Sulfonation Reactions Behind The Fate of White Wine's

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

H

OH
OH
metabolites

Article
Sulfonation Reactions behind the Fate of White Wine’s
Shelf-Life
Maria Nikolantonaki 1, * , Rémy Romanet 1,2 , Marianna Lucio 3 , Philippe Schmitt-Kopplin 3,4
and Régis Gougeon 1,2

1 Institut Universitaire de la Vigne et du Vin, UMR PAM Université de Bourgogne/Institut Agro Dijon,
Jules Guyot, Rue Claude Ladrey, BP 27877, 21078 Dijon, France; remy.romanet@u-bourgogne.fr (R.R.);
regis.gougeon@u-bourgogne.fr (R.G.)
2 DIVVA (Développement Innovation Vigne Vin Aliments) Platform/PAM UMR, IUVV, Rue Claude Ladrey,
BP 27877, CEDEX, 21078 Dijon, France
3 Research Unit Analytical BioGeoChemistry, Helmholtz Zentrum Muenchen, 85764 Neuherberg, Germany;
marianna.lucio@helmholtz-muenchen.de (M.L.); schmitt-kopplin@helmholtz-muenchen.de (P.S.-K.)
4 Analytical Food Chemistry, Technische Universität München, Alte Akademie 10, 85354 Freising, Germany
* Correspondence: maria.nikolantonaki@u-bourgogne.fr

Abstract: White wine’s oxidative stability after several years of bottle aging is synonymous to its
organoleptic quality. In order to gain control over the cascade of chemical reactions that are implicated
in that phenomenon, fourier transform ion cyclotron resonance mass spectrometry (FT-ICR-MS)-
based metabolomics and sensory evaluation were combined for the analysis of a vertical series of
white wines from different vineyard plots. Data mining using supervised cluster analysis allowed the
extraction of known and unknown sulfur- and nitrogen-containing molecular features, with oxidative
 stability molecular markers presenting an increased number of S and O atoms in their formulas. In

their majority, S-containing molecular features possessed between 4 to ~12 O atoms, indicating the
Citation: Nikolantonaki, M.;
relatively higher importance of sulfonation reactions as opposed to dimerization reactions. Molecular
Romanet, R.; Lucio, M.;
networking, based on sulfonation reaction transformations, evidences the importance of hitherto
Schmitt-Kopplin, P.; Gougeon, R.
unknown and/or minor sulfur dioxide binders (peptides, aldehydes, and polyphenols) on wine’s
Sulfonation Reactions behind the
oxidative stability.
Fate of White Wine’s Shelf-Life.
Metabolites 2022, 12, 323.
https://doi.org/10.3390/
Keywords: chardonnay; oxidation; oxidative stability; peptides; phenolic compounds; glutathione
metabo12040323

Academic Editor: Jose


Lorenzo Rodriguez
1. Introduction
Received: 8 March 2022 During bottle aging, wine is exposed to relatively low quantities of oxygen that are
Accepted: 29 March 2022 nevertheless sufficient to influence the outcomes of their shelf-life period. In particular,
Published: 2 April 2022 reactive oxygen species (ROS), formed by the reduction of molecular oxygen, modulate the
Publisher’s Note: MDPI stays neutral extent of different reactions involving volatile and nonvolatile components, resulting in the
with regard to jurisdictional claims in formation and/or degradation of a number of powerful aroma compounds, with major
published maps and institutional affil- consequences on the process of aromas’ oxidative evolution [1]. At this time, numerous
iations. authors have already reported studies of the mechanisms involved in wine oxidation, and
many of them were targeted at phenolic compounds, considered to be primary substrates
for oxidation [2]. However, other chemical reactions taking place during bottle aging do
not involve oxygen, meaning that, even in an environment completely devoid of oxygen, a
Copyright: © 2022 by the authors. certain form of aging will occur [3].
Licensee MDPI, Basel, Switzerland. Above all, these results emphasized the dependency of the oxygenation process on
This article is an open access article
the wine’s compositions, where different wines could exhibit different reactivity towards
distributed under the terms and
oxygen, thus stressing the need for analytical approaches that can provide comprehensive
conditions of the Creative Commons
pictures of subtle mechanistic variations among series of samples. In that respect, non-
Attribution (CC BY) license (https://
targeted mass spectrometry-based metabolomics have recently shown great potential in
creativecommons.org/licenses/by/
describing the evolution of chemical spaces involved in enological practices [4–7], but
4.0/).

Metabolites 2022, 12, 323. https://doi.org/10.3390/metabo12040323 https://www.mdpi.com/journal/metabolites


