Understanding the Relationship between Red Wine Matrix, Tannin

Nov 2, 2016 - To improve our understanding of the role that the enthalpy of interaction of tannin with a hydrophobic surface (tannin activity) has in ...
0 downloads 10 Views 724KB Size
Subscriber access provided by UNIV TORONTO

Article

Understanding the relationship between red wine matrix, tannin activity and sensory properties. Aude Annie Watrelot, Nadia K. Byrnes, Hildegarde Heymann, and James A. Kennedy J. Agric. Food Chem., Just Accepted Manuscript • DOI: 10.1021/acs.jafc.6b03767 • Publication Date (Web): 02 Nov 2016 Downloaded from http://pubs.acs.org on November 5, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Agricultural and Food Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29

Journal of Agricultural and Food Chemistry

1 2

Understanding the relationship between red wine matrix, tannin activity and sensory properties.

3 4 5

Aude A. Watrelot,abϯ* Nadia K. Byrnes,bϯ Hildegarde Heymann,b James A. Kennedy.bc

6 7 8 9 10 11

a

Department of Viticulture and Enology, California State University, 2360 East

Barstow Avenue, MS VR89, Fresno, CA 93740-8003, USA b

Department of Viticulture and Enology, University of California at Davis, One

Shields Ave., Davis, CA 95616-5270, USA c

Constellation Brands, Inc., 12667 Road 24, Madera, California, USA

12 13

ϯ

14

*corresponding author:

15

A. A. Watrelot

These authors contributed equally to this article

16

Department of Viticulture and Enology, University of California at Davis, One Shields Ave.,

17

Davis, CA 95616-5270, USA

18

Tel: 530-752-5054

19

Fax: 530-752-0382

20

E-mail: [email protected] 1 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 2 of 29

21

Abstract

22

One major red wine mouthfeel characteristic, astringency, is derived from grape-

23

extracted tannins and is considered to be a result of interaction with salivary proteins and the

24

oral mucosa. To improve our understanding of the role that the enthalpy of interaction of

25

tannin with a hydrophobic surface (tannin activity) has in astringency perception, a

26

chromatographic method was used to determine the tannin concentration and activity of 34

27

Cabernet Sauvignon wines, as well as sensory analysis done on 13 of those wines. In addition,

28

astringency-relevant matrix parameters (pH, titratable acidity, ethanol, glucose, and fructose)

29

were measured across all wines. Tannin activity was not significantly correlated with any

30

matrix variables and the perception of drying and grippy was not correlated with tannin

31

concentration and activity. However, ethanol content was well related to mouthfeel attributes

32

and appeared to drive perceived drying. Although fructose and glucose content were well

33

correlated, they did not drive the perception of sweetness, that is explained by the well-known

34

mixture suppression effect.

35

Keywords

36 37

Tannins, red wines, concentration, activity, hydrophobic interaction, pH, ethanol, residual

38

sugar.

39 40

41

42 2 ACS Paragon Plus Environment

Page 3 of 29

Journal of Agricultural and Food Chemistry

43

Introduction

44

In red wine, condensed tannins are the main components responsible for astringency.1

45

They are composed primarily of proanthocyanidins extracted from grape skin and seed during

46

fermentation and maceration. Proanthocyanidins are polymeric flavan-3-ols linked via

47

interflavan bonds between C4-C8 and C4-C6 and their constitutive units are mostly (+)-

48

catechin,

49

astringency response from wine tannins is considered to be a result of tannin interaction with

50

salivary proteins through hydrophobic and hydrogen bond interactions, and subsequent

51

protein aggregation and precipitation.4–7 Tannin structure modification(s) as a result of wine

52

production influence(s) these non-covalent interactions. The mean degree of polymerization

53

(mDP) of tannins, that is the average number of constitutive units, in red wine can vary from

54

monomers to polymers (up to 30 in grape skins). In model solutions monomers are perceived

55

to be more bitter than astringent, but as the mDP increases so does astringency perception.8,9

56

In addition to the level of polymerization, the tannin subunit composition is considered to be

57

an important variable in determining astringency.10 (-)-Epicatechin and (+)-gallocatechin are

58

known to increase the perception of astringency, as well as the proportion of galloylation in

59

contrast to (-)-epigallocatechin.9,10 In addition, winemaking processes and overall wine age

60

have an effect on astringency.11 The perception of tannins in red wine has also been shown to

61

vary with matrix composition, as explained below.

(-)-epicatechin,

(-)-epicatechin-3-O-gallate

and

(-)-epigallocatechin.2,3

The

62

The main components of red wine, in addition to polyphenols, are polysaccharides,

63

residual sugars, alcohol, and organic acids. During the crushing and pressing of grapes, plant

64

cells are disrupted and polysaccharides of cell walls become available to bind to polyphenols,

65

creating interactions that can influence the astringency of the wine.12 Among these reactions

66

are the non-covalent interactions between pectins (mostly rhamnogalacturonan II) and

3 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

67

Page 4 of 29

tannins,13–15 as well as mannoproteins from yeast cell walls and tannins.16,17

68

During fermentation, glucose and fructose are converted into ethanol and carbon

69

dioxide. Residual sugars found in finished red wines (mostly fructose)18 are responsible for

70

the sweetness, but as expected, no direct relationships with astringency have been found.

71

Rinaldi et al.19 have shown that an increase in fructose concentration in tannin-salivary

72

protein complexes leads to a reduction in the precipitation of salivary proteins. Ethanol in

73

wine has been shown to enhance bitterness20 and reduce astringency.21,22 The effect of ethanol

74

on tannin-salivary proteins complexes is not well understood, but current literature suggests

75

that it decreases the strength of interaction between tannin and protein.23

76

Wine pH depends on the total amount of acid present, the ratio of malic acid to tartaric

77

acid and the quantity of potassium.18 The organic acid type has been shown to not effect

78

astringency,24 while an increase in pH has been shown to decrease salivary protein

79

precipitation19 and astringency.25 Variation in pH, and the effect of titratable acidity on

80

tannin-salivary protein interactions is not well studied and understood.

