Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Comprehensive

Reviews
in
Food Science
and
Food Safety

Mechanisms
and Factors
for Edible Oil
Oxidation
Eunok Choe and David B. Min

ABSTRACT: Edible oil is oxidized during processing and storage via autoxidation and photosensitized oxidation, in
which triplet oxygen (3 O 2 ) and singlet oxygen (1 O 2 ) react with the oil, respectively. Autoxidation of oils requires radical
forms of acylglycerols, whereas photosensitized oxidation does not require lipid radicals since 1 O 2 reacts directly with
double bonds. Lipid hydroperoxides formed by 3 O 2 are conjugated dienes, whereas 1 O 2 produces both conjugated
and nonconjugated dienes. The hydroperoxides are decomposed to produce off-flavor compounds and the oil quality
decreases. Autoxidation of oil is accelerated by the presence of free fatty acids, mono- and diacylglycerols, metals such
as iron, and thermally oxidized compounds. Chlorophylls and phenolic compounds decrease the autoxidation of oil
in the dark, and carotenoids, tocopherols, and phospholipids demonstrate both antioxidant and prooxidant activity
depending on the oil system. In photosensitized oxidation chlorophyll acts as a photosensitizer for the formation
of 1 O 2 ; however, carotenoids and tocopherols decrease the oxidation through 1 O 2 quenching. Temperature, light,
oxygen concentration, oil processing, and fatty acid composition also affect the oxidative stability of edible oil.

Introduction
Oxidative stability of oils is the resistance to oxidation during
processing and storage (Guillen and Cabo 2002). Resistance to
oxidation can be expressed as the period of time necessary to attain the critical point of oxidation, whether it is a sensorial change
or a sudden acceleration of the oxidative process (Silva and others
2001). Oxidative stability is an important indicator to determine
oil quality and shelf life (Hamilton 1994) because low-molecularweight off-flavor compounds are produced during oxidation. The
off-flavor compounds make oil less acceptable or unacceptable
to consumers or for industrial use as a food ingredient. Oxidation
of oil also destroys essential fatty acids and produces toxic compounds and oxidized polymers. Oxidation of oil is very important
in terms of palatability, nutritional quality, and toxicity of edible
oils.
Different chemical mechanisms, autoxidation and photosensitized oxidation, are responsible for the oxidation of edible oils
during processing and storage depending upon the types of oxygen. Two types of oxygen can react with edible oils. One is called
atmospheric triplet oxygen, 3 O 2 , and the other is singlet oxygen,1 O 2 . 3 O 2 reacts with lipid radicals and causes autoxidation,
which is a free radical chain reaction. The chemical properties
of 3 O 2 to react with lipid radicals can be easily explained by the
molecular orbital of the oxygen as shown in Figure 1. The 3 O 2 in
the ground state with 2 unpaired electrons in the 2p antibond-

MS 20060292 Submitted 5/23/06, Accepted 7/13/06. Author Choe is with Dept.


of Food and Nutrition, The Inha Univ., Incheon, Korea. Author Min is with Dept.
of Food Science and Technology, The Ohio State Univ., Columbus, OH. Direct
inquiries to author Min (E-mail: min.2@osu.edu).


C

2006 Institute of Food Technologists

ing orbitals has a permanent magnetic moment with 3 closely


grouped energy states under a magnetic field and is called triplet
oxygen. 3 O 2 is a radical with 2 unpaired orbitals in the molecule.
It reacts with radical food compounds in normal reaction conditions according to spin conservation. Photosensitized oxidation
of edible oils occurs in the presence of light, sensitizers, and atmospheric oxygen, in which 1 O 2 is produced.
The electron configuration in the 2p antibonding orbitals of
1
O 2 is shown in Figure 2. Since 1 orbital of the 2p antibonding orbitals of1 O 2 has paired electrons and the other is completely empty, 1 O 2 has 1 energy level under a magnetic field
and is electrophilic. The nonradical electrophilic singlet oxygen
readily reacts with compounds with high electron densities, such
as the double bonds of unsaturated fatty acids. The 1 O 2 has an
energy of 93.6 kJ above the ground state of 3 O 2 (Korycka-Dahl
and Richardson 1978; Girotti 1998). High-energy 1 O 2 in solution is deactivated by transferring its energy to the solvent, and
its lifetime depends on the solvent. The lifetime of singlet oxygen
is about 2, 17, and 700 s in water, hexane, and carbon tetrachloride, respectively (Merkel and Kearns 1972; Long and Kearns
1975).
Oxidation of edible oils is influenced by an energy input such
as light or heat, composition of fatty acids, types of oxygen, and
minor compounds such as metals, pigments, phospholipids, free
fatty acids, mono- and diacylglycerols, thermally oxidized compounds, and antioxidants. Many efforts have been made to improve the oxidative stabilities of oils by systematic studies on the
effects of these factors.
This article reviews the reaction mechanism, kinetics, and
oxidation products of autoxidation and photosensitized oxidation. Factors affecting the oxidation of edible oils, free
fatty acids, mono- and diacylglycerols, metals, phospholipids,

