Subscriber access provided by GAZI UNIV

Article

Investigation of acyl migration in mono- and di-caffeoylquinic acids under aqueous basic, aqueous acidic and dry roasting conditions Sagar Deshpande, Rakesh Jaiswal, Marius Febi Matei, and Nikolai Kuhnert J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/jf5017384 • Publication Date (Web): 12 Aug 2014 Downloaded from http://pubs.acs.org on August 21, 2014

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

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

Page 1 of 43

Journal of Agricultural and Food Chemistry

1

Investigation of acyl migration in mono- and di-caffeoylquinic acids under aqueous

2

basic, aqueous acidic and dry roasting conditions

3 4

Sagar Deshpande, Rakesh Jaiswal, Marius Febi Matei and Nikolai Kuhnert*

5 6

School of Engineering and Science, Chemistry, Jacobs University Bremen, 28759 Bremen,

7

Germany

8 9 10 11 12

*Author to whom correspondence should be addressed Tel: 49 421 200 3120; Fax: 49 421

13

200 3229; E-mail: [email protected]

14 15 16 17 18 19 20 1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

21

Abstract

22

Acyl migration in chlorogenic acids describes the process of migration of cinnamoyl moieties

23

from one quinic acid alcohol group to another thus interconverting chlorogenic acid

24

regioisomers. It therefore constitutes a special case of a transesterification reaction. Acyl

25

migration constitutes an important reaction pathway in both coffee roasting and brewing

26

altering the structure of chlorogenic acid initially present in the green coffee bean. In this

27

contribution we describe detailed and comprehensive mechanistic studies comparing inter-

28

and intra-molecular acyl migration involving the seven most common chlorogenic acids in

29

coffee. We employ aqueous acidic and basic conditions mimicking the brewing of coffee

30

along with dry roasting conditions. We show that under aqueous basic conditions

31

intramolecular acyl migration is fully reversible with basic hydrolysis competing with acyl

32

migration. 3-caffeoylquinic acid was shown to be most labile to basic hydrolysis. We

33

additionally show that the acyl migration process is strongly pH dependant with increased

34

transesterification taking place at basic pH. Under dry roasting conditions acyl migration

35

competes with dehydration to form lactones. We argue that acyl migration precedes

36

lactonisation with 3-caffeoylquinic acid lactone being the predominant product.

37 38 39 40 41 42 43 2 ACS Paragon Plus Environment

Page 2 of 43

Page 3 of 43

Journal of Agricultural and Food Chemistry

44 45

46

47

Introduction

48

Coffee is one of the most valued agricultural commodities in terms of the economic aspects of

49

the exports from the developing coffee producing countries, accounting to ca. 8 million metric

50

tonnes per year. Approximately, 2.3 billion cups of coffee are consumed worldwide per day.1,

51

2

52

coffee) are the two types of coffee holding 70% and 30%, respectively of the total coffee

53

market in the world.3 Chlorogenic acids are present in the range of 6-12% of the dry weight of

54

the green coffee bean and constitute the most abundant class of secondary metabolites.4

55

Chlorogenic acids (CGAs) are a large group of esters formed between one or more cinnamic

56

acid derivatives and D-(-)-quinic acid. CGAs are classified on the basis of the number of the

57

cinnamoyl residues esterified with the quinic acid as well as the functional groups present on

58

the aromatic moiety of the cinnamoyl residues. Out of the total content of the CGAs in green

59

coffee, 5-O-caffeoylquinic acid (3) comprises about 50%. Other subclasses like

60

caffeoylquinic acids, dicaffeoylquinic acids, feruloylquinic acids and p-coumaroylquinic acids

61

contribute to a large extent to the other 50% of the total CGAs present in coffee. In total 45

62

CGAs have been reported in Arabica and 85 in Robusta green coffee beans.1, 2 CGAs are very

63

important plant secondary metabolites due to their pharmacological properties, such as

64

antioxidant property5, anti-hepatitis B virus activity6, antispasmodic activity4, anti-diabetic

65

activity7, inhibition of the HIV-1 integrase8,

66

carcinogenic compounds.4

Coffea arabica (known as Arabica coffee) and Coffea canephora (known as Robusta

9

and inhibition of the mutagenicity of

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 43

67

The various roasting conditions affect drastically the concentration and composition of CGAs

68

in coffee. The chemistry at elevated temperatures of CGAs has been described in detail with

69

epimerisation, dehydration and transesterifications (acyl migration a special case) dominating

70

the mechanistic spectrum.2 The evidence for transesterification phenomena was demonstrated

71

indirectly by the presence of a series of 1-substitited CGA derivatives in roasted coffee, which

72

are absent in the green bean. For every 1% of the dry matter of the total CGA content in the

73

green coffee beans, 8-10% of the original CGAs are transformed or decomposed into

74

respective cinnamic acid derivatives and quinic acid.1,2 The lightest drinkable roast (the so-

75

called ‘Cinnamon’ roast) involves roasting of green coffee beans at around 180 °C until the

76

coffee beans just encounter the ‘first crack’. It was reported by Clifford et al. 1 that during the

77

early stage of the roasting process, transformations such as isomerization (acyl migration) or

78

hydrolysis of the ester bond, take place in the CGAs. Later, chemical transformations like

79

decarboxylation of cinnamoyl moieties to produce a number of phenylindans, epimerization at

80

the quinic acid and dehydration to produce cyclohexenes and lactones take place.2, 3

81

Clifford et al.

82

and transesterification in 5-O-caffeoylquinic acid (3), 3-O-caffeoylquinic acid (2), 4-O-

83

caffeoylquinic acid (4), 3,4-di-O-caffeoylqunic acid (8), 3,5-di-O-caffeoylqunic acid (9), 4,5-

84

di-O-caffeoylqunic acid (10), 5-O-p-coumaroylquinic acid (13) and 5-O-feruloylqunic acid

85

(23) in which, the identification of some of the transformed products was based on the

86

putative conclusions acquired by analytical HPLC. Dawidowicz et al.12

87

transformation products of 5-O-caffeoylquinic acid (3) after five hours of reflux in an acid-

88

water solution including two water addition products. This study only incorporated 5-CQA

89

and the brewing time was not in line with common consumer practice. No comprehensive

90

mechanistic study has been previously reported which investigated the intra- versus inter-

91

molecular acyl migration under different conditions incorporating all major commercially

10, 11

also reported the basic hydrolysis induced intramolecular isomerization

4 ACS Paragon Plus Environment

found nine

Page 5 of 43

Journal of Agricultural and Food Chemistry

92

available regio-isomers of mono- and di-caffeoyl chlorogenic acids. Most importantly data

93

on acyl migration under roasting conditions are absent from the literature.

94

The complexity in the data interpretation for the structural analyses of the CGAs present in

95

the roasted coffee melanoidins arises from the regio- and stereoisomeric compounds in the

96

natural sources. For this reason, model roasting experiments on the commercially available

97

mono- and dicaffeoylquinic acids were attempted in the present work in order to study the

98

transformations taking place in CGAs during the early roasting stages. Also, isomerization

99

(acyl migration) was induced by both basic hydrolysis and simple hydrolysis (brewing) in

100

these reference standards to observe the isomeric transformations on the basis of relative

101

quantitation. Recently, tandem mass spectrometry has allowed accurate structural assignment

102

and identification of the CGA regioisomers. The advantage of multi-dimensional specificity

103

of LC-MSn enabled isomeric resolution and relative quantitation of the early roasting

104

transformations in the CGAs.13

105

Materials and Methods

106

Chemicals and materials

107

All the chemicals (analytical grade) were purchased from Sigma-Aldrich (Bremen, Germany).

