Open access peer-reviewed chapter

Salivary Protein-Tannin Interaction: The Binding behind Astringency

Written By

Alessandra Rinaldi and Luigi Moio

Submitted: 09 June 2020 Reviewed: 17 August 2020 Published: 05 October 2020

DOI: 10.5772/intechopen.93611

From the Edited Volume

Chemistry and Biochemistry of Winemaking, Wine Stabilization and Aging

Edited by Fernanda Cosme, Fernando M. Nunes and Luís Filipe-Ribeiro

Chapter metrics overview

1,201 Chapter Downloads

View Full Metrics

Abstract

Interactions between salivary proteins and tannins are at the basis of one of the main mechanisms involved in the perception of astringency. Astringency is a tactile sensation evoked in the mouth by plant polyphenol-derived products, such as red wine. It is generally recognised that tannins can provoke negative sensations such as shrinking, drawing, or puckering of the epithelium. On the other hand, the astringency of some red wines can be felt as pleasant mouth feelings of richness, fullness, mouth-coating, and velvet in the mouth. In this chapter, an overview of the research concerned with molecular and sensory mechanisms of astringency was updated. Because of many variables influence the perception of astringency, several methods have been developed to measure the intensity of the sensation. In this context, different indirect assessments were critically evaluated considering the pros and contras and correlated with sensory analysis. We focused the attention on the saliva precipitation index (SPI), based on the binding and precipitation of human saliva with grape and wine tannins, because it has been widely used for many applications in winemaking. A current great challenge is to have an in vitro measurement of astringency able to provide information on the fate of wine, from grape to bottle.

Keywords

  • astringency
  • salivary proteins
  • polyphenols
  • precipitation
  • methods

1. Introduction

The interaction between plant tannins and macromolecules such as proteins is at the basis of many processes involved in the industry, ecological and agricultural systems [1, 2, 3], and food and beverage sensory characteristics. The common factor is the binding between macromolecules and tannins that lead to: (i) the conversion of an animal hide into the leather (tanning or tannage); (ii) the plant defence strategies against pathogens [1, 4]; (iii) reduced palatability of high tannin feedings to both terrestrial and marine herbivores and then a reduced interference in the process of digestion [2, 5]; (iv) the perception of astringency in tannin-rich food and beverage [6].

In the tanning process, the tannins bind to the hide’s matrix, which is composed primarily of the protein collagen ordered in microcrystalline helical units. The purpose of tannage is primarily to increase the hydrothermal stability of the structure of collagen, secondarily to increase biological inertness, and finally, to improve the utility of the hide’s physical properties [7].

In higher plants, tannins are primarily reserved as a chemical defence against pathogens. The complex with macromolecules such as cellulose and pectin, send out the exo-enzymes capable of utilising cellulose or pectin, either as a carbon source or for branching cell wall barriers to more nutrient-cytoplasm, depriving of the substrate or binding sites to these substrates. Another important function of tannin complexes is to impede the decomposition of plant litter, also when the leaf is fallen. This provides the delay in decomposition, which allows a constant input or seasonally demanding input of nutrients to the soil [1].

In the other processes, proteins of animal or human saliva interact with tannins of the unripe fruit, forages, or vegetable-derivates such as red wine, tea, and chocolate. Tannin molecules can bind proteins or enzymes at the level of specific amino acids, and modify the folding, the molecular weight, and the core binding site, to form soluble complexes or precipitates, which can alter protein function or inhibit enzyme activity [8]. This binding is at the basis of the astringent sensation experienced when tannins precipitate salivary proteins, and as a result, they lose their ability to lubricate the epithelial membranes of the mouth [6]. This sensation in mouth discourages the animal from feeding the unripe fruit or high-tannin forages and determines the unpleasantness of consumers for some tannin-rich products. These are the reason why, in the last decades, the interest in astringency has been constantly increased in different research areas.

Advertisement

2. Perception of astringency

The term astringency derives from the Latin verb, ad-stringere that means tightly bind, strongly join. It refers to the propensity of vegetable tannins to complex with macromolecules, such as proteins and polysaccharides, and alkaloids. Bate-Smith [9] first speculated that astringent sensations were caused by the increase in friction between the mucosal surfaces, which resulted from a reduction in lubrication in the oral cavity as astringent compounds bound salivary proteins. The binding between polyphenols/salivary proteins forms soluble complexes and/or precipitates that can cause the rupture of the salivary pellicle [10], interact with oral cells [11], and stimulate and activate mechanoreceptors (MRs) hold in the mouth [12]. MRs are nerve endings that function like those of the skin, except that they have smaller receptive fields and lower activation thresholds [13]. They are selectively sensitive to different stimulus properties, such as particle size and/or mouth movements, and project such information to the central nervous system [14]. Besides, the activation of G-coupled proteins also seems to be involved in the perception of astringency, activating signal transduction pattern as that of taste recognition [15]. Some brain regions (hippocampus and anterior cingulate cortex) that have been shown to respond to basic tastes were activated by the intensity and pleasantness of astringency [16]. In particular, the right ventral anterior insula that responded to astringent stimuli contributed to the ability to recognise the qualitative features of astringency. The activation of the trigeminal nerve, chorda tympani, and brain regions involved in memory and emotions could explain astringency as a multi perceptual phenomenon.

Whilst the chemical definition of astringency is related to the ability of tannins to bind proteins, in sensory terms, it is described as different and concomitant feelings of drying, puckering, and roughing [17, 18]. Astringency can be defined as a tactile sensation, because: (i) it is perceived on non-gustatory surfaces such as on the soft palate, gingiva, lips [12], (ii) does not show adaptation but also (iii) increases upon repeated ingestion [19], leading to carry-over effects during the tasting. However, side tastes as bitterness, sourness, and sweetness can highly modulate the overall astringency [20]. The sensitivity of MRs to astringents as well as basic tastes may elucidate the complexity of red wine astringency, which has been described by 33 different subqualities [21]. Amongst these “hard,” “green,” and “rich” have been associated with bitterness, acidity, and high flavour concentration, respectively [22], “harsh,” “abrasive,” and “drying” have been found to define astringency as a negative sensation, whilst the “complex” and “mouth-coat” subqualities have been associated to a positive impact during tasting [21]. These subqualities were also associated with touch standards when utilised to describe the tactile astringent sensations in the mouth elicited by red wines [23, 24]. The qualitative traits of astringency as “soft”, “mouth-coat”, and “rich” represented the drivers of liking for Sangiovese wine [25]. Similarly, for Tannat [26], and Côtes du Rhône and Rioja appellations wines [27], the attribute “mouth-coat” contributed to the quality of the wine.

It is also true that the perception of astringency is mediated by psychological factors [28], but salivary protein composition [29] and tannin’s structure and composition [30, 31] represent the principal factors. In this regard, numerous reviews have been produced during the past years [32, 33, 34, 35, 36, 37, 38].

Advertisement

3. Salivary proteins

Saliva is a biological fluid primarily produced by the three pairs of “major” salivary glands (parotid, submandibular, and sublingual glands) in mouth and by the minor ones by 10% [39]. In the whole, saliva are presently more than 2000 different proteins and peptides [40, 41], which are the result of protein post-translational modifications before being secreted in the mouth [42]. Although saliva is predominantly a watery fluid (99.5%) with a complex mixture of proteins (0.3%; 1–2 mg/mL), ions and other organic compounds (0.2%) are also present. The whole saliva continuously baths the oral cavity and having a pH ranging from 6.2 to 7.4 acts as a buffering system. The saliva is continuously secreted (0.3–7 mL/min) and plays a role in protecting the tooth and mucosal integrity, in antibacterial and antiviral activity, digestion of food, speech, lubrication, taste, and represents a biomarker tool for some diseases [41, 43]. The main families of proteins include enzymes (amylase, carbohydrase, lipase), lactoferrin, high (M1), and low (M2) molecular-weight glycoproteins (mucins), peptides as agglutinins, immunoglobulins, proline-rich proteins (PRPs), cystatins, histatins and statherins [44].

There is evidence that saliva may affect the way we perceive the taste and mouthfeel of foods in various ways [45, 46, 47]. During the wine tasting, saliva transports and dissolves the stimuli substances [48]. Saliva constituents are of great importance for establishing protein-tannin interactions. In particular, the PRPs, histatins, mucin, amylase are the main salivary proteins involved in the binding with polyphenols eliciting astringency [49]. The differences between the binding of the same polyphenol to different proteins result from differences in the amino acid sequences [50].

The PRPs account for approximately 70% of the total secretory protein and are subdivided into acidic, basic, and glycosylated PRPs. They are characterised by an abundance of proline, glutamic acid/glutamine, and glycine [51]. The presence of these four amino acids, especially proline, which are the so-called alpha-helix breakers, enables the protein to form secondary structures, which assumes a random coils conformation in solution [10, 52]. This feature may allow PRPs to universally bind various types of polyphenols, mainly tannins with different sizes and structures. Some species, such as humans, rats, and mice, produce PRPs containing about 40% proline [53, 54]. However, some species produce salivary proteins, which are rich in proline but do not show a high affinity to tannins due to extensive glycosylation [54].

Parotid and submandibular secretions also contain several low molecular-weight histidine-rich peptides [55, 56]. Amongst 12 forms, the histatin 1, 3, and 5 are predominant and vary in size from 7 to 38 residues. These peptides show a high content of basic residues, such as lysine, arginine, and histidine [57]. They tightly bind tannins, even if some peptides are devoid of proline [58]. Conversely, others observed high tannin precipitation by histatins thanks to the interactions formed by basic residues and proline [59].

Amongst the low molecular weight salivary proteins, there is a selectivity in binding polyphenols (as PGG, procyanidin trimer, epicatechin, malvidin-3-glucoside): the acidic PRPs considerably form soluble and insoluble complexes with PGG and trimer but not with epicatechin; basic PRPs and glycosylated PRPs seem to not interact with trimer, whilst basic PRPs show a high affinity for epicatechin, malvidin-3-glucoside, and a mixture of both; the statherin shows no selectivity [60, 61].

Mucins are the major constituents of the viscous layer coating hard and soft tissues in the oral cavity. Mucins are generally composed of a peptide core (apomucin) enriched in serine, threonine, and proline residues and carbohydrate side chains (oligosaccharides) that are linked O-glycosidically to threonine or serine. M1 is a polymeric mucin due to the formation of disulfide linkages between cysteine residues in non-glycosylated domains, whilst M2 is a monomer [62]. Average proline content of 10% seems to be responsible for protein-phenol interactions [63].

Amylase is secreted mainly by the parotid gland in both glycosylated and non-glycosylated isoforms [64]. It is an enzyme capable of hydrolysing bonds within amylose and amylopectin and is composed mainly of amino acids like aspartic acid > glutamic acid > arginine [65]. However, amino acids as tyrosine and tryptophan seem to be crucial for interaction with polyphenols [66]. The non-glycosylated form of amylase contains 22 proline and 16 tryptophan amino acid residues in its sequence that enable the binding with polyphenols [50].

