You are on page 1of 10

Food Research International 140 (2021) 110049

Contents lists available at ScienceDirect

Food Research International


journal homepage: www.elsevier.com/locate/foodres

Monitoring the evolution of free and cysteinylated aldehydes from malt to


fresh and forced aged beer
P. Bustillo Trueba a, *, 1, B. Jaskula-Goiris a, 1, M. Ditrych a, W. Filipowska a, J. De Brabanter a,
G. De Rouck a, G. Aerts a, L. De Cooman a, J. De Clippeleer b, c
a
KU Leuven, Department of Microbial and Molecular Systems (M2S), Cluster for Bioengineering Technology (CBeT), Laboratory of Enzyme, Fermentation and Brewing
Technology (EFBT), Technology Campus Ghent, Gebroeders De Smetstraat 1, B-9000 Ghent, Belgium
b
Innovation centre for Brewing & Fermentation – IBF, Ghent University, Faculty of Bioscience Engineering, Department of Biotechnology, Valentin Vaerwyckweg 1, B-
9000 Ghent, Belgium
c
Innovation centre for Brewing & Fermentation – IBF, HOGENT University of Applied Sciences and Arts, Department of Life Sciences and Industrial Technology, Research
Centre AgroFoodNature, Valentin Vaerwyckweg 1, B-9000 Ghent, Belgium

A R T I C L E I N F O A B S T R A C T

Keywords: During storage, beer staling coincides with a gradual increase in the concentrations of aldehydes resulting in the
Cysteinylated aldehydes (2-substituted 1,3- appearance of undesirable flavours. Cysteinylated aldehydes, also referred to as 2-substituted 1,3-thiazolidine-4-
thiazolidine-4-carboxylic acids) carboxylic acids, have been proposed as potential precursors of this increase. This study aimed to further un­
Free aldehydes
derstand the origin of aldehydes in aged beer, by monitoring both free and cysteinylated aldehydes throughout
Quantification
Flavour instability
the brewing process, from the raw materials until the stored product. Quantification of free and cysteinylated
Malt aldehydes was performed for two different brews via headspace solid phase microextraction-gas chromatog­
Brewery samples raphy-mass spectrometry (HS-SPME-GC-MS) and ultra-high performance liquid chromatography-mass spec­
trometry (UHPLC-MS), respectively. All selected marker aldehydes were quantified in malt, wort, and the
resulting fresh and aged beer samples. Cysteinylated aldehydes were quantifiable in malt and up to the wort
boiling phase. The highest levels of free aldehydes were found in malt, whereas cysteinylated aldehydes showed
highest levels at mashing-in pointing to their formation during both malting and subsequent mashing-in. During
beer ageing, an increase in all free aldehydes was measured. In particular, a rise in 2-methylpropanal and furfural
is most striking. Although the presented experimental data obtained on malt and brewery samples do support the
concept of bound-state aldehydes, cysteinylated aldehydes cannot be consider as the cause of increasing levels of
staling aldehydes during beer ageing.

1. Introduction beer is characterised by a limited shelf life (Bamforth, 2011; Rodrigues


et al., 2011; Saison et al., 2009; Stewart, 2016). The chemical changes
The majority of food and beverage products deteriorate in their during beer ageing have been studied with regard to both the volatile
sensory properties over time, in particular flavour. Flavour instability is and non-volatile fractions, where changes in the concentrations of
a complex and highly relevant topic that has been affecting the food and different compounds have been identified, e.g. aldehydes, bitter acids,
beverage industry for decades (Kilcast & Subramaniam, 2011; Weer­ proteins, esters, etc. (Baert et al., 2012; Liégeois et al., 2002; Malfliet
awatanakorn et al., 2015). The flavour of food and beverages is expected et al., 2008; Vanderhaegen et al., 2006; Yu et al., 2014). Nevertheless,
to be pleasant, characteristic for the given product and consistent within beer staling has mainly been described by the loss of its pleasant
its expected shelf life to meet consumers’ expectations (Paternoster bitterness, due to iso-α-acids degradation, and the appearance of staling
et al., 2020; Stewart et al., 2018). Determining flavour stability is of off-flavours attributed to free aldehydes, due to their very low flavour
utmost importance for the entire food and beverage sector. In particular, threshold values (Caballero et al., 2012; Lehnhardt et al., 2018; Saison

* Corresponding author.
E-mail addresses: paula.bustillotrueba@kuleuven.be (P. Bustillo Trueba), barbara.jaskulagoiris@kuleuven.be (B. Jaskula-Goiris), maciej.ditrych@kuleuven.be
(M. Ditrych), weronika.filipowska@kuleuven.be (W. Filipowska), jos.debrabanter@kuleuven.be (J. De Brabanter), gert.derouck@kuleuven.be (G. De Rouck),
guido.aerts@kuleuven.be (G. Aerts), luc.decooman@kuleuven.be (L. De Cooman), jessika.declippeleer@ugent.be (J. De Clippeleer).
1
Shared co-first/equal authorship with Barbara Jaskula-Goiris.

https://doi.org/10.1016/j.foodres.2020.110049
Received 24 June 2020; Received in revised form 14 December 2020; Accepted 16 December 2020
Available online 8 January 2021
0963-9969/© 2020 Elsevier Ltd. All rights reserved.
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