108

Commercially available mono- and di-caffeoylquinic acids such as 5-O-caffeoylquinic acid

109

(3), 3-O-caffeoylquinic acid (neo-chlorogenic acid) (2), 4-O-caffeoylquinic acid (crypto-

110

chlorogenic acid) (4), 1,3-di-O-caffeoylquinic acid (cynarin) (5), 3,4-di-O-caffeoylquinic acid

111

(8), 3,5-di-O-caffeoylquinic acid (9), 4,5-di-O-caffeoylquinic acid (10) were purchased from

112

PhytoLab GmbH & Co. KG (Vestenbergsgreuth, Germany).

113

Hydrolysis by tetramethylammonium hydroxide (TMAH)

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

114

All the seven CGAs reference standards were treated with aqueous TMAH (25 g/L). Each

115

sample was diluted by 5 mL of aqueous TMAH and stirred at room temperature. pH of the

116

mixture was observed to be around 12. 1 mL solution from each sample was taken out at 2, 5,

117

10, 30 and 60 min time intervals. Each sample was saturated with brine and extracted twice

118

with ethyl acetate. Combined organic layers were concentrated in vacuo and each sample was

119

prepared in 1 mL of methanol for analysis by LC-MSn for intramolecular acyl migration.

120

To study the intermolecular acyl migration (cross-over experiment), 5-CQA (25 mg, 0.07062

121

mmol) was added to a round bottom flask containing ferulic acid (13.7 mg, 0.07062 mmol). 5

122

mL of 10 times diluted (25 g/litre) TMAH was added to the flask and the mixture was stirred

123

at room temperature. 1 mL samples were taken out from the flask at 2, 5, 10, 15 and 30 min

124

time intervals. Each sample was saturated with brine and extracted twice with ethyl acetate.

125

The combined organic layers were concentrated in vacuo and each sample was prepared in

126

methanol to be analysed by LC-MSn. The same procedure was repeated with p-coumaric acid

127

(11.57 mg, 0.07062 mmol) and 5-CQA (25 mg, 0.07062 mmol).

128

Model roasting

129

All the seven CGAs reference standards were heated at 180 °C for 12 min separately to study

130

the intramolecular acyl migration. Equimolar quantities of 5-CQA with p-coumaric acid and

131

5-CQA with ferulic acid were heated together at 180 °C for 12 min to study the intermolecular

132

acyl migration (cross-over experiment). All the samples were heated in a Buechi Glass Oven

133

B-585 and prepared in 1 mL methanol for LC-MSn analysis.

134

Brewing of CGAs (2-5 and 8-10)

135

Commercially available chlorogenic acids standards (each sample 500 µg) were infused in 3

136

mL of hot water each and stirred for 5 h under reflux. pH of each sample was determined to

6 ACS Paragon Plus Environment

Page 6 of 43

Page 7 of 43

Journal of Agricultural and Food Chemistry

137

be 5 with pH meter. The solvent was removed under low pressure and the samples were

138

dissolved in 1mL MeOH and used for LC-MSn.

139

LC-MSn

140

The 1100 series LC equipment (Agilent, Bremen, Germany) comprised a binary pump, an

141

auto sampler with a 100 µL loop, and a DAD detector with a light-pipe flow cell (recording at

142

254 and 320 nm and scanning from 200 to 600 nm). This was interfaced with an ion-trap mass

143

spectrometer fitted with an HCT Ultra ESI source (Bruker Daltonics, Bremen, Germany)

144

operating in full scan, auto MSn mode to obtain fragment ion m/z. Tandem mass spectra were

145

acquired in Auto-MSn mode (smart fragmentation) using a ramping of the collision energy.

146

Maximum fragmentation amplitude was set to 1 V, starting at 30% and ending at 200%. The

147

MS operating conditions (negative mode) had been optimized using 5-caffeoylquinic acid

148

with a capillary temperature of 365 oC, a dry gas flow rate of 10 L/min and a nebulizer

149

pressure of 10 psi.

150

HPLC

151

Separation was achieved on a 150 x 3 mm i.d., 5 µm diphenyl column, with a 5 mm x 3 mm

152

i.d. guard column (Varian, Darmstadt, Germany). Alternatively, separation was also achieved

153

on a 250 mm x 3 mm i.d., 5 µm C18-amide column, with a 5 mm x 3 mm i.d. guard column

154

of the same material (Varian, Darmstadt, Germany) for the cases of hydrolysis (brewing) of

155

reference standards experiments. Solvent A was water/formic acid (1000:0.05, v/v) and

156

solvent B was methanol. Solvents were delivered at a total flow rate of 500 µL/min. The

157

gradient profile was from 10 % B to 70 % B linearly in 60 min followed by 10 min isocratic,

158

and a return to 10 % B at 90 min and 10 min isocratic to re-equilibrate.

159

Preliminary assessment of data

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

160

All the data for the chlorogenic acids use the recommended IUPAC numbering system;14 the

161

same numbering system was adopted for chlorogenic acids, their cis-isomers, their acyl-

162

migration isomers and water addition products (Figure 1). The relative concentrations of the

163

transformed products are expressed here in terms of the peak areas obtained in their UV

164

chromatograms assuming the relative response factor in UV close to one, based on identical

165

absorptivity of all mono CGA.13,

166

accordingly.

167

Results and Discussion

168

Intramolecular acyl migration: hydrolysis by TMAH of 2-5 and 8-10

169

The most common CGAs in the green coffee bean 2-5 and 8-10 were selected and subjected

170

to aqueous TMAH, with samples taken at intervals from 2-60 minutes. Samples were directly

171

analyzed by LC-MS with compound identification and assignment achieved through

172

comparison of retention times and fragment spectra.2,

173

was carried out using HPLC-UV traces monitoring cinnamoyl absorption at 320 nm. The

174

concentration of the samples was chosen to be sufficiently low (1-1.5 mg/ mL) to prevent

175

intermolecular acyl migration or transesterification, yet corresponding to typical coffee brew

176

CGA levels (around 200 mg/cup = 1 mg/mL) therefore allowing observation of intra

177

molecular acyl migration exclusively. In the hydrolysis of 3-CQA (2), 5-CQA (3) and 4-CQA

178

(4), all the other mono-acyl derivatives were identified during the hydrolysis except for 1-

179

CQA (1). Also, we did not observe the formation of any di-acyl derivatives in the hydrolysis

180

of the mono-acyl quinic acids. This observation confirms that the acyl migration we observed

181

in this study was in fact an intramolecular process. Figure 2 represents the UV

182

chromatograms of the 5-CQA in basic solution at different time intervals. It should be noted

183

that the UV response in mono- and di-acylquinic acids has been used in the past as a reliable

15

In tables and figures, the peak area values are stated

18-23

Relative or absolute quantitation

8 ACS Paragon Plus Environment

Page 8 of 43

Page 9 of 43

Journal of Agricultural and Food Chemistry

184

relative response factor.1 Hanson et al.16 used radiolabelled quinic acid to investigate the acyl

185

migration pathway in cinnamoylqunic acids; in accordance with these findings, we assumed

186

that the mechanism of the acyl migration follows the 1,2-ortho ester intermediate formation.

187

This assumption is also supported by the study of Xie et al.17 The transformations of 5-CQA

188

(3), 4-CQA (4) and 3-CQA (2) with time are presented in Figure 3. From the results, we can

189

conclude that 5-CQA is much more stable than 4-CQA and 3-CQA and the order of the

190

stability is 5>4>3 in terms of the hydrolysis of the caffeoyl ester. This stability was observed

191

to provide the resistance to decomposition thus allowing 5-CQA to form the acyl migrated

192

products over longer hydrolysis durations. On the other hand, with 3-CQA being the least

193

stable of the three mono-acylquinic acids, it decomposes to form caffeic acid and quinic acid

194

even before acyl migration takes place.

