nonanoic Acid from Disposable Latex Gloves - ACS Publications


nonanoic Acid from Disposable Latex Gloves - ACS Publicationspubs.acs.org/doi/abs/10.1021/acs.jafc.7b02444Cachedby M MuÌ...

0 downloads 47 Views 681KB Size

Subscriber access provided by UNIV OF DURHAM

Article

Concentrations, Stability and Isolation of the Furan Fatty Acid 9-(3-Methyl-5Pentylfuran-2-yl)-Nonanoic Acid (9M5) from Disposable Latex Gloves Marco Müller, Melanie Hogg, Kerstin Ulms, and Walter Vetter J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b02444 • Publication Date (Web): 18 Aug 2017 Downloaded from http://pubs.acs.org on September 1, 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

1

Concentrations, Stability and Isolation of the Furan Fatty Acid 9-(3-Methyl-5-

2

Pentylfuran-2-yl)-Nonanoic Acid (9M5) from Disposable Latex Gloves

3 4

Marco Müller, Melanie Hogg, Kerstin Ulms and Walter Vetter*

5 6 7

University of Hohenheim, Institute of Food Chemistry, Department of Food Chemistry

8

(170b), D-70593 Stuttgart, Germany

9 10 11 12

* Corresponding author

13

Phone: +49 711 459 24016

14

Fax: +49 711 459 24377

15

Email: [email protected]

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

16

Page 2 of 28

ABSTRACT

17

Because of their antioxidant properties, furan fatty acids (furan-FA) are valuable

18

minor compounds with a widespread occurrence in all living matter. Unfortunately, pure

19

standards are not readily available, as they usually contribute only 1% to the lipid fraction. A

20

known exception of this is the milky fluid of Hevea brasiliensis, commonly known as latex,

21

in which the furan-FA 9-(3-methyl-5-pentylfuran-2-yl)-nonanoic acid (9M5) contributes with

22

about 90% to the triacylglycerides. In this study, we investigated the content of 9M5 in 30

23

different disposable latex gloves, which ranged from 0.7 to 8.2 mg/g glove. The light

24

degradability of 9M5 in latex gloves was investigated and different amounts of 9M5 in

25

disposable latex gloves were attributed to varying exposure time to light. Additionally, over

26

100 mg of the methyl or ethyl ester of 9M5 (purity >98%) could be extracted from disposable

27

latex gloves, employing cold extraction and silver ion chromatography. With this method,

28

standards for the quantitation of furan-FA are obtained easily and rapidly in all laboratories.

29 30

KEYWORDS

31

Furan fatty acids, disposable latex gloves, 9M5, Hevea brasiliensis, GC/MS

2 ACS Paragon Plus Environment

Page 3 of 28

32

Journal of Agricultural and Food Chemistry

INTRODUCTION

33

Furan fatty acids (furan-FA) are a group of naturally occurring fatty acids which are

34

characterized by a furan ring in the center of the molecule. The furan moiety is substituted

35

with a straight-chain (usually 9, 11 or 13 carbons) saturated acid in α-position and a short

36

alkyl chain (usually 3 or 5 carbons) in α'-position (Figure 1).1 A typical example for furan-FA

37

is 9-(3-methyl-5-pentylfuran-2-yl)-nonanoic acid (9M5), 1, which is substituted with one

38

methyl group at the β-position (M). Other naturally occurring furan-FA possess two methyl

39

groups at the β- and β'-positions (D), while furan-FA without methyl groups (F) are rarely

40

found in nature.2

41

Furan-FA are de novo synthesized by various plants and bacteria and can be

42

incorporated into the lipids of animals.1,3 Hence, they are found widespread in all living

43

matter, albeit at low concentrations.1,3–6 Because of their high radical scavenging activity,

44

they are potent antioxidants.7,8 Due to this property they are thought to protect

45

polyunsaturated fatty acids (PUFA) from oxidative stress and support them in their beneficial

46

effects on the prevention of cardiovascular diseases.1

47

Furan-FA are easily transformed into volatile compounds by exposure to light which

48

were found to cause off-flavor in spinach,9 soybean oil10 as well as dried herbs and

49

vegetables.11 In addition, the potential furan-FA metabolite 3-carboxy-4-methyl-5-propyl-2-

50

furanpropionic acid, 2, was found to be significantly high in blood plasma of patients with

51

type 2 diabetes.12,13 2 was also found to enhance type 2 diabetes by hampering the insulin

52

biosynthesis in β-cells.14 However, high amounts of fish in the diet only moderately increased

53

the plasma concentration of 2 and it could not be linked to harmful glucose metabolism.13

54

To gain further knowledge of the biological function of furan-FA, pure standards are

55

needed to perform tests concerning their biological activity. Unfortunately, isolation of furan-

56

FA from natural sources is non-economic because they usually contribute less than 1% to the 3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 28

57

total fatty acids of plant and animal lipids.1,2,15 Hitherto, the only relevant exception is the

