Long-Chain Fatty Acids Elicit a Bitterness-Masking ... - ACS Publications

Sep 12, 2015 - taste bitter.2 In food processing, bitter tastes can also be generated ... for the bitter taste receptors, T2Rs.8−11 .... overnight a...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIV OF NEBRASKA - LINCOLN

Article

Long-chain fatty acids elicit a bitterness-masking effect on quinine and other nitrogenous bitter substances by formation of insoluble binary complexes Kayako Ogi, Haruyuki Yamashita, Tohru Terada, Ryousuke Homma, Akiko Shimizu Ibuka, Etsuro Yoshimura, Yoshiro Ishimaru, Keiko Abe, and Tomiko Asakura J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.5b03193 • Publication Date (Web): 12 Sep 2015 Downloaded from http://pubs.acs.org on September 18, 2015

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 39

Journal of Agricultural and Food Chemistry

TITLE: Long-Chain Fatty Acids Elicit a Bitterness-Masking Effect on Quinine and Other Nitrogenous Bitter substances by Formation of Insoluble Binary Complexes

AUTHOR INFORMATUON Kayako Ogi,†,‡ Haruyuki Yamashita,†,‡ Tohru Terada,† Ryousuke Homma,† Akiko Shimizu-Ibuka,§ Etsuro Yoshimura,



Yoshiro Ishimaru,



Keiko Abe,† and Tomiko

Asakura*,†



Department of Applied Biological Chemistry, Graduate School of Agricultural and Life

Sciences, The University of Tokyo, 1-1-1, Yayoi, Bunkyo-ku, Tokyo 113-8657, Japan, and §

Faculty of Applied Life Sciences, Niigata University of Pharmacy and Applied Life

Sciences, 265-1 Higashijima, Akiha-ku, Niigata 956-8603, Japan.

1

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

ABSTRACT: We have previously found that fatty acids can mask the bitterness of certain

2

nitrogenous substances through direct molecular interactions. Using isothermal titration

3

calorimetry, we investigated the interactions between sodium oleate and 22 bitter

4

substances. The hydrochloride salts of quinine, promethazine and propranolol interacted

5

strongly with fatty acids containing 12 or more carbon atoms. The 1H-NMR spectra of

6

these substances obtained in the presence of the sodium salts of the fatty acids in dimethyl

7

sulfoxide, revealed the formation of hydrogen bonds between the nitrogen atoms of the

8

bitter substances and the carboxyl groups of the fatty acids. When sodium laurate and the

9

hydrochloride salt of quinine were mixed in water, an equimolar complex was formed as

10

insoluble heterogeneous needle-like crystals. These results suggested that fatty acids

11

interact directly with bitter substances through hydrogen bonds and hydrophobic

12

interactions to form insoluble binary complexes, which mask bitterness.

13 14

KEYWORDS: fatty acids, bitter substances, bitterness-masking, binary complex, quinine

15

2

ACS Paragon Plus Environment

Page 2 of 39

Page 3 of 39

Journal of Agricultural and Food Chemistry

16

INTRODUCTION

17

Taste comprises of five basic modalities: sweetness, sourness, bitterness, saltiness,

18

and umami. Bitter and sour tastes are generally undesirable1 and in particular, bitter tastes

19

are instinctively avoided because toxic substances often taste bitter.2 However, some

20

physiologically beneficial species such as flavonoids, polyphenols, and pharmaceutical

21

agents often taste bitter.2 In food processing, bitter tastes can also be generated during

22

fermentation and heating.3,4 Therefore, Bitterness-masking needs to be considered in food

23

preparation and pharmaceutical manufacturing.

24

Several bitterness-masking methods have been developed such as adding other

25

tastants and flavors to suppress bitter tastes.5-7 Bitterness can also be masked by applying

26

antagonists for the bitter taste receptors, T2Rs.8-11 Coating and encapsulating are often used

27

in the pharmaceutical industry to mask the bitterness of drugs.12-20 Cyclodextrin can

28

incorporate various substances to form an inclusion complex.21-23 Amino acid derivatives

29

are one example of low molecular weight bitterness-masking compounds.24 Zinc can also

30

mask the bitterness of quinine, tetralone, and denatonium benzoate.25 In many cases,

31

however, the masking mechanisms of these compounds have not been elucidated. For food

32

processing, these compounds must be harmless so identifying safe bitterness-masking

33

agents, originating from foods, is a desirable objective.

34

In a previous study,26 we found that certain cheeses contained bitterness-masking

3

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

35

substances. Their masking effect was attributed to the presence of fatty acids which could

36

mask the bitterness of some substances through direct interactions. However, there was

37

some specificity in the observed effects; for example, the bitterness of quinine was

38

suppressed, while the bitterness of caffeine was not.

39

We would investigate the bitterness-masking mechanism of fatty acids by defining

40

the interactions between the sodium salts of fatty acids and bitter substances using

41

isothermal titration calorimetry (ITC) and nuclear magnetic resonance (NMR)

42

spectroscopy. Initially, we will use ITC to screen various bitter substances to determine

43

which of them interact with the sodium salts of fatty acids. Sodium oleate will be used for

44

the screening because it has been shown to have the strongest bitterness-masking activity

45

against quinine hydrochloride (QHCl) in the sensory test. 26 We will then determine the

46

strength and mode of interactions between various sodium salts of fatty acids and the

47

screened bitter substances using ITC. Finally, the 1H-NMR spectra of the bitter substances

48

will be obtained in the absence and presence of the sodium salts of fatty acids to elucidate

49

the interactions with the bitter substances.

