Beef Protein-Derived Peptides as Bitter Taste Receptor T2R4 Blockers

This system is a completely automated taste analyzer equipped with seven sensors, BD, EB, JA,. 127. JG, KA, OA, and JE, based ... a Flexstation-3 micr...
1 downloads 4 Views 990KB Size
Subscriber access provided by Kaohsiung Medical University

Chemistry and Biology of Aroma and Taste

Beef Protein-Derived Peptides as Bitter Taste Receptor T2R4 Blockers Chunlei Zhang, Monisola Alashi, Nisha Singh, Kun Liu, Prashen Chelikani, and Rotimi E. Aluko J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.8b00830 • Publication Date (Web): 29 Apr 2018 Downloaded from http://pubs.acs.org on April 30, 2018

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 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 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.

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 42

Journal of Agricultural and Food Chemistry

Beef Protein-Derived Peptides as Bitter Taste Receptor T2R4 Blockers Chunlei Zhang,† Monisola A. Alashi,† Nisha Singh,‡,Φ Kun Liu,‡,Φ Prashen Chelikani,*,‡,Φ and Rotimi E. Aluko*,†,§,Φ



Department of Food and Human Nutritional Sciences, University of Manitoba, Winnipeg,

Manitoba, Canada R3T 2N2 ‡

Department of Oral Biology, University of Manitoba, Winnipeg, Manitoba, Canada R3E 0W2

§

Richardson Center for Functional Foods and Nutraceuticals, University of Manitoba, Winnipeg,

Manitoba, Canada R3T 2N2 Φ

Manitoba Chemosensory Biology (MCSB) research group, University of Manitoba, Winnipeg,

Manitoba, Canada R3E 0W2

Corresponding Authors *Phone: +1 204 474 9555. E-mail: [email protected] *Phone: +1 204 789 3539. E-mail: [email protected]

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

1

ABSTRACT: The aim of this work was to determine the T2R4 bitter taste receptor-blocking

2

ability of enzymatic beef protein hydrolysates and identified peptide sequences. Beef protein was

3

hydrolyzed with each of six commercial enzymes (alcalase, chymotrypsin, trypsin, pepsin,

4

flavourzyme, and thermoase). Electronic tongue measurements showed that the hydrolysates had

5

significantly (p < 0.05) lower bitter scores than quinine. Addition of the hydrolysates to quinine

6

led to reduced bitterness intensity of quinine with trypsin and pepsin hydrolysates being the most

7

effective. Addition of the hydrolysates to HEK293T cells that heterologously express one of the

8

bitter taste receptors (T2R4) showed alcalase, thermoase, pepsin and trypsin hydrolysates as the

9

most effective in reducing calcium mobilization. Eight peptides that were identified from the

10

alcalase and chymotrypsin hydrolysates also suppressed quinine-dependent calcium release from

11

T2R4 with AGDDAPRAVF and ETSARHL being the most effective. We conclude that short

12

peptide lengths or the presence of multiple serine residues may not be desirable structural

13

requirements for blocking quinine-dependent T2R4 activation.

14 15

KEYWORDS: bitterness, protein hydrolysates, bitter taste receptor, (T2R), peptides, quinine,

16

electronic tongue

17 18 19 20 21 22 23 2 ACS Paragon Plus Environment

Page 2 of 42

Page 3 of 42

Journal of Agricultural and Food Chemistry

24

INTRODUCTION

25

Human beings are naturally averse to bitter taste because of the non-pleasant oral sensation

26

coupled with the fact that some causative compounds are usually toxic and may be life-

27

threatening.1,2 Therefore, eliminating or reducing the bitter taste attribute of pharmaceutical and

28

food products could enhance organoleptic properties and consumer acceptance. Bitter taste is

29

sensed by 25 bitter taste receptors (T2Rs), which are responsible for human perception of

30

thousands of bitter-tasting substances.3-7 The genes that code for the T2R proteins, are referred to

31

as TAS2Rs (HUGO gene nomenclature). So far, only a few antagonists or blockers against

32

specific T2Rs have been identified to work at the receptor level to reduce bitterness intensity of a

33

limited number of food products8-13, and these compounds have been extensively reviewed

34

recently.14 Therefore, discovery of additional bitter blockers are required in order to expand

35

coverage of several other foods that require bitter taste masking, especially for improved

36

commercial success. A diversity of bitter taste-blocking compounds will also enable

37

determination of structural properties that enhance interactions with T2Rs to block activation by

38

bitter substances. However, due to the limited understanding of the mechanism of bitter taste

39

signal transduction, progress towards production and utilization of new blockers has been slow.

40

This is complicated by the fact that in addition to the oral cavity, T2Rs are also expressed in the

41

brain, skin, thyroid gland, large intestine, testis, nasal epithelium and human airway.11,15-19 These

42

extraoral T2Rs are believed to participate in various physiological functions and are considered

43

potential therapeutic targets for diseases such as those that affect the respiratory airway,

44

especially asthma-related dysfunctions.20-22

45

Bioactive peptides that are generated from enzymatic hydrolysis of food proteins are

46

increasingly gaining attention because they have been reported to possess multifunctional health-

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

47

promoting properties that include bitterness, antihypertensive, anti-inflammatory and anticancer

48

activities.23-26 However, there is limited information on bitterness-suppressing properties of food

49

protein hydrolysates and their constituent peptide chains. Previous studies have reported that

50

amino acids and peptides have the capacity to efficiently diminish bitter taste intensity. For

51

example, L-aspartyl-L-phenylalanine and L-ornithyl-L-alanine were reported to reduce the

52

bitterness of potassium chloride.27 In addition to protein hydrolysates, other compounds such as

53

simple nucleotides cytosine

54

demonstrated to cause a 40% and 60% reduction in bitterness of a 10 mM quinine solution,

55

respectively.28 However, quantitative data and the mechanisms of bitter taste suppression by

56

amino acids and peptides remain scarce. Beef proteins have been shown to generate (through

57

enzymatic hydrolysis) desirable flavor-promoting peptides.29,30 Thus, we envisaged that beef

58

protein could also be an excellent raw material to generate peptides with bitter taste-blocking

59

properties. Therefore, the aim of this work was to determine the T2R4 bitter taste receptor-

60

blocking ability of enzymatic meat protein hydrolysates followed by elucidation of the structural

61

and functional properties of the main peptides.

62

MATERIALS AND METHODS

monophosphate and 2-deoxyadenosine triphosphate were

63

Materials. Ground beef was purchased from a local market (Safeway, Winnipeg, Manitoba,

64

Canada). Chymotrypsin® (from bovine pancreas, EC 3.4.21.1), trypsin® (from porcine pancreas,

65

EC 3.4.21.4), Pepsin® (from porcine gastric mucosa, EC 3.4.23.1), Alcalase® (from fermentation

66

of Bacillus licheniformis, EC 3.4.21.62), and Flavourzyme® (from Aspergillus oryzae, EC

67

232.752.2) were all purchased from Sigma-Aldrich (St. Louis, MO, USA). Thermoase® (from

68

Bacillus stearothermophilus, EC 3.4.24.27) was a product of Amano Enzymes Inc. (Nagoya,

69

Japan). Electronic tongue instrument diagnostic solutions including hydrochloride (0.1 M HCl), 4 ACS Paragon Plus Environment

Page 4 of 42

Page 5 of 42

Journal of Agricultural and Food Chemistry

70

sodium chloride (0.1 M NaCl) and monosodium glutamate (0.1 M MSG) as well as the

71

calibration solution (1 M HCl) were purchased from Alpha M.O.S (Toulouse, France). Known

72

bitter score substances such as acetaminophen, caffeine monohydrate, quinine hydrochloride

73

(QHCl), leporamide hydrochloride and femotidine were purchased from MP Biomedicals (Solon,

74

OH, USA). Quinine and BCML (Nα,Nα-bis(carboxymethyl)-l-lysine) was from Sigma Aldrich

75

(St. Louis, MO, USA).

76

Preparation of beef protein hydrolysates (BPHs). Raw ground beef (approximately 250 g)

77

was evenly packed into aluminum foil plates, frozen at -20 °C for 24 h and then freeze-dried.

