You are on page 1of 48

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/227918544

Anthocyanins as food colorants – A review

Article  in  Journal of Food Biochemistry · September 1987


DOI: 10.1111/j.1745-4514.1987.tb00123.x

CITATIONS READS

214 2,385

4 authors, including:

Robert Jackman Rickey Y Yada

37 PUBLICATIONS   1,214 CITATIONS   
University of Guelph
182 PUBLICATIONS   3,502 CITATIONS   
SEE PROFILE
SEE PROFILE

Robert Alexander Speers


Dalhousie University
86 PUBLICATIONS   1,146 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Nanotechnological Examination of the SMA strain of Saccharomyces pastorianus View project

Garlic-containing Natural Health Products View project

All content following this page was uploaded by Robert Alexander Speers on 15 October 2017.

The user has requested enhancement of the downloaded file.


ANTHOCYANINS AS FOOD COLORANTS - A REVIEW
ROBERT L. JACKMA"
and
RICKEY Y. YADA

Department of Foai Science,


University of Guelph,
Guelph, Ontario N1G 2W1
and
MARVIN A. TUNG
and
R. ALEX SPEERS
Department of Food Science,
University of British Columbia
Suite 248-2357 Main Mall,
Vancouver, B. C. V6T 2A2

Received for Publication Gptember 8, 1986


Accepted for Publication January 12, 1987

ABSTRACT
A number of factors affecting anthocyanin stability and color are discussed in
this review. 13re anthocyanins are probably the most spectacular of plant
pigments since they are responsiblefor most of the red, purple and bluepigmen-
tation ofjlowers, h i t s and vegetables. However, because of their highly reac-
tive nature, anthocyanins readily degrade, or react with other constituents in the
media, to form colorless or brown colored compounds. lhe presence of an ox-
onium ion adjacent to carbon 2 makes the anthocyanins patticularly susceptible
to nucleophilic attack by such compounds as sulfur dioxide, ascorbic acid,
hydrogen peroxide and even water. Loss of anthocyanin pigmentation also oc-
curs in the presence of oxygen and various enzymes, and as a result of high
temperatureprocessing. A certain degree ofpigment stabilization may be confer-
red by acylation with various organic acidr, copigmentation, self-association
and/or metal chelation. In addition, p H has a marked effect on anthocyanin
stability, and on the color of media containing these pigments. A number of
anthocyanin-richsources have been investigatedfor their potential as commer-
cial pigment extracts. Although their application is primarily limited to acidic

'To whom correspondence should be directed.

Journal of Food Biochemistry ll(1987) 201-247. All Rights Reserved.


@Copyright 1987byFood & Nutrition Press, Inc., Wesrpon, Connecticut. 201
202 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

media, continued research on the chemistry of anthocyanins m a y lead to applica-


tion and stabilization of these pigments in a wider variety of food prducts.

INTRODUCTION
Anthocyanins comprise a group of naturally occurring pigments which are
responsible for the blue, red, purple, violet and magenta coloration of most
species in the plant kingdom. These polyphenolic substances are glycosides of
anthocyanins, polyhydroxy and polymethoxy derivatives of 2-phenylbenzo-
pyrylium or flavylium salts. The large number of glycosyl and acyl groups
which may bind to the sixteen different naturally occurring anthocyanidins
(Timberlake and Bridle 1975) has contributed to the more than 247 different an-
thocyanin pigments which have been reported in the literature. The structural
diversity of these pigments, in addition to their highly reactive and amphoteric
nature, has made their identification and quantitation tedious and difficult.
Qualitative and quantitative methods used for anthocyanin analysis have been
reviewed by several authors (Markakis 1974; Shrikhande 1976; Timberlake
1980; Francis 1982).
The safety of synthetic pigments has been in question for a number of years,
leading to progressive interest in the use of natural coloring compounds in food
processing and manufacture (Francis 1984). A major problem, however, is their
inherent instability, especially in complex systems such as food. The antho-
cyanins show greatest stability under acidic conditions, but are generally
unstable and degrade by one of several possible mechanisms to form first col-
orless, then insoluble brown colored products. These changes may occur under
normal conditions of processing and storage. Therefore, a knowledge of the fac-
tors governing the stability of anthocyanins, and the possible mechanisms by
which they may degrade, is important if these pigments are to be utilized as in-
gredients in the manufacture of products for which maximum color retention is
desired throughout their shelf-life.
A number of factors influence anthocyanin pigment stability including pH and
temperature, as well as the presence of ascorbic acid, sugars, metal ions and
copigments. Several comprehensive reviews have been published (Jurd 1972a;
Adams 1973a; Markakis 1974; Shrikhande 1976; Francis 1977; Eskin 1979;
Hrazdina 1974, 1981) which deal with the stability and color of anthocyanins as
influenced by these factors. This paper will review the factors known to in-
fluence anthocyanin color and degradation, and the attempts which have been
made to stabilize anthocyanin pigments in light of their possible use as colorants
in a variety of food systems.
ANTHOCYANINS AS FOOD COLORANTS -A REVlEW 203

FACTORS AFFECTING ANTHOCYANIN COLOR


AND STABILITY
Structural Effects
Anthocyanin pigments are virtually ubiquitous in the plant kingdom. They are
structurally characterized by possessing a C&c6 carbon skeleton, and because
of this, may be regarded as flavonoid compounds (Geissman 1962). However,
despite also having the same biosyntheticorigin as other natural flavonoids (Har-
borne 1967; Grisebach 1982), anthocyanins differ from these latter compounds
by strongly absorbing visible light (Brouillard 1982). The color and stability of
anthocyanins are functions of their inherent molecular structure, in addition to
various external factors.
The naturally occurring anthocyanins are glycosides of mainly six antho-
cyanidins, these being polyhydroxy and polymethoxy derivatives of
2-phenylbenzopyrylium (flavylium) salts (Fig. 1). These, however, cannot ex-
plain the infinite color variations observed in tissues of higher plants (Asen
1976). The high reactivity of the aglycone moiety is largely responsible for the
various structural modifications which anthocyanins undergo in acidic media
(Timberlake and Bridle 1967a; Brouillard 1982). The nature of constituent
sugars, acyl and methoxy groups has little effect on reactions which induce
structural modifications (Timberlake 1980), but their positions of attachment in
the pigment molecule may have a profound influence. According to Asen
(1976), the various structural modifications are generally obscure, and as yet,
illdefined, although they are assumed to be responsible for the range of colors
associated with anthyocyanins.
Naturally occurring anthocyanins are always glycosylated at the C-3position
(Harborne 1958a) (Fig. 1). The glycosyl moiety may be any of a number of dif-
ferent monosaccharides (mainly glucose, rhamnose, galactose, xylose or
arabinose), disaccharides (mainly rutinose, sambubiose or sophorose; lathyrose,
gentiobiose or laminariobioseoccur less frequently) or trisaccharides; however,
glucose is the most common. Since anthocyanidins are generally unstable and
less soluble in aqueous media than anthocyanins (Timberlake and Bridle 1966),
glycosylation is assumed to confer stability and solubility to the pigment
molecule (Harborne 1979). Loss of the glycosyl moiety at position 3 is accom-
panied by rapid decomposition of the aglycone in model systems, with irreversi-
ble loss of color (Jurd 1972a).
If a second site in the anthocyanin molecule is glycosylated, it is often located
at the 5 position (Brouillard 1982). Each glycosyl substitution is associated with
a characteristic bathochrornic shift, such that the difference in the absorption
maxima of anthocyanidin-3-glycosidesand anthocyanidin-3,5diglycosideshas
204 R.L. JACKMAN, R.Y. YADA, M.A. TIJNO AND R.A. SPEERS

T3’

Anion -

6H

Pelargonidin H H
Cyan id in OH H
Peon id in OCH3 H

Delphinidin OH OH

P e t un id in OCH3 OH

Malvidin OCH3 OCH3

FIG.1. STRUCTURES OF THE SIX COMMON ANTHOCYANIDINS IN WHICH


R3 I and RS r ARE HYDROXYL GROUPS
Structures of the corresponding anthocyanins contain moieties as indicated. The anthocyanins are
always glycosylated at position 3 (adapted from Timberlake 1980).

been used as a means for differentiation by spectral methods (Harborne 1958b).


Based on the position of attachment and number of sugars, Harborne (1963) has
classified the anthocyanins into 18 groups (i.e., 3-monosides, 3-biosides,
3-triosides, 3,5dimonosides, 3-bioside-5-monosides, 3-bioside-7-monosides,
etc.).
Besides the 3 and 5 positions, glycosylation may also occur in positions 7 , 3 I,
4 ’ and/or 5 ’ (Brouillard 1982). However, steric hindrance generally precludes
glycosylation at both the 3 ’ and 4’ positions of a pigment molecule. Brouillard
(1982) noted that no anthocyanin has been reported where glycosylation or
methylation occurs in all of positions 3, 5 , 7 and 4 ’. A free hydroxyl group at
any of the 5 , 7 or 4 ’ positions is essential for the formation of a quinoidal
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 205

(anhydrobase) structure (Brouillard 1982). This structure is derived from the


flavylium cation form of the anthocyanin, by loss of an acidic hydroxyl
hydrogen, generally above pH 3 (Fig. 2). The quinoidal structure of antho-
cyanins is primarily responsible for the pigmentation of flower (Jurd and Asen
1966; Asen et al. 1970, 1972, 1975, 1977) and fruit tissues (Scheffeldt and
Hrazdina 1978; Williams and Hrazdina 1979).

f--

t--

OH
C B
FIG.2. STRUCTURAL TRANSFORMATIONS OF ANTHOCYANINS
(ANTHOCYANIDIN-3-GLYCOSIDES)
At low pH, the flavylium cation (AH'), and its numerous resonance structures, is the dominant
form. With increases in pH, the carbinol pseudobase (B) forms with a proportional decrease in the
concentration of AH+.The colorless pseudobase (B) exists in tautomeric equilibrium with its
colorless chalcone (C). As the pH rises above 5 the blue quinoidal anhydrobase (A)begins to form,
resulting in a blue solution. R3I and R, t are normally H, OH or OCH, moieties; Gly represents
a glycosyl moiety (adapted from Brouillard and Delaporte 1977).

The flavylium cation, from which the quinoidal structure is derived, is


relatively stable under acidic conditions @H 1-3). This cation has been describ-
ed as a heterocyclic carboxonium cation, its positive charge delocalized over the
entire heterocyclic structure, giving rise to six contributing resonance forms
(Fig. 3) (Ingold 1969; Brouillard 1982). According to Ben& et al. (1967), the
highest partial positive charges occur at the 2 and 4 positions of the flavylium ca-
tion. The stability of the cation is highly dependent upon nucleophilic attack at
either position, by such compounds as water and sulfite ions (Ben& et al. 1%7;
Brouillard and El Hage Chahine 198Oa). In addition, protons of hydroxyl groups
at positions 5, 7 and 4 ' may easily be removed with only slight increases in pH.
206 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

woH
H

Ho
\
- H

HO'
O

\
-
e H

HO
/ /
o g

FIG. 3. THE SIX MAJOR CONTRIBUTING RESONANCE FORMS OF


5,7,4 '-TREIYDROXYFLAVYLIUMCATION
(adapted from Brouillard 1982).

The presence of the oxonium ion adjacent to the 2 position in the molecule is
responsible for the characteristic amphoteric nature of anthocyanins and their
ability to form salts with acids (Fuleki 1967). Due to this property, anthocyanins
exist as either an acid or a base, depending on the pH of the medium. This has
proven useful for electrophoretic separation of these pigments (Markakis 1960;
Von Elbe et al. 1969).
The structural changes of ordinary anthocyanins do not contribute much color
to systems of slightly acidic or neutral pH. Stabilizationof anthocyanin color and
structure is influenced by the presence of acyl groups linked to the sugar
moieties of the pigment molecule, in addition to a number of other factors.
Acylation of glycosylated anthocyanidins occurs primarily at the C-3 sugar
(Harborne 1964; Somers 1966a). Anthocyanins containing two or more acyl
groups have been discovered (Saito et al. 1972; Asen et al. 1972, 1977; Du and
Francis 1975; Asen 1976). These pigments display excellent color stability
throughout the entire pH range (Asen 1976).
Brouillard (1982) has suggested the presence of hydrophobic interactions bet-
ween the pyrylium ring and the aromatic moieties of the acyl groups protects the
pyrylium ring from nucleophilic attack (by water). Brouillard (1982) proposed
that one acyl group was situated above the pyrylium ring and the other beneath
it. Monoacylated anthocyanins have not displayed the color stability of antho-
cyanins containing more than one acyl group, suggesting that the presence of
two constituent acyl groups is required for good color retention in aqueous
media (Brouillard 1982). Deacylation results in fading that occurs immediately
after dissolution in slightly acidic or neutral media, similar to the behavior of
normal anthocyanins(Yoshitama 1978). Yoshitama and Hayashi (1974) reported
that, in neutral aqueous solutions, anthocyanins acylated with p-coumaric acid
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 207

were unstable in comparison to those acylated with caffeic acid. These authors
assumed the vicinal hydroxy grouping of the caffeic acid to be important in the
stability of the anthocyanins, especially above pH 3.
The pH dependent association of acyl groups with anthocyanins has led
Williams and Hrazdina (1979) and Scheffeldt and Hrazdina (1978) to assume the
mechanism of attachment to be through hydrogen bond formation. Additional
unsaturation at the 2 and 3 positions in the complex forming structure of the
flavonoid, electrostatic forces, and configurational or steric effects have also
been deemed important (Hrazdina 1981).
The main acylating groups tend to be the phenolic acids, pcoumaric ,caffeic,
ferulic or sinapic acids (Harbome 1964; Albach et al. 1965; Francis and Har-
borne 1966; Somers 1966a), but may sometimes be p-hydroxybenzoic (Shibata
and Yoshimata 1968), malonic (Bloom and Geissman 1973) or acetic acids
(Anderson et al. 1970). Harbome (1964) proposed that all anthocyanins are
ionically bound in the cell vacuole to aliphatic organic acids such as malonic,
malic or citric acid.
Under acidic conditions the color of anthocyanins is determined largely by the
degree of hydroxylation in the B-ring of the aglycone (Fig. l), the greater the
substitution the bluer the color (Asen 1976). Thus, glycosides of delphinidin are
bluer than those of cyanidin, which themselves are bluer than those of
pelargonidin. Methylation of one or more of these constituent hydroxyl groups
tends to have the opposite effect of hydroxylation, causing pigments to become
more red in color (Francis 1977). Aqueous extracts containing primarily
pelargonidin- and/or cyanidin-glycosides, therefore, tend to be orange-red,
those with glycosides of peonidin are deep red and those containing delphinidin-,
petunidin- and/or malvidin-glycosides display a bluish-red color (Francis 1977).
Hrazdina et al. (1970) have shown that a greater degree of methoxylation of the
anthocyanidin-3,5diglucosidesof grapes was associated with increased pigment
stability, but an increase in aglycone hydroxylation had the opposite effect.

pH Effects
pH has a marked influence on the color of anthocyanin solutions. Antho-
cyanins behave somewhat like pH indicators as a result of their amphoteric
nature. Below pH 3,anthocyanin solutions display their most intense red colora-
tion. When the pH of such solutions is raised, their red color normally fades to
the point where they appear colorless in the pH range of 4 to 5. Further increases
in pH give rise to anthocyanin solutions which are purple and blue, and these,
upon storage or heat treatment, have been observed to change in pigmentation
h m blue to yellow.
Prior to studies carried out by Brouillard and coworkers (Brouillard and
Dubois 1977; Brouillard and Delaporte 1977; Brouillard er al. 1978), the
208 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

observed changes in pigmentation with variations in pH were attributed to the


equilibrium reaction scheme presented in Eq. (1) (Jurd 1963).

