You are on page 1of 12

Article

Cite This: J. Agric. Food Chem. 2019, 67, 13694−13705 pubs.acs.org/JAFC

Development of a Solid-Phase Microextraction-Gas


Chromatography/Mass Spectrometry Method for Quantifying
Nitrogen-Heterocyclic Volatile Aroma Compounds: Application to
Spirit and Wood Matrices
Magali Picard,*,† Carla Garrouste,† Christelle Absalon,‡ Marie-Françoise Nonier,† Nathalie Vivas,†
and Nicolas Vivas†

Demptos Research Center, Centre d’Etude Structurale et d’Analyse des Molécules Organiques (CESAMO) and ‡CESAMO,
Institut des Sciences Moléculaires, Univ. Bordeaux, 351 Cours de la Libération, 33405 Talence, France
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via UNIVERSIDADE DO PORTO 01100 on September 28, 2023 at 10:45:17 (UTC).

*
S Supporting Information

ABSTRACT: Over wood aging, matured spirits developed a complex aromatic bouquet where roasted-like notes were often
perceived. Since many nitrogen heterocycles were related to these olfactory nuances, a headspace solid-phase microextraction
(HS-SPME) coupled to gas chromatography−mass spectrometry was developed and validated to quantify them in both spirit
and wood matrices. The various parameters affecting the extraction of the analytes from both spirit and wood samples were first
investigated (i.e., fiber coating phase, dilution, pH and volume sample, adding salt, extraction time and temperature, and
incubation time) to determine the best compromise for a single-run analysis of the whole set of studied compounds. Good
linearity (R2 > 0.99), repeatability, reproducibility, accuracy and low detection, and quantification limits were obtained, making
this analytical method a suitable tool for routine analysis of the selected nitrogen compounds. Fifteen pyrazines, three pyrroles,
and three quinolines were quantified in a series of oak wood and commercial spirit samples where some of them were identified
for the first time. The significant impact of some barrel features and the spirit in-wood maturation step on the N-heterocycle
profile in both matrices were finally discussed.
KEYWORDS: wood aging bouquet, spirit and oakwood matrices, roasted-like aromas, nitrogen-heterocyclic compounds,
headspace solid-phase microextraction (HS-SPME)

■ INTRODUCTION
Nowadays, a wide range of spirits are matured in oak casks,
containing heterocycles. Detected in coffee, black tea,
tobacco,12,13 as well as in whiskies,14 they are characterized
especially whiskies, brandies, Cognacs, Armagnacs, rums, and by medicinal and phenolic aromatic nuances, which are
liquors. In addition to the raw material and the typical particularly appreciated in peated whiskies.4 Indeed, peat can
expression of the terroir, it is widely accepted that wood aging be used as an external source of flavor, producing a smoke
accounts among the most important contributors of the quality “peak reek” aroma into the final spirit.15 Interestingly,
of distilled beverages, by increasing complexity and removing Nishimura and Masuda (1971)14 suggested that the presence
the immature character of the raw distillate.1,2 Indeed, over the of quinoline derivatives in whiskies can be attributed to peat
maturation time upon wood, major changes occur within the smoke since these compounds were only found in malt barley
volatile fraction of spirits, mainly driven by complex kilned with peat fire.
interactions with wood-derived components.2 The develop- All of these N-heterocycles belong to the Maillard reaction-
ment of this aromatic complexity and subtlety, also known as type volatile aroma compounds, the most important and
spirit aging bouquet, integrates a large palette of nuances where complex reaction in flavor chemistry occurring during the
roasted- and nutty-derived aromas are often cited in matured thermal food process. To date, some N-heterocyclic com-
spirit description.3,4 Pyrazines and pyrroles are potent and pounds of interest have been detected in spirit beverages. Kahn
characteristic volatile aroma compounds found in a wide range (1969)16 and Worben et al. (1971)17 identified several alkyl-
of raw and processed foods. They were identified in the pyrazines in Scotch whiskies, which was confirmed by
volatiles of roasted peanuts,5 cocoa,6 or coffee.7 Their sensory Nishimura and Masuda (1971)14 in Japanese whiskies and
properties are generally associated with pleasant roasted, nutty, later by Boothroyd et al. (2014)11 in a larger set of Scotch
cereal, licorice, coffee, and woody nuances8,9 (Table 1). Wang whiskies. Moreover, pyrazines were also found in other
et al. (1969)10 also postulated that pyrazines played an distillates, including Cognac,18 tequila,19 rum,20 and other
important role in the roasted barley flavor since their removal
resulted in a loss of the typical desirable roasted aroma. The Received: September 10, 2019
relevance of pyrazines to the nutty character of a new-make Revised: November 8, 2019
malt spirit was then confirmed by Boothroyd et al. (2014).11 Accepted: November 11, 2019
Quinoline derivatives constitute another type of nitrogen- Published: November 22, 2019

© 2019 American Chemical Society 13694 DOI: 10.1021/acs.jafc.9b05716


J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

Table 1. Odor Characteristic, Boiling Point, Decimal Logarithm of Octanol−Water Partition Coefficient (Log P) and Ions
Monitored in SIM Detection for each Nitrogen-Heterocyclic Compound Studied
compound aroma description molecular mass (g/moL) boiling point (°C)a log P specific ions (m/z)b
2-methylpyrazine nutty 94.1 137 0.2 67/94
2,3-dimethylpyrazine nutty, meaty 108.1 154 0.6 42/81/108
2,5-dimethylpyrazine nutty, meaty roast 108.1 154 0.6 42/81/108
2,6-dimethylpyrazine nutty, sweet, chocolate 108.1 154 0.6 42/81/108
2,3,5-trimethylpyrazine nutty, caramel, roasted peanut, burnt 122.2 171 1.9 42/81/122
2,3,5,6-tetramethylpyrazine walnut, green, cocoa 136.2 190 1.3 42/54/136
2-ethyl-3-methylpyrazine toasted nutty, peanut, cereal 122.2 170 1.1 67/94//121/122
2-ethyl-3,5(6)-dimethylpyrazine roasted, cocoa, peanut, coffee, woody 136.2 181 1.5 108/135/136
5-ethyl-2,3-dimethylpyrazine burnt, roasted cocoa 136.2 191 1.5 54/135/136
2,3-diethyl-5-methylpyrazine hazelnut, roasted, meaty 150.2 192 1.9 56/135/150
2-acetylpyrazine popcorn, roasted, hazelnut 122.1 189 0.2 80/94/122
2-acetyl-3-methylpyrazine nutty, hazelnut, roasted 136.2 165 0.2 43/93/136
2-acetyl-3,5(6)-dimethylpyrazine nutty, hazelnut, peanut 150.2 70 (7 mmHg) 1.1 107/108/150
pyrrole nutty, sweet 67.1 130 0.8 41/67
2-acetylpyrrole licorice, walnut 109.1 220 0.9 66/94/109
2-acetyl-1-methylpyrrole earthy, nutty, smoky 123.1 201 0.7 80/108/123
quinoline medicinal, tobacco, leather 129.1 237 2.0 102/129
2-methylquinoline medicinal 143.2 247 2.6 115/128/143
6-methylquinoline animal, leather, phenolic 143.2 260 2.6 115/143
a
Except for 2-acetyl-3,5(6)-dimethylpyrazine, all boiling points were obtained at 760 mm Hg. bQuantifier ions are in bold.

