Determination of Diacetyl in Beer by a Precolumn Derivatization-HPLC

Mar 11, 2017 - Diacetyl is an important flavoring compound in many foods, especially in beer. In the present study, we developed and validated a new p...
1 downloads 6 Views 553KB Size
Subscriber access provided by UB + Fachbibliothek Chemie | (FU-Bibliothekssystem)

Article

Determination of Diacetyl in Beer by a Pre-Column Derivatization-HPLC-UV Method Using 4-(2,3-dimethyl-6quinoxalinyl)-1,2-benzenediamine as a Derivatizing Reagent Ji-Yu Wang, Xin-Jie Wang, Xian Hui, Shui-Hong Hua, Heng Li, and Wen-Yun Gao J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b00990 • Publication Date (Web): 11 Mar 2017 Downloaded from http://pubs.acs.org on March 12, 2017

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 28

Journal of Agricultural and Food Chemistry

Determination of Diacetyl in Beer by a Pre-Column Derivatization-HPLC-UV Method Using 4-(2,3-dimethyl-6-quinoxalinyl)-1,2-benzenediamine as a Derivatizing Reagent

Ji-Yu Wang, Xin-Jie Wang, Xian Hui, Shui-Hong Hua, Heng Li*, and Wen-Yun Gao*

National Engineering Research Center for Miniaturized Detection Systems and College of Life Sciences, Northwest University, 229 North Taibai Road, Xi’an, Shaanxi 710069, China

*To whom correspondence should be addressed (Heng Li: Fax, +86 29 88303572; Tel, +86 29 88303446 ext. 834; E-mail: [email protected]. Wen-Yun Gao: Fax, +86 29 88303572; Tel, +86 29 88303446 ext. 832; E-mail: [email protected])

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

ABSTRACT: Diacetyl is an important flavoring compound in many foods, especially in beer.

2

In the present study, we developed and validated a new pre-column derivatization HPLC-UV

3

method for the determination of diacetyl using 4-(2,3-dimethyl-6-quinoxalinyl)-1,2-

4

benzenediamine as a novel derivatizing reagent. After derivatization with the reagent at a pH

5

value 4.0 at ambient temperature for 10 min, diacetyl was analyzed on an ODS column and

6

detected at 254 nm. The results show that the correlation coefficient of the method is 0.9991

7

in the range of 0.10 to 100.0 µM diacetyl, and the limit of detection is 0.02 µM. The method

8

was further evaluated in the analysis of beer samples with the recoveries ranging from 94.4%

9

to 102.6% and RSDs from 1.36% to 3.33%. The concentrations of diacetyl in 8 beer samples

10

were determined in the range of 0.19 to 0.42 µM. The method established in this study may be

11

well suitable for the determination of diacetyl in beer.

12

KEYWORDS: Diacetyl, determination, 4-(2,3-dimethyl-6-quinoxalinyl)-1,2-benzenediamine,

13

HPLC, pre-column derivatization

14 15 16 17 18 19 20 21 22 2

ACS Paragon Plus Environment

Page 2 of 28

Page 3 of 28

Journal of Agricultural and Food Chemistry

23

INTRODUCTION

24

2,3-Butanedione, also known as diacetyl or biacetyl, is a natural byproduct of fermentation

25

that is responsible for the aroma of many food products and beverages. It is also widely used

26

as a food additive for improving the flavor of popcorn, candy, chocolate, and roasted foods.1,2

27

However, the compound has some undesirable impacts on health safety and the flavor of wine

28

and beer. Research has shown that diacetyl may be harmful when inhaled over a long duration

29

and may cause various toxic responses, for example, lung disease, Alzheimer’s disease,

30

mutagenesis, and carcinogenesis.3,4 The odor threshold of diacetyl for wine can be up to 58.14

31

µM depending on the type of wine, 5 and this value is extremely low for beer (approximately

32

1.16 µM).6 If the concentration of diacetyl is higher than the sensory threshold, wine and beer

33

will smell and taste like spoiled food. Therefore, it is necessary to establish efficient and

34

practical methods to determine the concentration of this compound not only for health safety

35

but also for controlling the quality of various fermented foods such as wine and beer.

36

As a consequence of this concern, many methods have been developed for measuring

37

diacetyl concentrations.7,8 Among these methods, the pre-column derivatization HPLC is

38

currently the most popular way to quantify this compound in different samples, not only in

39

various food products such as wine,9-15 beer,9,16,17 coffee,18-19 soy sauce,13,19 honey,20,21

40

vinegar,22 baby food,23 soft drink,24 fructose agave syrups,25 and other foodstuffs,26 but also in

41

body liquids such as urine27-31 and plasma.32 The most frequently employed diacetyl

42

derivatization reagent is o-phenylenediamine (1, Figure 1) because it is readily available and

43

has a low price. It forms a quinoxaline derivative with diacetyl, which can be easily separated

44

and determined by HPLC equipped with various detectors such as UV, fluorescence, and 3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

45

MS.9-13,16,18-20,22-24,33 The drawback of 1 is that the derivatization reaction needs to be heated at