AH+ *
-H+

+H+
A 7 B (1)

(flavylium (quinoidal (carbinol


cation) anhydrobase) pseudobase)

In this scheme, under acidic conditions, there is an equilibrium between the


flavyium cation(AH+)and the carbinol pseudobase (B) form of the anthocyanin,
with the supposed existence of a transient species, the quinoidal anhydrobase(A)
structure obtained by the deprotonation of the flavylium cation (Sondheimer
1953; Timberlake and Bridle 1966; 1967a; Harper 1968; Jurd 1963, 1972a).
Pigment solutions above pH 7, held for an extended perior or at elevated
temperatures, were presumed to gradually change in pigmentation from blue to
yellow as an indirect result of the formation of a chalcone (C) structure via ring
fission of the anhydrobase (A)(Hrazdina 1974). The existence of the chalcone
structure was postulated by Markakis et al. (1957) and Adams (1972, 1973b) in
studies of the thermal degradation of anthocyanins. The chalcone has been
described as a colorless compound (Timberlake 1980); however, its ionized
form is associated with a pale yellow color (Brouillard 1982). This ionized
species is not particularly stable, and therefore, the observed yellow color tends
to disappear with time, depending on the solution pH (Brouillard 1982).
Hrazdina (1971) demonstrated the formation of 3,5-di-(O-&D-glucosy1)-7-
hydroxycoumarin from anthocyanidin-3,5-diglucosidesin heated aqueous solu-
tions at pH 3 to 7, presumably by structural transformation of the anhydrobase
(Fig. 4). However, the validity of the proposed mechanism is questioned since
the structural transformations of anthocyanins (Fig. 2) favor the formation of
chalcone at elevated temperatures (Brouillard and Delaporte 1977). Thus, direct
involvement of the chalcone, or its precursor, the carbinol base, is proposed in
the formation of coumarin derivatives from the anthocyanidin-3,5-diglucosides.
It should be noted that anthocyanidin-3-glucosides do not form coumarin
derivatives (Hrazdina and Franzese 1974).
The scheme illustrated in Eq. (1) was questioned by Brouillard and Dubois
(1977) on the basis of (a) the lack of evidence for anhydrobase formation below
an acidity threshold @H 3-4), and (b) the fact that the flavylium cationcarbinol
equilibrium was shown to exist even when the cation lacked a hydroxyl group.
Using chemical relaxation methds, Brouillard and Dubois (1977) showed that
equilibrium between the carbinol and quinoidal base forms of malvin occurred
exclusively by way of the flavylium cation, contrary to previous reports (Jurd
and Ckissmann 1963; Jurd and Asen 1966; Timberlake and Bridle 1967a;
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 209

FIG. 4. MECHANISM OF BREAKDOWN OF ANTHOCYANIDIN-3,S-DIGLUCOSIDES, AT


pH 3.7, TO (a) COUMARIN DERIVATIVES AND (b) A STRUCTURE RESEMBLING
THE B-RING OF THE PARENT ANTHOCYANIN
G1 represents a glucose moiety (adapted from Hrazdina 1971).

Harper and Chandler 1967; Jurd 1972a) that the quinoidal base did not hydrate.
Brouillard and Delaporte (1977) demonstrated that in acidic aqueous media
@H 1-6) at 25 OC,malvidin-3-glucosideexisted in four forms in equilibrium: the
quinoidal base (A), the flavylium cation (AH+), the carbinol pseudobase (B) and
the chalcone (C). The general transformation mechanisms involved are il-
lustrated in Fig. 2. The flavylium cation has been shown to yield the quinoidal
base through loss of a proton. The nucleophilic addition of water to this cation
yields the carbinol structure which exists in a ringchain watercatalyzed
tautomeric equilibrium with the chalcone form of the anthocyanin. The cation
deprotonation by the solvent is exothermic, whereas the cation hydration and
pyrylium ring opening are both endothermic and associated with positive en-
tropy changes (Brouillard and Delaporte 1977). Thus, any rise in temperature
will favor formation of the chalcone at the expense of the quinoidal, flavylium
and carbinol forms of a particular anthocyanin. This is consistent with earlier
work by Markakis et al. (1957) and Adams (1972, 1973b).
Timberlake (1980) and Brouillard (1982) suggested that the distributionof the
four different anthocyanin structures at a particular pH, under equilibrium con-
ditions, may lead to some interesting conclusions. Research by Brouillard and
Delaporte (1978) and Brouillard et al. (1982) has shown that in very acidic
media @H0.5), the red flavyliumcation exists as the only species at equilibrium
(Fig. 5 ) . With an increase in solution pH, both the concentration of this species
and the pigmentation of the solution decrease, as the cation hydrates to the col-
orless carbinol base.
The formation of colorless chalcone and blue quinoidal anhydrobase begins at
a pH slightly below that corresponding to the pK characteristic of the
equilibrium between the flavylium cation and carbinol base (when an equal
amount of each form is present). The proportions of each of the chalcone,
quinoidal base and carbinol base continue to increase with increasing pH
(generally to pH 4.5) at the expense of the red flavylium cation. In the pH range
210 R.L. JACKMAN, R.Y. YADA, M.A. TUNG ANDR.A. SPEERS

" 1 2 t 3 4 t 5 6
pK=2 60 pK=4 25
(AH': 8 ) (A $ AH')

PH
FIG. 5. DISTRIBUTION OF THE FOUR DIFFERENT ANTHOCYANIN STRUCTURES
WITH pH UNDER EQUILIBRIUM CONDITIONS
Malvidin-3-glucoside: 25 "C.AH+ = red flavylium cation, B = colorless carbinol pseudobase,
C = colorless cbalcone, A = blue quinoidal anhydrobase (adapted from Timberlake 1980).

of 4 to 5 , the concentrations of the two colored anthocyanin forms tend to be


very small, their total effect contributing little to the pigmentation of the solu-
tion.
Markakis (1960) showed that anthocyanins display minimum coloration at
their isoelectric point, or when the anthocyanin is predominantly in its carbinol
pseudobase form. The equilibrium between the red flavylium cation and blue
quinoidal base is associated with a second characteristicpK value (i.e., pH 4.25
in Fig. 5). Above pH 5, the blue quinoidal base is the only colored anthocyanin
form present; however, it is present at very low relative concentrations. It is evi-
dent that as the pH of a given system increases, the red pigmentation associated
with the anthocyanins is reduced. Unless some stabilizing factor is present to
augment the coloration of flavylium cation and quinoidal base forms (i.e., by
acylation, copigmentation, etc.) or at least maintaintheir concentrations, the use
of anthocyanins as food colorants would appear to be ineffective given that most
foods normally have a pH in the range of 3 to 7 (Timberlake 1980).
The rate of anthocyanin degradation has been shown by Meschter (1953) to be
pH dependent. Lowering the pH in the range of 5.0 to 1.O was shown to increase
the retention of anthocyanins in strawberryjuice, with maximum stability being
observed at pH 1.8 (Meschter 1953). Daravingas and Cain (1968) also observed
that the degradation of the major raspberry anthocyanin, cyanidin-3diglucoside,
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 211

was markedly retarded at low pH values. Tinsley and Bockian (1960)observed a


strong stabilizing effect on pelargonidin-3-glucosidein model systems heated in
air by lowering the pH.
Evidence had been presented earlier for the existence of an equilibrium bet-
ween hydronium ions and an anthocyanin, pelargonidin-3-glucoside,in the
forms of the red flavylium cation and the hydrated, colorless pseudobase in the
pH range of 1.2 to 4.0 (Sondheimer 1953).This equilibrium was later confirmed
by Lukton et al. (1956).
The rate of destruction of pelargonidin-3-glucosidein the presence of oxygen,
both in buffered solutions and strawberry juice at 45 "C,has been shown to be
directly related to the amount of pigment in the pseudobase form, but inversely
proportional to the amount of red flavylium cation (Lukton et af. 1956). Adams
(1972,1973b)similarly demonstrated accelerated anthocyanin destruction in the
presence of oxygen at pH 2 to 4.In contrast, anthocyanin destruction has been
shown to be virtually pH indendent in the range of pH 2.0to 4.5in the absence
of oxygen (Adams 1972, 1973b;Lukton et al. 1956).
Copigmentationand Condensation Reactions
Markakis (1974)noted that blue anthocyanin coloration of plant products was
not due to a pH effect, but rather, to copigmentationand metal chelation. Antho-
cyanins have been shown to form weak complexes with a number of compounds
such as gallotannins, other flavonoids (especially quercetin and ruth) and
polysaccharides (Asen et al. 1972)through what is referred to as intermolecular
copigmentation, or with substances which form part of an anthocyanin pigment
molecule through intramolecular copigmentation. It should be noted that the col-
or augmentation observed for natural anthocyanin containing systems with
gallotannin has recently been attributed to a copigmentationeffect, as well as to
the interaction of gallotannin with polyphenol oxidase to form a complex without
oxidase activity (Sakuraba and Ichinose 1982). Polyphenol oxidase is among
several enzymes known to cause a decolorizing effect on anthocyanins, a topic to
be discussed later.
The mechanism of association between anthocyanin and copigment is presum-
ed to involve hydrogen bonding (Timberlake and Bridle 1975;Asen 1976). Ac-
cording to Brouillard (1982),intramolecular copigmentation (i.e., acylation) is
more efficient in the stabilization of anthocyaninsthan intermolecular copigmen-
tation. Acylated anthocyanins generally display a lesser response to copigmenta-
tion reactions than the nonacylated pigments (Hrazdina 1981). The copigmenta-
tion effect may occur in conjunction with metal complexing, a special case of in-
tramolecular copigmentation (Jurd and Asen 1966;Van Teeling et al. 1971),but
according to Nakayama and Powers (1972),this is more often the exception than
the rule.
Complex formation, or copigmentation, results in a bathochromic shift in the
visible wavelength of maximum absorption, from red to blue (Asen et al. 1972).
212 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

Asen et al. (1972) reported that copigmentation caused a bathochromicshift with


both the flavylium cation and the quinoidal base, but that increases in tinctorial
strength were due only to stabilizationof the latter. This effect tended to be more
pronounced as anthocyanin concentration, and the ratio of copigment to antho-
cyanin, increased (Asen et al. 1972; Asen 1974, 1976; Timberlake 1980).
When pigment concentration becomes relatively high, anthocyanins
themselvesmay act as copigments (Nakayama and Powers 1972). Scheffeldt and
Hrazdina (1978) measured the color of solutions containing a constant amount of
the flavonol, rutin, as copigment over a wide range of anthocyanin concentra-
tions. They demonstrated that at low levels of anthocyanin (isolated from Vitis
spp.), color augmentationby copigmentation was intensified, but decreased with
corresponding increases in anthocyanin concentration. The reduced availability
of anthocyanin for copigmentation with rutin was attributed to stabilization of
the flavylium cation through self-association (Scheffeldt and Hrazdina 1978). A
reddish-brown precipitate, which developed in the acidified extract of
strawberry puree during storage, was presumed to be the result of polymerized
pelargonidin-3-glucoside(Erlandson and Wrolstad 1972). Spayd et al. (1984) at-
tributed increased turbidity and polymeric color, and a reduction in the propor-
tion of color due to tannin, to condensation of anthocyanins in apple and pear
juices blended with anthocyanin-containingfruit juices. They found that within a
given base juice, turbidity was highly correlated with polymeric color.
According to Timberlake and Bridle (1982a), the effects of self-association,
which tends to occur at low pH levels, and copigmentation, which is more pro-
minent at higher pH, may occur simultaneously as concentrationsof anthocyanin
are increased. The net result is that color may increase more than proportionally
to pigment concentration. Timberlake (1980) and Timberlake and Bridle (1982a)
have, therefore, stressed the importance of preventing the loss of anthocyanins
through dilution or degradation during processing, since this may result in a pro-
portionally greater loss of product color.
Aside from anthocyanin concentration (Scheffeldt and Hrazdina 1978; Asen
1976; Asen et al. 1971, 1972), a variety of other factors may influence the
copigmentation effect (Osawa 1982). The nature of the anthocyanin has been
shown to have a large effect on the bathochromic shift (bluing) associated with a
given copigment (Asen et al. 1972; Yazaki 1976; Williams and Hrazdina 1979).
Williams and Hrazdina (1979) noted that changes in tinctorial intensity of antho-
cyanin solutions were not uniform in the pH range of 1 to 7, but were influenced
by the substitution patterns of the B-ring and the nature of substituents in the 3
and 5 positions. The nature of the copigment (Asen et al. 1972) and its concen-
tration (Asen et al. 1971, 1972; Asen 1976) have also been shown to influence
the copigmentation effect.
The effect of pH on the color and stability of pure anthocyanins has been
covered previously in this review; however, its effect on copigmentation is no
ANTHOCYANINS AS FOODCOLORANTS - AREVLEW 213

less important (Asen er al. 1972; Yazaki 1976). pH affects not only the form of
anthocyanin which is available for participation in a copigmentation reaction,
but also the ionic nature of the copigment(s). Several other factors which may af-
fect copigmentation and the color augmentation associated wtih this
phenomenon have been discussed by Osawa (1982), and one is directed to his
review for more extensive coverage of the subject.
Besides the various flowers whose colors have been attributed to copigmenta-
tion (Saito et al. 1961, 1972; Yoshitama et al. 1975; Yoshitama and Abe 1976;
Yoshitama 1977; Asen et al. 1970, 1971, 1972, 1975, 1977), the tinctorial
strength and general color of fruit juices and red wines are also thought to be the
result of this phenomenon. Fruit juices obtained from enzymatically treated fruit
mashes have been described as being more highly colored than nonenzyme
treated press-juice. This is not surprising, given that various cellular constituents
such as flavonoids, alkaloids, amino acids, and nucleotides which act as
copigments to varying degrees (Asen et al. 1972; Timberlake and Bride 1977;
Osawa 1982) are decompartmentalized in the process, and thereby made
available for association with anthocyanin pigments. Co and Markakis (1%8)
identified a number of flavonoid compounds from extracts of strawberry fruit,
including catechin, quercetin-3-glucoside, and kaempferol-3-glucoside, all of
these having been identified as potential copigments by Asen et al. (1972).
Polymeric flavonoids and anthocyanins play an important role in the colora-
tion of grapes and red wines (Somers 1966b, 1967, 1971; McCloskey and
Yengoyan 1981). Sastry and Tischer (1952) and Somers (1971) found grape tan-
nins (condensed flavonoids) to have a protective effect on the anthocyanins of
Concord grape juice and red wines. Monomeric anthocyanins, initially responsi-
ble for the pigmentation of young red wines, have been shown to be progressive-
ly and irreversibly displaced during aging by more stable polymeric pigments
(Somers 1971; McCloskey and Yengoyan 1981). Somers (1971) has reported
such polymeric material in (aged) red wines to be less pH sensitive and relatively
resistant to decolorization by sulfur dioxide. Utilizing this latter property, a
method has been developed for determining the contribution of monomeric (and
also polymeric, by difference) anthocyanins to the overall color of red wines
(Somers and Evans 1974).
A number of condensation reactions have been demonstrated in studies with
synthetic flavylium salts. These salts readily condense with phloroglucinol (Jurd
and Waiss 1965) and catechin (Jurd 1%7) in aqueous solution to form flavanyl-
phloroglucinol and 4-flavanyl-flavyliumsalts, respectively. The linkages involv-
ed were similar to those involved in condensed tannins (Jurd 1967). In addition,
Jurd (1972b) showed synthetic flavylium salts to undergo dimerization in
aqueous media, the resulting dimeric pigments displaying similar spectral pro-
perties as those of their parent compounds. The dimerization reaction was
thought to involve initial condensationof the flavylium salts and its carbinol base
214 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

