Download as pdf or txt
Download as pdf or txt
You are on page 1of 150

Compositional factors affecting the mouth-feel of white wine

AUTHOR(S)

Richard Gawel

PUBLICATION DATE

27-11-2017

HANDLE

10536/DRO/DU:30105477

Downloaded from Deakin University’s Figshare repository

Deakin University CRICOS Provider Code: 00113B


Compositional Factors Affecting the Mouth-feel of White Wine

By

Richard Gawel

B.Sc. (Adel), Dip. Ed. (Adel), Grad. Dip. Wine (Roseworthy)

Submitted in fulfilment of the requirements for the degree of

Doctor of Philosophy

Deakin University

November 2017
Signature Redacted by Library
Signature Redacted by Library
Publications

This thesis is based on the following collection of published manuscripts.

Chapter 1. Gawel, R., Smith, P.A., Cicerale, S., and Keast, R.J. The
mouthfeel of white wine. Critical Reviews in Food Science
and Nutrition doi: 10.1080/10408398.2017.1346584

Chapter 2. Gawel, R., Van Sluyter, S., and Waters, E. J. (2007). The effect of
ethanol and glycerol on the body and other sensory characteristics
of Riesling wines. Australian Journal of Grape and Wine
Research 13: 38-45.

Chapter 3. Gawel, R., and Waters, E. J. (2008). The effect of glycerol on the
perceived viscosity of dry white table wine. Journal of Wine
Research 19: 109-114.

Chapter 4. Gawel, R., Van Sluyter, S.C., Smith, P.A., and Waters, E.J.
(2013). Effect of pH and alcohol on perception of phenolic
character in white wine. American Journal of Enology and
Viticulture 64: 425-429.

Chapter 5. Gawel, R., Schulkin, A., Smith, P.A., and Waters, E.J. (2014). The
effect of mixtures of caftaric acid and grape reaction product on
the in-mouth sensory character of model wine. Australian Journal
of Grape and Wine Research 20: 25-30.

Chapter 6. Gawel, R., Day, M., Van Sluyter, S.C., Holt, H., Waters, E.J., and
Smith, P.A. (2014). White wine taste and mouth-feel as affected
by juice extraction and processing. Journal of Agricultural and
Food Chemistry 62: 10008-10014.

Chapter 7. Gawel, R., Smith, P.A., and Waters, E.J. (2016). The influence of
polysaccharides on the taste and mouth-feel of white wine.
Australian Journal of Grape and Wine Research 22: 350-357.

iii
The journals in which the manuscripts were published are ranked below in the order
of impact factor (5 year average). Source: Journal citation report. Thomson Reuters
ISI (2016).

Journal Title Impact Factor

Critical Reviews in Food Science and Nutrition 6.705

Journal of Agricultural and Food Chemistry 3.504

Australian Journal of Grape and Wine Research 2.777

American Journal of Enology and Viticulture 2.184

Journal of Wine Research 0.750

iv
Panel of supervisors

Professor Russell Keast


School of Exercise and Nutrition Science
Deakin University

Dr Sarah Cicerale
School of Exercise and Nutrition Science
Deakin University

Dr Paul A. Smith
The Australian Wine Research Institute

v
Acknowledgements

To my late Polish Mum and Dad. Your sacrifices will never be forgotten.

And to my very much alive wife Simone who for 36 years has had an un-swerving but
equally nuanced confidence that I may finish something substantial before I die.

I also thank my supervisory team Professor Russell Keast, Dr Sarah Cicerale, and Dr
Paul Smith for their sage advice and Dr Liz Waters for her Zen like encouragement
and support.

vi
Thesis summary

White wine mouth-feel encompasses tactile, chemesthetic and taste attributes


including viscosity, astringency, oiliness, hotness and bitterness. These attributes are
potentially influenced by the major components of the wine matrix - pH, glycerol and
ethanol, and by other minor components that can be significantly influenced by
winemaking processes, notably phenolics and polysaccharides.

In an initial study, the widely accepted but ill-defined mouth-feel concept of white
wine ‘body’ was investigated. Specifically, the effect of ethanol and glycerol
concentrations typical of white wine were determined as these two compounds are
abundant in white wine and had the potential to contribute to perceived viscosity.
Higher ethanol concentrations generally resulted in higher perceived body in white
wine, but glycerols’ contribution to white wine body was unresolved. The
interpretation of the concept of ‘body’ by individual assessors was found to be largely
idiosyncratic, but wines perceived as being higher in ‘body’ were more likely to be
those that were more flavoursome and perceptively viscous but also less acidic in taste.

It was speculated that the inherent sweetness of glycerol may have been a confounded
factor in the first study. Two novel approaches were used in a follow up study that
eliminated this possibility: 1) by making low and high concentrations of glycerol equi-
sweet by adding a non-viscous intense artificial sweetener, and 2) by physiologically
inhibiting sweetness perception by prior oral exposure to the anti-saccharine agent
gymnemic acid extracted from the leaves of Gymnema sylvestre. After masking its
sweetness in these ways, glycerol was found not to contribute to the perceived
viscosity of white wine, a result which was in agreeance with previous studies by
others using less direct methods.

White winemakers have recently shown a greater interest in incorporating phenolics


into their white wines to produce more ‘textural’ wines. However, winemaker
perception of the influence of phenolics on mouth-feel is anecdotal. The total phenolic
pool from white wines were extracted and reconstituted into model and real white
wines adjusted to different pH and ethanol concentrations. Phenolics were found to
contribute to all the fundamental mouth-feel attributes- astringency, perceived

vii
viscosity, bitterness and hotness; results which have important and far ranging
implications for winemaking practice. The tastes and textures elicited by the presence
of phenolics were more pronounced at low alcohol and moderate pH levels suggesting
that phenolics have a greater effect on the mouth-feel of lighter bodied wines.

Following this discovery, the two most abundant phenolics in white wine, the tartaric
ester of caffeic acid (caftaric acid) and its glutathionyl derivative 2-S-glutathionyl
caffeoyl tartaric acid (GRP) were isolated from white wine using novel preparative
scale methods, and mixtures of these were assessed for mouth-feel in a white wine like
media. GRP suppressed astringency and added to oily mouth-feel, while caftaric acid
suppressed the burning sensation from GRP. Caftaric acid and GRP suppressed
textural characters elicited by each other, and the drying sensation produced by acidity.
The ability of caftaric acid and GRP to reduce burning and drying sensations from
alcohol and organic acids without adding to bitterness suggests that phenolics have the
potential to contribute positively to white wine mouth-feel which is counter to the
current paradigm.

The previous studies showed that total phenolic concentration influences all major
aspects of white wine mouth-feel, and that combination of phenolics can have complex
interactive effects on white wine mouth-feel elicited by phenolics and the wine matrix
itself. The methods used by winemakers to extract juice from white wine-grapes are
known to significantly affect the total concentration of phenolics and the phenolic
profile of white wine. So a study was conducted whereby the mouth-feel and taste
attributes of white wines using an extensive set of conventional juice extraction
processes were modelled on the concentrations of wine non-volatiles including
phenolics, polysaccharides, glycerol, pH, organic acids, sugars and ethanol.
Astringency was influenced more by pH than by total phenolic concentration. Higher
pH and phenolics were also associated with perceived viscosity, while both total
polysaccharides and glycerol did not influence the perceived viscosity of the wines.
While bitterness was positively associated with most phenolic classes, and particularly
with flavanol and flavanone concentration, only small increases in wine bitterness
were observed despite large variations in total wine phenolic concentration.

viii
The effect of polysaccharides on the mouth-feel of white wine, and their interaction
with pH, ethanol, and phenolics was determined. White wine polysaccharides defined
by their molecular weight and monosaccharide composition were added to white wine
of various pH, ethanol and phenolic concentrations and their taste and mouth-feel
assessed. The most significant finding of the work was that 13‒93 kDa MW
arabinogalactan proteins and/or small molecular weight mannoproteins at wine like
concentrations can reduce the perceived palate hotness elicited by alcohol, and
increase its perceived viscosity when at higher pH, which suggests that
polysaccharides may suppress alcohol hotness and therefore contribute to the quality
of higher alcohol white wines.

ix
Table of contents

Access to thesis i

Declaration ii

Publications iii

Panel of supervisors v

Acknowledgements vi

Thesis summary vii

Chapter 1. 1
Literature review.

Chapter 2. 47
The effect of ethanol and glycerol on the body and other sensory
characteristics of Riesling wines.

Chapter 3. 57
The effect of glycerol on the perceived viscosity of dry white table wine.

Chapter 4. 65
Effect of pH and alcohol on perception of phenolic character in white wine.

Chapter 5. 71
The effect of mixtures of caftaric acid and grape reaction product on the in-
mouth sensory character of model wine.

Chapter 6. 79
White wine taste and mouth-feel as affected by juice extraction and
processing.

Chapter 7. 87
The influence of polysaccharides on the taste and mouth-feel of white wine.

Chapter 8. 97
Concluding remarks and future perspectives.

Appendices.

I Gawel et al. (2017) The mouthfeel of white wine.


II Authorship declarations.
Chapter 1.

Literature review1

1.1 Introduction
1.2 Physiology of the mouth related to mouth-feel
1.3 Physiological basis for mouth-feel and bitterness
1.3.1 Astringency
1.3.2 Bitterness
1.3.3 Hotness and perceived viscosity
1.4 Compositional factors affecting white wine mouth-feel
1.4.1 Phenolic compounds
1.4.1.1 Non-flavonoids
1.4.1.2 Flavonoids
1.4.1.3 Yeast derived phenolic compounds
1.4.2 Polysaccharides
1.5 Influence of phenolics on white wine mouth-feel
1.5.1 General discussion
1.5.2 Total phenolics
1.5.3 Non-flavonoids
1.5.4 Flavonoids
1.5.5 Yeast derived phenolic compounds
1.5.6 Other considerations
1.6 Influence of polysaccharides on white wine mouth-feel
1.7 Influence of other wine components on white wine mouth-feel
1.7.1 Glycerol
1.7.2 Ethanol
1.7.3 Organic acids/pH
1.7.4 Dissolved carbon dioxide
1.7.5 Amino acids and peptides
1.8 Summary
1
This chapter appended with an abstracted version of Chapter 8 has been accepted for
publication by Critical Reviews in Food Science and Nutrition as Gawel, R., Smith, P.A.,
Cicerale, S. and Keast, R.J. As ‘White Wine Mouthfeel’.

1
1.1 Introduction

Many of the recognised great white wines of the world display a textural profile
or ‘mouth-feel’ that typifies the grape cultivars from which they were made, and
the methods used to vinify them. As examples, highly respected examples of wines
made from the cultivars Viognier and Pinot Gris are as much defined by their oily
mouth-feel as by varietal flavours, and winemakers worldwide invest in high-cost
processes to create the creamy texture that defines barrel fermented
Chardonnay wines (Robinson, 1999).

White wine elicits tactile sensations of viscosity and astringency/dryness, and the
chemosensory sensations of warmth, pricking and spritz (Jackson, 2014; Oberholster,
et al., 2009; Pickering and DeMiglio, 2008). Other mouth-feel attributes such as
metallic (Jones et al., 2008) and pungency (Gawel et al., 2013) that likely incorporate
aspects of bitter taste have also been reported. Mouth-feel has been defined as
‘the group of sensations characterised by a tactile response in the
mouth’ (Pickering and DeMiglio, 2008). However, considering recent findings
that chemically induced irritation and thermal sensations also contribute to white
wine mouth-feel, we propose a broader definition that wine mouth-feel comprises
‘the tactile, irritant and thermal sensations resulting from the activation of
chemosensory and somatosensory receptors within the oral cavity by wine chemical
stimuli’.

The compositional factors mostly cited as those contributing to the mouth-feel of white
wine are low molecular weight phenolic compounds, ethanol, glycerol, organic acids,
polysaccharides and dissolved carbon dioxide. The review will discuss 1) how these
compounds individually and through interaction affect the mouth-feel of white wine
and 2) the physiological basis for mouth-feel perception in white wine. The review is
scoped by studies specifically related to non-volatile compounds found in dry white
wine that have been shown to, or could potentially influence white wine mouth-feel.
Specifically, only reports of the influence of low molecular weight phenolic
compounds, and polysaccharides relevant to white wine are reviewed, and the effects
of compounds at concentrations typical of those found in dry white wine are

2
emphasised. While it is acknowledged that bitterness is a taste rather than a texture, it
has been included for discussion on the presumption that bitterness is likely to
modulate the perception of mouth-feel characters in white wine.

1.2 Physiology of the Mouth Related to Mouth-feel

The oral mucosa is the tissue membrane that lines the oral cavity. Among its varied
functions, it provides physical and immunological protection, as well as enabling the
sensations of touch, temperature and taste. The oral mucosa is structurally
heterogeneous as it must necessarily accommodate many physical functions. The
mucosa of the inside lips, cheek, tongue, and soft palate need to be elastic and
moveable as otherwise phonation would be impossible. Conversely the gum and hard
palate mucosa are significantly more rigid as they are routinely subjected to
considerable mechanical forces encountered when chewing food. The gum and hard
palate mucosa are more rigid because they are relatively thinner and are embedded
with a dense layer of keratin. In contrast, the thicker and more elastic mucosa of the
inner lips, cheek and soft palate are keratinised to a significantly lesser degree (Squier
and Kremer, 2001).

Trigeminal (5th cranial nerve) free nerve endings that extend into the middle and upper
layers of the oral epithelium are found throughout most of the oral cavity. Transient
receptor potential (TRP) channels that respond to combinations of heat, cold and other
chemical stimuli such as ethanol are located on these nerve endings (Clapham, 2003).
As the receptive elements for these chemical stimuli are located on the intracellular
domain of the TRP channel, the substance must firstly pass through a lipid bilayer
before the receptor can be activated (Furlan et al., 2014). Other receptors found on fast
conducting afferent nerve fibres that innervate the basal cells of the epithelium
function as mechanoreceptors (Watanabe, 2004). These receptors respond exclusively
to tactile stimuli such as providing information as to the velocity and position of the
food bolus necessary for swallowing, but logically may also signal stretching and
‘sticking’ sensations between oral surfaces experienced during wine consumption.

The oral mucosa is covered by a thin layer of adsorbed salivary proteins called the
acquired pellicle and on top of that, a significantly thicker film of bulk saliva.

3
Together, they play an important role in lubricating the soft tissues in the oral cavity,
but when disrupted by tasting wine may also influence the perception of some aspects
of mouth-feel. Details of the sources and composition of the salivary proteins that are
likely to interact with wine components and therefore influence mouth-feel have been
reviewed elsewhere (Gawel, 1998). Model studies using both hydrophilic and
hydrophobic surfaces (modelling the keratinised and non-keratinised mouth surfaces
respectively) suggest that salivary proteins rapidly form a thin boundary layer which
is capable of lubricating surfaces (Maheshwari and Dhathathreyan, 2006; Vitkov et al.,
2004). Two mechanisms for pellicle formation have been proposed: 1) the smaller
salivary proteins excreted from the parotid gland, Proline Rich Proteins (PRP),
statherin and histadin crosslink to form a pellicle which is then scaffolded by adsorbed
mucins excreted from the submandibular and sublingual glands (Proctor et al., 2005;
Yao et al., 2003; Berg et al., 2003), and 2) the smaller parotid proteins cross-link and
stabilise a previously established mucin layer hydrophobically bound to the epithelia
(Svendsen et al., 2006; Iontcheva et al., 2000). In either case, a thin lubricative
proteinaceous layer is formed as under both proposed scenarios, the oligosaccharide
side chains of the attached mucins can form outward facing ‘molecular brushes’ that
can repel opposing similar surfaces by osmotic pressure or steric effects (Cardenas et
al., 2007; Schwender et al., 2005). However, the salivary film covering the oral
epithelium at between 18 and 50m (Lee et al., 2002) is sufficiently thick to suggest
that mouth lubrication by saliva consists of a mixed form of hydrodynamic lubrication
resulting primarily from unattached mucins operating between opposing mouth
surfaces (Szabo et al., 2000; Gong and Osada, 1998) and thin film lubrication by the
pellicle at the oral surface.

1.3 Physiological Basis for Mouth-feel and Bitterness

1.3.1 Astringency

The dominant paradigm is that astringency is the perception of increased friction


between oral surfaces resulting from reduced salivary lubrication following an
interaction between salivary proteins and polyphenols (Gawel, 1998). The interaction
mostly involves both hydrogen bonding between amino acid carbonyl groups and the

4
hydroxyl groups on the polyphenol, and hydrophobic stacking of the benzoic ring with
the apolar face of amino acid residues on the protein (Baxter et al., 1997; Haslam and
Lilley, 1988; Luck et al., 1994; Spencer et al., 1988). PRP’s have mostly been the focus
of the studies as they are 1) the dominant class of proteins secreted by the parotid gland
and submandibular glands which provide the greatest volume of saliva (Walz et al.,
2006; Kauffman et al., 1991), and 2) have a strong binding affinity with wine
polyphenols due to their open and flexible structure that promotes hydrogen bonding
(Bacon and Rhodes, 2000; Luck et al., 1994).

However, a causal link between interactions of wine polyphenols and PRP’s found in
the bulk saliva leading to changes to its rheology and therefore astringency perception
is yet to be established. Fundamentally, the oral lubrication paradigm of astringency is
contingent upon mechanoreceptors of somatosensory nerves being activated during
oral exposure to astringents, but direct evidence of such activation is yet to be shown.
However, deactivation of the branch of the trigeminal nerve V that innervates oral
mechanoreceptors and chemoreceptors has been shown to result in a loss of
astringency perception (Schobel et al., 2014) suggesting that mechanoreception could
play a role in the perception of astringency.

Astringency perception may also involve chemothesis, whereby oral free nerve
endings with transient receptor potential channels V1 (TRPV1) are activated by
chemical or physical stimuli responsible for perceived irritation, hotness and coolness.
Kurogi et al. (2015) found that dimeric flavanols in tea activated TRPV1 channels in
sensory neurons. Consistent with the concept that astringency perception has a
chemosensory component, monomeric and polymeric phenolic compounds also
activate trigeminal ganglion neurons within a time frame consistent with the onset of
astringency perception (Schobel et al., 2014). Related to this is the observation that
ethanol, which elicits a drying mouth-feel (Jones et al., 2008) also perturbs the lipid
layer of oral epithelial cells in a similar fashion to polyphenols (Furlan et al., 2015).

Astringency can be perceived even when oral surfaces have been stripped of bulk
saliva (Nayak and Carpenter, 2008) which suggests that astringency perception may
also involve interactions between polyphenols and the proteins that comprise the
acquired pellicle, or other connective epithelial or membrane bound proteins (Coles et

5
al., 2010; Malone et al., 2003). Wine tannins are capable of binding directly to oral
epithelial cells (Payne et al., 2009) with an efficacy related to their degree of
polymerisation (Soares et al., 2016). Similarly, the perceived astringency of whey
proteins and the polysaccharide chitosan also correlate with their ability to bind with
oral epithelial cells (Ye et al., 2012; Malone et al., 2003). The exact mechanism of cell
binding, or of which receptor channels would relay information about the presence of
astringent compounds under these circumstances is unknown.

Anecdotally, astringency can be clearly perceived in a stationary mouth whereby there


is no movement between opposing oral surfaces. While this does not rule out the
influence of mechanoreception, as constriction of oral surfaces may be involved
(Verhagen and Engelen, 2006), it does strengthen the argument that astringency is
most likely the result of multimodal actions involving interactions with the bulk saliva,
acquired pellicle and oral epithelial cells leading to a general somatosensory response
most likely involving both chemoreception and mechanoreception.

1.3.2 Bitterness

Bitter taste is elicited by taste receptor cells located in taste buds that are embedded in
the epithelium of papillae on the tongue and soft palate. These taste receptor cells
contain sub-sets of the 25-member family of G protein-coupled receptors known as
T2Rs (Chandrashekar et al., 2000, Adler et al., 2000). The sub-structure of these
receptors has not been fully defined, but if they are heterodimeric like those of
sweetness and umami receptors, then there could be up to 325 functional bitterness
receptors in humans (DuBois, 2011).

Bitterness is unique amongst the tastes in that bitter tasting compounds display a high
level of structural diversity. This can be explained at a receptor level by 1) different
expression patterns of the 25 T2Rs across receptor cells (Chandrashekar et al., 2000)
which like the olfactory system may provide a mechanism for perceptual
discrimination between bitter compounds and 2) most of the T2Rs are broadly tuned
i.e. individual T2R’s can be activated by multiple structurally different bitterants, and
alternatively, the same bitterant can activate multiple receptor types (Meyerhof et al.,
2010). Broad tuning of T2R bitter receptors to white wine phenolic compounds has
recently been demonstrated (Soares et al., 2013). Furthermore, it is known that post-

6
receptor (cell level) transduction mechanisms are also involved in eliciting bitterness,
which in the case of white wine, may also include the direct activation of secondary
messenger systems following the passing of bitter hydrophobic compounds through
receptor cell membranes (Furlan et al., 2015).

1.3.3 Hotness and Perceived Viscosity

The perceived oral hotness, bitterness and dryness of white wine are influenced by its
ethanol content (Jones et al., 2008; Gawel et al., 2007). The physiological basis for
these effects is evidenced by the existence of the thermally responsive TRPV1
noiceptors embedded within the oral mucosa, and the bitter responsive T2R38
receptors located in the circumvallate papillae at back of the tongue which robustly
respond to the presence of ethanol (Allen et al., 2014; Trevisani et al., 2002).

The perceived viscosity of Newtonian fluids with low physical viscosity, i.e. white
wine has been defined as “an appraisal of the ease with which the liquid flows between
the upper surfaces of the tongue and the palate” (Van Vliet et al., 2009). While the
physical viscosities encountered in dry white wine are narrow (estimated by Kosmerl
et al., 2000 at 0.15 mPa/s), tasters can perceive changes in the oral viscosity of
commercial white wines at around 1/5th (0.027 mPa/s) of that physical range, and
viscous mouth-feel and physical viscosity were significantly correlated (Runnebaum
et al., 2011) suggesting that perceived viscosity is related to physical viscosity in white
wine. However, it is notable that higher pH (Gawel et al., 2013, 2014a), or higher
lactate (Runnebaum et al., 2011; Skogerson et al., 2009) which is a proxy for higher
pH in white wine has been shown to be strongly associated with higher perceived
viscosity. As pH is unlikely to affect physical viscosity, these results suggest that
physical viscosity of the wine is likely not solely responsible for the perception of oral
viscosity in white wine.

1.4 Compositional Factors Affecting White Wine Mouth-feel

1.4.1 Phenolic compounds

The phenolic compounds in white wine comprise a broad family of compounds that in
their basic form possess one or more benzenoid rings substituted by at least one

7
hydroxyl group. White wine phenolics can be categorised as either 1) non-flavonoids
or 2) flavonoids according to their benzoic ring structure. The structures, and reported
concentration ranges and thresholds of the phenolic compounds in white wine are
given in Figure 1 and Tables 1 and 2 respectively.

The major non-flavonoids in white wine are the hydroxycinnamic and hydroxybenzoic
acids and their derivatives. Hydroxybenzoic acids comprise of a single benzenoid ring
and hydroxyl group but are characterised by a further substitution with a carboxylic
acid group. The hydroxycinnamic acids are characterised by an ethylene group
between the benzene ring and the carboxylic acid group.

The flavonoids are more complex comprising a C15 skeleton with an aromatic (A) and
benzodihydropyran (C) ring bearing another aromatic ring (B) in position 2. The
flavonoid sub-groups are classified by the oxidation state of the C ring and individual
compounds within each group are differentiated by the number and location of either
hydroxyl or methoxyl groups, and glycosylation on the B ring. Flavonols and flavan-
3-ols are most significant classes of flavonoids in white wine in terms of concentration,
but others such as flavanonols and flavanones have consistently been found to be
present.

1.4.1.1 Non-flavonoids

The non-flavonoids in wine comprise the hydroxycinnamic acids and hydroxybenzoic


acids and their derivatives. The hydroxycinnamic acids in white wine occur mainly in
the form of tartaric acid esters, particularly those of caffeic (caftaric), p-coumaric
(coutaric) and ferulic (fertaric) acids. Free forms and ethyl and methyl esters are also
formed in lower concentrations by hydrolysis and pectinase catalysed esterification.
Caftaric acid is also the most abundant phenolic compound found in white juice (Ong
and Nagel, 1978). When the juice is subjected to oxidative must handling practices
e.g. polyphenol oxidase hyperoxidation, the caftaric acid quinone reacts with the grape
peptide glutathione under the action of polyphenol oxidase to form 2-S-glutathionyl
caftaric acid (Grape Reaction Product or GRP) (Cheynier et al., 1986; Singleton et al.,
1985). Other similar conjugates of p-coutaric acid and caffeic acid with cysteine and
glutamine are also formed in smaller amounts by the same mechanism (Cejudo-
Bastante et al., 2010; Cheynier et al., 1986).

8
The most prevalent hydroxybenzoic acids in white wine are gallic, gentisic, p-
hydroxybenzoic, syringic and salicylic acids (Monagas et al., 2005). They are mostly
found in their free form but ethyl, methyl and glucose esters have been reported in
Riesling wines (Baderschneider and Winterhalter, 2001).

1.4.1.2 Flavonoids

Flavonols are mostly confined to the epidermal skin tissues of the grape (Rodriguez
Montealegre et al., 2006). Their concentration in white grapes have been shown to
increase disproportionately to other skin phenolics following sun exposure (Friedel et
al., 2015) suggesting that they may play a protective role against grape tissue damage
by UV light. The dominant flavonols in white and wine are the 3-O-glycosides of
quercetin particularly the glycosidic and glucuronic forms (Castillo-Muñoz et al.,
2010) but myricetin, kaempferol and isorhamnetin glycosides are also present (Di
Lecce et al., 2014; Vilanova et al., 2009).

Flavanols are mostly found in the hypodermal layers of the skins and the parenchyma
of seeds and in conventional winemaking are extracted during a brief (0.5-24 hr) period
between grape crushing and draining when the skins and seeds are in contact the juice
prior to fermentation (Di Lecce et al., 2014). (+)-catechin and its distereoisomer (–)-
epicatechin are the dominant flavanols in white wine. They are dihyroxylated on the
C-3’ and C-4’ positions on the B ring. The trihydroxylated form, (-)-epigallocatechin
which is localised in skin, and those esterified with gallic acid notably (-)-
epigallocatechin gallate and (–)-epicatechin-3-O-gallate which are localised in seeds
are also present in white wine (Oszmianski and Sapis, 1989). Flavanol dimers and
trimers are also present in white wine but in lower concentrations (Ricardo da Silva et
al., 1993; Lea et al., 1979). They consist mainly of (+)-catechin and (-)-epicatechin
units linked by C4-C8 and/or C4-C6 bonds (Foo and Porter, 1980) but other dimers
containing gallocatechin and epigallocatechin units (de Pascual-Teresa et al., 2000;
Fulcrand et al., 1999) and gallic acid units have also been detected in white wine
(Ricardo da Silva et al., 1993).

Glycosylated flavanonols have been identified in the skin and stems, with the 3-O
rhamnosides of dihydroquercetin (astilbin) and dihydrokaempferol (engeletin) being
the most significant in white wine (Masa et al., 2007; Singleton and Trousdale, 1983).

9
Hydroxycinnamic acids and derivatives

Tart =

R1 R2 R3
Caffeic acid OH OH H
p-coumaric acid OH H H
Ferulic acid OH OCH3 H
Caftaric acid OH OH Tart
Coutaric acid OH H Tart
Fertaric acid OH OCH3 Tart
Caffeic acid ethyl ester OH OH C2CH3

2-S-glutathionyl caftaric acid (Grape Reaction Product, GRP)

10
Hydroxybenzoic acids and derivatives

R1 R2 R3 R4
Gallic acid OH OH OH H
Protocatechuic acid H OH OH H
Syringic acid OCH3 OH OCH3 H
Gallic acid ethyl ester OH OH OH CH2CH3

Flavonols and derivatives

R3 Glycosides
R1 R2 R3 Glucuronic Glucose Galactose Glucose-
acid O-
rhamnose
Quercetin OH H H * * *
Kaempferol H H H * *
Myricetin OH OH H *
Isorhamnetin OCH3 H H *

11
Flavanols and derivatives

Gallate =

R1 R2 R3
(+)- catechin OH H H
(-)-epicatechin H OH H
(-)-epicatechin-3-O-gallate H Gallate H
(-)-epigallocatechin H OH OH

Flavanol dimer B3

12
Flavononols

R1
Dihydroquercetin-3-O rhamnoside OH
Dihydrokaempferol 3-O-rhamnoside H

Flavanones (Naringenin)

Tyrosol

Figure 1: Structures of phenolic compounds reported in white wine

13
1.4.1.3 Yeast derived phenolic compounds

Tyrosol is formed from tyrosine by yeast during fermentation and as such its
concentration depends on yeast strain and on the initial concentration of sugars and
tyrosine in the must (Peña-Neira et al., 2000). The contribution of tyrosol to total white
wine phenolic concentration is unclear. Some have reported that tyrosol dominates the
profile (Peña-Neira et al., 2000), while others estimate that tyrosol comprises between
3% (Gawel et al. 2014a) and 10% (Myers and Singleton, 1979) of the total phenolic
content of white wine.

1.4.2 Polysaccharides

Polysaccharides in white wine arise from both grape and yeast activity during and after
fermentation. They are classified as macromolecules that range from 5 to 200 kDa,
making them as a group, the highest molecular weight compounds found in white wine
(Jones et al., 2008). The major classes of polysaccharides in white wine have been
defined by a combination of their monosaccharide composition and their probable
oenological source. 1) Mannoproteins are either released from yeast cell walls during
fermentation, and during the autolysis stage whereby non-active ‘spent’ yeast cells
(yeast lees) contact the wine during the maturation phase (Escot et al., 2001). They are
characterised by a high content of mannose relative to other monosaccharides
(Gonçalves et al., 2002), but constitute a broad molecular weight distribution which
suggests that several populations of mannnoproteins exist depending on their source.
2) Arabinogalactan proteins (AGP’s) are characterised by a high proportion of
arabinose and galactose residues and associated with a protein core comprising a high
proportion of hydroxyproline (Fincher and Stone, 1983). AGP’s exist in the grape
vacuole intracellular spaces and are therefore easily extracted during juicing and
during the early stages of fermentation (Guadalupe and Ayestarán, 2007), and 3)
Rhamnogalacturonan polysaccharides are released from pectins embedded within the
grape cell wall and are characterised by their low MW (<20kDa), a relatively high
proportion of galacturonic acid and rhamnose residues, and by the presence of the
diagnostic sugars fucose and apiose (Pellerin et al., 1996).

14
Table 1: Mean and concentration ranges of phenolic compounds in white wine

Phenolic Compound Mean Range Reference


(mg/L) (mg/L)

Hydroxybenzoic acids and derivatives


Gallic acid 3.3 0-8.4 Fernández-Pachón et al. (2006)
3.4 de Villiers et al. (2005)
4.9 0.7-18.5 Singleton and Truesdale (1983)
1.0 Betés-Saura et al. (1996)
1.2 Kallithraka et al. (2009)
1.9 0.8-3.2 Cejudo-Bastante et al. (2011a,b)
0.8 0.4-1.2 Gawel et al. (2014a)
Syringic acid 0.04 0-0.5 Fernández-Pachón et al. (2006)
0.4 Betés-Saura et al. (1996)
1.1 0.3-2.2 Gawel et al. (2014a)
Protocatechuic acid 1.2 0-4.7 Fernández-Pachón et al. (2006)
1.3 de Villiers et al. (2005)
1.2 Betés-Saura et al. (1996)
0.3 0-0.4 Gawel et al. (2014a)
Hydroxybenzoic ethyl esters 0.9 0.3-1.8 Gawel et al. (2014a)
Hydroxycinnamic acids and derivatives
Caffeic acid 1.3 0-2.7 Fernández-Pachón et al. (2006)
2.7 de Villiers et al. (2005)
0.2 0-0.5 Nicolini et al. (1991)
1.6 Betés-Saura et al. (1996)
1.6 0.3-4.3 Korenika et al. (2014)
0.6 Kallithraka et al. (2009)
1.4 0-4.0 Cheynier et al. (1986)
2.7 2.6-2.8 Cejudo-Bastante et al. (2011a,b)
Coumaric acid 0.8 0-3.0 Fernández-Pachón et al. (2006)
1.3 de Villiers et al. (2005)
0.2 Betés-Saura et al. (1996)
0.8 0.1-2.8 Korenika et al. (2014)
0.2 Kallithraka et al. (2009)
0.3 0-0.8 Cejudo-Bastante et al. (2011a,b)
Ferulic acid 0.4 de Villiers et al. (2005)
0.1 Betés-Saura et al. (1996)
0.6 0.4-1.3 Korenika et al. (2014)
0.3 Kallithraka et al. (2009)
0.9 0.9-1.0 Cejudo-Bastante et al. (2011a,b)
(cont)

15
Caftaric acid 17.8 3.9-53.5 Fernández-Pachón et al. (2006)
12.1 de Villiers et al. (2005)
11.9 0-44.8 Singleton and Truesdale (1983)
2.3 0.5-5.4 Nicolini et al. (1991)
39.4 Herrick and Nagel (1985)
43.7 12.8-109.5 Herrick and Nagel (1985)
12.6 Betés-Saura et al. (1996)
15.1 1.5-42.3 Korenika et al. (2014)
29.6 Kallithraka et al. (2009)
109.0 24-267 Herrick and Nagel (1985)
10.6 0-39.6 Cheynier et al. (1986)
41.3 12.7-70.7 Ricardo da Silva et al. (1993)
3.7 1.77-8.36 Cejudo-Bastante et al. (2011a,b)
7.8 0-31 Gawel et al. (2014a)
Coutaric acid 2.3 de Villiers et al. (2005)
3.8 0.2-14.5 Singleton and Truesdale (1983)
0.9 0.13-1.70 Nicolini et al. (1991)
3.2 Herrick and Nagel (1985)
5.9 1.1-12.4 Herrick and Nagel (1985)
16.3 Betés-Saura et al. (1996)
4.4 Kallithraka et al. (2009)
11.5 3.2-22.6 Ricardo da Silva et al. (1993)
3.6 0.40-8.64 Cejudo-Bastante et al. (2011a,b)
3.4 0.5-10.9 Gawel et al. (2014a)
Fertaric acid 5.4 Herrick and Nagel (1985)
2.5 0.6-4.2 Herrick and Nagel (1985)
0.2 Betés-Saura et al. (1996)
0.7 0-2.95 Korenika et al. (2014)
2.0 Kallithraka et al. (2009)
4.3 2.96-5.12 Cejudo-Bastante et al. (2011a,b)
1.6 0.1-3.9 Gawel et al. (2014a)
Grape Reaction Product (GRP) 3.4 0.5-11.0 Nicolini et al. (1991)
3.2 Betés-Saura et al. (1996)
16.3 0-49.2 Cheynier et al. (1986)
9.7 3.2-20.4 Ricardo da Silva et al. (1993)
12.2 8.2-19.4 Cejudo-Bastante et al. (2011a,b)
6.2 1.0-13.4 Gawel et al. (2014a)
Hydroxycinnamic acid ethyl 0.4 0.2-1.0 Gawel et al. (2014a)
esters 1.4 0-3.5 Fernández-Pachón et al. (2006)
Coutaric acid glycoside 5.4 0-32.9 Fernández-Pachón et al. (2006)
0.7 0.1-1.2 Nicolini et al. (1991)
(cont)

16
Total hydroxycinnamic acids 34.1 Betés-Saura et al. (1996)
63.1 38.2-101 Cejudo-Bastante et al. (2011a,b)
21.7 5.2-51.6 Gawel et al. (2014a)
Flavanols
Catechin 0.9 0-11.1 Fernández-Pachón et al. (2006)
7.2 de Villiers et al. (2005)
37.0 3-123.9 Singleton and Truesdale (1983)
9.1 Oberholster et al. (2008)
2.5 Betés-Saura et al. (1996)
9.8 Carando et al. (1999a)
17.7 Kallithraka et al. (2009)
4.4 1.7-9.8 Cejudo-Bastante et al. (2011a,b)
2.7 0-9.0 Gawel et al. (2014a)
28.3 2.3-28.7 Vitrac et al. (2002)
10.1 Gürbüz et al. (2007)
Epicatechin 0.8 0-10.4 Fernández-Pachón et al. (2006)
3.7 de Villiers et al. (2005)
27.3 0-80.8 Singleton and Truesdale (1983)
19.3 Oberholster et al. (2008)
4.1 Betés-Saura et al. (1996)
5.3 Carando et al. (1999a)
25.5 Kallithraka et al. (2009)
5.0 0-8.7 Cejudo-Bastante et al. (2011a,b)
4.8 0-21.3 Gawel et al. (2014a)
33.8 0.2-71.7 Vitrac et al. (2002)
2.6 Gürbüz et al. (2007)
Epicatechin gallate 2.4 0-8.0 Gawel et al. (2014a)
Epigallocatechin gallate 2.5 0-15.6 Fernández-Pachón et al. (2006)
1.0 0-6.0 Gawel et al. (2014a)
B1 dimer 1.2 0-8.6 Fernández-Pachón et al. (2006)
0.8-1.4 Carando et al. (1999b)
6.1 0.1-24.1 Ricardo da Silva et al. (1993)
B2 dimer 1.4 Betés-Saura et al. (1996)
0.5 0-1.5 Ricardo da Silva et al. (1993)
B3 dimer 0.9-2.2 Carando et al. (1999b)
1.2 Betés-Saura et al. (1996)
0.5 0-2.1 Ricardo da Silva et al. (1993)
B4 dimer 0.3-0.5 Carando et al. (1999b)
0.13 0-0.4 Ricardo da Silva et al. (1993)
Trimers 0.6-1.7 Carando et al. (1999b)
1.8 0-6.3 Ricardo da Silva et al. (1993)
Total dimers 5.3 0.3-2.2 Carando et al. (1999b)
Total trimers 1.6 0.6-1.7 Carando et al. (1999b)
(cont)

17
Total flavanols 20.2 13.9-22.6 Cejudo-Bastante et al. (2011a,b)
12.8 2.8-31.3 Gawel et al. (2014a)
11.0 3.4-17.3 Konitz et al. (2003)
Flavonols
Quercetin 1.4 0-3.4 Korenika et al. (2014)
1.5 0.1-3.2 Cejudo-Bastante et al. (2011a,b)
0.4 0-4.7 Gawel et al. (2014)
0.7 0-1.2 Simonetti et al. (1997)
Quercetin glucoside 0.4 de Villiers et al. (2005)
0.4 0-0.8 Korenika et al. (2014)
0.8 0.4-1.2 Cejudo-Bastante et al. (2011a,b)
3.4 0-30.3 Gawel et al. (2014)
Quercetin glucuronide 0.3 Betés-Saura et al. (1996)
1.5 1.1-1.9 Cejudo-Bastante et al. (2011a,b)
3.3 0-44.4 Gawel et al. (2014)
Quercetin rutinoside 0.4 0.2-0.9 Simonetti et al. (1997)
Kaempferol 0.04 0-0.3 Korenika et al. (2014)
0.40 0-0.9 Cejudo-Bastante et al. (2011a,b)
0.1 0.1-0.1 Simonetti et al. (1997)
Kaempferol glucoside 0.3 0.2-0.4 Cejudo-Bastante et al. (2011a,b)
Kaempferol galactoside 0.1 0.1-0.14 Cejudo-Bastante et al. (2011a,b)
Isorhamnetin 0.02 0-0.4 Korenika et al. (2014)
Myricetin 0.2 0.1-0.3 Simonetti et al. (1997)
Isorhamnetin glucoside 0.04 0.04-0.04 Cejudo-Bastante et al. (2011a,b)
Total Flavonols 51.7 39.5-80.0 Cejudo-Bastante et al. (2011a,b)
9.0 0.5-56.8 Gawel et al. (2014)
Flavanonols and Flavanones
Dihydroquercetin rhamnoside 1.1 0-10.3 Singleton and Truesdale (1983)
2.4 0-21.1 Gawel et al. (2014)
2.2 0.6-4.4 Vitrac et al. (2002)
Dihydrokaempferol rhamnoside 0.5 0-3.3 Singleton and Truesdale (1983)
Dihydromyricetin-rhamnoside 3.0 1.8-6.0 Vitrac et al. (2002)
Naringin 2.6 0.1-11.2 Gawel et al. (2014)
Other Phenolic Compounds
Tyrosol 2.5 0-11.9 Fernández-Pachón et al. (2006)
11.6 Betés-Saura et al. (1996)
2.8 2.2-3.6 Gawel et al. (2014)
13.3 6.3-22.7 Konitz et al. (2003)
Total Phenolics
Total flavonoids 168 128-222 Singleton et al. (1980)
Total non-flavonoids 91 66-112 Singleton et al. (1980)
Total Phenolics 251 164-316 Konitz et al. (2003)
258 224-328 Singleton et al. (1980)

18
Total Phenolics (cont) 279 Kallithraka et al. (2009)
168-404 Sokolowsky and Fischer (2012)
201 152-281 Cejudo-Bastante et al. (2011a,b)

1.5 Influence of phenolics on white wine mouth-feel

1.5.1 General discussion

The traditional paradigm of astringency i.e. decreased oral lubrication following


exposure to polyphenols depends on the polyphenol being capable of binding with a
salivary protein at different positions and aggregating multiple salivary proteins
(Baxter et al., 1997). The question as to whether the monomeric and dimeric
polyphenols that make up the majority of the phenolic profile of white wines have
sufficient numbers of hydrogen bonding sites and/or hydrophobic surfaces to interact
with salivary proteins has only recently been addressed. Recent studies have shown
that the basic proline rich protein (PRPb) amino acid probes IB-5 (Canon et al., 2011)
and IB9-37 (Canon et al., 2013; Cala et al., 2012) are able to bind to and aggregate
epicatechin gallate and epigallocatechin gallate (Pascal et al., 2009) molecules.
Flavanol dimers with extended structures have also been shown to act as bidentate
ligands for IB7-14 (Cala et al., 2010). Other monomeric non-flavanols such as
naringenin, apigenin and quercetin rhamnoside and rutinoside also have affinities for
IB-14 and entire PRPb (Plet et al., 2015). These studies show that many monomeric
and dimeric polyphenols found in white wine can form complexes with basic proline
rich proteins, and therefore have the potential to elicit astringency under the tactile
model.

With respect to the alternative paradigms for astringency perception, Soares et al.
(2016) found that in vivo, low molecular weight polyphenol fractions comprising
monomers and dimers did not bind well to oral epithelial cells regardless of the
simultaneous presence of salivary proteins, which under the epithelial binding model
of astringency would suggest that low molecular weight polyphenols do not contribute
to astringency. However, it should be noted that the influence of a salivary pellicle,
and epithelial bound salivary proteins typical of an oral surface were not modelled in
their experiment.

19
There is conflicting evidence as to whether monomers can elicit astringency under the
chemesthetic model. TRPV1 channels were not activated in response to a variety of
flavanol monomers (Kurogi et al., 2015; Carpenter, 2013), but galloylated flavanol
monomers have been shown to stimulate trigeminal neurons suggesting that they at
least can elicit astringency (Schobel et al., 2014). In conclusion, these differences may
be due to the widely different methods used to determine oral physiological response
to the phenolic stimulus.

1.5.2 Total phenolics

White wine phenolics are mostly monomeric and comprise over 80 compounds
spanning a range of phenolic classes defined by their ring structures, and within each
class they take on different forms based on their patterns of hydroxylation,
glycosylation, esterification or conjugation with amino acids. Despite this diversity,
the summed concentration of the phenolic species in white wine is relatively low
compared to those in red wines where polymeric flavanols derived from fermentation
on skins and seeds dominate their phenolic profile. However, as the sensory response
to mixtures of non-volatiles that elicit tastes and astringency are known to be at least
partially additive (Ferrer-Gallego et al., 2014; Keast et al., 2003), it is prudent to also
consider the effects of combined total of phenolic compounds on mouth-feel in white
wines.

The first study of its type in white wine used both conventional, and unconventional
winemaking methods more akin to those used in red winemaking to create extremes
in total phenolic concentration (Singleton et al., 1975). They found that white wine
bitterness was unrelated to total phenolic concentration which would suggest that
either 1) other (non-phenolic) compounds contribute to bitterness, or 2) that the
winemaking methods influenced the relative contribution of bitter phenolic
compounds (e.g. epicatechin and naringin) over non-bitter phenolic compounds (e.g.
caftaric acid and GRP) within the total pool of phenolics present in the wines. In a
broader study, Gawel et al. (2014a) found that a six-fold difference in total wine
phenolic concentration obtained by applying conventional winemaking techniques
increased bitterness only slightly, but the increases were strongly correlated with total
phenolic concentration.

20
The perceived viscosity of white wine has been associated with its total phenolic
concentration (Gawel et al., 2013; Okuda et al. 2007). In contrast, lower phenolic
concentration white wines made from hyperoxidised free run juice were not lower in
‘body’ than the control wines (Cejudo-Bastante 2011a,b). The differences in
conclusions may be explained by the expected differences in the range of total
phenolics, and from variations in other compositional factors such as pH and residual
sugar resulting from the different winemaking methods that were applied. To
overcome the limitations of correlative studies, addition studies have also been
conducted. White wines fortified with the total phenolic pool extracted from each of
three different varietal wines were consistently perceived to be more viscous,
astringent and bitter (Gawel et al., 2013). Total phenolics were also shown to enhance
the perceived hotness and astringency of model white wine, but only of those that were
initially low in hotness (i.e. low alcohol), and low in the astringent/drying character
(i.e. high pH).

1.5.3 Non-flavonoids

Benzoic acid derivatives when presented at two orders of magnitude higher than that
found in white wine elicit complex oral sensations of sourness, bitterness, astringency
and a ‘prickling’ (Peleg and Noble, 1995). The substitution pattern was found to affect
mouth-feel, with those substituted in the ortho position (salicylic and gentisic acids)
being astringent, and those with a greater number of hydroxyl groups more ‘prickling’
(Peleg and Noble, 1995). It is unclear if these results are relevant to white wine,
however as gentisic acid has also been reported to be astringent when tasted at white
wine like concentrations (Dadic and Belleau, 1973) it is possible that it and other
benzoic acids found in white wine may affect mouth-feel.

Many hydroxycinnamic acids and their ethyl and tartaric acid esters have been reported
to be astringent and bitter at concentrations higher than those observed in white wine
(Hufnagel and Hofmann, 2008). However, as the hydroxycinnamic acid
concentrations found in white wine are typically below their detection thresholds
(Okamura and Watanabe, 1981; Maga and Lorenz, 1973), it would suggest that they
do not impact on the taste or mouth-feel of white wine. Indeed, fortifying white wines
with a realistic level of caftaric acid did not produce perceptible changes to their taste

21
or mouth-feel (Sáenz-Navajas et al., 2012; Vérette et al., 1988). However, when in
model wine, wine-realistic levels of caftaric acid have been found to impact on acidity
(Vérette et al., 1988), and astringency (Gawel et al., 2014b). Similarly, wine realistic
concentrations of grape reaction product did not affect the palate properties when
added to a real white wine (Vérette et al., 1988), but GRP was found to increase the
intensity of oily mouth-feel and a burning aftertaste unrelated to ethanol when added
to model white wine (Gawel et al., 2014b).

1.5.4 Flavonoids

A rutinoside (di-glycoside) of quercetin has been suggested by Scharbert et al. (2004)


to be a powerful astringent at concentrations well below those of white wine. The
possible mouth-feel and taste impact of flavonol monoglycosides were demonstrated
when quercetin 3-O-glucoside, albeit at higher concentrations than found in white
wine, elicited astringency and bitterness (Ferrer-Gallego et al., 2016). A 3-O-
rhamnoside of quercetin has also been reported to be bitter and astringent at white wine
like concentrations (Dadic and Belleau, 1973).

More studies have been conducted on the taste and mouth-feel properties of flavanols
than any other phenolic class due to their occurrence in many different beverages
including wine, green and black tea, and cider. However, all studies to date have used
concentrations far higher than those observed in white wine (i.e. > 500 mg/L).
Therefore, the results of the studies should be considered only indicative of possible
sensory outcomes in white wine.

Catechin and its stereoisomer (-)-epicatechin are reported to be astringent and bitter
(Peleg et al. 1999; Thorngate and Noble 1995; Robichaud and Noble, 1990) with
epicatechin being more astringent and bitter than catechin (Ferrer-Gallego et al., 2014).
However, perceptual mapping of these simple monomers against recognised bitter and
astringent compounds suggests that they are more bitter than they are astringent
(Kielhorn and Thorngate, 1999).

Dihydroxylated and trihdroxylated flavanols esterified with gallic acid e.g. epicatechin
gallate and epigallocatechin gallate have also been deemed to be bitter and astringent
in taste tests on tea (Yu et al., 2014; Narukawa et al., 2010), with bitterness being later

22
confirmed physiologically by the observation that they activated the human bitter taste
receptor hTAS2R14 (Yamazaki et al., 2013).

Schobel et al. (2014) proposed that 3 hydroxyl groups on the B ring (galloylation) is a
necessary condition for robustly stimulating an oral chemosensory response indicative
of astringency. This would indicate that the dehydroxylated flavanols catechin and
epicatechin cannot elicit astringency. However, this conclusion is contingent on the
assumption that chemesthesis is the only astringency mechanism at play. Galloylation
also appears to favour astringency perception under the salivary protein interaction
model of astringency as trihydroxylation of the B-ring results in increased interactions
with poly-L-proteins (Poncet-Legrand et al., 2006), salivary mucins (Peleg et al., 1999)
and basic PRP’s (Cala et al., 2010). Glycosylated PRP’s which are considered
important in salivary lubrication are also precipitated by epigallocatechin gallate
(Pascal et al., 2008). In contrast, Ferrer-Gallego et al. (2015) noted that galloylated
forms were less astringent than dihydroxylated forms which they attribute to the ability
of basic PRP’s to simultaneously interact with fewer galloylated flavanol molecules.
Epigallocatechin gallate has been reported as being more astringent than other
flavanols monomers (Rossetti et al., 2009) and is the main contributor to astringency
in green tea (Yu et al., 2014). EGCG has been shown to increase the friction coefficient
of whole saliva while epicatechin had no effect, suggesting that esterification with
gallic acid and trihydroxylation of the B-ring may enhance the production of
astringency (Rossetti et al., 2009).

Compared to red wines, the flavanols in white wines have a low degree of
polymerisation (dP), typically ranging from monomers to trimers. It is now well
understood that the first stage of astringency perception under the tactile model
involves complex formation between proteins and flavanols, and this complexation
increases with the flavanols’ dP (Baxter et al., 1997). White wine flavanols (dP 1-3)
have been shown to be astringent (Peleg et al., 1999) with their astringency increasing
with dP. Consistent with this, flavanol dimers have been shown to form soluble
complexes with salivary proteins (Sarni-Manchado and Cheynier, 2002), and that they
can also activate TRPV1 channels associated with chemosensory perception (Kurogi
et al., 2015).