Advertisement

4. Tannins

Astringent wines are commonly defined as “tannic” because tannins are the main polyphenolic compounds involved in the sensation of astringency. Swain and Bate-Smith [67] provided the first useful phytochemical definition of tannin, being “water-soluble phenolic compounds, having molecular weights lying between 500 and 3000, which have the ability to precipitate alkaloids, gelatin, and other proteins”. Tannins can be classified in condensed tannins, phlorotannins, and hydrolysable tannins. Condensed tannins are large macromolecules that consist of two or more monomeric (+)-catechin or (−)-epicatechin units called procyanidins, whilst prodelphinidins consist of (+)-gallocatechin or (−)-epigallocatechin units. In plants, condensed tannins are found as oligomers (2–10 monomer units) or polymers (>10 monomer units). The number of monomer units in a polymer may be as high as 83 units [68]. The subunit composition varies amongst tannins from grape skins, seeds, and stems [69, 70, 71]. The phlorotannins are present in marine brown algae as polymers of phloroglucinol (1,3,5 trihydroxy-benzene) in different ranges of molecular sizes (126 Da–650 kDa). They are analogous to the terrestrial condensed tannins since they do not contain a carbohydrate core [72]. Hydrolysable tannins, structurally perhaps the most complex tannins, comprise three subclasses such as simple gallic acid, poly-galloyl esters of glucose (gallotannins), and esters of ellagic acid (ellagitannins). Derivatives of gallic acid contain one to five galloyl groups that can be esterified to either glucose (e.g., pentagalloyl glucose) or quinic acid (e.g., monogalloyl quinic acid). Gallotannis can contain six or more galloyl groups and can be characterised by having one or more digalloyl groups (e.g., hetpagalloyl glucose). Complex gallotannins have a higher capacity for precipitating proteins than simple galloyl glucoses [73].

Ellagitannins may be divided into six subgroups: hexahydroxydiphenoyl esters, dehydro-hexahydroxydiphenoyl esters and their modifications, nonahydroxytriphenoyl esters (e.g., vescalagin), flavonoellagitannins (e.g., acutissimin A), and oligomers with different degrees of oligomerisation and types of linkages [74].

Tannins are the main responsible for the qualitative aspects of astringency as well for the intensity of the sensation. Grape seed and skin tannins are felt astringent as the mean degree of polymerisation (mDP), and galloylation increased [75]. Their ability to precipitate proteins also increases with mDP up to a given degree of polymerisation [34, 76]. However, monomeric and dimeric flavan-3-ols can induce astringent and bitter sensations [77]. Galloylation of monomers/oligomers and polymers enhances protein precipitation, and its extent depends on the grape variety [78]. The presence of high galloylation seems to be responsible for the coarse perception [75], which in turn can be decreased by a high content of epigallocatechin units on the tannin molecule. On the contrary, it seems that the hydroxylation of B-ring seems to decrease velvety astringency and increase the perception of puckering and drying astringency of wine fractions [79]. Salivary proteins seem to have a higher affinity for condensed tannins than for hydrolysable tannins because of different structural flexibility, size, polarity, affinity constants, and presence of free galloyl groups [80, 81, 82, 83, 84]. Oakwood tannins were mainly associated with smooth and mouth-drying sensations at low concentrations [85]. Astringency subqualities such as mouth-coat, full-body, persistent were mainly associated with oak-derived tannin, whilst the velvet, soft, and satin terms were associated with the exotic wood-derived tannin [25].

4.1 Other stimuli

Compounds able to elicit sensations as tastes and mouth feelings are called stimuli. Chemically diverse astringents such as complex salts such as aluminium sulfate (alum), acids, and other phenolics, have also been shown to evoke astringency [17, 86]. Five organic acids and one inorganic elicited astringency and astringent subqualities [87], and dryness has also been reported [86, 88]. The addition of malic and lactic acid in red wine at the same pH did not differ significantly in astringency despite the difference in titratable acidity [89]. However, these acids were defined astringent in addition to their sour taste [90]. Wines more abundant in malic acid showed higher reactivity towards saliva proteins and then higher potential astringency than tartaric acid-rich wines at the same pH, probably due to different buffer capacities [91]. The astringency of acids is attributed either to the direct contribution of H+ ions or to the hydrogen bonding capabilities of the hydroxyl groups on the anion or un-dissociated acid [17]. Denaturation of proteins in the saliva could also affect the binding and dissociation of phenolic compounds and their precipitation. The intensity of astringency linearly increases as a function of pH reduction [19], implying significant precipitation of salivary proteins [92].

Anthocyanins, composed of a sugar bound to the anthocyanidin moiety (cyanidin, peonidin, delphinidin, petunidin, and malvidin), impart colour to the grapes and red wine and can be modified by different enological practices [93]. Controversial is the studies of the influence of anthocyanins on astringency. An anthocyanin fraction added in model wine solution was felt as “rough and chalk,” and slightly contributed to the overall astringency probably for contamination of the fraction with unknown phenolic compounds [94]. Successively, the isolated fractions of anthocyanidin–glucosides and anthocyanin coumarates did not influence astringency of wine solutions either the “coarse,” “chalk,” or “dry” astringent subqualities [95]. However, anthocyanins were able to interact with human salivary proteins forming soluble aggregates [96], and even precipitates, being the cinnamoylated the most reactive fraction (precipitation between 6.5 and 17.5%), also influencing the astringency perception [97]. Pyranoanthocyanins, anthocyanin-derived pigments that can form during red wine ageing, seems to be involved in astringency, since they are able to interact with salivary proteins by phenol, catechol or even flavanols structures, similarly to procyanidins [98].

Flavonols (kaempferol, quercetin, and myricetin) are present in grapes and wine as glycosides (sugar attached). In the plant, they act as a natural sunscreen in the skin of grape berries. In wine, they can be hydrolysed and act as cofactors for colour enhancement. Flavonol glycosides, such as 3-O-glucosides and 3-O-galactosides of quercetin, syringetin, and isorhamnetin, have been reported to be astringent at low detection threshold levels and characterised by a velvety astringency [99]. The addition of quercetin 3-O-glucoside (2 g/L) to wine increased astringency, leading to the formation of complexes with saliva at 200 μM [100]. However, such concentrations are not naturally present in red wine, in which quercetin 3-O-glucoside can range from 2 up to 34 mg/L, depending on the cultivar [101].

Many sensory active non-volatile compounds comprising hydroxybenzoic acids, hydroxycinnamic acids, flavon-3-ol glycosides, and dihydroflavon-3-ol rhamnosides were identified as the key inducers of the astringent mouthfeel of red wines using a molecular sensory approach [99]. The phenolic acids in wines, especially hydroxycinnamic and benzoic acid derivatives, have been reported to be more puckering astringent. These compounds have also been correlated with astringency in free-run and pressed wine [102]. The trans-p-coumaric, cis-aconitic, and trans-caftaric acids seem to participate in the astringency of Spanish wines [103].

Advertisement

5. Polyphenol-protein interactions

Given that the carbonyl function of salivary proteins is a very effective hydrogen bond acceptor [104], it would appear that it would play a significant role in bonding to polyphenols hydroxyls [10, 105]. Nowadays, the interaction between proteins and proanthocyanidins is widely recognised to be a combination of hydrogen bonding and hydrophobic effects in the acidic wine matrix. However, covalent bonding is also possible between proteins and polyphenols during oxidation [106] and nucleophilic addition processes [107]. In this chapter, we focused on the non-covalent binding involved in the astringent sensation.

Physico-chemical quantities (binding constants, stoichiometry, and atomic structure of complexes, driving forces for the association) have been utilised to understand the multifaceted sensation of astringency. Many techniques including circular dichroism (CD) [108], isothermal titration microcalorimetry (ITC) [109], fluorescence spectroscopy [50], dynamic light scattering (DLS) [110], and nuclear magnetic resonance (NMR) [111] have been employed to understand the formation mechanism of protein/polyphenol aggregates in solution. Generally, these studies focused on interactions between protein segment from human saliva PRPs proteins family and selected procyanidins, because it represents the easiest way to simulate such a complex phenomenon. They can reveal the hydrophobic interactions formed between the phenolic rings of the procyanidins and proline residues, and the hydrogen bonding between the hydroxyl groups on the phenolic B-ring and hydrogen acceptor sites of the peptide bond [52, 112]. The aggregation of procyanidin with peptide seems to be firstly mediated by hydrophobic forces, and then hydrogen bonding has been postulated to provide directional and robust bonding that stabilises the complex. The peptide is coated by polyphenols, which provides a crosslink between two or more peptides up to a critical point, after which precipitation begins. The stability of these complexes depends on the tannin dimension and number of free phenolic groups, as well as the nature of the protein involved [81, 109].

The driving factors that determine the binding between tannins and salivary proteins were identified to be the critical micelle concentration value (CMC), tannin structure preferences, and tannin colloidal state [113]. Below the values observed in wine (from 1.5 to 2.9 mM), procyanidins specifically interacted with peptide through hydrophilic recognition. A network of interactions can be formed depending on tannin conformation, and precipitation of the complex can occur, or if an intramolecular staking Π-Π of phenolic groups is preferred, the precipitation is not observed. Above these values, tannins spontaneously tend to form aggregates that, at first through specific interactions bind proteins, and then surrounded by the hydrophobic residues, stabilise the complex by hydrophobic bonding. To summarise, both hydrophilic and hydrophobic interactions contribute to form a complex network, which determines the precipitation of salivary proteins with tannins.

Advertisement

6. Assessments of astringency

A method for measuring astringency remains one of the great analytical challenges in wine chemistry and oenology. The interest in investigating the mechanisms and interactions between polyphenols and proteins can allow us to find the optimal way to simulate and evaluate what happens during the red wine tasting. Quite often, sophisticated techniques rely on the purification of both tannin and protein fractions, the extrusion from the wine content, and the omission of matrix components during reactions, and all contribute to send away astringency from the reality that is: wine polyphenols interacting with salivary proteins in mouth, causing drying sensations.

Several procedures have been carried out during the last decades for measuring tannins. Additionally, analyses of soluble (turbidimetric analysis) and insoluble (precipitation protein assays) protein-polyphenols complex have been developed for assessing astringency. The sensory analysis represents the human response as an analytical tool to evaluate wine perception. Many training and tasting sections are necessary over a long period involving a high number of tasters to form a reliable panel. In the case of astringency, it is complicated to discern amongst tastes and brings on fatigue. A method capable of estimating tannin palatability has to be the most objective as possible and must correlate with sensory data in order to reflect the real phenomenon of wine tasting.

6.1 Stimuli analysis: pros and contras

Amongst stimuli able to elicit astringency, tannins are the main compounds responsible for this sensation. Tannins are intrinsically amphiphilic molecules with high reactivity, have a diverse range of structures, and are often found in matrices with other phenolic molecules containing similar functional groups. Besides using sophisticated equipment and analytical techniques, there is also a great interest in a relatively simple method.

In the past, many colourimetric techniques were developed to analyse phenolics compounds spectrophotometrically. The first one used the Folin-Denis reagent [114], which was successively modified [115, 116], and lastly into the Folin-Ciocalteau assay [117]. However, they were not specific for tannins but detected any phenolic compound. More specific colour reactions were used to measure condensed tannins and their precursors. Depolymerisation in HCl and n-butanol of proanthocyanidins yield anthocyanidins that can be quantified spectrophotometrically [118, 119]. Others used vanillin reagent for flavanols [6, 120], or p-dimethylaminocinnamaldehyde for a more specificity and colour stability [121, 122]. Only the flavonoid-based condensed tannins can be detected with these reagents. As tannins can inhibit the catalytic activity of enzymes [6], many methods used the interaction with proteins in solution to measure the inhibition of different enzymes spectrophotometrically [123].

