Quantitation of Minor Components in Mango Juice ... - ACS Publications

ACS2GO © 2019. ← → → ←. loading. To add this web app to the home screen open the browser option menu and tap on Add to homescreen...
0 downloads 0 Views 985KB Size
Subscriber access provided by LAURENTIAN UNIV

Article

Quantitation of Minor Components in Mango Juice with Band-Selective Excitation NMR Spectroscopy Shoraku Ryu, Masanori Koda, Takuya Miyakawa, and Masaru Tanokura J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.7b03336 • Publication Date (Web): 04 Oct 2017 Downloaded from http://pubs.acs.org on October 8, 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 29

Journal of Agricultural and Food Chemistry

1

Quantitation of Minor Components in Mango Juice with Band-Selective

2

Excitation NMR Spectroscopy

3

4

Shoraku Ryu, Masanori Koda, Takuya Miyakawa and Masaru Tanokura*

5

6

Laboratory of Basic Science on Healthy Longevity, Department of Applied Biological

7

Chemistry, Graduate School of Agricultural and Life Sciences, The University of Tokyo,1-1-1

8

Yayoi, Bunkyo-ku, Tokyo 113-8657, Japan

9

10

*Corresponding

author

(Tel:

11

[email protected])

+81-3-5841-5165;

Fax:

12

13

1

ACS Paragon Plus Environment

+81-3-5841-8023;

E-mail:

Journal of Agricultural and Food Chemistry

14

ABSTRACT: 1H NMR-based metabolic analysis of foods has been widely applied. However,

15

dynamic range problems frequently impede its application because foodstuffs are composed of

16

various organic compounds in a wide range of concentrations. Band-selective excitation 1H

17

NMR spectroscopy has been found to be a useful tool for observing the minor components in

18

foods. Since quantitative information is important for metabolic analysis of foods and for

19

complete metabolome data, quantitation with the band-selective excitation 1H NMR method

20

was carefully investigated in the present study. As a result, the concentrations of minor

21

components in mango juice of the ‘Carabao’ cultivar were successfully quantitated by

22

band-selective excitation 1H NMR using standard curves that exhibited good linearity. The

23

band-selective excitation 1H NMR techniques was therefore effective for determining the

24

concentrations of minor components in foods.

25

KEYWORDS: NMR, band-selective excitation, quantitation, mango juice

2

ACS Paragon Plus Environment

Page 2 of 29

Page 3 of 29

Journal of Agricultural and Food Chemistry

26

INTRODUCTION

27

During the past several decades, metabolic profiling has become important in food science for

28

identifying and quantitating a number of metabolites, belonging to various classes of compounds.

29

For metabolic profiling, various methods of component analysis are employed, such as liquid

30

chromatography-mass spectrometry (LC-MS), gas chromatography-mass spectrometry

31

(GC-MS), capillary electrophoresis-time-of-flight mass spectrometry (CE-TOF-MS), and

32

nuclear magnetic resonance (NMR) spectroscopy. Among them, 1H NMR spectroscopy is

33

perceived as one of the most robust methods for metabolic profiling due to its multiple

34

advantages, such as it being non-destructive, comprehensive, quantitative, and fast.1-3 In the past

35

ten years, this technique has become an efficient method for combining statistical analysis with

36

origin authentication, varietal differentiation and quality evaluation of foods without a separation

37

procedure.4-11

38

Foodstuffs are composed of various organic compounds in a wide range of concentrations.