78

The freeze-dried beef was blended thereafter in a Waring blender to fine powder and defatted

79

repeatedly by mixing 100 g with 1 L food grade acetone. The mixture was continually stirred for

80

3 h in fume hood and decanted manually followed by two consecutive extractions of the

81

residues. The defatted beef (DB) was placed in aluminum foil plates and air-dried overnight in

82

the fume hood at room temperature. The DB was milled in the Waring blender into fine powder

83

and stored at -20 °C. DB was then mixed with water to prepare a (w/v) 5% suspension followed

84

by addition of an enzyme (1%, w/w, protein weight basis) to initiate protein hydrolysis. The

85

hydrolysis conditions (temperature and pH) of each enzyme were based on manufacturers’

86

instructions and literature information.31-33 For alcalase hydrolysis, the DB suspension was

87

heated to 55°C and adjusted to pH 8.0 using 2 M NaOH. The DB suspensions for trypsin,

88

chymotrypsin and thermoase hydrolysis were first heated to 37 °C and then adjusted to pH 8.0.

89

For flavourzyme and pepsin hydrolysis, the DB suspension was heated to 50 °C and 37 °C

90

followed by adjustment to pH 6.5 and 2.0, respectively. Each mixture was stirred continuously

91

for 4 h and the reaction terminated by heating at 95 °C for 15 min. The reaction mixtures were

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

92

thereafter centrifuged (3,270g at 4 °C) for 30 min and the resulting supernatants freeze-dried as

93

the BPHs and stored at -20 °C.

94

Determination of degree of hydrolysis (DH). The DH of various BPHs was determined by

95

the O-phthalic aldehyde (OPA) method, which was based on previous reports.34,35 The OPA

96

reagent, which was prepared fresh daily contained 6 mM OPA dissolved in 95% methanol and

97

5.7 mM DL-dithiothreitol in 0.1 M sodium tetraborate decahydrate with 2% (w/v) SDS. N-Gly-

98

Gly glycine solution was prepared as standard in 8 serial concentrations (0.05-0.4 mg/mL) while

99

DB and BPHs were prepared in distilled water at 0.25 mg/mL. Ten µL of the standard solutions,

100

DB or BPH were pipetted into microplate wells followed by addition of 200 µL of OPA reagent.

101

Absorbance of the standard and samples were then read at 340 nm and 37°C in a Synergy H4

102

multi-mode plate reader (Biotek Instruments, Winooski, Vermont, USA). The total amino groups

103

in the DB were determined using a sample that has been subjected to 6 M HCl hydrolysis at

104

110 °C for 24 h. The DH was calculated by the following equation: %‫ = ܪܦ‬ሾሺሺܰ‫ܪ‬ଶ ሻ஻௉ு ሻ − ሺܰ‫ܪ‬ଶ ሻ஽஻ /ሺሺܰ‫ܪ‬ଶ ሻ ்௢௧௔௟ − ሺܰ‫ܪ‬ଶ ሻ஽஻ ሻሿ ∗ 100

105

(NH2) BPH: Content of free amino groups in BPH

106

(NH2) DB: Content of free amino groups in DB

107

(NH2) Total: Content of free amino groups in acid hydrolyzed DB

108

Analysis of amino acid composition. The amino acid profiles of DB and BPH were

109

determined according the method of Bidlingmeyer,36 using an HPLC system to analyze amino

110

acid composition of samples that have been hydrolyzed with 6 M HCl for 24 h. The contents of

111

cysteine and methionine were measured after performic acid oxidation37 while tryptophan

112

content was determined after alkaline hydrolysis.38

113

Estimation of molecular weight distribution. Molecular weight distribution of BPHs was

6 ACS Paragon Plus Environment

Page 6 of 42

Page 7 of 42

Journal of Agricultural and Food Chemistry

114

determined based on the method described by He et al.,39 using an AKTA FPLC system (GE

115

Healthcare, Montreal, PQ) equipped with a Superdex Peptide12 10/300 GL column 154 (10 x

116

300 mm) and UV detector (λ = 214 nm). The column was calibrated using the following standard

117

proteins and an amino acid: cytochrome C (12,384 Da); aprotinin (6,512 Da); vitamin, 855 Da);

118

and glycine (75 Da). A 100 µL aliquot of the 5 mg/mL BPH sample (dissolved in 50 mM

119

phosphate buffer, pH 7.0 containing 0.15 M NaCl) was loaded onto the column and elution

120

performed at room temperature using the phosphate buffer at a flow rate of 0.5 mL/min. The

121

molecular weights (MW) of peptides in samples were estimated from a linear plot of log MW

122

versus elution volume of standards.

123

Estimation of bitter scores by electronic-tongue (e-tongue). Each BPH was dissolved in

124

distilled water to give 0.5, 1.0, 2.5, 5.0, and 10.0 mg/mL concentrations followed by filtration

125

first through a 0.45 µm filter disc and then a 0.2 µm filter. Bitter scores of filtered BPH solutions

126

(20 mL) were evaluated using the Astree II E-tongue system (Alpha M.O.S., Toulouse, France).

127

This system is a completely automated taste analyzer equipped with seven sensors, BD, EB, JA,

128

JG, KA, OA, and JE, based on the ChemFET technology (Chemical modified Field Effect

129

Transistor) to analyze liquid samples. Firstly, 0.01 M HCl was used to condition and calibrate

130

sensors and the reference electrode repeatedly until stable signals were obtained for all seven

131

sensors with minimal or no noise and drift. Secondly, diagnostic procedure was performed

132

repeatedly using 0.1 M HCl, NaCl, and MSG to ensure the sensors can identify distinctive tastes,

133

until the discrimination index achieved at least 0.94 on a principle component analysis map.

134

Thereafter, the partial least square (PLS) regression bitterness standard model was constructed

135

using several bitter taste compounds with known bitter taste scores determined from human

136

panelists, including 0.24 mM and 2.36 mM caffeine, 0.03 mM and 0.12 mM quinine, 0.44 mM

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

137

and 0.88 mM prednisolone, as well as 3.31 mM and 19.85 mM paracetamol (Figure S1).

138

Validation of the PLS model was achieved with 0.002 mM and 0.01 mM loperamide followed by

139

0.06 mM and 0.15 mM famotidine according to the manufacturer’s instructions.40 A series of

140

BPH concentrations (0.5 mg/mL, 1 mg/mL, 2.5 mg/mL, 5 mg/mL, and 10 mg/mL) were

141

prepared, filtered as indicated above and then used to determine the e-Tongue threshold. The 5

142

mg/mL sample was determined as the maximum strength useful to evaluate BPH bitterness

143

intensity and this concentration was subsequently used to obtain bitter scores. In order to

144

evaluate the bitterness suppressing ability of protein hydrolysates, the T2R4 agonist quinine and

145

BPH solutions were mixed to obtain 1 mM and 5 mg/mL concentrations, respectively. As a

146

positive control, the known T2R4 bitter taste blocker Nα,Nα-bis(carboxymethyl)-l-lysine (BCML)

147

with an IC50 of 59 nM for quinine were mixed together to obtain 59 nM (BCML) and 1 mM

148

(quinine) final concentrations, respectively. For all samples, triplicate analysis of each solution

149

was performed.

150

Determination of cellular calcium mobilization. Determination of the potential bitter taste-

151

activating or blocking activity of BPH was carried out by measuring intracellular Ca2+

152

mobilization using Fluo-4 NW calcium assay kit as previously described.3 Stable transfected

153

HEK293 cells expressing TAS2R4 and G-alpha 16/44 or HEK293T cells expressing only G-

154

alpha 16/44 were used as experimental and negative control group, respectively. Approx. 1×105

155

cells/ well plated in the 96-well BD-falcon biolux plate and then incubated at 37 °C in a CO2

156

incubator for 16 h. After incubation, the culture medium was substituted (for 40 min at 37 °C in

157

the CO2 incubator followed by 30 min incubation at room temperature) with Fluo-4 NW dye

158

solution, which contained lyophilized dye in 10 mL of assay buffer (1x Hanks’ balanced salt

159

solution, 20 mM HEPES) and 100 µL 2.5 mM probenecid added to prevent dye leakage from

8 ACS Paragon Plus Environment

Page 8 of 42

Page 9 of 42

Journal of Agricultural and Food Chemistry

160

cytosol. Calcium mobilization was measured in terms of relative fluorescence units (RFUs) using

161

a Flexstation-3 microplate reader (Molecular Devices, CA, USA) at 525 nm, following 494 nm

162

excitation. Based on the e-tongue data, BPH at 5 mg/mL, BCML at 59 nM and agonist quinine at

163

1 mM were used individually or in combination with peptides to determine T2R4 activation.

