Synephrine as a Specific Marker for Orange Consumption - Journal of

Taste-Active 3-(O-β-d-Glucosyl)-2-oxoindole-3-acetic Acids and Diarylheptanoids in Cimiciato-Infected Hazelnuts. Journal of Agricultural and Food...
0 downloads 0 Views 467KB Size
Article

Synephrine as a Specific Marker for Orange Consumption Matthias Bader, Tatjana Lang, Roman Lang, and Thomas Hofmann J. Agric. Food Chem., Just Accepted Manuscript • Publication Date (Web): 22 May 2017 Downloaded from http://pubs.acs.org on May 25, 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 24

Journal of Agricultural and Food Chemistry

1

Synephrine as a Specific Marker for Orange Consumption

2 3

Matthias Bader†,‡, Tatjana Lang†,‡, Roman Lang†,‡, and Thomas Hofmann†§

4 5 6



contributed equally

7 8



9

München, Lise-Meitner-Straße 34, D-85354 Freising, Germany,

Chair for Food Chemistry and Molecular Sensory Science, Technische Universität

10

§

11

85354 Freising, Germany.

Bavarian Center for Biomolecular Mass Spectrometry, Gregor-Mendel-Straße 4,

12 13 14 15 16 17 18 19

*

20

PHONE

+49-8161/71-2902

21

FAX

+49-8161/71-2949

22

E-MAIL

[email protected]

To whom correspondence should be addressed

23

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 24

24

ABSTRACT

25

To validate the suitability of synephrine, known as a highly abundant alkaloid in

26

oranges, as a dietary biomarker for orange consumption, a highly sensitive and

27

robust stable isotope dilution analysis (SIDA) as well as an ECHO method, using the

28

analyte itself as a pseudo-internal standard injected into the analysis run providing an

29

“echo peak” of the analyte, was developed to quantitate synephrine by LC-MS/MS in

30

citrus juices and human urine before and after ingestion of orange juice. A citrus juice

31

screening revealed high synephrine concentrations of 150 to 420 nmol/mL in orange

32

(n=11) and tangerine juices (n=2), while 20 to 100 times lower levels were found in

33

juice from grapefruit (n=14), lemon (n=5) and lime (n=4), respectively. Application of

34

the SIDA to quantitate synephrine in sulfatase/glucuronidase-treated urine samples

35

(n=10) after orange juice consumption showed an increase of synephrine from trace

36

levels (0.1 ±0.1 nmol/mL) in the 2-days washout phase to a maximum concentration

37

of 8.9 (±5.5) nmol/mL found 4 h after ingestion of orange juice. While proline betaine

38

has been recently reported as a dietary biomarker indicating the ingestion of any

39

citrus product and Chinese artichoke, respectively, synephrine may be used a

40

reliable additional biomarker with high specificity for orange and tangerine.

41 42

Keywords: Dietary biomarker, synephrine, ECHO quantitation, citrus, oranges

43 44 45

2 ACS Paragon Plus Environment

Page 3 of 24

Journal of Agricultural and Food Chemistry

46

INTRODUCTION

47

The beneficial impact of dietary patterns on human health is today investigated in

48

epidemiological and human intervention studies, mostly by using questionnaires to

49

assess types of food, quantities and regularity of intake. Although clear instructions

50

for the study subjects as well as physician involvement have been shown to improve

51

compliance to some extent, the subjective nature of self-reported dietary intake

52

assessment methods presents numerous challenges to obtaining accurate dietary

53

intake and nutritional status. Analytical monitoring of dietary biomarkers is considered

54

a suitable strategy to overcome this limitation as it holds promise to objectively

55

assess dietary consumption without the bias of self-reported dietary intake errors.