to yield a product which subsequently underwent a series of intermolecular


hydride ion transfer reactions (Jurd 1972b). These condensation reactions have
yet to be demonstrated for naturally occurring anthocyanins and/or antho-
cyanidins. Hydrolysis of constituents in the 3 or 3 and 5 positions would first be
necessary for the above reactions to take place (Hrazdina 1974). However, such
hydrolytic reactions may occur under the acidic conditions (i.e., pH 2-5) of fruit
juices and wines, making the proposed condensation reactions entirely feasible
in natural systems during processing and storage (Hrazdina 1974).
While many condensation reactions involving anthocyanins have been observ-
ed to enhance pigmentation, other condensation reactions have been shown to
reduce it. As indicated by Jurd (1972a), flavylium salts have been shown to
readily condense with amino acids, phioroglucinol, catechin and other
nucleophiles to yield 4-substituted flav-2enes which are reactive and may
undergo further changes depending on the nature of the 4-substituent (Jurd and
Waiss 1965; Jurd 1967). Ultimately, loss of anthocyanincolor would be observ-
ed.
The anthocyanin-ascorbicacid condensation is an example of a decolorizing
reaction (Jurd 1972a; Poei-Langston and Wrolstad 1981). The addition of ascor-
bic acid to anthwyaninsontainingjellies was shown by Freedman and Francis
(1984) to result in marked bleaching of these products. Sulfite bleaching of an-
thocyanins has also been attributed to nucleophilic attack of the flavylium cation
by the negatively charged bisulfite ion, yielding a chromen-4 (or 2)- sulfonic
acid (Jurd 1964a; Timberlake and Bride 1967b). Timberlake and Bridle (1968)
have shown flavylium salts containing methyl or phenyl moieties in position 4 to
be virtually unaffected by the bleaching effects of sulfur dioxide.
The phenomenon of copigmentation, and a number of related condensation
reactions, offers one explanation for the infinite color variations that occur in
plant tissues in the pH range where anthocyanins alone are known to be virtually
colorless (Asen et al. 1972). Although the stability of such colored complexes
has not been systematicallystudied under practical conditions (Marlrakis 1982a),
it appears as if they are not as labile as most of the uncondensed anthocyanins
(Timberlake 1980; Timberlake and Bridle 1982a; Ribereau-Gayon 1982).
Flavonols and flavones, as copigments, are always found in conjunction with an-
thocyanins in fruits and fruit juices (Hooper and Ayres 1950; Harborne 1959;
Chiriboga and Francis 1973), likely as a result of similar biosynthetic pathways
(Harborne 1967; Grisebach 1982).
The increasing application of enzymes to fruit mashes has resulted in increas-
ed copigmentation effects as a direct result of copigment decompartmentaliza-
tion, yielding more colorful and attractive fruit juices. Copigmentation has also
been shown to reduce the decolorizing effect of ascorbic acid (Shrikhande and
Francis 1974). Presumably, the condensation of anthocyanin with flavonol
prevents complex formation between anthocyanin and ascorbic acid.
ANTHOCYANINS AS FOOD COLORANTS - A REVEW 215

Metal Complexing
Anthocyanin solutions appear colorless under normal physiological conditions
@H 3 to 7) due to the relative instability of the anhydrobase. Copigmentation
and various other condensation reactions have been shown to act in stabilizing
anthocyanin pigments and their associated color; however, stabilization may
also result from complexing with iron, tin, aluminum, copper or various other
metal ions. According to Markakis (1974), these metal chelates and salts
(metallo-anthocyanins) normally cause a bathochromic displacement similar to
copigmentation, from red to stable blue and violet colors. Unlike copigmenta-
tion, however, only those anthocyanins with an orthodiphenolic group in the
B-ring are capable of complexing with metal ions (Geissman et al. 1953). It is
therefore possible to differentiate those common anthocyanins which possess
this group (cyanidin-, delphinidin- and petunidin-glycosides) from those which
do not, by the appearance of a bathochromic shift in the absorption maximum
upon addition of certain metals, i.e., iron or aluminum (Harborne 1973; Francis
1977).
Cyanidin-3,5diglucoside was shown by Bayer et al. (1966) not to form stable
or blue colored complexes with the divalent metals cobalt, nickel, calcium,
magnesium or barium in the pH range of 4 to 6. Iron and aluminum complexes,
however, were observed to be relatively stable with a deep blue color, similar to
chelates formed with tin, titanium, chromium, uranium and lead (Bayer et al.
1966). These latter metals tend to be present in only trace amounts, if at all, in
most plants, leading to a suggestion by Asen (1974, 1976)that it is unlikely that
metals other than iron or aluminum are involved in flower or fruit color. Pyysalo
(1973) showed cyanidin-3-galactosideand delphinidin to form intensely blue-
colored complexes in the presence of iron and tin ions. Cyanidin-3-glucoside
also formed a stable, colored complex in the presence of aluminum ions, with
maximum complex formation occurring at pH 5.5 (Jurd and Asen 1966).
pH has been shown to markedly affect the color of aluminum (Asen et al.
1969), iron and tin chelates (Pyysalo 1973). Within a narrow range @H 3.0 to
3.5) a slight pH change of only 0.1 unit may alter the color of aluminum chelates
from red to blue-violet (Asen et al. 1969). Complexes of cyanidin-3-glucoside
and aluminum were red below pH 3.0, changed to blue-violet up to pH 3.5, and
retained that color above pH 3.5 (Asen et al. 1969). Similar observations with
respect to color and pH sensitivity have been noted for iron chelates (Pyysalo
1973; Jurd and Asen 1966).
Despite the presence of metals in a number of isolated complex blue pigments
(Saito et al. 1961;Asen and Jurd 1967; Hayashi and Talc& 1970; Van Teeling
et al. 1971), Asen (1974, 1976) has suggested that metal chelates play only a
minor role in the color of plant products. However, a number of studies have
been carried out to investigate the stabilizing effects of metals under
216 R.L. JACKMAN, R.Y. YADA, M.A. TUN0 AND R.A. SPEERS

physiological conditions. Sistrunk and Cash (1970) were able to stabilize the col-
or of strawberry puree by the addition of stannic salts. Wrolstad and Erlandson
(1973) proposed that the stabilized red color of strawberryjuice was due to a tin-
cyanidin complex rather than to a tin complex with the major strawberry antho-
cyanin, pelargonidin-3-glucoside.Pelargonidin is incapable of complexing with
metals since it lacks an orthodiphenolic group (Geissman ef al. 1953). The
cyanidin was assumed to have been released from the colorless leucocyanidin
identified in strawberries by Co and Markakis (1968). A similar red complex of
tin and cyanidin has been reported in canned pears (Chandler and Clegg 197Oa,
b) . Chandler and Clegg (197Oc) noted that, despite the involvement of tin in this
complex, addition of stannous ions to susceptible pear puree prior to processing
partly or completely inhibited the development of red pigmentation. They noted
that the addition of other reducing agents such as SOz was also effective in in-
hibiting the discoloration.
The relatively high stability and tinctorial strength of the metal-anthocyanin
complexes have previously led to a proposal for their potential use as natural
food colorants (Jurd 1972a). The metals act by chelating the anthocyanins in a
stable quinoidal structure at pH values of 3 to 6 (Nakayama and Powers 1972),
the usual pH range of plant juices. Starr and Francis (1973), however, have
reported that although copper, iron, aluminum and tin provided greater protec-
tion, or stability, for anthocyanins of cranberry juice, the effect was not
beneficial since blue and brown discoloration also resulted from the
simultaneous formation of metal-tannin complexes. Segal and Oranescu (1978)
reported improved color stability in sour cherry, grape and strawberry juices to
which aluminum and tin salts were added. Coffey et al. (1981), investigatingthe
stability and color of purified cyanidin-3-glucosideand raspberryjuice extract in
the presence of tin (Sn2+)and aluminum (A13+)ions, concluded that although
metal complexing occurred, the pigment stability conferred by these complexes
was not sufficient to justify practical application.

Ascorbic Acid
Beattie et al. (1943) were among the first researchers to observe the concur-
rent disappearance of ascorbic acid and anthocyanin in stored fruit juices, and to
suggest a possible interaction between the two compounds. Similar observations
have since been reported by several investigators (Sondheimer and Kertesz
1952; Meschter 1953; Pratt et al. 1954; Markakis et al. 1957; Starr and Francis
1968). Studiesby Sondheimerand Kertesz (1953) suggested that maximal loss of
anthocyanins occurred under conditions most favorable to ascorbic acid oxida-
tion. This indicates the importance of oxidation products, rather than the ascor-
bic acid itself, in anthocyanin destruction. Although hydrogen peroxide is a pro-
duct of ascorbic acid oxidation, its formation in strawberry juice has not been
demonstrated. Despite this, Sondheimer and Kertesz (1952) demonstrated
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 217

significant losses in color due to the presence of added Hz02in strawberry juice.
Pratt et al. (1954)subsequently confmed the participation of ascorbic acid in
anthocyanin breakdown and suggested the participation of riboflavin in these
reactions.
Markakis et al. (1957)demonstrated a synergistic effect between ascorbic acid
and oxygen on pigment degradation. Starr and Francis (1968)similarly observed
that maximum pigment losses in cranberry cocktail occurred when high levels or
concentrations of oxygen and ascorbic acid were added to the juice.
An interaction between anthocyanin and ascorbic acid was noted in model
systems of strawberry anthocyanins, even in the absence of oxygen ( M a r M s et
al. 1957).Jurd (1972a)suggested a condensation reaction between anthocyanin
and ascorbic acid similar to that between flavylium salts and dimedone.
Degradation products formed from such a reaction were unstable and degraded
further to colorless compounds (Jurd 1972a). Poei-Langston and Wrolstad
(1981),using an anthocyanin-ascorbic acid-flavonol system, also presented
evidence that anthocyanin pigments and ascorbic acid degrade by a condensation
mechanism. Using polarographic methods, Harper (1968)showed that flavylium
salts were capable of accepting electrons at low negative potential. Shrikhande
and Francis (1974)subsequently proposed that anthocyanins, like copper ions
(Silverblatt et al. 1943),acted as possible initiators in the chain reactions involv-
ed in ascorbic acid degradation.
Work by Hooper and Ayres (1950),Timberlake (lW) , and Morton
Clegg
(1968) and Shrikhande and Francis (1974) indicated that flavonols retarded
ascorbic acid oxidation in anthocyanincontainingsystems. Harper et al. (1%9)
suggested that the antioxidant activity of flavonols was a function of their ability
to act as free radical acceptors. Their ability to act as copigments may also ex-
plain the observed antioxidant activity of these compounds.Theweak association
between flavonoid compounds and anthocyanins would inhibit, or at least
reduce, complex formation between ascorbic acid and anthocyanin, thereby
resulting in both pigment and ascorbic acid stabilization. Meschter (1953)
reported that dehydro-ascorbic acid had a decolorizing effect on anthocyanin
solutions as well, but noted that the rate was markedly slower than with equal
concentrations of ascorbic acid.
Wrolstad et d.(1970)found the effects of ascorbic acid on the color quality of
frozen strawberriesto be insignificantunder low temperature storage conditions.
Ascorbic acid was actually thought to stabilize anthocyanins at lower
temperatures, The accelerated destruction of anthocyanins in the presence of
ascorbic acid at high temperatureshas generally been attributed to the correspon-
ding degradation products of the ascorbic acid (Sondheimer and Kertesz 1953;
Wrolstad et al. 1970;Eskin 1979).Over a 24 h period at 50°C, Sistrunk and
Cash (1970)observed no effect of ascorbic acid (4 ppm added) on the redness
(CDM a) of strawberry puree. The color was claimed to be improved during the
first half of the holding period by the addition of ascorbic acid.
218 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

Sugars and Their Degradation Products


The effects of sugars and their degradation products have been investigated by
a number of researchers. Sugars generally accelerated the degradation of the ma-
jor strawberry anthocyanin, pelargonidin-3-glucoside (Meschter 1953;Mackin-
ney et al. 1955;Tinsley and Bockian 1960). Fructose, arabinose, lactose and
sorbose demonstrated far greater degradative effects on the pigment than
glucose, sucrose or maltose. Oxygen enhanced the degradative effects of all of
these sugars. According to Meschter (1953),in the presence of sugars, the rate
of pigment breakdown was associated with the rate at which the sugar was
degraded to furfural-type compounds. These compounds, furfural (derived
mainly from aldo-pentoses) and 5-hydroxymethylfurfural (HMF; formed from
keto-hexoses), may be formed via the Maillard reaction (Hodge 1953),or by the
oxidation of ascorbic acid (Kurata and Sakurai 1967; Sloan et al. 1969),
polyuronic acids (Starr and Francis 1968) or even anthocyanins themselves.
These degradation products may condense with anthocyanins to form complex
brown colored compounds (Debicki-Pospisil et al. 1983).
Using a model system of strawberry anthocyanins, Tinsley and Bockian
(1960)found that, by increasing the sugar concentration, the amount of ex-
tractable pigment was generally decreased. They attributed the loss in part to the
breakdown products furfural, HMF, and levulinic and formic acids. HMF
decomposes to levulinic and formic acids in the acidic environments of most
anthocyanin-containing fruits and their juices (Eskin 1979). In contrast to the
report by Calvi and Francis (1978)that the presence of the furyl aldehydes (fur-
fural and HMF) does not necessarily affect pigment degradation, a large number
of researchers (Meschter 1953; Markakis et al. 1957; Tinsley and Bockian
1960;Daravingas and Cain 1968;Segal and Negutz 1969;Debicki-Pospisil et
al. 1983)have found that furfural and HMF accelerate anthocyanindegradation,
even at low concentrations.
Daravingas and Cain (1968)demonstrated that sugar degradation products
were far more effective than their parent sugars in accelerating anthocyanin
breakdown. They investigated the effect of glucose and its degradation products
on the stability of the major black raspberry anthocyanin, cyanidin-3-
diglucoside. A minimum loss of pigment was observed in the presence of
glucose whereas a maximum loss was noted in the presence of furfural (Darav-
ingas and Cain 1968). Generally, the advanced sugar degradation products
(levulinic and formic acids) do not react nearly as rapidly with anthocyanins as
the parent furyl aldehydes (Mackinney et af. 1955; Markakis et al. 1957;
Tinsley and Bockian 1960;Daravingas and Cain 1965,1968;Debicki-Pospisil et
al. 1983). Debicki-Pospisil et al. (1983) described the accelerated pigment
degradation in the presence of furfural and HMF as being directly temperature
dependent, more pronounced in fruit juice and considerably decreased in
ANTHOCYAMNS AS FOOD COLORANTS - A REVIEW 219

nitrogen. Kinetic studies have indicated that the rate of pigment breakdown in
the presence of monosaccharides was first order (Tinsley and Bockian 1960;
Daravingas and Cain 1968).
Sulfur Dioxide
Sulfur dioxide, commonly used as a preservative in the food industry, has
been found to stabilize the anthocyanins of some berry fruits at low concentra-
tions, and is evidently more efficient in combination with ascorbic acid or rutin,
a copigment (Adams 1973a; Williams and Hrazdina 1979). Despite this stabiliz-
ing effect, however, the practical applicationof such a treatment is limited due to
the decolorizing effect of sulfite on the pigments (acidification to about pH 1 is
required for restoration of the red color)(Jurd 1964a). Jurd (1964a) suggested
that anthocyanins in the flavylium cation form participated in a reversible reac-
tion with bisulfite ions to form colorless flaven-4-sulfonic acid, a structure
analogous to the anthocyanic pseudobase (Fig. 6). Timberlake and Bridle
(196%) partially confirmed this hypothesis by demonstratingthe formation of a
complex between sulfur dioxide and the cations of flavylium salts, antho-
cyanidins, and anthocyanins in 1:1 molar ratios. Brouillard and E l Hage
Chahine (1980b) have shown that nucelophilic addition of hydrogen sulfite to the
flavylium salt of cyanin results in the formation of a highly stable, colorless
Meisenheimer-type adduct.
The formation constants for the anthocyanin-bisulfite complexes have been
published by Timberlake and Bridle (1967b) and Brouillard and El Hage
Chahine (1980b). The large values of these constants suggest that very small
amounts of free sulfur dioxide can decolorize significant quantities of antho-
cyanin (Brouillard and El Hage Chahine 198Ob; Markakis 1982a). Kinetic
studies have shown the anthocyanin-bisulfite complex to be very stable; the
bisulfite moiety presumably causes deactivation of the aglycone-sugar bond,
thereby preventing hydrolysis and subsequent formation of brown degradation
products (Admas 1972, 1973a). Timberlake and Bridle (1968) suggested that the
bisulfite ion binds at position 4, since flavylium salts containing methyl or
phenyl moieties at this position were essentially unaffected by sulfur dioxide.