58

milky fluid of Hevea brasiliensis, commonly known as latex. About 90% of the fatty acid

59

pattern of latex originated from the furan-FA 9M5, 1.16,17 Recently, isolation of pure 9M5, 1,

60

from liquid centrifuged latex was achieved by countercurrent chromatography (CCC).17 This

61

isolate is suited for the use as an analytical standard in furan-FA analysis15,18,19 and may also

62

serve as substrate for in vitro investigations of the relevance of furan-FA. However, the

63

isolation procedure is not only time consuming,17 but also CCC instruments and the milky

64

latex fluid are not readily available to all researchers interested in such studies.17

65

The goal of this study was to select a more convenient source and isolation procedure

66

for 9M5, 1, For this purpose, we studied its occurrence in disposable latex gloves, i.e. one

67

readily available industrial product made from latex. Disposable latex gloves are widely

68

distributed in medical and chemical laboratories, and samples from different producers were

69

screened for the presence of 9M5, 1, and possibly other furan-FA. We also studied the

70

stability of 9M5, 1, when disposable latex gloves were removed from the (light-protecting)

71

packaging and exposed to daylight. Finally, we present an easy and fast method for the

72

isolation of methyl and ethyl esters of 9M5, 1, from disposable latex gloves.

73 74

MATERIALS AND METHODS

75

Organic solvents and chemicals. Methanol (>99.8%) and n-hexane (>95%) were

76

from VWR Chemicals (Darmstadt, Germany), while ethanol, diethyl ether (both distilled

77

before use) and sulfuric acid were from Carl Roth (Karlsruhe, Germany). Silica gel 60 and

78

silver nitrate were from Sigma Aldrich (Taufkirchen, Germany). Sodium chloride and the

79

fatty acids palmitic acid (16:0; >98.5%), stearic acid (18:0; >98.5%), oleic acid (18:1∆9;

80

>99%), linoleic acid (18:2∆9,12; 99%) and nonadecanoic acid (19:0; >99%) were from Fluka

81

Chemicals (Taufkirchen, Germany). Myristic acid (14:0; >99%) was from Merck Chemicals

4 ACS Paragon Plus Environment

Page 5 of 28

Journal of Agricultural and Food Chemistry

82

(Darmstadt, Germany). A standard of 9M5, 1, was previously isolated from centrifuged latex

83

with CCC in our laboratory.17

84

Samples and standards. At least one disposable latex glove from 30 different

85

products was obtained from 14 producers (Table 1). The quantitation standard 19:0 (ISTD-1)

86

was prepared at a concentration of 10.1 mg/mL in n-hexane/diethyl ether (1:1, v/v). For the

87

second internal standard (14:0-EE, ISTD-2), 10 mg 14:0 were transesterified as described by

88

Wendlinger et al.15 with 2 mL ethanol containing sulfuric acid (1 vol%) instead of methanol.

89

The resulting ISTD-2 (14:0-EE, c = 5.61 mg/mL) was transferred into an amber 1.5 mL screw

90

cap vial and stored at -20 °C.

91

Gas chromatography with flame ionization detector (GC/FID) or mass

92

spectrometry (GC/MS). Samples were measured on an Autosystem XL GC/FID instrument

93

(Perkin Elmer, Rodgau, Germany) which was equipped with a 25 m x 0.53 mm i.d., 0.5 µm

94

DB 23-Megabore (50% cyanopropyl, 50% dimethyl polysiloxane) column (Agilent,

95

Waldbronn, Germany). Measurements were carried out with N2 (5.0) as carrier gas (constant

96

pressure, 15 kPa). Injections (1 µL) in split mode (split ratio 1:4.7, flow rate: 20 mL/min)

97

were performed at 205 °C. The GC oven program started for 1 min at 60 °C. Then the

98

temperature was first raised to 190 °C with a ramp of 20 °C/min and then to 220 °C with 4

99

°C/min. The final temperature of 230 °C was reached with another ramp of 20 °C/min and

100

held for 10 min.

101

GC/MS measurements were performed on a 6890/5973N GC/MS system (Agilent

102

Technologies, Santa Clara, CA), equipped with a 30 m x 0.25 mm i.d., 0.25 µm HP-5MS

103

(95% methyl, 5% phenyl polysiloxane) column (Agilent, Waldbronn, Germany) as recently

104

described (system 1).20 All measurements in full scan mode (m/z 60 - 600) were done with the

105

following oven program: The initial temperature of 60 °C was held for 1 min and was then

106

raised to 180 °C with a ramp of 13 °C/min and further increased to 250 °C with 3 °C/min. The

107

final temperature of 300 °C was reached with a ramp of 20 °C/min and held for 5 min, 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 28

108

resulting in a total run time of 41.06 min. For measurements in selected ion monitoring (SIM)

109

mode characteristic ions were divided into time windows according to Wendlinger et al.15

110

Response factors of fatty acid methyl esters (FAME) were also measured on another

