Está en la página 1de 12

Review

Melanoidins produced by the Maillard reaction: Structure and biological activity


He-Ya Wang

, He Qian, Wei-Rong Yao


School of Food Science and Technology, Jiangnan University, Wuxi 214122, Jiangsu, China
a r t i c l e i n f o
Article history:
Received 14 November 2010
Received in revised form 22 January 2011
Accepted 15 March 2011
Available online 21 March 2011
Keywords:
Maillard reaction
Melanoidins
Structural properties
Biological effects
Antioxidant activity
a b s t r a c t
Melanoidins are compounds generated in the late stages of the Maillard reaction from reducing sugars
and proteins or amino acids during food processing and preservation. Recently the effects of melanoidins
on human health and the chemical characterisation of the benecial components have gained a lot of
attention. Food melanoidins have been reported to be anionic, coloured compounds and some of their
key chromophores have been elucidated. The antioxidant activity and other biological effects of melanoi-
dins from real foods and model systems have been widely studied. Despite this, very few different mel-
anoidin structures have actually been described, and specic health effects have yet to be linked to
chemically distinct melanoidins. The variety of different Maillard reaction products formed during the
reaction, in conjunction with the difculty in purifying and identifying them, makes a thorough analysis
of melanoidins challenging. This review provides a comprehensive look at what is known to date about
melanoidin structure, the formation mechanism for these compounds, and the biological properties
related to the benecial health effects of melanoidins.
2011 Elsevier Ltd. All rights reserved.
Contents
1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573
2. Chemical and structural properties of melanoidins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
2.1. The negative charge of melanoidins. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
2.2. Molecular weight of melanoidins. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
2.3. Structural properties and the likely formation mechanism of melanoidins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 574
2.4. Quantification of model melanoidins. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 578
3. The biological properties of melanoidins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
3.1. Genotoxicity and cytotoxicity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
3.2. Antioxidative activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 579
3.3. Antimicrobial activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 580
3.4. Effect of melanoidins on xenobiotic enzymes activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 580
3.5. Prebiotic activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
3.6. Antihypertensive activity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
3.7. Other biological effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
3.8. The in vivo effects of melanoidins and the effects of melanoidins on foods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 581
4. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 582
1. Introduction
Melanoidins are heterogeneous, nitrogen-containing brown
pigments produced by the Maillard reaction (MR). They can absorb
light at wavelengths as high as 420 nm and are predominantly
0308-8146/$ - see front matter 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.foodchem.2011.03.075

Corresponding author. Tel.: +86 510 85328731; fax: +86 510 85329081.
E-mail address: heyawang@163.com (H.-Y. Wang).
Food Chemistry 128 (2011) 573584
Contents lists available at ScienceDirect
Food Chemistry
j our nal homepage: www. el sevi er . com/ l ocat e/ f oodchem
responsible for the characteristic brown colour of foods such as
coffee, cocoa, bread, malt, and honey (Lindenmeier, Faist, & Hofmann,
2002). The Maillard reaction is a chemical reaction that occurs be-
tween the carbonyl groups of reducing sugars and the amino
groups of amino acids, peptides or proteins. This non-enzymatic
browning reaction occurs during the preparation of many different
foods. Since its discovery in 1912, this reaction has been applied
widely in the development of colour, aroma and avour precursors
for food applications.
The Maillard reaction products (MRPs) are a particularly com-
plex mix of various compounds of different molecular weights.
They include not only aldehydes, ketones, dicarbonyls, acryl
amides, and heterocyclic amines, all of which contribute to avour,
but also melanoidins and advanced glycation endproducts (AGEs),
which are polymeric products formed at the advanced steps of MR.
Melanoidins are of particular interest to the food industry as the
brown colour imparted by these compounds is intimately associ-
ated in consumers minds with a high-grade product (Lindenmeier
et al., 2002) of desirable texture and avour (Andriot, Le Quere, &
Guichard, 2004).
The chemical characterisation of melanoidins and the health ef-
fects of chemically distinct melanoidins are hot research areas in
food science. Though the chemical structures and health effects
of melanoidins produced in both real foods and model systems
have been investigated for over 20 years, the health effects specic
to structurally distinct melanoidins are not well understood. The
complex array of melanoidins produced in the MR is strongly
dependent on the type of food, as well as the technological condi-
tions of the reaction (Martins & van Boekel, 2003) such as treating
temperature and time (Borrelli, Visconti, Mennella, Anese, &
Fogliano, 2002; Cosovic, Vojvodic, Boskovic, Plavsic, & Lee, 2010;
Lindenmeier & Hofmann, 2003; Ramonaityte, Kersiene, Adams,
Tehrani, & Kimpe, 2009; Verzelloni, Tagliazucchi, & Conte,
2010a), pH (Kim & Lee, 2008; Kwak, Lee, Murata, & Homma,
2005), solvent (Hofmann, 1998d), and the composition of amino
acids and reducing sugars (Kim & Lee, 2008). In this review, mela-
noidin structure, formation mechanism, and biological effects are
outlined to provide a comprehensive view of this complex com-
pound mixture.
2. Chemical and structural properties of melanoidins
Melanoidins are formed by cyclizations, dehydrations, retroald-
olizations, rearrangements, isomerizations, and condensations of
low molecular weight MRPs. The composition of melanoidin chem-
ical structures is relatively unknown, however, due to the com-
plexity of the products that are generated in the reaction
(Bekedam, Roos, Schols, Van Boekel, & Smit, 2008; Kim & Lee,
2009).
2.1. The negative charge of melanoidins
Although melanoidins are chemically diverse, many studies re-
port that melanoidins are negatively charged in both real foods and
in model systems. Melanoidins found in coffee were determined to
be negatively charged but heterogeneous with respect to their
polyanionic behaviour (Bekedam, De Laat, Schols, Van Boekel, &
Smit, 2007; Bekedam, Schols, van Boekel, & Smit, 2006): high
molecular weight (HMW) melanoidins in coffee were found to be
more negatively charged than low molecular weight (LMW) mela-
noidins (Bekedam, Roos et al., 2008). Chlorogenic acids have been
hypothesised to be the source of the negative charges in coffee
melanoidins (Bekedam, Schols, van Boekel, & Smit, 2008), though
melanoidins obtained from sugaramino acid model systems also
showed anionic characteristics in the absence of chlorogenic acids
(Cosovic et al., 2010; Morales, 2002). For example, Kwak et al.
(2005) showed that melanoidins prepared by reuxing glucose
and lysine could be separated into 14 bands over a pI range of
3.54.85, indicating that melanoidins were negatively charged at
neutral pH. Under these conditions, the type of amino acid present
during the reaction determined the anionic properties of the mel-
anoidins (Morales, 2002).
2.2. Molecular weight of melanoidins
It has been assumed that melanoidins are nitrogen-containing,
high molecular weight (HMW), coloured compounds (Hofmann,
1998e). More recent reports, however, have indicated that mela-
noidins include a LMW fraction, too (Cosovic et al., 2010; Faist,
Lindenmeier, Geisler, Erbersdobler, & Hofmann, 2002; Glosl et al.,
2004). In real foods, most of the melanoidins have been shown to
be HMW compounds. For example, the browning intensity of the
ethanol extracts of bread crust (Lindenmeier et al., 2002), manuka,
and dandelion honey (Brudzynski & Miotto, 2011a) increased with
increasing molecular weight upon heating, demonstrating that the
browning is due in large part to HMW melanoidins. In coffee, 59%
of the melanoidins are HMW (>1214 kDa). It has been found that
prolonged roasting leads predominantly to the formation of HMW
melanoidins (Bekedam, Loots, Schols, Van Boekel, & Smit, 2008;
Bekedam, Roos et al., 2008), which show a more intense brown col-
our than the LMW melanoidins (Bekedam et al., 2006, 2007). Sim-
ilarly, polymers of HMW melanoidins isolated from sweet wines
(>12 kDa) (Rivero-Perez, Perez-Magarino, & Gonzalez-San Jos,
2002), roasted malt (>60 kDa) (Faist, Lindenmeier et al., 2002)
and roasted cocoa beans (>5 kDa) (Summa et al., 2008), have been
shown to give rise to the brown colour of these foodstuffs.
Many studies have shown that melanoidins produced in model
MR systems heated for more than 24 h are predominantly of
molecular weights greater than 10 kDa (Cosovic et al., 2010; Ibarz,
Garvin, Garza, & Pagan, 2009; Morales, 2002; Morales & Jimnez-Prez,
2004; Ruan-Henares & Morales, 2007c). These studies suggested
that melanoidin pigments were predominantly HMW. Some other
reaction mixtures and conditions, however, such as glucose and
fructoseamino acids heated to 100 C for 2 h (Kim & Lee, 2008),
lactose and glycine under reuxing conditions for 5 h (Ramo-
naityte et al., 2009), and glucose and an alanine/glycine mixture
heated to 95 C for 4 h (Hofmann, 1998e), produced primarily
LMW coloured compounds (<3.5 kDa). However, the reaction be-
tween casein and glucose generates a sizable HMW fraction of pig-
mented melanoidins (Hofmann, 1998e).