50

The aim of this study is to elucidate which substructures of fatty acids and bitter

51

substances interact with each other, as well as their state after the interaction. These

52

findings will lead to clarification of the bitterness-masking mechanism on a molecular

53

basis. From this information, molecular models composed of fatty acids and bitter 4

ACS Paragon Plus Environment

Page 4 of 39

Page 5 of 39

Journal of Agricultural and Food Chemistry

54

substances can be constructed to elucidate the mode of interaction.

55

5

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 39

56

MATERIALS AND METHODS

57

Chemicals. These were obtained from commercial sources. The fatty acids, hexanoic,

58

octanoic, decanoic, linoleic, linolenic acids, sodium thiocyanate, taurine, acesulfame K,

59

saccharin sodium dehydrate, caffeine, thiamine hydrochloride, and amygdalin were

60

obtained from Wako Pure Chemical Industries Ltd. (Osaka, Japan); lauric acid,

61

N-phenylthiocarbamide, salicin, berberine hydrochloride monohydrate and quinine

62

hydrochloride dihydrate were obtained from Nacalai Tesque Inc. (Kyoto, Japan). The fatty

63

acids,

64

(±)-propranolol hydrochloride, promethazine hydrochloride, chloramphenicol, yohimbine

65

hydrochloride, colchicine, cromolyn disodium salt, limonin, and ouabain octahydrate were

66

obtained from Sigma-Aldrich Co. (Tokyo, Japan). Methyltryptophan and naringin were

67

obtained from Tokyo Chemical Industry Co., Ltd. (Tokyo, Japan). Acetic acid and

68

antipyrine were obtained from Kanto Chemical Co. Inc. (Tokyo, Japan), and glycerol

69

monolaurate was obtained from Chem-Implex International Inc. (Wood Dale, IL, USA).

70

Screening of Bitter Substances. ITC was performed using a MicroCal iTC200 instrument

71

(Malvern Instruments Japan Corp., Kobe, Japan). The fatty acids were neutralized and

72

solubilized by adding of an equimolar amount of sodium hydroxide. The interactions

73

between 22 bitter substances whose structures are shown in Supplemental Figure S1-1 and

74

sodium oleate were measured. The concentrations of the bitter substances and sodium

myristoleic,

palmitoleic,

oleic,

arachidonic

acids,

6

ACS Paragon Plus Environment

glycerol

monooleate,

Page 7 of 39

Journal of Agricultural and Food Chemistry

75

oleate were 1.5 mM and 0.5 mM, respectively. The compounds were dissolved in 5 mM

76

phosphate buffer (pH 7.0). The reference cell was filled with Milli-Q water (Millipore

77

Corp., Billerica, MA, USA). The bitter substances in buffer solutions were titrated into

78

sodium oleate dissolved in 5 mM phosphate buffer (pH 7.0) at a stirring rate of 1,000 rpm

79

and at 25 °C. Each titration was carried out with an initial injection volume of 0.4 µL

80

followed by 18 main injections of 2 µL each at 120 s intervals. The first titration (0.4 µL)

81

was excluded from the analysis. The data corresponding to the dilution of the ligand in the

82

buffer was subtracted from the titration data.

83

Measurement of Interactions between Fatty Acids and Bitter Substances. The

84

interactions between five bitter substances, QHCl, promethazine hydrochloride (PHCl),

85

propranolol hydrochloride (PrHCl), berberine hydrochloride (BHCl) and yohimbine

86

hydrochloride (YHCl), and 8 sodium salts of fatty acids (sodium hexanoate, sodium

87

octanoate, sodium decanoate, sodium laurate, sodium myristoleate, sodium oleate, sodium

88

linoleate, and sodium linolenate) and 2 monoacylglycerides (glycerol monolaurate and

89

glycerol monooleate) were measured. The concentrations of the fatty acids were 0.5 mM

90

and those of the bitter substances are given in Supplemental Table S1. The samples were

91

dissolved in 5 mM phosphate buffer (pH 7.0). Titrations were conducted as described

92

above.

93

The data were analyzed according to the “one set of sites” model provided in Origin 7

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

94

7.0 software for MicroCal iTC200. The binding constant (K) and enthalpy change (∆H) of

95

binding were obtained from the fitted curve. The entropy change (∆S) and free energy

96

change (∆G) of binding were obtained using the following equation:

97 98

∆G = ∆H − T∆S = −RT ln K, where R is the gas constant and T is the thermodynamic temperature.

99

Preparation of Crystals. QHCl (3 mmol, 1.08 g) was dissolved in 3 L of Milli-Q water. A

100

solution of sodium laurate (3 mmol, 0.67 g) in Milli-Q water (50 mL) was added to the

101

QHCl solution and was thoroughly stirred. The mixture was allowed to stand overnight at

102

room temperature. The precipitated crystals were filtered No. 5A (Toyo Rishi Kaisha Ltd.

103

Tokyo, Japan), rinsed with water and air-dried. Differential scanning calorimetry (DSC)

104

measurements were performed using a Shimazu DSC-60 (Kyoto, Japan) in the temperature

105

range from 5 to 60 °C at a heating rate of 5 °C/min under a nitrogen atmosphere.