195

Considering the mechanism of the acyl migration proceeding through an 1,2-ortho-ester

196

intermediate formation as 5-CQA ⇋ 4-CQA ⇋ 3-CQA, the reverse equilibrium, through

197

which the migrated acyl group reverts back to its original position appears to be slower. It is

198

difficult to comment on the thermodynamic equilibrium of the acyl migration because of the

199

ongoing simultaneous ester hydrolysis competing with the acyl migration process.

200

The equilibrium between 3-CQA (2) and 4-CQA (4) is readily achieved because the ortho-

201

ester intermediate is more stable due to the cis geometry (Figure 4). Compound 2 shows a

202

tendency to hydrolyse to generate caffeic acid and quinic acid rather than to undergo acyl

203

migration presumably due to the 1,3-syn-diaxial arrangement between the C1 hydroxyl group

204

and the C3 ester. The hydrogen bonding between C1-OH and the carbonyl oxygen on the ester

205

on C3 results in steric hindrance preventing the nucleophilic attack on the carbonyl carbon by

206

the C4 hydroxyl group and facilitates the hydrolysis of the ester bond thus dissociating the

207

caffeoyl moiety in basic conditions. At the same time this hydrogen bonding presumably

208

activates the ester at C3 for hydrolytic cleavage. 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 43

209

Similar to monocaffeoyl derivatives, dicaffeoyl quinic acids were subjected to aqueous

210

TMAH treatment and analysed in the same manner. Previous studies have shown that 1,5-

211

diCQA (7) was converted into 1,3-diCQA (5) and 5-CQA (3) rapidly and extensively by

212

TMAH treatment within one minute of hydrolysis.18, 19 In the present study however, basic

213

hydrolysis of 1,3-diCQA (5) did not show any presence of 1,5-diCQA (7) or 1,4-diCQA (6).

214

1,3-diCQA (5) decomposed largely into 1-CQA (1) rather than 3-CQA (2) in the ratio 2.2:1

215

after two min of base treatment. 1,3-diCQA (5) did not transform preferably into 3,5-diCQA

216

(9) whereas 3,4-diCQA (8) and 4,5-diCQA (10) were formed in very small quantities from (5)

217

(Figure 5). 1-CQA (1) was found to be present entirely as a decomposition product of 1,5-

218

diCQA (7) or 1,4-diCQA or (6) 1,3-diCQA (5) rather than a product of acyl migration as it

219

was not observed as an acyl migration product of any other mono- or di-acylated substrates;

220

this fact also supports the ortho-ester propagation of acyl migration process. Moreover, this

221

observation confirms the hydrolytic lability of the esters in the 3-acylated position.

222

In the case of hydrolysis of 4,5-diCQA (10), we observed that 4,5-diCQA transformed mainly

223

into 3,4-diCQA (8) after two min of base treatment. Additionally CQAs, in particular 5-CQA

224

(3), 4-CQA (4) and 3-CQA (2) were observed after two min reaction time. This trend

225

continued for five minutes during the base treatment, after which all of the CGA derivatives

226

were completely decomposed. 3,4-diCQA (8) was observed to be the least stable isomer

227

during the base treatment study. At two minutes, 8 was found to be in equilibrium with 3,5-

228

diCQA (9) however after five minutes of treatment, 3,5-diCQA (9) was observed to display a

229

slightly larger peak area than 3,4-diCQA (8). 3-CQA (2) was not observed in any sample

230

throughout the duration of the base treatment of 3,4-diCQA due to its tendency to decompose

231

rapidly by hydrolysis.

232

Observations based on the results from the base treatment on 3,5-diCQA (9) were quite

233

distinctive. After five minutes 3,5-diCQA (9) was observed to possess surprising stability to 10 ACS Paragon Plus Environment

Page 11 of 43

Journal of Agricultural and Food Chemistry

234

the hydrolysis of the ester as we did not detect any of the mono-CQA derivatives even after

235

60 min of base treatment (Figure 5). 3,5-diCQA (9) transformed mainly into 3,4-diCQA (8)

236

followed by 4,5-diCQA (10). The ratio of 9:8:10 remained approximately constant throughout

237

the 60 min of base treatment.

238

Intermolecular acyl migration (Transesterification): hydrolysis by TMAH (cross-over

239

experiment)

240

In these experiments, we investigated intermolecular acyl migration by carrying out cross-

241

over experiments, in which 5-CQA was reacted with the free acids like ferulic and p-coumaric

242

acids at 1:1 stoichiometry in presence of a base. The change of substituents hereby allows

243

easy identification of products by LC-MS based on the m/z value of the pseudomolecular ions

244

of the products. The intermolecular acyl migration was found to be simultaneously competing

245

with intramolecular acyl migration as well as hydrolysis of the CQA. The products identified

246

in the reaction of 5-CQA with ferulic acid and p-coumaric acid at 2, 5, 10, 15 and 30 min of

247

cross-over experiment are summarized in Table 1. Along with the transesterification products

248

of 5-CQA and respective free acid, formation of the cis-cinnamoyl isomers was also observed

249

in ferulic, caffeic and p-coumaric acids. All products were identified according to the

250

fragmentation schemes reported previously.18-21

251

In the case of the base treatment of an equimolar mixture of p-coumaric acid and 5-CQA (3),

252

the intramolecular acyl migration within 5-CQA seemed to be dominating the hydrolysis and

253

the intermolecular acyl migration. According to the peak areas observed, the formation of the

254

intermolecular acyl migrated species (transesters) was found to be least favoured (Figure 6).

255

For example, 3-CQA (2), 5-CQA (3) and 4-CQA (4) formed predominantly over caffeic acid

256

(32) during two minutes of TMAH treatment of 5-CQA. p-Coumaric acid (33) did not esterify

257

with the quinic acid generated from the hydrolysis of the 5-CQA (3) after two minutes hence,

258

no p-coumaroylquinic acids were observed. However, after 5 minutes of base treatment the 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 43

259

first transesterification product appeared in the form of 4-pCoQA (14) and after ten minutes

260

both 5-pCoQA (13) and 4-pCoQA (14) were observed. Unfortunately, due to very low

261

concentration we could not compare the peak areas of compounds 13 and 14 in the UV

262

chromatogram and hence cannot comment on the kinetics of the acyl migration. Although it

263

was clear that 4-pCoQA (14) was formed earlier than 5-pCoQA (13), it was not obvious

264

whether 13 was an acyl migration product of 14. Formation of intramolecular acyl migration

265

products takes place according to the conclusions established earlier in this study. After two

266

minutes the rate of hydrolysis of 5-CQA was very low and the amounts of 3-CQA (2), 5-CQA

267

(3) and 4-CQA (4) were the highest. Between five to ten minutes of basic hydrolysis,

268

equilibrium was reached where 5-CQA was found to be the predominant isomer. Formation

269

of the cis-cinnamoyl derivatives supposedly follow a water addition-elimination pathway to

270

the double bond in the cinnamoyl moiety during the basic hydrolysis as described earlier.25

271

When an equimolar mixture of ferulic acid (31) and 5-CQA (3) was treated with base we

272

observed three regio-isomers of feruloylquinic acid resulting from intermolecular

273

transesterification: 3-FQA (22), 5-FQA (23) and 4-FQA (24).22 The peak area of the

274

compounds 22-24 is very low compared to the intramolecular acyl migration products formed

275

in this experiment hence, they do not appear in the plot shown in Figure 6. The same was not

276

observed in the case of the p-coumaric acid experiment. After two minutes, 3-FQA (peak area

277

= 3215) was formed predominantly over 5-FQA and 4-FQA. 5-FQA (23) and 4-FQA (24)

278

showed negligible peak areas. Only 5-FQA remained stable enough to be detected after 10

279

and 15 min of basic hydrolysis. After 30 min all the transesters and the substrate 5-CQA was