46

approximately 40°C or higher for at least 10 min or longer to ensure its completion. If the

47

reaction is carried out at room temperature, then it has to be kept in the dark for at least a few

48

hours or even overnight.20,23,33 Both high reaction temperatures and long reaction times could

49

lower the accuracy of the determination, especially for wine and beer because of the

50

nonenzymatic oxidation of acetolactate to diacetyl.37,38 Other derivatization reagents used for

51

HPLC determination of diacetyl are analogues or derivatives of 1; however, these compounds

52

should also react with diacetyl under heating conditions, for example

53

5,6-diamino-2,4-hydroxypyrimidine (2, Figure 1, 60-80°C, 30 min),14,29 3,4-diaminopyridine

54

(3, 90°C, 2 hr),15 4-nitro-1,2-diaminobenzene (4, 45°C, 20 min),17 5,6-diamino-1-methyluracil

55

(5, 60% acetic acid, reflux for 24 hr),21 6-hydroxy-2,4,5-triaminopyrimidine (6, 60°C, 45

56

min),27,30 4-methoxy-1,2-diaminobenzene (7, 40°C, 4 hr or reflux in ethanol for 40 min),25,26,31

57

2,3-diaminonaphthalene (8, rt, overnight),32 4,5-dimethoxy-1,2-diaminobenzene (9, 60°C, 4

58

hr),28,34 1,2-diamino-4,5-methylenedioxybenzene (10, 60°C, 40 min),35 and rhodamine

59

B-hydrozine (11, 37°C, 3 hr).36 Therefore, it is necessary to develop novel reagents that can

60

quickly react with diacetyl under ambient conditions.

61

3,3΄-Diaminobenzidine (12) has been widely used in various scientific fields and is

62

currently utilized as a monomer to prepare high-temperature resistant synthetic resins and

63

fibers. However, in the analytical field, it has been employed only as a chromogenic reagent

64

to determine selenium.39,40 In this research, we found that the mono-quinoxaline derivative of

65

12, 4-(2,3-dimethyl-6-quinoxalinyl)-1,2-benzenediamine (13) (Figure 2) can react with

66

diacetyl quickly at room temperature and can be used to determine diacetyl concentrations in 4

ACS Paragon Plus Environment

Page 4 of 28

Page 5 of 28

Journal of Agricultural and Food Chemistry

67

beer with high sensitivity and accuracy. An additional advantage of the derivatization is that it

68

is almost independent of the acidity of the reaction mixture. Here, we will discuss the

69

experimental details.

70

MATERIALS AND METHODS

71

Chemicals and Reagents. Compounds 4, 12 and diacetyl are of analytical-grade and

72

were purchased from Fluka (Shanghai). HPLC-grade methanol was purchased from

73

Sigma-Aldrich (Beijing). Millipore water was obtained using a Milli-Q water purification

74

system. Compound 13 was synthesized in this lab from 12. All other chemicals and solvents

75

are of analytical-grade and were obtained from commercial sources. The stock solution of 100

76

mM diacetyl was prepared in Millipore water; 5 mM reagent 13 was prepared in 0.1 M HCl;

77

and 1.3 mM (200 mg/L) reagent 4 was prepared in methanol. The stock solutions were stored

78

in at 4°C before use.

79

Instrumentation. An Agilent 1200 HPLC system (Agilent Technologies, Shanghai)

80

equipped with a DAD detector and a Shim-pack VP-ODS column (250 × 4.6 mm, 4.6 µm,

81

Shimadzu, Japan) were used for the separation and the analysis. 1H-NMR spectra were

82

collected on a Varian Inova-600 MHz NMR spectrometer. HRESI-MS was performed with a

83

Thermo Fisher LTQ XL system.

84

Preparation of 13. One hundred and eight milligrams (0.3 mmol) of 12 was added to

85

a round bottom flask containing 10 mL of water. The mixture was stirred at room temperature

86

(rt) for approximately 20 min before 35 mg (0.4 mmol) diacetyl was added to the mixture.

87

After further stirred for half an hour, the mixture was basified with 1 M NaOH and extracted

88

with CHCl3 (15 mL × 3). The CHCl3 layers were pooled and dried over anhydrous MgSO4 for 5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

89

approximately 2 hr. The solid was then filtered out, and the organic solvent was evaporated

90

under reduced pressure. The dark red powder that was obtained was further isolated on a

91

silica gel column using CH3OH-CHCl3-diethylamine = 6-1-0.1 (V/V/V) as a developing

92

solvent. Two compounds were acquired, and their structures were elucidated by HRESI-MS

93

and 1H-NMR. The results showed that one product was the desired compound 13 and the

94

other was fully diacetylated compound 2,2΄,3,3΄-tetramethyl-6,6΄-biquinoxaline (14).

95

4-(2,3-Dimethyl-6-quinoxalinyl)-1,2-benzenediamine (13). Obtained as dark red

96

1 powder (45.2 mg, yield 57.1%). H-NMR (D2O, 600 MHz): 7.51 (1H, brs), 7.41 (1H, brs),

97

7.15 (2H, brs), 7.00 (2H, m), 2.38 (3H, s, Me), 2.31 (3H, s, Me). HRESI-MS (positive mode)

98

m/z: calcd for C16H18N4, [M+H]+ 265.1453, found 265.1461.