106

NMR Spectroscopy. All NMR experiments except for solid-state NMR were conducted

107

on a Varian Inova 500 NMR spectrometer (500 MHz) at 25 °C (Varian Inc, Palo Alto, CA,

108

USA). The bitter substances (32 µmol; QHCl, PHCl, and PrHCl) were measured in the

109

absence and presence of sodium laurate. The samples were dissolved in 600 µL of

110

DMSO-d6. The 1H-NMR measurement conditions were as follows: number of data points:

111

16384; acquisition time: 1.75 s; delay time: 5 s; number of scans: 16. The NOESY (nuclear

112

Overhauser enhancement spectroscopy) measurement conditions were as follows: number 8

ACS Paragon Plus Environment

Page 8 of 39

Page 9 of 39

Journal of Agricultural and Food Chemistry

113

of data points: 512 and 1024 for F1 and F2, respectively; digital resolutions of F1 and F2:

114

5.50 Hz and 8.79 Hz, respectively; delay time: 1.5 s; mixing time: 1.0 s; number of scans:

115

16. Solid-state NMR was conducted using an ECA 500 (500MHz) (JEOL, Tokyo, Japan) at

116

0 °C to stop the crystals from melting, using 4 mm of CP/MAS (Cross Polarization/ Magic

117

Angle Spinning) probe at 10,000 rpm. The measurement was as follows: number of data

118

points: 1024; digital resolution: 49.2 Hz; delay time: 5.0 s; number of scans: 2048.

119

Molecular Modeling. Initial models of conformations that were consistent with the

120

NOESY data of the 3 bitter substances in the absence and presence of sodium laurate were

121

manually constructed using GaussView 5.0 (Gaussian Inc. Wallingford, CT, USA). The

122

geometries were optimized in vacuo at the B3LYP/6-31G (d) level of theory, then, the

123

geometries were optimized in DMSO at the same level of theory. The polarizable

124

continuum model (PCM) was used to calculate solvent effects in DMSO. All calculations

125

were performed using the Gaussian 09 software package.27

126

9

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

127

RESULTS

128

ITC Measurements. The interactions between 22 bitter substances and sodium oleate

129

were studied using ITC (Supplemental Figure S1-2). Five bitter substances, QHCl, PHCl,

130

PrHCl, BHCl, and YHCl were found to interact with sodium oleate. Notably, each of these

131

substances bears a nitrogen atom (Figure 1). The interactions between these five substances

132

and six fatty acids were then studied (Table 1). The change in Gibbs free energy (∆G) was

133

negative for all the interactions detected, meaning that these interactions occurred

134

spontaneously. The change in enthalpy (∆H) contributed to ∆G in all cases and the change

135

in entropy (∆S) contributed in some cases (Table 1). In these cases, the titrated solutions

136

became turbid upon injecting the bitter substances, indicating that insoluble complexes had

137

been formed between the fatty acids and the bitter substances. Of the five bitter substances

138

examined, BHCl and YHCl exhibited relatively weak interactions with the fatty acids

139

(Table 2). Therefore, only the three which had strong interactions with the fatty acids,

140

QHCl, PHCl, and PrHCl, were examined.

141

No interactions were detected between fatty acids with 10 or fewer carbon atoms

142

and the three bitter substances. All these three substances showed the greatest binding

143

constants, upon binding with sodium oleate, which increased as the number of carbon

144

atoms in the alkyl chain of the fatty acids increased. The binding constant also decreased as

145

the number of double bonds in the fatty acids increased (Table 2). These results suggested

10

ACS Paragon Plus Environment

Page 10 of 39

Page 11 of 39

Journal of Agricultural and Food Chemistry

146

that the carbon chain length and the number of double bonds in the fatty acids affected

147

these interactions.

148

To examine the effect of the carboxyl group in the fatty acids, the interactions

149

between the bitter substances and glycerol monolaurate and glycerol monooleate were also

150

measured. However, no interactions between the bitter substances and these glycerol esters

151

were detected (Table 2).

152

Overall, carbon chains of a certain length and the presence of carboxyl groups in the

153

fatty acids were necessary to induce interactions with the bitter substances.

154

NMR Measurements. ITC measurements proved that fatty acids interacted directly with

155

some bitter substances. To examine the mechanism of these interactions, the 1H-NMR

156

spectra of three bitter substances (QHCl, PHCl, and PrHCl) were measured in the absence

157

and presence of an equimolar amount of sodium laurate. As these bitter substances are

158

insoluble in water when sodium laurate is added, DMSO-d6 was used as the solvent

159

instead.

160

The three bitter substances contained a nitrogen atom (Figure 1). The signals of the

161

protons on the hydrogen atoms located near the nitrogen atoms was shifted upfield in the

162

presence of sodium laurate (H2, H6, and H8 of QHCl; H9, H10, H11, H12, and H13 of

163

PHCl; H1″ and H3 of PrHCl) (Figure 2). The changes in the chemical shift (∆δ) of these

164

protons were more than 0.3 ppm for QHCl and PHCl and more than 0.2 ppm for PrHCl 11

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

165

(Supplemental Table S2). In the presence of glycerol monolaurate, no significant changes

166

were observed ( ∆δ < 0.02 ppm) (Supplemental Table S3).

167

The 1H-NMR spectra of QHCl were also measured in the presence of sodium

168

decanoate (10:0), sodium octanoate (8:0) and sodium hexanoate (6:0) which caused the

169

signals corresponding to H2, H6, and H8 of QHCl (Supplemental Table S4) to be shifted

170

upfield (∆δ > 0.3 ppm). These results suggested that the hydrogen bond between the

171

nitrogen atom of the bitter substances and the carboxyl group of the fatty acid was formed

172

in DMSO.

