Next Article in Journal
Effects of High Hydrostatic Pressure on Escherichia coli Ultrastructure, Membrane Integrity and Molecular Composition as Assessed by FTIR Spectroscopy and Microscopic Imaging Techniques
Next Article in Special Issue
Characterization of Four Popular Sweet Cherry Cultivars Grown in Greece by Volatile Compound and Physicochemical Data Analysis and Sensory Evaluation
Previous Article in Journal
Insecticidal Activities of Bark, Leaf and Seed Extracts of Zanthoxylum heitzii against the African Malaria Vector Anopheles gambiae
Previous Article in Special Issue
Evolution of the Aroma Volatiles of Pear Fruits Supplemented with Fatty Acid Metabolic Precursors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Applications of Solid-Phase Microextraction and Gas Chromatography/Mass Spectrometry (SPME-GC/MS) in the Study of Grape and Wine Volatile Compounds

by
Annarita Panighel
* and
Riccardo Flamini
Consiglio per la Ricerca e la Sperimentazione in Agricoltura-Centro di Ricerca per la Viticoltura (CRA-VIT), Viale XXVIII aprile 26, Conegliano (TV) 31015, Italy
*
Author to whom correspondence should be addressed.
Molecules 2014, 19(12), 21291-21309; https://doi.org/10.3390/molecules191221291
Submission received: 24 October 2014 / Revised: 4 December 2014 / Accepted: 8 December 2014 / Published: 18 December 2014
(This article belongs to the Special Issue Aromas and Volatiles of Fruits)

Abstract

:
Volatile compounds are responsible for the wine “bouquet”, which is perceived by sniffing the headspace of a glass, and of the aroma component (palate-aroma) of the overall flavor, which is perceived on drinking. Grape aroma compounds are transferred to the wine and undergo minimal alteration during fermentation (e.g., monoterpenes and methoxypyrazines); others are precursors of aroma compounds which form in winemaking and during wine aging (e.g., glycosidically-bound volatile compounds and C13-norisoprenoids). Headspace solid phase microextraction (HS-SPME) is a fast and simple technique which was developed for analysis of volatile compounds. This review describes some SPME methods coupled with gas chromatography/mass spectrometry (GC/MS) used to study the grape and wine volatiles.

1. Introduction

Wine aroma is formed by more than 800 volatile compounds and is characteristic of each product [1,2]. Often these compounds are present in very low concentration and are characterized by very low sensory thresholds (between ng/L and μg/L). Usually, the wine aroma profiling needs a sample preparation for isolation and concentration of the volatiles before performing gas chromatographic analysis. Several sample preparation methods for the analysis of grapes and wine were proposed: distillation [3,4], liquid–liquid extraction (LLE) [5,6], solid phase extraction (SPE) [7,8], dynamic headspace extraction [9], and headspace–solid phase microextraction (HS-SPME) [10,11,12,13].
SPME was developed in the 1990s by Pawliszyn and co-workers [14] and every year a thousand papers describing different aspects of this approach, and applications in different fields (chemical analysis, bioanalysis, food science, environmental science, and recently, pharmaceutical and medical sciences), are published [15]. This sample extraction technique was demonstrated to be rapid, simple, and reproducible, with no solvent use, and is suitable for the extraction and concentration of a high number of volatile and semi-volatile compounds from aqueous solutions [16,17]. Moreover, SPME needs a small sample volume and the coupling with gas chromatography and mass spectrometry (GC/MS) provides high sensitivity. For these reasons it has been used to study the volatile profile of many fruit varieties, vegetables, and beverages, including grapes and wine [18,19,20].
This paper reviews the main SPME-GC/MS applications developed to study the volatile and aroma compounds of grapes and wine.