111

6890/5973 GC/MS system equipped with a cool on-column inlet (Hewlett-Packard/Agilent,

112

Waldbronn, Germany) and a 15 m x 0.25 mm i.d., 0.1 µm Rtx-1 (100% dimethyl

113

polysiloxane) column (Restek, Bellefonte, PA) as recently described (system 2).20 All

114

measurements were performed in full scan mode (m/z 30 – 600). Additionally, response

115

factors were measured on a 5890 Series II GC/FID instrument (Hewlett Packard) equipped

116

with a 60 m Rtx 2330 column (biscyanopropyl cyanopropylphenyl polysiloxane, 0.25 mm

117

internal diameter, 0.1 µm film thickness) (Restek, Bellefonte, PA). Injections of 1 µL in split

118

mode (1:10) were done with a Hewlett-Packard 7673 autosampler.

119

Liquid extraction and determination of the fat content. Initial tests confirmed that

120

the extraction efficiency of 9M5, 1, with 25 mL n-hexane was >90%, which was considered

121

appropriate. Hence, disposable latex gloves were cut into pieces of ~1 cm2 and 1.5 g of these

122

pieces (~30% of one glove) were supplemented with 25 mL n-hexane and extracted by

123

ultrasonication for 5 min. The hexane extract was filtered and the residue on the filter was

124

washed twice with 5 mL n-hexane. The combined hexane phases were condensed to ~2 mL

125

with a rotary evaporator and transferred into a 10 mL volumetric flask. The volume was

126

adjusted to 10 mL with n-hexane. A 1 mL aliquot of this extract was used for gravimetric

127

determination of the (lipid) extract content, performed in a pre-weighed 1.5 mL screw cap vial

128

once the solvent was removed by a gentle stream of air at 40 °C. Due to logistic reasons, air

129

had to be used in this experiment instead of nitrogen. Because 9M5, 1, is sensitive to

130

oxidation, the variation of the sample weight under the chosen experimental conditions was

131

investigated in additional tests (Figure S1, Supporting Information).

132

Formation of methyl esters. Between 3 and 8 mL latex extract (corresponding with

133

~10 mg extract content) were placed in a 10 mL vial and 50 µL ISTD-1 (19:0) was added. 6 ACS Paragon Plus Environment

Page 7 of 28

Journal of Agricultural and Food Chemistry

134

The solvent was removed by a gentle stream of nitrogen at 40 °C and the sample was

135

transesterified as previously described.15 In short, 2 mL methanol containing 1 vol% sulfuric

136

acid was added, heated to 80 °C and extracted with 2 mL n-hexane. About 1 mL of the

137

resulting FAME solution was transferred into an amber 1.5 mL screw cap vial and stored at -

138

20 °C until analysis by GC/FID or GC/MS.

139

Determination of the GC/FID response factor of 9M5-ME. The GC/FID response

140

of FAMEs in the transesterified sample was found to be nearly the same for all conventional

141

fatty acids.21 The response factor of 9M5-ME was determined with a solution containing

142

methyl esters of palmitic acid (16:0), stearic acid (18:0), oleic acid (18:1∆9), linoleic acid

143

(18:2∆9,12) and 9M5, 1, all at the same concentration of 2 mg/mL. This solution was

144

measured three times on two different GC/FID systems each and on GC/MS system 2

145

operated in full scan mode.

146

Quantitation and quality control. Before GC/FID measurements all FAME solutions

147

of purified samples were spiked with 25 µL of ISTD-2 (14:0-EE) which was used to level out

148

differences in the injection volume. The peak area of 9M5-ME was additionally corrected by

149

its GC/FID response factor. A blank sample containing both internal standards ISTD-1 and

150

ISTD-2 was measured every day and the recovery of ISTD-1 (19:0) was generally >95%.

151

Therefore, the content of 9M5, 1, in the sample was quantified by means of ISTD-1. All peaks

152

in GC/FID chromatograms were verified by means of authentic standards. In addition, the

153

lipid extracts of eight samples #9, #10, #12, #13, #20, #22, #28 and #29 (Table 1) were

154

measured on GC/MS system 1.

155

Exposure of 1 (9M5, 1) to sunlight behind a window (inside). Five gloves from two

156

different producers each #13 and #30 (Table 1) were selected and one glove of each producer

157

was extracted as shown above and measured by GC/FID. The remaining four gloves from

158

each producer (eight total) were exposed to light by placing them on a windowsill inside the

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 28

159

laboratory, facing north (August 2016). Over the course of four days (including night- and

160

daytime), every 24 h one glove from each producer was removed and analyzed.

