Dietary Flavonoids Scavenge Hypochlorous Acid via Chlorination on

5 days ago - Dietary flavonoids are known as scavengers of reactive oxygen species such as hypochlorous acid. In spite of abundant scavenging capacity...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF WISCONSIN - MADISON

Functional Structure/Activity Relationships

Dietary Flavonoids Scavenge Hypochlorous Acid via Chlorination on A and C Rings as Primary Reaction Sites: Structure and Reactivity Relationship Xin Yang, Tian Wang, Jinlong Guo, Mingtai Sun, Ming Wah Wong, and Dejian Huang J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b06689 • Publication Date (Web): 22 Mar 2019 Downloaded from http://pubs.acs.org on March 24, 2019

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 32

1 2 3

Journal of Agricultural and Food Chemistry

Dietary Flavonoids Scavenge Hypochlorous Acid via Chlorination on A and C Rings as Primary Reaction Sites: Structure and Reactivity Relationship

4 5

Xin Yang,† Tian Wang,# Jinlong Guo,# Mingtai Sun,†,‡ Ming Wah Wong,*,# Dejian Huang*,†,§

6

†Food

7

gapore, 3 Science Drive 3, Singapore 117543, Republic of Singapore

8

‡Institute

9

#Department

Science and Technology Programme, Department of Chemistry, National University of Sin-

of Intelligent Machines, Chinese Academy of Sciences, Hefei, Anhui 230031, China of Chemistry, National University of Singapore, 3 Science Drive 3, Singapore, 117543,

10

Singapore

11

§National

12

Jiangsu 215123, China

University of Singapore (Suzhou) Research Institute, 377 Linquan Street, Suzhou,

13

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

14

Page 2 of 32

ABSTRACT

15

Dietary flavonoids are known as scavengers of reactive oxygen species such as hypochlorous acid.

16

In spite of abundant scavenging capacity data reported, few reports have addressed the relation-

17

ship between the scavenging capacity and structures of different flavonoids. We characterized the

18

reaction products of five flavonoids (apigenin, quercetin, naringenin, ampelopsin and epicatechin)

19

with hypochlorous acid and found that primary chlorination reaction occurred on A ring (C6 or

20

C8) and/or C rings but not on B rings. Correlation of the hypochlorous acid scavenging capacity

21

(IC50 values) and the structural features of flavonoids revealed that the hydroxyl groups in the A

22

ring and B ring can enhance the scavenging capacity whereas the C(2)C(3) double bond has nega-

23

tive impact to the HClO scavenging capacity. Combining the SAR analysis and chemical study, we

24

proposed that reaction mechanism between flavonoids and HClO should be electrophilic substi-

25

tution reaction. Density functional theory (DFT) results are in consistent with the selectivity of

26

chlorination on the flavonoids. Our findings highlight the importance of considering specific re-

27

active oxygen species when measuring radical scavenging capacity of dietary antioxidants.

28 29

KEYWORDS: hypochlorous acid scavenging capacity, flavonoids, antioxidant, density functional theory, structure-activity relationship.

2 ACS Paragon Plus Environment

Page 3 of 32

30

Journal of Agricultural and Food Chemistry

INTRODUCTION

31

Reactive oxygen species (ROS) in aerobic organisms play important roles in normal function of

32

organisms yet overproduction of them may lead oxidative stress,1 which is a causative factor of

33

several chronic diseases2-7 such as cancer, cardiovascular diseases,8-10 and aging itself.11-13 Among the

34

various ROS, hypochlorous acid is a weak acid (pKa 7.5) with HClO and ClO- both present amount

35

equal amount at physiological pH and can act as a weapon to deactivate pathogens in vitro and in

36

vivo.14 Hypochlorous acid is formed by oxidation of chloride by hydrogen peroxide in the presence

37

of myeloperoxidase (MPO) found in white blood cells such as neutrophils,15 monocytes, and macro-

38

phages.16 Hypochlorous acid is a non-specific oxidizing agent that can react rapidly with a wide

39

range of biomolecules including amino acids,17 proteins,18 carbohydrates,19 lipids,20 and DNA.21

40

Phytochemicals that can scavenge reactive oxygen species have believed to be able to provide

41

protective effects on oxidative stress and flavonoids are a major class of dietary antioxidants. Struc-

42

ture-activity relationship (SAR) for the ROS/RNS scavenging capacity of flavonoids were investi-

43

gated and it was suggested that increasing numbers of the phenolic groups enhance the antioxi-

44

dant or prooxidant activity in certain cases.22 In addition, C(2), C(3)-double bond in C ring allows

45

electron delocalization across three ring skeleton (A, B, C rings) and thus provides stabilization of

46

the aryloxyl radicals formed upon reaction with ROS.23 In biological systems, there are wide range

47

of reactive oxygen species, which have different reactivity patterns with dietary antioxidants. How-

48

ever, there is little study on product characterizations of flavonoids and specific reactive oxygen

49

species. Scavenging HClO is of special importance because of its relation to inflammation related

50

tissue damage. Binsack and coworkers pioneered the chemical characterization of the reaction

51

products of quercetin and rutin with HClO and found that the chlorination occurs at C6 and C8

52

positions of A rings. It is remarkable that the chlorinated products are more potent antioxidants 3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 32

53

than the parent compounds.24 Furthermore, Boersma et al studied the reaction products of isofla-

54

vones (genistein, biochanin-A, and daizein) and found that the chlorination occurred at A-ring.25

55

To further systematically uncover the structure and reactivity relationship between 24 types of fla-

56

vonoids and their scavenging capacity against individual reactive oxygen species. We measured the

57

HClO scavenging capacity of characterized the reaction products of flavonoids with different rep-

58

resentative structures (apigenin, quercetin, naringenin, ampelopsin and epicatechin) with hypo-

59

chlorous acid.