164

BPH or BCML was then mixed with quinine (1 mM final concentration) to obtain 5 mg/mL or

165

59 nM, respectively and used to determine inactivation of T2R4. The basal intracellular calcium

166

levels were measured for the first 20 s followed by addition of appropriate concentration of

167

ligands for another 120 s. To get the absolute RFUs, the basal RFU before adding the ligand,

168

which was labeled minimum value (Min) was deducted from the maximum RFU (Max) obtained

169

after stimulating with the ligand (absolute RFUs = Max – Min). Next, the signals from the

170

negative control group cells were subtracted from the observed signal of experimental group

171

cells to give ∆RFUs. Data were collected from two independent experiments, each done in

172

triplicate. PRISM software version 4.03 (GraphPad Software, San Diego) was used for data

173

analysis.

174

Separation of BPH by reversed-phase high-performance liquid chromatography (RP-

175

HPLC). Alcalase hydrolysate (AH) and chymotrypsin hydrolysate (CH), the two most active

176

bitter taste blockers were subjected to RP-HPLC separation on a 21 x 250 mm C12 preparative

177

column (Phenomenex Inc., Torrance, CA, USA) attached to a Varian 940-LC system (Agilent

178

Technologies, Santa Clara, CA, USA) according to the method of Girgih et al.41 Briefly, freeze-

179

dried AH or CH was dissolved in double distilled water that contained 0.1% trifluoroacetic acid

180

(TFA) as buffer A to give 100 mg/mL. After sequential filtration through 0.45 µm and 0.2 µm

181

filters, 4 mL of the sample solution was injected onto the C12 column. Fractions were eluted

182

from the column at a flow rate of 10 mL/min using a linear gradient of 0–100% buffer B

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

183

(methanol containing 0.1% TFA) over 60 min. Peptide elution was monitored at 214 nm, eluted

184

peptides were collected using an automated fraction collector every 1 min and pooled into four

185

fractions according to elution time. In this study, according to the distribution of peaks of

186

chromatograms, 4 fractions each were collected from AH (AH-F1, AH-F2, AH-F3, AH-F4) and

187

CH (CH-F1, CH-F2, CH-F3, CH-F4). The solvent in the pooled fractions was evaporated using a

188

vacuum rotary evaporator maintained at a temperature range between 35 and 45 oC and thereafter

189

the aqueous residue was freeze-dried. Fractionated peptides were analyzed for calcium

190

mobilization ability using the cell culture protocols described above. The most active fractions

191

(AH-F1 and CH-F4) were each subjected to a second round of RP-HPLC separation using

192

peptide load (400 mg), C12 preparative column, elution buffers, flow rate and detection

193

wavelength as used in the first round of separation. Elution was carried out with a linear gradient

194

of 0–35% buffer B in buffer A over 49 min. Eluted peptides were collected using an automated

195

fraction collector every 1 min and pooled into 4 AH-F1 fractions and 8 CH-F4 fractions. The

196

solvent in each fraction was removed under vacuum in a rotary evaporator and the aqueous

197

residues freeze-dried and used for calcium mobilization experiments as described above.

198

Peptide identification and sequencing. The most active fractions (AH-F1-3, CH-F4-3, CH-

199

F4-5) against T2R activation (calcium mobilization experiments) from the second RP-HPLC

200

separation peptide fractions were analyzed by tandem mass spectrometry. Briefly, a 10 ng/µL

201

aliquot of the sample (dissolved in an aqueous solution of 0.1% formic acid) was infused into an

202

Absciex QTRAP® 6500 mass spectrometer (Absciex Ltd., Foster City, CA, USA) coupled to an

203

electrospray ionization (ESI) source. Operating conditions were 5.5 kV ion spray voltage at 200

204

o

205

MS/MS spectra were analyzed using PEAKS 7.0 Studio software (Bioinformatics Solutions,

C, and 30 µL/min flow rate for 2 min in the positive ion mode with 2000 m/z scan maximum.

10 ACS Paragon Plus Environment

Page 10 of 42

Page 11 of 42

Journal of Agricultural and Food Chemistry

206

Waterloo, ON, Canada) to obtain peptide sequences. The identified peptides were chemically

207

synthesized (>95% purity) by Genscript Inc. (Piscataway, NJ, USA).

208

Statistical analysis. Data analyses were performed made using one-way analysis of variance

209

(ANOVA) with an IBM SPSS Statistical package (version 20). Mean values were compared

210

using the Duncan Multiple Range Test and significantly differences accepted at p < 0.05.

211

RESULTS AND DISCUSSION

212

Degree of hydrolysis. As shown in Table 1, the DH of the protein hydrolysates was enzyme-

213

dependent with the highest values for microbial enzymes (alcalase and flavourzyme). The high

214

DH for flavourzyme may have been due to the presence of endoproteases and exoproteases,42

215

which could have enhanced the rate of proteolysis. The results are similar to those previously

216

reported for bovine plasma proteins where the flavourzyme hydrolysates had higher DH values

217

than the alcalase hydrolysates.43 Similarly, Kane et al.44 reported a slightly higher DH for a

218

peanut hydrolysate produced from flavourzyme when compared to that of alcalase after

219

hydrolysis for 4 h. However, the results are different from those reported for tilapia muscle

220

protein hydrolysis where the flavourzyme produced lower DH than alcalase.34 An interesting

221

outcome is that the intestinal enzymes (trypsin and chymotrypsin) digested the meat proteins

222

more efficiently and produced hydrolysates with higher DH than the stomach enzyme, pepsin.

223

Amino acid composition. In comparison to the defatted beef protein, significant changes in

224

some of the amino acids were observed after enzymatic hydrolysis (Table 2). Glutamic acid +

225

glutamine (Glx) were the most abundant in the beef but there was no significant change in

226

content after protein hydrolysis. In contrast, the contents of aspartic acid + asparagine (Asx) and

227

arginine were significantly (p < 0.05) enhanced in the protein hydrolysates. However, the protein

228

hydrolysates had significantly reduced levels of cysteine, methionine and tryptophan. Kohl et 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

229

al.45 had previously shown that tryptophan is a T2R4 agonist. Therefore, protein hydrolysates

230

with lower tryptophan levels may provide a better source of weakly-acting T2R4 peptide

231

agonists or even antagonists when compared to those with higher levels. With respect to amino

232

acid groups, there were overall significant (p < 0.05) reductions in hydrophobic amino acids,

233

aromatic amino acids and sulfur-containing amino acids after the enzymatic protein hydrolysis.

234

But positively-charged amino acids were significantly (p < 0.05) higher in the protein

235

hydrolysates when compared to the defatted beef protein. A previous work has shown that the

236

peptide’s amino acid side groups may be more important than the amino and carbonyl groups

237

during T2R4 activation.45 Therefore, the results reflect the different proteolytic specificities of

238

the proteases used in this work, which provided a wide variation of peptides (different side

239

groups) that could function as T2R4 antagonists.

240

Gel-permeation chromatography. Peptide size distribution indicates the presence of mostly

241

peptides with the main peaks between 0.445-3.2 kDa (Figure 1). Alcalase hydrolysate (AH) had

242

the most uniform peptide distribution followed by chymotrypsin hydrolysate (CH) while the

243

remaining hydrolysates had big peptide peaks with estimated MW of 116-132 kDa. Alcalase is

244

an endopeptidase with very high degree of proteolysis, which may have contributed to the almost

245

normal distribution of low molecular weight peptides. However, unlike the other hydrolysates

246

the flavourzyme hydrolysate (FH) had a distinct big peak with ~60 Da estimated MW, which

247

most likely represents amino acids due to the presence of exopeptidase activity. Peptide size and

248

amino acid composition play an important role in taste. This is because previous studies have

249

reported that peptides of 0.36-2.10 kDa were primary contributors to bitterness of protein

250

hydrolysates, because smaller peptides failed to achieve the particular conformation required for

251

binding to taste receptors.46,47 The results suggest that peptide sizes identified from this work fall

12 ACS Paragon Plus Environment

Page 12 of 42

Page 13 of 42

252

Journal of Agricultural and Food Chemistry

within the 0.36-2.10 kDa range and can interact with bitter taste receptors.