81

It is well known that astringency perception is due to associations between salivary

82

proteins (proline-rich proteins) and tannins through hydrophobic interactions.26,27 Based upon

83

the ability of tannins to form hydrophobic interactions with proteins, a high performance

84

liquid chromatography (HPLC) method was developed to determine the activity of red wine

85

tannins.28 Tannin activity is defined as the enthalpy of interaction between wine tannins and a

86

hydrophobic surface. Astringency has been shown, in sensory analysis, to vary with regard to

87

tannin concentration as well as other matrix components such as ethanol content, residual

88

sugar and acidity, as explained above; however, these variables are not adequate predictors of

89

overall astringency perception. In addition, it has been suggested that the activity of tannins is

90

also involved in the perception of mouthfeel characteristics.29

4 ACS Paragon Plus Environment

Page 5 of 29

Journal of Agricultural and Food Chemistry

91

The purpose of this study was to improve our understanding of the relationship

92

between wine tannin concentration and activity relative to other astringency-modifying matrix

93

components. An important assumption made in this study was that by narrowing the focus on

94

recently bottled red wines cv. Cabernet Sauvignon primarily from California, the relationship

95

between winemaker expectations with regard to mouthfeel and corresponding tannin

96

chemistry and other matrix parameters would be simplified. The study was focused on tannin

97

concentration and activity as well as matrix parameters across 34 wines to determine how all

98

parameters might be related to sensory perception.

99 100

Materials and methods

101

Chemicals

102

All chemical solvents used were HPLC-grade. Acetonitrile and ortho-phosphoric acid

103

were purchased from VWR International (Radnor, PA). (-)-epicatechin (purity ≥ 90%) was

104

purchased from Sigma-Aldrich (St Louis, MO). All water was purified using an Ultrapure

105

purification system (Evoqua Corporation, Alpharetta, GA).

106

Wine samples

107

The 34 wines were sourced from Cabernet Sauvignon cultivar in California,

108

Washington, and Australia. Amongst the 34 wines, 27 were from 2012, 5 from 2013, 1 from

109

2010 and 2011 (Table 1 in Supporting Information). All wines were bottled between January

110

and June 2014. All wines were filtered using a 13 mm PTFE syringe filters (0.45 µm, Grace

111

Davison Discovery Science, Deerfield, IL, USA) prior to analysis.

112

Matrix parameters

5 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 6 of 29

113

A Winescan FT120 (FOSS, Eden Prairie, Minnesota, USA) was used to determine the

114

pH, ethanol concentration, titratable acidity (TA), fructose, and glucose concentrations of

115

wines in triplicate sampling from the same bottle.

116

Tannin activity and concentration

117

The HPLC method for measuring the activity of wine tannin has been described

118

previously.28,29 Briefly, the HPLC method used a polystyrene divinylbenzene reversed-phase

119

column (PLRP-S, 2.1 × 50 mm, 100 Å, 3 µm, Agilent Technologies, Santa Clara, CA)

120

protected with a guard column (PRP-1, 3 × 8 mm, Hamilton Company, Reno, NV), with DAD

121

detection at 280 nm. The mobile phases consisted of 1.5 % (w/w) ortho-phosphoric acid in

122

water (180 mM, mobile phase A) and 20 % (v/v) mobile phase A in acetonitrile (mobile phase

123

B) with a flow rate of 0.3 mL/min. The linear gradient was as follows (time in min (%B)): 0

124

(14), 12.6 (34), 13.3 (34), 15.05 (70), 16.8 (70), 19.6 (14), and 28 (14).

125

To determine thermodynamic information, samples were run at four column

126

temperatures (25-40 °C, 5 °C increments), and temperatures were converted to Kelvin for

127

calculations. Chromatograms at 280 nm were baseline-subtracted using a water as a blank

128

injection and were integrated as previously described.28,29 Briefly, a baseline was drawn at 0

129

mAU and with the resulting area clipped at 5 and 28 min for total tannin (TanninT); partial

130

tannin (TanninP), that corresponded to polymers, was the peak area eluting between 16.8 and

131

28 min. For each chromatogram an alternative retention factor for the tannin (kalt) was

132

calculated as follows:

 =

133

Tannin

Tannin − Tannin

The ln (kalt) is related to thermodynamic information as follows:

6 ACS Paragon Plus Environment

Page 7 of 29

Journal of Agricultural and Food Chemistry

 ( ) = −

ΔH° ΔS° + RT R

134

Where ∆H° and ∆S° are the enthalpy and the entropy (respectively) of the interaction,

135

R is the gas constant, T is the temperature of the experiment in Kelvin. The specific enthalpy

136

was calculated by the slope from the van’t Hoff plot (i.e. ln kalt versus the reciprocal of the

137

column temperature in Kelvin at each of the four temperatures).29 A purified grape skin tannin

138

isolate was used as an enthalpy of interaction control.

139

Tannin concentrations were determined by measuring total tannin peak area (TanninT)

140

at 280 nm and comparing this with an (-)-epicatechin quantitative standard. For comparison,

141

tannin concentrations were also determined by protein precipitation as previously described

142

by Harbertson et al. and Kennedy et al.30,31 In that case, the concentration of tannins was

143

expressed in (+)-catechin equivalents.

144

Sensory analysis

145

Panelists were recruited from UC Davis students, staff, university affiliates, and

146

community members to take part in a descriptive analysis panel on 13 of the Cabernet

147

Sauvignon wines (in bold in Supporting Information Table 1). Panelists were screened for age

148

(over 21), interest in wine, and availability to complete all the training and panel sessions.

149

During nine one-hour training sessions 12 panelists (5 men) sampled the study wines in

150

duplicate. In these sessions they generated terms describing the aroma/flavor, taste and

151

mouthfeel properties of the wines and came to a consensus about the references for aroma,

152

taste, and mouthfeel descriptors. During these training sessions panelists also came to

153

consensus regarding the definition of the mouthfeel attributes that were used in the final

154

evaluation. The definitions provided by Gawel and colleagues32 were used as a starting point

155

for these definitions. The final list (Table 2) contained 14 aroma/flavors, 4 tastes, and 5

7 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 8 of 29

156

mouthfeel terms. All aroma standards were presented to the panelists at the beginning of each

157

evaluation session. Taste and mouthfeel references were presented only at the first evaluation

158

session.

