Reaction Mechanisms of Metals with Hydrogen ... - ACS Publications

Apr 30, 2016 - acetaldehyde concentrations after the addition of Cu(II) (50 or 100 μM) to air-saturated ... the rapid formation of ∼1.4:1 H2S/Cu an...
4 downloads 0 Views 636KB Size
Subscriber access provided by University of Sussex Library

Article

Reaction Mechanisms of Metals with Hydrogen Sulfide and Thiols in Model Wine. Part 1: Copper Catalyzed Oxidation. Gal Y Kreitman, John C Danilewicz, David William Jeffery, and Ryan J. Elias J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b00641 • Publication Date (Web): 30 Apr 2016 Downloaded from http://pubs.acs.org on May 9, 2016

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

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

Page 1 of 43

Journal of Agricultural and Food Chemistry

Reaction Mechanisms of Metals with Hydrogen Sulfide and Thiols in Model Wine. Part 1: Copper Catalyzed Oxidation. Gal Y. Kreitman†, John C. Danilewicz§, David. W. Jeffery‡, Ryan J. Elias*,†



Department of Food Science, The Pennsylvania State University, University Park, Pennsylvania 16802, United States §

44 Sandwich Road, Ash, Canterbury, Kent CT3 2AF, United Kingdom



School of Agriculture, Food and Wine, Waite Research Institute, The University of Adelaide, PMB 1, Glen Osmond, South Australia 5064, Australia

* To whom correspondence should be addressed. Tel: +1 (814) 865-5371 Fax: +1 (814) 863-6132 E-mail: [email protected]

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

ABSTRACT

2

Sulfidic off-odors due to hydrogen sulfide (H2S) and low molecular weight thiols are

3

commonly encountered in wine production. These odors are usually removed by the process of

4

Cu(II) fining – a process that remains poorly understood. The present study aims to elucidate the

5

underlying mechanisms by which Cu(II) interacts with H2S and thiol compounds (RSH) under

6

wine-like conditions. Copper complex formation was monitored along with H2S, thiol, oxygen,

7

and acetaldehyde concentrations after addition of Cu(II) (50 or 100 μM) to air saturated model

8

wine solutions containing H2S, cysteine, 6-sulfanylhexan-1-ol, or 3-sulfanylhexan-1-ol (300 μM

9

each). The presence of H2S and thiols in excess to Cu(II) led to the rapid formation of ~1.4:1

10

H2S:Cu and ~2:1 thiol:Cu complexes, resulting in the oxidation of H2S and thiols, and reduction

11

of Cu(II) to Cu(I) which reacted with oxygen. H2S was observed to initially oxidize rather than

12

form insoluble copper sulfide. The proposed reaction mechanisms provide an insight into the

13

extent to which H2S can be selectively removed in the presence of thiols in wine.

14 15

KEYWORDS: H2S, thiols, copper, oxidation, wine aroma, reaction mechanism

2 ACS Paragon Plus Environment

Page 2 of 43

Page 3 of 43

16

Journal of Agricultural and Food Chemistry

INTRODUCTION

17

Volatile sulfur containing compounds (VSCs) have a major impact on the sensory quality of

18

wine.1–3 Typically, VSCs have exceedingly low aroma detection thresholds (i.e., μg/L to ng/L) and,

19

depending on their structure, can have beneficial or deleterious effects with respect to consumer

20

acceptance. Grape-derived varietal thiols, such as 3-sulfanylhexan-1-ol (3SH), 3-sulfanylhexyl

21

acetate (3SHA), and 4-methyl-4-sulfanypentan-2-one (4MSP), contribute pleasant aromas (e.g.,

22

grapefruit, passionfruit, and blackcurrant).4–6 On the other hand, the production of fermentation-

23

related VSCs, such as H2S, methanethiol (MeSH), and ethanethiol (EtSH), can result in the

24

development of undesirable odors, often described as rotten egg, putrefaction, sewage and burnt

25

rubber, that are obviously detrimental to wine quality.1,7,8 These odors are generally most evident at

26

low oxygen concentrations and are described to be sulfidic off-odors. Wines that display such odors

27

are described as having reductive character.