173

In addition, the 1H-NMR spectra were measured after changing the molar ratio of

174

sodium laurate to QHCl (0, 0.2, 0.5 and 0.8). The changes in the chemical shifts of H2, H6,

175

and H8 (∆δ) were plotted as a function of the molar ratio of sodium laurate (Figure 3). As

176

the chemical shifts depended on the molar ratio of sodium laurate, the exchange between

177

the free and complexed states was faster than the chemical shift difference between the two

178

states.

179

ITC measurements revealed that the length of the carbon chain and the number of

180

double bonds in the fatty acids affected the binding constants, which suggested that the

181

alkyl chains contributed to the interactions. To examine where the hydrophobic interactions

182

occurred, the NOESY spectra of the bitter substances were measured in the absence and

183

presence of sodium laurate (Figure 4). 12

ACS Paragon Plus Environment

Page 12 of 39

Page 13 of 39

Journal of Agricultural and Food Chemistry

184

In the present study, there were three possible locations for hydrophobic interactions:

185

the first was between the fatty acids and bitter substances; the second between different

186

molecules of the bitter substances; and the third between different molecules of the fatty

187

acids. Although no cross-peaks were observed between sodium laurate and QHCl, some

188

were observed between H2′ and H8′ of QHCl in the presence of sodium laurate (Figure 4).

189

Because these protons cannot interact with each other within a single molecule of QHCl,

190

intermolecular hydrophobic interactions between two molecules of QHCl occurred in the

191

presence of sodium laurate. However, in the case of PHCl and PrHCl, no interactions

192

between bitter substances and fatty acids and between different molecules of bitter

193

substances were observed (Supplemental Figure S3 and Table S5). Detailed analysis of the

194

cross-peaks between carbon chains was impossible because many signals overlapped.

195

However, because the length of the carbon chain affected the strength of the interactions in

196

the ITC experiment, we concluded that hydrophobic interactions existed primarily between

197

the carbon chains in an aqueous environment. Based on the NOESY results, we

198

constructed molecular models of the binary complexes formed between the bitter

199

substances and sodium laurate in DMSO (Figure 5).

200

Crystal Analysis of Sodium Laurate and QHCl. Needle-like crystals were observed

201

upon mixing an aqueous solution of sodium laurate and QHCl (Figure 6a). These crystals

202

were composed of an equimolar amount of laurate and quinine, as indicated by 1H-NMR. 13

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

203

The DSC curve of the crystal revealed a wide exothermic peak (Figure 6b). From CP/MAS

204

NMR spectra, it appeared that C2, C6 and C8 in QHCl resonated at chemical shifts of 62.3,

205

56.9 and 66.3 ppm, respectively, whereas each of the signals split into three peaks, with a

206

broadened peak width when the crystal was formed with lauric acid (Supplemental Figure

207

S2). This indicates that at least three states were involved in the crystal. In addition, peak

208

broadening of the methyl and methylene carbon atoms in lauric acid upon formation of

209

crystal demonstrated an incoherent microcrystalline structure. X-ray diffraction analysis

210

was applied to the ground crystal samples using the radial distribution function 4πr2ρ (r) to

211

obtain structural information. The sum of the data for QHCl and lauric acid, measured

212

independently, was different from the data for the ground crystal sample, implying that

213

specific interactions had been formed in the complex. The crystal seems to be in an

214

intermediate state regarding its amorphosity and crystal style (data not shown). These

215

results suggested that the crystal was composed of various conformations. The bitterness of

216

the crystal was significantly less than that of QHCl alone using sensory testing (data not

217

shown).

218

14

ACS Paragon Plus Environment

Page 14 of 39

Page 15 of 39

219

Journal of Agricultural and Food Chemistry

DISCUSSION

220

In a previous report,26 it was shown that sodium oleate suppressed the bitterness of

221

QHCl. Exothermic interactions were observed when QHCl was added to the fatty acid,

222

associated with masking the bitterness of QHCl by direct interaction with the fatty acids.

223

However, the detailed mechanism of the bitterness-masking reaction has not been

224

elucidated. In the present study, we found that two types of interaction, hydrogen bonds

225

and hydrophobic interactions were necessary to mask the bitterness of some compounds.

226

Specifically, NMR revealed that hydrogen bonding occurred between the nitrogen atom of

227

the bitter substances (N1 of QHCl, N11 of PHCl, and N4 of PrHCl) and the carboxyl group

228

of the fatty acids in DMSO (Figure 2). In addition, the ITC experiments in water indicated

229

that the strength of the interaction between the bitter substance and the fatty acid depended

230

on the length of the carbon chain of the fatty acid, demonstrating the involvement of

231

hydrophobic interactions.

232

In aqueous solutions, the bitter substances and fatty acids dissociated as weak bases

233

and acids, respectively.28 In neutral pH solutions, the dissociation equilibrium of the

234

long-chain fatty acids (RCOOH) shifted to the left of the equilibrium reaction presented in

235

Equation (I), given that the pKa values of oleic acid, linoleic acid, and linolenic acid are

236

9.85, 9.24, and 8.28, respectively.28

15

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 39

,

237 238

However, the dissociation equilibrium of nitrogenous bitter substances (BN) such as

239

quinine, promethazine, and propranolol shifted to the left side of Equation (II) given that

240

the pKa values of quinine,29 promethazine30 and propranolol31 are 8.3, 8.6 and 9.2,

241

respectively.