99

2,2΄,3,3΄-Tetramethyl-6,6΄-biquinoxaline (14). Obtained as dark red powder (18.4

100

mg, yield 24.6%). 1H-NMR (D2O, 600 MHz): 7.58 (2H, brs), 7.33 (2H, d, J = 6), 7.15 (2H, d,

101

J = 6), 2.41(6H, s, Me × 2), 2.36 (6H, s, Me × 2). HRESI-MS (positive mode) m/z: calcd for

102

C20H19N4, [M+H]+ 315.1623, found 315.1617.

103

Derivatization Procedure and Identification of the Derivative. Ten microliters of a

104

diacetyl solution (0.2 mM) was added to 50 µL of a solution of 13 (0.2 mM, containing 40%

105

methanol), and the resulting solution was mixed well using a vortex mixer. The total solution

106

was then kept at rt for 10 min, and 10 µL was used for HPLC analysis. All the samples were

107

filtered through a 0.22-µm filter membrane before they were injected into the HPLC system.

108

The derivative was then characterized by LC-MS.

109 110

Chromatographic Method. The mobile phase comprised methanol and water. The gradient used was as follows: 0 min, 60% MeOH; 10 min, 100% MeOH; and 15-17 min, 60% 6

ACS Paragon Plus Environment

Page 6 of 28

Page 7 of 28

Journal of Agricultural and Food Chemistry

111

MeOH. Both of the mobile-phase solvents were filtered with a 0.22-µm membrane before use.

112

The analysis was carried out at ambient temperature with an injection volume of 10 µL, a

113

flow rate of 0.7 mL/min, and UV detection at 254 nm. Each sample was injected in triplicate.

114

The retention times of compounds 13 and 14 were 7.55 and 13.03 min, respectively, and the

115

separation could be completed within 15 min.

116

Stability Test of Compound 14. Fifty microliters of a diacetyl solution (0.2 mM) was

117

added to 50 µL of a solution of 13 (1 mM, containing 40% methanol), and the resulting

118

solution was mixed well with vortex mixer. The total solution was then kept on the bench top

119

at rt for 4 days. Aliquots of 10 µL were used for HPLC analysis at different time points, and

120

the peak areas of 13 and the product 14 were recorded.

121

Analysis of Diacetyl in Beer Samples. Eight beers were purchased at local

122

supermarkets in Xi’an and stored at 4°C before determination. Then, 10 mL of each beer

123

sample was degassed with magnetic stirring at rt and 0.9 mL was transferred to a 2-mL

124

Eppendorf vial. Each vial was supplemented with 90 µL of a solution of 13 (0.5 mM), and 10

125

µL of Millipore water and homogenized with a vortex mixer. After been kept at rt for 10 min,

126

the resulting solutions were filtered through a 0.22-µm membrane and analyzed using the

127

HPLC system. For spiking experiments, the addition of 10 µL of Millipore water was

128

replaced by addition of 10 µL of a diacetyl standard solution. Each sample was measured in

129

triplicate.

130

RESULTS AND DISCUSSION

131 132

Preparation of Compound 13 and Identification of its Derivatives. The preparation of quinoxalines has been reported to require a high temperature, a strong acidic medium, and 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

133

a long time.41 However, a recent paper showed that vicinal diamines could react with

134

1,2-dicarbonyl compounds in water at rt.42 Thus, we carried out the reaction between 12 and

135

diacetyl using the reaction conditions with a 12 to diacetyl ratio approximately 1 to 1 (Figure

136

2). Because 12 possesses two sets of the reactive vicinal diamine moiety, we deduced that two

137

compounds, i.e., 13 and 14, could have formed. HPLC and LC-MS analyses of the reaction

138

mixture confirmed that these two compounds were produced (Figure 3). After isolation on a

139

silica gel column 13 and 14 were obtained and their structures were elucidated by interpreting

140

the 1H-NMR and HRESI-MS spectra. Then 13 was utilized to derivatize diacetyl and the

141

HPLC analysis of the reaction mixture showed that a new compound formed whose HPLC

142

behavior and HRESI-MS datum were identical with that of 14.

143

Optimization of Derivatization Conditions. To determine the optimum

144

derivatization conditions, the effects of four factors, including the concentration ratio of

145

compound 13 to diacetyl, the pH of the reaction mixture, the reaction time, and the reaction

146

temperature, on the resulting peak areas of the reaction products were investigated, and the

147

results are shown in Figure 4.