23
Table 2: Thresholds of Phenolic Compounds Found in White Wine

Compound Tc Attribute Method* Reference


(mg/L)
Flavanols
Catechin 119 astringency staircase Scharbert et al. (2004)
Catechin 290 bitterness triangle Hufnagel and Hofmann (2008)
Catechin 20 detection staircase Dadic and Belleau (1973)
Catechin 46 detection ascending limits Delcour et al. (1984)
Epicatechin 270 astringency staircase Scharbert et al. (2004)
Epicatechin 270 bitterness triangle Hufnagel and Hofmann (2008)
Epicatechin 270 astringency staircase Hufnagel and Hofmann (2008)
Epicatechin gallate 110 astringency staircase Scharbert et al. (2004)
Epigallocatechin 169 astringency staircase Scharbert et al. (2004)
Epigallocatechin
gallate 87 astringency staircase Scharbert et al. (2004)
Dimer B1 139 astringency staircase Hufnagel and Hofmann (2008)
Dimer B1 231 bitterness triangle Hufnagel and Hofmann (2008)
Dimer B2 110 astringency staircase Hufnagel and Hofmann (2008)
Dimer B2 280 bitterness triangle Hufnagel and Hofmann (2008)
Dimer B3 116 astringency staircase Hufnagel and Hofmann (2008)
Dimer B3 17 astringency ascending limits Delcour et al. (1984)
Dimer B3 289 bitterness triangle Hufnagel and Hofmann (2008)
Trimer C1 260 astringency staircase Hufnagel and Hofmann (2008)
Trimer C1 347 bitterness triangle Hufnagel and Hofmann (2008)
Trimers+tetramers 4 detection ascending limits Delcour et al. (1984)
Hydroxybenzoic acids and derivatives
Gallic acid 45 astringency staircase Glabasnia and Hofmann (2006)
Gallic acid 50 astringency staircase Hufnagel and Hofmann (2008)
Gallic acid 40 detection ascending limits Maga and Lorenz (1973)
Gallic acid 20 detection staircase Dadic and Belleau (1973)
Protocatechuic acid 32 astringency staircase Hufnagel and Hofmann (2008)
Protocatechuic acid 30 detection ascending limits Maga and Lorenz (1973)
Protocatechuic acid 20 detection staircase Dadic and Belleau (1973)
Syringic acid 52 astringency staircase Hufnagel and Hofmann (2008)
Syringic acid 240 detection ascending limits Maga and Lorenz (1973)
Gallic acid ee 37 astringency staircase Hufnagel and Hofmann (2008)

24
(cont)
Gallic acid ee 438 bitterness triangle Hufnagel and Hofmann (2008)
Protocatechuic acid ee 9 astringency staircase Hufnagel and Hofmann (2008)
Protocatechuic acid ee 182 bitterness triangle Hufnagel and Hofmann (2008)
Hydroxycinnamic acids and derivatives
Caffeic acid 13 astringency staircase Hufnagel and Hofmann (2008)
Caffeic acid 90 detection ascending limits Maga and Lorenz (1973)
Coumaric acid 23 astringency staircase Hufnagel and Hofmann (2008)
Coumaric acid 40 detection ascending limits Maga and Lorenz (1973)
Ferulic acid 13 astringency staircase Hufnagel and Hofmann (2008)
Ferulic acid 62 detection E679-91 Work and Camire (1996)
Caffeic acid ee 69 astringency staircase Hufnagel and Hofmann (2008)
Caffeic acid ee 229 bitterness triangle Hufnagel and Hofmann (2008)
Coumaric acid ee 27 astringency staircase Hufnagel and Hofmann (2008)
Coumaric acid ee 137 bitterness triangle Hufnagel and Hofmann (2008)
Ferulic acid ee 15 astringency staircase Hufnagel and Hofmann (2008)
Ferulic acid ee 158 bitterness triangle Hufnagel and Hofmann (2008)
Caftaric acid 5 astringency staircase Hufnagel and Hofmann (2008)
Caftaric acid <50 detection yes/no Okamura and Watanabe (1981)
Coutaric acid <25 detection yes/no Okamura and Watanabe (1981)
Flavonols
Quercetin 10 detection staircase Dadic and Belleau (1973)
Kaempferol 20 detection staircase Dadic and Belleau (1973)
Myricetin 10 detection staircase Dadic and Belleau (1973)
Quercetin glucoside 1 astringency staircase Hufnagel and Hofmann (2008)
Quercetin rhamnoside 9 detection staircase Dadic and Belleau (1973)
Quercetin galactoside 0.2 astringency staircase Hufnagel and Hofmann (2008)
Flavanonols
Dihydroquercetin
rhamnoside 1.7 detection staircase Hufnagel and Hofmann (2008)
Dihydrokaempferol
rhamnoside 2.1 detection staircase Hufnagel and Hofmann (2008)
Flavanones
Naringin 17 bitter Esaki et al. (1983)
Naringin 20 bitter Guadagni et al. (1973)
Other
Tyrosol 346 bitter ascending limits Takahashi et al. (1974)
* in water except Maga and Lorenz (1973) in 5% ethanol. Tc = threshold, ee = ethyl ester

25
1.5.5 Yeast derived phenolic compounds

The ‘tannin taste’ of Riesling wines from the same juice with different levels of must
solids correlated strongly with their tyrosol concentration, while being poorly
correlated with hydroxycinnamic acid and flavan-3-ol concentration (Konitz et al.,
2003). A flavanone, naringin has recently been tentatively identified in white wine
(Gawel et al., 2014b). As flavanones are only likely to be present in white wine at very
low concentrations a possible effect on white wine bitterness cannot be ruled out some
flavanone glycosides are intensely bitter when glycosylated in specific forms
(Horowitz and Gentili, 1969).

1.5.6 Other considerations

The perceived intensity of mixtures of similar tasting compounds including those that
are bitter, are known to be additive or sub-additive (Keast et al., 2003), and in the case
of some white wine phenolics, their effect on mouth-feel may even be hyper-additive
(Ferrer-Gallego et al., 2014). Therefore, even if the concentrations of individual
phenolic species are insufficient to induce a sensation in white wine, they may do so
in combination. Caftaric acid and its derivative GRP have been shown to be mutually
antagonistic with respect to astringency and burning after-taste, and were also found
to be able to suppress the astringent/drying sensation produced by the wine matrix
(Gawel et al., 2014b), showing the potential for phenolic compounds to interact in
complex ways with each other and other major components found in wine.

1.6 The influence of polysaccharides on white wine mouth-feel

Mixtures of mannoproteins and AGP’s at wine like concentrations have been shown
to enhance the perception of viscosity in model white wine (Gawel et al., 2016; Vidal
et al., 2004a). However, the effect of these polysaccharides on viscosity was shown to
depend on pH, with the increase in perceived viscosity only occurring in higher pH
wines (Gawel et al., 2016). Possible reasons of this pH effect include charge effects on
acidic polysaccharides, or changes to the ordering of bulk water in the environment
surrounding the polysaccharide – both of which could alter its hydration state and
therefore viscosity. The possible effect of polysaccharides on perceived viscosity has

26
also been demonstrated in real white wine as the ‘thickness’ of white Koshu wines was
shown to associate with the concentration of neutral polysaccharides (Okuda et al.,
2007). However, it was shown that doubling the total polysaccharide concentration of
white wine had a small but statistically insignificant effect on perceived viscosity
(Gawel et al., 2016). The different strength of conclusions could be due to
methodology. Okuda et al. (2007) used a correlative approach using Koshu (Vitis
vinifera x V. davidii) wines, while Gawel et al. (2016) reconstituted wines with
polysaccharides extracted from V. vinifera cv Chardonnay and Riesling.

The hotness of white wine was also reduced by realistic levels of white wine
polysaccharides, mostly attributed to medium molecular weight mannoproteins and
AGP’s in the 13-93 kDa MW range (Gawel et al., 2016). The physiological mechanism
for the suppression of alcohol hotness by polysaccharides is unknown but worthy of
further investigation given that increased flavour in white wine can be obtained by
picking grapes when riper, but this can incur an expense in wine quality as doing so
may result in excessive levels of palate hotness.

Wine polysaccharides may also influence astringency perception as they have been
shown to disrupt the interaction and aggregation between salivary proteins and
polymerised flavanols (Carvalho et al., 2006). This effect could be due to either
solubilisation following formation of ternary complexes between the protein,
polyphenol and polysaccharide, or by competition between the polysaccharide and
protein for the polyphenol (Luck et al., 1994). Such possible disruption has been shown
in a red wine context with lower perceived astringency of model wine containing
polymerised grape seed flavanols following addition of Rhamnogalacturonan-II (Vidal
et al., 2004b), and correlations between mannoprotein and AGP concentrations with
the astringency of Tempranillo red wines (Quijada-Morín et al., 2014).

However, studies using white wines have found little evidence that polysaccharides
influence their astringency. Additions of whole white wine polysaccharides at wine
realistic concentrations did not influence the astringent/drying character of white wine
(Gawel et al., 2016), and the astringency of white wines made using a broad spectrum
of white winemaking methods and grape cultivars were found to be unrelated to their
total polysaccharide concentration (Gawel et al., 2014a). The differences in the

27
observed effects of polysaccharides on astringency across studies could be due to the
difference in the pH of the wines that were used which could alter the charge properties
of the polysaccharides (Verhnet et al., 1996) and salivary proteins (McArthur et al.,
1995) and therefore their interaction with wine polyphenols, or to differences in the
relative concentrations of phenolic compounds to polysaccharides which could also
determine the type of interaction i.e. whether it be competitive or associative (Soares
et al., 2009).

The effect of polysaccharides on white wine bitterness and (presumably) related


characters such as metallic and pungency is unclear. In model systems, the bitterness
of polymerised grape seed tannin was reduced by a mixture of mannoproteins and
arabinogalactan-proteins (Vidal et al., 2004b) and whole polysaccharides from white
wine reduced the metallic character of a complex model white wine which included
ethanol, flavour compounds, glycerol, and white wine proteins (Jones et al., 2008).
However, others found that white wine polysaccharides did not affect the bitterness of
a high phenolic white wine, nor the bitterness elicited by ethanol in a simple model
white wine (Gawel et al., 2016). Further addition studies utilising real white wine are
clearly required to properly address the effect of polysaccharides on white wine
bitterness.

1.7 The influence of other components on white wine mouth-feel

1.7.1 Glycerol

Glycerol is the third most abundant compound in white wine after water and ethanol.
Glycerol is a product of yeast fermentation, and ranges between 5-10 g/L depending
on yeast strain and fermentation conditions (Nieuwoudt et al., 2002). In its pure form,
glycerol is clearly viscous suggesting that it may influence white wine viscosity.
However, the estimated viscosity difference threshold of glycerol in white wine is 26
g/L which suggests that it does not influence the perception of viscosity (Noble and
Bursick, 1984). Increasing glycerol by 5g/L in white wine (Gawel et al., 2007) and
model white wine (Jones et al., 2008) resulted in small increases in perceived viscosity.
However, in both these studies there was the possibility that the perception of
perceived viscosity was confounded with the inherent sweetness of glycerol. When the

28
sweet taste of glycerol was neutralised by blocking sweet taste receptors using the anti-
saccharin agent gymnemic acid, or was equalised by the addition of a non-viscous
artificial sweetener, then glycerol was found to have no effect on perceived viscosity
(Gawel and Waters, 2008). The sweetness of glycerol may also be a factor in the
reduced bitterness via mixture suppression as observed by Jones et al. (2008) in model
white wine.

1.7.2 Ethanol

Increased hotness in white wine due to ethanol can be perceived even within the
relatively narrow range of concentrations typically encountered in white wine (Gawel
et al., 2007). Higher ethanol white wines are also perceived to be slightly higher in
body/density (Nurgel and Pickering, 2005).

The intensity of astringent sensation declines in the presence of increasing amounts of


ethanol (DeMiglio and Pickering, 2008; Lea and Arnold 1978). Increasing ethanol
concentration results in reduced precipitation (Rinaldi et al., 2012) and interaction
strength (McRae et al., 2015) between salivary proteins and polymerised tannins. The
reduced impact of polyphenols on salivary proteins in the presence of ethanol has been
postulated to be due to the ability of ethanol to interfere with interactions between
protein H-acceptor sites and polyphenol hydroxyl groups, and by disrupting
hydrophobic interactions due to its influence on water cohesion (Pascal et al., 2008).
Ethanol may contribute directly to the observed reduction in astringency as it has been
shown to reduce the astringency of model wine devoid of phenolics (Fontoin et al.,
2008) which suggests that it may be able to modulate the drying sensation elicited by
organic acids. However, it is noteworthy that others have found ethanol to increase
dryness under similar conditions (Symoneaux et al., 2015; Jones et al., 2008).
Increased ethanol concentration may also improve the ‘quality’ of the astringent
sensation as it has shown to enhanced the production of sub-qualities 'velvety', and
'mouth-coat' (DeMiglio et al., 2002), and reductions in other possibly less desirable
characters of 'puckery', 'coarse' (Vidal et al., 2004b) and ‘grippy/adhesive’ (DeMiglio
and Pickering, 2008).

29
Ethanol is inherently bitter (Mattes and DiMeglio, 2001) and is known to stimulate the
bitter taste receptor TAS2R (Allen et al., 2014). Indeed, Fischer and Noble (1994)
found that ethanol in white wine was a stronger contributor to bitterness than the bitter
tasting flavanol catechin.

1.7.3 Organic acids/pH

White wines can vary in pH from between 3.1 to 4.0, but typically are within the range
of 3.3 to 3.5. The total acidity of white wine is mostly the combination of tartaric and
malic acid from the grape, and lactic acid in the case of wines that have undergone
malolactic fermentation. The fermentation product succinic acid which has been
described as salty and acidic (Kubıć ková and Grosch, 1998) and umami like (Rotzoll
et al., 2006; Kaneko et al., 2006) is also found in reasonable concentrations in white
wine.

Organic acids including those in white wine have been shown in model studies to elicit
mouth drying sensations (Hartwig and McDaniel, 1995). These sensations have been
attributed to hydrogen ion concentration rather than the total acidity or the anion
species involved (Sowalski and Noble, 1998; Lawless et al., 1996). Consistent with
these model studies, the astringency/dryness of white wines made using a broad set of
commercial juice extraction and handling methods consistently increased with
decreasing pH, but astringency was unrelated to total phenolic concentration (Gawel
et al., 2014a).

The increases in perceived astringency/dryness with decreasing pH may be explained


by reduced salivary viscosity (Veerman et al., 1989; Nordbo et al., 1984) possibly
resulting from greater interaction after charged salivary proteins neutralise at pH’s
closer to their isoelectric point. Alternative explanations could be an increased binding
of astringent compounds to oral epithelial cells which has been observed with both
phenolic (Payne et al., 2009) and non-phenolic compounds (Ye et al., 2012), or that
H+ ions which are strong disruptors of water structure may change the hydration
environment surrounding the solvated proteins embedded within salivary pellicle,
possibly leading to a feeling of palate dryness.

30
1.7.4 Dissolved carbon dioxide

Most bottled white wines contain some dissolved carbon dioxide as part of their
bottling specification with concentrations ranging from 0.8 to 1.8 g/L (Cocola, 2016).
Dissolved CO2 can create a mouth-feel character often described as ‘spritz’, which in
the case of still white wine is considered a negative quality attribute. To our
knowledge, no studies have been conducted on the effect of dissolved CO2 at white
wine concentrations on the perception of spritz or other mouth-feel characters.
However, in a closely analogous situation to white wine, 5 g/L of dissolved CO2 (a
level typical of semi sparkling wine) increased the perceived astringency of model
apple cider containing polyphenols (Symoneaux et al., 2015). Similar enhancing
effects on the astringency of model solutions by saturated solutions of CO2 have been
observed (Hewson et al., 2009). The increased astringency may be due to lower pH
levels resulting from the presence of carbonic acid formed by the dissociation of CO2
in solution, but given that both the perception of dissolved CO2 and astringency may
be of somatosensory origin, a direct effect of dissolved CO2 on perceived astringency
is possible.

1.7.5 Amino acids and peptides

It has been known for some time that many amino acids and peptides are bitter (Solms,
1969). The common structural features of bitter amino acids are that they have an L
configuration and a hydrophobic side chain. The bitterness of peptides can similarly
be equated to the average hydrophobicity of their amino acid components (Maga,
1990). White wine contains several hydrophobic peptides that are bitter (Desportes et
al., 2001). However, it is unknown if they are found in white wine in sufficient
concentrations to produce a bitter sensation either individually or in combination.
Peptides may impact on the fullness of white wine. Using an untargeted approach
Skogerson et al. (2009) found that the concentration of amino acids in white wine (both
individually and in total) were good predictors of whether a white wine was low,
medium or high in body.

31
1.8 Summary

White wine mouth-feel which encompasses the tactile, chemosensory and taste
attributes of perceived viscosity, astringency, hotness and bitterness is increasingly
being recognised as an important component of overall white wine quality. This review
summarises the physiological basis for the perception of white wine mouth-feel and
the direct and interactive effects of white wine composition, specifically those of low
molecular weight phenolic compounds, polysaccharides, pH, ethanol, glycerol,
dissolved carbon dioxide and peptides. Ethyl alcohol concentration and pH play a
direct role in determining most aspects of mouth-feel perception, and provide an
overall framework on which the other minor wine components can interact to influence
white wine mouth-feel. Phenolic compounds broadly impact on the mouth-feel by
contributing to its viscosity, astringency, hotness and bitterness. Their breadth of
influence likely results from their structural diversity which would allow them to
activate multiple sensory mechanisms involved in mouth-feel perception. Conversely,
polysaccharides have a small modulating effect on astringency and hotness perception,
and glycerol does not affect perceived viscosity within the narrow concentration range
found in white wine. Many of the major sensory attributes that contribute to the overall
impression of mouth-feel are elicited by more than one class compound suggesting
that different physiological mechanisms may be involved in the construct of mouth-
feel percepts.

Lastly, recent findings from receptor based studies, and the lack of evidence that oral
mechanoreceptors are activated following interactions of saliva with polyphenols
suggests that astringency is not exclusively, or perhaps not at all the result of oral
mechanoreception, but is more likely to result from a process involving
chemoreception or proteinaceous binding with the oral epithelium itself.

References

Adler, E., Hoon, M. A., Mueller, K. L., Chandrashekar, J., Ryba, N. J. P., and Zuker,
C. S. (2000). A novel family of mammalian taste receptors. Cell 100: 693-702.
Allen, A. L., McGeary, J. E., and Hayes, J. E. (2014). Polymorphisms in TRPV1 and
TAS2Rs associated with sensations from sampled ethanol. Alcoholism: Clinical and
Experimental Research 38: 2550-2560.

32
Bacon, J. R., and Rhodes, M. J. (2000). Binding affinity of hydrolyzable tannins to
parotid saliva and to proline-rich proteins derived from it. Journal of Agricultural and
Food Chemistry 48: 838-843.
Baderschneider, B., and Winterhalter, P. (2001). Isolation and characterization of
novel benzoates, cinnamates, flavonoids, and lignans from Riesling wine and
screening for antioxidant activity. Journal of Agricultural and Food Chemistry 49:
2788-2798.
Baxter, N. F., Lilley, T. H., Haslam, E., and Williamson, M. P. (1997). Multiple
interactions between polyphenols and a salivary proline-rich protein repeat results in
complexation and precipitation. Biochemistry 36: 5566-5577.
Berg, I. C. H., Rutland, M. W., and Arnebrant, T. (2003). Lubricating properties of the
initial salivary pellicle – an AFM study. Biofouling 19: 365-369.
Betés-Saura, C., Andrés-Lacueva, C., and Lamuela-Raventós, R. M. (1996). Phenolics
in white free run juices and wines from Penedes by high-performance liquid
chromatography: Changes during vinification. Journal of Agricultural and Food
Chemistry 44: 3040-3046.
Cala, O., Dufourc, E. J., Fouquet, E., Manigand, C., Laguerre, M., and Pianet, I.
(2012). The colloidal state of tannins impacts the nature of their interaction with
proteins: the case of salivary proline-rich protein/procyanidins binding. Langmuir 28:
17410-17418.
Cala, O., Pinaud, N., Simon, C., Fouquet, E., Laguerre, M., Dufourc, E. J., and Pianet,
I. (2010). NMR and molecular modeling of wine tannins binding to saliva proteins:
revisiting astringency from molecular and colloidal prospects. FASEB J 24: 4281-
4290.
Canon, F., Pate, F., Cheynier, V., Sarni-Manchado, P., Giuliani, A., Perez, J., Durand,
D., Li, J., and Cabane, B. (2013). Aggregation of the salivary proline-rich protein IB5
in the presence of the tannin EgCG. Langmuir 29: 1926-1937.
Canon, F., Ballivian, R., Chirot, F., Antoine, R., Sarni-Manchado, P., Lemoine, J., and
Dugourd, P. (2011). Folding of a salivary intrinsically disordered protein upon binding
to tannins. Journal of the American Chemical Society 133: 7847-7852.
Carando, S., Teissedre, P., Pascual-Martinez, L., and Cabanis, J. (1999a). Levels of
flavan-3-ols in French wines. Journal of Agricultural and Food Chemistry 47: 4161-
4166.
Carando, S., Teissedre, P. L., and Cabanis, J. C. (1999b). HPLC coupled with
fluorescence detection for the determination of procyanidins in white wines.
Chromatographia 50: 253-254.
Cardenas, M., Elofsson, U., and Lindh, L. (2007). Salivary mucin MUC5B could be
an important component of in vitro pellicles of human saliva: An in situ ellipsometry
and atomic force microscopy study. Biomacromolecules 8: 1149-1156.
Carpenter, G. H. (2013). Do transient receptor protein (TRP) channels play a role in
oral astringency? Journal of Texture Studies 44: 334-337.

33
Carvalho, E., Mateus, N., Plet, B., Pianet, I., Dufourc, E., and De Freitas, V. (2006).
Influence of wine polysaccharides on the interactions between condensed tannins and
salivary proteins. Journal of Agricultural and Food Chemistry 54: 8936-8944.
Castillo-Muñoz, N., Gómez-Alonso, S., García-Romero, E., and Hermosín-Gutiérrez,
I. (2010). Flavonol profiles of Vitis vinifera white grape cultivars. Journal of Food
Composition and Analysis 23: 699-705.
Cejudo-Bastante, M. J., Castro-Vázquez, L., Hermosín-Gutiérrez, I., and Pérez-
Coello, M. S. (2011a). Combined effects of prefermentative skin maceration and
oxygen addition of must on color-related phenolics, volatile composition, and sensory
characteristics of Airén wine. Journal of Agricultural and Food Chemistry 59: 12171-
12182.
Cejudo-Bastante, M. J., Hermosín-Gutiérrez, I., Castro-Vázquez, L., and Pérez-
Coello, M. S. (2011b). Hyperoxygenation and bottle storage of Chardonnay white
wines: Effects on color-related phenolics, volatile composition, and sensory
characteristics. Journal of Agricultural and Food Chemistry 59: 4171-4182.
Cejudo-Bastante, M. J., Pérez-Coello, M. S., and Hermosín-Gutiérrez, I. (2010).
Identification of new derivatives of 2-S-Glutathionylcaftaric acid in aged white wines
by HPLC-DAD-ESI-MSn. Journal of Agricultural and Food Chemistry 58: 11483-
11492.
Chandrashekar, J., Mueller, K. L., Hoon, M. A., Adler, E., Feng, L. X., Guo, W.,
Zuker, C. S., and Ryba, N. J. P. (2000). T2Rs function as bitter taste receptors. Cell
100: 703-711.
Cheynier, V. F., Trousdale, E. K., Singleton, V. L., and Salgues, M. J. (1986).
Characterization of 2-s-Glutathionylcaftaric acid and its hydrolysis in relation to grape
wines. Journal of Agricultural and Food Chemistry 34: 217-221.
Clapham, E. E. (2003). TRP channels as cellular sensors. Nature 426: 517-524.
Cocola, L. (2016). Validation and comparison of two tunable diode laser absorption
spectroscopy-based carbon dioxide sensors for bottled wine. American Journal of
Enology and Viticulture 67: 127-131.
Coles, J. M., Chang, D. P., and Zauscher, S. (2010). Molecular mechanisms of aqueous
boundary lubrication by mucinous glycoproteins. Current Opinion in Colloid and
Interface Science 15: 406-416.
Dadic, M., and Belleau, G. (1973). Polyphenols and beer flavor. Proceedings of the
American Society of Brewing Chemists 4: 107-114.
Delcour, J. A., Vandenberghe, M. M., Corten, P. F., and Dondeyne, P. (1984). Flavor
thresholds of polyphenolics in water. American Journal of Enology and Viticulture 35:
134-136.
de Pascual-Teresa, S., Rivas-Gonzalo, J. C., and Santos-Buelga, C. (2000).
Prodelphinidins and related flavanols in wine. International Journal of Food Science
& Technology 35: 33-40.
de Villiers, A., Majek, P., Lynen, F., Crouch, A., Lauer, H., and Sandra, P. (2005).
Classification of South African red and white wines according to grape variety based

34
on non-coloured phenolic content. European Food Research and Technology 221:
520-528.
DeMiglio, P., and Pickering, G. J. (2008). The influence of ethanol and pH on the taste
and mouth-feel sensations elicited by red wine. Journal of Food Agriculture and
Environment 6: 143-150.
DeMiglio, P., Reynolds, A., and Pickering, G. J. (2002). Effect of pH and ethanol on
the astringent sub-qualities of red wine. In: Proceedings of the International Baccus
to the Future Conference. pp. 31-52. Cullen, C.W., Pickering, G.J. and Phillips, R.
Eds., Brock University Press, St Catharines.
Desportes, C., Charpentier, M., Duteurtre, B., Maujean, A., and Duchiron, F. (2001).
Isolation, identification, and organoleptic characterization of low-molecular-weight
peptides from white wine. American Journal of Enology and Viticulture 52: 376-380.
Di Lecce, G., Arranz, S., Jáuregui, O., Tresserra-Rimbau, A., Quifer-Rada, P., and
Lamuela-Raventós, R. M. (2014). Phenolic profiling of the skin, pulp and seeds of
Albarino grapes using hybrid quadrupole time-of-flight and triple-quadrupole mass
spectrometry. Food Chemistry 145: 874-882.
DuBois, G. (2011). Validity of early indirect models of taste active sites and advances
in new taste technologies enabled by improved models. Flavour and Fragrance
Journal 26: 239-253.
Esaki, S., Tanaka, R., and Kamiya, S. (1983). Structure-taste relationships of flavanone
and dihydrochalcone glycosides containing (1→2) linked disaccharides. Agricultural
and Biological Chemistry 47: 1831-1834.
Escot, S., Feuillat, M., Dulau, L., and Charpentier, C. (2001). Release of
polysaccharides by yeasts and the influence of released polysaccharides on colur
stability and wine astringency. Australian Journal of Grape and Wine Research 7:
153-159.
Fernández-Pachon, M. S., Villaño, D., Troncoso, A. M., and Garcia-Parrilla, M. C.
(2006). Determination of the phenolic composition of sherry and table white wines by
liquid chromatography and their relation with antioxidant activity. Analytica Chimica
Acta 563: 101-108.
Ferrer-Gallego, R., Brás, N. F., Garcia-Estévez, I., Mateus, N., Rivas-Gonzalo, J. C.,
de Freitas, V., and Escribano-Bailón, M. T. (2016). Effect of flavonols on wine
astringency and their interaction with human saliva. Food Chemistry 209: 358-364.
Ferrer-Gallego, R., Quijada-Morín, N., Brás, N. F., Gómes, P., de Freitas, V., Rivas-
Gonzalo, J. C., and Escribano-Bailón, M. T. (2015). Characterization of sensory
properties of flavanols - A molecular dynamic approach. Chemical Senses 40: 381-
390.
Ferrer-Gallego, R., Hernández-Hierro, J. M., Rivas-Gonzalo, J. C., and Escribano-
Bailón, M. T. (2014). Sensory evaluation of bitterness and astringency sub-qualities of
wine phenolic compounds: synergistic effect and modulation by aromas. Food
Research International 62: 1100-1107.
Fincher, G. B., and Stone, B. A. (1983). Arabinogalactan-proteins: Structure,
biosynthesis and function. Annual Review of Plant Physiology 34: 47-70.

35
Fischer, U., and Noble, A. C. (1994). The effect of ethanol, catechin concentration,
and pH on sourness and bitterness of wine. American Journal of Enology and
Viticulture 45: 6-10.
Fontoin, H., Saucier, C., Teissedre, P. L., and Glories, Y. (2008). Effect of pH, ethanol
and acidity on astringency and bitterness of grape seed tannin oligomers in model wine
solution. Food Quality and Preference 19: 286-291.
Foo, L. Y., and Porter, L. J. (1980). The phytochemistry of proanthocyanidin
polymers. Phytochemistry 19: 1747-1754.
Friedel, M., Stoll, M., Patz, C. D., Will, F., and Dietrich, H. (2015). Impact of light
exposure on fruit composition of white 'Riesling' grape berries (Vitis vinifera L.). Vitis
54: 107-116.
Fulcrand, H., Remy, S., Souquet, J. M., Cheynier, V., and Moutounet, M. (1999).
Study of wine tannin oligomers by on-line liquid chromatography electrospray
ionisation mass spectrometry. Journal of Agricultural and Food Chemistry 47: 1023-
1028.
Furlan, A. L., Jobin, M.-L., Pianet, I., Dufourc, E. J., and Géan, J. (2015).
Flavanol/lipid interaction: a novel molecular perspective in the description of wine
astringency and bitterness and antioxidant action. Tetrahedron 71: 3143-3147.
Furlan, A. L., Castets, A., Nallet, F., Pianet, I., Grélard, A., Dufourc, E. J., and Géan,
J. (2014). Red wine tannins fluidify and precipitate lipid liposomes and bicelles. A role
for lipids in wine tasting? Langmuir 30: 5518-5526.
Gawel, R., Smith, P. A., and Waters, E. J. (2016). The influence of polysaccharides on
the taste and mouth-feel of white wine. Australian Journal of Grape and Wine
Research 22: 350-357.
Gawel, R., Day, M., Van Sluyter, S. C., Holt, H., Waters, E. J., and Smith, P. A.
(2014a). White wine taste and mouth-feel as affected by juice extraction and
processing. Journal of Agricultural and Food Chemistry 62: 10008-10014.
Gawel, R., Schulkin, A., Smith, P. A., and Waters, E. J. (2014b). Taste and textural
characters of mixtures of caftaric acid and Grape Reaction Product in model wine.
Australian Journal of Grape and Wine Research 20: 25-30.
Gawel, R., Van Sluyter, S. C., Smith, P. A., and Waters, E. J. (2013). Effect of pH and
alcohol on perception of phenolic character in white wine. American Journal of
Enology and Viticulture 64: 425-429.
Gawel, R., and Waters, E. J. (2008). The effect of glycerol on the perceived viscosity
of dry white table wine. Journal of Wine Research 19: 109-114.
Gawel, R., Van Sluyter, S., and Waters, E. J. (2007). The effects of ethanol and
glycerol on the body and other sensory characteristics of Riesling wines. Australian
Journal of Grape and Wine Research 13: 38-45.
Gawel, R. (1998). Red wine astringency: a review. Australian Journal of Grape and
Wine Research 4: 74-95.

36
Glabasnia, A., and Hofmann, T. (2006). Sensory-directed identification of taste-active
ellagitannins in American (Quercus alba L.) and European oak wood (Quercus robur
L.) and quantitative analysis in Bourbon whiskey and oak-matured red wines. Journal
of Agricultural and Food Chemistry 54: 3380-3390.
Gonçalves, F., Heyraud, A., De Pinho, M. N., and Rinaudo, M. (2002).
Characterization of white wine mannoproteins. Journal of Agricultural and Food
Chemistry 50: 6097-6101.
Gong, J., and Osada, Y. (1998). Gel friction: A model based on surface repulsion and
adsorption. Journal of Chemical Physics 109: 8062-8068.
Guadagni, D. G., Maier, V. P., and Turnbaugh, J. G. (1973). Effect of some citrus juice
constituents on taste thresholds for limonin and naringin bitterness. Journal of the
Science of Food and Agriculture 24: 1277-1288.
Guadalupe, Z., and Ayestarán, B. (2007). Polysaccharide profile and content during
the vinification and aging of Tempranillo red wines. Journal of Agricultural and Food
Chemistry 55: 10720-10728.
Gürbüz, O., Goçmen, D., Dagdelen, F., Gürsoy, M., Aydin, S., Sahin, I., Büyükuysal,
L., and Usta, M. (2007). Determination of flavan-3-ols and trans-resveratrol in grapes
and wine using HPLC with fluorescence detection. Food Chemistry 100: 518-525.
Hartwig, P., and McDaniel, M. R. (1995). Flavor characteristics of lactic, malic, citric
and acetic acids at various pH levels. Journal of Food Science 60: 384-388.
Haslam, E., and Lilley, T. H. (1988). Natural astringency in foodstuffs: A molecular
interpretation. Critical Revues in Food Science and Nutrition 27: 1-40.
Herrick, I. W., and Nagel, C. W. (1985) The caffeoyl tartrate content of White Riesling
wines from California, Washington, and Alsace. American Journal of Enology and
Viticulture 36: 95-97.
Hewson, L., Hollowood, T., Chandra, S., and Hort, J. (2009). Gustatory, olfactory and
trigeminal interactions in a model carbonated beverage. Chemosensory Perception 2:
94-107.
Horowitz, R. M., and Gentili, B. (1969). Taste and structure of phenolic glycosides.
Journal of Agricultural and Food Chemistry 17: 696-700.
Hufnagel, J. C., and Hofmann, T. (2008). Orosensory-directed identification of
astringent mouth-feel and bitter-tasting compounds in red wine. Journal of
Agricultural and Food Chemistry 56: 1376-1386.
Iontcheva, I., Oppenheim, F. G., Offner, G. D., and Troxler, R. F. (2000). Molecular
mapping of statherin- and histatin-binding domains in human salivary mucin MG1
(MUC5B) by the yeast two-hybrid system. Journal of Dental Research 79: 732-739.
Jackson, R. (2014). Wine Science: Principles and Applications. Academic Press, San
Diego.
Jones, P. R., Gawel, R., Francis, I. L., and Waters, E. J. (2008). The influence of
interactions between major white wine components on the aroma, flavour and texture
of model white wine. Food Quality and Preference 19: 596-607.

37
Kallithraka, S., Salacha, M. I., and Tzourou, I. (2009). Changes in phenolic
composition and antioxidant activity of white wine during bottle storage: Accelerated
browning test versus bottle storage. Food Chemistry 113: 500-505.
Kaneko, S., Kumazawa, K., Masuda, H., Henze, A., and Hofmann, T. (2006).
Molecular and sensory studies on the umami taste of Japanese green tea. Journal of
Agricultural and Food Chemistry 54: 2688-2694.
Kauffman, D. L., Bennick, A., Blum, M., and Keller, P. J. (1991). Basic proline-rich
proteins from human parotid saliva: Relationships of the covalent structures of ten
proteins from a single individual. Biochemistry 30: 3351-3356.
Keast, R. S. J., Bournazel, M. M. E., and Breslin, P. A. S. (2003). A psychophysical
investigation of binary bitter-compound interactions. Chemical Senses 28: 301-313.
Kielhorn, S., and Thorngate, J. H. (1999). Oral sensations associated with the flavan-
3-ols (+)-catechin and (-)-epicatechin. Food Quality and Preference 10: 109-116.
Konitz, R., Fruend, M., Seckler, J., Christmann, M., Netzel, M., Strass, G., Bitsch, R.,
and Bitsch, I. (2003). Einfluss der Mostvorklarung auf die sensorische qualitat von
Riesling weinen aus dem Rheingau. Mitteilungen Klosterneuburg 53: 166-183.
Korenika, A. J., Zulj, M. M., Puhelek, I., Plavsa, T., and Jeromel, A. (2014). Study of
phenolic composition and antioxidant capacity of Croatian macerated wines.
Mitteilungen Klosterneuburg 64: 171-180.
Kosmerl, T., Abramovic, H., and Klofutar, C. (2000). The rheological properties of
Slovenian wines. Journal of Food Engineering 46: 165-171.
Kubıć ková, J., and Grosch, W. (1998). Evaluation of flavour compounds of
Camembert cheese. International Dairy Journal 8: 11-16.
Kurogi, M., Kawai, Y., Nagatomo, K., Tateyama, M., Kubo, Y., and Saitoh, O. (2015).
Auto-oxidation products of epigallocatechin gallate activate TRPA1 and TRPV1 in
sensory neurons. Chemical Senses 40: 27-46.
Lawless, H. T., Horne, J., and Giasi, P. (1996). Astringency of organic acids is related
to pH. Chemical Senses 21: 397-403.
Lea, A. G. H., Bridle, P., Timberlake, C. F., and Singleton, V. L. (1979). The
procyanidins of white grapes and wines. American Journal of Enology and Viticulture
30: 289-300.
Lea, A. G. H., and Arnold, G. M. (1978). The phenolics of ciders: Bitterness and
astringency. Journal of the Science of Food and Agriculture 29: 478-483.
Lee, S. K., Lee, S. W., Chung, S. C., Kim, Y. K., and Kho, H. S. (2002). Analysis of
residual saliva and minor salivary gland secretions in patients with dry mouth.
Archives of Oral Biology 47: 637-641.
Luck, G., Liao, H., Murray, N. J., Grimmer, H. R., Warminski, E. E., Williamson, M.
P., Lilley, T. H., and Haslam, E. (1994). Polyphenol, astringency and proline rich
proteins. Phytochemistry 37: 357-371.
Maga, J. A. (1990). Compound structure versus bitter taste. In: Bitterness in foods and
beverages. pp. 35-48. Rouseff, R. L. Ed. Elsevier, Amsterdam.

38
Maga, J. A., and Lorenz, K. (1973). Taste threshold values for phenolic acids which
can influence flavor properties of certain flours, grains, and oilseeds. Cereal Science
Today 18: 326-328.
Maheshwari, R., and Dhathathreyan, A. (2006). Mucin at solution/air and
solid/solution interfaces. Journal of Colloid and Interface Science 293: 263-269.
Malone, M. E., Appelqvist, I. A. M., and Norton, I. T. (2003). Oral behaviour of food
hydrocolloids and emulsions. Part 1. Lubrication and deposition considerations. Food
Hydrocolloids 17: 763-773.
Masa, A., Vilanova, A., and Pomar, F. (2007). Varietal differences among the
flavonoid profiles of white grape cultivars studied by high-performance liquid
chromatography. Journal of Chromatography A 1164: 291-297.
Mattes, R. D., and DiMeglio, D. (2001). Ethanol perception and ingestion. Physiology
and Behavior 72: 217-229.
McArthur, C., Sanson, G. D., and Beal, A. M. (1995). Salivary proline-rich proteins in
mammals: Roles in oral homeostasis and counteracting dietary tannin. Journal of
Chemical Ecology 21: 663-691.
McRae, J. M., Ziora, Z. M., Kassara, S., Cooper, M. A., and Smith, P. A. (2015).
Ethanol concentration influences the mechanisms of wine tannin interactions with
poly(L-proline) in model wine. Journal of Agricultural and Food Chemistry 63: 4345-
4352.
Meyerhof, W., Batram, C., Kuhn, C., Brockhoff, A., Chudoba, E., Bufe, B.,
Appendino, G., and Behrens, M. (2010). The molecular receptive ranges of human
TAS2R bitter taste receptors. Chemical Senses 35: 157-170.
Monagas, M., Bartolomé, B., and Gomez-Cordovés, C. (2005). Updated knowledge
about the presence of phenolic compounds in wine. Critical Reviews in Food Science
and Nutrition 45: 85-118.
Myers, T. E., and Singleton, V. L. (1979). The nonflavonoid phenolic fraction of wine
and its analysis. American Journal of Enology and Viticulture 30: 98-102.
Narukawa, M., Kimata, H., Noga, C., and Watanabe, T. (2010). Taste characterisation
of green tea catechins. International Journal of Food Science & Technology 45: 1579-
1585.
Nayak, A., and Carpenter, G. H. (2008). A physiological model of tea-induced
astringency. Physiology and Behavior 95: 290-294.
Nicolini, G., Mattivi, F., and Dalla Serra, A. (1991). Iperossigenazione dei mosti:
Conseguenze analitiche e sensoriali su vini dell vendemmia 1989. Rivista di
Viticoltura e di Enologia 3: 45-56.
Nieuwoudt, H. H., Prior, B. A., Pretorius, I. S., and Bauer, F. F. (2002). Glycerol in
South African table wines. An assessment of its relationship to wine quality. South
African Journal of Enology and Viticulture 23: 22-30.
Noble, A. C., and Bursick, G. F. (1984). The contribution of glycerol to perceived
viscosity and sweetness in white wine. American Journal of Enology and Viticulture
35: 110-.

39
Nordbo, H., Darwish, S., and Bhatnagar, R. S. (1984). Salivary viscosity and
lubrication: Influence of pH and calcium. Scandinavian Journal of Dental Research
92: 306-314.
Nurgel, C., and Pickering, G. J. (2005). Contribution of glycerol, ethanol and sugar to
the perception of viscosity and density elicited by model white wines. Journal of
Texture Studies 36: 303-325.
Oberholster, A., Francis, I. L., Iland, P. G., and Waters, E. J. (2009). Mouth-feel of
white wines made with and without pomace contact and added anthocyanins.
Australian Journal of Grape and Wine Research 15: 59-69.
Okamura, S., and Watanabe, M. (1981). Determination of phenolic cinnamates in
white wine and their effect on wine quality. Agricultural and Biological Chemistry 45:
2063-2070.
Okuda, T., Fukui, M., Hisamoto, M., Iino, S., Hikawa, Y., Ogino, S., Takayanagi, T.,
and Yokotsuka, K. (2007). Relationships between macromolecules and thickness of
dry white wine. Journal of the American Society of Enology and Viticulture Japan 18:
15-21.
Ong, B. Y., and Nagel, C. W. (1978). Hydroxycinnamic acid-tartaric acid ester content
in mature grapes and during the maturation of White Riesling grapes. American
Journal of Enology and Viticulture 29: 277-281.
Oszmianski, J., and Sapis, J. C. (1989). Fractionation and identification of some low
molecular weight grape seed phenolics. Journal of Agricultural and Food Chemistry
37: 1293-1297.
Pascal, C., Pate, F., Cheynier, V., and Delsuc, M. A. (2009). Study of the interactions
between a proline-rich protein and a flavan-3-ol by NMR: Residual structures in the
natively unfolded protein provides anchorage points for the ligand. Biopolymers 91:
745-756.
Pascal, C., Poncet-Legrand, C., Cabane, B., and Verhnet, A. (2008). Aggregation of a
proline-rich protein induced by epigallocatechin gallate and condensed tannins: Effect
of protein glycosylation. Journal of Agricultural and Food Chemistry 56: 6724-6732.
Payne, C., Bowyer, P. K., Herderich, M., and Bastian, S. E. P. (2009). Interaction of
astringent grape seed procyanidins with oral epithelial cells. Food Chemistry 115: 551-
557.
Peleg, H., Gacon, K., Schlich, P., and Noble, A. C. (1999). Bitterness and astringency
of flavan-3-ol monomers, dimers and trimers. Journal of the Science of Food and
Agriculture 79: 1123-1128.
Peleg, H., and Noble, A. C. (1995). Perceptual properties of benzoic acid derivatives.
Chemical Senses 20: 393-400.
Pellerin, P., Doco, T., Vidal, S., P., W., Brillouet, J. M., and O’Neill, M. A. (1996).
Structural characterization of red wine rhamnogalacturonan II. Carbohydrate
Research 190: 183-197.
Peña-Neira, A., Hernández, T., Garcia-Vallejo, C., Estrella, I., and Suarez, J. A.
(2000). A survey of phenolic compounds in Spanish wines of different geographical
origin. European Food Research and Technology 210: 445-448.

40
Pickering, G. J., and DeMiglio, P. (2008). The wine wine mouth-feel wheel: A lexicon
for describing the oral sensations elicited by white wine. Journal of Wine Research 19:
51-67.
Plet, B., Delcambre, A., Chaignepain, S., and Schmitter, J. M. (2015). Affinity ranking
of peptide–polyphenol non-covalent assemblies by mass spectrometry approaches.
Tetrahedron 71: 3007-3011.
Poncet-Legrand, C., Edelmann, A., Putaux, J. L., Cartalade, D., Sarni-Manchado, P.,
and Vernhet, A. (2006). Poly (L-proline) interactions with flavan-3-ols units: Influence
of the molecular structure and the polyphenol/protein ratio. Food Hydrocolloids 20:
687-697.
Proctor, G. B., Hamdan, S., Carpenter, G. H., and Wilde, P. (2005). A statherin and
calcium enriched layer at the air interface of human parotid saliva. Biochemical
Journal 389: 111-116.
Quijada-Morín, N., Williams, P., Rivas-Gonzalo, J. C., Doco, T., and Escribano-
Bailón, M. T. (2014). Polyphenolic, polysaccharide and oligosaccharide composition
of Tempranillo red wines and their relationship with the perceived astringency. Food
Chemistry 154: 44-51.
Ricardo da Silva, J. M., Cheynier, V., Samsom, A., and Bourzeux, M. (1993). Effect
of pomace contact, carbonic maceration, and hyperoxidation on the procyanidin
composition of Grenache blanc wines. American Journal of Enology and Viticulture
44: 168-172.
Rinaldi, A., Gambuti, A., and Moio, L. (2012). Precipitation of salivary proteins after
the interaction with wine: the effect of ethanol, pH, fructose, and mannoproteins.
Journal of Food Science 77: 485-490.
Robichaud, J. L., and Noble, A. C. (1990). Astringency and bitterness of selected
phenolics in wine. Journal of the Science of Food and Agriculture 53: 343-353.
Robinson, J. (1999). Vines Grapes and Wines. Mitchell Beazley, London.
Rodrígues Montealegre, R., Romero Peces, R., Chacón Vozmediano, J. L., Martinez
Gascueña, J., and Carcia Romero, E. (2006). Phenolic compounds in skins and seeds
of ten grape Vitis vinifera varieties grown in warm climates. Journal of Food
Composition and Analysis 19: 687-693.
Rossetti, D., Bongaerts, J. H. H., Wantling, E., Stokes, J. R., and Williamson, A. M.
(2009). Astringency of tea catechins: More than an oral lubrication tactile percept.
Food Hydrocolloids 23: 1984-1992.
Rotzoll, N., Dunkel, A., and Hofmann, T. (2006). Quantitative studies, taste
reconstitution, and omission experiments on the key taste compounds in Morel
mushrooms (Morchella deliciosa fr.). Journal of Agricultural and Food Chemistry 54:
2705-2711.
Runnebaum, R. C., Boulton, R. B., Powell, R. L., and Heymann, H. (2011). Key
constituents affecting wine body - an exploratory study. Journal of Sensory Studies
26: 62-70.

41
Sáenz-Navajas, M. P., Avizcuri, J. M., Ferreira, V., and Fernández-Zurbano, P. (2012).
Insights on the chemical basis of the astringency of Spanish red wines. Food Chemistry
134: 1484-1493.
Sarni-Manchado, P., and Cheynier, V. (2002). Study of non-covalent complexation
between catechin derivatives and peptides by electrospray ionization mass
spectrometry. Journal of Mass Spectrometry 37: 609-616.
Scharbert, S., Holzmann, N., and Hofmann, T. (2004). Identification of the astringent
taste compounds in black tea by combining instrumental analysis and human
bioresponse. Journal of Agricultural and Food Chemistry 52: 3498-3508.
Schobel, N., Radtke, D., Kyereme, J., Wollmann, N., Cichy, A., Obst, K., Kallweit,
K., Kletke, O., Minovi, A., Dazert, S., Wetzel, C. H., Vogt-Eisele, A., Gisselmann, G.,
Ley, J. P., Bartoshuk, L. M., Spehr, J., Hofmann, T., and Hatt, H. (2014). Astringency
is a trigeminal sensation that involves the activation of G protein-coupled signaling by
phenolic compounds. Chemical Senses 39: 471-487.
Schwender, N., Huber, K., Al Marrawi, F., Hannig, M., and Ziegler, C. (2005). Initial
bioadhesion on surfaces in the oral cavity investigated by scanning force microscopy.
Applied Surface Science 252: 117-122.
Simonetti, P., Pietta, P., and Testolin, G. (1997). Polyphenol content and total
antioxidant potential of selected Italian wines. Journal of Agricultural and Food
Chemistry 45: 1152-1155.
Singleton, V. L., Salgues, M., Zaya, J., and Trousdale, E. (1985). Caftaric acid
disappearance and conversion to products of enzymic oxidation in grape must and
wine. American Journal of Enology and Viticulture 36: 50-56.
Singleton, V. L., and Trousdale, E. (1983). White wine phenolics - varietal and
processing differences as shown by HPLC. American Journal of Enology and
Viticulture 34: 27-34.
Singleton, V. L., Zaya, J., and Trousdale, E. (1980). White table wine quality and
polyphenol composition as affected by must SO2 content and pomace contact time.
American Journal of Enology and Viticulture 31: 14-20.
Singleton, V. L., Sieberhagen, H. A., de Wet, P., and van Wyk, C. J. (1975).
Composition and sensory qualities of wines prepared from white grapes by
fermentation with and without grape solids. American Journal of Enology and
Viticulture 26: 62-69.
Skogerson, K., Runnebaum, R., Wohlgemuth, G., de Ropp, J., Heymann, H., and
Fiehn, O. (2009). Comparison of gas chromatography-coupled time-of-flight mass
spectrometry and 1H nuclear magnetic resonance spectroscopy metabolite
identification in white wines from a sensory study investigating wine body. Journal of
Agricultural and Food Chemistry 57: 6899-6907.
Soares, S., Ferrer-Galego, R., Brandao, E., Silva, M., Mateus, N., and De Freitas, V.
(2016). Contribution of human oral cells to astringency by binding salivary
protein/tannin complexes. Journal of Agricultural and Food Chemistry 64: 7823-7828

42
Soares, S., Kohl, S., Thalmann, S., Mateus, N., Meyerhof, W., and De Freitas, V.
(2013). Different phenolic compounds activate distinct human bitter taste receptors.
Journal of Agricultural and Food Chemistry 61: 1525-1533.
Soares, S. I., Gonçalves, R. M., Fernandes, I., Mateus, N., and de Freitas, V. (2009).
Mechanistic approach by which polysaccharides inhibit alpha-amylase/procyanidin
aggregation. Journal of Agricultural and Food Chemistry 57: 4352-4358.
Sokolowsky, M., and Fischer, U. (2012). Evaluation of bitterness in white wine
applying descriptive analysis, time-intensity analysis, and temporal dominance of
sensations analysis. Analytica Chimica Acta 732: 46-52.
Solms, J. (1969). The taste of amino acids, peptides and proteins. Journal of
Agricultural and Food Chemistry 17: 686-688.
Sowalski, R. A., and Noble, A. C. (1998). Comparison of the effects of concentration,
pH and anion species on astringency and sourness of organic acids. Chemical Senses
23: 343-349.
Spencer, C. M., Cai, Y., Martin, R., Gaffney, S. H., Goulding, P. N., Magnolato, D.,
Lilley, T. H., and Haslam, E. (1988). Polyphenol complexation: Some thoughts and
observations. Phytochemistry 27: 2397-2409.
Squier, C. A., and Kremer, M. J. (2001). Biology of the oral mucosa and esophagus.
Journal of the National Cancer Institute, Monographs 29: 7-15.
Svendsen, I. E., Lindh, L., and Arnebrant, T. (2006). Adsorption behaviour and
surfactant elution of cationic salivary proteins at solid/liquid interfaces, studied by in
situ ellipsometry. Colloids and Surfaces B-Biointerfaces 63: 157-166.
Symoneaux, R., Le Quére, J. M., Baron, A., Bauduin, R., and Chollet, S. (2015).
Impact of CO2 and its interactions with the matrix components on sensory perception
in model cider. LWT-Food Science and Technology 63: 886-891.
Szabo, D., Akiyoshi, S., Matsunaga, T., Gong, J. P., and Osada, Y. (2000). Spreading
of liquids on gel surfaces. Journal of Chemical Physics 113: 8253-8259.
Takahashi, K., Tadenuma, M., Kitamoto, K., and Sato, S. (1974). L-Prolyl-L-leucine
anhydride: A bitter compound formed in aged sake. Agricultural and Biological
Chemistry 38: 927-932.
Thorngate, J. H., and Noble, A. C. (1995). Sensory evaluation of bitterness and
astringency of 3R(–)-epicatechin and 3S(+)-catechin. Journal of the Science of Food
and Agriculture 67: 531-535.
Trevisani, M., Smart, D., Gunthorpe, M. J., Tognetto, M., Barbieri, M., Campi, B.,
Amadesi, S., Gray, J., Jerman, J. C., Brough, S. J., Owen, D., Smith, G. D., Randall,
A. D., Harrison, S., Bianchi, A., Davis, J. B., and Geppetti, P. (2002). Ethanol elicits
and potentiates nociceptor responses via the vanilloid receptor-1. Nature Neuroscience
5: 546-551.
van Vliet, T., van Aken, G. A., de Jongh, H. H., and Hamer, R. J. (2009). Colloidal
aspects of texture perception. Advances in Colloid and Interface Science 150: 27-40.

43
Veerman, E. C. I., Valentijm-Benz, M., and Nieuw Amerongen, A. V. (1989).
Viscosity of human salivary mucins: Effect of pH and ionic strength and role of sialic
acid. Journal de Biologie Buccale 17: 297-306.
Vérette, E., Noble, A. C., and Somers, T. C. (1988). Hydroxy-cinnamates of Vitis
vinifera: Sensory assessment in relation to bitterness in white wines. Journal of the
Science of Food and Agriculture 45: 267-272.
Verhagen, J. V., and Engelen, L. (2006). The neurocognitive bases of human
multimodal food perception: Sensory integration. Neuroscience and Biobehavioral
Reviews 30: 613-650.
Verhnet, A., Pellerin, P., Prieur, C., Osmianski, J., and Moutounet, M. (1996). Charge
properties of some grape and wine polysaccharide and polyphenolic fractions.
American Journal of Enology and Viticulture 47: 25-30.
Vidal, S., Francis, L., Williams, P., Kwiatkowski, M., Gawel, R., Cheynier, V., and
Waters, E. (2004a). The mouth-feel properties of polysaccharides and anthocyanins in
a wine like medium. Food Chemistry 85: 519-525.
Vidal, S., Courcoux, P., Francis, L., Kwiatkowski, M., Gawel, R., Williams, P.,
Waters, E., and Cheynier, V. (2004b). Use of an experimental design approach for
evaluation of key wine components on mouth-feel perception. Food Quality and
Preference 15: 209-217.
Vilanova, M., Santalla, M., and Masa, A. (2009). Environmental and genetic variation
of phenolic compounds in grapes (Vitis vinifera) from northwest Spain. Journal of
Agricultural Science 147: 683-697.
Vitkov, L., Hannig, M., Nekrashevych, Y., and Krautgartner, W. D. (2004).
Supramolecular pellicle precursors. European Journal of Oral Sciences 112: 320-325.
Vitrac, X., Monti, J. P., Vercauteren, J., Deffieux, G., and Merillon, J. M. (2002).
Direct liquid chromatographic analysis of resveratrol derivatives and flavanonols in
wines with absorbance and fluorescence detection. Analytica Chimica Acta 458: 103-
110.
Walz, A., Stuhler, K., Wattenberg, A., Hawranke, E., Meyer, H. E., Schmalz, G.,
Bluggel, M., and Ruhl, S. (2006). Proteome analysis of glandular parotid and
submandibular-sublingual saliva in comparison to whole human saliva by two-
dimensional gel electrophoresis. Proteomics 6: 1631-1639.
Watanabe, I. (2004). Ultrastructures of mechanoreceptors in the oral mucosa.
Anatomical Science International 79: 55-61.
Yamazaki, T., Narukawa, M., Mochizuki, M., Misaka, T., and Watanabe, T. (2013).
Activation of the hTAS2R14 human bitter-taste receptor by (-)-epogallocatechin
gallate and (-)-epcatechin gallate. Bioscience, Biotechnology and Biochemistry 77:
1981-1983.
Yao, Y., Berg, E. A., Costello, C. E., Troxler, R. F., and Oppenheim, F. G. (2003).
Identification of protein components in human acquired enamel pellicle and whole
saliva using novel proteomics approaches. Journal of Biological Chemistry 278: 5300-
5308.