242

BNH+

BN +

H+

(II) ,

243

The ∆H values for the interactions between the fatty acids and bitter substances were

244

in the range of 10 to 40 kJ/mol for in all cases (Table 1). These results suggested that

245

hydrogen bonds existed between the fatty acids and bitter substances. Two possible

246

mechanisms of hydrogen bond formation are proposed: one involves interaction between

247

RCOOH and BN that comprise a unit of the binary complex; and the other occurs between

248

BH+ and RCOO−. The former hypothesis can explain the observation that long-chain fatty

249

acids interact with the bitter substances whereas short-chain fatty acids do not. The latter

250

hypothesis can account for the ionic attraction between oppositely charged species.

251

Of the twenty two bitter substances examined, those that exhibited strong

252

interactions with the fatty acids contained secondary or tertiary amino groups. Although

253

some bitter substances that did not interact with the fatty acid also had nitrogen atoms,

254

most of these nitrogen atoms were present in amide groups or in conjugated systems. The

255

rest were present in primary amino groups which might be unsuitable for forming 16

ACS Paragon Plus Environment

Page 17 of 39

256

Journal of Agricultural and Food Chemistry

hydrophobic interactions with the carbon chain of the fatty acids.

257

Regarding hydrophobic interactions, the contribution of the carbon chains was

258

obvious because the length of the carbon chain and the number of double bonds affected

259

the binding constants (Table 2). The NOESY data indicated the possibility of a stacking

260

effect between the aromatic rings of QHCl. However, the T∆S values varied widely and

261

were not correlated with ∆G, especially with sodium linoleate (Table 1). This phenomenon

262

was perhaps due to the conformational variation of the fatty acids or the decreased

263

solubility of the mixture of hydrochloride salts of the bitter substances and the sodium salts

264

of the fatty acids.

265

Based on these results, we propose that the bitterness of substances can be masked

266

with fatty acids by the formation of a binary complex between these species through a

267

hydrogen bond and a hydrophobic interaction (Figure 7). Specifically, the aggregation of a

268

unit of binary complex of fatty acids and bitter substances generated large insoluble

269

complexes. The formation of these complexes made it more difficult for the bitter

270

substances to interact with the bitter taste receptors, thereby masking the bitterness.

271

Particularly in the case of QHCl and sodium laurate, both substances aggregated because

272

of hydrophobic interactions between the carbon chains, forming needle-like crystals

273

(Figure 7). As suggested by DSC, the crystal was probably composed of quinine laurate in

274

various conformations with stacking interactions between the aromatic rings of QHCl. 17

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

275

In the present study, fatty acids have been found to interact directly with bitter

276

substances through hydrogen bonds and hydrophobic interactions. Moreover, both types of

277

interaction played important roles in masking bitterness. This study has been described

278

how direct interactions between bitterness-masking substances and bitter substances can

279

suppress bitterness. The results of this study should lead to the development of new method

280

for masking bitterness, which may be directly applicable to food processing. More research

281

would be needed to investigate the taste and acceptability of food containing fatty acids

282

and nitrogen-containing bitter compounds.

283

From this study, we can propose that a possible mechanism of how fatty acids can

284

mask the bitterness of bitter substances. Both hydrogen bonds and hydrophobic

285

interactions are essential for forming the insoluble binary complexes composed of bitter

286

substances and fatty acids. Nitrogen atoms in the bitter substances and fatty acids with

287

chain lengths of 12 carbon atoms or more are necessary for forming the binary complexes.

288

18

ACS Paragon Plus Environment

Page 18 of 39

Page 19 of 39

Journal of Agricultural and Food Chemistry

289

ASSOCIATED CONTENT

290

Supporting Information The supporting information is available for free of charge from the journal

291 292

website at http://pubs.acs.org.

293

The structures of screened substances and the isothermal titration profiles, solid-state

294

NMR spectra of crystal, lauric acid and QHCl, 1H-NMR assignments of QHCl, PHCl and

295

PrHCl in the absence and presence of sodium laurate, 1H-NMR assignments of QHCl,

296

PHCl and PrHCl in the absence and presence of glycerol monolaurate, 1H-NMR

297

assignments of QHCl in the absence and presence of sodium decanoate, sodium octanoate

298

and sodium hexanoate, NOESY assignments of QHCl, PHCl and PrHCl in the absence and

299

presence of sodium laurate, NOESY spectra of PHCl and PrHCl in the presence of sodium

300

laurate.

301

AUTHER INFORMATION

302

Corresponding Author *

303 304

Phone/Fax: +81-3-5841-1879; E-mail: [email protected]

Author Contributions ‡

305 306 307

These authors contributed equally to this work.

Notes The Authors declare no competing financial interest. 19

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

308

ACKNOWLEDGEMENTS

309

This work was supported by a Grant-in-Aid for Scientific Research 20259013 to T.A

310

from the Ministry of Education, Culture, Sports, Science and Technology of Japan. This

311

work was supported by the Council for Science, Technology and Innovation

312

(CSTI), Cross-Ministerial Strategic Innovation Promotion Program (SIP),

313

“Technologies for creating next-generation agriculture, forestry and fisheries”

314

ABBREVATIONS USED

315

ITC, isothermal titration calorimetry; 1H-NMR, proton nuclear magnetic resonance;

316

NOESY, nuclear Overhauser enhancement spectroscopy; DMSO, dimethyl sulfoxide;

317

T2Rs, taste receptors type 2; K, binding constant; ΔH, enthalpy change of binding; ΔS,

318

entropy change of binding; ΔG, free energy change of binding; R, the gas constant; T, the

319

thermodynamic temperature; QHCl, quinine hydrochloride; PHCl, promethazine

320

hydrochloride; PrHCl, propranolol hydrochloride; BHCl, berberine hydrochloride; YHCl,

321

yohimbine hydrochloride; CP/MAS, Cross Polarization/ Magic Angle Spinning; DSC,

322

differential scanning calorimetry; PCM, polarizable continuum model

323 324

20

ACS Paragon Plus Environment

Page 20 of 39

Page 21 of 39

Journal of Agricultural and Food Chemistry

325

REFERENCES

326

1. Chandrashekar, J.; Hoon, M. A.; Ryba, N. J. P.; Zuker, C. S. The receptors and cells for

327

328 329

mammalian taste. Nature 2006, 444, 288–294.