Oxygen and Hydrogen Peroxide


The deleterious effects of molecular oxygen on anthocyanins have been
observed by a number of researchers. Nebesky et al. (1949) referred to oxygen
and temperature as the most specific accelerating agents in the destruction of an-
thocyanins in juices from blueberry, currant, cherry, raspberry, grape and
strawberry fruits. Starr and Francis (1968) demonstrated the detrimental effect
of oxygen on cranberry anthocyanins. Daravingas and Cain (1968) showed that
anthocyanins of raspberries canned under nitrogen or vacuum were more
220 R.L.JACKMAN,R.Y.YADA,M.A.TUNGANDR.A.
SPEERS

FIG.6.REVERSIBLE FORMATION OF (a) ANTHOCYANINS (FLAVYLIUM SALTS)


WITH BISULFITE ION TO FORM (b) FLAVEN-4-SULFONICACID
Gly represents a glycosyl moiety (Jurd 1964b).

stable than those canned under air. Oxygen may cause its deleterious effects on
anthocyanins through a direct oxidative mechanism and/or through indirect ox-
idation whereby oxidized constituents of the medium are capable of reacting
with the anthocyanins to yield colorless or brown pigmented products.
Hydrogen peroxide has also been shown to be detrimental to anthocyanins.
Sondheimerand Kertesz (1952,1953) studied the kinetics of the oxidation of an-
thocyanins by H202 both in pure solutions of the major strawberry anthocyanin
@elargonidin-3-glucoside) and in strawberry juice. The authors postulated that
ascorbic acid-induced destruction of anthocyanins in strawberry products was
due to H202 formed during the aerobic oxidation of ascorbic acid. H202 has yet
to be detected in strawberry juice; however, it is presumed to react with antho-
cyanin pigments to produce breakdown products which lead to the formation of a
brown resinous precipitate in fruit juices (Sondheimer and and Kertesz 1952,
1953). Markakis et ul. (1957) recovered 85% of 14C-labelledanthocyanin from
such a precipitate.
Lukton et ul. (1956) suggested that precipitate and haze development in fruit
juices formed from direct oxidation of the pseudobase form of anthocyanins,
followed by hydrolysis and subsequent formation of insoluble compounds. Hraz-
dina (1970) showed that, upon nucleophilic attack of malvidin-3,5diglucoside
with H202 in aqueous solution, the heterocyclic ring was cleaved between the 2
and 3 positions to form 0-benzoyloxyphenyl acetic acid esters of the malvone
type or esters of trihydroxymadelic acid (Fig. 7 (b)). Similarly, various acylated
FIG.7. NUCLEOPHILIC AlTACK OF (a) FLAVYLIUM SALT BY H202TO FORM
(b) 0-BENZOYLOXYPHENYLACETIC ACID ESTERS OF THE MALVONE TYPE UNDER
ACIDIC CONDITIONS, AND
(c) 3,5-di-(O-&D-GLUCOSUL)-7-HYDROXYCOUMARIN UNDER NEUTRAL CONDITIONS
R3' and % I are normally H, OH or OCH3 moieties; Gly represents a glycosyl moiety (Hrazdinaand
Prazese 1974).
222 R.L. JACKMAN, R.Y. YADA, M.A.TUN0 AND R.A. SPEERS

anthocyanidin-3,5diglucosideswere oxidized with H202under acidic condi-


tions to acylated o-benzolyphenyl acetic acid esters (Hrazdina and Franzese
1974). However, when the same reaction was carried out under neutral condi-
tions the reaction product was 3-0-acyl-glucosyl-5-0-glucosyl-7-hydroxy
coumarin (Hrazdina and Franzese 1974)(Fig. 7(c)). The marked effect of consti-
tuents on the flavylium nucleus, and of pH and the nature of the solvent system,
is further demonstrated by the oxidation of 3-methylflavylium salts to 3-acetyl
benzofurans in aqueous methanolic solution (Jurd 1964b) and to enol benzoate in
acidic solution (Jurd 1966). The oxidation products may participate in further
degradation reactions.

Temperature
As with most reactions the rate of anthocyanin degradation, in natural and
model systems, was found to be influenced markedly by processing and storage
temperatures (Nebesky et al. 1949; Meschter 1953; Decareau et al. 1956;
Markakis et al. 1957; Ponting et al. 1960; Daravingas and Cain 1965; Segal and
Negutz 1969; Hrazdina et al. 1970; Adams and Ongley 1973; Polesello and
Bonzini 1977; Mishkin and Saguy 1982; Simard et al. 1982).
Meschter (1953) demonstrated a linear relationship between the log rate of an-
thocyanin color deterioration in strawberry preserves and storage temperature.
From these thermal destruction data, he showed that the time for 50% destruc-
tion of anthocyanin color in strawberry preserves was 1 h at 100°C; 240 h at
38 "C; and 1300 h at 20 "C.Similar results were obtained for pelargonidin-3-
glucoside in buffered solution by Markakis et al. (1957) who recommendedhigh
temperature-short time processing of fruit and vegetable products in order to
achieve maximum pigment retention. This recommendation has been supported
by Adams and Ongley (1973) who showed that pigment destruction in canned
strawberries was negligible during a high temperature sterilization process com-
pared to that occurring during slow cooling and subsequent ambient temperature
storage.
Keith and Powers (1965) reported the degradation rates of pelargonidh-3-
glucoside, rnalvidin-3-glucoside and petunidin-3-glucoside followed first order
kinetics. Pelargonidin-3-glucosidewas noted to be the most stable of the three.
Tinsley and Bockian (1960) had previously shown pelargonidin-3-glucoside
degradation followed first order kinetics in a model system; Wrolstad et al.
(1970) showed that cyanidin-3-glucoside degradation was also first order.
Several other researchers have indicated that for a number of different antho-
cyanins, in model and natural systems, degradation followed a first order reac-
tion mechanism (Meschter 1953; Markakis et al. 1957; Segal and Negutz 1969;
Tanchev 1983).
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 223

The dependence of anthocyanindegradation rate on temperaturemay be deter-


mined from first order and thermal destruction curves, which provide informa-
tion needed to calculate such kinetic parameters as activation energy (Ea), Qlo
and z values. These various parameters are related, as shown in Eq. (2)and (3)
(Lund 1975):
Ea = 2.303 R T Ti
Z
N 10°C
Z= (3)
log Qio
where Ea is the activation energy (cal/mol); R is the universal gas constant
(1.987cal/mol K); T is the temperature (K); TI is normally taken as the
temperature 10"Cabove T (K); l/zis the slope of the thermal destruction curve;
z is the the "C temperature change required to change the thermal destruction
time by a factor of 10 ("C);and Qlo is the ratio of reaction rate at (T + 10)"C to
that at T "C.By rearrangement of Eq. (2)and (3)Labuza and Riboh (1982)have
also shown that:
2.2 Ea
log Qio = (4)
T (T + 10)
The data provided by Meschter (1953)were used to calculate a z value of
25.6"C.According to Lund (1975),this value correspondsto an Ea of about 16
kcal/mol. Tanchev (1983),studying kinetic degradation of anthocyanin color (as
measured spectrophotometrically)in a variety of fruits stored at 10,20, 30 and
40°C, similarly reported an Ea of 16.6 f 0.4 kcal/mol, and z and Qlovalues of
24.5"C and 2.5 f 0.7, respectively. However, Tanchev (1983)reported an Ea
of 23.2 f 1.1 kcal/mol, and z and Qlovalues of 27.5 "C and 2.4 f 0.6,respec-
tively, for anthocyanin color in various fruits heated at processing temperatures
of 78, 88, 98 and 108°C.Using Eq. (2),(3)and (4), the z and Qlo values
reported by Tanchev (1983)correspond to the respective Ea values over the
temperature ranges studied. These results would indicate a definite change in Ea
with temperature, a phenomenon discussed by Labuza and Riboh (1982).
Markakis (1982a)and Mishkin and Saguy (1982)have also reported Ea values of
27 and 25 kcal/mol, respectively, for anthocyanin color degradation at process-
ing temperatures (i.e., > 70"C).
The above results suggest anthocyanin color to be more stable under process-
ing conditions than first thought. This may be the result of condensation reac-
tions, whereby the concentration of polymeric pigments increases with
temperature and storage time. Such polymeric pigments have been shown to
contribute to the coloration of grapes and red wines (Somers 1966b,1967,1971;
224 R.L.JACKMAN, R.Y.YADA, M.A. TUNG AND R.A. SPEERS

McCloskey and Yengoyan 1981) and Concord grape juice (Sastry and Tischer
1952), and their presence serves to protect the anthocyanins in these products. It
is very likely that polymeric pigments degrade differently from monomeric an-
thocyanins, as reflected in the respective Ea values for anthocyanin color
degradation under storage and processing conditions.
The rate of disappearance of monomeric anthocyanins of blueberries corres-
pond to a first order regression equation, i.e., seemed to follow zero order reac-
tion kinetics, when juice was stored at 4 and 10°C (Simard et al. 1982).
However, Simard et al. (1982) reported the rate of pigment disappearance to
correspond to a second order regression equation, i.e., appeared to be first
order, when juices were stored at 27 and 37 "C. This would again suggest that
anthocyanins undergo degradation by different mechanisms with increased pro-
cessing and storage temperatures. The possible influence of various constituents
in the media (Le., other flavonoids, protein, metal ions, etc.) cannot be
overlooked.
The large number of factors affecting anthocyanin degradation in natural
systems has prevented the elucidation of any single degradation scheme for these
pigments. It has been shown that pure anthocyanins, in the absence of added
compounds other than those acids and buffers used to adjust the pH, degrade by
a first order reaction mechanism (Adams 1973b). At least two pathways for an-
thocyanin degradation are possible. According to Markakis et al. (1957), the
heterocyclic ring of the colorless pseudobase opens to form a colorless chalcone
before hydrolysis of the glycosidic bond occurs (Fig. 8 (a)). This mechanism
was proposed in response to the lack of detectable anthocyanidin during the
breakdown of purified pelargonidin-3-glucoside (Markakis et al. 1957). The
degradation has been shown to be temperature dependent.
Adams (1972, 1973a,b) offered an alternative mechanism. Anthocyanins
heated in the pH range of 2 to 4 first undergo glycosidic hydrolysis followed by
conversion of anthocyanidin to a chalcone, which subsequently yields an
adiketone (Fig. 8(b)). Adams (1972, 1973b) contended that anthocyanidins
were generally unstable in the pH range 2 to 4, thus explaining why they had not
been detected under these conditions previously. The rate of formation of free
sugars was shown to closely follow the rate of disappearance of red color, pro-
viding evidence that glycosidic hydrolysis of anthocyanins was the main cause of
red color loss (Adams 1972, 1973b). It was assumed that further degradation of
the primary breakdown products, formed by either pathway proposed, led to
brown colored products (Lukton et al. 1956; Markakis et al. 1957; Erlandson
and Wrolstad 1972; Abers and Wrolstad 1979).
The mechanism of anthocyaninbreakdown, as mediated by increased process-
ing and storage temperatures, has been shown to be dependent on the nature of
the anthocyanin. Hrazdina (1971) identified a coumarin diglycoside as a com-
mon degradation product of malvidin-, cyanidin-, delphinidin-, petunidin- and
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 225

HO /

/- OH
OGly - H O Q
OH
Y H OGly
O W
OH
OGly (4

Gly * (b)

OH

OH

FIG. 8. MECHANISMS OF ANTHOCYANIN DEGRADATION AS PROPOSED BY


(a) Markakis et al. (1957),and (b) Adams (1972;1973a,b).

peonidin-3,5diglycosidesin the pH range 3 to 7. Temperature increases ac-


celerated the overall reaction (Hrazdina 1971). However, as previously noted,
anthocyanidm-3-glycosidesdid not form coumarin derivatives (Hrazdina and
Franzese 1974).
Markakis (1982a), supported by data presented by Brouillard (1982) and
Timberlake (1980), noted the importance of an equilibration time between the
chalcone and flavylium structures. The equilibrium reactions among the antho-
cyanic forms tend to be endothermic from left to right in the following scheme
(Brouillard 1982):

blue red colorless


quinoidal + flavylium A carbinol A
7
colorless
base cation base chalcone

The equilibrium is shifted toward the chalcone upon heating. However, rever-
sion back to the flavylium cation has been shown to be relatively slow. The red
flavylium cation has usually been taken as a measure of total anthocyanin con-
centration in quantitative analyses. Therefore, as Markakis (1982a) suggested
with reference to previous degradation studies, if insufficient time had been
given for the chalcone to flavylium transformation, the results of these studies
could have been inaccurate.