280

found to be decomposed completely as we could identify caffeic acid only. Formation of 3-

281

FQA was found to be kinetically favoured. This was found to be consistent with the fact that

282

during the first few minutes of the base treatment of 5-CQA to study intramolecular acyl

283

migration, 3-CQA dominates the acyl migration product spectrum. We also identified cis-112 ACS Paragon Plus Environment

Page 13 of 43

Journal of Agricultural and Food Chemistry

284

O-caffeoylquinic acid (47) in the EIC (extracted ion chromatogram) of m/z 353 and the UV

285

chromatogram on the basis of its early elution and fragmentation (Figure 7).23 Similar to 1-

286

cis-CQA (47), 4-cis-CQA (48) was also identified as another isomerised caffeoylquinic acid

287

derivative, both of which are assumed to be formed by an addition elimination of the water

288

molecule across the double bond in caffeic acid. 4-cis-CQA (48) was observed to be in

289

equilibrium with 4- trans-CQA (4) after two minutes but with an increase in the reaction time

290

only 4-CQA was observed to survive the base treatment. Unexpectedly, we identified two

291

chlorogenic acid lactones as minor by products, namely 3-CQL (37) and 4-CQL (38) in this

292

sample. Under basic conditions the dehydrated product of caffeoylquinic acid must be in

293

continuous equilibrium with caffeoylquinic acid. A single hetero-diacyl chlorogenic acid was

294

observed in the chromatogram in the form of caffeoyl-feruloylquinic acid (49) after 5 to 15

295

minutes of basic hydrolysis (Figure 6). Figure 7 shows the fragmentation pathway for

296

compound (49), in which it loses the ferulic acid moiety and undergoes simultaneous

297

dehydration to give m/z 335 as a base peak in MS2. Furthermore, in MS3 the dehydrated

298

caffeoylquinic acid entity undergoes decarboxylation to give a base peak at m/z 291 and also

299

shows the presence of ferulic acid as a secondary peak at m/z 193. This fragmentation

300

pathway for a caffeoyl-feruloylquinic acid was found to be inconsistent with the

301

fragmentation pathways of 1-caffeoyl-3-feruloylquinic acid, 3-feruloyl-5-caffeoylquinic acid,

302

cis-4-feruloyl-5-caffeoylquinic

303

feruloylquinic acid and cis-3-feruloyl-5-caffeoylquinic acid previously reported by Jaiswal et

304

al..21 Hence, the regio-chemistry of the acyl groups in caffeoyl-feruloylquinic acid (49)

305

remains unknown.

306

Intramolecular acyl migration: model roasting of 2-5 and 8-10

307

Compounds 3-CQA (2), 5-CQA (3), 4-CQA (4), 1,3-diCQA (5), 3,4-diCQA (8), 3,5-diCQA

308

(9) and 4,5-diCQA (10) were heated at 180 °C for 12 min separately to study the

acid,

4-feruloyl-5-caffeoylquinic

13 ACS Paragon Plus Environment

acid,

4-caffeoyl-5-

Journal of Agricultural and Food Chemistry

Page 14 of 43

309

intramolecular acyl migration in the absence of solvents in conditions closely mimicking

310

coffee roasting. From the data summarized in Table 2, we clearly see that only 4-CQA (4),

311

3,4-diCQA (8) and 3,5-diCQA (9) undergo transformations to generate various dehydrated

312

products mainly in the form of caffeoyl quinic acid lactones. In the case of mono-acylated

313

chlorogenic acid reference standards, only 4-CQA undergoes acyl migration with

314

simultaneous dehydration. 3-CQL (37) was found to be the predominant transformation

315

product in the heat treatment of 4-CQA. 5-caffeoyshikimic acid was also identified but was

316

found to be the least favoured dehydration product after 4-CQL. In this experiment, we

317

observed that 3-CQL (37) was forming predominantly over 4-CQL (38) irrespective of the

318

substrate, possibly because of the additional stability gained by the equatorial position that 3-

319

CQL assumes in the inverted chair conformation required for stereoelectronic reasons for

320

lactonization. Therefore we suggest that the dehydration processes such as lactone and

321

shikimic acid formation at the quinic acid moiety follows acyl migration. i.e. in the simulated

322

roasting environment, acyl migration takes place before dehydration at the quinic acid moiety.

323

Additionally, lactonization dominates over the alternative shikimate formation in the model

324

roasting of all the substrates (Table 2). 1,3-diCQA (5) did not undergo noticeable

325

transformation by the heat treatment whereas 3,4-diCQA (8) generated most of the

326

transformation products among all four di-CQAs. It is likely that the observed products 3-

327

CQL (37) and 4-CQL (38) resulting from 3,4-diCQA (8) were formed by lactonization and

328

loss of one of the two acyl moieties following the order of the processes as dissociation first

329

and dehydration at the quinic acid moiety later; attributed to the fact that both of the caffeoyl

330

lactones possess similar peak areas. 3,4-di-O-caffeoyl-1,5-quinide (46)

331

formed in the heat treatment of 3,4-diCQA. The presence of the two different peaks eluting at

332

51.2 and 52.4 min having the same fragmentation pattern as 3,4-diCQL (46) suggested the

333

presence of a cis isomer of compound 46 (Figure 8). Both isomers of 3,4-diCQL (46)

14 ACS Paragon Plus Environment

was preferably

Page 15 of 43

Journal of Agricultural and Food Chemistry

334

generate a base peak at m/z 335 with virtually non-existent secondary peaks confirming the

335

regio-chemistry of 3,4-diCQL.

336

Among the unchanged substrates throughout the heat treatment at 180 °C for 12 min such as,

337

5-CQA, 3-CQA, 1,3-diCQA and 4,5-diCQA the stability of 5-CQA (3) and 1,3-diCQA (5) to

338

high temperatures was confirmed by heating them at 200 °C for 10 min.

339

Transesterification (Intermolecular acyl migration): model roasting (cross-over

340

experiment)

341

In these sets of experiments, we explored the possibilities of transesterification or

342

intermolecular acyl migration by carrying out cross-over experiments, in which 5-CQA was

343

subjected to heat treatment in the presence of free acids like ferulic acid and p-coumaric acid

344

at 1:1 stoichiometry. In the first case an equimolar mixture of pCoA and 5-CQA (3) was

345

heated together. It was observed that 5-CQA (3) mostly remained unchanged by the heat

346

treatment with the peak area corresponding to 5-CQA accounting for five times the peak area

347

of all the transformation products combined (Table 3). Again caffeoyl quinic acid lactones

348

dominated as the main transformation products. Since it was established earlier by Clifford et

349

al.3 that acyl migration takes place before dehydration in chlorogenic acid when there is still

350

some water present in the sample, it can be assumed that the 5-acyl group migrates to

351

positions C3 and C4 first and then both products undergo dehydration to yield the

352

corresponding lactones in the form of 3-CQL and 4-CQL. 4-CQL (38) was formed in larger

353

quantities if compared to 3-CQL (37). Only a single transesterification product was identified

354

in the form of 4-pCoQA (14), which formed during model roasting (Table 3 and Figure 8). It

355

is speculated that the formation of the 1,5-quinide takes place earlier making C5 on quinic acid

356

unavailable for the possible condensation with p-coumaric acid hence, we do not observe 5-

357

pCoQA (13) but 4-pCoQA (14).