159

Panelists’ evaluation sessions were performed in isolated, temperature-controlled (20

160

°C) tasting booths. All wines and attribute references were presented under white light in

161

black, pear-shaped ISO glasses (ISO 1977) covered with clear plastic petri dishes. Prior to

162

assessing any of the wines, panelists were given a quiz on the aroma references. In a booth

163

identical to those used for evaluation, a full set of aroma references were presented in random

164

order with blinding codes and panelists were asked to match the blinded attribute references

165

to the appropriate attribute names. Each time the panelists completed the aroma quiz they

166

were presented with the same aroma references, though different blinding codes were used

167

and the presentation order was changed.

168

Panelists were presented with four or five wines during this portion of the evaluation

169

session, presented in randomized blocks and labeled with random three-digit codes. Each

170

sample consisted of 20 ml of wine presented at room temperature (20°C). Panelists evaluated

171

each wine monadically were asked during their 1-minute break between samples to cleanse

172

their palate with the filtered water (Arrowhead, Nestle, Stamford, CT) and unsalted top saltine

173

crackers (Nabisco, Mondelez Ltd, East Hanover, NJ) provided. All wine was expectorated.

174

Panelists rated the wine for each of the aroma attributes before putting the wine in their

175

mouths. Panelists held the wine in their mouth for 30 seconds, during which they rated the

176

taste and flavor attributes. After expectorating the wine, panelists rated the mouthfeel

177

attributes. Rating was performed using a computer and mouse on a 10-cm line scale ranging

178

from “none” on the left end of scale to “a lot” on the right end of the scale. This scale was

179

marked only with tick marks at the ends. Panelists evaluated all of the wine samples in

180

triplicate, completing in total nine evaluation sessions (13 wines × 3 replicates, 4 to 5 wines in 8 ACS Paragon Plus Environment

Page 9 of 29

Journal of Agricultural and Food Chemistry

181

each of 9 sessions). Data was collected and compiled using Fizz version 2.45A (Biosystèmes,

182

Couternon, France).

183

Statistical analysis

184

Exploratory data analysis of the chemical data was conducted using Principal

185

Components Analysis (PCA) on the correlation matrix of the averaged data set. Descriptive

186

Analysis (DA) data were analyzed by three-way multivariate analysis of variance

187

(MANOVA) with fixed-effect model for the wine, judge, and replication effects with all two-

188

way interactions. Following this, three-way univariate analyses of variance (ANOVAs) for

189

each attribute were used with a fixed-effect model for the wine, judge, and replication effects

190

with all two-way interactions. For those attributes that showed significant main effects of

191

wine as well as a significant wine-judge interaction, a pseudo-mixed model was used.33

192

Canonical variate analysis (CVA) was used to visualize the data, with 95% confidence circles

193

(Chatfield and Collins 1980).34 Partial Least Squares Regression (PLSR) was used to correlate

194

the chemical and the sensory data sets to one another. The data was standardized prior to

195

analysis and the leave-one-out cross-validation method was used.

196

All statistical analyses and graphs were prepared using RStudio,35 with the

197

SensoMineR36, car37, agricolae38, candisc39, and pls40 packages. All statistics were interpreted

198

using alpha = 0.05.

199

200

Results

201

Matrix variables, tannin concentration and activity

202

Summary statistics of all wine matrix parameters are shown in Table 1. Full chemical

203

data is shown in Supporting Information Table 1. The tannin concentration determined by 9 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 10 of 29

204

HPLC varied across the 34 wines from 2.75 to 6.16 g/L (mean ± standard error: 4.18 ± 0.76),

205

while tannin activity varied between -4.82 and -1.43 kJ/mol (-2.74 ± 0.69). Overall, the tannin

206

concentration values measured by the protein precipitation method varied from 0.35 to 1.34

207

g/L (0.66 ± 0.21) and were lower than when measured by HPLC, with an average of 4.18 g/L

208

by HPLC versus 0.66 g/L by protein precipitation (Table 1; Supporting Information Table 1).

209

The two methods were well correlated with a r of 0.84 (p = 0.00). Among the 34 Cabernet

210

Sauvignon wines, matrix variables varied as follows. The ethanol concentration varied from

211

13.04 to 17.05 v/v %, with an average of 14.44 (± 0.81). The TA varied from 4.48 to 6.22 g/L

212

(5.26 ± 0.40) and the pH varied from 3.61 to 4.02 (3.76 ± 0.11). The fructose concentration

213

varied from 1.00 to 3.85 g/L with an average of 2.16 (± 0.70) and the glucose concentration

214

was much lower, varying from not detectable to 1.75 g/L (0.24 ± 0.62). As observed in the

215

Supporting Information Table 1, the wine containing the highest ethanol concentration had the

216

highest tannin concentration, the lowest titratable acidity and one of the highest pH values.

217

The fructose and the glucose concentrations of this wine were slightly higher than the average

218

(3.16 g/L and 1.16 g/L, respectively). In contrast, the wine with the least ethanol

219

concentration did not contain extreme amounts of other matrix variables.

220

In order to understand the relationships between matrix parameters and tannin activity,

221

a principal component analysis (PCA) was performed across the 34 wines (Figure 1). In total,

222

66.8% of the variance was explained by the first two components (PC1 = 34.5%, PC2 =

223

32.3%). Regarding the loading plot, ethanol concentration was situated quite close to pH and

224

tannin activity but this last parameter was not significantly correlated with any matrix

225

variables. Fructose and glucose were positively correlated (r = 0.88, p = 0.00) and ethanol

226

concentration was negatively correlated with titratable acidity (r = -0.38, p = 0.03) and with

227

pH (r = 0.76, p = 0.00). Tannin concentration was significantly correlated with ethanol

228

concentration (r = 0.39, p = 0.02), fructose (r = 0.46, p = 0.01), glucose (r = 0.40, p = 0.02), 10 ACS Paragon Plus Environment

Page 11 of 29

229

Journal of Agricultural and Food Chemistry

and pH (r = 0.56, p = 0.00).