Metabolites 2022, 12, 323 2 of 10

very few non-targeted studies have actually analyzed the evolution of wine during bottle
aging [8–11] and even fewer have done so in relation to its oxidative evolution [12]. A
recent study from our research group showed that controlled electrochemical oxidation
of chardonnay wines could reveal the diversity of sulfur-containing compounds, which
are most sensitive to oxidative degradation, and the nitrogen-containing compounds
resulting from oxidation [13]. Moreover, the application of a multidisciplinary approach
that combined sensory evaluation, chemical and metabolomic analyses, and investigating
the oxygen transfer through the bottleneck/stopper elucidated the importance of the
glass/cork interface [12]. It shows unambiguously that the transfer of oxygen at the
interface between the cork stopper and the glass bottleneck must be considered a potentially
significant contributor to the oxidation state during bottle aging, leading to a notable
modification of a wine’s sulfur–nitrogen chemical signature.
Sulfur- and nitrogen-containing compounds constitute the antioxidant molecular
capital of white wines in the bottle and determine their aging potential [14,15]. These
compounds are grape, yeast, and bacteria metabolism derivatives and act as natural an-
tioxidants while reacting as sacrificial nucleophiles with quinones and protecting wines
against oxidative spoilage [16–19]. In addition, sulfites are present in the most widely
used additive as antioxidant in winemaking. High-resolution mass spectrometry-based
metabolomic studies could easily fingerprint wines according to the SO2 concentration
added to must at pressing or to wine at bottling [20,21]. These results thus bring unprece-
dented molecular scale information about the sulfur-related chemistry that is involved
immediately at the beginning of the winemaking process in the must and that is carried
out throughout bottle aging. It is shown that a great diversity of metabolite classes are
actually impacted by sulfur dioxide additions to must or wine, including amino acids,
carbohydrates, phenolic compounds, and indoles. S-sulfonated products of flavanols,
indoles, and amino acids have been recently reported in bottle-aged wines from different
cultivars, while the physicochemical parameters controlling sulfonation reactions remain
today poorly understood [21,22].
All these studies highlighted the extent to which the chemistry that could be involved
in wine’s oxidative stability during bottle aging is still unknown. The main scope of the
present study was to examine the importance of key chemical reactions such as sulfonation
in wine’s oxidative stability during bottle aging. In particular, in this work, we aimed to
compare the metabolic profile of white wines, sealed with a natural cork, sensory evaluated
as oxidized or good (not oxidized) after several years of bottle aging under cellar conditions.
To obtain a high level of variability, the experimental design included wines from the same
winery from two vintages and five vineyard plots. An investigation of known and unknown
sulfonation reaction products was performed by means of in silico deconjugation.

2. Results and Discussion


2.1. Multivariate FT-ICR-MS-Based Statistical Analysis of Wines
Unsupervised PCA was initially conducted to obtain an overview of FT-ICR-MS data
from wines exhibiting oxidative (O) or not (G) sensory characters. As shown in Figure 1,
samples were mostly discriminated according to the vintage along with the first component
and according to plots along with the second component. Interestingly, for each plot within
a given vintage, G and O samples appeared grouped, although this was to a lesser extent
for the 2005 vintage. Yet, to maximize the discrimination among the classes and explore
the potential molecular marker candidates in detail, hierarchical cluster analysis readily
revealed G and O subclusters within each vintage (Figure 2A). The analysis readily revealed
two distinct clusters of O and G wines, each including the two vintages. The extensive
chemical dissimilarity between the two main clusters (O and G) is depicted in Figure 2B,C.
The van Krevelen diagrams and ring charts display the most representative molecular
formulas (correlation coefficient r > 0.8) for each cluster. G-related molecular markers were
dominated by sulfur-containing molecules (CHOS and CHONS) with contributions to the
total intensity of assigned molecular formulas of up to 55% and 39%, respectively. On
Metabolites 2022, 12, x FOR PEER REVIEW 3 of 10

Metabolites 2022, 12, x FOR PEER REVIEW 3 of 10

related molecular markers were dominated by sulfur-containing molecules (CHOS and


CHONS) with contributions to the total intensity of assigned molecular formulas of up to
Metabolites 2022, 12, 323 markers were dominated by sulfur-containing molecules (CHOS and CHONS) with3 con- of 10
55% and 39%, respectively. On the other hand, O-related molecular markers were
tributions to the total intensity of assigned molecular formulas of up to 55% and 39%,
characterized by a significantly higher contribution of CHON formulas with 31% of the
respectively. On the other hand, O-related molecular markers were characterized by a
total intensity of assigned molecular formulas (28% for G) and lower contributions of
significantly
the other hand,higher contribution
O-related molecular of CHON
markers formulas
were(27% with 31% ofby
characterized the significantly
total intensityhigher
of as-
sulfur-containing formulas, in particular CHOS and 28% foraCHOS and CHONS
signed molecular
contribution formulas (28% for G) and lower contributions of sulfur-containing for-
respectively).ofThis
CHONresultformulas with 31% of
is in accordance the recent
with total intensity
published of assigned molecular
data reporting the
mulas,
formulas in particular
(28%of CHOS
forCHOS
G) andand (27%
lower and 28% for CHOS and CHONS respectively). This result
consumption thecontributions
formation of of CHON
sulfur-containing
and CHONS formulas, in particular
compounds after
is in accordance
CHOS withfor
recent
(27%electrochemical
and 28% CHOS published
and CHONS datarespectively).
reporting theThisconsumption is inofaccordance
result wine CHOS and the
with
controlled simulation of oxidative reactions in white [13]. Annotated
formation
recent of CHON
published data and CHONSthe
reporting compounds
consumption afterofcontrolled
CHOS electrochemical simulation
elemental formulas from identified discriminant mass peaksand the formation
revealed of CHON
that markers for G
of oxidative
and reactions in after
CHONS compounds whitecontrolled
wine [13].electrochemical
Annotated elemental formulas
simulation from identified
of oxidative reactions
wines are found in the areas of sulfonated polyphenols and amino acids/peptides [23].
discriminant
in white wine mass peaks revealed
[13]. Annotated that markers
elemental formulasfor fromG wines are found
identified in the areas
discriminant mass of sul-
peaks
However, O-related markers appear mostly located in the area of polyphenols (Figure
fonated polyphenols and amino acids/peptides [23]. However, O-related
revealed that markers for G wines are found in the areas of sulfonated polyphenols and markers appear
2B,C).
mostlyacids/peptides
amino located in the area
[23]. of polyphenols
However, (Figure
O-related 2B,C).appear mostly located in the area
markers
of polyphenols (Figure 2B,C).