148

Different concentrations of 13 (0.1-1.0 mM) were reacted with 0.1 mM diacetyl at rt

149

for 10 min (pH 4.0). The results (Figure 4A) indicate when the ratio reaches 4 to 1, the

150

maximum production of 14 could be detected in the reaction mixture. Therefore, the ratio of

151

13 to diacetyl was set to 5 to 1 in the following experiments to ensure the sensitivity of the

152

determination. In the test of the optimum acidity, we found (Figure 4B) that the pH values

153

ranging from 1-10 did not affect the formation of 14. Since the pH value of the reaction

154

mixture resulted from 90 µL of a solution of 13 (0.5 mM) and 0.9 mL beer is approximately 8

ACS Paragon Plus Environment

Page 8 of 28

Page 9 of 28

Journal of Agricultural and Food Chemistry

155

4.0, we selected this acidity for the derivatization procedure. From the data shown in Figure

156

4C, we were able to determine that rt would be the best choice for the derivatization of

157

diacetyl with 13. Figure 4D depicted the influence of the reaction time on the production of

158

14. The results revealed that the reaction went to its end within 5 min. To ensure the

159

completeness of the derivatization and the accuracy of the analysis, we selected 10 min as the

160

best time duration.

161

Based on the above assays, we decided the optimum conditions of the derivatization

162

reaction between the reagent 13 and diacetyl were as follows: ratio of 13 to diacetyl: 5 to 1;

163

reaction acidity: pH 4.0; reaction temperature: rt; and reaction time: 10 min. Due to the

164

relatively poor water solubility of the product 14, we added methanol to the reaction mixture

165

to avoid its precipitation. We found that 20% methanol in the derivatization solution was

166

enough to keep the compound soluble.

167

Stability of Compound 14. The stability test results showed that there were no

168

obvious changes in the peak area of 14 after it has been kept on the bench top at room

169

temperature for 4 days, indicating their stability under testing conditions.

170

Validation of the Method. Totally ten samples at diacetyl concentrations between 0.1

171

and 200 µM were derivatized with 13 separately and measured and the linear calibration

172

curve was fitted. We then evaluated the sensitivity of the procedure by determining the limit

173

of detection (LOD) at a signal-to-noise ratio of 3 and the limit of quantification (LOQ) at a

174

signal-to-noise ratio of 10.43 Subsequently, we assessed the reproducibility of the method by

175

determining the relative standard deviation (RSD) with a diacetyl concentration of 10 µM. As

176

the mean value of six measurements, RSD of method reached 2.83%. The linear calibration 9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

177

ranges, regression equations, regression coefficient (R2), LOD, LOQ, and RSD of the new

178

method were calculated and are listed in Table 1.

179

Application of Compound 13 to Beer Analysis. Based on the above validation, we

180

evaluated the accuracy of the developed method by determining the concentrations of diacetyl

181

in different beer samples using the standard addition method. Three different concentrations

182

of standard diacetyl were added to the beer samples. Three replicates were used for each

183

concentration, and each sample was injected in six replicates. The recovery and RSD values

184

were calculated and are listed in Table 2. The representative chromatograms of the blank of

185

compound 13, the standard diacetyl derivatized with 13, the beer sample without 13

186

derivatization, and 13 derivatized diacetyl in the beer sample before and after spiking are

187

shown in Figure 5. The data exhibit that the recoveries of diacetyl are between 94.4% and

188

102.6% and the RSDs are between 1.36% and 3.33%, relating to the different beer samples.

189

Therefore, the new method established in this study is well suitable for the quantification of

190

diacetyl in beer samples.

191

Comparison of Compounds 4 and 13. Up to now, quite a few derivatizing reagents

192

have been employed to measure diacetyl in various samples by HPLC methods (Figure 1).

193

Among all the HPLC procedures using UV detection, the method utilizing 4 as a derivatizing

194

reagent displays the lowest LOD and LOQ (Tables 1).8,17 To compare 4 and 13, we validated

195

the published method in the derivatization of standard diacetyl and in the determination of

196

beer samples and the results were listed in Tables 1 and 2, respectively. Combining these

197

validation data and the optimized derivatization conditions for each reagent, we would

198

conclude that: 10

ACS Paragon Plus Environment

Page 10 of 28

Page 11 of 28

199

Journal of Agricultural and Food Chemistry

i)

Compound 13 is more reactive than 4 because 13 could derivatize diacetyl under

200

milder conditions (rt, 5min at pH from 1-10) than 4 (45°C, 20 min at pH from 1-3).

201

This could be owing to the structural difference of the two compounds. Nitro

202

group possesses strong electron withdrawing effect and its existence can largely

203

reduce the nucleophilicity of the amino groups in compound 4; whereas the

204

quinoxalinyl moiety of 13 shows mainly its conjugation effect which can enhance

205

to some extent the reactivity of its amino groups.

206

ii) Although there is no significant difference in the recoveries and RSD of the two

207

methods (Table 2), compound 13 is more suitable than 4 for quantification of

208

diacetyl because the slope of the calibration curve of 13 is about three times

209

bigger than that of 4; whereas 4 is a better option in the qualification of diacetyl

210

because it shows lower LOD than 13 (Table 1). The larger slope of the method

211

using 13 should result from the bigger conjugation system of this compound, but

212

the reason that the method using 4 shows a lower LOD remains ambiguous.