60

EXPERIMENTAL SECTION

61

Materials and Instruments. Hypochlorous acid (15%) was obtained from Merck & Co., Inc and

62

its concentration quantified using UV-VIS spectroscopic method. Fluorescein disodium were ob-

63

tained from Aldrich (Singapore, Singapore). Flavonoids (7,8-dihydroxyflavone, baicalein, luteolin,

64

scutellarein, wogonin, fisetin, kaempferol, morin, myricetin, quercetin, 3,3',4'-trihydroxyflavone,

65

3,5,7,8,3',4'-hexahydroxyflavone, alpinetin, eriodictyol, liquiritigenin, hesperetin, naringenin, pi-

66

nocembrin, ampelopsin, taxifolin, catechin, epicatechin, epigallocatechin, genistein) were ob-

67

tained from Nanjing Plant Origin Biological Technology Co., Ltd. A Synergy HT microplate fluo-

68

rescence reader (Bio-Tek Instruments, Inc., Winooski, VT) was used for measurement of fluores-

69

cence with an excitation filter of 485 ± 20 nm and an emission filter of 530 ± 25 nm. The plate reader

70

was controlled by software KC4 3.0 (revision 29). Sample dilution was accomplished by a Precision

71

X automatic pipetting system managed by precision power software (version 1.0) (Bio-Tek Instru-

72

ments, Inc.). The 96-well polystyrene microplates and the covers were purchased from VWR Inter-

73

national Inc (Bridgeport, NJ). All aqueous solutions were prepared with 18.2 MΩ·cm ultrapure wa-

74

ter obtained by a Millipore water purification system. 1H NMR spectra were measured using a

75

Bruker AVANCE I 300 NMR spectrometer, and mass spectrometry was performed on a Thermo 4 ACS Paragon Plus Environment

Page 5 of 32

Journal of Agricultural and Food Chemistry

76

Scientific LCQ Fleet ion trap mass spectrometer in ES positive mode. Merck F254 silica gel-60

77

plates were chose to do thin-layer chromatography (TLC) and silica gel-60 (230–400 mesh) was

78

selected as solid phrases for column chromatography.

79

All flavonoids were dissolved in DMF to make 5.0 mM stock solution for storage at freezer (-20

80

oC).

They were diluted with 75 mM potassium phosphate buffer (pH 7.4) to 50 µM for measure-

81

ment. Hypochlorous acid was prepared by adjusting the pH of a 12-15% (m/v) solution of NaClO

82

to 6.2 with dropwise addition of 10% sulfuric acid. The concentration of HClO was further deter-

83

mined spectrophotometrically at 235 nm using the molar absorption coefficient of 100 (M−1cm−1).26

84

A 0.1 mM stock solution of fluorescein was stored in fridge (4 oC) and diluted to working solutions

85

(0.1 μM) immediately before the determination of HClO scavenging capacity.

86

Experimental Procedure for Hypochlorous acid Scavenging Assay. The HClO scavenging

87

capacity was measured by monitoring the effect of flavonoids on HClO induced bleaching of flu-

88

orescein by slide modification of a protocol previously described.27, 28 Reaction mixtures contained

89

the following reactants at the indicated final concentrations (final volume of 200 μL): 25 μL sam-

90

ples at five different concentrations, 150 μL fluorescein (0.1 μM), and 25 μL HClO (200 μM). The

91

fluorimetric assays were performed at 37 °C in the microplate reader, at the emission wavelength

92

at 528 ± 20 nm with excitation at 485 ± 20 nm. The mixture in a total volume was incubated for 20

93

min in a plate reader kept at 37 °C. The IC50 (50% inhibition concentration) values of the scavengers

94

were determined using fluorescein. The HClO scavenging capacity was calculated (Support infor-

95

mation Figure S-12 and S-13). The oxidation products of fluorescein was studied in the LC-MS (Fig-

96

ure S-14). The effects of flavonoids and flavonoids chlorinated products were discussed using lute-

97

olin as an example (Figure S-15).

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 32

98

Characterizations of Reaction Product by LC-MS. The LC–MS system was equipped with a

99

C18 column (Phenomenex, Luna 5µ C18, 250 × 4.6 mm) for analysis of the reaction products be-

100

tween flavonoids and HClO. The samples (30 μL) were filtered through 0.2 μm membrane (Merck

101

Millipore, USA) before being injected into the HPLC system. The column was equilibrated with

102

100% mobile phase A (water) for 10 min before gradient elution of mobile phase A ratio from 100 to

103

0% in 30 min in linear increase at a flow rate of 1.0 mL/min, and then remain 100% mobile phase B

104

(methanol) for the following 10 min. The Waters 2998 Photodiode Array (PDA) Detector was connected

105

to the HPLC system with detection wavelengths from 190 and 800 nm. Bruker AmaZon-X is applied for

106

LC-MS and LC-MS2 to characterize unknown compounds. The LC conditions for LC−MS analysis were

107

similar to those mentioned above, except detection wavelengths were set at 280 and 350 nm. All mass

108

spectra were acquired in both positive and negative ion mode using electrospray ionization. The parent

109

ion was selected with a width of ±2.5 Da and fragmented with 50% setting.