Grape and Wine Volatile Compounds

Since early 80’s a great number of studies of grape and wine volatiles have been performed and the main compounds identified are listed in Table 1. Principal in grape are monoterpenes, C13-norisoprenoids, benzene compounds, C6 aldehydes, and alcohols [21,22,23]. These compounds are present in berry skin and pulp in both free (volatile) and glycosidically-bound (non-volatile) form. Free volatile compounds directly contribute to grape and wine aroma while glycosides are flavorless compounds which can act as aroma precursors for enzymatic and acid hydrolysis occurring in winemaking and during wine storage [24].
In general, monoterpenes have floral and citrus notes; C13-norisoprenoids such as β-damascenone and β-ionone are characterized by “fruity-flowery”, “honey-like”, “sweet” and “violet” notes, respectively [25].
Main grape and wine volatile benzenoids are aldehydes and alcohols, such as benzaldehyde, phenylacetaldehyde (hyacinth and rose-like odor [26]), vanillin, benzyl alcohol, and 2-phenylethanol [24], probably formed from L-phenylalanine via the shikimic pathway [27]. The structures of the main grape aroma compounds are shown in Figure 1, Figure 2 and Figure 3.
Table 1. Principal aroma compounds identified in grapes and wine [20,23,24,28,29].
Table 1. Principal aroma compounds identified in grapes and wine [20,23,24,28,29].
TerpenoidsBenzenoidsSulfur compounds
linaloolzingeronemethyl mercaptan
nerolzingerolethyl mercaptan
geraniolacetophenonedimethyl sulfide
citronellolvanillin diethyl sulfide
α-terpineolmethyl salicylatedimethyl disulfide
cis/trans ocimenol eugenoldiethyl disulfide
cis/trans linalooloxide (furanic form)cis/trans isoeugenolmethyl thioacetate
cis/trans linalooloxide (pyranic form)2-phenylethanolethyl thioacetate
hydroxycitronellol benzyl alcohol2-mercaptoethanol
8-hydroxydihydrolinalool acetovanillone2-(methylthio)-1-ethanol
7-hydroxygeraniolbenzaldehyde3-(methylthio)-1-propanol
7-hydroxynerol 4-hydroxybenzaldehyde4-(methylthio)-1-butanol
cis/trans 8-hydroxylinalool2,4-dimethylbenzaldehyde2-furanmethanethiol
diendiol I phenylacetaldehydebenzothiazole
endiolsyringaldehydethiazole
diendiol II coniferaldehyde5-(2-hydroxyethyl)-4-methylthiazole
neroloxidesinapaldehyde4-methyl-4-mercaptopentan-2-one
2-exo-hydroxy-1,8-cineolpropriosyringone3-mercaptohexanol acetate
1,8-cineolpropriovanillonecis/trans 2-methylthiophan-3-ol
cis/trans 1,8-terpinesyringol2-methyltetrahydrothiophen-3-one
p-menthenediol Iconiferyl alcoholcis/trans 2-methyltetrahydrothiophen-3-ol
(E)-geranic acidvanillic alcohol3-mercaptohexan-1-ol
(E)-2,6-dimethyl -6-hydroxyocta-2,7-dienoic acidsinapic alcohol3-mercaptohexyl acetate
(E)- and (Z)-sobrerol o-cymene4-mercapto-4-methylpentan-2-ol
cis/trans rose oxidep-cymene3-mercapto-3-methylbutan-1-ol
lilac alcohols guaiacol
triol 4-ethylguaiacol
hotrienol 4-vinylguaiacol
myrcenol 4-ethylphenol
limonene4-vinylphenol
β-phellandrenemethyl anthranilate
β-ocimene2'-aminoacetophenone
wine lactone
Aliphatic alcoholsAcidsSesquiterpenes
1-butanolisobutyric acidrotundone
2-nonanolisovaleric acidfarnesol
3-methyl-1-butanol acetic acidgermacrene D
2-methyl-1-butanol butyric acidγ-cadinene
isobutanolhexanoic acidα-ylangene
1-pentanoloctanoic acidα-farnesene
1-hexanoldecanoic acidβ-farnesene
1-octanolhexadecanoic acidnerolidol
(E)-3-hexen-1-oloctadecanoic acid
(Z)-3-hexen-1-ol
4-methyl-3-penten-1-ol
(E)-2-hexen-1-ol
1-octen-3-ol
2-ethyl-1-hexanol
furfuryl alcohol
6-methyl-5-hepten-2-ol
Carbonyl compoundsEstersNorisoprenoids
acetaldehydeethyl 2-methylpropanoate TDN (1,1,6-trimethyl-1,2-dihydronaphthalene)
isobutyraldehydeethyl 2-methylbutanoateβ-damascone
2-methylbutanalethyl 3-methylbutanoateβ-damascenone
isovaleraldehydeethyl 2-hydroxypropanoatevomifoliol
1-octen-3-oneethyl 3-hydroxybutanoatedihydrovomifoliol
(E)-2-heptenalethyl 4-hydroxybutanoate3-hydroxy-β-damascone
methionaldiethyl succinate3-oxo-α-ionol
(E)-2-octenaldiethyl malate3-hydroxy-7,8-dihydro-β-ionol
hexanalethyl butanoateα-ionol
(E)-2-hexenalethyl hexanoateβ-ionol
(Z)-3-hexenalethyl octanoateα-ionone
(Z)-2-nonenalethyl decanoateβ-ionone
furfuralethyl benzoateactinidols
5-methylfurfuralisoamyl octanoatevitispiranes
1H-pyrrole-2-carboxyaldehydeethyl furoateRiesling acetal
geranialethyl dihydrocinnamatehydroxy-megastigmen-2-one
neralethyl cinnamatehydroxy-megastigmen-3-one
acetoin methyl vanillate4-oxo-isophorone
diacetylethyl vanillateβ-isophorone
glyoxalethyl acetate4-oxo-2,3-dehydro-β-ionol
methylglyoxalisobutyl acetateβ-cyclocitral
glycolaldehydeisoamyl acetate
hydroxypropandialethyl 2-phenylacetate
2,4-nonadienalhexyl acetate
2,6-nonadienal
LactonesNitrogen compounds
γ-butyrolactone3-isobutyl-2-methoxypyrazine
γ-hexalactone3-sec-butyl-2-methoxypyrazine
γ-nonalactone3-isopropyl-2-methoxypyrazine
γ-decalactone3-ethyl-2-methoxypyrazine
cis/trans oak lactone
sotolon
Twenty-four aldehydes (mostly alkyls) were identified in wine; they can form during winemaking or be present in the grapes [28]. For example, glyoxal, methylglyoxal, hydroxypropandial and glycolaldehyde form by microorganisms such as Saccharomyces cerevisiae or Leuconostoc oenos, or can form in grapes as a consequence of Botrytis cinerea grape attack [30,31]. In general, C6 aldehydes (hexanal, (E)-2-hexenal, and (Z)-3-hexenal) and (Z)-2-nonenal are responsible for the green, herbaceous, and sometimes bitter aroma of wines [21]. They are mainly due to the enzymatic cleavage of oxidized linoleic and linolenic acid during grape crushing before fermentation [29].
Figure 1. Principal monoterpenes identified in grapes. In brackets, the odor descriptor is reported [24].
Figure 1. Principal monoterpenes identified in grapes. In brackets, the odor descriptor is reported [24].
Molecules 19 21291 g001
Figure 2. Principal C13-norisoprenoids identified in grapes. In brackets, the odor descriptor is reported [24].
Figure 2. Principal C13-norisoprenoids identified in grapes. In brackets, the odor descriptor is reported [24].
Molecules 19 21291 g002
Figure 3. Structures of “foxy-smelling” compounds and alkyl methoxypyrazines in grapes. In brackets, the odor descriptor is reported [25].
Figure 3. Structures of “foxy-smelling” compounds and alkyl methoxypyrazines in grapes. In brackets, the odor descriptor is reported [25].
Molecules 19 21291 g003
Main wine volatiles are ethyl esters, acetates, and higher alcohols. Esters are produced by the yeasts during fermentation. Principal are ethyl esters characterized by fruity and floral notes, such as ethyl hexanoate, ethyl octanoate, ethyl decanoate, ethyl dodecanoate, isoamyl acetate, hexyl acetate, and 2-phenylethyl acetate [32,33]. Contents of hexanoic, octanoic, and decanoic acid in wine depend on the yeast strain, fermentation conditions and grape must composition [27]. Higher alcohols are formed by the yeast sugar metabolism (anabolic pathway) as well as via the catabolic or Ehrlich pathway of amino acids [27,34,35]. Rapp and Versini reported that a higher alcohols concentration below 300 mg/L is desirable for the aroma complexity of wine whereas a concentration exceeding 400 mg/L can have a detrimental effect [36].
Volatile phenols, such as 4-vinylphenol (aroma descriptors: spicy, pharmaceutical), 4-vinylguaiacol (smoke, phenolic), 4-ethylphenol (horse stable, medicinal), and 4-ethyl guaiacol (spice, phenolic), are produced during alcoholic fermentation by some microorganisms (Brettanomyces yeasts and bacteria) by decarboxylation of hydroxycinnamic acids present in must [32,33,37].

2. SPME-GC/MS Methods

2.1. Analysis of PFBOA-Derivatives

In general, carbonyl compounds contribute to the wine aroma even though they are present in low levels. GC/MS analysis of the O-pentafluorobenzyl (PFB) derivatives formed by reaction with O-(2,3,4,5,6-pentafluorobenzyl)-hydroxylamine (PFBOA) by recording in singular-ion-monitoring (SIM) the mass spectrum base peak signal at m/z 181 (characteristic of PFB-oximes) is a selective and sensitive method. On the other hand, derivatization increases the complexity of analyte peaks in the chromatogram due to formation of two oximes for each carbonyl group (isomers E and Z, except for formaldehyde) [38]. By this method several studies of carbonyl compounds in hydro-alcoholic matrices, such as wine, model wine solutions, and spirits, were performed [30,31,39,40,41,42,43,44,45,46,47].
Malolactic fermentation (MLF) is an important oenological process performed after the alcoholic fermentation for improving the organoleptic characteristics and the microbiological stability of wine [46]. The process is carried out by lactic acid bacteria: it can occur naturally or be induced by the inoculum of commercial bacteria strains. Conversion of L(−)-malic into L(+)-lactic acid decreases wine acidity [30]; usually, the inoculation of selected bacteria strains enables control over the process [48].
With MLF, together with the other fermentative compounds (esters, sulfur and nitrogen compounds, volatile phenols, and volatile fatty acids), also the carbonyl profile changes, increasing the aromatic complexity of wine [42,49].
HS-SPME and GC/MS analysis of PFBOA derivatives was performed to study the carbonyl compounds in Merlot wines after MLF [46]. In general, PFBOA derivatives are characterized by higher volatility, thermal stability, and affinity for the SPME fiber with respect to the corresponding carbonyl compounds [20]. SPME was performed by using a polydimethylsiloxane/divinylbenzene (PDMS/DVB) fiber; GC/MS analysis was performed by electron impact ionization (EI) and chemical ionization in both positive (PICI) and negative (NCI) mode [50]. NCI coupled to SPME provided lower detection limits for isobutyraldehyde, 2-methylbutanal, isovaleraldehyde, (E)-2-hexenal, 1-octen-3-one, (E)-2-heptenal, methional, (E)-2-octenal, phenylacetaldehyde, and (E)-2-nonenal [50]. PICI was used for determination of the principal carbonyl compounds in wine, such as acetaldehyde, diacetyl, and acetoin [46]. By using methane as reagent gas, abundant formation of acetaldehyde derivative [M+H]+ ion at m/z 240, diacetyl mono-derivative [M+H]+ ion at m/z 282, and o-chlorobenzaldehyde derivative [M+H]+ ion at m/z 336 (the internal standard), was observed. For acetoin abundant formation of [M+H-18]+ ion at m/z 266 was observed. These signals were used for the quantification of compounds. Experimental PICI conditions used are reported in Table 2 [46].
Table 2. Experimental conditions used for PFBOA derivatization and SPME-GC/MS analysis (ion trap and positive chemical ionization) of the main wine carbonyl compounds. Adapted from Flamini et al., 2005 [46].
Table 2. Experimental conditions used for PFBOA derivatization and SPME-GC/MS analysis (ion trap and positive chemical ionization) of the main wine carbonyl compounds. Adapted from Flamini et al., 2005 [46].
Sample volume100 μL
Vial volume4 mL
Derivatization conditions200 μL IS o-chlorobenzaldehyde, 3.4 mg/L in ethanol/water solution; 1 mL of PFBOA 2 g/L aqueous solution, volume adjusted to 2 mL with water
SPME fiber65-μm poly(ethylene glycol)/divinylbenzene (PEG/DVB)
Addition to the sample50 mg NaCl
Sample heating 50 °C for 20 min under stirring
Extraction temperature and time 50 °C for 5 min
Desorption temperature and time240 °C for 1 min
Fiber cleaning250 °C for 5 min
GC columnHP-INNowax (30 m × 0.25 mm i.d; 0.25-μm film thickness)
Carrier gasHelium, column headpressure 16 psi
InjectorT = 240 °C, sample volume 0.5 μL, splitless injection
Oven program60 °C for 5 min, 3 °C/min to 210 °C, held 5 min
MS-IT conditionsPICI mode using methane as reagent gas (flow 1 mL/min), ion source at 200 °C, damping gas 0.3 mL/min, simultaneous SCAN (range m/z 40–660, 1.67 scan/s) and MS/MS
CID experimentsCollision gas He, excitation voltage 225 mV
QuantitativeRecorded signals at m/z 240 for acetaldehyde, m/z 282 for diacetyl, m/z 266 for acetoin, m/z 336 for o-chlorobenzaldehyde (I.S.)

