Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Caramelization in Foods: A Food Quality and Safety Perspective

lı and Vural Gökmen, Hacettepe University, Beytepe Campus, Ankara, Turkey
Tolgahan Kocadag
© 2018 Elsevier Inc. All rights reserved.

Introduction 1
Caramelization Reactions 1
Flavor Development 4
Browning Development 6
Potential Toxicants Formed During Caramelization 7
Imidazoles 7
Furan and Derivatives 9
a-Dicarbonyl Compounds 9
References 10

Introduction

Caramelization is a nonenzymatic browning reaction of sugars providing a caramel-like flavor during high temperature treatment of
foods. Degradation of sugars is catalyzed by amino acids during the Maillard reaction, which is characterized by nitrogen containing
low and high molecular weight compounds. At elevated temperatures, both reactions proceed together in a way that one affects the
other.
The Maillard reaction may take place under milder conditions, but sugars are caramelized at temperatures above 120 C. Food
products reach these elevated temperatures during processes like roasting (180–240 C) and baking (160–240 C). After drying of
the outer layers of bakery products during heating, the surface can reach temperatures above 100 C where the browning and flavor
development begin (Mogol and Gökmen, 2014a,b). Caramelization reactions are also observed in jams, canned fruit products, fruit
juices and concentrates, soft drinks, honey, and sugar syrups during thermal treatment or storage. At milder temperatures, pH below
three and above nine are often required to observe reactions in moderately higher rates (Kroh, 1994).
Food caramels are commercially available as caramel color, burnt sugars, aromatic caramel, and caramelized sugar syrups. Burnt
sugars, aromatic caramel, and caramelized sugar syrups are produced without using additives, while additives are used for caramel
color production. Caramel color is used as a colorant in food and beverages, and it is not usually a flavoring.

Caramelization Reactions

The main reactions of sugar degradation are schematized in Fig. 1. The acyclic forms of sugars are very reactive, and more acyclic
forms are found by increasing temperature. Isomerization, epimerization, dehydration, and oxidation reactions of the cyclic
reducing sugar are initiated by ring opening followed by enolization (Speck, 1958). Aldose and ketose sugars isomerize via their
common intermediate, 1,2-enediol. This isomerization also involves epimerization, that is, the change of the configuration of C-
2 in aldoses and C-3 in ketoses. Glucose, fructose, and mannose are found in equilibrium in alkaline solutions (Fig. 2). Similarly,
galactose, tagatose, and talose may interconvert each other. These aldose–ketose isomerization and epimerization reactions are
known as the Lobry de Bruyn–Alberda van Ekenstein transformation (rearrangement) (LdB-AvE) (Angyal, 2001). Within these reac-
tions, aldose–ketose interconversion predominates, and to a lesser extent epimerization is observed. Although the LdB-AvE rear-
rangement is observed faster in alkaline media, it can also occur limitedly in acidic media (Zhao et al., 2007). In addition to the
alkaline conditions required for the isomerization, a considerable amount of glucose–fructose interconversion is observed during
the heating of sugar-rich low-moisture foods and dry heating of sugars at elevated temperatures (Kocada glı and Gökmen, 2016a,b).
Aldose–ketose isomerization is substantially catalyzed in the presence of alkali and alkaline earth metal cations (Naþ, Kþ, Mg2þ,
Caþ2) by increasing the rate of chain opening (Speck, 1958).
Aldose–ketose isomerization (LdB-AvE) is also observed in the reducing end of the oligosaccharides, such as maltose (4-O-a-D-
glucopyranosyl-D-glucose) to maltulose (4-O-a-D-glucopyranosyl-D-fructose) and lactose (4-O-b-D-galactopyranosyl-D-glucose) to
lactulose (4-O-b-D-galactopyranosyl-D-fructose). Maltose/maltulose ratio can be used as a marker of thermal input in certain cereal
products such as cookies, crackers, breakfast cereals (Garcı ́a-Baños et al., 2004a), pasta (Garcı ́a-Baños et al., 2004b), and infant
formula (Morales et al., 2004). Lactulose is an indigestible disaccharide with laxative and prebiotic properties. It can be formed
during heating infant formula, milk, and milk products and can also be used as an ingredient. Isomerization reaction of lactose
to lactulose during thermal treatment of milk is quantitatively more important than the Maillard reaction for the degradation of
lactose (Berg, 1993).
In an acidic medium, enolization and subsequent b-elimination of 1–3 water molecules (dehydration) predominates. In an
alkaline medium, enolization is followed by fragmentation reactions (including fragmentation after dehydrations and

1
2 Caramelization in Foods: A Food Quality and Safety Perspective

Isomerization
e.g. glucose-fructose

Epimerization
e.g. glucose-mannose
Fragmentation Shorter chain
Acyclic reducing Enolization
Enediols carbonyl compounds
sugars (e.g. 1,2-enediol, Dehydration (e.g. methylglyoxal)
2,3-enediol)
Deoxyosones
Dehydration Furan derivatives
Mutarotation

Oxidation (e.g. 5-hydroxymethyl-2-furfural)


e.g. glucosone, glyoxal
Intramolecular glycosidic
bond formation
Anhydrosugars (e.g. levoglucosan)

Glycoside formation
Cyclic reducing sugars Oligomers
(α- and β- anomers)
Dehydration
Furan and pyran derivatives (e.g. 5-hydroxymethyl-2-furfural)
Hydrolysis

Oligosaccharides
Polysaccharides
Figure 1 Main reactions of sugar degradation.

D-glucose 1,2-enediol D-mannose

D-fructose 2,3-enediol D-psicose

Figure 2 Lobry de Bruyn–Alberda van Ekenstein transformation.

isomerizations) via splitting of carbon bonds, and subsequently recombination is observed via aldol addition. Enediols are
oxidized via transition metal catalysis, such as Cu2þ, and also in the presence of oxygen.
During heating or drying of fruit products, heating of sugar with or without additives, and baking of sourdough, many dehydra-
tion products of sugars may form because of the acidic conditions of the products and high temperature. Enediols are also the key
intermediates for the dehydration of sugars under acidic conditions. Dehydration refers to the removal of a water molecule by
b-elimination mechanism from the carbon skeleton of a carbohydrate. Heating accelerates the dehydration especially under low
moisture conditions and even in milder pH conditions because of an increment in acyclic forms. The dehydration may also occur
Caramelization in Foods: A Food Quality and Safety Perspective 3