et al., 2009; Vanderhaegen et al., 2006). Although very low amounts of 89-0), ethanol (≥99.2%, CAS 64-17-5), dehydrated calcium chloride
aldehydes are already found in fresh beer, the appearance of charac­ (CaCl2 ≥ 95.0%, CAS 10043-52-4), and (L)-Lactic acid (≥90.0%, CAS
teristic beer staling on storage coincides with their increase in levels 50-21-5) were obtained from Merck (Darmstad, Germany). Ultrapure
(Hashimoto, 1966). Several more recent studies also point to aldehydes water was obtained by a Synergy 185 system from Millipore (mQ, 18.2
and their potential role as the major source of beer staling (Baert et al., MΩ⋅cm at 25 ◦ C) (Molsheim, France). LCMS/MS grade acetonitrile was
2018; Gibson et al., 2018; Jaskula-Goiris et al., 2019; Lehnhardt et al., obtained from Biosolve Chemie (Dieuze, France). Formic Acid (≥98%,
2018; Wietstock et al., 2016). Moreover, staling aldehydes still emerge CAS 64-18-6) LCMS/MS grade was acquired from Merck (Darmstad,
in beers despite of improved storage conditions, such as lightproof Germany). Nitrogen (N28), helium (P0252), and methane gas (N46)
packaging, exclusion of oxygen to a minimum and low temperature were obtained from Air Liquide (Liège, Belgium). The derivatization
preservation (Fratianni, 2001; Jaskula-Goiris et al., 2011; Malfliet et al., reagent o-(2,3,4,5,6-pentafluorobenzyl)hydroxylamine (PFBHA, ≥ 99%,
2008). Various pathways for the appearance of aldehydes have been CAS 57981-02-9) was obtained from Sigma-Aldrich (St.-Louis, MO,
proposed (Baert et al., 2012), however, despite extensive previous USA). Deuterated 2-methylbutanal (2-methylbutanal-d10, 100 atom %D,
research, elucidation of the mechanism of the aldehyde increase during ≈200 mg in 1.5 mL of CDCl3) was synthesized upon ordering by Mer­
beer ageing has not yet been accomplished. caChem (Nijmegen, the Netherlands. Deuterated benzaldehyde (benz­
Aldehydes are highly reactive molecules, thus, able to react with a aldehyde-d6, ≥ 98%, 98 atom %D, CAS 17901-93-8) was purchased
variety of compounds, e.g. amino acids, proteins, and sulphur com­ from Sigma-Aldrich (St. Louis, MO, USA).
pounds (Solomons et al., 2013). In particular the reaction of aldehydes
with cysteine and/or bisulphite leading to the formation of bound-state 2.2. Standard malt analyses
aldehydes has recently been proposed as a potential source of the
aldehyde increase during beer ageing as a result of subsequent aldehyde Four industrial malts, two Pilsner malts (malt A (Fermentis, Marcq-
release from their bound state(s) (Baert et al., 2018, 2015a, 2015b; en-Barœul, France) and malt B (Holland malt, The Netherlands)), and
Kaneda et al., 1996; Lehnhardt et al., 2018). As the presence of cysteine two specialty malts (Munich and Biscuit (Dingemans, Stabroek,
and bisulphite in intermediate brewery products is well-known (Baert Belgium)) were used in this study. Evaluation of malt quality parameters
et al., 2015a; Guido, 2016; Matsui & Kitabatake, 1984), the formation of comprised different analyses according to the EBC (Analysis committee
a bound-state between the aldehyde and cysteine or bisulphite is feasible of the EBC, 2018) and ASBC (ASBC Methods of Analysis, 2004): malt
(Baert et al., 2018, 2015a, 2015b, 2012). Due to their non-volatile moisture (EBC 4.1), extract (EBC 4.4), extract difference (EBC 4.5), pH
character, these bound-state aldehydes would neither be removed by (EBC 4.6), visual colour (EBC 4.7), soluble nitrogen (EBC 4.9), total
evaporation during wort boiling, nor be reduced by yeast during protein (ASBC malt chapter 8A), photometric colour (EBC 8.3), boiling
fermentation (Saison et al., 2010), and, as a result, they may end up in wort colour (EBC 8.3.2), and free amino nitrogen (FAN, EBC 8.8). The
the final beer. Subsequently, during beer ageing, release of staling al­ malt standard analyses were carried-out using the Anton Paar DMA
dehydes from their bound-state can be expected, due to beer charac­ 5000 Alcolyser (Anton Paar, Austria), Anton Paar Automated Micro
teristics, i.e. pH of the final beer (Bustillo Trueba et al., 2018a, 2018b), Viscometer (Anton Paar, Austria), UV–VIS spectrophotometer Cary 100
or beer storage conditions, e.g. storage at increased temperature (Fra­ UV–Vis (Agilent Technologies, Santa Clara, USA), Kjeldahl Digest Sys­
tianni, 2001; Jaskula-Goiris et al., 2019). tem K-437 (Buchi, Switzerland), and Kjeldahl Distillation Unit K-350
The aim of the present study is to assess for the first time the potential (Buchi, Switzerland). Malt quality parameters are summarised in
relevance of bound-state aldehydes, in particular cysteinylated alde­ Table 2.
hydes, in relation to beer flavour instability by monitoring the evolution
of free and cysteinylated aldehydes in malt, throughout the brewing, 2.3. Preparation of pilot-scale beers
fermentation and maturation processes, as well as in the resulting fresh
and forced aged beer samples. Two different experimental brews were prepared in the 5 hL pilot
brewery (Meura, Belgium) of KU Leuven Technology Campus Ghent
2. Material and methods (Ghent, Belgium): the brew of a top fermented blond beer (beer A) and a
top fermented amber beer (beer B). The applied mashing ratio was 2.2 L
2.1. Chemicals brewing water/kg milled malt for both beers. For the top fermented
beer, 87 kg of Pilsner malt A (Fermentis, Marcq-en-Barœul, France), and
The carbonyl compounds 2-methylpropanal (2MP ≥ 99%, CAS 78- for the amber beer, 54 kg of Pilsner malt B1 (Holland malt, The
84-2), 2-methylbutanal (2MB ≥ 95%, CAS 96-17-3), 3-methylbutanal Netherlands), 25 kg of Munich malt2 (Dingemans, Stabroek, Belgium)
(3MB ≥ 98%, CAS 590-86-3), methional (METH ≥ 95%, CAS 3268-49- and 8 kg of Biscuit malt (Dingemans, Stabroek, Belgium), were respec­
3), phenylacetaldehyde (PHEN ≥ 98%, CAS 122-78-1), hexanal (HEX tively mixed with 1.9 hL of reverse osmosis water. Brewing water was
≥ 98%, CAS 66-25-1), and furfural (FURF ≥ 99%, CAS 98-01-1) were enriched with 80 mg/L Ca2+ (CaCl2). The same mashing protocol was
purchased from Acros Organics (Geel, Belgium). The amino acid L- applied for both brews: 30 min at 63 ◦ C; 15 min at 72 ◦ C; 1 min at 78 ◦ C,
cysteine (CYS ≥ 98%, CAS 52-90-4) was acquired from Sigma Aldrich with a temperature rise of 1 ◦ C/min. The pH of the mash was adjusted to
(St.-Louis, MO, USA). The cysteinylated aldehydes, 2-isopropylthiazoli­ pH 5.3 with lactic acid (30% v/v). The wort was filtered using a mem­
dine-4-carboxylic acid (2MP-CYS ≥ 99%, CAS 14347-75-2), 2-(sec- brane assisted thin bed filter (Meura 2001, Meura, Belgium); a sparging
butyl)thiazolidine-4-carboxylic acid (2MB-CYS ≥ 99%, CAS 1214831- rate of 2.33 L/kg was applied. At the onset of boiling, the wort was
88-5), 2-isobutylthiazolidine-4-carboxylic acid (3MB-CYS ≥ 98%, CAS adjusted to 13 ◦ P (Beer A) and 19 ◦ P (Beer B), respectively. Boiling was
215669-71-9), 2-(2-(methylthio)ethyl)thiazolidine-4-carboxylic acid carried out for 60 min. Hopping of the beers was done with hop pellets.
(METH-CYS ≥ 99%, CAS 53943-83-2), 2-benzylthiazolidine-4-carbox­ Beer A was first hopped with the Magnum hop variety (α-acids 13.0%
ylic acid (PHEN-CYS ≥ 94%, CAS 50739-30-5), 2-pentylthiazolidine-4- (w/w); 50.5 g/hL, Geert Clarebout, Heuvelland, Belgium), and second
carboxylic acid (HEX-CYS ≥ 92%, CAS 69588-05-2) and 2-(furan-2-yl) hopping in the whirlpool was done with Tettnanger (α-acids 3.0% (w/
thiazolidine-4-carboxylic acid (FURF-CYS ≥ 99%, CAS 72678-98-9) w); 100 g/hL, Geert Clarebout, Heuvelland, Belgium) and Saaz (α-acids
were synthesized according to Ershov’s methodology (Ershov et al., 2.5% (w/w); 120 g/hL, Geert Clarebout, Heuvelland, Belgium). For beer
2014) as described by (Bustillo Trueba et al., 2019). Table 1 shows the
molecular and structural formula of the compounds under study. Citric
acid monohydrate (≥99.5%, CAS 5949-29-1), boric acid (≥99.5%, CAS 2
Munich malt was not available for analysis due to the complete use during
10043-35-3), tri-sodium phosphate dodecahydrate (≥98%, CAS 10101- the brewing process

2
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

Table 1
Molecular and structural formula of representative free and cysteinylated beer staling aldehydes (the latter also referred to as 2-substituted 1,3-thiazolidine-4-car­
boxylic acids).
Aldehyde Abbreviation Molecular Molecular Structure Cysteinylated aldehyde Abbreviation Molecular Molecular Structure
Formula Formula

2-Methylpropanal 2MP C4H8O 2-isopropylthiazolidine-4- 2MP-CYS C7H13NO2S


carboxylic acid

2-Methylbutanal 2MB C5H10O 2-(sec-butyl) thiazolidine-4- 2MB-CYS C8H15NO2S


carboxylic acid
3-Methylbutanal 3MB C5H10O 2-isobutylthiazolidine-4- 3MB-CYS C8H15NO2S
carboxylic acid

Methional METH C4H8OS 2-(2-(methylthio)ethyl) METH-CYS C7H13NO2S2


thiazolidine-4-carboxylic acid
Phenylacetaldehyde PHEN C8H8O 2-benzylthiazolidine-4- PHEN-CYS C10H11NO2S
carboxylic acid

Furfural FURF C5H4O2 2-(furan-2-yl) thiazolidine-4- FURF-CYS C8H9NO3S


carboxylic acid

Hexanal HEX C6H12O 2-pentylthiazolidine-4- HEX-CYS C9H17NO2S


carboxylic acid
trans-2-Nonenal T2N C9H16O (E)-2-(oct-1-en-1-yl) T2N-CYS C12H21NO2S
thiazolidine-4-carboxylic acid