2.2. Aroma Compounds and Wine Aging

Aging in wooden barrels is a process used to stabilize the color and to improve limpidity and the sensorial characteristics of wines. During aging many compounds are transferred from the wood to the wine: polyphenols, lactones, coumarins, polysaccharides, hydrocarbons and fatty acids, terpenes, C13-norisoprenoids, steroids, carotenoids, and furan compounds. Wood volatiles such as cis- and trans-β-methyl-γ-octalactones (oak lactones), syringaldehyde, vanillin, coniferaldehyde, sinapaldehyde and their alcohols, propiosyringone and propiovanillone, hydroxy-megastigmen-2-one and hydroxy-megastigmen-3-one [51], furfural, 5-methyl furfural, guaiacol, eugenol, 4-ethylphenol and 4-ethyl guaiacol [52,53], furfuryl alcohol, β-ionone, γ-noanalactone, and acetovanillone [54] confer the typical organoleptic characteristics of aged wines.
In making barrels for wine aging, oak (Quercus sessilis, Q. petraea, Q. robur, Q. peduncolata, Q. alba) is the wood more often used but other species, such as acacia (Robinia pseudoacacia), chestnut (Castanea sativa), cherry (Prunus avium), and mulberry (Morus alba and Morus nigra) are also being considered [55].
SPME-GC/MS was used to study the evolution of wine aroma during aging in 225-L barrels (barriques) made with these wood types. Experimental conditions used are reported in Table 3; main compounds identified are reported in Table 4 [55].
Table 3. SPME-GC/MS conditions used to study the evolution of volatile compounds of Raboso Piave wine during aging in five different types of wood barrels [55].
Table 3. SPME-GC/MS conditions used to study the evolution of volatile compounds of Raboso Piave wine during aging in five different types of wood barrels [55].
SPME fiber65-μm carbowax/divinylbenzene (CAR/DVB)
Sample volume10 mL
Vial volume20 mL
Addition to the sample3 g NaCl
Sample heating 70 °C for 10 min
Extraction temperature and time 70 °C for 30 min
Desorption temperature and time230 °C fo 5 min
Fiber cleaning10 min
GC columnHP-INNowax (30 m × 0.25 mm i.d; 0.25 μm film thickness)
Injection Splitless
Oven program40 °C for 5 min, 3 °C/min to 230 °C, held 10 min
MS conditionsionization energy 70 eV, acquisition SIM mode
Wines aged in acacia, chestnut and oak wood showed higher contents of vanillin and eugenol and the acacia-aged sample showed an increase of 4-ethylguaiacol. Mulberry-aged wine had a significant decrease of 4-ethylguaiacol and increase of 4-ethylphenol; the wine aged in cherry barrel already showed high levels of 4-ethylguaiacol after three months of aging.
Table 4. Principal compounds studied in Raboso Piave wines aged nine months in 225 L barrels of acacia, cherry, chestnut, mulberry, and oak (nd: not detected; tr: trace (<0.01 ppm) [55]).
Table 4. Principal compounds studied in Raboso Piave wines aged nine months in 225 L barrels of acacia, cherry, chestnut, mulberry, and oak (nd: not detected; tr: trace (<0.01 ppm) [55]).
BarrelMonths of AgingCompounds mg/L
Furfural5-Methylfurfural4-EthylguaiacolEugenol4-EthylphenolVanillin
Acacia30.02 ± 0.010.03 ± 0.012.24 ± 0.210.009 ± 0.0010.67 ± 0.070.09 ± 0.03
60.04 ± 0.010.03 ± 0.012.94 ± 0.140.015 ± 0.0010.92 ± 0.080.16 ± 0.01
90.03 ± 0.010.03 ± 0.013.25 ± 0.670.021 ± 0.0051.29 ± 0.410.31 ± 0.07
Cherry3Ndnd3.01 ± 1.130.008 ± 0.0041.00 ± 0.440.08 ± 0.04
6Trnd3.13 ± 0.260.009 ± 0.0011.04 ± 0.060.10 ± 0.01
9Ndnd2.79 ± 0.510.007 ± 0.0010.86 ± 0.180.12 ± 0.03
Chestnut30.04 ± 0.020.03 ± 0.012.53 ± 0.430.024 ± 0.0040.84 ± 0.200.45 ± 0.06
60.04 ± 0.010.02 ± 0.022.30 ± 0.120.035 ± 0.0030.74 ± 0.080.60 ± 0.02
90.07 ± 0.010.04 ± 0.011.84 ± 0.180.026 ± 0.0020.64 ± 0.040.43 ± 0.03
Mulberry3Trnd2.69 ± 0.750.004 ± 0.0011.06 ± 0.260.09 ± 0.03
6Trnd2.72 ± 0.440.006 ± 0.0011.27 ± 0.260.08 ± 0.02
9Trtr1.84 ± 0.200.006 ± 0.0011.19 ± 0.070.08 ± 0.01
Oak30.18 ± 0.080.14 ± 0.042.51 ± 0.140.009 ± 0.0010.90 ± 0.070.27 ± 0.04
60.56 ± 0.160.19 ± 0.052.08 ± 0.020.012 ± 0.0030.75 ± 0.050.34 ± 0.08
90.60 ± 0.060.32 ± 0.042.90 ± 0.750.018 ± 0.0051.06 ± 0.360.36 ± 0.09