Other methods, based on the acid-catalysed condensation reactions with benzyl mercaptan (thiolysis) and phloroglucinol (phloroglucinolysis), can determine both the chain length (mDP) and composition by HPLC [124, 125]. Most of our current knowledge about the general composition and structure of grape and wine tannins have been obtained by depolymerisation [126]. Poor yields due to reaction product instability, reactions with non-proanthocyanidin compounds, and side reactions also contribute negatively to the utility of thiolytic methods [124]. The problem with phloroglucinolysis, on the other hand, is that it produces low yields, and only a fraction of the tannin is converted to known flavan-3-ol products [127]. Normal-phase HPLC (NP-HPLC) method has also been developed to quantify the proanthocyanidins into low and high molecular-weight polymers [128]. A simple method based on Fourier transform mid-infrared (FT-MIR) spectroscopy combined with multivariate data analysis, was successfully used to measure the tannin concentration of 86 red wines, previously purified by solid-phase extraction (SPE) [129].

6.2 Precipitation assays: pros and contras

Protein precipitation assays are of particular interest because the interaction of proteins with tannins can be used to model astringency perception [130]. The ability of gelatin to precipitate phenols, including tannins, has been observed since 1934 [131]. The same phenomenon was observed when hide powder or polyvinylpyrrolidone were used in high concentrations [132]. Bate-Smith [130] noted that protein of skin differed from proteins of saliva, which caused the “puckery” sensation induced by tannin. For measuring the relative astringency of tannins, a spectrophotometric technique based on the precipitation of the haemoglobin with tannin was then introduced [130]. Similarly, another spectrophotometric technique measured the inhibition of β-glucosidase after the precipitation with tannic acid and condensed tannins [133]. Alternatively, Hagerman and Butler [134] used bovine serum albumin (BSA) as a precipitant agent, which was successively taken by Harbertson et al. [135] for wine analysis. Glories [136] proposed the gelatin index, in which tannins were precipitated by gelatin protein. This procedure required the measure of proanthocyanidin concentration before and after precipitation with an excess of gelatin. Besides, gelatin is a heterogeneous mixture of proteins, and its composition may change amongst the different commercial products, leading to a source of variability and imprecision of data. For this, some researchers replaced gelatin with ovalbumin [137]. Another tannin assay used the methylcellulose to precipitate tannin (MCP) [138, 139]. The MCP tannin assay is based on the formation of an insoluble polymer-tannin complex, which can be separated by centrifugation. The total phenolic content (absorbance at 280 nm) is measured in control and treated samples. However, if the assays utilise synthetic agent or protein different from saliva, the binding reaction seems not to reproduce the physiological conditions during the wine tasting, because the binding affinity of the protein is not comparable to that of salivary protein. In the case of bovine serum albumin, it has been shown that the salivary protein has a higher affinity for tannin than BSA. In fact, in the presence of an excess of BSA, the tannin preferentially bound the salivary protein. Other proteins, including dietary proteins, may not complex any tannin in the presence of the salivary tannin-binding protein [8]. The use of salivary proteins has been proposed to represent the model system for astringency better. In precipitation assays, fractionated [8, 140] or whole [141, 142] human saliva has been used. Mixing whole saliva and grape polyphenols give rise to a “soft cloudy” precipitate, which gathered after centrifugation on the bottom of the tube so that the supernatant was easily recovered without disturbing this pellet. The binding reaction was performed at 25°C, and the complex formed was successively precipitated by centrifugation at 4°C in order to stop further reactions. The induced precipitation allowed to separate the proteins bound to polyphenols from whose remained in the solution that not reacted with them. Both the nature of condensed tannin [141] and salivary proteins [142] involved in the precipitation were analysed. In both works, the sodium dodecyl sulphate-polyacrylamide gel electrophoresis (SDS-PAGE) of human saliva was carried out, Sarni-Manchado et al. [141], analysed the tannins in the supernatant and pellet. In contrast, Gambuti et al. [142], analysing the supernatant, revealed the proteins mainly reactive with polyphenols by comparison with the control saliva. Evidence of the qualitative and quantitative changes in salivary protein profile after tasting tannin solutions and wines was also made by HPLC [143]. Interactions and precipitation of low molecular weight salivary proteins with procyanidins confirmed the involvement of different families of salivary proteins in the development of astringency [144]. The use of salivary proteins involves the collection of human saliva from different healthy volunteers according to a specific protocol, and it must take into account the salivary flow to limit the effect of individual differences in astringency perception due to subjects’ saliva characteristics [145].

6.3 Nephelometry: pros and contras

Nephelometry is a method that allows a direct estimation of the amount of protein/tannin complexes by measuring the scattered light in the solution that results from the gradual formation of a cloudy precipitate corresponding to the soluble aggregate. Chapon [146] proposed this technique by studying the interactions between beer polyphenols and proteins involved in the colloidal instability of beer. Similarly, the haze formed between salivary proteins and polyphenols represents the first step in the development of astringency and can be measured with a turbidimeter [147, 148]. A continuous flow method was also used to study the interactions between grape extracts and wine with BSA at different concentrations [149]. Globular proteins and PRPs were used to measure a relative tannin specific activity of procyanidin oligomers from grape seeds [30], and PRPs showed the strongest affinity. Human salivary proteins have been considered as the most suitable model proteins. For this reason, in turbidity measurement, whole human saliva [148] and mucin, a high molecular weight salivary protein [150], were used as model proteins for astringency assessment. Based on polyphenol/mucin reactivity, a micro-plate assay was also developed [151]. Tannic acid [150], grape seed extracts [151], wine extracts [63], tannin fractions added to model solutions [152] were analysed by nephelometry. The turbidity of the solution, formed by the tannin-protein aggregates, linearly correlated with astringency. However, no direct analysis of wines was carried out. Lastly, instead, wine samples were analysed trough nanotechnology such as localised surface plasmon resonance (LSPR) combined with surface imprinted polymers, as a measure of the interactions of polyphenol with salivary protein and then astringency [153].

6.4 Sensory analysis: pros and contras

The sensory analysis represents the human response to wine tasting. A sensory panel can provide information about the sensory properties of a product, but significant training is required before the panel becomes a reliable sensory instrument. Astringency is a difficult sensory attribute to evaluate, owing to particular characteristics of the sensation. Generally, it is evaluated by tasting but can suffer from individual subjectivity. The feeling can take over 15 seconds to develop fully and is known to build in intensity and become increasingly difficult to clear from the mouth over repeated exposures [19, 154]. Carry-over effects can occur. When wines or tannic solutions are evaluated by a well-trained panel using established sensory methodologies, the panel leader can expect to obtain reliable information about the intensity in the perceived astringency of the samples. Screening, selection, training, and panel maintenance are exercises that help the panel attain proficiency before sample evaluation. Classical methodologies widely applied are descriptive and rating sensory analyses. The first helps to distinguish between samples by a qualitative description of their sensory properties [75] and the second permits to scale samples according to the intensity of the perception. However, time-intensity (TI) is a temporal methodology widely used. This method consists of recording one by one the intensity evolution of given attributes [155]. However, TI showed some limitations because it is time-consuming due to the evaluation of only a few attributes at the same time [156]. Furthermore, carry-over effects can overcome when assessing the temporal perception of an attribute [157]. To overcome these drawbacks, Pineau et al. [156] developed a new method called temporal dominance of sensations (TDS), which consists of identifying and rating sensations perceived as dominant until the perception ends. Before the development of this method, a similar experimental approach was successfully used to describe the temporality of sensations in wines [158]. Astringency, a dynamic sensation, takes many seconds to develop after the basic tastes, and the duration depends on the wine. Notwithstanding, TDS can be difficult when panellists had select the dominant attribute and score its intensity, but proper training can overcome this problem [159].

It is also essential to discuss and familiarise with the terms associated with astringency. A vocabulary of 33 terms has been proposed by a combined panel of experienced tasters and winemakers to describe the mouthfeel characteristics of red wines [160]. The check-all-that-apply (CATA) question that consists of a list of subqualities from which the panellists have to select all the options they consider appropriate to that wine has been utilised for the characterisation of the astringency subqualities of Tannat wine [161]. Recently, a sensory method that combines CATA approach and training in astringency subqualities with touch-standards resulted very useful for investigating the astringency characteristics of red wines [24, 25, 162]. In any case, intense training is necessary to distinguish astringency from other tastes, especially bitterness, and to reveal the different qualitative attributes. Fatigue and loss of stimuli memory may occur, particularly with panellists who are unfamiliar with astringency, and when too many samples are presented. Training is also expensive and time-consuming. However, it is necessary to investigate the astringency subqualities of red wines. Sensory analysis is of fundamental importance, but in some cases, it is not possible to perform, so the replacement with an analytical instrument able to measure astringency could help in research as well as in the winery.

6.5 Correlation between sensory and analytical analysis

Because astringency is one of the main attributes for wine quality, winemakers are interested in an analytical and objective method to evaluate it. No method can substitute entirely sensory analysis, but a method that results in a reproducible index has to correlate quite well with it. A statistically significant correlation between the sensorial and analytical methods is necessary.

The gelatin index has represented the almost widely analytical method for estimating astringency in red wine [136]. Besides, it furnished only approximate results [137]. Successively, a positive correlation (R2 = 0.56) between the gelatin index and time-intensity data was obtained only at a low concentration of polyphenols utilising 29 wines judged by 10 panellists [163]. A method that used the ovalbumin in alternatively to gelatin as a precipitation agent was proposed to determine astringency [137]. Ten wines were tested by 10 expert enologists evaluating the astringency on a scale from 1 to 100. The method resulted in more reproducible than the gelatin index and was positively correlated (R2 = 0.77) with sensory analysis. This method was also used to assess the astringency of Greek wines, and a good correlation was found (R2 = 0.93) [164]. Another predictive model for astringency estimation was based on phenolic compounds and colour analysis of 34 wines by 12 judges on a 9-point intensity scale [165]. Multiple regression generated three possible models to predict astringency from analytical data, the most simple depended on total phenolics and co-pigmented anthocyanins, besides the predicted astringency plotted versus observed astringency resulted in low but acceptable correlation from a sensory perspective.

Monteleone et al. [150] proposed a predictive model by measuring the polyphenol-mucin reactivity in which the capability of polyphenolic extracts to induce astringency was estimated on their ability to develop turbidity in the in vitro assay. They found a linear relation between astringency perceived by 30 trained judges and the mucin index for tannic acid model solutions (R2 = 0.993) grape seed extracts (R2 = 0.996), and phenolic extracts (R2 = 0.95) [63].

In a study by Kennedy et al. [166], 40 red wines were evaluated by a panel consisting of three winemakers and two enologists for the astringency intensity scored from zero to 10. The aim was to correlate astringency and tannin concentration measured by different analytical methods: absorption at 280 nm, phloroglucinolysis, gel chromatography, and BSA protein precipitation. The analytical method having the strongest correlations with perceived astringency was the protein precipitation one (R2 = 0.82). Protein precipitation represents the method the most similar to the physiological response to astringent stimuli and can be used as an in vitro tool for understanding how tannin can modulate astringency perception. Generally, it was assumed that the most suitable proteins for evaluating astringency are the salivary PRPs. However, other proteins in whole human saliva were preferentially precipitated by increasing tannin solutions [142]. Successively, the percentage decrease of two salivary proteins after the precipitation with tannins, measured by electrophoresis, represented an indicator of the reactivity of tannin. The saliva precipitation index (SPI) was well correlated with the sensory evaluation of the astringency of 57 red wines (R2 = 0.97) made by 18 trained assessors [167].