213

In summary, we set up in this study a novel derivatization procedure for the measurement

214

of diacetyl in beer by HPLC employing 13 as a new derivatizing reagent. The advantage of

215

the method is that the derivatization reaction could be carried out quickly under mild

216

conditions (rt, 5 min) and in a wide pH range (1-10). Moreover, the process also exhibits

217

effective chromatographic separation, satisfactory linearity, and excellent repeatability. The

218

mild conditions for diacetyl derivatization can efficiently prevent the formation of extra

219

diacetyl under heating condition owing to nonenzymatic oxidation of the precursor of diacetyl

220

in the samples and thus ensure the accuracy of the determination. These conditions effectively 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

221

simplify the operation and expedite the measurement as well.

222 223

Funding

224

This work was supported by the National Science Foundation of China (Grants 21172179,

225

21402152), the Program for Changjiang Scholars and Innovative Research Team in

226

University (No. IRT_15R55), and the Scientific Research Project of Education Department of

227

Shaanxi Provincial Government (No. 15JK1710).

228

Acknowledgments

229

The authors gratefully acknowledge Dr. Xinfeng Zhao and Dr. Chaoni Xiao of the College of

230

Life Sciences, Northwest University for their kind help with the LC-MS and 1H-NMR

231

analyses.

232

Notes

233

The authors declare no competing financial interest.

234 235

REFERENCES

236

[1] Cheng, H. Volatile flavor compounds in yogurt: A review. Crit. Rev. Food Sci. Nutr.

237

2010, 50, 938−950.

238

[2] Lozano, P. R.; Miracle, E. R.; Krause, A. J.; Drake, M.; Cadwallader, K. R. Effect of cold

239

storage and packaging material on the major aroma components of sweet cream butter. J.

240

Agri. Food Chem. 2007, 55, 7840−7846.

241 242

[3] Kovacic, P.; Cooksy, A. L. Role of diacetyl metabolite in alcohol toxicity and addiction via electron transfer and oxidative stress. Arch. Toxicol. 2005, 79, 123–128. 12

ACS Paragon Plus Environment

Page 12 of 28

Page 13 of 28

243

Journal of Agricultural and Food Chemistry

[4] Stoner, G. D.; Shimkin, M. B.; Kniazeff, A. J.; Weisburger, J. H.; Weisburger, E. K.;

244

Gori, G. B. Test for carcinogenicity of food additives and chemotherapeutic agents by the

245

pulmonary tumor response in strain a mice. Cancer Res. 1973, 33, 3069–3085.

246

[5] Martineau, B.; Acree, T. E.; Henick-Kling, T. Effect of wine type on the detection

247

threshold for diacetyl. Food Res. Int. 1995, 28, 139−143.

248

[6] Rodrigues, P. G.; Rodrigues, J. A.; Barros, A. A.; Lapa, R. A. S.; Lima, J. L. F. C.; Cruz,

249

J. M.; Ferreira, A. A. Automatic flow system with voltammetric detection for diacetyl

250

monitoring during brewing process. J. Agri. Food Chem. 2002, 50, 3647−3653.

251 252 253 254 255

[7] Krogerus, K.; Gibson, B. R. 125th anniversary review: diacetyl and its control during brewery fermentation. J. Inst. Brew. 2013, 119, 86−97. [8] Shibamoto, T. Diacetyl: Occurrence, Analysis, and Toxicity. J. Agric. Food Chem. 2014, 62, 4048−4053. [9] Barros A.; Rodrigues, J. A.; Almeida, P. J.; Oliva-Teles, M. T. Determination of glyoxal,

256

methylglyoxal, and diacetyl in selected beer and wine by HPLC with UV

257

spectrophotometric detection, after derivatization with o-phenylenediamine. J. Liq.

258

Chrom. & Rel. Technol. 1999, 22, 2061–2069.

259

[10] De Revel, G.; Pripis-Nicolau, L.; Barbe, J.-C.; Bertrand, A. The detection of α-dicarbonyl

260

compounds in wine by formation of quinoxaline derivatives. J. Sci. Food Agric. 2000, 80,

261

102–108.

262

[11] Da Silva-Ferreira, A. C.; Reis, S.; Rodrigues, C.; Oliveira, C.; Guedes de Pinho, P.

263

Simultaneous determination of ketoacids and dicarbonyl compounds, key Maillard

264

intermediates on the generation of aged wine aroma. J. Food Sci. 2007, 72, S314−S318. 13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

265

[12] Ramos, R. M.; Grosso Pacheco, J.; Moreira Gonçalves, L.; Valente, I. M.; Rodrigues, J.

266

A.; Araújo Barros, A. Determination of free and total diacetyl in wine by HPLC−UV

267

using gas-diffusion microextraction and pre-column derivatization. Food Control 2012,

268

24, 220−224.

269

[13] Santos, C. M.; Valente, I. M.; Gonçalves, L. M.; Rodrigues, J. A. Chromatographic

270

analysis of methylglyoxal and other α-dicarbonyls using gas-diffusion microextraction.

271

Analyst, 2013, 138, 7233–7237.