Melanoidins produced in the Maillard reactions between pro-
teins and sugars in real foods are predominantly HMW com-
pounds. Considering that the molecular weight of melanoidins
produced from MRs is highly dependant on the heating intensity,
and HMW melanoidins are produced at longer reaction times
(>24 h), it is possible that in the initial stages of the MR, LMW chro-
mophoric melanoidins are formed, which subsequently polymerise
or cross-link with other MRPs to produce HMW melanoidins dur-
ing the later stages of the MR.
2.3. Structural properties and the likely formation mechanism of
melanoidins
Although numerous attempts have been undertaken to isolate
and purify melanoidins from foods such as bread crust (Lindenme-
ier & Hofmann, 2003; Lindenmeier et al., 2002) and coffee
(Bekedam et al., 2006; Bekedam, Loots et al., 2008; Bekedam, Roos
et al., 2008; Gniechwitz et al., 2008), as well as from sugaramino
acid model systems (Ames, Bailey, & Mann, 1999; Frank &
Hofmann, 2000; Hofmann, 1998b, 1998c; Tressl, Wondrak, Kruger,
574 H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584
& Rewicki, 1998), little is known about the overall structural prop-
erties of melanoidins (Lindenmeier et al., 2002).
A few preliminary studies on melanoidin structure have been
conducted, though, and HMW melanoidin structures from water-
free model systems (glucosegluten heated to 150 C for 45 min
and glucoseglycine/glutamic acid heated to 125 C for 2 h)
(Adams, Abbaspour Tehrani, Kersiene, Venskutonis, & De Kimpe,
2003) and foodstuffs (Adams, Borrelli, Fogliano, & De Kimpe,
2005) have been reported. The melanoidins from model systems,
bread crust, coffee, and tomato sauce, elucidated from the volatile
proles after thermal degradation, were shown to be comprised
primarily of furans accompanied by carbonyl compounds, pyrroles,
pyrazines and pyridines.
LMW chromogenic compounds and coloured substructures of
melanoidin-type compounds (Fig. 1AD), produced by reuxing
neutral aqueous solutions of xylose/glucose and alanine/proline/
lysine for less than 5 h, were shown to consist of a similar chemical
mixture, that contained furan, pyrrole and pyrrolinone structures
and derivatives (Fig. 1AD) (Frank & Hofmann, 2000; Hofmann,
1998b, 1998c).
Blue and red pigments with polymerising activities have been
isolated from alkalic aqueous xyloseglycine, glucoseglycine
and xyloseb-alanine reactions conducted under nitrogen (pH
8.1, 60% ethanol, 26.5 C, 48 h). The blue pigments (pyrrolopyrrole
dimer) were determined to have a novel chemical structure con-
sisting of two pyrrolopyrrole rings combined with a methine
bridge (Fig. 1E), (Hayase et al., 2008; Sasaki, Shirahashi, Nishiyama,
Watanabe, & Hayase, 2006), and the red pigments (pyrrolopyrrole
azepine) were shown to have unique structures of pyrrolopyrrole,
pyrrole and azepin rings (Hayase, Usui, & Watanabe, 2006; Shirahashi,
Watanabe, & Hayase, 2009) (Fig. 1F). Yellow compounds (acetyl
pyridinone and acetyl azepinone), consisting of either a ring of
pyridinone (Fig. 1G) or azepinone (Fig. 1H), were also isolated from
the acidic xyloseglycine and glucoseglycine systems (reuxing
for 2 h at pH 5), respectively (Ames, Bailey et al., 1999). Moreover,
Tressl et al. (1998) reported an N-substituted pyrrole polymer
chromophore with extraordinary polycondensation activity
(Fig. 1I), formed from neutral pentose and methylamine in an
aqueous MR model system (120 C, 1 h). The polymerising activi-
ties of these LMW colourants suggests they are important reaction
intermediates in the formation of melanoidins (Hayase et al., 2006;
Tressl et al., 1998).
In model MR systems, pH appears to be an important factor
that inuences the structure of the chromophoric melanoidins,
while the reaction temperature and time dictate the molecular
weight of melanoidins. Three theories on HMW melanoidin for-
mation have been described. One theory is that HMW coloured
structures are formed by polymerisation of LMW Maillard reac-
tion intermediates, such as furans, pyrroles, pyrrolopyrroles,
and/or their derivatives, in the later stages of the reaction (Hayase
et al., 2006; Tressl et al., 1998). The polymerising activities of the
LMW colourants mentioned above suggest these compounds
could function as reaction intermediates during polycondensation,
making the molecular weight of the nal melanoidins dependant
on the heating time.
Hofmann (1998a, 1998e) has postulated that HMW melanoi-
dins are derived from cross-linking chromophoric LMW MRPs
and reactive amino acid side chains such as lysine, arginine, or cys-
teine. Several melanoidin-amino acid chromores have been de-
scribed. An intense red-brown chromophoric compound with the
substructure of furylimidazolidine dimer linked with two argi-
nine moieties (Fig. 2A) was detected in a heated solution of N
a
-
acetyl-L-arginine with glyoxal and furan-2-carboxaldehyde (90 C
for 80 min, at pH 7.0), an MR model system that contains arginines
(Hofmann, 1998a). Brownorange melanoproteins were detected
in heated aqueous reactions of casein and furan-2-carboxaldehyde
(70 C for 1 h, pH 7.0). These compounds were formed by linking
trifuryl-pyrrolinone chromophores (Fig. 1C) to the e-amino group
of a lysine residue (Fig. 2B) (Hofmann, 1998c). Another melanoidin
structure, protein-bound pyrrolinone reductonyl-lysine, also
known as pronyl-L-lysine (Fig. 2C), has been identied in bread
crust, malt and MRPs from an aqueous model system (100 C for
25 min, pH 5.5) containing the lysine side chain (Lindenmeier &
Hofmann, 2003; Lindenmeier et al., 2002). This compound is
responsible for the dark colour of these foods.
The chromophores of melanoproteins usually contain pyrroles,
imidazoles, and their nitrogen-containing derivatives. The nitrogen
atoms of these structures can be cross-linked to proteins via the
amino and guanidine groups of amino acids, as well as to the
furan substructures produced upon heating the carbohydrates
(Hofmann, 1998a, 1998c).
The third and nal theory proposes that the melanoidin skele-
ton is built primarily from sugar degradation products branched
with amino compounds, such as amino acids. It has been demon-
strated in aqueous systems that melanoidins are formed by aldol
condensations of highly reactive a-dicarbonyl compounds, which
are the main intermediates during the early stages of the MR,
and partially branched by amino compounds (Kroh, Fiedler, &
Wagner, 2008). In water-free reaction conditions, the transglycosy-
lation reactions that form the carbohydrate-skeleton of melanoi-
dins have been shown to take place during the early stages of
MRs (Cammerer, Jalyschko, & Kroh, 2002a). When using di- or oli-
gosaccharides as carbonyl components, the glycosidic bonds re-
main unchanged both in aqueous and water-free systems, though
a small fraction of the glycosidic bonds could be broken (Cammerer
et al., 2002a).
Therefore, the aldol condensations of a-dicarbonyl compounds
in aqueous conditions, in addition to the transglycosylation reac-
tions of saccharides in water-free conditions produced the postu-
lated carbohydrate-based skeleton shown in Fig. 3A. Amino acids
could react with the unsaturated carbonyl structures of this struc-
ture to form melanoidins branched with amino compounds
(Fig. 3B) (Cammerer, Jalyschkov, & Kroh, 2002b). Cosovic et al.
(2010) investigated the adsorption behaviour of melanoidins pre-
pared from glucose and amino acids via an electrochemical meth-
od. The observed similarity of the adsorption isotherm for
melanoidins prepared from glucose compared with those prepared
from glucose and amino acids supports the idea that melanoidins
have a sugar derived backbone.
Some of the important reactions that lead to the formation of
melanoidins are illustrated in Fig. 4. In the early stages of the
MR, reducing sugars react with amino compounds to form a Schiff
base adduct which is then stabilized by an Amadori rearrange-
ment. During the intermediate stages, a-dicarbonyls, aldehydes,
furaldehydes and furanone, are generated, which rapidly react
with each other in an aldol-type condensation. Also, degradation
of aldose sugars leads to the formation of furaldehydes and fura-
nones too, such as furan-2-carboxaldehyde and furan-2-aldehyde
(Marko et al., 2003; Shirahashi et al., 2009). A range of reactions
takes place in the advanced stages of the MR, including cyclisa-
tions, dehydrations, retro-adolisations, rearrangements, isomeri-
sations and additional condensations. Ultimately, these
intermediate products react with amino acids to form LMWs,
leading to the production of HMWs by polymerisation. HMW
MRPs can also be produced via the modication of proteins by
the intermediate products. Melanoidins with browning and AGEs
with uorescence are two main components of advanced MRPs,
but the choromophores of them are much different. It has been
reported that uorescence can be observed in the MR before
browning occurs, even though AGEs and melanoidins are gener-
ated at similar stages in the MR (Chuyen, 2006). Thus, the forma-
tion of LMW and HMW melanoidins depends on the degree of
H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584 575
polymerisation among the LMW intermediates and/or the degree
of cross-linking between LMW chromophores and non-coloured
HMW biopolymers.