Figure 1. General chemical structures of the nitrogen-heterocyclic compounds studied. R = H, CH3, or CH2CH3.

types of liquors.21 In addition, several researches also noted the profiles, the various parameters were optimized in each matrix
presence of pyrazines and pyrroles in wood,22 wood smoke.23 independently. To our knowledge, it is the first time that a
Kim et al. (1974)24 and Maga and Chen (1985)25 identified study was designed to develop such an analytical approach,
eight alkyl- and acetyl-pyrazines present at various concen- well-suited for each matrix. It offered the opportunity to
trations depending on the wood source. However, despite simultaneously examine, from both qualitative and quantitative
there is a wealth of information on the occurrence of nitrogen- standpoints, a large set of the targeted nitrogen-heterocyclic
heterocyclic aroma compounds in food, they have received compounds covering a wide range of physico-chemical
little attention in the spirit aging context, regarding the wide properties and characterized by roasted- and nutty-derived
range of data collected on other wood-derived volatile aroma notes (Table 1).
compounds such as whisky lactones, volatile phenols, phenolic,
and furanic aldehydes.26−28
Besides, the complexity of spirit and wood-derived volatile
■ MATERIALS AND METHODS
Chemicals and Reference Compounds. Absolute ethanol
aroma compound profiles represents a current challenge from (purity >99.8%) was obtained from Merck (Darmstadt, Germany)
an analytical standpoint and, more particularly, when and used as a solvent for preparing the stock solutions of analytes as
compounds at low concentration have to be not only identified well as the wood solutions. Ultrapure water (18.2 M Ω.cm) was
but also quantified. Headspace solid-phase microextraction obtained from a Milli-Q purification system (Millipore, Saint-
(HS-SPME) is considered one of the most effective sample- Quentin-en-Yvelines, France), and sodium chloride (99%) was
supplied by Sigma-Aldrich (Fontenay-sous-bois, France).
enrichment technique as it is solvent-free, accurate, sensitive, The general chemical formulas of the twenty-one studied N-
and easy-to-automate.29 The present study was undertaken to heterocyclic compounds are presented in Figure 1. They are derived
gain knowledge on the composition of spirit and wood samples from either pyrazine, pyrrole, or quinoline chemical backbones. All
in terms of nitrogen-heterocyclic compounds. Indeed, although standard compounds were obtained from Sigma-Aldrich (Fontenay-
some quantitative analyses of N-heterocycles in whiskies and sous-bois, France): 2-methylpyrazine (CAS registry no. 109-08-0,
Chinese liquors were reported,30,31 no previous study purity 99%), 2,3-dimethylpyrazine (CAS registry no. 5910-89-4,
investigated in-depth how maturation in wood barrels impacts purity 95%), 2,5-dimethylpyrazine (CAS registry no. 123-32-0, purity
on the spirit-related roasted and peated volatile aroma 98%), 2,6-dimethylpyrazine (CAS registry no. 108-50-9, purity 98%),
2,3,5-trimethylpyrazine (CAS registry no. 14667-55-1, purity 99%),
compounds. A HS-SPME extraction method, coupled to 2,3,5,6-tetramethylpyrazine (CAS registry no. 1124-11-4, purity 98%)
GC/MS analysis, was therefore developed and validated by 2-ethyl-3-methylpyrazine (CAS registry no. 15707-23-0, purity 98%),
assaying twenty-one N-heterocyclic metabolites, to address in a 2-ethyl-3,5(6)-dimethylpyrazine (mixture of isomers, CAS registry no.
single-analytical run their spirit and wood full profiling. As 27043-05-6, purity 95%), 5-ethyl-2,3-dimethylpyrazine (CAS registry
spirit and wood matrices had different volatile and nonvolatile no. 15707-34-3, purity 96%), 2,3-diethyl-5-methylpyrazine (CAS

13695 DOI: 10.1021/acs.jafc.9b05716


J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

registry no. 18138-04-0, purity 99%), 2-acetylpyrazine (CAS registry 40, 50, and 60 min). The incubation time was also optimized by
no. 22047-25-2, purity 99%), 2-acetyl-3-methylpyrazine (CAS registry evaluating extraction efficiency after 5, 10, and 15 min.
no. 23787-80-6, purity 98%), 2-acetyl-3,5(6)-dimethylpyrazine HS-SPME Optimized Procedure. The 2 mL initial samples were
(mixture of isomers, CAS registry no. 54300-08-2, purity 98%), diluted 10-fold in MQ water to reach a final volume of 20 mL, and pH
pyrrole (CAS registry no. 109-97-7, purity 98%), 2-acetylpyrrole was adjusted to 4 and 7 for wood and spirit solutions, respectively,
(CAS registry no. 1072-83-9, purity 98%), 2-acetyl-1-methylpyrrole using 1 M HCl and 1 M and 10 M NaOH solutions. To a 20 mL
(CAS registry no. 932-16-1, purity 98%), quinoline (CAS registry no. headspace vial containing 3 g of sodium chloride was added either a
91-22-5, purity 98%), 2-methylquinoline (CAS registry no. 91-63-4, 10 mL spirit or a 5 mL wood-diluted sample spiked with 10 or 5 μL of
purity 95%), 6-methylquinoline (CAS registry no. 91-62-3, purity internal standard 3,4-dimethylphenol solution (100 mg/L in alcoholic
98%), and 3,4-dimethylphenol (CAS registry no. 95-65-8, purity solution), respectively. The vial was then tightly sealed with a PTFE-
98%). lined cap, and the solution was homogenized with a vortex shaker
Spirit and Wood Samples. Method Optimization and before loading onto an auto-sampling device. In the optimized
Validation. For spirit analysis, the method was optimized and conditions, extractions were performed using the CAR-PDMS-coated
validated using a commercial whisky (40% of ethanol) spiked with the fiber and the extraction program consisted in swirling the vial for 5
studied nitrogen heterocycles at 500 μg/L. Concerning the wood min at 60 °C under regular agitation (200 rpm, 10 s intervals), then
analysis, a nontoasted oak chip sample (25 g) was extracted in a 500 inserting the fiber into the headspace for 30 min at 60 °C as the
mL of hydroalcoholic solution (50% of ethanol) during 12 h at room solution was swirled again, and finally transferring the fiber to the
temperature under magnetic stirring. The oakwood mixture was then injector for desorption at 250 °C for 5 min.
filtered using a filter paper, and the filtrate was spiked with 500 μg/L GC−MS Instrumentation and Analytical Conditions. Chro-
of the studied volatile aroma compounds. matographic analyses were carried out using a Trace GC Ultra system
Method Application. Once the analytical parameters were (Thermo Fisher Scientific, CA, USA) coupled to an ISQ quadrupole
optimized, a series of six different commercial oak chips (Oak In mass spectrometer, equipped with a Triplus autosampler (Thermo
Wine, Arobois, Gangnac-Sur-Cere, France) were analyzed. They were Fisher Scientific) and an Optima Waxplus capillary column (30 m ×
issued from French (Quercus petraea and Quercus robur) and 0.25 mm ID, 0.25 μm film thickness, Macherey-Nagel, Düren,
American (Quercus alba) oak wood species and were toasted at Germany). The carrier gas was Helium N55 at a constant flow of 1.2
three temperature levels in an industrial kiln, based on the cooper’s mL min−1. The oven temperature was raised from 40 °C (1 min) to
skills (light: 120−130 °C; medium: 160−170 °C, and heavy: 210− 160 °C at a rate of 5 °C/min and then ramped at a rate of 15 °C/min
220 °C). Oak chips solutions were prepared in the same way that of to 260 °C (final isotherm for 5 min). The ion source was set at 200
during the method optimization step (25 g in a 500 mL water/ethanol °C and the transfer line between GC and MS at 250 °C. Detection
solution, 50/50; v/v) and were analyzed using the validated and quantitation were carried out in electron ionization (70 eV) and
methodology. The analytical method was also applied to 41 the selected-ion monitoring (SIM) mode, based both on the retention
commercial spirit samples (alcohol between 40 and 70%), issued time of each compound and the selection of specific ions (two or
from various brands, and including 6 Armagnacs, 3 Cognacs, 3 three ions for each analyte: one or two for confirmation and one for
brandies, 10 rums, and 19 whiskies. Six of them were non matured quantification, on the basis of the best measured signal-to-noise ratio;
upon wood, while the others covered a wide range of maturation Table 1). The 3,4-dimethylphenol internal standard was quantified
periods in wood barrels (from 1 to 48 years, Table 5). Each sample with an ion mass-to-charge (m/z) of 122, and the qualifier ion was an
was analyzed in duplicate for both spirit and wood matrices. m/z of 107. Before the first daily analysis, a blank test was performed
Preliminary Optimization of Several Analytical Parameters to check possible carryover between samples.
during the Extraction Step. In food analysis, various factors are Method Validation. The method linearity was evaluated both in
reported to influence the effectiveness of SPME in concentrating whisky and wood matrices used in the method optimization through
analytes.29,32 In this study, two kinds of parameters were more 10 calibration levels covering a 0.5−1000 μg/L range of
particularly studied in relation to either the sample (pH, volume, concentrations for the various analytes. The performance of the
dilution factor, and sodium chloride content) or the fiber (coating, proposed optimized method was evaluated in terms of limits of
adsorption temperature and time, and incubation time) (Table S1, detection (LOD) and quantification (LOQ), defined as the
Supporting Information). The optimized SPME parameters were concentrations giving signal-to-noise ratios of 3 and 10, respectively.
selected to achieve the highest extraction efficiency, by comparing the Based on linear regression analysis, LOD and LOQ were calculated
peak area of analytes obtained under each test condition. All from the calibration line, at low concentrations and using the
experiments were performed in duplicate. following formulas: LOD = 3Sa/b and LOQ = 10Sa/b, where Sa is the
SPME Fiber Coatings. Several fibers (Supelco, Bellefonte, PA, standard deviation of the response, and b the slope of the calibration
USA) coated with various stationary phases and film thicknesses were curve.33 To evaluate repeatability (intraday precision), 10 replicates of
compared: polydimethylsiloxane 100 μm (PDMS), carboxen-poly- (i) the same spirit and (ii) wood solutions were spiked at 50 μg/L
dimethylsiloxane 85 μm (CAR-PDMS), and divinylbenzene-carbox- with reference standards. The reproducibility of the method (interday
en-polydimethylsiloxane 50/30 μm (DVB/CAR/PDMS). They were precision) was determined by analyzing 12 replicates of the same
conditioned prior to use by insertion into the GC injector, as whisky and wood solution spiked at 50 μg/L over a period of 1
recommended by the manufacturer. month. The accuracy of the analytical method was evaluated by
Sample Preparation. The impact of the dilution was first calculating the recoveries on a non matured spirit and a light-toasted
investigated by comparing results obtained from the solution sample oakwood extract solution using a standard addition technique.
diluted by factors 2, 4, and 10. The pH was then modified by adding Accuracy is reported for each assay as the percentage recovery of a
directly in 25 mL samples, a few drops of 1 M HCl and 1 and 10 M known amount of analyte added to the spirit and wood samples. Prior
NaOH solutions, to compare the relative efficiency of extractions at to sample preparation, each matrix was spiked with three different
pH 2.5, 4, 7, and 9, respectively. The optimization of sample concentration levels: 10 μg/L (low level), 100 μg/L (medium level),
preparation also focused on the volume phase for which three and 500 μg/L (high level).
volumes (3, 5, and 10 mL) were tested. The effect of salt addition on Statistical Analysis. For each optimization step, data were
the extraction was finally studied, by increasing the sodium chloride statistically analyzed with one-way analysis of variance (ANOVA) to
content in the sample (0, 1.5, and 3 g). determine any significant differences. The F-ratio served as a marker
Fiber Adsorption Parameters. Both the temperature and time of of the discrimination ability, and the statistically significant level was
HS-SPME significantly affect the equilibrium during extraction and set at 5% (p value <0.05).
thus influence the compound extraction. Operating conditions were All quantitative data obtained on the set of wood samples were
optimized while setting SPME extractions of samples at different analyzed using a two-way ANOVA (factor 1: wood species, factor 2:
adsorption temperatures (40, 50, 60, and 70 °C) and times (20, 30, toasting degree) to correlate N-heterocycles concentrations to wood