2. Drewnowski, A.; Gomez-Carneros, C. Bitter taste, phytonutrients, and the consumer: a review. Am. J. Clin. Nutr. 2000, 72, 1424–1435.

330

3. Degenhardt, A. G.;.Hofmann, T. Bitter-tasting and kokumi-enhancing molecules in

331

thermally processed avocado (Persea americana MILL.). J. Agric. Food Chem,

332

2010, 58, 12906–12915.

333 334

4. Maehashi, K.; Huang, L. Bitter peptides and bitter taste receptors. Cell. Mol.Life Sci. 2009, 66, 1661–1671.

335

5. Nakamura, T.; Tanigake, A.; Miyanaga, Y.; Ogawa, T.; Akiyoshi, T.; Matsuyama, K.;

336

Uchida, T. The effect of various substances on the suppression of the bitterness of

337

quinine-human gustatory sensation, binding, and taste sensor studies. Chem.

338

Pharm. Bull. 2002, 50, 1589–1593.

339

6. Wilkie, L. M.; Capaldi Phillips, E. D.; Wadhera, D. Sodium chloride suppresses

340

vegetable bitterness only when plain vegetables are perceived as highly bitter.

341

Chem. Percept. 2014, 7, 10–22.

21

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

342

7. Squier, C. A.; Mantz, M. J.; Wertz, P. W. Effect of menthol on the penetration of

343

tobacco carcinogens and nicotine across porcine oral mucosa ex vivo. Nicotine Tob.

344

Res. 2010, 12, 763–767.

345

8. Slack, J. P.; Brockhoff, A.; Batram, C.; Menzel, S.; Sonnabend, C.; Born, S.; Galindo, M.

346

M.; Kohl, S.; Thalmann, S.; Ostopovici-Halip, L.; Simons, C. T.; Ungureanu, I.;

347

Duineveld, K.; Bologa, C. G.; Behrens, M.; Furrer, S.; Oprea, T. I.; Meyerhof, W.

348

Modulation of bitter taste perception by a small molecule hTAS2R antagonist. Curr.

349

Biol. 2010, 20, 1104–1109.

350

9. Roland, W. S. U.; Gouka, R. J.; Gruppen, H.; Driesse, M.; Buren, L., van; Smit, G.;

351

Vincken, J. P. 6-Methoxyflavanones as bitter taste receptor blockers for hTAS2R

352

antagonist. PLoS One 2014, 9, e94451.

353

10. Greene, T. A.; Alarcon, S.; Thomas, A.; Berdougo, E.; Doranz, B. J.; Breslin, P. A. S.;

354

Rucker, J. B. Probenecid inhibits the human bitter taste receptor TAS2R16 and

355

suppresses bitter perception of salicin. PLoS one 2011, 6, e20123.

356

11. Brockhoff, A.; Behrens, M.; Roudnitzky, N.; Appendino, G.; Avonto, C.; Meyerhof, W.

357

Receptor agonism and antagonism of dietary bitter compounds. J. Neurosci. 2011, 31,

358

14775–14782.

22

ACS Paragon Plus Environment

Page 22 of 39

Page 23 of 39

359 360

361 362

Journal of Agricultural and Food Chemistry

12. Bora, D.; Borude, P.; Bhise, K. Taste masking by spray-drying technique. AAPS PharmSciTech. 2008, 9, 1159–1164.

13. Joshi, S.; Petereit, H. Film coatings for taste masking and moisture protection. Int. J. Pharm. 2013, 457, 395–406.

363

14. Yan, Y.; Woo, J. S.; Kang, J. H.; Yong, C. S.; Choi, H. Preparation and evaluation of

364

taste-masked donepezil hydrochloride orally disintegrating tablets. Biol. Pharm.

365

Bull. 2010, 33, 1364–1370.

366

15. Khan, S.; Kataria, P.; Nakhat, P.; Yeole, P. Taste masking of ondansetron

367

hydrochloride by polymer carrier system and formulation of rapid-disintegrating tablets.

368

AAPS Pharm. Sci. Tech. 2007, 8, E127–E133.

369

16. Li, S. P.; Martellucci, S. A.; Bruce, R. D.; Kinyon, A. C.; Hay, M. B.; Higgins III, J. D.

370

Evaluation of the film-coating properties of a hydroxyethyl cellulose/hydroxypropyl

371

methylcellulose polymer system. Drug Dev. Ind. Pharm. 2002, 28, 389–401.

372 373

17. Elder, D. Pharmaceutical applications of ion-exchange resins. J. Chem. Educ. 2005, 82, 575.

23

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 39

374

18. Samprasit, W.; Akkaramongkolporn, P.; Ngawhirunpat, T.; Rojanarata, T.; Opanasopit,

375

P. Formulation and evaluation of meloxicam oral disintegrating tablet with

376

dissolution enhanced by combination of cyclodextrin and ion exchange resins. Drug

377

Dev. Ind. Pharm. 2014, 1–11.