39

Minor components at relatively low concentrations in foods frequently contribute to the food

40

characterization, such as geographic origins, botanical origins, and food processing. For example,

41

minor components in mango juice are useful for discriminating the cultivars.7 In honey and olive

42

oil minor components are used for identifying botanical and geographic origins, respectively.12–

43

16

Despite the importance of minor components in foods, they may not be detected in 1H NMR

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

44

spectra if they are in the baseline or below the lowest dynamic range of the A/D converter

45

because the major components (e.g., sugars, lipids, ethanol, and acetic acid) existing at relatively

46

high concentrations define the maximum of the dynamic range. This problem frequently

47

impedes the application of NMR in metabolic profiling of foods. In a previous study, we

48

investigated the band-selective excitation of 1H NMR as an efficient approach to overcoming the

49

dynamic range problem in NMR,17 and found it to be a useful tool for providing information

50

about minor components in mango juice and used it to successfully discriminate the cultivars of

51

mango fruits.7 In addition, there are several reports for the band-selective applications in the field

52

of NMR food analysis.12,18–22

53

A thorough metabolic profiling analysis of foods requires both the identification and

54

quantitation of the organic compounds because the concentration information is important for a

55

complete metabolome of foods. However, in a band-selective excitation 1H NMR spectrum, the

56

ratio between the signal intensities may not directly reflect the concentration ratio because of the

57

influence of relaxation times and coupling constants.21,23 There are very few studies focusing on

58

quantitative characteristics for band-selective excitation.22 For this reason, it is still unclear

59

whether band-selective excitation 1H NMR is suitable for quantitation of the minor components

60

in foods.

61

In the present study, the quantitative characteristics were statistically compared between

4

ACS Paragon Plus Environment

Page 4 of 29

Page 5 of 29

Journal of Agricultural and Food Chemistry

62

band-selective excitation 1H NMR and non-selective 1H NMR. Moreover, the technique was

63

successfully used to quantitate the minor components in mango juice of the ‘Carabao’ cultivar

64

and provide concentration data for those components.

65

66

MATERIALS AND METHODS

67

Materials and Sample Preparation. Mangos of the ‘Carabao’ cultivar were purchased at

68

a local market, and deuterium oxide (D2O, 99.9% D) was obtained from Shoko Co., Ltd. (Tokyo,

69

Japan). All other chemicals were of spectroscopic grade and were purchased from Wako Pure

70

Chemical Industries, Ltd. (Osaka, Japan) and Nacalai Tesque Inc. (Kyoto, Japan). Mango juice

71

was prepared according to a previous report6 and was stored at −20 °C until it was used for

72

NMR analyses. For the NMR analyses, 70 µL of D2O was added to 630 µL of mango juice, then

73

the sample was transferred to a 5 mm NMR tube (Tokyo Chemical Industry Co., Ltd., Tokyo,

74

Japan). A sealed capillary tube with 0.83 mM [2,2,3,3-d4] sodium 3-(trimethylsilyl) propanoic

75

acid (TSP-d4), 0.72 mM hydroxymethylfurfural (HMF) and 5.5 mM MnCl2 was coaxially

76

inserted into the NMR tube as an external reference for concentration determination. To

77

investigate the quantitative properties of band-selective excitation, alanine was chosen since it is

78

found in mango juice (the target sample in this study) and in many other foods. For quantitation

79

of mango juice components, histidine, shikimic acid, trigonelline, uridine and adenosine were

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

80

used as standard solutions dissolved in D2O at five concentrations (0.2, 0.5, 1, 2 and 5 mM).

81

These standard solutions (700 µL) were transferred to a 5 mm NMR tube for NMR

82

measurements.

83

NMR Spectroscopy. All NMR spectra were obtained at 20 °C on a Unity INOVA-500

84

spectrometer (Agilent Technologies, Santa Clara, CA). The water signal was suppressed by a

85

presaturation method. The selective excitation was achieved using DPFGSE24 combined with

86

Re-Burp.25 The pulse lengths of Re-Burp were 2.0 and 2.9 ms for covering 6.0−11.0 ppm and

87

−0.2 to 3.1 ppm, respectively. The spin-lattice relaxation times (T1) were measured by inversion

88

recovery experiments and the delay time (d1) was determined with the following equation.5

89

d1 ≥ 5 × T1 − aq

90

where aq is acquisition time. The acquisition parameters of the 1H NMR spectra were as

91

follows: spectral width, 8 000 Hz; number of data points, 32 k; acquisition time, 2.048 s; number

92

of scans, 128. The receiver gain was set at 18 dB for non-selective and 60 dB for band-selective

93

excitation 1H NMR. The measurements for alanine were performed in triplicate.