(04_7)
(04_7)
(04_7)
(04_7)
(04_1)(04_1)
(05_2) (05_2) (04_1) (04_1)
(05_4)
(05_4) (05_2)
(05_2)
(05_4) (05_7)
(04_11) (04_11)
(05_4) (05_7) (05_7)
(04_10)
(05_8) (04_9) (04_11) (04_11)
(04_10)
(05_7) (05_8)
(04_9)
(04_9) (04_10)
(05_8)
(04_9) (04_10)
(05_8)

(05_9)

(05_9)

(05_9)

(05_9)

Figure 1. PCA of FT-ICR-MS data for the 20 bottles of wine from 2004 (04) and 2005 (05) vintages
Figure1.1.PCA
Figure PCAofofFT-ICR-MS
FT-ICR-MS data
data forfor
thethe
20 20 bottles
bottles of wine
of wine fromfrom
20042004 (04) 2005
(04) and and (05)
2005vintages
(05) vintages
from
from different vineyard plots (1; 2; 4; 7; 8; 9; 10; 11), qualified sensory as oxidized (yellow) or good
from different
different vineyard
vineyard plots plots
(1; 2; (1;
4; 7; 2;
8; 4;
9; 7;
10;8; 9;
11), 10; 11), qualified
qualified sensory sensory
as as
oxidizedoxidized
(yellow)(yellow)
or good or good
(blue).
(blue).
(blue).

Figure 2. 2. Characteristic(-)ESI
(-)ESI FT-ICR-MS-derived molecular formulas elaborated by hierarchical
Figure 2.Characteristic
Figure Characteristic (-)ESI FT-ICR-MS-derived
FT-ICR-MS-derived molecular
molecularformulas elaborated
formulas by hierarchical
elaborated cluster
by hierarchical
cluster
analysis. analysis. (A)
(A) Hierarchical Hierarchical cluster
cluster analysis analysis of the assigned (-)ESI FT-ICR-MS-derived
cluster analysis. (A) Hierarchical clusterofanalysis
the assigned
of the(-)ESI FT-ICR-MS-derived
assigned molecular formulas
(-)ESI FT-ICR-MS-derived molec-
observed
ular formulasin wines from 2004
observed (04) from
in wines and 2005
2004(05)
(04)vintages
and 2005and (05)different
vintagesvineyard plots vineyard
and different (1; 2; 4; 7;plots
8; 9;
(1; 11),
10; 2; 4;according
7; 8; 9; 10;to
11), according
their oxidativeto character
their oxidative character
(O: oxidized; G: (O: oxidized;
good). (B) VanG:Krevelen
good). (B) Van Krevelen
diagrams depict
the most representative characteristic molecular formulas for the two main subclusters. Van Krevelen
diagrams depict the most representative characteristic molecular formulas for the two main sub-
clusters. Van Krevelen plots were colored according to molecular classes, i.e., CHO (grey), CHON
(orange), CHOS (green), CHONS (red). The bubble area depicts the relative mass peak intensity
within the respective subcluster. (C) Ring charts displaying the allocation of molecular classes.
Metabolites 2022, 12, 323 4 of 10

The van Krevelen diagrams (H/C versus O/C) were generally used to identify the
structural propertiesplots
of were
molecular markers
colored according in FT-ICR-MS
to molecular research,
classes, i.e., CHO as only
(grey), CHON theCHOS
(orange), molecular
(green),
formula was given. CHONS
Upon(red). The bubble area depicts the relative mass peak intensity within the respective subclus-
comparing the ratios of H/C and O/C for CHOS- and CHONS-
ter. (C) Ring charts displaying the allocation of molecular classes.
containing compounds, products on the right-hand side of the van Krevelen diagram with
higher O/C ratios, G samples allowed
The van Krevelen the highest
diagrams number
(H/C versus O/C) of were oxygen
generally atoms
used toto support
identify the
structural properties of molecular markers in FT-ICR-MS
sulfonation (-SO3H) and nitration (-NO3H) reactions, resulting good markers (Figure 2). research, as only the molecular
formula was given. Upon comparing the ratios of H/C and O/C for CHOS- and CHONS-
In addition to van Krevelen diagrams,products
containing compounds, plotting on elemental
the right-handcontributions as bardiagram
side of the van Krevelen charts with
also
provided useful overview
higher O/C analyses. In particular,
ratios, G samples allowed the comparing
highest number the ofoxygen class todistribu-
oxygen atoms support
sulfonation (-SO H) and nitration (-NO H) reactions,
tions of G and O groups of molecular markers gave useful additional information. Figure
3 3 resulting good markers (Figure 2).
In addition to van Krevelen diagrams, plotting elemental contributions as bar charts also
3A shows that the number of CHO
provided useful formulas
overview analyses.bearing 6 to
In particular, 11 O atoms
comparing increased
the oxygen sharply
class distributions
with oxidation. However,
of G andthe number
O groups of molecular
of molecular markers formulae containing
gave useful additional up to 2 atoms
information. Figure 3A of
shows that the number of CHO formulas bearing 6 to 11
sulfur (S) was higher in G samples (Figure 3B). Moreover, in their majority, S-containing O atoms increased sharply with
oxidation. However, the number of molecular formulae containing up to 2 atoms of
molecular features possessing
sulfur (S) wasbetween
higher in G4samples
to ~12(Figure
O atoms appearedinto
3B). Moreover, bemajority,
their favored for the G
S-containing
state of aged wines,molecular
indicating thepossessing
features relatively higher
between 4 to importance
~12 O atoms appeared of sulfonation
to be favored reactions
for the G
state of aged wines, indicating the relatively higher importance
as opposed to dimerization reactions, which is linked to wine’s oxidative stability during of sulfonation reactions
as opposed to dimerization reactions, which is linked to wine’s oxidative stability during
aging (Figure 3C). aging (Figure 3C).