under alkaline condition during heating, followed by extensive fragmentation. Dehydration of sugars produces several reactive
deoxyosones, and more stable furan and pyran derivatives, which are important for the formation of caramel color and flavor.
3-Deoxyosone, having an a-dicarbonyl structure, is formed by the removal of a water molecule from the third carbon of an aldo-
hexose or a ketohexose (Fig. 3). When hexoses are glucose and fructose, 3-deoxyglucosone (3-deoxy-D-erythro-hexos-2-ulose) is
formed. Similarly, galactose yields 3-deoxygalactosone (3-deoxy-D-threo-hexos-2-ulose) (Hellwig et al., 2010). Dehydration from
C-4 of 3-deoxyglucosone produces 3,4-dideoxyglucosone-3-ene (3,4-dideoxy-D-glycero-hex-2-enos-2-ulose). Dehydration from
a hemiacetal structure of 3,4-dideoxyglucosone-3-ene results in 5-hydroxymethyl-2-furfural (HMF, 5-hydroxymethylfuran-2-
carbaldehyde). HMF is a rather stable product, and it is often used as an indicator of the thermal load applied to foods (Gökmen
and Senyuva, 2006; Gökmen et al., 2007, 2008; Göncüo glu and Gökmen, 2013). However, HMF might also behave as a reactive
intermediate, especially at elevated temperatures, because of its carbonyl group (Nikolov and Yaylayan, 2011a,b; Gökmen et al.,
2012).
Fructose gives rise to the formation of a wide variety of products because of its ability to be involved in both 1,2- and 2,3-
enolization. Dehydration of fructose from C-1 through 2,3-enediol yields 1-deoxyglucosone (1-Deoxy-D-erythro-hexo-2,3-
diulose) (Fig. 3). 1-Deoxyglucosone is the key reactive intermediate in the formation of many other reactive intermediates and
important aroma compounds (Hirsch et al., 1995; Voigt and Glomb, 2009). Isomerization of 1-deoxyglucosone along the carbon
chain is an important aspect in the formation of a wide range of products that are important for color and flavor development
(Belitz et al., 2009).
Glucose and fructose oxidation produces glucosone (D-arabino-hexos-2-ulose) by removal of two protons, especially by transi-
tion metal catalysis and in the presence of oxygen. Nonoxidative degradation of glucosone yields C-5 intermediates, which are pen-
tosone, 3-deoxypentosone, and 1-deoxypentosone (Gobert and Glomb, 2009).
Shorter chain a-dicarbonyl compounds, a-hydroxyaldehydes, and a-hydroxyketones are formed by the cleavage of the carbon
skeleton, especially through deoxyosones. Alkaline conditions accelerate the formation of fragments. These shorter chain carbonyl
compounds are of importance because they substantially contribute to browning, and involve in the formation of flavor
compounds (Hollnagel and Kroh, 1998).
Typical products of the fragmentation with one carbon is formaldehyde; with two carbons are glyoxal, 2-hydroxyacetaldehyde
(glycolaldehyde), and acetaldehyde; with three carbons are 2,3-dihydroxypropanal (glyceraldehyde), 1,3-dihydroxyacetone,
hydroxyacetone (acetol), methylglyoxal (pyruvaldehyde), 2-hydroxypropanal (lactaldehyde), hydroxymethylglyoxal; with four
carbons are 2,3-butanedione (diacetyl), 1-hydroxybutan-2,3-dione, and with five carbons is 2,3-pentanedione (Fig. 4) (Smuda
and Glomb, 2013). These can be formed by direct degradation of the carbon chain (after isomerization and/or dehydration)
and also by condensation of the fragments. The mechanisms largely depend on the reaction conditions, and the source of carbon
may come from different regions of the carbon skeleton. Some of these a-dicarbonyl compounds may also form as secondary lipid
oxidation products and in ascorbic acid degradation (Fujioka and Shibamoto, 2004; Schulz et al., 2007).
In addition to the formation of short-chain carbonyl compounds, carboxylic acids are also formed. Formic acid and acetic acid
are formed from 3-deoxyglucosone and 1-deoxyglucosone, respectively. The main mechanism is hydrolytic b-dicarbonyl cleavage
(Davidek et al., 2005; Smuda and Glomb, 2013), and to a lesser extent oxidative cleavage (Davidek et al., 2006).

D-glucose 1,2-enediol D-fructose 2,3-enediol enol form 1-deoxyglucosone

5-hydroxymethyl-2-furfural

3-deoxyglucosone 3-deoxyglucosone 3,4-dideoxyglucosone-3-ene


(enol form)
Figure 3 Dehydration of glucose and fructose.
4 Caramelization in Foods: A Food Quality and Safety Perspective

glyoxal 2-hydroxyacetaldehyde acetaldehyde 2,3-dihydroxypropanal

1,3-dihydroxyacetone hydroxyacetone methylglyoxal 2-hydroxypropanal

hydroxymethylglyoxal 2,3-butanedione 1-hydroxybutan-2,3-dione 2,3-pentanedione


Figure 4 Fragments with carbonyl moieties formed in caramelization.

Glycosidic bonds of oligosaccharides are cleaved by acid catalysis. Hydrolysis of the glycosidic bond is affected especially by pH
and temperature, and also depends on the structure. Reducing oligosaccharides, such as maltose and lactose, may participate in
browning reactions by producing various reactive intermediates (Hollnagel and Kroh, 1998). Degradation of polysaccharides might
also be important at very high temperatures as in the case of coffee roasting.
Hydrolysis and decomposition of sucrose in sweet bakery products are of importance as inversion products glucose and fructose
participate in nonenzymatic browning reactions. Moreover, cleavage of a glycosidic bond of sucrose releases glucose and fructofur-
anosyl cation (Fig. 5) (Perez Locas and Yaylayan, 2008). At elevated temperatures and under dry heating conditions, the fructofur-
anosyl cation produces HMF via elimination of a proton and dehydration (Perez Locas and Yaylayan, 2008). HMF formation with
this pathway is more efficient than dehydration of 3-deoxyglucosone to HMF through acyclic intermediates (Perez Locas and
Yaylayan, 2008). Another possible pathway for HMF formation is dehydration of fructose over cyclic intermediates (Antal et al.,
1990; Mayes et al., 2014; Kocada glı and Gökmen, 2016a,b).
Reactions of sugars are not only limited to the decomposition reactions but inter- and intramolecular glycoside bond formations
are also observed. Oligomers with up to six carbohydrate units may form during caramelization of glucose, fructose, and sucrose via
glycosidic bonding (Golon and Kuhnert, 2012; Golon and Kuhnert, 2013). These oligomers are also found in their dehydrated (up
to eight H2O loss) and hydrated forms.
Anhydrosugars (sugar anhydrites) are formed from aldohexoses by intramolecular condensation of the hemiacetal and other
hydroxyl groups. Levoglucosan (1,6-anhydro-b-D-glucopyranose or b-glucosan) is abundantly formed during heating of glucose,
starch, and cellulose at high temperatures (Fig. 6). To a limited extent, its furanose form (1,6-anhydro-b-D-glucofuranose) may
also form (Velisek, 2014).
Intramolecular glycosidic bond formation is not observed in fructose, but difructose dianhydrides (Fig. 6) and their glycosylated
derivatives are formed via fructofuranosyl cation (Audemar et al., 2017). Glycosidic bond formation between fructofuranosyl cation
and fructose produces fructodisaccharides (fructobioses), which then form a tricyclic structure via intramolecular glycosidic
bonding. Isomerization in the molecule gives various diastereomers and further glycosylation may give glycosylated derivatives.
These difructose dianhydrides and their glycosylated derivatives are prebiotic compounds found in caramel (Mellet and Garcia
Fernandez, 2010). Isomers of difructose dianhydrides are also important as they can be used for the authenticity of caramel and
for the detection of addition of caramel in foods (Ratsimba et al., 1999).