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 43

358

Model roasting experiment between 5-CQA (3) and ferulic acid (31) generated a larger

359

number of transformation products compared to the number of transformation products

360

detected from the same experiment with pCoA (33) and 5-CQA (3). Similar trend in terms of

361

number of the transformed products was observed in the cross-over experiment by TMAH

362

treatment. The list of the transformation products in this experiment looks very similar to the

363

list in Table 3. An additional two products are found to be the acyl migration products of 5-

364

CQA. The acyl group in 5-CQA (3) was found to presumably migrate intramolecularly to

365

produce 3-CQA (2) and 4-CQA (4) but the corresponding caffeoyllactones 3-CQL (37) and 4-

366

CQL (38) were formed more in quantity. The ratio of 3-CQL (37) to 3-CQA (2) was 11:1

367

whereas, the ratio of 4-CQL (38) to 4-CQA (4) was found to be 1:1.3 since, 4-CQL formed in

368

higher quantity if compared to 3-CQL judged by comparison of the corresponding peak areas

369

(Table 3). In this experiment it was observed that 5-CQA (3) remained unchanged to even

370

greater extent than in previous experiments by the heat treatment since the peak area under 5-

371

CQA alone is more than 16 times the sum of peak areas of all the transformation products

372

combined (Table 3). 5-FQA (23) was formed in considerable quantity confirming

373

intermolecular acyl migration between ferulic acid and 5-CQA (3) (Figure 9).

374

Intramolecular acyl migration: brewing of CGAs

375

In these experiments we investigated acyl migration products of mono- and di-acyl CGAs (2-

376

5 and 8-10) formed during the brewing process (Table 4). Typically, the pH of a coffee brew

377

is between 4.7-5.1. In our case the pH of the model brew was measured at 5.0. We observed

378

that hot water also serves as a reactive reagent other than just as a simple solvent in coffee

379

brewing similar to previous work on tea fermentation, where water was shown to be the key

380

reagent in thearubigin formation.24 Apart from the acyl migration products and trans-cis

381

isomerization (cis-caffeoylquinic acids) products; the resulting chromatograms showed

16 ACS Paragon Plus Environment

Page 17 of 43

Journal of Agricultural and Food Chemistry

382

transformation products referred here as hydroxy-dihydrocaffeoylquinic acids arising through

383

conjugate addition of water molecule to the olefinic cinnamoyl moiety.25

384

Similar to the model roasting experiment, 5-CQA (3) did not show any acyl migrated

385

products in hydrolysis (brewing). The acyl moiety in 4-CQA (4) and 3-CQA (2) did not

386

migrate to C5 of the quinic acid but acyl moiety interchange between C3 and C4 was observed

387

due to the stability of the ortho-ester intermediate arising from the cis geometry (Figure 4).

388

Acyl migration to C5 from C3 and C4 was found to be highly pH dependent as we only

389

observed it in case of basic hydrolysis.

390

Cynarin (1,3-diCQA) did not show any transformation products under acidic brewing

391

conditions after 5 h. of reflux. Among the remaining di-acylated reference standards 3,4-di-

392

acylated esters were preferably formed irrespective of the substrate. This observation can

393

possibly be attributed to the fact that the parallel aromatic rings allow for

394

interactions providing added stability to the 3,4-diCQA in the minimum energy chair

395

conformation of qunic acid moiety. By comparing the peak areas in the UV chromatogram it

396

was observed that the decomposition of the di-CQAs to produce mono-CQAs was taking

397

place in minute quantity as compared to the basic hydrolysis experiment to study

398

intermolecular acyl migration. In the case of mono-acylated chlorogenic acids, up to 1.5-2.0%

399

of the chlorogenic acids were transformed into their hydroxylated derivatives and the

400

diacylated chlorogenic acids up to 4-4.5% if the relative peak areas in EIC are considered.25

401

but their peak areas in UV chromatograms were found to be negligible. The structures for the

402

water addition compounds can be found in in our previous publication25 and are not further

403

commented on here.

404

In this work we investigated acyl migration under three conditions relevant to the processing

405

of coffee, aqueous basic and acidic and dry roasting conditions. Under all three conditions

406

acyl migration was observed, however only dominating as a pathway under aqueous basic 17 ACS Paragon Plus Environment

-

stacking

Journal of Agricultural and Food Chemistry

Page 18 of 43

407

conditions. Under aqueous acidic and dry roasting conditions acyl migration products were in

408

contrast observed as minor products. Acyl migration in aqueous solution is under typical

409

chlorogenic acids concentrations in a coffee brew predominantly a reversible intramolecular

410

process, showing a strong pH dependency. We could furthermore show that the acyl

411

migration phenomena occurs in dry roasting before dehydration takes place in the quinic acid

412

moiety. Acyl migration is facilitated in presence of the liquid media, as the esters present on

413

C3 position of the quinic acid are prone to hydrolysis of the ester bond, rather than undergoing

414

acyl migration in any experimental condition. In cross-over experiments (TMAH and model

415

roasting), with 5-CQA (3) ferulic acid if compared to p-coumaric acid showed increased

416

reactivity generating a large number of transesterification products. A significant proportion

417

of 3-CQA and 4-CQA quantities and their derivatives in a cup of coffee must be assumed to

418

be a result of acyl migration occurring in roasting and brewing processes originating from 5-

419

CQA.

420

Acknowledgements

421

The authors gratefully acknowledge the financial support from Kraft Food (now Mondelez

422

International) and Jacobs University Bremen gGmbH. We acknowledge support from Dr.

423

Frank Ullrich during the project. Furthermore we acknowledge excellent technical support by

424

Ms. Anja Müller.

425

Supporting Information Available:

426

Table S1: Initial concentrations of the reference standards.

427

Figure S1: Structures identified after brewing of the reference standards.

428

Figure S2: Amount of the transformation products after base hydrolysis for different time

429

intervals of 3,4- and 4,5-diCQA reference standards. 18 ACS Paragon Plus Environment

Page 19 of 43

430

Journal of Agricultural and Food Chemistry

This material is available free of charge via the Internet at http://pubs.acs.org

431 432 433 434 435 436

437

Figure Legends

438

Figure 1. Structure of mono- and di-caffeoylquinic, p-coumaroylquinic and feruloylquinic

439

acids

440

Figure 2. UV Chromatograms (318-322 nm) at 2, 5, 10 and 30 min of basic hydrolysis of 5-

441

CQA

442

Figure 3. Amount of the transformation products after basic hydrolysis for different time

443

intervals of 5-CQA, 4-CQA and 3-CQA

444

Figure 4. Mechanism of the acyl migration through an ortho-ester intermediate formation

445

Figure 5. Amount of the transformation products after basic hydrolysis for different time

446

intervals of di-acylated reference standards

447

Figure 6. Comparison between the peak areas of compounds formed during TMAH treatment

448

of 5-CQA with p-coumaric acid and 5-CQA with ferulic acid

449

Figure 7. EIC and fragmentation patterns for 1-cis-caffeoylquinic acid (47) at m/z 353 and

450

caffeoyl-feruloylquinic acid (49) at m/z 529 in transesterification induced by TMAH 19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

451

Figure 8. EIC and fragmentation patterns for ion at m/z 497 observed during model roasting

452

Figure 9. MS3 and MS4 of 4-pCoQA (14) and 5-FQA (23) respectively observed during

453

cross-over experiment by model roasting

20 ACS Paragon Plus Environment

Page 20 of 43

Page 21 of 43

454

Journal of Agricultural and Food Chemistry

References

455

1. Clifford, M.N. Chlorogenic acids and other cinnamates - nature, occurrence, dietary

456

burden, absorption and metabolism. J. Sci. Food Agric. 2000, 80, 1033-1043.

457

2. Jaiswal, R.; Matei, M.F.; Golon, A.; Witt, M.; Kuhnert, N. Understanding the fate of

458

chlorogenic acids in coffee roasting using mass spectrometry based targeted and non-targeted

459

analytical strategies. Food Funct. 2012, 3, 976-984.

460

3. Crozier, A.; Jaganath, I.B.; Clifford, M.N. Dietary phenolics: chemistry, bioavailability and

461

effects on health. Nat. Prod. Rep. 2009, 26, 1001-1043.