230 231

Sensory analysis

232

Table 2 shows the attributes and attribute definitions and references that were

233

determined during the descriptive panel. Three-way MANOVA showed significant

234

differences between the wines. ANOVA of the attributes generated in descriptive analysis

235

showed that the outdoors aroma and flavor, vanilla oak aroma and flavor, fresh berry flavor,

236

green pepper flavor, tamarind flavor, sweetness, and the mouthfeel attributes grippy, drying,

237

and viscosity were significant for the wines. The means and least significant difference values

238

for these attributes are shown in Supporting Information Table 2.

239

CVA showed that three dimensions were most appropriate to represent the results

240

from descriptive analysis (Figure 2). The first two dimensions of CVA accounted for 78.7%

241

of the variance ratio (Figure 2a). The third dimension accounted for another 16.3% (Figure

242

2b). The first dimension was characterized on the positive side by fresh berry flavor and sweet

243

taste, and on the negative side by outdoors flavor and aroma and the mouthfeel attributes

244

grippy and drying. The second dimension was characterized by vanilla oak flavor and aroma

245

(negative) and green pepper flavor (positive). The third dimension was characterized by

246

tamarind flavor and the mouthfeel attributes drying and viscosity at the positive end.

247

There was a high level of overlap in the 95% confidence circles, as shown in Figure 2a

248

and 2b, indicating that the panelists did not perceive large differences between these wines.

249

Along the first dimension wines 3 and 13 were the most different, differentiated along this

250

axis by the mouthfeel characteristics of grippy and drying (high for wine 13) and tamarind

251

and fresh berry flavors (higher for wine 3). While wine 3 had the highest rating of sweet taste

252

and highest fructose concentration, these were not significantly higher than the other wines, 11 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 12 of 29

253

and the perceived sweetness in the wines was not correlated to either fructose or glucose

254

concentration in the wines. Along the second dimension, wines were distinguished by green

255

pepper flavor (wines 7 and 8) and vanilla oak flavor (wine 5), though none of these wines

256

were significantly different than the other wines for these attributes (Table 2). Wines 7 and 8

257

were the Australian wines, both from the same producer, which may account for the

258

similarities to one another and for the differences from the other wines in the study.

259

In Figure 2b, there was clear separation of the tamarind and fresh berry flavors along

260

the third dimension, which in Figure 2a appears to overlap. On this plot, it can be seen that

261

wine 3 was significantly higher in the fresh berry flavor than the other wines, while wines 9,

262

10, and 11 were characterized more by the tamarind flavor. Again however, there was not

263

much distinction between most of the wines along the third dimension. The wine that was the

264

most different from the rest of the wines in the study was wine 13. This wine was rated the

265

highest for both mouthfeel characteristics of grippy and drying. This was not unexpected, as

266

the tannin concentration in this wine was the highest of all of the wines, as was the tannin

267

activity. Interestingly however, the perception of drying and grippy was not correlated with

268

the tannin concentration or tannin activity. Instead, perception of drying showed a moderate

269

correlation with ethanol concentration, suggesting that the ethanol concentration was driving

270

the perception of the drying sensation. Overall, the mouthfeel characteristic drying was

271

significantly correlated with grippy (r = 0.66, p = 0.01) and viscosity (r = 0.61, p = 0.03),

272

however grippy and viscosity were not significantly correlated.

273 274

Regressing sensory on chemical variables

275

Figure 3 shows the output of PLSR of sensory attributes on the chemical variables.

276

Collectively, the first three model components account for roughly 80% of the total variance 12 ACS Paragon Plus Environment

Page 13 of 29

Journal of Agricultural and Food Chemistry

277

of the predictor matrix (chemical variables, Component 1: 40.82% , Component 2: 23.12%,

278

Component 3 – not shown: 17.00%). The model did not improve by adding additional

279

components. Overall, the chemical variables predicted the sensory variables moderately well,

280

however only the sensory variables viscosity, grippy, drying (all mouthfeel), outdoors flavor,

281

chemical aroma, and fresh berry flavor were sufficiently explained by the model.

282

Correlations between the sensory and chemical variables were observed. pH and

283

titratable acidity were related to chemical aroma (r = 0.79, p = 0.00 and r = -0.68, p = 0.01,

284

respectively). Titratable acidity was also related to the fresh berry flavor (r = 0.6, p = 0.03).

285

Outdoors aroma was correlated to fructose concentration (r = 0.59, p = 0.03) and tannin

286

activity was related to tamarind flavor (r = 0.57, p = 0.04). The only mouthfeel attribute to be

287

significantly related to any chemical variables was drying, which was related to ethanol

288

concentration (r = 0.60, p = 0.03).

289 290

Discussion

291

In an effort to reduce variability in the samples, we chose to focus on wines, primarily

292

from California of similar vintage. Doing so, we assumed that mouthfeel and taste aspects

293

(bitterness, astringency, sweetness, etc.) of perception would be similarly balanced. In the

294

present study, the relationships between tannin activity and other chemical components

295

related to mouthfeel were first investigated for 34 Cabernet Sauvignon wines produced in

296

various regions. Subsequently, 13 of the wines were characterized using sensory descriptive

297

analysis. In the methodology used here, the dynamic system (gradient by HPLC) and the use

298

of a hydrophobic surface to determine tannin activity (non-covalent interactions between

299

tannins and a hydrophobic surface) did not involve precipitation/aggregation. The astringency

300

mouthfeel has been described as the non-covalent interactions between tannins and salivary 13 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 14 of 29

301

proteins through hydrogen bonds and hydrophobic interactions followed by aggregation and

302

complex precipitation, leading to the drying sensation.41 No significant correlation between

303

tannin activity, matrix variables and with any mouthfeel attributes was observed in this study

304

which could be explained by the chemical method used, where the tannin activity correspond

305

to the first step of interactions (hydrophobic interaction between tannins and a hydrophobic

306

surface) and not necessarily to the protein precipitation that is known to lead to the in-mouth

307

perceptions. Alternatively a more complex set of matrix interactions with tannins, insufficient

308

tannin activity variation or missing analytically-relevant information (e.g.: polysaccharides)

309

may have influenced these results.