Flavor Development

Major flavor compounds formed during caramelization of sugars are oxygen-containing heterocyclic compounds, furanones and
pyranones, and also carbocyclic compounds, cyclopentenolones. In these groups of molecules, caramel aroma appears if a planar
vicinal enol-oxo configuration is found (Fig. 7) (Belitz et al., 2009). A volatile fraction of caramel constitutes low molecular weight
compounds derived from degradation of sugars. These compounds may also play a role in the aroma of baked and roasted foods via
Caramelization in Foods: A Food Quality and Safety Perspective 5

glucose + fructose

sucrose

dry

fructofuranosyl cation fructofuranose

fructose

2,5-anhydro-D-mannose (enol)

2,3-dihydrofuran 5-hydroxymethyl-2-furfural
Figure 5 Degradation of sucrose and fructose to HMF. Adapted from Antal Jr., M.J., Mok, W.S., Richards, G.N., 1990. Mechanism of formation of
5-(hydroxymethyl)-2-furaldehyde from d-fructose and sucrose. Carbohydr. Res. 199, 91–109 and Locas, C.P., Yaylayan, V.A., 2008. Isotope labeling
studies on the formation of 5-(hydroxymethyl)-2-furaldehyde (HMF) from sucrose by pyrolysis-GC/MS. J. Agric. Food Chem. 56, 6717–6723.

1,6-anhydro- -D-glucopyranose difructose dianhydrate


Figure 6 Levoglucosan and difructose dianhydride

Figure 7 Enol-oxo configuration found in compounds with caramel aroma.

caramelization reactions; however, many characteristic aroma compounds with relatively low odor thresholds are formed via
involvement of amino acids in the Maillard reaction. Increment in fragmentation reactions during the Maillard reaction may
also provide short-chain carbonyl compounds for aldol reactions required for the formation of compounds with a caramel flavor.
One of the important compounds is furaneol (4-hydroxy-2,5-dimethyl-2H-furan-3-one) (Fig. 8) with an intensive caramel fla-
vor and a relatively low odor threshold (Zabetakis et al., 1999). It is found in caramel, biscuits, coffee, and dark beer. It can be
formed via aldol condensation of hydroxyacetone and methylglyoxal followed by cyclization and dehydration (Velisek, 2014).
It is mainly formed from L-rhamnose (6-deoxy-L-mannose) (Haleva-Toledo et al., 1997; Illmann et al., 2009). Another furanone
6 Caramelization in Foods: A Food Quality and Safety Perspective

furaneol sotolon abhexon acetylformoin

maltol isomaltol cyclotene


Figure 8 Compounds with distinct caramel flavor.

with a higher odor threshold, norfuraneol (4-hydroxy-5-methyl-2H-furan-3-one), can be formed from the furanose form of 1-
deoxyglucosone upon dehydration and splitting of formaldehyde (Velisek, 2014).
Sotolon (caramel furanone, 3-hydroxy-4,5-dimethyl-5H-furan-2-one) and abhexon (maple furanone, 5-ethyl-3-hydroxy-4-
methyl-5H-furan-2-one) are powerful aroma compounds with relatively low odor thresholds (Maarse, 1991). Sotolon can be
produced from glycolaldehyde with 2,3-butanedione, and abhexon from glycolaldehyde and 2,3-pentanedione via aldol conden-
sation (Belitz et al., 2009). Sotolon has caramel, maple syrup, and a burnt sugar odor at low concentrations, and fenugreek or curry
odor at high concentrations (Tokitomo et al., 1980; Girardon et al., 1986). Abhexon has an intense maple syrup and typical protein
hydrolyzate aroma. Both sotolon and abhexon can be formed from different precursors in different food products (Velisek, 2014).
Acetylformoin (2,4-dihydroxy-2,5-dimethyl-2H-furan-3-one) is another important furanone with a caramel flavor (Cutzach
et al., 1999). It is formed from the dehydration of hexoses through 1-deoxyglucosone (Hofmann, 1998). Acetylformoin is found
in many tautomeric forms in its acyclic structure (Goto et al., 1963). It is very reactive and unstable because of its reductone structure
in the acyclic form. Acetylformoin is also an intermediate in furaneol formation (Hofmann and Schieberle, 2000). It can be found in
aromatic caramel and burnt sugar (Paravisini et al., 2012).
Maltol is found in biscuits, roasted coffee, chocolate, caramel, malt, and beer. It has a high odor threshold and may not always
contribute to the aroma of food (Belitz et al., 2009). It is one of the important odorants in burnt sugar caramels (Paravisini et al.,
2012). Maltol is formed when glucose has a 4-O glycoside linkage, like in disaccharides maltose and lactose, and its formation from
monosaccharides is negligible (Yaylayan and Mandeville, 1994). Isomaltol (2-acetyl-3-hydroxyfuran) also has a caramel-like flavor,
and it is formed from hexoses, and especially from 4-O-substituted glucoses at higher concentrations.
Cyclopentenolones are alicyclic compounds formed from fragments of sugar degradation and have a distinct caramel flavor.
Cyclotene (2-hydroxy-3-methyl-2-cyclopenten-1-one) has an intense burnt sugar and caramel flavor, and it can be formed from
hydroxyacetone and 2-hydroxypropanal (Velisek, 2014).
The volatile fraction of sugar degradation contains many more compounds with different aroma properties and odor thresholds.
Among others, diacetyl, 2,3-pentanedione, acetoin, furan, 2-methylfuran, 3-methylfuran, furfural, 2-acetylfuran, and 2-
hydroxyacetylfuran are the constituents of the volatile fraction in thermally treated foods. Although they can be directly formed
from sugar backbone during caramelization, carbon atoms from amino acid fragmentation may also be involved, and the quanti-
tative distribution largely depends on the conditions and food product composition (Kerler et al., 2010).