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

169

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


Figure 1 --- Molecular orbital of triplet oxygen, 3 O2

Molecular
*

Atomic

2px

2py

2pz

Atomic

2pz

2py

2px

Ener gy

2S

2S

1S

Figure 2 --- Electronic configuration of 2p antibonding


orbital of singlet oxygen, 1 O2

1S

Table 1 --- Hydroperoxides of fatty acids by autoxidationa


Fatty acid

Hydroperoxides at

Relative amount (%)

Oleic acid

pigments, thermally oxidized compounds, and antioxidants, are


also discussed.

C8
C9
C10
C11

2628
2225
2224
2628

Linoleic acid

C9
C13

4853
4853

Mechanisms of Autoxidation in Edible Oil

Linolenic acid

C9
C12
C13
C16

2835
813
1013
2835

Autoxidation of oils, free radical chain reaction, includes initiation, propagation, and termination steps:
a Frankel

Initiation
Propagation
Termination

RH
R + 3 O2
ROO + RH
ROO + R
R + R

R + H
ROO
ROOH + R
ROOR
RR
(R : lipid alkyl)

Autoxidation of oils requires fatty acids or acylglycerols to be


in radical forms. Fatty acids or acylglycerols are in nonradical
singlet states, and the reaction of fatty acids with radical state
atmospheric 3 O 2 is thermodynamically unfavorable due to electronic spin conservation (Min and Bradley 1992). The hydrogen
atom in the fatty acids or acylglycerols in edible oil is removed
and lipid alkyl radicals are produced in the initiation step. Heat,
metal catalysts, and ultraviolet and visible light can accelerate
free radical formation of fatty acids or acylglycerols. The energy
required to remove hydrogen from fatty acids or acylglycerols
is dependent on the hydrogen position in the molecules. The
hydrogen atom adjacent to the double bond, especially hydrogen attached to the carbon between the 2 double bonds, is removed easily. Hydrogen at C11 of linoleic acid is removed at
170

(1985).

50 kcal/ mol. The energy required to remove hydrogen in C8 and


C14 of linoleic acid is 75 kcal/ mol and the homolytic dissociation
energy between hydrogen and C17 or C18 is about 100 kcal/mol
(Min and Boff 2002). The double bond adjacent to the carbon radical in linoleic acid shifts to the more stable next carbon and from
the cis to the trans form. Autoxidation of linoleic and linolenic
acids produces only conjugated products. The hydroperoxide positional isomers formed in the autoxidation of oleic, linoleic, and
linolenic acids are shown in Table 1.
The lipid alkyl radical reacts with 3 O 2 and forms lipid peroxy
radical, another reactive radical. Reaction between lipid alkyl
radical and 3 O 2 occurs very quickly at normal oxygen pressure and, consequently, the concentration of lipid alkyl radical
is much lower than that of lipid peroxy radical (Aidos and others 2002). The lipid peroxy radical abstracts hydrogen from other
lipid molecules and reacts with the hydrogen to form hydroperoxide and another lipid alkyl radical. These radicals catalyze the
oxidation reaction, and autoxidation is called the free radical
chain reaction. Figure 3 shows the hydroperoxide formation in

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Mechanisms and factors for edible oil oxidation