272

[14] Hurtado-Sanchez, M. C.; Espinosa-Mansilla, A.; Rodríguez-Caceres, M. I.; Duran-Meras,

273

I. Evaluation of Liquid Chromatographic Behavior of Lumazinic Derivatives, from α‑

274

Dicarbonyl Compounds, in Different C18 Columns: Application to Wine Samples Using

275

a Fused-Core Column and Fluorescence Detection. J. Agric. Food Chem. 2014, 62,

276

97−106.

277

[15] Rodríguez-Cáceres, I.; Palomino-Vasco, M.; Mora-Diez, N.; Acedo-Valenzuela, I. Novel

278

HPLC- Fluorescence methodology for the determination of methylglyoxal and glyoxal.

279

Application to the analysis of monovarietal wines “Ribera del Guadiana”. Food Chem.

280

2015, 187,159–165.

281

[16] Grosso Pacheco, J.; Maria Valente, I.; Moreira Goncalves, L.; Jorge Magalhas, P.;

282

António Rodrigues, J.; Araújo Barros A. Development of a membraneless extraction

283

module for the extraction of volatile compounds: Application in the chromatographic

284

analysis of vicinal diketones in beer. Talanta 2010, 81, 372–376.

285 286

[17] Li, P.; Zhu, Y.; He, S.; Fan, J.; Hu, Q.; Cao, Y. Development and validation of high-performance liquid chromatography method for the determination of diacetyl in beer 14

ACS Paragon Plus Environment

Page 14 of 28

Page 15 of 28

Journal of Agricultural and Food Chemistry

287

using 4-nitro-o-phenylenediamine as the derivatization reagent. J. Agric. Food Chem.

288

2012, 60, 3013−3019.

289

[18] Daglia, M.; Patetti, A.; Aceti, C.; Sordelli, B.; Spini, V.; Gazzani, G. Isolation and

290

determination of α-dicarbonyl compounds by RP-HPLC-DAD in green and roasted

291

coffee. J. Agric. Food Chem. 2007, 55, 8877−8882.

292

[19] Papetti, A.; Mascherpa, D.; Gazzani, G. Free α-dicarbonyl compounds in coffee, barley

293

coffee and soy sauce and effects of in vitro digestion. Food Chem. 2014,164, 259–265.

294

[20] Marceau, E.; Yaylayan, V. A. Profiling of α-Dicarbonyl Content of Commercial Honeys

295

from Different Botanical Origins: Identification of 3,4-Dideoxyglucoson-3-ene (3,4-DGE)

296

and Related Compounds. J. Agric. Food Chem. 2009, 57, 10837–10844.

297

[21] Daniels, B. J.; Prijic, G.; Meidinger, S.; Loomes, K. M.; Stephens, J. M.; Schlothauer, R.

298

C.; Furkert, D. P.; Brimble, M. A. Isolation, Structural Elucidation, and Synthesis of

299

Lepteridine from Manuka (Leptospermum scoparium) Honey. J. Agric. Food Chem. 2016,

300

64, 5079−5084.

301

[22] Papetti, A.; Mascherpa, D.; Marrubini, G.; Gazzani, G. Effect of In Vitro Digestion on

302

Free α-Dicarbonyl Compounds in Balsamic Vinegars. J. Food Sci. 2013, 78, C514–C519.

303

[23] Kocadagli, T.; Gokmen, V. Investigation of α‑Dicarbonyl Compounds in Baby Foods by

304

High-Performance Liquid Chromatography Coupled with Electrospray Ionization Mass

305

Spectrometry. J. Agric. Food Chem. 2014, 62, 7714–7720.

306

[24] Gensberger, S.; Glomb, M. A.; Pischetsrieder, M. Analysis of sugar degradation products

307

with alpha-dicarbonyl structure in carbonated soft drinks by UHPLC-DAD-MS/MS. J.

308

Agric. Food Chem. 2013, 61, 10238−10245. 15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

309

[25] Corrales Escobosa, A. R.; Gomez Ojeda, A.; Wrobel, K.; Alcazar Magana, A.; Wrobel,

310

K.; Methylglyoxal is associated with bacteriostatic activity of high fructose agave syrups.

311

Food Chem. 2014, 165, 444–450.

312 313 314

[26] Maroulis, A. J.; Voulgarofwjlos A. N.; Hadjiantoniou-Maroulis, C. P. Fluorometric determination of biacetyl. Talanta 1985, 32, 504–506. [27] Espinosa-Mansilla, A.; Duran-Meras, I.; Salinas F. High-Performance Liquid

315

Chromatographic-Fluorometric Determination of Glyoxal, Methylglyoxal, and Diacetyl

316

in Urine by Prederivatization to Pteridinic Rings. Anal. Biochem. 1998, 255, 263–273.

317

[28] Akira, K.; Matsumoto, Y.; Hashimoto T. Determination of urinary glyoxal and

318

methylglyoxal by high-performance liquid chromatography. Clin. Chem. Lab Med. 2004,

319

42, 147–153.

320

[29] Espinosa-Mansilla, A.; Duran-Meras, I.; Canada Canada, F.; Marquez, P.

321

High-performance liquid chromatographic determination of glyoxal and methylglyoxal in

322

urine by prederivatization to lumazinic rings using in serial fast scan fluorimetric and

323

diode array detectors. Anal. Biochem. 2007, 371, 82–91.