13696 DOI: 10.1021/acs.jafc.9b05716


J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

Figure 2. Impact of (a) fiber coating phase, (b) sample dilution, (c) sample pH, and (d) sample volume on SPME extraction efficiency in spirit and
wood matrices. Letters indicate significant differences at p value <0.05.

specificities and identify interactions between the considered factors. of each parameter was studied individually, in the two distinct
A one-way ANOVA, at a significance level of 95%, was applied to the matrices (Figures 2 and 3).
two global clusters of spirit samples (matured and non matured) for During the first steps of the development process (SPME
all the volatile aroma compound concentrations. A one-way ANOVA fiber selection and optimization of the sample preparation), the
test was also specifically performed on the matured and non matured
program of the headspace SPME was set on the following
samples within each kind of spirit clusters (i.e., Armagnac, Cognacs,
and brandies/rums/whiskies). All statistical analyses were conducted parameters: incubation/extraction temperature: 40 °C; in-
using XLSTAT Sensory software (Addinsoft, Paris, France, version cubation time: 15 min; extraction time: 30 min.
2019.2.2). Selection of the Fiber Coating. Since SPME is by nature an


equilibrium technique, the aim is to reach the equilibrium
between the sample matrix and the coating of the SPME device
RESULTS AND DISCUSSION
as quickly as possible. Because both matrix and coating are
Development of the HS-SPME Analytical Method: competing for analytes, the affinity of coating for targeted
Preliminary Optimizations. Although pyrazines, pyrroles, analytes is crucial in SPME sampling. The extraction of solutes
and quinolines belong to the N-heterocyclic chemical family, from the aqueous phase into the SPME fiber is controlled
they present a wide range of chemical properties such as among others by the partition coefficients (usually known as
boiling points and polarities (log P), which may directly affect log Ko/w or log P). To investigate the extraction yields of the
their extraction (Table 1). Therefore, it was necessary to studied analytes, three fiber coatings were tested: PDMS,
determine the best compromise within the optimized CAR/PDMS, and DVB/CAR/PDMS. PDMS fiber is the most
conditions for getting in a single-run analysis the whole set useful coating for nonpolar analytes, whereas CAR/PDMS and
of studied heterocyclic compounds. In this optimization stage, DVB/CAR/PDMS fibers provide the best extraction efficien-
the impact of various operating conditions on HS-SPME cies for a wide range of analytes with different polarities and
performance was evaluated, including fiber coating phase, molecular weights.32 For this first-optimization step, the
sample preparation (pH, volume, dilution, salt addition), and parameters related to the sample preparation were first fixed
fiber adsorption parameters (extraction temperature and time, as follows for both matrices: pH 4; 10 mL volume; 10-fold
incubation time). dilution; 3 g of sodium chloride addition. The comparison of
For better clarity, N-heterocycles were classified in three absolute peak areas between the experiments revealed that
groups according to similarities into their chemical structures: CAR/PDMS was the most efficient coating phase for pyrazines
alkypyrazines and pyrrole (group 1), acetylpyrazines and and pyrroles as it presented the significant greatest specificity
acetylpyrroles (group 2), and quinolines (group 3). The effect in both matrices (p value <0.05; Figure 2a). Better results were
13697 DOI: 10.1021/acs.jafc.9b05716
J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

Figure 3. Effect of (a) adsorption temperature and (b) fiber exposure time on SPME extraction efficiency in spirit and wood matrices.