Enzymatic Degradation
There are a number of enzymes, endogenous in most plant tissues, which have
been implicated in the decoloration of anthocyanins. Generally, these have been
identified as either glycosidases (Huang 1955; Forsyth and Quensel 1957) or
226 R.L. JACKMAN, R.Y.YADA,M.A. TUNG AND R.A. SPEERS

phenolases (Peng and Markakis 1963), although Grommeck and Markakis


(1964) have also reported a peroxidase-catalyzed anthocyanin degradation.
Huang (1955, 1956a) investigated the decolorizing action of several crude
fungal enzyme preparations on solutions of pure anthocyanin within a pH range
of 3.0to 4.5. Enzymatic decoloration was attributed to the action of glycosidases
(anthocyanases) which hydrolyzed the anthocyanin to anthocyanidin and sugar,
with the destabilized aglycone spontaneously degrading to colorless derivatives
(Huang 1955, 1956a; Forsyth and Quesnel 1957). The transformation of
liberated aglycone to colorless derivatives and the enzymatic deglycosylationof
anthocyanins by fungal anthocyanase was found to follow first order kinetics at
pH 3.95 and 30°C (Huang 1956b).
A fungal anthocyanase preparation was used by Yang and Steele (1958) to
remove excess anthocyanin from concentrated products (blackberry jams and
jellies) where pigmentation was too dark and unattractive. Huang (1955) had
suggested the use of such enzyme preparations in the manufacture of white wines
from mature red grapes. Wagenknecht et al. (1960) attributed the loss of red col-
or associated with scald in sour cherries to endogenous anthocyanase activity.
Harborne (1%5) utilized fl-glucosidase and anthocyanase preparations to
characterize various flavonoid glycosides. The rate of deglycosylation has been
claimed to be useful in distinguishing 0- from C-glycosides and in characteriz-
ing flavonoids containing more than one sugar moiety (Harborne 1965).
Tanchev et al. (1%9) reported the destruction of anthocyanins in raspberry
and strawberryjuices which had been treated with several commercial pectolytic
enzyme preparations. The observed loss of color in these juices may have been
the result of contaminating glycosidase (or phenolase) activity present in the
commercial preparations. In most cases secondary activity in commercial en-
zyme preparations is desirable, and costs of purification often preclude the total
removal of such activity.
Phenolases are virtually ubiquitous in the plant kingdom, catalyzing the oxida-
tion of phenolic compounds such as catechol and other o-dihydroxyphenols to
oquinones, which may subsequently (a) react with one another to form brown
colored, higher molecular weight polymers; (b) react with amino acids and/or
proteins to form macromolecular complexes which are also brown in color; or
(c) oxidize compounds of lower oxidation-reduction potentials (Mathew and
Parpia 1971). Anthocyanins serve as poor substrates for the enzyme (Markakis
1982a). Peng and Markakis (1963) showed that, although the two major antho-
cyanins isolated from red tart cherries proved to be poor substrates for
mushroom phenolase, rapid decolorization occurred in the presence of catechol.
On the bssis of this observation, these authors proposed a scheme of sequential
reactions to explain the effect of mediating phenols (Fig. 9). The catechol is ox-
idized by phenolase to 0-benzoquinonewhich itself then oxidizes anthocyanin to
a colorless product (Peng and Markakis 1963). A similar coupled oxidation
ANTHOCYANINSAS FOOD COLORANTS - A REVIEW 221

mechanism was offered as explanation for the accelerating effect which


chlorogenic acid and catechin had on the degradation of anthocyanins by
phenolases in eggplant (Sakamura and Obata 1963; Sakamura et al. 1965) and
sweet cherries (Pifferi and Cultrera 1974). In a recent study investigating the
role of strawberry phenolase in anthocyanin degradation, Wesche-Ebeling
(1984) suggested that strawberry anthocyanins were destroyed either through
direct oxidation by quinones formed from catechin by endogenous phenolase, or
through copolymerization of the anthocyanins into brown polymeric pigments
(tannins) formed from catechin-quinone polymerization.

Degraded Anthocyanin

02YaY Enzymic Nonenzvmic

FIG.9. COUPLED OXIDATION MECHANISM INVOLVED IN PHENOLASE


DECOLORIZATION OF ANTHOCYANINS
(Peng and Markakis 1%3).

Pifferi and Cultrera (1974) showed ascorbic acid to have a protective effect on
anthocyanins and their associated color by being preferentially oxidized by the
oquinone formed through the enzymatic oxidation of the mediating phenol. An-
thocyanin was not destroyed as long as ascorbic acid was present in the system.
Markakis (1982a) suggested that this was analogous to the inhibition of
phenolase-catalyzed browning where ascorbic acid reduces the oquinone before
it can polymerize.
Pifferi and Cultrera (1974) isolated two predominant phenolases from sweet
cherries which displayed maximum activity toward a catechol substrate at pH
4.2 and 6.5, respectively. However, maximum pigment destruction was observ-
ed at pH 5.6, and not at the pH optimum of either of the isolated phenolase
preparations. According to these authors, the anhydrobase of the anthocyanin
was, therefore, more susceptible to the direct or indirect action of phenolase,
than the flavylium cation. Sakuraba and Ichinose (1982) suggested that the rate
228 R.L. JACKMAN, R.Y. YADA,M A . TUNG AND R.A.SPEERS

of anthocyanin decoloration by phenolases is dependent upon the hydroxylation


pattern of the B-ring, and the extent of glycosylation. In addition to eggplant,
sweet cherries and strawberries, phenolase has been shown to be responsible for
anthocyanin decoloration in sour cherries (Van Buren et d.1960; Shrikhande
1973), grapes (Segal and Segal 1969), and in flowers and leaves of several other
higher plants (Proctor and Creasy 1969; Schumacker and Bastin 1974).
Several methods for inhibiting or destroying anthocyanindegrading enzymes
have been attempted. Siege1 er al. (1971) and Shrikhande (1973) used steam
blanching of sour cherries for 45-60 s prior to freezing to provide minimal an-
thocyanin destruction during subsequent thawing and blending in air. Goodman
and Markakis (1965) inhibited phenolase activity in sour cherry juice by addition
of 30 ppm SOz. They reported that at low pH similar inhibition could be ac-
complished by addition of even lower SO2concentrations. NaZSzO5,dithiothrei-
tol, phenylhydrazine and cysteine were effective inhibitors of phenolase in Con-
cord grape juice (Cash er al. 1976). Sakuraba and Ichinose (1982) reported that
gallotannin deactivated phenolase through a direct condensation mechanism.
Anthocyanin stabilization was further affected by copigmentation, where excess
gallotannin and anthocyanin complexed through hydrophobic interaction
(Sakuraba and Ichinose 1982).

STABILIZATION OF ANTHOCYANINS
AND THEIR USE As FOOD COLORS
Color has been regarded as one of the most important quality attributes of food
products often forming the basis of consumer acceptability (Clydesdale 1978).
With this in mind, food manufacturers often require that artificial dyes and/or
natural colors be added to their products to ensure desirable/acceptable ap-
pearance throughout their shelf-life. Concern over the toxicological safety of the
synthetic dyes used in the food industry has led to much interest in the develop-
ment of stable, natural pigments for use as food ingredients. Many of the syn-
thetic colorants on the market have a natural counterpart which may be used in
their place without sacrificing product quality. In addition to green and blue,
however, the color red has no natural pigment which has gained wide acceptance
for use in the food industry.
Anthocyanins are probably the best known of the natural food pigments,
rendering fruits and vegetables red, purple and blue as a direct result of their
presence in these tissues (Shrikhande 1976). However, as stated previously, an-
thocyanins are relatively unstable, complex compounds, a fact which has limited
their wide use as food additives. In addition, their commercial unavailability and
the difficulty and costs involved in their extraction and purification from natural
sources have contributed further to restricting their application as food colorants
(Shrikhande 1976).
The classification of anthocyanins as natural food color additives when ex-
tracted from a "natural " source exempts these compounds from rigorous and
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 229

costly toxicological testing, which synthetic dyes must undergo prior to their
clearance as safe food ingredients. Francis (1977) has regarded this double stan-
dard of safety, one for natural compounds and another for synthetic compounds,
as illogical. However, it has provided much imptus for the development of
alternative, natural coloring agents suitable for addition to a variety of food pro-
ducts. A number of highly pigmented sources have been utilized for extraction
of anthocyanins.
Both a spraydried powder and a concentrated solution containing grape an-
thocyanins are presently being marketed from Italy under the trade name of
Enocolor (formerly Enocianina)(Anonymous 1981). These products are dark
reddish-blue, due largely to high concentrations of malvine chloride, and have
been used in cosmetics, pharmaceuticals, wines and various other beverages.
Lancrenon (1978) claimed that these contained excessive amounts of impurities.
However, these impurities may be removed by passing clarified grape extracts
through resins selective for adsorbing polyphenols (Lancrenon 1978; Codounis
et al. 1983).
Peterson and Jaffe (1969) patented a process by which anthocyanins were ex-
tracted from grape pomace with an alcoholic solvent containing 200-2000 ppm
SOz. These researchers found that the addition of SO2to the extraction solvent
resulted in an increased pigment yield. Palamidis and Markakis (1975)
demonstrated that, besides containing anthocyanin of higher purity than a hot
water extract, anthocyanin extracted from grape wine pomace with 500 ppm SOz
was more stable as a colorant for a carbonated beverage. Philip (1974) used an
alcoholic solvent containing 0.1-1 .O% tartaric acid to extract anthocyaninsfrom
grape skins or pomace. Calvi and Francis (1978) recovered anthocyaninpigment
from Concord (Vitis labncscu) grapejuice lees by traditional acid-alcoholextrac-
tion of the filter-press cake. Main et ul. (1978) succeeded in spraydrying ex-
tracts from Concord grape juice sludge, and Clydesdale et al. (1978)
demonstrated their potential for use in dry pack food products such as powdered
beverage mix and gelatin desserts.
A recent United States patent was granted to Langston (1985) for the produc-
tion of a highly ‘colored’ monomeric anthocyanin pigment from grape pomace.
The pigment extract, devoid of sugars, organic acids, polymerized anthocyanin
and other water soluble material, was obtained using an aqueous extraction sol-
vent containing bisulfite, followed by chromatographic separationon a nonionic
adsorbant. Vitis spp. extracts contain relatively high concentrations of malvidin-
diglycosides (Koeppen and Basson 1966; Singleton and Esau 1969). Hrazdina et
ul. (1970) have shown these anthocyaninsto be more stable to oxidation and heat
than other anthocyanins which are commonly found in extracts of other fruits.
Chiriboga and Francis (1970) used cranberry (Vuccinium macrucurpun)
pomace as a source for anthocyanins, with the alcoholic extract subsequently
purified using ion-exchange resins. The freezedried product was suitable for
fortifying the color of pale cranberry juice cocktail (Chiriboga and Francis
230 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

1973). Main et al. (1978) successfully spraydried the purified extract from
cranberry press-cake, but found its application in beverage mix and gelatin
dessert dry pack foods to be limited by an astringent flavor which it imparted to
these products (Clydesdale et al. 1979a). Approximately 40% of the red antho-
cyanin pigments of cranberries, monogalactosides and monoarabinosides of
cyanidin and peonidin (Zapsalis and Francis 1%5), remain in the press cake
following extraction (Staples and Francis 1%8).
Esselen and Sammy (1973, 1975) prepared an aqueous extract of the antho-
cyanins of roselle (Hibiscussub&r@) calyces. The major pigments in roselle
are cyanidin and delphinidin derivatives (Du and Francis 1973) which produce a
hue similar to that of amaranth (Red No. 2)(Esselen and Sammy 1973). A con-
centrate of the aqueous roselle extract was very stable in frozen storage, and
when used in jams and jellies, pigmentation and flavor were well preserved at 35
and 84°F over the course of six months storage (Esselen and Sammy 1973,
1975). Clydesdale er al. (1979b) prepared a spraydried powdered pigment ex-
tract from a roselle liquid concentrate. They found, however, that despite
displaying good stability in reconstituted beverage mix and gelatin dessert dry
pack foods, the roselle powder imparted an unacceptable flavor to the products.
The leaves and fruit skins of cherry-plum (Pncnus cerusiferu), an ornamental
plant, have been investigated as a source for anthocyanins (Baker et ul. 1974a,
b). The cherry-plum anthocyanins, which include the cyanidin and peonidin
3-glucosides and 3-rutinosides (Tanchev and Vasilev 1973), were extracted us-
ing acidified ethanol. These were shown by Baker et al. (1974b) to have accep-
table organoleptic properties and to be nonmutagenic.
Blueberries (Vuccinium spp.) are a very rich source of anthocyanins, contain-
ing malvidin, delphinidin and petunidin 3-galactosides and 3-arabinosides.
Although too expensive to be grown solely for pigment extraction, waste pro-
ducts from frozen blueberry operations may provide an economically feasible
source of natural pigment (Shewfelt and Ahmed 1977). Red cabbage has also
been suggested as a potential commercial source of anthocyanins, containing
cyanidin-3,5-diglucosideand cyanidin-3-sophoroside-5-glucosideacylated with
sinapic acid (Tanchev and Timberlake 1969). Shewfelt and Ahmed (1978) ex-
tracted anthocyanins from blueberries and red cabbage using a SO2 extraction
procedure and a methanol extractionhon exchange purification procedure. The
freeze-dried extracts from both procedures showed greater color stability in dry
soft drink mixes than in reconstituted beverages. This is in agreement with
Erlandson and Wrolstad (1972) who reported strawberry anthocyanins to be
relatively stable in a low moisture environment, i.e., powder as opposed to
reconstituted beverage. Shewfelt and Ahmed (1977) found methanolic extracts
of red cabbage to have greater purity and tinctorial strength than anthocyanins
extracted using SOz. They suggested that methanol-extracted powders could
provide possible replacements for the red synthetic colorants (Red No. 2 and
ANTHOCYANINS AS FOOD COLORANTS - A REVEW 23 1

Red No.40)if pigment losses were minimized in the procedure and a protective
agent added to retard ascorbic acid degradation.
The tropical red berry known as the "mimclefirtit"(Synsepulum dulciJicwn)
contains a taste modifier that has been investigated as a potential sweetener;
however, a byproduct of taste modifier production is anthocyanin pigment
(Buckmire and Francis 1976). Buckmire and Francis (1978) prepared a pigment
extract from the miracle fruit using the procedure of Chiriboga and Francis
(1970), and found it imparted an orange-red color to a carbonated beverage. The
anthocyanin stability was comparable to that of anthocyanins from other
sources. Cyanidin-3-glycosides have been reported as the major anthocyanin
present in miracle fruit extracts (Buckmire and Francis 1976).
Francis (1975) has suggested that berries of Vibernwn dentutzun, which con-
tain anthocyanin pigments amounting to about 1% of the berry fresh weight,
could be used as a commercial source of anthocyanins. The major pigments of
Vibemum dentuzum berries are all cyanidin derivatives, and therefore, extracts
tend to be orange-red in color (Francis 1975).
Martinez and Valle (1981) investigated the use of duhat (Sysyium cumini
Linn.) anthocyanins as a food colorant. Characterization of duhat fruit antho-
cyanins has not been reported. A crude anthocyanin extract of duhat fruits had
acceptable color properties when compared with red synthetic colorants, but a
detectable after-taste was imparted to colored fruit drinks (Martinez and Valle
1981).
Bilberries (Vucciniwn myrtillus) of the Central Massif have been used by a
factory in France to manufacture high grade anthocyanin used for phar-
maceutical purposes (Anonymous 1975). Although detailed information regar-
ding the identity of anthocyanins of this fruit is lacking (Timberlake and Bridle
1982b), Suomalainen and Keranen (1961) noted the presence of glucose and
arabinose derivatives of cyanidin, delphinidin, petunidin and malvidin.
Lowry and Chew (1974) reported the use of anthocyanin extracts from the
dried flowers of Clitoreu ternutiu as a colorant for Malaysian ricecakes. Supris-
ingly, the Clitoreu flower contained only glycosides of delphinidin, the least
stable of the common anthocyanidins (Lowry and Chew 1974). Stability of the
anthocyanin extract was thought to be connected with adsorptionof the pigments
onto the starch gel of the glutinous rice (Lowry and Chew 1974).
Markakis (1982b) mentioned having seen liquid and spraydried anthocyanin
preparations originating from olive skins in Oreece. Codounis et ul. (1983)
described a process by which anthocyanin pigments from the effluents of olive-
oil extracting plants were extracted and purified. The anthocyanins of the ripe
olive fruit have been identified as glycosides of cyanidin and peonidin, often
acylated with caffeic acid, rendering the fruit dark blue or purple in color
(Timberlake and Bridle 1982b).
A Japanese patent has been obtained (Kikuchi et ul. 1977) for a process by
232 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