310

Red wines contain many grape-derived compounds (e.g., polyphenols, organic acids,

311

sugars) as well as fermentation-derived compounds (e.g. ethanol) that contribute to the

312

mouthfeel characteristics of the wine.18 In contrast to the literature where the ethanol has

313

been previously related to an increase of the bitterness and burning sensation of wines20,42 and

314

to a decrease of astringency intensity,43,44 in this study, the ethanol concentration was well

315

correlated to the pH and to mouthfeel attributes such as grippy, drying and viscosity. In the

316

case of 13 Cabernet Sauvignon mostly from the 2012 vintage and from California, the ethanol

317

concentration was related to an increase in grippy, drying sensation and of the viscosity

318

perception. A principal purpose of wine production is to produce a “balanced” wine, meaning

319

that an increase in tannin concentration would be associated with an increase in ethanol for

320

attenuating the wine astringency that would result from tannin concentration alone. This was

321

shown in the present study, where a positive correlation between ethanol and tannin

322

concentration was observed. The pH has been found to be positively related to tannin

323

concentration and ethanol concentration as well as to mouthfeel attributes. This result is in

324

agreement with Kallithraka et al.24 who showed that a decrease in the pH increased the

325

maximum intensity and the total duration of astringency in model solutions and red wine as 14 ACS Paragon Plus Environment

Page 15 of 29

Journal of Agricultural and Food Chemistry

326

well as with Obreque-Slier et al.45 Similarly, an increase in wine pH has been shown to lead to

327

a decrease in salivary protein precipitation,19 even if in our sensory analysis, pH was not

328

clearly related to any taste or mouthfeel.

329

When present, fructose was found to be the most significant residual sugar in this set

330

of wines. Some researchers have found that an increase in fructose concentration decreased

331

the perceived astringency and decreased salivary protein precipitation,19,46 while others have

332

shown little effect on mouthfeel dryness.42 In our study, the fructose was correlated to glucose

333

in 34 wines and not to any other matrix variables, suggesting that for Cabernet Sauvignon

334

wines, residual sugars are not the main drivers of mouthfeel perception. Also in the

335

descriptive analysis on 13 wines, it is possible, that due to well-known mixture suppression

336

effects, which could arise from the perception of bitterness47–49 or astringency50 in the wines,

337

the perception of sweetness was suppressed, accounting for the lack of correlation between

338

fructose or glucose and sweet taste in the wines.

339

The lack of a relationship between the tannin concentration and activity and sensory

340

perception of astringency observed in this study could be due to a set of wines that were too

341

perceptually similar for participants to distinguish between. It is also possible that interactions

342

between aroma, flavor, and astringent compounds in the wines altered the perception of

343

astringency in these wines. Literature suggests that aromatic compounds and phenolic

344

compounds in wines and other foods and beverages may interact to suppress the perception of

345

each sensation.51-53 It is also possible that there are more complex matrix effects impacting the

346

perception of astringency in these wines. As documented by Ares and colleagues, the

347

presence of astringent compounds can suppress the perception of sweetness in wines.

348

Additionally, the presence of certain aromas, such as fruity or berry aromas, can enhance the

349

perceptually sweetness of a food or beverage without any increase in the concentration of

350

sweet tastant.54 It is possible that the interaction between the cherry and fresh fruit/berry 15 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 16 of 29

351

aromas and flavors, as well as the perceived sweetness, present in these wines directly or

352

indirectly impacted the perception of astringent qualities in this sample set of wines.

353

Additional work in the area of mixture suppression within the wine matrix is needed to fully

354

clarify the impact that the various components and tannin activity have on the perception of

355

astringency in red wines.

356

Acknowledgements

357 358

The authors thank all of the wineries that contributed wines to this study.

359

Funding sources

360 361

We thank the American Vineyard Foundation (AVF) for project funding (UGMVE # 2015-

362

1691).

363

364

Supporting Information

365

Tannin concentration and activity data and matrix variables values (pH, ethanol, titratable

366

acidity, fructose and glucose) of 34 red wines (cv. Cabernet Sauvignon). Means and least

367

significant difference values for the descriptive analysis attributes of red wines. This material

368

is available free of charge via the Internet at http://pubs.acs.org.

369 370

References

16 ACS Paragon Plus Environment

Page 17 of 29

371

Journal of Agricultural and Food Chemistry

1.

372 373

Wine Polyphenols. Aust. J. Grape Wine Res. 2001, 7 (1), 33–39. 2.

374 375

3.

4.

J.-M.;

Cheynier,

V.;

Brossaud,

F.;

Moutounet,

M.

Polymeric

Prinz, J. F.; Lucas, P. W. Saliva Tannin Interactions. J. Oral Rehabil. 2000, 27 (11), 991–994.

5.

380 381

Souquet,

Proanthocyanidins from Grape Skins. Phytochemistry 1996, 43 (2), 509–512.

378 379

Prieur, C.; Rigaud, J.; Cheynier, V.; Moutounet, M. Oligomeric and Polymeric Procyanidins from Grape Seeds. Phytochemistry 1994, 36 (3), 781–784.

376 377

Brossaud, F.; Cheynier, V.; Noble, A. C. Bitterness and Astringency of Grape and

Freitas, V. de; Mateus, N. Nephelometric Study of Salivary Protein–tannin Aggregates. J. Sci. Food Agric. 2002, 82 (1), 113–119.

6.

Noble, A. C. Astringency and Bitterness of Flavonoid Phenols. In Chemistry of Taste:

382

Mechanisms, Behaviors, and Mimics; Given, P., Paredes, D., Eds.; Amer Chemical

383

Soc: Washington, 2002; Vol. 825, pp 192–201.

384

7.

385

Bajec, M. R.; Pickering, G. J. Astringency: Mechanisms and Perception. Crit. Rev. Food Sci. Nutr. 2008, 48 (9), 858–875.

386

8.

Noble, A. Bitterness in Wine. Physiol. Behav. 1994, 56 (6), 1251–1255.

387

9.

Vidal, S.; Francis, L.; Guyot, S.; Marnet, N.; Kwiatkowski, M.; Gawel, R.; Cheynier,

388

V.; Waters, E. J. The Mouth-Feel Properties of Grape and Apple Proanthocyanidins in

389

a Wine-like Medium. J. Sci. Food Agric. 2003, 83 (6), 564–573.

390

10.