44
Ye, A., Zheng, T., Ye, J. Z., and Singh, H. (2012). Potential role of the binding of
whey proteins to human buccal cells on the perception of astringency in whey protein
beverages. Physiology and Behavior 106: 645-650.
Yu, P., Yeo, A. S., Low, M., and Zhou, W. (2014). Identifying key non-volatile
compounds in ready-to-drink green tea and their impact on taste profile. Food
Chemistry 155: 9-16.

45
Chapter 2.

The effect of ethanol and glycerol on the body and other sensory
characteristics of Riesling wines

Richard Gawel, Steven C. Van Sluyter and Elizabeth J. Waters

The Australian Wine Research Institute, PO Box 197, Glen Osmond, 5064, Australia

Australian Journal of Grape and Wine Research – 2007; 13: 38-45.


DOI: 10.1111/j.1755-0238.2007.tb00070.x

47
38 Sensory effects of ethanol and glycerol in Riesling wines Australian Journal of Grape and Wine Research 13, 38–45, 2007

The effects of ethanol and glycerol on the body and other


sensory characteristics of Riesling wines
RICHARD GAWEL1,2,3, STEVEN VAN SLUYTER1 and ELIZABETH J.WATERS1,2
1
The Australian Wine Research Institute, PO Box 197, Glen Osmond, SA 5064, Australia
2
Cooperative Research Centre for Viticulture, PO Box 154, Glen Osmond, 5064, Australia
3
Corresponding Author: Mr Richard Gawel, facsimile: +61 8 8303 6601, email richard.gawel@awri.com.au

Abstract
The effect of ethanol and glycerol concentration on the body, sweetness, acidity, aroma and flavour
intensity, and perceived viscosity and hotness of three Riesling wines was assessed. The ethanol and
glycerol contents of the wines were adjusted by addition to give three realistic levels (5.2, 7.2, 10.2 g/L
glycerol and 11.6, 12.6 and 13.6 v/v ethanol). The nine treatment combinations (3 glycerol × 3 ethanol)
were rated on the above attributes by a panel of trained tasters. Increased alcohol levels resulted in
increased perceived hotness in all wines, and in higher body and perceived viscosity in two of the three
wines. The effect of increasing glycerol content was less consistent with only one of each of the three wines
showing increased viscosity and body. However, the mean viscosity ratings given to wines with 10 g/L
glycerol was higher than at 5 g/L at all alcohol levels and for all wines, suggesting that differences in
glycerol concentration typically displayed between dry white table wines can affect their perceived
viscosity. Neither alcohol nor glycerol consistently affected sweetness, acidity, aroma or flavour intensity.
Higher ratings of the abstract term ‘body’ were most commonly associated with higher ratings of flavour
and/or perceived viscosity, suggesting that for the majority of tasters, these two attributes contributed to
their interpretation of the term ‘body’. Perceived hotness was not an important component of body, while
the role of acidity in body perception was taster dependent.

Keywords: white wine, wine body, glycerol, alcohol level, perceived viscosity, alcohol hotness

Introduction equivalent understanding of ‘body’ in a group of Char-


The terms ‘body’ and ‘fullness’ are frequently used to donnay wines, and considered the feature important in
describe the in-mouth impression of both red and white distinguishing between the wines.
table wines. Wines are routinely categorised as being light, It has long been speculated that alcohol strongly con-
medium or full-bodied – presumably as wines of different tributes to palate fullness in white wine (Amerine and
style appeal to different market segments, and are con- Roessler 1983). Pickering et al. (1998) were the first to for-
sumed in different social and culinary contexts. However, mally test this premise. They found that the perceived
despite its widespread use and application, there appears density of a de-alcoholised wine generally increased with
to be a lack of common understanding within the wine increasing alcohol over a 14% v/v range, while its per-
trade as to what sensory aspects contribute to wine body. ceived viscosity was highest at 10% ethanol. Later work
Beers are also routinely categorised by fullness, and using model wines showed a positive monotonic effect of
attempts have been made to define body in the context of alcohol content on both perceived viscosity and density
this beverage. Langstaff et al. (1991a) considered fullness over the same alcohol range (Nurgel and Pickering 2005),
as being one of three primary mouth-feel classes, with vis- further supporting the existence of a positive relationship
cosity and density being contributing sub-qualities. They between alcohol content and fullness in white wine.
defined viscosity as the ‘degree to which beer resists flow Glycerol is a major product of yeast fermentation and
under an applied force in the mouth’, and density as ‘the is reported to range up to 9.9 g/L in Australian white table
perceived density or weight of beer in the mouth’. Despite wines (Rankine and Bidson 1971), and 9.36 g/L in South
these attempts to define and classify fullness in beer, a African dry white wines (Nieuwoudt et al. 2002). In its
number of issues remain. Most importantly, there appears pure form glycerol is a viscous liquid at room temperature.
to be no agreed position on the necessary conditions for Therefore it is reasonable to assume that it contributes to
‘fullness’ in either beer or wine. That is, what attributes, the perceived viscosity and fullness of dry white wines.
if missing, would preclude a wine from being full-bodied? However, Noble and Bursick (1984) estimated that an
Despite the apparent lack of agreement on what consti- additional 26 g/L of glycerol is required before an increase
tutes body in wine, Gawel (1997) showed that experi- in white wine viscosity is just noticeable. Based on this
enced wine tasters with extensive practical training had an result, it is unlikely that glycerol concentration influences
Gawel, van Sluyter & Waters Sensory effects of ethanol and glycerol in Riesling wines 39

the perceived viscosity of dry white wine. However, hotness, acidity, sweetness, aroma and flavour intensity)
recently Nurgel and Pickering (2005) reported enhanced using the same scale. In addition, assessors were asked to
perceived viscosity of a model wine upon increasing its select which wine of each pair they considered to be fuller
glycerol concentration from 10 to 25 g/L. in body. Immediately following each tasting the panellists
The contribution of ethanol to wine sensory properties discussed their results. The panel moderator used these
extends beyond that of possibly enhancing fullness. discussions to consolidate a common understanding of
Ethanol affects the headspace concentrations of many each of the attributes excluding body. Assessors were
wine volatiles (Guth and Sies 2002), and also contributes given overall panel feedback in the form of a written
to sweetness (Scinska et al. 2000). Furthermore, ethanol- report giving collated results at the following session.
induced palate warmth and perceived viscosity may While the feedback included the identity of the wines, the
indirectly affect both aroma and flavour perception (see panel moderator intentionally avoided making any sug-
Delwiche 2004 for a review). gestions as to any possible effects of the chemical modifi-
This paper investigates the effect of realistic levels of cations to the wines. To illustrate the concept of viscosity,
ethanol and glycerol on the body, viscosity, hotness, aroma a control white wine was compared with the same wine
and taste intensity of dry white table wine. Furthermore, to which a commercially available food grade thickener
it attempts to explore assessor interpretation of the con- was added. Only opinions as to perceived viscosity were
cept of wine body as distinct from perceived viscosity and discussed during this session.
density.
Formal assessment
Methods Three commercial South Australian Riesling wines were
Tasting panel selected on the basis of their low alcohol (< 12% v/v) con-
A panel of 10 volunteer assessors comprising one female tent (Table 1). Alcohol level was measured using NIR
and nine male employees of the Australian Wine Research (AWRI Analytical Services) and glycerol concentration
Institute was convened. All but one taster had at least two was determined enzymatically as previously described
years general wine tasting experience as part of their pro- (Nieuwoudt et al. 2002). Each wine was either diluted
fession, but none had participated in previous training with mineral water or fortified with food grade 96% v/v
specifically relating to wine body. ethanol (Tarac Technologies, SA) to 11.6, 12.6 and 13.6%
v/v, and glycerol (Symex, Vic) from a 50% w/v stock
Assessor training solution in water to 5.2, 7.2 and 10.2 g/L. These additions
Training consisted of three, forty-minute sessions per resulted in nine (three ethanol × three glycerol) treatment
week over four weeks. The purpose of training was to (1) combinations for each of the three wines. Flavour dilution
accommodate assessor views on which attributes in- effects were likely to be insignificant as the maximum
fluence body in white wine, (2) ensure that there was no additions of water and ethanol were 0.3% and 1.8%
redundancy in the selected attributes, (3) arrive at a broad respectively.
definition of wine body and definitions of the attributes
contributing to body, and 4) select and refine an appro- Table 1. Composition of the base wines.
priate scale.
Initial training consisted of presenting sets of four or Wine Alcohol Glycerol Fermentable pH Titratable
sugars acidity
five young commercial Australian and New Zealand dry % v/v g/L g/L g/L
white table wines of varieties Riesling, Chardonnay,
Viognier, Semillon, Sauvignon Blanc and Pinot Gris. The A 11.9 5.1 1.1 2.96 7.3
assessors discussed and justified whether the wines were B 11.7 5.3 2.5 3.07 6.0
light, medium or full-bodied. These discussions gave rise C 11.9 5.3 1.3 3.24 6.9
to a list of attributes that were thought by the assessors to
affect perceived fullness in white table wines. The formal assessment was conducted in tasting booths
In later sessions, assessors were randomly presented with 30 mL of each sample being presented at a constant
with pairs of Riesling wines, one being a control and the room temperature in black ISO wine tasting glasses.
other the same wine adjusted by either (1) distilled water During each session, all nine treatment combinations of a
addition of 10% (v/v), or (2) ethanol addition of 1.0, 1.5% particular wine were presented in a completely ran-
or 2.0% (v/v), or (3) glycerol addition of 2, 4 or 5 g/L, or domised order to each assessor. The assessors rated the
(4) combinations of glycerol and ethanol given in (2) and intensity of aroma, body and acidity using a nine point
(3). On one occasion, a commercial wine that had previ- category scale developed on the advice of the assessors
ously been identified as full-bodied was blended (50:50 during training. Word anchors were applied to the odd-
v/v) with another that had previously been identified as numbered categories. For the body attribute, these were
light-bodied. This blend was either presented with the light, light-medium, medium, medium-full and full. For all
light-bodied or the full-bodied component of the blend. In other attributes the word anchors describing intensity
all cases assessors were asked to rate the fullness of each were low, low-medium, medium, medium-high and high.
wine of the pair on a nine point category scale with word After surrendering their ballots and waiting between one
anchors on every second scale point. They also rated the and three minutes, the assessors re-tasted the wines and
intensity of the previously selected attributes (viscosity, rated viscosity, hotness, flavour and sweetness using the
40 Sensory effects of ethanol and glycerol in Riesling wines Australian Journal of Grape and Wine Research 13, 38–45, 2007

same scale. This protocol was subsequently repeated, three Correlations between body and the predictor variables
times a week over three weeks. In total, data for the nine were determined using Spearmans rank correlation co-
treatments by three wines by three tasting replicates was efficient, and individual judge ratings of body were
collected. modelled on the predictor variables of viscosity, flavour,
Directional difference tests were also employed to acidity, hotness and sweetness using ordinal logistic re-
enable direct comparison of the fullness of wines with gression. To assist in the interpretation of the individual
increased alcohol and glycerol levels. These tests were con- models, points on the nine point category scale for body
ducted 6 months after the rating sessions described above. were combined into three categories, 1–3 being categorised
The panel consisted of 12 assessors, five of whom had pre- as light-bodied, 4–6 as medium bodied and 7–9 as full-
viously participated in the body/fullness test described bodied. All analysis was performed using MINITAB
above. During the recruitment process potential assessors Release 14.13 (Minitab Inc). Directional difference test
were asked if they were familiar with the term body and data with pooled replicates was analysed using tables
would be confident of their ability to use this term in a derived from the binomial distribution. As overall con-
sensory test. All samples were prepared on the morning of fidence in completing a task is known to be highly
the tasting with the exception of day two, where wines individualistic, the confidence ratings provided by each
from day one were used. In this instance wines from day assessor were normalised by subtracting the grand mean
one were blanketed with nitrogen gas and kept at 4°C confidence rating of that individual from each of their
overnight. The treatments were: wine A or wine B with confidence ratings. Due to the inherent variability in sen-
(a) ethanol addition of 2%, (b) glycerol addition of 5 g/L sory ratings, a significance level of 10% was considered
or (c) ethanol addition of 2% and glycerol addition of 5 appropriate and used throughout.
g/L. Each of these treatments was paired with its respec-
tive base wine (i.e. either wine A or wine B) in a direc- Results
tional difference test protocol. Attribute selection and definition
Wine (30 mL) was presented to the assessors under the The commercial wines used for training were selected on
conditions described above. Replicate tastings were con- the basis of expected differences in body resulting from
ducted over three consecutive days, whereby six pairs of variations in grape variety, alcohol content and oak treat-
wines were presented in a randomised order balanced ment. After tasting these wines and discussing their per-
across the twelve assessors. Tasters were required to taste ceptions, the assessors agreed on the following concepts
the pairs of samples in the order in which they were pre- and definitions: (1) body was defined as ‘the overall
sented and indicate which of the pair was fullest in body. impression of weight or substantiveness of the wine in the
Consistent with the previous trial, a definition of full- mouth’; (2) the terms ‘body’ and ‘fullness’ were syn-
ness/body was not provided. The assessors were also asked onyms; (3) viscosity was defined as ‘the amount of force
to indicate how confident they were in their selection that must be applied to move the wine around in the
using a five point category scale with the following verbal mouth’; (4) viscosity was deemed to be a single physical
descriptors on each point: 1. not at all confident, 2. a low property of the liquid phase of the wine equating to its
degree of confidence, 3. moderately confident, 4. a high ‘thinness/thickness’, and; (5) aroma and flavour were
degree of confidence, 5. a very high degree of confidence. defined as the wine’s fruitiness perceived orthonasally and
Assessors were asked to rinse their mouth with water and retronasally, respectively.
were forced to wait 30 seconds before tasting the next set. The assessors also agreed that wine body was an
abstract concept rather than a single sensory attribute. The
Statistical analysis consensus was that wine body may be influenced by the
Category scales were used to collect attribute intensity and attributes of flavour intensity, viscosity, sweetness, acidity
body ratings. As such the data generated were considered and heat, so these were included in the study. Overall
to be ordinal scale variables. For this reason, non-para- aroma intensity was included to investigate the effect of
metric statistical analysis methods were chosen in favour alcohol and glycerol level on perceived aroma rather than
of parametric methods such as analysis of variance on wine body per se.
(ANOVA) and multiple regression. The attribute ratings
were modelled on alcohol and glycerol content using Effect of alcohol and glycerol
ordinal logistic regression using a logit function as the Alcohol level had a significant positive effect on the body
integrator. The data were summarised as mean ratings, of wines B and C (Table 2 and Figure 1). Wine A showed
with rating variability represented as standard errors. some evidence of increasing body with alcohol content at

Table 2. Significance (P values) of the effect of ethanol and glycerol concentration on intensity of sensory attributes.

Wine Body Viscosity Flavour Hotness Acidity Sweetness Aroma


Ethanol Glycerol Ethanol Glycerol Ethanol Glycerol Ethanol Glycerol Ethanol Glycerol Ethanol Glycerol Ethanol Glycerol
A 0.81 0.44 0.78 0.06 0.41 0.67 0.01 0.14 0.59 0.61 0.84 0.73 0.82 0.43
B 0.09 0.07 0.12 0.46 0.81 0.40 0.01 0.70 0.06 0.86 0.48 0.71 0.48 0.07
C 0.10 0.51 0.18 0.94 0.38 0.20 0.01 0.76 0.41 0.50 0.83 0.59 0.75 0.26
Gawel, van Sluyter & Waters Sensory effects of ethanol and glycerol in Riesling wines 41

6.0
Wine A 6.0

Mean viscosity rating


5.5 Wine A
5.5
Mean body rating

5.0 5.0
4.5
4.5
4.0
4.0
3.5
3.5 3.0
11.6 12.6 13.6
3.0
Alcohol (%v/v)
11.6 12.6 13.6

5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol


Alcohol (%v/v)
5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol
6.0
Wine B

Mean viscosity rating


6.0
5.5
Wine B
Mean body rating

5.5
5.0
5.0
4.5
4.5
4.0
4.0
3.5
3.5
3.0
3.0
11.6 12.6 13.6 11.6 12.6 13.6

Alcohol (%v/v) Alcohol (%v/v)


5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol 5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol

6.0 6.0
Mean viscosity rating
Wine C Wine C
Mean body rating

5.5 5.5

5.0 5.0

4.5 4.5

4.0 4.0

3.5 3.5

3.0 3.0
11.6 12.6 13.6 11.6 12.6 13.6
Alcohol (%v/v) Alcohol (%v/v)
5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol 5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol

Figure 1. The effect of increasing ethanol and glycerol concentration Figure 2. The effect of increasing ethanol and glycerol concentration
in Wine A, Wine B or Wine C on the mean body rating. Two standard in Wine A, Wine B or Wine C on the mean perceived viscosity rating.
errors of ratings from three replicates and ten assessors are shown Two standard errors of ratings from three replicates and ten
only to illustrate rating variability. assessors are shown only to illustrate rating variability.

the highest glycerol concentration (Figure 1). Directional differences, the assessors indicated a general lack of con-
difference tests also showed that enhanced levels of alco- fidence in their responses. This, together with the fact that
hol increased the body of wine B, as did combined higher the glycerol content did not affect the fullness of wine C
levels of alcohol and glycerol in wine A (Table 3). at any alcohol level (Figure 1), suggests that the effect of
The effect of glycerol concentration on body was also glycerol on wine body in the range used in this study was,
wine-dependent. The body of wine B was positively at best, subtle.
affected by glycerol content, as was wine A at the highest The effect of alcohol on perceived viscosity was not
alcohol level (Table 2, Figure 1). When compared directly, statistically significant for any of the three wines (Table 2).
higher glycerol in wine B was seen to have increased the However, there was a consistent increase in perceived vis-
body of that wine, but in wine A, a combination of higher cosity with increasing alcohol in wine B and C (Figure 2).
alcohol and glycerol was most suggestive of increased For both wines, 13.6% v/v alcohol resulted in significantly
body (Table 3). It is noteworthy that despite the observed higher viscosity than 11.6% v/v but only at the highest

Table 3. Perceived differences in the body of Riesling wines with varying levels of alcohol and glycerol.

Treatment Control + 2% P Control + 5 g/L P Control + 2 % ethanol P


ethanol glycerol + 5 g/L glycerol
Wine A
No. selected as higher in body 16 20 0.309 18 18 0.566 14 22 0.121
Assessor confidence* 0.174 0.024 –0.057 –0.320 –0.586 0.294

Wine B
No. selected as higher in body 13 23 0.066 13 23 0.066 15 21 0.203
Assessor confidence* 0.045 0.073 –0.178 –0.149 0.295 0.040

* Overall degree of assessor confidence in their evaluation of the relative body of Riesling wines which vary in alcohol and glycerol concentration. A positive
(negative) value indicates that on average the assessors were more (less) confident in their choice of the fullest bodied wine.
42 Sensory effects of ethanol and glycerol in Riesling wines Australian Journal of Grape and Wine Research 13, 38–45, 2007

5.0
Combined Wines
Assessor interpretation of body
Mean hotness rating

4.5 Table 4 shows the level of significance of the regression


coefficients when body category ratings (light, medium
4.0
and full-bodied) were modelled on the ratings for poten-
3.5
tial predictor variables for individual assessors. The
3.0 models provided a good fit between wine body and the
11.6 12.6 13.6
Alcohol (%v/v) predictor variables of flavour, viscosity, hotness, acidity
5.2 g/L Glycerol 7.2 g/L Glycerol 10.2 g/L Glycerol
and sweetness for all assessors. The coefficients for vis-
cosity and flavour were significant for five of the ten asses-
Figure 3. The effect of increasing ethanol and glycerol concentration sors who indicated a positive relationship between these
in all wines on the mean perceived hotness rating. Two standard two attributes and overall body. Viscosity and flavour also
errors of ratings from three replicates, ten assessors and three wines displayed the strongest correlations with body (Table 5).
are shown only to illustrate rating variability.
For two assessors, higher acidity was also positively
associated with fullness, however a negative association
glycerol concentration.
was observed between body and acidity for four of the
Higher glycerol concentrations favoured perceived vis-
assessors. Hotness was not associated with fullness in any
cosity in wine A (Table 2 and Figure 2). At the highest
instance, and a minority of judges provided evidence for
alcohol level, the perceived viscosity of wine B with 10.2
a positive association between fullness and sweetness.
g/L glycerol was higher than at either 5.2 or 7.2 g/L. No
clear trend in viscosity with glycerol was seen in wine C.
A robust effect of in-mouth heat with increasing
Discussion
alcohol was seen for all three wines. As the effects were
The term ‘fullness’ is a commonly used term to indicate
independent of wine, the pooled results are given in
the style of wine. However there appears to be little agree-
Figure 3. In contrast, glycerol content did not have a con-
ment on an appropriate definition of ‘fullness’. In partic-
sistent or significant effect on the heat elicited by any of
ular, different schools of thought exist as to whether full-
the wines. However, at the lowest alcohol level, glycerol
ness is a single sensory attribute or whether it is more
seemed to have a slight depressive effect on alcohol
abstract in nature. Langstaff et al. (1991a) have suggested
elicited palate heat. Neither sweetness, aroma or flavour
that viscosity and density are components of fullness in
intensity were affected by the addition of either glycerol or
beer, a classification later extended to wine (Pickering et
ethanol (Table 2).

Table 4. Weightings of predictor variables of white wine ‘body’ by assessor.

Assessor 1 2 3 4 5 6 7 8 9 10

Perceived viscosity +++ + +++ + +


(0.37) (0.65) (0.44) (0.60) (0.58)
Flavour ++ ++ + +++ ++
(0.49) (0.40) (0.62) (0.52) (0.57)
Hotness

Acidity +++ – –– – –– +++


(0.15) (1.87) (1.68) (1.21) (2.06) (0.49)
Sweetness ++ +++
(0.32) (0.54)
P (model fit) 0.000 0.057 0.016 0.006 0.000 0.027 0.060 0.072 0.061 0.023

+ + + , + + , + indicates significant positive impact on body at the 1, 5 and 10% level respectively.
– – , – indicates significant negative impact on body at the 5 and 10% level respectively.
Odds ratio of rating as low bodied compared with rating medium or high bodied are given in brackets.

Table 5. Spearman rank correlation coefficients between attributes for all wines.
Body Acidity Flavour Hotness Viscosity Sweetness

Body 1.00
Acidity 0.12 ns 1.00
Flavour 0.28 ** 0.11 ns 1.00
Hotness 0.15 # 0.01 ns 0.16 # 1.00
Viscosity 0.42 *** 0.09 ns 0.25 ** 0.40 *** 1.00
Sweetness –0.07 ns –0.28 ** 0.10 ns 0.33 *** 0.01 ns 1.00

#, *, **, *** indicates significance at 10%, 5%, 1% and 0.1% respectively, ns= not significant.
Gawel, van Sluyter & Waters Sensory effects of ethanol and glycerol in Riesling wines 43

al. 1998, Nurgel and Pickering, 2005). Viscosity was Effect of alcohol and glycerol
defined by Langstaff et al. (1991a) as ‘the degree to which The increased body with increasing alcohol levels seen in
the beer resists flow under an applied force in the mouth’, two of the three wines (Table 3 and Figure 1, wines B and
while ‘density’ was defined as the ‘perceived density or C) is consistent with the long held premise that body in
weight of beer in the mouth’. However, various observa- white wine is influenced by alcohol content. Nurgel and
tions suggest that there is some redundancy in these Pickering (2005) also reported increases in density (a term
terms. Using these definitions, Langstaff et al. (1991b) equivalent to the term body used here) with increased
found that the perceived viscosity and density of beer was alcohol content over the same range of alcohol levels.
highly correlated to physical viscosity, and also to each However, as was observed here, commercially significant
other. Nurgel and Pickering (2005) applied the Langstaff differences in alcohol content produced relatively small
et al. (1991a) definitions of density and viscosity to model changes in density when white wine was used as a base
wine solutions with varying alcohol and glycerol levels. and flavour effects were removed (Pickering et al. 1998).
Their results indicated that the perceived changes in the Here, the differences in mean ratings of body resulting
intensity of these attributes were also highly correlated. from a 2% alcohol increase were around a third of a point
It is also noteworthy that they used the same sensory on a nine point scale. This suggests that within the range
standard to represent both viscosity and density further of 11.7 to 13.7% v/v, alcohol alone only has a small effect
suggesting a lack of orthogonality between these attrib- on white wine body. This seems to contradict the com-
utes. However, this probable lack of orthogonality does monly held view that alcohol level is important to fullness
not necessarily imply that palate density and viscosity are of white wine. Wines with high alcohol levels are often
completely synonymous. Another possibility is that high more flavoursome, possibly a result of them being pro-
viscosity is a necessary condition for denseness in wine, duced from riper grapes. It is conceivable therefore that
but that other sensory attributes such as flavour also con- perceptions regarding the role of alcohol may have been
tribute to it. influenced by the more intense flavours typically found in
The term ‘density’ (Langstaff et al. 1991a) is equivalent wines of higher alcoholic strength. Langstaff et al. (1991b)
to the broad definition of ‘body’ decided upon by the also noted that the fullness of commercial beers were only
assessors in this study. A question arises. Is ‘body’, a con- moderately correlated with their alcohol content suggest-
crete or an abstract sensory attribute? A concrete sensory ing that factors other than alcohol were likely to have
attribute has been defined as one that can be clearly illus- influenced beer fullness.
trated using a single reference standard, while abstract The main effect of alcohol on perceived viscosity was
attributes cannot be adequately illustrated by any single or not statistically significant for any of the three wines (Table
set of reference standards due to their multidimensional 2). However, Figure 2 shows a consistent increase in per-
nature (Gawel 1997). The assessors used in this study ceived viscosity with increasing alcohol for wines B and C.
agreed that body was not singular in nature. Langstaff et Nurgel and Pickering (2005) reported monotonic increases
al. (1991a) found that the ‘taste’ standards that best illus- in perceived viscosity of model wine with increasing alco-
trated the different mouthfeel sensations in beer including hol level over the alcohol range of 0 to 15%. A positive
palate density were different stylistic examples of beers effect of alcohol on the perceived viscosity of white wine
themselves. This outcome also suggests that body is an in the range of 7–14% v/v has also been reported
abstract rather than concrete sensory attribute. (Pickering et al. 1998).
Clapperton (1973) considered ‘body’ in beer to Figure 2 shows that at 13.6% v/v ethanol, increasing
include ‘flavour fullness’. This notion equates well with glycerol concentration from 5.2 to 10.2 g/L produced a
the views of assessors used in this study regarding the small but significant increase in perceived viscosity in two
nature of white wine body. During the initial discussion of the three wines. Noble and Bursick (1984) estimated
session they suggested that flavour intensity was an that a glycerol addition in the order of 26 g/L is required
important feature of wine body. This assertion was vindi- to elicit a just perceptible increase in the viscosity of a
cated in part by the observed moderate but significant light-bodied dry wine containing 4.8 g/L glycerol. Using
association between body and flavour (Table 5), and the the Weber fraction calculated from the results of Noble
low odds ratios of the flavour coefficient when it was and Bursick (1984), the difference threshold for perceived
regressed against body ratings (Table 4). viscosity in the Riesling wines in the present study with a
The assessors agreed upon a definition of viscosity base concentration of 5.2 g/L should be in the order of 28
which was effectively identical to that of Langstaff et al. g/L. The greatest difference in glycerol concentrations in
(1991a) for beer. Initial discussions indicated that most of the present study was 5 g/L. Therefore, the data of Noble
the assessors considered viscosity to be a singular sensory and Bursick (1984) would suggest that any viscosity
attribute and one of the components of white wine body. differences resulting from the addition of glycerol should
The highest observed correlation with body ratings were be undetectable in these wines. One possible reason for
with those of perceived viscosity (Table 5) which was con- this apparent discrepancy is that the addition of glycerol
sistent with the views expressed in the initial discussions. made the Riesling wines perceptively sweeter, which some
Other researchers have also found strong correlations assessors may have associated with viscosity. However as
between perceived density and viscosity in beer (Langstaff the differences in sweetness produced by glycerol addition
et al. 1991b), white wine (Pickering et al. 1998) and were insignificant (Table 2), and sweetness and perceived
model wine (Nurgel and Pickering, 2005). viscosity were uncorrelated (Table 5), it is unlikely that the
44 Sensory effects of ethanol and glycerol in Riesling wines Australian Journal of Grape and Wine Research 13, 38–45, 2007

differences in perceived viscosity were due to confound- was evaluated by Clapperton (1974). He found that the
ing with sweetness. addition of the buttery compound diacetyl increased the
It is shown here that the effect of glycerol on viscosity body of ale, but decreased the body of lager beer. He
is dependent on the wine to which it was added. The attributed this to the fact that the characteristic flavour of
effect of glycerol addition was greater in wine C than of ale was retained more strongly than that of lager follow-
either wine A or B (Figure 2). All the wines had similar ing the addition. This suggests that the type of flavour may
sugar content and identical alcohol levels (Table 1), so the be important in fullness evaluation. That is, the presence
differential effects of glycerol addition on the wines of typical or expected flavours may increase fullness, while
remains unexplained. a greater intensity of an atypical flavour may not.
Alcohol level significantly increased the palate heat of While perceived viscosity and flavour intensity appeared
all three wines (Table 2). While the palate warming effect to be the most consistent predictors of body, other con-
of alcohol is well known, it appears that little research has tributors to fullness perception were idiosyncratic. Higher
been conducted into the chemesthetic aspect of alcohol. acidity ratings by four assessors were associated with lower
Clapperton (1974) reported that his subjects described body, while for another two assessors, increased acidity
alcoholic solutions as both drying and warming. However was associated with higher body. A possible reason for
the ability of assessors to detect different levels of hotness these assessor differences is that some may have inter-
due to alcohol was not tested. preted high acidity as contributing to a varietal citrus
There was some evidence that glycerol suppressed the flavour and hence body, or that the positive role of acid-
alcoholic heat at the lower alcohol level (Figure 3). Berg ity on body may have simply been a previously learned
et al. (1955b) also noted that sucrose sweetness tended to response.
raise the alcohol difference thresholds in water, indicating Higher sweetness appeared to be a factor in the per-
a form of sweetness suppression on alcohol perception. ception of body by two assessors. Although sweetness dif-
However, these authors did not indicate what specific ferences in the wines would be expected to be small
aspect of alcohol perception was affected by sweetness. because the differences in the wine’s sugar and glycerol
Alcohol-elicited sweetness would be more difficult to levels were sub-threshold (Berg et al. 1955a, Noble and
detect in the presence of sucrose, while alcohol-elicited Bursick 1984), it is also possible that they had an additive
bitterness would be suppressed by the sweetness of effect on overall sweetness perception. The apparent rela-
sucrose. Both explain the raised alcohol detection thresh- tionship between sweetness and body could conceivably
old. However, a suppressive effect of carbohydrate vis- have resulted from these assessors perceptually associat-
cosity on palate heat is also plausible as it is known that ing sweetness perceived in the wines with fullness. This
they partly share the same somatosensory pathway (Rolls association could result from the assessors’ previous expe-
et al. 2003). Further work needs to be done to elucidate riences of sweeter beverages being more physically viscous
the effect of both tastes and textures on alcohol-induced due to their sugar content. Some assessors may also have
hotness in wine. associated sweetness with flavour. Hort and Hollowood
Adding the sweet tasting alcohol and glycerol (Scinska (2004) found that for all but their most experienced asses-
et al. 2000) did not increase the perceived sweetness of sors, sucrose sweetness was a key driver of fruit flavour
any wine. It is plausible that the contribution to sweetness intensity of flavoured aqueous solutions. Presumably this
by these substances was effectively suppressed by the high association derives from our common experiences of con-
acidity of these wines. Similarly, neither aroma nor flavour currently experiencing fruit flavours and sweetness when
intensity was consistently affected by either glycerol or consuming ripe fruits. There is some evidence of this
ethanol level. Across a wider concentration range than association, with small increases in mean flavour intensity
used here, ethanol was shown to suppress the perceived being noted with increasing glycerol concentration at all
aroma and flavour intensity of wine volatiles (Guth and alcohol levels (data not shown). In fact, some winemakers
Sies 2002). However, consistent with the result here, believe that residual sweetness in white wine can
ethanol levels slightly above those used here did not affect enhance the impression of its fruitiness and fullness. While
headspace volatilities (Conner et al. 1998, Escalona et al. experienced and trained tasters can easily separate
1999). flavour sweetness from sugar sweetness, many naive
tasters do not.
Assessor interpretation of body Lastly, the association between sweetness and body
Table 4 shows that for half the assessors, both flavour and may be the result of a common response. That is, sweet-
perceived viscosity were positively associated with body. ness and body may have responded to changes in some
The small odds ratios imply large increases in the proba- other variable. Sweetness was significantly correlated with
bility of a wine being rated in a higher body class with hotness (Table 5), a character which was almost certainly
each unit increase in perceived flavour or viscosity (Table due to alcohol (Figure 3). As ethanol is itself sweet
4). That is, for half the assessors, perceived viscosity and (Scinska et al. 2000), it is likely that increased ethanol was
flavour were important components of body. It is worth the cause of the positive relationship between body and
noting that any difference in flavour perception was due sweetness.
to the influence of the addition of alcohol and glycerol as
the concentrations of wine volatiles was the same across Conclusion
all treatments. The role of flavour in fullness perception Ethanol and glycerol levels in realistic ranges had a small
Gawel, van Sluyter & Waters Sensory effects of ethanol and glycerol in Riesling wines 45

but inconsistent positive effect on the body and viscosity Escalona, H., Piggott, J.R., Conner, J.M. and Paterson, A. (1999) Effect
of Riesling wines. The perceived hotness of the wines was of ethanol strength on the volatility of higher alcohols and aldehy-
des. Italian Journal of Food Science 11, 241–248.
strongly influenced by alcohol level, while sweetness,
Gawel, R. (1997) The use of language by trained and untrained expe-
flavour and aroma intensity were relatively unaffected by rienced wine tasters. Journal of Sensory Studies 12, 267–284.
either glycerol or alcohol. Assessors given a broad defini- Guth, H. and Sies, A. (2002) Flavour of wines: towards an under-
tion of wine fullness were idiosyncratic with regard to standing by reconstitution experiments and an analysis of ethanol’s
what features of the wine contributed to its fullness. effect on odour activity of key components. Proceedings of the
Eleventh Australian Wine Industry Technical Conference, 7–11
However, flavour and perceived viscosity were most fre-
October 2001, Adelaide, Australia (Australian Wine Industry Tech-
quently and most strongly correlated with the fullness of nical Conference Inc.: Adelaide, SA) pp. 128–139.
these wines. Hort, J. and Hollowood, T.A. (2004) Controlled continuous flow deliv-
ery system for investigating taste-aroma interactions. Journal of
Acknowledgements Agricultural and Food Chemistry 52, 4834–4843.
Langstaff, S.A., Guinard, J.-X. and Lewis, M.J. (1991a) Sensory eval-
This project was supported by Australia’s grapegrowers
uation of the mouthfeel of beer. American Society of Brewing
and winemakers through their investment body, the Chemists Journal 49, 54–59.
Grape and Wine Research and Development Corporation, Langstaff, S.A., Guinard, J.-X. and Lewis, M.J. (1991b) Instrumental
with matching funds from the Federal Government, and evaluation of the mouthfeel of beer and correlation with sensory
by the Cooperative Research Centre for Viticulture. We evaluation. Journal of the Institute of Brewing 97, 427–433.
Noble,A. C. and Bursick, G. F. (1984) The contribution of glycerol to
also thank the tasters for contributing their time and skill,
perceived viscosity and sweetness in white wine. American Journal
and Dr Mark Sefton and Dr Leigh Francis for critically of Enology and Viticulture 35, 110–112.
reviewing the manuscript. Nieuwoudt, H.H., Prior, B.A., Pretorius, I.S. and Bauer, F.F. (2002)
Glycerol in South African table wines: an assessment of its rela-
tionship to wine quality. South African Journal of Enology and
References Viticulture 23, 22–30.
Amerine, M.A. and Roessler, E.B. (1983) Wines: Their Sensory Nurgel, C. and Pickering, G. (2005) Contribution of glycerol, ethanol
Evaluation. (W.H. Freeman & Co: San Francisco). and sugar to the perception of viscosity and density elicited by model
Berg, H. W., Filipello, F., Hinreiner, E. and Webb, A.D. (1955a) white wines. Journal of Texture Studies 36, 303–325.
Evaluation of thresholds and minimum difference concentrations for Pickering, G.J., Heatherbell, D.A., Vanhanen, L.P. and Barnes, M.F.
various constituents of wines I. Water solutions of pure substances. (1998) The effect of ethanol concentration on the temporal per-
Food Technology 9, 23–26. ception of viscosity and density in white wine. American Journal of
Berg, H. W., Filipello, F., Hinreiner, E. and Webb, A.D. (1955b) Enology and Viticulture 49, 306–318.
Evaluation of thresholds and minimum difference concentrations Rankine, B.C. and Bidson, D.A. (1971) Glycerol in Australian wines
for various constituents of wines. II. Sweetness: the effect of ethyl and factors influencing its formation. American Journal of Enology
alcohol, organic acids and tannin. Food Technology 9, 138–140. and Viticulture 22, 6–12.
Clapperton, J.R. (1974) Profile analysis and flavour discrimination in Rolls, E.T., Verhagen, J.V. and Kadohisa, M. (2003) Representations of
beer. Journal of the Institute of Brewing 80, 164–173. the texture of food in the primate orbitofrontal cortex: Neurons
Clapperton J.R. (1973) Derivation of a profile method for sensory responding to viscosity, grittiness, and capsaicin. Journal of
analysis of beer flavour. Journal of the Institute of Brewing 79, Neurophysiology 90, 3711–3724.
495–508. Scinska, A., Koros, E., Habrat, B., Kukwa, A., Kostowski, W. and
Conner, J.M., Birkmyre, L., Paterson, A. and Piggott, J.R. (1998) Bienkowski, P. (2000) Bitter and sweet components of ethanol taste
Headspace concentrations of ethyl esters at different alcoholic in humans. Drug and Alcohol Dependence 60, 199–206.
strengths. Journal of the Science of Food and Agriculture 77,
121–126.
Delwiche, J. (2004) The impact of perceptual interactions on perceived
Manuscript received: 17 July 2006
flavor. Food Quality and Preference 15, 137–146. Revised manuscript received: 16 January 2997
Chapter 3.

The effect of glycerol on the perceived viscosity of dry white


table wine

Richard Gawel and Elizabeth J. Waters

The Australian Wine Research Institute, PO Box 197, Glen Osmond, 5064, Australia

Journal of Wine Research 2008; 19: 109-114.


DOI:10.1080/09571260802622191

57
Journal of Wine Research, 2008, Vol. 19, No. 2,
pp. 109 – 114

The Effect of Glycerol on the Perceived Viscosity


of Dry White Table Wine

RICHARD GAWEL and ELIZABETH J. WATERS


Original manuscript received, 1 June 2008
Revised manuscript received, 16 October 2008

ABSTRACT The effect of glycerol on the perceived viscosity of dry white wines was investigated using
direct paired comparison sensory methods. Before tasters assessed wines for viscosity, the natural
sweetness of glycerol needed to first be masked. This was achieved using two novel methods—prior
oral exposure to the anti-sweetness agent Gymnema sylvestre, and by sweetness equalisation using
a non-viscous high potency artificial sweetener. After masking its sweetness, the addition of 6 g/L of
glycerol did not increase the perceived oral viscosity of dry white wine suggesting that glycerol does
not play a role in the perception of dry white wine viscosity. Therefore, palate viscosity in dry white
wine cannot be enhanced by employing traditional winemaking approaches that elevate glycerol levels.

Introduction
Glycerol is the most abundant compound in dry white table wines after water and ethyl
alcohol. Around 5 to 10 g/L of glycerol is produced by yeast during primary fermenta-
tion, with the concentration depending on yeast strain and fermentation temperature
(Nieuwoudt et al., 2002; Erasmus et al., 2004). Pure glycerol is around 600 times more
viscous than water (Aguirre et al., 1989) which explains why wines supplemented with
glycerol in the laboratory become physically more viscous (Noble and Bursick, 1984).
Although increased glycerol levels in white wine predictably make them more viscous,
there is debate whether these increases are significant enough to be perceptible to wine
tasters. The perceived viscosities of model wines that differed in glycerol concentration
by 5 g/L are differentiable (Nurgel and Pickering, 2005). However, varying the glycerol
concentration within the wine realistic range of 5.2 to 10.2 g/L did not significantly affect
the perceived viscosity of two dry Riesling wines, and only marginally increased the per-
ceived viscosity of a third wine (Gawel et al., 2007). Similarly, the difference threshold of
glycerol in white wine was estimated by Noble and Bursick (1984) to be around 26 g/L
suggesting that glycerol induced increases in the physical viscosity of wine are largely
imperceptible. However, it is worth noting that this threshold value was not obtained

Richard Gawel, The Australian Wine Research Institute, PO Box 197, Glen Osmond, SA 5064, Australia
(E-mail: richard.gawel@awri.com.au).

ISSN 0957-1264 print/ISSN 1469-9672 online/08/020109-6 # 2008 Taylor & Francis


DOI: 10.1080/09571260802622191
110 RICHARD GAWEL AND ELIZABETH J. WATERS

by directly comparing wines with different glycerol concentrations, but by adding a taste-
less food gum to a wine in order to determine the difference in physical viscosity that was
just noticeable to tasters. From this Noble and Bursick determined the change in glycerol
concentration that was required to produce an equivalent change in physical viscosity.
Presumably the wines were not directly compared as glycerol’s distinct sweetness would
have acted as a cue biasing their experimental outcomes.
To make direct comparisons of perceived viscosity in wines of different glycerol levels
possible, glycerol’s sweetness first needs to be suppressed. One possible strategy involves
prior oral exposure to the leaf extract of the sub-tropical climber, Gymnema sylvestre
(GS). Tasting this ‘tea’ containing the anti-saccharine agent gymnemic acid (GA) sig-
nificantly reduces the sweetness of both carbohydrate and non-carbohydrate sweet-
eners (Stocklin, 1969; Warren et al., 1969). The ability of GA to block the sweetness
of a wide variety of structurally diverse sweeteners suggests that it functions as an
antagonist at the receptor level. Consequently, GA is also likely to block glycerol sweet-
ness. If this is the case, it would allow glycerol induced changes to oral viscosity to be
assessed using direct forced choice sensory methods. An alternative approach to using
GS involves modifying the sweetness of all the samples being compared so as to break
the nexus between glycerol concentration and sweetness. This could be achieved by
supplementing both the samples being compared in a forced choice procedure with a
non-viscous high potency artificial sweetener. The physical viscosity of the samples
would not be affected by the addition, but the ability of the tasters to associate per-
ceived sweetness with glycerol level would be severely hindered.
We aimed to determine whether differences in the glycerol concentration within the
range typically encountered in dry white wine were sufficient to induce perceptible
increases in oral viscosity. As a forced choice test method was used, it was necessary
to mask the natural sweetness of glycerol using the two abovementioned approaches.

Materials and Methods


General
All tastings were conducted in white booths using black ISO tasting glasses. The wines
were prepared two hours prior to the tasting and were served at room temperature
(238C + 28C). The volume of sample tasted was not standardised as it was assumed
that each individual taster would sample a similar volume of control and treated
wine when comparing samples. The Gymnema sylvestra extract (Medicine Garden
Australia) was prepared two hours prior to the tasting by adding 50 g of dried and
powdered Gymnema sylvestra leaves to 1 L of boiling Milli-Q water, cooling to room
temperature and filtering using filter paper. Table 1 gives the alcohol, glycerol and
residual sugar concentrations of the Riesling wines used (Gawel et al., 2007).

Table 1. Composition of the base wines used


Wine Glycerol (g/L) Alcohol (% v/v) Sugars (g/L)

A 5.3 12.0 1.1


B 6.2 13.2 0.7
C 6.0 12.5 2.8
D 6.0 11.9 1.6
THE EFFECT OF GLYCEROL 111

Sweetness Blocking Using Gymnema sylvestre Extract


Ten volunteer tasters (five males, five females, average age 38) were first screened for a
positive anti-sweetness response to GS. Tasters rinsed their mouths for 15 seconds with
GS extract followed by water. They then immediately tasted three wines comprising a
dry white wine with 10 g/L added glycerol and two controls and indicated which was
the sweetest. These steps were repeated to give triplicate presentations. A positive
response to GS was considered to occur if the taster 1) failed to identify the glycerol
sample as the sweetest in at least two of three sets, or 2) identified it in two sets but
also indicated that they were only guessing. Based on these criteria, five males and
three females (average age 37) were retained. All had difference testing experience
and five had previously been involved in tasting trials involving wine viscosity. The
tasters were presented with triplicate directional difference tests; a dry Riesling white
wine (Wine A, Table 1) and the same wine with 6 g/L added glycerol. The assessors
rinsed their mouths with the GS extract for 15 seconds followed by water. They
waited another 15 seconds before tasting the randomised pair and stating which
sample they perceived as the most viscous. Specifically, the tasters were asked to
taste each pair of samples and state which of the pair required a greater amount of
force to move the wine around in the mouth.

Sweetness Modification Using Saccharine


Thirty tasters experienced in directional difference testing were presented with three
paired comparisons in a single session. Each pair comprised a different dry Riesling
white wine (Wines B, C and D, Table 1) and the same wine with 6 g/L of added gly-
cerol. Some 80 mM/L of saccharine (Hermesetas, Switzerland) was added to the
control to make it approximately equi-sweet to the glycerol treatment. This concen-
tration was chosen on the advice of three experienced wine tasters following informal
bench-top tasting. Smaller amounts of saccharine (ranging from 16 to 20 mM/L)
were also randomly added to the glycerol treatment to further mask differences in
sweetness intensity and persistence so as to further confuse the tasters. The tasters
selected the most orally viscous wine of each pair. The serving order of the pairs was
balanced across the panel, and the position of the glycerol modified sample was
randomised within each pair. The data were analysed by the binomial test.

Results
Assessing the Perceived Viscosity of Glycerol Following Blocking Its Sweetness
Using Gymnema sylvestre
The glycerol treatment was judged to be more viscous in only 11 of the 24 presenta-
tions. Therefore, there was insufficient evidence to suggest that the addition of 6 g/L
of glycerol significantly increased the oral viscosity of this wine (p ¼ 0.839).

Assessing the Perceived Viscosity of Glycerol Following Modifying Its Sweetness


by Adding Saccharin
The results of the directional difference tests are given in Table 2. There was insufficient
evidence to suggest that an additional 6 g/L of glycerol resulted in a significant increase
in orally perceived viscosity of any of the three wines used.
112 RICHARD GAWEL AND ELIZABETH J. WATERS

Table 2. Number of tasters indicating the wine higher in


perceived viscositya
Wine Control Wine Control þ 6 g/L Glycerol Significance

B 20 10 ns
C 15 15 ns
D 16 14 ns

Note: ns: not significant at 5% level.


a
Sweetness of the wines modified by saccharine addition. See Methods and Materials.

Discussion
To our knowledge this is the first time that the effect of glycerol on the perceived
viscosity of wine has been determined directly using forced choice sensory methods.
Paired comparison tests have an advantage over other sensory methods in that they
are scale free and the samples being presented act as their own frame of reference.
GS was shown to be an effective masking agent of glycerol sweetness for the majority
of tasters used here. However, as its effectiveness was not uniform across tasters,
prior screening of panellists was necessary. The GS extract also had an unpleasant
bitter taste which necessitated that the tasters rinsed their mouths with water and
waited for the bitterness to subside. However, the persistent effect of GS was found
to be advantageous as tasters could condition their mouths before subsequently
tasting both samples in quick succession. Saccharine addition was a simple and effective
way of modifying the sweetness of the samples. While the additions were determined
informally using bench-top tasting, the fact that the samples were approximately
equi-sweet and varied in an unpredictable fashion appeared to be sufficient to break
the association between glycerol concentration and sample sweetness. Furthermore,
there was no evidence that saccharine contributed bitterness to any of the samples.
This may have been due to the low concentration of saccharine used, or through
masking by other tastes present in the wine.
Despite its viscosity when in its pure form, glycerol does not appear to affect the
perceived viscosity of dry white wine (Table 2). Other findings support this result.
Recently we found that increasing the glycerol levels of Riesling wines by 5 g/L mar-
ginally increased the oral viscosity of only one of three wines, and this effect was seen
only when alcohol levels were elevated by alcohol addition (Gawel et al., 2007). Here,
increased glycerol did not enhance the perceived viscosity of any of the three wines
regardless of the fact that they varied in alcohol content by 2% v/v. Similarly,
Nurgel and Pickering (2005) found that statistically significant increases in the
perceived viscosity of model wine only occurred at glycerol concentrations above
25 g/L. Noble and Bursick (1984) estimated that a difference in white wine viscosity
of 0.14 mPa s (equating to 25.7 g/L of glycerol) was just detectable. Perceived oral
viscosity has been demonstrated to be proportional to the third root of its physical
viscosity (Christensen and Casper, 1987). When this relationship is applied to the
physical viscosity data of Noble and Bursick (1984), we estimate that the perceived
viscosity of white wine with a glycerol concentration at the high end of the range
found in that wine type (12.5 g/L) would only be 1% higher than the perceived
viscosity of an equivalent wine that is low in glycerol (5 g/L). Clearly then, differ-
ences in oral viscosity would be hard to perceive in the 5 to 10 g/L range typical of
these wines.
THE EFFECT OF GLYCEROL 113

The inability of tasters to detect any change to the oral viscosity of dry white wines
resulting from increased glycerol may simply reflect the insignificant effect that glycerol
has on the physical viscosity of wine. The physical viscosity of the wines used in this
study was not measured. However, others have shown that increasing glycerol in
both white wine (Noble and Bursick, 1984; Nurgel and Pickering, 2005) and in
water (Yanniotis et al., 2007) results in only minor increases in their physical viscosity.
Lastly, inspection of the data of Kosmerl et al. (2000), revealed that while alcohol and
residual sugar significantly contributed to the physical viscosity of commercial dry
wines (i.e. those with less than 7 g/L sugar), glycerol did not.
Significantly increasing the physical viscosity of the matrix has been shown to reduce
flavour intensity (Hollowood et al., 2002). However, Gawel et al. (2007) recently
showed that modifying the glycerol level of Riesling wines within a wine realistic
range did not affect the perceived flavour or other taste characteristics of the wines.
This null result was probably due to the fact that when present at concentrations typi-
cally found in dry wines, glycerol does not significantly alter the physical viscosity of the
system.
In conclusion, glycerol does not appear to contribute to the perceived oral viscosity
of dry white table wines when found in the concentration range typically encountered
in this wine style.

Acknowledgements
This project was supported by Australia’s grapegrowers and winemakers through their
investment body, the Grape and Wine Research and Development Corporation, with
matching funds from the Federal Government. We also thank the Brooke Travis, Leigh
Francis and Belinda Bramley for conducting the sensory trials, and the tasters for
volunteering their time and skill.

References
AGUIRRE , A., MENDOZA , B., LEVINE , M.J., HATTON , M.N. and DOUGLAS , W.H. (1989) In
vitro characterization of human salivary lubrication, Archives of Oral Biology, 34, 675– 677.
CHRISTENSEN , C.M. and CASPER , L.M. (1987) Oral and nonoral perception of solution
viscosity, Journal of Food Science, 52, 445– 447.
ERASMUS , D.J., CLIFF , M. and VAN VUUREN , H.J.J. (2004) Impact of yeast strain on the
production of acetic acid, glycerol, and the sensory attributes of icewine, American
Journal of Enology and Viticulture, 55, 371– 378.
GAWEL , R., VAN SLUYTER , S. and WATERS , E.J. (2007) The effects of ethanol and glycerol on
the body and other sensory characteristics of Riesling wines, Australian Journal of Grape and
Wine Research, 13, 38 – 45.
HOLLOWOOD , T.A., LINFORTH , R.S.T. and TAYLOR , A.J. (2002) The effect of viscosity on the
perception of flavour, Chemical Senses, 27, 583– 591.
KOSMERL , T., ABRAMOVIC , H. and KLOFUTAR , C. (2000) The rheological properties of
Slovenian wines, Journal of Food Engineering, 46, 165– 171.
NIEUWOUDT , H.H., PRIOR , B.A., PRETORIUS , I.S. and BAUER , F.F. (2002) Glycerol in South
African table wines. An assessment of its relationship to wine quality, South African Journal
of Enology and Viticulture, 23, 22 – 30.
NOBLE , A.C. and BURSICK , G.F. (1984) The contribution of glycerol to perceived viscosity and
sweetness in white wine, American Journal of Enology and Viticulture, 35, 110.
NURGEL , C. and PICKERING , G. (2005) Contribution of glycerol, ethanol and sugar to the
perception of viscosity and density elicited by model white wines, Journal of Texture
Studies, 36, 303– 325.
114 RICHARD GAWEL AND ELIZABETH J. WATERS

STOCKLIN , W. (1969) Chemistry and physiological properties of Gymnemic Acid, the antisac-
charine principle of the leaves of Gymnema sylvestre, Journal of Agricultural and Food Chemistry,
17, 704– 708.
WARREN , R.P., WARREN , R.M. and WENINGER , M.G. (1969) Inhibition of the sweet taste by
Gymnema sylvestre, Nature, 223, 94 – 95.
YANNIOTIS , S., KOTSERIDIS , G., ORFANIDOU , A. and PETRAKI , A. (2007) Effect of ethanol,
dry extract and glycerol on the viscosity of wine, Journal of Food Engineering, 81, 399– 403.
Chapter 4.