which a spraydried anthocyanin+ontaining food colorant was obtained from


purple corn (Maiz morado). Cyanidin-3-glucosidehas been identified as the ma-
jor anthocyanin in the cobs and kernels of purple corn (Timberlake and Bridle
1982b).
A U.S.patent was applied for by Asen et al. (1978) for the use of an antho-
cyanin, peonidin-3-(dicaffeyl-sophoroside)-5-glucoside, which produces a wide
range of colors, and is stable in foods and beverages over the pH range of 2.0 to
8.0. The pigment is found in "Heavenly Blue " morning glory flowers (Ipomea
tricolor), and is extracted/purified by means of three column chromatographic
steps followed by ethyl ether precipitation (Asen et al. 1977).
Timberlake (1980) stated that the anthocyanin preparations extracted from the
various sources mentioned were likely of varied quality, containing partially
degraded, condensed and/or polymerized anthocyanins, in addition to various
tannins and copigments. Given the nature of these preparations, the pigment ex-
tracts would be expected to suffer the same fate as the original pigments of the
foodstuff (Timberlake 1980). The addition of anthocyanin extracts to food
systems may encourage reactions with endogenous constituents, quite possibly
resulting in a product of reduced quality. However, by carefully controlling pro-
cessing and storage conditions, the high quality of products containing naturally
occurring anthocyanin pigments may be maintained, thereby providing the con-
sumer with pleasing, colorful products. A knowledge of the behavior of these
pigments under a wide variety of conditions and in a variety of food systems is
required if commercially available extracts are to be used to advantage.
A number of synthetic flavylium salts have been suggested as potential color
additives (Jurd 1964c; Timberlake and Bridle 1%8). Jurd (1964c) proposed the
use of synthetic benzopyrylium dyes, structurally related to natural anthocyanin
pigments, as replacements for some of the common "coal tar " dyes. Four yellow
and two red benzopyrylium dyes were claimed by Jurd (19%) to be easier and
less costly to synthesize than the "coal tar " dyes. They displayed intense colora-
tion and were stable in acidic media. However, their stability to SOz and ascor-
bic acid was low. Timberlake and Bridle (1968) improved the stability of the
synthetic flavylium salts of Jurd (19%) to SO2 and ascorbic acid, by
substituting a methyl or phenyl group at position 4 of the flavylium molecule.
The synthetic flavylia all require extensive toxicological testing prior to their
clearance as food color additives. Thus, the use of natural pigments as food col-
orants is desirable, despite their shortcomings.
According to Hrazdina (1981), anthocyanins may be successfully used in high
acid foods such as soft drinks, jams and jellies, and in red wines where they play
an important role in the ageing process. "Beyond this, we will have to live with
their limitations and accept the deterioration of color upon processing and
storage of the anthocyanin-containingfood products " (Hrazdina 1981). The pro-
curement of intensely colored fruits and vegetables is necessary to ensure that if
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 233

some of the pigment were destroyed through processing or storage, a sufficient


portion would remain to impart the characteristic color of the product (Markakis
1974; Timberlake and Bridle 1982a).
Few substanceshave been found which serve as practical stabilizing agents for
the anthocyanins. The use of sulfite as a stabilizing agent is limited not only by
its decolorizing effects (which may or may not be reversible), but also by its
characteristic undesirable odor and flavor which may be detected at very low
levels of use in certain products, i.e., at SO ppm in berry juices (Jackman et al.
1984). Cysteine inhibited anthocyanin degradation in Concord grapejuice below
a temperature of 75 "C, and in addition, protected amrbic acid from oxidation
(Skalski and Sistrunk 1973). Cysteine was presumed to act as a reducing agent,
as well as an enzyme inhibitor for polyphenol oxidas (Skalski and Sistrunk
1973). In an investigationinto the effects of 19 differentadditives on color reten-
tion in strawberryjuice and buffered pigment solutions, Markakis et d.(1957)
found only thiourea, propyl gallate and quercetin to show beneficial effects.
Recently, Maccarone et al. (1985) improved the stability of the red color of
pigmented orange fruit juice by pasteurizing with microwaves and adding tar-
taric acid and glutathione which acted as mildiy acidic and antioxidant agents,
respectively. The addition of phenolic compounds, such as rutin and caffeic
acid, also markedly stabilized color, presumably by means of copigmentation
(Maccarone et al. 1985).
It is important that investigations into the chemistry of anthocyanins continue
in order that we may exploit natural media in which these pigments are modified
to advantage (Timberlake 1980). This will increase our knowledge of the
behavior of endogenous pigments, and extracts obtained from natural sources
which may then be manipulated in such a way as to produce pigments with
desired properties. As Timberlake (1980) has stated, "there is still a lot to learn
about the mechanism of anthocyanin color augmentation and stabilization, but if
sufficient effort is devoted to unravelling these effects, there would seem to be
every prospect of developing pigments of improved characteristics, suitable as
color additives. "

REFERENCES
ABERS, J.E. and WROLSTAD, R.E. 1979. Causative factors of color
deterioration in strawberry preserves during processing and storage. J. Food
Sci. 44,75-78, 81.
ADAMS, J.B. 1972. Changes in the polyphenols of red fruits during processing
- the kinetics and mechanism of anthocyanin degradation. Campden Food
Pres. Res. Assoc. Tech. Bull. p. 22, Chipping Campden, Gloucestershire,
Eng .
234 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

ADAMS,J.B. 1973a. Colour stability of red fruits. Food Manuf., Feb., 19-20,
41.
ADAMS, J.B. 1973b. Thermal degradation of anthocyanins with particular
reference to the 3-glycosides of cyanidin. I. In acidified aqueous solution at
100OC. J. Sci. Food Agric. 24,747-762.
ADAMS,J.B. and ONGLEY, M.H. 1973. The degradation of anthocyanins in
canned strawberries. I. The effect of various processing parameters on the
retention of pelargonidin-3-glucoside.J. Food Technol. 8, 139-145.
ALBACH, R.F., WEBB, A.D. and KEPNER, R.E. 1965. Structures of
acylated anthocyanin pigments in Vitis vinifera variety Tinta Pinheira. II.
Position of acylation. J. Food Sci. 30, 620-626.
ANDERSON, D.W., GUEFFROY, D.E., WEBB. A.D. and KEPNER, R.E.
1980. Identification of acetic acid as an acylating agent of anthocyanin
pigments in grapes. Phytochem. 9, 1579-1583.
ANON. 1975. Vignes et vins. Numero special consacre aux anthocyanes du
raisin et du vin, 25. Cited by Timberlake, C.F. 1980. Anthocyanins-
occurrence, extraction and chemistry. Food Chem. 5, 69-80.
ANON. 1981. Enocolor. ?‘he importance of the natural colour, Reggiana Anto-
ciani, Industria Coloranti Naturali, Italy.
ASEN, S. 1974. Factors affecting flower colour. Acta Hortic. 41, 57-68.
ASEN, S . 1976. Known factors responsible for infinite flower color variations.
Acta Hortic. 63, 217-223.
ASEN, S. and JURD, L. 1967. The constitution of crystalline, blue cornflower
pigment. Phytochem. 6, 577-584.
ASEN, S.,NORRIS, K.H. and STEWART, R.N. 1969. Absorption spectra and
color of aluminum-cyanidin 3-glucoside complexes as influenced by pH.
Phytochem. 8, 653-659.
ASEN, S.,NORRIS, K.H. and STEWART, R.N. 1971. Effect of pH and con-
centration of the anthocyanin-flavonol co-pigment complex on the color of
‘Better Times’ roses. J. Am. SOC. Hort. Sci. 96, 770-773.
ASEN, S.,STEWART, R.N. and NORRIS, K.H. 1972. Co-pigmentation of an-
thocyanins in plant tissues and its effect on color. Phytochem. 11, 1139-1 144.
ASEN, S.,STEWART, R.N. and NORRIS, K.H. 1975. Anthocyanin, flavonol
copigments, and pH responsible for Larkspur flower color. Phytochem. 14,
2677-2682.
ASEN, S., STEWART, R.N. and NORRIS, K.H. 1977. Anthocyanin and pH
involved in the color of ‘Heavenly Blue’ morning glory. Phytochem. 16,
1118-1 119.
ASEN, S . , STEWART, R.N. and NORRIS, K.H. 1978. Naturally occurring
colorant for food and beverages. US. Pat. Appl. 910,152.
ASEN, S., STEWART, R.N., NORRIS, K.H. and MASSIE, D.R. 1970. A
stable blue nonmetallic co-pigment complex of delphanin and
C-glycosylflavones in Prof. Blaauw Iris. Phytochem. 9, 619-627.
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 235

BAKER, C.H., JOHNSTON, M.R. and BARBER, W.D. 1974a. Cherry-plum


pigment as a natural food colorant: Instrumental evaluation. Food Prod. Dev.
8(3), 83-84, 86-87.
BAKER, C.H., JOHNSON, M.R. and BARKER, W.D. 1974b. Cherry-plum
pigment as a natural red food colorant: Sensory evaluation and mutagenicity
tests. Food Prod. Dev. 8(4), 65-70.
BAYER, F., FINK, A., NETHER, K. and WEGMA", K. 1966. Complex
formation and flower colors. Angew. Chem. Inter. Edit. 5,791-798. Cited by
Asen, S. 1976. Known factors responsible for infinite flower color variations.
Acta Hort. 63, 217-223.
BEATTIE, H.G., WHEELER, K.A. and PEDERSON, C.S. 1943. Changes oc-
curring in fruit juices during storage. Food Res. 8, 395404.
BENDZ, G., MARTENSSON, 0. and NILSSON, E. 1967. Studies of
flavylium compounds. I. Some flavylium compounds and their properties.
Ark. Kemi 27(7), 65-77.
BLOOM, M. and GEISSMAN, T.A. 1973. Malonic acid: The acyl moiety of
the Mimulus luteus anthocyanin. Phytochem. 12, 2005-2006.
BROUILLARD, R. 1982. Chemical structure of anthocyanins. In Anthocyanins
as Food Colors, (P.Markakis, ed.) pp. 1-40, Academic Press, New York.
BROUILLARD, R. and DELAPORTE, B. 1977. Chemistry of anthocyanin
pigments. 2. Kinetic and thermodynamic study of proton transfer, hydration,
and tautomeric reactions of malvidin 3-glucoside. J. Am. Chem. Soc. 99,
8461-8468.
BROUILLARD, R. and DELAPORTE, B. 1978. Etude par relaxation chimique
des equilibres de transfert de proton, d'hydration et de tautomerie prototropi-
que de l'anthocyane majoritaire de Vitis vinifem, la glucoside-3malvidine. In
Protons and Ions Involved in Fast Dynamic Phenomena, (P.Lazlo, ed.) pp.
403-412, Elsevier Scientific Publishing Co., Amsterdam.
BROUILLARD, R. and DUBOIS, J.E. 1977. Mechanism of the structural
transformations of anthocyanins in acidic media. J. Am. Chem. Soc. 99,
1359-1364.
BROUILLARD, R. and EL HAGE CHAHINE, J.M. 198Oa. Chemistry of an-
thocyanin pigments. 5 . Stability of colored forms and reactivity with sulfur
dioxide. Bull. Liaison - Groupe Polyphenols 9, 77-88.
BROUILLARD, R. and EL HAGE CHAHINE, J.M. 198Ob. Chemistry of an-
thocyanin pigments. 6. Kinetic and thermodynamic study of hydrogen sulfite
addition to cyanin. Formation of a highly stable Meisenheimer-type adduct
derived from a 2-phenylbenzopyrylium salt. J. Am. Chem. Soc. 202,
5375-5378.
BROUILLARD, R.,DELAPORTE, B. and DUBOIS, J.E. 1978. Chemistry of
anthocyanin pigments. 3. Relaxation amplitudes in pH-jump experiments. J.
Am. Chem. Soc. 100, 6202-6205.
236 R.L.JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

BROUILLARD, R., IACOBUCCI, G.A. and SWEENY, J.G. 1982. Chemistry


of anthocyanin pigments. 9. UV-visible spectrophotometric determination of
the acidity constants of apigeninidin and three related 3-deoxyflavylium salts.
J. Am. Chem. Soc. 104, 7585-7590.
BUCKMIRE, R.E. and FRANCIS, F.J. 1976. Anthacyanins and flavonols of
miracle fruit, Synsepulum dufcgcum, Schum. J. Food Sci. 42, 1363-1365.
BUCKMIRE, R.E. and FRANCIS, F.J. 1978. Pigments of miracle fruit,
Synsepulum dufcgcum, Schum, as potential food colorants. J. Food Sci. 43,
908-91 1.
CALVI, J.P. and FRANCIS, F.J. 1978. Stability of Concord grape (V.
labwcu) anthocyanins in model systems. J. Food Sci. 43, 1448-1456.
CASH, J.N., SISTRUNK, W.A. and STUTTE, C.A. 1976. Characteristics of
Concord grape polyphenoloxidase involved in juice color loss. J. Food Sci.
41, 1398-1402.
CHANDLER, B.V. and CLEGG, K.M. 197Oa. Pink discoloration in canned
pears. I. Role of tin in pigment formation. J. Sci. Food Agric. 21, 315-319.
CHANDLER, B.V. and CLEGG, K.M. 1970b. Pink discoloration in canned
pears. II. Measurement of potential and developed color in pear samples. J.
Sci. Food Agric. 21, 319-323.
CHANDLER, B.V. and CLEGG, K.M. 197Oc. Pink discoloration in canned
pears. III. Inhibition by chemical additives. J. Sci. Food Agric. 21, 323-328.
CHIRIBOGA, C. and FRANCIS, F.J. 1970. An anthocyanin recovery system
from cranberry pomace. J. Am. Soc. Hort. Sci. 95, 233-236.
CHIRIBOGA, C. and FRANCIS, F.J. 1973. Ion exchange purified anthocyanin
pigments as a colorant for cranberry juice cocktail. J. Food Sci. 38,464467.
CLEGG, K.M. and MORTON, A.D. 1968. The phenolic compounds of black-
currant juice and their protective effect on ascorbic acid. J. Food Technol. 3,
277-284.
CLYDESDALE, F.M. 1978. Colorimetry - methodology and applications.
CRC Crit. Rev. Food Sci. Nutr. 10, 243-301.
CLYDESDALE, F.M., MAIN, J.H. and FRANCIS, F.J. 1979a. Cranberry
pigments as colorants for beverages and gelatin desserts. J . Food Protect. 42,
196-201.
CLYDESDALE, F.M., MAIN, J.H. and FRANCIS, F.J. 1979b. Roselle
(Hibiscussubdunflu L.) anthocyanins as colorants for beverages and gelatin
desserts. J. Food Protect. 42, 204-207.
CLYDESDALE, F.M., MAIN, J.H., FRANCIS, F.J. and DAMON JR., R.A.
1978. Concord grape pigments as colorants for beverages and gelatin desserts.
J. Food Sci. 43, 1687-1692, 1697.
CO, H. and MARKAKIS, P. 1%8. Flavonoid compounds in the strawberry
fruit. J. Food Sci. 33, 281-283.
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 231

CODOUNIS, M., KATSABOXAKIS, K. and PAPANICOLAOU, D. 1983.


