Influence of Chemical Species on Polyphenol-Protein Interactions

Mar 11, 2019 - ... techniques such as SDS−Polyacrylamide gel electrophoresis (SDS-PAGE) and cell cultures using a cell-based model of the oral epith...
0 downloads 0 Views 697KB Size
Subscriber access provided by UNIV OF TEXAS DALLAS

Food and Beverage Chemistry/Biochemistry

Influence of Chemical Species on PolyphenolProtein Interactions related to wine astringency Alba María Ramos-Pineda, Guy H. Carpenter, Ignacio García-Estévez, and María Teresa Escribano-Bailon J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.9b00527 • Publication Date (Web): 11 Mar 2019 Downloaded from http://pubs.acs.org on March 12, 2019

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 30

Journal of Agricultural and Food Chemistry

Influence of Chemical Species on Polyphenol-Protein Interactions related to Wine Astringency

A.M. Ramos-Pinedaa, G.H. Carpenterb, I. García Estéveza*, M.T. Escribano-Bailóna

aGrupo

de Investigación en Polifenoles (GIP), Facultad de Farmacia, University of

Salamanca, Salamanca, Spain bSalivary

Research Unit, King's College London Dental Institute, Guy's Hospital, London,

SE1 9RT, United Kingdom

*Corresponding author: Ignacio García Estévez Phone: +34 923 294 537 e-mail: [email protected]

1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 30

1

ABSTRACT

2

One of the most accepted mechanisms of astringency consists of the interaction between

3

polyphenols and some specific salivary proteins. This work aims to obtain further insights

4

into the mechanisms leading to a modulation of astringency elicited by polyphenols. The

5

effect of the presence of different chemical species (present in food and beverages as food

6

additives) on the polyphenol-protein interaction has been evaluated by means of techniques

7

such as SDS−Polyacrylamide gel electrophoresis (SDS-PAGE) and cell cultures using a

8

cell-based model of the oral epithelium. Results obtained showed that several chemicals,

9

particularly sodium carbonate, seem to inhibit polyphenol binding to salivary proteins and

10

to oral epithelium. These results point out that polyphenol-saliva protein interactions can be

11

affected by some food additives, what can help to better understand changes in astringency

12

perception.

13 14 15

KEYWORDS: Astringency, Tannin, Wine, Salivary proteins, SDS-PAGE, Epithelial cell

16 17 18

2 ACS Paragon Plus Environment

Page 3 of 30

Journal of Agricultural and Food Chemistry

20

INTRODUCTION

21

Wine polyphenols have attracted great interest in wine research and industry. They have a

22

number of important functions in wine, contributing to nutrition and organoleptic properties

23

such as color, taste or mouthfeel. Astringency is one of the main sensory attributes in red

24

wines. It can be defined as a drying or puckering feeling in the mouth1 and plays a key role

25

in the mouthfeel of wines. Even more, it is considered a quality parameter and it is an

26

important property and character of a balanced wine. Condensed tannins, also called

27

proanthocyanidins, have been the tannins generally considered as the main responsible for

28

astringency in wine. Proanthocyanidins are polymers composed of monomeric units of

29

(epi)catechin and (epi)gallocatechin.2 These compounds are extracted from both grape

30

seeds and skins during red winemaking. Moreover, they can be found in commercially

31

available grape seed extracts (GSEs), added to red wine during production in order improve

32

their mouthfeel and structure or even to promote color stability.3

33

It is well known the ability of food tannins to interact with some specific salivary proteins,

34

namely proline-rich proteins (PRPs), leading to protein-tannin aggregates that could

35

precipitate in the oral cavity, which is one of the main accepted mechanisms to explain oral

36

astringency.4 Salivary proteins (SP) have been grouped into different classes according to

37

their structure and characteristics, namely, α-amylases, mucins, carbonic anhydrases,

38

statherins, P-B peptide, histatins, cystatins and proline rich proteins (PRPs). PRPs, which

39

are in turn divided into acidic (aPRPs), basic (bPRPs) and glycosylated (gPRPs),5 have

40

been specially studied regarding protein-tannin complexation. It has been suggested that

41

their affinity for tannins is associated with their particular structure and composition, rich in

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 4 of 30

42

proline (25-40% of total residues) and with a remarkable content of glycine and

43

glutamine.6,7

44

Several types of tannin-protein interaction have been described, being hydrogen bonds and

45

hydrophobic interactions the dominant ones.4,8 These interactions can be affected by several

46

factors, like environmental factors (including solvent composition, ionic strength, pH and

47

temperature) or the presence of other substances such us acids, sugars, ethanol and others.