94

NMR Data Processing and Quantitation of Minor Components. Free induction

95

decades (FIDs) were converted to spectra using uniform processing and phasing parameters

96

(exponential apodization with 0.25 Hz line broadening). The baseline was carefully calibrated

97

manually. The signal of alanine at 1.47 ppm was chosen as the reference for chemical shifts,

6

ACS Paragon Plus Environment

Page 6 of 29

Page 7 of 29

Journal of Agricultural and Food Chemistry

98

which was the shift of the alanine signal when the signal of external TSP was set to 0 ppm in the

99

absence of MnCl2. The signals of HMF and TSP were used as concentration standards for the

100

low-field and high-field regions, respectively. Their integral values were set to 1 and the relative

101

integral values were calculated for all other observed signals. The standard curves, which

102

characterize the correlation between integral values and concentrations, were determined using

103

the standard solutions. Experimental concentrations of mango juice components were

104

determined by the integral value of each signal in the low-field region and the corresponding

105

standard curve according to the following equation.

106

Experimental Concentration =

107

(  )(





!" #"

 !" #"

) × Theoretical Concentration''

108

where Integral and Integralstandard represent the integral value of each signal and methyl proton

109

signal of standard compounds (HMF or TSP), respectively, protons and protonsstandard are the

110

number of protons constituting each signal and the methyl proton signal of the standard

111

compounds (HMF or TSP), respectively, and Theoretical Concentrationstandard is 0.83 mM. The

112

concentrations of the minor components appearing in the low field region were determined

113

using the standard curves. The components observed in the high field region, such as choline,

114

GABA, citric acid, quinic acid, and alanine, were quantitated by non-selective excitation 1H

115

NMR spectra.

116

Statistical Analysis. The linearity of standard curves was evaluated with the coefficient of 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

117

determination R2, and the linear regression parameters (slope and y-intercept) were calculated.

118

According to guidance from the International Conference on Harmonisation (ICH),26 the

119

detection limit (DL) and quantitation limit (QL) were estimated using the following equations.

120

QL = 10σ/S

121

DL = 3.3σ/S

122

where σ represents the standard deviation of the response calculated using regression analysis,

123

and S is the slope of the calibration curve.

124

125

RESULTS AND DISCUSSION

126

Quantitation of Alanine as a Standard Sample. The standard curves for quantitation

127

were generated with the alanine solution. The spectra and the integrals of the β-proton signals

128

with band-selective excitation and non-selective excitation are shown in Figure 1. The relative

129

integral value of β-proton signals (1.47 ppm) was calculated in comparison to the value of the

130

TSP signal in 5.5 mM MnCl2 (0.355 ppm), which was set to 1 (Figure 1). MnCl2 was added to

131

shorten the relaxation time and recycling delay. The half-width of the methyl proton signal of

132

TSP-d4 increased by 11.5 times in the presence of 5.5 mM MnCl2. The standard curve for

133

band-selective excitation 1H NMR exhibited good linearity based on the coefficient of

134

determination, with R2 values of 0.9997, which is almost the same value as that of the

8

ACS Paragon Plus Environment

Page 8 of 29

Page 9 of 29

Journal of Agricultural and Food Chemistry

135

non-selective 1H NMR (R2 = 0.9999) (Figure 1B). However, the ratio of integral value to

136

concentration (Int/Con slope) for those measured with the non-selective and the band-selective

137

1

138

indicating an excellent consistency between the theoretical and the experimental concentrations.