Progress in extraction and purification of anthocyanin pigments from the ef-
fluents of olive+il extracting plants. In Progress in Food Engineering. Solid
Extruction, Isolation and PuriJication, Texturizution, (C. Cantarelli and C.
Peri, eds.) pp. 567-572, Forster Publishing Ltd., Kusnacht, Switzerland.
COFFEY, D.G., CLYDESDALE, F.M., FRANCIS, F.J. and DAMON JR.,
R.A. 1981. Stability and complexation of cyanidin-3-glucosideand raspberry
juice extract in the presence of selected cations. J. Food Protect. 44,516-523.
DARAVINGAS, G. and CAIN, R.F. 1965. Changes in the anthocyanin pig-
ments of raspberries during processing and storage. J. Food Sci. 3 0 , 4 0 0 4 5 .
DARAVINGAS, G. and CAIN, R.F.1968. Thermal degradation of black rasp-
berry anthocyanin pigments in model systems. J. Food Sci. 33, 138-142.
DEBICKI-POSPISIL, J., LOVRIC, T., TRINAJSTIC, N. and SABLJIC, A.
1983. Anthocyanin degradation in the presence of furfural and 5-hydroxy-
methylfurfural. J. Food Sci. 48, 411416.
DECAREAU, R.V., LIVINGSTON, G.E. and FELLERS, C.R. 1956. Color
changes in strawberry jellies. Food Technol. 10, 125-128.
DU, C.T. and FRANCIS, F.J. 1973. Anthocyanins of roselle (Hibiscus sab-
dariffu L.). J. Food Sci. 38, 810-812.
DU, C.T. and FRANCIS, F.J. 1975. Anthocyanins of garlic (Allium sativum
L.). J. Food Sci. 40,1101-1102.
ERLANDSON, J.A. and WROLSTAD, R.E. 1972. Degradation of antho-
cyanins at limited water concentration. J. Food Sci. 37, 592-595.
ESKIN, N.A.M. 1979. Plant Pigments, Flavors and Textures: The Chemistry
and Biochemistry of Selected Compounds, pp. 2842, Academic Press, New
York.
ESSELEN, W.B. and SAMMY, G.M. 1973. Roselle - a natural red colorant
for foods? Food Prod. Dev. 7(1), 80, 82, 86.
ESSELEN, W.B. and SAMMY, G.M. 1975. Applications for roselle as a red
food colorant. Food Prod. Dev. 9(8), 37-38, 40.
FORSYTH, W.G.C. and QUESNEL, V.C. 1957. Cacao polyphenolic
substances. 4. The anthocyanin pigments. Biochem. J. 65, 177-179.
FRANCIS, F.J. 1975. Anthocyanins as food colors. Food Technol. 29,52,54.
FRANCIS, F.J. 1977. Anthocyanins. In Current Aspects of Food Colorants,
(T.E. Furia, ed.) pp. 19-27, CRC Press, Cleveland, OH.
FRANCIS, F.J. 1982. Analysis of anthocyanins. In Anthocyanins As Food Col-
ors, (P. Markakis, d.)pp. 182-208, Academic Press, New York.
FRANCIS, F.J. 1984. Future trends. In Developments in Food Colors-2, (J.
Walford, ed.) pp. 233-247, Applied Science Publishers, New York.
FRANCIS, F.J. and HARBORNE, J.B. 1966. Anthocyanins of the garden
huckleberry, Solanum guineese. J. Food Sci. 31, 524-528.
238 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

FREEDMAN, L. and FRANCIS, F.J. 1984. Effect of ascorbic acid on color of


jellies. J. Food Sci. 49, 1212-1213.
FULEKI, T. 1967. Development of Quantitative Methodr for Individual Antho-
cyanins in Cranberry and Cranberry Products, pp. 1-286, Ph.D. thesis,
University of Massachusetts, Amherst, MA.
GEISSMAN, T.A. (4.)1962. l?re Chemistry of Flavonoid Compounds, pp.
1-666, The MacMillan Co., New York.
GEISSMAN, T.A., JORGENSON, E.C. and HARBORNE, J.B. 1953. The ef-
fect of aluminum chloride on absorption spectra of anthocyanins. Chem. Ind.,
Dec., 1389.
GOODMAN, L.P. and MARKAKIS, P. 1965. Sulfur dioxide inhibition of an-
thocyanin degradation by phenolase. J. Food Sci. 30, 135-137.
GRISEBACH, H. 1982. Biosynthesis of anthocyanins. In Anthocyanins as
Food Colors, (P. Markairis, ed.) pp. 69-92, Academic Press, New York.
GROMMECK, R. and MARKAKIS, P. 1964. The effect of peroxidase on an-
thocyanin pigments. J. Food Sci. 29, 53-57.
HARBORNE, J.B. 1958a. The chromatographic identification of anthocyanin
pigments. J. Chromatog. 1, 473488.
HARBORNE, J.B. 1958b. Spectral methods of characterizing anthocyanins.
Biochem. J. 70, 22-28.
HARBORNE, J.B. 1959. The chromatography of flavonoid pigments. J.
Chromatog. 2, 581-604.
HARBORNE, J.B. 1963. Distribution of anthocyanins in higher plants. In
Chemical Plant Tawnomy, (T. Swain, ed.) pp. 359-388, Academic Press,
New York.
HARBORNE, J.B. 1964. Plant polyphenols - XI.The structure of acylated an-
thocyanins. Phytochem. 3, 151-160.
HARBORNE, J.B. 1965. Plant polyphenols - XIV. Characterization of
flavonoid glycosides by acidic and enzymic hydrolyses. Phytochem. 4,
107- 120.
HARBORNE, J.B. 1967. Comparative Biochemistry of l?re Flavonoids, pp.
1-36, Academic Press, London, England.
HARBORNE, J.B. 1973. Phytochem’cal Methods, pp. 38-88, Chapman &
Hall Ltd., London, England.
HARBORNE, J.B. 1979. Variation in and functional significance of phenolic
conjugation in plants. In Recent Advances in Phytochemistry, VoZ. 12.
Biochemistry of Plant Phenolics, (T. Swain, J.B. Harborne and C.F. Van
Sumere, 4 s . ) pp. 457-474, Plenum Press, New York.
HARPER, K.A. 1968. Structural changes of flavylium salts. IV.Polarographic
and spectrometric examination of pelargonidin chloride. Aust. J. Chem. 21,
221-227.
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 239

HARPER, K.A. and CHANDLER, B.V. 1%7. Structural changes of flavylium


salts. Aust. J. Chem. 20,745-756.
HARPER, K.A., MORTON, A.D. and ROLFE, E.J. 1969.The phenolic com-
pounds of blackcurrent juice and their protective effect on ascorbic acid. III.
The mechanism of ascorbic acid oxidation and its inhibition by flavonoids. J.
Food T~hn01.4, 255-267.
HAYASHI, K. and TAKEDA, K. 1970.Studies on anthocyanins. LXII. Further
purification and component analysis of commelinin showing the presence of
magnesium in the blue complex molecule. Roc. Jap. Acad. 46, 535-540.
HODGE,J .E . 1953.Chemistry of browning reactions in model systems. J . Agr .
Food Chem. 2, 928-943.
HOOPER, F.C. and AYRES, A.D. 1950.The enzymatic degradation of ascor-
bic acid. Part I. - The inhibition of the enzymatic oxidation of ascorbic acid
by substances occurring in blackcurrents. J. Sci. Food Agric. 1, 5-8.
HRAZDINA, G. 1970. Oxidation of the anthocyanidin-3,5diglucosideswith
H,Oz: The structure of malvone. Phytochem. 9, 1647-1652.
HRAZDINA, G. 1971. Reactions of the anthocyanidin-3,5diglucosides:For-
mation of 3,5di-(O-~-D-glucosyl)-7-hydroxycoumarin.Phytochem. 10,
1125-1130.
HRAZDINA, G. 1974.Reactions of anthocyanins in food products. Lebensm.
-Wiss. u. Technol. 7(4), 103-108.
HRAZDINA, G. 1981.Anthocyaninsand their role in food products. Lebensm.
-Wiss. u. Technol. 24(6), 283-286.
HRAZDINA, G.,BORZELL, A.J. and ROBINSON, W.B. 1970. Studies on
the stability of the anthocyanidin-3,5diglucosides.Am. J. Enol. Vitic. 21,
201-204.
HRAZDINA, G. and FRANZESE, A.J. 1974.Oxidation products of acylated
anthocyanins under acidic and neutral conditions. Phytochem. 13,231-234.
HUANG, H.T. 1955. Decolorization of anthocyanins by fungal enzymes. J.
Agr. Food Chem. 3, 141-146.
HUANG, H.T. 19%. Enzymatic identification of the anthocyanin pigment of
blackberry. Nature 277,39.
HUANG, H.T. 1956b.The kinetics of the decolorization of anthocyanins by
fungal “anthocyanase”.J. Am. Chem. Soc. 78,2390-2393.
INGOLD, C.K. 1969.Structure and Mechanism in Organic Chemistry, p. 847,
Cornell University Press, Ithaca, New York.
JACKMAN, R.L., SPEERS, R.A. and TUNG, M.A. 1984.Unpublished data.
University of British Columbia, Vancouver, BC.
JURD, L. 1%3. Anthocyanins and related compounds. I. Structural transforma-
tions of flavylium salts in acidic solutions. J. Org. Chem. 28,987-991.
240 R.L. JACKMAN, R.Y. YADA, M.A. TUNGAhDR.A. SPEERS

JURD, L. 1964a. Reactions involved in sulfite bleaching of anthocyanins. J.


Food Sci. 29, 16-19.
JURD, L. 1964b. Anthocyanins and related compounds. III. Oxidation of
substituted flavylium salts to 2-phenylbenzofurans. J. Org. Chem. 29,
2602-2605.
JURD, L. 1964c. Some benzopyrylium compounds potentially useful as color
additives for fruit drinks and juices. Food Technol. 18, 559-561.
JURD, L. 1966. Anthocyanins and related compounds X. Peroxide oxidation
products of 3-alkylflavylium salts. Tetrahedron 22, 2913-2921.
JURD, L. 1967. Anthocyanins and related compounds XI. Catechin-flavylium
salt condensation reactions. Tetrahedron 23, 1057-1064.
JURD, L. 1972a. Some advances in the chemistry of anthocyanin-type plant
pigments. In Ihe Chemistry of Pkant Pigments, (C.O. Chichester, ed.) pp.
123-142, Academic Press, New York.
JURD, L. 1972b. Anthocyanins and related compounds XVI. The dimerization
of flavylium salts in aqueous solutions. Tetrahedron 28, 493-504.
JURD, L. and ASEN, S. 1966. The formation of metal and "copigment " com-
plexes of cyanidin 3-glucoside. Phytochem. 5, 1263-1271.
JURD, L. and OEISSMAN, T.A. 1963. Anthocyanins and related compounds.
II. Structural transformations of some anhydro bases. J. Org. Chem. 28,
2394-2397.
JURD, L. and WAISS JR., A.C. 1965. Anthocyanins and related compounds
VI. Flavylium salt-phloroglucinol condensation products. Tetrahedron 21,
1471-1483.
KEITH, E.S. and POWERS, J.J. 1965. Polarographic measurement and ther-
mal decomposition of anthocyanin compounds. J. Agr. Food Chem. 13,
577-579.
KIKUCHI, K., CHIBA, A., MIYAKE, K., NAKAI, T. and TOKUDA, M.
1977. Anthocyanin food coloring agent from purple corn. Japan. Kokai 77,
130, 824.
KOEPPEN, B.H. and BASSON, D.S. 1966. The anthocyanin pigments of
Barlinka grapes. Phytochem. 5, 183-187.
KURATA, T. and SAKURAI, Y. 1967. Degradation of L-ascorbic acid and
mechanism of non-enzymic browning reaction. Part II. Non-oxidative
degradation of L-ascorbic acid including the formation of 3deoxy-L-
pentosone. Agr. Biol. Chem. 31, 170-176.
LABUZA, T.P. and RIBOH, D. 1982. Theory and application of Arrhenius
kinetics to the prediction of nutrient losses in foods. Food Technol. 36(10),
66-74.
LANCRENON, X. 1978. Recent trends in the manufacturingof natural red col-
ors. Process Biochem. 16(10), 16.
ANTHOCYANINS AS FOOD COLORANTS -A REVIEW 241

LANGSTON, M.S.K. 1985. Anthocyanin colorant from grape pomace. U.S.


Patent 4,500,556.
LOWRY, J.B. and CHEW, L. 1974.On the use of extracted anthocyanin as a
food dye. E o n . Bot. 28, 61-62.
LUKTON, A,, CHICHESTER, C.O. and MACKINNEY, G. 1956. The
breakdown of strawberry anthocyanin pigment. Food Technol. 10,427-432.
LUND, D.B. 1975. Heat processing. In PrincipZes of Food Science. Part ZZ.
Physical Principles of Food Preservation, (O.R. Fennema, ed.) pp. 31-92,
Marcel Dekker, New York.
MACCARONE, E., MACCARRONE, A. and RAPISARDA, P. 1985. Sta-
bilization of anthocyanins of blood orange fruit juice. J. Food Sci. 50,
901-904.
MACKINNEY, G.,LUKTON, A. and CHICHESTER, C.O. 1955.Strawberry
preserves by a low temperature process. Food Technol. 9, 324-326.
MAIN, J.H., CLYDESDALE, F.M. and FRANCIS, F.J. 1978. Spray drying
anthocyanin concentrates for use as food colorants. J. Food Sci. 43,
1693-1694, 1697.
MARKAKIS, P. 1960. Zone electrophoresis of anthocyanins. Nature 187,
1092-1093.
MARKAKIS, P. 1974. Anthocyanins and their stability in foods. CRC Crit.
Rev. Food Technol. 4, 437-456.
MARKAKIS, P. 1982a.Stability of anthocyanins in foods. In Anthocyanins As
Food Colors, (P. Markakis, ed.) pp. 163-181,Academic Press, New York.
MARKAKIS, P. 1982b. Anthocyanins as food additives. In Anthocyanins As
Food Colors, (P. Markakis, ed.) pp. 245-253,Academic Press, New York.
MARKAKIS, P., LMNGSTON, G.E. and FELLERS, C.R. 1957. Quan-
titative aspects of strawberry pigment degradation. Food Res. 22, 117-129.
MARTINEZ, S.B. and DEL VALLE, M.J. 1981. Storage stability and sensory
quality of duhat (Sysyim cumini Linn.) anthocyanins as a food colorant. UP
Home &on. J. 9, 7-10.
MATHEW, A.G. and PARPIA, H.A.B. 1971.Food browning as a polyphenol
reaction. Adv. Food Res. 19, 75-145.
MCCLOSKEY, L.P. and YENGOYAN, L.S. 1981. Analysis of anthocyanins
in Vitis viniferu wines and red color versus aging by HPLC and spec-
trophotometry. Am. J. Enol. Vitic. 32, 257-261.
MESCHTER, E.E. 1953. Effects of carbohydrates and other factors on
strawberry products. J. Agr. Food Chem. 1 , 574-579.
MISHKIN, M. and SAGUY, I. 1982. Thermal stability of pomegranate juice.
Z. Lebensm. u. Forsch. 175, 410-412.
NAKAYAMA, T.O.M. and POWERS, J.J. 1972.Absorption spectra of antho-
cyanin in vivo. Adv. Food Res., Supp. 3, 193-199.
242 R.L.JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