253

Prediction of bitter score from e-tongue. Electronic-tongues have been widely used to

254

detect bitter taste of samples, but also can determine the suppression ability of bitter taste

255

modifiers, such as high potency sweeteners suppressing bitter taste of quinine hydrochloride, and

256

acesulfame K and citric acid suppressing the bitter taste of epinephrine.48,49 Quinine can activate

257

multiple T2Rs50 and therefore, is a suitable compound to test for antagonists or bitter taste

258

suppressors. Figure 2 shows e-tongue bitterness scores for individual beef protein hydrolysates

259

and their combinations with quinine. Quinine had the strongest bitterness intensity while the

260

inverse agonist for T2R4 (BCML) had the least (p < 0.05). Among the hydrolysates, AH had the

261

least bitterness intensity (p < 0.05) while there were no significant differences between the

262

remaining five hydrolysates. The lower bitterness intensity of the AH may be attributed to two

263

main factors. First, is the higher DH (compared to the other hydrolysates), which could have

264

enhanced further proteolysis to split very bitter and larger peptides to produce smaller peptides

265

with reduced bitter-taste. This is consistent with the lower molecular weight profile (~445 Da) of

266

the AH peptides when compared to the other hydrolysates (Figure 1). The possibility of using

267

extensive proteolysis to decrease protein hydrolysate bitterness has also been discussed.51 Second,

268

the sum of hydrophobic amino acids (HAA) and positively-charged amino acids (PCAA) has

269

been reported to be important for enhancing bitterness intensity of peptides.52 AH and FH had

270

the lowest sum of HAA + PCAA but it can be concluded that FH has less content of peptides and

271

more free amino acids due to the exopeptidase activity in flavourzyme. Therefore, the lower

272

bitterness intensity of AH may also be due to the reduced contents of PCAA and HAA. The

273

results are consistent with previous works that have reported enzymatic food protein

274

hydrolysates usually have bitter taste (Humiski and Aluko).47,53,54 In the presence of BCML and

13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

275

each of the protein hydrolysates, bitterness intensity of quinine was significantly (p < 0.05)

276

decreased. BCML produced the most decrease in quinine bitterness intensity followed by TH and

277

PH. The structural basis for the bitter taste blocking efficiency of the hydrolysates is difficult to

278

determine because of the presence of a large pool of peptides. However, it is pertinent to note

279

that the TH had the lowest content of HAA while PH had the least content of negatively-charged

280

amino acids.

281

Inhibition of quinine-dependent T2R4 activation by protein hydrolysates. In previous

282

studies that examined small molecular weight bitter taste modifiers, TAS2R4 gene was

283

heterologously expressed in HEK293T cells and intracellular calcium mobilization assay has

284

been applied to measure the bitterness inhibitory ability of the compounds.8-10 In this assay, high

285

intracellular calcium mobilization means that the T2Rs are activated more intensely, while lower

286

calcium releases suggest weak activation. Results from the calcium mobilization assay indicate

287

highest activity for quinine, CH and FH (Figure 3). In contrast, TH, PH, TMH and AH were

288

significantly less effective in inducing intracellular calcium mobilization in T2R4 expressing

289

cells. The low calcium mobilization ability of AH is consistent with the lowest bitterness

290

intensity among the hydrolysates as determined by e-tongue. However, the calcium mobilization

291

ability of the remaining five hydrolysates was not related to the e-tongue data. The uniform

292

peptide size distribution in AH when compared to the other hydrolysates with non-uniform size

293

distribution may have contributed to this discrepancy. The results suggest that high molecular

294

weight peptides (more abundant in FH, CH, TH, PH and TMH) behave differently from smaller

295

molecular weight peptides (more abundant in AH) with respect to electrochemical signaling

296

properties in the e-tongue and biological interactions with the T2R4. Moreover, previous reports

297

have suggested the e-tongue instrument appears to have difficulty in sensing organic bitter taste

14 ACS Paragon Plus Environment

Page 14 of 42

Page 15 of 42

Journal of Agricultural and Food Chemistry

298

substances, such as amino acids and peptides; therefore, a wider variation in peptide size could

299

have exacerbated this deficiency.55,56 Because the e-tongue responses varied from those of the

300

calcium mobilization assay, the least effective (AH) and most effective (CH) in causing calcium

301

release from T2R4 cells were chosen for peptide identification through RP-HPLC separation to

302

obtain 4 fractions each (Figure 4). Moreover, protein hydrolysates contain a wide range of

303

peptides and measured properties such as bitterness intensity only reflect net activity. Therefore,

304

the protein hydrolysates could still contain individual peptides with T2R4 antagonistic property

305

that was masked by other peptides during the e-tongue experiments, hence the need for peptide

306

fractionation and sequence identification. The AH-F1 was the most effective in blocking

307

quinine-dependent calcium release in the T2R4 expressing HEK293T cells while the four CH

308

fractions behaved similarly to each other. However, CH-F4 was chosen for further peptide

309

fractionation due to higher abundance when compared to CH-F1, CH-F2 and CH-F3. AH-F1

310

separation on the RP-HPLC column also led to 4 isolated fractions with the AH-F1-3 producing

311

the most significant (p < 0.05) blockage of quinine-dependent calcium release from T2R4

312

expressing HEK293T cells (Figure 5). In contrast, RP-HPLC separation of CH-F4 yielded 8

313

fractions, all which produced significant (p < 0.05) reductions in calcium release when added to

314

quinine (Figure 6). However, CH-F4-1, CH-F4-3, CH-F4-4 and CH-F4-5 had the most

315

reductions in calcium release. Based on absolute values of the decrease in calcium release, AH-

316

F1-3, CH-F4-3 and CH-F4-5 were chosen for peptide identification and amino acid sequencing.

317

Inhibition of quinine-dependent T2R4 activation by synthesized peptides. Four peptides

318

were identified from AH-F1-3, one from CH-F4-3 and three from CH-F4-5 with sequences

319

shown in Table 3 and supplementary Figures S1-S8. Threonine, serine, methionine, leucine, and

320

alanine were the most common amino acids in the peptides. With the exception of AAMY and

15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

321

AAYM, the ratio of hydrophobic amino acids in each identified peptides was less than 50%,

322

which suggest that hydrophilic characteristics may have contributed to the bitter taste-

323

suppressing ability of these peptides. However, it has been suggested that hydrophobic properties

324

can enhance peptide entry into target organs through cell membrane lipid bilayer,57 which

325

contributes to enhanced bioactivity inside the cells. Besides, some hydrophobic compounds such

326

as D-Tryptophan benzyl ester and N,N-Dibenzyl-L-serine methyl ester, were reported to have

327

high predicted antagonistic binding affinity to T2R4,8 implying that it is possible for

328

hydrophobic substances to inhibit bitter taste receptors. Thus, peptides with high contents of

329

hydrophobic amino acids may also have ability to suppress bitterness.

330

All peptides showed significantly (p < 0.05) lower calcium mobilization than quinine

331

(Figure 7). Peptides ETSARHL, ETCL, AGDDAPRAVF and AAMY showed less calcium

332

mobilization (∆RFU) upon co-incubation with quinine in HEK293T cells expressing T2R4

333

indicating that these four peptides may have triggered calcium mobilization through a T2R4-

334

dependent pathway. Next, competition assays on T2R4 were pursued using three of the peptides

335

to obtain their inhibitory concentrations (IC50). Result showed that these peptides inhibited

336

quinine response in a concentration dependent manner, with AGDDAPRAVF showing a lower

337

IC50 value of 85 µM among the three peptides analyzed (Figure 8).