462

4. Farah, A.; De Paulis, T.; Trugo, L.C.; Martin, P.R. Effect of roasting on the formation of

463

chlorogenic acid lactones in coffee. J. Agric. Food Chem. 2005, 53, 1505-1513.

464

5. Kweon, M.H.; Hwang, H.J.; Sung, H.C. Identification and antioxidant activity of novel

465

chlorogenic acid derivatives from bamboo (Phyllostachys edulis). J. Agric. Food Chem. 2001,

466

49, 4646-4655.

467

6. Wang, G.; Shi, L.; Ren, Y.; Liu, Q.; Liu, H.; Zhang, R.; Li, Z.; Zhu, F.; He, P.; Tang, W.;

468

Tao, P.; Li, C.; Zhao, W.; Zuo, J. Anti-hepatitis B virus activity of chlorogenic acid, quinic

469

acid and caffeic acid in vivo and in vitro. Antiviral Res. 2009, 83, 186-190.

470

7. Hemmerle, H.; Burger, H.; Below, P.; Schubert, G.; Rippel, R.; Schindler, P.W.; Paulus, E.;

471

Herling, A.W. Chlorogenic acid and synthetic chlorogenic acid derivatives: novel inhibitors

472

of hepatic glucose-6-phosphate translocase. J. Med. Chem. 1997, 40, 137-145.

473

8. Robinson, W.E.; Reinecke, M.G.; AbdelMalek, S.; Jia, Q.; Chow, S.A. Inhibitors of HIV-1

474

replication that inhibit HPV integrase. Proc. Natl. Acad. Sci. U. S. A. 1996, 93, 6326-6331.

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

475

9. Kwon, H.C.; Jung, C.M.; Shin, C.G.; Lee, J.K.; Choi, S.U.; Kim, S.Y.; Lee, K.R. A new

476

caffeoyl quinic acid from Aster scaber and its inhibitory activity against human

477

immunodeficiency virus-1 (HIV-1) integrase. Chem. Pharm. Bull. 2000, 48, 1796-1798.

478

10. Clifford, M.N.; Kellard, B.; Birch, G.G. Characterisation of chlorogenic acids by

479

simultaneous isomerisation and transesterification with tetramethylammonium hydroxide.

480

Food Chem. 1989, 33, 115-123.

481

11. Clifford, M.N.; Kellard, B.; Birch, G.G. Characterization of caffeoylferuloylquinic acids

482

by simultaneous isomerization and transesterification with tetramethylammonium hydroxide.

483

Food Chem. 1989, 34, 81-88.

484

12. Dawidowicz, A.L.; Typek, R. Thermal stability of 5-O-caffeoylquinic acid in aqueous

485

solutions at different heating conditions. J. Agric. Food Chem. 2010, 58, 12578-12584.

486

13. Trugo, L.C.; Macrae, R. A study of the effect of roasting on the chlorogenic acid

487

composition of coffee using HPLC. Food Chem. 1984, 15, 219-227.

488

14. Anonymous IUPAC Commission on the Nomenclature of Organic Chemistry (CNOC)

489

and IUPAC-IUB Commission on Biochemical Nomenclature (CBN). Nomenclature of

490

cyclitols. Recommendations, 1973. Biochem J. 1976, 153, 23-31.

491

15. Clifford, M.N.; Williams, T.; Bridson, D. Chlorogenic acids and caffeine as possible

492

taxonomic criteria in Coffea and Psilanthus. Phytochemistry 1989, 28, 829-838.

493

16. Hanson, K.R. Chlorogenic acid biosynthesis Chemical synthesis and properties of the

494

mono-O-cinnamoylquinic acids. Biochemistry 1965, 4, 2719-2731.

495

17. Xie, C.; Yu, K.; Zhong, D.; Yuan, T.; Ye, F.; Jarrell, J.A.; Millar, A.; Chen, X.

496

Investigation of isomeric transformations of chlorogenic acid in buffers and biological 22 ACS Paragon Plus Environment

Page 22 of 43

Page 23 of 43

Journal of Agricultural and Food Chemistry

497

matrixes by ultraperformance liquid chromatography coupled with hybrid quadrupole/ion

498

mobility/orthogonal acceleration time-of-flight mass spectrometry. J. Agric. Food Chem.

499

2011, 59, 11078-11087.

500

18. Clifford, M.N.; Johnston, K.L.; Knight, S.; Kuhnert, N. Hierarchical scheme for LC-MSn

501

identification of chlorogenic acids. J. Agric. Food Chem. 2003, 51, 2900-2911.

502

19. Clifford, M.N.; Knight, S.; Kuhnert, N. Discriminating between the six isomers of

503

dicaffeoylquinic acid by LC-MSn. J. Agric. Food Chem. 2005, 53, 3821-3832.

504

20. Jaiswal, R.; Matei, M.F.; Ullrich, F.; Kuhnert, N. How to distinguish between

505

cinnamoylshikimate esters and chlorogenic acid lactones by liquid chromatography-tandem

506

mass spectrometry. J. Mass Spectrom. 2011, 46, 933-942.

507

21. Jaiswal, R.; Sovdat, T.; Vivan, F.; Kuhnert, N. Profiling and characterization by LC-MSn

508

of the chlorogenic acids and hydroxycinnamoylshikimate esters in Mate (Ilex

509

paraguariensis). J. Agric. Food Chem. 2010, 58, 5471-5484.

510

22. Kuhnert, N.; Jaiswal, R.; Matei, M.F.; Sovdat, T.; Deshpande, S. How to distinguish

511

between feruloyl quinic acids and isoferuloyl quinic acids by liquid chromatography/tandem

512

mass spectrometry. Rapid Commun. Mass Spectrom. 2010, 24, 1575-1582.

513

23. Clifford, M.N.; Kirkpatrick, J.; Kuhnert, N.; Roozendaal, H.; Salgado, P.R. LC-MSn

514

analysis of the cis isomers of chlorogenic acids. Food Chem. 2008, 106, 379-385.

515

24. Kuhnert, N.; Drynan, J.W.; Obuchowicz, J.; Clifford, M.N.; Witt, M. Mass spectrometric

516

characterization of black tea thearubigins leading to an oxidative cascade hypothesis for

517

thearubigin formation. Rapid Commun. Mass Spectrom. 2010, 24, 3387-3404.

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

518

25. Matei, M.F.; Jaiswal, R.; Kuhnert, N. Investigating the chemical changes of chlorogenic

519

acids during coffee brewing: conjugate addition of water to the olefinic moiety of chlorogenic

520

acids and their quinides. J. Agric. Food Chem. 2012, 60, 12105-12115.