324

[30] Hurtado-Sanchez, M. C.; Espinosa-Mansilla, A.; Rodrıguez-Caceres, M. I.;

325

Martın-Tornero E.; Duran-Meras, I. Development of a method for the determination of

326

advanced glycation end products precursors by liquid chromatography and its application

327

in human urine samples. J. Sep. Sci. 2012, 35, 2575–2584.

328

[31] Gómez Ojeda, A.; Wrobel, K.; Corrales Escobosa, A. R.; Garay-Sevilla, E.; Wrobel, K.

329

High-performance liquid chromatography determination of glyoxal, methylglyoxal, and

330

diacetyl in urine using 4-methoxy-o-phenylenediamine as derivatizing reagent. Anal. 16

ACS Paragon Plus Environment

Page 16 of 28

Page 17 of 28

331 332

Journal of Agricultural and Food Chemistry

Biochem. 2014, 449, 52–58. [32] Yamada, H.; Miyata S.; Igaki, N.; Yataben, H.; Miyauchill, Y.; Ohara, T.; Sakai, M.;

333

Shoda, H.; Oimomi, M.; Kasuga, M. Increase in 3-Deoxyglucosone Levels in Diabetic

334

Rat Plasma. J. Biol. Chem. 1994, 269, 20275–20280.

335 336 337

[33] Henning, C.; Liehr, K.; Girndt, M.; Ulrich, C.; Glomb, M. A. Extending the Spectrum of Dicarbonyl Compounds in Vivo. J. Biol. Chem. 2014, 289, 28676–28688. [34] McLellan, A. C.; Thomalley, P. J. Synthesis and chromatography of 1,2-diamino-4,5

338

dimethoxybenzene, 6,7-dimethoxy-2-methylquinoxaline and

339

6,7-dimethoxy-2,3-dimethylquinoxaline for use in a liquid chromatographic fluorimetric

340

assay of methylglyoxal. Anal. Chim. Acta 1992, 263,137–142.

341

[35] Hara, S.; Yamaguchi, M.; Takemori, Y.; Yoshitake, T.; Nakamura, M.

342

1,2-Diamino-4,5-methylenedioxybenzene as a highly sensitive fluorogenic reagent for

343

α-dicarbonyl compounds. Anal. Chim. Acta 1988, 215, 267–276.

344 345 346

[36] Li, X.; Duerkop, A.; Wolfbeis, O. S. A Fluorescent Probe for Diacetyl Detection. J. Fluoresc. 2009, 19, 601–606. [37] Park, S. H.; Xing, R.; Whitman, W. B. Nonenzymatic acetolactate oxidation to diacetyl

347

by flavin, nicotinamide and quinone coenzymes. Biochim. Biophys. Acta 1995, 1245,

348

366–370.

349 350 351 352

[38] Chai, P.-H.; Li, S.; Wu, H.-P.; Shi, Y.; Dan L.-P. Determination of the flavor matter diacetyl in beer. Indust. Microbiol. 200l, 31, 31–33. [39] Yan, H.; Luo, Y.; Li, Z.; Liu, W.; Lu, J. Optimization conditions on determining contents of Se in Yunnan Puer tea. Shipin Keji 2010, 35, 263–266. 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

353

[40] Ramadan, A.A.; Mandil, H.; Shikh-Debes, A. Differential pulse anodic stripping

354

voltammetric determination of selenium(IV) at a gold electrode modified with 3,3'-

355

diaminobenzidine. Intl. J. Pharm. Pharmaceu. Sci. 2014, 6, 148–153.

356

[41] Liu, J. H.; Wu, A. T.; Huang, M. H.; Wu, C. W.; Chung, W. S. The syntheses of

357

pyrazino-containing sultines and their application in Diels-Alder reactions with

358

electron-poor olefins and [60]fullerene. J. Org. Chem. 2000, 65, 3395–3403.

359

[42] Delpivo, C.; Micheletti, G.; Boga, C. A Green Synthesis of Quinoxalines and

360 361 362

2,3-Dihydropyrazines. Synthesis 2013, 45, 1546–1552. [43] Miller, J. C.; Miller, J. N. Statistic for Analytical Chemistry; Ellis Horwood: Chichester, U.K., 1985.

363

364

365

366

367

368

369

370

371

372 18

ACS Paragon Plus Environment

Page 18 of 28

Page 19 of 28

Journal of Agricultural and Food Chemistry

373

Figure captions

374

Figure 1. Chemical structures of the reagents used for the pre-column derivatization HPLC

375

determination of diacetyl.

376

Figure 2. Reaction of diacetyl with compound 12 (1 : 1) in acidic water and with 13 (1 : 5) in

377

acidic aqueous MeOH at room temperature.

378

Figure 3. HPLC and LC-MS profiles of the reaction of compound 12 and diacetyl. Line 1,

379

compound 12; line 2, compound 12 to diacetyl ratio is 1 : 1. Reaction conditions: pH 3.0, rt

380

for 10 min.