48

8–10

49

suggesting a negative linear relation between them.11 It seems that acidic conditions could

50

result on increase in undissociated phenol groups susceptible to be involved in hydrogen

51

bonds with salivary proteins, which results in an increase in astringency perception. This

52

explanation appears to be consistent with other studies where the impact of pH on tannin-

53

induced astringency was assessed by a trained sensory panel.12,13 Further, ionic strength and

54

ethanol content have been investigating regarding their strong incident on tannin

55

aggregation and hence on tannin-induced astringency.14

56

However, astringency is a very complex process that is not fully understood yet. Some

57

authors have proposed that astringency results from the contribution of several

58

mechanisms, such as loss of saliva lubricity,15 alteration of the mucosal pellicle,16

59

activation of specific taste receptors,17 direct interaction between tannins and oral

60

epithelium,18 or even several of these mechanisms occurring simultaneously.19 In the last

61

years, a lot of studies have been published trying to elucidate the mechanisms behind oral

62

astringency,20–22 however, only very few authors have been focused on the contribution of

63

the oral epithelial cells to astringency.16,18,23 The first study addressing this approach was

64

carried out by Payne and co-workers,18 demonstrating that procyanidins present in wine

A number of authors have studied the relationship between pH and astringency,

4 ACS Paragon Plus Environment

Page 5 of 30

Journal of Agricultural and Food Chemistry

65

react with oral epithelial cells in vitro. After these results, subsequent studies have

66

examined the role of oral epithelium considering also the effect of salivary proteins on

67

these interactions.16,23 Recently, Ployon and co-workers16 have even suggested that PRPs

68

play a protective role by scavenging tannins, blocking their access to the mucosal pellicle,

69

highlighting the importance of considering more than one mechanism when studying

70

astringency.

71

Apart from that, red wine is usually a palate-pleasing accompaniment to other foods. A

72

pairing selection can transform wine sensory components, sometimes with a positive

73

impact enhancing specific attributes, but others with a negative impact.24 Flavour

74

perception can be also strongly affected by the oral physiology, and specifically by the

75

interaction of some of these compounds with saliva.25 Therefore, we could consider the

76

effect of certain ingredients and additives present in foods on the perception of wine

77

astringency.

78

Based on these considerations, the present work aims to go further into depth on knowledge

79

of the mechanisms involved in food oral astringency, considering both salivary proteins and

80

epithelial cells, and its modulation by the presence of other substances used as specific

81

additives in food and beverages.

82

MATERIALS AND METHODS

83

Chemicals (Food additives)

84

12 food additives with different applications were used for the interaction assays: (1)

85

ascorbic acid (VWR-BDH Prolabo Leuven, Belgium), (2) ammonium bicarbonate (MP

86

Biomedicals Europe, Illkirch, France), (3) citric acid (Thermo Fisher Scientific, Pittsburgh, 5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 30

87

PA, USA), (4) calcium carbonate (Sigma-Aldrich, Gillingham, UK), (5) calcium chloride

88

(Sigma-Aldrich, Gillingham, UK), (6) EDTA (Sigma-Aldrich, Gillingham, UK), (7)

89

glycine (Sigma-Aldrich, Gillingham, UK), (8) zinc sulfate (Sigma-Aldrich, Gillingham,

90

UK), (9) sodium carbonate (VWR-BDH Chemicals UK), (10) sodium chloride (Sigma-

91

Aldrich, Gillingham, UK), (11) sodium bicarbonate (Fisher Scientific, Loughborough,

92

UK), (12) magnesium sulfate (Sigma-Aldrich, Gillingham, UK).