B, Mosaic (α-acids 12.0% (w/w); 49.9 g/hL, harvest 2017, YCH hops, 2.6. Quantification of free aldehydes via HS-SPME-GC-MS in malt, wort
Yakima, WA, United States) was used as first hopping, and second and beer
hopping was done with Strisselspalt (α-acids 12.0% (w/w); 111.1 g/hL,
harvest 2017, Comptoir Commercial Agricole, Sombreffe, Belgium) at The following aldehydes under study are considered markers for beer
the end of the boil and Aramis (α-acids 5.8% (w/w); 34.5 g/hL, harvest ageing: 2-methylpropanal (2MP), 2-methylbutanal (2MB), 3-methylbu­
2017, Comptoir Commercial Agricole, Sombreffe, Belgium) in the tanal (3MB), methional (METH), phenylacetaldehyde (PHEN), furfural
whirlpool. Wort clarification was performed using a whirlpool. Freshly (FURF), hexanal (HEX), and trans-2-nonenal (T2N). Samples for free
propagated yeast and dry yeast (BE-256, Fermentis, Marcq-en-Barœul, aldehydes analysis were prepared under nitrogen gas in an anaerobic
France) were used in a concentration of 80 g/hL for beer A and B, cabinet (Whitley Anaerobic Workstation A100 model B, Don Whitley
respectively. Fermentation was carried out at 24 ◦ C for 7 days. Lagering Scientific Ltd., West Yorkshire, UK) in order to avoid oxidation. Malt
at 0 ◦ C was performed in kegs for 14 days for beer A, and 22 days for beer extract was prepared by weighing 1 g of finely milled malt in 100 g
B. The beers were filtered using 1 μm cellulose sheets (BECOPAD 350- N2–flushed Milli-Q water in a 100 mL Duran glass bottle, and by sub­
Eaton, Groot-Bijgaarden, Belgium). The beers were carbonated up to sequent agitation for 15 min (200 rpm) at ambient temperature. In order
5.6 g CO2/L and bottled in 25cL brown glass bottles using a 6 head to also avoid light-induced degradation, the bottles were covered with
rotating counter pressure filler with double pre-evacuation and high aluminium foil. After sedimentation, an aliquot of 10 mL of supernatant
pressure injection of hot water before bottle sealing (CIMEC, Italy). of malt extract was transferred to a 20 mL amber glass headspace vial
During the entire brewing process, samples were taken for the quanti­ and sealed with a magnetic bimetal crimp cap prior to the analysis.
fication of free and cysteinylated aldehydes at the following steps: onset Considering brewery samples (wort, fresh and aged beer samples), an
of mashing (sample n◦ 1), end of mashing (2), wort filtration (3), onset of aliquot of 4 mL was transferred to a 20 mL amber vial and sealed with a
boiling (4), end of boiling (5), wort clarification (6), wort cooling (7), magnetic bimetal crimp cap for analysis. Wort samples (from onset
onset of fermentation (after 24 h) (8), onset of maturation (9), end of mashing up to pitching wort (sample points 1–8, see section 2.3)) and
maturation (before beer filtration) (10), and freshly bottled beer (11). samples taken during fermentation (sample points 9–10, see section 2.3),
were 10-fold diluted with N2-flushed Milli-Q water prior to transfer.
2.4. Forced ageing of beer samples Before transfer to the GC vials, fresh and aged beers were initially
degassed inside a nitrogen atmosphere cabinet by pouring out in a glass
Bottled beers were forced aged up to 3 months in the dark at 30 ◦ C in recipient. Calibration curves based on N2 flushed Milli-Q water in the
a thermostatically controlled room. The following beer samples were case of malt and wort, and N2 flushed Milli-Q water with addition of
taken during storage at different time intervals: 1 week (sample n◦ 12), 2 ethanol (in accordance with the alcohol percentage of the beer analysed)
weeks (13), 1 month (14), 2 months (15), and 3 months (16). in the case of beer, were used for quantification of all samples. Linear
calibration fittings with correlation coefficients R ≥ 0.9900 were ob­
2.5. Standard beer analyses tained for the assessment of free aldehydes concentrations. The free
aldehydes were then quantified according to Baert et al. (2015a, 2015b),
Aiming at basic characterisation of the obtained pilot-scale beers, using headspace solid-phase microextraction gas chromatography-mass
alcohol, density, apparent, real and original extract, as well as apparent spectrometry (HS-SPME-GC-MS).
and real degree of fermentation were determined with the DMA 5000
Alcolyser (Anton Paar, Austria). Moreover, the pH of the beers and the
2.7. Quantification of cysteinylated aldehydes via LC-MS in malt, wort
oxygen levels were determined using a P600 pH-meter (Consort,
and beer
Turnhout, Belgium) and Haffmans Inpack TPO/CO2 Meter c-TPO (Pen­
tair, UK), respectively. Table 3 covers the beer analyses.
The cysteinylated aldehydes, as summarised in Table 1, were quan­
tified in malt, wort and beer samples according to our previously pub­
lished work (Bustillo Trueba et al., 2019), by applying ultra-high-