A 6
B 250
G G
Number of molecular formulae

Number of molecular formulae

5 O O
200
4
150
3
100
2

1 50

0 0
1 2 3 4 5 6 7 8 9 10 11 12 1 2 3 4
Number of O in CHO Number of S in CHOS and CHONS

C
45
G
Number of molecular formulae

40
O
35
30
25
20
15
10
5
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14
Number of O in CHOS and CHONS

Figure 3. Counts of elemental formulas for both good (G) and oxidized (O) wines, with increasing
Figure 3. Counts of elemental formulas for both good (G) and oxidized (O) wines, with increasing
numbers of O atoms in CHO (A), with increasing numbers of S atoms in CHOS and CHONS (B), and
numbers of O atoms in CHO
with (A),numbers
increasing with increasing
of O atoms innumbers
CHOS and of S atoms
CHONS (C). in CHOS and CHONS (B),
and with increasing numbers of O atoms in CHOS and CHONS (C).
2.2. Compositional Networks for Metabolite Identification
In order to give more insights into the chemical mechanisms implicated in wine’s
2.2. Compositional Networks for Metabolite Identification
oxidative stability during bottle aging, an alternative approach exploiting the exact mass
information
In order to give provided into
more insights by FT-ICR-MS through the
the chemical building of compositional
mechanisms implicated networks was
in wine’s
applied in the present study. Compositional networks enable mass coverage beyond clas-
oxidative stability during bottle aging, an alternative approach exploiting the exact mass
sical annotations and thus the identification of unknowns. Exact mass from FT-ICR-MS
information provided by FT-ICR-MS through the building of compositional networks was
applied in the present study. Compositional networks enable mass coverage beyond clas-
sical annotations and thus the identification of unknowns. Exact mass from FT-ICR-MS
data thus provides the chemical basis on which we can construct sample-related metabolic
Metabolites 2022, 12, 323 5 of 10

data thus provides the chemical basis on which we can construct sample-related metabolic
pathways in a data-driven network-based approach. By displaying data points as nodes
that conditionally become connected (edges) to each other, networks have been proven
ideal for efficient modeling and visualization of multivariate data [24]. The total network
of G markers appeared clustered according to chemical families, where CHONS-containing
compounds (in red) and CHOS (in green) were present in condensed areas of the network
(Figure 4). However, CHO- (in blue) and CHON- (in orange) containing compounds
were distributed across the network. Using mass difference network analysis, it became
possible to make assumptions about the structure of yet-unknown markers, depending
on the identity of their m/z connectivity. For the present study, connections (edges) were
referred to mass differences specific to nucleophilic addition reactions of some chemical
families, including known phenols, sulfites (HSO3 − ), carbonyls, and amino acids—peptide
compounds with unknowns in G-related molecular markers. In the presented composi-
tional network, edges are colored according to the chemical families mentioned above. The
clustering of G markers shows clearly that CHONS-containing compounds resulted from
the sulfonation reaction essentially of amino acids and peptides, while CHOS-containing
compounds are the reaction products of essentially polyphenol and carbonyl sulfonation
reactions. Among connected masses, 172 were connected by targeted grape- and oak wood-
derived polyphenols, 504 were connected to carbonyls, 310 were connected to sulfites, and
145 were connected to amino acid and peptides. In all targeted groups of chemical families,
some of the selected masses exhibited very high connectivity (>10), thus suggesting its key
role in wine’s oxidative stability during bottle aging. In detail, among targeted polyphenols,
globally non-carboxyl monophenols, tyrosol, and hydroxytyrosol (formed during yeast
fermentation from tyrosine), followed by main grape-derived phenolic acids (caffeic, ferulic,
and coumaric acid) showed higher connectivity than flavanols (catechin and epicatechin).
Sulfonation reactions occur under wine’s acidic conditions from the very beginning of
winemaking until/during bottle aging. In wine matrices (pH = 3.0–4.0), sulfites are present
in the form of bisulfite ions (HSO3 − ), which possibly react with wine-relevant nucleophiles,
mainly polyphenols, amino acids, sugars, keto acids, and other unknown compounds. The
reactivity/behavior of HSO3 − changes according to the metabolic environment, which
changes constantly during aging. The few known sulfite binders are just the tip of the
iceberg, and their presence can only partially explain the variability related to the binding
strength of sulfites on different wines [25]. Indeed, sulfonation reactions, even for com-
pounds known as strong binders such as acetaldehyde, can be reversible [26], indicating
the potential release of free sulfur dioxide during bottle aging. This result highlights the
importance of identifying the hitherto-unknown and minor binders in order to better
estimate the antioxidant capacity of bound sulfur dioxide. In the present study, 310 links
were found between annotated peaks from global wine FT-ICR-MS data on the basis of
sulfonation reaction transformations. The mass difference corresponds exactly to the mass
difference between the initial targeted compound and the sulfonation product.
Further investigations were conducted to putatively identify the compounds corre-
sponding to sulfonation products. The mass differences (∆ppm) between measured and
exact theoretical masses were calculated to support the compound identification. A total
of 6 compounds corresponding to 164 nodes from the network were putatively identified
as known analogues of polyphenols, carbonyls, and amino acids—peptide compounds
with a very good error (∆ppm < 1). Indeed, the unidentified masses at m/z 217.0175;
m/z 273.0074; m/z 369.0284; m/z 124.9914; m/z 305.0812; and m/z 386.0332 assigned to
[C8 H10 O5 S]− ; [C10 H10 O7 S]− ; [C15 H14 O9 S]− ; [C2 H6 O4 S]− ; [C2 H6 O4 S]− ; [C11 H18 O6 N2 S]− ;
and [C10 H17 O9 N3 S2 ]− formulae could be identified as tyrosol, ferulic acid, catechin, ac-
etaldehyde, γ-glutamyl-cysteine, and glutathione sulfonation products, respectively. In
general, the sulfonation products of the above-mentioned compounds were quantified in all
analyzed wines from 2004 and 2005 vintages, and the box plots of their relative abundances
are reported in Figure 5. All 6 sulfonation products were detected in all wines, presenting
relative abundances in a vintage-dependent manner. It should be pointed out that tyrosol,
compounds with unknowns in G-related molecular markers. In the presented composi-
tional network, edges are colored according to the chemical families mentioned above.
The clustering of G markers shows clearly that CHONS-containing compounds resulted
from the sulfonation reaction essentially of amino acids and peptides, while CHOS-con-
taining
Metabolites 2022, 12, 323 compounds are the reaction products of essentially polyphenol and carbonyl sul- 6 of 10
fonation reactions. Among connected masses, 172 were connected by targeted grape- and
oak wood-derived polyphenols, 504 were connected to carbonyls, 310 were connected to
sulfites, and 145 were connected
acetaldehyde, to amino
and catechin acidwere
sulfones andsignificantly
peptides. more
In allabundant
targetedingroups of the
wines from
chemical families, some
2005 of the
vintage, selected
whereas massesglutathione
sulfonated exhibitedwas very high
more connectivity
abundant (>10),
in wines fromthus
the 2004
vintage. No significant difference was observed for ferulic acid and
suggesting its key role in wine’s oxidative stability during bottle aging. In detail, among γ-glutamyl-cysteine
derivatives. Considering the quantitative ratio among analyzed vintages between the
targeted polyphenols, globally non-carboxyl monophenols, tyrosol, and hydroxytyrosol
sulfonated polyphenols (tyrosol and catechin) and the glutathione, we can formulate the
(formed duringhypothesis
yeast fermentation from tyrosine), followed by main grape-derived phe-
that sulfonation reactions are highly dependent on the chemical environment as
nolic acids (caffeic,
well ferulic, and coumaric acid)
as other physicochemical showed
parameters (i.e.,higher
oxygenconnectivity than
intake, wine pH) flavanols
that occur during
(catechin and epicatechin).
bottle aging.