Effect of pH and alcohol on perception of phenolic character in


white wine

Richard Gawel, Steven C. Van Sluyter, Paul A. Smith and Elizabeth J. Waters

The Australian Wine Research Institute, PO Box 197, Glen Osmond, 5064, Australia

American Journal of Enology and Viticulture – 2013; 64: 425-429.


DOI: 10.5344/ajev.2013.13016

65
Effect of pH and Alcohol on Perception of
Phenolic Character in White Wine
Richard Gawel,1* Steven C. Van Sluyter,1,2 Paul A. Smith,1
and Elizabeth J. Waters1,3

Abstract: The in-mouth perception of textures of white wine arising from the interactions among white wine
phenolics, pH, and alcohol level was evaluated. Phenolics were extracted from white wines and added back to white
wines that were adjusted to different pH and ethanol concentrations within wine realistic ranges. Adding phenolics to
a white wine at pH 3.3 significantly increased its astringency, but the same addition did not contribute to the higher
astringency elicited by the same wine when adjusted to pH 3.0. Higher phenolics generally increased bitterness and
viscosity, but the effect depended on the source of the phenolics. Wines with added phenolics were generally perceived
to be hotter, and significantly so when the wine was low in alcohol. The combined effect of phenolic content and
alcohol concentration on astringency and bitterness was additive, suggesting that alcohol directly contributes to these
attributes in white wines. Overall, the tastes and textures produced by white wine phenolics were more pronounced
in wines with lower alcohol levels.
Key words: white wine phenolics, astringency, bitterness, hotness, viscosity

The conventional New World approach to white wine- presence of phenolics, such as bitterness, perceived viscosity,
making typically involves using techniques and processes and drying sensations. Aqueous ethanol is bitter at winelike
that minimize the concentration of phenolic compounds in concentrations (Scinska et al. 2000), and ethanol has been
the finished wine. The acceptance of grape processing and shown to elicit metallic tastes and palate roughness in model
winemaking strategies that limit phenolic concentration prob- wine systems devoid of phenolics (Jones et al. 2008). The
ably arose from the notion that their presence masks varietal perceived viscosity of white wine increased with the addition
flavors, accelerates oxidation, and can result in palate hard- of ethanol (Gawel et al. 2007a), and in-mouth sensations of
ness. Despite this thinking, many winemakers have recently dryness have been associated with low pH in model systems
demonstrated a greater willingness to incorporate phenolics (Rubico and McDaniel 1992, Lawless et al. 1996).
into their white wines, referring to what they perceive as Studies on the influence of alcohol and acidity on the per-
greater fullness and enhanced palate texture. However, few ception of phenolic character have been confined to red wine,
studies have explored the role of white wine phenolics in and specifically to their effect on astringency (Kallithraka et
white wine taste and texture or how alcohol and acidity influ- al. 1997, Fontoin et al. 2008). However, it is unclear whether
ence these perceptions. Early work demonstrated that white these interactions also apply to the tastes and textures in
wines vinified in ways that produced higher phenolics were white wines, as white wines are typically lower in pH and
more bitter and astringent (Singleton et al. 1975). However, alcohol and have different phenolic profiles than those of red
higher phenolic concentrations in white wines also enhance wines (Waterhouse 2002). This work investigated how white
palate fullness in white wines (Cejudo-Bastante et al. 2011). wine phenolics contribute to the perceived textures and tastes
Alcohol and acidity are known to contribute directly to the of white wine and how alcohol and pH when in typical white
tastes and mouthfeel attributes normally associated with the wine ranges influence these perceptions. This was achieved
by examining the contribution of phenolics taken from a
single white wine to its astringency, bitterness, and hotness
1
Australian Wine Research Institute, P.O. Box 197, Glen Osmond, Adelaide, and the influence of pH and ethanol concentration on these
SA 5064, Australia; 2(currently) Macquarie University, North Ryde, NSW
2109, Australia; and 3(currently) Grape and Wine Research and Development parameters. These results were generalized by assessing the
Corporation, P.O. Box 610, Kent Town, SA 5071, Australia. effect on white wine mouthfeel of phenolics taken from three
*Corresponding author (email: richard.gawel@awri.com.au) wines (Fiano, Viognier, and Gewürztraminer) displaying dif-
Acknowledgments: The work was funded by Australia’s grapegrowers and ferent phenolic profiles.
winemakers through their investment body, the Grape and Wine Research
Development Corporation, with matching funds from the Australian govern- Materials and Methods
ment and Pernod Ricard Pacific. The AWRI is a member of the Wine Industry
Cluster, located in the Waite precinct. Collection and analysis of phenolic extracts. Four liters of
Publication costs of this article defrayed in part by page fees. wine was slowly introduced to the top reservoir of a 1 m x 2.5
Manuscript submitted Jan 2013, revised May 2013, accepted Jul 2013 cm column packed with Amberlite FPX-66 resin (Rhom and
Copyright © 2013 by the American Society for Enology and Viticulture. All Haas, Midland, MI) and allowed to flow through under grav-
rights reserved. ity. The resin was washed with 2 column volumes of MilliQ
doi: 10.5344/ajev.2013.13016 water before being exhaustively eluted with 96% v/v ethanol
425
Am. J. Enol. Vitic. 64:4 (2013)
426 – Gawel et al.

under gravity. The ethanol was removed under vacuum at at room temperature (22°C ± 1°C) and under amber lighting.
30°C, after which 50 mL of MilliQ water was added and the Perceived astringency, bitterness, hotness, and viscosity were
sample freeze-dried and stored at -80°C before immediate use. rated on a 15 cm partially structured line scale with the word
The phenolic composition of the extracts and the wines from anchors “low” and “high” at 1.5 cm and 13.5 cm, respectively.
which the extracts were taken was determined by HPLC using Panelists rinsed their mouths with water and waited 30 sec be-
a method described elsewhere (Smith and Waters 2012). The fore tasting the following sample. The samples were presented
peak identities were confirmed by retention time of known in triplicate over two tasting sessions conducted one week
standards and by liquid chromatography-mass spectrometry. apart. An entire set of treatment combinations was randomly
Study 1: Effect of pH, alcohol, and phenolics on bit- presented each week, with the third replicate randomly split
terness, astringency, and hotness. Sample preparation. A across the two tasting sessions.
one-year-old Riesling (Lengs and Cooter, Watervale, South Study 2: Effect of alcohol level and different phenolics
Australia) was selected as the base wine due to its low alcohol on mouthfeel. Sample preparation. Two base wines were
concentration (11.4% v/v), moderate pH (3.3), and low residual used for study 2: a one-year-old Riesling (Lengs and Cooter)
sugar (1.3 g/L). A 3 x 2 full factorial design was applied: with 12.5% v/v alcohol and an unoaked Chardonnay (Chapel
two ethanol levels (11.4% and 12.6%), two pH levels (3.0 and Hill, McLaren Vale, South Australia) with 13.7% v/v alco-
3.3), and phenolic addition or no phenolic addition. The two hol. Phenolic extracts were taken from three varietal wines:
alcohol levels were the original alcohol content of the base Viognier (Clonakilla, Murumbateman, New South Wales),
wine (11.4% v/v) and 12.6% v/v following the addition of 96% Fiano (Coriole Vineyards, McLaren Vale, South Australia),
v/v food grade ethanol. The two pH treatments comprised the and Gewürztraminer (Furst, Alsace, France) using the method
original wine pH of 3.3, and a pH of 3.0 achieved by adding described above. A 2 x 2 x 4 full factorial design was applied:
concentrated sulfuric acid. The alcohol and pH levels were two base wines x two ethanol levels (native alcohol concen-
selected to represent a range found in light-bodied dry white tration and 1% v/v ethanol addition) x three phenolic sources
wines. The “added phenolic” treatment was produced by add- (Viognier, Fiano, Gewürztraminer) plus a control (no phenolic
ing half of the phenolic extract taken from the same base wine addition). The method of phenolic extraction and the levels
to the same volume of base wine from which the phenolics used were the same as in study 1.
had been extracted. The percentage increase in the phenolic Sensory analysis. Four females and seven males with
level of the wine resulting from this addition was estimated similar levels of wine experience to those in study 1 were
by spectroscopy at 280 nm (Somers and Ziemelis 1985) to be trained in a similar manner, except that they were also trained
30%, which was confirmed gravimetrically by weighing the to rate their perception of viscosity using a 3 g/L carboxy-
dried eluate collected during consecutive extractions in which methylcellulose standard. The tastings were conducted using
the flow-through wine was reintroduced to the resin and the the protocol described previously, except that each of three
phenolics subsequently eluted using 96% ethanol. The con- tasting replicates was presented in a random manner to panel-
centration range of phenolics used was within that expected ists within a single tasting session.
when applying conventional white wine production processes Statistical analysis. The sensory data for both studies
(Cejudo-Bastante et al. 2011, Sims et al. 1995). was collected using FIZZ sensory data acquisition software
Sensory analysis. The panel comprised four female and (version 2.46; Biosystèmes, Couternon, France). ANOVA
nine male employees of the Australian Wine Research In- were conducted with panelists, phenolics, pH, and alcohol
stitute. All panelists had at least two to three years general as factors in study 1 and panelists, source of phenolics, and
wine-tasting experience as part of their profession, but none alcohol level as factors in study 2. Panelists were consid-
had previously participated in tastings specifically related to ered to be random factors. In study 1, means were considered
white wine mouthfeel. All the panelists were experienced in as statistically different if their difference was greater than
using unstructured line scales to assess sensory intensities. two standard errors. In study 2, Tukey’s tests were used to
Panelist training consisted of two 45-min sessions. During compare treatment means with the controls. Analyses were
the first session, panelists discussed the mouthfeel attributes conducted using Minitab 14 (Minitab Inc., State College, PA).
of six dry white wines previously deemed by a panel of ex-
perienced wine tasters to display different in-mouth textures. Results and Discussion
These included Australian Riesling, Chardonnay, and Fiano Phenolic extraction. High-performance liquid chroma-
wines, an Alsatian Gewürztraminer, and Italian wines made tography analysis showed that the phenolic extracts had phe-
from the Vermentino, Verdicchio, and Greco varieties. The nolic profiles similar to those of the wines from which they
panelists discussed the concepts of bitterness, astringency, were derived (data not shown). However, for all three wines,
hotness, and viscosity in the context of the wines presented. To the contributions of the early eluting hydrophilic phenolics
further demonstrate these attributes, panelists were exposed to such as the benzoic acids were lower in the extracts than in
a commercial Riesling wine to which 1% v/v ethanol (for hot- the wines themselves. Based on the proportion of peak area
ness), or 15 mg/L quinine sulfate (bitterness), or 200 mg/L of measured at 320 nm, depending on the wine, between 32%
commercial grape skin tannin (astringency) had been added. and 50% of these compounds either were not retained by
Sample assessment was conducted in tasting booths with the resin or were washed off during water rinsing aimed at
30 mL of each sample presented in ISO wine-tasting glasses removing sugars and organic acids. However, the phenolic

Am. J. Enol. Vitic. 64:4 (2013)


Effect of pH and Alcohol on Phenolics in White Wine – 427

extracts contained similar proportions of the abundant white the same increase in astringency. The increase in perceived
wine phenolics representing the major phenolic classes to the astringency when pH was reduced can be attributed either to
wines from which they were obtained: caftaric acid, grape an accentuation of astringency elicited by the phenolics in
reaction product, coumaric acid (hydroxycinnamates), tyrosol the base wine or to a direct effect of pH itself. Both effects
(phenols), catechin (flavan-3-ols), quercetin (flavonols), and have been extensively reported. The astringency of catechin
dihydroquercetin rhamnoside (flavanonol glycosides). (Kallithraka et al. 1997), tannic acid (Peleg et al. 1998), whole
Astringency. In study 1, adding phenolics to the base grape seed tannin (Fia et al. 2008), and flavanol oligomers
wine significantly increased its astringency (Table 1). How- extracted from grape seed (Fontoin et al. 2008) all increased
ever, the strength of the effect depended on pH (Figure 1A). with decreasing pH. Both organic and inorganic acids have
While the lower pH control wine was perceived as more also been shown to be astringent-like in the absence of phe-
astringent than the higher pH control wine, phenolic addi- nolics (Hartwig and McDaniel 1995). The direct effect of
tion at the lower pH had little further effect on astringency. pH on astringency has been proposed to result from either
However, at the higher pH of 3.3, phenolic addition increased direct changes to salivary rheology (Nordbo et al. 1984) or
astringency to the same level as the low pH condition, sug- enhancing interactions between lubricating salivary proteins
gesting that, for this particular wine, a 30% increase in phe- and phenolics (Obreque-Slier et al. 2012).
nolic concentration and a 0.3 unit decrease in pH resulted in Ethanol concentration did not affect perceived astringency
in study 1 (Figure 1A), but it had an enhancing effect in
study 2 (Table 2, Figure 2A). These results are contrary to
Table 1 Study 1: pH, ethanol, and phenolic addition effects in previous findings whereby the astringency elicited by both
a single base wine. Significance (p values) of ANOVA factors
between sensory attributes and composition. red wine and oligomeric seed tannins decreased in the pres-
Astringency Bitterness Hotness
ence of higher alcohol (Gawel et al. 2007b, Fontoin et al.
2008). Ethanol is thought to reduce the astringency elicited by
Phenolics (P) 0.046 0.217 <0.001
Ethanol (E) 0.524 0.079 <0.001
polymeric phenolics by disrupting hydrophobic interactions
pH 0.074 0.001 0.004
between salivary protein acceptor sites and the polyphenol
PxE 0.819 0.972 0.012 hydroxyl groups (Pascal et al. 2008). The neutral or enhanc-
E x pH 0.302 0.183 0.984 ing effect of ethanol on perceived astringency observed in
P x pH 0.095 0.770 0.384 these two studies may be attributed to either an insufficient
E x pH x P 0.916 0.144 0.021 number of hydroxyl groups present on monomeric white wine
phenolic species or the ability of ethanol to dehydrate the oral
epithelia resulting in the perception of palate dryness.
Bitterness. While not significant, adding phenolics to a
Riesling base wine (study 1) resulted in an increase in bitter-
ness under all combinations of alcohol and pH (Figure 1B). In
study 2, the overall effect of phenolic addition on bitterness
was significant (Table 2), with the phenolics extracted from
the Gewürztraminer resulting in a consistently significant in-
crease in bitterness in both base wines (Figure 2B). Single-
ton et al. (1975) also found that white wines that were more
phenolic were perceived to be consistently more bitter (albeit
not statistically significant). The lack of a consistently strong
effect of phenolics on bitterness across the various studies
may be the result of a naturally large, genetically determined
variation in bitterness perception among individuals (Prescott
et al. 2004).

Table 2 Study 2: ethanol and phenolic fractions effects in two


base wines. Significance (p values) of ANOVA factors between
sensory attributes and composition.
Astringency Bitterness Hotness Viscosity
Riesling
Phenolics (P) 0.001 <0.001 0.163 0.037
Ethanol (E) 0.028 0.012 <0.001 0.034
PxE 0.864 0.929 0.950 0.779
Chardonnay
Phenolics 0.120 <0.001 0.684 0.091
Figure 1 Mean intensity of astringency (A), bitterness (B), and hotness Ethanol 0.010 0.003 <0.001 0.091
(C) of a white wine at two pH values, two alcohol levels, and with and
PxE 0.302 0.797 0.637 0.340
without added phenolics (study 1). Error bars indicate two standard errors.

Am. J. Enol. Vitic. 64:4 (2013)


428 – Gawel et al.

Higher alcohol generally resulted in lower bitterness (Table sweetness also varies depending on concentration, then this
1, Figure 1B). This moderating effect of alcohol on phenolic may explain the observed differences in results of the two
bitterness is inconsistent with studies involving oligomeric studies, as the alcohol levels used differed by up to 3.3% v/v.
flavanols added to model wine (Fontoin et al. 2008) and cat- The lack of interactive effects between ethanol and phenolics
echin added to dealcoholized wine (Fischer and Noble 1994). on bitterness (Table 1, Table 2) in both studies also lends
The reason for this effect is unclear, but may be attributed support to a direct effect of ethanol on bitterness perception.
here to the use of a complex mixture of phenolics taken from Lower pH resulted in an increase in perceived bitterness
wine that contained a relatively low concentration of flavanols of both the base wine and those wines with added phenolics
compared to the earlier studies. Ethanol has been reported to (Table 1). Similarly, Fischer and Noble (1994) found that cat-
be bittersweet (Scinska et al. 2000). As such, it has the poten- echin was perceived as slightly more bitter when presented at
tial both to add to and to suppress phenolic bitterness. Other the lower end of the pH range used here. On the other hand,
compounds that elicit both bitterness and sweetness do so in the bitterness of larger oligomeric phenolics was unaffected
a relative way on a concentration-dependent basis (Fujimaru by pH (Fontoin et al. 2008).
et al. 2012). If the relative intensity of ethanol bitterness to Hotness. In study 1, phenolics accentuated perceived hot-
ness, particularly at lower alcohol levels (Table 1). Adding
30% phenolics increased the perceived hotness of the low
alcohol wine to around that of the high alcohol (+1.3% v/v)
wine without phenolics. The same effect of phenolics and
alcohol on perceived hotness was observed in study 2 (Figure
2). These results suggest that perceived hotness in white wine
may result from its phenolic content as well as its alcohol
content. The notion that phenolics may contribute to trigemi-
nal sensations related to heat is not unique to wine. While
not found in wine, the phenolic compounds oleocanthal and
zingerone have been directly attributed to the back palate heat
and pungency of extra virgin olive oil and ginger, respectively
(Andrewes et al. 2003, Nomura 1917).
In study 1, the increase in perceived hotness due to phe-
nolic addition was less pronounced at the higher alcohol level
(Figure 1C). A similar masking of phenolic induced hotness
by ethanol was observed in study 2. Here, perceived hotness
increased when phenolics from three different wines were
added to a lower alcohol Riesling base wine (12.5% v/v) and
a 1% v/v ethanol addition increased perceived hotness in an
additive fashion. However, in the Chardonnay base wine that
had been fortified to an alcohol level of 14.5%, perceived hot-
ness was unchanged following the same phenolic additions
(Figure 2C). These results have practical ramifications. The
alcohol levels of wines used in study 1 (11.4 to 12.6%) were
representative of light-bodied white wine. Adding 30% more
phenolics increased the hotness of all wines within this range,
but the effect was more pronounced in the lower alcohol wine.
Similarly, study 2 showed a consistent contribution of phe-
nolics to perceived hotness, but as alcohol levels increased,
the effect became smaller to the point whereby no effect
of phenolics on hotness was observed in a 14.5% v/v wine.
While it was demonstrated that phenolics have the capacity
to contribute to hotness in white wine, it seems likely that in
fuller bodied styles that are typically higher in alcohol, the
impact of phenolics on hotness is likely to be minor. It is in
these higher alcohol styles where winemakers are more likely
Figure 2 Mean intensity of (A) astringency, (B) bitterness, (C) hotness, to use methods that increase phenolic concentrations, such as
and (D) viscosity in Riesling (12.5% v/v) and Chardonnay (13.5% v/v) skin contact and pressing addition.
base wines supplemented with phenolics from Fiano (+F), Gewürztra- In study 1, alcohol concentration influenced the perception
miner (+G), and Viognier (+V) wines, with and without the addition of
of hotness and bitterness elicited by phenolics taken from a
1% ethanol (+Et). C indicates no phenolic addition (study 2). Error bars
indicate two standard errors. * indicates significantly different from the single Riesling wine. The generality of this effect was in-
respective control using Tukey’s test (p = 0.05). vestigated in study 2 in which 1% v/v ethanol was added to

Am. J. Enol. Vitic. 64:4 (2013)


Effect of pH and Alcohol on Phenolics in White Wine – 429

two base wines that had been supplemented with the addition Hartwig, P., and M.R. McDaniel. 1995. Flavor characteristics of
of phenolics from three varietal white wines. Study 2 also lactic, malic, citric and acetic acids at various pH levels. J. Food
Sci. 60:384-388.
considered the effect of alcohol and phenolics on perceived
viscosity. Jones, P.R., R. Gawel, I.L. Francis, and E.J. Waters. 2008. The inf lu-
ence of interactions between major white wine components on the
Viscosity. Adding phenolics from the Alsatian Gewürz-
aroma, f lavour and texture of model white wine. Food Qual. Pref.
traminer to both base wines resulted in higher perceived vis- 19:596-607.
cosity, which was further enhanced by the addition of alcohol. Kallithraka, S., J. Bakker, and M.N. Clifford. 1997. Effect of pH on
While increased perceived viscosity of white wine due to astringency in model solutions and wines. J. Agric. Food Chem.
alcohol has previously been reported (Gawel et al. 2007a), the 45:2211-2216.
positive effect of phenolics on white wine viscosity observed Lawless, H.T., J. Horne, and P. Giasi. 1996. Astringency of organic
here requires further investigation. acids is related to pH. Chem. Senses 21:397-403.
Nomura, H. 1917. The pungent principles of ginger. Part I. A new
Conclusions ketone, zingerone (4-hydroxy-3-methoxyphenylethyl methyl ketone)
Phenolics contribute to the perception of astringency, hot- occurring in ginger. J. Chem. Soc. 111:769-776.
ness, viscosity, and bitterness in white wine. Hotness and bit- Nordbo, H., S. Darwish, and R.S. Bhatnagar. 1984. Salivary viscos-
terness also depended on both pH and ethanol concentration. ity and lubrication: Inf luence of pH and calcium. Scand. J. Dental
Notably, the tastes and textures elicited by the presence of Res. 92:306-314.
phenolics are more pronounced at lower alcohol levels and at Obreque-Slier, E., A. Pena-Neira, and R. Lopez-Solis. 2012. Interac-
moderate pH levels, indicating that the presence of phenolics tions of enological tannins with the protein fraction of saliva and
astringency perception are affected by pH. LWT-Food Sci. Tech.
is more important in the context of lighter bodied wines. The 45:88-93.
consistently observed additive effect of ethanol indicates that
Pascal, C., C. Poncet-Legrand, B. Cabane, and A. Verhnet. 2008.
it contributes directly to the tastes and textures that are nor- Aggregation of a proline-rich protein induced by epigallocatechin
mally attributed to phenolic content and as such contributes gallate and condensed tannins: Effect of protein glycosylation. J.
significantly to white wine style. Agric. Food Chem. 56:6724-6732.
Peleg, H., K.K Bodine, and A.C. Noble. 1998. The inf luence of acid
Literature Cited on astringency of alum and phenolic compounds. Chem. Senses
Andrewes, P., J.L. Busch, T. de Joode, A. Groenewegen, and H. Al- 23:371-378.
exandre. 2003. Sensory properties of virgin olive oil polyphenolic Prescott, J., J. Soo, H. Campbell, and C. Robells. 2004. Responses of
identification of deacetoxy-ligtroside aglycone as a key contributor PROP taster groups to variations in sensory qualities within foods
to pungency. J. Agric. Food Chem. 51:1415-1420. and beverages. Physiol. Behav. 82:459-469.
Cejudo-Bastante, M.J., L. Castro-Vázquez, I. Hermosin-Gutiérrez, and Rubico, S.M., and M.R. McDaniel. 1992. Sensory evaluation of acids
M.S. Pérez-Coello. 2011. Combined effects of prefermentation skin by free-choice profiling. Chem. Senses 17:273-289.
maceration and oxygen addition of must on color-related phenolics,
Scinska, A., E. Koros, B. Habrat, A. Kukwa, W. Kostowski, and P.
volatile composition, and sensory characteristics of Airen white
Bienkowski. 2000. Bitter and sweet components of ethanol taste in
wine. J. Agric. Food Chem. 59:12171-12182.
humans. Drug Alcohol Depend. 60:199-206.
Fia, G., C. Dinella, M. Bertuccioli, and E. Monteleone. 2008. Predic-
Sims, C.A., J.S. Eastridge, and R.P. Bates. 1995. Changes in phenols,
tion of grape polyphenol astringency by means of a f luorimetric
color, and sensory characteristics of Muscadine wines by pre- and
micro-plate assay. Food Chem. 113:325-330.
post-fermentation additions of PVPP, casein, and gelatin. Am. J.
Fischer, U., and A.C. Noble. 1994. The effect of ethanol, catechin Enol. Vitic. 46:155-158.
concentration, and pH on sourness and bitterness of wine. Am. J.
Singleton, V.L., H.A. Sieberhagen, P. de Wet, and C.J. van Wyk.
Enol. Vitic. 45:6-10.
1975. Composition and sensory qualities of wines prepared from
Fontoin, H., C. Saucier, P.L. Teissedre, and Y. Glories. 2008. Effect white grapes by fermentation with and without grape solids. Am.
of pH, ethanol and acidity on astringency and bitterness of grape J. Enol. Vitic. 26:62-69.
seed tannin oligomers in model wine. Food Qual. Pref. 19:286-291.
Smith, P.A., and E.J. Waters 2012. Identification of the major drivers
Fujimaru, T., J.H. Park, and J. Lim. 2012. Sensory characteristics of ‘phenolic’ taste in white wines. Final report, Australian Wine
and relative sweetness of Tagatose and other sweeteners. J. Food Research Institute 0901. Grape and Wine Research and Development
Sci. 77:S323-S328. Corporation, Kent Town, Australia.
Gawel, R., S. Van Sluyter, and E.J. Waters. 2007a. The effects of Somers, T.C., and G. Ziemelis. 1985. Spectral evaluation of total
ethanol and glycerol on the body and other sensory characteristics phenolic components in Vitis vinifera grapes and wines. J. Sci. Food
of Riesling wines. Aust. J. Grape Wine Res. 13:38-45. Agric. 36:1275-1284.
Gawel, R., I.L. Francis, and E.J. Waters. 2007b. Statistical correla- Waterhouse, A. 2002. Wine phenolics. Ann. N.Y. Acad. Sci. 957:21-36.
tions between the in-mouth textural characteristics and the chemical
composition of Shiraz wines. J. Agric. Food Chem. 55:2683-2687.

Am. J. Enol. Vitic. 64:4 (2013)


Chapter 5.

The effect of mixtures of caftaric acid and grape reaction


product on the in-mouth sensory character of model wine

Richard Gawel, Alex Schulkin, Paul A. Smith and Elizabeth J. Waters

The Australian Wine Research Institute, PO Box 197, Glen Osmond, 5064, Australia

Australian Journal of Grape and Wine Research – 2014; 20: 25-30.


DOI: 10.1111/ajgw.12056

71
Gawel et al. Palate effects of caftaric acid and Grape Reaction Product 25

Taste and textural characters of mixtures of caftaric acid and


Grape Reaction Product in model wine
R. GAWEL, A. SCHULKIN, P.A. SMITH and E.J. WATERS*
The Australian Wine Research Institute, PO Box 197, Glen Osmond, Adelaide, SA 5064, Australia
Corresponding author: Mr Richard Gawel, email richard.gawel@awri.com.au
* Present address: Grape and Wine Research and Development Corporation, PO Box 610, Kent Town,
SA 5071, Australia.

Abstract
Background and Aims: Caftaric acid and Grape Reaction Product (GRP) are abundant phenolic substances in white
wine. Reports of their sensory effects are confined to individual threshold assessments and anecdotal evaluation of
their contribution to mouthfeel. This study profiles the taste and texture of mixtures of caftaric acid and GRP when
presented at concentrations typical of white wine.
Methods and Results: Mixtures of caftaric acid and GRP, isolated from white wine by countercurrent chromatog-
raphy, were added to a model wine using a 3 × 3 full factorial design, and their taste and texture were profiled by
trained assessors. Grape Reaction Product suppressed astringency and added to oily mouthfeel, while caftaric acid
suppressed the burning sensation from GRP. Neither GRP nor caftaric acid elicited significant bitterness.
Conclusions: Caftaric acid and GRP suppressed textural characters elicited by each other, and also by acidity, without
contributing significantly to bitterness.
Significance of Study: The ability of caftaric acid and GRP to reduce burning and drying sensations without adding
to bitterness suggests that they can potentially contribute positively to white wine texture. Oxidative juice handling
may contribute to oily mouthfeel that typifies some wine styles.

Keywords: astringency, bitterness, caftaric acid, Grape Reaction Product, mouthfeel, white wine

Introduction substrates, oxygen, caftaric acid and glutathione, with the latter
The phenolic profiles of white wines are influenced by the being the limiting factor (Rigaud et al. 1991). Sulfur dioxide
juice extraction and handling processes of the early stages of (SO2), copper ions and flavanols, however, can interfere with
winemaking. Harvesting techniques, crushing and draining the conversion of caftaric acid to GRP. Sulfur dioxide interferes
methods, prefermentation skin contact, and the incorporation with the conversion by reducing the caftaric acid quinone back
of press fractions influence both the total amount and the types to caftaric acid, while Cu(II)-catalysed auto-oxidation reactions
of phenolic substances extracted into the juice and the resultant reduce the glutathione substrate by converting it to its cor-
finished wines (Darias-Martín et al. 2000, Cejudo-Bastante responding disulfide (Kachur et al. 1998). The presence of
et al. 2011). flavanols also affects the reaction probably through a process of
Of the 80 or more phenolic substances identified in white coupled oxidation that results in the regeneration of caftaric acid
wine, caftaric acid (caffeoyl tartaric acid) and its derivative 2-S- from its quinone (Rigaud et al. 1991).
glutathionyl caftaric acid [Grape Reaction Product (GRP)] are The relative ratio of GRP to caftaric acid in white wine can
among the most abundant (Baderschneider and Winterhalter be increased by winemaking practices that involve oxidative
2001). Caftaric acid is predominantly found in the pulp of white must handling, and by limiting the addition of SO2. Oxygen
grapes, while GRP derives from the reaction between the exposure and resultant modification of the concentration of
o-quinones formed by the oxidation of caftaric acid with the caftaric acid and GRP can occur incidentally during the trans-
tripeptide glutathione that also resides in the pulp. Browning in port of machine-harvested grapes that have undergone partial
white wine is caused by coupled oxidation reactions between juicing or while conducting standard winemaking processes,
o-quinones and flavanols (Cheynier et al. 1988). Therefore, the such as crushing, draining and pressing (Cheynier et al. 1991).
scavenging of o-quinones by glutathione to produce colourless The concentration of phenolic substances and the ratio of
GRP can result in wines that have less propensity to brown caftaric acid to GRP can also be reduced following the deli-
enzymatically. While GRP may be oxidised by excess caftaric berate extensive oxidation (hyperoxidation) of juice prior
acid quinones following depletion of glutathione, it is not itself to fermentation (Cheynier et al. 1991, Cejudo-Bastante et al.
a substrate for further oxidation via the most significant oxida- 2011).
tive pathway that involves polyphenol oxidase (Rigaud et al. Despite their major presence in white wines and the avail-
1991). ability of winemaking practices to change their level and ratio,
The reaction balance between GRP and caftaric acid depends the contribution of GRP and caftaric acid to white wine taste
on the presence of polyphenol oxidase, the availability of its properties is not clear. Early taste threshold and difference tests

doi: 10.1111/ajgw.12056
© 2013 Australian Society of Viticulture and Oenology Inc.
26 Palate effects of caftaric acid and Grape Reaction Product Australian Journal of Grape and Wine Research 20, 25–30, 2014

suggest that neither caftaric acid nor GRP when tasted in isola- Table 1. Taste and textural attributes and their definitions used to
tion are likely to affect the taste of white wine. The detection describe mixtures of caftaric acid and Grape Reaction Product.
threshold of caftaric acid in water is reported to be 50 mg/L that
is higher than the typical concentration of this compound in Attribute Definition
white wine (Okamura and Watanabe 1981). Tasters could not
detect a taste difference resulting from the addition of five times Astringency Drying, mouth puckering sensation or after
the base concentration of caftaric acid (150 mg/L) or two times expectorating includes furry, chalky and grippy
the base concentration of GRP (50 mg/L) to a Chardonnay wine Viscosity Body, weight or thickness of the wine in the mouth
(Vérette et al. 1988). Caftaric acid at a suprathreshold level, Hotness Level of warmth or heat
however, is reported to be both bitter and astringent (Dadic
Bitter Bitter taste and/or after taste
and Belleau 1973). While the taste quality of GRP is yet to be
established, it is likely that caftaric acid and its derivative GRP Oily Feeling of oil in the mouth, mouth coating, a lingering
share some taste qualities. It is well established that combining oily feel
compounds with the same taste quality results in a perceived Acid AT Acid/sourness tasted after expectorating
overall taste intensity that is a partial summation of the intensity Acidity Sour/ acid taste, tangy, tart, the taste of lemon
of the components (Keast et al. 2003). Therefore, in order to
Metallic Metallic taste, spoon in the mouth, steely taste
assess the possibility that mixtures of caftaric acid and GRP may
have a sensory effect on white wine, we profiled the in-mouth Prickling Prickly sensation on the tongue and after expectorating
sensory characters of mixtures of these compounds at a concen- Hotness AT Warmth or heat after expectorating
tration and proportion that could realistically be achieved in Burning AT Chilli like burning sensation, different from hotness
wines.
AT, aftertaste.
Methods
Preparative isolation of caftaric acid and GRP
Caftaric acid and GRP were separated from other wine phenolic Smith and Waters (2012). The identity of the phenolic sub-
substances by high-speed countercurrent chromatography stances in the retained fractions was determined by their
(HSCCC). This process separates compounds on the basis of UV-spectra and confirmed by mass spectroscopy. The poss-
their hydrophobicity through high-speed mixing and separation ible presence of high molecular mass impurities (>1 kDa)
between immiscible aqueous and organic liquid phases. Phe- was assessed with size exclusion chromatography (SEC)
nolic substances were extracted from 36 L of 1-year-old white [Phenomenex Biosep Sec-S-2000 column (Phenomenex, Tor-
wines made from pressings juice from Chardonnay (18 L), rance, CA, USA), 1 mL/min NaNO3, 40°C, 5 mg/mL, 25 mL
Viognier (4.5 L) and from Riesling (13.5 L). The wines were injection] with refractive index and UV detection at 280 nm.
loaded onto a 500-mL column packed with Amberlite FPX66 The possible presence of organic acids, monomeric sugars, glyc-
resin (Dow, Camberwell, Vic., Australia). The column was erol and alcohols in the fractions was assessed by the method of
washed with 2 L of MilliQ water (EMD Millipore, Billerica, MA, Nissen et al. (1997).
USA) before eluting the phenolic substances with 2 L of 96% The model wine used in this study comprised 10% v/v
v/v ethanol. The eluate was chilled at 4°C overnight before ethanol in MilliQ water saturated with potassium hydrogen
being dried under vacuum at 35°C. tartrate and adjusted to pH 3.5 with tartaric acid. Caftaric acid
The phenolic substances were fractionated with a Quattro and GRP were added to model wine 2 h prior to tasting on the
Mk II HSCCC device (AECS-QuikPrep Ltd, Bridgend, Wales) basis of their reported or estimated extinction coefficients
equipped with a 500-mL stainless steel preparative scale coil (Cheynier et al. 1986) to 30 and 60 mg/L. A full factorial design
(2.16 mm i.d. tubing). Elution of the phenolic substances was was used, i.e. all possible combinations of 0, 30 and 60 mg/L
monitored with two single wavelength ultra violet (UV) detec- caftaric acid and 0, 30 and 60 mg/L GRP were added to the
tors (UPC-900 Amersham Biosciences, Amersham, England for model wine.
A280 and a GBC LC 1210, Braeside, Vic., Australia for A320).
Fractions (10 mL) were collected (Amersham Pharmacia Biotec,
Uppsala, Sweden). Sensory assessment
In a two-stage separation, GRP was isolated first from the Eight female and six male tasters with previous experience
whole phenolic substances extract with a butanol : water (1:1) both in taste and in-mouth sensory profiling of white wines
system. The phenolic substances from 6 L of wine were dis- and phenolic fractions taken from white wines were trained for
solved in 50 mL of a mixture of the mobile and stationary phase this study over six sessions. Training began by familiarising
before being injected into HSCCC (800 rpm, 2 mL/min of the tasters with the palate sensory attributes commonly identified
aqueous layer as the mobile phase, 28–32°C), during which the in white wines (bitter, astringent, viscous and hotness) using
early-eluting GRP fraction was collected. The stationary phase four commercial samples made from cultivars normally high in
was recovered and dried under vacuum at less than 30°C to phenolic substances. Two or three examples of model wines
recover the remaining phenolic substances. These were dis- containing either a high (60 mg/L) or low (30 mg/L) concen-
solved in 20 mL of a mixture of mobile and stationary phase of tration of GRP and caftaric acid, or their mixtures were pre-
a hexane : ethyl acetate : methanol : water (3:7:3:7) system sented over the following three training sessions in order to
whereby a caftaric acid fraction was collected under the same generate and define taste and texture attributes associated with
chromatographic conditions of the previous step. The two-step the samples. The list of attributes and their definitions as
separation process was repeated five times, and the fractions agreed upon by the tasting panel is given in Table 1. Aqueous
collected from each run were pooled to obtain sufficient mate- solutions of aluminium potassium sulfate (0.5 g/L), quinine
rial for sensory assessment. sulfate (15 mg/L) and carboxymethycellulose (3 g/L) were also
The composition of the phenolic substances of the pooled presented to demonstrate astringency, bitterness and viscosity,
fractions was assessed by HPLC using the method described in respectively. A pungent, extra virgin olive oil was used to

© 2013 Australian Society of Viticulture and Oenology Inc.


Gawel et al. Palate effects of caftaric acid and Grape Reaction Product 27

illustrate the concept of ‘burning aftertaste’. The standard was using Fisher’s least significant difference. A Type 1 error of
chosen on the basis that the compound responsible for its 10% (P < 0.10) criteria for a difference or interaction was used
pungency is a monomeric phenolic substance that manifests throughout.
itself as an ‘aftertaste’ located at the back of the throat
(Cicerale et al. 2009). The tasting panel accepted the olive oil
standard as being a good facsimile of the burning aftertaste Results and discussion
sensation perceived in the training samples. In order to famil- Characterisation of fractions
iarise the tasters with the use of the rating scale and to identify Figure 1 shows the HSCCC separation chromatograms meas-
any attribute redundancies, model wine, high and low in GRP, ured at A320 and fraction cuts to obtain GRP and caftaric acid.
and high in caftaric acid, was presented in duplicate in the The HPLC chromatograms of the phenolic fractions used in the
final two training sessions that occurred over consecutive sensory trial are given in Figure 2. Size exclusion chromatogra-
days. phy showed that the GRP fraction contained no high molecular
The test samples were assessed in triplicate with the 15 test mass polar compounds, such as polysaccharides and proteins,
samples presented each session over three sessions in a com- that had the potential to co-elute with GRP during separation by
pletely randomised order within the session. The intensity of the countercurrent chromatography. However, while SEC showed
selected attributes was assessed using a 15 cm partially struc- that there were some low molecular mass (<1 kDa) compounds
tured line scale with ‘low’ and ‘high’ anchor points placed at 1.5 present in the sample, reverse phase C18 chromatography sub-
and 13.5 cm along the scale. Samples (40 mL) were presented in sequently showed organic acids, sugars, glycerol and alcohols in
black-tasting glasses at room temperature (23 ± 1°C). A 2-min the GRP fraction were all present at a concentration below their
rest was enforced between samples, and an additional 8-min limit of detection and therefore were below their reported
rest was enforced after the third sample. sensory threshold.

Statistical analysis Bitterness


Where zero ratings for a particular attribute were given to all Adding caftaric and GRP had no effect on the perceived bitter-
samples on all tasting occasions by an assessor, their ratings ness or acidity of the model wine (Table 2). The taste of caftaric
were excluded from analysis as they were considered either acid has been described anecdotally as being bitter (Maga and
insensitive to the attribute or they did not understand it. The Lorenz 1973), and its non-esterified form (caffeic acid) as acidic
ratings from at least 11 assessors remained for each attribute. (Vérette et al. 1988). To date, however, all evaluations of the
A three-way analysis of variance of each tasting attribute was sensory effect of caftaric acid and GRP have been confined to
conducted with assessors, caftaric acid concentration and detection threshold and overall difference testing. Evaluation of
GRP concentration as treatment effects, with assessors being specific taste and textural attributes of either of these com-
treated as random effects. Means separation was conducted pounds has not been attempted. Nagel and Graber (1988) could

Figure 1. High-speed counter-


current chromatography of
phenolic substances extracted
from white wine: (a)
butanol : water (1:1) system
with the shaded area being the
fraction cut for Grape Reaction
Product; and(b) hexane : ethyl
acetate : methanol : water
(3:7:3:7) system with the
shaded area being the fraction
cut for caftaric acid.

© 2013 Australian Society of Viticulture and Oenology Inc.


28 Palate effects of caftaric acid and Grape Reaction Product Australian Journal of Grape and Wine Research 20, 25–30, 2014

not find a consistent relationship between a GRP concentration highest concentration and when GRP was absent (Figure 3).
and bitterness in white wine. Notably, most combinations of GRP and caftaric acid reduced
the (acid- and alcohol-induced) astringency of the model
Astringency wine (Figure 3), with GRP being the significant main effect
The model wine was perceived to be astringent in its own right. (Figure 4). Naish et al. (1993) reported that a compound related
Its astringency can logically be attributed to its acid or ethanol to caftaric acid – the quinic acid ester of caffeic acid – also
content as they both can produce in-mouth ‘dryness’, a term suppressed astringency. While the mechanism that underlies the
that is also frequently used to describe the astringent quality reduction of wine matrix-induced astringency by these phenolic
resulting from when phenolic substances interact with proteins substances remains unclear, their lack of contribution to astrin-
either found in saliva or bound to the oral mucosa (Payne et al. gency could be explained by an insufficient number of phenol
2008). rings and hydroxyl groups being present to enable them to
Under the phenolic–protein binding model of astringency, it either hydrophobically interact, or to hydrogen bond with sali-
may be expected that the phenolic substances GRP and caftaric vary or oral bound proteins (Rawel et al. 2006).
acid would be astringent. Adding GRP and caftaric acid to the
model wine, however, did not contribute to its astringency Burning aftertaste
except in the instance where caftaric acid was added at the Grape Reaction Product significantly increased ‘burning after-
taste’ in the absence of caftaric acid (Figure 3). Caftaric acid,
however, reduced the intensity of the ‘burning aftertaste’ pro-
duced by GRP with higher concentrations of caftaric acid
resulting in a greater suppressive effect (Figure 3). The exact
nature of the back of the throat burning sensation is unknown,
however, small molecular mass phenolic substances that inter-
act with the receptors in the throat to produce irritation are
known. For example, the phenolic aldehyde oleocanthal that
is responsible for the back of the throat burning sensation in
extra virgin olive oil binds to specifically tuned receptors that
are located in the human pharynx (Peyrot des Gachons
et al. 2011). ‘Burning aftertaste’ was not positively associated
with either in-mouth hotness (r = −0.331) or hot aftertaste
(r = −0.395), which suggests that it is a sensory character that
is conceptually different to that of hotness. However, as the
model wine control also elicited a burning aftertaste, this sug-
gests that the tasters could not totally differentiate between
alcohol- and phenolic-induced hotness on the aftertaste.
Recently, both ethanol and a phenolic methyl ester found in
white wine have been shown to activate the same nonselec-
tive sensory channel responsible for oral irritation (TRPA1)
(Komatsu et al. 2012, Son et al. 2012) that provides some
physiological evidence for a phenolic-derived chemosensory
effect and for the suppression of alcohol hotness by low
molecular mass phenolic substances.

Oiliness and viscosity


The oily mouthfeel of the model wine increased consistently
with GRP concentration (Figure 4). While it is possible that the
textures described by the panel as ‘oily’ and ‘viscous’ were
synonymous (as their ratings were moderately correlated,
r = 0.53, P = 0.141), the observation that GRP addition did not
Figure 2. High-performance liquid chromatography traces of frac- significantly affect perceived viscosity (P = 0.380) suggests that
tions selected for sensory assessment: (a) Grape Reaction Product these two textures are sensorially distinct. It is unlikely that GRP
and (b) caftaric acid. at the concentration used in this study could instigate the

Table 2. Significance (P values) of sensory effects of caftaric acid and Grape Reaction Product addition to model wine.

Astringency Viscosity Oily Bitter Metallic Prickling Hotness Acidity Acid Hotness Burning
AT AT AT

Caftaric acid 0.09 0.42 0.19 0.51 0.62 0.37 0.96 0.29 0.19 0.47 0.03
Grape Reaction 0.10 0.38 0.09 0.39 0.65 0.71 0.86 0.32 0.75 0.71 0.65
Product
Interaction 0.31 0.09 0.20 0.14 0.76 0.27 0.99 0.61 0.30 0.87 0.78

Bold value indicates a significant effect (P < 0.1). AT, aftertaste.

© 2013 Australian Society of Viticulture and Oenology Inc.


Gawel et al. Palate effects of caftaric acid and Grape Reaction Product 29

Figure 3. Mean intensity rating of (a) astringency and (b) burning


aftertaste of mixtures of caftaric acid and Grape Reaction Product.
Mixtures of (a) caftaric acid (0, 30 and 60 mg/L) contained 0 ( ); 30
( ); and 60 ( ) mg/L Grape Reaction Product and (b) mixtures of
Grape Reaction Product (0, 30 and 60 mg/L) contained 0 ( ); 30 ( );
and 60 ( ) mg/L caftaric acid. Means with different letters are sig-
nificantly different (P < 0.1).

amount of intermolecular interaction necessary to cause an


increase in perceived viscosity given that the highly viscous
kosmotropic glycerol could not invoke an increase in perceived
viscosity in white wine even when added at a concentration 150 Figure 4. Mean intensity rating of (a) astringency, (b) burning after-
times greater than the highest concentration of GRP used here taste and (c) oily mouthfeel of mixtures of caftaric acid and Grape
(Gawel and Waters 2008). Reaction Product. The mixtures contained 0 ( ), 30 ( ) and 60 ( )
mg/L of either caftaric acid or Grape Reaction Product, and the
An alternative mechanism for oiliness perception involves
intensity rating for each concentration of caftaric acid and Grape
possible changes to the structure of the thin layer of absorbed Reaction Product was, respectively, the average of the values for the
proteins that coat oral surfaces that is referred to as the salivary three concentrations of Grape Reaction Product and caftaric acid.
pellicle. Most of this layer is thought to comprise mucins and Means with different letters are significantly different (P < 0.1).
other glycosylated proteins that by extending their hydrated
hydrophilic side chains out from adjacent oral surfaces result
in steric repulsion between them (Zappone et al. 2007). instead to changes in the microrheology and morphology of
Grape Reaction Product has two terminal carboxyl groups that the salivary pellicle (Aguirre et al. 1989, Dickinson and Mann
could allow cross-linking via hydrogen bonding between the 2006).
oligosaccharide side-chains extending from the adjacent inter-
acting mouth surfaces. Any disruption to the microrheology of
these surfaces, following such cross-linking, would be signalled Overall sensory impact
by free nerve fibres embedded in the mucosa as a tactile Using duo-trio difference testing Vérette et al. (1988) reported
change. While this explanation is speculative, there is growing that caftaric acid and GRP, when individually presented, were
evidence that textural perception of aqueous systems contain- undetectable in white wine when present in the upper range of
ing phenolic substances cannot simply be attributed to the concentration found in white wine. The effect of a small number
rheology of the liquid being tasted or to changes to the of tasters and an insensitive test method, however, may have
rheology of the pool of unbound saliva following tasting but contributed to their null results.

© 2013 Australian Society of Viticulture and Oenology Inc.


30 Palate effects of caftaric acid and Grape Reaction Product Australian Journal of Grape and Wine Research 20, 25–30, 2014

Recently, Sáenz-Navajas et al. (2012) found that the sensory Dadic, M. and Belleau, G. (1973) Polyphenols and beer flavor. Proceedings
effects of binary mixtures of structurally similar quercetin of the American Society of Brewing Chemists 4, 107–114.
Darias-Martín, J.J., Rodríguez, O., Díaz, E. and Lamuela-Raventós, R.M.
glycosides were also not additive. Instead, the existence of one (2000) Effect of skin contact on the antioxidant phenolics in white wine.
quercetin glycoside in a mixture was able to completely sup- Food Chemistry 71, 483–487.
press the sensory effect of another. Here, the structurally similar Dickinson, M.E. and Mann, B. (2006) Nanomechanics and morphology of
phenolic substances, caftaric acid and GRP, were also shown to salivary pellicle. Journal of Materials Research 21, 1996–2002.
both suppress and contribute to several taste and textural attrib- Gawel, R. and Waters, E.J. (2008) The effect of glycerol on the perceived
viscosity of dry white table wine. Journal of Wine Research 19, 109–
utes in model wine. The suppression, however, of the astringent 114.
and heat-related sensations arising from alcohol and acidity by Kachur, A.V., Koch, C.J. and Biaglow, J.E. (1998) Mechanism of copper-
small molecular mass phenolic substances has not previously catalyzed oxidation of glutathione. Free Radical Research 28, 259–269.
been reported. Keast, R.S.J., Bournazel, M.M.E. and Breslin, P.A.S. (2003) A
psychophysical investigation of binary bitter-compound interactions.
Chemical Senses 28, 301–313.
Conclusion Komatsu, T., Uchida, K., Fujita, F., Zhou, Y.M. and Tominaga, M. (2012)
Phenolic substances in white wine have traditionally been asso- Primary alcohols activate human TRPA1 channel in a carbon chain length-
ciated with the negative palate characteristics of bitterness and dependent manner. European Journal of Physiology 463, 549–559.
Maga, J.A. and Lorenz, K. (1973) Taste threshold values for phenolic acids
astringency (Singleton et al. 1975). The results reported here which can influence flavor properties of certain flours, grains, and
demonstrate that two of the major phenolic substances, caftaric oilseeds. Cereal Science Today 18, 326–328.
acid and GRP, are neither bitter nor astringent at a concentration Nagel, C.W. and Graber, W.R. (1988) Effect of must oxidation on quality of
typically encountered in white wine. The textural effect of GRP white wines. American Journal of Enology and Viticulture 39, 1–4.
Naish, M., Clifford, M.N. and Birch, G.G. (1993) Sensory astringency of
and caftaric acid was not additive. On the contrary, for some 5-O-caffeoylquinic acid, tannic acid and grape seed tannin by a time-
textures, they were antagonistic with one suppressing the tex- intensity procedure. Journal of the Science of Food and Agriculture 61,
tural contribution of the other. Furthermore, in combination, 57–64.
they suppressed perceived astringency produced by acid and/or Nissen, T.L., Schulze, U., Nielsen, J. and Villadsen, J. (1997) Flux distribu-
alcohol. While GRP also contributed to a burning aftertaste, it tions in anaerobic, glucose-limited continuous cultures of Saccharomyces
cerevisiae. Microbiology (Reading, England) 143, 203–218.
also increased palate oiliness that is considered to be a positive Okamura, S. and Watanabe, M. (1981) Determination of phenolic
characteristic in some wine styles. The observed contrasting cinnamates in white wine and their effect on wine quality. Agricultural
sensory impacts of GRP and caftaric acid suggest that white wine and Biological Chemistry 45, 2063–2070.
texture may be manipulated by juice handling techniques, as Payne, C., Bowyer, P.K., Herderich, M. and Bastian, S.E.P. (2008) Interac-
tion of astringent grape seed procyanidins with oral epithelial cells. Food
the subsequent concentration of either GRP or caftaric acid in Chemistry 115, 551–557.
wine can be increased at the expense of the other through Peyrot des Gachons, C., Kunitoshi, U., Bryant, B., Shima, A., Sperry, J.B.,
managing oxygen exposure of the juice. Dankulich-Nagrudny, L., Tominaga, M., Smith, A.B., Beauchamp, G.K.
and Breslin, P.A.S. (2011) Unusual pungency from extra virgin olive oil Is
attributable to restricted spatial expression of the receptor of oleocanthal.
Acknowledgements The Journal of Neuroscience 31, 999–1009.
The work was funded by Australia’s grapegrowers and wine- Rawel, H.M., Frey, S.K., Meidtner, K., Kroll, J. and Schweigert, F.J. (2006)
makers through their investment body, the Grape and Wine Determining the binding affinities of phenolic compounds to proteins by
Research Development Corporation, with matching funds from quenching of the intrinsic tryptophan fluorescence. Molecular Nutrition
and Food Research 50, 705–713.
the Australian government, and by Pernod Ricard Pacific. The Rigaud, J., Cheynier, V., Souquet, J.-M. and Moutounet, M. (1991) Influ-
Australian Wine Research Institute is a member of the Wine ence of must composition on phenolic oxidation kinetics. Journal of the
Industry Cluster, located in the Waite Precinct of The University Science of Food and Agriculture 57, 55–63.
of Adelaide. Sáenz-Navajas, M.-P., Avizcuri, J.-M., Ferreira, V. and Fernández-Zurbano,
P. (2012) Insights on the chemical basis of the astringency of Spanish red
wines. Food Chemistry 134, 1484–1493.
References Singleton, V.L., Sieberhagen, H.A., de Wet, P. and van Wyk, C.J. (1975)
Aguirre, A., Mendoza, B., Levine, M.J., Hatton, M.N. and Douglas, W.H. Composition and sensory qualities of wines prepared from white grapes by
(1989) In vitro characterization of human salivary lubrication. Archives of fermentation with and without grape solids. American Journal of Enology
Oral Biology 34, 675–677. and Viticulture 26, 62–69.
Baderschneider, B. and Winterhalter, P. (2001) Isolation and characteriza- Smith, P.A. and Waters, E.J. (2012) Identification of the major drivers of
tion of novel benzoates, cinnamates, flavonoids, and lignans from Riesling phenolic taste in white wines. Grape and Wine Research and Development
wine and screening for antioxidant activity. Journal of Agricultural and Corporation Report AWRI 0901 (Grape and Wine Research and Develop-
Food Chemistry 49, 2788–2798. ment Corporation: Adelaide, SA, Australia).
Cejudo-Bastante, M.J., Hermosín-Gutiérrez, I., Castro-Vázquez, L. and Son, H.J., Kim, M.J., Park, J.-H., Ishii, S., Misaka, T. and Rhyu, M.-R. (2012)
Pérez-Coello, M.S. (2011) Hyperoxygenation and bottle storage of Char- Methyl syringate, a low molecular weight phenolic ester, as an activator of
donnay white wines: effects on color-related phenolics, volatile composi- the chemosensory ion channel TRPA1. Archives of Pharmacal Research
tion, and sensory characteristics. Journal of Agricultural and Food 35, 2211–2218.
Chemistry 59, 4171–4182. Vérette, E.N., Noble, A.C. and Somers, T.C. (1988) Hydroxycinnamates of
Cheynier, V., Owe, C. and Rigaud, J. (1988) Oxidation of grape juice phe- Vitis vinifera: sensory assessment in relation to bitterness in white wines.
nolic compounds in model solutions. Journal of Food Science 53, 1729– Journal of the Science of Food and Agriculture 45, 267–272.
1732. Zappone, B., Ruths, M., Green, G.W., Jay, G.D. and Israelachvili, J.N. (2007)
Cheynier, V., Souquet, J.-M., Samson, A. and Moutounet, M. (1991) Adsorption, lubrication, and wear of lubricin on model surfaces: polymer
Hyperoxidation: influence of various oxygen supply levels on oxidation brush-like behavior of a glycoprotein. Biophysical Journal 92, 1693–
kinetics of phenolic compounds and wine quality. Vitis 30, 107–115. 1708.
Cheynier, V.F., Trousdale, E.K., Singleton, V.L., Salgues, M.J. and Wylde, R.
(1986) Characterization of 2-s-glutathionylcaftaric acid and its hydrolysis
in relation to grape wines. Journal of Agricultural and Food Chemistry 34,
217–221.
Cicerale, S., Breslin, P.A.S., Beauchamp, G.K. and Keast, R.S.J. (2009) Manuscript received: 25 January 2013
Sensory characterization of the irritant properties of oleocanthal, a natural Revised manuscript received: 5 June 2013
anti-inflammatory agent in extra virgin olive oils. Chemical Senses 34,
333–339. Accepted: 20 August 2013

© 2013 Australian Society of Viticulture and Oenology Inc.


