters are significantly different (P < 0.05). Bars with the same letters such as “a”

429

and “ab” are not significantly different.

430 431 432 433

Reference

434

1. Pace-Asciak, C. R.; Hahn, S.; Diamandis, E. P.; Soleas, G.; Goldberg, D. M. The red

435

wine phenolics trans-resveratrol and quercetin block human platelet aggregation and

436

eicosanoid synthesis: implications for protection against coronary heart disease. Clin.

437

Chim. Acta 1995, 235, 207-219.

24

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

438

2. Kaidama, W. M.; Gacche, R. N. Anti-inflammatory activity of quercetin in acute and

439

chronic phases of inflammation in guinea pigs. Am. J. Phytomedicine Clin. Ther.

440

2015, 3, 129-136.

441

3. Lu, J.; Papp, L. V.; Fang, J.; Rodriguez-Nieto, S.; Zhivotovsky, B.; Holmgren, A.

442

Inhibition of mammalian thioredoxin reductase by some flavonoids: implications for

443

myricetin and quercetin anticancer activity. Cancer Res. 2006, 66, 4410-4418.

444

4. Graefe, E. U.; Wittig, J.; Mueller, S.; Riethling, A. -K.; Uehleke, B.; Drewelow, B.;

445

Pforte, H.; Jacobasch, G.; Derendorf, H.; Veit, M. Pharmacokinetics and

446

bioavailability of quercetin glycosides in humans. J. Clin. Pharmacol., 2001, 41,

447

492-499.

448 449

5. Zhong, Y.; Shahidi, F. Lipophilized Epigallocatechin Gallate (EGCG) Derivatives as Novel Antioxidants. J. Agric. Food Chem. 2011, 59, 6526-6533.

450

6. Zhong, Y.; Shahidi, F. Lipophilised epigallocatechin gallate (EGCG) derivatives and

451

their antioxidant potential in food and biological systems. Food Chem. 2012, 131,

452

22-30.

453

7. Zhong, Y.; Chiou, Y., -S.; Pan, M. -H.; Ho, C. -T.; Shahidi, F. Protective effects of

454

epigallocatechin gallate (EGCG) derivatives on azoxymethane-induced colonic

455

carcinogenesis in mice. J. Funct. Foods 2012, 4, 323-330.

456 457

458 459

8. Zhou, D. -Y.; Sun, Y. -X.; Shahidi, F. Preparation and antioxidant activity of tyrosol and hydroxytyrosol esters. J. Funct. Foods 2017, 37, 66-73. 9. Pokorski, M.; Marczak, M.; Dymecka, A.; Suchocki, P. Ascorbyl palmitate as a carrier of ascorbate into neural tissues. J. Biomed. Sci. 2003, 10, 193-198. 25

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

Journal of Agricultural and Food Chemistry

460

10. Wong, J. M. W.; de Souza, R.; Kendall, C. W. C.; Emam, A.; Jenkins, D. J. A. Colonic

461

health: fermentation and short chain fatty acids. J. Clin. Gastroenterol. 2006, 40,

462

235-243.

463 464

11. Brüssow, H.; Parkinson, S. J. You are what you eat. Nat. Biotechnol. 2014, 32, 243245.

465

12. Tapiero, H.; Ba, G. N.; Couvreur, P.; Tew, K. D. Polyunsaturated fatty acids (PUFA)

466

and eicosanoids in human health and pathologies. Biomed. Pharmacother. 2002, 56,

467

215-222.

468

13. Biasutto, L.; Marotta, E.; Marchi, U. D.; Zoratti, M.; Paradisi, C. Ester-based precursors

469

to increase the bioavailability of quercetin. J. Med. Chem., 2007, 50, 241-253.

470

14. Wanasundara, U. N.; Shahidi, F. Concentration of omega 3-polyunsaturated fatty acids

471

of seal blubber oil by urea complexation: optimization of reaction conditions. Food

472

Chem. 1999, 65, 41-49.

473

15. Tetko, I. V.; Bruneau, P. Application of ALOGPS to predict 1-octanol/water

474

distribution coefficients, logP, and logD
of AstraZeneca in-house database. J. Pharm.

475

Sci. 2004, 93, 3101-3110.

476

16. Miller, N.; Rice-Evans, C.; Davies, M.; Gopinathan, V.; Milner, A. A novel method for

477

measuring antioxidant capacity and its application to monitoring the antioxidant

478

status in premature neonates. Clin. Sci. 1993, 84, 407-412.