110

Preparative scale isolation of reaction products of flavonoids and hypochlorous acid.

111

Flavonoids (100 mg, luteolin, apigenin, quercetin, naringenin, ampelopsin and epicatechin) were

112

dissolved in acetone/water (1:1) solution and reacted with HClO (pH 6.2) in at molecular ratio of 1:

113

0.5. After evaporation of the solvent in vacuo, the mixture was subjected to silica column chroma-

114

tography to separate the products.

115

Density Functional Theory (DFT) Study. A first principles study of flavonoids: luteolin, quer-

116

cetin, naringenin, ampelopsin and epicatechin were carried out to investigate their differences in

117

the relevant electronic properties. All calculations were performed using Density Functional The-

118

ory (DFT) methods implemented in the Gaussian 09 package.29 Both geometry optimization and

119

electronic structures are calculated using B3LYP hybrid function with 6-311+g(d,p) basis sets.30-32

6 ACS Paragon Plus Environment

Page 7 of 32

Journal of Agricultural and Food Chemistry

120

The HOMO (highest occupied molecular orbital) and HOMO-1 (the second highest occupied mo-

121

lecular orbital) of all flavonoids were displayed on Gauss View 5.0.33

122

The condensed Fukui functions were calculated by taking the finite difference approximations

123

from population analysis of atoms in molecules, depending on the direction of electron transfer,

124

through the following formula:34

125

𝑓𝑘 + = 𝑞𝑘(𝑁 + 1) − 𝑞𝑘(𝑁) [for nucleophilic attack] (1)

126

𝑓𝑘 − = 𝑞𝑘(𝑁) − 𝑞𝑘(𝑁 − 1) [for electrophilic attack] (2)

127

𝑓𝑘 0 = [𝑞𝑘(𝑁 + 1) − 𝑞𝑘(𝑁 − 1)]/2 [for radical attack] (3)

128

where qk is the gross charge of atom k in the molecule.

129

Statistical Analysis. Descriptive statistical analyses were performed using Origin 8.0 for calcu-

130

lating the means and the standard error of the mean. Results were expressed as the mean ± stand-

131

ard deviation (SD).

132

RESULTS AND DISCUSSION

133

Hypochlorous acid Scavenging Capacity of Phenolic Compounds. This study on the SAR of

134

the scavenging of HClO by flavonoids was started by characterizing the scavenging capacity of

135

substituted phenols which shared a key group (phenolic moieties) with flavonoids. The concentra-

136

tion of phenol that can inhibit half of fluorescein oxidation (IC50) is 18.7 µM (1/IC50 =0.047 ± 0.0035).

137

Catechol, resorcinol and hydroquinone were tested to determine the effect of OH-substitution on

138

the scavenging capacity of phenol (Figure 1). Substitution position and the number of OH are the

139

key factors increases the scavenging capacity (1/IC50) of phenols. Catechol is more active than phe-

140

nol, resorcinol shows better scavenging capacity than pyrogallol.

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 32

141

The phenomenon that hydroxyl substituent at different positions results in dissimilarity in the

142

activity could be explained by the electronic effect. As shown in the Figure 2, the electron donating

143

or withdrawing effect was described by Hammett σ with negative value and positive value respec-

144

tively35. In a conjugated system, hydroxyl group acts as electron donor and enhances the electronic

145

density at the ortho and para positions while decrease the electron density at the meta position. As

146

such all the potential reaction sites in the phloroglucinol are activated whereas the other phenols,

147

some potential reaction sites in the aromatic ring were inactivated due to the electron withdrawing

148

effect by OH groups at the meta position. During the chlorination of these phenols, chlorine atom

149

in hypochlorous acid acts as electrophilic center and prefers to attack the electron-rich site in the

150

aromatic ring (as shown in Scheme 1). This nicely fits with the ranking order of scavenging capacity

151

against hypochlorous acid observed for the catechol, pyrogallol, resorcinol and phloroglucinol.

152

Thus, nucleophilic substitution reaction is suitable to describe the initial reaction between phenols

153

and HClO. With this understanding in mind, we examined the flavonoids structure and hypo-

154

chlorous acid scavenging capacity relationship.

155

Hypochlorous acid Scavenging Capacity of Flavonoids: Importance of B-ring HO groups as

156

shown in Figure 3, luteolin (3,5,7,3’,4’-pentahydroxyflavone), is a very good scavenger of HClO (IC50

157

= 2.39 μM). In sharp contrast, the homolog, apigenin (one HO less in B ring) has much lower HClO

158

scavenging capacity with IC50 > 9.375 μM). Similar trend is observed in other flavonols (quercetin

159

vs kaempferol), flavanones (eriodictyol vs naringenin), flavanonols (ampelopsin vs taxifolin), and

160

flavanol (epicatechin vs epigallocatechin). This observation agreed nicely with the theory pro-

161

posed by Cao, Sofic and Prior 22 that more OH groups in the aromatic ring result in better antioxi-

162

dant activity. 22

8 ACS Paragon Plus Environment

Page 9 of 32

Journal of Agricultural and Food Chemistry

163

HClO scavenging capacity of flavonoids: importance of HO groups in A ring. As shown in Figure

164

4, the extra OH group in the C6 of the A ring result in the increase of scavenging capacity from

165

IC50> 9.375 μM of apigenin to 3.99 μM of scutellarein). This trend was observed in flavonols and

166

flavanones.