161

Silver ion mini-column chromatography. FAMEs obtained from the latex extracts

162

were subjected to silver ion mini-column chromatography according to Wendlinger et al.15 In

163

short, ~450 mg silica gel containing 20% silver nitrate deactivated with 1% water was placed

164

into a Pasteur pipette whose conical exit was packed with glass wool. The column was

165

conditioned with ~5 mL n-hexane. Then the lipid extract (obtained from 2.5 g disposable

166

latex glove sample) was placed on the column and fractionated with three solvents. Fraction

167

(i) was collected with 15 mL n-hexane/diethyl ether (99.5:0.5, v/v), fraction (ii) was eluted

168

with 15 mL n-hexane/diethyl ether (97:3, v/v) and fraction (iii) with 15 mL n-hexane/diethyl

169

ether (80:20, v/v). The solvents of all fractions were then removed by a gentle stream of

170

nitrogen and adjusted to exactly 1 mL with n-hexane. All fractions were measured on GC/MS

171

system 1.

172

Isolation of 9M5, 1, from latex gloves. About 50 g disposable latex gloves (sample

173

#30) (Table 1) were cut into pieces of ~1 cm² and placed in an amber 1 L-glass bottle. After

174

the addition of 800 mL n-hexane, extraction was performed by ultrasonication (5 min). The

175

extract was filtered and the residue was washed twice with 50 mL n-hexane. The solvent was

176

removed and the residue was taken up in 50 mL methanol containing 1 vol% sulfuric acid (or

177

ethanol for the extraction of 9M5-EE). The solution was refluxed for 2 h at 80 °C. Then 30

178

mL water and 30 mL saturated NaCl solution were added and fatty acid methyl (ethyl) esters

179

were extracted 3x with 50 mL n-hexane. The solvent was removed by a rotary evaporator at

180

40 °C and 230 mbar and the residue taken up in 1 mL n-hexane. Column chromatography as

181

described before was performed on a large scale. About 5 g silica gel containing 20% silver

182

nitrate deactivated with 1% water was placed in a glass column (inner diameter ~1 cm) and

183

fractionation was done by eluting the sample with 70 mL n-hexane/diethyl ether (99.5:0.5,

8 ACS Paragon Plus Environment

Page 9 of 28

Journal of Agricultural and Food Chemistry

184

v/v) (fraction 1), 50 mL n-hexane/diethyl ether (97:3, v/v) (fraction 2) and 50 mL n-

185

hexane/diethyl ether (80:20, v/v) (fraction 3).

186

Statistical analysis. After verifying normal distribution with the Anderson-Darling-

187

test (p > 0.05), all results were analyzed using univariate ANOVA tests (α = 0.05) as

188

integrated in Excel and statistical significance was assumed for p > 0.05. Statistical outliers

189

were detected with the David-Hartley-Pearson-test (α = 0.05) and not considered for ANOVA

190

tests. Correlation of the lipid extract and the amount of 9M5, 1, in disposable latex gloves was

191

calculated with the Pearson correlation coefficient (PCC) in Excel.

192 193

RESULTS AND DISCUSSION

194

Peak identification in latex gloves. The hexane extracts of disposable latex gloves

195

consisted of 0.6-1.6% of the initial sample weight (mean value: 1.1%) (Table 1). In either

196

case, 9M5, 1, was detected as prominent peak in all samples. In addition to 9M5, 1, the fatty

197

acid pattern of the samples was characterized by 18:2∆9,12; 18:0; 18:1∆9; 16:0; 18:3∆9,12,15

198

and 20:0 as their methyl esters (Figure 2A), which was verified by GC/MS analysis (Figure

199

2B) of eight arbitrarily selected samples #9, #10, #12, #13, #20, #22, #28 and #29 (Table 1).

200

FAMEs were identified by means of the correct retention time and the molecular ions relative

201

to authentic external reference standards. Specifically, m/z 270 (16:0-ME), m/z 292 (18:3n-3-

202

ME), m/z 294 (18:2n-6-ME), m/z 296 (18:1n-9-ME), m/z 298 (18:0-ME) and m/z 326 (20:0-

203

ME). Likewise, 9M5-ME was identified by GC/MS data including the characteristic

204

molecular ion at m/z 322, the McLafferty ion at m/z 109, and the base peak at m/z 165).2

205

Additional GC/MS measurements in SIM mode (Figure 3A) allowed to identify traces of the

206

furan-FA 9M3, 3, 9D5, 4, and 11M5, 5, in two samples #9 and #22 (Table 1) by means of

207

their characteristic MS molecular ion, McLafferty ion, and base peak (Figures 3B-D).2

208

Unfortunately, the producer country of the disposable latex gloves was only listed on four

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 28

209

samples, but the two samples which featured the minor furan-FA were both produced in

210

Malaysia.