Advertisement

7. The saliva precipitation index (SPI)

The SPI represents a useful tool to assess the physiological response to astringents, measuring the astringency of red wine indirectly. This index evaluated the precipitation of salivary proteins occurring during the tasting of an astringent stimulus. The SPI, analysing the salivary protein pattern by SDS-PAGE electrophoresis, has been improved considering the in-mouth temperature (37°C) for the binding reaction, the choice of resting saliva, and the ratio saliva:wine. The excess of saliva with respect to wine (2:1) in a static environment permits to measure the binding capacity of tannins better [167]. Successively, to reduce the time and solvents, the chip electrophoresis replaced the SDS-PAGE, providing similar results [168]. In the last years, the SPI has been used for different technological practices proving useful information for winemakers and enologists to manage the style and quality of red wines.

7.1 Applications of SPI in winemaking

7.1.1 Enological practices

In winemaking, the clarification process is fundamental to stabilise and clarify the wine by adding exogenous proteins into wine [169]. Proteins used for fining interact with wine tannins by a mechanism similar to that occurring during the tasting. The interaction protein-tannin, binding, and precipitation determine a decrease in polyphenolic compounds responsible for the sensation of astringency [170]. The SPI was used to evaluate the efficacy of the fining of different proteins at different concentrations in Aglianico [171], and Sangiovese wines [172]. In Aglianico, the gelatin (animal protein) and patatin (plant protein) showed similar efficacy in diminishing wine polyphenols reactive towards salivary proteins, and then astringency, whilst in Sangiovese it depended on the polyphenolic content of the wine. The information provided by SPI was useful to understand that each wine, with peculiar polyphenolic composition, should be treated maintaining the ratio anthocyanins and tannins such as to assure a modulation of astringency and at the same time a correct evolution of the colour during ageing.

A common practice is the utilisation of enological tannins as a substitute for oak barrels to improve colour stability and taste and is authorised by the International Organisation of the Vine and Wine (OIV) for musts and wines clarification [173]. Commercial preparations of tannins of different origins showed different abilities in precipitating salivary proteins: condensed tannins resulted in higher SPI and astringency than hydrolysable tannins. The addition of tannins in wines modify the astringency or not depending on the wine phenolic content. The SPI was useful to understand the effect of tannins addition on wine astringency in order not to compromise overall wine quality [83]. Similarly, after a moderate oxidation (21 mg/L of oxygen equivalent), the addition of 2 g/L of enological tannins did not result in an increase in the reactivity of wine tannins towards salivary proteins after 30 days of treatment. This effect was also shown in the oxidation process in the presence of acetaldehyde [174]. The SPI seems to be sensitive to reaction-products such as polymers of flavanols and anthocyanins formed directly or via a molecular bridge (e.g., acetaldehyde) [31, 175], and new-formed proanthocyanidins [93, 176]. This may explain why during the oxidation of red wines, the SPI followed a different trend from BSA reactive tannins [174, 177, 178].

7.1.2 Ageing

The decrease of astringency with time has been shown to depend on the reduced concentration of tannins due to precipitation [31, 68], but the trend is not strictly related to the age of wine [179]. The astringency of red wine decreases during ageing because of the changes in the structure of tannins due to cleavage reactions generating low molecular weight species [31], polymerisation without the participation of anthocyanins and subsequent precipitation [95], direct or indirect condensation with anthocyanins [180], and the formation of flavan-3-ol sulfonates by SO2 [181]. Wine becomes soft and mellow for the decline of tannin mean degree of polymerisation [182], velvet and mouth-coating for the formation of the polymeric pigments [24], or satin for lower content of flavans and astringent tannins (measured by SPI), and higher formation of polymers [183] after ageing. Studies on Sangiovese wine revealed that the astringency profile changed from an unripe, dry astringency towards rich, full-body, and mouth-coating sensations after about 2 years of ageing [184]. However, pucker sensations can appear if the oxidation is excessive [24, 25]. Astringency subqualities have been able to discriminate wines of different denominations with a chemical age of 3–5 years, more than other wine parameters [25]. Red wine benefits of a moderate oxygenation during ageing favouring changes in tannin structures that, affecting their reactivity towards proteins, can modulate wine astringency. The SPI was utilised to objectively evaluate changes in astringency as a function of oxygen uptake before and after bottling [185]. Although conflicting results were reported for astringency after micro-oxygenation of wines, a significant variation of wine reactivity towards salivary proteins and, then, in wine astringency was observed after 42 months of ageing in bottle only in low pH wines. Moreover, oxygen permeating towards closures determined changes in wine phenolics detectable only using SPI. It was significantly lower when the bottles were sealed with closures at high oxygen transfer rate (OTR). Such differences were not perceived by sensory analysis, demonstrating that SPI can be more sensitive in revealing slight differences in the reactivity of tannins. Lastly, the effect of ageing on the precipitation of salivary proteins is a function of ageing time, wine pH and phenolic composition, and oxygen level in red wine. The decisive role of pH on wine astringency has been confirmed in a recent work of Forino et al. [92], in which the SPI was used to measure wines with different pH levels (3.7–3.2) obtained by adding strong acids or bases, which made the wine unsafe to taste. The binding and precipitation of wine tannins with saliva proteins was favoured at low pH values, and this effect was dominant with respect to the tannins content. Previously, the tartaric acid addition in wine, modifying the pH, resulted in high SPI [186], due to the increase of tannins in the phenolate form, and therefore to an increase of hydrogen bonding with salivary proteins. It is also likely that at low pH increases the accessibility of the binding sites leading to enhanced Van der Waal interactions and hydrogen bonding between proteins and polyphenols [187]. However, other parameters, such as ethanol, fructose, and mannoproteins have been shown to influence astringency and SPI [186]. The effect of mannoproteins on the inhibition of salivary protein precipitation was also showed in Aglianico and Sangiovese wines after 12 months of ageing. The sensory analysis confirmed a reduction in wine astringency. Some mannoproteins interact with tannins forming higher molecular weight structures that prevent the binding with salivary proteins, and thus are not able to elicit astringency [94]. Mannoproteins can also act as steric stabilisers limiting the binding with tannins [112]. Wine polysaccharides inhibit tannin-salivary proteins interaction by a mechanism that involves the formation of protein-tannin complex firstly, probably ruled by hydrophobic interactions and stabilised by hydrogen bonds, and then the polysaccharides can act by a ternary mechanism through the encapsulation of this complex, increasing its solubility. However, the efficiency depends on the polarity of both salivary proteins and tannins [188]. Beyond the molecular mechanism, mannoproteins can highly influence the qualitative sensory perception of astringency, conferring positive subqualities of astringency to red wines [162].

Advertisement

8. Conclusions

Astringency is still a complex phenomenon, and despite the many efforts from researchers, it is not fully understood. However, the different in vitro assessments have been shown to be useful in evaluating the wine astringency. They could replace the sensory evaluation when there is no possibility of tasting wines: for low sample availability, when tasting is not permitted (as in the pandemic period due to Covid-19) or unsafe, or when too many samples must be tasted. An analytical method for astringency may be potentially useful not only in research purposes but also in the optimisation of the winemaking process and may help wine producers to improve wine quality.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Nomenclature

BSAbovine serum albumin
CATAcheck-all-that-apply
CDcircular dichroism
CMCcritical micelle concentration
DLSdynamic light scattering
FT-MIRFourier transform mid-infrared
OIVInternational Organisation of the Vine and Wine
ITCisothermal titration microcalorimetry
LSPRlocalised surface plasmon resonance
M1high molecular-weight mucin
M2low molecular-weight mucin
mDPmean degree of polymerisation
MRsmechanoreceptors
MCPmethylcellulose precipitable-tannin
NP-HPLCnormal-phase HPLC
NMRnuclear magnetic resonance
OTRoxygen transfer rate
PRPsproline-rich proteins
SPIsaliva precipitation index
SDS-PAGEsodium dodecyl sulphate-polyacrylamide gel electrophoresis
SPEsolid-phase extraction
TDStemporal dominance of sensations
TItime-intensity