13

12

10

13

(CH2)7COOH

CH3(CH2)4

12

10

(CH2)7COOH

CH3(CH2)4

-H
13

12

10

13

(CH2)7COOH

12

10

12

12

10

10

(CH2)7COOH

+3O2, H

13

(CH2)7COOH CH3(CH2)4

CH3(CH2)4

CH3(CH2)4

+3O2, H

13

10

(CH2)7COOH

13

(CH2)7COOH

12

CH3(CH2)4

CH3(CH2)4

HOO

Figure 3 --- Hydroperoxide


formation in the
autoxidation of linoleic
acid

-H

CH3(CH2)4

13

13-Hydroperoxide

12

10

OOH

(CH2)7COOH

9-Hydroperoxide

the autoxidation of linoleic acid. The rates for the formation of


lipid peroxy radical and hydroperoxide depend only on oxygen
availability and temperature (Velasco and others 2003). When
radicals react with each other, nonradical species are produced
and the reaction stops.
The primary oxidation products, lipid hydroperoxides, are relatively stable at room temperature and in the absence of metals.
However, in the presence of metals or at high temperature they are
readily decomposed to alkoxy radicals and then form aldehydes,
ketones, acids, esters, alcohols, and short-chain hydrocarbons.
The most likely pathway of hydroperoxide decomposition is a
homolytic cleavage between oxygen and the oxygen bond, in
which alkoxy and hydroxy radicals are produced. The activation
energy to cleave the oxygenoxygen bond is 46 kcal/mol lower
than that to cleave the oxygenhydrogen bond (Hiatt and others
1968). The alkoxy radical then undergoes homolytic -scission of
the carboncarbon bond and produces oxo-compounds and saturated or unsaturated alkyl radicals (Figure 4). After electron rear-

rangement, the addition of hydroxyl radical, or hydrogen transfer,


the ultimate secondary lipid oxidation products are mostly lowmolecular-weight aldehydes, ketones, alcohols, and short-chain
hydrocarbons, as shown in Table 2.
The time for secondary product formation from the primary
oxidation product, hydroperoxide, varies with different oils. Secondary oxidation products are formed immediately after hydroperoxide formation in olive and rapeseed oils. However, in
sunflower and safflower oils, secondary oxidation products are
formed when the concentration of hydroperoxides is appreciable
(Guillen and Cabo 2002).
Most decomposition products of hydroperoxides are responsible for the off-flavor in the oxidized edible oil. Aliphatic carbonyl
compounds have more influence on the oxidized oil flavor due
to their low threshold values. Threshold values for hydrocarbons,
alkanals, 2-alkenals, and trans, trans-2,4-alkadienals are 90 to
2150, 0.04 to1, 0.04 to 2.5, and 0.04 to 0.3 ppm, respectively
(Frankel 1985). Hexanal (23.5%) and 2-decenal (34.3%), and

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

171

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


R1

Figure 4 --- Mechanisms of hydroperoxide


decomposition to form secondary oxidation
products

R2
O2

CH

R2

CH

CH

R1

OOH

OH

B
R2

CH

CH

A
CH

R1

O
CH

R2

OH

CH

CH

R2
OH

CH

+ OHC

B1
R2

R1

R2

CH

CH

R1

R 3H

A1

CHO

OH

A2

R3

R 2H

R 3H
R3

OH
B2

R2

CH

CHO

R2

CH

CH

2-heptenal (29.5%) and trans-2-octenal (18.1%), were the major


volatile compounds detected by the solid-phase microextraction
method in soybean and corn oils (peroxide value of 5), respectively (Steenson and others 2002). Pentane, hexanal, propenal,
and 2,4-decadienal were present in high amounts in canola oil
stored uncovered at 60 C (Vaisey-Genser and others 1999).
Frankel (1985) reported that trans, cis-2,4-decadienal was the
most significant compound in determining the oxidized flavor
of oil, followed by trans, trans-2,4-decadienal, trans, cis-2,4heptadienal, 1-octen-3-ol, butanal, and hexanal. Hexanal, pentane, and 2,4-decadienal were suggested and used as indicators
to determine the extent of the oil oxidation (Warner and others
1978; Przybylski and Eskin 1995; Choe 1997; Heinonen and others 1997). Trans-2-hexenal, and trans, cis, trans-2,4,7-decatrienal
and 1-octen-3-one, were reported to give grass-like and fish-like
flavor in oxidized soybean oil, respectively (Min and Bradley
1992). No single flavor compound is mainly responsible for the
oxidized flavor of vegetable oils.
172

Singlet Oxygen Formation and Mechanisms


of Photosensitized Oxidation of Edible Oil
Oil oxidation is accelerated by light, especially in the presence
of sensitizers such as chlorophylls. Sensitizers in singlet state absorb light energy very rapidly, in picoseconds, and become excited. Excited singlet sensitizers can return to their ground state
via emission of light, internal conversion, or intersystem crossing
(Figure 5). Fluorescence and heat are produced by emission of
light and internal conversion, respectively. Intersystem crossing
results in excited triplet state of sensitizers.
Excited triplet sensitizers may accept hydrogen or an electron
from the substrate and produce radicals (type I) as shown in
Figure 6. Excited triplet sensitizers react with 3 O 2 and produce superoxide anion by electron transfer. Superoxide anion produces
hydrogen peroxide, one of the reactive oxygen species by spontaneous dismutation, and the reaction of hydrogen peroxide with
superoxides results in singlet oxygen formation by HaberWeiss
reaction in the presence of transition metals such as iron or copper

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Mechanisms and factors for edible oil oxidation