28

The accumulation of sulfidic off-odors is a common problem for winemakers and is usually

29

remedied by splash racking in order to volatilize and/or oxidize VSCs or, classically, by the use of

30

copper fining.2,7,9 In this latter practice, Cu(II) is added as its sulfate or citrate salt whereby it is

31

assumed to remove H2S by forming a highly insoluble colloidal CuS precipitate (Figure 1),9,10 which

32

can be subsequently removed from the wine by racking and/or filtration. The mechanism for copper

33

fining remains poorly understood and there are known disadvantages to the process. In the case of

34

disulfides, thioacetates, and cyclic sulfur compounds, which can also contribute unpleasant sulfidic

35

off-odors, copper fining is ineffective due to the absence of a free thiol group.2,7 Copper fining can

36

also cause significant losses of beneficial thiol compounds (e.g. 3SH, 3SHA, 4MSP) that are

37

important to the varietal character of a wine.11 Furthermore, other thiols could interfere with the fining

38

process by competing for Cu(II) given that the average combined concentration of cysteine (Cys), N-

39

acetylcysteine and homocysteine is reported to be ca. 20 µM in a number of white wines, while the 3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 43

40

average concentration of glutathione (GSH) is reported to be ca. 40 µM in wines made from

41

Sauvignon blanc.12–15 These nonvolatile thiols would be in large molar excess to the exogenous

42

copper (3–6 µM) used in a fining operation, and would far exceed the concentration of H2S (ca. 300

43

nM)16 when copper fining is considered. Furthermore, a recent study by Clark et al.17 demonstrated

44

the practical difficulty of removing CuS from wine, even with filtration, as the precipitate may not be

45

observed.10 This lack of precipitate formation would leave residual copper in wine that can contribute

46

to a series of redox-mediated reactions in the post-bottling period, as elaborated below.

47

After bottling, the concentration of sulfidic off-odors can increase, especially under reductive

48

conditions when oxygen exposure is limited such as when screw cap closures are used.11,18,19

49

Although the causative mechanism remains unclear, wine appears to contain precursors that are able

50

to produce H2S and MeSH.20,21 The formation of H2S from the Strecker degradation of Cys has been

51

previously reported,22 while some have suggested that H2S may be formed by the direct reduction of

52

sulfate or sulfite.19 It has also been shown that thiols can be reversibly bound by iron and copper,23,24

53

and that wines containing higher copper concentrations can accumulate sulfidic off-odors during

54

bottle aging.11,25 While transition metals are known to be essential for catalyzing oxidation reactions

55

in wine,26 Cu, Fe, Mn, Zn, and Al have more recently been shown to synergistically affect the

56

evolution of VSCs under anaerobic storage conditions.25

57

In order to understand how wines develop sulfidic off-odors during storage, it is essential to

58

understand how H2S and thiols react in the presence of oxygen and transition metals prior to bottling.

59

The identification of reaction products may then allow potentially troublesome precursors to be

60

targeted. Recent studies in this area have advanced our general mechanistic understanding of iron-

61

catalyzed wine oxidation; however, the role of copper remains poorly understood. The goal of this

62

present study is to determine the underlying mechanism of Cu-catalyzed H2S and thiol oxidation

63

under wine conditions. 4 ACS Paragon Plus Environment

Page 5 of 43

64

Journal of Agricultural and Food Chemistry

MATERIALS AND METHODS

65

Chemicals. 4-Methylcatechol (4-MeC), L-cysteine (Cys), monobromobimane (MBB), 5,5-

66

dimethyl-1-pyrroline N-oxide (DMPO), bathocuproinedisulfonic acid (BCDA) disodium salt, 6-

67

sulfanylhexan-1-ol (6SH), and diethylenetriaminepentaacetic acid (DTPA) were obtained from

68

Sigma-Aldrich (St. Louis, MO). 2,4-Dinitrophenylhydrazine (DNPH) was purchased from MCB

69

laboratory chemicals (Norwood, OH) and L-tartaric acid, 3SH, and 5,5’-dithiobis(2-nitrobenzoic

70

acid) (DTNB) were obtained from Alfa Aesar (Ward Hill, MA). Cupric sulfate pentahydrate was

71

purchased from EMD Chemicals (Gibbstown, NJ), TRIS hydrochloride from J.T. Baker (Center

72

Valley, PA), and sodium hydrosulfide hydrate (as a source of H2S) was purchased from Acros

73

Organics (Geel, Belgium). Water was purified through a Millipore Q-Plus system (Milipore Corp.,

74

Bedford, MA). All other chemicals and solvents were of analytical or HPLC grade, and solutions

75

were prepared volumetrically, with the balance made up with Milli-Q water unless specified

76

otherwise.