Chapter 6.

White wine taste and mouth-feel as affected by juice extraction


and processing

Richard Gawel, Martin Day, Steven C. Van Sluyter, Helen Holt


and Elizabeth J. Waters

The Australian Wine Research Institute, PO Box 197, Glen Osmond, 5064, Australia

Journal of Agricultural and Food Chemistry – 2014; 62: 10008-10014.


DOI: 10.1021/jf503082v

79
Article

pubs.acs.org/JAFC

White Wine Taste and Mouthfeel As Affected by Juice Extraction and


Processing
Richard Gawel,* Martin Day, Steven C. Van Sluyter,† Helen Holt, Elizabeth J. Waters,‡ and Paul A. Smith
Australian Wine Research Institute, P.O. Box 197, Glen Osmond, Adelaide, South Australia 5064, Australia
*
S Supporting Information

ABSTRACT: The juice used to make white wine can be extracted using various physical processes that affect the amount and
timing of contact of juice with skins. The influence of juice extraction processes on the mouthfeel and taste of white wine and
their relationship to wine composition were determined. The amount and type of interaction of juice with skins affected both
wine total phenolic concentration and phenolic composition. Wine pH strongly influenced perceived viscosity, astringency/
drying, and acidity. Despite a 5-fold variation in total phenolics among wines, differences in bitter taste were small. Perceived
viscosity was associated with higher phenolics but was not associated with either glycerol or polysaccharide concentration.
Bitterness may be reduced by using juice extraction and handling processes that minimize phenolic concentration, but lowering
phenolic concentration may also result in wines of lower perceived viscosity.
KEYWORDS: white wine, mouthfeel, phenolics, polysaccharides, bitterness, hyperoxidation, skin contact

■ INTRODUCTION
Differences in wine sensory profiles resulting from winemaking
colorless derivative, grape reaction product (GRP), and to
remove other phenolics that can potentially contribute to wine
practices are often referred to as “wine styles” and are bitterness by complexation and precipitation.7
important as they define the context in which the wine is The ability to influence wine composition by applying
best consumed. White wines are usually broadly classified on different juice extraction and handling methods partly arises
the basis of their flavor intensity, acidity, and sweetness, but from differences in the relative concentrations of phenolic types
they can also be differentiated on other taste and mouthfeel found in the pulp, skins, and seeds of the grape. Pulp cell
attributes such as perceived viscosity, hotness, astringency vacuoles contain relatively low concentrations of phenolics
(dryness), and bitterness. These variations in white wine compared to skins and seeds and are mostly represented by
mouthfeel have been attributed to phenolics,1 glycerol,2 hydroxycinnamic acids, whereas seeds contain significant
ethanol,3,4 sugar,4 and polysaccharides.2,5 amounts of flavanols and skins contain flavanols, flavononols,
The same batch of white grapes can be made into wines with flavonols, and hydroxycinnamic acids.8
distinguishably different mouthfeel characteristics by employing Consequently, juice extraction methods that favor interaction
conventional methods of juice extraction and processing prior between grape skins and juice such as prefermentation skin
to fermentation.1 The majority of white wine by volume is contact or pressing promote higher concentrations of
made from juice that is obtained by draining crushed grapes flavonoids in wine.9 Derivatives of these grape phenolics may
under the action of gravity immediately following destemming form or be degraded during fermentation and subsequent
and crushing. However, this basic method of juice production storage. Phenolic acids can undergo esterification and form
can be varied to affect changes in wine style. When a flavorsome phenolic ethyl esters, and these and other grape-derived
wine with a viscous mouthfeel (i.e., a fuller bodied style) is (cinnamyltartaric) esters may hydrolyze under wine acidic
sought, winemakers may allow grape skins to stay in contact conditions.10,11 Yeasts undertaking primary fermentation are
with juice prior to draining (skin contact) or they may include thought to metabolize the grape amino acid tyrosine into the
the juice pressed off the wet skins that remain after draining phenolic tyrosol.9
(pressings), as both methods result in more flavorsome wines Glycerol is produced by yeast during primary fermentation
with higher phenolic concentration.6 Conversely, juices with and is the most abundant compound in dry white wines after
lower phenolic concentrations that are destined for the water and ethyl alcohol.12 Its concentration in wine depends on
production of less intensely flavored wines with a “thinner” yeast strain and fermentation parameters, and it has the effect of
mouthfeel (i.e., lighter bodied styles) can be obtained by increasing the physical viscosity of white wine.13 Consequently,
applying pressure to whole uncrushed bunches (whole bunch glycerol can potentially contribute to the perception of white
pressing). The phenolic content of juice can be further wine viscosity or fullness. White wine polysaccharides are
decreased by hyperoxidation whereby nonsulfited juice is derived from yeast cell walls during fermentation and
saturated with oxygen prior to fermentation, resulting in the
condensation and precipitation of phenolics during fermenta- Received: February 11, 2014
tion and lees settling.1 Hyperoxidation is used commercially to Revised: September 23, 2014
improve the oxidative stability of wine by promoting the Accepted: September 24, 2014
conversion of caftaric acid to its more oxidatively stable and Published: September 24, 2014

© 2014 American Chemical Society 10008 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014
Journal of Agricultural and Food Chemistry Article

subsequent settling or from grape cell walls prior to or during represent the tastes and mouthfeel characteristics of the training wines
fermentation. Whereas polysaccharides have been shown to were ‘astringency (drying)’, drying or mouth puckering; ‘viscosity’,
contribute to the perceived viscosity of model wine,2,5 their role body, weight, or thickness; ‘hotness’, warmth perceived after
in the mouthfeel and taste of white wine, where they have the expectorating; ‘acidity’, sourness/tartness while in the mouth;
‘sweet’, sweetness while in the mouth; ‘bitter’, bitterness in the
potential to form complex colloids with other macromolecules mouth. Bitterness and acidity after expectorating were also selected by
and phenolics, is less clear. the panel as relevant descriptors, but were subsequently found to
A variety of juice extraction and handling options are used in correlate strongly with in-mouth sensations so they were deemed
the production of white wines: whole bunch pressing, free run redundant.
juice, pressing fraction juice, prefermentation skin contact, and Standards for ‘astringency (drying)’ (0.5 g/L aluminum potassium
juice hyperoxidation. A simultaneous comparison of the effects sulfate, Ajax Finechem, Sydney, NSW, Australia), ‘viscosity’ (3 g/L
of these winemaking options on white wine composition and carboxymethyl cellulose, sodium salt, Sigma-Aldrich, St. Louis, MO,
their effect on the broad range of mouthfeel attributes USA), “bitter” (15 mg/L quinine sulfate, Sigma-Aldrich), ‘hotness’
encountered in white wine has not been conducted to date. (20% v/v ethanol, Tarac Industries, Nuriootpa, SA, Australia), and
‘acidity’ (2 g/L tartaric acid, Ajax Finechem, Sydney, Australia) in
This study aimed first to investigate the effects of the water were also presented for discussion during training. The tasters
commonly applied juice extraction and processing methods were familiarized with the use of the unstructured line scale used in the
used in white winemaking on wine phenolic composition and, formal assessment by rating six of the experimental wines in duplicate
second, to examine the relationships between phenolic on the chosen attributes. The line scale was 15 cm long with “low” and
composition and other aspects of the wine matrix, pH, total “high” anchor points placed at 1.5 and 13.5 cm along the scale.
polysaccharides, glycerol and alcohol concentrations, on the The perceived intensity of the taste and mouthfeel attributes of all
mouthfeel profile of white wine. wines (3 grape sources × 6 treatments × 2 fermentation replicates)


were assessed in triplicate using the line scale used in training. Twenty
wines were tasted per day over six sessions that extended over 14 days.
MATERIALS AND METHODS Thirty milliliters of wine was tasted from clear ISO tasting glasses
Grape Sources and Winemaking. Handpicked bunches (3 t) of under natural lighting and at room temperature (22 ± 1 °C). The
Riesling from Eden Valley, South Australia (harvested February 25, tasters were forced to rest for 1 min between wines and for 10 min
2010), Viognier from the Adelaide Hills, South Australia (harvested after every fourth wine. Sensory assessment was conducted 1 month
March 11, 2010), and Chardonnay from Lyndoch, South Australia after bottling. Basic wine analysis was conducted at the time of
(harvested February 11, 2010) were sourced (hereafter referred to as bottling, and polysaccharide concentration and phenolic profiles were
“grape sources”). Potassium metabisulfite was added at picking at 50 evaluated 4 months after bottling.
mg/L on the basis of an estimated yield of 700 L/t. The grapes were Wine Analysis. Phenolics. Total phenolics (TP) and total
transported for a maximum of 2 h in 0.5 t bins before being chilled hydroxycinnamates (HCA) were determined on the basis of
overnight. The grapes were subsequently destemmed and crushed absorbance at 280 and 320 nm.15 Absorbance at 370 nm (A370) was
using a Diemme 4t/h destemmer-crusher. An amount of 1.75 t of the also measured to assess the concentration of flavonol compounds. As
resultant must from each grape source was pressed by a 2.5 t capacity the variation in TP, HCA, and A370 between replicate ferments was low
Willmes membrane press (Willmes “Merlin”, 1 t whole bunch and 2.5 (CV < 2.5%), one fermentation replicate from each of the three grape
t crushed capacity) to yield three fractions: free run juice (FR) at <0.5 sources was further analyzed for phenolic composition by HPLC with
bar pressure, light pressings (LP) at 0.5−1.0 bar, and hard pressings an Agilent 1100 chromatography system under the following
(HP) at 1.0−2.0 bar. conditions. Tandem (2×) Phenomenex Gemini C6-Phenyl columns
A portion of the FR juice was extensively hyperoxidized (HOX) by (2.1 mm × 150 mm, 3 μm) were used with 20 μL injections of wine at
sparging with 100% oxygen at 8−10 L/min until reaching 40 mg/L 45 °C. Gradient composition and flow rates were as follows: 2% formic
free dissolved oxygen (Nomasense PSt3, Nomacorc LLC, Zebulon, acid in water (v/v, solvent A) and 2% formic acid in methanol (v/v,
NC, USA). Skin-contacted (SC) treatments were produced by solvent B); 0.5% B from 0 to 5 min, to 20% B from 5 to 70 min, to
blending free run and press fractions obtained from must held in 98% B from 70 to 110 min (0−110 min, 0.16 mL/min), 98% B from
press for 60 h at 5 °C and pressed as described above. Whole bunch 110 to 115 min (0.13 mL/min), and 0.5% B from 110 to 120 min
pressed juice (WB) was also obtained by pressing 0.5 t of whole (0.16 mL/min).
bunches at 1.0−2.0 bar. The phenolic compounds from a mixture of wines from the three
The juices were cold settled with the aid of pectolytic enzymes grape sources were identified using HPLC-ESI-QTOF. Identification
(Novozyme, Vinoclear Classic, Dittingen, Switzerland) for 24−48 h was made on the basis of accurate masses (mass difference <5 mDa)
and acid adjusted using tartaric acid. Duplicate 30 L fermentation and isotope ratios (mSigma <20) of the respective [M − H]− ions to
vessels were filled with each juice and inoculated with Saccharomyces the assigned elemental composition as well as their MS/MS and UV
cereviseae strain EC1118. The ferments proceeded at 0.5−1.0 Baumé spectra. The identities of some phenolic compounds were confirmed
per day until dryness. Thereafter, 60 mg/L SO2 was added, and the by comparison with the MS/MS and UV spectra of corresponding
wines were cold stabilized by seeding with 4 g/L potassium bitartrate reference compounds (Supportying Information). For quanitification,
at 4 °C. Protein stability was achieved by adding 1 g/L sodium benzoic acids and flavanols were expressed in gallic acid equivalents,
bentonite to the juice prior to fermentation and by adding additional hydroxycinnamates and GRP as ferulic acid equivalents, and flavones
sodium bentonite after cold stabilization at a rate determined by the 80 and flavanonols as quercetin-3-glucoside equivalents.
°C/2 h test.14 The stabilized wines were sparged with nitrogen and Polysaccharides and Glycerol. For polysaccharide determination,
argon to attain a dissolved oxygen level of <0.6 mg/L, bottled in 375 240 mL of wine was reduced to 60 mL under vacuum at 30 °C (in
mL antique green bottles under Sarin-tin screw caps, and stored at 15 triplicate) and 180 mL of ethanol added (96% v/v, Tarac Industries,
°C prior to analysis and sensory assessment. Nuriootpa, SA, Australia). The precipitate was removed by
Sensory Assessment of Mouthfeel and Taste. Eleven tasters centrifugation and washed twice with ethanol, dissolved in Milli-Q
(five males and six females) were used. All but one had experience in water, and extensively dialyzed at 23 °C against Milli-Q water for 36 h
profiling the mouthfeel and tastes in white wine. Taster training was (3.5 kDa MW cutoff tubing, Thermo Scientific, Rockford, IL, USA)
conducted over five sessions to generate and define mouthfeel and before being freeze-dried and weighed. Size exclusion chromatography
taste attributes appropriate to the wines and to identify attribute (Phenomenex BioSep-SEC-S 2000, 300 × 7.8 mm, 40 μL injection, 40
redundancies. Tasters were trained using a subset of the wines used in °C, 0.1 M NaNO3 as mobile phase, 1 mL/min. RI and UV−vis
the study and a high-phenolic white wine made by fermenting juice in detection at 280 nm. MW estimated using Pullulan standards 100, 48,
the presence of skins. The attribute descriptions agreed upon to 23.7, and 12.2 kDa, Showa Denko, Japan) was used to confirm the

10009 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014


Journal of Agricultural and Food Chemistry Article

Table 1. Wine Compositiona


Glu+Fru ethanol glycerol
treatment (g/L) pH TA (g/L) (% v/v) (g/L) PS (mg/L) TP A280 (au) HCA A320 (au) A370 (au)
WB 0.55 (0.09) 3.06 (0.02) 7.4 (0.3) 13.1 (0.1) 6.47 (0.16) 154 (1) 0.655 (0.132) 2.147 (0.308) 0.513 (0.033)
HOX 0.77 (0.08) 3.04 (0.05) 7.4 (0.3) 13.2 (0.1) 6.41 (0.13) 160 (2) 0.527 (0.096) 1.591 (0.194) 0.530 (0.052)
FR 0.79 (0.10) 3.11 (0.03) 7.3 (0.3) 13.4 (0.1) 6.43 (0.06) 154 (2) 1.569 (0.266) 3.300 (0.470) 0.598 (0.042)
LP 0.45 (0.08) 3.17 (0.01) 7.2 (0.2) 13.2 (0.1) 6.79 (0.17) 155 (1) 2.281 (0.315) 3.270 (0.578) 0.678 (0.054)
HP 0.55 (0.09) 3.18 (0.02) 7.3 (0.2) 13.2 (0.1) 6.81 (0.27) 146 (2) 2.643 (0.234) 3.283 (0.486) 0.716 (0.059)
SC 0.58 (0.04) 3.15 (0.06) 7.6 (0.2) 13.0 (0.2) 6.71 (0.08) 167 (3) 4.029 (0.394) 5.499 (0.819) 0.919 (0.087)
P 0.008 n/ab n/a <0.001 0.003 0.029 <0.001 <0.001 <0.001
a
Two standard errors are shown in parentheses (n = 6). bn/a indicates juices were acid adjusted with tartaric acid so effects cannot be attributed to
treatment.

absence of low molecular weight carbohydrates and low molecular


weight phenolic compounds in the lyophilized material. Glycerol
concentration was determined by HPLC as previously described.16
Statistical Analysis. Wine Composition. The effects of juice
extraction and handling methods (hereafter referred to as wine-
making) on the concentrations of total polysaccharides, glycerol, TP,
HCA, and A370 were analyzed by ANOVA with winemaking as factors
with fermentation replicates nested within them. Winemaking was
considered fixed and fermentation replicates, random factors.
Concentrations of phenolic classes, benzoic acids, tyrosol and
resveratrols, hydroxycinnamic acids, GRP, flavanols, flavonols,
flavononols, and flavanones, were analyzed by ANOVA with
replications provided by grape sources.
Sensory Attribute Ratings. Taster ratings were analyzed by
ANOVA with winemaking and assessors as factors, with fermentation
replicates nested within winemaking treatments. Assessors and
fermentation replicates were treated as random factors and wine-
making as fixed factors.
Relationships between Wine Composition and Sensory Ratings.
Mean sensory intensities were modeled on wine pH, titratable acidity,
ethanol, total polysaccharides, glycerol, and glucose and fructose
concentrations, on the concentrations of major phenolic classes, and Figure 1. Difference in total phenolics, total hydroxycinnamates, and
on TP, HCA, and A370 using orthogonal scores partial least-squares A370 of wines as a percentage of wines made from free run juice. HOX,
regression (PLS). hyperoxidised free run; WB, whole bunch press; FR, free run; LP, light
ANOVA were conducted using MINITAB 14 (Minitab Inc., State pressings; HP, hard pressings; SC, 60 h skin contact prefermentation.
College, PA, USA). PLS regression analysis was conducted using Error bars represent 2 standard errors.
Unscrambler 10.2 (CAMO Software AS, Oslo, Norway).

■ RESULTS AND DISCUSSION


Winemaking. The experimental winemaking protocol
The whole bunch pressed (WB) wines contained signifi-
cantly lower TP than the free run wines (Figure 1), which
resulted from lower concentrations of benzoic acids, hydrox-
produced wines with analytical parameters generally expected ycinnamic acids (Figure 2A), and flavanols (Figure 2B). Lower
of white wine (Table 1). Although juices from press fractions concentrations of flavanols were expected in the whole bunch
are more likely to be selected for hyperoxidization in pressed wines as the process results in less disruption to the
commercial practice than free run juices, the inclusion of a berry than do crushing and draining. The lower HCA in the
hyperoxidized free run treatment was produced in this study whole bunch pressed wines may be explained by less contact
with the intention of producing wines with phenolic with the two possible sources of HCA, grape pulp solids and
concentrations at the lower end that may be found in skins.17
commercial practice. Conversely, the extended period of skin The hard pressing (HP) wines in this study did not differ in
contact (60 h at 5 °C) was applied to produce phenolic their concentrations of either HCA or their oxidative products
concentrations typical of high-phenolic white wines. It should (GRP) (Figure 2A) compared with the free run wines, which
also be noted that as the juices were pH adjusted by tartaric suggests either that they were not released from the skins
acid addition, the differences in pH and TA among wines given during pressing or that they were subsequently removed
in Table 1 and the perceived acidity and sweetness of the wines possibly via fining type actions of other macromolecules
do not necessarily reflect the effect of the winemaking released during pressing or following other condensation
treatments applied. reactions during fermentation.
Phenolic Composition. The free run (FR) treatment Hyperoxidation of free run juice (HOX) resulted in
represents a logical “control” treatment because on a volume significantly lower TP and HCA compared to the free run
basis it is the most widely used method of white wine control1 (Figure 1). The reduction in TP resulted from
production. Juice extraction and handling significantly affected significantly reduced concentrations of hydroxycinnamic acids,
TP, HCA, and A370 (p < 0.05) as well as benzoic acid, flavanols, and flavanonols (Figure 2). The higher concentration
hydroxycinnamic acid, flavanol, flavanonol, flavonol, flavanone, of GRP (albeit not statistically significant) in the hyperoxidized
and phenol concentrations (P < 0.1) (Figures 1 and 2). wines relative to the free run wines suggests that some caftaric
10010 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014
Journal of Agricultural and Food Chemistry Article

interaction between juice and skins, and therefore both would


be expected to result in greater extraction of phenolics from the
skins.19,20 However, the wines made from the juices produced
from pressing skins and those made by skin contact had
significantly different phenolic profiles (Figure 2). Specifically,
extensive prefermentation skin contact resulted in wines with
significantly higher flavanonol concentrations than the free run
wines, but the light and hard pressing wines did not differ in
this regard, whereas the pressing wines contained significantly
higher concentrations of flavanones than did either the free run
or skin contacted wines. The different phenolic profiles
between the pressings and skin contact wines suggest that
mechanisms other than simple extraction were involved during
skin contact and pressing. The skin contact juices were in
contact with skins for a substantially longer period than those
of pressings juices, which may have favored enzymatically
catalyzed oxidative processes such as the C-4 hydroxylation and
dehydrogenation of flavanols into flavanonols.21 The lower
proportional increase in flavanones in the skin-contacted wines
compared with that of the pressing wines is worthy of further
investigation given that these compounds are known to be the
primary contributors of bitterness in other foods and
beverages.22 In citrus, flavanones are enzymatically biotrans-
formed to flavanonols via flavonol intermediates.23 The
concomitant lower concentrations of flavanones and higher
concentrations of flavanonols and flavonols observed in the
skin-contacted wines compared with the pressing wines may
suggest that extensive skin contact could similarly involve
enzymatic transformation of flavanones into flavononols.
Keeping juice in contact with skins prior to fermentation has
been shown to increase total hydroxycinnamates when done
under conditions similar to those of this study, that is, at low
temperature and under the action of pectolytic enzymes.24,25
However, the contrasting effect of skin contact and pressing
with respect to wine HCA concentration is worthy of further
investigation.
The tyrosol concentration was higher in the pressings and
wines than the free run wines. This was most likely the result of
greater extraction of tyrosine from the skins, which is
metabolized by yeast during primary fermentation to produce
Figure 2. Difference in concentrations of major phenolic classes as a tyrosol.26
percentage of those in the corresponding free run wines: (A) Glycerol and Polysaccharide Concentration. Wines
nonflavonoids; (B) flavonoids. Error bars represent 2 standard errors. made from pressings and skin-contacted juice had significantly
∗ indicates significant difference from free run control. higher (p = 0.003) glycerol concentrations that those made
from whole bunch pressed or free run juice (Table 1).
acid was converted to GRP. Similarly, the significantly lower However, the range in mean glycerol concentrations among
concentrations of flavanols in the hyperoxidized wines treatments was <10% of the range of glycerol concentrations
compared with the free run wines may be attributed to that have been reported in surveys of commercial dry white
polyphenol oxidase catalyzed oxidation.18 Reduced flavanonol wines.12,27 The small practical differences in glycerol concen-
concentration resulting from hyperoxidation has not previously tration between wines from juices obtained from different juice
been reported. However, given that flavanonols and flavanols extraction methods are expected as it is known that glycerol
both contain catechol 3′,4′-dihydroxyl electron-donating production depends primarily on yeast strain and fermentation
groups, their reduction may attributable to the same oxidative conditions,28 which were the same across all winemaking
mechanism. The hyperoxidation regimen used in this study treatments.
could be considered to be extreme in that the juice was The total polysaccharide concentrations of wine made
extensively saturated using pure oxygen. Whether flavanonol primarily from juice deriving mostly from the grape pulp
concentration is similarly affected when a more commercially (whole bunch pressed, free run, and hyperoxidized free run)
representative oxygenation regimen is applied is yet to be were similar. Skin-contacted wines were significantly higher in
tested. total polysaccharides, possibly due to skin-bound pectins being
The wines made from pressings (LP and HP) and skin- released during maceration, adding to the polysaccharides
contacted juice (SC) were higher in TP than those made from already present in the juice (Table 1). Wines made from hard
free run juices (Figure 1). The pressing and skin contact pressings contained significantly lower polysaccharide concen-
treatments were similar in that they both involved some form of trations than the other treatments, which suggest that
10011 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014
Journal of Agricultural and Food Chemistry Article

polysaccharides may form condensation products with same wine that had been previously been adjusted to pH 3.0.31
phenolics under the action of hard pressing. However, consistent with the results of this study, lower pH
Mouthfeel and Taste Effects of Juice Extraction and wines were generally perceived to be more astringent than
Relationships to Wine Composition. The mean sensory higher pH wines regardless of their phenolic content.31
attribute intensity ratings with treatments ordered from lowest Astringency can be modulated by viscosity,32 which could
to highest mean TP are given in Figure 3. ‘Astringency explain why the wines perceived as more viscous were rated as
less astringent (Figure 3). However, as perceived viscosity was
positively correlated with total phenolic concentration, which
under the classic protein−phenolic precipitation model would
be expected to increase astringency, it is likely that variations in
astringency/drying in these wines were due to factors other
than total phenolic concentration.
Perceived viscosity was positively associated with TP, A370,
and pH (p < 0.001, Table 2). PLS analysis also suggests that
both pH and phenolics contribute to perceived viscosity
(Figure 4). Oral perception of viscosity has previously been
attributed to pH-induced changes in oral surface rheology.33
Adding whole phenolics from white wine has also recently been
Figure 3. Mean ratings of mouthfeel and taste attribute ratings by juice shown to increase the perceived viscosity of white wine.31
extraction and handling treatment. Bars represent LSD 5%. Perceived viscosity was most strongly correlated with benzoic
acid ethyl ester (r = 0.48, p = 0.032), cinnamic acid ethyl ester
(drying)’ generally decreased with increasing TP. Under the (r = 0.57, p = 0.009), and flavonol ethyl ester (r = 0.42, p =
classical protein−phenolic binding model of astringency, wines 0.063) concentrations. Phenolic acid ethyl esters are formed
with higher phenolics should be more, rather than less, during the course of alcoholic fermentation and are ubiquitous
astringent. However, it may be that the monomeric phenolics in alcohol beverages including wine,34 but as both their
in white wine may not have a sufficient number of hydrogen formation through esterification and degradation by hydrolysis
bonding sites that are required to extensively cross-link salivary are pH dependent, their role in perceived viscosity remains
proteins with phenolics and therefore elicit astringency. speculative.
Analogously, chlorogenic acid, which is structurally similar to Despite the potential for glycerol and polysaccharides to
and contains the same number of hydrogen bonding sites as increase the perceived viscosity by increasing physical
caftaric acid, was also found not to elicit astringency in model viscosity,13,35 neither glycerol nor polysaccharide concentra-
studies.29 Organic acids are capable of producing drying tions were positively associated with perceived viscosity in these
astringent-like sensations in their own right with pH being a wines (Table 2; Figure 4). Whereas this may have been due to
greater influence than titratable acidity.30 The acidity the small relative differences in glycerol and polysaccharide
parameters of pH and titratable acidity were very strongly concentrations among treatments (Table 1), various studies
correlated with ‘astringency (drying)’, with lower pH (r = have suggested that perceived viscosity in white wine or model
−0.87, p < 0.001) and higher TA (r = 0.53, p < 0.001) wines wine is either unaffected or nominally affected by glycerol or
being more ‘astringent (drying)’. These correlations and PLS polysaccharide concentrations within those typical of dry white
results (Figure 4) strongly suggest that the astringent (drying) wines2,4,13,36
sensation perceived in these wines resulted from acidity rather ‘Hotness’ was not significantly associated with any composi-
than phenolic concentration. Adding white wine phenolics to a tional variable, including ethanol concentration (Table 2). In
white wine of pH 3.3 increased its astringency, but the same other studies, adding either ethanol,3 or phenolics31 derived
addition of phenolics did not increase the astringency of the from white wines made white wine perceptively hotter. The
lack of a correlation between ‘hotness’ and ethanol concen-
tration may have resulted either from the narrow range of
ethanol concentration among samples or by suppression by
other wine components.
Juice extraction and handling had a significant effect on
bitterness (p = 0.013; Figure 3), with bitterness increasing with
total phenolic concentration (r = 0.460, p = 0.034; Table 2).
PLS analysis also suggested that phenolics contributed to
bitterness in these wines (Figure 4). Bitterness was strongly
associated with flavanol concentration (r = 0.722, p < 0.001).
Flavanols have been shown to elicit bitterness,37 and their role
in bitterness perception has recently been confirmed following
the discovery that they activate known bitter receptors on the
tongue.38 Benzoic acids and their derivatives were also
positively associated with bitterness (r = 0.417, p = 0.034),
Figure 4. Partial least-squares regression correlation loadings of which has been observed in red wines. 39 Flavanone
sensory attributes and total phenolics (TP), total hydroxycinnamates concentrations were related to greater bitterness (r = 0.470, p
(HCA), A370, total polysaccharides (PS), glucose and fructose (Glu = 0.018), which is consistent with findings related to other
+Fru), titratable acidity (TA), phenolic classes, pH, ethanol and foodstuffs.23 However, as flavanone concentration was strongly
glycerol concentrations. correlated with flavanol concentration (r = 0.730, p < 0.001),
10012 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014
Journal of Agricultural and Food Chemistry Article

Table 2. Correlations between Wine Composition and Mouthfeel and Taste Attributesa
Glu+Fru pH TA ethanol glycerol PS TP HCA A370
astringency(drying) 0.29 −0.87*** 0.53*** −0.21 0.16 0.39* −0.39* −0.09 −0.15
viscosity −0.20 0.70*** −0.34* −0.02 0.02 −0.34* 0.55*** 0.32* 0.53***
hotness −0.07 −0.03 −0.07 0.25 0.24 −0.14 −0.11 −0.06 −0.02
acidity 0.25 −0.74*** 0.59*** −0.10 0.23 0.39* −0.27 −0.06 −0.13
bitter −0.25 0.12 −0.12 −0.16 0.03 0.16 0.46** 0.39* 0.31
sweet 0.02 0.73*** −0.62*** 0.39* −0.29 −0.53** −0.03 −0.24 −0.12
a
*, **, and *** indicate significance at p < 0.05, 0.01, and 0.001, respectively.

the interpretation that flavanones contribute to bitterness


remains speculative. Hydroxycinnammic acids in white wine did
■ AUTHOR INFORMATION
Corresponding Author
not influence bitterness perception even when added to *(R.G.) Phone: +61 8 3136630. Fax: +61 8 3136600. E-mail:
concentrations well above that found in white wines,40 which richard.gawel@awri.com.au.
is inconsistent with the positive association (r = 0.417, p =
0.034) between summed concentrations of hydroxycinnamic Present Addresses

acids and their derivatives and bitterness observed in this study. (S.C.V.S.) Department of Biological Sciences, Macquarie
One possible reason is that the previous studies assessed the University, New South Wales 2109, Australia.

effect of individual hydroxycinnamics, whereas in this study (E.J.W.) Australian Grape and Wine Authority, Kent Town,
bitterness was equated to the summed presence of free South Australia 5067, Australia.
hydroxycinnamic acids and their tartaric acid and ethyl esters. Funding
Prefermentation skin contact resulted in wines with The work was funded by Australia’s grapegrowers and
significantly higher TP than the pressing wines, but they winemakers through their investment body, the Australian
tended to be less bitter. This may have been the result of Grape and Wine Authority, with matching funds from the
different phenolic profiles (Figure 2) as some phenolic classes Australian government, and by Orlando Wines.
may elicit more bitterness than others.41 Alternatively, the Notes
lower bitterness in the skin-contacted wines may be the result The authors declare no competing financial interest.


of masking by acidity or polysaccharides (Table 1).
Sweetness perception and glucose and fructose concen- REFERENCES
trationd were unrelated, which could be due to the narrow
(1) Cejudo-Bastante, M. J.; Castro-Vázquez, L.; Hermosín-Gutiérrez,
range of concentrations (0.2−1.0 g/L) and masking by acidity
I.; Pérez-Coello, M. S. Combined effects of preferementative skin
(r sweetness with TA = −0.62, p < 0.001). Whereas maceration and oxygen addition of must on color-related phenolics,
polysaccharide concentration was negatively associated with volatile composition, and sensory characteristics of Airén white wine. J.
sweetness (Table 2; Figure 4), a direct effect cannot be Agric. Food Chem. 2011, 59, 12171−12182.
attributed as polysaccharide concentration was correlated with (2) Jones, P. R.; Gawel, R.; Francis, I. L.; Waters, E. J. The influence
pH (r = 0.67, p < 0.001). However, there is a potential for of interactions between major white wine components on the aroma,
polysaccharides to act as sweetness antagonists as their flavour and texture of model white wine. Food Quality Pref. 2008, 19,
extensive and complex structures provide the possibility that 596−607.
they have glycophores of size and shape similar to those of (3) Gawel, R.; Van Sluyter, S.; Waters, E. J. The effects of ethanol and
small molecular weight sweet compounds in white wine. glycerol on the body and other sensory characteristics of Riesling
wines. Aust. J. Grape Wine Res. 2007, 13, 38−45.
In conclusion, different methods of juice extraction and
(4) Nurgel, C.; Pickering, G. Contribution of glycerol, ethanol and
handling techniques resulted in significant differences in overall sugar to the perception of viscosity and density elicited by model white
phenolic content and phenolic profiles, but only small wine. J. Texture Stud. 2005, 36, 303−323.
differences in glycerol and polysaccharide concentrations. (5) Vidal, S.; Courcoux, P.; Francis, L.; Kwiatkowski, M.; Gawel, R.;
White wine bitterness was nominally affected by overall Williams, P.; Waters, E.; Cheynier, V. Use of an experimental design
phenolic concentration and by the relative concentrations of approach for evaluation of key wine components on mouth-feel
phenolic compounds present. The perception of palate perception. Food Quality Pref. 2004, 15, 209−217.
astringency or drying sensation could best be accounted for (6) Darias-Martín, J.; Díaz-González, D.; Díaz-Romero, C. Influence
by pH differences rather than phenolics and, therefore, could be of two pressing processes on the quality of must in white wine
influenced by winemaking methods used to modify acidity. pH production. J. Food Eng. 2004, 63, 335−340.
(7) Cheynier, V.; Souquet, J.-M.; Samson, A.; Moutounet, M.
in combination with phenolic content influenced perceived
Hyperoxidation: influence of various oxygen supply levels on oxidation
viscosity, with higher pH and more phenolic wines having more kinetics of phenolic compounds and wine quality. Vitis 1991, 30, 107−
apparent viscosity, which is an important attribute when fuller 115.
bodied styles of white wine are required.


(8) Di Lecce, G.; Arranz, S.; Jáuregui, O.; Tresserra-Rimbau, A.;
Quifer-Rada, P.; Lamuela-Raventós, R. M. Phenolic profiling of the
ASSOCIATED CONTENT skin, pulp and seeds of Albarino grapes using hybrid quadrupole time-
*
S Supporting Information
of-flight and triple-quadrupole mass spectrometry. Food Chem. 2014,
145, 874−882.
Chromatographic and spectral characteristics of pure reference (9) Trousdale, E. K.; Singleton, V. L. Astilbin and engeletin in grape
phenolic compounds analyzed by HPLC-DAD; phenolic and wine. Phytochemistry 1983, 22, 619−620.
compounds determined by HPLC-ESI-QTOF-MS/MS. This (10) Edwards, T. L.; Singleton, V. L.; Boulton, R. Formation of ethyl
material is available free of charge via the Internet at http:// esters of tartaric acid during wine aging: chemical and sensory effects.
pubs.acs.org. Am. J. Enol. Vitic. 2004, 36, 118−124.

10013 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014


Journal of Agricultural and Food Chemistry Article

(11) Ramey, D. D.; Ough, C. S. Volatile ester hydrolysis or formation (33) Nordbo, H.; Darwish, S.; Bhatnagar, R. S. Salivary viscosity and
during storage of model solutions and wines. Am. J. Enol. Vitic. 1980, lubrication: Influence of pH and calcium. Scand. J. Dental Res. 1984,
28, 928−934. 92, 306−314.
(12) Nieuwoudt, H. H.; Prior, B. A.; Pretorious, I. S.; Bauer, F. F. (34) Hashizume, K.; Ito, T.; Shimohashi, M.; Kokita, A.; Tokiwano,
Glycerol in South African table wines: an assessment of its relationship T.; Okuda, M. Taste-guided fractionation and instrumental analysis of
to wine quality. S. Afr. J. Enol. Vitic. 2002, 23, 22−30. hydrophobic compounds in saki. Biosci., Biotechnol., Biochem. 2012, 76,
(13) Runnenbaum, R. C.; Boulton, R. B.; Powell, R. L.; Heymann, H. 1291−1295.
Key constituents affecting wine body − an exploratory study. J. Sensory (35) Weaver, L.; Yu, L. P.; Rollings, J. E. Weighted intrinsic viscosity
Stud. 2011, 26, 62−70. relationships for polysaccharide mixtures in dilute aqueous solutions. J.
(14) Pocock, K. F.; Waters, E. J. Protein haze in bottled white wines: Appl. Polym. Sci. 1988, 35, 1631−1637.
how well do stability tests and bentonite fining trials predict haze (36) Gawel, R.; Waters, E. J. The effect of glycerol on the perceived
formation during storage and transport? Aus. J. Grape Wine Res. 2008, viscosity of dry white table wine. J. Wine Res. 2008, 19, 109−114.
12, 212−220. (37) Narukawa, M.; Noga, C.; Ueno, Y.; Sato, T.; Misaka, T.;
(15) Somers, T. C.; Ziemelis, G. Spectral evaluation of total phenolic Watanabe, T. Evaluation of the bitterness of green tea catechins by a
components in Vitis vinifera grapes and wines. J. Sci. Food Agric. 1985, cell-based assay with the human bitter taste receptor hTAS2R39.
36, 1275−1284. Biochem. Biophys. Res. Commun. 2011, 405, 620−625.
(16) Nissen, T. L.; Schulze, U.; Nielsen, J.; Villadsen, J. Flux (38) Soares, S.; Kohl, S.; Thalmann, S.; Mateus, N.; Meyerhof, W.;
distribution in anaerobic, glucose-limited continuous cultures of De Freitas, V. Different phenolic compounds activate distinct human
Saccharomyces cerevisiae. Microbiology 1997, 143, 203−218. bitter taste receptors. J. Agric. Food Chem. 2013, 61, 1525−1533.
(17) Somers, T. C.; Pocock, K. F. Phenolic assessment of white (39) Hufnagel, J. C.; Hofmann, T. Orosensory-directed identification
musts: varietal differences in free-run juices and pressings. Vitis 1991, of astringent mouthfeel and bitter-tasting compounds in red wine. J.
30, 189−201. Agric. Food Chem. 2008, 56, 1376−1386.
(18) Guyot, S.; Cheynier, V.; Souquet, J.-M.; Moutounet, M. (40) Verétte, E.; Noble, A. C.; Somers, T. C. Hydroxycinnamates of
Influence of pH on the enzymatic oxidation of (+)-catechin in model Vitis vinifera: sensory assessment in relation to bitterness in white
systems. J. Agric. Food Chem. 1995, 43, 2458−2462. wines. J. Sci. Food Agric. 1988, 45, 267−272.
(19) Singleton, V. L.; Zaya, J.; Truesdale, E. White table wine quality (41) Maga, J. A.; Lorenz, K. Taste threshold values for phenolic acids
and polyphenol composition as affected by must SO2 content and which can influence flavor properties of certain flours, grains and
pomace contact time. Am. J. Enol. Vitic. 1980, 31, 14−20. oilseeds. Cereal Sci. Today 1973, 18, 326−330.
(20) Darias-Martín, J. J.; Rodríguez, O.; Díaz, E.; Lamuela-Raventós,
R. M. Effect of skin contact on the antioxidant phenolics in white wine.
Food Chem. 2000, 71, 483−487.
(21) Matsuda, M.; Otsuka, Y.; Jin, S.; Wasaki, J.; Watanabe, J.;
Watanabe, T.; Osaki, M. Biotransformation of (+)-catechin into
taxifolin by a two-step oxidation: primary stage of (+)-catechin
metabolism by a novel (+)-catechin-degrading bacteria, Burkholderia
sp. KTC-1, isolated from tropical peat. Biochem. Biophys. Res. Commun.
2008, 366, 414−419.
(22) Horowitz, R. M.; Gentili, B. Taste and structure in phenolic
glycosides. J. Agric. Food Chem. 1969, 17, 696−700.
(23) Moriguchi, T.; Kita, M.; Tomono, Y.; Endo-Inagaki, T.; Omura,
M. Gene expression in flavonoid biosynthesis: correlation with
flavonoid accumulation in developing citrus fruit. Physiol. Plant.
2001, 111, 66−74.
(24) Fernández-Zurbano, P.; Ferreira, V.; Peña, C.; Escudero, A.;
Cacho, J. Effects of maceration time and pectolytic enzymes added
during maceration on the phenolic composition of must. Food Sci.
Technol. Int. 1999, 5, 319−325.
(25) Hernanz, D.; Recamales, A. F.; Gonzáles-Miret, M. L.; Gómez-
Míguez, M. J.; Vicario, I. M.; Heredia, F. J. Phenolic composition of
white wines with a prefermentative maceration at experimental and
industrial scale. J. Food Eng. 2007, 80, 327−335.
(26) Singleton, V. L.; Trousdale, E. White wine phenolics: varietal
and processing differences as shown by HPLC. Am. J. Enol. Vitic. 1983,
34, 27−34.
(27) Rankine, B. C.; Bridson, D. A. Glycerol in Australian wines and
factors influencing its formation. Am. J. Enol. Vitic. 1971, 22, 6−12.
(28) Erasmus, D. J.; Cliff, M.; Van Vuuren, H. J. J. Impact of yeast
strain on the production of acetic acid, glycerol and the sensory
attributes of icewine. Am. J. Enol. Vitic. 2004, 55, 371−378.
(29) Naish, M.; Clifford, M. N.; Birch, G. G. Sensory astringency of
5-O-caffeoylquinic acid, tannic acid and grape seed tannin by a time
intensity procedure. J. Sci. Food Agric. 1993, 61, 57−64.
(30) Lawless, H. T.; Horne, J.; Giasi, P. Astringency of organic acids
is related to pH. Chem. Senses 1996, 21, 397−403.
(31) Gawel, R.; Van Sluyter, S. C.; Smith, P. A.; Waters, E. J. Effect of
pH and alcohol on perception of phenolic character in white wine. Am.
J. Enol. Vitic. 2013, 64, 425−429.
(32) Peleg, H.; Noble, A. C. Effect of viscosity, temperature and pH
on astringency in cranberry juice. Food Qual. Pref. 1999, 10, 343−347.

10014 dx.doi.org/10.1021/jf503082v | J. Agric. Food Chem. 2014, 62, 10008−10014


Chapter 7.

The influence of polysaccharides on the taste and mouth-feel of


white wine

Richard Gawela, Paul A. Smitha and Elizabeth J. Watersb

a
The Australian Wine Research Institute, PO Box 197, Glen Osmond, 5064,
Australia
b
Wine Australia, Hackney Road, Kent Town, 5071, Australia

Australian Journal of Grape and Wine Research, – 2016;.22: 350-357


DOI: 10.1111/ajgw.12222

87
350 Effect of polysaccharides on white wine mouthfeel Australian Journal of Grape and Wine Research 22, 350–357, 2016

Influence of polysaccharides on the taste and mouthfeel of white wine

R. GAWEL1, P. A. SMITH1 and E. J. WATERS1,2


1
The Australian Wine Research Institute, Glen Osmond, SA 5064, Australia; 2 Wine Australia, Kent Town, SA 5071, Australia
Corresponding author: Richard Gawel, email richard.gawel@awri.com.au

Abstract
Background and Aims: Mouthfeel attributes of white wine contribute to its style and therefore to the context in which it is most
appropriately consumed. The effect of polysaccharides on the mouthfeel of white wine and their interaction with pH, ethanol and
phenolic substances was determined.
Methods and Results: White wine polysaccharides and polysaccharide fractions, defined by molecular mass and composition,
were added to either white or model white wine of variable pH and concentration of alcohol and phenolic substances to assess their
effect on white wine taste and mouthfeel. A higher concentration of polysaccharides reduced perceived palate hotness. The effect
of polysaccharides on mouthfeel and taste was independent of the concentration of phenolic substances. A medium molecular
mass polysaccharide fraction of 13–93 kDa, containing arabinogalactan protein and mannoprotein reduced palate hotness and
increased viscosity at higher pH.
Conclusions: Polysaccharides had a relatively small effect on mouthfeel and taste compared with the effect of wine pH and eth-
anol. Medium molecular mass polysaccharides were mostly responsible for the difference in perceived hotness and viscosity.
Significance of the Study: Palate hotness is a negative characteristic in white table wines. Winemaking practices that increase
the concentration of arabinogalactan protein and low molecular mass mannoprotein could assist in masking palate hotness in
white wine. Management of pH and alcohol can also be used to significantly influence the taste and mouthfeel in white wine.

Keywords: ethanol, mouthfeel, phenolic substances, polysaccharides, white wine

Introduction Polysaccharides comprise a diverse family of compounds in


The style of dry white wine is influenced by the flavour inten- terms of their molecular mass (MM) and monosaccharide com-
sity, acidity and mouthfeel attributes of the wine. Wine mouth- position, and have previously been grouped based on their
feel is ill-defined but most likely encompasses the tactile and monosaccharide structure and their source in wine.
chemosensory attributes of viscosity, astringency/dryness and Arabinogalactan proteins (AGPs) are characterised by a high
hotness (Pickering and Demiglio 2008). proportion of arabinose and galactose monosaccharide residues
Perceived viscosity is a primary contributor to the concept and are associated with a hydroxyproline-rich protein core
of white wine body (Skogerson et al. 2009) and has been asso- (Fincher et al. 1983). They are localised in a highly soluble form
ciated with the concentration of phenolic substances (PS) within the grape intracellular spaces and are extracted into
(Cejudo-Bastante et al. 2011) and higher wine pH (Gawel wine during the processing of grapes into juice and during the
et al. 2014). Neither ethanol nor glycerol, which on a concen- early stages of fermentation (Guadalupe and Ayestaran
tration basis are significant aspects of the wine matrix, contrib- 2007). Mannoprotein polysaccharides are extracted from yeast
ute significantly to the perception of white wine viscosity cell walls either during alcoholic fermentation or later by enzy-
despite both being viscous in their pure form (Nurgel and Pick- matic action during wine contact with yeast lees (Escot et al.
ering 2005, Gawel and Waters 2008). 2001). They present a broad molecular mass distribution repre-
While some mouthfeel attributes such as perceived viscos- sented by several populations and are characterised by a high
ity are determinants of white wine body and hence overall style mannose residue content relative to that of other sugars
rather than quality, others such as perceived palate hotness and (Goncalves et al. 2002). Rhamnogalacturonan polysaccharides
bitterness are considered negative quality traits as they detract are released from pectins embedded within the grape cell wall
from palate ‘structure’ and complexity by masking other as- presumably during the period of white winemaking when
pects of mouthfeel and flavour. Both palate hotness and bitter- grape skins and grape cell solids are in contact with the juice
ness in white wine can in part be attributed to an excessive prior to fermentation. Rhamnogalacturonans are low MM
concentration of ethyl alcohol and PS (Fischer and Noble compared with that of the other polysaccharide classes
1994, King and Heymann 2013). (<20 kDa) and are characterised by a high proportion of
The impact of astringent/drying sensations on white wine galacturonic acid and rhamnose residues, and by the presence
quality remains controversial. While the existence of these at- of the diagnostic sugars fucose and apiose (Pellerin et al. 1996).
tributes in light bodied, delicate white wines most likely The diversity and complexity of wine polysaccharide struc-
detracts from their overall quality, in other styles, it is possible tures provides them with the capacity to form intermolecular
that they may contribute positively by enhancing the associations with other wine components responsible for wine
perception of the palate weight or body. The astringent/drying- mouthfeel either through hydrogen bonding or hydrophobic
like mouthfeel displayed by white wine results from a high interactions. For example, the perception of bitterness, astrin-
concentration of PS (Singleton et al. 1975) or from low wine gency and hotness caused by white wine PS may be modulated
pH (Lawless et al. 1996). following hydrophobic interactions between the protein core of

doi: 10.1111/ajgw.12222
© 2016 Australian Society of Viticulture and Oenology Inc.
Gawel et al. Effect of polysaccharides on white wine mouthfeel 351

some polysaccharides and the PS ring, or by hydrogen bonding wines were mixed with 20% mass/vol Amberlite FPX-66 resin
between monosaccharide residues on the polysaccharide and (Rohm and Hass, Sarl, Germany) and strirred for 2 h before be-
PS hydroxyl groups (Vidal et al. 2004a, Coupland and Hayes ing evaporated by 3/4 under vacuum at <40°C. Four volumes
2014). Furthermore, wine polysaccharides carry negative of cold ethanol (96% v/v, Tarac Industries, Nurioopta, SA, Aus-
charges at wine pH and are therefore highly hydrated (Vernhet tralia) were added, and after 16 h at 4°C, the precipitate was re-
et al. 1996). The partial immobilisation of water around moved by centrifugation. The precipitates were washed twice
the large polysaccharide molecules may disrupt the surround- with 96% v/v ethanol before being recovered by centrifuga-
ing local water–water and water–ethanol structures, which tion, dissolved in water and extensively dialysed against water
in turn could affect the viscosity of the solution and the per- [3.5 kDa molecular mass (MM) cutoff, Thermo Scientific,
ception of alcohol hotness (Chinachoti 1995). Furthermore, Rockford, IL, USA] for 36 h at 22°C before being freeze dried.
these changes to the bulk structure of aqueous ethanol are
likely to be influenced by pH via ionisation of carboxyl
groups on the uronic acid residues contained on AGPs and Base wine preparation. A high-PS Riesling wine (Grape
rhamnogalacturonans. source: Eden Valley, SA, Australia, alcohol 12.1% v/v, residual
Previous studies have reported the effects of polysaccha- sugar 0.4 g/L) was produced by fermenting 80 L of juice ob-
rides and their interactions with other wine components on tained from hard press (200 kPa) of grape skins obtained from
various mouthfeel attributes of model wine. Vidal et al. machine-harvested grapes. The grape solids were removed
(2004b) considered the effect of polysaccharide fractions and from the pressings juice by settling under gravity at 4°C for
polymerised PS extracted from red wine on the fullness, astrin- 3 days with the aid of pectic enzyme preparation (3 mL/hL)
gency and bitterness of a model wine, while Jones et al. (2008) (Ultrazyme CPL, Novozyme) and fermented at 15°C with Sac-
assessed the influence of white wine polysaccharides and white charomyces cerevisiae yeast (EC1118 strain: Lallemand, Montréal,
wine proteins on the fullness and hotness of a model system QC, Canada). The finished wine was bentonite fined (1 g/L) to
representing a white wine. Neither of these studies, however, remove proteins and cold stabilised at 4°C. A corresponding
considered the effects of monomeric PS typical of white wine, low PS wine was produced by passing the high PS Riesling wine
nor of wine pH, both of which have more recently been shown through a column (2.5 cm × 50 cm) of Amberlite FPX-66 resin
to significantly affect the mouthfeel of white wine (Gawel et al. (Rhom and Hass). The polysaccharide concentration of the
2013, 2014). In this study, we investigate the interactive effects high and low PS base wines was determined gravimetrically af-
of realistic levels of white wine polysaccharides and polysaccha- ter extraction and purification as described earlier. The concen-
ride fractions of wine on the taste and mouthfeel of wine of tration of PS of the two base wines was assessed using
varying concentration of PS and alcohol and of pH. absorbance at 280 nm (Table 1), and their PS composition
was determined using a previously described HPLC method
Methods (Gawel et al. 2014).
We designed and executed two studies: the first investigated
the effect of white wine polysaccharides on the taste and
mouthfeel of white wine of variable concentration of PS; and Sensory assessment. Eight tasters (five male and three fe-
the second study considered the effect of polysaccharide molec- male, mean age 39) with experience and previous training in
ular mass fractions and their interaction with pH and alcohol assessing overall flavour and texture of white wines were used.
concentration on the taste and texture of model white wine. The tasters were trained for 3 h in the concepts of perceived vis-
cosity, bitterness, astringency and palate hotness using com-
Study 1. Interaction between polysaccharides and PS in white mercial wines of the cultivars Chardonnay, Riesling, Fiano,
wine Pinot Gris and Vermentino that ranged in alcohol concentra-
Polysaccharide extraction. Polysaccharides were extracted tion from 11.6 to 14.2% v/v and contained a fourfold difference
from two commercial dry white wines (Riesling, Watervale, in PS. The concept of palate hotness was further reinforced by
SA, Australia and Chardonnay, McLaren Vale, SA, Australia). exposing tasters to a 12% v/v alcohol white wine and to the
Both wines were produced from grape juices that had been same wine that had been fortified with 0.5 and 1.0% v/v etha-
clarified prior to fermentation with pectic enzyme preparations nol. Furthermore, they compared the viscosity, bitterness and
(3 mL/hL) (Ultrazyme CPL, Novozyme, Bagsvaerd, Denmark) astringency of the control wine with the same wine to which
and had been rendered protein stable using bentonite fining 2 g/L of apple pectin (Sigma-Aldrich, St Louis, MO, USA) was
following fermentation. The Chardonnay wine had spent added for viscosity, 15 mg/L quinine sulfate (Sigma-Aldrich)
4 months on yeast lees in tank, while the Riesling wine was re- for bitterness and 200 mg/L of commercial tannin (Tanin
moved from its lees shortly after the end of fermentation. The Galalcool, Laffort, Bordeaux, France) for astringency.