References

  1. 1. Zucker WV. Tannins: Does structure determine function? An ecological perspective. The American Naturalist. 1983;121(3):335-365. DOI: 10.1086/284065
  2. 2. Stern JL, Hagerman AE, Steinberg PD, Mason PK. Phlorotannin-protein interactions. Journal of Chemical Ecology. 1996;22(10):1877-1899. DOI: 10.1007/BF02028510
  3. 3. Aerts RJ, Barry TN, McNabb WC. Polyphenols and agriculture: Beneficial effects of proanthocyanidins in forages. Agriculture, Ecosystems and Environment. 1999;75(1-2):1-12. DOI: 10.1016/S0167-8809(99)00062-6
  4. 4. Bernays EA, Driver GC, Bilgener M. Herbivores and plant tannins. In: Advances in Ecological Research. Vol. 19. London, United Kingdom: Academic Press; 1989. pp. 263-302. DOI: 10.1016/S0065-2504(08)60160-9
  5. 5. Mehansho H, Butler LG, Carlson DM. Dietary tannins and salivary proline-rich proteins: Interactions, induction and defense mechanisms. Annual Review of Nutrition. 1987;7:423-440
  6. 6. Goldstein JL, Swain T. Changes in tannins in ripening fruits. Phytochemistry. 1963;2(4):371-383. DOI: 10.1016/S0031-9422(00)84860-8
  7. 7. Harlan JW, Feairheller SH. In: Friedman M, editor. Protein Crosslinking. Advances in Experimental Medicine and Biology. Chemistry of the Crosslinking of Collagen during Tanning. Boston, MA: Springer; 1977. DOI: 10.1007/978-1-4684-3282-4_27
  8. 8. Austin PJ, Suchar LA, Robbins CT, Hagerman AE. Tannin-binding proteins in saliva of deer and their absence in saliva of sheep and cattle. Journal of Chemical Ecology. 1989;15:1335-1347. DOI: 10.1007/BF01014834
  9. 9. Bate-Smith EC. Leuco-anthocyanins 1. Detection and identification of anthocyanidins formed from leuco-anthocyanins in plant tissues. The Biochemical Journal. 1954;58(1):122
  10. 10. Hagerman AE, Butler LG. The specificity of proanthocyanidin-protein interactions. The Journal of Biological Chemistry. 1981;256:4494-4497
  11. 11. Soares S, Brandão E, Guerreiro C, Mateus N, de Freitas V, Soares S. Development of a new cell-based oral model to study the interaction of oral constituents with food polyphenols. Journal of Agricultural and Food Chemistry. 2019;67(46):12833-12843. DOI: 10.1021/acs.jafc.9b05575
  12. 12. Breslin P, Gilmore M, Beauchamp G, Green B. Psychophysical evidence that oral astringency is a tactile sensation. Chemical Senses. 1993;18(4):405-417. DOI: 10.1093/chemse/18.4.405
  13. 13. Trulsson M, Essick GK. Low-threshold mechanoreceptive afferents in the human lingual nerve. Journal of Neurophysiology. 1997;77(2):737-748. DOI: 10.1152/jn.1997.77.2.737
  14. 14. Chen J, Engelen L. Food Oral Processing: Fundamentals of Eating and Sensory Perception. Oxford, United Kingdom: Wiley-Blackwell; 2012
  15. 15. Schöbel N, Radtke D, Kyereme J, Wollmann N, Cichy A, Obst K, et al. Astringency is a trigeminal sensation that involves the activation of G protein-coupled signaling by phenolic compounds. Chemical Senses. 2014;39(6):471-487. DOI: 10.1093/chemse/bju014
  16. 16. Kishi M, Sadachi H, Nakamura J, Tonoike M. Functional magnetic resonance imaging investigation of brain regions associated with astringency. Neuroscience Research. 2017;122:9-16. DOI: 10.1016/j.neures.2017.03.009
  17. 17. Lawless HT, Corrigan CJ, Lee CB. Interactions of astringent substances. Chemical Senses. 1994;19:141-154. DOI: 10.1093/chemse/19.2.141
  18. 18. ASTM. Standard definitions of terms relating to sensory evaluation of materials and products. In: Kuznicki JT, Rutkiewic AF, Johnson RA, editors. Annual Book of ASTM Standards. Philadelphia, PA: American Society for Testing and Materials; 2004. pp. 5-15
  19. 19. Guinard JX, Pangborn RM, Lewis MJ. The time-course of astringency in wine upon repeated ingestion. American Journal of Enology and Viticulture. 1986;37(3):184-189
  20. 20. Fleming EE, Ziegler GR, Hayes JE. Salivary protein levels as a predictor of perceived astringency in model systems and solid foods. Physiology & Behavior. 2016;163:56-63. DOI: 10.1016/j.physbeh.2016.04.043
  21. 21. Gawel R, Iland P, Francis I. Characterising the astringency of red wine: A case study. Food Quality and Preference. 2001;12(1):83-94. DOI: 10.1016/S0950-3293(00)00033-1
  22. 22. King MC, Cliff MA, Hall J. Effectiveness of the ‘Mouth-feel Wheel’ for the evaluation of astringent subqualities in British Columbia red wines. Journal of Wine Research. 2003;14(2-3):67-78. DOI: 10.1080/09571260410001677932
  23. 23. De Miglio P, Pickering GJ. The influence of ethanol and pH on the taste and mouthfeel sensations elicited by red wine. Journal of Food, Agriculture and Environment. 2008;6:143-150
  24. 24. Rinaldi A, Moio L. Effect of enological tannin addition on astringency subqualities and phenolic content of red wines. Journal of Sensory Studies. 2018;33(3):e12325. DOI: 10.1111/joss.12325
  25. 25. Rinaldi A, Moine V, Moio L. Astringency subqualities and sensory perception of Tuscan Sangiovese wines. OENO One. 2020;54(1):75-85. DOI: 10.20870/oeno-one.2020.54.1.2523
  26. 26. Vidal L, Antúnez L, Rodríguez-Haralambides A, Giménez A, Medina K, Boido E, et al. Relationship between astringency and phenolic composition of commercial Uruguayan Tannat wines: Application of boosted regression trees. Foodservice Research International. 2018;112:25-37. DOI: 10.1016/j.foodres.2018.06.024
  27. 27. Sáenz-Navajas M, Ballester J, Pêcher C, Peyron D, Valentin D. Sensory drivers of intrinsic quality of red wines. Foodservice Research International. 2013;54(2):1506-1518. DOI: 10.1016/j.foodres.2013.09.048
  28. 28. Martens M. A philosophy for sensory science. Food Quality and Preference. 1999;10(4-5):233-244. DOI: 10.1016/S0950-3293(99)00024-5
  29. 29. Lu Y, Bennick A. Interaction of tannin with human salivary proline-rich proteins. Archives of Oral Biology. 1998;43(9):717-728. DOI: 10.1016/S0003-9969(98)00040-5
  30. 30. De Freitas V, Mateus N. Structural features of procyanidin interactions with salivary proteins. Journal of Agricultural and Food Chemistry. 2001;49(2):940-945. DOI: 10.1021/jf000981z
  31. 31. Cheynier V, Dueñas-Paton M, Salas E, Maury C, Souquet JM, Sarni-Manchado P, et al. Structure and properties of wine pigments and tannins. American Journal of Enology and Viticulture. 2006;57(3):298-305
  32. 32. Gawel R. Red wine astringency: A review. Australian Journal of Grape and Wine Research. 1998;4(2):74-95. DOI: 10.1111/j.1755-0238.1998.tb00137.x
  33. 33. de Freitas V, Mateus N. Protein/polyphenol interactions: Past and present contributions. Mechanisms of astringency perception. Current Organic Chemistry. 2012;16(6):724-746. DOI: 10.2174/138527212799958002
  34. 34. Bajec MR, Pickering GJ. Astringency: Mechanisms and perception. Critical Reviews in Food Science and Nutrition. 2008;48:858-875. DOI: 10.1080/10408390701724223
  35. 35. McRae JM, Kennedy JA. Wine and grape tannin interactions with salivary proteins and their impact on astringency: A review of current research. Molecules. 2011;16(3):2348-2364. DOI: 10.3390/molecules16032348
  36. 36. Ma W, Guo A, Zhang Y, Wang H, Liu Y, Li H. A review on astringency and bitterness perception of tannins in wine. Trends in Food Science and Technology. 2014;40(1):6-19. DOI: 10.1016/j.tifs.2014.08.001
  37. 37. Soares S, Brandão E, Mateus N, de Freitas V. Sensorial properties of red wine polyphenols: Astringency and bitterness. Critical Reviews in Food Science and Nutrition. 2017;57(5):937-948. DOI: 10.1080/10408398.2014.946468
  38. 38. Li SY, Duan CQ . Astringency, bitterness and colour changes in dry red wines before and during oak barrel aging: An updated phenolic perspective review. Critical Reviews in Food Science and Nutrition. 2019;59(12):1840-1867. DOI: 10.1080/10408398.2018.1431762
  39. 39. Humphrey SP, Williamson RT. A review of saliva: Normal composition, flow, and function. Journal of Prosthetic Dentistry. 2001;85(2):162-169. DOI: 10.1067/mpr.2001.113778
  40. 40. Scarano E, Fiorita A, Picciotti PM, Passali GC, Calò L, Cabras T, et al. Proteomics of saliva: Personal experience. ACTA Otorhinolaryngologica Italica. 2010;30(3):125-130
  41. 41. Bandhakavi S, Stone MD, Onsongo G, Van Riper SK, Griffin TJ. A dynamic range compression and three-dimensional peptide fractionation analysis platform expands proteome coverage and the diagnostic potential of whole saliva. Journal of Proteome Research. 2009;8(12):5590-5600. DOI: 10.1021/pr900675w
  42. 42. Dawes C. Salivary flow patterns and the health of hard and soft oral tissues. Journal of the American Dental Association. 2008;139:18S-24S. DOI: 10.14219/jada.archive.2008.0351
  43. 43. Schipper RG, Silletti E, Vingerhoeds MH. Saliva as research material: Biochemical, physicochemical and practical aspects. Archives of Oral Biology. 2007;52(12):1114-1135. DOI: 10.1016/j.archoralbio.2007.06.009
  44. 44. Helmerhorst EJ, Oppenheim FG. Saliva: A dynamic proteome. Journal of Dental Research. 2007;86(8):680-693. DOI: 10.1177/154405910708600802
  45. 45. Christensen CM. Role of saliva in human taste perception. In: Meiselman HL, Rivlin RS, editors. Clinical Measurement of Taste and Smell. New York: Macmillan Publishing Co; 1985. pp. 414-428
  46. 46. Fischer U, Boulton RB, Noble AC. Physiological factors contributing to the variability of sensory assessments: Relationship between salivary flow rate and temporal perception of gustatory stimuli. Food Quality and Preference. 1994;5:55-64. DOI: 10.1016/0950-3293(94)90008-6
  47. 47. Spielman AI. Interaction of saliva and taste. Journal of Dental Research. 1990;69:838-848. DOI: 10.1177/00220345900690030101
  48. 48. Matsuo R. Role of saliva in the maintenance of taste sensitivity. Critical Reviews in Oral Biology and Medicine. 2000;11(2):216-229. DOI: 10.1177/10454411000110020501
  49. 49. Lamy E, Torregrossa AM, Castelo PM, Silva FC. Saliva in ingestive behavior research: Association with oral sensory perception and food intake. In: Tvarijonaviciute A, Martínez-Subiela S, López-Jornet P, Lamy E, editors. Saliva in Health and Disease. Cham: Springer; 2020. pp. 23-48. DOI: 10.1007/978-3-030-37681-9_2
  50. 50. Soares S, Mateus N, De Freitas V. Interaction of different polyphenols with bovine serum albumin (BSA) and human salivary α-amylase (HSA) by fluorescence quenching. Journal of Agricultural and Food Chemistry. 2007;55:6727-6735. DOI: 10.1021/jf070905x