77

Model wine experiments. Model wine was prepared by dissolving tartaric acid (5 g/L) in

78

water, followed by the addition of ethanol to yield a final concentration of 12% v/v. The solution was

79

adjusted to pH 3.6 with sodium hydroxide (10 M) and brought to volume with water. For H2S and

80

Cys, an aqueous stock solution of each (0.5 M) was freshly prepared, whereas 6SH and 3SH were

81

added directly by syringe during experimentation (Figure 2). An aqueous stock solution of Cu(II)

82

sulfate (0.1 M) was prepared freshly. In certain experiments, 4-MeC (1 mM) was added prior to the

83

addition of H2S and thiol compounds, and Cu(II). H2S, Cys, 6SH, or 3SH were added to air saturated

84

model wine (1 L, 300 μM) followed by thorough mixing. Cu(II) was added to H2S, Cys, and 6SH (50

85

μM) or 3SH (100 μM) and thoroughly mixed. For mixed H2S and Cys system, H2S (100 µM) and Cys

86

(400 µM) were added to air saturated model wine (1 L), followed by the addition of Cu(II) (100 µM)

87

and thorough mixing. The solution was immediately transferred to 60 mL glass Biological Oxygen 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 43

88

Demand (B.O.D.) bottles (Wheaton, Millville, NJ), allowing the solution to overflow, and bottles

89

were capped immediately with ground glass stoppers, thereby eliminating headspace. The glass

90

reservoir of the B.O.D. bottles was topped off with water daily. The bottles were stored in the dark at

91

ambient temperature. One B.O.D. bottle was sacrificed per time point per replicate and used for

92

further analyses. All experiments were conducted in triplicate and had their own series of sacrificial

93

bottles.

94

For experiments focusing on 6SH-disulfide formation, one experiment was prepared as

95

described above and followed over time. For additional experiments for deciphering immediate

96

disulfide generation, model wine (3 mL) containing 6SH (600 μM) in a glass test tube was

97

deoxygenated for 2 min under argon with stirring. After sparging, Cu(II) was added at varying

98

concentrations (50, 100, or 200 µM) under argon and reacted with stirring for 5 minutes. The solution

99

was then immediately analyzed to determine 6SH and 6SH-disulfide concentrations (described

100

below). In experiments involving 4-MeC or DMPO, these compounds were dissolved directly into

101

model wine to achieve a final concentration of 1 mM prior to addition of Cu(II) (100 µM).

102

Determination of oxygen consumption. Prior to the experiment, 60 mL glass B.O.D. bottles

103

containing PSt3 oxidots (Nomacorc LLC, Zublon, NC) were filled with air saturated model wine for

104

a minimum of 2 hours to allow the oxidots to equilibrate. One B.O.D. bottle was used as a model

105

wine control (i.e., did not contain a treatment) and two other bottles were used as technical duplicates

106

to determine oxygen concentration for each treatment replicate (3 treatment replicates total). Thus,

107

immediately after the addition of Cu(II) solution, the model wine used for equilibration was discarded

108

and the respective treatment solution was instantly transferred into the bottles. Oxygen readings were

109

taken per time point using NomaSense O2 P6000 meter (Nomacorc LLC, Zublon, NC), and data were

110

normalized to the model wine reference sample. Starting oxygen concentrations were approximately

111

7 mg/L (~220 µM) in all solutions. 6 ACS Paragon Plus Environment

Page 7 of 43

Journal of Agricultural and Food Chemistry

112

Cu-complex formation and dissolution. 6SH-Cu(I) complex was prepared by adding Cu(II)

113

(100 µM) to model wine (1 L) containing 6SH (400 µM). The immediately formed precipitate was

114

vacuum filtered with a 0.45 µm nylon membrane (Wheaton, Millville, NJ), washed with water

115

followed by ethyl acetate in order to remove residual disulfide, and dried under vacuum. In an

116

anaerobic chamber (95% Ar, 5% H2), ~1 mg of the solid was added to water containing approximately

117

5× molar excess of BCDA. This mixture was stirred for approximately 30 min until all of the solid

118

dissolved. 6SH, 6SH-disulfide, and Cu(I) concentrations were measured as described below.