3
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

Table 2 Table 3
Malt quality parameters and concentrations of free and cysteinylated aldehydes Standard characterisation of fresh pilot-scale top fermented beers A (blond beer)
for two Pilsner malts, A and B1, and the specialty malt B2 Biscuit. and B (amber beer).
Pilsner - Pilsner - Biscuit - Beer A Beer B
malt A malt B1 malt B2
Alcohol (% v/v) 5.50 8.43
Moisture (%) 4.0 4.3 4.2 Density (g/cm3) 1.01 1.01
Extract (% D.M.) 80.9 81.0 78.6 Apparent extract (◦ P) 2.99 2.56
Extract difference 0.1 0.7 1.1 Real extract (◦ P) 5.03 5.48
(%) Original extract (◦ P) 13.33 17.86
Total protein (% D. 9.9 10.0 8.4 App. degree of fermentation (%) 77.61 85.65
M.) Real degree of fermentation (%) 64.32 71.32
Soluble N in wort 75.6 76.1 60.5 O2 (µg/L) 186 106
(mg/100 mL) pH 4.38 4.30
Kolbach index (%) 47.5 42.6 39.7
FAN (mg/L) 136.0 163.0 28.0 Beer quality parameters are expressed as mean values (n = 2).
pH 6.0 6.0 5.6
Visual colour (EBC) 7.0 4.4 47.0 performance liquid chromatography-mass spectrometry (UHPLC-MS).
Photometric colour 5.8 4.3 42.6
For the extraction of cysteinylated aldehydes from malt, 5 g of finely
(EBC)
Congress boiling 11.6 6.8 48.0 milled malt was weighed in 100 g of Milli-Q water, followed by agitation
wort colour (EBC) at 200 rpm for 5 min at ambient temperature (Grant-Bio Multifuncional
Viscosity 10% (cP) 1.5 1.6 1.5 PSU20-I Orbital Shaker, Cambridge, UK). The extraction bottles were
Mean ± Mean ± SD Mean ± SD covered with aluminium foil. Initially, centrifugation was performed at
SD
Aldehydes (µg/kg 2MP 945.8 ± 1003.8 ± 52521.9 ±
9000 rpm during 10 min at 10 ◦ C (Hettich Zentrifugen Universal 320R
DM) 131.5 182.8 3243.1 (Tuttlingen, Germany). Malt extract as well as brewery samples were
2MB 880.4 ± 940.0 ± 46181.0 ± centrifuged for a second time in an Eppendorf centrifuge 5424R (Aar­
89.7 161.4 867.0 schot, Belgium) for 5 min at 10 ◦ C and 11,000 rpm. Samples were
3MB 2143.9 ± 2432.8 ± 56174.1 ±
filtered with a Hamilton Gastight syringe #1001 of 1 mL (PTFE Luer
342.3 496.0 1576.8
METH 189.6 ± 256.7 ± 1018.0 ± Lock) and a SPARTAN HPLC Syringe filter (regenerated cellulose, 13
31.7 33.0 63.0 mm syringe filter, 0.2 μm pore size). An aliquot of 1 mL was transferred
PHEN 1011.6 ± 1139.3 ± 7795.1 ± into a 2 mL amber LC vial closed with a screw cap with PTFE septum.
82.8 150.9 341.3 Calibration curves were prepared by dissolving the authentic synthe­
FURF 316.7 ± 317.5 ± 6331.7 ±
2.1 134.1 616.5
sized cysteinylated aldehydes in an aqueous buffer solution at pH 9.0
HEX 268.3 ± 234.1 ± 3187.8 ± (Carmody, 1961) and by diluting to different concentrations ranging
70.7 4.4 32.4 from 1 to 1000 µg/L (Bustillo Trueba et al., 2019). For each cysteiny­
T2N 131.0 ± 122.5 ± 360.3 ± lated aldehyde linear calibration fittings were obtained. Accepted cor­
43.6 1.0 14.7
relation coefficients were of R ≥ 0.995.
Cysteinylated 2MP-CYS 206.5 ± 172.5 ± 148.0 ± 4.9
aldehydes (µg/kg 26.0 17.4
DM)
2MB-CYS 114.3 ± 99.9 ± 49.4 ± 4.0
2.8. Statistical analysis
10.5 17.0
3MB-CYS 552.5 ± 587.8 ± 122.9 ± Data analysis was carried out using the statistical software SPSS
96.5 140.9 36.8 (Statistical Package for Social Science, version 26, IBM Corp., Armonk,
METH-CYS 130.3 ± 85.2 ± –
NY, USA). For analysis of the statistical significance, the T-test was
7.9 14.9
PHEN-CYS 269.3 ± 251.2 ± 108.7 ± applied. Samples were considered significantly different when P-values
27.1 19.0 20.8 were lower than 0.05. All samples were analysed in triplicate, allowing
FURF-CYS 35.6 ± 2.8 12.8 ± 0.3 27.1 ± 8.1 the calculation of mean values and standard deviations.
HEX-CYS 76.6 ± 5.4 27.9 ± 3.0 17.6 ± 2.4
Sum Sum Sum
Aldehydes (µg/kg Strecker 5171.3 5772.5 163690.2 3. Results and discussion
DM) degradation
Maillard 316.7 317.5 6331.7 3.1. Free and cysteinylated aldehydes in malt samples
reaction
Lipid 399.3 356.5 3548.1
oxidation
One of the paramount challenges that the brewing industry faces, is
TOTAL 5887.3 6446.5 173570.0 to prolong the shelf life of beer. Malt, the major raw material used in
Cysteinylated Strecker 1272.9 1196.6 428.9 beer production, is known to have a positive impact on beer flavour
aldehydes (µg/kg degradation stability, for instance due to the presence of antioxidants (Boivin, 2001;
DM)
Coghe et al., 2004a, 2004b; Guido et al., 2007; Mikyška & Hrabak, 2002;
Maillard 35.6 12.8 27.1
reaction Vaughan et al., 2005; Yu et al., 2014). However, research has also shown
Lipid 76.6 27.9 17.6 a clear correlation between the rate of beer ageing and malt quality
oxidation parameters, e.g. free aldehyde levels, free amino nitrogen (FAN) and
TOTAL 1385.1 1237.3 473.6 malt filterability (De Clippeleer et al., 2010a, 2010b; Jaskula-Goiris
Malt quality parameters are expressed as mean values (n = 2). The individual et al., 2011), pointing to this raw material as a critical factor in rela­
aldehyde levels are expressed as mean values (µg/kg DM) ± standard deviation; tion to beer flavour stability. Therefore, in this paper we aimed at a
(n = 3). Classification of free aldehydes under study by their nature: I) Strecker detailed evaluation of different industrial malt samples by comparison
degradation aldehydes: 2MP (2-methylpropanal), 2MB (2-methylbutanal), 3MB of standard malt quality parameters and their content in free and cys­
(3-methylbutanal), METH (methional) and PHEN (phenylacetaldehyde); II) teinylated aldehydes (Table 2).
Maillard reaction aldehydes: FURF (furfural); and III) Lipid oxidation aldehydes:
Comparing the standard malt quality characteristics, we can pinpoint
HEX (hexanal) and T2N (trans-2-nonenal). Cysteinylated aldehydes (-CYS) are
colour and FAN as the parameters differing the most between the Pilsner
indicated in the table by using the same abbreviations for free aldehydes.
malt samples A and B1 versus the specialty malt B2 (Table 2). A well-