Edges Nodes
Carbonyls CHO
Polyphenols CHOS
Sulfites CHONS
Amino acids - peptides CHON

Figure 4. Full compositional network built from the CHONS chemical space of G wines, with nodes
colored according to molecular classes, i.e., CHO (blue), CHOS (green), CHONS (red), and CHON
(orange), and edges colored according to mass differences specific to nucleophilic addition reactions of
known phenolic, sulfites (HSO3 − ), carbonyls, and amino acids—peptide compounds with unknowns
in G-related molecular markers.
es 2022, 12, x FOR PEER
Metabolites 2022, REVIEW
12, 323 7 of 10
7 of 10

4,00E+08
4.0 x 108
3,50E+08
3.5 x 108 d
Abudance relative (a.u.)

3.0 x 108
3,00E+08

2,50E+08
2.5 x 108

2.0 x 108
2,00E+08 g b
1.5 x 108
1,50E+08 c
1.0 x 108
1,00E+08 a
5.0 x 107 a eh e a f e
5,00E+07 h
0.0 x100
0,00E+00
1 1
2004 2005

Tyrosol sulfonate Ferulic acid 4-sulfonate


Catechin sulfonate Acetaldehyde sulfonate
S-sulfonate-γ-glutamyl-cysteine S-sulfonate glutathione

Figure 5. BoxFigure
plots of
5. the
Boxrelative
plots ofabundances
the relative of putativelyofidentified
abundances putativelysulfonated analogues analogues
identified sulfonated of tyro- of
sol, ferulic acid, catechin,
tyrosol, ferulicacetaldehyde,
acid, catechin, γ-glutamyl-cysteine, and glutathione
acetaldehyde, γ-glutamyl-cysteine, related to the
and glutathione good
related to the
oxidative state (G) oxidative
good of aged wines from
state (G) 2004wines
of aged and 2005
fromvintages. Means
2004 and 2005 not sharing
vintages. Meansanot
letter are signif-
sharing a letter are
icant (p < 0.05).
significant (p < 0.05).