167

The loss of the OH group in the C5 position of the A ring results in a significant decrease in the

168

scavenging capacity of the flavonoids. Moreover, OH group in the 7 sites also play a positive role in

169

increasing the HClO scavenging capacity. Similarly, the loss of the hydroxyl group in the 5 site of

170

the naringenin resulting in the increased IC50 from 1.97 μM to 6.55 μM (liquiritigenin). Overall,

171

the hydroxyl groups exhibit positive effect both in A-ring and B-ring, then enhance the electron

172

density on the aromatic ring, resulting in better attraction of electrophilic attack by hypochlorous

173

acid.

174

Effects of C(2)C(3) double bonds in C-ring. The C(2)C(3) double bond in the C seems to play

175

a negative role in the scavenging capacity as illustrated by two pairs of compounds, luteolin vs

176

eridictyol and apigenin vs naringenin (Figure 5), with the latter pair having greater impacts. This

177

results were unconventional because C(2)=C(3) allows electron delocalization across the molecule

178

for stabilization of the aryloxyl radicals, which has been proposed previously Rice-Evans, et al. 36.

179

However, the conjugated system linked by C(2)-C(3) double bond may lead to more electron flow

180

into electrophilic center in ketone group of C-ring. In the case of HClO, the reaction does not

181

involve radicals and the electron delocalization may play a negative role.

182

Characterization of reaction products of flavonoids with HClO. To shed some light on the

183

reasons behind the observed trend of IC50, we characterized the primary reaction products of se-

184

lected flavonoids and HClO by LC-MS(n). To ensure the reaction only involve HClO, we conducted

185

the reaction under weakly acidic conditions (pH 6.20), which flavonoids are not deprotonated. As 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 32

186

HClO is a potent oxidant and, if it is in large excess, it further oxidize the primary oxidation prod-

187

ucts of the flavonoids. To avoid such complication and better characterize the primary products of

188

HClO and flavonoids, we intentionally conduct the reaction with HClO as the limiting reagent.

189

The reaction products of the flavone (apigenin), flavonol (quercetin), flavanone (naringenin) and

190

flavanonol (ampelopsin) were summarized in the Figure 6A. For apigenin, three chlorination prod-

191

ucts were observed, they are AP-1, AP-2 and AP-3 formed at different retention times (27.68, 28.51

192

and 32.43 min) with similar UV-VIS spectral characteristics to that of apigenin. These products

193

were firstly characterized by their mass spectra (Table S1) and fragments analysis (Figure S7). These

194

compounds exhibit mass charge ratio (anionic mode) of 303.68 (AP-1), 337.29 (AP-2), and

195

371.51(AP-3) and are respectively mono, di, and trichlorinated apigenin which has m/z of 269.24.

196

To confirm the exact structures of the products, they were isolated and analyzed by 1H NMR (Figure

197

S1), the NMR spectra shown that monochlorinated apigenin (AP-1) was modified at the C(3) site

198

of the C-ring, and dichlorinated apigenin (AP-2) was modified at the C(3) site of the C-ring and

199

C(8) site of the A-ring whereas trichlorinated apigenin (AP-3) occurred at C(3) site of the C-ring

200

and C(8) as well as C(6) site of the A-ring.

201

Similarly, three more lipophilic chlorination products of quercetin, QU-1, QU-2 and QU-3 were

202

detected by HPLC at 25.63, 26.86, and 29.12 min following exposure of quercetin to HClO (Figure

203

6B). The MS signals of chlorinated products were recorded in the Table S1, and the fragmentation

204

analysis was shown in the Figure S8. The chlorinated sites of in the quercetin were revealed by 1H

205

NMR at C(8) and C(6) site of the A-ring (Figure S2). This result matched the conclusion proposed

206

by Binsack and coworkers.24 It is apparent that C(8) and C(6) are both sites for chlorination the

207

isomeric products were observed.

10 ACS Paragon Plus Environment

Page 11 of 32

Journal of Agricultural and Food Chemistry

208

Similarly, three chlorinated products of naringenin, NA-1, NA-2 and NA-3, were detected by

209

HPLC with retention times of 13.88, 14.23 and 16.26 min. Their structures are shown in Figure 6C.

210

Moreover, the fragmental analyses of NA-1 or NA-2 and NA-3 (Figure S9) show that the monochlo-

211

rination and dichlorination only happened at C(6) or C(8) site. The dichlorinated product NA-3

212

was characterized by 1H NMR (Figure S3). Remarkably, C(3) was not chlorinated because, from 1H

213

NMR spectrum, the hydrogen in the C(3) site remained. This is in contrast to flavone apigenin.

214

The reaction products between flavanonol and HClO was studied using ampelopsin (AM) as an

215

example. Three chlorinated products AM-1, AM-2 and AM-3 were observed by HPLC with reten-

216

tion times of 20.21, 21.93 and 24.28 min respectively (Figure 7D). LC-MS results shown that the

217

molecular ion (M-1, anionic mode) of 353.18, 353.18 and 387.02 respectively. From the 1H NMR spec-

218

tra of the monochlorinated product AM-1 and AM-2 (Figure S4), the C(8) or C(6) were the reaction

219

sites in the chlorination. The MS fragmentation patterns of AM-1, AM-2, or AM-3 (Figure S10)

220

suggested that only A-ring is the reactive sites in the chlorination.