211

Amount of 9M5, 1, in disposable latex gloves. Surprisingly, the GC/FID response of

212

9M5-ME (injected in the same concentration of 2 mg/mL) was considerably lower than those

213

of conventional FAMEs (Figure 4). GC/FID measurements on two instruments showed that

214

the response factor of 9M5-ME was only 40% of the response of conventional FAMEs (16:0-

215

ME, 18:0-ME, 18:1∆9-ME, 18:2∆9,12-ME), which had very similar GC/FID response factors

216

(all >85% compared to 16:0). Standard deviations for the calculated response factor for 9M5-

217

ME were 0.05,

243

ANOVA) was observed for the ratio of 18:2∆9,12 to 18:0 within samples of producers A (p =

244

0.068), B (p = 0.12), C (p = 0.33) and D/E (p = 0.08), for the ratio of 18:2∆9,12 to 18:1∆9 in

245

samples of producer C (p = 0.16) and for gloves of single producers (p = 0.17) as well as for

246

the ratio of 18:2∆9,12 to 16:0 for producer B (p = 0.69) and for gloves of single producers (p

247

= 0.06). Although no general statistical significance was found, the data strongly indicated

248

that the content of 9M5, 1, in the disposable latex gloves was lower and more variable than in

249

fresh liquid latex.16,17,22

250

Processing of crude latex and/or storage could reduce the content of 9M5, 1, but not

251

the content of the more stable conventional fatty acids. For instance, Englert et al.17 noted that

252

9M5, 1, was instable at pH 12 (~30% transformed in 12 h). Thus, partial post-harvest

253

transformation of 9M5, 1, at pH 10 in technical (stabilized) liquid latex could not be excluded.

254

Liengprayoon et al.23 noted that some lipids (saturated and unsaturated fatty acids) are

255

retained in the dry rubber and varied shares of them may have an effect on the plasticizing

256

properties of the product. Hence, 9M5, 1, could have an effect on the product quality of latex.

257

Next to the pH value, the crucial processes could be varying exposure of 9M5, 1, to light.

258

Further investigations were therefore carried out to study the stability of 9M5, 1, in disposable

259

latex gloves when exposed to natural sunlight. 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 28

260

Exposure of disposable latex gloves to light. The stability of 9M5, 1, in disposable latex

261

gloves when exposed to daylight inside a room behind a window for 1-4 days (24-96 h) was

262

studied with two types of samples in which 9M5, 1, initially contributed with 68% (sample

263

#30) (Table 1) and 53% to the total fatty acids (sample #13) (Table 1). The contribution of

264

9M5, 1, to the total fatty acids decreased every 24 h (Figure 6). Due to the changing

265

meteorological conditions during the exposure time (Table S1, Supporting Information) a

266

kinetic evaluation was not possible. For instance, higher decomposition rates in the first 24

267

hours (28% and 23%, respectively) (Figures 6A and 6B) were likely due to the higher light

268

intensity on day 1. Nevertheless, our data verifies the sensitivity of 9M5, 1, when exposed to

269

light. Varied exposure to light during production and storage of the disposable latex gloves

270

could also partly explain the varying contribution of 9M5, 1, to the fatty acids in individual

271

samples, which was lower than in the raw material (~90%).16,17,22

272

Additional peaks were not detected in the GC/FID chromatograms of the extracts after

273

exposure to sunlight. Thus, degradation products of 9M5, 1, may either be volatile as

274

suggested previously9,10 for other furan-FA and/or polar compounds that were not extracted or

275

accessible to GC analysis.

276

Isolation of 9M5-ME by silver ion chromatography. Separation of other FAMEs

277

from 9M5-ME in the extract of 2.5 g disposable latex gloves from sample #30 (Table 1) was

278

achieved with silver ion chromatography. Fraction 1 contained the methyl esters of saturated

279

fatty acids (16:0-ME, 18:0-ME and 20:0-ME) (Figure 7A), whereas the methyl esters of

280

unsaturated fatty acids (18:1∆9-ME and 18:2∆9,12-ME, 18:3∆9,12,15-ME) eluted into

281

fraction 3 (Figure 7C). Hence, 9M5-ME in fraction 2 could be obtained with a purity of ~98%

282

(Figure 7B). The yield was ~5 mg (~97% from the calculated content of 9M5, 1, of 5.1 mg in

283

2.5 g disposable latex gloves, Table 1).

284

Functionality of the method was confirmed by the extraction of 50 g disposable latex

285

gloves from sample #30 (Table 1) (corresponding with ~12 gloves) which provided 101 mg 12 ACS Paragon Plus Environment

Page 13 of 28

Journal of Agricultural and Food Chemistry

286

9M5-ME after column chromatography. The yield was still very high (99% of the calculated

287

content of 9M5, 1, vs. 97% in the analytical batch) and the purity was ~99%.

288

Furan-FA analysis is currently hampered by the lack of authentic reference standards

289

for quantitation and method verification.24 Recently, we suggested the use of the ethyl ester of

290

9M5 (9M5-EE) as internal standard for the quantitative analysis of furan-FA in fish and butter

291

samples.2,15,18 This standard can easily be prepared when transesterification is performed with

292

ethanol instead of methanol. In this way, 82.1 mg 9M5-EE (purity ~99%) were obtained from

293

50 g latex from sample #30 (Table 1).