3. Materials and Methods


3. Materials and Methods
3.1. Chemicals
3.1. Chemicals MS-grade methanol was obtained from Fisher Scientific (Loughborough, UK) and ultra-
MS-grade
puremethanol
water (18.2was obtained
MΩ·cm) from Fisher
was obtained Scientific
from a Milli-Q (Loughborough,
system UK)Germany).
(Merck; Darmstadt, and
ultrapure water (18.2 MΩ·cm) was obtained from a Milli-Q system (Merck; Darmstadt,
3.2. Experimental Design
Germany).
The experimental design included 20 chardonnay wines from five different vineyard
plots located in Burgundy—North East France, from vintages of 2004 and 2005, produced
3.2. Experimental Design
according to standard winemaking procedures (grapes were destemmed and crushed
The experimental
before beingdesign
put intoincluded 20 chardonnay
tanks; alcoholic wines
fermentation from five
was brought to different
dryness andvineyard
followed by
plots locatedmalolactic
in Burgundy—North Eastaged
fermentation) and France,
in oakfrom
woodvintages of 2004
barrels for and 2005,
nine months produced
at the same winery.
according toWines
standard
werewinemaking procedures
bottled after finning using(grapes
natural were destemmed
cork and and conditions.
aged in cellar crushed be-They
were
fore being put all tanks;
into analyzed in June fermentation
alcoholic 2015. was brought to dryness and followed by
malolactic fermentation) and aged in oak wood barrels for nine months at the same win-
3.3. Sensory Analysis
ery. Wines were bottled after finning using natural cork and aged in cellar conditions.
A panel of six trained judges performed sensory analyses of the wines. All panelists
They were all analyzed in June 2015.
had extensive experience in wine tasting and participated regularly in sensory panels with
chardonnay wines. All of the assessments were performed at room temperature (18 ± 1 ◦ C)
3.3. Sensory Analysis
in individual booths under daylight lighting. Fifty milliliters of wine were poured in
A panelstandard
of six trained
ISO 3591judges performed
XL5-type tasting sensory analyses
glasses with glass of the wines.
covers All by
identified panelists
three-digit
random codes and assessed within 15 min of pouring. Each wine was
had extensive experience in wine tasting and participated regularly in sensory panels with submitted to the
chardonnay wines. All of the assessments were performed at room temperature (18 ± 1 or
panelists just after the bottle was opened. They were asked to evaluate the presence
absence of oxidation flavors (G: no oxidative character; O: oxidative character recognized).
°C) in individual booths under daylight lighting. Fifty milliliters of wine were poured in
standard ISO 3591
3.4. XL5-type
Fourier tasting
Transform glassesResonance
Ion Cyclotron with glass covers
Mass identified
Spectrometry by three-digit
Analysis ran-
and Data Processing
dom codes and assessed
Analyses within 15 minusing
were realized of pouring. Each
a Bruker wineIon
SolariX was submitted
Cyclotron to the pan-
Resonance Fourier
elists just after the bottle was opened. They were asked to evaluate the presence
Transform Mass Spectrometer (FT-ICR-MS) (BrukerDaltonics GmbH; Bremen, or absence
Germany)
of oxidationequipped
flavors (G: noaoxidative
with character; O: oxidative
12 Tesla superconducting character
magnet (Magnex recognized).
Scientific Inc.; Yarnton, GB,
USA) and an APOLO II ESI source (BrukerDaltonics GmbH; Bremen, Germany) in nega-
tive ion mode.
3.4. Fourier Transform Samples were
Ion Cyclotron diluted
Resonance at 1/20
Mass into methanol
Spectrometry and and
Analysis injected
Dataat flow rate of
Processing
Analyses were realized using a Bruker SolariX Ion Cyclotron Resonance Fourier
Transform Mass Spectrometer (FT-ICR-MS) (BrukerDaltonics GmbH; Bremen, Germany)
equipped with a 12 Tesla superconducting magnet (Magnex Scientific Inc.; Yarnton, GB,
Metabolites 2022, 12, 323 8 of 10

120 µL/h into the electrospray. The broadband spectra were acquired between 147 m/z
and 1000 m/z mass range in 2 MW FIDs and summed over 300 scans. All samples were
analyzed randomly. The MS was first calibrated using arginine ion clusters (57 nmol/mL
in methanol). Next, raw spectra were further internally calibrated using a reference list
including known wine markers and ubiquitous fatty acids to achieve the best possible mass
accuracy and precision among the samples [4].

3.5. Data Mining


Evaluation of statistical significance was conducted by one-way analysis of variance
(ANOVA) followed, if significant (p < 0.05), by a posteriori Duncan test. Differences
between groups were considered significant when p < 0.05. Principal Component Analysis
(PCA) and Hierarchical Cluster Analysis (HCA) were obtained with Perseus software
(v.1.5.1.6 http://www.perseus-framework.org, accessed on 1 January 2022, Max Planck
Institute of Biochemistry, Germany) [27]. The clustering was performed using a Pearson’s
correlation. FT-ICR-MS data were analyzed with DataAnalysis (v.4.3, Bruker Daltonik
GmbH; Mannheim, Germany). Features (m/z peaks) were filtered according to S/N higher
than 10 and an absolute intensity above 2 × 106 . Features alignment was made with the
Matrix Generator software (v.0.4, Helmholtz Zentrum Muenchen) with a mass accuracy
window of 1 ppm and filtered to keep features present in 90% of samples. Elementary
formula assignment of features was realized with in-house software NetCalc 2015 (v.1.1a,
Helmholtz Zentrum Muenchen) [24]. A visual compounds–targets network was built by
Cytoscape Version 3.9.0. Annotation of the correlated features were realized using online
databases KEGG (KEGG: Kyoto Encyclopedia of Genes and Genomes, n.d.) and Metlin
(Metlin, n.d.), and the online tool Oligonet [28].