119

Spectrophotometric measurements of thiols. UV-vis spectra were recorded on an Agilent

120

8453 UV-Vis spectrophotometer (Agilent, Santa Clara, CA). Determination of Cu binding to H2S and

121

thiols was determined by measurement over 200-700 nm. The concentration of H2S, Cys, 6SH, and

122

3SH was determined using Ellman’s reagent (DTNB).27 An aliquot of sample (100 μL) diluted with

123

model wine (900 μL) was treated with a solution of DTNB (400 μL, 2 mM) in phosphate buffer (10

124

mM, pH 7.0) followed by addition of TRIS-phosphate buffer (100 μL, 1 M, pH 8.1). The mixture was

125

left at ambient temperature for 30 min before the absorbance was measured at 412 nm against a blank

126

consisting of model wine, DTNB solution, and TRIS-phosphate buffer in the proportions specified

127

above.

128

Spectrophotometric measurement of Cu(I)-BCDA. Cu(I) concentration was analyzed

129

using the BCDA assay.28 Treatment and standard solutions consisted of excess Cys (5 mM) to ensure

130

Cu(I) remained in its reduced state. An external standard curve of the Cu(I)-BCDA complex was

131

prepared in model wine, and absorbance values were recorded at 484 nm against a model wine blank.

132

HPLC analyses of thiols. MBB derivatization was used to determine each H2S and Cys

133

concentrations in the mixed system based on a modification of a previous method.29 MBB reagent

134

(40 mM) was prepared anaerobically by dissolving the solid in acetonitrile. Aliquots of the reagent 7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 43

135

were stored at -80 °C. Briefly, a sample aliquot (70 μL) was mixed with an equal volume of TRIS-

136

HCl buffer (100 mM) containing DTPA (0.1 mM) at pH 9.5, followed by the immediate addition of

137

MBB (10 μL; 40 mM). The reaction was allowed to proceed aerobically at room temperature in the

138

dark for 30 min before the addition of sulfuric acid (50 μL, 200 mM) and 6SH-bimane internal

139

standard (50 μL). 6SH-bimane was prepared following a sulfide-dibimane synthesis described

140

previously.29 Samples were filtered through PTFE syringe tip filters (0.45 μm, 13 mm filter diameter;

141

AcrodiscTM, Ann Arbor, MI) prior to analysis by HPLC-MS/MS.

142

Quantitative analysis was performed with a Shimadzu LC-VP series HPLC (Columbia, MD)

143

interfaced to a Waters Quattro micro triple quadrupole mass spectrometer (Milford, MA) that was

144

operated with MassLynx software. Bimane adducts were separated on a ZORBAX Eclipse Plus C18

145

column (2.1 x 150 mm, 5 μm) with a guard column of the same material at a flow rate of 0.2 mL/min

146

with mobile phases consisting of 0.1% v/v formic acid (A) and 0.1% v/v formic acid in acetonitrile

147

(B) and a linear gradient according to the following program: 0 min, 2% B; 9 min, 50% B; 14 min,

148

100% B; 18 min, 100% B; 19 min, 2% B; 26 min, 2% B.

149

Detection of bimane adducts was performed using negative ion electrospray ionization (ESI-

150

) with multiple reaction monitoring (MRM) (Figures S1-S3). The ESI capillary spray voltage was set

151

to 4 kV, the sample cone voltage was set to 25 V, and the source temperature was 120 °C. The

152

desolvation gas flow was 450 L/h and collision energy was set to 20 eV. The mass transition of

153

sulfide-dibimane was monitored at m/z 413→191, cysteine-bimane was monitored at m/z 310→223,

154

and the internal standard 6SH-bimane was monitored at m/z 323.2→222.2. An external standard curve

155

was prepared for sulfide-dibimane and Cys-bimane and data were normalized to the 6SH-bimane

156

internal standard.

8 ACS Paragon Plus Environment

Page 9 of 43

Journal of Agricultural and Food Chemistry

157

For experiments involving 6SH and its disulfide, quantitative analysis was performed using

158

the HPLC system described above and UV detection at 210 nm with external standard calibration

159

curves. Separation was achieved at a flow rate of 0.2 mL/min with mobile phases consisting of 0.1%

160

v/v formic acid (A) and 0.1% v/v formic acid in acetonitrile (B) and a linear gradient according to the

161

following program: 0 min, 5% B; 20 min, 95% B; 28 min, 95% B; 28.1 min, 5% B; 38 min, 5% B.