2.3. “Foxy Smelling Compounds” and 3-Alkyl-2-Methoxypyrazines in Grape Juice

2'-Aminoacetophenone (o-AAP) is the main compound identified as the cause of the aging note—the so-called “hybrid note”, “foxy-smelling” or “American character”—typical of V. labruscana grapes, even though it was also found in some V. vinifera wines such as Müller-Thurgau, Riesling, and Silvaner [56]. This note is variously described as “acacia blossom,” “naphthalene note,” “furniture polish,” “fusel alcohol,” and “damp cloth,” and causes a considerable number of wine rejections. The formation of o-AAP in grape is promoted by several factors, such as reduced nitrogen fertilization in combination with hot and dry summers, and the risk increases in wines made with grapes harvested early. The phytohormone indole-3-acetic acid (IAA) is the principal precursor of o-AAP through non-enzymatic processes [57,58]. Also, methyl anthranilate (MA) contributes to the typical foxy taint of wines made with American and wild vine grapes, although it was also found in some V. vinifera white wines in concentrations of up to 0.3 μg/L [59].
For analysis of o-AAP in wine a direct-immersion SPME method by using a DVB/CAR/PDMS fiber and GC/MS, was proposed [60]; instead, analysis of MA in grape juice was performed by using a PDMS fiber [61].
3-Alkyl-2-methoxypyrazines are present in the grape skin, pulp, and bunch stems. These compounds contribute to the aroma of wines by conferring vegetative, herbaceous, bell pepper, or earthy notes. They are characterized by very low sensory thresholds: for 3-isobutyl-2-methoxypyrazine (IBMP), 3-sec-butyl-2-methoxypyrazine (SBMP),and 3-isopropyl-2-methoxypyrazine (IPMP) are between 1 and 2 ng/L in water. The level of IBMP in wine may be 10 times the sensory threshold, whereas SBMP and IPMP are normally comparable to their sensory thresholds [62,63,64,65,66,67,68].
In general, the climate influences the biosynthesis of methoxypyrazines, and higher contents were found in grapes from cooler regions. For example, it was observed that 3-isobutyl-2-methoxypyrazine decreases dramatically during ripening. Methoxypyrazines may be also influenced by the light exposure: in general berries exposed to more sunlight have lower contents [67].
SPME-GC/MS of 3-alkyl-2-methoxypyrazines in grape juice and wine was performed by using DVB/CAR/PDMS, PDMS/DVB and CAR/PDMS fibers [69,70,71,72]. Also, a SPME-GC/MS and multiple mass spectrometry (MS/MS) method for simultaneous determination of o-AAP, MA, and the main four 3-alkyl-2-methoxypyrazines [ethylmethoxypyrazine (ETMP), IPMP, SBMP, and IBMP] in grape juice was proposed [73]. Quantitation of methoxypyrazines was performed on the signal area of MS/MS ions at m/z 119 (for ETMP), m/z 109 (IPMP), m/z 81 (IBMP), m/z 81 (SBMP) using 2-ethoxy-3-isopropylpyrazine as internal standard. For quantification of o-AAP and MA, as internal standard 2,4-dichloroaniline was used.
The optimized HS-SPME experimental conditions are described in Table 5 and GC/MS conditions in Table 6 [73].
Table 5. HS-SPME conditions used for simultaneous analysis of “foxy smelling compounds” (o-AAP and MA) and 3-alkyl-2-methoxypyrazines (ETMP, IPMP, IBMP, and SBMP) in grape juice [73].
Table 5. HS-SPME conditions used for simultaneous analysis of “foxy smelling compounds” (o-AAP and MA) and 3-alkyl-2-methoxypyrazines (ETMP, IPMP, IBMP, and SBMP) in grape juice [73].
SPME fiber50/30 μm divinylbenzene/CarboxenTM/polydimethylsiloxane (DVB/CAR/PDMS)
Sample volume10 mL
Vial volume20 mL
Addition to the sample3 g NaCl
Extraction temperature and time50 °C for 30 min
Desorption temperature and time250 °C fo 5 min
Fiber cleaning10 min
Table 6. GC/MS conditions used for simultaneous analysis of o-AAP, MA and 3-alkyl-2-methoxypyrazines in grape juice [73]. MW: molecular weight.
Table 6. GC/MS conditions used for simultaneous analysis of o-AAP, MA and 3-alkyl-2-methoxypyrazines in grape juice [73]. MW: molecular weight.
GC columnHP-5ms: (5%-phenyl) methylpolysiloxane (30 m × 0.25mm i.d; 0.25-μm film thickness)
Carrier gasHelium at constant flow 1.2 mL/min
Injector250 °C
Oven program40 °C for 5 min, 5 °C/min to 230 °C, held 3 min
MSD conditionsIonization energy 70 eV, transfer line temperature 280 °C, ion source 250 °C, ion trap in MS/MS mode
IT-MS/MS
Precursor ionMS/MS signal
AnalyteMWGC retention time (min)m/z
3-ethyl-2-methoxypyrazine138.1715.10138119
3-isopropyl-2-methoxypyrazine152.2016.46137109
3-isobutyl-2-methoxypyrazine166.2218.9012481
3-sec-butyl-2-methoxypyrazine166.2219.1413881
2-ethoxy-3-isopropylpyrazine (IS)166.2218.45166123
methyl anthranilate151.1623.71151TIC
2'-aminoacetophenone135.1622.59135TIC
2,4-dichloroaniline (IS)162.0223.35161TIC

2.4. Volatile Phenols in Wine

4-Ethylphenol (4-EP) and 4-ethylguaiacol (4-EG) are associated with wine defects that can form during winemaking or, more commonly, wine aging. These compounds are characterized by sensorial characteristics described as “stable”, “animal” and “phenolic” and their presence is particularly detrimental for the product [74,75,76]. They are produced by winery contaminants, such as Brettanomyces and Dekkera yeasts, through processes of decarboxylation and reduction of ferulic and p-coumaric acids present in the grape [75]. Sensory thresholds of 4-EP and 4-EG in wine are 440 μg/L and 33 μg/L, respectively [77].
For their determination several sensitive SPME-GC/MS methods were proposed [20,78,79]. Martorell et al., proposed the use of two PDMS 100 μm fibers: for 4-EP the method had a limit of detection (LOD) and of quantification (LOQ) of 2 μg/L and 5 μg/L, respectively, for 4-EG 1 μg/L and 5 μg/L, respectively [80]. An optimized method for analysis of ethylphenols and vinylphenols in white and red wines was developed by the use of StableFlex Carbowax/DVB (CW/DVB 70 μm) and polyacrilate (PA 85 μm) fibers [79].
4-EG and 4-EP in wine were also analyzed by a multiple-headspace SPME method. By performing three consecutive extractions of the sample with a CW/DVB fiber, the possible matrix effects were minimized by providing a LOD of 0.06 μg/L for both 4-EG and 4-EP [81].

2.5. Higher Alcohols and Esters in Wine

By using a PDMS 100-μm fiber, effective methods for analysis of higher alcohols and aliphatic esters in wine were performed [82,83]. This coating fiber showed high affinity for non-polar compounds such as ethyl esters and acetates [84,85,86], while CW/DVB fiber is suitable for more polar compounds, such as 1-hexanol, hexen-1-ol, 1-octanol, and monoterpenols [84].
Antalick et al., developed a SPME and GC/MS-SIM method which provided simultaneous determination of 32 esters in wine in concentration between ng/L and mg/L [87]. Seven different fibers were tested: DVB/CAR/PDMS 50/30 μm, CAR/PDMS 85 μm, PDMS 100 μm, PDMS/DVB 65 μm, PA 85 μm, CW/DVB 70 μm, and polyethyleneglycol (PEG) 60 μm. PDMS was the most efficient in extracting the less polar and less volatile compounds; for more volatile esters the best coating was CAR/PDMS, and aromatic esters were better recovered by CW/DVB. In general, the PDMS fiber showed high efficiency for all compounds and provided LOQs between 0.4 ng/L and 4.0 μg/L.
Recent applications showed that the tri-phase fiber DVB/CAR/PDMS provides extraction of the highest number of wine volatiles, including ethyl esters (56% of the compounds identified), alcohols, and acids [2,88].

2.6. Wine Volatile Sulfur Compounds

Various sulfur compounds are present in wine, such as thiols, sulphides, thioesters, and heterocyclic compounds. Thiol and thio-type compounds are in general associated with flavor defects of wine and they are classified as “light” (boiling point < 90 °C) and “heavy” (b.p. > 90 °C) compounds [25,29,89,90,91]. Sulfur compounds can be formed through several enzymatic and non-enzymatic processes, such as yeast fermentation and chemical, photochemical, and thermal reactions occurring in winemaking and during wine storage [89,90]. Most prevailing are ethylmercaptan (EtSH; onion as aroma descriptor), dimethyl sulfide (DMS; grassy/truffle-like note), 2-furanmethanethiol (FFT; roasted coffee), diethyl sulfide (DES; cooked vegetables, onion, garlic), dimethyl disulfide (DMDS; cooked gabbage, intense onion), diethyl disulfide (DEDS; garlic, burnt rubber), methyl thioacetate (MTA), ethyl thioacetate (ETA), 2-mercaptoethanol (ME; burnt rubber), 2-(methylthio)-1-ethanol (MTE; cauliflower), 3-(methylthio)-1-propanol (MTP; sweet, potato), 4-(methylthio)-1-butanol (MTB; earthy-like scent), benzothiazole (BT; rubber), and 5-(2-hydroxyethyl)-4-methylthiazole (HMT) [29,91]. Also, 2-methyl-3-furanthiol (MF; cooked meat), a very odoriferous compound with an odor threshold of 0.4–1.0 ppt, was found [92].
A SPME method for analysis of 13 volatile sulfur compounds in wine (i.e., DMS, EtSH, DES, MTA, ETA, ME, DMDS, DEDS, BT, HMT, MTB, MTP, and MTE) with b.p. ranging from 35 °C to 231 °C by using a CAR/PDMS/DVB 50:30 μm 2 cm length fiber, was developed [90]. By addition of MgSO4 (1.0 M) to increase ionic strength of solution and performing the extraction at 35 °C, the method showed a high sensitivity for all the analytes.
3-mercaptohexan-1-ol (3-MH; passion fruit, grapefruit), 3-mercaptohexyl acetate (3-MHA; Riesling type note, passion fruit, box tree), and 4-methyl-4-mercaptopentan-2-one (4–MP; blackcurrant or box tree note) are tropical fruit scenting volatiles present in wines at ng/L level [29]. A SPME-GC/MS method for analysis of these compounds by using a CAR/PDMS/DVB fiber was proposed [93]. HS-SPME conditions were optimized by performing the extraction of wine adjusted to pH 7 at 40 °C for 40 min [93].
A method for analysis of thiols in wine by synthesis of the pentafluorobenzyl derivatives was also proposed. Derivatization was performed directly on the PDMS/DVB fiber (65 μm), LODs achieved were 0.05 ng/L, 0.03 ng/L, 0.11 ng/L, 0.5 ng/L, and 0.8 ng/L for FFT, 3-MHA, MF, 4-MP, and 3-MH, respectively [94].