56

The analytical compliance control of study subjects is, however, still limited by

57

the lack of suitable biomarker molecules reflecting wider aspects of diet. For

58

example, N-methylpyridinium ions have been identified as a urinary dietary biomarker

59

indicating the intake of roasted coffee1 and the analysis of alkylresorcinols in plasma

60

was reported to be suitable for measuring the consumption of whole grains.2,3 Very

61

recently, quantitative high-throughput LC-MS/MS screening of a variety of foods and

62

beverages confirmed extraordinarily high levels of proline betaine (1; figure 1), also

63

known as stachydrin, in citrus fruits as well as in Chinese artichoke,4 thus confirming

64

earlier reports.5-7 While the intake of other foods did not significantly affect the urinary

65

excretion profiles, consumption of Chinese artichoke resulted in similarly high

66

abundance of 1 in urine as found for citrus juice.4 As elevated urinary levels of 1 were

67

found to be also indicative for the consumption of Chinese artichoke, proline betaine

68

cannot be considered a citrus-specific dietary biomarker.

69

In comparison, the alkaloid synephrine (2; figure 1), which has been found to

70

act on α-, β1-, β2- and β3-adrenoceptors,8 has been reported as a trace amine present 3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 24

71

in bitter oranges,8-10 sweet oranges,11-13 tangerines,12-14 and lemons,12,13 while other

72

foods seem not to contain significant amounts.15 This alkaloid has just recently been

73

reported in orange honey and has been proposed as an authenticity marker of

74

orange honey.16 While the (R)-configured synephrine is the naturally occurring

75

enantiomer in fruits, traces of the (S)-isomer have been detected in thermally

76

processed juices and marmalades.14 The isomeric phenylephrine, coined m-

77

synephrine (3), seems not to occur naturally in citrus fruits and was only found in a

78

few dietary supplements.17-21

79

The objective of the present study was to develop a precise, sensitive and

80

accurate LC-MS/MS method using an internal standard and an ECHO standard,22-24

81

respectively, to enable a fast quantitation of 2 and 3 in various citrus fruits and dietary

82

supplements, and to investigate its suitability as a candidate citrus-specific urinary

83

biomarker.

84 85 86

MATERIALS AND METHODS

87 88

Chemicals. Synephrine

(2),

phenylephrine

(3,

as

hydrochloride salt),

d3-

89

phenylephrine (d3-3), ammonium acetate, β-glucuronidase (from Helix pomatia, Type

90

HP-2, ≥ 100000 U/mL), sulfatase (from Helix pomatia, Type H-1, ≥ 10000 U/g), d4-

91

methanol and methanol (LC-MS grade) was purchased from Sigma-Aldrich

92

(Steinheim, Germany). Water for LC-MS separation was purified with an integral 5

93

system (Millipore, Schwalbach, Germany). Citrus samples and juices were obtained

94

from a local supermarket (Freising, Germany). Citrus fruit (~2 kg) were individually

4 ACS Paragon Plus Environment

Page 5 of 24

Journal of Agricultural and Food Chemistry

95

squeezed, combined and homogenized. Fresh juices were frozen at -20°C until

96

analysis.

97

Urine Collection After Food Ingestion. Urine samples (frozen at -80°C) were

98

taken from a human intervention study conducted in 2016.4 In brief, 27 healthy

99

volunteers (age 24-32 years), with no reports on food-specific allergic reactions and

100

drug intake, were asked to abstain from citrus products for at least two days (wash-

101

out phase) and, then, to collect a morning urine sample. Ten of the volunteers were

102

then asked to consume orange juice (250 mL), which was shown to contain 68 µmol

103

(11.5 mg) synephrine (2). At least three further urine samples were collected on that

104

day (day 0), then morning urine on the next day (day 1) and a sample the afternoon,

105

finally a morning urine sample each on day two and three. For each specimen, the

106

time-point of sampling was recorded to calculate the time passed after the ingestion

107

of the food to enable plotting of synephrine/creatinine (pmol/µmol) versus time after

108

ingestion (h). Individual data on urinary creatinine levels were taken from a recent

109

report.4 The participants were allowed their own individual eating habits but were

110

asked to abstain from any citrus product until after the final urine sample was

111

collected.