139

In contrast, the Exp Con/Theo Con slope was calculated as 1.41 indicating a worse consistency

140

in quantitation with band-selective excitation 1H NMR. Because the intensity of 1H NMR

141

signals should be proportional to the absolute quantity of hydrogen atoms, these results indicated

142

that the signal intensity was attenuated during band-selective excitation, which leads to an

143

inadequate spectrum for direct quantitative analysis. On the other hand, the high coefficient of

144

determination (R2) showed that band-selective excitation 1H NMR can be reliably utilized to

145

quantitate components by plotting standard curves for the individual protons.

H NMR were different. The Exp Con/Theo Con slope for the non-selective 1H NMR was 1.02,

146

According to the ICH guidance, we calculated DL and QL using σ values and the slope of

147

standard curve and compared the results between the non-selective and the band-selective

148

excitation 1H NMR spectra. The σ values of the series of three measurements calculated with

149

regression analysis were 0.028 and 0.020 mM for non-selective and band-selective excitation 1H

150

NMR, respectively. The DL and QL values were 0.241 and 0.729 mM for non-selective 1H

151

NMR spectra, respectively. On the other hand, the DL and QL values for band-selective

152

excitation were 0.120 and 0.379 mM, respectively. These results showed that band-selective

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

153

excitation 1H NMR could have 2-fold higher sensitivity, which indicates it could be used to

154

detect and quantitate components that are half as concentrated as components non-selective 1H

155

NMR could detect.

156

The Influence of Relaxation Delay Time on the Quantitation. For precise quantitation

157

with 1H NMR, a long relaxation delay times (d1) are usually used to allow the equilibrium

158

magnetization state to be re-established before the next pulse cycle begins. Consequently, the

159

experimental time is considerably lengthened. To decrease the time needed for a quantitative

160

experiment, we examined whether the similar good linearity could be obtained with a short d1

161

when using standard curves.

162

The ε-proton signal of histidine (Hε) at 8.64 ppm was chosen for its long T1 value (7.75 s).

163

First, we plotted the standard curves with band-selective excitation with d1 value of 38 s

164

(>5×T1−aq); a 128-scan experiment with d1 = 38 s takes 113 minutes to run. The curve

165

displayed a linearity with the R2 value of 0.9993. Next, we compared that curve with the

166

standard curve obtained when the d1 value was set to 2 s, which is the typical measurement

167

condition for 1H NMR. A 128-scan experiment with a d1 = 2 s takes only 36 min to run. In the

168

standard curve with a 2 s delay time, the ratio of integral to concentration was obviously lower

169

than the ratio in the experiment with a long enough delay time (≥5×T1−aq) (Figure 1C) because

170

2 s is an insufficient relaxation time during scans. Despite the attenuation of the signal intensities,

10

ACS Paragon Plus Environment

Page 10 of 29

Page 11 of 29

Journal of Agricultural and Food Chemistry

171

a good linearity (R2 = 0.9953) was also seen in the standard curve in band-selective excitation 1H

172

NMR with a 2 s relaxation time. These results demonstrated that the shorter relaxation delay is

173

also applicable for quantitation with a specific standard curve. This method using shorter delay

174

times would help reduce the experimental time in the case of long T1 if the S/N ratios are

175

sufficiently high.

176

Quantitation of Minor Components of Mango Juice. Figure 2A shows the expanded 1H

177

NMR spectra of the mango juice from 5.8 to 10.0 ppm. Only a few signals were observed in this

178

region even with 1024 scans. When using band-selective excitation, the receiver gain could be

179

increased and many signals were then detected with just 256 scans (Figure 2B). In our previous

180

study, the minor components of mango juice were assigned with band-selective excitation 1D

181

and 2D NMR spectra.7 In the present study, five mango fruits of the ‘Carabao’ cultivar were

182

measured with band-selective excitation 1H NMR spectra and non-selective 1H NMR spectra,

183

which contained signals of many major components such as amino acids (histidine, alanine,

184

γ-aminobutyric acid (GABA), etc.), organic acids (citric acid, shikimic acid, and quinic acid),

185

alkaloids (trigonelline), nucleosides (adenosine and uridine), and others (choline, etc.) (Figure 3).