479

17. Ambigaipalan, P.; Al-Khalifa, A. S.; Shahidi, F. Antioxidant and angiotensin I

480

converting enzyme (ACE) inhibitory activities of date seed protein hydrolysates

481

prepared using Alcalase, Flavourzyme and Thermolysin. J. Funct. Foods 2015, 18,

26

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

482 483 484

1125-1137. 18. Oh, W. Y.; Shahidi, F. Lipophilization of resveratrol and effects on antioxidant activities. J. Agric. Food Chem. 2017, 65, 8617-8625.

485

19. Bukhari, S. B.; Memon, S.; Tahir, M. M.; Bhanger, M. I. Synthesis, characterization

486

and investigation of antioxidant activity of cobalt-quercetin complex. J. Mol. Struct.

487

2008, 892, 39-46.

488

20. Galati, G.; Moridani, M. Y.; Chan, T. S.; O’Brien, P. J. Peroxidative metabolism of

489

apigenin and naringenin versus luteolin and quercetin: glutathione oxidation and

490

conjugation. Free Radic. Biol. Med. 2001, 30, 370-382.

491

21. Moalin, M.; van Strijdonck, G. P. F.; Beckers, M.; Hagemen, G. J.; Borm, P. J.; Bast,

492

A.; Haenen, G. R. M. M. A planar conformation and the hydroxyl groups in the b

493

and c rings play a pivotal role in the antioxidant capacity of quercetin and quercetin

494

derivatives. Molecules 2011, 16, 9636-9650.

495

22. Richards, S. A.; Hollerton, J. C. 2-D proton-carbon (multiple bond) correlated

496

spectroscopy. In Essential practical NMR for organic chemistry, John Wiley & Sons,

497

Ltd.: West Sussex, United Kingdom, 2011; pp. 74-135.

498

23. Ayouni, L.; Cazorla, G.; Chaillou, D.; Herbreteau, B.; Rudaz, S.; Lantéri, P.; Carrupt

499

P. A. Fast determination of lipophilicity by HPLC. Chromatographia 2005, 62, 251-

500

255.

501

24. Giaginis, C.; Tsantili-Kakoulidou, A. Current state of the art in HPLC methodology for

502

lipophilicity assessment of basic drugs. A Review. J. Liq. Chromatogr. Relat.

503

Technol. 2008, 31, 79-96.

504

25. Finizio, A.; Vighi, M.; Sandroni
 D. Determination of n-octanol/water partition

27

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

Journal of Agricultural and Food Chemistry

505

coefficient (Kow) of critical review and comparison of methods. Chemosphere 1997,

506

34, 131-161.

507 508

26. Bruice, P., Y. Electronic structure and bonding, acids and bases. In Organic chemistry, 6.; Prentice hall: Glenview, IL, USA, 2011; pp. 25-33.

509

27. Mainini, F.; Contini, A.; Nava, D.; Corsetto, P. A.; Rizzo, A. M.; Agradi, E.; Pini, E.

510

Synthesis, molecular characterization and preliminary antioxidant activity evaluation

511

of quercetin fatty esters. J. Am. Oil Chem.’ Soc. 2013, 90, 1751-1759.

512 513

28. Bors, W.; Heller, W.; Saran, H. Flavonoids as antioxidants: determination of radical scavenging efficiencies. Methods Enzymol. 1990, 186, 343-355.

514

29. Pulido, R.; Bravo, L.; Saura-Calixto, F. Antioxidant activity of dietary polyphenols as

515

determined by a modified ferric reducing/antioxidant power assay. J. Agric. Food

516

Chem. 2000, 48, 3396-3402.

517

30. Rice-Evans, C. A.; Miller, N. J.; Paganga, G. Structure-antioxidant activity

518

relationships of flavonoids and phenolic acids. Free Radic. Biol. Med. 1996, 20, 933–

519

956.

520

31. Prior, R. L., Wu, X. and Schaich, K. Standardized methods for the determination of

521

antioxidant capacity and phenolics in foods and dietary supplements. J. Agric. Food

522

Chem., 2005, 53, 4290-4302.

523 524

32. Shahidi, F. and Zhong, Y. Measurement of antioxidant activity. J. Funct. Foods, 2015, 18, 757-781.

525

33. Sun, Y., Zhou, D. and Shahidi, F. (2018). Antioxidant properties of tyrosol and

526

hydroxytyrosol saturated fatty acid esters. Food Chem., 2018, 245, 1262-1268.

28

ACS Paragon Plus Environment