Table 1. General composition of the wines tasted (Study 1).

Addition of polysaccharides pH TA (g/L) Polysaccharides (mg/L) Alcohol (v/v) Phenolic substances (A280)

Before
Low PS wine 3.40 4.4 188 12.0 0.57
High PS wine 3.35 6.3 192 12.0 1.97
After
Low PS wine 3.39 4.2 338 12.0 0.58
High PS wine 3.35 6.3 342 12.0 1.93

PS, phenolic substances; TA, titratable acidity.

© 2016 Australian Society of Viticulture and Oenology Inc.


352 Effect of polysaccharides on white wine mouthfeel Australian Journal of Grape and Wine Research 22, 350–357, 2016

Table 2. Descriptors and definitions used in the formal sensory descriptive P-10, MM = 12 200, P-20, MM = 23 700, P-50, MM = 48 000,
analysis of white wine polysaccharide fractions in model wine. P-100, MM = 100 000, P-200, MM = 186 000, P-400
MM = 380 000) (Shodex P82 kit, Showa Denko, Tokyo,
Attribute Definition Japan). Fractions from individual runs were combined into
three fractions based on their MM distribution and on the
Flavour Intensity of overall flavour
criteria that 75% of the total peak area belonged to one of
Sweet Intensity of the taste of sucrose
the following: high (HMM) (>93 kDa, peak 234 kDa), me-
Acidity Intensity of acid taste perceived in the mouth or after
dium (MMM) (13–93 kDa, peak 31 kDa) and low (LMM)
expectorating.
(5–13 kDa, peak 8.9 kDa). The consolidated fractions were
Viscosity Perception of the body, weight or thickness of the wine
desalted by dialysing against water for 36 h (3.5 kDa MM
in the mouth. Low = watery, thin mouth feel.
cutoff, Thermo Scientific), freeze dried and stored at 20°C.
High = oily, thick mouth feel.
The monosaccharide composition of the fractions was deter-
Astringency Intensity of the drying and mouth-puckering sensation
mined by the method of Ruiz-Garcia et al. (2014). Briefly,
in the mouth. Low = coating teeth; Medium = mouth
polysaccharides were hydrolysed with concentrated H2SO4
coating and drying; High = puckering, lasting
and derivatised with 1-phenyl-3-methyl-5-pyrazolone
astringency.
(PMP). The PMP-monosaccharide derivatives were quantified
Hotness Intensity of heat perceived in the mouth: Low = warm;
by HPLC (C18 Kinetex, 2.6 μm, 100 Å, 100 × 3.0 mm) using
High = hot. Including hot aftertaste and the burning
commercial standards (Sigma-Aldrich). The protein concen-
sensation in the mouth or after expectorating.
tration of the fractions was determined using a fluorescence-
Bitter Intensity of bitter taste perceived in the mouth or after
based method as per the manufacturers’ instructions (EZQ
expectorating.
protein quantitation kit, Invitrogen, Mt Waverley, Vic.,
Australia). The contribution of each fraction to the total poly-
saccharides was determined gravimetrically.
Following the discussion, the tasters agreed on the definition of
the attributes given in Table 2. A practice session whereby
tasters rated the intensity of the attributes of six of the treat- Model wine preparation. Four model wines (pH 3.3 and
ments and of the two controls on the same 15 cm partially 3.6 × alcohol 11.5 and 13.5% v/v) were prepared by
structured line scale used in the formal tasting completed the adjusting the ethanol concentration and pH of a saturated
training regime. aqueous solution of potassium bitartrate (Sigma-Aldrich)
One hour prior to tasting, 150 mg/L of each of the Riesling with the addition of 96% v/v food grade ethanol (Tarac
and Chardonnay polysaccharides were dissolved in the high Industries) and saturated tartaric acid (Ajax Finechem,
and low PS base wines (Table 1). Samples were presented in Sydney, NSW, Australia). The model wines were flavoured
30 mL aliquots in 3-digit-coded, covered, ISO standard wine by adding the compound mixture representing a cool climate
glasses at 22–24°C, in isolated booths under sodium lighting. Chardonnay wine-like flavour (Jones et al. 2008) plus an
A replicate set of the treatment combinations and controls with additional 50 mg/L of isoamyl acetate to increase ‘ripe fruity’
no added polysaccharides were presented in each of four tast- character. The LMM, MMM and HMM polysaccharide frac-
ing sessions. The tasting order was randomised. A water tions (150 mg/L) were dissolved in the model wines 2 h prior
mouthwash and 90 s rest period was enforced between sam- to sensory analysis.
ples, and a 30 min rest period between every six samples. The
perceived astringency, bitterness, hotness and viscosity of the Sensory assessment. Twelve tasters (four male, eight fe-
samples were rated on a 15 cm partially structured line scale male, mean age 39) were used. Eight tasters were the same as
with the word anchors ‘low’ and ‘high’ at 1.5 cm and 13.5 cm those used in Study 1, and the remaining four were experi-
points, respectively. enced in taste and texture profiling of white wine. Taste stan-
dards were used in three training sessions to familiarise tasters
Study 2. Interaction between polysaccharide fractions and pH with the attributes of interest [1 g/L tartaric acid (Ajax
and alcohol in model wine Finechem) for acidity, 15 mg/L quinine sulfate for bitterness,
Polysaccharide extraction and fractionation. A 2010 0.5 g/L aluminium potassium sulfate (Ajax Finechem) for as-
Chardonnay wine (20 L) from the Barossa Valley region was tringency, 2 g/L apple pectin (Sigma-Aldrich) for viscosity and
stripped of its proteins using a strong cation exchange resin 15.5% v/v aqueous ethanol (Tarac Industries) for hotness]. In
(Macro-Prep High S resin, Bio-Rad, Hercules, CA, USA) and addition to familiarisation of attributes, the tasters also prac-
PS (Amberlite FPX-66 resin, Rohm and Hass) as described by ticed rating the intensity of the attributes of several test sam-
Van Sluyter et al. (2009), and the polysaccharides extracted as ples. A subset of the test samples was also rated in duplicate
described earlier. The polysaccharides were fractionated on on the agreed attributes using the same scaling method used
the basis of MM: 200 mg of polysaccharides were dissolved in the formal taste sessions as a final practice prior to the formal
in 0.1 mol/L NaNO3 (Sigma-Aldrich) and loaded onto a evaluation of the test samples.
900 × 25 mm column packed with Sephacryl 300 resin The set of treatments was tasted in triplicate using a
(Pharmacia, Uppsala, Sweden), and eluted with 0.1 mol/L completely randomised design over four sessions conducted
NaNO3 at 1 mL/min, with 10 mL fractions collected from 2.5 1 week apart. Twelve samples were tasted in each session with
to 7 h. The process was repeated 20 times to obtain sufficient an enforced 90s rest between samples and a 15min rest every
quantity for sensory assessment. Every second fraction col- third sample. The tasting conditions and intensity scale used
lected during each run was analysed by size exclusion chroma- were the same as those used in Study 1.
tography using a Biosep SEC-2000 300 × 7.8 mm column
(Phenomenex, Torrance, CA, USA) (20 μL injection, 1 mL/min, Data collection and analysis
0.1 mol/L NaNO3 mobile phase, 40°C). Molecular mass calibra- Sensory data were acquired with Fizz sensory software, version
tion was conducted with Pullulan standards (P-5, MM = 5800, 2.46 (Biosystèmes, Couternon, France). Panel performance

© 2016 Australian Society of Viticulture and Oenology Inc.


Gawel et al. Effect of polysaccharides on white wine mouthfeel 353

was assessed using Fizz, Senstools (OP&P, Utrecht, The


Netherlands) and PanelCheck (Nofima Mat, Ås, Norway) soft-
ware. The tasters employed in both studies were found to be
performing to an acceptable standard. Attribute ratings were
analysed using ANOVA with tasters considered to be random
factors and all other variables fixed factors. In Study 1 where
all treatments were assessed together with their respective con-
trols, a two-sample t-test was applied to the attribute ratings.
Analyses were conducted using MINITAB 14.13 (Minitab, State
College, PA, USA).

Results and discussion

Polysaccharide and base wine composition


Study 1. Spectroscopic analysis showed that treating the high
PS base wine with FPX-66 resin was effective in producing a Figure 2. Size exclusion chromatogram of the three polysaccharide fractions,
low PS wine (Table 1). The resin treatment resulted in a signif- high molecular mass (HMM, >93 kDa) ( ), medium MM (MMM,
icant reduction in all classes of PS (hydroxycinnamic acids, 13–93 kDa) ( ) and low MM (LMM, 5–13 kDa): ( ) and of the extracted
flavanols, flavonols and flavonol glycosides) (HPLC analysis – polysaccharides prior to fractionation (Study 2) ( )
data not shown). The treatment was less effective in removing
the highly hydrophilic hydroxybenzoic acids. The molecular Sensory effects
mass distribution of the two polysaccharide preparations used Thickness/viscosity/body. The addition of polysaccharides
is given in Figure 1 and is consistent with that reported by Jones at a concentration typical of that of white wine resulted in a
et al. (2008). small increase in perceived viscosity with the effect being more
pronounced in the wine containing the higher concentration of
Study 2. The polysaccharide fractions were separated based PS (Figure 4). Okuda et al. (2007) also found that the perceived
on their molecular mass (Figure 2). The monosaccharide pro- ‘body’ of Koshu white wine was strongly correlated with its
files of the fractions (Figure 3) indicated that the HMM fraction concentration of polysaccharides and PS. Increases in solution
(MM > 93 kDa), which comprised 45% of the polysaccharides viscosity is contingent upon the concentration of polysaccha-
by mass, consisted primarily of mannoproteins, as mannans rides being sufficiently high as to enable the formation of
comprised 90% of the total residues. The MMM fraction (MM entangled molecular networks (Hollowood et al. 2002). The
13–93 kDa) which comprised 34% of the total by mass, was observed increase in perceived viscosity due to polysaccharide
more diverse in its monosaccharide composition with man- addition is unlikely to be due to this phenomenon as only a
nose, galactose and arabinose being the most predominant, low concentration of polysaccharides was tasted. Furthermore,
suggesting it was a mixture of smaller MM mannoproteins as all the predominant wine polysaccharides carry negative
and AGPs. The LMM fraction (5–13 kDa) had a similar mono- charges at wine pH (Vernhet et al. 1996), polymer interaction
saccharide profile to that of the MMM fraction but also is likely to be suppressed due to electrostatic repulsion. A signif-
contained the monosaccharide fucose, which is specific to RG- icant interactive effect, however, between the polysaccharide
2. This fraction made up 21% of the polysaccharides by mass. fraction and pH on perceived viscosity was found (Study 2).
The LMM distribution of this fraction suggests that the poly- At low wine pH (3.3), polysaccharides had no significant effect
saccharides could be subunits of larger mannoproteins and on perceived viscosity, but at high pH (3.6), the MMM fraction
AGPs caused by enzymatic activity arising from the action caused a significant increase in perceived viscosity (Table 3,
of exogenous pectic enzyme preparations used during the Figure 5). Using a model wine of pH 3.6, Vidal et al. (2004b)
juice clarification stage of winemaking or by acid hydrolysis. showed that a mixture of mannoproteins and AGP, like the
The protein concentration of the fractions was negligible,
ranging from 0.11 to 0.25% m/m.

Figure 3. Monosaccharide composition of the three polysaccharide fractions


Figure 1. Size exclusion chromatogram of the polysaccharides added to the isolated and used in Study 2. Mannose ( ), rhamnose ( ), glucuronic acid ( ),
Chardonnay ( ) and Riesling ( ) wines (Study 1). galacturonic acid ( ), glucose ( ), galactose ( ), arabinose ( ) and fucose ( ).

© 2016 Australian Society of Viticulture and Oenology Inc.


354 Effect of polysaccharides on white wine mouthfeel Australian Journal of Grape and Wine Research 22, 350–357, 2016

Figure 4. Effect of adding white wine polysaccharides (150 mg/L) on the taste
and mouthfeel relative to that of control white wines containing high ( ) and low
( ) concentration of phenolic substances (n = 2 polysaccharides) (Study 1).
Error bars represent two standard errors.

MMM fraction used in this study, also increased the perceived


viscosity of model wine. The monosaccharide composition
and size distribution of the MMM fraction strongly suggest that
it contained arabinogalactan proteins, which are generally neg-
atively charged at wine pH (Vernhet et al. 1996). At high pH,
these would be more ionised favouring hydration by free un-
bound water. This structure-forming activity could explain
the small increase in perceived viscosity. It is noteworthy that
the LMM fraction did not affect changes to perceived viscosity Figure 5. Effect of the addition of 150 mg/L of high (>93 kDa), medium
in the same way as the MMM fraction despite both containing (13–93 kDa) and low (5–13 kDa) molecular mass (MM) polysaccharide fractions
uronic acid residues. The difference in viscosity effects resulting on the mean sensory intensity ratings of model wine of (a) pH 3.6 (■) and 3.3
(●) and (b) alcohol concentration of 13.5% (▲) and 11.5% (♦) (Study 2).
from the addition of these two fractions may be attributed to Error bars represent two standard errors. * indicates it is significantly different
the relative size of both polysaccharides. Electrostatic effects from the control.
are expected to be more diffuse in larger polysaccharides, so
consequently the water structure surrounding them is domi- variation in the perceived viscosity of those white wines could
nated by water–water hydrogen bonding in the first surround- also be attributed to a difference in pH. Similarly, Runnebaum
ing hydration shell. Conversely, smaller charged ions with a et al. (2011) found a strong positive association between lactic
weak charge density are known to be net disruptors of water acid concentration and the perceived ‘body’ of commercial
structure (Nose et al. 2004), which may explain the lower vis- white wines. Lactic acid in wine arises from the conversion of
cosity produced by the LMM fraction at the higher pH. the stronger malic acid during malolactic fermentation by lactic
The greatest effect on perceived viscosity was that of pH it- acid bacteria and as such is accompanied by an increase in pH.
self (P < 0.001) (Figure 5a). The pH of the model wines used The concept that modification to the hydrogen-bonded net-
in this study were adjusted by the addition of tartaric acid to work of water molecules (water structure), rather than molec-
the potassium salt of tartaric acid replicating the major acid ular entanglement, may be responsible for the perception of
equilibrium in wine. Using white wines made from the same higher viscosity is strengthened by the observation that the
juice but prepared with different juice extraction and handling higher pH model wine was perceived to be significantly more
methods, Gawel et al. (2014) reported that most of the viscous than that of the lower pH model wine regardless of

Table 3. P values of main effects and interactions of polysaccharide type, alcohol concentration and pH (Study 2).

Main effects and interactions Astringency Viscosity Hotness Bitterness Acidity Aroma Flavour

Main effects
Polysaccharide type 0.646 0.216 0.037 0.471 0.746 0.826 0.086
Alcohol 0.004 0.359 <0.001 0.001 <0.001 0.707 0.543
pH <0.001 <0.001 0.104 <0.001 <0.001 0.199 <0.001
Interactions
Polysaccharide type*Alc 0.137 0.072 0.438 0.878 0.983 0.663 0.542
Polysaccharide type*pH 0.985 0.039 0.506 0.542 0.661 0.347 0.927
Alcohol*pH 0.129 0.449 0.685 0.971 0.297 0.636 0.735
Polysaccharide type*Alc*pH 0.596 0.110 0.199 0.300 0.450 0.414 0.575

Bold indicates significance P < 0.05.

© 2016 Australian Society of Viticulture and Oenology Inc.


Gawel et al. Effect of polysaccharides on white wine mouthfeel 355

the presence of polysaccharide fractions. The dicarboxyl ion of these polymerised PS compared with the simple monomeric
tartaric acid is a much stronger water structure maker than that PS of white wine may explain the apparent contradictory find-
of its corresponding carboxyl group (Nose et al. 2004, ings with regard to the effect of polysaccharides on bitterness
Chmielewska et al. 2007) which could explain why the model produced by PS. Indeed, Guadalupe and Ayestaran (2008)
wine was more viscous at higher pH. found that mannoprotein addition did not modify the concen-
tration of free monomeric PS in red wine, which is consistent
with the suggestion that the small number of hydroxyl binding
Hotness. The addition of 150 mg/L of polysaccharides sites on monomeric PS limit their ability to bind to polysaccha-
reduced the perceived palate hotness of both the high and rides. An alternative hypothesis is that the non-bitter PS that
low PS Riesling wines (Figure 4). Consistent with this finding, comprise the majority of the pool of PS in white wine may out-
the addition of LMM and MMM polysaccharide fractions compete bitter-tasting PS for binding sites on polysaccharides.
at the same concentration also reduced the perception of The recent finding that flavanols, which are known bitter
alcohol hotness in model wine (Figure 5b). The effect of principles in white wine, have an adsorption capacity similar
polysaccharides on hotness was independent of alcohol to that of the far more abundant and non-bitter-tasting
concentration (Table 3). The addition of a similar concentra- hydroxycinnamics for the polysaccharide β-glucan supports
tion of polysaccharides to model wine by Jones et al. (2008) this hypothesis (Wang et al. 2013). Polysaccharide fractions
did not affect its perceived hotness. The ineffectiveness of also did not significantly affect the bitterness elicited by alcohol
polysaccharides, however, to reduce hotness in this instance in the model wines (Table 3), a result consistent with Jones et al.
may have been due to the elevated concentration of glycerol (2008) who also found no significant effect of the addition of
and proteins that were also added and which interacted sig- polysaccharides on model wine bitterness.
nificantly with polysaccharides with respect to the perception It is not possible to assess the interactive effect of polysac-
of hotness. Sulfur dioxide, which has a trigeminal component, charides and PS on white wine astringency as despite the large
was also added in this earlier study, which also may have differences in the concentration of PS between the low and
masked the perception of palate hotness. The perception of eth- high PS wines, they were found not to differ significantly in as-
anol hotness is thought to involve stimulation of free nerve tringency (Figure S1), suggesting that PS were not responsible
endings that extend into the surface of the oral mucosa. Etha- for the perception of astringency. In Study 2 where the model
nol via its hydroxyl group has significant potential for hydro- wines did not contain PS, the intensity of perceived acidity
gen bonding with polysaccharides, which could explain the and astringency was strongly correlated (r = 0.88, P < 0.001)
reduction in alcohol hotness in the presence of polysaccharides. suggesting that the astringent sensation was most likely
The finding that polysaccharides reduce alcohol-induced confused with the drying sensation elicited by organic acids
hotness is of practical significance to white wine quality. (Lawless et al. 1996). The addition of polysaccharide fractions
Greater flavour intensity in wine is often achieved by harvest- did not affect the intensity of the drying sensation in these
ing grapes at a greater maturity, but this also results in wines model wines (Table 3). Others using RG-2 at a concentration
of higher alcoholic strength and enhanced perceived hotness. an order of magnitude greater than that of white wine
Therefore, any increase in wine quality resulting from greater (500 mg/L) found that it can reduce the intensity of the drying
flavour intensity can be offset by increased palate hotness and sensation produced in model wine containing only organic
a corresponding loss of palate structure. acids and alcohol (Vidal et al. 2004b).
The perceived acidity and sweetness of the white wines
(Study 1) and the acidity of the model wines (Study 2) were
Astringency, bitterness and acidity. Polysaccharides did unaffected by addition of polysaccharides (Figure 4, Table 3),
not significantly reduce the bitterness or astringency of the high which suggests that despite their saccharine nature, polysac-
PS white wine (Study 1). Low molecular mass PS found in charides are unable to interact with sweetness receptors, nor
white wines have been associated with bitterness and astrin- do they mask wine acidity.
gency (Singleton et al. 1975, Hufnagel and Hofmann 2008).
Polysaccharides are capable of inhibiting salivary protein–
polymerised PS interactions responsible for red wine astrin- Aroma and flavour. The addition of polysaccharides or
gency either by encapsulating and forming ternary complexes polysaccharide fractions (Figure 4, Table 3) did not affect the
with proteins and polymerised PS (Luck et al. 1994) or by overall perceived intensity of aroma or flavour in either the real
forming polymerised PS–polysaccharide complexes, which in- or flavoured model wines. Many studies have reported reduced
terfere with the ability of polymerised PS to interact with sali- volatility of wine flavour compounds in the presence of wine
vary proteins (Poncet-Legrand et al. 2007). Polysaccharides polysaccharides, which has been attributed to lessened solva-
may also modulate bitterness perception. All the major classes tion and an associated increase in hydrophobic interaction
of PS found in white wine are capable of forming hydrogen- between the hydrophobic aroma compounds to the hydropho-
bonded complexes with polysaccharides (Wang et al. 2013). bic core of proteoglycans (Mitropoulou et al. 2011). The volatil-
The formation of these large complexes may sterically hinder ity of aroma and flavour compounds, however, appear only to
the interaction of bitter-tasting PS groups with taste receptor be reduced at a high concentration of polysaccharides and spe-
sites. Despite the ability of polysaccharides to form complexes cifically at the concentration (known as c*) at which polysac-
with bitter-tasting PS, the addition of polysaccharides did not charides begin to form entangled polymer networks that
reduce the bitterness of the high PS white wine used in this appear as viscous fluids (Hollowood et al. 2002). The c* for
study (Figure 4). This may be explained by the possible exis- wine polysaccharides is unknown, however, it is unlikely that
tence of non-PS bitter compounds that are incapable of forming networks were formed as previous studies involving structur-
complexes with polysaccharides at wine pH (Fabre et al. 2014). ally diverse polysaccharides demonstrated that entanglement
Vidal et al. (2004b) found that the bitterness of polymeric typically occurs at a concentration order of magnitude higher
flavanols extracted from grape seeds was inhibited by a mixture than that used in this study (Gonzales-Tomas et al. 2004,
of wine mannoproteins and AGPs. The multidentate nature of Garrec and Norton 2012).

© 2016 Australian Society of Viticulture and Oenology Inc.


356 Effect of polysaccharides on white wine mouthfeel Australian Journal of Grape and Wine Research 22, 350–357, 2016

General discussion authors also thank the AWRI sensory team for the training of
The pH and alcohol concentration of the model wine signifi- the expert tasters and the conduct of the sensory trials.
cantly affected the majority of its taste and mouthfeel attri-
butes. Higher pH model wines were perceived to be References
significantly more bitter and viscous and less astringent, while Cejudo-Bastante, M.J., Castro-Vázquez, L., Hermosín-Gutiérrez, I.
higher alcohol wines were perceived to be significantly more and Pérez-Coello, M.S. (2011) Combined effects of prefermentative
bitter, astringent and hot (Table 3). These effects are well skin maceration and oxygen addition of must on color-related phe-
known and, in addition to those already discussed, can be ex- nolics, volatile composition, and sensory characteristics of Airén
white wine. Journal of Agricultural and Food Chemistry 59,
plained by masking effects by acidity on bitterness (Moskowitz 12171–12182.
1971) and by the inherent bitterness, hotness and drying na- Chinachoti, P. (1995) Carbohydrates – functionality in foods. American
ture of alcohol (Gawel et al. 2013). In contrast, polysaccharides Journal of Clinical Nutrition 61, 922–929.
produced a significant effect on palate hotness and viscosity Chmielewska, A., Wypych-Stasiewicz, A. and Bald, A. (2007) Visco-
metric studies of aqueous solutions of monosodium salts of dicarbox-
only at higher pH, which suggests that alcohol level and pH
ylic acids. Journal of Molecular Liquids 136, 11–14.
were more influential in affecting the overall taste and mouth- Coupland, J.N. and Hayes, J.E. (2014) Physical approaches to masking
feel of model wines than were polysaccharides. Similarly, only bitter taste: lessons from food and pharmaceuticals. Pharmaceutical
slightly suppressive effects on most taste and mouthfeel attri- Research 31, 2921–2939.
butes were observed following the addition of polysaccharides Escot, S., Feuillat, M., Dulau, L. and Charpentier, C. (2001) Release of
polysaccharides by yeasts and the influence of released polysaccha-
to white wine (Study 1).
rides on colour stability and wine astringency. Australian Journal of
The relative influence of wine matrix composition and of Grape and Wine Research 7, 146–152.
polysaccharides has significant winemaking implications. The Fabre, S., Absalon, C., Pinaud, N., Venencie, C., Teissedre, P.L., Fou-
pH and alcohol concentration of a wine can be strongly influ- quet, E. and Pianet, I. (2014) Isolation, characterization, and determi-
enced using several standard winemaking strategies, including nation of a new compound in red wine. Analytical and Bioanalytical
Chemistry 406, 1201–1208.
the selection of harvest date, or by adding sugar, acid or water Fincher, G.B., Stone, B.A. and Clarke, A.E. (1983) Arabinogalactan-
to the juice, or by promoting malolactic fermentation. In con- proteins: structure, biosynthesis and function. Annual Review of
trast, achieving a significant increase in polysaccharide concen- Plant Physiology 34, 47–70.
tration in wine appears more difficult. White juice extraction Fischer, U. and Noble, A.C. (1994) The effect of ethanol, catechin con-
methods, including pre-fermentation skin maceration, skin centration, and pH on sourness and bitterness of wine. American
Journal of Enology and Viticulture 45, 6–10.
and whole-bunch pressing, yielded differences only in the con- Garrec, D.A. and Norton, I.T. (2012) The influence of hydrocolloid hy-
centration of polysaccharides of the order of 15% (Gawel et al. drodynamics on lubrication. Food Hydrocolloids 26, 389–397.
2014). Although further increases in polysaccharides via higher Gawel, R. and Waters, E.J. (2008) The effect of glycerol on the per-
mannoprotein concentration can be achieved later in the ceived viscosity of dry white table wine. Journal of Wine Research
winemaking process by maintaining wine on yeast lees (Escot 19, 109–114.
Gawel, R., Van Sluyter, S.C., Smith, P.A. and Waters, E.J. (2013) Effect
et al. 2001), the effect of increased mannoproteins on mouth- of pH and alcohol on perception of phenolic character in white wine.
feel is unclear as they were present in all the fractions tested. American Journal of Enology and Viticulture 64, 425–429.
The 13–93 kDa mannoproteins, however, are of particular in- Gawel, R., Day, M., Van Sluyter, S.C., Holt, H., Waters, E.J. and Smith,
terest as they partly comprised the fraction that most consis- P.A. (2014) White wine taste and mouthfeel as affected by juice ex-
tently reduced palate hotness. The other component of the traction and processing. Journal of Agricultural and Food Chemistry
62, 10008–10014.
sensorially significant MMM fraction were AGPs. In red wine, Goncalves, F., Heyraud, A., Norberta de Pinho, M. and Rinaudo, M.
these polysaccharides are quickly extracted during the early (2002) Characterization of white wine mannoproteins. Journal of Ag-
stages of fermentation, but then reduce significantly thereafter ricultural and Food Chemistry 50, 6097–6101.
(Guadalupe and Ayestaran 2007). The reason for the loss of Gonzales-Tomas, L., Carbonell, I. and Costell, E. (2004) Influence of
AGPs during fermentation is unclear but may involve a type, concentration and flow behaviour of hydrocolloid solutions on
aroma perception. European Food Research and Technology 218,
fining-like effect with tannins or by adsorption onto yeast cell 248–252.
walls. Whether the concentration of AGP follows the same pat- Guadalupe, Z. and Ayestaran, B. (2007) Polysaccharide profile and con-
tern during white wine ferments is unknown, but an under- tent during the vinification and aging of Tempranillo red wines. Jour-
standing of the dynamics of AGP during fermentation could nal of Agricultural and Food Chemistry 55, 10720–10728.
lead to better retention of AGPs in wine and a corresponding Guadalupe, Z. and Ayestaran, B. (2008) Effect of commercial
mannoprotein on polysaccharide, polyphenolic, and color composi-
decrease in palate hotness. tion in red wines. Journal of Agricultural and Food Chemistry 56,
9022–9029.
Hollowood, T.A., Linforth, R.S.T. and Taylor, A.J. (2002) The effect
Conclusion of viscosity on the perception of flavour. Chemical Senses 27,
Winemaking practices that affect the pH and alcohol concen- 583–591.
tration of white wine can significantly influence mouthfeel, Hufnagel, J.C. and Hofmann, T. (2008) Orosensory-directed iden-
and those that enhance the concentration of MMM tification of astringent mouthfeel and bitter-tasting compounds
arabinogalactan proteins and MMM mannoproteins could be in red wine. Journal of Agricultural and Food Chemistry 56,
1376–1386.
used to mask negatively perceived hotness derived from etha- Jones, P.R., Gawel, R., Francis, I.L. and Waters, E.J. (2008) The influ-
nol and increase the viscosity of higher pH white wines. The ence of interactions between major white wine components on the
effect of polysaccharides on white wine mouthfeel and taste aroma, flavour and texture of model wine. Food Quality and Prefer-
was small compared with that of the wine matrix components ence 19, 596–607.
of alcohol and pH. King, E. and Heymann, H. (2013) The effect of reduced alcohol on the
sensory profiles and consumer preferences of white wine. Journal of
Sensory Studies 29, 33–42.
Acknowledgements Lawless, H.T., Horne, J. and Giasi, P. (1996) Astringency of organic acids
is related to pH. Chemical Senses 21, 397–403.
This work was supported by Australia’s grapegrowers and Luck, G., Liao, H., Murray, N.J., Grimmer, H.R., Warminiski, E.E. and
winemakers through their investment body Wine Australia, Williamson, M.P. (1994) Polyphenols, astringency and proline-rich
with matching funds from the Australian Government. The proteins. Phytochemistry 37, 357–371.

© 2016 Australian Society of Viticulture and Oenology Inc.


Gawel et al. Effect of polysaccharides on white wine mouthfeel 357

Mitropoulou, A., Hatzidimitriou, E. and Paraskevopoulou, A. (2011) Van Sluyter, S.C., Marangon, M., Stranks, S.D., Neilson, K.A.,
Aroma release of a model wine solution as influenced by the presence Hayasaka, Y., Haynes, P.A., Menz, R.I. and Waters, E.J. (2009) Two-
of non-volatile components. Effect of commercial tannin extracts, step purification of pathogenesis-related proteins from grape juice
polysaccharides and artificial saliva. Food Research International 44, and crystallization of thaumatin-like proteins. Journal of Agricultural
1561–1570. and Food Chemistry 57, 11376–11382.
Moskowitz, H.R. (1971) Intensity scales for pure tastes and for taste Vernhet, A., Pellerin, P., Prieur, C., Osmianski, J. and Moutounet, M.
mixtures. Perception and Psychophysics 9, 51–56. (1996) Charge properties of some grape and wine polysaccharide
Nose, A., Hojo, M. and Ueda, T. (2004) Effects of salts, acids, and phe- and polyphenolic fractions. American Journal of Enology and Viticul-
nols on the hydrogen-bonding structure of water-ethanol mixtures. ture 47, 25–29.
Journal of Physical Chemistry B 108, 798–804. Vidal, S., Francis, L., Williams, P., Kwiatkowski, M., Gawel, R.,
Nurgel, C. and Pickering, G.J. (2005) Contribution of glycerol, ethanol Cheynier, V. and Waters, E. (2004a) The mouth-feel properties of
and sugar to the perception of viscosity and density elicited by model polysaccharides and anthocyanins in a wine like medium. Food
wines. Journal of Texture Studies 36, 303–325. Chemistry 85, 519–525.
Okuda, T., Fukui, M., Hisamoto, M., Iino, S., Hikawa, Y., Ogino, S., Vidal, S., Courcoux, P., Francis, L., Kwiatkowski, M., Gawel, R.,
Takayanagi, T. and Yokotsuka, K. (2007) Relationship between Williams, P., Waters, E. and Cheynier, V. (2004b) Use of an ex-
macromolecules and thickness in dry white wine. Journal of perimental design approach for evaluation of key wine compo-
the American Society of Enology and Viticulture (Japan) nents on mouth-feel perception. Food Quality and Preference
18, 15–21. 15, 209–217.
Pellerin, P., Doco, T., Vidal, S., Williams, P., Brillouet, J.M. and Wang, Y., Liu, J., Chen, F. and Zhao, G. (2013) Effect of molecu-
O’Neill, M.A. (1996) Structural characterization of red wine lar structure of polyphenols on their noncovalent interactions
rhamnogalacturonan II. Carbohydrate Research 190, 183–197. with oat β-glucan. Journal of Agricultural and Food Chemistry
61, 4533–4538.
Pickering, G.J. and Demiglio, P. (2008) The white wine mouthfeel
wheel: a lexicon for describing the oral sensations elicited by white
wine. Journal of Wine Research 19, 51–67.
Poncet-Legrand, C., Doco, T., Williams, P. and Vernhet, A. (2007) Inhi- Manuscript received: 31 August 2015
bition of grape seed tannin aggregation by wine mannoproteins: ef-
fect of polysaccharide molecular weight. American Journal of Revised manuscript received: 16 October 2015
Enology and Viticulture 58, 87–91. Accepted: 20 October 2015
Ruiz-Garcia, Y., Smith, P.A. and Bindon, K.A. (2014) Selective extrac-
tion of polysaccharide affects the adsorption of proanthocyanidin by
grape cell walls. Carbohydrate Polymers 114, 102–114.
Runnebaum, R.C., Boulton, R.B., Powell, R.L. and Heymann, H. (2011) Supporting information
Key constituents affecting wine body–an exploratory study. Journal
of Sensory Studies 26, 62–70.
Additional Supporting Information may be found in the online
Singleton, V.L., Sieberhagen, H.A., de Wet, P. and van Wyk, C.J. (1975) version of this article at the publisher’s web-site: http://
Composition and sensory qualities of wines prepared from white onlinelibrary.wiley.com/doi/10.1111/ajgw12222/abstract
grapes by fermentation with and without grape solids. American
Journal of Enology and Viticulture 26, 62–69.
Skogerson, K., Runnebaum, R., Wohlgemuth, G., de Ropp, J.,
Heymann, H. and Fiehn, O. (2009) Comparison of gas chromatogra- Figure S1. Mean sensory intensity ratings of high ( ) and low
phy coupled time-of-flight mass spectrometry and 1H nuclear mag-
netic resonance spectroscopy metabolite identification in white ( ) phenolic control wines (Study 1). Error bars represent two
wines from a sensory study investigating wine body. Journal of Agri- standard errors.
cultural and Food Chemistry 57, 6899–6907.

© 2016 Australian Society of Viticulture and Oenology Inc.


Chapter 8.

CONCLUDING REMARKS AND FUTURE DIRECTIONS

The overall aim of this thesis was to identify the aspects of white wine composition
that directly, or by interaction, influence white wine mouth-feel, and to gain insight
into the mechanisms that underpin mouth-feel perception in white wine. The outcomes
of this thesis have provided an understanding of the relative contribution to white wine
mouth-feel of the major classes of non-volatile compounds, including those
responsible for the basic wine matrix. Alcohol, pH and phenolics were found to
influence most aspects of white wine mouth-feel while the role of polysaccharides in
white wine appear to be mostly confined to moderating the negative quality attribute
of perceived hotness. Conversely, the effect of glycerol concentration on mouth-feel
perception of dry white wine appears negligible.

The findings have significant winemaking implications. The pH and alcohol level of
white wine are mostly determined by the maturity of the grape at harvest, but are
usually modified by winemaking practice to ensure that the wine is microbiologically
and chemically stable and has a balanced taste that meets the overall style specification
expected by the consumer. The results of the work reported in this thesis suggest that
modulating alcohol content and acidity can be used to vary the perception of white
wine style, i.e. light - full body (Gawel et al. 2007), and that the perception of viscosity
and the astringent/drying sensation in white wine can be substantially modified by
simply changing its pH (Gawel et al. 2013, Gawel et al. 2014a).

The possible enhancing role of glycerol in the perceived viscosity and body of white
wine has long been controversial. Glycerol is an important compound in white wine
quantitatively speaking and its pure form it is also clearly viscous. Early work
presented as part of this thesis (Gawel, et al. 2007) suggested that glycerol
concentrations at the upper end of concentrations found in white wine might affect
perceived viscosity of white wine. However, in a following study whereby the
confounding factor of sweetness was negated, glycerol was found not to affect the
perception of viscosity in white wine (Gawel and Waters 2008). Together with this
result and those of others using less direct methods (Noble and Bursick 1984), and the

97
knowledge that glycerol concentration in white wine cannot easily be varied beyond a
‘background level’ by winemaking practice strongly suggests that glycerol does not
affect the perceived viscosity of white wine. However, in response to consumer
demands, recently researchers have successfully produced ‘fruit forward’ wines from
riper grapes but with a lower alcohol content. The lower alcohol levels in these wines
have been achieved by redirecting carbon metabolism of the yeast Saccharomyces
cerevisiae from ethanol production to glycerol production (Goold et al. 2017). If these
findings translate to commercial practice, some future white wines will likely have
substantially higher glycerol concentrations than those currently produced so further
work on the effect of glycerol on white wine mouth-feel may be needed.

For over half a decade, the conventional wisdom amongst many makers of white wine
has been that low phenolic concentration in wine equates to greater purity of varietal
character and lower bitterness and ‘coarseness’ (Gawel et al. 2008). More recently
though, winemakers have considered the idea that phenolics could add to the
complexity of white wine by contributing to its mouth-feel (McLean 2005). The
eminent finding of this thesis that phenolics have an overarching effect on all major
aspects of white wine mouth-feel, affecting the perception of astringency, viscosity,
oiliness, hotness and bitterness supports this contemporary view. Furthermore, the
knowledge that total phenolic concentration and the phenolic profile of white wine can
be substantially modified by applying standard winemaking practice (Gawel et al.
2014a) indicates that winemakers can influence white wine mouth-feel by how they
manage phenolics.

While most winemakers know how to ‘dial up’ or ‘dial down’ total phenolic
concentration in their wines, and some are also aware of how they can influence the
phenolic makeup of their wines by their production decisions, knowledge as to how
they could specifically influence mouth-feel via phenolic management is lacking. Such
knowledge would present a powerful tool as it would allow winemakers to better
construct wines with desirable mouth-feels and modulate their overall style.

Practically, the demonstration that phenolic analogues of similar structure can be


antagonistic with respect to perception of some aspects of white wine mouth-feel
(Gawel et al. 2014b) draws into question the assumption that all phenolics are

98
necessarily perceptually additive, and consequently the usefulness of ‘total phenolic’
measures as predictors of coarseness and bitterness in white wine. In the work
presented here, white wine bitterness was found to be insensitive to the large variations
in total phenolic concentration created from the entire range of juice extraction
methods used in conventional white wine making using three cultivars (Gawel et al.
2014a). This supports the notion that ‘total phenolic’ concentration is only a fair
indicator of white wine bitterness, and that more specific analytical measures such as
flavanol and flavanone concentrations be further investigated as possible proxies for
white wine bitterness. The work also suggests that other non-phenolic compounds
could also be responsible for the bitterness perceived in some white wines.

From a theoretical perspective, the observation that two phenolic compounds that
shared a common phenolic moiety suppressed each other in eliciting astringency
(Gawel et al. 2014b) suggests that some form of competitive binding to receptors is
involved (Schobel et al. 2014). While it could be argued that the observed suppression
is simply a matter of two compounds adding along the same compressed
psychophysical function as occurs in taste (Bartoshuk 1975), the observation that
suppression occurs supports a receptor based model of astringency perception.
However, the narrow scope of the study did not allow for the psychophysical functions
of the compounds to be defined, and from a practical perspective did not suggest the
extent or circumstances in which phenolic compounds containing common moieties
may suppress each other, or whether phenolic compounds from different classes could
also be suppressive.

The work presented in this thesis only represents a partial picture of the interactions
between white wine components in the percept of mouth-feel, in that at most, three
levels of each component were statistically crossed which did not allow any models
describing their interaction to be proposed. Looking ahead, a worthy goal of further
research conducted from the results of this thesis would be to provide practically useful
models which could be used to predict the mouth-feel profiles of white wine from
knowledge of their composition. Ideally, the models would incorporate the smallest
number of non-redundant explanatory (compositional) variables that could be used to
both explain and predict the mouth-feel attributes perceived in white wine.

99
A prerequisite for the construction of such models has been a knowledge of the
psychophysical functions (or at least their form) of each of the major components i.e.
ethanol, organic acids/pH, phenolics and polysaccharides that make up the ‘mixture’
that is white wine. Surprisingly, even knowledge of the mathematical relationships
between pH and perceived astringency/dryness, and that between hotness and ethanol
concentration appear to be lacking. The development of such models will be
challenging. Firstly, obtaining enough material for large scale sensory studies of this
sort whether it be by extraction and purification or by synthesis is notoriously difficult,
expensive or both. The creation of quantitative models requires collecting ratio scaled
data, which is unfeasible using the accepted but fatiguing method of Magnitude
Estimation, and although a straightforward line scale with ratio properties suited to
measuring chemosensory responses has been developed and validated (Green et al.
1996), its utility in quantifying the intensities of the diverse group of attributes, not all
of which are of chemesthetic origin, that comprise mouth-feel needs to be first tested

Finally, there is the question of which phenolic and polysaccharide compounds, or


even which broad compound classes should be included when modelling mouth-feel
attributes on composition. Including all candidate compounds (e.g. the phenolic class
alone could contribute 80 compounds of varied benzoic ring structures and
glycosylation and esterification patterns) in any theoretically plausible mixture model
is impractical, and perhaps even impossible.

So how could any such mixture model be simplified inasmuch that it is practical but
also theoretically valid? Research in the taste modality provides some guidance in this
matter. In that modality, similar tasting substances are known to be additive (Bartoshuk
1975, Keast et al. 2003) and that in circumstances whereby compound concentrations
are correlated as is the case of phenolic compounds from the same class (Gawel et al.
2014a), then they add along an averaged psychophysical function (Frijters and
DeGraaf 1987). If this were shown to be the case of phenolic compounds and their
relationship to mouth-feel, then the summed concentrations of phenolic compounds
with similar mouth-feel attributes could be modelled rather than the impractically
larger number of individual compounds, and that a single indicative compound could
be used to quantify the combined effect of multiple members of the same class of
compounds.

100
However, a further question arises as to which phenolics have the same mouth-feel
attributes. The work presented here and elsewhere has shown that changes to the
substitution pattern of phenolics i.e. hydroxylation, esterification, glycosylation
(Chapter 1), and derivitisation with glutathione (Gawel et al. 2014b) can significantly
alter the mouth-feel of phenolic compounds. These general results indicate that
classifying phenolic compounds only on their ‘text book’ phenolic ring structure is
likely to be inadequate, and that a broad set of structure-function relationship studies
of phenolic compounds would be needed to derive a general set of rules which could
be used to create a general classification of phenolic compounds based on their sensory
properties.

A cursory analysis of the literature shows that the concentrations of individual


phenolics reported in white wine are well below that of their reported detection
thresholds (Chapter 1 and Gawel 1998). This indicates that the reported impacts of
white wine phenolics on mouth-feel (Gawel et al. 2013, Cejudo-Bastante et al. 2011)
must be the result of sensory additivity of the phenolic compounds involved. The
possibility that sub-threshold additivity occurs in mouth-feel is not without precedent,
as the taste and olfactory modalities also function hypo-additively (Delwiche and
Heffelfinger 2005). However, whether multicomponent mixtures of compounds that
elicit qualitatively similar mouth-feel sensations can be perceived even when their
concentrations are below threshold, and the form of interaction that applies in this
situation still needs to be established.

The result that certain polysaccharides in white wine can suppress the negative quality
aspect of palate hotness elicited by ethanol (Gawel et al. 2016) is important from the
perspective of winemaking practice and may explain why some wines do not appear
‘hot’ on the palate although they contain elevated levels of alcohol. The result is also
important as it suggests that distinct physiological systems can interact to mediate
mouth-feel sensations. However, the physiological mechanisms and related
psychophysical rules that govern the interactions between these sensory systems are
unknown but could involve changes to the structure of free and bound water contained
in the saliva adjacent to oral surfaces.

101
Finally, the overall perception of mouth-feel in white wine is a response to the
chemosensory, tactile and taste systems (Chapter 1). A full investigation into the nature
of sensory interactions across these systems was beyond the scope of this thesis, but a
large body of work in the related field of gustation which also involve interactions
between different sensory modalities suggest that the overall percepts of flavour and
taste cannot be understood by simply investigating its contributing parts. Prescott
(1999) has gone so far to suggest that flavour might be considered as a functionally
distinct modality cognitively constructed from the integration of the distinct
physiological sensory systems of olfaction and taste, the latter of which is itself may
be considered an integrated construct of sensory systems in the mouth (Green 2002).
It has been known for some time that the non-volatiles in food (Delwiche 2004), and
in white wine (Lubbers et al. 1994, 2001) can influence flavour perception. A few
studies have also demonstrated the reverse case of retronasally perceived volatile
compounds influencing mouthfeel perception (Symoneaux et al. 2015; Sáenz-Navagas
et al., 2010; Jones et al. 2008). Further work needs to be conducted to establish whether
the interaction between volatile and non-volatile components in white wine is of a
symmetric or asymmetric nature.

All the studies presented as part of this thesis reported the perception of mouth-feel in
white wine at ambient temperature. Given that temperature is likely to affect the
physical viscosity of wine, and that there is a likely overlap in response by oral
transient receptor potential channels to temperature and to irritation elicited by ethanol
and phenolics, future work should include wine temperature as an experimental
variable.

Forty years ago, Van Bursick and Erickson (1977) noted that olfactory, taste and
trigeminal information converge in the solitary nucleus, and suggested that the
trigeminal system may serve to bind the physiologically distinct olfactory and
gustatory systems into a single perceptual system when eating and drinking. Under a
similar more integrated interpretation, ‘mouth-feel’ might also be considered a sensory
construct, described by the accepted term ‘body’. While the concept of white wine
‘body’ was considered and explored early in this research (Gawel et al. 2007), perhaps
the concept of ‘body’ should be revisited in future research given that it has been long
used by experts to relate information about white wine to consumers. However,

102
regardless of how the sensory concept of mouth-feel is interpreted, there is little doubt
that its measurement and analysis provides unique challenges due to the heterogeneity
but physical proximity of the receptive systems involved, and what is a fundamental
but poorly understood influence of saliva which likely mediates all aspects of mouth-
feel perception.

References

Bartoshuk, L. M. (1975). Taste mixtures: Is mixture suppression related to


compression? Physiology and Behavior 14: 643-649.
Cejudo-Bastante, M. J., Hermosín-Gutiérrez, I., Castro-Vázquez, L., and Pérez-
Coello, M. S. (2011). Hyperoxygenation and bottle storage of Chardonnay white
wines: Effects on color-related phenolics, volatile composition, and sensory
characteristics. Journal of Agricultural and Food Chemistry 59: 4171-4182.
Delwiche, J., and Heffelfinger, A. L. (2005). Cross-modal additivity of taste and smell.
Journal of Sensory Studies 20: 512-525.
Frijters, J. E. R., and DeGraaf, C. (1987). The equiratio taste mixture model
successfully predicts the sensory response to the sweetness intensity of complex
mixtures of sugars and sugar alcohols. Perception 5: 615-628.
Gawel, R., Smith, P. A., and Waters, E. J. (2016). Influence of polysaccharides on the
taste and mouth-feel of white wine. Australian Journal of Grape and Wine Research
22: 350-357.
Gawel, R., Day, M., Van Sluyter, S. C., Holt, H., Waters, E. J., and Smith, P. A.
(2014a). White wine taste and mouth-feel as affected by juice extraction and
processing. Journal of Agricultural and Food Chemistry 62: 10008-10014.
Gawel, R., Schulkin, A., Smith, P. A., and Waters, E. J. (2014b). Taste and textural
characters of mixtures of caftaric acid and Grape Reaction Product in model wine.
Australian Journal of Grape and Wine Research 20: 25-30.
Gawel, R., Van Sluyter, S. C., Smith, P. A., and Waters, E. J. (2013). Effect of pH and
alcohol on perception of phenolic character in white wine. American Journal of
Enology and Viticulture 64: 425-429.
Gawel, R., and Waters, E. J. (2008). The effect of glycerol on the perceived viscosity
of dry white table wine. Journal of Wine Research 19: 109-114.
Gawel, R., Dimanin, P. A. G., Francis, I. L., Waters, E. J., Herderich, M. J., and
Pretorius, I. S. (2008). Coarseness in white table wine. Australian and New Zealand
Wine Industry Journal 23: 19-22.
Gawel, R., Van Sluyter, S., and Waters, E. J. (2007). The effects of ethanol and
glycerol on the body and other sensory characteristics of Riesling wines. Australian
Journal of Grape and Wine Research 13: 38-45.

103
Gawel, R. (1998). Red wine astringency: a review. Australian Journal of Grape and
Wine Research 4: 74-95.
Goold, H. D., Kroukamp, H., Williams, T. C., Paulsen, I.T., Varela, C., and Pretorius,
I. S. (2017). Yeast’s balancing act between ethanol and glycerol production in low-
alcohol wines. Microbiological Technology 10: 264-278.
Green, B. G. (2002). Studying taste as a cutaneous sense. Food Quality and Preference
14: 99-109.
Green, B. G., Dalton, P., B., C., Schaffer, G., Rankin, K., and Higgins, J. (1996).
Evalutating the 'labelled magnitude scale' for measuring sensation of taste and smell.
Chemical Senses 21: 323-334.
Keast, R. S. J., Bournazel, M. M. E., and Breslin, P. A. S. (2003). A psychophysical
investigation of binary bitter-compound interactions. Chemical Senses 28: 301-313.
McLean, H. (2005). Managing phenolics in white wine. In: Advances in Tannin
Management, pp 36-39. Allen, M., Dundon, C., Francis, M. E., Howell, G., and Wall,
G., Eds., Australian Society of Viticulture and Oenology Inc., Adelaide, South
Australia.
Noble, A. C., and Bursick, G. F. (1984). The contribution of glycerol to perceived
viscosity and sweetness in white wine. American Journal of Enology and Viticulture
35: 110-112.
Prescott, J. (1999). Flavour as a psychological construct: implications for perceiving
and measuring the sensory quality of foods. Food Quality and Preference 10: 349-356.
Sáenz-Navajas, M. P., Campo, E., Fernández-Zurbano, P., Valentin, D., and Ferreira,
V. (2010). An assessment of the effects of wine volatiles on the perception of taste and
astringency in wine. Food Chemistry 121: 1139-1149.
Schobel, N., Radtke, D., Kyereme, J., Wollmann, N., Cichy, A., Obst, K., Kallweit,
K., Kletke, O., Minovi, A., Dazert, S., Wetzel, C. H., Vogt-Eisele, A., Gisselmann, G.,
Ley, J. P., Bartoshuk, L. M., Spehr, J., Hofmann, T., and Hatt, H. (2014). Astringency
is a trigeminal sensation that involves the activation of G protein-coupled signaling by
phenolic compounds. Chemical Senses 39: 471-487.
Symoneaux, R., Le Quére, J. M., Baron, A., Bauduin, R., and Chollet, S. (2015).
Impact of CO2 and its interactions with the matrix components on sensory perception
in model cider. LWT-Food Science and Technology 63: 886-891.
Van Bursick, R. L., and Erickson, R. P. (1977). Odorant responses in taste neurons of
the rat NTS. Brain Research 135: 287-303.