NEBESKY, E.A., ESSELEN, W.B. JR., MCCONNELL, J.E. W. and


FELLERS, C.R. 1949. Stability of color in fruit juices. Food Res. 14,
26 1-274.
OSAWA, Y. 1982. Copigmentation of anthocyanins. In Anthocyanins As Food
Colors, (P. Markakis, ed.) pp. 41-86, Academic Press, New York.
PALAMIDIS, N. and MARKAKIS, P. 1975. Stability of grape anthocyanin in a
carbonated beverage. J. Food Sci. 40, 1047-1049.
PENG, C.Y. and MARKAKIS, P. 1963. Effect of phenolase on anthocyanins.
Nature 199, 597-598.
PETERSON, R.J. and JAFFE, E.B. 1969. Berry and fruit treatment process for
pigment and flavor extraction. U.S. Patent 3,484,254.
PHILIP, T. 1974. An anthocyanin recovery system from grape wastes. J. Food
Sci. 39, 859.
PIFFERI, P.G. and CULTRERA, R. 1974. Enzymatic degradation of antho-
cyanins: The role of sweet cherry polyphenol oxidase. J. Food Sci. 39,
786-791.
POEI-LANGSTON, M.S.and WROLSTAD, R.E. 1981. Color degradation in
an ascorbic acid-anthocyanin-flavonol model system. J. Food Sci. 46,
1218-1236.
POLESELLO, A. and BONZINI, C. 1977. Observations on pigments of sweet
cherries and on pigment stability during frozen storage. I. Anthocyanin com-
position. Confructa 22, 170-175.
PONTING, J.D., SANSHUCK, D.W. and BREKKE, J.E. 1960. Color
measurement and deterioration in grape and berry juices and concentrates.
Food Res. 25, 471478.
PRATT, D.E., BALKCOM, C .M., POWERS, J. J. and MILLS, L.W. 1954.
Interaction of ascorbic acid, riboflavin, and anthocyanin pigments. J. Agr.
Food Chem. 2, 367-372.
PROCTOR, J.T.A. and CREASY, L.L. 1969. An anthocyanindecolorizing
system in florets of Cichon'm intybus. Phytochem. 8, 1401-1403.
PYYSALO, H. 1973. The dependence of the colour of anthocyanins on the pH
of aqueous solutions. An absorptiometric study of complexes of cyanidin-3-
galactoside and delphinidin with iron and tin. Suomen Kemistilehti B 46,
103- 105.
RIBEREAU-GAYON, P. 1982. The anthocyanins of grapes and wines. In An-
thocyaninsAs Food Colors, (P. Markakis, ed.) pp. 209-244, Academic Press,
New York.
SAITO, N., MITSUI,S. andHAYASHI, K. 1%1. Anthocyanins. XXXV. Fur-
ther analysis of organic and inorganic components in crystalline protocyanin.
Proc. Japan Acad. 37, 485490.
ANTHOCYANLNS AS FOOD COLORANTS - A ReVIEW 243

SAITO, N., OSAWA, Y. and HAYASHI, K. 1972. Isolation of a blue-violet


pigment from the flowers of Platycoabn grandiflorum. Bot. Mag. Tokyo 85,
105-1 10.
SAKAMURA, S. and OBATA, Y. 1963. Anthocyanase and anthocyanins in
eggplant, Solanum meZongena L. Part II. Isolation and identification of
chlorogenic acid and related compounds from eggplant. Agr. Biol. Chem.
27(2), 121-127.
SAKAMURA, S., WATANABE, S. and OBATA, T. 1965. Anthocyanase and
anthocyanins occurring in the eggplant (Solanurn molengena). III. Oxidative
decolorization of the anthocyanin by polyphenol oxidase. Agr. Biol. Chem.
29(3), 181-190.
SAKURABA, H. and ICHINOSE, H. 1982. Deactivation of polyphenol oxidase
and stabilization of anthocyanins with tannins. Agric. Chem. Soc. Jap. 56,
5 17-524.
SASTRY, L.V.L. and TISCHER, R.G. 1952. Stability of the anthocyanin
pigments in Concord grade juice. Food Technol. 6, 264-268.
SCHEFFELDT, P. and HRAZDINA, 0. 1978. Co-pigmentation of antho-
cyanins under physiological conditions. J. Food Sci. 43, 517-520.
SCHUMACKER, R. and BASTIN, M. 1974. Anthocyanin decolorizing system
in leaves of Perilla naukinensis. Bull. Soc. R. Sci. Liege 34, 42. Cited by
Shrikhande, A.J. 1976. Anthocyanins in foods. CRC Crit. Rev. Food Sci.
Nutr. 7, 193-218.
SEGAL, B. and NEGUTZ, G. 1969. Zur thermischen zersetzung von kera-
cyanin. N a h m g 13, 531-535.
SEGAL, B. and ORANESCU, E. 1978. The effect of aluminum(3+) and
tin(2+) on the juice color of anthocyanin rich fruit juices. Bul. Univ. Galati,
Fasc. 6(1), 53-62.
SEGAL, B. and SEGAL, R.M. 1969. Enzymic degradation of grape antho-
cyanins. Rev. Ferment. Ind. Aliment. 24, 22-24.
SHEWFELT, R.L. and AHMED, E.M. 1977. Anthocyanin extracted froh red
cabbage shows promise as coloring for dry beverage mixes. Food Prod. Dev.
11(4), 52, 57-59, 62.
SHEWFELT, R.L. and AHMED, E.M. 1978. Enhancement of powdered soft
drink mixes with anthocyanin extracts. J. Food Sci. 43, 435438.
SHIBATA, M. and YOSHIMATA, K. 1968. On anthocyanins found in the
flowers of larkspur. Kumamoto J. Sci. 9,49. Cited by Shirkhande, A.J. 1976.
Anthocyanins in foods. CRC Crit. Rev. Food Sci. Nutr. 7, 193-218.
SHRIKHANDE, A.J. 1973. The Isolation and Identification of Anthocyanin
and Flavonol Pigments of Sour Cherries and Some Studies On me Antho-
cyanin Stability In Model Systems, Ph.D. thesis, University of Massachusetts,
Amherst, MA.
244 R.L.JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

SHRIKHANDE, A.J. 1976. Anthocyanins in foods. CRC Crit. Rev. Food Sci.
Nutr. 7, 193-218.
SHRIKHANDE,A.J. and FRANCIS, F.J. 1974. Effect of flavonols on ascorbic
acid and anthocyanin stability in model systems. J. Food Sci. 39, 904-906.
SIEGEL, A., MARKAKIS, P. and BEDFORD, C.L. 1971. Stabilizationof an-
thocyanins in frozen tart cherries by blanching. J. Food Sci. 36, 962-963.
SILVERBLATT, E., ROBINSON, A.L. and KING, C.G. 1943. The kinetics
of the reaction between ascorbic acid and oxygen in the presence of copper
ion. J. Am. Chem. Soc. 65, 137-141.
SIMARD, R.E., BOURZEIX, M. and HEREDIA, N. 1982. Factors influen-
cing color degradation in blueberry juice. Lebensm. -Wiss. u. Technol. 25,
177- 180.
SINGLETON, V .L . and ESAU ,P. L . 1969. Phenolic Substunces In Grapes and
Wines and l k i r Signflcance, pp. 1-282, Academic Press, New York.
SISTRUNK, W.A. and CASH, J.N. 1970. The effect of certain chemicals on
the color and polysaccharides of strawberry puree. Food Technol. 24,
169-173.
SKALSKI, C. and SISTRUNK, W.A. 1973. Factors influencing color degrada-
tion in Concord grape juice. J. Food Sci. 38, 1060-1062.
SLOAN, J.L., BILLS, D.D. and LIBBEY, L.M. 1969. Heat-induced com-
pounds in strawberries. J. Agr. Food Chem. 27, 1370-1372.
SOMERS, T.C. 1966a. Grape phenolics: The anth~yaninsof Vitis vinifea,
variety Shiraz. J. Sci. Food Agric. 27, 215-219.
SOMERS, T.C. 1966b. Wine tannins - isolation of condensed flavonoid
pigments by gel-filtration. Nature 209, 368-370.
SOMERS, T.C. 1967. Resolution and analysis of total phenolic constituents of
grape pigment. J. Sci. Food Agric. 28, 193-1%.
SOMERS, T.C. 1971. The polymeric nature of wine pigments. Phytochem. 10,
2175-2 186.
SOMERS, T.C. and EVANS, M.E. 1974. Wine quality: Correlations with color
density and anthocyanin equilibria in a group of young red wines. J . Sci. Food
A@. 25, 1369-1379.
SONDHEIMER, E. 1953. On the relation between spectral changes and pH of
the anthocyanin pelargonidin 3-monoglucoside. J. Am. Chem. Soc. 75,
1507-1508.
SONDHEIMER, E. and KERTESZ, Z.I. 1952. The kinetics of the oxidation of
strawberry anthocyanin by hydrogen peroxide. Food Res. 27,288-297.
SONDHEIMER, E. and KERTESZ, Z.I. 1953. Participationof ascorbic acid in
the destruction of anthocyanin in strawberry juice and model systems. Food
Res. 218, 475-479.
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 245

SPAYD, S.E., NAGEL, C.W., HAYRYNEN, L.D. and DRAKE, S.R. 1984.
Color stability of apple and pear juices blended with fruit juices containing an-
thocyanins. J. Food Sci. 49, 411-414.
STAPLES, L.C. and FRANCIS, F.J. 1968. Colorimetry of cranberry cocktail
by wide range spectrophotometry. Food Technol. 22, 611-614.
STARR, M.S. and FRANCIS, F.J. 1968. Oxygen and ascorbic acid effect on
the relative stability of four anthocyanin pigments in cranberry juice. Food
T~hn01.22, 91-93.
STARR, M.S. and FRANCIS, F.J. 1973. Effect of metallic ions on color and
pigment content of cranberry juice cocktail. J. Food Sci. 38, 1043-1046.
SUOMALAINEN, H. and KERANEN, A.J.A. 1961. The frrst anthocyanins
appearing during the ripening of blueberries. Nature 292,498499.
TANCHEV, S . 1983. Kinetics of thermal degradation of anthocyanins. In Basic
Studies In Food Science, Volume 2. Proceedings of the Sixth International
Congress of Food Science and Technology, (J.V. McLoughlin and B.M.
McKenna, eds.) p. 96, Boole Press, Dublin, Ireland.
TANCHEV, S.S. and TIMBERLAKE, C.F. 1%9. The anthocyanins of red
cabbage (Brassicu oleruceu). Phytochem. 8, 1825-1827.
TANCHEV, S.S. and VASILEV, V.N. 1973. Gradinar. Lozar. Nauka 20,
23-28. Cited by Timberlake, C.F. and Bridle, P. 1982b. Distribution of an-
thocyanins in food plants. In Anthocyanins as Food Colors, (P. Markakis, ed.)
pp. 125-162, Academic Press, New York.
TANCHEV, S., VLADIMIROV, G. and IONCHEVA, N. 1%9. Effect of
some pectolytic enzymes on the destruction of anthocyanins. Nauchni
Trudove, Vissh Inst. Khranit. Vkusova Promyshl. 26, 77-82.
TIMBERLAKE, C.F. 1960. Metallic components of fruit juices. III. Oxidation
and stability of ascorbic acid in model systems resembling blackcurrant juice.
J. Sci. Food Agric. 11, 258-267.
TIMBERLAKE, C .F . 1980. Anthocyanins-occurence, extraction and
chemistry. Food Chem. 5, 69-80.
TIMBERLAKE, C.F. and BRIDLE, P. 1966. Spectral studies of anthocyanin
and anthocyanidin equilibria in aqueous solution. Nature 222, 158-159.
TIMBERLAKE, C.F. and BRIDLE, P. 1967a. Flavylium salts, anthocyanidins
and anthocyanins. I. Structural transformations in acid solutions. J. Sci. Food
Agric. 18,473-478.
TIMBERLAKE, C.F. and BRIDLE, P. 1967b. Flavylium salts, anthocyanidins
and anthocyanins. II. Reactions with sulphur dioxide. J. Sci. Food Agric. 18,
479-485.
TIMBERLAKE, C.F. and BRIDLE, P. 1968. Flavylium salts resistant to
sulphur dioxide. Chem. Ind., a t . , 1489.
246 R.L. JACKMAN, R.Y. YADA, M.A. TUNG AND R.A. SPEERS

TIMBERLAKE, C.F. and BRIDLE, P. 1975. The anthocyanins. In Ihe


Flavonoids, (J.B. Harbome, T.J. Mabry and H. Mabry, 4 s . ) pp. 215-266,
Chapman & Hall Ltd., London.
TIMBERLAKE, C.F. and BRIDLE, P. 1977.Anthocyanins: Colour augmen-
tation with catechin and acetaldehyde. J. Sci. Food Agric. 28, 539-544.
TIMBERLAKE, C.F. and BRIDLE, P. 1982a.Colour in beverages. In Sensory
Quality In Foods and Beverages: DeJinition, Measurement and Control,
(A.A. Williams and R.K. Atkin, eds.) pp. 140-154,Ellis Horwood Limited,
Chichester, England.
TIMBERLAKE, C.F. and BRIDLE, P. 1982b.Distribution of anthocyanins in
food plants. In Anthocyanins as Food Colors, (P. Markakis, ed.) pp. 125-162,
Academic Press, New York.
TINSLEY, I.J. and BOCKIAN, A.H. 1960. Some effects of sugars on the
breakdown of pelargonidin-3-glucosidein model systems at 90°C. Food Res.
25, 161-173.
VAN BUREN, J.P., SCHEINER, D.M. and WAGENKNECHT, A.C. 1960.
An anthocyanindecolorizing system in sour cherries. Nature 185, 165-166.
VAN TEELING, C.G., CANSFIELD, P.E. and GALLOP, R.A. 1971.An an-
thocyanin complex isolated from the syrup of canned blueberries. J . Food Sci.
36, 1061-1063.
VON ELBE, J.H., BIXBY, D.G. and MOORE, J.D. 1%9. Electrophoretic
comparison of anthocyanin pigments in eight varieties of sour cherries. J.
Food Sci. 34, 113-115.
WAGENKNECHT, A.C., SCHEINER, D.M. and VAN BUREN, J.P. 1960.
Anthocyanase activity and its possible relation to scald in sour cherries. Food
Technol. 14, 47-49.
WESCHE-EBELING, P.A.E. 1984.PuriJicationOfStrawberryPolyphenol Ox-
iduse and Its Role In Anthocyanin Degradation, pp. 1-180, Ph.D. thesis,
Oregon State University, Corvallis, OR.
WILLIAMS, M. and HRAZDINA, G. 1979.Anthocyanins as food colorants:
Effect of pH on the formation of anthocyanin-rutin complexes. J. Food Sci.
44,6648.
WROLSTAD, R.E. and ERLANDSON, J.A. 1973.Effect of metal ions on the
color of strawberry puree. J. Food Sci. 38, 460463.
WROLSTAD, R.E., PUTMAN, T.P. and VARSEVELD, G.W. 1970.Color
quality of frozen strawberries: Effect of anthocyanin, pH, total acidity and
ascorbic acid variability. J. Food Sci. 35,448-452.
YANG, H.Y. and STEELE, W.F. 1958.Removal of excess anthocyanin pig-
ment by enzyme. Food Technol. 12, 517-519.
YAZAKI, Y . 1976.Co-pigmentationand the color change with age in petals of
Fuchsia hydrida. Bot. Mag. Tokyo 89,45-57.
ANTHOCYANINS AS FOOD COLORANTS - A REVIEW 247

YOSHITAMA, K. 1977. An acylated deiphinidin 3-rutinoside-5,3'5'-


triglucoside from Lobelia erinus. Phytochem. 16, 1857-1858.
YOSHITAMA, K. 1978. Blue and purple anthocyanins isolated from the
flowers of Tradescantia reflexa. Bot. Mag. Tokyo 91, 207-212.
YOSHITAMA, K. and ABE, K. 1976. Chromatographic and spectral
characterization of 3 '-glycosylation in anthocyanidins. Phytochem. 16,
591-593.
YOSHITAMA, K. and HAYASHI, K. 1974. Concerning the structure of
cinerarin, a blue anthocyanin from garden cineraria. Studies on anthocyanins,
LXVI. Bot. Mag. Tokyo 87, 33-40.
YOSHITAMA, K., HAYASHI, K., ABE, K. and KAKISAWA, H. 1975. Fur-
ther evidence for the glycoside structure of cinerarin. Studieson anthocyanins,
LXVII. Bot. Mag. Tokyo 88, 213-217.
ZAPSALIS, C. and FRANCIS, F.J. 1965. Cranberry anthocyanins. J. Food
Sci. 30, 396-399.

View publication stats

You might also like