  51. 51. Hay DI, Bennick A, Schlesinger DH, Minaguchi K, Madapallimattam G, Schluckebier SK. The primary structures of six human salivary acidic proline-rich proteins (PRP-1, PRP-2, PRP-3, PRP-4, PIF-s and PIF-f). The Biochemical Journal. 1988;255(1):15-21. DOI: 10.1042/bj2550015
  52. 52. Murray NJ, Williamson MP. Conformational study of a salivary proline-rich protein repeat sequence. European Journal of Biochemistry. 1994;219(3):915-921. DOI: 10.1111/j.1432-1033.1994.tb18573.x
  53. 53. Mehansho H, Hagerman A, Clements S, Butler L, Rogler J, Carlson DM. Modulation of proline-rich protein biosynthesis in rat parotid glands by sorghums with high tannin levels. Proceedings of the National Academy of Sciences. 1983;80(13):3948-3952. DOI: 10.1073/pnas.80.13.3948
  54. 54. Mole S, Butler LG, Iason G. Defense against dietary tannin in herbivores: A survey for proline rich salivary proteins in mammals. Biochemical Systematics and Ecology. 1990;18(4):287-293. DOI: 10.1016/0305-1978(90)90073-O
  55. 55. Oppenheim FG, Yang YC, Diamond RD, Hyslop D, Offner GD, Troxler RF. The primary structure and functional characterisation of the neutral histidine-rich polypeptide from human parotid secretion. The Journal of Biological Chemistry. 1986;261(3):1177-1182
  56. 56. Troxler RF, Offner GD, Xu T, Vanderspek JC, Oppenheim FG. Structural relationship between human salivary histatins. Journal of Dental Research. 1990;69(1):2-6. DOI: 10.1177/00220345900690010101
  57. 57. Inzitari R, Cabras T, Rossetti DV, Fanali C, Vitali A, Pellegrini M, et al. Detection in human saliva of different statherin and P-B fragments and derivatives. Proteomics. 2006;6:6370-6379. DOI: 10.1002/pmic.200600395
  58. 58. Wróblewski K, Muhandiram R, Chakrabartty A, Bennick A. The molecular interaction of human salivary histatins with polyphenolic compounds. European Journal of Biochemistry. 2001;268(16):4384-4397. DOI: 10.1046/j.1432-1327.2001.02350.x
  59. 59. Baxter NJ, Lilley TH, Haslam E, Williamson MP. Multiple interactions between polyphenols and a salivary proline-rich protein repeat result in complexation and precipitation. Biochemistry. 1997;36(18):5566-5577. DOI: 10.1021/bi9700328
  60. 60. Soares S, Mateus N, De Freitas V. Interaction of different classes of salivary proteins with food tannins. Foodservice Research International. 2012;49(2):807-813. DOI: 10.1016/j.foodres.2012.09.008
  61. 61. Soares S, Santos Silva M, García-Estévez I, Brandão E, Fonseca F, Ferreira-da-Silva F, et al. Effect of malvidin-3-glucoside and epicatechin interaction on their ability to interact with salivary proline-rich proteins. Food Chemistry. 2019;276:33-42. DOI: 10.1016/j.foodchem.2018.09.167
  62. 62. Tabak LA. Structure and function of human salivary mucins. Critical Reviews in Oral Biology and Medicine. 1990;1(4):229-234. DOI: 10.1177/10454411900010040201
  63. 63. Condelli N, Dinnella C, Cerone A, Monteleone E, Bertuccioli M. Prediction of perceived astringency induced by phenolic compounds II: Criteria for panel selection and preliminary application on wine samples. Food Quality and Preference. 2006;17:96-107. DOI: 10.1016/j.foodqual.2005.04.009
  64. 64. Kauffman DL, Watanabe S, Evans JR, Keller PJ. The existence of glycosylated and non-glycosylated forms of human submandibular amylase. Archives of Oral Biology. 1973;18(9):1105-1113. DOI: 10.1016/0003-9969(73)90084-8
  65. 65. Muus J. The amino acid composition of human salivary Amylase1. Journal of the American Chemical Society. 1954;76(20):5163-5165
  66. 66. Gyémánt G, Zajácz Á, Bécsi B, Ragunath C, Ramasubbu N, Erdődi F, et al. Evidence for pentagalloyl glucose binding to human salivary α-amylase through aromatic amino acid residues. Biochimica et Biophysica Acta (BBA) - Proteins and Proteomics. 2009;1794:291-296. DOI: 10.1016/j.bbapap.2008.10.012
  67. 67. Swain T, Bate-Smith EC. Flavonoid compounds. In: Florkin M, Mason HS, editors. Comparative Biochemist, a Comprehensive Treaty, Vol. III: Constituents of Life-Part A. New York: Academic Press; 1962. pp. 755-809
  68. 68. Es-Safi N, Fulcrand H, Cheynier V, Moutounet M. Studies on the acetaldehyde-induced condensation of (−)-epicatechin and malvidin 3-o-glucoside in a model solution system. Journal of Agricultural and Food Chemistry. 1999;47(5):2096-2102. DOI: 10.1021/jf9806309
  69. 69. Prieur C, Rigaud J, Cheynier V, Moutounet M. Oligomeric and polymeric procyanidins from grape seeds. Phytochemistry. 1994;36:781-784. DOI: 10.1016/S0031-9422(00)89817-9
  70. 70. Souquet JM, Cheynier V, Brossaud F, Moutounet M. Polymeric proanthocyanidins from grape skins. Phytochemistry. 1996;43:509-512. DOI: 10.1016/0031-9422(96)00301-9
  71. 71. Souquet JM, Labarbe B, Le Guernevé C, Cheynier V, Moutounet M. Phenolic composition of grape stems. Journal of Agricultural and Food Chemistry. 2000;48:1076-1080. DOI: 10.1021/jf991171u
  72. 72. Targett NM, Arnold TM. Minireview—Predicting the effects of brown algal phlorotannins on marine herbivores in tropical and temperate oceans. Journal of Phycology. 1998;34:195-205. DOI: 10.1046/j.1529-8817.1998.340195.x
  73. 73. McManus JP, Davis KG, Beart JE, Gaffney SH, Lilley TH, Haslam E. Polyphenol interactions. Part 1. Introduction; some observations on the reversible complexation of polyphenols with proteins and polysaccharides. Journal of the Chemical Society, Perkin Transactions. 1985;2:1429-1438. DOI: 10.1039/P29850001429
  74. 74. Salminen JP, Karonen M, Sinkkonen J. Chemical ecology of tannins: Recent developments in tannin chemistry reveal new structures and structure–activity patterns. Chemistry--A European Journal. 2011;17(10):2806-2816. DOI: 10.1002/chem.201002662
  75. 75. Vidal S, Francis IL, Guyot S, Marnet N, Kwiatkowski M, Gawel R, et al. The mouthfeel properties of grape and apple proanthocyanidins in a wine-like medium. Journal of the Science of Food and Agriculture. 2003;83:564-573. DOI: 10.1002/jsfa.1394
  76. 76. Sun B, Md S, Leandro C, Caldeira I, Duarte FL, Spranger I. Reactivity of polymeric proanthocyanidins toward salivary proteins and their contribution to young red wine astringency. Journal of Agricultural and Food Chemistry. 2013;61(4):939-946. DOI: 10.1021/jf303704u
  77. 77. Hufnagel JC, Hofmann T. Orosensory-directed identification of astringent mouthfeel and bitter-tasting compounds in red wine. Journal of Agricultural and Food Chemistry. 2008;56(4):1376-1386. DOI: 10.1021/jf073031n
  78. 78. Rinaldi A, Jourdes M, Teissedre P, Moio L. A preliminary characterisation of Aglianico (Vitis vinifera L. cv.) grape proanthocyanidins and evaluation of their reactivity towards salivary proteins. Food Chemistry. 2014;164:142-149. DOI: 10.1016/j.foodchem.2014.05.050
  79. 79. Gonzalo-Diago A, Dizy M, Fernández-Zurbano P. Taste and mouthfeel properties of red wines proanthocyanidins and their relation to the chemical composition. Journal of Agricultural and Food Chemistry. 2013;61(37):8861-8870. DOI: 10.1021/jf401041q
  80. 80. Hagerman AE, Rice ME, Ritchard NT. Mechanisms of protein precipitation for two tannins, pentagalloyl glucose and epicatechin16(4→8) catechin (procyanidin). Journal of Agricultural and Food Chemistry. 1998;46(7):2590-2595. DOI: 10.1021/jf971097k
  81. 81. Charlton AJ, Baxter NJ, Khan ML, Moir AJG, Haslam E, Davies AP, et al. Polyphenol/peptide binding and precipitation. Journal of Agricultural and Food Chemistry. 2002;50(6):1593-1601. DOI: 10.1021/jf010897z
  82. 82. Hofmann T, Glabasnia A, Schwarz B, Wisman KN, Gangwer KA, Hagerman AE. Protein binding and astringent taste of a polymeric procyanidin, 1,2,3,4,6-penta-O-galloyl-β-D-glucopyranose, castalagin, and grandinin. Journal of Agricultural and Food Chemistry. 2006;54(25):9503-9509. DOI: 10.1021/jf062272c
  83. 83. Rinaldi A, Gambuti A, Moine-Ledoux V, Moio L. Evaluation of the astringency of commercial tannins by means of the SDS–PAGE-based method. Food Chemistry. 2010;122(4):951-956. DOI: 10.1016/j.foodchem.2010.03.105
  84. 84. Dobreva MA, Green RJ, Mueller-Harvey I, Salminen J, Howlin BJ, Frazier RA. Size and molecular flexibility affect the binding of ellagitannins to bovine serum albumin. Journal of Agricultural and Food Chemistry. 2014;62(37):9186-9194. DOI: 10.1021/jf502174r
  85. 85. Stark T, Wollmann N, Wenker K, Lösch S, Glabasnia A, Hofmann T. Matrix-calibrated LC-MS/MS quantitation and sensory evaluation of oak ellagitannins and their transformation products in red wines. Journal of Agricultural and Food Chemistry. 2010;58(10):6360-6369. DOI: 10.1021/jf100884y
  86. 86. Lee CB, Lawless HT. Time-course of astringent sensations. Chemical Senses. 1991;16(3):225-238. DOI: 10.1093/chemse/16.3.225
  87. 87. Corrigan Thomas CJ, Lawless HT. Astringent subqualities in acids. Chemical Senses. 1995;20(6):593-600. DOI: 10.1093/chemse/20.6.593
  88. 88. Rubico SM, McDaniel MR. Sensory evaluation of acids by free-choice profiling. Chemical Senses. 1992;17(3):273-289. DOI: 10.1093/chemse/17.3.273
  89. 89. Kallithraka S, Bakker J, Clifford M. Red wine and model wine astringency as affected by malic and lactic acid. Journal of Food Science. 1997;62(2):416-420. DOI: 10.1111/j.1365-2621.1997.tb04016.x
  90. 90. Corrigan Thomas CJ, Lawless HT. Astringent subqualities in acids. Chemical Senses. 1995;20(6):593-600. DOI: 10.1093/chemse/20.6.593
  91. 91. Picariello L, Rinaldi A, Martino F, Petracca F, Moio L, Gambuti A. Modification of the organic acid profile of grapes due to climate changes alters the stability of red wine phenolics during controlled oxidation. Vitis. 2019;58(5):127-133. DOI: 10.5073/vitis.2019.58.special-issue.127-133
  92. 92. Forino M, Picariello L, Rinaldi A, Moio L, Gambuti A. How must pH affects the level of red wine phenols. Lebensmittel-Wissenschaft & Technologie. 2020;129:109546. DOI: 10.1016/j.lwt.2020.109546
  93. 93. He F, Liang N, Mu L, Pan Q, Wang J, Reeves MJ, et al. Anthocyanins and their variation in red wines I. Monomeric anthocyanins and their colour expression. Molecules. 2012;17(2):1571-1601. DOI: 10.3390/molecules17021571
  94. 94. Vidal S, Francis L, Williams P, Kwiatkowski M, Gawel R, Cheynier V, et al. The mouthfeel properties of polysaccharides and anthocyanins in a wine like medium. Food Chemistry. 2004;85(4):519-525. DOI: 10.1016/S0308-8146(03)00084-0
  95. 95. Vidal S, Francis L, Noble A, Kwiatkowski M, Cheynier V, Waters E. Taste and mouthfeel properties of different types of tannin-like polyphenolic compounds and anthocyanins in wine. Analytica Chimica Acta. 2004;513(1):57-65. DOI: 10.1016/j.aca.2003.10.017