104
APPENDIX I

Gawel et al. (2017) The mouthfeel of wine wine

doi:10.1080/10408398.2017.1346584
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION
https://doi.org/10.1080/10408398.2017.1346584

The mouthfeel of white wine


Richard Gawela, Paul A. Smitha, Sara Ciceraleb, and Russell Keastb
a
Australian Wine Research Institute, Paratoo Road, Urrbrae, Australia; bDeakin University Faculty of Health, School of Exercise and Nutrition Sciences,
Burwood, Australia

ABSTRACT KEYWORDS
White wine mouthfeel which encompasses the tactile, chemosensory and taste attributes of perceived White wine; mouthfeel;
viscosity, astringency, hotness and bitterness is increasingly being recognized as an important component texture; phenolics;
of overall white wine quality. This review summarizes the physiological basis for the perception of white polysaccharides; alcohol;
wine mouthfeel and the direct and interactive effects of white wine composition, specifically those of low acidity
molecular weight phenolic compounds, polysaccharides, pH, ethanol, glycerol, dissolved carbon dioxide,
and peptides. Ethyl alcohol concentration and pH play a direct role in determining most aspects of
mouthfeel perception, and provide an overall framework on which the other minor wine components can
interact to influence white wine mouthfeel. Phenolic compounds broadly impact on the mouthfeel by
contributing to its viscosity, astringency, hotness and bitterness. Their breadth of influence likely results
from their structural diversity which would allow them to activate multiple sensory mechanisms involved
in mouthfeel perception. Conversely, polysaccharides have a small modulating effect on astringency and
hotness perception, and glycerol does not affect perceived viscosity within the narrow concentration
range found in white wine. Many of the major sensory attributes that contribute to the overall impression
of mouthfeel are elicited by more than one class compound suggesting that different physiological
mechanisms may be involved in the construct of mouthfeel percepts.

Introduction polysaccharides, and dissolved carbon dioxide. The review


will discuss (1) how these compounds individually and
Many of the recognized great white wines of the world display a
through interaction affect the mouthfeel of white wine and
textural profile or ‘mouthfeel’ that typifies the grape cultivars
(2) the physiological basis for mouth-feel perception in white
from which they were made and the methods used to vinify
wine. The review is scoped by studies specifically related to
them. As examples, highly respected examples of wines made
nonvolatile compounds found in dry white wine that have
from the cultivars Viognier and Pinot Gris are as much defined
been shown to, or could potentially influence white wine
by their oily mouthfeel as by varietal flavors, and winemakers
mouthfeel. Specifically, only reports of the influence of low
worldwide invest in high cost processes to create the creamy
molecular weight phenolic compounds, and polysaccharides
texture that defines barrel fermented Chardonnay wines
relevant to white wine are reviewed, and the effects of com-
(Robinson, 1999).
pounds at concentrations typical of those found in dry white
White wine elicits tactile sensations of viscosity and astrin-
wine are emphasized. While it is acknowledged that bitterness
gency/dryness, and the chemosensory sensations of warmth,
is a taste rather than a texture, it has been included for discus-
pricking and spritz (Jackson, 2014; Oberholster et al., 2009;
sion on the presumption that bitterness is likely to modulate
Pickering and DeMiglio, 2008). Other mouthfeel attributes
the perception of mouthfeel characters in white wine.
such as metallic (Jones et al., 2008) and pungency (Gawel et al.,
2013) that likely incorporate aspects of bitter taste have also
been reported. Mouthfeel has been defined as ‘the group of sen-
sations characterized by a tactile response in the mouth’ (Pick-
Physiology of the mouth related to mouthfeel
ering and DeMiglio, 2008). However, considering recent The oral mucosa is the tissue membrane that lines the oral cav-
findings that chemically induced irritation and thermal sensa- ity. Among its varied functions, it provides physical and immu-
tions also contribute to white wine mouthfeel, we propose a nological protection, as well as enabling the sensations of
broader definition that wine mouth-feel comprises ‘the tactile, touch, temperature and taste. The oral mucosa is structurally
irritant and thermal sensations resulting from the activation of heterogeneous as it must necessarily accommodate many phys-
chemosensory and somatosensory receptors within the oral ical functions. The mucosa of the inside lips, cheek, tongue,
cavity by wine chemical stimuli’. and soft palate need to be elastic and moveable as otherwise
The compositional factors mostly cited as those contribut- phonation would be impossible. Conversely the gum and hard
ing to the mouthfeel of white wine are low molecular palate mucosa are significantly more rigid as they are routinely
weight phenolic compounds, ethanol, glycerol, organic acids, subjected to considerable mechanical forces encountered when

CONTACT Richard Gawel richard.gawel@awri.com.au Australian Wine Research Institute, Paratoo Road, Urrbrae, 5064, Australia.
© 2017 Taylor & Francis Group, LLC
2 R. GAWEL ET AL.

chewing food. The gum and hard palate mucosa are more rigid Physiological basis for mouthfeel and bitterness
because they are relatively thinner and are embedded with a
Astringency
dense layer of keratin. In contrast, the thicker and more elastic
mucosa of the inner lips, cheek and soft palate are keratinized The dominant paradigm is that astringency is the percep-
to a significantly lesser degree (Squier and Kremer, 2001). tion of increased friction between oral surfaces resulting
Trigeminal (5th cranial nerve) free nerve endings that from reduced salivary lubrication following an interaction
extend into the middle and upper layers of the oral epithe- between salivary proteins and polyphenols (Gawel, 1998).
lium are found throughout most of the oral cavity. Tran- The interaction mostly involves both hydrogen bonding
sient receptor potential (TRP) channels that respond to between amino acid carbonyl groups and the hydroxyl
combinations of heat, cold and other chemical stimuli such groups on the polyphenol, and hydrophobic stacking of the
as ethanol are located on these nerve endings (Clapham, benzoic ring with the apolar face of amino acid residues on
2003). As the receptive elements for these chemical stimuli the protein (Baxter et al., 1997; Haslam and Lilley, 1988;
are located on the intracellular domain of the TRP channel, Luck et al., 1994; Spencer et al., 1988). Proline rich proteins
the substance must firstly pass through a lipid bilayer (PRP’s) have mostly been the focus of the studies as they
before the receptor can be activated (Furlan et al., 2014). are (1) the dominant class of proteins secreted by the
Other receptors found on fast conducting afferent nerve parotid gland and submandibular glands which provide the
fibers that innervate the basal cells of the epithelium func- greatest volume of saliva (Walz et al., 2006; Kauffman
tion as mechanoreceptors (Watanabe, 2004). These recep- et al., 1991), and (2) have a strong binding affinity with
tors respond exclusively to tactile stimuli such as providing wine polyphenols due to their open and flexible structure
information as to the velocity and position of the food that promotes hydrogen bonding (Bacon and Rhodes, 2000;
bolus necessary for swallowing, but logically may also signal Luck et al., 1994). However, a causal link between interac-
stretching and ‘sticking’ sensations between oral surfaces tions of wine polyphenols and PRP’s found in the bulk
experienced during wine consumption. saliva leading to changes to its rheology and therefore
The oral mucosa is covered by a thin layer of adsorbed astringency perception is yet to be established. Fundamen-
salivary proteins called the acquired pellicle and on top of tally, the oral lubrication paradigm of astringency is contin-
that, a significantly thicker film of bulk saliva. Together, they gent upon mechanoreceptors of somatosensory nerves
play an important role in lubricating the soft tissues in the being activated during oral exposure to astringents. To our
oral cavity, but when disrupted by tasting wine may also knowledge, direct evidence of such activation is yet to be
influence the perception of some aspects of mouthfeel. shown. However, deactivation of the branch of the trigemi-
Details of the sources and composition of the salivary pro- nal nerve V that innervates oral mechanoreceptors and che-
teins that are likely to interact with wine components and moreceptors has been shown to result in a loss of
therefore influence mouthfeel have been reviewed elsewhere astringency perception (Schobel et al., 2014) suggesting that
(Gawel, 1998). Model studies using both hydrophilic and mechanoreception could play a role in the perception of
hydrophobic surfaces (modelling the keratinized and non- astringency.
keratinized mouth surfaces, respectively) suggest that sali- Astringency perception may also involve chemothesis,
vary proteins rapidly form a thin boundary layer which is whereby oral free nerve endings with transient receptor
capable of lubricating surfaces (Maheshwari and Dhatha- potential channels V1 (TRPV1) are activated by chemical or
threyan, 2006; Vitkov et al., 2004). Two mechanisms for pel- physical stimuli responsible for perceived irritation, hotness
licle formation have been proposed: (1) the smaller salivary and coolness. Kurogi et al. (2015) found that dimeric flava-
proteins excreted from the parotid gland, PRP-1, statherin nols in tea activated TRPV1 channels in sensory neurons.
and histadin crosslink to form a pellicle which is then scaf- Consistent with the concept that astringency perception has
folded by adsorbed mucins excreted from the submandibular a chemosensory component, monomeric and polymeric phe-
and sublingual glands (Proctor et al., 2005; Yao et al., 2003; nolic compounds also activate trigeminal ganglion neurons
Berg et al., 2003), and (2) the smaller parotid proteins cross- within a time frame consistent with the onset of astringency
link and stabilize a previously established mucin layer perception (Schobel et al., 2014). Related to this is the obser-
hydrophobically bound to the epithelia (Svendsen et al., vation that ethanol, which elicits a drying mouth-feel (Jones
2006; Iontcheva et al., 2000). In either case, a thin lubricative et al., 2008) also perturbs the lipid layer of oral epithelial
proteinaceous layer is formed as under both proposed sce- cells in a similar fashion to polyphenols (Furlan et al., 2015).
narios, the oligosaccharide side chains of the attached Astringency can be perceived even when oral surfaces have
mucins can form outward facing ‘molecular brushes’ that been stripped of bulk saliva (Nayak and Carpenter, 2008) which
can repel opposing similar surfaces by osmotic pressure or suggests that astringency perception may also involve interac-
steric effects (Cardenas et al., 2007; Schwender et al., 2005). tions between polyphenols and the proteins that comprise the
However, the salivary film covering the oral epithelium at acquired pellicle, or other connective epithelial or membrane
between 18 and 50 mm (Lee et al., 2002) is sufficiently thick bound proteins (Coles et al., 2010; Malone et al., 2003). Wine
to suggest that mouth lubrication by saliva consists of a tannins are capable of binding directly to oral epithelial cells
mixed form of hydrodynamic lubrication resulting primarily (Payne et al., 2009) with an efficacy related to their degree of
from unattached mucins operating between opposing mouth polymerization (Soares et al., 2016).
surfaces (Szabo et al., 2000; Gong and Osada, 1998) and thin Similarly, the perceived astringency of whey proteins and
film lubrication by the pellicle at the oral surface. the polysaccharide chitosan also correlate with their ability to
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 3

bind with oral epithelial cells (Ye et al., 2012; Malone et al., the upper surfaces of the tongue and the palate” (Van
2003). The exact mechanism of cell binding, or of which recep- Vliet et al., 2009). While the physical viscosities encoun-
tor channels would relay information about the presence of tered in dry white wine are narrow (estimated by Kosmerl
astringent compounds under these circumstances is unknown. et al., 2000 at 0.15 mPa/s), tasters are able to perceive
Anecdotally, astringency can be clearly perceived in a sta- changes in the oral viscosity of commercial white wines at
tionary mouth whereby there is no movement between oppos- around 1/5th (0.027 mPa/s) of that physical range, and
ing oral surfaces. While this does not rule out the influence of viscous mouthfeel and physical viscosity were significantly
mechanoreception, as constriction of oral surfaces may be correlated (Runnebaum et al., 2011) suggesting that per-
involved (Verhagen and Engelen, 2006), it does strengthen the ceived viscosity is related to physical viscosity in white
argument that astringency is most likely the result of multi- wine. However, it is notable that higher pH (Gawel et al.,
modal actions involving interactions with the bulk saliva, 2013, 2014a), or higher lactate (Runnebaum et al., 2011;
acquired pellicle and oral epithelial cells leading to a general Skogerson et al., 2009) which is a proxy for higher pH in
somatosensory response most likely involving both chemore- white wine has been shown to be strongly associated with
ception and mechanoreception. higher perceived viscosity. As pH is unlikely to affect
physical viscosity, these results suggest that physical vis-
cosity of the wine is likely not solely responsible for the
Bitterness
perception of oral viscosity in white wine.
Bitter taste is elicited by taste receptor cells located in taste buds
that are embedded in the epithelium of papillae on the tongue
and soft palate. These taste receptor cells contain subsets of the Compositional factors affecting white wine mouthfeel
25 member family of G protein-coupled receptors known as Phenolic compounds
T2Rs (Chandrashekar et al., 2000, Adler et al., 2000). The sub-
structure of these receptors has not been fully defined, but if The phenolic compounds in white wine comprise a broad fam-
they are heterodimeric like those of sweetness and umami ily of compounds that in their basic form possess one or more
receptors, then there could be up to 325 functional bitterness benzenoid rings substituted by at least one hydroxyl group.
receptors in humans (DuBois, 2011). White wine phenolics can be categorized as either (1) nonflavo-
Bitterness is unique amongst the tastes in that bitter tasting noids or (2) flavonoids according to their benzoic ring struc-
compounds display a high level of structural diversity. This can ture. The structures, reported concentration ranges and sensory
be explained at a receptor level by (1) different expression pat- thresholds of the phenolic compounds in white wine are given
terns of the 25 T2Rs across receptor cells (Chandrashekar et al., in Figure 1, and Tables 1 and 2, respectively.
2000) which like the olfactory system may provide a mecha- The major nonflavonoids in white wine are the hydroxycin-
nism for perceptual discrimination between bitter compounds namic and hydroxybenzoic acids and their derivatives. Hydrox-
and (2) most of the T2Rs are broadly tuned, i.e., individual ybenzoic acids comprise of a single benzenoid ring and
T2R’s can be activated by multiple structurally different bitter- hydroxyl group but are characterized by a further substitution
ants, and alternatively, the same bitterant can activate multiple with a carboxylic acid group. The hydroxycinnamic acids are
receptor types (Meyerhof et al., 2010). Broad tuning of T2R bit- characterized by an ethylene group between the benzene ring
ter receptors to white wine phenolic compounds has recently and the carboxylic acid group.
been demonstrated (Soares et al., 2013). Furthermore, it is The flavonoids are more complex comprising a C15
known that postreceptor (cell level) transduction mechanisms skeleton with an aromatic (A) and benzodihydropyran (C)
are also involved in eliciting bitterness, which in the case of ring bearing another aromatic ring (B) in position 2. The
white wine, may also include the direct activation of secondary flavonoid subgroups are classified by the oxidation state of
messenger systems following the passing of bitter hydrophobic the C ring and individual compounds within each group
compounds through receptor cell membranes (Furlan et al., are differentiated by the number and location of either
2015). hydroxyl or methoxyl groups, and glycosylation on the B
ring. Flavonols and flavan-3-ols are most significant classes
of flavonoids in white wine in terms of concentration, but
Hotness and perceived viscosity others such as flavanonols and flavanones have consis-
The perceived oral warmth/hotness, bitterness, and dryness tently been found to be present.
of white wine are influenced by its ethanol content (Jones
et al., 2008; Gawel et al., 2007). The physiological basis for Nonflavonoids
these effects is evidenced by the existence of the thermally The hydroxycinnamic acids in white wine occur mainly in
responsive TRPV1 noiceptors embedded within the oral the form of tartaric acid esters, particularly those of caffeic
mucosa and the bitter responsive T2R38 receptors located in (caftaric), p-coumaric (coutaric), and ferulic (fertaric) acids.
the circumvallate papillae at back of the tongue robustly Free forms and ethyl and methyl esters are also formed in
respond to the presence of ethanol (Allen et al., 2014; Trevi- lower concentrations by hydrolysis and pectinase catalyzed
sani et al., 2002). esterification. Caftaric acid is also the most abundant phe-
The perceived viscosity of Newtonian fluids with low nolic compound found in white juice (Ong and Nagel,
physical viscosity, i.e., white wine has been defined as “an 1978). When the juice is subjected to oxidative must han-
appraisal of the ease with which the liquid flows between dling practices, e.g., polyphenol oxidase hyperoxidation, the
4 R. GAWEL ET AL.

caftaric acid quinone reacts with the grape peptide glutathi-


one under the action of polyphenol oxidase to form
2-S-glutathionyl caftaric acid (Grape Reaction Product or
GRP) (Cheynier et al., 1986; Singleton et al., 1985). Other
similar conjugates of p-coutaric acid and caffeic acid with
cysteine and glutamine are also formed in smaller amounts

Figure 1. Structures of phenolic compounds reported in white wine.


CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 5

Silva et al., 1993; Lea et al., 1979). They consist mainly of


(C)-catechin and (¡)-epicatechin units linked by C4–C8
and/or C4–C6 bonds (Foo and Porter, 1980) but other
dimers containing gallocatechin and epigallocatechin units
(de Pascual-Teresa et al., 2000; Fulcrand et al., 1999) and
gallic acid units have also been detected in white wine
(Ricardo da Silva et al., 1993).
Glycosylated flavanonols have been identified in the
skin and stems, with the 3-O rhamnosides of dihydro-
quercetin (astilbin) and dihydrokaempferol (engeletin)
being the most significant in white wine (Masa et al.,
2007; Singleton and Trousdale, 1983). Flavanonol concen-
trations in 94 commercial white wines were found to
range between 1 and 13 mg/L (Vitrac et al., 2002).

Other phenolic compounds


Tyrosol is formed from tyrosine by yeast during fermenta-
tion and as such its concentration depends on yeast strain
and on the initial concentration of sugars and tyrosine in
the must (Pe~ na-Neira et al., 2000). The contribution of
tyrosol to total white wine phenolic concentration is
Figure 1. (Continued) unclear. Some have reported that tyrosol dominates the pro-
file (Pe~na-Neira et al., 2000), while others estimate that
by the same mechanism (Cejudo-Bastante et al., 2010; tyrosol comprises between 3% (Gawel et al. 2014a) and 10%
Cheynier et al., 1986). (Myers and Singleton, 1979) of the total phenolic content
The most prevalent hydroxybenzoic acids in white wine are of white wine.
gallic, gentisic, p-hydroxybenzoic, syringic, and salicylic acids
(Monagas et al., 2005). They are mostly found in their free
form but ethyl, methyl, and glucose esters have been reported Polysaccharides
in Riesling wines (Baderschneider and Winterhalter, 2001). Polysaccharides in white wine arise from both grape and
yeast activity during and after fermentation. They are classi-
Flavonoids fied as macromolecules that range from 5 to 200 kDa,
Flavonols are mostly confined to the epidermal skin tissues making them as a group, the highest molecular weight com-
of the grape (Rodrıguez Montealegre et al., 2006). Their pounds found in white wine (Jones et al., 2008). The major
concentration in white grapes have been shown to increase classes of polysaccharides in white wine have been defined
disproportionately to other skin phenolics following sun by a combination of their monosaccharide composition and
exposure (Friedel et al., 2015) suggesting that they may their probable oenological source: (1) Mannoproteins are
play a protective role against grape tissue damage by UV either released from yeast cell walls during fermentation,
light. The dominant flavonols in white and wine are the 3- and during the autolysis stage whereby nonactive ‘spent’
O-glycosides of quercetin particularly the glycosidic and yeast cells (yeast lees) contact the wine during the matura-
glucuronic forms (Castillo-Mu~ noz et al., 2010) but myrice- tion phase (Escot et al., 2001). They are characterized by a
tin, kaempferol, and isorhamnetin glycosides are also pres- high content of mannose relative to other monosaccharides
ent (Di Lecce et al., 2014; Vilanova et al., 2009). (Gonçalves et al., 2002), but constitute a broad molecular
Flavanols are mostly found in the hypodermal layers of weight distribution which suggests that several populations
the skins and the parenchyma of seeds and in conven- of mannnoproteins exist depending on their source, (2) ara-
tional winemaking are extracted during a brief (0.5–24 hr) binogalactan proteins (AGP’s) are characterized by a high
period between grape crushing and draining when the proportion of arabinose and galactose residues and associ-
skins and seeds are in contact the juice prior to fermenta- ated with a protein core comprising a high proportion of
tion (Di Lecce et al., 2014). (C)-catechin and its dister- hydroxyproline (Fincher and Stone, 1983). AGP’s exist in
eoisomer (¡)-epicatechin are the dominant flavanols in the grape vacuole intracellular spaces and are therefore eas-
white wine. They are dihyroxylated on the C-30 and C-40 ily extracted during juicing and during the early stages of
positions on the B ring. The trihydroxylated form, fermentation (Guadalupe and Ayestaran, 2007), and (3)
(¡)-epigallocatechin which is localized in skin, and those Rhamnogalacturonan polysaccharides are released from pec-
esterified with gallic acid notably (¡)-epigallocatechin gal- tins embedded within the grape cell wall and are character-
late and (¡)-epicatechin-3-O-gallate which are localized ized by their low MW (<20 kDa), a relatively high
in seeds are also present in white wine (Oszmianski and proportion of galacturonic acid and rhamnose residues, and
Sapis, 1989). Flavanol dimers and trimers are also present by the presence of the diagnostic sugars fucose and apiose
in white wine but in lower concentrations (Ricardo da (Pellerin et al., 1996).
6 R. GAWEL ET AL.

Table 1. Mean and concentration ranges of phenolic compounds in white wine.

Mean concentration (mg/L) Concentration range (mg/L) References

Nonflavonoids
Hydroxybenzoic acids
Gallic acid 3.3 0–8.4 Fernandez-Pachon et al. (2006)
3.4 0.7–18.5 de Villiers et al. (2005)
4.9 0.8–3.2 Singleton and Trousdale (1983)
1.0 0.4–1.2 Betes-Saura et al. (1996)
1.2 Kallithraka et al. (2009)
1.9 Cejudo-Bastante et al. (2011a,b)
0.8 Gawel et al. (2014a)
Syringic acid 0.04 0–0.5 Fernandez-Pachon et al. (2006)
0.4 0.3–2.2 Betes-Saura et al. (1996)
1.1 Gawel et al. (2014a)
Protocatechuic acid 1.2 0–4.7 Fernandez-Pachon et al. (2006)
1.3 0–0.4 de Villiers et al. (2005)
1.2 Betes-Saura et al. (1996)
0.3 Gawel et al. (2014a)
Hydroxybenzoic ethyl esters 0.9 0.3–1.8 Gawel et al. (2014a)
Hydroxycinnamic acids
Caffeic acid 1.3 0–2.7 Fernandez-Pachon et al. (2006)
2.7 0–0.5 de Villiers et al. (2005)
0.2 0.3–4.3 Nicolini et al. (1991)
1.6 0–4.0 Betes-Saura et al. (1996)
1.6 2.6–2.8 Korenika et al. (2014)
0.6 Kallithraka et al. (2009)
1.4 Cheynier et al. (1986)
2.7 Cejudo-Bastante et al. (2011a,b)
Coumaric acid 0.8 0–3.0 Fernandez-Pachon et al. (2006)
1.3 0.1–2.8 de Villiers et al. (2005)
0.2 0–0.8 Betes-Saura et al. (1996)
0.8 Korenika et al. (2014)
0.2 Kallithraka et al. (2009)
0.3 Cejudo-Bastante et al. (2011a,b)
Ferulic acid 0.4 de Villiers et al. (2005)
0.1 0.4–1.3 Betes-Saura et al. (1996)
0.6 0.9–1.0 Korenika et al. (2014)
0.3 Kallithraka et al. (2009)
0.9 Cejudo-Bastante et al. (2011a,b)
Tartaric acid esters
Caftaric acid 17.8 3.9–53.5 Fernandez-Pachon et al. (2006)
12.1 0–44.8 de Villiers et al. (2005)
11.9 0.5–5.4 Singleton and Trousdale (1983)
2.3 12.8–109.5 Nicolini et al. (1991)
39.4 1.5–42.3 Herrick and Nagel (1985)
43.7 24–267 Herrick and Nagel (1985)
12.6 0–39.6 Betes-Saura et al. (1996)
15.1 12.7–70.7 Korenika et al. (2014)
29.6 1.77–8.36 Kallithraka et al. (2009)
109.0 0–31 Herrick and Nagel (1985)
10.6 Cheynier et al. (1986)
41.3 Ricardo da Silva et al. (1993)
3.7 Cejudo-Bastante et al. (2011a,b)
7.8 Gawel et al. (2014a)
Coutaric acid 2.3 de Villiers et al. (2005)
3.8 0.2–14.5 Singleton and Trousdale (1983)
0.9 0.13–1.70 Nicolini et al. (1991)
3.2 1.1–12.4 Herrick and Nagel (1985)
5.9 3.2–22.6 Herrick and Nagel (1985)
16.3 0.40–8.64 Betes-Saura et al. (1996)
4.4 0.5–10.9 Kallithraka et al. (2009)
11.5 Ricardo da Silva et al. (1993)
3.6 Cejudo-Bastante et al. (2011a,b)
3.4 Gawel et al. (2014a)
Fertaric acid 5.4 Herrick and Nagel (1985)
2.5 0.6–4.2 Herrick and Nagel (1985)
0.2 0–2.95 Betes-Saura et al. (1996)
0.7 2.96–5.12 Korenika et al. (2014)
2.0 0.1–3.9 Kallithraka et al. (2009)
4.3 Cejudo-Bastante et al. (2011a,b)
1.6 Gawel et al. (2014a)
Other hydroxycinnamic acid derivatives
Glutathione (GRP) 3.4 0.5–11.0 Nicolini et al. (1991)
3.2 0–49.2 Betes-Saura et al. (1996)
(Continued on next page)
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 7

Table 1. (Continued )
Mean concentration (mg/L) Concentration range (mg/L) References

16.3 3.2–20.4 Cheynier et al. (1986)


9.7 8.2–19.4 Ricardo da Silva et al. (1993)
12.2 1.0–13.4 Cejudo-Bastante et al. (2011a,b)
6.2 Gawel et al. (2014a)
Other amino acid derivatives 3.6 0.3–7.3 Cheynier et al. (1986)
1.3 Gawel et al. (2014a)
Ethyl esters 0.4 0.2–1.0 Gawel et al. (2014a)
1.4 0–3.5 Fernandez-Pach on et al. (2006)
Coutaric acid glycoside 5.4 0–32.9 Fernandez-Pach on et al. (2006)
0.7 0.1–1.2 Nicolini et al. (1991)
Total hydroxycinnamic acids 34.1 Betes-Saura et al. (1996)
63.1 38.2–101 Cejudo-Bastante et al. (2011a,b)
21.7 5.2–51.6 Gawel et al. (2014a)
Total nonflavonoids 91 66–112 Singleton et al. (1980)
Flavonoids
Flavan-3-ols
Epicatechin 0.8 0–10.4 Fernandez-Pach on et al. (2006)
3.7 0–80.8 de Villiers et al. (2005)
27.3 0–8.7 Singleton and Trousdale (1983)
19.3 0–21.3 Oberholster et al. (2009)
4.1 0.2–71.7 Betes-Saura et al. (1996)
5.3 Carando et al. (1999a)
25.5 Kallithraka et al. (2009)
5.0 Cejudo-Bastante et al. (2011a,b)
4.8 Gawel et al. (2014a)
33.8 Vitrac et al. (2002)
2.6 G€urb€uz et al. (2007)
Catechin 0.9 0–11.1 Fernandez-Pach on et al. (2006)
7.2 3–123.9 de Villiers et al. (2005)
37.0 1.7–9.8 Singleton and Trousdale (1983)
9.1 0–9.0 Oberholster et al. (2009)
2.5 2.3–28.7 Betes-Saura et al. (1996)
9.8 Carando et al. (1999a)
17.7 Kallithraka et al. (2009)
4.4 Cejudo-Bastante et al. (2011a,b)
2.7 Gawel et al. (2014a)
28.3 Vitrac et al. (2002)
10.1 G€urb€uz et al. (2007)
Epicatechin gallate 2.4 0–8.0 Gawel et al. (2014a)
Epigallocatechin gallate 2.5 0–15.6 Fernandez-Pachon et al. (2006)
1.0 0–6.0 Gawel et al. (2014a)
Flavan-3-ol oligomers
B1 dimer 1.2 0–8.6 Fernandez-Pachon et al. (2006)
6.1 0.8–1.4 Carando et al. (1999b)
0.1–24.1 Ricardo da Silva et al. (1993)
B2 dimer 1.4 0–1.5 Betes-Saura et al. (1996)
0.5 Ricardo da Silva et al. (1993)
B3 dimer 0.9–2.2 Carando et al. (1999b)
1.2 0–2.1 Betes-Saura et al. (1996)
0.5 Ricardo da Silva et al. (1993)
B4 dimer 0.3–0.5 Carando et al. (1999b)
0.13 0–0.4 Ricardo da Silva et al. (1993)
Trimers 1.8 0.6–1.7 Carando et al. (1999b)
0–6.3 Ricardo da Silva et al. (1993)
Total dimers 5.3 0.3–2.2 Carando et al. (1999b)
Total trimers 1.6 0.6–1.7 Carando et al. (1999b)
Total flavan-3-ols 20.2 13.9–22.6 Cejudo-Bastante et al. (2011a,b)
12.8 2.8–31.3 Gawel et al. (2014a)
11.0 3.4–17.3 Konitz et al. (2003)
Flavonols
Quercetin (Q) 1.4 0–3.4 Korenika et al. (2014)
1.5 0.1–3.2 Cejudo-Bastante et al. (2011a,b)
0.4 0–4.7 Gawel et al. (2014a)
0.7 0–1.2 Simonetti et al. (1997)
Q glucoside 0.4 0–0.8 de Villiers et al. (2005)
0.4 0.4–1.2 Korenika et al. (2014)
0.8 0–30.3 Cejudo-Bastante et al. (2011a,b)
3.4 Gawel et al. (2014a)
Q-glucuronide 0.3 1.1–1.9 Betes-Saura et al. (1996)
1.5 0–44.4 Cejudo-Bastante et al. (2011a,b)
3.3 Gawel et al. (2014a)
Q-rutinoside 0.4 0.2–0.9 Simonetti et al. (1997)
(Continued on next page)
8 R. GAWEL ET AL.

Table 1. (Continued )
Mean concentration (mg/L) Concentration range (mg/L) References

Kaempferol (K) 0.04 0–0.3 Korenika et al. (2014)


0.40 0–0.9 Cejudo-Bastante et al. (2011a,b)
0.1 0.1–0.1 Simonetti et al. (1997)
K-glucoside 0.3 0.2–0.4 Cejudo-Bastante et al. (2011a,b)
K-galactoside 0.1 0.1–0.14 Cejudo-Bastante et al. (2011a,b)
Isorhamnetin (I) 0.02 0–0.4 Korenika et al. (2014)
Myricetin 0.2 0.1–0.3 Simonetti et al. (1997)
I-glucoside 0.04 0.04–0.04 Cejudo-Bastante et al. (2011a,b)
Total flavonols 51.7 39.5–80.0 Cejudo-Bastante et al. (2011a,b)
9.0 0.5–56.8 Gawel et al. (2014a)
Other phenolic compounds
Tyrosol 2.5 0–11.9 Fernandez-Pachon et al. (2006)
11.6 2.2–3.6 Betes-Saura et al. (1996)
2.8 6.3–22.7 Gawel et al. (2014a)
13.3 Konitz et al. (2003)
Total flavonoids 168 128–222 Singleton et al. (1980)
Flavanonols
Dihydroquercetin rhamnoside (astilbin) 1.1 0–10.3 Singleton and Trousdale (1983)
2.4 0–21.1 Gawel et al. (2014a)
2.2 0.6–4.4 Vitrac et al. (2002)
Dihydrokaempferol rhamnoside (engeletin) 0.5 0–3.3 Singleton and Trousdale (1983)
Dihydromyricetin-rhamnoside 3.0 1.8–6.0 Vitrac et al. (2002)
Flavanones 2.6 0.1–11.2 Gawel et al. (2014a)
Total phenolics 251 164–316 Konitz et al. (2003)
258 224–328 Singleton et al. (1980)
279 168–404 Kallithraka et al. (2009)
201 152–281 Sokolowsky and Fischer (2012)
Cejudo-Bastante et al. (2011a,b)

Influence of white wine composition on mouthfeel not contribute to astringency. However, it should be noted
that the influence of a salivary pellicle, and epithelial bound
Phenolic compounds
salivary proteins typical of an oral surface were not mod-
General discussion elled in their experiment.
The traditional paradigm of astringency i.e. decreased oral There is conflicting evidence as to whether monomers can
lubrication following exposure to polyphenols depends on the elicit astringency under the chemesthetic model. TRPV1 chan-
polyphenol being capable of binding with a salivary protein at nels were not activated in response to a variety of flavanol
different positions and aggregating multiple salivary proteins monomers (Kurogi et al., 2015; Carpenter, 2013), galloylated
(Baxter et al., 1997). The question as to whether the monomeric flavanol monomers have been shown to stimulate trigeminal
and dimeric polyphenols that make up the majority of the phe- neurons suggesting that they at least can elicit astringency
nolic profile of white wines have sufficient numbers of hydro- (Schobel et al., 2014). In conclusion, these differences may be
gen bonding sites and/or hydrophobic surfaces to interact with due to the widely different methods used to determine oral
salivary proteins has only recently been addressed. Recent stud- physiological response to the phenolic stimulus.
ies have shown that the basic proline rich protein (PRPb)
amino acid probes IB-5 (Canon et al., 2011) and IB9-37 (Canon Total phenolics
et al., 2013; Cala et al., 2012) are able to bind to and aggregate White wine phenolics are mostly monomeric and comprise
epicatechin gallate and epigallocatechin gallate (Pascal et al., over 80 compounds spanning a range of phenolic classes
2009) molecules. Flavanol dimers with extended structures defined by their ring structures, and within each class they take
have also been shown to act as bidentate ligands for IB7-14 on different forms based on their patterns of hydroxylation,
(Cala et al., 2010). Other monomeric nonflavanols such as nar- glycosylation, esterification, or conjugation with amino acids.
ingenin, apigenin and quercetin rhamnoside and rutinoside Despite this diversity, the summed concentration of the pheno-
also have affinities for IB14 and entire PRPb (Plet et al., 2015). lic species in white wine is relatively low compared to those in
These studies show that many monomeric and dimeric poly- red wines where polymeric flavanols derived from fermentation
phenols found in white wine can form complexes with basic on skins and seeds dominate their phenolic profile. However, as
proline rich proteins, and therefore have the potential to elicit the sensory response to mixtures of nonvolatiles that elicit
astringency under the tactile model. tastes and astringency are known to be at least partially additive
With respect to the alternative paradigms for astringency (Ferrer-Gallego et al., 2014; Keast et al., 2003), it is prudent to
perception, Soares et al. (2016) found that in vivo, low also consider the effects of combined total of phenolic com-
molecular weight polyphenol fractions comprising mono- pounds on mouthfeel in white wines.
mers and dimers did not bind well to oral epithelial cells The first study of its type in white wine used both conven-
regardless of the simultaneous presence of salivary proteins, tional, and unconventional winemaking methods more akin to
which under the epithelial binding model of astringency those used in red winemaking to create extremes in total phe-
would suggest that low molecular weight polyphenols do nolic concentration (Singleton et al., 1975). They found that
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 9

Table 2. Thresholds of phenolic compounds in white wine.

Threshold
Compound (mg/L) Attribute Medium Method Reference

Flavanols
Catechin 119 Astringency Water Staircase Scharbert et al. (2004)
Catechin 290 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Catechin 20 Detection 5% ethanol Staircase Dadic and Belleau (1973)
Catechin 46 Detection Water Ascending limits Delcour et al. (1984)
Dimer B1 139 Astringency Water Staircase Hufnagel and Hofmann (2008)
Dimer B1 231 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Dimer B2 110 Astringency Water Staircase Hufnagel and Hofmann (2008)
Dimer B2 280 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Dimer B3 116 Astringency Water Staircase Hufnagel and Hofmann (2008)
Dimer B3 17 Astringency Water Ascending limits Delcour et al. (1984)
Dimer B3 289 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Epicatechin 270 Astringency Water Staircase Scharbert et al. (2004)
Epicatechin 270 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Epicatechin 270 Astringency Water Staircase Hufnagel and Hofmann (2008)
Epicatechin gallate 110 Astringency Water Staircase Scharbert et al. (2004)
Epigallocatechin 169 Astringency Water Staircase Scharbert et al. (2004)
Epigallocatechin gallate 87 Astringency Water Staircase Scharbert et al. (2004)
Trimer C1 260 Astringency Water Staircase Hufnagel and Hofmann (2008)
Trimer C1 347 Bitterness Water Triangle Hufnagel and Hofmann (2008)
TrimersCtetramers 4 Detection Water Ascending limits Delcour et al. (1984)
Hydroxybenzoic Acids
Gallic acid 45 Astringency Water Staircase Glabasnia and Hofmann (2006)
Gallic acid 50 Astringency Water Staircase Hufnagel and Hofmann (2008)
Gallic acid 40 Detection Water Ascending limits Maga and Lorenz (1973)
Gallic acid 20 Detection 5% ethanol Staircase Dadic and Belleau (1973)
Gallic acid 50 Difference Beer Staircase Dadic and Belleau (1973)
Protocatechuic acid 32 Astringency Water Staircase Hufnagel and Hofmann (2008)
Protocatechuic acid 30 Detection Water Ascending limits Maga and Lorenz (1973)
Protocatechuic acid 20 Detection 5% ethanol Staircase Dadic and Belleau (1973)
Syringic acid 52 Astringency Water Staircase Hufnagel and Hofmann (2008)
Syringic acid 240 Detection Water Ascending limits Maga and Lorenz (1973)
Gallic acid ethyl ester 37 Astringency Water Staircase Hufnagel and Hofmann (2008)
Gallic acid ethyl ester 438 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Protocatechuic acid ethyl ester 9 Astringency Water Staircase Hufnagel and Hofmann (2008)
Protocatechuic acid ethyl ester 182 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Flavanonols
Dihydroquercetin rhamnoside 1.7 Detection Water Staircase Hufnagel and Hofmann (2008)
Dihydrokaempferol rhamnoside 2.1 Detection Water Staircase Hufnagel and Hofmann (2008)
Hydroxycinnamic Acids
Caffeic acid 13 Astringency Water Staircase Hufnagel and Hofmann (2008)
Caffeic acid 90 Detection Water Ascending limits Maga and Lorenz (1973)
Coumaric acid 23 Astringency Water Staircase Hufnagel and Hofmann (2008)
Coumaric acid 40 Detection Water Ascending limits Maga and Lorenz (1973)
Ferulic acid 13 Astringency Water Staircase Hufnagel and Hofmann (2008)
Ferulic acid 62 Detection Water E679-91 Work and Camire (1996)
Caffeic acid ethyl ester 69 Astringency Water Staircase Hufnagel and Hofmann (2008)
Caffeic acid ethyl ester 229 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Coumaric acid ethyl ester 27 Astringency Water Staircase Hufnagel and Hofmann (2008)
Coumaric acid ethyl ester 137 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Ferulic acid ethyl ester 15 Astringency Water Staircase Hufnagel and Hofmann (2008)
Ferulic acid ethyl ester 158 Bitterness Water Triangle Hufnagel and Hofmann (2008)
Caftaric acid 5 Astringency Water Staircase Hufnagel and Hofmann (2008)
Caftaric acid <50 Detection Water Yes/no Okamura and Watanabe (1981)
Coutaric acid <25 Detection Water Yes/no Okamura and Watanabe (1981)
Flavonols
Quercetin 10 Detection 5% ethanol Staircase Dadic and Belleau (1973)
Kaempferol 20 Detection 5% ethanol Staircase Dadic and Belleau (1973)
Myricetin 10 Detection 5% ethanol Staircase Dadic and Belleau (1973)
Quercetin glucoside 1 Astringency Water Staircase Hufnagel and Hofmann (2008)
Quercetin rhamnoside 9 Detection Water Staircase Dadic and Belleau (1973)
Quercetin galactoside 0.2 Astringency Water Staircase Hufnagel and Hofmann (2008)
Other
Naringin 17 Bitter Water Esaki et al. (1983)
Naringin 20 Bitter Water Guadagni et al. (1973)
Tyrosol 346 Bitter Water Ascending limits Takahashi et al. (1974)

white wine bitterness was unrelated to total phenolic concen- compounds (e.g., epicatechin and naringin) over nonbitter phe-
tration which would suggest that either (1) other (nonphenolic) nolic compounds (e.g., caftaric acid and GRP) within the total
compounds contribute to bitterness, or (2) that the winemaking pool of phenolics present in the wines. In a broader study,
methods influenced the relative contribution of bitter phenolic Gawel et al. (2014a) found that a 6-fold difference in total wine
10 R. GAWEL ET AL.

phenolic concentration obtained by applying conventional concentrations well below those of white wine. The possible
winemaking techniques increased bitterness only slightly, but mouthfeel and taste impact of flavonol monoglycosides were
the increases were strongly correlated with total phenolic demonstrated when quercetin 3-O-glucoside, albeit at higher
concentration. concentrations than found in white wine, elicited astringency
The perceived viscosity of white wine has been associated and bitterness (Ferrer-Gallego et al., 2016). A 3-O-rhamnoside
with its total phenolic concentration (Gawel et al., 2013; Okuda of quercetin has also been reported to be bitter and astringent
et al. 2007). In contrast, lower phenolic concentration white at white wine like concentrations (Dadic and Belleau, 1973).
wines made from hyperoxidized free run juice were not lower More studies have been conducted on the taste and mouth-
in ‘body’ than the control wines (Cejudo-Bastante et al. 2011a, feel properties of flavanols than any other phenolic class due to
b). The differences in conclusions may be explained by the their occurrence in many different beverages including wine,
expected differences in the range of total phenolics, and from green and black tea, and cider. However, all studies to date
variations in other compositional factors such as pH and resid- have used concentrations far higher than those observed in
ual sugar resulting from the different winemaking methods white wine (i.e., >500 mg/L). Therefore, the results of the stud-
that were applied. To overcome the limitations of correlative ies should be considered only indicative of possible sensory
studies, addition studies have also been conducted. White wines outcomes in white wine.
fortified with the total phenolic pool extracted from each of Catechin and its stereoisomer (¡)-epicatechin are reported
three different varietal wines were consistently perceived to be to be astringent and bitter (Peleg et al., 1999; Thorngate and
more viscous, astringent and bitter (Gawel et al., 2013). Total Noble, 1995; Robichaud and Noble, 1990) with epicatechin
phenolics were also shown to enhance the perceived hotness being more astringent and bitter than catechin (Ferrer-Gallego
and astringency of model white wine, but only of those that et al., 2014). However, perceptual mapping of these simple
were initially low in hotness (i.e. low alcohol), and low in the monomers against recognized bitter and astringent compounds
astringent/drying character (i.e., high pH). suggests that they are more bitter than they are astringent
(Kielhorn and Thorngate, 1999).
Nonflavonoids Dihydroxylated and trihdroxylated flavanols esterified with
Benzoic acid derivatives when presented at two orders of mag- gallic acid, e.g., epicatechin gallate and epigallocatechin gallate
nitude higher than that found in white wine elicit complex oral have also been deemed to be bitter and astringent in taste tests
sensations of sourness, bitterness, astringency and ‘prickling’ on tea (Yu et al., 2014; Narukawa et al., 2010), with bitterness
(Peleg and Noble, 1995). The substitution pattern was found to being later confirmed physiologically by the observation that
affect mouthfeel, with those substituted in the ortho position they activated the human bitter taste receptor hTAS2R14
(salicylic and gentisic acids) being astringent, and those with a (Yamazaki et al., 2013).
greater number of hydroxyl groups more ‘prickling’ (Peleg and Schobel et al. (2014) proposed that 3 hydroxyl groups on the
Noble, 1995). It is unclear if these results are relevant to white B ring (galloylation) is a necessary condition for robustly stim-
wine, however as gentisic acid has also been reported to be ulating an oral chemosensory response indicative of astrin-
astringent when tasted at white wine like concentrations (Dadic gency. This would indicate that the dehydroxylated flavanols
and Belleau, 1973) it is possible that it and other benzoic acids catechin and epicatechin cannot elicit astringency. However,
found in white wine may affect mouthfeel. this conclusion is contingent on the assumption that chemes-
Many hydroxycinnamic acids and their ethyl and tartaric thesis is the only astringency mechanism at play. Galloylation
acid esters have been reported to be astringent and bitter also appears to favour astringency perception under the salivary
at concentrations higher than those observed in white protein interaction model of astringency as trihydroxylation of
wine (Hufnagel and Hofmann, 2008). However, as the hydrox- the B-ring results in increased interactions with poly-L-proteins
ycinnamic acid concentrations found in white wine are (Poncet-Legrand et al., 2006), salivary mucins (Peleg et al.,
typically below their detection thresholds (Okamura and Wata- 1999) and basic PRP’s (Cala et al., 2010). Glycosylated PRP’s
nabe, 1981; Maga and Lorenz, 1973) it would suggest that they which are considered important in salivary lubrication are also
do not impact on the taste or mouthfeel of white wine. Indeed, precipitated by epigallocatechin gallate (Pascal et al., 2008). In
fortifying white (Verette et al., 1988) and red (Saenz-Navajas contrast, Ferrer-Gallego et al. (2015) noted that galloylated
et al., 2012) wines with realistic levels of caftaric acid did not forms were less astringent than dihydroxylated forms which
produce perceptible changes to their taste or mouthfeel. How- they attribute to the ability of basic PRP’s to simultaneously
ever, when in model wine, wine-realistic levels of caftaric acid interact with fewer galloylated flavanol molecules.
has been found to impact on acidity (Verette et al., 1988), and Epigallocatechin gallate has been reported as being more
astringency (Gawel et al., 2014b). Similarly, wine realistic con- astringent than other flavanols monomers (Rossetti et al., 2009)
centrations of grape reaction product did not affect the palate and is the main contributor to astringency in green tea (Yu
properties when added to a real white wine (Verette et al., et al., 2014). EGCG has been shown to increase the friction
1988), but GRP was found to increase the intensity of oily coefficient of whole saliva while epicatechin had no effect, sug-
mouthfeel and a burning aftertaste unrelated to ethanol when gesting that esterification with gallic acid and trihydroxylation
added to model white wine (Gawel et al., 2014b). of the B-ring may enhance the production of astringency (Ros-
setti et al., 2009).
Flavonoids Compared to red wines, the flavanols in white wines have a
A rutinoside (di-glycoside) of quercetin has been suggested by low degree of polymerization (dP), typically ranging from
(Scharbert et al., (2004) to be a powerful astringent at monomers to trimers. It is now well understood that the first
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 11

stage of astringency perception under the tactile model involves The hotness of white wine was also reduced by realistic lev-
complex formation between proteins and flavanols, and this els of white wine polysaccharides, mostly attributed to medium
complexation increases with the flavanols’ degree of polymeri- molecular weight mannoproteins and AGP’s in the 13-93 kDa
zation (Baxter et al., 1997). White wine flavanols (dP 1-3) have MW range (Gawel et al., 2016). The physiological mechanism
been shown to be astringent (Peleg et al., 1999) with their for the suppression of alcohol hotness by polysaccharides is
astringency increasing with dP. Consistent with this, flavanol unknown but worthy of further investigation given that
dimers have been shown to form soluble complexes with sali- increased flavour in white wine can be obtained by picking
vary proteins (Sarni-Manchado and Cheynier, 2002), and that grapes when riper, but this can incur an expense in wine quality
they can also activate TRPV1 channels associated with chemo- as doing so may result in excessive levels of palate hotness.
sensory perception (Kurogi et al., 2015). Wine polysaccharides may also influence astringency per-
ception as they have been shown to disrupt the interaction and
Other phenolic compounds aggregation between salivary proteins and polymerized flava-
The ‘tannin taste’ of Riesling wines from the same juice with nols (Carvalho et al., 2006). This effect could be due to either
different levels of must solids correlated strongly with their solubilization following formation of ternary complexes
tyrosol concentration, while being poorly correlated with between the protein, polyphenol and polysaccharide, or by
hydroxycinnamic acid and flavan-3-ol concentration (Konitz competition between the polysaccharide and protein for the
et al., 2003). A flavanone, naringin has recently been tentatively polyphenol (Luck et al., 1994). Such possible disruption has
identified in white wine (Gawel et al., 2014a). While flavanones been shown in a red wine context with lower perceived astrin-
are only likely to be present in white wine at very low concen- gency of model wine containing polymerized grape seed flava-
trations a possible effect on white wine bitterness cannot be nols following addition of Rhamnogalacturonan-II (Vidal et al.,
ruled out some flavanone glycosides are intensely bitter when 2004b), and correlations between mannoprotein and AGP con-
glycosylated in specific forms (Horowitz and Gentili, 1969). centrations with the astringency of Tempranillo red wines
(Quijada-Morın et al., 2014).
However, studies using white wines have found little evi-
Other considerations dence that polysaccharides influence their astringency. Addi-
The perceived intensity of mixtures of similar tasting com- tions of whole white wine polysaccharides at wine realistic
pounds including those that are bitter, are known to be additive concentrations did not influence the astringent/drying charac-
or sub-additive (Keast et al., 2003), and in the case of some ter of white wine (Gawel et al., 2016), and the astringency of
white wine phenolics, their effect on mouthfeel may even be white wines made using a broad spectrum of white winemaking
hyperadditive (Ferrer-Gallego et al., 2014). Therefore, even if methods and grape cultivars were found to be unrelated to their
the concentrations of individual phenolic species are insuffi- total polysaccharide concentration (Gawel et al., 2014a). The
cient to induce a sensation in white wine, they may do so in differences in the observed effects of polysaccharides on astrin-
combination. Caftaric acid and its derivative GRP have been gency across studies could be due to the difference in the pH of
shown to be mutually antagonistic with respect to astringency the wines that were used which could alter the charge proper-
and burning aftertaste, and were also found to be able to sup- ties of the polysaccharides (Verhnet et al., 1996) and salivary
press the astringent/drying sensation produced by the wine proteins (McArthur et al., 1995) and therefore their interaction
matrix (Gawel et al., 2014b), showing the potential for phenolic with wine polyphenols, or to differences in the relative concen-
compounds to interact in complex ways with each other and trations of phenolic compounds to polysaccharides which could
other major components found in wine. also determine the type of interaction, i.e., whether it be com-
petitive or associative (Soares et al., 2009).
The effect of polysaccharides on white wine bitterness and
Polysaccharides
(presumably) related characters such as metallic and pungency
Mixtures of mannoproteins and AGP’s at wine like concentra- is unclear. In model systems, the bitterness of polymerized
tions have been shown to enhance the perception of viscosity grape seed tannin was reduced by a mixture of mannoproteins
in model white wine (Gawel et al., 2016; Vidal et al., 2004a). and arabinogalactan-proteins (Vidal et al., 2004b) and whole
However, the effect of these polysaccharides on viscosity was polysaccharides from white wine reduced the metallic character
shown depend on pH, with the increase in perceived viscosity of a complex model white wine which included ethanol, flavor
only occurring in higher pH wines (Gawel et al., 2016). Possible compounds, glycerol, and white wine proteins (Jones et al.,
reasons of this pH effect include charge effects on acidic poly- 2008). However, others found that white wine polysaccharides
saccharides, or changes to the ordering of bulk water in the did not affect the bitterness of a high phenolic white wine, nor
environment surrounding the polysaccharide—both of which the bitterness elicited by ethanol in a simple model white wine
could alter its hydration state and therefore viscosity. The pos- (Gawel et al., 2016). Further addition studies utilizing real white
sible effect of polysaccharides on perceived viscosity has also wine are clearly required to properly address the effect of poly-
been demonstrated in real white wine as the ‘thickness’ of white saccharides on white wine bitterness.
Koshu wines was shown to associate with the concentration of
neutral polysaccharides (Okuda et al., 2007). However, it was
Glycerol
shown that doubling the total polysaccharide concentration of
white wine had a small but statistically insignificant effect on Glycerol is the third most abundant compound in white wine
perceived viscosity (Gawel et al., 2016). after water and ethanol. Glycerol is a product of yeast
12 R. GAWEL ET AL.