4. Conclusions
In the present study, (-)ESI FT-ICR-MS-based metabolomics were combined with
sensory analysis to characterize the chemical diversity of compounds related to white
wine’s oxidative stability during bottle aging. Molecular markers of wine’s oxidative
stability were first described by a specific CHONS chemical space in accordance with
already published results indicating the increase in this specific molecular fingerprint
during aging. Our findings based on the comparison of the ratios of H/C and O/C for
CHOS- and CHONS-containing compounds evidenced that, in their majority, S-containing
molecular features possessing O4∼12 number appeared to be related to the good state of
aged wines, indicating the importance of sulfonation reactions instead of dimerization
reactions on wine’s oxidative stability during aging. Moreover, molecular networking
provided a more complete overview of the sulfonation molecular targets during wine aging.
Indeed, based on sulfonation reaction transformations, we could report the importance of
hitherto unknown and/or minor sulfur dioxide binders on wine’s oxidative stability.

Author Contributions: M.N., R.G. and P.S.-K. conceived the presented idea; M.N., R.R. and M.L.
developed the methodology and performed the experimentations; M.N. wrote, reviewed, and edited
the manuscript with support from R.G., M.L. and P.S.-K.; M.L., R.R. and M.N. verified the analytical
methods and proceeded to data acquisition. All authors discussed the results and contributed to the
final manuscript. All authors have read and agreed to the published version of the manuscript.
Funding: This work was supported by the Regional Council of Bourgogne-Franche-Comté, and the
“Fonds Européen de Developpement Régional (FEDER)”.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available in article.
Conflicts of Interest: The authors declare no conflict of interest.
Metabolites 2022, 12, 323 9 of 10