112

Quantitation of Synephrine (2) and Phenylephrine (3). Standard Solutions:

113

synephrine (2; 5 mg) and phenylephrine (3; 5 mg as hydrochloride salt) were

114

individually dissolved in d4-methanol (2 mL each) and aliquots (600 µL) used to

115

accurately determine the concentration by means of quantitative

116

spectroscopy.25 Aliquots of these stock solutions were combined to yield an analyte

117

stock with 100 nmol/mL of each analyte.

1

H NMR

118

Matrix Calibrations, Precision and Bias: The analyte stock was serially diluted

119

with water to yield working solution containing 100, 50, 25, 12.5, 6.25, 3.13, 1.56, 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 24

120

0.78, and 0.39 nmol/mL of synephrine and phenylephrine, respectively. All of these

121

dilutions (100 µL) were further spiked into matrix (blank urine or blank lemon juice,

122

900 µL each) to yield matrix standard with final analyte concentrations 10000, 5000,

123

2500, 1250, 1250, 625, 312.5, 156.3, 78.1, and 39.1 nmol/L in matrix. Aliquots of

124

these matrix calibrations standards (100 µL) were mixed with the internal standard

125

working solution of d3-phenylephrine (6000 nmol/L, 100 µL) and an aqueous

126

ammonium acetate solution (10 mmol/L, 800 µL). The standards were injected in

127

replicates (n=3) and calibration curves constructed from area ratios and

128

concentration ratios. Precision and accuracy values of the matrix standards (cf. table

129

1) are given from the back-calculated standards and calibration curve.

130

Quality control samples were prepared (triplicates) similarly in urine (2500 and

131

156 nmol/L) and lemon juice (2500 and 313 nmol/L). Precision was calculated as

132

relative standard deviation of the mean concentration of the replicates. Bias was

133

calculated as the ratio of the calculated and the nominal value.

134

Enzymatic Hydrolysis of Urine Samples: The enzyme suspension consisted of

135

sulfatase (100 mg) and glucuronidase (suspension, 0.1 mL) in water (1 mL). To

136

liberate synephrine from phase I/II conjugates, urine samples (100 µL) were mixed

137

with the enzyme suspension (50 µL) and incubated for 4 h at 37 °C. After hydrolysis,

138

the internal standard solution (6.0 mmol/L, 50 µL) was added. After vortexing (5 s),

139

acetonitrile/methanol (9/1, v/v; 800 µL) was added, the suspension mixed and

140

centrifuged (10 min, 4 °C, 13200 rpm). The supernatant was decanted into new

141

Eppendorf caps and dried under a stream of nitrogen at 37 °C. The samples were

142

dissolved in aqueous ammonium acetate (10 mmol/L, 50 µL) and analyzed by means

143

of LC-MS/MS.

144

Statistics. Data handling was done using Microsoft Excel 2016 and GraphPad

145

Prism versions 5.00 and 7.00 for Windows (GraphPad Software, La Jolla California 6 ACS Paragon Plus Environment

Page 7 of 24

Journal of Agricultural and Food Chemistry

146

USA, www.graphpad.comGraphpad Prism). Synephrine/creatinine ratio (pmol/µmol)

147

in spot urine samples were determined for the segments 0-2, 2-4, 4-8, 8-12, 12-24

148

and 24-36 h to calculate means and standard deviations. Synephrine/creatinine

149

values of morning urine prior to the food intervention study (blank) were subtracted

150

from the morning urines collected after orange juice consumption, followed by

151

Wilcoxon

152

synephrine/creatinine value in the morning urines taken before and after food

153

consumption.

signed

rank

test

to

evaluate

significant

differences

between

154

Instrumental Analysis with HPLC-MS/MS. The chromatographic system was

155

a Shimadzu Nexera X2 UPLC (Shimadzu, Duisburg, Germany), comprising an

156

Autosampler (SIL 30AC, kept at 15 °C), pumps (2×LC30AD), degasser (DGU 20

157

A5R), column oven (CTO 30A, kept at 40°C) and communication device (CBM 20A).