162

For experiments involving dissolution of 6SH-Cu complex with BCDA, the same

163

chromatographic conditions described for 6SH and its disulfide were followed. However, the BCDA

164

peak could not be resolved from that of 6SH at 210 nm, therefore detection of 6SH was performed

165

using ESI+ with selective ion monitoring (SIM) at m/z 135 with an external calibration curve. The

166

ESI capillary spray voltage was set at 4 kV, the sample cone voltage was set to 25 V and the source

167

temperature was 120 °C. The desolvation gas flow was 650 L/h.

168

HPLC analysis of catechols. For experiments containing 4-MeC, quantitative analysis was

169

performed with the HPLC system described above and UV detection at 280 nm with an external

170

standard calibration curve. 4-MeC was separated on an Ultra Aromax column (2.1 x 150 mm, 5 μm)

171

with a guard column of the same material at a flow rate of 0.2 mL/min with mobile phases consisting

172

of 0.1% v/v formic acid (A) and 0.1% v/v formic acid in acetonitrile (B) and a linear gradient

173

according to the following program: 0 min, 30% B; 3 min, 30% B; 12 min, 100% B; 20 min, 100%

174

B; 20.1 min, 30% B; 25 min, 30% B. The putative formation of oxidation products including catechol-

175

thiol adducts and condensed units was monitored both at 280 nm and with negative ion ESI-MS (total

176

ion chromatogram m/z 100-1000).

177

HPLC analysis of acetaldehyde. Acetaldehyde was measured in model wine treatment

178

solutions as its 2,4-dinitrophenylhydrazone (DNPH) derivative by HPLC as described previously30

179

with the following modification: the sample was centrifuged at 15000 × g at 4 °C for 10 min. The

180

supernatant was then transferred to an HPLC vial for further analysis. 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 43

181

Copper determination. For each given time point, samples were mixed in B.O.D. bottles and

182

then filtered through a 0.45 um PTFE syringe filter. The resulting filtrate (5 mL) was digested by the

183

addition of 30% hydrogen peroxide (3 mL) and sulfuric acid (100 μL) based on modification of

184

previous reported methodology.31 The samples were heated in a convection oven at 110 °C overnight

185

before being reconstituted to 5 mL with 0.1 M nitric acid. Samples were analyzed by inductively

186

coupled plasma optical emission spectroscopy (Agilent 700 Series, Santa Clara, CA) using a

187

vertically aligned torch and with monitoring at 324.7 nm.

188

EPR analysis. Loss of the electron paramagnetic resonance (EPR) signal for active Cu(II)

189

(0.5 mM) in model wine was monitored after the metal solution was mixed with the respective H2S

190

and thiol treatments (1.5 mM). Samples were transferred to a cuvette and snap frozen in liquid

191

nitrogen. Continuous wave EPR spectra were acquired on a Bruker ESP300 X-band spectrometer

192

(Billerica, MA) equipped with a ER 041MR microwave bridge and a Bruker ER 4102ST resonator.

193

Temperature was controlled by a variable temperature helium flow cryostat (ER 4112-HV, Oxford

194

Instruments, Abingdon, UK). Data acquisition and control of experimental parameters were

195

performed using the EWWIN 2012 software package. Instrument settings were as follows:

196

temperature, 100 K; microwave power, 2 mW; modulation frequency, 9480 MHz; modulation

197

amplitude, 20 dB; scan range, 2000 G.

198 199

RESULTS

200

The reactivity of Cu(II) with H2S, which is the primary target of Cu fining, and the following

201

three thiols was investigated under wine conditions (Figure 2): (1) Cys, which also represented homo-

202

Cys and Cys derivatives, (2) 6SH to represent primary thiols, and (3) 3SH to represent secondary

203

thiols. With H2S Cu(II) addition resulted in an immediate uptake of ~1.4 (72 µM) mole equivalents 10 ACS Paragon Plus Environment

Page 11 of 43

Journal of Agricultural and Food Chemistry

204

of H2S, the remainder was then fully consumed within 72 h. However, with the thiols, the immediate

205

uptake increased to approximately two equivalents (Figure 3), with initial consumption of 101 and

206

121 µM for Cys and 6SH, respectively, the remainder then being fully consumed within 48 h. The

207

varietal thiol 3SH reacted in the same manner but more slowly, with 2 mole equivalent of 3SH (210

208

µM) consumed relative to Cu(II) added after 2 hours, and was not fully reacted after 168 h (Figure

209

3).