294

The presented data shows that disposable latex gloves contain high amounts of 9M5,

295

1, which can easily be extracted with n-hexane. For quality assurance reasons, disposable

296

latex gloves should not be used during analyses on furan-FA. Yet, the simple extraction and

297

isolation protocol described in this study allows access to sufficient amounts of 9M5-ME and

298

9M5-EE in high purity which can be used as internal standard for furan-FA quantitation. Only

299

5 mg 9M5-EE extracted and isolated from 2.5 g disposable latex gloves will be enough for the

300

use as internal standard in ~5,000 quantitative analyses on fatty acids.

301 302

AUTHOR INFORMATION

303

Corresponding Author

304

* (W.V.) E-mail: [email protected] Phone: +49 711 459 24016. Fax +49 711

305

459 309 24377.

306

Funding

307

No funding was obtained for performing this study.

308

Notes

309

The authors declare no competing interests.

310

13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 28

311

ABBREVIATIONS USED. 9M5, 9-(3-methyl-5-pentylfuran-2-yl)-nonanoic acid; 16:0,

312

palmitic acid; 18:0, stearic acid; 18:1∆9, oleic acid; 18:2∆9,12, linoleic acid; 18:3∆9,12,15, α-

313

linoleic acid; 20:0, arachidic acid; furan-FA, furan fatty acid; FAME, fatty acid methyl ester;

314

ISTD, internal standard; SIM, selected ion monitoring;

315

316

SUPPORTING INFORMATION. Response factors of stearic acid (18:0), oleic acid

317

(18:1∆9), linoleic acid (18:2∆9,12) and the furan fatty acid 9-(3-methyl-5-pentylfuran-2-yl)-

318

nonanoic acid (9M5), 1, compared to palmitic acid (16:0) measured on two different GC/FID

319

systems and weight reduction of the lipid extract of disposable latex gloves when evaporated

320

with air over the period of 10 minutes in two samples. This material is available free of charge

321

via the Internet at http://pubs.acs.org.

322 323

REFERENCES

324

(1)

325

acids responsible for the cardioprotective effects of a fish diet? Lipids, 2005, 40, 755–771

326

(2)

327

Determination of furan fatty acids in food samples. J. Am. Oil Chem. Soc., 2012, 89,1501–

328

1508

329

(3)

330

marine bacterium, Shewanella putrefaciens. Biochim. Biophys. Acta, Lipids Lipid Metab.,

331

1997, 1346, 253–260

332

(4)

333

systems. Phytochem Rev., 2016, 15, 121–127

Spiteller, G. Furan fatty acids: Occurrence, synthesis, and reactions. Are furan fatty

Vetter, W.; Laure, S.; Wendlinger, C.; Mattes, A.; Smith, A. W. T.; Knight, D. W.

Shirasaka, N.; Nishi, K.; Shimizu, S. Biosynthesis of furan fatty acids (F-acids) by a

Mawlong, I.; Sujith Kumar, M. S.; Singh, D. Furan fatty acids. Their role in plant

14 ACS Paragon Plus Environment

Page 15 of 28

Journal of Agricultural and Food Chemistry

334

(5)

Hannemann, K.; Puchta, V.; Simon, E.; Ziegler, H.; Ziegler, G.; Spiteller, G. The

335

common occurrence of furan fatty acids in plants. Lipids, 1989, 24, 296–298

336

(6)

337

fatty acids (F-acids). Lipids, 1988, 23, 1032–1036

338

(7)

339

occuring furan fatty acids. Biol. Pharm. Bull., 1996, 19, 1607–1610

340

(8)

341

Antioxidant effect of naturally occurring furan fatty acids on oxidation of linoleic acid in

342

aqueous dispersion. J. Am. Oil Chem. Soc., 1990, 67, 858–862

343

(9)

344

Z. Lebensm.-Unters. Forsch., 1998, 206, 108–113

345

(10)

346

Soc., 1997, 74, 323–326

347

(11)

348

degradation products in dried herbs and vegetables. Chimia. 2002, 56, 263–265

349

(12)

350

B.; Wei, D. W.; Cox, B. Rapid elevation in CMPF may act as a tipping point in diabetes

351

development. Cell Rep., 2016, 14, 2889–2900

352

(13)

353

Mykkanen, H.; Poutanen, K.; Schwab, U.; Uusitupa, M. CMPF does not associate with

354

impaired glucose metabolism in individuals with features of metabolic syndrome. PLoS One,

355

2015, 10, 1-8

Gorst-Allman, C. P.; Puchta, V.; Spiteller, G. Investigations of the origin of the furan

Okada, Y.; Kaneko, M.; Okajima, H. Hydroxyl radical scavenging activity of naturally

Okada, Y.; Okajima, H.; Konishi, H.; Terauchi, M.; Ishii, K.; Liu, I.-M.; Watanabe, H.

Masanetz, C.; Guth, H.; Grosch, W. Fishy and hay-like off-flavours of dry spinach.

Guth, H.; Grosch, W. Stability of furanoid fatty acids in soybean oil. J. Am. Oil Chem.