186

We quantitated the minor components of mango juice with band-selective excitation 1H NMR

187

spectra and other minor components with non-selective 1H NMR spectra. Considering their low

188

concentration in mango juice, the delay time was set to be > 5×T1−aq in all the experiments,

11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

189

where T1 was the longitudinal relaxation time of the histidine Hε proton, which has the

190

longest T1 observed in the spectrum. Although some amine protons from compounds such

191

as glutathione and glutamine were clearly observed in the low-field region, it was difficult

192

to quantitate those chemicals because of the exchangeability of their protons.27,28

193

HMF was used as the external standard for the down field region (6.0−11.0 ppm). The

194

standard curves were prepared for five standard chemicals (trigonelline, histidine, uridine,

195

shikimic acid, and adenosine). All chemicals showed good linearity for their standard curves

196

(Figure 5 and Table 1). The concentrations of the minor components were determined using the

197

standard curves. Because the signals of citric acid overlapped with other signals, we applied

198

curve fitting to calculate the signal intensity. The results of the quantitation are summarized in

199

Figure 5 and Table 2. Because the signals of shikimic acid were observed in both the down-field

200

(6.59 ppm) and up-field regions (2.76 ppm), the concentrations of shikimic acid calculated from

201

both the non-selective and band-selective excitation 1H NMR spectra of mango juice were

202

compared. The quantitative values were 6.1 ± 0.4 mM and 5.6 ± 0.6 mM (mean ± standard

203

deviation) for non-selective and band-selective excitation 1H NMR, respectively (Table 2 and

204

Figure 5), indicating that the results of quantitation using standard curves by band-selective

205

excitation 1H NMR are consistent with the results using non-selective 1H NMR spectra.

206

In conclusion, the present study has indicated that band-selective excitation 1H NMR

12

ACS Paragon Plus Environment

Page 12 of 29

Page 13 of 29

Journal of Agricultural and Food Chemistry

207

spectroscopy is an efficient technique for determining the concentrations of minor components

208

in complex mixtures using standard curves. Even in the case of a short delay time (for example,

209

2 s), the quantitative analysis can be carried out if the S/N ratio is sufficient. A previous report on

210

the metabolic profiling of mango juices suggested that the concentration variations of minor

211

components might cause the differences between groups rather than certain marker

212

components;7 however, it was unclear whether signal intensity had any correlation with actual

213

concentration of each component. The present study has elucidated the signal intensity is

214

linearly correlated with the concentration, which has revealed that band-selective excitation 1H

215

NMR spectra exactly provides quantitative information for metabolic profiling.

216

13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

217

ASSOCIATED CONTENT

218

Supporting Information

219

The Supporting Information is available free of charge on the ACS Publications website.

220

Calibration data for working curves of alanine; and calibration data for standard curves

221

of the Hε Proton of Histidine (8.64 ppm) (PDF)

222

223

AUTHOR INFORMATION

224

Corresponding Author

225

*Tel: +81-3-5841-5165; Fax: +81-3-5841-8023; E-mail: [email protected]

226

ORCD

227

Masaru Tanokura: 0000-0001-5072-2480

228

Notes

229

The authors declare no competing financial interest.

230

14

ACS Paragon Plus Environment

Page 14 of 29

Page 15 of 29

Journal of Agricultural and Food Chemistry

231

REFERENCES

232

(1) Wishart, D. S. Metabolomics: applications to food science and nutrition research. Trends

233

Food Sci. Technol. 2008, 19, 482‒493.

234

(2) Cevallos-Cevallos, J. M.; Reyes-De-Corcuera, J. I.; Etxeberria, E.; Danyluk, M. D.; Rodrick,

235

G. E. Metabolomic analysis in food science: a review. Trends Food Sci. Technol. 2009, 20, 557‒

236

566.