Quijada-Morín, N.; Regueiro, J.; Simal-Gándara, J.; Tomás, E.; Rivas-Gonzalo, J. C.;

391

Escribano-Bailón, M. T. Relationship between the Sensory-Determined Astringency

392

and the Flavanolic Composition of Red Wines. J. Agric. Food Chem. 2012, 60 (50),

393

12355–12361.

394 395

11.

Chira, K.; Pacella, N.; Jourdes, M.; Teissedre, P.-L. Chemical and Sensory Evaluation of Bordeaux Wines (Cabernet-Sauvignon and Merlot) and Correlation with Wine Age.

17 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

396 397

Page 18 of 29

Food Chem. 2011, 126 (4), 1971–1977. 12.

Quijada-Morín, N.; Williams, P.; Rivas-Gonzalo, J. C.; Doco, T.; Escribano-Bailón, M.

398

T. Polyphenolic, Polysaccharide and Oligosaccharide Composition of Tempranillo Red

399

Wines and Their Relationship with the Perceived Astringency. Food Chem. 2014, 154,

400

44–51.

401

13.

Carvalho, E.; Mateus, N.; Plet, B.; Pianet, I.; Dufourc, E.; De Freitas, V. Influence of

402

Wine Pectic Polysaccharides on the Interactions between Condensed Tannins and

403

Salivary Proteins. J Agric Food Chem 2006, 54 (23), 8936–8944.

404

14.

Watrelot, A. A.; Le Bourvellec, C.; Imberty, A.; Renard, C. M. G. C. Interactions

405

between Pectic Compounds and Procyanidins Are Influenced by Methylation Degree

406

and Chain Length. Biomacromolecules 2013, 14, 709–718.

407

15.

Watrelot, A. A.; Le Bourvellec, C.; Imberty, A.; Renard, C. M. G. C. Neutral Sugar

408

Side Chains of Pectins Limit Interactions with Procyanidins. Carbohydr. Polym. 2014,

409

99, 527–536.

410

16.

Escot, S.; Feuillat, M.; Dulau, L.; Charpentier, C. Release of Polysaccharides by Yeasts

411

and the Influence of Released Polysaccharides on Colour Stability and Wine

412

Astringency. Aust. J. Grape Wine Res. 2001, 7 (3), 153–159.

413

17.

Mazauric, J.-P.; Salmon, J.-M. Interactions between Yeast Lees and Wine Polyphenols

414

during Simulation of Wine Aging: I. Analysis of Remnant Polyphenolic Compounds in

415

the Resulting Wines. J. Agric. Food Chem. 2005, 53 (14), 5647–5653.

416

18.

Conde, C.; Silva, P.; Fontes, N.; Dias, A. C. P.; Tavares, R. M.; Sousa, M. J.; Agasse,

417

A.; Delrot, S.; Geros, H. Biochemical Changes throughout Grape Berry Development

418

and Fruit and Wine Quality. Food. 2007, 1, 1-22.

419 420

19.

Rinaldi, A.; Gambuti, A.; Moio, L. Precipitation of Salivary Proteins After the Interaction with Wine: The Effect of Ethanol, pH, Fructose, and Mannoproteins. J.

18 ACS Paragon Plus Environment

Page 19 of 29

Journal of Agricultural and Food Chemistry

421 422

Food Sci. 2012, 77 (4), C485–C490. 20.

423 424

Fischer, U.; Noble, A. The Effect of Ethanol, Catechin Concentration, and Ph on Sourness and Bitterness of Wine. Am. J. Enol. Vitic. 1994, 45 (1), 6–10.

21.

Vidal, S.; Courcoux, P.; Francis, L.; Kwiatkowski, M.; Gawel, R.; Williams, P.;

425

Waters, E.; Cheynier, V. Use of an Experimental Design Approach for Evaluation of

426

Key Wine Components on Mouth-Feel Perception. Food Qual. Prefer. 2004, 15 (3),

427

209–217.

428

22.

Casassa, L. F.; Larsen, R. C.; Beaver, C. W.; Mireles, M. S.; Keller, M.; Riley, W. R.;

429

Smithyman, R.; Harbertson, J. F. Sensory Impact of Extended Maceration and

430

Regulated Deficit Irrigation on Washington State Cabernet Sauvignon Wines. Am. J.

431

Enol. Vitic. 2013, 64 (4), 505–514.

432

23.

McRae, J. M.; Ziora, Z. M.; Kassara, S.; Cooper, M. A.; Smith, P. A. Ethanol

433

Concentration Influences the Mechanisms of Wine Tannin Interactions with Poly(l-

434

Proline) in Model Wine. J. Agric. Food Chem. 2015, 63 (17), 4345–4352.

435

24.

436 437

Affected by Malic and Lactic Acid. J. Food Sci. 1997, 62 (2), 416–420. 25.

438 439

Kallithraka, S.; Bakker, J.; Clifford, M. N. Red Wine and Model Wine Astringency as

Kallithraka, S.; Bakker, J.; Clifford, M. N. Effect of pH on Astringency in Model Solutions and Wines. J. Agric. Food Chem. 1997, 45 (6), 2211–2216.

26.

Pascal, C.; Poncet-Legrand, C.; Imberty, A.; Gautier, C.; Sarni-Manchado, P.;

440

Cheynier, V.; Vernhet, A. Interactions between a Non Glycosylated Human Proline-

441

Rich Protein and Flavan-3-Ols Are Affected by Protein Concentration and

442

Polyphenol/Protein Ratio. J. Agric. Food Chem. 2007, 55 (12), 4895–4901.

443

27.

Poncet-Legrand, C.; Edelmann, A.; Putaux, J.-L.; Cartalade, D.; Sarni-Manchado, P.;

444

Vernhet, A. Poly(l-Proline) Interactions with Flavan-3-ols Units: Influence of the

445

Molecular Structure and the Polyphenol/Protein Ratio. Food Hydrocoll. 2006, 20 (5),

19 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

446 447

687–697. 28.

448 449

Page 20 of 29

Barak, J. A.; Kennedy, J. A. HPLC Retention Thermodynamics of Grape and Wine Tannins. J. Agric. Food Chem. 2013, 61, 4270–4277.

29.

Revelette, M. R.; Barak, J. A.; Kennedy, J. A. High-Performance Liquid

450

Chromatography Determination of Red Wine Tannin Stickiness. J. Agric. Food Chem.