Sigrist, I. A.; Manzardo, G. G.; Amadò, R. Furan fatty acid photooxidative

Liu, Y.; Prentice, K. J.; Eversley, J. A.; Hu, C.; Batchuluun, B.; Leavey, K.; Hansen, J.

Lankinen, M. A.; Hanhineva, K.; Kolehmainen, M.; Lehtonen, M.; Auriola, S.;

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 28

356

(14)

Prentice, K. J.; Luu, L.; Allister, E. M.; Liu, Y.; Jun, L. S.; Sloop, K. W.; Hardy, A.

357

B.; Wei, L.; Jia, W. The furan fatty acid metabolite CMPF is elevated in diabetes and induces

358

beta cell dysfunction. Cell Metab., 2014, 19, 653–666

359

(15)

360

samples from the German market. J. Agric. Food Chem., 2014, 62, 8740–8744

361

(16)

362

brasiliensis latex. Lipids, 1978, 13, 905–907

363

(17)

364

Hevea brasiliensis latex employing the combined use of pH-zone-refining and conventional

365

countercurrent chromatography. J. Sep. Sci., 2016, 39, 490–495

366

(18)

367

lipids and the cholesteryl ester fraction of fish liver. Food Anal. Methods, 2016, 9, 459–468

368

(19)

369

acids in fish from a zero discharge aquaculture system. J. Nutr. Food Sci., 2016, 2, 8–14

370

(20)

371

(Capsicum annuum). J. Agric. Food Chem., 2016, 64, 6306–6311

372

(21)

373

homologous series. Chromatographia, 2001, 54, 511–517

374

(22)

375

Hevea brasiliensis latex. Phytochemistry, 2011, 72, 1902–1913

376

(23)

377

Development of a new procedure for lipid extraction from Hevea brasiliensis natural rubber.

378

Eur. J. Lipid Sci. Technol., 2008, 110, 563–569

Wendlinger, C.; Vetter, W. High concentrations of furan fatty acids in organic butter

Hasma, H.; Subramaniam, A. The occurrence of a furanoid fatty acid in Hevea

Englert, M.; Ulms, K.; Wendlinger, C.; Vetter, W. Isolation of a furan fatty acid from

Wendlinger, C.; Hammann, S.; Vetter, W. Detailed study of furan fatty acids in total

Vetter, W.; Ulms, K.; Wendlinger, C.; van Rijn, J. Novel non-methylated furan fatty

Krauss, S.; Hammann, S.; Vetter, W. Phytyl fatty acid esters in the pulp of bell pepper

Kállai, M.; Veres, Z.; Balla, J. Response of flame ionization detectors to different

Liengprayoon, S.; Sriroth, K.; Dubreucq, E.; Vaysse, L. Glycolipid composition of

Liengprayoon, S.; Bonfils, F.; Sainte-Beuve, J.; Sriroth, K.; Dubreucq, E.; Vaysse, L.

16 ACS Paragon Plus Environment

Page 17 of 28

Journal of Agricultural and Food Chemistry

379

(24)

Vetter, W.; Wendlinger, C. Furan fatty acids - valuable minor fatty acids in food. Lipid

380

Technol. 2013, 25, 7–10

17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

381

CAPTIONS TO FIGURES

382

Figure 1. Chemical structures of (9-(3-methyl-5-pentylfuran-2-yl)-nonanoic acid (9M5)), 1,

383

(3-carboxy-4-methyl-5-propyl-2-furanpropanoic acid), 2, (9-(3-methyl-5-propylfuran-2-yl)-

384

nonanoic (9M3)), 3, (9-(3,4-dimethyl-5-pentylfuran-2-yl)-nonanoic acid (9D5)), 4, and (11-

385

(3-methyl-5-pentylfuran-2-yl)-undecanoic acid (11M5)), 5.

Page 18 of 28

386 387

Figure 2. Extract of (A) the GC/FID chromatogram and (B) the GC/MS chromatogram

388

(system 1) of the lipid extract from disposable latex gloves, #28 (Table 1). All fatty acids

389

were measured as methyl esters.

390 391

Figure 3. Extract of (A) GC/MS-SIM chromatogram (system 1) and (B-D) mass spectra of

392

the furan fatty acids (9-(3-methyl-5-propylfuran-2-yl)-nonanoic (9M3)), 3, (9-(3,4-dimethyl-

393

5-pentylfuran-2-yl)-nonanoic acid (9D5)), 4, and (11-(3-methyl-5-pentylfuran-2-yl)-

394

undecanoic acid (11M5), 5, detected in disposable latex gloves from Malaysia, #9 and #22

395

(Table 1).

396 397

Figure 4. GC/FID chromatogram of a standard mix containing methyl esters of palmitic acid

398

(16:0), stearic acid (18:0), oleic acid (18:1∆9), linoleic acid (18:2∆9,12) and 9-(3-methyl-5-

399

pentylfuran-2-yl)-nonanoic acid (9M5), 1, in the same concentration (c = 2 mg/mL).