4
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

known consequence of the application of higher temperatures (during similar profile in that the aldehydes that exhibit higher concentrations in
malt kilning and wort boiling) and resulting Maillard reactions is a their free form also appear at higher concentrations in their cysteiny­
higher degree of malt colour and wort boiled colour, respectively (Coghe lated form.
et al., 2004a, 2004b). The used Biscuit malt B2 is characterised by a
visually determined malt colour of 47.0 EBC, a photometrically deter­
mined malt colour of 42.6 EBC, and a colour of 48.0 EBC for malt- 3.2. Free and cysteinylated aldehydes during brewing and beer ageing
derived Congress boiled wort, due to exposure to a considerable de­
gree of malt roasting. Conversely, due to less heat-load during kilning 3.2.1. Brew A
and therefore less Maillard reactions, the Pilsner malts A and B1 are only
moderately coloured with a visual malt colour of 7.0 EBC, a photometric 3.2.1.1. Free and cysteinylated aldehydes: From malt to onset mashing. As
malt colour of 5.8 EBC and a colour of 11.6 EBC for Congress boiled wort discussed in section 3.1., high concentrations of free aldehydes were
for malt A, versus 4.4 EBC, 4.3 EBC, and 6.8 EBC for malt B1, respec­ observed in the malt samples. Comparison of the concentrations of free
tively. The difference in colour between the Pilsner malts A and B1 re­ aldehydes quantified in malt (µg/kg DM) and the recalculated concen­
flects the somewhat higher kilning-off temperature to which malt A was trations at the onset of mashing (based on 2.2 L/kg mashing ratio,
exposed to during the malting process compared to malt B1. A high FAN expressed as µg/kg DM) is shown in Fig. 1A. In comparison with malt A,
value can indicate strong malt modification, i.e. proteolysis (Jaskula- significantly lower concentrations of PHEN, HEX, T2N and 2MB were
Goiris et al., 2011). However, as a consequence of high kilning and present at the onset of mashing. On the other hand, METH and FURF
roasting temperatures, more Maillard reactions involving free amino showed higher concentrations at the onset of the mash. No significant
acids and peptides are expected, resulting in a decreased FAN level. This differences were observed for 2MP and 3MB. Differences in the aldehyde
is in particular observed for malt B2, but also for Pilsner malt A levels observed between the malt and at onset of mashing, could be
compared to malt B1, which corresponds to the applied kilning tem­ related to adsorption to insoluble material, evaporation (depending on
peratures and the observed colour of these malt samples. Specialty malt the nature of the compound) or even binding to cysteine or the forma­
B2 is characterised by a FAN level of 28.0 mg/L in contrast to the tion of another type of bound-state aldehydes. When comparing the
significantly higher FAN levels of Pilsner malts A and B1, with 136.0 cysteinylated aldehydes (or non-volatile fraction) in malt and at the
mg/L and 163.0 mg/L, respectively. Conversion of amino acids and onset of mashing (Fig. 1B), the levels of most cysteinylated aldehydes
peptides, with the concomitant lower FAN values, results in Strecker (2MP-CYS, 2MB-CYS, 3MB-CYS, PHEN-CYS, and HEX-CYS) were
degradation aldehydes and Maillard reaction products. In this respect, significantly higher at the onset of mashing in comparison to the levels
analysis of free and cysteinylated aldehydes was carried out (Table 2). quantified in malt A. However, METH-CYS and FURF-CYS showed lower
High concentrations of free aldehydes were observed in the malt sam­ concentrations at the onset of mashing compared to the malt. These
ples under study (Table 2), which is in agreement with previous data findings suggest that both free and cysteinylated aldehydes from malt
from literature (Cramer et al., 2005; De Clippeleer et al., 2010a, 2010b; are extracted into the brewing water at the onset of mashing, and that
Ditrych et al., 2019; Liégeois & Collin, 2003). In particular, specialty free aldehydes, either extracted or produced de novo from e.g. amino
malt B2 showed the highest concentration of free aldehydes, probably acids via the well-known Strecker degradation, are bound to cysteine at
due to the application of higher kilning temperatures and longer heating mashing-in.
periods, with a total marker aldehydes level of 173,570 µg/kg malt. In
comparison, Pilsner malts A and B1 show in total 5887 and 6446 mg of 3.2.1.2. Free and cysteinylated aldehydes: During brewing until aged beer.
marker aldehydes per kg of malt, respectively. All quantified aldehydes As explained in section 2.3. and 2.4., samples from the onset of mashing
are present to a larger extent in specialty malt B2, in particular the until the final beer as well as during forced ageing of the beer samples,
Strecker aldehydes 2MP, 2MB and 3MB, for which the highest levels were analysed for their free and cysteinylated aldehyde content. This
were noted (Table 2). These levels of 2MP and 2MB are up to 50 times evolution is presented in Fig. 2 for brew A. All free aldehydes were
higher, and for 3MB up to 23 times higher than what is present in the detected at all the sampling points under study (Fig. 2A). The highest
Pilsner malts. Also furfural is present at a 20 times higher level in malt concentration levels were observed at the onset of mashing, suggesting
B2. When comparing the levels of free aldehydes between the Pilsner that malt is the major source of aldehydes introduced into the brewing
malts A and B1, slightly higher concentrations of free aldehydes are process. In particular, the Strecker degradation aldehydes 3MB, fol­
found in malt B1, however, no significant differences were observed (P- lowed by 2MP and 2MB, showed the highest concentrations at the
value ≥ 0.05). In these Pilsner malt samples, 3MB was found to be pre­ beginning of wort production, i.e. onset of mashing. During mashing,
sent in the highest concentration, followed by PHEN, 2MP, and 2MB. some increase in 2MP is noticed in addition to a remarkable decrease in
Similar findings regarding the levels of the Strecker degradation alde­ T2N, (most probably due to binding to insoluble material). When
hydes 3MB, PHEN, 2MP and 2MB in pilsner malt were previously ob­ comparing the behaviour of free aldehydes till the end of boiling, a
tained by Jaskula-Goiris et al. (2011). For the cysteinylated aldehydes in general trend shows a clear decrease in all Strecker degradation alde­
the Pilsner malts, the same trend was observed as for the free aldehydes, hydes (2MP, 2MB, 3MB, METH and PHEN) and the lipid oxidation
i.e. 3MB-CYS is present in the highest concentration, followed by PHEN- aldehyde HEX. The Strecker aldehydes in the highest concentrations
CYS, 2MP-CYS, and 2MB-CYS (Table 2). The specialty malt B2 showed (2MP, 2MB and 3MB) decreased more rapidly, when compared to other
lower amounts of cysteinylated aldehydes compared to the Pilsner malts aldehydes, which is in accordance with their relatively low boiling
A and B1 with the highest level for 2MP-CYS, followed by 3MB-CYS, points (Ditrych et al., 2019). The lipid oxidation aldehyde T2N clearly
PHEN-CYS, and 2MB-CYS (Table 2). Clearly, the lower amounts of increased during wort filtration, possibly due to lipoxygenase activity in
cysteinylated aldehydes in malt B2 are in agreement with its lower level the spent grains. During the subsequent wort boiling step, the concen­
of FAN. tration of T2N remained stable, likely due to its relatively high boiling
When the levels of both free and cysteinylated aldehydes are point (Lide, 2005). A different behaviour is observed for FURF, for
compared for the respective malt samples A, B1, and B2 (as presented in which the concentration at the onset of mashing initially remained
the Supplementary material 1), malt B2 clearly shows the highest levels stable until the onset of boiling, after which levels increased as a result of
of free but the lowest levels of corresponding cysteinylated aldehydes, the applied heat load during wort boiling. This is not surprising since
compared to Pilsner malts A and B1. This behaviour was expected this aldehyde is a Maillard reaction product, as described in literature
because of the high kilning temperatures applied during the malting (Baert et al., 2012; De Schutter et al., 2008, 2009; Vanderhaegen et al.,
process for malt B2 and, therefore, the lower availability of residual 2006). The concentration of FURF also further increased during the
amino acids, including cysteine. Furthermore, malts A and B1 show a following wort clarification step where the heat load remains high. After

5
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

Fig. 1. Comparison of free and cysteinylated aldehydes in Pilsner malt A, and in onset mashing samples for brew A: 2MP (2-methylpropanal), 2MB (2-methyl­
butanal), 3MB (3-methylbutanal), METH (methional) and PHEN (phenylacetaldehyde), FURF (furfural), HEX (hexanal) and T2N (trans-2-nonenal), 2-isopropylthia­
zolidine-4-carboxylic acid (2MP-CYS), 2-(sec-butyl) thiazolidine-4-carboxylic acid (2MB-CYS), 2-isobutylthiazolidine-4-carboxylic acid (3MB-CYS), 2-(2-(methylthio)
ethyl) thiazolidine-4-carboxylic acid (METH-CYS), 2-benzylthiazolidine-4-carboxylic acid (PHEN-CYS), 2-(furan-2-yl) thiazolidine-4-carboxylic acid (FURF-CYS), and
2-pentylthiazolidine-4-carboxylic acid (HEX-CYS). Results are expressed as mean values (n = 3). Error bars represent standard deviation. (NS: no significant
differences).