In the production of real foods such as bread, coffee, roasted
malt, cocoa and biscuits, the MR often takes place at lower mois-
ture contents, at temperatures higher than 150 C, and usually for
N
OH
CH
3
R O
O
O
O
N
HO
H
3
C
R
O
A (Difuryl-pyrrolinone) B ( Pyrrolyl-furanone)
R: CH(CH
3
)COOH R: CH(CH
3
)COOH
O
CHO
N
O
O
O
R
*
O
N
O
O
O
O
C (Trifuryl-pyrrolinone) D (Trifuryl- pyrroloxazolidine)
R: CH(CH
3
)COOH (CH
2
)
4
CH(NH
2
)COOH
N
N
N
N
R2
CH
2
COOH
CH
2
COOH
CH
2
COOH
CH
2
COOH
R1
N
N
N
N
R1
CH
2
COOH
CH
2
COOH
R2
CH
2
COOH
CH
2
COOH
CH
2
OH
E (Pyrrolopyrrole dimer) F (Pyrrolopyrrole azepine)
R1: CH
2
CHOHCH
2
OH; R1: CH
2
CHOHCH
2
OH;
R2: CH
2
CHOHCH
2
OH, CHOH)
2
CH
2
OH R2: (CHOH)
3
CH
2
OH
N
O
H
HO
CH
3
O
N
O
CH
3
HO
HO
H
O
N
R
n
G (Acetyl pyridinone) H (Acetyl azepinone) I (Pyrrole polymer)
R CH
3
Fig. 1. Some LMW compounds with chromophores or substructures for melanoidins

Assymmetric carbon for isomer.
576 H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584
reaction times of less than 2 h. Under these conditions, the oligo-
and polysaccharides preferentially react as complete molecules at
the reducing end. Additional side chains could thus be formed by
transglycosylation reactions, producing HMW melanoidins with
carbohydrate skeletons (Cammerer et al., 2002a). Alternatively,
sugar or carbohydrate degradation products can react with pro-
tein to form HMW melanoidins with a protein skeleton. LMW
melanoidins can also be formed by the reaction of sugar or carbo-
hydrate degradation products with amino acids. In foods with
higher moisture content such as beer, sweet wine and grape syr-
up, the MR often takes place during the fermentation and storing,
usually at a temperature lower than 50 C for a period of more
than 30 days. Under these conditions, LMW melanoidins produced
by the reaction of sugars and amino acids at lower temperatures
and long reaction times allow for extensive polymerisation and
formation HMW melanoidins. Thus in food melanoidins, HMW
compounds are usually more enriched compared to LMW
compounds.
Melanoidin composition in real foods is more complex than in
model systems due to the more diverse pool of reactants (Adams
et al., 2005; Bekedam, Roos et al., 2008). Phenolics are constituents
of coffee melanoidins (Bekedam et al., 2006; Bekedam, Loots et al.,
2008; Gniechwitz et al., 2008), largely due to the incorporation of
chlorogenic acids (Bekedam, Roos et al., 2008; Bekedam, Schols
et al., 2008). Arabinogalactan proteins (AGP) were present as
AGPmelanoidin complexes in the HMW fraction of coffee mela-
noidins, too (Bekedam et al., 2007; Bekedam, Loots et al., 2008;
Bekedam, Roos et al., 2008). Similarly, total melanoidins and
HMW melanoidins of grape syrup and sweet wine have been
shown to be dependent on the presence or absence of polyphenols,
as the darkest pigments were formed in the presence of polyphe-
nols (Ortega-Heras & Gonzalez-Sanjose, 2009).
HN
N
N
HN
O
N
N
O
O
N
H
H
N
N
H
N
H
H
N
O
O
R1 R2
O
O
R3 R4
O
CHO
N
O
O
O
N
H
H
N
H
N
O
O
R2 R1
O
A (Furyl-imidazolidine dimer-protein) B (Trifuryl-pyrrolinone-protein)
N
HO
CH
3
N
H
H
N
H
N
O
O
R2 R1
O
H
3
C
O
OH
C (Pronyl-protein)
Fig. 2. The proposed melanoidin cross-linking structures of chromophoric LMW with HMW protein.
H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584 577
2.4. Quantication of model melanoidins
Although it is difcult to quantify melanoidin composition from
a Maillard reaction given the complexity of these reactions, quan-
tication is critical to optimising the browning in processed foods.
The extinction coefcient can be used to spectrophotometrically
estimate browning (absorbance values), as related to the melanoi-
din concentration (Martins & van Boekel, 2003). According to the
LambertBeer equation A = ecl, there is a direct linear relation
between absorbance (A) and concentration (c) (Martins & van Boe-
kel, 2003). In order to quantify melanoidins using this equation, the
extinction coefcient must be known.
The extinction coefcient of the melanoidins does not appear to
vary with pH, temperature or reaction time (Martins & van
Boekel, 2003). In a study of Brands, Wedzicha, and van Boekel
(2002), the extinction coefcient of melanoidins at 420 nm
(e
420
) produced from the glucosecasein and fructosecasein
reactions (120 C, 90 min, pH 6.8) were similar: 0.48 0.05 and
OH
OH
O-R OH
O
R-O O
HO
OH
OH
O-R OH
HO
HN
R-O O
HO
CH
COOH R'
A (Oligodeoxyhexosuloses) B (Oligodeoxyhexosuloses amine)
+ Amino acid ( -)
Fig. 3. Proposed carbohydrate-based melanoidin structure R: H, Glc, (Glc)
n
.
Aldose sugar
N-substitued
glycosylamine
-H
2
O
Amadori products
Ketoamine
Dicarbonyl
compounds
Aldehydes
Furaldehyde
and furanone
Aldol condensation
products
LMW advanced MRPs
HMW advanced MRPs
+ Protein
polymeration
+ Amino acids
+ Amino
compound
Fig. 4. Formation of advanced MRPs from the Maillard reaction.
578 H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584
0.53 0.04 l mmol
1
cm
1
, respectively. The e
420
of nondialyzable
(>3.5 kDa) glucose-glycine melanoidins (both at 120 C, pH 6.8
and at 100 C, pH 5.5) was found to be 1.01 0.02 l mmol
1
cm
1
and those of HMW (>12.4 kDa) melanoidins from glucose/fruc-
tose-glycine/asparagine/lysine basic reactions (at 145 C for 2 h)
fell between 0.500.76 l mmol
1
cm
1
(Kim & Lee, 2009; Martins
& van Boekel, 2003). Moreover, Mundt and Wedzicha (2003) deter-
mined a e
470
of 0.48 0.02 l mmol
1
cm
1
using the kinetic model
for melanoidins produced from the fructoseglycine reaction,
which is half the value determined for melanoidins from the glu-
cose-glycine reaction (0.96 0.04 l mmol
1
cm
1
).
Bekedam et al. (2006) demonstrated that use of the specic
extinction coefcient (K) was preferable to the molar extinction
coefcient (e), since the molecular weight of melanoidins is un-
known and probably variable. When using K, the concentration
parameter used in the law of LambertBeer is expressed in
l g
1
cm
1
(Bekedam et al., 2006). Deduced from the slope of the
curve representing the absorbance vs. melanoidin concentration,
the K values of glucose-amino acids melanoidins ranged from
0.225 to 4.315 l g
1
cm
1
for glucose-tryptophan and glucose-ly-
sine melanoidins, respectively (Ruan-Henares & Morales, 2007c).
3. The biological properties of melanoidins
Melanoidins appear in several heat-treated foods and are preva-
lent in our diet. Thus, the benecial and detrimental biological ef-
fects of melanoidins are of great interest.
3.1. Genotoxicity and cytotoxicity
It has been reported that LMW brown MRPs isolated from dry-
heated glucoseglycine systems (125 C, 2 h) (Glosl et al., 2004),
and HMW brown MRPs from aqueous sugar-lysine model systems
(121 C, 1 h, pH 9.0) (Jing & Kitts, 2000) have genotoxic and cyto-
toxic effects on cells in high doses. Bread crust and coffee melanoi-
dins were also shown to impair the viability of lung epithelial H358
cells in high doses (Bartling, Rehbein, Somoza, Silber, & Simm,
2005).