378

19. Yewale, C. P.; Rathi, M. N.; Kore, G. G.; Jadhav, G. V.; Wagh, M. P. Formulation

379

and

development

of

taste

masked

fast-disintegrating

tablets

(FDTs)

of

380

Chlorpheniramine maleate using ion-exchange resins. Pharm. Dev. Technol. 2013,

381

18, 367–376.

382

20. Prajapati, S. T.; Patel, P. B.; Patel, C. N. Formulation and evaluation of sublingual

383

tablets containing Sumatriptan succinate. Int. J. Pharm. Investig. 2012, 2, 162–168.

384

21. Binello, A.; Cravotto, G.; Nano, G.; Spagliardi, P. Synthesis of chitosan-cyclodextrin

385

adducts and evaluation of their bitter-masking properties. Flavour Fragr. J. 2004,

386

19, 394–400.

387

22. Izutani, Y.; Kanaori, K.; Imoto, T.; Oda, M. Interaction of gymnemic acid with

388

cyclodextrins analyzed by isothermal titration calorimetry, NMR and dynamic

389

light scattering. FEBS J. 2005, 272, 6154–6160.

24

ACS Paragon Plus Environment

Page 25 of 39

390 391

Journal of Agricultural and Food Chemistry

23. Szejtli, J.; Szente, L. Elimination of bitter, disgusting tastes of drugs and foods by cyclodextrins. Eur. J. Pharm. Biopharm. 2005, 61, 115–125.

392

24. Tokuyama, E.; Shibasaki, T.; Kwabe, H.; Mukai, J.; Okada, S.; Uchida, T. Bitterness

393

suppression of BCAA solutions by L-ornithine. Chem. Pharm. Bull. 2006, 54,

394

1288–1292.

395

25. Keast, R. S. J.; Breslin, P. A. S Bitterness suppression with zinc sulfate and

396

Na-cyclamate: a model of combined peripheral and central neural approaches to

397

flavor modification. Pharm. Res. 2005, 22, 1970–1977.

398

26. Homma, R.; Yamashita, H.; Funaki, J.; Ueda, R.; Sakurai, T.; Ishimaru, Y.; Abe, K.;

399

Asakura, T. Identification of bitterness-masking compounds from cheese. J. Agric.

400

Food Chem. 2012, 60, 4492–4499.

401

27. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman,

402

J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato,

403

M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.;

404

Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.;

405

Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A. Jr.; Peralta, J. E.;

406

Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.;

25

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

407

Keith, T.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.;

408

Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.;

409

Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.;

410

Yazyev, O. ; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.;

411

Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich,

412

S.; Daniels, A. D.; Farkas, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J.;

413

Gaussian, Inc., Wallingford CT, USA, 2013.

414 415

28. Kanicky, J. R.; Shah, D. O. Effect of degree, type, and position of unsaturation on the pKa of long-chain fatty acids. J. Colloid Interface Sci, 2002, 256, 201-207.

416

29. Geiser, L.; Henchoz, Y.; Galland, A.; Carrupt, P.; Veuthey, J. Determination of pKa

417

values by capillary zone electrophoresis with a dynamic coating procedure. J. Sep.

418

Sci. 2005, 28, 2374–2380.

419 420

30. Magnussen, M. P. The effect of ethanol on the gastrointestinal absorption of drugs in the rat. Acta Pharmacol. Toxicol. 1968, 26, 130–144.

421

31. Krämer, S. D.; Braun, A.; Jakits-Deiser, C.; Wunderli-Allenspach, H. Towards the

422

predictability of drug-lipid membrane interactions: the pH-dependent affinity of

26

ACS Paragon Plus Environment

Page 26 of 39

Page 27 of 39

Journal of Agricultural and Food Chemistry

423

propranolol to phosphatidylinositol containing liposomes. Pharm. Res. 1998, 15,

424

739–744.

425

27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 28 of 39

426

FIGURE CAPTIONS

427

Figure 1. Chemical structures of five bitter substances showing exothermic interactions

428

with sodium oleate.

429

Figure 2. 1H-NMR spectra of (1) QHCl, (2) PHCl, and (3) PrHCl. a) Spectra of bitter

430

substances in the presence of sodium laurate; b) spectra of bitter substances alone; c)

431

spectra of bitter substances in the presence of glycerol monolaurate. Characteristic protons

432

are indicated by arrows. The broadening of the OH and perturbation of H9 in QHCl and

433

the broadening of OH and NH in PrHCl resulted from the exchange of OH or NH protons

434

with those in the carboxyl moiety in sodium laurate. *Signals of sodium laurate,

435

of glycerol monolaurate.

436

Figure 3. Titration curve corresponding to QHCl (H6, H8, and H2) upon addition of

437

sodium laurate. The change in the chemical shift (∆δ) in 32 µmol of QHCl is plotted as a

438

function of the molar ratio of sodium laurate (0, 0.2, 0.5, and 0.8).

439

Figure 4. NOESY spectra of QHCl in the presence of sodium laurate. 1H-NMR spectra of

440

QHCl/sodium laurate is displayed on X–Y axis. Characteristic cross-peaks are indicated by

441

red circles. *Signals of sodium laurate;

442

Figure 5. Structures of three bitter substances and their laurates



signals of aromatic protons in QHCl.

28

ACS Paragon Plus Environment



signals

Page 29 of 39

Journal of Agricultural and Food Chemistry

443

Carbon atoms are presented in gray, oxygen in red, sulfur in yellow, nitrogen in blue, and

444

chloride in light green.