24 h of fermentation, FURF disappeared almost completely, most all aged beer samples at relatively constant concentrations (0.7–0.9 µg/
probably due to the reducing capacity of the yeast (Saison, 2009; Saison L). These findings are supported by data presented by Baert et al.
et al., 2010; Vanderhaegen et al., 2004). In contrast, a clear increase for (2015b) in beer pH model systems which describe the relative stability
PHEN, 3MB, and 2MB was observed after the first 24 h of fermentation, of cysteinylated aldehydes, e.g. 3MB-CYS, due to the aliphatic nature of
which could be related to yeast metabolism of amino acids. the 2-substitution. Although increasing levels of FURF were found dur­
During further maturation, free aldehydes were found at marginal ing beer ageing, FURF-CYS appeared only in limited amounts after 2
levels, below their limit of detection (LOD). When beer A, a top fer­ months of forced ageing (1.5 µg/L). This observation suggests that the
mented blond beer with 5.5% (v/v) alcohol (Table 3), was subsequently FURF increase over beer storage is not related to its bound form with
subjected to forced ageing, levels of all free aldehydes (Fig. 2D, 2E and cysteine, but more likely to the presence and conversion of Maillard
2F) proved to increase, but in particular FURF and 2MP (the most typical reaction intermediates such as 3-deoxyosones, as suggested earlier by
marker aldehydes of beer ageing) showed a noticeable increase. The Baert et al. (2012). Moreover, these results are also supported by the
increase in free aldehydes during beer ageing can be explained by de instability of the cysteinylated form of furfural (FURF-CYS), caused by
novo formation and/or release from a bound-state precursor form (Baert the aromatic character of furfural, presented in the work of Baert and co-
et al., 2018, 2015a, 2015b, 2012). workers in model solutions at brewing pH values (Baert et al., 2015b).
Cysteinylated aldehydes were quantified for the first time during the All of the other cysteinylated aldehydes were below their limit of
entire brewing process and subsequent beer ageing (Fig. 2B and 2C). The detection in forced aged beer samples.
highest concentrations of all cysteinylated aldehydes were observed at
the beginning of wort production, i.e. the onset of mashing. In addition, 3.2.2. Brew B
3MB-CYS followed by 2MP-CYS showed the highest concentrations of all
cysteinylated aldehydes under study. These results can be explained by 3.2.2.1. Free and cysteinylated aldehydes: During brewing until aged beer.
the equilibrium between the cysteinylated form of these aldehydes and Beer B is a top fermented amber beer of 8.5% (v/v) alcohol (Table 3).
their corresponding free form (Bustillo Trueba et al. 2019, 2018a, Results on free and cysteinylated aldehydes in wort and fresh and forced
2018b), since also the highest concentrations of free aldehydes were aged beers are presented in Fig. 3. Free aldehydes were quantifiable
observed for 3MB and 2MP (Fig. 2A), at the onset of mashing. A decrease throughout the entire process, i.e. from the onset of mashing until the
in all cysteinylated aldehydes was found from the onset of mashing until forced aged beer samples (Fig. 3A, 3D, 3E and 3F). Overall, compared to
the onset or the end of wort boiling, depending on the nature of the 2- brew A, in brew B higher levels of free aldehydes were found, which is in
substitution in the thiazolidine ring. In particular, the compounds accordance with the significant proportion of specialty malt B2 used in
showing the highest concentrations (3MB-CYS and 2MP-CYS) decreased this brew, and its high content in aldehydes. In analogy with brew A, the
more rapidly compared to those present at a lower range. As described highest concentrations of free aldehydes were measured at the initial
previously, levels in free aldehydes were relatively low from the end of stage of wort production, i.e. the onset of mashing, and, typically, a
wort boiling on, except for FURF. Therefore, lower amounts of alde­ decrease throughout the brewing process until the end of beer matura­
hydes are available for binding with cysteine. Moreover, due to heat, tion is again observed for all free aldehydes under study. The Strecker
degradation of the cysteinylated forms may occur (Bustillo Trueba et al., aldehyde 2MP was found to be present in the highest concentration,
2018a, 2018b), followed by evaporation of the free aldehyde when followed by 2MB and 3MB, i.e. the same Strecker aldehydes as previ­
applying increased temperatures from mashing-in until the end of ously observed in brew A (Fig. 2A). In brew B, relatively high levels of
boiling and wort clarification. Unambiguous detection of most of the HEX were noticed during the process, which corresponds to the high
cysteinylated aldehydes was not achieved from the end of wort boiling levels quantified in the applied specialty malt. Furthermore, the Mail­
throughout fermentation and maturation until the final fresh beer. lard reaction product FURF again showed a remarkable increase in
However, in forced aged beer samples, an increase in 2MP-CYS was concentration when heat load was applied. Interestingly, an increase
observed (up to 10 µg/L), which is in line with the significant increase in was again observed during the first 24 h of fermentation for the alde­
free 2MP found during beer ageing. Furthermore, 3MB-CYS was found in hydes 3MB, 2MB, and PHEN, which is related to the amino acid

6
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

Fig. 2. Evolution of free and cysteinylated staling aldehydes for brew A, from the onset of mashing until the fresh and forced aged beer (up to 3 months at 30 ◦ C): 2-
methylpropanal (2MP), 2-methylbutanal (2MB), 3-methylbutanal (3MB), methional (METH), phenylacetaldehyde (PHEN), furfural (FURF), hexanal (HEX), trans-2-
nonenal (T2N), 2-isopropylthiazolidine-4-carboxylic acid (2MP-CYS), 2-(sec-butyl) thiazolidine-4-carboxylic acid (2MB-CYS), 2-isobutylthiazolidine-4-carboxylic
acid (3MB-CYS), 2-(2-(methylthio)ethyl) thiazolidine-4-carboxylic acid (METH-CYS), 2-benzylthiazolidine-4-carboxylic acid (PHEN-CYS), 2-(furan-2-yl)
thiazolidine-4-carboxylic acid (FURF-CYS), and 2-pentylthiazolidine-4-carboxylic acid (HEX-CYS). Results are expressed as mean values (n = 3). Error bars represent
standard deviation.

metabolism of yeast. As for beer A, the lowest concentrations of free at the onset of mashing were the cysteinylated forms of the Strecker
aldehydes were observed in the final fresh beer B. However, levels of degradation aldehydes 2MP, 2MB and 3MB. All three aldehydes were
free aldehydes were somewhat higher in the fresh beer B compared to also most abundant in their free form, and, as a consequence, results are
beer A, which may be the result of higher levels of aldehydes introduced in accordance with the equilibrium mechanism proposed by Bustillo
by the specialty malt B2. Forced beer ageing showed the expected in­ Trueba et al. (2019, 2018a, 2018b). Furthermore, the same trend as for
crease in all free aldehydes (Fig. 3D, 3E and 3F). Most apparent was the brew A was noticed, i.e. a decrease throughout the brewing process for
increase in both FURF and 2MP as also observed for brew A. In partic­ all cysteinylated aldehydes under study, with quantification achievable
ular, FURF showed a high increase up to 600 µg/L after 3 months of till the onset or the end of wort boiling, depending on the nature of the
storage at 30 ◦ C. cysteinylated aldehyde. For cysteinylated aldehydes, unambiguous
In accordance with brew A, the corresponding cysteinylated alde­ detection was not accomplished for all compounds in samples taken
hydes were detected and quantified in the first steps of the brewing after wort boiling. In the fresh beer B, however, the following cys­
process (Fig. 3B). The compounds present in the highest concentrations teinylated aldehydes were found: PHEN-CYS (which was detected but

7
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

Fig. 3. Evolution of free and cysteinylated staling aldehydes for brew B, from the onset of mashing until the fresh and forced aged beer (up to 3 months at 30 ◦ C): 2-
methylpropanal (2MP), 2-methylbutanal (2MB), 3-methylbutanal (3MB), methional (METH), phenylacetaldehyde (PHEN), furfural (FURF), hexanal (HEX), trans-2-
nonenal (T2N), 2-isopropylthiazolidine-4-carboxylic acid (2MP-CYS), 2-(sec-butyl) thiazolidine-4-carboxylic acid (2MB-CYS), 2-isobutylthiazolidine-4-carboxylic
acid (3MB-CYS), 2-(2-(methylthio)ethyl) thiazolidine-4-carboxylic acid (METH-CYS), 2-benzylthiazolidine-4-carboxylic acid (PHEN-CYS), 2 (furan-2-yl)
thiazolidine-4-carboxylic acid (FURF-CYS), and 2-pentylthiazolidine-4-carboxylic acid (HEX-CYS). Results are expressed as mean values (n = 3). Error bars represent
standard deviation.

below the limit of quantification), METH-CYS and HEX-CYS at 6.2 µg/L 4. Conclusions
and 3.7 µg/L, respectively. Furthermore, some increase in 2MP-CYS was
observed as a function of forced beer ageing (Fig. 3C). As mentioned For the first time, the behaviour of both free aldehydes and their
previously, this data is in agreement with the increase observed for free corresponding cysteinylated forms, also referred to as 2-substituted 1,3-
2MP over beer ageing and, supported by the data on the stability of 2MP- thiazolidine-4-carboxylic acids, was monitored throughout the entire
CYS as reported by Baert et al. (2015a, 2015b). In contrast, the con­ brewing process, from the malt until the final beer, including beer
centrations of HEX-CYS, METH-CYS and PHEN-CYS did not change ageing. In malt, quantification of all selected free and cysteinylated al­
during beer ageing. dehydes was achieved. Clearly, the presence of high levels of free al­
dehydes in malt results in higher concentrations of the corresponding
cysteinylated aldehydes. Comparison between malt samples and brew­
ery samples taken at the onset of mashing showed higher concentrations