221

The reaction between flavanol and HClO seems to be a different pattern. Following addition of

222

HClO to epicatechin, HPLC analysis revealed one major product (RT: 13.88 min) that was very close

223

to epicatechin (RT: 14.23 min) and photodiode array analysis of the peak identified similar spectral

224

characteristics (Figure 7). In addition, two more products appeared at 16.26 min and 18.28 min,

225

indicating products of high polarity. Fragment analysis was undertaken to identify the structures

226

of these reaction products with mass of 323.71 Da (CA-1), 307.71 Da (CA-2) and 425.61 Da (CA-3)

227

respectively. The hydrogen atom in C(6) or C(8) is the most reactive site for chlorination, resulting

228

in the formation major reaction product CA-1, whose structure was further confirmed by 1H NMR

229

(Figure S5). Then further chlorination happened in the rest potential sites of the A-ring and C-ring

230

(CA-3). However, the formation of CA-2 observed in 16.26 min indicated a different mechanism 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 32

231

which the OH group in the C-ring was substituted by chloride (protonation of HO group followed

232

and nucleophilic substitution by Cl-). The structures of CA-1, CA-2 and CA-3 were confirmed by

233

the fragmental analysis in Figure S11.

234

Reaction mechanisms of flavonoids with HClO. From the results shown in Figure 6 and 7, it

235

is apparent that there is a common features of the reaction between HClO and different types of

236

flavonoids. Chlorination happens on A ring and C ring but not on B ring. We propose a reaction

237

mechanism, electrophilic aromatic substitution, as described in the Scheme 2. The electrophile (Cl

238

atom in HClO) reacts with the electron-rich carbon (A or C ring) of the flavonoids. However it is

239

not obvious, judging from the flavonoid structure, which carbon in the flavonoid ring skeleton is

240

most likely nucleophile. Therefore, we seek DFT calculation to gain some insights.

241

As widely-used quantum chemical-descriptors, the HOMO-LUMO energies lay an essential role

242

in quantitative structure-activity and structure-property relationship (QSAR/QSPR) studies of

243

wide range of chemical reactions.37,38 The HOMO-LUMO frontier orbital compositions for flavo-

244

noids (apigenin, quercetin, naringenin, ampelopsin and epicatechin) calculated at the DFT/6-

245

311G(d,p) level are depicted in Figure 8 (a), (b), (c), (d) and (e) respectively. The highest and second

246

highest occupied MO’s (HOMO and HOMO-1) were executed to describe the orientation of elec-

247

trophilic aromatic substitution and reveal the principle under the reactive site for the reaction be-

248

tween flavonoids and HClO.

249

The HOMO electron densities of apigenin are spread mainly over C(3), C(6) and C(8) in the A-

250

ring, which means that these sites may be the reactive site for the electrophilic attack, because the

251

electron donating ability is positively related with the HOMO energy level and the higher HOMO

252

energy means the stronger electron donation ability. The high density of negative charge distribu-

253

tion in the C(3) accounts for its high reactivity with HClO. As for quercetin, the OH group in the 12 ACS Paragon Plus Environment

Page 13 of 32

Journal of Agricultural and Food Chemistry

254

C(3) is not a good leaving group and thus prevent chlorination happening at C(3) site. Instead, C(8)

255

and C(6) in the A-ring of the flavonol were the preferred chlorination sites, which is consistent

256

with experimental results.

257

The HOMO electron density distribution localizes dominantly on the C(6) and C(8) sites of the

258

A-ring, and the low electron density on the C(2)-C(3) may due to the missing of double bond in

259

the pyrone moiety. This hinders electron delocalization between B ring and A ring.39 Therefore,

260

C(6) and C(8) at the A-ring are the only possible functional sites for the chlorination.

261

From the frontier molecular orbitals (FMO) compositions of ampelopsin and epicatechin, it can

262

be observed that the electron density for HOMO is distributed more in B-ring compared with A

263

and C-rings. But the HOMO-1 orbital of epicatechin and the HOMO-2 orbital of ampelopsin were

264

characterized by a charge distribution mainly on A-ring. This indicated that the C(6) and C(8) sites

265

on the A-ring still have potential to be the electrophilic reaction sites.

266

Fukui indices, the derivatives of the electron density of a molecule with respect to the number of

267

electrons,40 can be employed as indicators of chemical reactivity to evaluate the regions of the

268

molecule to undergo nucleophilic, electrophilic and radical attack.41 The study of the charge distri-

269

bution was carried out by considering two types of charge analysis, natural population analysis

270

(NPA) and electrostatic potential analysis (ESP). The analysis of Fukui indices of flavonoids was

271

conducted under the same level B3LYP/6-311+G(d,p), the Fukui indices result was summarized in

272

Table S2 and principal chemical reactivity sites was showed in Figure 9. According to the Fukui

273

indices, the electrophilic attack in apigenin can be presented in the C(3) atom, the preferred site

274

for electrophilic attack in quercetin, naringenin, ampelopsin and epicatechin molecule was the C(6)

275

and C(8) of the ring A. The conclusions are also linear with DFT calculation and experimental

276

results. 13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 32

277

Based on our results of HClO scavenging capacity and the structure of flavonoids, it becomes

278

apparent that the hydroxyl groups in the carbon skeleton of flavonoids can enhance the reactivity

279

by donating the electron and increasing the electron density of the chlorination sites. In addition,

280

the double bond in the C(2) and C(3) amplifies the electron-withdrawing effect of ketone in the C-

281

ring, resulting in decreasing HClO scavenging capacity. Meanwhile, under conditions of a large

282

molar excess of hypochlorite (>10-fold molar excess of HClO), oxidation of flavonoid may also oc-

283

cur in addition to chlorination on B ring. 25, 42, 43

284

Flavonoids represent major group of polyphenolic antioxidants, which are known radical scaven-

285

gers due to their possession of multiple phenolic groups capable of quenching peroxyl radicals in

286

radical chain reactions.44, 45 In contrast, our results are in agreement with previous work,24, 25 which

287

suggest that the flavonoids may react other biologically relevant reactive oxygen species, such as

288

HClO, through the aromatic carbons. It is therefore important to elucidate the reaction products

289

and the mechanisms between different ROS and flavonoids in order to better understand the struc-

290

ture and activity relationship of polyphenolic antioxidants.