237

(3) Miyakawa, T.; Liang, T.; Tanokura, M. NMR-based metabolomics of foods. In Genomics,

238

Proteomics and Metabolomics in Nutraceuticals and Functional Foods, 2nd edition; Bagchi, D.,

239

Swaroop, A., Bagchi, M., Eds.; John Wiley & Sons, Ltd.; Hoboken, NJ, 2015; pp. 379‒387.

240

(4) Hu, F.; Furihata, K.; Ito-Ishida, M.; Kaminogawa, S.; Tanokura, M. Nondestructive

241

observation of bovine milk by NMR spectroscopy: Analysis of existing states of compounds and

242

detection of new compounds. J. Agric. Food Chem. 2004, 52, 4969‒4974.

243

(5) Hu, F.; Furihata, K.; Kato, Y.; Tanokura, M. Nondestructive quantification of organic

244

compounds in whole milk without pretreatment by two-dimensional NMR spectroscopy. J.

245

Agric. Food Chem. 2007, 55, 4307‒4311.

246

(6) Wei, F.; Furihata, K.; Koda, M.; Hu, F.; Miyakawa, T.; Tanokura, M. Roasting process of

247

coffee beans as studied by nuclear magnetic resonance: Time course of changes in composition.

248

J. Agric. Food Chem. 2012, 60, 1005‒1012.

15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

249

(7) Koda, M.; Furihata, K.; Wei, F.; Miyakawa, T.; Tanokura, M. Metabolic discrimination of

250

mango juice from various cultivars by band-selective NMR spectroscopy. J. Agric. Food Chem.

251

2012, 60, 1158‒1166.

252

(8) Hong, Y. S. NMR-based metabolomics in wine science. Magn. Reson. Chem. 2011, 49, S13‒

253

S21.

254

(9) Kodani, Y.; Miyakawa, T.; Komatsu, T.; Tanokura, M. NMR-based metabolomics for

255

simultaneously evaluating multiple determinatns of primary beef quality in Japanese Black

256

cattle. Sci. Rep. 2017, 7, 1297.

257

(10) Liang, T.; Wei, F.; Lu, Y.; Kodani, Y.; Nakada, M.; Miyakawa, T.; Tanokura, M.

258

Comprehensive NMR analysis of compositional changes of black garlic during thermal

259

processing. J. Agric. Food Chem. 2015, 63, 683‒691.

260

(11) Wei, F.; Furihata, K.; Miyakawa, T.; Tanokura, M. A pilot study of NMR-based sensory

261

prediction of roasted coffee bean extracts. Food Chem. 2014, 152, 363‒369.

262

(12) Schievano, E.; Peggion, E.; Mammi, S. 1H nuclear magnetic resonance spectra of

263

chloroform extracts of honey for chemometrics determination of its botanical origin. J. Agric.

264

Food Chem. 2010, 58, 57‒65.

265

(13) Belton, P. S.; Colquhoun, I. J.; Kemsley, E. K.; Delgadillo, I.; Roma, P.; Dennis, M. J.;

266

Sharman, M.; Holmes, E.; Nicholson, J. K.; Spraul, M. Application of chemometrics to the H-1

16

ACS Paragon Plus Environment

Page 16 of 29

Page 17 of 29

Journal of Agricultural and Food Chemistry

267

NMR spectra of apple juices: discrimination between apple varieties. Food Chem. 1998, 61,

268

207‒213.

269

(14) Bertelli, D.; Lolli, M.; Papotti, G.; Bortolotti, L.; Serra, G.; Plessi, M. Detection of honey

270

adulteration by sugar syrups using one-dimensional and two-dimensional high-resolution

271

nuclear magnetic resonance. J. Agric. Food Chem. 2010, 58, 8495‒8501.

272

(15) Sacchi, R.; Mannina, L.; Fiordiponti, P.; Barone, P.; Paolillo, L.; Patumi, M.; Segre, A.