210

EPR analysis showed that Cu(II) was immediately reduced to Cu(I) due to loss of

211

paramagnetic Cu(II) signal by Cys, 6SH and H2S; again, 3SH reacted more slowly (Figure 4A), with

212

Cu(II) reduction being complete after 2 h (data not shown). The apparent formation of a Cu(I)

213

complex was observed by UV spectroscopy (Figure 4B). Absorbance increased markedly from 200-

214

400 nm by the addition of H2S and Cys to model wine containing Cu(II), but did not produce a distinct

215

absorbance maximum above 220 nm. In contrast, 6SH showed a maximum at 353 nm, and 3SH had

216

absorbance maxima at 282 and 311 nm (Figure 4B).

217

The addition of Cu(II) to H2S in model wine resulted in a clear golden colored solution that

218

yielded a green/black precipitate over time, whereas a haze that developed with the three thiol

219

treatments (Cys, 6SH, 3SH) aggregated to form a fine white/yellow precipitate. This was particularly

220

evident for 6SH, as essentially all the Cu(I) complex was removed by filtration (0.45 µm) from 5 to

221

45 min after mixing (Figure 5A). Filtration at earlier time points and measurement of residual copper

222

remaining in solution confirmed that the 6SH aggregate formed rapidly and could be removed from

223

solution by filtration after 5 min (Figure 5B). However, at the last time point, copper had been released

224

from the insoluble Cu(I) complex in a copper form that could not be removed by a 0.45 µm filter.

225

3SH reacted in the same manner, but more slowly. For the H2S treatment, ca. 60% of the copper was

226

removed by filtration within 5 min and up to 24 h. After 72 h, there was a green-black precipitate.

227

Approximately 90% of copper was then removed from solution (Figure 5B). 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 43

228

The aggregate initially formed from the reaction between Cu(II) and 6SH on drying gave a

229

fine powder, which was solubilized in water containing BCDA (a Cu(I) selective chelator28). The

230

insoluble Cu(I)-complex dissolved as BCDA displaced the thiolate ligand, yielding 1.17 ± 0.02 mM

231

Cu(I), as determined by UV spectrophotometry, and 1.17 ± 0.13 mM 6SH was released, as determined

232

by HPLC-MS, giving a ~1:1 Cu(I):6SH molar ratio with minimal disulfide formation (data not

233

shown).

234

When H2S (75 µM) and Cys (468 µM) were added together to model wine in the presence of

235

Cu(II), ca. 53 and 135 µM of H2S and Cys, respectively, were consumed within 5 min (Figure 6).

236

Together this gives 189 µM of sulfhydryl compounds consumed with added 100 µM Cu(II) which

237

translates to a ~2:1 binding ratio of H2S + Cys:Cu(II). Subsequent reaction resulted in complete loss

238

of H2S within 40 min and Cys after 48 h. While a visible precipitate was observed at the end of the

239

reaction (74 h), it was not observed to the same extent as was the case with H2S alone.

240

The 6SH/Cu(II) system was used to monitor disulfide formation under argon. Addition of

241

Cu(II) at 50, 100, and 200 µM resulted in disulfide generation of 19.7 ± 3.6, 43.4 ± 3.1, and 98.2 ±

242

3.6 µM, respectively (data not shown). In addition, the oxidation of 6SH (240 μM), in the presence

243

of 50 µM Cu(II) was monitored over time in air saturated model wine (Figure 7). After 262 h, 231 ±

244

2.5 µM of the thiol reacted and 116 ± 2.7 µM disulfide was produced. Approximately 69 ± 8.0 µM

245

O2 was consumed in this reaction (Figure 7), giving an O2:thiol molar reaction ratio of ~1:3.3.

246

To further examine the mechanism of disulfide formation using 6SH as a model, an attempt

247

was made to intercept potential intermediate thiyl radicals with the o-quinone-producing 4-MeC, and

248

the radical trap DMPO. However, no change in disulfide formation was observed by HPLC upon

249

addition of Cu(II) (100 µM) to model wine containing 6SH (600 µM) and 4-MeC or DMPO (1.0 mM)

250

under anaerobic conditions (data not shown).

12 ACS Paragon Plus Environment

Page 13 of 43

Journal of Agricultural and Food Chemistry

251

Oxygen consumption was also measured in model wines containing the H2S and thiol

252

treatments, as well as a combination treatment consisting of Cys+H2S (Figure 8). Minimal O2 uptake

253

(