Jing and Kitts (2000) observed that HMW MRPs from fructose
lysine reactions with lower absorption values (at 420 nm) were
more cytotoxic than those from glucoselysine reactions, which
suggested that the melanoidin component of MRPs might not be
responsible for the observed toxicity. It has been reported that
brown glucose-glycine MRPs and cocoa melanoidin fractions are
not mutagenic to Salmonella typhimurium strains TA98, TA100
and TA102 (Ames test) at low or high concentrations (Glosl et al.,
2004; Summa et al., 2008). It has also been shown that MRPs con-
taining melanoidins produced from proteolyzed products of the
dry and wet glutenglucose model systems (Borrelli et al., 2003),
the caseinsugar system (Jing & Kitts, 2002), and bread and biscuit
(Borrelli et al., 2003; Martin et al., 2009) displayed no cytotoxicity
towards hepatocytes, HepG2 or Caco-2 cells. In conclusion, though
brown MRPs containing melanoidins may show modest genotoxic
and cytotoxic effects in high doses and after long incubation times
(Bartling et al., 2005; Glosl et al., 2004; Jing & Kitts, 2000), the low
concentrations found in foods do not pose a health risk.
3.2. Antioxidative activity
The benecial antioxidative effects of melanoidins are currently
gaining much attention. Studies have primarily focused on the
antioxidant properties of melanoidins or brown MRPs from differ-
ent model systems (Hayase et al., 2006; Jing & Kitts, 2000; Morales
& Jimnez-Prez, 2004) and melanoidin fractions in foods such as
coffee (Daglia et al., 2008; Gmez-Ruiz, Ames, & Leake, 2008;
Morales & Jimnez-Prez, 2004), bread crust (Lindenmeier et al.,
2002; Michalska, Amigo-Benavent, Zielinski, & del Castillo, 2008;
Somoza et al., 2005), beer (Morales & Jimnez-Prez, 2004), biscuit
(Martin et al., 2009), roasted barley (Papetti et al., 2006), roasted
cocoa (Summa et al., 2008), vinegar (Tagliazucchi, Verzelloni, &
Conte, 2010a; Xu, Tao, & Ao, 2007), honey (Brudzynski & Miotto,
2011a) and meat avouring (Wang, YU, & Qian, in press; Wang,
YU, QIAN, & YAO, 2010). These studies demonstrated that brown
chromophoric MRPs, independent of the reaction conditions, dis-
played in vitro antioxidant protective effects on lipids (Borrelli
et al., 2002; Tagliazucchi, Verzelloni, & Conte, 2010b; Verzelloni,
Tagliazucchi, & Conte, 2010b; Wang et al., 2010), rat microsomes
(Daglia et al., 2008; Papetti et al., 2006), hepatocytes (Lindenmeier
et al., 2002; Valls-Bells et al., 2004), HepG2 cells (Martin et al.,
2009) and human lymphocytes (Wang et al., 2010) against oxida-
tion challenge. An in vivo study (Somoza et al., 2005) showed that
bread crust, caraffa malt, or pronyl bovine serum albumin (which
contained pyrrolinone reductonyl-lysine, a food melanoidin) de-
creased oxidative stress levels and thiobarbituric acid reactive sub-
stances levels, and increased tocopherol in the plasma of rats.
Although the specic components responsible for the protective
antioxidant behaviour observed for melanoidins are not yet
known, melanoidins are thought to be important in preventing oxi-
dative damage and diseases related to free radicals.
The antioxidant properties of melanoidins have been partly as-
cribed to the metal chelating capacity of these compounds
(Verzelloni et al., 2010b). The anionic nature of melanoidins en-
ables them to chelate transition metals (Daglia et al., 2008; Gu
et al., 2010; Morales, Fernandez-Fraguas, & Jimenez-Perez, 2005;
Ruan-Henares & de la Cueva, 2009; Tagliazucchi et al., 2010b;
Verzelloni et al., 2010b). The nitrogen atoms in melanoidins were
proposed to be responsible for the chelation of copper ion (Cosovic
et al., 2010). Hashiba (1986) reported that the ketone and/or hy-
droxyl groups of the pyranone or pyridone components may act
as chelation donor groups in melanoidins. Though no relationship
was observed between browning and the iron binding ability of
melanoidins in model systems (Morales et al., 2005), Wen et al.
(2005) showed that the zinc-chelating activity of brewed coffee de-
creased as the intensity of the roasting increased, suggesting that
the chromophoric groups of melanoidins may not be the main
co-ordination sites for metal complexation, but the metal chelating
activity of melanoidins may be due to other, as yet unknown,
structures of them.
The other main mechanism for the antioxidant activity of mel-
anoidins is their radical-scavenging activity. It has been reported
that melanoidin fractions in food and model systems can scavenge
a variety of reactive oxygen species. Some studies have shown a
positive correlation between the radical-scavenging activity and
colour or heated degree of melanoidins, while others have not.
A strong linear relationship between the peroxyl (Gmez-Ruiz
et al., 2008; Morales & Jimnez-Prez, 2004), 2,2
0
-azobis-2-
methyl-propanimidamide, dihydrochloride (AAPH) (Brudzynski &
Miotto, 2011b) and 2,2-azobis (3-ethylbenzothiazoline-6-sul-
phonic acid) (ABTS) (Ruan-Henares & Morales, 2007c) radical
scavenging activity and the colour intensity of the melanoidin frac-
tions (derived from either model systems or commercial samples)
has been established. The ABTS radical-scavenging activity of the
non-phenolic fraction of brewed coffee and bread extract was ob-
served to increase with prolonged roasting or baking (Michalska
et al., 2008; Sacchetti, Di Mattia, Pittia, & Mastrocola, 2009), and
the hydroxyl radical scavenging activity of vinegar melanoidins in-
creased during the decoction, storing and aging of the vinegar (Xu
et al., 2007). These results indicate that the brown compounds
formed during the heating of foods or model systems, which are
primarily melanoidins, are the compounds most likely responsible
for the radical-scavenging activity (Gmez-Ruiz et al., 2008).
H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584 579
Other studies have reported very different results. For example,
no positive linear correlation was found between the hydroxyl rad-
ical-scavenging activity and browning (Morales, 2005; Morales &
Jimnez-Prez, 2004). In two studies, the antiradical activity mea-
sured from ABTS and N,N-dimethyl phenylenediamine dihydro-
chloride (DMPD) assays of coffee melanoidin fractions was in fact
observed to decrease as the intensity of the roasting increased
(Borrelli et al., 2002; Rivero-Perez et al., 2002). The conicting data
on the radical-scavenging activity of melanoidins may be a result
of the complicated components of melanoidin fractions, as well
as the assays used to measure these activities.
Although the antiradical activity of MRPs from both model sys-
tems and browned foods has been measured several times, most
investigations reported only the activity of the bulk browned water
extracts or the HMW fraction of the extracts, rather than correlat-
ing these effects with chemically distinct melanoidins (Lindenmeier
et al., 2002). Although the HMW fractions isolated from model sys-
tems (Jing & Kitts, 2000) and food (Daglia et al., 2004) exhibited
signicant antioxidative activity when compared with the LMW
fractions, some studies (Delgado-Andrade, Ruan-Henares, &
Morales, 2005; Ruan-Henares & Morales, 2007b, 2007c; Taglia-
zucchi et al., 2010a; Vignoli, Bassoli, & Benassi, 2011) support the
hypothesis that the bounded melanoidin compounds (BMCs), for
instance polyphenols and LMW MRPs, which are linked non-cova-
lently to the melanoidin skeleton, contribute to the antioxidant
activity, too. The BMC fraction of brewed coffee (Delgado-Andrade
et al., 2005) and glucoseamino acid MRPs (Ruan-Henares & Mor-
ales, 2007b, 2007c) demonstrated higher antiradical activity than
pure melanoidins. Tagliazucchi et al. (2010a), however, reported
that the antioxidant properties of the BMC fraction of HMW mela-
noidins found in balsamic vinegar was always lower than that of
the pure melanoidins to which they are linked.
The different ndings in these reports may be due to the variety
of structural properties and different concentrations of BMCs in
different systems. Polyphenols, which are BMCs found in commer-
cial foods such as coffee, are partially lost during heating (Delgado-
Andrade et al., 2005). Thus, melanoidins are probably the predominant
contributors to the antioxidant activity in severely heat-treated
samples (Daglia et al., 2004; Jing & Kitts, 2000). One possible expla-
nation for the conicting reports mentioned above is that the BMCs
linked to melanoidins or other components in the melanoidin frac-
tion have antioxidant activity to some degree, but their structures
may be changed during heating, which induces variability in the
relationship between radical-scavenging activity and melanoidin
content in different systems. The assays used to measure the rad-
ical-scavenging activity could also account for the discrepancies
in the above studies. The variety of different chelation sites poten-
tially present in melanoidins, as well as non-chromophore struc-
tures such as phenolic residues, may also introduce some variety
among the measured scavenging activity of different radicals by
melanoidins.
3.3. Antimicrobial activity
Ruan-Henaress group studied the antimicrobial activity of
melanoidins against pathogenic microorganisms and found mela-
noidin fractions from both aqueous model systems (100 C, 24 h,
pH 7) and foods (coffee, beer, sweet wine) exhibited higher antimi-
crobial activity towards gram-positive microorganisms (e.g. Staph-
ylococcus aureus) compared to gram-negative microorganisms (e.g.