291

CONCLUSIONS

292

In summary, our results have illustrated that hypochlorous acid was quenched by wide range of

293

flavonoids through electrophilic reaction with the flavonoids, particularly at A ring, in which the

294

hydroxyl groups can enhance the scavenging capacity of the flavonoids. From our results, fla-

295

vanones, particularly eriodictyol (with IC50 of 1.2 µM) have high scavenging activity. The double

296

bond in the C ring has negative impact to the scavenging capacity because it helps to channel

297

electron towards electron-withdrawing ketone group on C(4). It is unintuitive that the chlorination

298

does not happen on B rings, which are the reaction sites for radicals (such as peroxyl radical). As 14 ACS Paragon Plus Environment

Page 15 of 32

Journal of Agricultural and Food Chemistry

299

Prior has observed that ROS scavenging activity of fruits and vegetables are dependent on the spe-

300

cific ROS and recommended to use multiple ROS scavenging capacity plot (spider web plot) to

301

comprehensively characterize the antioxidant capacity of foods.46 Therefore, it is imperative to es-

302

tablish the structure and reactivity relationship between flavonoids and individual biologically rel-

303

evant reactive oxygen species in order to understand the chemical reasons of the different ROS

304

scavenging capacity of fruits and vegetables and their health promotion property.

305

ASSOCIATED CONTENT

306

Supporting Information

307

Additional information as noted in the text (Figures S1−S15) is available free of charge via the Inter-

308

net at http://pubs.acs.org.

309

Corresponding Author

310

*Telephone: 65-6516-8821. Fax: 65-6775-7895. E-mail: [email protected].

311

ACKNOWLEDGMENT

312

Authors thank Singapore Ministry of Education for financial support (grant no: MOE2014-T2-1-

313

134).

314

ABBREVIATIONS

315

ROS, reactive oxygen species; RNS, reactive nitrogen species; DFT, density functional theory; MPO,

316

myeloperoxidase; SAR, structure-activity relationship; FMO, frontier molecular orbitals.

317

REFERENCES

318

1.

319

Houstis, N.; Rosen, E. D.; Lander, E. S., Reactive oxygen species have a causal role in multiple forms of insulin resistance. Nature 2006, 440, 944. 15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

320

2.

321

Page 16 of 32

Halliwell, B., Free radicals, antioxidants, and human disease: curiosity, cause, or consequence? The lancet 1994, 344, 721-724.

322

3.

Harman, D., Free radical theory of aging: the “free radical” diseases. Age 1984, 7, 111-131.

323

4.

Hensley, K.; Carney, J.; Mattson, M.; Aksenova, M.; Harris, M.; Wu, J.; Floyd, R.; Butterfield,

324

D., A model for beta-amyloid aggregation and neurotoxicity based on free radical generation by

325

the peptide: relevance to Alzheimer disease. Proceedings of the National Academy of Sciences

326

1994, 91, 3270-3274.

327

5.

328 329

free radical production. Pediatr. Res. 1988, 23, 143. 6.

330 331

7.

8.

Totter, J. R., Spontaneous cancer and its possible relationship to oxygen metabolism. Proceedings of the National Academy of Sciences 1980, 77, 1763-1767.

9.

336 337

Olanow, C., An introduction to the free radical hypothesis in Parkinson's disease. Ann. Neurol. 1992, 32.

334 335

Slater, T. F., Free radical mechanisms in tissue injury. In Cell Function and Disease, Springer: 1988; pp 209-218.

332 333

Saugstad, O. D., Hypoxanthine as an indicator of hypoxia: its role in health and disease through

Ames, B. N., Endogenous oxidative DNA damage, aging, and cancer. Free Radic. Res. Commun. 1989, 7, 121-128.

10.

Ray, G.; Batra, S.; Shukla, N. K.; Deo, S.; Raina, V.; Ashok, S.; Husain, S. A., Lipid

338

peroxidation, free radical production and antioxidant status in breast cancer. Breast Cancer Res.

339

Treat. 2000, 59, 163-170.

340

11.

Harman, D., Free radical theory of aging. Mutation Research/DNAging 1992, 275, 257-266.

341

12.

Beckman, K. B.; Ames, B. N., The free radical theory of aging matures. Physiol. Rev. 1998, 78,

342 343

547-581. 13.

Harraan, D., Aging: a theory based on free radical and radiation chemistry. 1955. 16 ACS Paragon Plus Environment

Page 17 of 32

344

14.

345 346

Morris, J. C., The acid ionization constant of HClO from 5 to 35. The Journal of Physical Chemistry 1966, 70, 3798-3805.

15.