Browning Development

The nature of the pigment molecules responsible for the color development during heating sugars is largely unknown, similar to
melanoidins formed in the Maillard reaction. During heating and storage of several food products, the Maillard reaction is respon-
sible for nonenzymatic browning via catalysis of sugar dehydration and fragmentation by involvement of amino acids and incor-
poration of nitrogen to the polymers. At low pH values, amino compounds get protonated, making them less susceptible to
participate in carbonyl-amine condensation, which is the first step of the Maillard reaction, because of the decrement in the nucle-
ophilic strength of the amino group (Martins et al., 2000; Martins and Van Boekel, 2005). Browning in dessert wine, where these
conditions are observed, is favored via caramelization reactions (Kroh, 1994). Removal of amino acids from citrus juices is known
to decrease undesired nonenzymatic browning during thermal treatment and storage (Bharate and Bharate, 2014). However,
browning still progresses to a limited extent via caramelization reactions and ascorbic acid degradation. Caramelization reactions
of fructose might account for 10%–36% of the browning observed during heating fructose–lysine aqueous mixture (Ajandouz and
Puigserver, 1999; Ajandouz et al., 2001). However, it is generally hard to make a conclusion to distinguish between caramelization
and the Maillard reactions, especially at very high temperatures, for the contribution to browning.
Caramelization in Foods: A Food Quality and Safety Perspective 7

During dry heating of solid sugars, reactivity is observed after melting or glass transition temperatures because transition from
solid to liquid brings mobility to the molecules, which is required to observe any reaction at a considerable rate (Roos et al., 2013).
Therefore the caramelization and browning in sugar solids requires higher temperatures. Melting temperatures of fructose, glucose,
and sucrose are 127, 158, and 190 C, respectively (Roos and Drusch, 2016). Increasing temperature increases the kinetic energy of
the molecules and thus caramelization reactions get faster. In parallel with that, more acyclic sugar molecules are found at higher
temperatures, and thus fragmentation reactions provide an intensive dark color development. The increments in short-chain
a-dicarbonyl compounds, a-hydroxycarbonyl compounds, and furan derivatives are associated with an increment in browning
intensity (Hollnagel and Kroh, 1998; Kroh et al., 2008).
The presence of alkali metal cations (like NaCl, KCl, CaCl2) enhances the browning during caramelization of sugars. Color inten-
sity increases with higher concentrations of NaCl and KCl in cereal model systems and breakfast cereals (Moreau et al., 2009b,
2011). Browning provided by NaCl is linked neither to the hygroscopic behavior of NaCl nor to the physical state of the cereal
model systems (Moreau et al., 2009a). In cookies, the presence of NaCl, KCl, and CaCl2 increases 2-furfural and 5-
hydroxymethyl-2-furfural in parallel with browning (Kocada glı and Gökmen, 2016c). Contribution of fructose to browning devel-
opment is generally higher than glucose. Heating glucose with alkali metal cations increases the rate of isomerization of glucose to
fructose; therefore more furan derivatives are formed from fructose, which might be associated with the colored pigment formation
(Hollnagel and Kroh, 1998; Mayes et al., 2015; Kocada glı and Gökmen, 2016a).
Alkaline pH is also necessary to increase browning intensity during heating of sugars and sugar-rich foods. Higher pH values also
increase the rate of browning by increasing enolization, as temperature does. Enhancement of browning at alkaline conditions is
related to the increasing rate of fragmentation reactions.
Caramel colors are produced by controlled heating of sugars at temperatures above 120 C, and pressure can also be applied.
Sucrose, glucose, fructose, sugar syrups, and starch are the main sources of raw material. Caramel colors are classified into four cate-
gories depending on the additives used (FAO, JECFA) (Table 1). The additives are used as catalysts to promote browning and to
achieve desired properties (such as colloidal charge) for specific food products. Ammonia-sulfite caramel is the darkest colored
caramel, and it is commonly used in soft drinks. Nitrogen-containing heterocyclic compounds pyrroles, pyridines, pyrazines,
and imidazoles are formed in the case of using ammonium salts, which are characteristic to the Maillard reaction.

Potential Toxicants Formed During Caramelization


Imidazoles
4(5)-Methylimidazole (4-MI) and 2-methylimidazole (2-MI) are found in class III and IV caramel colors, and 2-acetyl-4-
tetrahydroxybutylimidazole (THI) is found only in class III caramel color (EFSA, 2011). 4-MI and 2-MI are classified as “possibly
carcinogenic to humans (Group 2B)” by the International Agency for Research on Cancer (IARC, 2013). This classification is based
on the sufficient evidence in experimental animals for the carcinogenicity of both compounds (IARC, 2013). THI is a potent immu-
nosuppressant, which means it reduces the strength of the body’s immune system (EFSA, 2011). Maximum residue limits have been
set for 4-MI as 200 mg kg1 in class III and 250 mg kg1 in class IV, and for THI as 10 mg kg1 on equivalent color basis by the
European Union. 4-MI can be present in caramel colors up to 1000 mg kg1, being higher in class IV, and varies largely (Hengel
and Shibamoto, 2013). Concentration of THI can be up to 50 mg kg1 (Elsinghorst et al., 2013). 2-MI is generally found in lower
amounts.
The imidazole ring is formed via reactions of reactive carbonyl compounds and ammonia that is found in class III and IV caramel
colors. 4-MI has been proposed to be formed from methylglyoxal in the presence of ammonia (Fig. 9) (Moon and Shibamoto,
2011). THI is formed via iminofructosamine, which is formed in the reaction of hexoses with ammonia (Fig. 10) (Kroeplien

Table 1 Classification of caramel colors

Name Plain Caustic sulfite Ammonia Ammonium sulfite


Class Class I, E150a Class II, E150b Class III, E150c Class IV, E150d

Additive NaOH, KOH, Ca(OH)2, Na2CO3, H2SO4, Na2SO3, K2SO3, NaOH, NH4OH, (NH4)2CO3, NH4HCO3, NH4OH, (NH4)2CO3, NH4HCO3,
(promoter) K2CO3, H2SO4, acetic acid, KOH, Ca(OH)2 ammonium phosphate, ammonium phosphate,
citric acid Na2CO3, NaOH, KOH, (NH4)2SO3, NH4HSO3,
Ca(OH)2 Na2CO3, NaOH, KOH,
Ca(OH)2, H2SO4, Na2SO3,
NaHCO3, K2SO3, KHCO3
Color golden-yellow reddish-brown brown-yellow dark brown-grey
Colloidal weak negative positive negative
charge
Uses alcoholic beverages, coffee spirits, vermouth, brandies, beer, vinegar, acidic foods, soft drinks, bakery products,
extract, bakery products, bakery products, sauces, bakery products, confectionary
confectionary confectionary confectionary
8 Caramelization in Foods: A Food Quality and Safety Perspective

formamide acetamide

methylglyoxal

acetaldehyde
formaldehyde

hydroxyacetone

2-aminopropanal formamide
4(5)-methylimidazole

2-aminopropanal formamide
Figure 9 Formation of 4(5)-methylimidazole. Adapted from Moon, J.K., Shibamoto, T., 2011. Formation of carcinogenic 4(5)-methylimidazole in
maillard reaction systems. J. Agric. Food Chem. 59, 615–618.