Table 2 --- Secondary oxidation products of fatty acid methyl ester by autoxidationa
Class

Oleic acid

Linoleic acid

Linolenic acid

Aldehydes

Octanalb
Nonanal
2-Decenal
Decanal

Pentanal
Hexanal
2-Octenal
2-Nonenal
2,4-Decadienal

Propanal
Butanal
2-Butenal
2-Pentenal
2-Hexenal
3,6-Nonadienal
Decatrienal

Carboxylic acid

Methyl heptanoate
Methyl octanoate
Methyl 8-oxooctanoate
Methyl 9-oxononanoate
Methyl 10-oxodecanoate
Methyl 10-oxo-8-decenoate
Methyl 11-oxo-9-undecenoate

Methyl heptanoate
Methyl octanoate
Methyl 8-oxooctanoate
Methyl 9-oxononanoate
Methyl 10-oxodecanoate

Methyl heptanoate
Methyl octanoate
Methyl nonanoate
Methyl 9-oxononanoate
Methyl 10-oxodecanoate

Alcohol

1-Heptanol

1-Pentanol
1-Octene-3-ol

Hydrocarbons

Heptane
Octane

Pentane

a Frankel

Ethane
Pentane

(1985).

Figure 5 --- Excitation and deactivation of sensitizer


(Sen; sensitizers, ISC; intersystem crossing)

(Kellogg and Fridovich 1975):


+

O
2 + O2 + 2H H2 O2 + O2

1
H2 O2 + O
2 HO + OH + O2

The excitation energy of triplet sensitizers can be transferred onto adjacent 3 O 2 to form 1 O 2 by triplettriplet annihilation, and the sensitizers return to their ground singlet state
(type II). Kochevar and Redmond (2000) reported that a sensitizer molecule may generate 103 to 105 molecules of 1 O 2 before
becoming inactive.
The rate of the type I or II processes depends on the kinds of
sensitizers and substrates (Chen and others 2001), and concentrations of substrates and oxygen (Foote 1976). Phenols or amines,
which are readily oxidized, or readily reducible quinones favor
the type I process. On the other hand, olefins, dienes, and aromatic compounds, which are not so readily oxidized or reduced,
more often favor type II. Photosensitized oxidation of edible oil
follows the singlet oxygen oxidation pathway. 1 O 2 was suggested
to be involved in the initiation of the oil oxidation (Rawls and Van
Santen 1970; Lee and Min 1988).
1
O 2 either reacts chemically with other molecules or transfers
its energy to them. When 1 O 2 reacts with unsaturated fatty acids,

mostly allyl hydroperoxides are formed by ene reaction (Gollnick


1978), as shown in Figure 7.
Electrophilic 1 O 2 can directly react with high-electron-density
double bonds without the formation of alkyl radical, and form
hydroperoxides at the double bonds. When hydroperoxide is
formed, double bond migration and trans fatty acid occur, producing both conjugated and nonconjugated hydroperoxides, as
shown in Table 3. Production of nonconjugated hydroperoxides
is not observed in the autoxidation. Figure 8 shows the oxidation
pathway of linoleic acid by 1 O 2 .
Hydroperoxides formed by 1 O 2 oxidation are decomposed by
the same mechanisms for the hydroperoxides formed by 3 O 2
in autoxidation. 1 O 2 oxidation produces more 2-decenal and
octane, two of the decomposition products of hydroperoxides,
in oleate than autoxidation does (Frankel 1985). The contents
of octanal and 10-oxodecanoate in autoxidized oleate were
higher than those of 1 O 2 -oxidized oleate. 2-Heptenal and 2butenal were noticeable in 1 O 2 -oxidized linoleic and linolenic
acids, whereas they were negligible in autoxidized linoleic and
linolenic acids. Heptenal was formed only in 1 O 2 -oxidized soybean oil in the presence of chlorophyll and light (Min and others
2003).
A beany flavor, which is a unique and undesirable flavor in
soybean oil with low peroxide value, has been a problem for the

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

173

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


3

Sen*

Figure 6 --- Reaction of triplet sensitizers with


substrates

Type I

Type II

+ 3O2

+ RH

R + Sen H ( or

R+ + Sen- )