400 401

Figure 5. Share of fatty acids stearic acid (18:0), oleic acid (18:1∆9) and palmitic acid (16:0)

402

compared to linoleic acid (18:2∆9,12) in disposable latex gloves from different producers.

403

Producers D and E and those with only one sample (single gloves) were summed up in one

404

group each. 18 ACS Paragon Plus Environment

Page 19 of 28

Journal of Agricultural and Food Chemistry

405 406

Figure 6. Percentage contribution of 9-(3-methyl-5-pentylfuran-2-yl)-nonanoic acid (9M5), 1,

407

to the total fatty acids in disposable latex gloves exposed to daylight from (A) samples #30

408

and (B) #13 (Table 1) at different time points.

409 410

Figure 7. GC/MS chromatograms (system 1) in full scan mode after silver ion mini

411

chromatography of (A) fraction 1, (B) fraction 2 and (C) fraction 3 of the lipid extract of

412

disposable latex gloves, #30 (Table 1).

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 20 of 28

Table 1: Share of Lipid Extract per Glove, Contribution of (9-(3-methyl-5-pentylfuran-2-yl)nonanoic acid (9M5)), 1, of the Total Fatty Acids, and Amount of 9M5 in 30 Different Disposable Latex Gloves from 14 Different Producers.

Number

Product

1 A1 2 A2 3 A3 4 A4 5 A5 6 A6 7 A7 8 A8 9 B1 10 B2 11 B3 12 B4 13 B5 14 C1 15 C2 16 C3 17 C4 18 D1 19 D2 20 E1 21 E2 22 F1 23 G1 24 H1 25 I1 26 J1 27 K1 28 L1 29 M1 30 N1 Mean/Median:

Share of lipid extract per glove [%]

Contribution of 9M5 to the total fatty acids [%]

Amount of 9M5 [mg/g glove]

1.4 1.3 1.2 1.2 0.8 0.6 1.6 1.2 1.6 1.2 1.0 1.0 0.9 1.5 1.4 1.1 0.9 0.9 0.9 1.1 0.9 1.6 1.5 1.2 1.2 1.0 1.0 0.9 0.8 0.8 1.1/1.1

67.2 67.2 65.4 58.1 52.3 57.4 51.9 37.9 57.6 74.7 39.4 70.3 52.8 54.9 87.0 47.3 54.7 85.1 66.9 68.2 68.4 59.2 68.7 76.0 60.5 58.3 57.3 56.1 53.0 67.9 61.4/58.8

2.6 2.5 1.9 1.6 0.8 1.1 2.0 0.7 2.9 3.2 0.6 2.9 0.8 2.0 8.2 1.0 1.5 4.4 1.8 2.0 1.7 3.8 2.8 3.8 1.6 2.0 1.7 2.2 1.1 2.1 2.2/2.0

20 ACS Paragon Plus Environment

Page 21 of 28

Journal of Agricultural and Food Chemistry

Figure 1 

′ ′



9

1

M

5

2

3

4

5

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 28

Figure 2

22 ACS Paragon Plus Environment

Page 23 of 28

Journal of Agricultural and Food Chemistry

Figure 3 Irel

9M5 (1)

A

9M3 (3) 9D5 (4) 18.0

Irel

19.0

20.0

21.0

22.0

23.0

11M5 (5) 24.0

25.0

137

9M3 (3)

[min]

B

294 109

100

Irel

180

140

260

220

300

9D5 (4)

m/z

C 179

123

100

Irel

336

140

11M5 (5)

180

220

260

165

m/z

D

109

100

300

350

140

180

220

260

300

340

380

m/z

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 28

Figure 4

24 ACS Paragon Plus Environment

Page 25 of 28

Journal of Agricultural and Food Chemistry

share of fatty acid compared to 18:2 9,12 [%]

Figure 5 100 120

18:0 18:1∆9 18:1 16:0

80

60

40

20

0 A

B

C

D/E

single gloves

different producers

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 28

Figure 6

Contribution of 9M5 to total fatty acids [%]

80

60

A 68

49

40

40 34 20

22

0 0

24

48 hours in daylight [h]

72

96

Contribution of 9M5 to total fatty acids [%]

80

B

60 53 40 41 36 31

28

72

96

20

0 0

24

48 hours in daylight [h]

26 ACS Paragon Plus Environment

Page 27 of 28

Journal of Agricultural and Food Chemistry

Figure 7 Irel

Fraction 1

A

18:0

16:0

20:0 14.0 Irel

18.0

20.0

Fraction 2

14.0 Irel

16.0

22.0

24.0

B

9M5

16.0

Fraction 3

18.0

20.0

22.0

[min]

24.0

[min]

C

18:2∆9,12

18:1∆9

18:3∆9,12,15 14.0

16.0

18.0

20.0

22.0

24.0

[min]

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 28

TABLE OF CONTENTS GRAPHIC

28 ACS Paragon Plus Environment