338

It has been observed in several soybean protein-derived peptides that leucine residue at

339

C-terminal was responsible for bitterness; treatment with a carboxypeptidase led to a marked

340

reduction in bitterness intensity.58,59 Hence the peptides AGDDAPRAVF, AAMY, VSSY,

341

AAYM should have less bitter taste compared to other identified peptides with leucine residue at

342

C-terminal. But the results obtained in this work do not agree with such hypothesis because the

343

leucine-containing peptides (especially ETCL), had similar or even lower calcium release ability

16 ACS Paragon Plus Environment

Page 16 of 42

Page 17 of 42

Journal of Agricultural and Food Chemistry

344

than peptides that did not contain leucine (Figure 7). Addition of each peptide to quinine led to

345

significant reductions (except VSSY) in calcium release from the T2R4-expressing cells with

346

AGDDAPRAVF being the most effective blocker. The two peptides with multiple numbers of

347

serine residues were not very effective, which suggests that this amino acid may not be an

348

important structural requirement for bitter taste-blocking peptides. However, the number of

349

peptides studied in this work is too small to estimate structure-function properties. But it is

350

important to point out that the two most active suppressors of quinine bitter taste are also the

351

longest peptide chains, which could indicate an importance for the number of amino acids.

352

In conclusion, this work has revealed the potential to use beef protein hydrolysates and

353

derived peptides as bitter taste T2R4 receptor blockers. Recent studies have revealed that T2Rs

354

are also expressed in the gastrointestinal tract, enteroendocrine STC-1 cells, respiratory system,

355

male reproductive system, central nervous system and several tissues.60-62 This development

356

suggests that T2Rs may possess more potential important physiological functions other than

357

bitter taste sensation. For example, most drugs have bitter taste, which may be responsible for

358

some of the observed off-target (or negative) effects since these drugs can activate T2Rs in non-

359

target parts of the body. Therefore, in addition to enhancing eating quality of foods, potent bitter

360

taste-suppressing peptides may find additional uses as agents to reduce off-target effects of

361

certain drugs. Moreover, since the mechanism of signal transduction of T2Rs in different cell

362

types are not completely understood, peptides with inhibitory ability may be used to explore the

363

signaling pathways in different cell types. Since only one T2R was studied in this work, future

364

research works will determine the inhibitory effects of these food protein-derived hydrolysates

365

and peptides against multiple T2Rs.

366

17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

367

ACKNOWLEDGMENTS

368

The authors are grateful for the financial support provided by Alberta Agriculture and Forestry

369

funding reference number 2015P002R, Edmonton, Alberta, Canada. We acknowledge the

370

support of the Natural Sciences and Engineering Council of Canada (NSERC), funding reference

371

numbers RGPIN-249890-13 and RGPIN-2014-4099. Cette recherche a été financée par le

372

Conseil de recherches en sciences naturelles et en génie du Canada (CRSNG), numéro de

373

référence RGPIN-249890-13 et RGPIN-2014-4099. We thank Appalaraju Jaggupilli for work

374

with the HEK293T-T2R4 stable cell line.

375 376 377 378 379 380 381 382 383 384 385 386 387 388 389

18 ACS Paragon Plus Environment

Page 18 of 42

Page 19 of 42

390 391

Journal of Agricultural and Food Chemistry

REFERENCES (1) Levit, A.; Nowak, S.; Peters, M.; Wiener, A.; Meyerhof, W.; Behrens, M.; Niv, M. Y.

392

The bitter pill: Clinical drugs that activate the human bitter taste receptor TAS2R14.

393

FASEB J. 2014, 28, 1181–1197.

394 395 396

(2) Meyerhof, W. Elucidation of mammalian bitter taste. Rev. Physiol. Biochem. Pharmacol. 2005, 154, 37–72. (3) Upadhyaya, J.; Pydi, S. P.; Singh, N.; Aluko, R. E.; Chelikani, P. Bitter taste receptor

397

T2R1 is activated by dipeptides and tripeptides. Biochem. Biophys. Res. Commun. 2010,

398

398, 331–335.

399

(4) Meyerhof, W.; Batram, C.; Kuhn, C.; Brockhoff, A.; Chudoba, E.; Bufe, B.; Appendino,

400

G.; Behrens, M. The molecular receptive ranges of human TAS2R bitter taste receptors.

401

Chemical Senses 2010, 35, 157–170.

402

(5) Lossow, K.; Hübner, S.; Roudnitzky, N.; Slack, J. P.; Pollastro, F.; Behrens, M.;

403

Meyerhof, W. Comprehensive analysis of mouse bitter taste receptors reveals different

404

molecular receptive ranges for orthologous receptors in mice and humans. J. Biol. Chem.

405

2016, 291, 15358–15377.

406 407 408

(6) Wiener, A.; Shudler, M.; Levit, A.; Niv, M. Y. BitterDB: a database of bitter compounds. Nucleic Acids Res., 2012, 40, D413–D419. (7) Sainz, E.; Cavenagh, M. M.; Gutierrez, J.; Battey, J. F.; Northup, J. K.; Sullivan, S. L.

409

Functional characterization of human bitter taste receptors. Biochem. J. 2007, 403, 537–

410

43.

411

(8) Pydi, S. P.; Sobotkiewicz, T.; Billakanti, R.; Bhullar, R. P.; Loewen, M. C.; Chelikani, P.

412

Amino acid derivatives as bitter taste receptor (T2R) blockers. J. Biol. Chem. 2014, 289,

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

413 414

25054–25066. (9) Pydi, S. P.; Jaggupilli, A.; Nelson, K. M.; Abrams, S. R.; Bhullar, R. P.; Loewen, M. C.;

415

Chelikani, P. Abscisic acid acts as a blocker of the bitter taste G protein-coupled receptor

416

T2R4. Biochem. 2015, 54, 2622–2631.

417

(10) Roland, W. S. U.; Gouka, R. J.; Gruppen, H.; Driesse, M.; Van Buren, L.; Smit, G.;

418

Vincken, J. P. 6-Methoxyflavanones as bitter taste receptor blockers for hTAS2R39.

419

PLoS One 2014, 9(4), e94451.

420

(11) Liszta, K. I.; Ley, J. P.; Liedera, B.; Behrens, M.; Stöger, B.; Reiner, A.; Hochkogler, C.

421

M.; Köck, E.; Marchiori, A.; Hans, J.; Widder, S.; Krammer, G.; Sanger, G. J.; Somoza,

422

M. M.; Meyerhof, W.; Somoza, V. Caffeine induces gastric acid secretion via bitter taste

423

signaling in gastric parietal cells. PNAS, 2017, 114, E6260-E6269.

424

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

425

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

426

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

427

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

428

2010, 20, 1104–1109.

429

(13) Brockhoff, A.; Behrens, M.; Massarotti, A.; Appendino, G.; Meyerhof, W. Broad tuning

430

of the human bitter taste receptor hTAS2R46 to various sesquiterpene lactones, clerodane

431

and labdane diterpenoids, strychnine, and denatonium. J. Agric. Food Chem. 2007, 55,

432

6236-6243.

433

(14) Jaggupilli, A.; Howard, R.; Upadhyaya, J. D.; Bhullar, R. P.; Chelikani, P. Bitter taste

434

receptors: Novel insights into the biochemistry and pharmacology. Int. J. Biochem. Cell

435

Biol., 2016, 77, 184-196.

20 ACS Paragon Plus Environment

Page 20 of 42

Page 21 of 42

436

Journal of Agricultural and Food Chemistry

(15) Wu, S. V.; Rozengurt, N.; Yang, M.; Young, S. H.; Sinnett-Smith, J.; Rozengurt, E.

437

Expression of bitter taste receptors of the T2R family in the gastrointestinal tract and

438

enteroendocrine STC-1 cells. Proc. Natl. Acad. Sci. 2002, 99, 2392–2397.

439 440 441

(16) Clark, A. A.; Liggett, S. B.; Munger, S. D. Extraoral bitter taste receptors as mediators of off-target drug effects. FASEB J. 2012, 26, 4827-4831. (17) Avau, B.; Rotondo, A.; Thijs, T.; Andrews, C. N.; Janssen, P.; Tack, J.; Depoortere, I.

442

Targeting extra-oral bitter taste receptors modulates gastrointestinal motility with effects

443

on satiation. Scientific Reports 2015, 5, Article number: 15985.

444

(18) Clark, A. A.; Dotson, C. D.; Elson, A. E. T.; Voigt, A.; Boehm, U.; Meyerhof, W.;

445

Steinle, N. I.; Munger, S. D. TAS2R bitter taste receptors regulate thyroid function.

446

FASEB J. 2015, 29, 164–172.

447

(19) Wölfle, U.; Elsholz, F. A.; Kersten, A.; Haarhaus, B.; Müller, W. E.; Schempp C. M.