381

Figure 4. Optimization of derivatization conditions. Effects of the (A) concentration ratio of

382

compound 13 to diacetyl, (B) derivatization pH, (C) derivatization temperature, and (D)

383

derivatization time on the peak areas of the product 14. Reaction conditions: (A) 0.1 mM

384

diacetyl, pH 4.0, rt, 10 min; (B) 0.1 mM diacetyl, 0.5 mM 13, rt, 10 min; (C) 0.1 mM diacetyl,

385

0.5 mM 13, pH 4.0, 10 min; (D) 0.1 mM diacetyl, 0.5 mM 13, pH 4.0, rt.

386

Figure 5. HPLC chromatograms obtained from (A) the blank of compound 13 (line 1) and

387

compound 13 derivatized standard diacetyl (line 2), (B) beer sample without derivatization,

388

(C) beer sample blank, and (D) beer sample spiked with diacetyl standard (2.0 µM). HPLC

389

conditions: column, Shim-pack VP-ODS column (250 × 4.6 mm inner diameter, Shimadzu,

390

Japan); UV detection, λ = 254 nm; gradient, 0 min, 60% MeOH; 10 min, 100% MeOH;

391

15-17min, 60% MeOH; flow rate, 0.7 mL/min; temperature, ambient.

392 393 394 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

395

Tables Table 1. Linear Calibration Ranges, Regression Equations, and Detection Limits (LOD and LOQ) of the Methods Using Compounds 4 or 13 as Derivatizing Reagents.

a

parameters

4d

4e

13

calibration range (µM)

0.058-116

0.075-120

0.10-100

regression equation, ya

330507x +337

112348x +319

411610x + 481

R2

0.9992

0.999

0.9992

RSD (%) (n = 6)

1.76

2.06

2.83

LOD (µM)b

0.009

0.01

0.02

LOQ (µM)c

0.03

0.05

0.10

x, concentration of diacetyl (µM); y, peak area of the products. bS/N = 3, per 10 µL injection

volume. cS/N = 10, per 10 µL injection volume. ddata extracted from [17]. edata obtained in this study under the optimized conditions for reagent 4.17

20

ACS Paragon Plus Environment

Page 20 of 28

Page 21 of 28

Journal of Agricultural and Food Chemistry

Table 2. The Recoveries of Diacetyl from Different Beer Samples by the Methods Using Compounds 4 or 13 as Derivatization Reagents samples

lager Ae

lager Be

lager Ce

lager De

lager Ee

lager Fe

initiala

spiked

totalb

foundc

recovery

RSDd

(µM)

(µM)

(µM)

(µM)

(%)

(%)

0.20

0.383

0.193

96.5

3.29

0.50

0.663

0.473

94.6

3.17

2.00

2.154

1.964

98.2

2.32

0.20

0.569

0.199

99.5

3.24

0.50

0.872

0.502

100.4

2.33

2.00

2.302

1.932

96.6

1.98

0.50

0.782

0.472

94.4

3.02

2.00

2.352

2.042

102.1

2.65

5.00

5.335

5.025

100.5

1.67

0.50

0.762

0.512

102.4

2.98

2.00

2.286

2.036

101.8

1.36

5.00

5.225

4.975

99.5

2.03

0.50

0.897

0.477

95.4

3.02

2.00

2.414

1.994

99.7

2.89

5.00

5.235

4.815

96.3

2.56

0.50

0.812

0.512

102.4

3.33

2.00

2.302

2.002

100.1

2.79

0.19

0.37

0.31

0.25

0.42

0.30

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

lager Ge

lager He

lager Af

lager Bf

a

0.26

0.22

0.21

0.37

Page 22 of 28

5.00

5.190

4.890

97.8

1.99

0.50

0.773

0.513

102.6

3.10

2.00

2.270

2.010

100.5

1.98

5.00

5.345

5.085

101.7

2.22

0.50

0.696

0.476

95.2

3.01

2.00

2.132

1.912

95.6

2.95

5.00

5.260

5.040

100.8

1.89

0.20

0.407

0.197

98.5

2.19

0.50

0.685

0.475

95.0

2.86

2.00

2.114

1.904

95.2

3.11

0.20

0.575

0.205

102.5

3.14

0.50

0.849

0.479

95.8

1.66

2.00

2.342

1.972

98.6

2.18

Mean value of diacetyl content in beer before spiking (n=6). bMean value of diacetyl content

in beer after spiking (n=6). cThe total measured diacetyl content minus the initial measured diacetyl content. dMean value of six determinations. eCompound 13 as a derivatizing reagent. f

Compound 4 as a derivatizing reagent.

22

ACS Paragon Plus Environment

Page 23 of 28

Journal of Agricultural and Food Chemistry

Figure graphics

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

24

ACS Paragon Plus Environment

Page 24 of 28

Page 25 of 28

Journal of Agricultural and Food Chemistry

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

26

ACS Paragon Plus Environment

Page 26 of 28

Page 27 of 28

Journal of Agricultural and Food Chemistry

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

TOC graphic

28

ACS Paragon Plus Environment

Page 28 of 28