451

2014, 62, 6626–6631.

452

30.

Harbertson, J. F.; Kennedy, J. A.; Adams, D. O. Tannin in Skins and Seeds of Cabernet

453

Sauvignon, Syrah, and Pinot Noir Berries during Ripening. Am. J. Enol. Vitic. 2002, 53

454

(1), 54–59.

455

31.

Kennedy, J. A.; Ferrier, J.; Harbertson, J. F.; Gachons, C. P. des. Analysis of Tannins in

456

Red Wine Using Multiple Methods: Correlation with Perceived Astringency. Am. J.

457

Enol. Vitic. 2006, 57 (4), 481–485.

458

32.

Gawel, R.; Oberholster, A.; Leigh Francis, I. A “Mouth-Feel Wheel”:terminology for

459

Communicating the Mouth-Feel Characteristics of Red Wine. Aust. J. Grape Wine Res.

460

2000, 6, 203–207.

461

33.

Gay, C. Invitation to Comment. Food Qual. Prefer. 1998, 9 (3), 166.

462

34.

Chatfield, C.; Collins, A. J. Introduction to Multivariate Analysis; Springer US: Boston,

463 464

MA, 1980. 35.

465

R core Team. R: A Language and Environment for Statistical Computing.; R Foundation for Statistical Computing: Vienna, Austria, 2014.

466

36.

Husson, F.; Le, S.; Cadoret, M. Sensory Data Analysis with R; 2014.

467

37.

Fox, J.; Weisberg, S. An {R} Companion to Applied Regression, Second Edition.;

468

Thousand Oaks CA: Sage, 2016.

469

38.

de Mendiburu, F. Statistical Procedures for Agricultural Research; 2016.

470

39.

Friendly, M.; Fox, J. Visualizing Generalized Canonical Discriminant and Canonical

20 ACS Paragon Plus Environment

Page 21 of 29

Journal of Agricultural and Food Chemistry

471 472

Correlation Analysis; 2016. 40.

473 474

Mevik, B.-H.; Wehrens, R.; Liland, K. H. Partial Least Squares and Principal Component Regression; 2015.

41.

McRae, J. M.; Falconer, R. J.; Kennedy, J. A. Thermodynamics of Grape and Wine

475

Tannin Interaction with Polyproline: Implications for Red Wine Astringency. J. Agric.

476

Food Chem. 2010, 58 (23), 12510–12518.

477

42.

Villamor, R. R.; Evans, M. A.; Ross, C. F. Effects of Ethanol, Tannin, and Fructose

478

Concentrations on Sensory Properties of Model Red Wines. Am. J. Enol. Vitic. 2013,

479

64 (3), 342–348.

480

43.

Vidal, S.; Francis, L.; Noble, A.; Kwiatkowski, M.; Cheynier, V.; Waters, E. Taste and

481

Mouth-Feel Properties of Different Types of Tannin-like Polyphenolic Compounds and

482

Anthocyanins in Wine. Anal. Chim. Acta 2004, 513 (1), 57–65.

483

44.

Fontoin, H.; Saucier, C.; Teissedre, P.-L.; Glories, Y. Effect of pH, Ethanol and Acidity

484

on Astringency and Bitterness of Grape Seed Tannin Oligomers in Model Wine

485

Solution. Food Qual. Prefer. 2008, 19 (3), 286–291.

486

45.

Obreque-Slier, E.; Peña-Neira, Á.; López-Solís, R. Interactions of Enological Tannins

487

with the Protein Fraction of Saliva and Astringency Perception Are Affected by pH.

488

LWT - Food Sci. Technol. 2012, 45 (1), 88–93.

489

46.

Symoneaux, R.; Chollet, S.; Bauduin, R.; Le Quéré, J. M.; Baron, A. Impact of Apple

490

Procyanidins on Sensory Perception in Model Cider (Part 2): Degree of Polymerization

491

and Interactions with the Matrix Components. LWT - Food Sci. Technol. 2014, 57 (1),

492

28–34.

493

47.

Green, B. G.; Lim, J.; Osterhoff, F.; Blacher, K.; Nachtigal, D. Taste Mixture

494

Interactions: Suppression, Additivity, and the Predominance of Sweetness. Physiol.

495

Behav. 2010, 101 (5), 731–737.

21 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

496

48.

497 498

Kroeze, J. H. A.; Bartoshuk, L. M. Bitterness Suppression as Revealed by Split-Tongue Taste Stimulation in Humans. Physiol. Behav. 1985, 35 (5), 779–783.

49.

499 500

Page 22 of 29

McBurney, D. H.; Bartoshuk, L. M. Interactions between Stimuli with Different Taste Qualities. Physiol. Behav. 1973, 10 (6), 1101–1106.

50.

Ares, G.; Barreiro, C.; Deliza, R.; Gámbaro, A. Alternatives to Reduce the Bitterness,

501

Astringency and Characteristic Flavour of Antioxidant Extracts. Food Res. Int. 2009,

502

42 (7), 871–878.

503

51.

Jung, D.-M.; de Ropp, J. S.; Ebeler, S. E. Study of Interactions between Food Phenolics

504

and Aromatic Flavors Using One- and Two-Dimensional 1H NMR Spectroscopy. J.

505

Agric. Food Chem. 2000, 48 (2), 407–412.

506

52.

Goldner, M. C.; di Leo Lira, P.; van Baren, C.; Bandoni, A. Influence of Polyphenol

507

Levels on the Perception of Aroma in Vitis Vinifera Cv. Malbec Wine. South Afr. J.

508

Enol. Vitic. 2011, 32 (1), 21–27.

509

53.

510 511

Symoneaux, R.; Guichard, H.; Le Quere, JM.; Baron, A.; Chollet, S. Could cider aroma modify cider mouthfeel properties? Food Qual. Prefer. 2015, 45, 11-17.

54.

Sáenz-Navajas, M.-P.; Campo, E.; Avizcuri, J. M.; Valentin, D.; Fernández-Zurbano,

512

P.; Ferreira, V. Contribution of Non-Volatile and Aroma Fractions to in-Mouth Sensory

513

Properties of Red Wines: Wine Reconstitution Strategies and Sensory Sorting Task.