obtained for quinolines when polyphasic coating fibers were follows for both matrices: pH 4; 10 mL volume; 3 g of sodium
used, but no significant differences were observed between chloride. From our results and also on the basis of previously
CAR/PDMS and DVB/CAR/PDMS phases. Interestingly, it published data,37 decreasing the ethanol content had a
was reported that combining polar and nonpolar coatings on beneficial effect on extraction, with significant higher area
the same fiber (typically a mixed CAR/PDMS coating phase) responses found in samples diluted 10 times (p value < 0.01)
improves the extraction efficiency of polar organic analytes in spirit and wood matrices, respectively (Figure 2b). Thus, a
within a polar matrix, while the simultaneous extraction of 10-fold dilution was adopted for the next experiments.
interferents is not possible.34 These results were also confirmed Influence of Sample pH. Once the fiber coating and the
by previous studies where the CAR/PDMS coating fiber was dilution factor had been selected, the influence of matrix pH
chosen in heterocyclic compounds quantitation for wine35 or on extraction was studied. In this experiment and for both
meat36 applications. Altogether, the CAR/PDMS fiber was matrices, the sample volume and the salt content were always
selected for subsequent studies. kept at 10 mL and 3 g, respectively. Since the N-heterocyclic
Dilution Effect. The percentage of the protic solvent is compounds are weak basic compounds, the pH of the sample
usually the most influencing variable in SPME-based analysis potentially impacts their extraction. Indeed, some previous
since a high content may dramatically reduce fiber extraction studies adjusted the pH solution in developing SPME
efficiency and also damage the polymer-coating phase of the analytical method.35,38 For both matrices, the least efficient
SPME fiber. As ethanol is one of the main components in conditions were found at pH 2.5 where the response areas for
spirit, the effect of the ethanol content on the aroma the whole set of analytes were significantly lower than in other
compound extraction must be taken into account. Thus, the pH conditions. It was especially verified for pyrazine and
high ethanol concentration (40−50%, v/v) of both spirit and pyrrole derivatives for which the response factor was divided
wood extract solutions required sample dilution. Three ethanol by a factor 2 to 7 (p value <0.01; Figure 2c). A slight but not
level contents were compared, and the initial spirit and wood significant increase in peak areas was obtained at pH higher
samples were diluted in water by 2, 4, and 10 times. The other than 4 in spirit and wood matrices, meaning that pH had less
parameters related to the sample preparation were fixed as effect on extraction than other parameters. However, it was
13698 DOI: 10.1021/acs.jafc.9b05716
J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

Table 2. Validation Parameters (Linearity, LOD and LOQ, Repeatability, and Reproducibility) of the Optimized SPME-GC/
MS Method Applied to Wood Matrix
intraday precision interday precision
linearity sensitivity (repeatability) at 50 μg/L (reproducibility) at 50 μg/L
concentration LOD LOQ
compound range (μg/L) R2 slope intercept (μg/L) (μg/L) RSD (%) RSD (%)
2-methylpyrazine 4.9−994 0.991 0.1698 +0.0005 0.069 0.233 4.5 14.4
2,3-dimethylpyrazine 5.2−1038 0.991 0.2647 +0.0892 0.117 0.391 4.5 15.0
2,5-dimethylpyrazine 5.1−1029 0.991 0.3608 −0.0373 0.153 0.511 7.9 13.9
2,6-dimethylpyrazine 4.6−923 0.993 0.5399 +0.1082 0.185 0.617 6.1 14.7
2,3,5- 4.7−951 0.997 0.2662 +0.0477 0.066 0.221 7.4 12.1
trimethylpyrazine
2,3,5,6- 4.7−940 0.992 0.9977 +0.1137 0.373 1.242 8.5 18.9
tetramethylpyrazine
2-ethyl-3- 5.5−1095 0.991 0.6354 +0.2159 0.303 1.011 3.8 16.6
methylpyrazine
2-ethyl-3,5- 6.2−1230 0.991 0.9808 +0.2461 0.517 1.722 6.2 16.4
dimethylpyrazine
2-ethyl-3,6- 6.2−1230 0.991 1.9527 +0.7053 1.055 3.518 6.7 14.5
dimethylpyrazine
5-ethyl-2,3- 4.9−994 0.990 0.0509 +0.0193 0.022 0.076 5.0 13.8
dimethylpyrazine
2,3-diethyl-5- 5.9−1182 0.992 2.8401 +1.0429 1.411 4.705 4.9 12.8
methylpyrazine
2-acetylpyrazine 6.4−1276 0.995 0.648 −0.0057 0.025 0.083 9.4 13.7
2-acetyl-3- 5.7−1132 0.993 0.3378 +0.050 0.140 0.468 7.6 11.1
methylpyrazine
2-acetyl-3,5- 5.1−1025 0.994 0.4795 +0.0338 0.167 0.557 9.4 9.1
dimethylpyrazine
2-acetyl-3,6- 5.1−1025 0.992 0.1379 −0.0095 0.055 0.184 8.9 11.1
dimethylpyrazine
pyrrole 4.9−982 0.991 0.0762 −0.002 0.031 0.101 9.0 17.0
2-acetylpyrrole 4.8−958 0.997 0.078 −0.0008 0.018 0.060 5.4 17.1
2-acetyl-1- 6.2−1238 0.991 0.8506 +0.2395 0.445 1.092 8.0 12.9
methylpyrrole
quinoline 4.6−918 0.992 0.5974 −0.3586 0.879 2.929 9.0 18.2
2-methylquinoline 6.9−1393 0.991 0.2228 −0.3005 0.540 1.802 9.2 16.3
6-methylquinoline 4.9−990 0.993 0.4795 −0.3923 0.722 2.408 9.6 16.6

noticed a significant increase in absolute areas for quinolines at and pyrrole (p value <0.05) as well as, even not significant, a
pH 4 in the wood matrix as well as a better resolution of trend to a better extraction for acetylpyrazines, acetylpyrroles,
chromatographic peaks in the spirit matrix at pH 7. and quinolines (Figure 2d).
Accordingly, we chose to buffer the sample pH to 4 and 7 Salt Addition. Most studies have shown that the addition of
for wood and spirit matrices, respectively. a salt (usually sodium chloride) enhances the ionic strength
Sample Volume. As previously described, the sample and thus reduces the solubility of analytes, which are more
volume is an important SPME parameter to be optimized easily retained.29 To check this “salting out” effect, 10 and 5
since it is directly related to the method sensitivity.39 To mL samples for the spirit and wood matrix, respectively,
address the impact of this parameter on compounds extraction without or with sodium chloride (1.5 and 3 g) were compared.
performance, three volumes (3, 5, and 10 mL) were compared As expected, results highlighted that, for all the analytes studied
in both matrices. This step was performed in the optimized in both matrices, responses clearly increased with the sodium
conditions previously determined, and the added content of chloride content (p value <0.001), with the highest peak areas
sodium chloride was fixed at 3 g. Results obtained in the spirit obtained in the most saturated sample (data not shown). A 3 g
matrix highlighted that a 3 mL sample volume led to addition of sodium chloride was thus adopted for the study.
contrasting results between analytes. Indeed, the best Adsorption Temperature and Time. Once the best
extraction efficiency was observed for alkylpyrazines and conditions for the fiber coating phase and the sample
pyrrole, whereas lowest responses were obtained for the preparation were identified and adopted, the next steps of
other analytes. Moreover, a 10 mL sample provided the best the SPME optimization method were focused on the
sensitivity for quinoline compounds where the abundances adsorption time and temperature. Time and temperature are
increased by a factor of 6 compared to the other conditions (p parameters closely related to each other.40 Indeed, higher
value <0.001). Consequently, considering the extraction temperatures increase the partial vapor pressure of analytes and
efficiency for all the set of compounds and as alkylpyrazines enable shorter exposure time mainly for highly volatile
and pyrrole were more easily detectable, a volume of 10 mL components. However, an excessive increase in temperature
was set for the spirit samples. On the other hand, a 5 mL can cause premature desorption of analytes with the coating
volume appeared to be the best choice for wood extracts since phase beginning to lose its ability to absorb them. Practically,
a significant gain in efficiency was observed for alkylpyrazines the effect of temperature was investigated by analyzing spiked
13699 DOI: 10.1021/acs.jafc.9b05716
J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