273

Characterization of Italian extra virgin olive oils using H-1-NMR spectroscopy. J. Agric. Food

274

Chem. 1998, 46, 3947‒3951.

275

(16) Beretta, G.; Caneva, E.; Regazzoni, L.; Bakhtyari, N. G.; Facino, R. M. A solid-phase

276

extraction procedure coupled to 1H NMR, with chemometric analysis, to seek reliable markers

277

of the botanical origin of honey. Anal. Chim. Acta 2008, 620, 176‒182.

278

(17) Bodenhausen, G.; Freeman, R.; Morris, G. A. Simple pulse sequence for selective

279

excitation in Fourier-transformation NMR. J. Magn. Reson. 1976, 23, 171‒175.

280

(18) Koda, M.; Furihata, K.; Wei, F.; Miyakawa, T.; Tanokura, M. NMR-based metabolic

281

profiling of rice wines by F2-selective total correlation spectra. J. Agric. Food Chem. 2012, 60,

282

4818‒4825.

283

(19) Dugo, G.; Rotondo, A.; Mallamace, D.; Cicero, N.; Salvo, A.; Rotondo, E.; Corsaro, C.

284

Enhanced detection of aldehydes in Extra-Virgin Olive Oil by means of band selective NMR

17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

285

spectroscopy. Phys. A 2015, 420, 258‒264.

286

(20) Papaemmanouil, C.; Tsiafoulis, C. G.; Alivertis, D.; Tzamaloukas, O.; Miltiadou, D.;

287

Tzakos, A. G.; Gerothanassis, I. P. Selective one-dimensional total correlation spectroscopy

288

nuclear magnetic resonance experiments for a rapid identification of minor components in the

289

lipid fraction of milk and dairy products: Toward spin chromatography? J. Agric. Food Chem.

290

2015, 63, 5381‒5387.

291

(21) Sandusky, P.; Raftery, D. Use of selective TOCSY NMR experiments for quantifying minor

292

components in complex mixtures: Application to the metabonomics of amino acids in honey.

293

Anal. Chem. 2005, 77, 2455‒2463.

294

(22) Rastrelli, F.; Schievano, E.; Bagno, A.; Mammi, S. NMR quantification of trace

295

components in complex matrices by band-selective excitation with adiabatic pulses. Magn.

296

Reson. Chem. 2009, 47, 868‒872.

297

(23) Hajduk, P. J.; Horita, D. A.; Lerner, L. E. Theoretical-analysis of relaxation during shaped

298

pulses. I. The effects of short T1 and T2. J. Magn. Reson., Ser. A 1993, 103, 40‒52.

299

(24) Hwang, T. L.; Shaka, A. J. Water suppression that works - excitation sculpting using

300

arbitrary wave - forms and pulsed - field gradients. J. Magn. Reson., Ser. A 1995, 112, 275‒279.

301

(25) Geen, H.; Freeman, R. Band - selective radiofrequency pulses. J. Magn. Reson. 1991, 93,

302

93‒141.

18

ACS Paragon Plus Environment

Page 18 of 29

Page 19 of 29

Journal of Agricultural and Food Chemistry

303

(26)

U.S.

Food

and

Drug

Administration.

Guidances

304

(http://www.fda.gov/Drugs/GuidanceComplianceRegulatoryInformation/Guidances/default.htm

305

) (18 September 2017).

306

(27) Hwang, T. L.; van Zijl, P. C. M.; Mori, S. Accurate quantitation of water-amide proton

307

exchange rates using the Phase-Modulated CLEAN chemical EXchange (CLEANEX-PM)

308

approach with a Fast-HSQC (FHSQC) detection scheme. J. Biomol. NMR 1998, 11, 221‒226.