514

Anal. Chim. Acta 2012, 732, 64–72.

22 ACS Paragon Plus Environment

Page 23 of 29

Journal of Agricultural and Food Chemistry

Figure captions.

Figure 1. PCA loadings plot for chemical variables on 34 Cabernet Sauvignon wines. Figure 2. CVA biplots for sensory analysis on 13 Cabernet Sauvignon wines. Wines, coded CS1-CS13, are shown in the left pane of the figure. In the right pane, only significant attributes are displayed (alpha = 0.05). “ar” indicates the attribute is an aroma attribute, “fl” indicates that the attribute is a flavor attribute, and “mf” indicates that the attribute is a mouthfeel attribute. CV1 versus CV2 displayed in a) and CV1 versus CV3 displayed in b). Figure 3. PLSR correlation plots for the first and second model components on 13 Cabernet Sauvignon wines. Predictors (chemical variables) are shown in black italic font, while the predicted variables (sensory attributes) are shown in red font.

23 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 24 of 29

Figure 1.

Figure 2. a)

b) 24 ACS Paragon Plus Environment

Page 25 of 29

Journal of Agricultural and Food Chemistry

Figure 3.

25 ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Page 26 of 29

Table captions. Table 1. Summary statistics for the 34 Cabernet Sauvignon wines used in this study. Full chemical data available in Supporting Information Table 1. TA is titratable acidity. Table 2. Attributes and attribute definitions and references generated during descriptive analysis on 13 wines.

Table 1.

Tannin Concentration (HPLC ; g/L)

Tannin Concentration (Protein precipitation ; g/L)

Tannin activity (kJ/mol)

Ethanol (v/v %)

TA (g/L)

pH

Fructose (g/L)

Glucose (g/L)

Minimum - Maximum

2.75 –6.16

0.35 – 1.34

- 1.43 – -4.82

13.04 – 17.05

4.48 – 6.22

3.61 – 4.02

1.00 – 3.85

-0.60 – 1.75

Mean ± standard error

4.18 ± 0. 76

0.66 ± 0.20

-2.74 ± 0.69

14.44 ± 0.81

5.26 ± 0.40

3.76 ± 0.11

2.16 ± 0.70

0.24 ± 0.62

26

ACS Paragon Plus Environment

Page 27 of 29

Journal of Agricultural and Food Chemistry

Table 2. Group

Attribute Alcohol Cedar Cherry

Chocolate Citrus Cooked, dried fruit

Fresh fruit/berry Aroma/flavor Green pepper

Outdoors

Spiced Stemmy

Tamarind Vanilla oak Bitter Savory Taste

Sour Sweet Tingling/pricking

Mouthfeel

Reference/Definition 1mL vodka (Ketel One) 2mL Cedar liquid (cedar sheets extracted in alcohol) 2g frozen dark sweet cherries (Woodstock), 2g canned Bing cherries (Bada Bing), 2g canned cherry pie filling (Duncan Hines), 2g powdered cherry candy (Wonka) 4g 100% unsweetened cocoa powder (Ghirardelli Chocolate Company) 0.05g orange rind (Nugget Markets, Davis, CA), 1.75g grapefruit rind (Nugget Markets, Davis, CA), 1g blackberry preserves (Safeway Inc.), 1g currant jelly (The J. M. Smucker Co.), 1mL prune juice (Sunsweet), 1g dried cherries (Safeway Inc.), 1g freeze-dried raspberries (Just Raspberries) 1g blueberry (Nugget Markets, Davis, CA), 1g blackberry (Nugget Markets, Davis, CA), 1g strawberry (Nugget Markets, Davis, CA), 1g raspberry (Nugget Markets, Davis, CA) 0.5g serrano pepper (Nugget Markets, Davis, CA), 0.5g jalapeno pepper (Nugget Markets, Davis, CA), 1g green bell pepper Nugget Markets, Davis, CA) 0.25g leather shoe lace (Kiwi), 1g grated Burr oak branch, 0.1g tobacco (Malboro) 0.5g dirt from 771 Pole Line Rd, 0.5g dirt and leaf matter from Mace Ranch Park 0.5g pumpkin pie spice (McCormick & Co., Inc.), 0.1g ground ginger (McCormick & Co., Inc.), 0.25g freshly ground pepper (McCormick & Co., Inc.), 2g table grape stems(Nugget Markets, Davis, CA) 2g tamarind (Melissa’s/World Variety Produce), 0.1g tobacco (Malboro) 5 High Vanilla oak stave (EvOak, Oak Solutions, Napa, CA) 1.5g/L caffeine (Sigma-Aldrich) in filtered water 1.6g/L MSG (monosodium glutamate, Accent Flavor Enhancer) in filtered water 1.25g/L L-(+)-tartaric acid, FCC, FG (SigmaAldrich) in filtered water 20g sucrose (C&H) in filtered water 15mL club soda (Canada Dry) Definition: Tingling is low on the scale while pricking is a higher intensity. Light, diffuse pins 27

ACS Paragon Plus Environment

Journal of Agricultural and Food Chemistry

Warm/hot

Grippy

Drying

Viscosity

Page 28 of 29

and needles sensation on the tongue (tingling). Deeper, more localized needle prick sensation on the tongue (pricking). 20% alcohol (Ketel One). Definition: Warm is lower intensity than hot. This scale represents the sensation of alcohol in the mouth 5g/L VF tannin (Biotan, tanin proanthocyanidique, Laffort) in filtered water. Definition: Lack of slip of tongue with mouth surfaces. Requires movement to be felt. 5g/L grape tannins (Biotan, tanin proanthocyanidique, Laffort) in filtered water. Definition: Feeling of water leaving the mouth. Does not require movement to be felt and is perceived on all surfaces of the mouth. 15 mL nonfat milk (Lucerne) Definition: The thinness/thickness of a solution when moved in the mouth.

28 ACS Paragon Plus Environment

Page 29 of 29

Journal of Agricultural and Food Chemistry

Table of Content Graphic.

For Table of Contents Only. Understanding the relationship between red wine matrix, tannin activity and sensory properties. Aude A. Watrelot, Nadia K. Byrnes, Hildegarde Heymann, James A. Kennedy

29 ACS Paragon Plus Environment