References
1. Lopes, P.; Silva, M.A.; Pons, A.; Tominaga, T.; Lavigne, V.; Saucier, C.; Darriet, P.; Teissedre, P.L.; Dubourdieu, D. Impact of oxygen
dissolved at bottling and transmitted through closures on the composition and sensory properties of a Sauvignon blanc wine
during bottle storage. J. Agric. Food Chem. 2009, 57, 10261–10270. [CrossRef] [PubMed]
2. Waterhouse, A.L.; Laurie, V.F. Oxidation of wine phenolics: A critical evaluation and hypotheses. Am. J. Enol. Vitic. 2006, 57,
306–313.
3. Ugliano, M.; Kwiatkowski, M.; Travis, B.; Francis, I.L.; Waters, E.J.; Herderich, M.; Pretorius, I.S. Post-bottling management of
oxygen to reduce off-flavour formation and optimize wine style. Aust. N. Z. Wine Ind. J. 2009, 24, 24–28.
4. Gougeon, R.D.; Lucio, M.; De Boel, A.; Frommberger, M.; Hertkorn, N.; Peyron, D.; Chassagne, D.; Feuillat, F.; Cayot, P.;
Voilley, A.; et al. Expressing Forest Origins in the Chemical Composition of Cooperage Oak Woods and Corresponding Wines by
Using FTICR-MS. Chem. Eur. J. 2009, 15, 600–611. [CrossRef] [PubMed]
5. Arbulu, M.; Sampedro, M.C.; Gómez-Caballero, A.; Goicolea, M.A.; Barrio, R.J. Untargeted metabolomic analysis using liquid
chromatography quadrupole time-of-flight mass spectrometry for non-volatile profiling of wines. Anal. Chim. Acta 2015, 858,
32–41. [CrossRef] [PubMed]
6. Cuadros-Inostroza, A.; Giavalisco, P.; Hummel, J.; Eckardt, A.; Willmitzer, L.; Peña-Cortés, H. Discrimination of Wine Attributes
by Metabolome Analysis. Anal. Chem. 2010, 82, 3573–3580. [CrossRef]
7. Nikolantonaki, M.; Julien, P.; Coelho, C.; Roullier-Gall, C.; Ballester, J.; Schmitt-Kopplin, P.; Gougeon, R.D. Impact of Glutathione
on Wines Oxidative Stability: A Combined Sensory and Metabolomic Study. Front. Chem. 2018, 6, 182. [CrossRef]
8. Arapitsas, P.; Scholz, M.; Vrhovsek, U.; Di Blasi, S.; Bartolini, A.B.; Masuero, D.; Perenzoni, D.; Rigo, A.; Mattivi, F. A Metabolomic
Approach to the Study of Wine Micro-Oxygenation. PLoS ONE 2012, 7, e37783. [CrossRef]
9. Arapitsas, P.; Speri, G.; Angeli, A.; Perenzoni, D.; Mattivi, F. The influence of storage on the “chemical age” of red wines.
Metabolomics 2014, 10, 816–832. [CrossRef]
10. Roullier-Gall, C.; Witting, M.; Moritz, F.; Gil, R.B.; Goffette, D.; Valade, M.; Schmitt-Kopplin, P.; Gougeon, R.D. Natural
oxygenation of Champagne wine during ageing on lees: A metabolomics picture of hormesis. Food Chem. 2016, 203, 207–215.
[CrossRef]
11. Roullier-Gall, C.; Boutegrabet, L.; Gougeon, R.D.; Schmitt-Kopplin, P. A grape and wine chemodiversity comparison of different
appellations in Burgundy: Vintage vs terroir effects. Food Chem. 2014, 152, 100–107. [CrossRef] [PubMed]
12. Karbowiak, T.; Crouvisier-Urion, K.; Lagorce, A.; Ballester, J.; Geoffroy, A.; Roullier-Gall, C.; Chanut, J.; Gougeon, R.D.;
Schmitt-Kopplin, P.; Bellat, J.-P. Wine aging: A bottleneck story. NPJ Sci. Food 2019, 3, 14. [CrossRef] [PubMed]
13. Roullier-Gall, C.; Kanawati, B.; Hemmler, D.; Druschel, G.K.; Gougeon, R.D.; Schmitt-Kopplin, P. Electrochemical triggering of
the Chardonnay wine metabolome. Food Chem. 2019, 286, 64–70. [CrossRef] [PubMed]
14. Romanet, R.; Coelho, C.; Liu, Y.; Bahut, F.; Ballester, J.; Nikolantonaki, M.; Gougeon, D.R. The Antioxidant Potential of White
Wines Relies on the Chemistry of Sulfur-Containing Compounds: An Optimized DPPH Assay. Molecules 2019, 24, 1353. [CrossRef]
15. Romanet, R.; Sarhane, Z.; Bahut, F.; Uhl, J.; Schmitt-Kopplin, P.; Nikolantonaki, M.; Gougeon, R.D. Exploring the chemical space
of white wine antioxidant capacity: A combined DPPH, EPR and FT-ICR-MS study. Food Chem. 2021, 355, 129566. [CrossRef]
16. Jordão, A.M.; Ricardo-Da-Silva, J.M.; Laureano, O.; Mullen, W.; Crozier, A. Effect of ellagitannins, ellagic acid and
volatile compounds from oak wood on the (+)-catechin, procyanidin B1 and malvidin-3-glucoside content of model
wines. Aust. J. Grape Wine Res. 2008, 14, 260–270. [CrossRef]
17. Nikolantonaki, M.; Waterhouse, A.L. A Method to Quantify Quinone Reaction Rates with Wine Relevant Nucleophiles: A Key to
the Understanding of Oxidative Loss of Varietal Thiols. J. Agric. Food Chem. 2012, 60, 8484–8491. [CrossRef]
18. Bahut, F.; Liu, Y.; Romanet, R.; Coelho, C.; Sieczkowski, N.; Alexandre, H.; Schmitt-Kopplin, P.; Nikolantonaki, M.; Gougeon, R.D.
Metabolic diversity conveyed by the process leading to glutathione accumulation in inactivated dry yeast: A synthetic media
study. Food Res. Int. 2019, 123, 762–770. [CrossRef]
19. Bahut, F.; Romanet, R.; Sieczkowski, N.; Schmitt-Kopplin, P.; Nikolantonaki, M.; Gougeon, R.D. Antioxidant activity from
inactivated yeast: Expanding knowledge beyond the glutathione-related oxidative stability of wine. Food Chem. 2020, 325, 126941.
[CrossRef]
20. Roullier-Gall, C.; Hemmler, D.; Gonsior, M.; Li, Y.; Nikolantonaki, M.; Aron, A.; Coelho, C.; Gougeon, R.D.; Schmitt-Kopplin, P.
Sulfites and the wine metabolome. Food Chem. 2017, 237, 106–113. [CrossRef]
21. Arapitsas, P.; Ugliano, M.; Perenzoni, D.; Angeli, A.; Pangrazzi, P.; Mattivi, F. Wine metabolomics reveals new sulfonated products
in bottled white wines, promoted by small amounts of oxygen. J. Chromatogr. A 2016, 1429, 155–165. [CrossRef] [PubMed]
22. Arapitsas, P.; Guella, G.; Mattivi, F. The impact of SO2 on wine flavanols and indoles in relation to wine style and age. Sci. Rep.
2018, 8, 858. [CrossRef] [PubMed]
23. Gougeon, R.D.; Lucio, M.; Frommberger, M.; Peyron, D.; Chassagne, D.; Alexandre, H.; Feuillat, F.; Voilley, A.; Cayot, P.;
Gebefugi, I.; et al. The chemodiversity of wines can reveal a metabologeography expression of cooperage oak wood.
Proc. Natl. Acad. Sci. USA 2009, 106, 9174–9179. [CrossRef] [PubMed]
24. Tziotis, D.; Hertkorn, N.; Schmitt-Kopplin, P. Kendrick-Analogous Network Visualisation of Ion Cyclotron Resonance Fourier
Transform Mass Spectra: Improved Options for the Assignment of Elemental Compositions and the Classification of Organic
Molecular Complexity. Eur. J. Mass Spectrom. 2011, 17, 415–421. [CrossRef] [PubMed]
Metabolites 2022, 12, 323 10 of 10

25. Jackowetz, J.N.; de Orduña, R.M. Survey of SO2 binding carbonyls in 237 red and white table wines. Food Control 2013, 32,
687–692. [CrossRef]
26. Tachtalidou, S.; Sok, N.; Denat, F.; Noret, L.; Schmit-Kopplin, P.; Nikolantonaki, M.; Gougeon, R.D. Direct NMR evidence for the
dissociation of sulfur-dioxide-bound acetaldehyde under acidic conditions: Impact on wines oxidative stability. Food Chem. 2022,
373, 131679. [CrossRef]
27. Tyanova, S.; Temu, T.; Sinitcyn, P.; Carlson, A.; Hein, M.Y.; Geiger, T.; Mann, M.; Cox, J. The Perseus computational platform for
comprehensive analysis of (prote)omics data. Nat. Methods 2016, 13, 731. [CrossRef]
28. Liu, Y.; Forcisi, S.; Lucio, M.; Harir, M.; Bahut, F.; Deleris-Bou, M.; Krieger-Weber, S.; Gougeon, R.D.; Alexandre, H.; Schmitt-
Kopplin, P. Digging into the low molecular weight peptidome with the OligoNet web server. Sci. Rep. 2017, 7, 11692. [CrossRef]

You might also like