  96. 96. Ferrer-Gallego R, Soares S, Mateus N, Rivas-Gonzalo J, Escribano-Bailón MT, de Freitas V. New anthocyanin–human salivary protein complexes. Langmuir. 2015;31(30):8392-8401. DOI: 10.1021/acs.langmuir.5b01122
  97. 97. Paissoni MA, Waffo-Teguo P, Ma W, Jourdes M, Rolle L, Teissedre P. Chemical and sensorial investigation of in-mouth sensory properties of grape anthocyanins. Scientific Reports. 2018;8:17098. DOI: 10.1038/s41598-018-35355-x
  98. 98. García-Estévez I, Cruz L, Oliveira J, Mateus N, de Freitas V, Soares S. First evidences of interaction between pyranoanthocyanins and salivary proline-rich proteins. Food Chemistry. 2017;228:574-581. DOI: 10.1016/j.foodchem.2017.02.030
  99. 99. Hufnagel JC, Hofmann T. Orosensory-directed identification of astringent mouthfeel and bitter-tasting compounds in red wine. Journal of Agricultural and Food Chemistry. 2008;56(4):1376-1386. DOI: 10.1021/jf073031n
  100. 100. Ferrer-Gallego R, Brás NF, García-Estévez I, Mateus N, Rivas-Gonzalo JC, de Freitas V, et al. Effect of flavonols on wine astringency and their interaction with human saliva. Food Chemistry. 2016;209:358-364. DOI: 10.1016/j.foodchem.2016.04.091
  101. 101. Gambuti A, Picariello L, Rinaldi A, Forino M, Blaiotta G, Moine V, et al. New insights into the formation of precipitates of quercetin in Sangiovese wines. Journal of Food Science and Technology. 2020;57(7):2602-2611. DOI: 10.1007/s13197-020-04296-7
  102. 102. Rinaldi A, Louazil P, Iturmendi N, Moine V, Moio L. Effect of marc pressing and geographical area on Sangiovese wine quality. Lebensmittel-Wissenschaft & Technologie. 2020;118:108728. DOI: 10.1016/j.lwt.2019.108728
  103. 103. Sáenz-Navajas M, Avizcuri J, Ferreira V, Fernández-Zurbano P. Insights on the chemical basis of the astringency of Spanish red wines. Food Chemistry. 2012;134(3):1484-1493. DOI: 10.1016/j.foodchem.2012.03.060
  104. 104. Luck G, Liao H, Murray NJ, Grimmer HR, Warminski EE, Williamson MP, et al. Polyphenol, astringency and proline rich proteins. Phytochemistry. 1994;37:357-371
  105. 105. Haslam E. Polyphenol-protein interactions. The Biochemical Journal. 1974;139:285-288
  106. 106. Haslam E. Natural polyphenols (vegetable tannins) as drugs: Possible modes of action. Journal of Natural Products. 1996;59(2):205-215. DOI: 10.1021/np960040+
  107. 107. Beart JE, Lilley TH, Haslam E. Plant polyphenols—Secondary metabolism and chemical defence: Some observations. Phytochemistry. 1985;24(1):33-38. DOI: 10.1016/S0031-9422(00)80802-X
  108. 108. Jöbstl E, O’Connell J, Fairclough JPA, Williamson MP. Molecular model for astringency produced by polyphenol/protein interactions. Biomacromolecules. 2004;5(3):942-949. DOI: 10.1021/bm0345110
  109. 109. Frazier RA, Papadopoulou A, Mueller-Harvey I, Kissoon D, Green RJ. Probing protein−tannin interactions by isothermal titration microcalorimetry. Journal of Agricultural and Food Chemistry. 2003;51(18):5189-5195. DOI: 10.1021/jf021179v
  110. 110. Poncet-Legrand C, Edelmann A, Putaux J, Cartalade D, Sarni-Manchado P, Vernhet A. Poly(L-proline) interactions with flavan-3-ols units: Influence of the molecular structure and the polyphenol/protein ratio. Food Hydrocolloids. 2006;20(5):687-697. DOI: 10.1016/j.foodhyd.2005.06.009
  111. 111. Gonçalves R, Mateus N, Pianet I, Laguerre M, de Freitas V. Mechanisms of tannin-induced trypsin inhibition: A molecular approach. Langmuir. 2011;27(21):13122-13129. DOI: 10.1021/la202280c
  112. 112. Poncet-Legrand C, Gautier C, Cheynier V, Imberty A. Interactions between flavan-3-ols and poly(L-proline) studied by isothermal titration calorimetry: Effect of the tannin structure. Journal of Agricultural and Food Chemistry. 2007;55(22):9235-9240. DOI: 10.1021/jf071297o
  113. 113. Cala O, Pinaud N, Simon C, Fouquet E, Laguerre M, Dufourc EJ, et al. NMR and molecular modeling of wine tannins binding to saliva proteins: Revisiting astringency from molecular and colloidal prospects. The FASEB Journal. 2010;24(11):4281-4290. DOI: 10.1096/fj.10-158741
  114. 114. Folin O, Denis W. On phosphotungstic-phosphomolybdic compounds as colour reagents. The Journal of Biological Chemistry. 1912;12:239-243
  115. 115. Swain T, Hillis WE. The phenolic constituents of Prunus domestica. I.—The quantitative analysis of phenolic constituents. Journal of the Science of Food and Agriculture. 1959;10(1):63-68. DOI: 10.1002/jsfa.2740100110
  116. 116. Laurent S. Étude comparative de différentes méthodes d'extraction et de dosage des tannins chez quelques ptéridophytes. Archives Internationales de Physiologie et de Biochimie. 1975;83(4):735-752. DOI: 10.3109/13813457509081892
  117. 117. Singleton VL, Rossi JA. Colorimetry of total phenolics with phosphomolybdic-phosphotungstic acid reagents. American Journal of Enology and Viticulture. 1965;16(3):144-158
  118. 118. Bate-Smith E. Phytochemistry of proanthocyanidins. Phytochemistry. 1975;14(4):1107-1113. DOI: 10.1016/0031-9422(75)85197-1
  119. 119. Porter LJ, Hrstich LN, Chan BG. The conversion of procyanidins and prodelphinidins to cyanidin and delphinidin. Phytochemistry. 1985;25(1):223-230. DOI: 10.1016/S0031-9422(00)94533-3
  120. 120. Broadhurst RB, Jones WT. Analysis of condensed tannins using acidified vanillin. Journal of the Science of Food and Agriculture. 1978;29(9):788-794. DOI: 10.1002/jsfa.2740290908
  121. 121. McMurrough I, McDowell J. Chromatographic separation and automated analysis of flavanols. Analytical Biochemistry. 1978;91(1):92-100. DOI: 10.1016/0003-2697(78)90819-9
  122. 122. Vivas N, Glories Y, Lagune L, Cédric S, Augustin M. Estimation du degré de polymérisation des procyanidines du raisin et du vin par la méthode au ρ-dimethylaminocinnamaldéhyde. OENO One. 1994;28(4):319. DOI: 10.20870/oeno-one.1994.28.4.1138
  123. 123. Schofield P, Mbugua D, Pell A. Analysis of condensed tannins: A review. Animal Feed Science and Technology. 2001;91(1-2):21-40. DOI: 10.1016/S0377-8401(01)00228-0
  124. 124. Matthews S, Mila I, Scalbert A, Donnelly DM. Extractable and non-extractable proanthocyanidins in barks. Phytochemistry. 1997;45(2):405-410. DOI: 10.1016/S0031-9422(96)00873-4
  125. 125. Kennedy JA, Jones GP. Analysis of proanthocyanidin cleavage products following acid-catalysis in the presence of excess phloroglucinol. Journal of Agricultural and Food Chemistry. 2001;49(4):1740-1746. DOI: 10.1021/jf001030o
  126. 126. Lorrain B, Ky I, Pechamat L, Teissedre P. Evolution of analysis of polyhenols from grapes, wines, and extracts. Molecules. 2013;18(1):1076-1100. DOI: 10.3390/molecules18011076
  127. 127. McRae JM, Falconer RJ, Kennedy JA. Thermodynamics of grape and wine tannin interaction with polyproline: Implications for red wine astringency. Journal of Agricultural and Food Chemistry. 2010;58(23):12510-12518. DOI: 10.1021/jf1030967
  128. 128. Kennedy JA, Waterhouse AL. Analysis of pigmented high-molecular-mass grape phenolics using ion-pair, normal-phase high-performance liquid chromatography. Journal of Chromatography. A. 2000;866(1):25-34. DOI: 10.1016/S0021-9673(99)01038-9
  129. 129. Fernández K, Agosin E. Quantitative analysis of red wine tannins using fourier-transform mid-infrared spectrometry. Journal of Agricultural and Food Chemistry. 2007;55(18):7294-7300. DOI: 10.1021/jf071193d
  130. 130. Bate-Smith E. Haemanalysis of tannins: The concept of relative astringency. Phytochemistry. 1973;12(4):907-912. DOI: 10.1016/0031-9422(73)80701-0
  131. 131. Nierenstein M. The Natural Organic Tannins. London: J&A Churchill Ltd; 1934
  132. 132. Swain T. Tannins and lignins. In: Rosenthal GA, Janzen DH, editors. Herbivores: Their Interactions with Secondary Plant Metabolites. New York: Academic Press; 1979. pp. 657-682
  133. 133. Goldstein JL, Swain T. The inhibition of enzymes by tannins. Phytochemistry. 1965;4:185-192
  134. 134. Hagerman AE, Butler LG. Protein precipitation method for the quantitative determination of tannins. Journal of Agricultural and Food Chemistry. 1978;26(4):809-812. DOI: 10.1021/jf60218a027
  135. 135. Harbertson JF, Picciotto EA, Adams DO. Measurement of polymeric pigments in grape berry extract sand wines using a protein precipitation assay combined with bisulfite bleaching. American Journal of Enology and Viticulture. 2003;54(4):301-306
  136. 136. Glories Y. La couleur des vins rouges. 1°2° partie. Connaissance Vigne Vin. 1984;18:253-271
  137. 137. Llaudy MC, Canals R, Canals J, Rozés N, Arola L, Zamora F. New method for evaluating astringency in red wine. Journal of Agricultural and Food Chemistry. 2004;52(4):742-746. DOI: 10.1021/jf034795f
  138. 138. Sarneckis CJ, Dambergs RG, Jones P, Mercurio M, Herderich MJ, Smith PA. Quantification of condensed tannins by precipitation with methyl cellulose: Development and validation of an optimised tool for grape and wine analysis. Australian Journal of Grape and Wine Research. 2008;12(1):39-49. DOI: 10.1111/j.1755-0238.2006.tb00042.x
  139. 139. Mercurio MD, Dambergs RG, Herderich MJ, Smith PA. High throughput analysis of red wine and grape phenolics adaptation and validation of methyl cellulose precipitable tannin assay and modified Somers color assay to a rapid 96 well plate format. Journal of Agricultural and Food Chemistry. 2007;55(12):4651-4657. DOI: 10.1021/jf063674n
  140. 140. Yan Q, Bennick A. Identification of histatins as tannin-binding proteins in human saliva. The Biochemical Journal. 1995;311(1):341-347. DOI: 10.1042/bj3110341
  141. 141. Sarni-Manchado P, Cheynier V, Moutounet M. Interactions of grape seed tannins with salivary proteins. Journal of Agricultural and Food Chemistry. 1999;47(1):42-47. DOI: 10.1021/jf9805146
  142. 142. Gambuti A, Rinaldi A, Pessina R, Moio L. Evaluation of Aglianico grape skin and seed polyphenol astringency by SDS–PAGE electrophoresis of salivary proteins after the binding reaction. Food Chemistry. 2006;97(4):614-620. DOI: 10.1016/j.foodchem.2005.05.038
  143. 143. Kallithraka S, Bakker J, Clifford MN. Evidence that salivary proteins are involved in astringency. Journal of Sensory Studies. 1998;13(1):29-43. DOI: 10.1111/j.1745-459X.1998.tb00073.x
  144. 144. Soares S, Vitorino R, Osório H, Fernandes A, Venâncio A, Mateus N, et al. Reactivity of human salivary proteins families toward food polyphenols. Journal of Agricultural and Food Chemistry. 2011;59(10):5535-5547. DOI: 10.1021/jf104975d