93

Grape Seed Tannins

94

Berries of the Zalema variety were harvested at maturity and seeds were separated from the

95

skins and pulp. Grape seeds were lyophilized and ground to obtain homogeneous powder.

96

The resulting powder was extracted three times with ethanol/water (75:25 v/v) for 30 min in

97

and ultrasonic bath. The solution was then centrifuged, and the supernatant was evaporated

98

to remove the organic solvents. The aqueous grape seed extract (GSE) was frozen and

99

freeze-dried. The resulting GSE powder was stored at 4 °C until analysis.

100

The content of monomeric and oligomeric procyanidins was determined by HPLC-DAD-

101

MS analysis. The mean degree of polymerization (mDP) was determined by acid-catalysis

102

reaction in the presence of phloroglucinol according to a method described in the

103

literature.26 The composition of the GSE is shown in Table SI-1 (Supplementary file). The

104

mDP of the total flavanols of GSE was 2.05. The content of galloylated flavanols was

105

16.67%, while the non-galloylated represent 83.33%. GSE is composed mainly of

106

monomers and dimers, with a percentage of 33.85% and 47.20%, respectively.

107

Saliva collection and treatment

6 ACS Paragon Plus Environment

Page 7 of 30

Journal of Agricultural and Food Chemistry

108

Unstimulated whole mouth saliva (WMS) was collected from five healthy individuals (25-

109

45 years) who had no history of disorders in oral perception and were not taking any

110

medication. Saliva samples were taken between 10 and 12 am at least 1 h after consuming

111

any food. Samples were collected by expectorating saliva into an ice-cooled tube. All

112

samples were pooled and centrifuged at 4000 g for 20 min at 4 °C to remove any insoluble

113

material. The resulting supernatant was aliquoted and immediately frozen at – 80 °C, which

114

is referred to as whole saliva (WS).27,28

115

Binding assay (saliva-polyphenols)

116

Interaction mixtures were prepared in a final volume of 20 µL, containing 11.5 µL of whole

117

saliva (WS), GSE (final concentration 0.6 mg/mL) and/or chemicals (final concentration 10

118

mM). Binding assays were performed in Eppendorf tubes at room temperature during 15

119

min. The mixture was then centrifuged at 13000 g for 5 min. The pellet was discarded and

120

lithium dodecyl sulfate (LDS) sample buffer (Invitrogen, Paisley, UK) was added to the

121

supernatant before loading the samples into the gel (1:4 LDS:sample). Control solutions of

122

WS with Dulbecco’s Phosphate Buffered Saline (DPBS) (Sigma-Aldrich, Gillingham, UK)

123

and WS with GSE were also included.

124

SDS-PAGE electrophoresis

125

Sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE) of samples was

126

carried out under non-reducing conditions (no dithiothreitol (DTT) and no boiling of the

127

samples). 15 µL of each sample was loaded (per well) into a NuPAGE Novex Bis-Tris 4-

128

12% resolving gels (Invitrogen, Paisley, UK). Electrophoresis was performed on a Bio-Rad

129

MiniProtean Cell electrophoresis apparatus according to manufacturer’s instructions in

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 30

130

NuPAGE MES-SDS running buffer (Invitrogen, Paisley, UK). Resolved proteins were

131

stained using a solution of 0.2% w/v Coomassie Brilliant Blue (CBB) R Staining Solution

132

(Sigma-Aldrich, Poole, Dorset, UK) in 25% methanol and 10% acetic acid for 1h. Gels

133

were then destained in 10% acetic acid overnight. The electrophoretic gels were scanned

134

using a white light transilluminator screen in a GelDocXR + Imaging System (Bio-Rad

135

Laboratories, Hercules, CA, U.S.A.). Bands were analyzed using Image Lab software (v.

136

5.1, Bio-Rad Laboratories, Hercules, CA, U.S.A.). Molecular masses were estimated using

137

protein markers (SeeBlue Plus2 Pre-Stained Standard, Invitrogen, Paisley, UK).