3. Conclusions

Grape aroma is composed of a hundred compounds and wine’s volatile profile also includes a number of fermentative compounds. SPME coupled to GC/MS showed to be effective for studying several classes of these analytes without solvent use. It often resulted in a high-sensitive technique for quantitative analysis of compounds for which the standard is available with high reproducibility. Moreover, the use of a multiphase fiber coupled to MS and multivariate data analysis allows sampling automation and statistical treatment of fragment abundances for the identification of compounds [95,96].
On the other hand, different to most of the sample preparation methods performed by liquid-liquid extraction and SPE, the selectivity of SPME fiber often changes dramatically for the different analytes. As a consequence, it is rarely possible to perform the semi-quantitative profiling of the sample on the internal standard signal, which is particularly useful in the characterization of grape varieties and the monitoring of the winemaking processes.
Logically, by increasing the standards commercially available new SPME-GC/MS applications are developed, and methods for the profiling of specific classes of grape aroma compounds, such as terpenols and norisoprenoids, could be particularly useful.

Author Contributions

A.P. and R.F. wrote the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Weldegergis, B.T.; Croucha, A.M.; Górecki, T.; de Villiers, A. Solid phase extraction in combination with comprehensive two-dimensional gas chromatography coupled to time-of-flight mass spectrometry for the detailed investigation of volatiles in South African red wines. Anal. Chim. Acta 2011, 701, 98–111. [Google Scholar] [CrossRef] [PubMed]
  2. Sagratini, G.; Maggi, F.; Caprioli, G.; Cristalli, G.; Ricciutelli, M.; Torregiani, E.; Vittori, S. Comparative study of aroma profile and phenolic content of Montepulciano monovarietal red wines from the Marches and Abruzzo regions of Italy using HS-SPME-GC-MS and HPLC-MS. Food Chem. 2012, 132, 1592–1599. [Google Scholar] [CrossRef]
  3. Blanch, G.P.; Reglero, G.; Herraiz, M. Rapid extraction of wine aroma compounds using a new simultaneous distillation-solvent extraction device. Food Chem. 1996, 56, 439–444. [Google Scholar] [CrossRef]
  4. Bosch-Fusté, J.; Riu-Aumatell, M.; Guadayol, J.M.; Caixach, J.; López-Tamames, E.; Buxaderas, S. Volatile profiles of sparkling wines obtained by three extraction methods and gas chromatography-mass spectrometry (GC-MS) analysis. Food Chem. 2007, 105, 428–435. [Google Scholar] [CrossRef]
  5. Mayr, C.M.; Geue, J.P.; Holt, H.E.; Pearson, W.P.; Jeffery, D.W.; Francis, I.L. Characterization of the key aroma compounds in Shiraz wine by quantitation, aroma reconstitution, and omission studies. J. Agric. Food Chem. 2014, 62, 4528–4536. [Google Scholar] [CrossRef] [PubMed]
  6. Mamede, M.; Pastore, G.M. Study of methods for the extraction of volatile compounds from fermented grape must. Food Chem. 2006, 96, 586–590. [Google Scholar] [CrossRef]
  7. Mateo, J.J.; Gentilini, N.; Huerta, T.; Jiménez, M.; di Stefano, R. Fractionation of glycoside precursors of aroma in grape and wine. J. Chromatogr. A 1997, 778, 219–224. [Google Scholar] [CrossRef] [PubMed]
  8. Cabrita, M.J.; Costa Freitas, A.M.; Laureano, O.; Borsa, D.; di Stefano, R. Aroma compounds in varietal wines from Alentejo, Portugal. J. Food Comp. Anal. 2007, 20, 375–390. [Google Scholar] [CrossRef]
  9. Rosillo, L.; Salinas, M.R.; Garijo, J.; Alonso, G.L. Study of volatiles in grapes by dynamic headspace analysis: Application to the differentiation of some Vitis vinifera varieties. J. Chromatogr. A 1999, 847, 155–159. [Google Scholar] [CrossRef]
  10. Marengo, E.; Aceto, M.; Maurino, V. Classification of Nebbiolo-based wines from Piedmont (Italy) by means of solid-phase microextraction-gaschromatography-mass spectrometry of volatile compounds. J. Chromatogr. A 2001, 943, 123–137. [Google Scholar] [CrossRef]
  11. Sanchéz-Palomo, E.; Díaz-Maroto, M.C.; Pérez-Coello, S. Rapid determination of volatile compounds in grapes by HS-SPME coupled with GC-MS. Talanta 2005, 66, 1152–1157. [Google Scholar] [CrossRef] [PubMed]
  12. Coelho, E.; Rocha, S.M.; Delgadillo, I.; Coimbra, M.A. Headspace-SPME applied to varietal volatile components evolution during Vitis vinifera L. cv. “Baga” ripening. Anal. Chim. Acta 2006, 563, 204–214. [Google Scholar] [CrossRef]
  13. Verzera, A.; Ziino, M.; Scacco, A.; Lanza, C.M.; Mazzaglia, A.; Romeo, V.; Condurso, C. Volatile compound and sensory analysis for the characterization of an Italian white wine from “Inzolia” grapes. Food Anal. Methods 2008, 1, 144–151. [Google Scholar] [CrossRef]
  14. Arthur, C.L.; Pawliszyn, J. Solid-phase microextraction with thermal desorption using silica optical fibers. Anal. Chem. 1990, 62, 2145–2148. [Google Scholar] [CrossRef]
  15. Bojko, B.; Cudjoe, E.; Gómez-Ríos, G.A.; Gorynski, K.; Jiang, R.; Reyes-Garcés, N.; Risticevic, S.; Silva, É.A.S.; Togunde, O.; Vuckovic, D.; et al. SPME-Quo vadis? Anal. Chim. Acta 2012, 750, 132–151. [Google Scholar] [CrossRef]
  16. Harmon, A.D. Solid-phase microextraction for the analysis of flavors. In Techniques for Analyzing Food Aroma; Marsili, R., Ed.; Marcel Decker, Inc.: New York, NY, USA, 1997; pp. 81–112. [Google Scholar]
  17. Yu, Y.-J.; Lu, Z.-M.; Yu, N.-H.; Xu, W.; Li, G.-Q.; Shi, J.-S.; Xu, Z.-H. HS-SPME/GC-MS and chemometrics for volatile composition of Chinese traditional aromatic vinegar in the Zhenjiang region. J. Inst. Brew. 2012, 118, 133–141. [Google Scholar] [CrossRef]
  18. Vas, G.; Vékey, K. Solid-phase microextraction: A powerful sample preparation tool prior to mass spectrometric analysis. J. Mass Spectrom. 2004, 39, 233–254. [Google Scholar] [CrossRef] [PubMed]
  19. Castro, R.; Natera, R.; Durán, E.; García-Barroso, C. Application of solid phase extraction techniques to analyse volatile compounds in wines and other enological products. Eur. Food Res. Technol. 2008, 228, 1–18. [Google Scholar] [CrossRef]
  20. Flamini, R. Volatile and aroma compounds in wines. In Mass Spectrometry in Grape and Wine Chemistry; Flamini, R., Traldi, P., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010; pp. 117–162. [Google Scholar]