Escherichia coli) (Ruan-Henares & de la Cueva, 2009; Ruan-
Henares & Morales, 2008b). The HMW fractions of coffee and
biscuit melanoidins with darker colour displayed higher antimicro-
bial activity against E. coli than LMW fractions (Ruan-Henares &
Morales, 2008a), and melanoidins from more strongly heated sam-
ples such as roasted coffee, displayed a higher activity (Ruan-
Henares & Morales, 2006). It was also determined that the BMC
fraction contributed to the antimicrobial activity of melanoidins.
Strong antimicrobial activity was observed in both pure coffee
melanoidins, as well as the BMCs linked to these melanoidins
(Ruan-Henares & Morales, 2006), though pure melanoidins iso-
lated from a glucose-amino acid model system were observed to
exert a lower antimicrobial activity than that of the BMC fraction
(Ruan-Henares & Morales, 2007b, 2007c). Therefore, pure mela-
noidins do have antimicrobial activity and the chromophore may
be responsible for it.
Antimicrobial activity was determined to be bacteriostatic at
low concentrations of melanoidins, but bactericidal at higher con-
centrations (Ruan-Henares & de la Cueva, 2009; Ruan-Henares &
Morales, 2008a). The dependence of the microbicidal mechanism
on the melanoidin concentration could be related to their
metal-chelating properties. Three different mechanisms for the
antimicrobial activity of melanoidins have been proposed. At low
concentrations, melanoidins exert a bacteriostatic activity mainly
mediated by the chelation of iron from the culture medium. In bac-
terial strains able to produce siderophores for iron acquisition, the
chelation of the siderophoreFe
3+
complex by melanoidins has
been observed, which could decrease the virulence of pathogenic
bacteria. Finally, melanoidins at high concentrations can disrupt
both the outer and inner membranes by chelating Mg
2+
ions from
the outer membrane, leading to a destabilization of the inner
membrane (Ruan-Henares & de la Cueva, 2009; Ruan-Henares
& Morales, 2008a). The higher antimicrobial activity exhibited to-
wards gram-positive microorganisms was ascribed to the absence
of a protective outer membrane, which makes them more suscep-
tible to antimicrobial substances (Ruan-Henares & de la Cueva,
2009).
3.4. Effect of melanoidins on xenobiotic enzymes activity
Most chemopreventive, nonendogenously formed agents are
processed by enzyme systems that modulate phase I and phase II
enzymes, which are called xenobiotic enzymes. A decrease in
phase I enzyme activity associated with an increase in phase II
enzyme activity is effective in contributing to chemopreventive
potential (Wilkinson & Clapper, 1997).
Bread crust and pronyl-bovine serum albumin containing large
amounts of pronyl-L-lysine were shown to enhance phase II en-
zymes, e.g., UDP-glucuronyl-tansferase (UDP-GT) and glutathione
S-transferase (GST), in rats (Faist, Hofmann et al., 2002; Somoza
et al., 2005). Melanoidin fractions from bread crust induced a sig-
nicant increase in GST activity (Faist, Hofmann et al., 2002;
Lindenmeier et al., 2002), but lead to a decrease in the activity of
phase I NADPH cytochrome c reductase (CCR) in Caco-2 cells
(Lindenmeier et al., 2002), indicative of their protective effects.
Melanoidin fractions from biscuits, however, were observed to de-
crease GST activity (Borrelli et al., 2003) and roasted malt melanoi-
dins were observed to elevate CCR activity (Faist, Lindenmeier
et al., 2002). Similarly, the effect of model melanoidin fractions
on xenobiotic enzymes is not consistent too (Van Nguyen, 2006).
Thus, the effects of melanoidin fractions on phase I and II enzymes
are dependant on their source.
The relationship between the melanoidin phase I and II en-
zyme-modulating activity and melanoidin colour intensity is not
known, so the xenobiotic enzyme-modulating activity described
above can not be linked to specic choromphoric or structural
properties of melanoidins. It is known that MRPs contain high con-
centrations of Ne-carboxymethyllysine (CML), a compound known
to induce CCR and inhibit GST (Faist, Hofmann et al., 2002; Faist,
Lindenmeier et al., 2002), which may partially explain the effects
of melanoidin fractions on CCR and GST (Faist, Lindenmeier et al.,
2002). However, this xenobiotic enzyme-modulating activity
580 H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584
cannot be entirely explained by the presence of CML, as it seems
various reaction products formed during the heat-treatment of
food and model systems, such as MRPs and native polyphenols,
also contribute to this effect, which could help to explain the dis-
crepancies in the above investigations.
3.5. Prebiotic activity
Bidobacteria and Lactobacilli in the gastrointestinal tract have
been shown to confer benecial health effects. Food components
that promote the growth of benecial versus pathogenic bacteria
are known as prebiotics (Borrelli & Fogliano, 2005). Data from
in vitro and in vivo studies suggest that most melanoidins escape
digestion in the upper gastrointestinal tract and are mainly recov-
ered in the faeces (Borrelli & Fogliano, 2005; Faist & Erbersdobler,
2001). Similar to dietary bre, Borrelli and Fogliano (2005) demon-
strated that melanoidin fractions from bread crust appeared to
selectively enhance the growth of Bidobacteria in the gut, and
potentially have a prebiotic activity similar to that of dietary bres.
However, an in vitro study (Ames, Wynne, Hofmann, Plos, &
Gibson, 1999) investigating the effects of the melanoidin fractions
from glucose-lysine MRPs (reuxing at pH 5 for 2 h) on colonic
bacteria demonstrated that melanoidins could cause a non-specic
increase in all anaerobic bacteria of the human large-bowel,
including Bacteroides, Clostridia, Bidobacteria and Lactobacilli.
Thus, it may be difcult to ascribe prebiotic effects to melanoidins.
3.6. Antihypertensive activity
The potential antihypertensive activity of pure melanoidins iso-
lated from coffee, beer, sweet wine (Ruan-Henares & Morales,
2007a), and puried model melanoidins (Ruan-Henares &
Morales, 2007c) has been evaluated by investigating the in vitro
ACE-inhibitory activity. The in vitro ACE-inhibitory activity of food
melanoidins was shown to be similar to that reported for well-
known antihypertensive peptides (Ruan-Henares & Morales,
2007a). The ACE-inhibitory activity of coffee melanoidins was sig-
nicantly higher with coffee that had been heated for prolonged
periods of time, which indicated that the melanoidins are likely
responsible for this activity, though the ACE-inhibitory activity
has also been linked to the presence of BMCs as well (Ruan-
Henares & Morales, 2007a). Ruan-Henares and Morales (2007c)
also demonstrated that the in vitro ACE-inhibitory activity of pure
melanoidins extracted from aqueous glucoseamino acids systems
(100 C, 24 h, pH 7) was related to melanoidin structure, and the
chromophores of melanoidins were not involved.
3.7. Other biological effects
Hiramoto et al. (2004) investigated the effects of food melanoi-
din fractions on the prevention of Helicobacter pylori (H. pylori)
infection. Urease proteins located on the extracellular surface of
H. pylori are adhesions that target gastric surfaces. These proteins
play an important role in infection and colonisation of the host.
It was found that the HMW MRPs of alkali lactose/glucoseprotein
systems (85 C, 4 h) and brown fractions isolated from the crust of
an English mufn inhibited the binding of urease to gastric mucin
by competing with gastric mucin for urease-binding sites. More-
over, in an animal model and in human subjects infected with
H. pylori, these melanoidin fractions were shown to suppress
H. pylori colonisation (Hiramoto et al., 2004). This indicated that
food melanoidins may have antibiotic activities and might prevent
H. pylori infection.
Synthesised and puried chromophoric MRPs, identied in
thermal food processing products, were reported to inhibit the
growth of human carcinoma cells in vitro by interfering with the
proliferation of gastric carcinoma cells, causing cell cycle arrest
and apoptosis (Marko et al., 2003). These results suggest that some
melanoidins may inhibit the growth of human tumour cells.
HMW MRP fractions from coffee were demonstrated to have
antiadhesive properties and inhibited the ability of Streptococcus
mutans (S. mutans) to adsorb to saliva-coated hydroxyapatite
beads. This is considered an important virulence property of S. mu-
tans and a cause of tooth decay. The antiadhesive activity may be
largely due to small molecules that occur naturally in coffee (e.g.
chlorogenic acid and trigonelline), however, so it was proposed
that coffee could minimise S. mutans colonisation and might be
effective in preventing S. mutans-induced tooth decay (Daglia
et al., 2002).
Takahama and Hirota (2008) reported that in addition to chlor-
ogenic acid and its isomers, coffee melanoidins also reacted with
salivary nitrite and SCN

at pH 2, mimicking the mixture of coffee,


saliva, and gastric juices in the gastric lumen. This produces nitric
oxide (NO), which may inhibit microbial growth, and regulate
mucosal ow, mucosal formation and gastric mobility.