NH3 NH3

D-glucose
2-acetyl-4-tetrahydroxybutylimidazole
Figure 10 Formation of 2-acetyl-4-tetrahydroxybutylimidazole. Adapted from Kroeplien, U., Rosdorfer, J., Van Der Greef, J., Long, R.C., Goldstein,
J.H., 1985. 2-acetyl-4(5)-(1,2,3,4-tetrahydroxybutyl)imidazole: detection in commercial caramel color III and preparation by a model browning
reaction. J. Org. Chem. 50, 1131–1133.

et al., 1985). Although their presence is commonly of concern in caramel colors, they may also form in thermally processed foods
during the Maillard reaction (Mottier et al., 2017). The source of nitrogen or ammonia in this case is the Strecker degradation of
amino acids (Yaylayan and Haffenden, 2003).
Divalent cations Ca2þ and Mg2þ have been shown to decrease the formation of 4-MI (Wu et al., 2016). The levels of sulfide play
an important role for the formation of 4-MI during class IV caramel color production, and addition of appropriate levels of sulfide
decreases 4-MI formation (Lee et al., 2013). Due to the wide usage of caramel colors in several food products, controlling the expo-
sure to imidazoles and other toxicants is of importance from the point of view of food safety.
Caramelization in Foods: A Food Quality and Safety Perspective 9

Furan and Derivatives


IARC has classified furan as “possibly carcinogenic to humans (Group 2B)” (IARC, 1995). The precursors of furan are sugars, amino
acids, polyunsaturated fatty acids, carotenoids, and ascorbic acid in foods (Fig. 11) (Yaylayan, 2006). The reactivity of sugars for
furan formation is in the order of erythrose > ribose > sucrose > glucose ¼ fructose (Perez Locas and Yaylayan, 2004). It is
a very volatile compound, and its presence in foods highly depends on the characteristics of foods. In caramel colors, its concen-
tration can be ranged from 52 to 177 mg kg1 (EFSA, 2011).
The main furan compound formed during heating of hexoses is HMF, and various other furan compounds is also formed to
a lesser extent as mentioned above. HMF can reach extremely high concentrations up to 30 g kg1 in caramel products (EFSA,
2011), and its levels might exceed 1 g kg1 in dried fruits, roasted coffee, malt, vinegar, and various thermally treated sugar-rich
foods (Morales, 2008). Its metabolite 5-sulfoxymethyl-2-furfural is of serious concern because of its genotoxicity (Capuano and
Fogliano, 2011). Formation of furan derivatives highly depends on the thermal load applied, and therefore limiting thermal input
is one of the main strategies that could be applied. Formulation changes, such as choice of sugar, is also important to mitigate furan
derivatives (Şenyuva and Gökmen, 2007; Gökmen et al., 2008; Kocada glı et al., 2012; Van der Fels-Klerx et al., 2014; Taş and
Gökmen, 2016).

a-Dicarbonyl Compounds
a-Dicarbonyl compounds are highly reactive compounds, and therefore their presence might be a concern both in foods and in vivo.
In addition to their cytotoxic effects, they may also react with free or bound lysine and arginine, yielding advance glycation end
products, which are related to various health consequences (Henle, 2007; Hellwig and Henle, 2014). a-Dicarbonyl compounds
found in foods may cause dicarbonyl stress mainly in the gastrointestinal lumen (Degen et al., 2013, Degen et al., 2014; Rabbani

Amino acids
(serine, cysteine, alanine, threonine, and aspartic acid)

aldol condensation

Carbohydrates
(including ascorbic acid)

aldotetrose derivative

4-hydroxy-2-butenal

[O]

PUFA
Figure 11 Furan formation pathways. Adapted from Perez Locas, C., Yaylayan, V.A., 2004. Origin and mechanistic pathways of formation of the
parent furan: a food toxicant. J. Agric. Food Chem. 52, 6830–6836.
10 Caramelization in Foods: A Food Quality and Safety Perspective

Figure 12 Reaction of methylglyoxal with ()-epicatechin via electrophilic aromatic substitution.

and Thornalley, 2015). It has been shown that dietary exposure to glyoxal has tumor growth–promoting properties in the small
intestine in mice (Svendsen et al., 2016). Moreover, a-dicarbonyl compounds play a role in the formation of other toxigenic
compounds during food processing, such as acrylamide, furan, heterocyclic aromatic amines, and 4(5)-methylimidazole.
The major a-dicarbonyl compound found in foods is 3-deoxyglucosone, among others (Degen et al., 2012; Gensberger et al.,
2012, 2013; Kocada glı and Gökmen, 2014). The concentration of 3-deoxyglucosone is generally quite higher than HMF. Thermal
input, sugar composition, and pH are critical factors for the formation of a-dicarbonyl compounds. Phenolic compounds are
known to trap a-dicarbonyl compounds in foods and in physiological conditions in vitro (Totlani and Peterson, 2005; Lo et al.,
2006). a-Dicarbonyl compounds react with certain positions on the phenol ring via electrophilic aromatic substitution as shown
for ()-epicatechin in Fig. 12.