347 348

Journal of Agricultural and Food Chemistry

Kettle, A.; Winterbourn, C., Myeloperoxidase: a key regulator of neutrophil oxidant production. Redox Report 1997, 3, 3-15.

16.

Van der Veen, B. S.; de Winther, M. P.; Heeringa, P., Myeloperoxidase: molecular mechanisms

349

of action and their relevance to human health and disease. Antioxidants & redox signaling 2009,

350

11, 2899-2937.

351

17.

352 353

Hawkins, C.; Pattison, D.; Davies, M. J., Hypochlorite-induced oxidation of amino acids, peptides and proteins. Amino Acids 2003, 25, 259-274.

18.

Hawkins, C. L.; Davies, M. J., Hypochlorite-induced damage to proteins: formation of nitrogen-

354

centred radicals from lysine residues and their role in protein fragmentation. Biochem. J 1998,

355

332, 617.

356

19.

357 358

Patel, K.; Mehta, H.; Srivastava, H., Kinetics and mechanism of oxidation of starch with sodium hypochlorite. J. Appl. Polym. Sci. 1974, 18, 389-399.

20.

Hazell, L.; Van den Berg, J.; Stocker, R., Oxidation of low-density lipoprotein by hypochlorite

359

causes aggregation that is mediated by modification of lysine residues rather than lipid

360

oxidation. Biochem. J 1994, 302, 297.

361

21.

Spencer, J. P.; Whiteman, M.; Jenner, A.; Halliwell, B., Nitrite-induced deamination and

362

hypochlorite-induced oxidation of DNA in intact human respiratory tract epithelial cells. Free

363

Radical Biol. Med. 2000, 28, 1039-1050.

364

22.

365 366 367

Cao, G.; Sofic, E.; Prior, R. L., Antioxidant and prooxidant behavior of flavonoids: structureactivity relationships. Free Radical Biol. Med. 1997, 22, 749-760.

23.

Santos, M. R.; Mira, L., Protection by flavonoids against the peroxynitrite-mediated oxidation of dihydrorhodamine. Free Radical Res. 2004, 38, 1011-1018. 17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

368

24.

Binsack, R.; Boersma, B. J.; Patel, R. P.; Kirk, M.; White, C. R.; Darley‐Usmar, V.; Barnes, S.;

369

Zhou, F.; Parks, D. A., Enhanced antioxidant activity after chlorination of quercetin by

370

hypochlorous acid. Alcoholism: Clinical and Experimental Research 2001, 25, 434-443.

371

25.

372 373

Boersma, B. J.; Patel, R. P.; Kirk, M.; Jackson, P. L.; Muccio, D.; Darley-Usmar, V. M.; Barnes, S., Chlorination and nitration of soy isoflavones. Arch. Biochem. Biophys. 1999, 368, 265-275.

26.

Payá, M.; Halliwell, B.; Hoult, J., Interactions of a series of coumarins with reactive oxygen

374

species: scavenging of superoxide, hypochlorous acid and hydroxyl radicals. Biochem.

375

Pharmacol. 1992, 44, 205-214.

376

Page 18 of 32

27.

Gomes, A.; Fernandes, E.; Silva, A. M.; Santos, C. M.; Pinto, D. C.; Cavaleiro, J. A.; Lima, J.

377

L., 2-Styrylchromones: novel strong scavengers of reactive oxygen and nitrogen species. Biorg.

378

Med. Chem. 2007, 15, 6027-6036.

379

28.

380 381

Robaszkiewicz, A.; Bartosz, G., Estimation of antioxidant capacity against peroxynitrite and hypochlorite with fluorescein. Talanta 2010, 80, 2196-2198.

29.

Frisch, M.; Trucks, G.; Schlegel, H.; Scuseria, G.; Robb, M.; Cheeseman, J.; Scalmani, G.;

382

Barone, V.; Mennucci, B.; Petersson, G., Gaussian 09, revision D. 01. In Gaussian, Inc.,

383

Wallingford CT: 2009.

384

30.

385 386

behavior. Physical Review A 1988, 38, 3098-3100. 31.

387 388

Becke, A. D., Density-functional exchange-energy approximation with correct asymptotic

Lee, C.; Yang, W.; Parr, R. G., Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Physical Review B 1988, 37, 785-789.

32.

Vosko, S. H.; Wilk, L.; Nusair, M., Accurate spin-dependent electron liquid correlation energies

389

for local spin density calculations: a critical analysis. Canadian Journal of physics 1980, 58,

390

1200-1211.

391

33.

Dennington, R.; Keith, T.; Millam, J., GaussView, version 5. 2009. 18 ACS Paragon Plus Environment

Page 19 of 32

392

34.

393 394

35.

36.

37.

Zhou, Z.; Parr, R. G., Activation hardness: new index for describing the orientation of electrophilic aromatic substitution. J. Am. Chem. Soc. 1990, 112, 5720-5724.

38.

401 402

Rice-Evans, C. A.; Miller, N. J.; Paganga, G., Structure-antioxidant activity relationships of flavonoids and phenolic acids. Free Radical Biol. Med. 1996, 20, 933-956.

399 400

Leo, A.; Hoekman, D., Exploring QSAR.:. Fundamentals and applications in chemistry and biology. An American Chemical Society Publication: 1995; Vol. 1.

397 398

Yang, W.; Mortier, W. J., The use of global and local molecular parameters for the analysis of the gas-phase basicity of amines. J. Am. Chem. Soc. 1986, 108, 5708-5711.