521

522

24 ACS Paragon Plus Environment

Page 24 of 43

Page 25 of 43

Journal of Agricultural and Food Chemistry

Table 1. Compounds identified after base treatment of CGA with p-coumaric acid and CGA with ferulic acid for various time intervals 5-CQA + p-coumaric acid 5-CQA + ferulic acid Reaction Product RT (min) m/z (M-H) Reaction Product RT (min) m/z (M-H) time(min) number time(min) number 2

33 32 37 38 2 3 4

26.70 19.30 25.70 29.40 11.50 17.10 20.80

162.6 178.6 334.8 334.9 352.8 352.8 352.8

5

37 38 14 1 2 3 4 40 32 33

25.80 29.80 27.60 10.70 12.10 17.30 21.70 30.60 19.50 26.60

335.0 335.1 337.0 353.1 353.0 353.1 353.1 367.1 178.6 162.6

2 37 38

12.20 25.90 29.80

353.1 335.1 335.0

10

2

31 34 22 23 24 47 2 48 3 4 37 38 32 35

27.3 29.6 19.2 24.1 26.4 3.0 11.2 13.1 16.9 20.5 23.6 25.7 7.4 19.2

192.6 192.6 366.7 366.9 366.9 352.8 352.9 352.8 352.8 352.8 334.9 334.9 178.6 178.6

5

23 40 1 37 38 2 3

26.9 28.7 11.10 25.9 29.6 12.5 17.7

367.1 367.1 353.1 335.1 335.1 353.1 353.0

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

15

30

1 3 4 33 40 13 14 32

10.80 17.50 21.90 26.40 30.60 24.30 28.40 19.70

353.1 353.0 353.1 162.9 367.1 336.8 337.0 178.6

38 2 3 32 4

29.80 12.40 17.70 19.60 22.00

335.1 353.1 353.0 178.9 352.8

3 4 33

18.00 22.60 26.60

353.1 352.8 162.9

4 49

22.2 30.8

353.0 529.0

10

37 1 2 3 4 23 40 49

25.9 10.9 12.2 17.6 22.1 26.8 28.9 31.0

335.1 353.0 353.0 353.0 353.1 367.1 367.1 529.0

15

37 2 3 4 23 40 49

26.2 12.4 17.7 22.3 27.2 28.9 31.2

335.1 353.0 353.0 353.1 367.1 367.1 529.0

30

40 3

28.9 18.0

367.1 353.0

26 ACS Paragon Plus Environment

Page 26 of 43

Page 27 of 43

Journal of Agricultural and Food Chemistry

Table 2. Compounds identified after heating (model roasting) reference standards Starting material 3

Compound 5-CQA

Product 3

RT(min) 23.7

Peak Area(UV) 11035

2

3-CQA

2

18.9

6034

4

4-CQA

37 38 41 47 48 52 48

31.8 35.1 32.9 44.9 52.3 46.3 49.5

2418 1543 396 210 197 80 521

5

1,3-diCQA

5

30.2

16245

8

3,4-diCQA

37 38 46 46-cis

31.7 35.2 51.2 52.4

98 74 7099 896

9

3,5-diCQA

46 46-cis

56.2 56.9

3034 119

10

4,5-diCQA

10

46.3

26560

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 43

Table 3. Compounds identified after heating (Model roasting) of 5-CQA (3) with p-coumaric acid and 5-CQA (3) with ferulic acid 5-CQA (3) and p-coumaric acid Product RT (min) Peak area number 25.9 1093 37 29.6 1538 38 28.0 330 14 36.9 145 47 39.8 642 52 41.6 2137 52 17.5 30503 3

5-CQA (3) and ferulic acid Product RT (min) Peak area number 26.0 484 37 29.4 535 38 16.8 44224 3 12.2 45 2 21.9 421 4 26.8 484 23 36.8 176 47 39.9 138 52 41.9 500 52

28 ACS Paragon Plus Environment

Page 29 of 43

Journal of Agricultural and Food Chemistry

Table 4. Compounds identified after hydrolysis of reference standards (Brewing of CGAs) Starting material Product name

RT (Min) Peak area(UV)

5-CQA(3)

5-CQA cis-5-CQA 5-hCQA I 5-hCQA II

20.1 23.0 7.3 7.9

17610 347 NA NA

4-CQA(4)

3-CQA 4-CQA cis-3-CQA cis-4-CQA 4-hCQA I 4-hCQA II

13.1 20.6 11.9 16.5 6.9 7.9

17576 6098 17 179 NA NA

3-CQA(2)

3-CQA 4-CQA cis-3-CQA 3-hCQA I + II

13.1 20.6 11.9 5.6

639 954 150 17

1, 3-diCQA(5)

1, 3-diCQA

22.9

3422

3, 4-diCQA(8)

3,4-diCQA cis-4,5-diCQA I 4,5-diCQA cis-3,4-diCQA I cis-3,4-diCQA II 3-CQA 4-CQA 5-CQA 3-C-4-hCQA

35.2 38.0 36.9 34.2 35.9 12.3 19.3 15.6 26.4

2558 99 520 393 345 20 27 2 57

3, 5-diCQA(9)

3,4-diCQA 3,5-diCQA 4,5-diCQA 3-C-5-hCQAI 3-hC-5-CQA I 3-hC-cis-5-CQA 3-hC-5-CQA II 3-CQA 5-CQA

36.5 37.3 41.4 23.8 27.8 28.3 31.7 12.9 20.3

1406 1057 1107 14 22 4 4 80 126

4, 5-diCQA(10)

3,4-diCQA

43.7

1825

29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

3,5-diCQA 4,5-diCQA cis-4,5-diCQA I cis-4,5-diCQA II 3-CQA 4-CQA 5-CQA 4-hC-5-CQA 3-hC-5-CQA II 3-C-5-hCQA II

42.6 45.7 47.2 52.0 16.8 27.9 23.1 34.7 35.5 63.1

512 2148 36 7 19 14 18 18 18 23

NA- Insignificant peak area

30 ACS Paragon Plus Environment

Page 30 of 43

Page 31 of 43

Journal of Agricultural and Food Chemistry

HO

R'1O

O C

OR5 OR4 R' O 1

O C

OR5 OR4

O C CH HC

HO

O C CH HC

HO

O C CH

HC

OR3

OR1 OR3

OH Q

Number 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17

S

O

OH

OH

OH

C

pCo

F

HO C CH O CH CH3 HO

Q R1 R3 R4 R5 R'1 1-O-caffeoylquinic acid (1-CQA) C H H H 3-O-caffeoylquinic acid (3-CQA) H C H H 5-O-caffeoylquinic acid (5-CQA) H H H C 4-O-caffeoylquinic acid (4-CQA) H H C H 1,3-di-O-caffeoylqunic acid (1,3-diCQA) C C H H 1,4-di-O-caffeoylqunic acid (1,4-diCQA) C H C H 1,5-di-O-caffeoylqunic acid (1,5-diCQA) C H H C 3,4-di-O-caffeoylqunic acid (3,4-diCQA) H C C H 3,5-di-O-caffeoylqunic acid (3,5-diCQA) H C H C 4,5-di-O-caffeoylqunic acid (4,5-diCQA) H H C C 1-O-p-coumaroylqunic acid (1-pCoQA) pCo H H H 3-O-p-coumaroylqunic acid (3-pCoQA) H pCo H H 5-O-p-coumaroylqunic acid (5-pCoQA) H H H pCo 4-O-p-coumaroylqunic acid (4-pCoQA) H H pCo H 1,3-di-O-p-coumaroylqunic acid (1,3-dipCoQA) pCo pCo H H 1,4-di-O-p-coumaroylqunic acid (1,4-dipCoQA) pCo H pCo H 1,5-di-O-p-coumaroylqunic acid (1,5-dipCoQA) pCo H H pCo -