References

Ajandouz, E.H., Puigserver, A., 1999. Nonenzymatic browning reaction of essential amino acids: effect of ph on caramelization and Maillard reaction kinetics. J. Agric. Food Chem.
47, 1786–1793.
Ajandouz, E.H., Tchiakpe, L.S., Dalle Ore, F., Benajiba, A., Puigserver, A., 2001. Effects of ph on caramelization and Maillard reaction kinetics in fructose-lysine model systems.
J. Food Sci. 66, 926–931.
Angyal, S.J., 2001. The Lobry de Bruyn-Alberda van Ekenstein transformation and related reactions. In: Stütz, A.E. (Ed.), Glycoscience: Epimerisation, Isomerisation and Rear-
rangement Reactions of Carbohydrates. Springer Berlin Heidelberg, Berlin, Heidelberg.
Antal Jr., M.J., Mok, W.S., Richards, G.N., 1990. Mechanism of formation of 5-(hydroxymethyl)-2-furaldehyde from D-fructose and sucrose. Carbohydr. Res. 199, 91–109.
Audemar, M., Atencio-Genes, L., Ortiz Mellet, C., Jérôme, F.G., Fernandez, J.M., De Oliveira Vigier, K., 2017. Carbon dioxide as a traceless caramelization promotor: preparation of
prebiotic difructose dianhydrides (DFAS)-enriched caramels from D-fructose. J. Agric. Food Chem. 65, 6093–6099.
Belitz, H.D., Grosch, W., Schieberle, P., 2009. Food Chemistry. Springer-Verlag, Berlin.
Berg, H.E., 1993. Reactions of Lactose during Heat Treatment of Milk: A Quantitative Study. PhD Thesis. Wageningen University.
Bharate, S.S., Bharate, S.B., 2014. Non-enzymatic browning in citrus juice: chemical markers, their detection and ways to improve product quality. J. Food Sci. Technol. 51,
2271–2288.
Capuano, E., Fogliano, V., 2011. Acrylamide and 5-hydroxymethylfurfural (HMF): a review on metabolism, toxicity, occurrence in food and mitigation strategies. LWT-Food Sci.
Technol. 44, 793–810.
Cutzach, I., Chatonnet, P., Henry, R., Dubourdieu, D., 1999. Identifying new volatile compounds in toasted oak. J. Agric. Food Chem. 47, 1663–1667.
Davidek, T., Devaud, S., Robert, F., Blank, I., 2005. The effect of reaction conditions on the origin and yields of acetic acid generated by the Maillard reaction. Mail. React. Chem.
Interface Nutr. Aging, Dis. 1043, 73–79.
Davidek, T., Robert, F., Devaud, S., Vera, F.A., Blank, I., 2006. Sugar fragmentation in the Maillard reaction cascade: formation of short-chain carboxylic acids by a new oxidative
alpha-dicarbonyl cleavage pathway. J. Agric. Food Chem. 54, 6677–6684.
Degen, J., Hellwig, M., Henle, T., 2012. 1,2-Dicarbonyl compounds in commonly consumed foods. J. Agric. Food Chem. 60, 7071–7079.
Degen, J., Vogel, M., Richter, D., Hellwig, M., Henle, T., 2013. Metabolic transit of dietary methylglyoxal. J. Agric. Food Chem. 61, 10253–10260.
Degen, J., Beyer, H., Heymann, B., Hellwig, M., Henle, T., 2014. Dietary influence on urinary excretion of 3-deoxyglucosone and its metabolite 3-deoxyfructose. J. Agric. Food
Chem. 62, 2449–2456.
EFSA Panel on Food Additives and Nutrient Sources Added to Food (Ans), 2011. Scientific opinion on the re-evaluation of caramel colours (E 150 a,b,c,d) as food additives. EFSA
J. 9.
Elsinghorst, P.W., Raters, M., Dingel, A., Fischer, J., Matissek, R., 2013. Synthesis and application of 13C-labeled 2-acetyl-4-((1r,2s,3r)-1,2,3,4-tetrahydroxybutyl)imidazole (THI),
an immunosuppressant observed in caramel food colorings. J. Agric. Food Chem. 61, 7494–7499.
Fujioka, K., Shibamoto, T., 2004. Formation of genotoxic dicarbonyl compounds in dietary oils upon oxidation. Lipids 39, 481–486.
Garcı ́a-Baños, J.L., Villamiel, M., Olano, A.N., Rada-Mendoza, M., 2004a. Study on nonenzymatic browning in cookies, crackers and breakfast cereals by maltulose and furosine
determination. J. Cereal Sci. 39, 167–173.
Garcı ́a-Baños, J.L., Corzo, N., Sanz, M.L., Olano, A., 2004b. Maltulose and furosine as indicators of quality of pasta products. Food Chem. 88, 35–38.
Gensberger, S., Mittelmaier, S., Glomb, M.A., Pischetsrieder, M., 2012. Identification and quantification of six major alpha-dicarbonyl process contaminants in high-fructose corn
syrup. Anal. Bioanal. Chem. 403, 2923–2931.
Gensberger, S., Glomb, M.A., Pischetsrieder, M., 2013. Analysis of sugar degradation products with a-dicarbonyl structure in carbonated soft drinks by UHPLC-DAD-MS/MS.
J. Agric. Food Chem. 61, 10238–10245.
Girardon, P., Sauvaire, Y., Baccou, J.C., Bessiere, J.M., 1986. Identification of 3-hydroxy-4,5-dimethyl-2(5h)-furanone in aroma of fenugreek seeds (Trigonella-foenum-graecum l).
Lebensmittel-Wissenschaft Technol. 19, 44–46.
Gobert, J., Glomb, M.A., 2009. Degradation of glucose: reinvestigation of reactive a-dicarbonyl compounds. J. Agric. Food Chem. 57, 8591–8597.
Gökmen, V., Senyuva, H.Z., 2006. Improved method for the determination of hydroxymethylfurfural in baby foods using liquid chromatography-mass spectrometry. J. Agric. Food
Chem. 54, 2845–2849.
Gökmen, V., Acar, Ö.C., Köksel, H., Acar, J., 2007. Effects of dough formula and baking conditions on acrylamide and hydroxymethylfurfural formation in cookies. Food Chem. 104,
1136–1142.
Caramelization in Foods: A Food Quality and Safety Perspective 11