Table 3. Validation Parameters (Linearity, LOD and LOQ, Repeatability, and Reproducibility) of the Optimized SPME-GC/
MS Method Applied to Spirit Matrix
intraday precision interday precision
linearity sensitivity (repeatability) at 50 μg/L (reproductibility) at 50 μg/L
concentration LOD LOQ
compound range (μg/L) R2 slope intercept (μg/L) (μg/L) RSD (%) RSD (%)
2-methylpyrazine 0.49−497 0.998 0.3108 −0.0009 0.25 0.84 6.6 15.7
2,3-dimethylpyrazine 0.52−519 0.990 0.4026 +0.0652 0.117 0.391 6.6 15.1
2,5-dimethylpyrazine 0.51−514 0.995 0.7312 −0.0164 0.093 0.309 6.9 18.4
2,6-dimethylpyrazine 0.46−461 0.994 0.9474 +0.1203 0.123 0.411 7.0 18.3
2,3,5- 0.47−475 0.992 0.4589 +0.0317 0.067 0.225 8.4 18.2
trimethylpyrazine
2,3,5,6- 0.47−470 0.994 0.6773 +0.070 0.266 0.888 9.0 13.8
tetramethylpyrazine
2-ethyl-3- 0.55−548 0.992 0.9239 +0.0754 0.163 0.546 7.5 17.4
methylpyrazine
2-ethyl-3,5- 0.62−615 0.993 1.2768 +0.1144 0.236 0.788 9.7 15.8
dimethylpyrazine
2-ethyl-3,6- 0.62−615 0.993 3.0134 +0.2965 0.560 1.866 9.6 14.8
dimethylpyrazine
5-ethyl-2,3- 0.49−497 0.990 0.0928 +0.0109 0.016 0.053 9.2 16.7
dimethylpyrazine
2,3-diethyl-5- 0.59−591 0.994 4.1585 +0.6307 0.845 2.818 9.8 15.7
methylpyrazine
2-acetylpyrazine 0.64−638 0.993 0,0286 +0.0016 0.022 0.087 8.3 16.9
2-acetyl-3- 0.57−566 0.998 0.5623 +0.0172 0.054 0.180 8.4 16.5
methylpyrazine
2-acetyl-3,5- 0.51−512 0.995 0.7744 +0.0028 0.092 0.306 7.5 15.8
dimethylpyrazine
2-acetyl-3,6- 0.51−512 0.997 0.2159 −0.0034 0.021 0.069 7.2 12.2
dimethylpyrazine
pyrrole 0.49−491 0.995 0.1452 +0.0131 0.017 0.058 5.7 18.7
2-acetylpyrrole 0.48−479 0.990 0.0146 +0.0084 0.03 0.09 6.7 16.7
2-acetyl-1- 0.62−619 0.997 1.534 +0.0778 0.188 0.626 8.0 16.4
methylpyrrole
quinoline 0.46−459 0.995 2.235 +0.0186 0.216 0.722 5.9 12.5
2-methylquinoline 0.69−696 0.993 2.252 −0.3119 0.559 1.865 9.2 16.1
6-methylquinoline 0.49−495 0.994 2.3402 −0.2367 0.381 1.271 8.5 17.2

spirit and wood matrices at 40, 50, 60, and 70 °C, with a roles in both matrices, with an equilibrium reached after 30
constant incubation time set at 15 min and extraction time at min of exposure. The same response profile was obtained in
30 min (Figure 3a,b). As can be observed in both matrices, the both matrices for quinoline derivatives over the time range,
extraction performance for alkylpyrazines and pyrrole, with the highest peak areas obtained at 30 min followed by a
acetylpyrazines and acetylpyrroles, increased from 40 to 60 decrease until 60 min. Consequently, a 30 min fiber exposure
°C before decreasing at 70 °C. Concerning quinoline time was adopted to achieve the best performance analysis of
compounds, a global increase in the response peak areas was the twenty-one compounds studied.
observed over the extraction temperature range. Thus, an Incubation Time. The time period before inserting the fiber
adsorption temperature of 60 °C seemed like a relevant into the sample was finally optimized. Several incubation times
compromise for the whole set of studied N-heterocyclic (5, 10, and 15 min) were thus examined. Although no analyte
compounds. The exposure time is also an important parameter, loss was observed, increasing the incubation time to 10 or 15
which influences the partition of solutes between sample min did not give any significant efficiency gain (data not
solution and fiber coating, inducing a direct influence on the shown). Therefore, as a compromise between quick analysis
analysis time and performance. Although a longer time favors run and good method sensitivity, a 5 min incubation time was
the analyte occupation of more sites on the fiber, an excessive selected for subsequent experiments.
time may sometimes cause desorption.29 Validation of the Analytical Method. All validation
To optimize the SPME extraction period, a range of parameters, such as linearity, detection and quantification
extraction times (20−60 min) were tested at 60 °C, with a limits, precision, and accuracy, were evaluated under optimum
previous incubation time set at 15 min. Figure 3c,d presents selected conditions.
the different extraction profiles of N-heterocyclic compounds Calibration. Both spirit and wood matrices were spiked with
in spirit and wood matrices, respectively. For alkylpyrazines the studied analytes at various concentrations to obtain 10
and pyrrole in the spirit matrix, a slight increase in peak areas calibration levels (from 0.5 to 500 μg/L for the spirit matrix
was observed between 20 and 30 min, followed by a decrease and from 5 to 1000 μg/L for the wood matrix). The calibration
until 60 min (Figure 3c), whereas a constant decrease was curves were plotted as the relative peak area ratio (analyte vs
observed over the studied times in wood samples (Figure 3d). internal standard) and as a function of the added compound
A similar trend was found for acetylpyrazines and acetylpyr- concentration. From the data obtained, the optimized method
13700 DOI: 10.1021/acs.jafc.9b05716
J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