448

Expression and functional activity of the bitter taste receptors TAS2R1 and TAS2R38 in

449

human keratinocytes. Skin Pharmacol. Physiol. 2015, 28, 137-146.

450

(20) Deshpande, D. A.; Wang, W. C. H.; McIlmoyle, E. L.; Robinett, K. S.; Schillinger, R. M.;

451

An, S. S.; Liggett, S. B. Bitter taste receptors on airway smooth muscle bronchodilate by

452

localized calcium signaling and reverse obstruction. Nature Med. 2010, 16, 1299–1304.

453

(21) Zhang, C. H.; Lifshitz, L. M.; Uy, K. F.; Ikebe, M.; Fogarty, K. E.; ZhuGe, R. The

454

cellular and molecular basis of bitter tastant-induced bronchodilation. PLoS Biology 2013,

455

11, e1001501.

456

(22) Shaik, F. A.; Singh, N.; Arakawa, M.; Duan, K.; Bhullar, R. P.; Chelikani, P. Bitter taste

457

receptors: Extraoral roles in pathophysiology. Int. J. Biochem. Cell Biol., 2016, 77, 197-

458

204.

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

459

(23) Kannan, A.; Hettiarachchy, N. S.; Marshall, M.; Raghavan, S.; Kristinsson, H. Shrimp

460

shell peptide hydrolysates inhibit human cancer cell proliferation. J. Sci. Food Agric.

461

2011, 91, 1920–1924.

462

(24) Ryu, B.; Qian, Z. J.; Kim, S. K. SHP-1, a novel peptide isolated from seahorse inhibits

463

collagen release through the suppression of collagenases 1 and 3, nitric oxide products

464

regulated by NF-??B/p38 kinase. Peptides 2010, 31, 79–87.

465 466 467 468 469

(25) Udenigwe, C. C.; Aluko, R. E. Food protein-derived bioactive peptides: Production, processing, and potential health benefits. J. Food Sci. 2012, 77, R11-R24. (26) Karametsi, K.; Kokkinidou, S.; Ronningen, I.; Peterson, D. G. Identification of bitter peptides in aged Cheddar cheese. J. Agric. Food Chem. 2014, 62, 8034-8041. (27) Fuller, W.; Qurtz, R. J. Composition comprising bitter- or metallic-tasting eatable and

470

tastand-comprising (hydroxy-, amino- or mercapto)-substituted di:carboxylic acid or

471

derivatives, especially di:peptide containing aspartic acid. US Patent 5643955, 1997.

472

(28) Ley, J. P. Masking bitter taste by molecules. Chem. Percept. 2008, 1, 58–77.

473

(29) Akitomi, H.; Tahara, Y.; Yasuura, M.; Kobayashi, Y.; Ikezaki, H.; Toko, K.

474

Quantification of tastes of amino acids using taste sensors. Sens. Actuators B Chem. 2013,

475

179, 276–281.

476 477 478 479 480 481

(30) Williams, P. G. Nutritional composition of red meat. Nutr. Dietetics 2007, 64(Suppl. 4), S113–S119. (31) Hou, H.; Li, B.; Zhao, X. Enzymatic hydrolysis of defatted mackerel protein with low bitter taste. J. Ocean Univ. China 2011, 10, 85–92. (32) Lin, S.; Tian, W.; Li, H.; Cao, J.; Jiang, W. Improving antioxidant activities of whey protein hydrolysates obtained by thermal preheat treatment of pepsin, trypsin, alcalase

22 ACS Paragon Plus Environment

Page 22 of 42

Page 23 of 42

482 483 484 485

Journal of Agricultural and Food Chemistry

and flavourzyme. Int. J. Food Sci. Technol. 2012, 47, 2045–2051. (33) Nchienzia, H. A.; Morawicki, R. O.; Gadang, V. P. Enzymatic hydrolysis of poultry meal with endo- and exopeptidases. Poultry Sci. 2010, 89, 2273–2280. (34) Charoenphun, N.; Cheirsilp, B.; Sirinupong, N.; Youravong, W. Calcium-binding

486

peptides derived from tilapia (Oreochromis niloticus) protein hydrolysate. Eur. Food Res.

487

Technol. 2013, 236, 57–63.

488 489 490

(35) Nielsen, P. M.; Petersen, D.; Dambmann, C. Improved method for determining food protein degree of hydrolysis. J. Food Sci. 2001, 66, 642–646. (36) Bidlingmeyer, B. A.; Cohen, S. A.; Tarvin, T. L. Rapid analysis of amino acids using pre-

491

column derivatization. J. Chrom. B: Biomedical Sciences and Applications 1984, 336,

492

93–104.

493

(37) Gehrke, C. W.; Wall, L. L.; Absheer, J. S.; Kaiser, F. E.; Zumwalt, R. W. Sample

494

preparation for chromatography of amino-acids: acid hydrolysis of proteins. J. Assoc.

495

Official Anal. Chem. 1985, 68, 811–821.

496 497 498

(38) Landry, J.; Delhaye, S. Simplified procedure for the determination of tryptophan of foods and feedstuffs from barytic hydrolysis. J. Agric. Food Chem. 1992, 40, 776–779. (39) He, R.; Alashi, A.; Malomo, S. A.; Girgih, A. T.; Chao, D.; Ju, X.; Aluko, R. E.

499

Antihypertensive and free radical scavenging properties of enzymatic rapeseed protein

500

hydrolysates. Food Chem. 2013, 141, 153-159.

501 502 503 504

(40) Alpha MOS. Astree electrochemical sensor technology—Technical note: T-SAS-04. 2004. (41) Girgih, A. T.; Udenigwe, C. C.; Aluko, R. E. Reverse-phase HPLC separation of hemp seed (Cannabis sativa L.) protein hydrolysate produced peptide fractions with enhanced

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

505

antioxidant capacity. Plant Foods Hum. Nutr. 2013, 68, 39–46.

506

(42) Chuesiang, P.; Sanguandeekul, R. Protein hydrolysate from tilapia frame: antioxidant and

507

angiotensin I converting enzyme inhibitor properties. Int. J. Food Sci. Technol. 2015, 50,

508

1436–1444.

509

(43) Wanasundara, P. K. J. P. D.; Amarowicz, R.; Pegg, R.; Shand, P. J. Preparation and

510

characterization of hydrolyzed proteins from defibrinated bovine plasma. J. Food Sci.

511

2002, 67, 623-630.

512

(44) Kane, L. E.; Davis, J. P.; Oakes, A. J.; Dean, L. L.; Sanders, T. H. Value-added

513

processing of peanut meal: enzymatic hydrolysis to improve functional and nutritional

514

properties of water soluble extracts. J. Food Biochem. 2012, 36, 520-531.

515

(45) Kohl, S.; Behrens, M.; Dunkel, A.; Hofmann, T.; Meyerhof, W. Amino acids and

516

peptides activate at least five members of the human bitter taste receptor family. J. Agric.

517

Food Chem. 2013, 61, 53−60.

518 519 520 521 522

(46) Maehashi, K.; Huang, L. Bitter peptides and bitter taste receptors. Cell. Mol. Life Sci. 2009, 66, 1661–1671. (47) Kim, M. R.; Choi, S. Y.; Lee, C. H. Molecular characterization and bitter taste formation of tryptic hydrolysis of 11S glycinin. J. Microbiol. Biotechnol. 1999, 9, 509–513. (48) Rachid, O.; Simons, F. E. R.; Rawas-Qalaji, M.; Simons, K. J. An electronic tongue:

523

Evaluation of the masking efficacy of sweetening and/or flavoring agents on the bitter

524

taste of epinephrine. AAPS PharmSciTech. 2010, 11, 550–557.

525

(49) Wu, X.; Onitake, H.; Haraguchi, T.; Tahara, Y.; Yatabe, R.; Yoshida, M.; Toko, K.

526

Quantitative prediction of bitterness masking effect of high-potency sweeteners using

527

taste sensor. Sens. Actuators B Chem. 2016, 235, 11–17.

24 ACS Paragon Plus Environment

Page 24 of 42

Page 25 of 42

Journal of Agricultural and Food Chemistry

528

(50) Upadhyaya, J. D.; Chakraborty, J.; Shaik, F. A.; Jaggupilli, A.; Bhullar, R. P.; Chelikani,

529

P. The pharmacochaperone activity of quinine on bitter taste receptors. PLoS One, 2016,

530

11, e0156347.