8
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

of free aldehydes in malt samples. Conversely, cysteinylated aldehydes of Brewing Chemists, 73(1), 100–108. https://doi.org/10.1094/ASBCJ-2015-0109-
01.
appeared to be present in higher concentrations in samples taken at the
Baert, J. J., De Clippeleer, J., Hughes, P. S., De Cooman, L., & Aerts, G. (2012). On the
onset of mashing when compared to the malt samples, pinpointing to origin of free and bound staling aldehydes in beer. Journal of Agricultural and Food
additional binding of free aldehydes to cysteine as a result of mashing. Chemistry, 60(46), 11449–11472. https://doi.org/10.1021/jf303670z.
Free aldehydes could be quantified in all intermediate brewery samples Baert, J. J., De Clippeleer, J., Jaskula-Goiris, B., Van Opstaele, F., De Rouck, G., Aerts, G.,
& De Cooman, L. (2015b). Further elucidation of beer flavor instability: The
and aged beer samples. For all aldehydes, a decrease along the brewing potential role of cysteine-bound aldehydes. Journal of the American Society of Brewing
process was observed, resulting in fresh beer samples showing the lowest Chemists, 73(3), 243–252. https://doi.org/10.1094/ASBCJ-2015-0531-01.
levels in free aldehydes. Aged beer samples showed an increase in all Bamforth, C. W. (2011). 125th Anniversary Review: The non biological instability of
beer. Journal of the Institute of Brewing, 57(3), 81–90. https://doi.org/10.1002/
free aldehydes (in particular in furfural and 2-methylpropanal) for both j.2050-0416.2011.tb00496.x.
brews. The use of specialty malt in the brewing process is reflected in Boivin, P. (2001). Pro- and anti-oxidant enzymatic activity in malt. Cerevisia, 26(2),
higher concentrations of free aldehydes, both in wort samples as well as 109–115.
Bustillo Trueba, P., De Clippeleer, J., De Rouck, G., Aerts, G., & De Cooman, L. (2018a).
in fresh beers. In particular, furfural was highly abundant in the roasted Study of cysteine-bound aldehydes in model solutions as a cause of flavour
malt B2, and strong increases in furfural were found in the beer derived instability: Influence of pH and temperature. Poster presentation session at the 13th
thereof upon ageing. Analogist to the free aldehydes, a decrease in trends in brewing symposium.
Bustillo Trueba, P., De Clippeleer, J., Van der Eycken, E., Guevara Romero, J. S., De
cysteinylated aldehydes was noticed during the brewing process. Un­ Rouck, G., Aerts, G., & De Cooman, L. (2018b). Influence of pH on the stability of 2-
equivocal quantification of cysteinylated aldehydes in brewery samples substituted 1,3-thiazolidine-4-carboxylic acids in model solutions. Journal of the
was only achieved until either the start or the end of wort boiling American Society of Brewing Chemists, 76(4), 272–280. https://doi.org/10.1080/
03610470.2018.1546094.
(depending on the nature of the cysteinylated aldehyde). Throughout
Bustillo Trueba, P., Jaskula-Goiris, B., De Clippeleer, J., Goiris, K., Praet, T.,
beer ageing, only the cysteine adduct of 2-methylpropanal showed an Sharma, U. K., Van der Eycken, E., Sanders, M. G., Vincken, J.-P., De Brabanter, J.,
increase in levels. Even though cysteinylated aldehydes could not be De Rouck, G., Aerts, G., & De Cooman, L. (2019). Validation of an ultra-high-
identified as the direct cause of increase in free aldehydes during beer performance liquid chromatography-mass spectrometry method for the
quantification of cysteinylated aldehydes and application to malt and beer samples.
ageing, the results presented in this study support the concept of for­ Journal of Chromatography A, 1604, Article 460467. https://doi.org/10.1016/j.
mation of bound-state aldehydes in malting and brewing. chroma.2019.460467.
Caballero, I., Blanco, C. A., & Porras, M. (2012). Iso-α-acids, bitterness and loss of beer
quality during storage. Trends in Food Science and Technology, 26(1), 21–30. https://
Funding doi.org/10.1016/j.tifs.2012.01.001.
Carmody, W. R. (1961). An easily prepared wide range buffer series. Journal of Chemical
The authors wish to thank “Internal Funds KU Leuven” (project: C24/ Education, 38(11), 559–560.
Coghe, S., Derdelinckx, G., & Delvaux, F. R. (2004a). Effect of non-enzymatic browning
15/020) for financial support. on flavour, colour and antioxidative activity of dark specialty malts - A review.
Monatsschrift Für Brauwissenschaft, 57(5–6), 25–38.
CRediT authorship contribution statement Coghe, S., Martens, E., D’Hollander, H., Dirinck, P. J., & Delvaux, F. R. (2004b). Sensory
and instrumental flavour analysis of wort brewed with dark specialty malts. Journal
of the Institute of Brewing, 110(2), 94–103. https://doi.org/10.1002/j.2050-
P. Bustillo Trueba: Conceptualization, Methodology, Investigation, 0416.2004.tb00188.x.
Writing - review & editing. B. Jaskula-Goiris: Supervision, Writing - Cramer, A.-C.-J., Mattinson, D. S., Fellman, J. K., & Baik, B.-K. (2005). Analysis of
volatile compounds from various types of barley cultivars. Journal of Agricultural and
review & editing. M. Ditrych: Conceptualization, Methodology, Inves­ Food Chemistry, 53, 7526–7531. https://doi.org/10.1021/jf0506939.
tigation. W. Filipowska: Conceptualization, Methodology, Investiga­ De Clippeleer, J., De Rouck, G., De Cooman, L., & Aerts, G. (2010a). Influence of the
tion. J. De Brabanter: Supervision. G. De Rouck: Supervision. G. hopping technology on the storage-induced appearance of staling aldehydes in beer.
Journal of The Institute of Brewing, 116(4), 381–398.
Aerts: Supervision. L. De Cooman: Supervision, Writing - review & De Clippeleer, J., Van Opstaele, F., Vercammen, J., Francis, G. J., De Cooman, L., &
editing. J. De Clippeleer: Supervision, Writing - review & editing. Aerts, G. (2010b). Real-time profiling of volatile malt aldehydes using selected ion
flow tube mass spectrometry. LCGC North America, 28(5), 386–395.
De Schutter, D. P., Saison, D., Delvaux, F., Derdelinckx, G., Rock, J. M., Neven, H., &
Declaration of Competing Interest Delvaux, F. R. (2008). Release and evaporation of volatiles during boiling of
unhopped wort. Journal of Agricultural and Food Chemistry, 56(13), 5172–5180.
The authors declare that they have no known competing financial https://doi.org/10.1021/jf800610x.
De Schutter, D. P., Saison, D., Delvaux, F. R. F., Derdelinckx, G., & Delvaux, F. R. F.
interests or personal relationships that could have appeared to influence (2009). The chemistry of aging beer. In Beer in Health and Disease. Prevention, Issue
the work reported in this paper. 36, 375–388.
Ditrych, M., Filipowska, W., De Rouck, G., Jaskula-Goiris, B., Aerts, G., Andersen, M. L.,
& De Cooman, L. (2019). Investigating the evolution of free staling aldehydes
Acknowledgments throughout the wort production process. BrewingScience, 72, 10–17. https://doi.
org/10.23763/BrSc18-21ditrych.
The authors wish to thank Jasper De Buyse for his support with the Ershov, A. Y., Nasledov, D. G., Lagoda, I. V., & Shamanin, V. V. (2014). Synthesis of 2-
substituted (2R,4R)-3-(3-mercapto-propionyl)thiazolidine-4-carboxylic acids.
brewing trials, and Agata Soszka and Johanna Schlich for their technical Chemistry of Heterocyclic Compounds, 50(7), 1119–1126. https://doi.org/10.1002/
support. cbic.200800164.
Fratianni, A. J. (2001). Results of the effect of storage temperature on packaged beer
flavour. Master Brewers Association of the Americas Technical Quarterly, 38(3),
Appendix A. Supplementary material 159–161.
Gibson, B., Aumala, V., Heiniö, R. L., Mikkelson, A., & Honkapää, K. (2018). Differential
Supplementary data to this article can be found online at https://doi. evolution of Strecker and non-Strecker aldehydes during aging of pale and dark
beers. Journal of Cereal Science, 83(August), 130–138. https://doi.org/10.1016/j.
org/10.1016/j.foodres.2020.110049. jcs.2018.08.009.
Guido, L. F., Curto, A. F., Boivin, P., Benismail, N., Gonçalves, C. R., & Barros, A. A.
References (2007). Correlation of malt quality parameters and beer flavor stability: Multivariate
analysis. Journal of Agricultural and Food Chemistry, 55(3), 728–733. https://doi.org/
10.1021/jf0623079.
Analysis committee of the EBC. (2018). Analytica-EBC (4th ed.). Brauerei-und Getränke.
Guido, L. F. (2016). Sulfites in beer: Reviewing regulation, analysis and role. Scientia
American Society of Brewing Chemists (2004). ASBC Methods of Analysis (14th ed.). St.
Agricola, 73(2), 189–197. https://doi.org/10.1590/0103-9016-2015-0290.
Paul, MN, U.S.A.
Hashimoto, N. (1966). Studies on volatile carbonyl compounds in beer. Report of the
Baert, J. J., De Clippeleer, J., Bustillo Trueba, P., Jaskula-Goiris, B., De Rouck, G.,
Research Laboratories of Kirin Brewery Co., Ltd.
Aerts, G., & De Cooman, L. (2018). Exploring aldehyde release in beer by 4-vinyl­
Jaskula-Goiris, B., De Causmaecker, B., De Rouck, G., Aerts, G., Paternoster, A., Braet, J.,
pyridine and the effect of cysteine addition on the beer’s pool of bound aldehydes.
& De Cooman, L. (2019). Influence of transport and storage conditions on beer
Journal of the American Society of Brewing Chemists, 76(4), 257–271. https://doi.org/
quality and flavour stability. Journal of the Institute of Brewing, 125(1), 60–68.
10.1080/03610470.2018.1518639.
https://doi.org/10.1002/jib.535.
Baert, J. J., De Clippeleer, J., De Cooman, L., & Aerts, G. (2015a). Exploring the binding
behavior of beer staling aldehydes in model systems. Journal of the American Society