  21. Garcóa, E; Chacón, J.L.; Martínez, J.; Izquierdo, P.M. Changes in volatile compounds during ripening in grapes of Airén, Macabeo and Chardonnay white varieties grown in La Mancha Region (Spain). Food Sci. Technol. Int. 2003, 9, 33–41. [Google Scholar] [CrossRef]
  22. Noguerol-Pato, R.; González-Barreiro, C.; Cancho-Grande, B.; Santiago, J.L.; Martínez, M.C.; Simal-Gándara, J. Aroma potential of Brancellao grapes from different cluster positions. Food Chem. 2012, 132, 112–124. [Google Scholar] [CrossRef]
  23. Marais, J. Terpenes in the aroma of grapes and wines: A review. S. Afr. J. Enol. Vitic. 1983, 4, 49–58. [Google Scholar]
  24. Flamini, R. Grape aroma compounds: Terpenes, C13-norisoprenoids, benzene compounds, and 3-alkyl-2-methoxypyrazines. In Mass Spectrometry in Grape and Wine Chemistry; Flamini, R., Traldi, P., Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2010; pp. 97–116. [Google Scholar]
  25. Riberéau-Gayon, P.; Glories, Y.; Maujean, A.; Dubourdieu, D. Varietal Aroma. In Handbook of Enology: The Chemistry of Wine. Stabilization and Treatments, 2nd ed.; John Wiley & Sons Ltd: Chichester, UK, 2006; Volume 2, pp. 205–230. [Google Scholar]
  26. Wang, Y.; Kays, S.J. Contribution of volatile compounds to the characteristic aroma of baked Jewel sweetpotatos. J. Am. Soc. Hortic. Sci. 2000, 125, 638–643. [Google Scholar]
  27. Robinson, A.L.; Boss, P.K.; Solomon, P.S.; Trengove, R.D.; Heymann, H.; Ebeler, S.E. Origins of grape and wine aroma. Part 1. Chemical Components and Viticultural Impacts. Am. J. Enol. Vitic. 2014, 65, 1–24. [Google Scholar] [CrossRef]
  28. Riberéau-Gayon, P.; Glories, Y.; Maujean, A.; Dubourdieu, D. Alcohols and other volatile compounds. In Handbook of Enology: The Chemistry of Wine. Stabilization and Treatments, 2nd ed.; John Wiley & Sons Ltd: Chichester, UK, 2006; Volume 2, pp. 51–64. [Google Scholar]
  29. Bakker, J.; Clarke, R.J. Volatile components. In Wine Flavour Chemistry, 2nd ed.; John Wiley & Sons, Ltd: Chichester, UK, 2012; pp. 155–238. [Google Scholar]
  30. Flamini, R.; de Luca, G.; di Stefano, R. Changes in carbonyl compounds in Chardonnay and Cabernet Sauvignon wines as a consequence of malolactic fermentation. Vitis 2002, 41, 107–112. [Google Scholar]
  31. Flamini, R.; Dalla Vedova, A. Glyoxal/glycolaldehyde: A redox system involved in malolactic fermentation of wine. J. Agric. Food Chem. 2003, 51, 2300–2303. [Google Scholar] [CrossRef] [PubMed]
  32. Baumes, R. Wine Aroma Precursors. In Wine Chemistry and Biochemistry; Victoria Moreno-Arribas, M., Carmen Polo, M., Eds.; Springer Science + Business Media, LLC: New York, NY, USA, 2009; pp. 251–274. [Google Scholar]
  33. Gambetta, J.M.; Bastian, S.E. P.; Cozzolino, D.; Jeffery, D.W. Factors influencing the aroma composition of Chardonnay wines. J. Agric. Food Chem. 2014, 62, 6512–6534. [Google Scholar] [CrossRef] [PubMed]
  34. Bell, S.J.; Henschke, P.A. Implications of nitrogen nutrition for grapes, fermentation and wine. Aust. J. Grape Wine Res. 2005, 11, 242–295. [Google Scholar] [CrossRef]
  35. Swiegers, J.H.; Bartowsky, E.J; Henschke, P.A.; Pretorius, I.S. Yeast and bacterial modulation of wine aroma and flavour. Aust. J. Grape Wine Res. 2005, 11, 139–173. [Google Scholar] [CrossRef]
  36. Rapp, A.; Versini, G. Influence of nitrogen compounds in grapes on aroma compounds in wine. In Proceedings of the International Symposium on Nitrogen in Grapes and Wine, Seattle, DC, USA, 18–19 June 1991; American Society of Enology and Viticulture: Davis, CA, USA, 1991; pp. 156–164. [Google Scholar]
  37. Riberéau-Gayon, P.; Glories, Y.; Maujean, A.; Dubourdieu, D. Phenolic compounds. In Handbook of Enology: The Chemistry of Wine. Stabilization and Treatments, 2nd ed.; John Wiley & Sons Ltd: Chichester, UK, 2006; Volume 2, pp. 141–203. [Google Scholar]
  38. Cancilla, D.A.; Que Hee, S.S. O-(2,3,4,5,6-Pentafluorophenyl)methylhydroxylamine hydrochloride: A versatile reagent for the determination of carbonyl-containing compounds. J. Chromatogr. A 1992, 627, 1–16. [Google Scholar] [CrossRef]
  39. Vanderlinde, R.; Bertrand, A.; Segur, M.C. Dosage des aldehydes dans les eaux-de-vie. In Proceedings of the 1er Symposium Scientifique International du Cognac, “Elaboration et connaissance des spiritueux”, Cognac, France, 11–15 May 1992; Lavoiser TEC & DOC: Paris, France, 1992; pp. 506–511. [Google Scholar]
  40. Vidal, J.P.; Mazerolles, G.; Estreguil, S.; Cantagrel, R. Analyse quantitative de la fraction carbonylée volatile des eaux-de-vie de Cognac. In Proceedings of the 1er Symposium Scientifique International du Cognac, “Elaboration et connaissance des spiritueux”, Cognac, France, 11–15 May 1992; Lavoiser TEC & DOC: Paris, France, 1992; pp. 529–537. [Google Scholar]
  41. De Revel, G.; Bertrand, A. A method for the detection of carbonyl compounds in wine: Glyoxal and methylglyoxal. J. Sci. Food Agric. 1993, 61, 267–272. [Google Scholar] [CrossRef]
  42. De Revel, G.; Bertrand, A. Dicarbonyl compounds and their reduction products in wine. Identification of wine aldehydes. In Trends in Flavour Research; Maarse, H., van der Heij, D.G., Eds.; Elsevier Science B.V.: Amsterdam, The Netherlands, 1994; pp. 353–361. [Google Scholar]
  43. Guillou, I.; Bertrand, A.; de Revel, G.; Barbe, J.C. Occurrence of hydroxypropanedial in certain musts and wines. J. Agric. Food Chem. 1997, 45, 3382–3386. [Google Scholar] [CrossRef]
  44. Flamini, R.; Tonus, T.; Dalla Vedova, A. A GC-MS method for determining acetaldehyde in wines. Riv. Vitic. Enol. 2002, 2/3, 15–21. [Google Scholar]
  45. Flamini, R.; Dalla Vedova, A.; Panighel, A. Study of carbonyl compounds in some Italian marc distillate (grappa) samples by synthesis of O-(2,3,4,5,6-pentafluorobenzyl)-hydroxylamine derivatives. Riv. Vitic. Enol. 2005, 1, 51–63. [Google Scholar]
  46. Flamini, R.; Dalla Vedova, A.; Panighel, A.; Perchiazzi, N.; Ongarato, S. Monitoring of the principal carbonyl compounds involved in malolactic fermentation of wine by solid-phase microextraction and positive ion chemical ionization GC/MS analysis. J. Mass Spectrom. 2005, 40, 1558–1564. [Google Scholar] [CrossRef] [PubMed]
  47. López-Vázquez, C.; Orriols, I.; Perelló, M.-C.; de Revel, G. Determination of aldehydes as pentafluorobenzyl derivatives in grape pomace distillates by HS-SPME-GC/MS. Food Chem. 2012, 130, 1127–1133. [Google Scholar] [CrossRef]
  48. Bauer, R.; Dicks, L.M.T. Control of Malolactic Fermentation in Wine. A Review. S. Afr. J. Enol. Vitic. 2004, 25, 74–88. [Google Scholar]