309

(28) Keun, H. C.; Ebbels, T. M. D.; Antti, H.; Bollard, M. E.; Beckonert, O.; Scholotterbeck, G;

310

Senn, H.; Niederhauser, U.; Holmes, E.; Lindon, J. C.; Nicholson, J. K. Analytical

311

reproducibility in H-1 NMR-based metabonomic urinalysis. Chem. Res. Toxicol. 2002, 15,

312

1380‒1386.

313

19

ACS Paragon Plus Environment

(Drugs).

URL

Journal of Agricultural and Food Chemistry

314

FIGURE CAPTIONS

315

Figure 1. (A) The spectra and (B) standard curves of alanine for non-selective (black line, ▲)

316

and band-selective excitation (red line, ♦) 1H NMR spectra. (C) The influence of relaxation

317

delay (d1) on the quantitative nature of the integral of the signal at 8.64 ppm (Hε proton) of

318

histidine in standard samples. The inset circle indicates the 1H position used for quantitation. The

319

relaxation delay was 2 s (■) and 38 s (♦)

320

321

Figure 2. Expanded 1H NMR spectra of mango juice (5.8̶− ̶ 10 ppm) observed with (A)

322

non-selective excitation (number of scans, 1024; receiver gain, 18 dB) and (B) band-selective

323

excitation (number of scans, 256; receiver gain, 60 dB)

324

325

Figure 3. (A) Low-field region of the band-selective excitation 1H NMR spectra of mango juice

326

from five different fruits (‘Carabao’ cultivar). (B) High-field region of the non-selective 1H

327

NMR spectra of mango juice from five different fruits (‘Carabao’ cultivar).

328

329

Figure 4. Standard curves for minor components in the low-field region (6.0−11.0 ppm): ○,

330

adenosine; □, trigonelline; ∆, histidine; +, uridine; and *, shikimic acid.

331

20

ACS Paragon Plus Environment

Page 20 of 29

Page 21 of 29

Journal of Agricultural and Food Chemistry

332

333

Figure 5. The concentrations of minor components in mango juice (A) at concentrations lower

334

than 5 mM and (B) at concentrations higher than 5 mM. For shikimic acid. The labels “nonsel”

335

and “sel” represent the quantitative values determined by the signals of non-selective and

336

band-selective excitation 1H NMR spectra, respectively.

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 22 of 29

Table 1. Calibration Data for Minor Components in the Low-Field Region (6.0−11.0

ppm). Analyte

Int/Con slope Int/Con intercept

R2

T1

trigonelline (8.83 ppm)

3.03

0.032

1.0000

4.59

histidine (8.64 ppm) uridine (7.86 ppm)

1.42 1.65

0.075 0.047

0.9994 0.9997

7.75 1.45

shikimic acid (6.59 ppm) adenosine (6.08 ppm)

1.51 1.61

0.035 0.035

0.9996 0.9998

3.77 2.87

22

ACS Paragon Plus Environment

Page 23 of 29

Journal of Agricultural and Food Chemistry

Table 2. Concentrations of Minor Components in Mango Juice.

Component

Average concentration (SD*) (mM)

trigonelline

0.08 (0.02)

histidine

0.12 (0.03)

uridine

0.09 (0.01)

adenosine choline GABA

0.12 (0.01) 0.29 (0.04) 1.12 (0.22)

quinic acid alanine citric acid shikimic acid (nonsel) shikimic acid (sel)

2.63 (0.38) 0.92 (0.44) 36.5 (6.08) 6.11 (0.37) 5.63 (0.62)

*

SD stands for standard deviation.

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 1

24

ACS Paragon Plus Environment

Page 24 of 29

Page 25 of 29

Journal of Agricultural and Food Chemistry

Figure 2

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3

26

ACS Paragon Plus Environment

Page 26 of 29

Page 27 of 29

Journal of Agricultural and Food Chemistry

Figure 4

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 5

28

ACS Paragon Plus Environment

Page 28 of 29

Page 29 of 29

Journal of Agricultural and Food Chemistry

Table of Contents Graphic

29

ACS Paragon Plus Environment