showed linear functions throughout the concentration range in Table 4. F-Ratio Values Obtained in a Two-Way ANOVA
both matrices, with coefficients of determination ranging Performed on the N-Heterocycles Mean Concentrations
between 0.990 and 0.998 (Tables 2 and 3). Obtained in the Wood Samples. Factor 1 Referred to the
Sensitivity. Limits of detection (LOD) and quantification Oak Wood Species (American and French Oak Woods) and
(LOQ) varied according to the analyte. In the wood matrix, Factor 2 to the Toasting Levels (Light, Medium and
they were found between 0.03 and 1.41 μg/L and from 0.06 to Heavy)a
4.71 μg/L, respectively (Table 2). In the spirit matrix, LODs
interaction of
and LOQs were between 0.02 and 0.85 μg/L and from 0.05 to factors factors
2.82 μg/L, respectively (Table 3). In both mixtures, 2,3-
compound oak toasting oak x toasting
diethyl-5-methylpyrazine presented the highest detection and
quantification limits. Moreover, LODs and LOQs determined 2,6-dimethylpyrazine 0.43 56.06*** 2.67
in wood were slightly higher than those found in spirit 5-ethyl-2,3- 56.31*** 47.31*** 11.30**
dimethylpyrazine
probably due to differences in the matrix composition. In the 2-acetylpyrazine 11.47 81.76*** 3.46
study of complex mixtures, characterization and quantitation of 2-acetyl-3-methylpyrazine 3.12 3.49 11.81
analytes are often impaired by matrix effects, which directly 2-acetyl-3,5- 14.52** 0.14 5.92
affect the release and thus the extraction efficiency of volatile dimethylpyrazine
aroma compounds. Headspace partitioning differences are 2-acetyl-3,6- 8.60* 16.09** 8.69**
observed according to the hydrophilic−hydrophobic ratio dimethylpyrazine
within the sample. Other major sources of matrix effects were pyrrole 212.0*** 121.29*** 82.26***
also previously reported, such as the retentive effect of the 2-acetylpyrrole 8.53* 7.37* 5.38*
nonvolatile fraction and the volatile fraction composition.41,42 2-acetyl-1-methylpyrrole 0.004 6.05 2.74
Precision and Accuracy. Repeatability and reproducibility quinoline 2.74 4.29 18.12
were evaluated by relative standard deviations. To evaluate the 2-methylquinoline 0.002 2.07 0.40
repeatability, 10 identical spiked samples (wood and spirit, at 6-methylquinoline 40.07*** 1.88 5.12
50 μg/L) were analyzed. The coefficients obtained were below a
Asteriks *, **, and *** indicate significant differences at p values
10% for all the compounds. Concerning reproducibility for <0.05, <0.0,1 and <0.001, respectively.
each matrix, the relative standard deviations were calculated for
12 independent assays, run under the same analytical tryptophan under roasting conditions, their formation is
conditions and over a 1 month period. Results deviations assumed to proceed via alkylated indoles by ring enlargement
were below 20%, ranging from 9.1% (2-acetyl-3,5-dimethylpyr- reactions or intramolecular cyclizations. To our knowledge, no
azine) to 18.9% (2,3,5,6-tetramethylpyrazine) in the wood quantitative data for the current set of N-heterocyclic markers
sample (Table 2) and from 12.2% (2-acetyl-3,6-dimethylpyr- had been published in wood. Their average concentrations
azine) to 18.7% (pyrrole) in the spirit sample (Table 3). The ranged from 2.04 to 14.51 μg/L (corresponding to 0.04 to 0.29
accuracy of the analytical method was evaluated by calculating μg/g of dry wood) for alkylpyrazines (2,6-dimethylpyrazine
the individual analyte recoveries at three different spiking levels and 5-ethyl-2,3-dimethylpyrazine), 0.79 to 4.69 μg/L (i.e.,
in each matrix. The recovery rate of the method was 0.016 to 0.094 μg/g of dry wood) for acetylpyrazines, 0.62 to
satisfactory since it ranged from 70.5 to 120.5% for all the 9.27 μg/L (i.e., 0.012 to 0.19 μg/g of dry wood) for pyrrole,
studied compounds in wood (Table S2) and from 78.4 to 9.09 to 41.77 μg/L (i.e., 0.18 to 0.84 μg/g of dry wood) to
121.6% in spirit (Table S3). acetylpyrroles derivatives and from 10.85 to 18.52 μg/L (i.e.,
Proof Applicability: Quantitation of Nitrogen-Heter- 0.22 to 0.37 μg/g of dry wood) for quinolines (Figure 4). The
ocyclic Compounds in Wood and Spirit Matrices. analysis of the variance pointed out the influence of the wood
Influence of Oak Species and Toasting Conditions on N- species and its toasting level on many N-heterocyclic markers.
Heterocycles Profiling in Wood. Thirteen N-heterocyclic Concerning the wood effect, significantly higher contents of 5-
compounds were detected in oakwood samples. 2,3,5,6- ethyl-2,3-dimethylpyrazine, 2-acetyl-3,5-dimethylpyrazine, 2-
Tetramethylpyrazine was found at lower concentrations than acetyl-3,6-dimethylpyrazine, pyrrole, 2-acetylpyrrole, and 6-
its LOQ, and thus, 12 compounds were quantified (Table 4). methylquinoline were found in French oak compared to
Albeit previous studies had reported the presence of some American oak. Moreover, these results highlighted that the
alkylpyrazines in oak and wood smoke,9,25 neither 2,3,5,6- toasting level also significantly impacted on total N-
tetramethylpyrazine and 5-ethyl-2,3-dimethylpyrazine nor 2- heterocyclic aroma compounds concentrations (Figure 4).
acetyl-1-methylpyrrole and quinolines derivatives were identi- More specifically, alkylpyrazines, acetylpyrazines, pyrrole, and
fied so far in the wood matrix. All the studied N-heterocycles acetylpyrrole contents increased with the toasting level,
originated from the Maillard reaction, which is one of the most whereas no particular trend was observed for quinolines in
important routes to aromas in heated foods. Typically, in the our conditions. The components mainly affected by the
Maillard reaction, the carbonyl group of a reducing sugar reacts toasting degree were 2,6-dimethylpyrazine, 5-ethyl-2,3-dime-
with the amino group of an amino acid, leading to a complex thylpyrazine, 2-acetylpyrazine, and pyrrole (p value <0.001),
network of further chemical reactions and the release of followed by 2-acetyl-3,6-dimethylpyrazine (p value <0.01) and
multiple volatile aroma compounds. More particularly, after 2-acetylpyrrole (p value <0.05) (Table 4). These results
the initial Maillard reaction step, pyrazines and pyrroles were suggested that the toasting intensity had a clear effect on the
issued from reductones via Strecker degradation and Amadori level of nitrogen-heterocyclic compounds in wood, with
rearrangement, followed by further various condensation and medium and heavy toasting contributing to increase the
cyclisation mechanisms.43,44 Baltes and Knoch (1993)45 occurrence of the assayed volatile aroma compounds. These
proposed, meanwhile, a synthesis pathway for alkylquinolines findings were in agreement with other results applied to food
formation. After an initial Maillard reaction between sugars and analysis and covering the same range of temperature. It was
13701 DOI: 10.1021/acs.jafc.9b05716
J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

Figure 4. Distribution of nitrogen-heterocyclic compound concentrations in wood samples according to their origin species (French vs American)
and toasting level (light, medium, and heavy). Alkylpyrazines group includes the two pyrazines quantified in the wood matrix (i.e. 2,6-
dimethylpyrazine and 5-ethyl-2,3-dimethylpyrazine). Letters indicate significant differences at p value <0.05.

Table 5. Mean Concentrations (μg/L, ± SD) of Targeted Nitrogen-Heterocyclic Compounds in the Series of 41 Commercial
Spirits Analyzeda
kind of spirit maturity alkylpyrazines acetylpyrazines pyrrole acetylpyrroles quinolines
Armagnac (n = 1) non matured 11.7 ± 1.6 12.4 ± 2.1 n.d. 23.7 ± 5.6 3.4 ± 0.7
brandy (n = 2) non matured 12.7 ± 2.6 0.5 ± 0.2 0.7 ± 0.3 5.9 ± 1.3 4.3 ± 1.6
rum (n = 2) non matured 27.8 ± 2.6 7.6 ± 2.6 n.d. 4.5 ± 0.3 2.1 ± 0.6
whisky (new-make malt; n = 1) non matured 16.8 ± 0.1 5.1 ± 0.4 7.8 ± 0.3 22.7 ± 2.2 5.1 ± 0.8
mean concentrations in non matured spirits (n = 17.3 ± 2.7 7.2 ± 2.1 b 4.3 ± 0.3 b 13.3 ± 2.3 b 5.1 ± 0.8
6)
Armagnac (n = 5) matured (15−48 11.9 ± 3.1 27.1 ± 4.2 32.6 ± 8.6 35.9 ± 9.9 7.2 ± 3.7
years)
brandy (n = 1) matured (7 years) 6.1 ± 0.4 1.6 ± 2.6 13.7 ± 4.3 9.3 ± 2.5 10.5 ± 4.1
Cognac (n = 3) matured (3−10 years) 11.3 ± 2.0 29.8 ± 4.3 8.0 ± 3.1 17.4 ± 5.6 1.6 ± 0.5
rum (n = 8) matured (1−12 years) 24.6 ± 3.3 35.1 ± 9.5 39.2 ± 9.7 20.5 ± 7.5 7.9 ± 3.6
whisky (n = 18) matured (3−21 years) 12.7 ± 2.4 39.3 ± 10.7 40.7 ± 12.1 45.2 ± 12.7 16.9 ± 5.8
mean concentrations in matured spirits (n = 35) 13.3 ± 2.3 26.3 ± 6.6 a 26.8 ± 7.5 a 25.7 ± 7.9 a 8.6 ± 3.6
a
n.d.: not detected. Letters indicate significant differences between the two clusters (non matured and matured spirits) at p value <0.05.