OH

cis-C

Name and abbreviation

R3 -

31 ACS Paragon Plus Environment

S R4 -

R5 -

Journal of Agricultural and Food Chemistry

18 19 20 21 22 23 24 25 26 27 28 29 30 36 37 38 39 40 41 42 43 44 45 46 47 48 Q-

3,4-di-O-p-coumaroylqunic acid (3,4-dipCoQA) H pCo pCo H 3,5-di-O-p-coumaroylqunic acid (3,5-dipCoQA) H pCo H pCo 4,5-di-O-p-coumaroylqunic acid (4,5-dipCoQA) H H pCo pCo 1-O-feruloylquinic acid (1-FQA) F H H H 3-O-feruloylquinic acid (3-FQA) H F H H 5-O-feruloylquinic acid (5-FQA) H H H F 4-O-feruloylquinic acid (4-FQA) H H F H 1,3-di-O-feruloylqunic acid (1,3-diFQA) F F H H 1,4-di-O-feruloylqunic acid (1,4-diFQA) F H F H 1,5-di-O-feruloylqunic acid (1,5-diFQA) F H H F 3,4-di-O-feruloylqunic acid (3,4-diFQA) H F F H 3,5-di-O-feruloylqunic acid (3,5-diFQA) H F H H 4,5-di-O-feruloylqunic acid (4,5-diFQA) H H F F 1-O-caffeoyl-1,5-quinide (1-CQL) C H H L 3-O-caffeoyl-1,5-quinide (3-CQL) H C H L 4-O-caffeoyl-1,5-quinide (4-CQL) H H C L 5-O-p-coumaroyl-methylquinate H H H pCo Me 5-O-caffeoyl-methylquinate H H H C Me 5-O-caffeoylshikimic acid (5-CSA) cis-5-O-caffeoylshikimic acid (5-cis-CSA) 4,5-di-O-caffeoylshikimic acid (4,5-diCSA) 3,5-di-O-caffeoylshikimic acid (4,5-diCQA) 3,4,5-tri-O-caffeoylshikimic acid (3,4,5-triCSA) 3,4-di-O-caffeoyl-1,5-quinide (3,4-diCQL) H C C L cis-1-O-caffeoylquinic acid (1-cis-CQA) cis-C H H H cis-4-O-caffeoylquinic acid (4-cis-CQA) H H cis-C H quinic acid, Ccaffeic acid, pCop-coumaric acid, F-

32 ACS Paragon Plus Environment

H H H C C -

H C H cis-C C C H C C C ferulic acid,

Page 32 of 43

L-

lactone,

Me-

methyl

Page 33 of 43

Journal of Agricultural and Food Chemistry

Figure 1. Structure of mono and di caffeoylquinic, p-coumaroylquinic and feruloylquinic acids

33 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Intens. [mAU] 100 75 50 25 0

2 Min

5 Min 100 75 50 25 0

2

3

2

3

Page 34 of 43

4 32

4 32

10 Min

4 2

3

20 10 32 0 3

30 Min 10.0 7.5 5.0

2 4

32

0 10.0

12.5

15.0

17.5

20.0

22.5

25.0

27.5

30.0 Time [min]

Figure 2. UV Chromatograms (318-322 nm) at 2, 5, 10 and 30 min of basic hydrolysis of 5CQA

34 ACS Paragon Plus Environment

Page 35 of 43

Journal of Agricultural and Food Chemistry

3000

A Peak area (mAU)

2500 3-CQA 2000

4-CQA 5-CQA

1500 1000 500 0 2 Min.

5 Min. 10 Min. Time (Min)

30 Min.

30

Peak area (mAU)

25

B 3-CQA

20

4-CQA 5-CQA

15 10 5 0 2 Min.

5 Min. 10 Min. Time (Min)

30 Min.

8 Peak area (mAU)

7

C

6 5

3-CQA

4

4-CQA

3

5-CQA

2 1 0 2 Min.

5 Min. 10 Min. Time (Min)

30 Min.

Figure 3. Amount of the transformation products after basic hydrolysis for different time intervals of A. 5-CQA; B. 4-CQA; and C. 3-CQA 35 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

O

O O

O

HO OH

OH

5-CQA (3)

H

O

-H2O

R O

Page 36 of 43

O O

HO OH

OH

O

R

OH

O O

HO OH

OH

O

R

H OH

4-CQA (4)

trans-ortho-ester intermediate

-H2O

O

OH

O OH

HO OH

O

O

R 3-CQA (2)

OH O

HO OH

R

O O

cis-ortho-ester intermediate

Figure 4. Mechanism of the acyl migration through an ortho-ester intermediate formation

36 ACS Paragon Plus Environment

Page 37 of 43

Journal of Agricultural and Food Chemistry

900

A

Peak area (mAU)

800

1, 3-diCQA

700

3, 4-diCQA

600

4, 5-diCQA

500

3-CQA

400

5-CQA

300

4-CQA

200

1-CQA

100 0 2 Min.

5 Min. Duration (Min)

300

10 Min.

30 Min.

B

250 Peak area (mAU)

3, 5-diCQA 200

3, 4-diCQA

150

4, 5-diCQA

100 50 0 2 Min.

5 Min.

10 Min.

30 Min.

60 Min.

Duration (Min)

Figure 5. Amount of the transformation products after basic hydrolysis for different time intervals of A. 1,3-diCQA; B. 3,5-diCQA

37 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

12000

A

3-CQA

10000

Peak Area (mAU)

Page 38 of 43

5-CQA 4-CQA

8000

Caffeic acid 6000 4000 2000 0 2

5

10

15

30

Duration (Min)

16000 14000

B

3-CQA 5-CQA

10000

4-CQA

Peak Area (mAU)

12000

4-cis-CQA 8000

1-CQA

6000

Caffeic acid

4000 2000 0 2

5

10 Duration (Min)

15

30

Figure 6. Comparison between the peak areas of compounds formed during TMAH treatment of A. 5-CQA (3) with p-coumaric acid (pCoA); B. 5-CQA with ferulic acid (FA)

38 ACS Paragon Plus Environment

Page 39 of 43

Journal of Agricultural and Food Chemistry

Intens. 7 x10

3

EIC 353.0

4

3 2

2

48

47 1 0

5

[%]

10

15

20

25

Time [min]

MS2(353)

47

191

100 0 172.5

126.7

100

MS3(353>191)

110.7 0 170.6

100

0

110

154.6

126.6

110.6 120

MS4(353 >191>173)

130

140

150

160

170

180

190

Intens. x107

m/z

EIC 529.0

0.8 0.6 0.4 49

0.2 0.0

22

24

26

28

30

32

39 ACS Paragon Plus Environment

34

Time [min]

Journal of Agricultural and Food Chemistry

[%] 49

Page 40 of 43

MS2(529.0)

334.9

100 290.9 0

MS3(529>335)

290.9 100 148.9

192.9

0

MS4(529>335>291) 148.9

100 0

50

100

150

200

250

300

350

400

450

500

m/z

Figure 7. EIC and fragmentation patterns for 1-cis-caffeoylquinic acid (47) at m/z 353 and caffeoyl-feruloylquinic acid (49) at m/z 529 in transesterification induced by TMAH

40 ACS Paragon Plus Environment

Page 41 of 43

Journal of Agricultural and Food Chemistry

Intens. x10 6

EIC at m/z 497

46

4

46-cis

2

0

42

44

46

48

50

52

54

56

[%] 100

MS2(497)

335.1

46/ 46-cis

Time [min]

0

MS3(497>335)

160.9 100 135.3

0

MS4(497>335>161) 133

100 0

100

200

300

400

500

m/z

Figure 8. EIC and fragmentation patterns for ion at m/z 497 observed during model roasting

41 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

[%] 100

MS, 28.0min

337

4-pCoQA (14)

Page 42 of 43

0

MS2(337), 28.0min

173 100 0

MS3(337>173), 28.0min 110.9 93.1

100 0

100

154.7 150

200

250

300

350

[%]

0 100

127 110.7

0

m/z

MS3(367 >191), 26.8min

173

0 100

450

MS2(366.7), 26.8min

191.1

5-FQA (23) 100

400

MS4(367>191 >173), 26.9min 109

100

125

150

175

200

225

250

275

300

m/z

Figure 9. MS3 and MS4 of 4-pCoQA (14) and 5-FQA (23) respectively observed during cross-over experiment by model roasting

42 ACS Paragon Plus Environment

Page 43 of 43

Journal of Agricultural and Food Chemistry

Table of Content

43 ACS Paragon Plus Environment

Investigation of acyl migration in mono- and dicaffeoylquinic acids under aqueous basic, aqueous acidic, and dry roasting conditions.

Acyl migration in chlorogenic acids describes the process of migration of cinnamoyl moieties from one quinic acid alcohol group to another, thus inter...
749KB Sizes 1 Downloads 5 Views