  49. Lerm, E.; Engelbrecht, L.; du Toit, M. Malolactic Fermentation: The ABC’s of MLF. S. Afr. J. Enol. Vitic. 2010, 31, 186–212. [Google Scholar]
  50. Zapata, J.; Mateo-Vivaracho, L.; Cacho, J.; Ferreira, V. Comparison of extraction techniques and mass spectrometric ionization modes in the analysis of wine volatile carbonyls. Anal. Chim. Acta 2010, 660, 197–205. [Google Scholar] [CrossRef] [PubMed]
  51. Pérez-Coello, M.S.; Sanz, J.; Cabezudo, M.D. Determination of volatile compounds in hydroalcoholic extracts of French and American oak wood. Am. J. Enol. Vitic. 1999, 50, 162–165. [Google Scholar]
  52. Garde-Cedàn, T.; Lorenzo, C.; Carot, J.M.; Jabaloyes, J.M.; Esteve, M.D.; Salinas, M.R. Statistical differentiation of wines of different geographic origin and aged in barrel according to some volatile components and ethylphenols. Food Chem. 2008, 111, 1025–1031. [Google Scholar] [CrossRef]
  53. Dìaz-Maroto, M.C.; Sànchez-Palomo, E.; Pérez-Coello, M.S. Fast screening method for volatile compounds of oak wood used for aging wines by headspace SPME-GC-MS (SIM). J. Agric. Food Chem. 2004, 52, 6857–6861. [Google Scholar] [CrossRef] [PubMed]
  54. Carillo, J.D.; Garrido-Lòpez, A.; Tena, M.T. Determination of oak volatile compounds in wine by headspace solid-phase microextraction and gas chromatography-mass spectrometry. J. Chromatogr. A 2006, 1102, 25–36. [Google Scholar] [CrossRef] [PubMed]
  55. De Rosso, M.; Panighel, A.; Dalla Vedova, A.; Stella, L.; Flamini, R. Changes in chemical composition of a red wine aged in acacia, cherry, chestnut, mulberry, and oak wood barrels. J. Agric. Food Chem. 2009, 57, 1915–1920. [Google Scholar] [CrossRef] [PubMed]
  56. Rapp, A.; Versini, G.; Ullemeyer, H. 2-aminoacetophenone: causal component of “untypical aging flavor” (“naphtalene note”, “hybrid note”) of wine. Vitis 1993, 32, 61–62. [Google Scholar]
  57. Hoenicke, K.; Borchert, O.; Grüning, K.; Simat, T.J. “Untypical aging off-flavor” in wine: synthesis of potential degradation compounds of indole-3-acetic acid and kynurenine and their evaluation as precursors of 2-aminoacetophenone. J. Agric. Food Chem. 2002, 50, 4303–4309. [Google Scholar] [CrossRef] [PubMed]
  58. Hoenicke, K.; Simat, T.J.; Steinhart, H.; Christoph, N.; Geßner, M.; Köhler, H.-J. “Untypical aging off-flavor” in wine: formation of 2-aminoacetophenone and evaluation of its influencing factors. Anal. Chim. Acta 2002, 458, 29–37. [Google Scholar] [CrossRef]
  59. Rapp, A.; Versini, G. Methylanthranilate (“foxy taint”) concentrations of hybrid and Vitis vinifera wines. Vitis 1996, 35, 215–216. [Google Scholar]
  60. Fan, W.; Tsai, I.-M.; Qian, M.C. Analysis of 2-aminoacetophenone by direct-immersion solid-phase microextraction and gas chromatography-mass spectrometry and its sensory impact in Chardonnay and Pinot gris wines. Food Chem. 2007, 105, 1144–1150. [Google Scholar] [CrossRef]
  61. Massa, M.J.; Robacker, D.C.; Patt, J. Identification of grape juice aroma volatiles and attractiveness to the Mexican fruit fly (Diptera: Tephritidae). Florida Entomologist 2008, 91, 266–276. [Google Scholar] [CrossRef]
  62. Lacey, M.; Allen, M.S.; Harris, R.L.N.; Brown, W.V. Methoxypyrazines in Sauvignon blanc grapes and wines. Am. J. Enol. Vitic. 1991, 42, 103–108. [Google Scholar]
  63. Allen, M.S.; Lacey, M.J.; Harris, R.L.N.; Brown, W.V. Contribution of Methoxypyrazines to Sauvignon blanc Wine Aroma. Am. J. Enol. Vitic. 1991, 42, 109–112. [Google Scholar]
  64. Allen, M.S.; Lacey, M.J. Methoxypyrazine grape flavour: Influence of climate, cultivar and viticulture. Die Wein-Wiss. 1993, 48, 211–213. [Google Scholar]
  65. Hashizume, K.; Umeda, N. Methoxypyrazine content of Japanese red wines. Biosci. Biotechnol. Biochem. 1996, 60, 802–805. [Google Scholar] [CrossRef]
  66. Hashizume, K.; Samuta, T.J. Green odorants of grape cluster stem and their ability to cause a wine stemmy flavor. J. Agric. Food Chem. 1997, 45, 1333–1337. [Google Scholar] [CrossRef]
  67. Hashizume, K.; Samuta, T. Grape maturity and light exposure affect berry methoxypyrazine concentration. Am. J. Enol. Vitic. 1999, 50, 194–198. [Google Scholar]
  68. Roujou de Boubee, D.; Cumsille, A.M.; Pons, D.; Dubordieu, D. Location of 2-methoxy-3-isobutylpyrazine in Cabernet Sauvignon grape bunches and its extractability during vinification. Am. J. Enol. Vitic. 2002, 53, 1–5. [Google Scholar]
  69. Kotseridis, Y.S.; Spink, M.; Brindle, I.D.; Blake, A.J.; Sears, M.; Chen, X.; Soleas, G.; Inglis, D.; Pickering, G.J. Quantitative analysis of 3-alkyl-2-methoxypyrazines in juice and wine using stable isotope labelled internal standard assay. J. Chromatogr. A 2008, 1190, 294–301. [Google Scholar] [CrossRef] [PubMed]
  70. Sala, C.; Mestres, M.; Martì, M.P.; Busto, O.; Guasch, J. Headspace solid-phase microextraction analysis of 3-alkyl-2- methoxypyrazines in wines. J. Chromatogr. A 2002, 953, 1–6. [Google Scholar] [CrossRef] [PubMed]
  71. Galvan, T.L.; Kells, S.; Hutchison, W.D. Determination of 3-alkyl-2-methoxypyrazines in Lady beetle-Infested wine by Solid-Phase Microextraction headspace sampling. J. Agric. Food Chem. 2008, 56, 1065–1071. [Google Scholar] [CrossRef] [PubMed]
  72. Ryan, D.; Watkins, P.; Smith, J.; Allen, M.; Marriott, P. Analysis of methoxypyrazines in wine using headspace solid phase microextraction with isotope dilution and comprehensive two-dimensional gas chromatography. J. Sep. Sci. 2005, 28, 1075–1082. [Google Scholar] [CrossRef] [PubMed]
  73. Panighel, A.; Dalla Vedova, A.; de Rosso, M.; Gardiman, M.; Flamini, R. A solid-phase microextraction gas chromatography/ion trap tandem mass spectrometry method for simultaneous determination of “foxy smelling compounds” and 3-alkyl-2-methoxypyrazines in grape juice. Rapid Commun. Mass Spectrom. 2010, 24, 2023–2029. [Google Scholar] [CrossRef] [PubMed]
  74. Chatonnet, P.; Boidron, J.-N.; Pons, M. Elevage des vins rouges en fûts de chêne: Évolution de certains composés volatils et de leur impact aromatique. Sci. Aliment. 1990, 10, 565–587. [Google Scholar]
  75. Chatonnet, P.; Dubourdieu, D.; Boidron, J.-N.; Pons, M. The origin of ethylphenols in wines. J. Sci. Food Agric. 1992, 60, 165–178. [Google Scholar] [CrossRef]