HMW brown MRPs obtained from dry glucose-arginine model
systems (125 C, 3 h) were tested for their ability to inuence the
contractility of gastric smooth muscles. Brown MRPs at low con-
centrations provoked concentration-dependent contraction,
whereas brown MRPs at high concentrations induced muscle
relaxation (Argirova, Stefanova, Krustev, & Turiiski, 2010).
3.8. The in vivo effects of melanoidins and the effects of melanoidins on
foods
At present, not much is known about the bioavailability of mel-
anoidins. The amount of protein-bound LMW melanoidins ab-
sorbed and excreted in the urine of rats was found at levels
ranging from 1% to 5% (Faist & Erbersdobler, 2001), which sug-
gested that melanoidins escape digestion in the upper gastrointes-
tinal tract and are mainly recovered in the faeces. Finot and
Magnenat (1981) investigated the metabolic transit of LMW and
HMW fractions of casein/glycine
14
C-glucose mixture MRPs in
rats. They observed that 61% of the LMW fractions (<10 kDa) of
the casein
14
C-glucose mixture were excreted in the faeces. For
the HMW fractions (>10 kDa), however, 87% of the ingested radio-
labeled casein
14
C-glucose and 93% of the ingested radiolabeled
glycine
14
C-glucose were excreted in the faeces, respectively.
These results indicated that HMW melanoidins are absorbed to a
much lesser extent than LMW compounds.
On the other hand, much research has been done investigating
the digestibility and bioavailability of AGEs (Somoza, 2005), which
are advanced MRPs formed in reaction pathways that are analo-
gous to the reaction pathways that form melanoidins. It is known
that dietary AGEs, particularly LMW AGEs, can be absorbed by hu-
mans and rats, with elevated levels observed in the plasma and or-
gans (Chuyen, 2006).
It can be concluded that about 30% of LMW melanoidins and/or
their intestinal degradation products are absorbed and are likely to
be taken up in the blood stream (Chuyen, 2006). The low dose of
absorbed melanoidins is not likely to be genotoxic or cytotoxic to
humans. Owning to the much lower dose of digestible melanoidins
compared with in vitro studies, however, they may exert limited
in vivo benecial effects, such as antioxidative activities, modula-
tion of phase I and phase II enzymes and antihypertensive activity.
The more important in vivo benecial effects can probably be as-
cribed to the indigestible melanoidins. As they pass through the
intestine, they will behave as antioxidants, display antimicrobial
activity against pathogenic microorganisms and prebiotic activity,
provided they survive gastric conditions for a long enough period
of time.
H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584 581
Melanoidins are produced during the processing and storage of
foods. In foods, the antioxidant properties of melanoidins can inhi-
bit the oxidation of unsaturated lipids and functional food ingredi-
ents, such as vitamins, polyphenols and avonoids. Moreover, the
antimicrobial activity can inhibit the growth of microorganisms
and prevent the spoilage and deterioration of foods. Thus, melanoi-
dins may preserve the quality and safety of foods.
4. Conclusions
As mentioned above, the Maillard reaction occurs during the
preparation of food and leads to the production of melanoidins.
Large quantities of melanoidins are consumed by eating cooked,
fried and roasted foods. Food melanoidins are generally anionic
chromophoric compounds and can be either LMW pigments,
HMW compounds generated by the polymerisation of LMW chro-
mophores, or non-coloured high molecular weight biopolymers
cross-linked to LMW chromophores. Melanoidins, particularly
LMW melanoidins, are digestible to some extent. Absorbed mela-
noidins may have benecial in vivo effects, such as antioxidant
activity and phase I and II enzyme-modulating activity. However,
the indigestible melanoidins are more likely to be benecial to hu-
man health, benecial to human health, displaying in vivo antiox-
idant, antimicrobial and prebiotic activity in the intestine.
Melanoidins in food may also serve as preservatives and protect
the quality of food during storage. It is usually not known which
specic effect is assigned to a particular melanoidin, due to the
large amount of MRPs formed during the Maillard reaction and
the difculties in purifying and identifying these compounds. To
better understand the benets of melanoidins in healthy foods,
these benecial compounds need to be chemically characterised
and categorised according to their structure-specic health effects.
Such information will enable the optimisation of food processing
technologies to enhance the production of benecial melanoidins.
In addition, the in vivo biological effects of these compounds
should be more extensively studied in order to acquire more infor-
mation about their healthful effects.
Acknowledgements
Authors are grateful for the nancial support of the project JUS-
RP11121 of Jiangnan University (JUSRP11121).
References
Adams, A., Abbaspour Tehrani, K., Kersiene, M., Venskutonis, R., & De Kimpe, N.
(2003). Characterization of model melanoidins by the thermal degradation
prole. Journal of Agricultural and Food Chemistry, 51, 43384343.
Adams, A., Borrelli, R. C., Fogliano, V., & De Kimpe, N. (2005). Thermal degradation
studies of food melanoidins. Journal of Agricultural and Food Chemistry, 53,
41364142.
Ames, J. M., Bailey, R. G., & Mann, J. (1999). Analysis of furanone, pyranone, and new
heterocyclic coloured compounds from sugar-glycine model Maillard systems.
Journal of Agricultural and Food Chemistry, 47, 438443.
Ames, J. M., Wynne, A., Hofmann, A., Plos, S., & Gibson, G. R. (1999). The effect of a
model melanoidin mixture on faecal bacterial populations in vitro. British
Journal of Nutrition, 82, 489495.
Andriot, I., Le Quere, J. L., & Guichard, E. (2004). Interactions between coffee
melanoidins and avour compounds: Impact of freeze-drying (method and
time) and roasting degree of coffee on melanoidins retention capacity. Food
Chemistry, 85, 289294.
Argirova, M., Stefanova, I., Krustev, A., & Turiiski, V. (2010). Testing biological
activity of model Maillard reaction products: Studies on gastric smooth muscle
tissues. Amino Acids, 38, 797803.
Bartling, B., Rehbein, G., Somoza, V., Silber, R. E., & Simm, A. (2005). Maillard
reaction product-rich food impair cell proliferation and induce cell death
in vitro. Signal Transduction, 5, 303313.
Bekedam, E. K., De Laat, M. P., Schols, H. A., Van Boekel, M. A., & Smit, G. (2007).
Arabinogalactan proteins are incorporated in negatively charged coffee brew
melanoidins. Journal of Agricultural and Food Chemistry, 55, 761768.
Bekedam, E. K., Loots, M. J., Schols, H. A., Van Boekel, M. A., & Smit, G. (2008).
Roasting effects on formation mechanisms of coffee brew melanoidins. Journal
of Agricultural and Food Chemistry, 56, 71387145.
Bekedam, E. K., Roos, E., Schols, H. A., Van Boekel, M. A. J. S., & Smit, G. (2008). Low
molecular weight melanoidins in coffee brew. Journal of Agricultural and Food
Chemistry, 56, 40604067.
Bekedam, E. K., Schols, H. A., van Boekel, M. A. J. S., & Smit, G. (2006). High molecular
weight melanoidins from coffee brew. Journal of Agricultural and Food Chemistry,
54, 76587666.
Bekedam, E. K., Schols, H. A., van Boekel, M. A. J. S., & Smit, G. (2008). Incorporation
of chlorogenic acids in coffee brew melanoidins. Journal of Agricultural and Food
Chemistry, 56, 20552063.
Borrelli, R. C., & Fogliano, V. (2005). Bread crust melanoidins as potential prebiotic
ingredients. Molecular Nutrition and Food Research, 49, 673678.
Borrelli, R. C., Mennella, C., Barba, F., Russo, M., Russo, G. L., Krome, K., et al. (2003).
Characterization of coloured compounds obtained by enzymatic extraction of
bakery products. Food and Chemical Toxicology, 41, 13671374.
Borrelli, R. C., Visconti, A., Mennella, C., Anese, M., & Fogliano, V. (2002). Chemical
characterization and antioxidant properties of coffee melanoidins. Journal of
Agricultural and Food Chemistry, 50, 65276533.
Brands, C. M. J., Wedzicha, B. L., & van Boekel, M. A. J. S. (2002). Quantication of
melanoidin concentration in sugarcasein systems. Journal of Agricultural and
Food Chemistry, 50, 11781183.
Brudzynski, K., & Miotto, D. (2011a). The recognition of high molecular weight
melanoidins as the main components responsible for radical-scavenging
capacity of unheated and heat-treated canadian honeys. Food Chemistry, 125,
570575.
Brudzynski, K., & Miotto, D. (2011b). The relationship between the content of
Maillard reaction-like products and bioactivity of Canadian honeys. Food
Chemistry, 124, 869874.
Cammerer, B., Jalyschko, W., & Kroh, L. W. (2002a). Intact carbohydrate structures as
part of the melanoidin skeleton. Journal of Agricultural and Food Chemistry, 50,
20832087.
Cammerer, B., Jalyschkov, V., & Kroh, L. W. (2002b). Carbohydrate structures as part
of the melanoidin skeleton. International Congress Series, 1245, 269273.