158

The UPLC was connected to an AB Sciex 5500 Qtrap mass spectrometer (Sciex,

159

Darmstadt, Germany) operating in positive electrospray mode. Analyst 1.6.2 was

160

used for instrument control. Curtain Gas was 40, collision gas “medium”, ion spray

161

voltage -5.5 kV, source temperature 500 °C, nebulizer gas 50 and heater gas 60.

162

Resolution was set to “unit”. Dwell time for each mass transition was 20 ms. At least

163

two mass transitions per compound were recorded. The samples were separated on

164

a Kinetex Phenyl-F5 column (1.7 µm, 100×2.1 mm, Phenomenex, Aschaffenburg,

165

Germany) with ammonium acetate in water (10 mmol/L; eluent A) and ammonium

166

acetate in methanol (10 mmol/L; eluent B) at a flow rate of 300 µL/min. After

167

injection, eluent B was increased from 3 – 80% in 3 min with a non-linear gradient

168

(curve 8) followed by 1 min of isocratic elution. The starting conditions were re-

169

established within 0.5 min and equilibration was 2.5 min prior to the next injection.

170

The column effluent was diverted to the MS source between1 – 5 min. 1 min after

171

sample injection, the ECHO standard (100 pmol/mL, 2 µL) was injected. 7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 24

172 173 174

RESULTS AND DISCUSSION

175 176

Method

Development

for

Quantitation

of

Synephrine

(2)

and

177

Phenylephrine (3). The method development involved the preparation of standard

178

solutions of the analytes synephrine (2) and phenylephrine (3) in deuterium oxide

179

and exact quantification of the concentration by quantitative 1H NMR to obtain stock

180

solutions which were diluted appropriately. Detection of the analytes was facilitated

181

by tuning the ion source and ion path parameters of the MS/MS system in positive

182

electrospray for optimized collision induced dissociation. One quantifier and two

183

qualifier mass transitions were recorded for each analyte (table 1).

184

The chromatographic separation of the stable isotope dilution analysis (SIDA)

185

using HPLC-MS/MS was performed on a phenyl-F5 column using d3-phenylephrine

186

(d3-3) as the internal standard for quantitation of analytes 2 and 3. In comparison,

187

quantitation was performed by means of the ECHO technique using purified

188

synephrine (2) as the pseudo-internal ECHO standard.23,24 A time interval of 1 min

189

after sample injection, a defined amount of the purified analyte (2) was put onto the

190

column by means of a second injection to navigate the ECHO peak, which served as

191

a pseudo-internal standard and was used to normalize the analyte signal, in close

192

proximity to the analyte, that is ~1.3 and ~0.4 min after synephrine (2) and

193

phenylephrine (3), respectively (figure 2). The total run-time was 6.5 min including

194

equilibration time and retention times were stable in both citrus juice and urine matrix

195

(cf. table 1). The method therefore was fast enough to handle the samples from the

196

orange juice drinking study in a reasonable time. 8 ACS Paragon Plus Environment

Page 9 of 24

Journal of Agricultural and Food Chemistry

197

Calibration curves were calculated and recovery experiments were performed

198

in analyte-free lemon-juice and urine to evaluate and compare the quantitation

199

methods using the SIDA and the ECHO method, respectively. The analysis of

200

analyte-free lemon juice spiked with 2 and 3 (figure 2, A) and authentic orange juice

201

(B) showed well shaped and resolved peaks of the analytes, the internal standard

202

and the ECHO standard, respectively. The lower end of the calibrated range of 2 was

203

78 nM in urine and 156 nM in juice matrix, and 156 nM and 78 nM in urine and juice

204

for 3, respectively (table 1), providing sufficient sensitivity for analysis of trace

205

amounts in urine and citrus juice within our studies. The replicate analysis of lemon

206

juice spiked with two different concentrations of the analytes 2 and 3 gave precision

207

values