assumed that temperature accounts among the key factors formation rate of the resulted aroma compounds. Accordingly,
conditioning the kinetics of the Maillard reaction and thus the previous studies related to roasted coffee or sesame seeds46,47
13702 DOI: 10.1021/acs.jafc.9b05716
J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

reported, within a constant roasting time, a continuous grapes; cluster 2: rums made from molasses, and cluster 3:
increase in pyrazines and pyrroles derivatives when the process whiskies made from barley). Because the former results of the
temperature was increased up to 250 °C. Strong interactions current study had shown that these N-heterocycles were
between the wood species and the toasting degree were also aromatic chemical markers also well present in oak wood, these
observed for pyrrole (p value <0.001), 5-ethyl-2,3-dimethyl- latter findings highlighted the essential role played by the
pyrazine and 2-acetyl-3,6-dimethylpyrazine (p value <0.01), maturation step in modifying their content in spirits by
and 2-acetylpyrrole (p value <0.05) (Table 4). This suggests transferring them from the oak wood barrel to the raw distillate
that the relationship between increasing the toasting level and over the aging period.
the concentration of N-heterocyclic volatile aroma compounds The current-optimized SPME-GC/MS methodology was
also depends on the type of wood since medium or heavy able to quantify in both wood and spirit matrices and in a
toasted-French oak shows a higher release of these molecular single chromatographic run and at low level, twenty-one
markers compared to American oak. In the basic fraction of a nitrogen-heterocyclic compounds characterized by roasted-
charred American oak wood (Quercus alba), Maga (1987)23 and nutty-derived notes. As a result, it was shown that the
identified several pyrazines, which are considered as important barrel aging process played an important role in the
sensory contributors. He also postulated that these compounds modulation of the N-heterocycle profiling in both wood and
were generated from nitrogen-containing macromolecules by spirit extracts since the geographical origin and the toasting
the charring process and could be transferred into alcoholic level of wood as well as the spirit maturation in woods barrels
beverages over the maturation period in wood barrels. Taking significantly affected the contents of alkylpyrazines, acetylpyr-
into account the recent identification of melanoidins in oak azines, pyrrole, and acetylpyrroles in spirit samples. Never-
wood,48 the difference currently observed between American theless, further research work is still required to better clarify
and French oakwood species, on the release of nitrogen the impact of the barrel toasting process on nitrogen-
heterocycles after the toasting step, might be thus related to heterocyclic compound profiling at different depths in wood,
their initial content in available nitrogen precursors. as well as the evolution of these volatile roasted-aroma
N-Heterocycle Profiling in Commercial Spirits and compounds in a spirit matured upon various wood barrel
Correlation with Wood Maturation. The set of 21 nitrogen- conditions. With regard to their common chemical structure,
heterocyclic compounds was quantified in a series of 41 non the sensory interactions between the nitrogen heterocylic
matured or matured spirits. The average concentrations of the compounds and their role in enhancing nutty and roasted
volatile aroma compounds are presented in Table 5. Several notes in the spirit aging bouquet represent another area for
alkylpyrazines (such as 2-methylpyrazine, 2,3-dimethylpyra- future investigations.
zine, 2,5-dimethylpyrazine, 2,6-dimethylpyrazine, 2,3,5-trime-
thylpyrazine, 2-ethyl-3-methylpyrazine, 2-ethyl-3,6-dimethyl-
pyrazine, and 2,3,5,6-tetramethylpyrazine) were previously

*
ASSOCIATED CONTENT
S Supporting Information
detected in whisky and rum and quantified from few to The Supporting Information is available free of charge at
several hundred micrograms per liter (μg/L).17,30 Moreover, https://pubs.acs.org/doi/10.1021/acs.jafc.9b05716.
quinoline 2- and 6-methylquinolines were previously identified Sample and SPME parameters studied during the
in whisky.49 However, to our knowledge, it was the first time optimization stage of the analytical method are
that 2-ethyl-3,5-dimethylpyrazine, 5-ethyl-2,3-dimethylpyra- presented in Table S1, N-heterocyclic volatile aroma
zine, 2,3-diethyl-5-methylpyrazine, acetylpyrazine, and acetyl- compounds recoveries obtained in a light-toasted
pyrrole derivatives were reported in spirit beverages. oakwood and non matured spirit using the optimized
Alkylpyrazines, acetylpyrazines, pyrrole, acetylpyrroles, and SPME-GC/MS is presented in Tables S2 and S3,
quinolines were detected in non matured spirits with global respectively (PDF)


concentrations ranging from 7.3 μg/L for pyrrole to 17.3 μg/L
for alkylpyrazines. Within the non matured sample cluster,
AUTHOR INFORMATION
alkylpyrazines and pyrrole were found in higher concentrations
in rums and new-make whiskies. As a consequence, the impact Corresponding Author
of the raw material could not be excluded to explain their *E-mail: picard@demptos.fr. Phone: +33 (0) 5 40 00 38 01.
occurrence in the final spirit. For the other assayed ORCID
compounds, no significant differences were observed according Magali Picard: 0000-0001-9083-8250
to the kind of spirit. This above finding suggested that some of Notes
them could either be initially present in the raw material or The authors declare no competing financial interest.


formed at high temperature from precursors during the spirit
production process and prior to barrel aging (that is, during ABBREVIATION USED
the distillation step in the spirit production process, malting HS-SPME, headspace solid-phase microextraction; GC/MS,
and/or peating step(s) in the whisky elaboration).10,50 gas chromatography−mass spectrometry; LOD, limit of
Interestingly, the ANOVA test applied to mean N-heterocycle detection; LOQ, limit of quantification


concentrations in the two global spirit clusters (matured vs non
matured in wood barrels) revealed that matured spirits had REFERENCES
significantly higher concentrations of acetylpyrazines, pyrrole,
(1) Maga, J. A. The contribution of wood to the flavor of alcoholic
and acetylpyrroles than non matured ones (p value <0.05; beverages. Food Rev. Int. 1989, 5, 39−99.
Table 5). The same significant trend (p value <0.05) between (2) Mosedale, J. R.; Puech, J. L. Wood maturation of distilled
non matured and matured samples was observed when spirits beverages. Trends Food Sci Technol. 1998, 9, 95−101.
were clustered according to the nature of the raw material (that (3) Jolly, N. P.; Hattingh, S. A brandy aroma wheel for South African
is, cluster 1: Armagnacs, brandies, and Cognacs made from brandy. S. Afr. J. Enol. Vitic. 20101, 22, 16−21.