531 532

(51) Sun, X. D. Enzymatic hydrolysis of soy proteins and the hydrolysates utilization. Int. J. Food Sci. Technol. 2011, 46, 2447–2459.

533

(52) Tamura, M.; Mori, N.; Miyoshi, T.; Koyama, S.; Kohri, H.; Okai, H. Practical debittering

534

using model peptides and related compounds. Agric. Biol. Chem. 1990, 54, 41–51.

535 536 537 538 539

(53) Humiski, L.; Aluko, R. E. Physicochemical and bitterness properties of enzymatic pea protein hydrolysates. J. Food Sci. 2007, 72, S605-S611. (54) Geisenhoff, H. Bitterness of soy protein hydrolysates according to molecular weight of peptides, MSc thesis dissertaion, Iowa State Univeristy, 2009. (55) Miyanaga, Y.; Tanigake, A.; Nakamura, T.; Kobayashi, Y.; Ikezaki, H.; Taniguchi, A.;

540

Uchida, T. Prediction of the bitterness of single, binary- and multiple-component amino

541

acid solutions using a taste sensor. Int. J. Pharm. 2002, 248, 207–218.

542

(56) Rudnitskaya, A.; Nieuwoudt, H. H.; Muller, N.; Legin, A.; Du Toit, M.; Bauer, F. F.

543

Instrumental measurement of bitter taste in red wine using an electronic tongue. Anal.

544

Bioanal. Chem. 2010, 397, 3051–3060.

545 546 547

(57) Sarmadi, B. H.; Ismail, A. Antioxidative peptides from food proteins: A review. Peptides 2010, 31, 1949–1956. (58) Edens, L.; Dekker, P.; Van Der Hoeven, R.; Deen, F.; De Roos, A.; Floris, R.

548

Extracellular prolyl endoprotease from Aspergillus niger and its use in the debittering of

549

protein hydrolysates. J. Agric. Food Chem. 2005, 53, 7950–7957.

550

(59) FitzGerald, R. J.; O’Cuinn, G. Enzymatic debittering of food protein hydrolysates.

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

551 552

Biotechnol. Adv. 2006, 24, 234-237. (60) Singh, N.; Vrontakis, M.; Parkinson, F.; Chelikani, P. Functional bitter taste receptors are

553

expressed in brain cells. Biochem. Biophys. Res. Commun. 2011, 406, 146–151.

554

(61) Xu, J.; Cao, J.; Iguchi, N.; Riethmacher, D.; Huang, L. Functional characterization of

555 556

bitter-taste receptors expressed in mammalian testis. Mol. Hum. Reprod. 2013, 19, 17–28. (62) Jaggupilli, A.; Singh, N.; Upadhyaya, J.; Sikarwar, A. S.; Arakawa, M.; Dakshinamurti,

557

S.; Chelikani, P. Analysis of the expression of human bitter taste receptors in extraoral

558

tissues. Mol. Cell. Biochem. 2017, 426, 137–147.

559

26 ACS Paragon Plus Environment

Page 26 of 42

Page 27 of 42

Journal of Agricultural and Food Chemistry

Figure Captions Figure 1. Molecular weight distribution of peptides present in enzymatic beef protein hydrolysates based on elution volume from a calibrated Superdex12 10/300 GL gel permeation column. Figure 2. Estimated e-tongue bitter scores of individual beef protein hydrolysates and their ability to suppress quinine bitterness intensity. AH, alcalase hydrolysate; CH, chymotrypsin hydrolysate; PH, pepsin hydrolysate; TH, trypsin hydrolysate; TMH, thermoase hydrolysate; FH, flavourzyme hydrolysate. BCML (Nα,Nα-bis(carboxymethyl)-l-lysine) was used as a positive control. Bars with different letters have significantly different (p < 0.05) mean values. Bars labelled with uppercase or lowercase letters represent hydrolysate only or hydrolysate + quinine treatments, respectively. Figure 3. Calcium mobilization in T2R4 expressing HEK293T cell system after treatment with beef protein hydrolysates (5 mg/mL) and quinine (1 mM). AH, alcalase hydrolysate; CH, chymotrypsin hydrolysate; PH, pepsin hydrolysate; TH, trypsin hydrolysate; TMH, thermoase hydrolysate; FH, flavourzyme hydrolysate. Bars with different letters have significantly different (p < 0.05) mean values. ∆RFU: changes in relative fluorescence unit (test cells-control cells). Figure 4. Calcium mobilization in T2R4 expressing HEK293T cell system after treatment with quinine (1 mM), beef protein hydrolysate peptide fractions (5 mg/mL) or a quinine (1 mM) solution that contained 5 mg/ml protein hydrolysate peptide fractions from first round of RPHPLC separation. AH-F, alcalase hydrolysate fractions; CH-F, chymotrypsin hydrolysate fractions. Bars with different letters have significantly different (p < 0.05) mean values. Inset shows the RP-HPLC separation and fraction collection. Bars labelled with uppercase or lowercase letters represent hydrolysate only or hydrolysate + quinine treatments, respectively.

27 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 5. Calcium mobilization in T2R4 expressing HEK293T cell system after treatment with quinine (1 mM), beef protein hydrolysate peptide fractions (5 mg/mL) or a quinine (1 mM) solution that contained 5 mg/ml alcalase hydrolysate peptide fractions (AH-F) from second round of RP-HPLC separation. Bars with different letters have significantly different (p < 0.05) mean values. Inset shows the RP-HPLC separation and fraction collection. Bars labelled with uppercase or lowercase letters represent hydrolysate only or hydrolysate + quinine treatments, respectively. Figure 6. Calcium mobilization in T2R4 expressing HEK293T cell system after treatment with quinine (1 mM), beef protein hydrolysate peptide fractions (5 mg/mL) or a quinine (1 mM) solution that contained 5 mg/ml alcalase hydrolysate peptide fractions (CH-F) from second round of RP-HPLC separation. Bars with different letters have significantly different (p < 0.05) mean values. Inset shows the RP-HPLC separation and fraction collection. Bars labelled with uppercase or lowercase letters represent hydrolysate only or hydrolysate + quinine treatments, respectively. Figure 7. Calcium mobilization in HEK293T cells stably expressing T2R4 and treated with different peptides. A. The T2R4 expressing cells were treated with quinine (1 mM), synthesized peptides (1mM) or a quinine solution that contained the synthesized peptides. The calcium responses of cells treated with buffer are used as control. Statistically significant values are shown by asterisk (*p < 0.05) and (***p < 0.001). B. Raw traces for calcium mobilization analysis showing decrease in calcium release upon stimulation with different peptides (Pep 1-8) with quinine. The black arrows at 20 sec indicates the addition of the compounds. Raw traces for both HEK293T-T2R4 stable cells and untransfected HEK293T cells are shown. The changes in fluorescence by the calcium sensitive dye Fluo-4NW are measured as relative fluorescence unit

28 ACS Paragon Plus Environment

Page 28 of 42

Page 29 of 42

Journal of Agricultural and Food Chemistry

(∆RFU) on the Y-axis for a total of 180 seconds (X-axis) using a Flex Station 3 plate reader. Figure 8. Peptides and quinine competition calcium mobilization assay on T2R4. HEK293T cells stably expressing T2R4 were treated with 1 mM quinine and with increasing concentrations of peptides ETSARHL, AGDDAPRAVF and AAMY ranging from 0.015- 1 mM. Changes in intracellular calcium were measured (∆RFUs), and IC50 values were calculated using Graph Pad Prism 4.0. Data were collected from two-three independent experiments carried out in triplicate.