  76. Rodrigues, N.; Gonçalves, G.; Pereira-da-Silva, S.; Malfeito-Ferreira, M.; Loureiro, V. Development and use of a new medium to detect yeast of the genera Dekkera/Brettanomyces sp. J. Appl. Microbiol. 2001, 90, 588–599. [Google Scholar] [CrossRef] [PubMed]
  77. Carrillo, J.D.; Tena, M.T. Determination of ethylphenols in wine by in situ derivatisation and headspace solid-phase microextraction gas chromatography mass spectrometry. Anal. Bioanal. Chem. 2007, 387, 2547–2558. [Google Scholar] [CrossRef] [PubMed]
  78. Monje, M.-C.; Privat, C.; Gastine, V.; Nepveu, F. Determination of ethylphenol compound in wine by headspace solid-phase microextraction in conjunction with gas chromatography and flame ionization detection. Anal. Chim. Acta 2002, 458, 111–117. [Google Scholar] [CrossRef]
  79. Castro Mejías, R.; Natera Marín, R.; García Moreno, M.d.V.; García Barroso, C. Optimisation of headspace solid-phase microextraction for the analysis of volatile phenols in wine. J. Chromatogr. A 2003, 995, 11–20. [Google Scholar] [CrossRef] [PubMed]
  80. Martorell, N.; Martì, M.P.; Mestres, M.; Busto, O.; Guasch, J. Determination of 4-ethylguaiacol and 4-ethylphenol in red wines using headspace-solid-phase microextraction-gas chromatography. J. Chromatogr. A 2002, 975, 349–354. [Google Scholar] [CrossRef] [PubMed]
  81. Pizarro, C.; Pérez-del-Notario, N.; Gonzàlez-Sàiz, J.M. Determination of Brett character responsible compounds in wines by using multiple headspace solid-phase microextraction. J. Chromatogr. A 2007, 1143, 176–181. [Google Scholar] [CrossRef] [PubMed]
  82. Francioli, S.; Guerra, M.; López-Tamames, E.; Guadayoi, J.M.; Caixach, J. Aroma of sparkling wines by headspace/solid phase microextraction and gas chromatography/mass spectrometry. Am. J. Enol. Vitic. 1999, 50, 404–408. [Google Scholar]
  83. Demyttenaere, J.C.R.; Dagher, C.; Sandra, P.; Kallithraka, S.; Verhé, R.; de Kimpe, N. Flavour analysis of Greek white wine by solid-phase microextraction-capillary gas chromatography-mass spectrometry. J. Chromatogr. A 2003, 985, 233–246. [Google Scholar] [CrossRef] [PubMed]
  84. Bonino, M.; Schellino, R.; Rizzi, C.; Aigotti, R.; Delfini, C.; Baiocchi, C. Aroma compounds of an Italian wine (Ruchè) by HS-SPME analysis coupled with GC/IT/MS. Food Chem. 2003, 80, 125–133. [Google Scholar] [CrossRef]
  85. Vianna, E.; Ebeler, S. Monitoring ester formation in grape juice fermentations using solid phase microextraction coupled with gas chromatography-mass spectrometry. J. Agric. Food Chem. 2001, 49, 589–595. [Google Scholar] [CrossRef] [PubMed]
  86. Martí, M.P.; Mestres, M.; Sala, C.; Busto, O.; Guasch, J. Solid-phase microextraction and gas chromatography olfactometry analysis of successively diluted samples. A new approach of the aroma extract dilution analysis applied to the characterization of wine aroma. J. Agric. Food Chem. 2003, 51, 7861–7865. [Google Scholar]
  87. Antalick, G.; Perello, M.-C.; de Revel, G. Development, validation and application of a specific method for the quantitative determination of wine esters by headspace-solid-phase microextraction-gas chromatography-mass spectrometry. Food Chem. 2010, 121, 1236–1245. [Google Scholar] [CrossRef]
  88. Welke, J.E.; Zanus, M.; Lazarotto, M.; Schmitta, K.G.; Zini, C.A. Volatile characterization by multivariate optimization of headspace-solid phase microextraction and sensorial evaluation of Chardonnay base wines. J. Braz. Chem. Soc. 2012, 23, 678–687. [Google Scholar] [CrossRef]
  89. Mestres, M.; Busto, O.; Guasch, J. Analysis of organic sulfur compounds in wine aroma. J. Chromatogr. A 2000, 881, 569–581. [Google Scholar] [CrossRef] [PubMed]
  90. Fedrizzi, B.; Magno, F.; Moser, S.; Nicolini, G.; Versini, G. Concurrent quantification of light and heavy sulphur volatiles in wine by headspace solid-phase microextraction coupled with gas chromatography/mass spectrometry. Rapid Commun. Mass Spectrom. 2007, 21, 707–714. [Google Scholar] [CrossRef] [PubMed]
  91. Flamini, R. Analysis of aroma compounds in wine. In Hyphenated Techniques in Grape and Wine Chemistry; Flamini, R., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2008; pp. 173–225. [Google Scholar]
  92. Bouchilloux, P.; Darriet, P.; Dubourdieu, D. Identification d’un thiol fortement odorant, le 2-methyl-3-furanthiol, dans les vins. Vitis 1998, 37, 177–180. [Google Scholar]
  93. Fedrizzi, B.; Versini, G.; Lavagnini, I.; Nicolini, G.; Magno, F. Gas chromatography-mass spectrometry determination of 3-mercaptohexan-1-ol and 3-mercaptohexyl acetate in wine. A comparison between Solid Phase Extraction and Headspace Solid Phase Microextraction methods. Anal. Chim. Acta 2007, 596, 291–297. [Google Scholar] [CrossRef] [PubMed]
  94. Mateo-Vivaracho, L.; Ferreira, V.; Cacho, J. Automated analysis of 2-methyl-3-furanthiol and 3-mercaptohexyl acetate at ng/L level by headspace solid-phase microextracion with on—Fibre derivatisation and gas chromatography-negative chemical ionization mass spectrometric determination. J. Chromatogr. A 2006, 1121, 1–9. [Google Scholar] [CrossRef] [PubMed]
  95. Kinton, V.R.; Collins, R.J.; Kolahgar, B.; Goodner, K.L. Fast analysis of beverages using mass spectral based chemical sensor. Gerstel AppNote 2003, 4, 1–10. [Google Scholar]
  96. Cozzolino, D.; Smyth, H.E.; Lattey, K.A.; Cynkar, W.; Janik, L.; Dambergs, R.G.; Francis, I.L.; Gishen, M. Combining mass spectrometry based electronic nose, visible-near infrared spectroscopy and chemometrics to assess the sensory properties of Australian Riesling wines. Anal. Chim. Acta 2006, 563, 319–324. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Panighel, A.; Flamini, R. Applications of Solid-Phase Microextraction and Gas Chromatography/Mass Spectrometry (SPME-GC/MS) in the Study of Grape and Wine Volatile Compounds. Molecules 2014, 19, 21291-21309. https://doi.org/10.3390/molecules191221291

AMA Style

Panighel A, Flamini R. Applications of Solid-Phase Microextraction and Gas Chromatography/Mass Spectrometry (SPME-GC/MS) in the Study of Grape and Wine Volatile Compounds. Molecules. 2014; 19(12):21291-21309. https://doi.org/10.3390/molecules191221291

Chicago/Turabian Style

Panighel, Annarita, and Riccardo Flamini. 2014. "Applications of Solid-Phase Microextraction and Gas Chromatography/Mass Spectrometry (SPME-GC/MS) in the Study of Grape and Wine Volatile Compounds" Molecules 19, no. 12: 21291-21309. https://doi.org/10.3390/molecules191221291

Article Metrics

Back to TopTop