Gökmen, V., Acar, Ö.C., Serpen, A., Morales, F.J., 2008. Effect of leavening agents and sugars on the formation of hydroxymethylfurfural in cookies during baking. Eur. Food Res.
Technol. 226, 1031–1037.
Gökmen, V., Kocadaglı, T., Göncüoglu, N., Mogol, B.A., 2012. Model studies on the role of 5-hydroxymethyl-2-furfural in acrylamide formation from asparagine. Food Chem. 132,
168–174.
Golon, A., Kuhnert, N., 2012. Unraveling the chemical composition of caramel. J. Agric. Food Chem. 60, 3266–3274.
Golon, A., Kuhnert, N., 2013. Characterisation of “caramel-type” thermal decomposition products of selected monosaccharides including fructose, mannose, galactose, arabinose
and ribose by advanced electrospray ionization mass spectrometry methods. Food Funct. 4, 1040–1050.
Göncüoglu, N., Gökmen, V., 2013. Accumulation of 5-hydroxymethylfurfural in oil during frying of model dough. J. Am. Oil Chem. Soc. 90, 413–417.
Goto, R., Miyagi, Y., Inokawa, H., 1963. Syntheses and structures of acetylformoin and its related compounds. I. Bull. Chem. Soc. Jpn. 36, 147–151.
Haleva-Toledo, E., Naim, M., Zehavi, U., Rouseff, R.L., 1997. 4-hydroxy-2,5-dimethyl-3(2H)-furanone formation in buffers and model solutions of citrus juice. J. Agric. Food Chem.
45, 1314–1319.
Hellwig, M., Henle, T., 2014. Baking, ageing, diabetes: a short history of the Maillard reaction. Angew. Chem. Int. Ed. 53, 10316–10329.
Hellwig, M., Degen, J., Henle, T., 2010. 3-Deoxygalactosone, a “new” 1,2-dicarbonyl compound in milk products. J. Agric. Food Chem. 58, 10752–10760.
Hengel, M., Shibamoto, T., 2013. Carcinogenic 4(5)-methylimidazole found in beverages, sauces, and caramel colors: chemical properties, analysis, and biological activities.
J. Agric. Food Chem. 61, 780–789.
Henle, T., 2007. Dietary advanced glycation end products – a risk to human health? A call for an interdisciplinary debate. Mol. Nutr. Food Res. 51, 1075–1078.
Hirsch, J., Mossine, V.V., Feather, M.S., 1995. The detection of some dicarbonyl intermediates arising from the degradation of Amadori compounds (The Maillard reaction).
Carbohydr. Res. 273, 171–177.
Hofmann, T., 1998. Acetylformoin – a chemical switch in the formation of colored maillard reaction products from hexoses and primary and secondary amino acids. J. Agric. Food
Chem. 46, 3918–3928.
Hofmann, T., Schieberle, P., 2000. Acetylformoin – an important progenitor of 4-hydroxy-2,5-dimethyl-3(2H)-furanone and 2-acetyltetrahydropyridine during thermal food pro-
cessing. In: Rothe, M. (Ed.), Flavour 2000 – Perception, Release, Evaluation, Formation, Acceptance, Nutrition/health, Germany.
Hollnagel, A., Kroh, L.W., 1998. Formation of alpha-dicarbonyl fragments from mono- and disaccharides under caramelization and Maillard reaction conditions. Zeitschrift Fur
Lebensmittel-Untersuchung Und-Forschung a-Food Res. Technol. 207, 50–54.
Illmann, S., Davidek, T., Gouézec, E., Rytz, A., Schuchmann, H.P., Blank, I., 2009. Generation of 4-hydroxy-2,5-dimethyl-3(2H)-furanone from rhamnose as affected by reaction
parameters: experimental design approach. J. Agric. Food Chem. 57, 2889–2895.
Kerler, J., Winkel, C., Davidek, T., Blank, I., 2010. Basic chemistry and process conditions for reaction flavours with particular focus on Maillard-type reactions. In: Food Flavour
Technology. Wiley-Blackwell, Oxford, UK.
Kocadaglı, T., Gökmen, V., 2014. Investigation of alpha-dicarbonyl compounds in baby foods by high-performance liquid chromatography coupled with electrospray ionization mass
spectrometry. J. Agric. Food Chem. 62 (31), 7714–7720.
Kocadaglı, T., Gökmen, V., 2016a. Effect of sodium chloride on a-dicarbonyl compound and 5-hydroxymethyl-2-furfural formations from glucose under caramelization conditions:
a multiresponse kinetic modeling approach. J. Agric. Food Chem. 64, 6333–6342.
Kocadaglı, T., Gökmen, V., 2016b. Multiresponse kinetic modelling of Maillard reaction and caramelisation in a heated glucose/wheat flour system. Food Chem. 211, 892–902.
Kocadaglı, T., Gökmen, V., 2016c. Effects of sodium chloride, potassium chloride, and calcium chloride on the formation of alpha-dicarbonyl compounds and furfurals and the
development of browning in cookies during baking. J. Agric. Food Chem. 64, 7838–7848.
Kocadaglı, T., Göncüoglu, N., Hamzalıoglu, A., Gökmen, V., 2012. In depth study of acrylamide formation in coffee during roasting: role of sucrose decomposition and lipid oxidation.
Food Funct. 3, 970–975.
Kroeplien, U., Rosdorfer, J., Van Der Greef, J., Long, R.C., Goldstein, J.H., 1985. 2-acetyl-4(5)-(1,2,3,4-tetrahydroxybutyl)imidazole: detection in commercial caramel color III and
preparation by a model browning reaction. J. Org. Chem. 50, 1131–1133.
Kroh, L.W., 1994. Caramelisation in food and beverages. Food Chem. 51, 373–379.
Kroh, L.W., Fiedler, T., Wagner, J., 2008. Alpha-dicarbonyl compounds-key intermediates for the formation of carbohydrate-based melanoidins. Ann. N. Y. Acad. Sci. 1126,
210–215.
Lee, K.G., Jang, H., Shibamoto, T., 2013. Formation of carcinogenic 4(5)-methylimidazole in caramel model systems: a role of sulphite. Food Chem. 136, 1165–1168.
Lo, C.Y., Li, S.M., Tan, D., Pan, M.H., Sang, S.M., Ho, C.T., 2006. Trapping reactions of reactive carbonyl species with tea polyphenols in simulated physiological conditions. Mol.
Nutr. Food Res. 50, 1118–1128.
Maarse, H., 1991. Volatile Compounds in Foods and Beverages. Marcel Dekker, New York.
Martins, S.I.F.S., Van Boekel, M. a. J.S., 2005. Kinetics of the glucose/glycine maillard reaction pathways: influences of ph and reactant initial concentrations. Food Chem. 92,
437–448.
Martins, S.I.F.S., Jongen, W.M.F., Van Boekel, M. a. J.S., 2000. A review of Maillard reaction in food and implications to kinetic modelling. Trends Food Sci. Technol. 11, 364–373.
Mayes, H.B., Nolte, M.W., Beckham, G.T., Shanks, B.H., Broadbelt, L.J., 2014. The alpha–bet(a) of glucose pyrolysis: computational and experimental investigations of
5-hydroxymethylfurfural and levoglucosan formation reveal implications for cellulose pyrolysis. ACS Sustain. Chem. Eng. 2, 1461–1473.
Mayes, H.B., Nolte, M.W., Beckham, G.T., Shanks, B.H., Broadbelt, L.J., 2015. The alpha–bet(a) of salty glucose pyrolysis: computational investigations reveal carbohydrate
pyrolysis catalytic action by sodium ions. ACS Catal. 5, 192–202.
Mellet, C.O., Garcia Fernandez, J.M., 2010. Difructose dianhydrides (DFAs) and DFA-enriched products as functional foods. Top. Curr. Chem. 294, 49–77.
Mogol, B.A., Gökmen, V., 2014a. Computer vision-based analysis of foods: a non-destructive colour measurement tool to monitor quality and safety. J. Sci. Food Agric. 94,
1259–1263.
Mogol, B.A., Gökmen, V., 2014b. Mitigation of acrylamide and hydroxymethylfurfural in biscuits using a combined partial conventional baking and vacuum post-baking process:
preliminary study at the lab scale. Innov. Food Sci. Emerg. Technol. 26, 265–270.
Monographs, I.A.R.C., 1995. Dry Cleaning, Some Chlorinated Solvents and Other Industrial Chemicals, vol. 63. International Agency for Research on Cancer (IARC).
Monographs, I.A.R.C., 2013. Some Chemicals Present in Industrial and Consumer Products, Food and Drinking-water, 4-Methylimidazole and 2-methylimidazole, vol. 101.
International Agency for Research on Cancer (IARC).
Moon, J.K., Shibamoto, T., 2011. Formation of carcinogenic 4(5)-methylimidazole in maillard reaction systems. J. Agric. Food Chem. 59, 615–618.
Morales, F.J., 2008. Hydroxymethylfurfural (hmf) and related compounds. In: Stadler, R.H., Lineback, D.R. (Eds.), Process-induced Food Toxicants. John Wiley & Sons, Inc., New
Jersey.
Morales, V., Olano, A., Corzo, N., 2004. Ratio of maltose to maltulose and furosine as quality parameters for infant formula. J. Agric. Food Chem. 52, 6732–6736.
Moreau, L., Bindzus, W., Hill, S., 2009a. Influence of sodium chloride on color development of cereal model systems through changes in glass transition temperature and water
retention. Cereal Chem. 86, 232–238.
Moreau, L., Lagrange, J., Bindzus, W., Hill, S., 2009b. Influence of sodium chloride on colour, residual volatiles and acrylamide formation in model systems and breakfast cereals.
Int. J. Food Sci. Technol. 44, 2407–2416.
Moreau, L., Bindzus, W., Hill, S., 2011. Influence of salts on starch degradation: part II – salt classification and caramelisation. Starch-Starke 63, 676–682.
Mottier, P., Mujahid, C., Tarres, A., Bessaire, T., Stadler, R.H., 2017. Process-induced formation of imidazoles in selected foods. Food Chem. 228, 381–387.
Nikolov, P.Y., Yaylayan, V.A., 2011a. Reversible and covalent binding of 5-(hydroxymethyl)-2-furaldehyde (HMF) with lysine and selected amino acids. J. Agric. Food Chem. 59,
6099–6107.
12 Caramelization in Foods: A Food Quality and Safety Perspective