138

Cell culture

139

TR146 epithelial cells (ATCC, Middlesex, U.K.) were used in this study as an in vitro

140

model of human buccal epithelium.29 TR146 cells were grown in Dulbecco’s modified

141

Eagle medium (DMEM)/F12 (1:1, v/v) supplemented with 15% Foetal Bovine Serum

142

(FBS) and 1% penicillin/streptomycin (P/S) (Gibco® by Life Technologies). Cells were

143

cultured in T75 flasks and the medium was changed every two days. Cells were dissociated

144

with trypsin enzyme and subcultures were prepared at 80% confluence using Trypsin-

145

EDTA. Culture conditions were maintained at 37 ° and 7.5% CO2.

146

Interaction assays with oral cells

147

TR146 cells were cultured into 96 well flat and grown to confluence before its use in an

148

assay. The cell monolayers were washed twice with Dulbecco’s Phosphate Buffered Saline

149

(DPBS) in order to remove residual growth medium. Stock solutions of each chemical

150

compound were prepared in DPBS. Interaction assays were performed following two

151

different protocols:

8 ACS Paragon Plus Environment

Page 9 of 30

Journal of Agricultural and Food Chemistry

152

Protocol A (Figure 1A): Cells were incubated for 2h with whole saliva diluted into growth

153

medium (1:1, v/v) (50 µL/well) in order to deposit a mucosal pellicle on the cell’s surface.

154

30

155

DPBS (lines C to E, Fig. 1A) and the different chemical solutions (columns 1-12, Fig. 1A)

156

were added in triplicate (final concentration GSE: 0.6 mg/mL, chemicals 1-12: 10mM) into

157

a final volume of 45 µL (30 µL buffer + 15 µL GSE + chemicals), incubating for 15 min.

158

After two washes with DPBS (in order to remove the unbound material) the DMACA assay

159

was done. Different control conditions were also tested, in triplicate, such as the

160

GSE/DPBS and saliva/DPBS with or without oral cells (lines A and B, Fig. 1A). The

161

control condition named as C corresponds with GSE + DPBS with oral cells, while CS

162

corresponds with GSE + saliva with oral cells.

163

Protocol B (Figure 1B): Whole saliva was mixed with GSE and the chemical solutions,

164

incubating during 15 min (30 µL/well saliva or DPBS + 15 µL GSE + chemicals), before

165

adding to the cells. 45 µL/well of the mixture was added in triplicate (lines C-E: DPBS +

166

GSE + chemicals; lines F-H: saliva + GSE + chemicals, see Fig. 1B) to attain the final

167

concentration GSE: 0.6 mg/mL, chemicals 1-12: 10mM and left in contact with the

168

monolayers for 15 min. After two washes with DPBS, the DMACA assay was performed.

169

The same control conditions as in Protocol A were tested.

170

DMACA bioassay

171

4-(dimethylamino)cinnamaldehyde (DMACA) was purchased from Sigma-Aldrich as a

172

chromogenic reagent for indoles and flavanols. DMACA assay was performed since this

173

compound reacts selectively with catechins and procyanidins to form a blue-green product,

174

allowing us to visualize and quantify the amount of procyanidins that remains bound to the

After incubation, samples were washed twice with DPBS. GSE (lines F to H, Fig. 1A) or

9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 30

175

epithelial cells.18,31 A 0.1% of DMACA solution was prepared in acidified methanol (0.75

176

M H2SO4)18 and added to the 96 well plates after the interaction assays. Cells were

177

incubated with 70 µL/well of the reagent for 20 min at room temperature. Finally, the

178

absorbance of the wells at 655 nm was determined in a microtiter plate reader.

179

Statistical analysis

180

In order to determine statistical significance of the differences between the absorbance

181

values obtained for the interaction assays, data were evaluated by a t-student test for

182

independent samples, using the software packing for Windows IBM SPSS 23 (SPSS, Inc.

183

Chicago, IL, USA). Differences were considered to be statistically significant when p