29 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Table 1. Degree of Hydrolysis of Enzymatic Beef Protein Hydrolysates enzyme

degree of hydrolysis (%)

Alcalase

®

35.57 ± 0.01

Chymotrypsin

®

25.83 ± 0.02

Trypsin®

24.18 ± 0.02

Pepsin®

8.60 ± 0.02

Flavourzyme® Thermoase

®

47.02 ± 0.02 18.26 ± 0.01

30 ACS Paragon Plus Environment

Page 30 of 42

Page 31 of 42

Journal of Agricultural and Food Chemistry

Table 2. Amino Acid Composition (%) of Defatted Beef Protein (DBP) and the Enzymatic Hydrolysates* AA Asx Thr Ser

DBP 9.98 4.63 4.27

AH 10.41 4.45 4.21

CH 10.25 4.67 4.32

PH 9.99 4.26 4.13

TH TMH 10.19 10.48 4.43 4.73 4.07 4.41

FH 10.46 4.36 4.35

Mean±SD 10.23 ± 0.22 4.48 ± 0.18 4.25 ± 0.13

P-Value 0.02 0.11 0.71

Glx Pro Gly Ala Cys Val Met Ile Leu Tyr Phe His Lys Arg Trp

15.51 4.79 5.46 6.01 1.03 4.55 2.63 4.21 7.86 3.35 4.22 4.13 8.85 6.29 1.02

15.95 5.07 6.50 6.32 0.89 4.39 1.81 3.91 7.56 3.10 3.92 4.03 8.98 6.59 0.67

16.18 4.28 4.34 5.87 0.98 4.57 2.29 4.24 7.95 3.49 4.31 4.26 9.30 6.52 0.88

14.30 5.58 7.33 6.31 0.89 4.89 1.88 4.30 7.12 2.93 4.15 4.42 8.75 6.80 0.75

16.31 4.74 5.56 6.09 0.83 4.34 1.85 3.97 7.56 2.96 3.77 4.40 9.66 7.29 0.66

15.89 4.44 5.15 5.98 0.91 4.64 1.89 4.35 7.87 3.43 4.27 4.05 9.20 6.31 0.92

17.86 4.18 4.57 6.51 0.82 4.57 2.02 4.18 8.07 2.72 3.82 5.05 7.73 6.88 0.59

15.94 ± 0.90 4.72 ± 0.53 5.58 ± 1.15 6.18 ± 0.24 0.89 ± 0.06 4.57 ± 0.20 1.96 ± 0.18 4.16 ± 0.18 7.69 ± 0.35 3.11 ± 0.30 4.04 ± 0.23 4.37 ± 0.37 8.94 ± 0.67 6.72 ± 0.37 0.75 ± 0.13

0.23 0.75 0.82 0.14 0.002 0.84 0.001 0.51 0.28 0.10 0.12 0.18 0.76 0.04 0.004

HAA PCAA NCAA AAA SCAA BCAA

39.66 19.27 25.49 8.59 3.66 16.62

37.64 19.61 26.36 7.69 2.70 15.86

38.86 20.08 26.43 8.68 3.27 16.76

38.80 19.97 24.21 7.83 2.76 16.31

36.75 21.34 26.50 7.38 2.68 15.87

38.70 19.45 26.37 8.62 2.80 16.85

37.48 19.66 28.32 7.14 2.84 16.81

38.04 ± 0.87 20.02 ± 0.69 26.37 ± 1.30 7.90 ± 0.64 2.84 ± 0.22 16.41 ± 0.47

0.01 0.05 0.16 0.04 0.001 0.32

*AH, alcalase hydrolysate; CH, chymotrypsin hydrolysate; PH, pepsin hydrolysate; TH, trypsin hydrolysate; TMH, thermoase hydrolysate; FH, flavourzyme hydrolysate HAA: hydrophobic amino acids (alanine, valine, isoleucine, leucine, tyrosine, phenylalanine, tryptophan, proline, methionine and cysteine); PCAA: positively charged amino acids (histidine, lysine, arginine); NCAA: negatively charged amino acids (Asx = asparagine + aspartic acid, Glx = glutamine + glutamic acid); AAA: aromatic amino acids (phenylalanine, tryptophan, tyrosine) SCAA: Sulphur-containing amino acids (cysteine, methionine); BCAA: Branched-chain amino acid (valine, isoleucine, leucine)

31 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 32 of 42

Table 3. Peptides identified from T2R4 inhibitory alcalase hydrolysate (AH) and chymotrypsin hydrolysate (CH) RP-HPLC fractions peptide Source AH-F1-3

Obs (m/z) 233.1

Z 2

suggested peptide TMTL (Pep 1)

parent protein

position

Versican core protein

f529-532

calculated mol. wt. (Da) 446.6

AH-F1-3

233.1

2

ETCL (Pep 2)

Coagulation factor XIII, B polypeptide

f1540-1545

446.5

AH-F1-3

306.2

2

SSMSSL (Pep 3)

Cardiomyopathy associated protein 1

f1540-1545

592.72

AH-F1-3

407

2

ETSARHL (Pep Myosin class II heavy chain 4)

f23-29

794.85

CH-F4-3

509

2

AGDDAPRAVF (Pep 5)

Alpha-actin-2, Alpha-actin-1, f24-33 Alpha-cardiac actin

1000.08

CH-F4-5

228.1

2

AAMY (Pep 6)

DDR2 protein FDPS protein

f368-371 f280-283

436.56

CH-F4-5

228.1

2

VSSY (Pep 7)

Desmin, Fibrillin-1 Glucagon

f20-23 f308-311 f107-110

436.43

CH-F4-5

228.1

2

AAYM (Pep 8)

KRT5 protein

f282-285

436.56

32 ACS Paragon Plus Environment

Page 33 of 42

Journal of Agricultural and Food Chemistry

Figure 1

33 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 2

34 ACS Paragon Plus Environment

Page 34 of 42

Page 35 of 42

Journal of Agricultural and Food Chemistry

Figure 3

35 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Absorbance at 214nm (mAU)

7,500

Chromatogram of AH in red Chromatogram of CH in black

6,000

4,500

3,000

1,500

F1 F1

0 0

6

F2 F2 12

18

F3 F3 24

F4 F4 30

36

42

48

54

60

Run time (min)

Figure 4

36 ACS Paragon Plus Environment

Page 36 of 42

Page 37 of 42

Journal of Agricultural and Food Chemistry

Absorbance at 214nm (mAU)

1000 900 800 700 600 500 400 300 200 100 0

F1 F2 F3

F4

-100 0

6

12

18

24

30

36

42

48

Run time (min)

Figure 5

37 ACS Paragon Plus Environment

Absorbance at 214nm (mAU)

Journal of Agricultural and Food Chemistry

1,000 900 800 700 600 500 400 300 200 100

F1 F2 F3 F4 F5 F6 F7 F8 0 0

6

12

18

24

30

36

42

48

Run time (min)

Figure 6

38 ACS Paragon Plus Environment

Page 38 of 42

Page 39 of 42

Journal of Agricultural and Food Chemistry

A

B

Calcium mobilized (∆RFU)

300 200 100 300 200 100

A. Quinine B. Quinine+Pep 1 C. Pep 1

A

A

A. Quinine B. Quinine+Pep 4 C. Pep 4

A. Quinine B. Quinine+Pep 3 C. Pep 3

A. Quinine B. Quinine+Pep 2 C. Pep 2

A

A

B

B C

B C

C A. Quinine B. Quinine+Pep 1 C. Pep 1

A. Quinine B. Quinine+Pep 2 C. Pep 2

A

A

B

B C

B

C A. Quinine B. Quinine+Pep 4 C. Pep 4

A. Quinine B. Quinine+Pep 3 C. Pep 3

A B

HEK 293T–T2R4

HEK 293T

A B

C

C

C

30 60 90 120 30 60 90 120 30 60 90 120 30 60 90 120

Time (sec) 300 200 100 300 200 100

A. Quinine B. Quinine+Pep 6 C. Pep 6

A. Quinine B. Quinine+Pep 5 C. Pep 5

A

A

B

A

C A. Quinine B. Quinine+Pep 6 C. Pep 6

A

B

B

C

C

A

HEK 293T–T2R4

C

C A. Quinine B. Quinine+Pep 5 C. Pep 5

A. Quinine B. Quinine+Pep 8 C. Pep 8

B

B

B C

A. Quinine B. Quinine+Pep 7 C. Pep 7

A

A. Quinine B. Quinine+Pep 7 C. Pep 7 A B C

A. Quinine B. Quinine+Pep 8 C. Pep 8 A

HEK 293T

B C

30 60 90 120 30 60 90 120 30 60 90 120 30 60 90 120

Time (sec)

Figure 7 39 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Figure 8

40 ACS Paragon Plus Environment

Page 40 of 42

Page 41 of 42

Journal of Agricultural and Food Chemistry

TOC graphic

41 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

TOC Graphic 183x153mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 42 of 42