Nikolov, P.Y., Yaylayan, V.A., 2011b. Thermal decomposition of 5-(hydroxymethyl)-2-furaldehyde (HMF) and its further transformations in the presence of glycine. J. Agric. Food
Chem. 59, 10104–10113.
Paravisini, L., Gourrat-Pernin, K., Gouttefangeas, C., Moretton, C., Nigay, H., Dacremont, C., Guichard, E., 2012. Identification of compounds responsible for the odorant properties
of aromatic caramel. Flavour Fragr. J. 27, 424–432.
Perez Locas, C., Yaylayan, V.A., 2004. Origin and mechanistic pathways of formation of the parent furan: a food toxicant. J. Agric. Food Chem. 52, 6830–6836.
Perez Locas, C., Yaylayan, V.A., 2008. Isotope labeling studies on the formation of 5-(hydroxymethyl)-2-furaldehyde (HMF) from sucrose by pyrolysis-GC/MS. J. Agric. Food Chem.
56, 6717–6723.
Rabbani, N., Thornalley, P.J., 2015. Dicarbonyl stress in cell and tissue dysfunction contributing to ageing and disease. Biochem. Biophys. Res. Commun. 458, 221–226.
Ratsimba, V., Fernández, J.M.G.A., Defaye, J., Nigay, H., Voilley, A., 1999. Qualitative and quantitative evaluation of mono- and disaccharides in d-fructose, d-glucose and sucrose
caramels by gas–liquid chromatography–mass spectrometry: di-d-fructose dianhydrides as tracers of caramel authenticity. J. Chromatogr. A 844, 283–293.
Roos, Y.H., Drusch, S., 2016. Chapter 2-physical state and molecular mobility. In: Phase Transitions in Foods, second ed. Academic Press, San Diego.
Roos, Y.H., Karel, M., Labuza, T.P., Levine, H., Mathlouthi, M., Reid, D., Shalaev, E., Slade, L., 2013. Melting and crystallization of sugars in high-solids systems. J. Agric. Food
Chem. 61, 3167–3178.
Schulz, A., Trage, C., Schwarz, H., Kroh, L.W., 2007. Electrospray ionization mass spectrometric investigations of alpha-dicarbonyl compounds – probing intermediates formed in
the course of the nonenzymatic browning reaction of L-ascorbic acid. Int. J. Mass Spectrom. 262, 169–173.
Şenyuva, H.Z., Gökmen, V., 2007. Potential of furan formation in hazelnuts during heat treatment. Food Addit. Contam. 24, 136–142.
Smuda, M., Glomb, M.A., 2013. Fragmentation pathways during Maillard-induced carbohydrate degradation. J. Agric. Food Chem. 61, 10198–10208.
Speck Jr., J.C., 1958. The Lobry de Bruyn-Alberda van Ekenstein transformation. Adv. Carbohydr. Chem. 13, 63–103.
Svendsen, C., Høie, A.H., Alexander, J., Murkovic, M., Husøy, T., 2016. The food processing contaminant glyoxal promotes tumour growth in the multiple intestinal neoplasia (min)
mouse model. Food Chem. Toxicol. 94, 197–202.
Taş, N.G., Gökmen, V., 2016. Effect of alkalization on the Maillard reaction products formed in cocoa during roasting. Food Res. Int. 89, 930–936.
Tokitomo, Y., Kobayashi, A., Yamanishi, T., Muraki, S., 1980. Studies on the sugary flavor of raw cane sugar. 3. Key compound of the sugary flavor. Proc. Jpn. Acad. Ser. B-Phys.
Biol. Sci. 56, 457–462.
Totlani, V.M., Peterson, D.G., 2005. Reactivity of epicatechin in aqueous glycine and glucose maillard reaction models: Quenching of c(2), c(3), and c(4) sugar fragments. J. Agric.
Food Chem. 53, 4130–4135.
Van Der Fels-Klerx, H.J., Capuano, E., Nguyen, H.T., Mogol, B.A., Kocadaglı, T., Taş, N.G., Hamzalıoglu, A., Van Boekel, M.A.J.S., Gökmen, V., 2014. Acrylamide and 5-
hydroxymethylfurfural formation during baking of biscuits: Nacl and temperature-time profile effects and kinetics. Food Res. Int. 57, 210–217.
Velisek, J., 2014. The Chemistry of Food, UK. John Wiley & Sons.
Voigt, M., Glomb, M.A., 2009. Reactivity of 1-deoxy-d-erythro-hexo-2,3-diulose: a key intermediate in the Maillard chemistry of hexoses. J. Agric. Food Chem. 57, 4765–4770.
Wu, X., Yu, D., Kong, F., Yu, S., 2016. Effects of divalent cations on the formation of 4(5)-methylimidazole in fructose/ammonium hydroxide caramel model reaction. Food Chem.
201, 253–258.
Yaylayan, V.A., 2006. Precursors, formation and determination of furan in food. J. für Verbraucherschutz und Lebensmittelsicherheit 1, 5–9.
Yaylayan, V.A., Haffenden, L.J.W., 2003. Mechanism of imidazole and oxazole formation in [C-13-2]-labelled glycine and alanine model systems. Food Chem. 81, 403–409.
Yaylayan, V.A., Mandeville, S., 1994. Stereochemical control of maltol formation in Maillard reaction. J. Agric. Food Chem. 42, 771–775.
Zabetakis, I., Gramshaw, J.W., Robinson, D.S., 1999. 2,5-dimethyl-4-hydroxy-2H-furan-3-one and its derivatives: analysis, synthesis and biosynthesis – a review. Food Chem. 65,
139–151.
Zhao, H., Holladay, J.E., Brown, H., Zhang, Z.C., 2007. Metal chlorides in ionic liquid solvents convert sugars to 5-hydroxymethylfurfural. Science 316, 1597–1600.

You might also like