395 396

Journal of Agricultural and Food Chemistry

Karelson, M.; Lobanov, V. S.; Katritzky, A. R., Quantum-chemical descriptors in QSAR/QSPR studies. Chem. Rev. 1996, 96, 1027-1044.

39.

Trouillas, P.; Marsal, P.; Siri, D.; Lazzaroni, R.; Duroux, J.-L., A DFT study of the reactivity of

403

OH groups in quercetin and taxifolin antioxidants: the specificity of the 3-OH site. Food Chem.

404

2006, 97, 679-688.

405

40.

406 407

reactivity. J. Am. Chem. Soc. 1984, 106, 4049-4050. 41.

408 409

Ayers, P. W.; Parr, R. G., Variational principles for describing chemical reactions: the Fukui function and chemical hardness revisited. J. Am. Chem. Soc. 2000, 122, 2010-2018.

42.

410 411

Parr, R. G.; Yang, W., Density functional approach to the frontier-electron theory of chemical

Krych-Madej, J.; Stawowska, K.; Gebicka, L., Oxidation of flavonoids by hypochlorous acid: reaction kinetics and antioxidant activity studies. Free Radical Res. 2016, 50, 898-908.

43.

Pannala, A. S.; Chan, T. S.; O'Brien, P. J.; Rice-Evans, C. A., Flavonoid B-ring chemistry and

412

antioxidant activity: fast reaction kinetics. Biochem. Biophys. Res. Commun. 2001, 282, 1161-

413

1168.

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

414

44.

Matsuo, M.; Matsumoto, S.; Iitaka, Y.; Niki, E., Radical-scavenging reactions of vitamin E and

415

its model compound, 2, 2, 5, 7, 8-pentamethylchroman-6-ol, in a tert-butylperoxyl radical

416

generating system. J. Am. Chem. Soc. 1989, 111, 7179-7185.

417

45.

Wright, J. S.; Johnson, E. R.; DiLabio, G. A., Predicting the activity of phenolic antioxidants:

418

theoretical method, analysis of substituent effects, and application to major families of

419

antioxidants. J. Am. Chem. Soc. 2001, 123, 1173-1183.

420 421

46.

Page 20 of 32

Prior, R. L., Oxygen radical absorbance capacity (ORAC): New horizons in relating dietary antioxidants/bioactives and health benefits. J. Funct. Foods 2015, 18, 797-810.

422

20 ACS Paragon Plus Environment

Page 21 of 32

Journal of Agricultural and Food Chemistry

Figure 1. Hypochlorous acid scavenging capacity (1/IC50) of phenol, catechol, pyrogallol, resorcinol, phloroglucinol.

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 2. The electronic influence (Hammett σ) of substitution of an OH group in an aromatic ring. The more negative the σ is the greater the electron donating effect. The effect at the ortho and para position is electron donating and at the meta position electron withdrawing.

22 ACS Paragon Plus Environment

Page 22 of 32

Page 23 of 32

Journal of Agricultural and Food Chemistry

Scheme 1. The proposed mechanism (nucleophilic substitution reaction) for the reaction between phenol and hypochlorous acid.

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 32

Figure 3. In impact of OH group at B ring on the hypochlorous acid scavenging capacity (IC50) of flavonoids.

24 ACS Paragon Plus Environment

Page 25 of 32

Journal of Agricultural and Food Chemistry

Figure 4. Impact of A-ring OH group on the hypochlorous acid scavenging capacity (IC50) of flavonoids.

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 32

Figure 5. The hypochlorous acid scavenging capacity (IC50) of flavonoids from the different subclasses: flavones, flavonols, flavanones, and flavanonols.

26 ACS Paragon Plus Environment

Page 27 of 32

Journal of Agricultural and Food Chemistry

Figure 6. HPLC and LC-MS analysis of reaction products formed between flavonoids and HClO. All flavonoids (150 μM) was reacted with HClO (50 μM), and the reaction products were separated by reverse-phase HPLC. Analysis of the reaction products of flavonoids with HClO. (A) Analysis of the reaction products of apigenin with HClO. (B) Analysis of the reaction products of quercetin with HClO. (C) Analysis of the reaction products of naringenin with HClO. (D) Analysis of the reaction products of ampelopsin with HClO. AP, apigenin; QU, quercetin; NA, naringenin; AM, ampelopsin.

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 7. Analysis of the reaction products of epicatechin with HClO.

28 ACS Paragon Plus Environment

Page 28 of 32

Page 29 of 32

Journal of Agricultural and Food Chemistry

Scheme 2. The proposed mechanism (electrophilic aromatic substitution) for the reaction between flavonoids and hypochlorous acid. A is the reaction between flavones and HClO, B is the reaction between flavonols, monochlorinated flavones and HClO, and C is the reaction between flavanones and HClO.

29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 32

Figure 8. Charge distribution of FMOs (HOMO and HOMO-1) in the optimized molecule of flavonoids: (a) apigenin, (b) quercetin, (c) naringenin, (d) ampelopsin and (e) epicatechin, calculated under the B3LYP/6-311+G(d,p) level.

30 ACS Paragon Plus Environment

Page 31 of 32

Journal of Agricultural and Food Chemistry

Figure 9. Preferred sites for flavonoids in electrophilic reaction with HClO according to Fukui induces: (a) apigenin, (b) quercetin, (c) naringenin, (d) ampelopsin and (e) epicatechin.

31 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Table of Contents (TOC)

32 ACS Paragon Plus Environment

Page 32 of 32