13703 DOI: 10.1021/acs.jafc.9b05716


J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

(4) Lee, K. Y. M.; Paterson, A.; Piggott, J. R.; Richardson, G. D. (27) Maga, J. A. Oak lactones in alcoholic beverages. Food Rev. Int.
Origins of flavour in whiskies and a revised flavour wheel: a review. J. 1996, 12, 105−130.
Inst. Brew. 2001, 107, 287−313. (28) Mosedale, J. R. Effects of oak wood on the maturation of
(5) Mason, M. E.; Johnson, B.; Hamming, M. Flavor components of alcoholic beverages with particular reference to whisky. Forestry 1995,
roasted peanuts. Some low molecular weight pyrazines and pyrrole. J. 68, 203−230.
Agric. Food Chem. 1966, 14, 454−460. (29) Wardencki, W.; Michulec, M.; Curyło, J. A review of theoretical
(6) Aprotosoaie, A. C.; Luca, S. V.; Miron, A. Flavor chemistry of and practical aspects of solid-phase microextraction in food analysis.
cocoa and cocoa products - an overview. Compr. Rev. Food Sci. Food Int. J. Food Sci. Tech. 2004, 39, 703−717.
Saf. 2016, 15, 73−91. (30) Miller, G. H. Malting. In Whisky Science; Springer: Cham,
(7) Caporaso, N.; Whitworth, M. B.; Cui, C.; Fisk, I. D. Variability of Switzerland, 2019, 83−119.
single bean coffee volatile compounds of Arabica and robusta roasted (31) Wu, J. F.; Xu, Y. Comparison of pyrazine compounds in seven
coffees analysed by SPME-GC-MS. Food Res. Int. 2018, 108, 628− Chinese liquors using headspace solid-phase micro-extraction and
640. GC-nitrogen phosphourus detection. Food Sci. Biotechnol. 2013, 22,
(8) Maga, J. A. Pyrroles in foods. J. Agric. Food Chem. 1981, 29, 1−6.
691−694. (32) Balasubramanian, S.; Panigrahi, S. Solid-phase microextraction
(9) Maga, J. A. Pyrazine update. Food Rev. Int. 1992, 8, 479−558. (SPME) techniques for quality characterization of food products: a
(10) Wang, P. S.; Kato, H.; Fujimaki, M. Studies on flavor review. Food Bioprocess Tech. 2011, 4, 1−26.
components of roasted barley. Part III. The major volatile basic (33) Shrivastava, A.; Gupta, V. B. Methods for the determination of
compounds. Agric. Biol.Chem. 1969, 33, 1775−1781. limit of detection and limit of quantitation of the analytical methods.
(11) Boothroyd, E.; Linforth, R. S. T.; Jack, F.; Cook, D. J. Origins of Chron. Young Sci. 2011, 2, 21.
the perceived nutty character of new-make malt whisky spirit. J. Inst. (34) Spietelun, A.; Pilarczyk, M.; Kloskowski, A.; Namieśnik, J.
Brew. 2014, 120, 16−22. Current trends in solid-phase microextraction (SPME) fibre coatings.
(12) Vitzthum, O. G.; Werkhoff, P.; Hubert, P. New volatile Chem. Soc. Rev. 2010, 39, 4524−4537.
constituents of black tea aroma. J. Agric. Food Chem. 1975, 23, 999− (35) Burin, V. M.; Marchand, S.; de Revel, G.; Bordignon-Luiz, M.
1003. T. Development and validation of method for heterocyclic
(13) Demole, E.; Berthet, D. A Chemical Study of Burley Tobacco compounds in wine: optimization of HS-SPME conditions applying
Flavour (Nicotiana tabacum L.). I. Volatile to medium-volatile a response surface methodology. Talanta 2013, 117, 87−93.
constituents bp≤ 84°/0.001 Torr. Helv. Chim. Acta 1972, 55, 1866− (36) Lorenzo, J. M. Influence of the type of fiber coating and
1882. extraction time on foal dry-cured loin volatile compounds extracted by
(14) Nishimura, K.; Masuda, M. Minor constituents of whisky fusel solid-phase microextraction (SPME). Meat Sci. 2014, 96, 179−186.
oils. 1. Basic, phenolic and lactonic compounds. J. Food Sci. 1971, 36, (37) Slaghenaufi, D.; Perello, M. C.; Marchand, S.; de Revel, G.
819−822. Quantification of megastigmatrienone, a potential contributor to
(15) Boothroyd, E. L. Investigation of the congeners responsible for tobacco aroma in spirits. Food Chem. 2016, 203, 41−48.
nutty/cereal aroma character in new make malt whisky; University of (38) Sala, C.; Mestres, M.; Marti, M. P.; Busto, O.; Guasch, J.
Nottingham, 2013. Headspace solid-phase microextraction analysis of 3-alkyl-2-methox-
(16) Kahn, J. H. Compounds identified in whisky, wine, and beer: a ypyrazines in wines. J. Chromatogr. A 2002, 953, 1−6.
tabulation. J. Assoc. Off. Anal. Chem. 1969, 52, 1166−1178. (39) Pawliszyn, J. Theory of solid-phase microextraction. In
(17) Worben, H. J.; Timmer, R.; Ter Heide, R.; De Valois, P. J. Handbook of Solid Phase Microextraction; Elsevier: 2012, 13−59,
Nitrogen compounds in rum and whiskey. J. Food Sci. 1971, 36, 464− DOI: 10.1016/B978-0-12-416017-0.00002-4.
465. (40) Mestres, M.; Marti, M. P.; Busto, O.; Gausch, J. Analysis of low-
(18) Léauté, R. Distillation in alambic. Am. J. Enol. Vitic. 1990, 41, volatility organic sulphur compounds microextraction and gas
90−103. chromatography. J. Chromatogr. A 2000, 881, 583−590.
(19) Benn, S. M.; Peppard, T. L. Characterization of tequila flavor by (41) Muñoz-González, C.; Martín-Á lvarez, P. J.; Moreno-Arribas, M.
instrumental and sensory analysis. J. Agric. Food Chem. 1996, 44, V.; Pozo-Bayón, M. Á . Impact of the nonvolatile wine matrix
557−566. composition on the in vivo aroma release from wines. J. Agric. Food
(20) Jouret, C.; Pace, E.; Parfait, A. Analytical differentiation of Chem. 2013, 62, 66−73.
agricultural and industrial rums using alkylpyrazines. Ann. Falsif. (42) Robinson, A. L.; Ebeler, S. E.; Heymann, H.; Boss, P. K.;
Expert. Chim. Toxicol. 1994, 87, 85−90. Solomon, P. S.; Trengove, R. D. Interactions between wine volatile
(21) Fan, W.; Xu, Y.; Zhang, Y. Characterization of pyrazines in compounds and grape and wine matrix components influence aroma
some Chinese liquors and their approximate concentrations. J. Agric. compound headspace partitioning. J. Agric. Food Chem. 2009, 57,
Food Chem. 2007, 55, 9956−9962. 10313−10322.
(22) Chatonnet, P.; Cutzach, I.; Pons, M.; Dubourdieu, D. (43) Van Boekel, M. A. J. S. Formation of flavour compounds in the
Monitoring toasting intensity of barrels by chromatographic analysis Maillard reaction. Biotechnol. Adv. 2006, 24, 230−233.
of volatile compounds from toasted oak wood. J. Agric. Food Chem. (44) Vernin, G.; Parkanyi, C. Mechanisms of formation of
1999, 47, 4310−4318. heterocyclic compounds in Maillard and pyrolysis reactions. In The
(23) Maga, J. A. The flavor chemistry of wood smoke. Food Rev. Int. chemistry of heterocyclic flavoring and aroma compounds: Horwood
1987, 3, 139−183. Publishers: Chichester Ellis, United-Kingdom, 1982, 151−207.
(24) Kim, K.; Kurata, T.; Fujimaki, M. Identification of flavor (45) Baltes, W.; Knoch, E. Model reactions on roast aroma
constituents in carbonyl, non-carbonyl neutral and basic fractions of formation. XIII. The formation of some uncommon N-heterocyclic
aqueous smoke condensates. Agric. Biol. Chem. 2014, 38, 53−63. compounds and furans after roasting of tryptophan with reducing
(25) Maga, J. A.; Chen, Z. Pyrazine composition of wood smoke as sugars and sugar degradation products. Food Chem. 1993, 46, 343−
influenced by wood source and smoke generation variables. Flavour 349.
Fragr. J. 1985, 1, 37−42. (46) Silwar, R.; Lüllmann, C. Investigation of aroma formation in
(26) Nonier, M. F.; Vivas, N.; Vivas De Gaulejac, N.; Absalon, C.; Robusta coffee during roasting. Café Cacao Thé 1993, 37, 145−152.
Soulié, P.; Fouquet, E. Pyrolysis−gas chromatography/mass spec- (47) Xu-Yan, D.; Ping-Ping, L.; Fang, W.; Mu-lan, J.; Ying-Zhong,
trometry of Quercus sp. wood: application to structural elucidation of Z.; Guang-Ming, L.; Hong, C.; Yuan-Di, Z. The impact of processing
macromolecules and aromatic profiles of different species. J. Anal. on the profile of volatile compounds in sesame oil. Eur. J. Lipid Sci.
Appl. Pyrol. 2006, 75, 181−193. Technol. 2012, 114, 277−286.

13704 DOI: 10.1021/acs.jafc.9b05716


J. Agric. Food Chem. 2019, 67, 13694−13705
Journal of Agricultural and Food Chemistry Article

(48) Nonier, M. F.; Vivas, N.; de Gaulejac, N. V.; Mouche, C.;


Huguet, C. R.; Daugey, N. Purification and partial characterization of
melanoidins fractions from toasted oak heartwood, comparison with
melanoidins from roasted coffee. J. Food Res. 2018, 7, 37−58.
(49) Viro, M. Heterocyclic nitrogen compounds in whisky and beer.
Chromatographia 1984, 19, 448−451.
(50) Deki, M.; Yoshimura, M. Studies on the volatile components of
peated malt. III. Identification of acidic and basic components. Chem.
Pharm. Bull. 1974, 22, 1760−1764.

13705 DOI: 10.1021/acs.jafc.9b05716


J. Agric. Food Chem. 2019, 67, 13694−13705

You might also like