fermentation, and ranges between 5 and 10 g/L depending on Organic acids/pH


yeast strain and fermentation conditions (Nieuwoudt et al.,
White wines can vary in pH from between 3.1 and 4.0, but typi-
2002). In its pure form, glycerol is clearly viscous suggesting
cally are within the range of 3.3 to 3.5. The total acidity of white
that it may influence white wine viscosity. However, the esti-
wine is mostly the combination of tartaric and malic acid from
mated viscosity difference threshold of glycerol in white wine is
the grape, and lactic acid in the case of wines that have undergone
26 g/L which suggests that it does not influence the perception
malolactic fermentation. The fermentation product succinic acid
of viscosity (Noble and Bursick, 1984). Increasing glycerol by
which has been described as salty and acidic (Kubı́ ckova and
5 g/L in white wine (Gawel et al., 2007) and model white wine
Grosch, 1998) and umami like (Rotzoll et al., 2006; Kaneko et al.,
(Jones et al., 2008) resulted in small increases in perceived vis-
2006) is also found in reasonable concentrations in white wine.
cosity. However, in both these studies there was the possibility
Organic acids including those in white wine have been
that the perception of perceived viscosity was confounded with
shown in model studies to elicit mouth drying sensations
the inherent sweetness of glycerol. When the sweet taste of
(Hartwig and McDaniel, 1995). These sensations have been
glycerol was neutralized by blocking sweet taste receptors using
attributed to hydrogen ion concentration rather than the total
the anti-saccharin agent gymnemic acid, or was equalized by
acidity or the anion species involved (Sowalski and Noble,
the addition of a nonviscous artificial sweetener, then glycerol
1998; Lawless et al., 1996). Consistent with these model studies,
was found to have no effect on perceived viscosity (Gawel and
the astringency/dryness of white wines made using a broad set
Waters, 2008). The sweetness of glycerol may also be a factor in
of commercial juice extraction and handling methods consis-
the reduced bitterness via mixture suppression as observed by
tently increased with decreasing pH, but astringency was unre-
Jones et al. (2008) in model white wine.
lated to total phenolic concentration (Gawel et al., 2014a).
The increases in perceived astringency/dryness with decreas-
ing pH may be explained by reduced salivary viscosity
Ethanol (Veerman et al., 1989; Nordbo et al., 1984) possibly resulting
from greater interaction after charged salivary proteins neutral-
Increased hotness in white wine due to ethanol can be per-
ize at pH’s closer to their isoelectric point. Alternative explana-
ceived even within the relatively narrow range of concentra-
tions could be an increased binding of astringent compounds
tions typically encountered in white wine (Gawel et al., 2007).
to oral epithelial cells which has been observed with both
Higher ethanol white wines are also perceived to be slightly
phenolic (Payne et al., 2009) and nonphenolic compounds
higher in body/density (Nurgel and Pickering, 2005).
(Ye et al., 2012), or that HC ions which are strong disruptors of
The intensity of astringent sensation declines in the presence
water structure may change the hydration environment sur-
of increasing amounts of ethanol (DeMiglio and Pickering,
rounding the solvated proteins embedded within salivary pelli-
2008; Lea and Arnold 1978). Increasing ethanol concentration
cle, possibly leading to a feeling of palate dryness.
results in reduced precipitation (Rinaldi et al., 2012) and inter-
action strength (McRae et al., 2015) between salivary proteins
and polymerized tannins. The reduced impact of polyphenols Carbon dioxide
on salivary proteins in the presence of ethanol has been postu-
lated to be due to the ability of ethanol to interfere with interac- Most bottled white wines contain some dissolved carbon diox-
tions between protein H-acceptor sites and polyphenol ide as part of their bottling specification with concentrations
hydroxyl groups, and by disrupting hydrophobic interactions typically in the range 0.5 to 1 g/L. Dissolved CO2 can create a
due to its influence on water cohesion (Pascal et al., 2008). mouthfeel character often described as ‘spritz,’ which in the
Ethanol may contribute directly to the observed reduction in case of still white wine is considered a negative quality attribute.
astringency as it has been shown to reduce the astringency of To our knowledge, no studies have been conducted on the
model wine devoid of phenolics (Fontoin et al., 2008) which effect of dissolved CO2 at white wine concentrations on the per-
suggests that it may be able to modulate the drying sensation ception of spritz or other mouthfeel characters. However, in a
elicited by organic acids. However, it is noteworthy that others closely analogous situation to white wine, 5 g/L of dissolved
have found ethanol to increase dryness under similar condi- CO2 (a level typical of semi sparkling wine) increased the per-
tions (Symoneaux et al., 2015; Jones et al., 2008). Increased eth- ceived astringency of model apple cider containing polyphenols
anol concentration may also improve the ‘quality’ of the (Symoneaux et al., 2015). Similar enhancing effects on the
astringent sensation as it has shown to enhanced the produc- astringency of model solutions by saturated solutions of CO2
tion of subqualities ‘velvety,’ and ‘mouth-coat’ (DeMiglio et al., have been observed (Hewson et al., 2009). The increased astrin-
2002), and reductions in other possibly less desirable characters gency may be due to lower pH levels resulting from the pres-
of ‘puckery,’ ‘coarse’ (Vidal et al., 2004b) and ‘grippy/adhesive’ ence of carbonic acid formed by the dissociation of CO2 in
(DeMiglio and Pickering, 2008). solution, but given that both the perception of dissolved CO2
Ethanol is inherently bitter (Mattes and DiMeglio, 2001) and and astringency may be of somatosensory origin, a direct effect
is known to stimulate the bitter taste receptor TAS2R (Allen of dissolved CO2 on perceived astringency is possible.
et al., 2014). Indeed, the difference in ethanol concentrations
encountered in white wine at around 4% contributed more to
Amino acids and peptides
the bitterness of model white wine than the flavanol catechin at
concentrations the high end of the range found in red wine It has been known for some time that many amino acids and
(Fischer and Noble, 1994). peptides are bitter (Solms, 1969). The common structural
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 13

features of bitter amino acids are that they have an L configura- would allow winemakers to better construct wines with desir-
tion and a hydrophobic side chain. The bitterness of peptides able mouthfeels and modulate their overall style.
can similarly be equated to the average hydrophobicity of their Looking ahead, a worthy goal of further research would be
amino acid components (Maga, 1990). White wine contains to provide theoretically based mixture models that can be
several hydrophobic peptides that are bitter (Desportes et al., applied in practical situations to predict the mouthfeel profiles
2001). However, it is unknown if they are found in white wine of white wine from knowledge of their composition. A prereq-
in sufficient concentrations to produce a bitter sensation either uisite for the construction of mixture models in other modali-
individually or in combination. Peptides may impact on the ties such as olfaction and taste has been requisite on a
fullness of white wine. Using an untargeted approach Skoger- knowledge of the psychophysical functions of each of the com-
son et al. (2009) found that the concentration of amino acids in ponents that make up the ‘mixture’ that is white wine. Develop-
white wine (both individually and in total) were good predic- ing these models will be challenging. Apart from the practical
tors of whether a white wine was low, medium or high in body. difficulties of obtaining sufficient pure material for large scale
sensory studies, the sensory assessment of all combinations of
compounds that could possibly impact on mouthfeel is unfeasi-
Practical implications and future directions
ble. So how could any such mixture model be simplified inas-
This review aimed to identify the aspects of white wine composi- much that it is practical but also theoretically valid? Research
tion that directly, or by interaction influence white wine mouth- in the taste modality provides some guidance in this matter. In
feel, and to elucidate the possible mechanisms that underpin that modality, similar tasting substances are known to be addi-
mouthfeel perception in white wine. Alcohol, pH and phenolics tive (Keast et al., 2003; Bartoshuk, 1975) and that in circum-
were found to influence most aspects of white wine mouthfeel stances whereby compound concentrations are correlated (as is
while the role of polysaccharides in white wine appeared to be the case with phenolics), then they add along an averaged psy-
mostly confined to moderating the negative quality attribute of chophysical function (Frijters and DeGraaf, 1987). If this were
perceived hotness and the effect of glycerol concentration on shown to be true of phenolic compounds and their relationship
mouthfeel perception of dry white wine appears to be negligible. to mouthfeel, then the summed concentrations of phenolic
These conclusions have significant implications for winemak- compounds with similar mouthfeel attributes could be mod-
ing practice. The pH and alcohol level of white wine are mostly elled rather than the impractically larger number of individual
determined by the maturity of the grape at harvest, but are rou- compounds, and that a single indicative compound could be
tinely modified by winemakers primarily to ensure that the wine used to quantify the combined effect of multiple members of
is microbiologically and chemically stable and that it has a bal- the same class of compounds. However, a further question
anced taste that meets expectations of the consumer. The litera- arises as to which phenolics have the same mouthfeel attributes.
ture suggests that modulating ethanol content and (particularly) Small changes to the substitution pattern of phenolics, i.e.,
pH are the front-line tools that can be used by winemakers to hydroxylation, esterification, glycosylation, and derivitization
creating wines with specific mouthfeel attributes. with glutathione can significantly alter the mouthfeel of pheno-
The knowledge that glycerol which is the third most abun- lic compounds indicating that classifying phenolic compounds
dant compound in white wine cannot be varied sufficiently by based on their ring structure is likely to be inadequate. There-
traditional winemaking practice to influence mouthfeel sug- fore, a broad set of structure-function relationship studies of
gests that no further work be conducted. However, in response phenolic compounds is needed to derive a general set of rules
to consumer demands for ‘fruit forward’ wines from riper which could be used to create a general classification of pheno-
grapes but with a lower alcohol content, new technologies have lic compounds based on their sensory properties.
emerged whereby carbon metabolism of the yeast Saccharomy- The overall perception of mouthfeel in white wine is a
ces cerevisiae has been directed away from ethanol in favor of response to the chemosensory, tactile and taste systems. A large
glycerol production (reviewed in Goold et al., 2017), which if body of work in the related field of gustation which also involve
translated to commercial practice could potentially increase the interactions between different sensory modalities suggest that
glycerol concentration of some white wines. the overall percepts of flavor and taste cannot be understood by
The conventional wisdom amongst many white wine mak- simply investigating its contributing parts. Prescott (1999) pro-
ers has been that low phenolic concentration in wine equates to posed that flavor might be considered as a functionally distinct
greater purity of varietal character and lower bitterness and modality cognitively constructed from the integration of the
‘coarseness’ (Singleton et al., 1975). More recently though, distinct physiological sensory systems of olfaction and taste,
winemakers have considered incorporating phenolics into their the latter of which is itself may be considered an integrated
wines to add complexity by the way of mouthfeel (McLean, construct of sensory systems in the mouth (Green, 2002). It has
2005). The review suggests that phenolics have an overarching been known for some time that the nonvolatiles food (Delwiche
effect on all major aspects of white wine mouthfeel, affecting 2004), and in white wine (Lubbers et al. 1994, 2001) can influ-
the perception of astringency, viscosity, oiliness, hotness and ence flavor perception. A few studies have also demonstrated
bitterness. However, while most winemakers know how to the reverse case of retronasally perceived volatile compounds
influence total phenolic concentration in their wines, and some influencing mouthfeel perception (Oladokun et al., 2017;
are also aware of how they can influence the phenolic makeup Symoneaux et al. 2015; Saenz-Navajas et al., 2010; Jones et al.
of their wines by their production decisions, knowledge as to 2008). Further work needs to be conducted to establish whether
how they can influence mouthfeel via phenolic management is the interaction between volatile and nonvolatile components in
lacking. Such knowledge would present a powerful tool as it white wine is of a symmetric or asymmetric nature. Forty years
14 R. GAWEL ET AL.

ago, Van Bursick and Erickson (1977) proposed that the tri- Betes-Saura, C., Andres-Lacueva, C., and Lamuela-Ravent os, R. M. (1996).
geminal system acts to bind the physiologically distinct olfac- Phenolics in white free run juices and wines from Penedes by high-per-
tory and gustatory systems into a single perceptual system formance liquid chromatography: Changes during vinification. J. Agric.
Food Chem. 44:3040–3046.
known as flavor. Under a similar integrated interpretation, Cala, O., Dufourc, E. J., Fouquet, E., Manigand, C., Laguerre, M., and Pia-
mouthfeel in white wine should also be considered as a multidi- net, I. (2012). The colloidal state of tannins impacts the nature of their
mensional sensory construct and future research be conducted interaction with proteins: the case of salivary proline-rich protein/pro-
accordingly. cyanidins binding. Langmuir 28:17410–17418.
The measurement of mouthfeel provides unique challenges Cala, O., Pinaud, N., Simon, C., Fouquet, E., Laguerre, M., Dufourc, E. J.,
and Pianet, I. (2010). NMR and molecular modeling of wine tannins
due to the heterogeneity but physical proximity of the receptive binding to saliva proteins: revisiting astringency from molecular and
systems involved and that the physiological mechanisms that colloidal prospects. FASEB J. 24:4281–4290.
govern perception of mouthfeel attributes are unclear. Astrin- Canon, F., Pate, F., Cheynier, V., Sarni-Manchado, P., Giuliani, A., Perez,
gency is a case in point. For over a half a century the dominant J., Durand, D., Li, J., and Cabane, B. (2013). Aggregation of the salivary
paradigm for astringency has been that of polyphenol-salivary proline-rich protein IB5 in the presence of the tannin EgCG. Langmuir
29:1926–1937.
tannin interaction leading the changes to salivary rheology that Canon, F., Ballivian, R., Chirot, F., Antoine, R., Sarni-Manchado, P.,
are detected by oral mechanoreceptors. Yet while a substantial Lemoine, J., and Dugourd, P. (2011). Folding of a salivary intrinsically
body of work has identified the basic mechanisms of salivary disordered protein upon binding to tannins. J. Am. Chem. Soc.
protein and wine polyphenol binding (some) leading to a 133:7847–7852.
change to salivary rheology, it is yet to be demonstrated how Carando, S., Teissedre, P., Pascual-Martinez, L., and Cabanis, J. (1999a). Levels
of flavan-3-ols in French wines. J. Agric. Food Chem. 47:4161–4166.
these changes in the bulk phase of saliva translate to astrin- Carando, S., Teissedre, P. L., and Cabanis, J. C. (1999b). HPLC coupled
gency perception. More recent findings using in vitro receptor with fluorescence detection for the determination of procyanidins in
based studies indicate that astringency is a chemoreceptive pro- white wines. Chromatographia 50:253–254.
cess which must therefore necessarily involve binding of astrin- Cardenas, M., Elofsson, U., and Lindh, L. (2007). Salivary mucin MUC5B
gent molecules to the oral mucosa. However, the issue of could be an important component of in vitro pellicles of human saliva:
An in situ ellipsometry and atomic force microscopy study. Biomacro-
possible interaction between polyphenols and salivary proteins molecules 8:1149–1156.
in the bulk saliva or salivary pellicle preceding the binding have Carpenter, G. H. (2013). Do transient receptor protein (TRP) channels
not been adequately addressed in these studies. Therefore, a play a role in oral astringency? J. Texture Stud. 44:334–337.
more comprehensive and integrated approach involving inves- Carvalho, E., Mateus, N., Plet, B., Pianet, I., Dufourc, E., and De Freitas, V.
tigation of interaction of astringent substances with the salivary (2006). Influence of wine polysaccharides on the interactions between con-
densed tannins and salivary proteins. J. Agric. Food Chem. 54:8936–8944.
phase, mucosal binding mechanisms, and crucially, which Castillo-Mu~ noz, N., Gomez-Alonso, S., Garcıa-Romero, E., and Hermosin-
receptive fields are innervated resulting from changes to sali- Gutierrez, I. (2010). Flavonol profiles of Vitis vinifera white grape culti-
vary rheology and/or mucosal binding is required if astringency vars. J. Food Composit. Anal. 23:699–705.
perception is to be fully understood. Cejudo-Bastante, M. J., Castro-Vazquez, L., Hermosın-Gutierrez, I., and
Perez-Coello, M. S. (2011a). Combined effects of prefermentative skin
maceration and oxygen addition of must on color-related phenolics,
Acknowledgments volatile composition, and sensory characteristics of Airen wine. J. Agric.
Food Chem. 59:12171–12182.
The work was funded by Australia’s grapegrowers and winemakers Cejudo-Bastante, M. J., Hermosın-Gutierrez, I., Castro-Vazquez, L., and
through their investment body, Wine Australia, with matching funds from Perez-Coello, M. S. (2011b). Hyperoxygenation and bottle storage of
the Australian government. Chardonnay white wines: Effects on color-related phenolics, volatile com-
position, and sensory characteristics. J. Agric. Food Chem. 59:4171–4182.
Cejudo-Bastante, M. J., Perez-Coello, M. S., and Hermosın-Gutierrez, I.
References (2010). Identification of new derivatives of 2-S-Glutathionylcaftaric
acid in aged white wines by HPLC-DAD-ESI-MSn. J. Agric. Food
Adler, E., Hoon, M. A., Mueller, K. L., Chandrashekar, J., Ryba, N. J. P., Chem. 58:11483–11492.
and Zuker, C. S. (2000). A novel family of mammalian taste receptors. Chandrashekar, J., Mueller, K. L., Hoon, M. A., Adler, E., Feng, L. X., Guo,
Cell 100:693–702. W., Zuker, C. S., and Ryba, N. J. P. (2000). T2Rs function as bitter taste
Allen, A. L., McGeary, J. E., and Hayes, J. E. (2014). Polymorphisms in receptors. Cell 100:703–711.
TRPV1 and TAS2Rs associated with sensations from sampled ethanol. Cheynier, V. F., Trousdale, E. K., Singleton, V. L., and Salgues, M. J. (1986).
Alcohol: Clin. Exp. Res. 38:2550–2560. Characterization of 2-s-Glutathionylcaftaric acid and its hydrolysis in
Bacon, J. R., and Rhodes, M. J. (2000). Binding affinity of hydrolyzable tan- relation to grape wines. J. Agric. Food Chem. 34:217–221.
nins to parotid saliva and to proline-rich proteins derived from it. J. Clapham, E. E. (2003). TRP channels as cellular sensors. Nature 426:517–524.
Agric. Food Chem. 48:838–843. Coles, J. M., Chang, D. P., and Zauscher, S. (2010). Molecular mechanisms
Baderschneider, B., and Winterhalter, P. (2001). Isolation and characteri- of aqueous boundary lubrication by mucinous glycoproteins. Curr.
zation of novel benzoates, cinnamates, flavonoids, and lignans from Opin. Colloid Interf. Sci. 15:406–416.
Riesling wine and screening for antioxidant activity. J. Agric. Food Dadic, M., and Belleau, G. (1973). Polyphenols and beer flavor. Proc. Am.
Chem. 49:2788–2798. Soc. Brewing Chem. 4:107–114.
Bartoshuk, L. M. (1975). Taste mixtures: Is mixture suppression related to Delcour, J. A., Vandenberghe, M. M., Corten, P. F., and Dondeyne, P.
compression? Physiol. Behav. 14:643–649. (1984). Flavor thresholds of polyphenolics in water. Am. J. Enol. Viti-
Baxter, N. F., Lilley, T. H., Haslam, E., and Williamson, M. P. (1997). Mul- cult. 35:134–136.
tiple interactions between polyphenols and a salivary proline-rich pro- Delwiche, J. (2004). The impact of perceptual interactions on perceived fla-
tein repeat results in complexation and precipitation. Biochemistry vor. Food Qual. Prefer. 15:137–146.
36:5566–5577. DeMiglio, P., and Pickering, G. J. (2008). The influence of ethanol and pH
Berg, I. C. H., Rutland, M. W., and Arnebrant, T. (2003). Lubricating proper- on the taste and mouthfeel sensations elicited by red wine. J. Food
ties of the initial salivary pellicle—an AFM study. Biofouling 19:365–369. Agric. Environ. 6:143–150.
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 15

DeMiglio, P., Pickering, G.J. and Reynolds, A.G. (2002). Astringent description of wine astringency & bitterness and antioxidant action.
sub-qualities elicited by red wine: the role of ethanol and pH. In: Tetrahedron 71:3143–3147.
Proceedings of the International Bacchus to The Future Conference. Furlan, A. L., Castets, A., Nallet, F., Pianet, I., Grelard, A., Dufourc, E. J.,
May 23rd–25th 2002, St Catharines, ON, pp. 31–52. Cullen, C. W., and Gean, J. (2014). Red wine tannins fluidify and precipitate lipid lipo-
Pickering, G. J. and Phillips, R., Eds., Brock University Press, St somes and bicelles. A role for lipids in wine tasting? Langmuir
Catharines. 30:5518–5526.
de Pascual-Teresa, S., Rivas-Gonzalo, J. C., and Santos-Buelga, C. (2000). Gawel, R., Smith, P. A., and Waters, E. J. (2016). The influence of polysac-
Prodelphinidins and related flavanols in wine. Int. J. Food Sci. Technol. charides on the taste and mouth-feel of white wine. Aust. J. Grape
35:33–40. Wine Res. 22:350–357.
Desportes, C., Charpentier, M., Duteurtre, B., Maujean, A., and Duchiron, Gawel, R., Day, M., Van Sluyter, S. C., Holt, H., Waters, E. J., and Smith, P.
F. (2001). Isolation, identification, and organoleptic characterization of A. (2014a). White wine taste and mouthfeel as affected by juice extrac-
low-molecular-weight peptides from white wine. Am. J. Enol. Viticult. tion and processing. J. Agric. Food Chem. 62:10008–10014.
52:376–380. Gawel, R., Schulkin, A., Smith, P. A., and Waters, E. J. (2014b). Taste and
de Villiers, A., Majek, P., Lynen, F., Crouch, A., Lauer, H., and Sandra, P. textural characters of mixtures of caftaric acid and Grape Reaction
(2005). Classification of South African red and white wines according Product in model wine. Aust. J. Grape Wine Res. 20:25–30.
to grape variety based on non-coloured phenolic content. Eur. Food Gawel, R., Van Sluyter, S. C., Smith, P. A., and Waters, E. J. (2013). Effect
Res. Technol. 221:520–528. of pH and alcohol on perception of phenolic character in white wine.
Di Lecce, G., Arranz, S., Jauregui, O., Tresserra-Rimbau, A., Quifer-Rada, P., Am. J. Enol. Viticult. 64:425–429.
and Lamuela-Raventos, R. M. (2014). Phenolic profiling of the skin, pulp Gawel, R., and Waters, E. J. (2008). The effect of glycerol on the perceived
and seeds of Albarino grapes using hybrid quadrupole time-of-flight and viscosity of dry white table wine. J. Wine Res. 19:109–114.
triple-quadrupole mass spectrometry. Food Chem. 145:874–882. Gawel, R., Van Sluyter, S., and Waters, E. J. (2007). The effects of ethanol
DuBois, G. (2011). Validity of early indirect models of taste active sites and and glycerol on the body and other sensory characteristics of Riesling
advances in new taste technologies enabled by improved models. Fla- wines. Aust. J. Grape Wine Res. 13:38–45.
vour Fragrance J. 26:239–253. Gawel, R. (1998). Red wine astringency: A review. Aust. J. Grape Wine Res.
Esaki, S., Tanaka, R., and Kamiya, S. (1983). Structure-taste relationships 4:74–95.
of flavanone and dihydrochalcone glycosides containing (1!2) linked Glabasnia, A., and Hofmann, T. (2006). Sensory-directed identification of
disaccharides. Agric. Biol. Chem. 47:1831–1834. taste-active ellagitannins in American (Quercus alba L.) and European
Escot, S., Feuillat, M., Dulau, L., and Charpentier, C. (2001). Release of pol- oak wood (Quercus robur L.) and quantitative analysis in Bourbon
ysaccharides by yeasts and the influence of released polysaccharides on whiskey and oak-matured red wines. J. Agric. Food Chem. 54:3380–
colur stability and wine astringency. Aust. J. Grape Wine Res. 7:153– 3390.
159. Gonçalves, F., Heyraud, A., De Pinho, M. N., and Rinaudo, M. (2002).
Fernandez-Pachon, M. S., Villa~ no, D., Troncoso, A. M., and Garcia-Par- Characterization of white wine mannoproteins. J. Agric. Food Chem.
rilla, M. C. (2006). Determination of the phenolic composition of 50:6097–6101.
sherry and table white wines by liquid chromatography and their rela- Gong, J., and Osada, Y. (1998). Gel friction: A model based on surface
tion with antioxidant activity. Anal. Chim. Acta. 563:101–108. repulsion and adsorption. J. Chem. Phys. 109:8062–8068.
Ferrer-Gallego, R., Bras, N. F., Garcıa-Estevez, I., Mateus, N., Rivas-Gon- Goold, H. D., Kroukamp, H., Williams, T. C., I.T., P., Varela, C., and Petor-
zalo, J. C., de Freitas, V., and Escribano-Bailon, M. T. (2016). Effect of ius, I. S. (2017). Yeast’s balancing act between ethanol and glycerol pro-
flavonols on wine astringency and their interaction with human saliva. duction in low-alcohol wines. Microbiol. Technol. 10:264–278.
Food Chem. 209:358–364. Green, B. G. (2002). Studying taste as a cutaneous sense. Food Qual. Prefer.
Ferrer-Gallego, R., Quijada-Morın, N., Bras, N. F., Gomes, P., de Freitas, 14:99–109.
V., Rivas-Gonzalo, J. C., and Escribano-Bail on, M. T. (2015). Charac- Guadagni, D. G., Maier, V. P., and Turnbaugh, J. G. (1973). Effect of some
terization of sensory properties of flavanols - A molecular dynamic citrus juice constituents on taste thresholds for limonin and naringin
approach. Chem. Senses 40:381–390. bitterness. J. Sci. Food Agric. 24:1277–1288.
Ferrer-Gallego, R., Hernandez-Hierro, J. M., Rivas-Gonzalo, J. C., and Guadalupe, Z., and Ayestaran, B. (2007). Polysaccharide profile and con-
Escribano-Bailon, M. T. (2014). Sensory evaluation of bitterness and tent during the vinification and aging of Tempranillo red wines. J.
astringency sub-qualities of wine phenolic compounds: synergistic Agric. Food Chem. 55:10720–10728.
effect and modulation by aromas. Food Res. Int. 62:1100–1107. G€urb€uz, O., Goçmen, D., Dagdelen, F., G€ ursoy, M., Aydin, S., Sahin, I.,
Fincher, G. B., and Stone, B. A. (1983). Arabinogalactan-proteins: Struc- B€ ukuysal, L., and Usta, M. (2007). Determination of flavan-3-ols and
uy€
ture, biosynthesis and function. Ann. Rev. Plant Physiol. 34:47–70. trans-resveratrol in grapes and wine using HPLC with fluorescence
Fischer, U., and Noble, A. C. (1994). The effect of ethanol, catechin concentra- detection. Food Chem. 100:518–525.
tion, and pH on sourness and bitterness of wine. Am. J. Enol. Viticult. 45:6– Hartwig, P., and McDaniel, M. R. (1995). Flavor characteristics of lactic,
10. malic, citric and acetic acids at various pH levels. J. Food Sci. 60:384–
Fontoin, H., Saucier, C., Teissedre, P. L., and Glories, Y. (2008). Effect of 388.
pH, ethanol and acidity on astringency and bitterness of grape seed Haslam, E., and Lilley, T. H. (1988). Natural astringency in foodstuffs: A
tannin oligomers in model wine solution. Food Qual. Prefer. 19:286– molecular interpretation. Crit. Revues Food Sci. Nutr. 27:1–40.
291. Herrick, I. W., and Nagel, C. W. (1985). The caffeoyl tartrate content of
Foo, L. Y., and Porter, L. J. (1980). The phytochemistry of proanthocyani- White Riesling wines from California, Washington, and Alsace. Am. J.
din polymers. Phytochemistry 19:1747–1754. Enol. Viticult. 36:95–97.
Friedel, M., Stoll, M., Patz, C. D., Will, F., and Dietrich, H. (2015). Impact Hewson, L., Hollowood, T., Chandra, S., and Hort, J. (2009). Gustatory,
of light exposure on fruit composition of white ‘Riesling’ grape berries olfactory and trigeminal interactions in a model carbonated beverage.
(Vitis vinifera L.). Vitis 54:107–116. Chemosens Percept. 2:94–107.
Frijters, J. E. R., and DeGraaf, C. (1987). The equiratio taste mixture model Horowitz, R. M., and Gentili, B. (1969). Taste and structure of phenolic
successfully predicts the sensory response to the sweetness intensity of glycosides. J. Agric. Food Chem. 17:696–700.
complex mixtures of sugars and sugar alcohols. Perception 5:615–628. Hufnagel, J. C., and Hofmann, T. (2008). Orosensory-directed identifica-
Fulcrand, H., Remy, S., Souquet, J. M., Cheynier, V., and Moutounet, M. tion of astringent mouthfeel and bitter-tasting compounds in red wine.
(1999). Study of wine tannin oligomers by on-line liquid chromatogra- J. Agric. Food Chem. 56:1376–1386.
phy electrospray ionisation mass spectrometry. J. Agric. Food Chem. Iontcheva, I., Oppenheim, F. G., Offner, G. D., and Troxler, R. F. (2000).
47:1023–1028. Molecular mapping of statherin- and histatin-binding domains in
Furlan, A. L., Jobin, M.-L., Pianet, I., Dufourc, E. J., and Gean, J. (2015). human salivary mucin MG1 (MUC5B) by the yeast two-hybrid system.
Flavanol/lipid interaction: a novel molecular perspective in the J. Dental Res. 79:732–739.
16 R. GAWEL ET AL.

Jackson, R. (2014). Wine Science: Principles and Applications. Academic McArthur, C., Sanson, G. D., and Beal, A. M. (1995). Salivary proline-rich
Press, San Diego. proteins in mammals: Roles in oral homeostasis and counteracting die-
Jones, P. R., Gawel, R., Francis, I. L., and Waters, E. J. (2008). The tary tannin. J. Chem. Ecol. 21:663–691.
influence of interactions between major white wine components on McLean, H. (2005). Managing phenolics in white wine. In: Advances in
the aroma, flavour and texture of model white wine. Food Qual. Tannin Management, pp. 36–39. Allen, M., Dundon, C., Francis, M.E.,
Prefer. 19:596–607. Howell, G., and Wall, G., Eds., Australian Society of Viticulture and
Kallithraka, S., Salacha, M. I., and Tzourou, I. (2009). Changes in phenolic Oenology Inc., Adelaide.
composition and antioxidant activity of white wine during bottle storage: McRae, J. M., Ziora, Z. M., Kassara, S., Cooper, M. A., and Smith, P. A.
Accelerated browning test versus bottle storage. Food Chem. 113:500–505. (2015). Ethanol concentration influences the mechanisms of wine tan-
Kaneko, S., Kumazawa, K., Masuda, H., Henze, A., and Hofmann, T. nin Interactions with poly(L-proline) in model wine. J. Agric. Food
(2006). Molecular and sensory studies on the umami taste of Japanese Chem. 63:4345–4352.
green tea. J. Agric. Food Chem. 54:2688–2694. Meyerhof, W., Batram, C., Kuhn, C., Brockhoff, A., Chudoba, E., Bufe, B.,
Kauffman, D. L., Bennick, A., Blum, M., and Keller, P. J. (1991). Basic pro- Appendino, G., and Behrens, M. (2010). The molecular receptive ranges
line-rich proteins from human parotid saliva: Relationships of the of human TAS2R bitter taste receptors. Chem. Senses 35:157–170.
covalent structures of ten proteins from a single individual. Biochemis- Monagas, M., Bartolome, B., and Gomez-Cordoves, C. (2005). Updated
try 30:3351–3356. knowledge about the presence of phenolic compounds in wine. Crit.
Keast, R. S. J., Bournazel, M. M. E., and Breslin, P. A. S. (2003). A psycho- Rev. Food Sci. Nutr. 45:85–118.
physical investigation of binary bitter-compound interactions. Chem. Myers, T. E., and Singleton, V. L. (1979). The nonflavonoid phenolic frac-
Senses 28:301–313. tion of wine and its analysis. Am. J. Enol. Viticult. 30:98–102.
Kielhorn, S., and Thorngate, J. H. (1999). Oral sensations associated with Narukawa, M., Kimata, H., Noga, C., and Watanabe, T. (2010). Taste char-
the flavan-3-ols (C)-catechin and (¡)-epicatechin. Food Qual. Prefer. acterisation of green tea catechins. Int. J. Food Sci. Technol. 45:1579–
10:109–116. 1585.
Konitz, R., Fruend, M., Seckler, J., Christmann, M., Netzel, M., Strass, G., Nayak, A., and Carpenter, G. H. (2008). A physiological model of tea-
Bitsch, R., and Bitsch, I. (2003). Einfluss der Mostvorkl€arung auf die induced astringency. Physiol. Behav. 95:290–294.
sensorische qualit€at von Riesling weinen aus dem Rheingau. Mitteilun- Nicolini, G., Mattivi, F., and Dalla Serra, A. (1991). Iperossigenazione dei
gen Klosterneuburg 53:166–183. mosti: Conseguenze analitiche e sensoriali su vini dell vendemmia
Korenika, A. J., Zulj, M. M., Puhelek, I., Plavsa, T., and Jeromel, A. (2014). 1989. Rivista di Viticoltura e di Enologia 3:45–56.
Study of phenolic composition and antioxidant capacity of Croatian Nieuwoudt, H. H., Prior, B. A., Pretorius, I. S., and Bauer, F. F. (2002).
macerated wines. Mitteilungen Klosterneuburg 64:171–180. Glycerol in South African table wines. An assessment of its relationship
Kosmerl, T., Abramovic, H., and Klofutar, C. (2000). The rheological prop- to wine quality. S. Afric. J. Enol. Viticult. 23:22–30.
erties of Slovenian wines. J. Food Eng. 46:165–171. Noble, A. C., and Bursick, G. F. (1984). The contribution of glycerol to per-
Kubı́ ckova, J., and Grosch, W. (1998). Evaluation of flavour compounds of ceived viscosity and sweetness in white wine. Am. J. Enol. Viticult.
Camembert cheese. Int. Dairy J. 8:11–16. 35:110–112.
Kurogi, M., Kawai, Y., Nagatomo, K., Tateyama, M., Kubo, Y., and Saitoh, Nordbo, H., Darwish, S., and Bhatnagar, R. S. (1984). Salivary viscosity and
O. (2015). Auto-oxidation products of epigallocatechin gallate activate lubrication: Influence of pH and calcium. Scand. J. Dental Res. 92:306–
TRPA1 and TRPV1 in sensory neurons. Chem. Senses 40:27–46. 314.
Lawless, H. T., Horne, J., and Giasi, P. (1996). Astringency of organic acids Nurgel, C., and Pickering, G. J. (2005). Contribution of glycerol, ethanol
is related to pH. Chem. Senses 21:397–403. and sugar to the perception of viscosity and density elicited model
Lea, A. G. H., Bridle, P., Timberlake, C. F., and Singleton, V. L. (1979). The white wines. J. Texture Stud. 36:303–325.
procyanidins of white grapes and wines. Am. J. Enol. Viticult. 30:289– Oberholster, A., Francis, I. L., Iland, P. G., and Waters, E. J. (2009). Mouth-
300. feel of white wines made with and without pomace contact and added
Lea, A. G. H., and Arnold, G. M. (1978). The phenolics of ciders: Bitterness anthocyanins. Aust. J. Grape Wine Res. 15:59–69.
and astringency. J. Sci. Food Agric. 29:478–483. Okamura, S., and Watanabe, M. (1981). Determination of phenolic cinna-
Lee, S. K., Lee, S. W., Chung, S. C., Kim, Y. K., and Kho, H. S. (2002). Anal- mates in white wine and their effect on wine quality. Agric. Biol. Chem.
ysis of residual saliva and minor salivary gland secretions in patients 45:2063–2070.
with dry mouth. Arch Oral Biol. 47:637–641. Okuda, T., Fukui, M., Hisamoto, M., Iino, S., Hikawa, Y., Ogino, S.,
Lubbers, S., Verret, C., and Voilley, A. (2001). The effect of glycerol on the Takayanagi, T., and Yokotsuka, K. (2007). Relationships between mac-
perceived aroma of a model wine and white wine. Food Sci. Technol. romolecules and thickness of dry white wine. J. Am. Soc. Enol. Viticult.
34:262–265. Jpn. 18:15–21.
Lubbers, S., Voilley, A., Feuillat, M., and Charpentier, C. (1994). Influence Oladokun, O., James, S., Cowley, T., Dehrmann, F., Smart, K., Hort, J.,
of mannoproteins from yeast on the aroma intensity of a model wine. and Cook, D. (2017). Perceived bitterness character of beer in rela-
Food Sci. Technol. 27:108–114. tion to hop variety and the impact of hop aroma. Food Chem.
Luck, G., Liao, H., Murray, N. J., Grimmer, H. R., Warminski, E. E., Wil- 230:215–224.
liamson, M. P., Lilley, T. H., and Haslam, E. (1994). Polyphenol, astrin- Ong, B. Y., and Nagel, C. W. (1978). Hydroxycinnamic acid-tartaric acid
gency and proline rich proteins. Phytochemistry 37:357–371. ester content in mature grapes and during the maturation of White
Maga, J. A. (1990). Compound structure versus bitter taste. In: Bitterness in Riesling grapes. Am. J. Enol. Viticult. 29:277–281.
Foods and Beverages. pp. 35–48. Rouseff, R. L. Eds. Elsevier, Amsterdam. Oszmianski, J., and Sapis, J. C. (1989). Fractionation and identification of
Maga, J. A., and Lorenz, K. (1973). Taste threshold values for phenolic some low molecular weight grape seed phenolics. J. Agric. Food Chem.
acids which can influence flavor properties of certain flours, grains, and 37:1293–1297.
oilseeds. Cereal Sci. Today. 18:326–328. Pascal, C., Pate, F., Cheynier, V., and Delsuc, M. A. (2009). Study of the
Maheshwari, R., and Dhathathreyan, A. (2006). Mucin at solution/air and interactions between a proline-rich protein and a flavan-3-ol by NMR:
solid/solution interfaces. J. Colloid Interface Sci. 293:263–269. Residual structures in the natively unfolded protein provides anchorage
Malone, M. E., Appelqvist, I. A. M., and Norton, I. T. (2003). Oral behav- points for the ligand. Biopolymers 91:745–756.
iour of food hydrocolloids and emulsions. Part 1. Lubrication and Pascal, C., Poncet-Legrand, C., Cabane, B., and Verhnet, A. (2008). Aggre-
deposition considerations. Food Hydrocoll. 17:763–773. gation of a proline-rich protein induced by epigallocatechin gallate and
Masa, A., Vilanova, A., and Pomar, F. (2007). Varietal differences among condensed tannins: Effect of protein glycosylation. J. Agric. Food Chem.
the flavonoid profiles of white grape cultivars studied by high-perfor- 56:6724–6732.
mance liquid chromatography. J. Chromatogr. A 1164:291–297. Payne, C., Bowyer, P. K., Herderich, M., and Bastian, S. E. P. (2009). Inter-
Mattes, R. D., and DiMeglio, D. (2001). Ethanol perception and ingestion. action of astringent grape seed procyanidins with oral epithelial cells.
Physiol Behav. 72:217–229. Food Chem. 115:551–557.
CRITICAL REVIEWS IN FOOD SCIENCE AND NUTRITION 17

Peleg, H., Gacon, K., Schlich, P., and Noble, A. C. (1999). Bitterness and Schobel, N., Radtke, D., Kyereme, J., Wollmann, N., Cichy, A., Obst, K.,
astringency of flavan-3-ol monomers, dimers and trimers. J. Sci. Food Kallweit, K., Kletke, O., Minovi, A., Dazert, S., Wetzel, C. H., Vogt-
Agric. 79:1123–1128. Eisele, A., Gisselmann, G., Ley, J. P., Bartoshuk, L. M., Spehr, J., Hof-
Peleg, H., and Noble, A. C. (1995). Perceptual properties of benzoic acid mann, T., and Hatt, H. (2014). Astringency is a trigeminal sensation
derivatives. Chem. Sens. 20:393–400. that involves the activation of G protein-coupled signaling by phenolic
Pellerin, P., Doco, T., Vidal, S. P. W., Brillouet, J. M., and O’Neill, M. A. compounds. Chem. Senses 39:471–487.
(1996). Structural characterization of red wine rhamnogalacturonan II. Schwender, N., Huber, K., Al Marrawi, F., Hannig, M., and Ziegler, C.
Carbohydr. Res. 190:183–197. (2005). Initial bioadhesion on surfaces in the oral cavity investigated by
Pe~na-Neira, A., Hernandez, T., Garcia-Vallejo, C., Estrella, I., and scanning force microscopy. Appl. Surf. Sci. 252:117–122.
Suarez, J. A. (2000). A survey of phenolic compounds in Spanish Simonetti, P., Pietta, P., and Testolin, G. (1997). Polyphenol content and
wines of different geographical origin. Eur. Food Res. Technol. total antioxidant potential of selected Italian wines. J. Agric. Food
210:445–448. Chem. 45:1152–1155.
Pickering, G. J., and DeMiglio, P. (2008). The wine wine mouth-feel wheel: Singleton, V. L., Salgues, M., Zaya, J., and Trousdale, E. (1985). Caftaric
A lexicon for describing the oral sensations elicited by white wine. J. acid disappearance and conversion to products of enzymic oxidation in
Wine Res. 19:51–67. grape must and wine. Am. J. Enol. Viticult. 36:50–56.
Plet, B., Delcambre, A., Chaignepain, S., and Schmitter, J.-M. (2015). Affin- Singleton, V. L., and Trousdale, E. (1983). White wine phenolics—varietal
ity ranking of peptide–polyphenol non-covalent assemblies by mass and processing differences as shown by HPLC. Am. J. Enol. Viticult.
spectrometry approaches. Tetrahedron 71:3007–3011. 34:27–34.
Poncet-Legrand, C., Edelmann, A., Putaux, J.-L., Cartalade, D., Sarni-Man- Singleton, V. L., Zaya, J., and Trousdale, E. (1980). White table wine qual-
chado, P., and Vernhet, A. (2006). Poly (L-proline) interactions with ity and polyphenol composition as affected by must SO2 content and
flavan-3-ols units: Influence of the molecular structure and the poly- pomace contact time. Am. J. Enol. Viticult. 31:14–20.
phenol/protein ratio. Food Hydrocoll. 20:687–697. Singleton, V. L., Sieberhagen, H. A., de Wet, P., and van Wyk, C. J. (1975).
Prescott, J. (1999). Flavour as a psychological construct: implications for Composition and sensory qualities of wines prepared from white
perceiving and measuring the sensory quality of foods. Food Qual. Pre- grapes by fermentation with and without grape solids. Am. J. Enol. Viti-
fer. 10:349–356. cult. 26:62–69.
Proctor, G. B., Hamdan, S., Carpenter, G. H., and Wilde, P. (2005). A sta- Skogerson, K., Runnebaum, R., Wohlgemuth, G., de Ropp, J., Heymann,
therin and calcium enriched layer at the air interface of human parotid H., and Fiehn, O. (2009). Comparison of gas chromatography-coupled
saliva. Biocheml. J. 389:111–116. time-of-flight mass spectrometry and 1H nuclear magnetic resonance
Quijada-Morın, N., Williams, P., Rivas-Gonzalo, J. C., Doco, T., and Escri- spectroscopy metabolite identification in white wines from a sensory
bano-Bailon, M. T. (2014). Polyphenolic, polysaccharide and oligosac- study investigating wine body. J. Agric. Food Chem. 57:6899–6907.
charide composition of Tempranillo red wines and their relationship Soares, S., Ferrer-Galego, R., Brandao, E., Silva, M., Mateus, N., and De
with the perceived astringency. Food Chem. 154:44–51. Freitas, V. (2016). Contribution of human oral cells to astringency by
Ricardo da Silva, J. M., Cheynier, V., Samsom, A., and Bourzeux, M. binding salivary protein/tannin complexes. J. Agric. Food Chem..
(1993). Effect of pomace contact, carbonic maceration, and hyperoxi- Soares, S., Kohl, S., Thalmann, S., Mateus, N., Meyerhof, W., and De Frei-
dation on the procyanidin composition of Grenache blanc wines. Am. tas, V. (2013). Different phenolic compounds activate distinct human
J. Enol. Viticult. 44:168–172. bitter taste receptors. J. Agric. Food Chem. 61:1525–1533.
Rinaldi, A., Gambuti, A., and Moio, L. (2012). Precipitation of salivary pro- Soares, S. I., Gonçalves, R. M., Fernandes, I., Mateus, N., and de Freitas, V.
teins after the interaction with wine: the effect of ethanol, pH, fructose, (2009). Mechanistic approach by which polysaccharides inhibit alpha-
and mannoproteins. J. Food Sci. 77:C485–490. amylase/procyanidin aggregation. J. Agric. Food Chem. 57:4352–4358.
Robichaud, J. L., and Noble, A. C. (1990). Astringency and bitterness of Sokolowsky, M., and Fischer, U. (2012). Evaluation of bitterness in white
selected phenolics in wine. J. Sci. Food Agric. 53:343–353. wine applying descriptive analysis, time-intensity analysis, and tempo-
Robinson, J. (1999). Vines Grapes and Wines. Mitchell Beazley, London. ral dominance of sensations analysis. Anal. Chim. Acta. 732:46–52.
Rodrıgues Montealegre, R., Romero Peces, R., Chac on Vozmediano, J. L., Solms, J. (1969). The taste of amino acids, peptides and proteins. J. Agric.
Martinez Gascue~ na, J., and Carcia Romero, E. (2006). Phenolic com- Food Chem. 17:686–688.
pounds in skins and seeds of ten grape Vitis vinifera varieties grown in Sowalski, R. A., and Noble, A. C. (1998). Comparison of the effects of con-
warm climates. J. Food Composit. Anal. 19:687–693. centration, pH and anion species on astringency and sourness of
Rossetti, D., Bongaerts, J. H. H., Wantling, E., Stokes, J. R., and William- organic acids. Chem. Senses 23:343–349.
son, A. M. (2009). Astringency of tea catechins: More than an oral Spencer, C. M., Cai, Y., Martin, R., Gaffney, S. H., Goulding, P. N., Magno-
lubrication tactile percept. Food Hydrocoll. 23:1984–1992. lato, D., Lilley, T. H., and Haslam, E. (1988). Polyphenol complexation:
Rotzoll, N., Dunkel, A., and Hofmann, T. (2006). Quantitative studies, Some thoughts and observations. Phytochemistry 27:2397–2409.
taste reconstitution, and omission experiments on the key taste com- Squier, C. A., and Kremer, M. J. (2001). Biology of the oral mucosa and
pounds in Morel mushrooms (Morchella deliciosa fr.). J. Agric. Food esophagus. J. Natl. Cancer Inst. Monogr. 29:7–15.
Chem. 54:2705–2711. Svendsen, I. E., Lindh, L., and Arnebrant, T. (2006). Adsorption behaviour
Runnebaum, R. C., Boulton, R. B., Powell, R. L., and Heymann, H. (2011). and surfactant elution of cationic salivary proteins at solid/liquid interfa-
Key constituents affecting wine body—an exploratory study. J. Sens. ces, studied by in situ ellipsometry. Colloids Surf. B-Biointerf. 63:157–166.
Stud. 26:62–70. Symoneaux, R., Le Quere, J. M., Baron, A., Bauduin, R., and Chollet, S.
Saenz-Navajas, M. P., Avizcuri, J. M., Ferreira, V., and Fernandez-Zur- (2015). Impact of CO2 and its interactions with the matrix components
bano, P. (2012). Insights on the chemical basis of the astringency of on sensory perception in model cider. LWT – Food Sci. Technol.
Spanish red wines. Food Chem. 134:1484–1493. 63:886–891.
Saenz-Navajas, M. P., Campo, E., Fernandez-Zurbano, P., Valentin, D., Szabo, D., Akiyoshi, S., Matsunaga, T., Gong, J. P., and Osada, Y. (2000).
and Ferreira, V. (2010). An assessment of the effects of wine volatiles Spreading of liquids on gel surfaces. J. Chem. Phys. 113:8253–8259.
on the perception of taste and astringency in wine. Food Chem. Takahashi, K., Tadenuma, M., Kitamoto, K., and Sato, S. (1974). L-Prolyl-
121:1139–1149. L-leucine anhydride: A bitter compound formed in aged sake. Agric.
Sarni-Manchado, P., and Cheynier, V. (2002). Study of non-covalent com- Biol. Chem. 38:927–932.
plexation between catechin derivatives and peptides by electrospray Thorngate, J. H., and Noble, A. C. (1995). Sensory evaluation of bitterness
ionization mass spectrometry. J. Mass Spectrom. 37:609–616. and astringency of 3R(¡)-epicatechin and 3S(C)-catechin. J. Sci. Food
Scharbert, S., Holzmann, N., and Hofmann, T. (2004). Identification of Agric. 67:531–535.
the astringent taste compounds in black tea by combining instru- Trevisani, M., Smart, D., Gunthorpe, M. J., Tognetto, M., Barbieri, M.,
mental analysis and human bioresponse. J. Agric. Food Chem. Campi, B., Amadesi, S., Gray, J., Jerman, J. C., Brough, S. J., Owen, D.,
52:3498–3508. Smith, G. D., Randall, A. D., Harrison, S., Bianchi, A., Davis, J. B., and
18 R. GAWEL ET AL.

Geppetti, P. (2002). Ethanol elicits and potentiates nociceptor Vitkov, L., Hannig, M., Nekrashevych, Y., and Krautgartner, W. D. (2004).
responses via the vanilloid receptor-1. Nat. Neurosci. 5:546–551. Supramolecular pellicle precursors. Eur. J. Oral Sci. 112:320–325.
Van Bursick, R. L., and Erickson, R. P. (1977). Odorant responses in taste Vitrac, X., Monti, J. P., Vercauteren, J., Deffieux, G., and Merillon, J. M.
neurons of the rat NTS. Brain Res. 135:287–303. (2002). Direct liquid chromatographic analysis of resveratrol deriva-
Van Vliet, T., van Aken, G. A., de Jongh, H. H., and Hamer, R. J. (2009). Colloi- tives and flavanonols in wines with absorbance and fluorescence detec-
dal aspects of texture perception. Adv. Colloid Interface Sci. 150:27–40. tion. Anal. Chim. Acta. 458:103–110.
Veerman, E. C. I., Valentijm-Benz, M., and Nieuw Amerongen, A. V. Walz, A., St€uhler, K., Wattenberg, A., Hawranke, E., Meyer, H. E., Schmalz,
(1989). Viscosity of human salivary mucins: Effect of pH and ionic G., Bl€uggel, M., and Ruhl, S. (2006). Proteome analysis of glandular
strength and role of sialic acid. J. Biol. Buccale 17:297–306. parotid and submandibular-sublingual saliva in comparison to whole
Verette, E., Noble, A. C., and Somers, T. C. (1988). Hydroxy-cinnamates of human saliva by two-dimensional gel electrophoresis. Proteomics
Vitis vinifera: Sensory assessment in relation to bitterness in white 6:1631–1639.
wines. J. Sci. Food Agric. 45:267–272. Watanabe, I. (2004). Ultrastructures of mechanoreceptors in the oral
Verhagen, J. V., and Engelen, L. (2006). The neurocognitive bases of mucosa. Anatom. Sci. Int. 79:55–61.
human multimodal food perception: Sensory integration. Neurosci. Work, T. M., and Camire, M. E. (1996). Phenolic acid detection thresholds
Biobehav. Rev. 30:613–650. in processed potatoes. Food Qual. Prefer. 7:271–274.
Verhnet, A., Pellerin, P., Prieur, C., Osmianski, J., and Moutounet, M. Yamazaki, T., Narukawa, M., Mochizuki, M., Misaka, T., and Watanabe, T.
(1996). Charge properties of some grape and wine polysaccharide and (2013). Activation of the hTAS2R14 human bitter-taste receptor by
polyphenolic fractions. Am. J. Enol. Viticult. 47:25–30. (¡)-epogallocatechin gallate and (¡)-epcatechin gallate. Biosci., Bio-
Vidal, S., Francis, L., Williams, P., Kwiatkowski, M., Gawel, R., Chey- technol. Biochem. 77:1981–1983.
nier, V., and Waters, E. (2004a). The mouth-feel properties of pol- Yao, Y., Berg, E. A., Costello, C. E., Troxler, R. F., and Oppenheim, F. G.
ysaccharides and anthocyanins in a wine like medium. Food Chem. (2003). Identification of protein components in human acquired
85:519–525. enamel pellicle and whole saliva using novel proteomics approaches. J.
Vidal, S., Courcoux, P., Francis, L., Kwiatkowski, M., Gawel, R., Williams, Biol. Chem. 278:5300–5308.
P., Waters, E., and Cheynier, V. (2004b). Use of an experimental design Ye, A., Zheng, T., Ye, J. Z., and Singh, H. (2012). Potential role of the bind-
approach for evaluation of key wine components on mouth-feel per- ing of whey proteins to human buccal cells on the perception of astrin-
ception. Food Qual. Prefer. 15:209–217. gency in whey protein beverages. Physiol. Behav. 106:645–650.
Vilanova, M., Santalla, M., and Masa, A. (2009). Environmental and Yu, P., Yeo, A. S., Low, M., and Zhou, W. (2014). Identifying key non-vola-
genetic variation of phenolic compounds in grapes (Vitis vinifera) from tile compounds in ready-to-drink green tea and their impact on taste
northwest Spain. J. Agric. Sci. 147:683–697. profile. Food Chem. 155:9–16.
APPENDIX II

Authorship Declarations
Signature Redacted by Library
Signature Redacted by Library

Signature Redacted by Library


Signature Redacted by Library
Signature Redacted by Library
Signature Redacted by Library
Signature Redacted by Library

Signature Redacted by Library

Signature Redacted by Library


Signature Redacted by Library
Signature Redacted by Library

Signature Redacted by Library

Signature Redacted by Library


Signature Redacted by Library
Signature Redacted by Library

Signature Redacted by Library

Signature Redacted by Library

Signature Redacted by Library


Signature Redacted by Library
Signature Redacted by Library

Signature Redacted by Library


1

You might also like