  145. 145. Dinnella C, Recchia A, Vincenzi S, Tuorila H, Monteleone E. Temporary modification of salivary protein profile and individual responses to repeated phenolic astringent stimuli. Chemical Senses. 2010;35(1):75-85. DOI: 10.1093/chemse/bjp084
  146. 146. Chapon L. Nephelometry as a method for studying the relations between polyphenols and proteins. Journal of the Institute of Brewing. 1993;99(1):49-56. DOI: 10.1002/j.2050-0416.1993.tb01146.x
  147. 147. de Freitas V, Mateus N. Nephelometric study of salivary protein-tannin aggregates. Journal of the Science of Food and Agriculture. 2002;82(1):113-119. DOI: 10.1002/jsfa.1016
  148. 148. Horne J, Hayes J, Lawless HT. Turbidity as a measure of salivary protein reactions with astringent substances. Chemical Senses. 2002;27(7):653-659. DOI: 10.1093/chemse/27.7.653
  149. 149. Carvalho E, Mateus N, De Freitas V. Flow nephelometric analysis of protein–tannin interactions. Analytica Chimica Acta. 2004;513(1):97-101. DOI: 10.1016/j.aca.2003.10.010
  150. 150. Monteleone E, Condelli N, Dinnella C, Bertuccioli M. Prediction of perceived astringency induced by phenolic compounds. Food Quality and Preference. 2004;15(7-8):761-769. DOI: 10.1016/j.foodqual.2004.06.002
  151. 151. Fia G, Dinnella C, Bertuccioli M, Monteleone E. Prediction of grape polyphenol astringency by means of a fluorimetric micro-plate assay. Food Chemistry. 2009;113(1):325-330. DOI: 10.1016/j.foodchem.2008.07.058
  152. 152. Rébénaque P, Rawyler A, Boldi M, Deneulin P. Comparison between sensory and nephelometric evaluations of tannin fractions obtained by ultrafiltration of red wines. Chemosensory Perception. 2015;8(1):33-43. DOI: 10.1007/s12078-015-9175-x
  153. 153. Guerreiro JRL, Teixeira N, de Freitas V, Sales MGF, Sutherland DS. A saliva molecular imprinted localised surface plasmon resonance biosensor for wine astringency estimation. Food Chemistry. 2017;233:457-466. DOI: 10.1016/j.foodchem.2017.04.051
  154. 154. Lyman BJ, Green BG. Oral astringency: Effects of repeated exposure and interactions with sweeteners. Chemical Senses. 1990;15(2):151-164. DOI: 10.1093/chemse/15.2.151
  155. 155. Lee WE, Pangborn RM. Time-intensity: The temporal aspects of sensory perception. Food Technology. 1986;40(11):71-82
  156. 156. Pineau N, Schlich P, Cordelle S, Mathonnière C, Issanchou S, Imbert A, et al. Temporal dominance of sensations: Construction of the TDS curves and comparison with time–intensity. Food Quality and Preference. 2009;20(6):450-455. DOI: 10.1016/j.foodqual.2009.04.005
  157. 157. Clark CC, Lawless HT. Limiting response alternatives in time-intensity scaling: An examination of the halo-dumping effect. Chemical Senses. 1994;19(6):583-594. DOI: 10.1093/chemse/19.6.583
  158. 158. Pessina R, Patron C, Pineau N, Piombino P, Moio L, Schlich P. Measuring temporality of sensations in wine. In: European Conference on Sensory Science of Food and Beverages “A Sense of Identity”. 2004. pp. 26-29
  159. 159. Schlich P. Temporal dominance of sensations (TDS): A new deal for temporal sensory analysis. Current Opinion in Food Science. 2017;15:38-42. DOI: 10.1016/j.cofs.2017.05.003
  160. 160. Gawel R, Oberholster A, Il F. A ‘Mouth-feel Wheel’: Terminology for communicating the mouthfeel characteristics of red wine. Australian Journal of Grape and Wine Research. 2000;6(3):203-207. DOI: 10.1111/j.1755-0238.2000.tb00180.x
  161. 161. Vidal L, Antúnez L, Giménez A, Medina K, Boido E, Ares G. Sensory characterisation of the astringency of commercial Uruguayan Tannat wines. Foodservice Research International. 2017;102:425-434. DOI: 10.1016/j.foodres.2017.09.022
  162. 162. Rinaldi A, Coppola M, Moio L. Aging of Aglianico and Sangiovese wine on mannoproteins: Effect on astringency and colour. Lebensmittel-Wissenschaft & Technologie. 2019;105:233-241. DOI: 10.1016/j.lwt.2019.02.034
  163. 163. Goldner M, Zamora M. Effect of polyphenol concentrations on astringency perception and its correlation with gelatin index of red wine. Journal of Sensory Studies. 2010;25(5):761-777. DOI: 10.1111/j.1745-459X.2010.00304.x
  164. 164. Kallithraka S, Kim D, Tsakiris A, Paraskevopoulos I, Soleas G. Sensory assessment and chemical measurement of astringency of Greek wines: Correlations with analytical polyphenolic composition. Food Chemistry. 2011;126(4):1953-1958. DOI: 10.1016/j.foodchem.2010.12.045
  165. 165. Cliff M, Brau N, King MC, Mazza G. Development of predictive models for astringency from anthocyanin, phenolic and color analyses of British Columbia red wines. Journal International des Sciences de la Vigne et du Vin. 2002;36:21-30
  166. 166. Kennedy JA, Ferrier J, Harbertson JF, des Gachons CP. Analysis of tannins in red wine using multiple methods: Correlation with perceived astringency. American Journal of Enology and Viticulture. 2006;57(4):481-485
  167. 167. Rinaldi A, Gambuti A, Moio L. Application of the SPI (saliva precipitation index) to the evaluation of red wine astringency. Food Chemistry. 2012;135(4):2498-2504. DOI: 10.1016/j.foodchem.2012.07.031
  168. 168. Rinaldi A, Iturmendi N, Gambuti A, Jourdes M, Teissedre P, Moio L. Chip electrophoresis as a novel approach to measure the polyphenols reactivity toward human saliva. Electrophoresis. 2014;35(11):1735-1741. DOI: 10.1002/elps.201300622
  169. 169. Ribéreau-Gayon P, Glories Y, Maujean A, Dubourdieu D. The chemistry of wine stabilisation and treatments. In: Handbook of Enology. Chichester, United Kingdom: John Wiley and Sons; 2006
  170. 170. Maury C, Sarni-Manchado P, Lefebvre S, Cheynier V, Moutounet M. Influence of fining with different molecular weight gelatins on proanthocyanidin composition and perception of wines. American Journal of Enology and Viticulture. 2001;52(2):140-145
  171. 171. Gambuti A, Rinaldi A, Moio L. Use of patatin, a protein extracted from potato as alternative to animal proteins in fining of red wine. European Food Research and Technology. 2012;235(4):753-765. DOI: 10.1007/s00217-012-1791-y
  172. 172. Rinaldi A, Errichiello F, Moio L. Alternative fining of Sangiovese: Effect on polyphenolic and sensory characteristics. Australian Journal of Grape and Wine Research. 2020
  173. 173. OIV. International Code of Oenological Practices. Paris, France: Organisation Internationale de la Vigne et du Vin; 2012. p. 1
  174. 174. Sheridan MK, Elias RJ. Exogenous acetaldehyde as a tool for modulating wine color and astringency during fermentation. Food Chemistry. 2015;177:17-22. DOI: 10.1016/j.foodchem.2014.12.077
  175. 175. Saucier C, Little D, Glories Y. First evidence of acetaldehyde-flavanol condensation products in red wine. American Journal of Enology and Viticulture. 1997;48(3):370-373
  176. 176. Fulcrand H, Atanasova V, Salas E, Cheynier V. The fate of anthocyanins in wine: Are there determining factors? In: Red Wine Color. ACS Symposium Series. Washington, USA: American Chemical Society; Vol. 886. 2004. pp. 68-88. DOI: 10.1021/bk-2004-0886.ch006
  177. 177. Picariello L, Gambuti A, Petracca F, Rinaldi A, Moio L. Enological tannins affect acetaldehyde evolution, colour stability and tannin reactivity during forced oxidation of red wine. International Journal of Food Science and Technology. 2018;53(1):228-236. DOI: 10.1111/ijfs.13577
  178. 178. Gambuti A, Picariello L, Rinaldi A, Moio L. Evolution of Sangiovese wines with varied tannin and anthocyanin ratios during oxidative aging. Frontiers in Chemistry. 2018;6(63):1-11. DOI: 10.3389/fchem.2018.00063
  179. 179. McRae JM, Dambergs RG, Kassara S, Parker M, Jeffery DW, Herderich MJ, et al. Phenolic compositions of 50 and 30 year sequences of Australian red wines: The impact of wine age. Journal of Agricultural and Food Chemistry. 2012;60(40):10093-10102. DOI: 10.1021/jf301571q
  180. 180. Weber F, Greve K, Durner D, Fischer U, Winterhalter P. Sensory and chemical characterisation of phenolic polymers from red wine obtained by gel permeation chromatography. American Journal of Enology and Viticulture. 2013;64(1):15-25. DOI: 10.5344/ajev.2012.12074
  181. 181. Ma L, Watrelot AA, Addison B, Waterhouse AL. Condensed tannin reacts with SO2 during wine aging, yielding flavan-3-ol sulfonates. Journal of Agricultural and Food Chemistry. 2018;66(35):9259-9268. DOI: 10.1021/acs.jafc.8b01996
  182. 182. Chira K, Pacella N, Jourdes M, Teissedre P. Chemical and sensory evaluation of Bordeaux wines (cabernet-sauvignon and merlot) and correlation with wine age. Food Chemistry. 2011;126(4):1971-1977. DOI: 10.1016/j.foodchem.2010.12.056
  183. 183. Gambuti A, Picariello L, Rinaldi A, Ugliano M, Moio L. Impact of 5 years bottle aging under controlled oxygen exposure on sulfur dioxide and phenolic composition of tannin-rich red wines. OENO One. 2020
  184. 184. Rinaldi A, Gonzalez A, Moio L. The sensory profile of astringency: Application on Sangiovese wines. In: Proceedings of 11th International Symposium of Oenology of Bordeaux and 11th Edition of the Symposium In Vino Analytica Scientia (ŒnoIVAS 2019). France: Bordeaux; 25-28 June 2019
  185. 185. Gambuti A, Siani T, Picariello L, Rinaldi A, Lisanti MT, Ugliano M, et al. Oxygen exposure of tannins-rich red wines during bottle aging. Influence on phenolics and color, astringency markers and sensory attributes. European Food Research and Technology. 2017;243(4):669-680. DOI: 10.1007/s00217-016-2780-3
  186. 186. Rinaldi A, Gambuti A, Moio L. Precipitation of salivary proteins after the interaction with wine: The effect of ethanol, pH, fructose, and mannoproteins. Journal of Food Science. 2012;77(4):C485-C490. DOI: 10.1111/j.1750-3841.2012.02639.x
  187. 187. Poncet-Legrand C, Cartalade D, Putaux J, Cheynier V, Vernhet A. Flavan-3-ol aggregation in model ethanolic solutions: Incidence of polyphenol structure, concentration, ethanol content, and ionic strength. Langmuir. 2003;19(25):10563-10572. DOI: 10.1021/la034927z
  188. 188. Brandão E, Silva MS, García-Estévez I, Williams P, Mateus N, Doco T, et al. The role of wine polysaccharides on salivary protein-tannin interaction: A molecular approach. Carbohydrate Polymers. 2017;177:77-85. DOI: 10.1016/j.carbpol.2017.08.075

Written By

Alessandra Rinaldi and Luigi Moio

Submitted: 09 June 2020 Reviewed: 17 August 2020 Published: 05 October 2020