445

Figure 6. (a) Needle-like crystals formed between sodium laurate and QHCl. (b) DSC

446

thermal curve of the crystal.

447

Figure 7. Possible mechanism of interaction between sodium laurate and QHCl. Sodium

448

laurate and QHCl interact weakly via step 1; dissociation equilibrium of QHCl as weak

449

bases and sodium laurate as weak acids. Hydrogen bonding and insoluble binary

450

complexes occur via step 2; For fatty acids with a chain length of 12 or more carbon atoms,

451

hydrogen bond formation between quinine and laurate. The complex model is presented in

452

the lower image. QHCl molecules are colored blue, and sodium laurate is shown in orange.

453

The binary complexes are connected to each other by hydrophobic interactions, and grow

454

in a single direction, forming of the needle-like crystals.

29

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 30 of 39

TABLES

Table 1. Thermodynamic parameters for association between five bitter substances and sodium salts of fatty acids. QHCl a

Lauric acid (12:1) Myristoleic acid (14:1) Palmitoleic acid (16:1) Oleic acid (18:1) Linoleic acid (18:2) Linolenic acid (18:3)

∆H d

b

PHCl

PrHCl

BHCl

YHCl

e

−25.5±3.5

−20.3±2.6

−16.7±2.3

-

-

−3.5±3.5

0.7±2.6

1.3±2.5

-

-

c

∆G

−22.0±0.1

−21.0±0.0

−18.0±0.1

-

-

∆H

−10.8±1.4

−12.6±0.7

-

-

-

T∆S

13.7±1.9

11.0±0.6

-

-

-

∆G

−24.5±0.5

−23.6±0.1

-

-

-

∆H

−28.0±2.6

−21.8±2.5

−24.4±1.4

-

-

T∆S

−0.3±2.8

5.2±3.0

1.3±1.5

-

-

∆G

−27.6±0.3

−27±0.6

−25.7±0.1

-

-

∆H

−17.7±0.4

−10.7±0.2

−6.4±0.1

−11.5±0.2

−10.0±0.5

T∆S

11.5±0.4

18.3±0.3

22.0±0.1

15.7±0.2

14.6±0.4

∆G

−29.2±0.0

−29.0±0.3

−28.4±0.3

−27.2±0.0

−24.6±0.2

∆H

−27.3±3.1

−21.0±2.3

−28±1.7

-

-

T∆S

1.3±3.1

7.5±2.4

−2.1±1.8

-

-

∆G

−28.6±0.1

−28.6±0.1

−25.8±0.1

-

-

∆H

−42.3±2.5

−30.4±3.2

−28.2±3.8

−34.5±0.8

−7.8±1.5

T∆S

−15.1±2.5

−2.6±3.2

−2.6±4.0

8.8±8.2

−17.4±1.7

∆G

−27.2±0.0

−27.8±0.0

−25.6±0.2

−25.6±0.1

−25.1±0.3

T ∆S

a

Enthalpy change ∆H (kJ/mol), bentropy change ∆S (kJ/mol・K), cGibbs free energy ∆G (kJ/mol), and dT thermodynamic temperature (K). e“-” means not determined by

calculation according to “one set of site” model using the titration curve. Mean and standard deviation values from the fitted data of three independent experiments.

30

ACS Paragon Plus Environment

Page 31 of 39

Journal of Agricultural and Food Chemistry

Table 2. Binding constant K (× ×103・M−1) derived from ITC profiles for the association between five bitter substances and nine sodium salts of fatty acids and two glycerol esters.

QHCl

PHCl

PrHCl

BHCl

YHCl

a

Octanoic acid (8:0)

-

-

-

-

-

Decanoic acid (10:0)

-

-

-

-

-

Lauric acid (12:0)

7.2±0.2

4.9±0.0

1.4±0.1

Myristoleic acid (14:1)

19.9±4.0

13.7±0.3

-

-

-

Palmitoleic acid (16:1)

69.2±6.9

54.8±14.0

32.2±2.2

-

-

Oleic acid (18:1)

129±1.2

123.0±14.3

94.8±9.1

59.5±4.8

20.4±1.5

Linoleic acid (18:2)

101.1±3.4

100.7±4.9

33.4±0.8

-

-

Linolenic acid (18:3)

60.4±0.7

74.2±0.5

31.2±2.2

31.1±1.0

25.8±3.6

Glycerol monolaurate

-

-

-

-

-

Glycerol monooleate

-

-

-

-

-

Hexanoic acid (6:0)

a

“-” means not determined by calculation according to the “one set of site” model using the

titration curve. Mean and standard deviation values obtained from the fitted data of three independent experiments

31

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

FIGURE GRAPHICS

Figure 1.

32

ACS Paragon Plus Environment

Page 32 of 39

Page 33 of 39

Journal of Agricultural and Food Chemistry

Figure 2.

33

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 3.

34

ACS Paragon Plus Environment

Page 34 of 39

Page 35 of 39

Journal of Agricultural and Food Chemistry

Figure 4.

35

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 5.

QHCl

Quinine Laurate

PHCl

Promethazine Laurate

PrHCl

Propranolol Laurate

36

ACS Paragon Plus Environment

Page 36 of 39

Page 37 of 39

Journal of Agricultural and Food Chemistry

Figure 6.

37

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 7.

38

ACS Paragon Plus Environment

Page 38 of 39

Page 39 of 39

Journal of Agricultural and Food Chemistry

GRAPHICS FOR TABLE OF CONTENTS

39

ACS Paragon Plus Environment