Chuyen, N. V. (2006). Toxicity of the AGEs generated from the Maillard reaction: On
the relationship of food-AGEs and biological-AGEs. Molecular Nutrition and Food
Research, 50, 11401149.
Cosovic, B., Vojvodic, V., Boskovic, N., Plavsic, M., & Lee, C. (2010). Characterization
of natural and synthetic humic substances (melanoidins) by chemical
composition and adsorption measurements. Organic Geochemistry, 41, 200205.
Daglia, M., Papetti, A., Aceti, C., Sordelli, B., Gregotti, C., & Gazzani, G. (2008).
Isolation of high molecular weight components and contribution to the
protective activity of coffee against lipid peroxidation in a rat liver
microsome system. Journal of Agricultural and Food Chemistry, 56, 1165311660.
Daglia, M., Racchi, M., Papetti, A., Lanni, C., Govoni, S., & Gazzani, G. (2004). In vitro
and ex vivo antihydroxyl radical activity of green and roasted coffee. Journal of
Agricultural and Food Chemistry, 52, 17001704.
Daglia, M., Tarsi, R., Papetti, A., Grisoli, P., Dacarro, C., Pruzzo, C., et al. (2002).
Antiadhesive effect of green and roasted coffee on streptococcus mutans
adhesive properties on saliva-coated hydroxyapatite beads. Journal of
Agricultural and Food Chemistry, 50, 12251229.
Delgado-Andrade, C., Ruan-Henares, J. A., & Morales, F. J. (2005). Assessing the
antioxidant activity of melanoidins from coffeebrews by different antioxidant
methods. Journal of Agricultural and Food Chemistry, 53, 78327836.
Faist, V., & Erbersdobler, H. F. (2001). Metabolic tansit and in vivo effects of
melanoidins and precursor compounds deriving from the Maillard reaction.
Annals of Nutrition and Metabolism, 45, 112.
Faist, V., Hofmann, T., Zill, H., Baynes, J. W., Thorpe, S. R., Sebekova, K., et al. (2002).
Effects of dietary Ne-carboxymethyllysine on expression of the
biotransformation enzyme, glutathione-S-transferase, in the rat. International
Congress Series, 1245, 313320.
Faist, V., Lindenmeier, M., Geisler, C., Erbersdobler, H. F., & Hofmann, T. (2002).
Inuence of molecular weight fractions isolated from roasted malt on the
enzyme activities of NADPH-cytochrome C-reductase and glutathione-S-
transferase in Caco-2 cells. Journal of Agricultural and Food Chemistry, 50,
602606.
Finot, P. A., & Magnenat, E. (1981). Metabolic transit of early and advanced Maillard
products. Progress in Food and Nutrition Science, 5, 193207.
Frank, O., & Hofmann, T. (2000). Characterization of key chromophores formed by
nonenzymatic browning of hexoses and L-alanine by using the colour activity
concept. Journal of Agricultural and Food Chemistry, 48, 63036311.
Glosl, S., Wagner, K. H., Draxler, A., Kaniak, M., Lichtenecker, S., Sonnleitner, A., et al.
(2004). Genotoxicity and mutagenicity of melanoidins isolated from a roasted
glucose-glycine model in human lymphocyte cultures, intestinal Caco-2 cells
and in the salmonella typhimurium strains TA98 and TA102 applying the Ames
test. Food and Chemical Toxicology, 42, 14871495.
Gniechwitz, D., Reichardt, N., Ralph, J., Blaut, M., Steinhart, H., & Bunzel, M. (2008).
Isolation and characterisation of a coffee melanoidin fraction. Journal of the
Science of Food and Agriculture, 88, 21532160.
Gmez-Ruiz, J., Ames, J., & Leake, D. (2008). Antioxidant activity and protective
effects of green and dark coffee components against human low density
lipoprotein oxidation. European Food Research and Technology, 227, 10171024.
Gu, F. L., Kim, J. M., Abbas, S., Zhang, X. M., Xia, S. Q., & Chen, Z. X. (2010). Structure
and antioxidant activity of high molecular weight Maillard reaction products
from caseinglucose. Food Chemistry, 120, 505511.
582 H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584
Hashiba, H. (1986). Oxidative browning of Amadori compounds. Colour formation
by iron with Maillard reaction products. In M. Fujimaki, H. Kato, & M. Namiki
(Eds.), Amino-carbonyl reactions in food and biological systems (pp. 155164).
Amsterdan: Elsevier.
Hayase, F., Usui, T., Ono, Y., Shirahashi, Y., Machida, T., Ito, T., et al. (2008). Formation
mechanisms of melanoidins and uorescent pyridinium compounds as
advanced glycation end products. Annals of the New York Academy of Sciences,
1126, 5358.
Hayase, F., Usui, T., & Watanabe, H. (2006). Chemistry and some biological effects of
model melanoidins and pigments as Maillard intermediates. Molecular Nutrition
& Food Research, 50, 11711179.
Hiramoto, S., Itoh, K., Shizuuchi, S., Kawachi, Y., Morishita, Y., Nagase, M., et al.
(2004). Melanoidin, a food protein-derived advanced Maillard reaction product,
suppresses helicobacter pylori in vitro and in vivo. Helicobacter, 9, 429435.
Hofmann, T. (1998a). 4-alkylidene-2-imino-5-[4-alkylidene-5-oxo-1, 3-imidazol-2-
inyl] aza-methylidene-1, 3- imidazolidinea novel coloured substructure in
melanoidins formed by Maillard reactions of bound arginine with glyoxal and
furan-2-carboxaldehyde. Journal of Agricultural and Food Chemistry, 46,
38963901.
Hofmann, T. (1998b). Identication of novel coloured compounds containing
pyrrole and pyrrolinone structures formed by Maillard reactions of pentoses
and primary amino acids. Journal of Agricultural and Food Chemistry, 46,
39023911.
Hofmann, T. (1998c). Studies on melanoidin-type colourants generated from the
Maillard reaction of protein-bound lysine and furan-2-carboxaldehyde-
chemical characterisation of a red coloured domaine. Zeitschrift fr
Lebensmitteluntersuchung und -Forschung A, 206, 251258.
Hofmann, T. (1998d). Studies on the inuence of the solvent on the contribution of
single Maillard reaction products to the total colour of browned pentose/
alanine solutions-A quantitative correlation using the colour activity concept.
Journal of Agricultural and Food Chemistry, 46, 39123917.
Hofmann, T. (1998e). Studies on the relationship between molecular weight and the
colour potency of fractions obtained by thermal treatment of glucose/amino
acid and glucose/protein solutions by using ultracentrifugation and colour
dilution techniques. Journal of Agricultural and Food Chemistry, 46, 38913895.
Ibarz, A., Garvin, A., Garza, S., & Pagan, J. (2009). Toxic effect of melanoidins from
glucose-asparagine on trypsin activity. Food and Chemical Toxicology, 47,
20712075.
Jing, H., & Kitts, D. D. (2000). Comparison of the antioxidative and cytotoxic
properties of glucose-lysine and fructose-lysine Maillard reaction products.
Food Research International, 33, 509516.
Jing, H., & Kitts, D. D. (2002). Chemical and biochemical properties of caseinsugar
Maillard reaction products. Food and Chemical Toxicology, 40, 10071015.
Kim, J.-S., & Lee, Y.-S. (2008). Effect of reaction pH on enolization and racemization
reactions of glucose and fructose on heating with amino acid enantiomers and
formation of melanoidins as result of the Maillard reaction. Food Chemistry, 108,
582592.
Kim, J.-S., & Lee, Y.-S. (2009). Enolization and racemization reactions of glucose and
fructose on heating with amino-acid enantiomers and the formation of
melanoidins as a result of the Maillard reaction. Amino Acids, 36, 465474.
Kroh, L. W., Fiedler, T., & Wagner, J. (2008). Dicarbonyl compoundskey
intermediates for the formation of carbohydrate-based melanoidins. Annals of
the New York Academy of Sciences, 1126, 210215.
Kwak, E. J., Lee, Y. S., Murata, M., & Homma, S. (2005). Effect of pH control on the
intermediates and melanoidins of nonenzymatic browning reaction.
Lebensmittel-Wissenschaft und-Technologie, 38, 16.
Lindenmeier, M., Faist, V., & Hofmann, T. (2002). Structural and functional
characterization of pronyl-lysine, a novel protein modication in bread crust
melanoidins showing in vitro antioxidative and phase I/II enzyme modulating
activity. Journal of Agricultural and Food Chemistry, 50, 69977006.
Lindenmeier, M., & Hofmann, T. (2003). Inuence of baking conditions and
precursor supplementation on the amounts of the antioxidant pronyl-Llysine
in bakery products. Journal of Agricultural and Food Chemistry, 52, 350354.
Marko, D., Habermeyer, M., Kemeny, M., Weyand, U., Niederberger, E., Frank, O.,
et al. (2003). Maillard reaction products modulating the growth of human
tumor cells in vitro. Chemical Research in Toxicology, 16, 4855.