9
P. Bustillo Trueba et al. Food Research International 140 (2021) 110049

Jaskula-Goiris, B., De Causmaecker, B., De Rouck, G., De Cooman, L., & Aerts, G. (2011). Saison, D. (2009). Aged beer flavour: Role of carbonyl compounds and the impact of
Detailed multivariate modeling of beer staling in commercial pale lagers. Brewing yeast reducing activity. In PhD Dissertation.
Science, 64(11–12), 119–139. Saison, D., De Schutter, D. P., Delvaux, F., & Delvaux, F. R. (2009). Determination of
Kaneda, H., Takashio, M., Osawa, T., Kawakishi, S., Koshino, S., & Tamaki, T. (1996). carbonyl compounds in beer by derivatisation and headspace solid-phase
Analysis of aldehyde-bisulfites in beer by HPLC-fluorescence detection with post- microextraction in combination with gas chromatography and mass spectrometry.
column derivatization. Journal of Food Science, 61(1), 105–108. https://doi.org/ Journal of Chromatography A, 1216(26), 5061–5068. https://doi.org/10.1016/j.
10.1111/j.1365-2621.1996.tb14736.x. chroma.2009.04.077.
Kilcast, D., & Subramaniam, P. (2011). Food and beverage stability and shelf life. Food Saison, D., De Schutter, D. P., Vanbeneden, N., Daenen, L., Delvaux, F., & Delvaux, F. R.
and Beverage Stability and Shelf Life. https://doi.org/10.1533/9780857092540.1.3. (2010). Decrease of aged beer aroma by the reducing activity of brewing yeast.
Lehnhardt, F., Gastl, M., & Becker, T. (2018). Forced into aging: Analytical prediction of Journal of Agricultural and Food Chemistry, 58(5), 3107–3115. https://doi.org/
the flavor-stability of lager beer. A review. Critical Reviews in Food Science and 10.1021/jf9037387.
Nutrition, 59, 1–12. https://doi.org/10.1080/10408398.2018.1462761. Solomons, T. W. G., Fryhle, C., & Snyder, S. (2013). Organic chemistry (11th ed.).
Lide, D. R. (2005). CRC Handbook of Chemistry and Physics (W. M. Haynes (Ed.); 97th Stewart, G. G. (2016). Beer shelf life and stability. In The Stability and Shelf Life of Food
ed.). (2nd ed.). Food Science, Technology and Nutrition. https://doi.org/10.1016/b978-
Liégeois, C., & Collin, S. (2003). Contribution of malt kilning to the cardboard flavour of 0-08-100435-7.00010-1.
aged beers. 29th European Brewery Convention. Stewart, G. G., Russell, I., & Anstruther, A. (2018). Handbook of brewing (3rd ed.). CRC
Liégeois, C., Meurens, N., Badot, C., & Collin, S. (2002). Release of deuterated (E)-2- Press. https://doi.org/10.1038/213765a0.
nonenal during beer aging from labeled precursors synthesized before boiling. Vanderhaegen, B., Neven, H., Verachtert, H., & Derdelinckx, G. (2006). The chemistry of
Journal of Agricultural and Food Chemistry, 50(26), 7634–7638. https://doi.org/ beer aging - A critical review. Food Chemistry, 95(3), 357–381. https://doi.org/
10.1021/jf020617v. 10.1016/j.foodchem.2005.01.006.
Malfliet, S., Van Opstaele, F., De Clippeleer, J., Syryn, E., Goiris, K., De Cooman, L., & Vanderhaegen, B., Neven, H., Verstrepen, K. J., Delvaux, F. R., Verachtert, H., &
Aerts, G. (2008). Flavour instability of pale lager beers: Determination of analytical Derdelinckx, G. (2004). Influence of the brewing process on furfuryl ethyl ether
markers in relation to densory ageing. Journal of the Institute of Brewing, 114(2), formation during beer aging. Journal of Agricultural and Food Chemistry, 52(22),
180–192. https://doi.org/10.1002/j.2050-0416.2008.tb00324.x. 6755–6764. https://doi.org/10.1021/jf0490854.
Matsui, S., & Kitabatake, K. (1984). Fluorimetric determination of cysteine in beer by Vaughan, A., O’Sullivan, T., & Sinderen, D. (2005). Enhancing the microbiological
High-Performance Liquid Chromatography with precolumn derivatisation. Journal of stability of malt and beer - A review. Journal of the Institute of Brewing, 111(4),
the Institute of Brewing, 90, 20–23. 355–371. https://doi.org/10.1002/j.2050-0416.2005.tb00221.x.
Mikyška, A., & Hrabak, M. (2002). The role of malt and hop polyphenols in beer quality, Weerawatanakorn, M., Wu, J. C., Pan, M. H., & Ho, C. T. (2015). Reactivity and stability
flavour and haze stability. Journal of the Institute of Brewing, 108(1), 78–85. https:// of selected flavor compounds. Journal of Food and Drug Analysis, 23(2), 176–190.
doi.org/10.1002/j.2050-0416.2002.tb00128.x. https://doi.org/10.1016/j.jfda.2015.02.001.
Paternoster, A., Jaskula-Goiris, B., Buyse, J., Perkisas, T., Springael, J., Braet, J., De Wietstock, P. C., Kunz, T., & Methner, F. J. (2016). Relevance of oxygen for the formation
Rouck, G., & De Cooman, L. (2020). The relationship between flavour instability, of Strecker aldehydes during beer production and storage. Journal of Agricultural and
preference and drinkability of fresh and aged beer. Journal of the Institute of Brewing, Food Chemistry, 64(42), 8035–8044. https://doi.org/10.1021/acs.jafc.6b03502.
126, 59–66. https://doi.org/10.1002/jib.582. Yu, J., Huang, S., Dong, J., Fan, W., Huang, S., Liu, J., Chang, Z., Tian, Y., Hao, J., &
Rodrigues, J. A., Barros, A. S., Carvalho, B., Brandão, T., & Gil, A. M. (2011). Probing Hu, S. (2014). The influence of LOX-less barley malt on the flavour stability of wort
beer aging chemistry by nuclear magnetic resonance and multivariate analysis. and beer. Journal of the Institute of Brewing, 120(2), 93–98. https://doi.org/10.1002/
Analytica Chimica Acta, 702(2), 178–187. https://doi.org/10.1016/j. jib.122.
aca.2011.06.042.

10

You might also like