Martin, M. A., Ramos, S., Mateos, R., Ruan-Henares, J. A., Morales, F. J., Bravo, L.,
et al. (2009). Biscuit melanoidins of different molecular masses protect human
HEPG2 cells against oxidative stress. Journal of Agricultural and Food Chemistry,
57, 72507258.
Martins, S. I. F. S., & van Boekel, M. A. J. S. (2003). Melanoidins extinction coefcient
in the glucose/glycine Maillard reaction. Food Chemistry, 83, 135142.
Michalska, A., Amigo-Benavent, M., Zielinski, H., & del Castillo, M. D. (2008). Effect of
bread making on formation of Maillard reaction products contributing to the
overall antioxidant activity of rye bread. Journal of Cereal Science, 48, 123132.
Morales, F. J. (2002). Application of capillary zone electrophoresis to the study of
food and food-model melanoidins. Food Chemistry, 76, 363369.
Morales, F. J. (2005). Assessing the non-specic hydroxyl radical scavenging
properties of melanoidins in a Fenton-type reaction system. Analytica Chimica
Acta, 534, 171176.
Morales, F. J., Fernandez-Fraguas, C., & Jimenez-Perez, S. (2005). Iron-binding ability
of melanoidins from food and model systems. Food Chemistry, 90, 821827.
Morales, F., & Jimnez-Prez, S. (2004). Peroxyl radical scavenging activity of
melanoidins in aqueous systems. European Food Research and Technology, 218,
515520.
Mundt, S., & Wedzicha, B. L. (2003). A kinetic model for the glucosefructose
glycine browning reaction. Journal of Agricultural and Food Chemistry, 51,
36513655.
Ortega-Heras, M., & Gonzalez-Sanjose, M. L. (2009). Binding capacity of brown
pigments present in special spanish sweet wines. LWT - Food Science and
Technology, 42, 17291737.
Papetti, A., Daglia, M., Aceti, C., Quaglia, M., Gregotti, C., & Gazzani, G. (2006).
Isolation of an in vitro and ex vivo antiradical melanoidin from roasted barley.
Journal of Agricultural and Food Chemistry, 54, 12091216.
Ramonaityte, D. T., Kersiene, M., Adams, A., Tehrani, K. A., & Kimpe, N. D. (2009). The
interaction of metal ions with Maillard reaction products in a lactose-glycine
model system. Food Research International, 42, 331336.
Rivero-Perez, M. D., Perez-Magarino, S., & Gonzalez-San Jos, M. L. (2002). Role of
melanoidins in sweet wines. Analytica Chimica Acta, 458, 169175.
Ruan-Henares, J. A., & de la Cueva, S. P. (2009). Antimicrobial activity of coffee
melanoidinsa study of their metal-chelating properties. Journal of Agricultural
and Food Chemistry, 57, 432438.
Ruan-Henares, J. A., & Morales, F. J. (2006). A new application of a commercial
microtiter plate-based assay for assessing the antimicrobial activity of Maillard
reaction products. Food Research International, 39, 3339.
Ruan-Henares, J. A., & Morales, F. J. (2007a). Angiotensin-I converting enzyme
inhibitory activity of coffee melanoidins. Journal of Agricultural and Food
Chemistry, 55, 14801485.
Ruan-Henares, J. A., & Morales, F. J. (2007b). Effect of in vitro enzymatic digestion
on antioxidant activity of coffee melanoidins and fractions. Journal of
Agricultural and Food Chemistry, 55, 1001610021.
Ruan-Henares, J. A., & Morales, F. J. (2007c). Functional properties of melanoidins:
In vitro antioxidant, antimicrobial and antihypertensive activities. Food Research
International, 40, 9951002.
Ruan-Henares, J. A., & Morales, F. J. (2008a). Antimicrobial activity of melanoidins
against Escherichia Coli is mediated by a membrane-damage mechanism. Journal
of Agricultural and Food Chemistry, 56, 23572362.
Ruan-Henares, J. A., & Morales, F. J. (2008b). Microtiter plate-based assay for
screening antimicrobial activity of melanoidins against E. coli and S. aureus. Food
Chemistry, 111, 10691074.
Sacchetti, G., Di Mattia, C., Pittia, P., & Mastrocola, D. (2009). Effect of roasting
degree, equivalent thermal effect and coffee type on the radical scavenging
activity of coffee brews and their phenolic fraction. Journal of Food Engineering,
90, 7480.
Sasaki, S., Shirahashi, Y., Nishiyama, K., Watanabe, H., & Hayase, F. (2006).
Identication of a novel blue pigment as a melanoidin intermediate in the
Dxylose-glycine reaction system. Bioscience, Biotechnology, and Biochemistry, 70,
25292531.
Shirahashi, Y., Watanabe, H., & Hayase, F. (2009). Identication of red pigments
formed in a Dxylose-glycine reaction system. Bioscience, Biotechnology, and
Biochemistry, 73, 22872292.
Somoza, V. (2005). Five years of research on health risks and benets of Maillard
reaction products: An update. Molecular Nutrition and Food Research, 49,
663672.
Somoza, V., Wenzel, E., Lindenmeier, M., Grothe, D., Erbersdobler, H. F., & Hofmann,
T. (2005). Inuence of feeding malt, bread crust, and a pronylated protein on the
activity of chemopreventive enzymes and antioxidative defense parameters
in vivo. Journal of Agricultural and Food Chemistry, 53, 81768182.
Summa, C., McCourt, J., Cmmerer, B., Fiala, A., Probst, M., Kun, S., et al. (2008).
Radical scavenging activity, anti-bacterial and mutagenic effects of cocoa bean
Maillard reaction products with degree of roasting. Molecular Nutrition & Food
Research, 52, 342351.
Tagliazucchi, D., Verzelloni, E., & Conte, A. (2010a). Contribution of melanoidins to
the antioxidant activity of traditional balsamic vingar during aging. Journal of
Food Biochemistry, 34, 10611078.
Tagliazucchi, D., Verzelloni, E., & Conte, A. (2010b). Effect of dietary melanoidins on
lipid peroxidation during simulated gastric digestion: Their possible role in the
prevention of oxidative damage. Journal of Agricultural and Food Chemistry, 58,
25132519.
Takahama, U., & Hirota, S. (2008). Reduction of nitrous acid to nitric oxide by coffee
melanoidins and enhancement of the reduction by thiocyanate: Possibility of its
occurrence in the stomach. Journal of Agricultural and Food Chemistry, 56,
47364744.
Tressl, R., Wondrak, G. T., Kruger, R. P., & Rewicki, D. (1998). New melanoidin-like
Maillard polymers from 2-deoxypentoses. Journal of Agricultural and Food
Chemistry, 46, 104110.
Valls-Bells, V., Torres, M., Muiz, P., Boix, L., Gonzlez-Sanjose, M., & Codoer-
Franch, P. (2004). The protective effects of melanoidins in adriamycin-induced
oxidative stress in isolated rat hepatocytes. Journal of the Science of Food and
Agriculture, 84, 17011707.
Van Nguyen, C. (2006). Toxicity of the ages generated from the maillard reaction:
On the relationship of food-ages and biological-ages. Molecular Nutrition and
Food Research, 50, 11401149.
Verzelloni, E., Tagliazucchi, D., & Conte, A. (2010a). Changes in major antioxidant
compounds during aging of traditional balsamic vinegar. Journal of Food
Biochemistry, 34, 152171.
Verzelloni, E., Tagliazucchi, D., & Conte, A. (2010b). From balsamic to healthy:
Traditional balsamic vinegar melanoidins inhibit lipid peroxidation during
simulated gastric digestion of meat. Food and Chemical Toxicology, 48,
20972102.
H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584 583
Vignoli, J. A., Bassoli, D. G., & Benassi, M. T. (2011). Antioxidant activity, polyphenols,
caffeine and melanoidins in soluble coffee: The inuence of processing
conditions and raw material. Food Chemistry, 124, 863868.
Wang, H. Y., YU, X. Y. & Qian, H. (2011). Antioxidant capacitiy of meat avouring
on human lymphocytes. Journal of Food Science and Biotechnology, In Press,
Corrected Proof.
Wang, H. Y., YU, X. Y., QIAN, H., & YAO, W. R. (2010). Study on separation of
antioxidant components in meat avouring and their antioxidant capacities.
Food Science, 31, 5162.
Wen, X., Enokizo, A., Hattori, H., Kobayashi, S., Murata, M., & Homma, S. (2005).
Effect of roasting on properties of the zinc-chelating substance in coffee brews.
Journal of Agricultural and Food Chemistry, 53, 26842689.
Wilkinson, J. t., &Clapper, M. L. (1997). Detoxication enzymes and chemoprevention.
Proceedings of the Society for Experimental Biology and Medicine, 216, 192200.
Xu, Q., Tao, W., & Ao, Z. (2007). Antioxidant activity of vinegar melanoidins. Food
Chemistry, 102, 841849.
584 H.-Y. Wang et al. / Food Chemistry 128 (2011) 573584

También podría gustarte