O2

+ O2
+ 3 O2

+ RH

ROOH

CH3(CH2)6

O2 + Sen+

(CH2)6COOH
1
O2

ROOH

CH3(CH2)6

(CH2)6COOH

OOH

(CH2)6COOH

CH3(CH2)6
1

(CH2)6COOH

CH3(CH2)6

O2

Figure 7 --- Formation of


allyl hydroperoxide of
oleic acid by ene
reaction

OOH
(CH2)6COOH

CH3(CH2)6
O
O

last 70 years and studied extensively both nationally and inter- Table 3 --- Hydroperoxides of fatty acids by singlet oxygen oxidation
nationally (Chang and others 1966; Smouse and Chang 1967).
Relative
2-Pentylfuran and pentenylfuran were reported to be responsible
Fatty acid
Hydroperoxides at amount (%)a Type of acid
for the beany flavor (Chang and others 1966, 1983; Smouse and
Chang 1967; Ho and others 1978; Smagula and others 1979), and Oleic acid
C9
48
1
O 2 was involved in their formation from linoleic and linolenic
C10
52
acids present in soybean oil, as shown in Figure 9 and 10 (Min and
C9
32
Conjugated
others 2003). Min and others (2003) strongly indicated that the Linoleic acid
C10
17
Nonconjugated
reversion flavor of soybean oil can be decreased or eliminated by
C12
17
Nonconjugated
removing chlorophylls from the oil during processing. The soyC13
34
Conjugated
bean oil industry currently removes effectively chlorophylls from
soybean oil using bleaching materials during the refining process,
C9
23
Conjugated
and the beany flavor is no longer a serious problem in soybean Linolenic acid
C10
13
Nonconjugated
oil.
C12
C13
C15
C16

Factors Affecting the Oxidation of Edible Oil


The oxidation of oil is influenced by the fatty acid composition of the oil, oil processing, energy of heat or light, the
174

a Frankel

(1985).

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

12
14
13
25

Conjugated
Conjugated
Nonconjugated
Conjugated

Mechanisms and factors for edible oil oxidation

CH3(CH2)4

OOH

O2

(CH2)7COOH

CH3(CH2)4

(CH2)7COOH

(CH2)7COOH

CH3(CH2)4

Figure 8 --- Hydroperoxide


formation in linoleic acid
oxidation by singlet
oxygen

Conjugated
1

OOH

O2
O
CH3(CH2)4

(CH2)7COOH
CH3(CH2)4

CH(CH2)6COOH

Nonconjugated
1

OOH

O2
CH3(CH2)4

O
H

(CH2)7COOH

(CH2)7COOH

CH3(CH2)3CH

Nonconjugated
OOH
1

OOH

O2
O
CH3(CH2)4

(CH2)7COOH
(CH2)7COOH

CH3(CH2)4

Conjugated

concentration and type of oxygen, and free fatty acids, mono- and
diacylglycerols, transition metals, peroxides, thermally oxidized
compounds, pigments, and antioxidants. These factors interactively affect the oxidation of oil and it is not easy to differentiate
the individual effect of the factors.
Fatty acid composition of oils

Oils that are more unsaturated are oxidized more quickly than
less unsaturated oils (Parker and others 2003). As the degree of
unsaturation increases, both the rate of formation and the amount
of primary oxidation compounds accumulated at the end of the
induction period increase (Martin-Polvillo and others 2004). Soybean, safflower, or sunflower oil (iodine values > 130) stored in
the dark showed a significantly (P < 0.05) shorter induction period than coconut or palm kernel oil whose iodine value is less
than 20 (Tan and others 2002). High oleic and high stearic oils
by gene silencing of the oilseeds or hydrogenated soybean oil
showed higher autoxidative stability (Evans and others 1973; Liu
and others 2002). The autoxidation rate greatly depends on the
rate of fatty acid or acylglycerol alkyl radical formation, and the
radical formation rate depends mainly on the types of fatty acid
or acylglycerol. The relative autoxidation rate of oleic, linoleic,
and linolenic acids was reported as 1:40 to 50:100 on the basis
of oxygen uptake (Min and Bradley 1992).
The difference in1 O 2 oxidation rate among fatty acids is lower
than that for autoxidation. The reaction rates between 1 O 2 and
stearic, oleic, linoleic, and linolenic acids are 1.2 104 , 5.3
104 , 7.3 104 , and 10.0 104 M1 s1 , respectively (VeverBizet and others 1989). The relative reaction rates of 1 O 2 with
oleic, linoleic, and linolenic acids are 1.0:1.4:1.9, respectively.
Soybean oil reacts with 1 O 2 at a rate of 1.4 105 M1 s1 in
methylene chloride at 20 C (Lee and Min 1991). The type of
polyunsaturated fatty acids, nonconjugated or conjugated dienes
or trienes, has little effect on the reaction between the lipid and
1
O 2 (Rahmani and Saari Csallani 1998).

Oil processing

The oil-processing method affects the oxidative stability of an


oil. Crude soybean oil was the most stable to the oxidation followed by deodorized, degummed, refined, and bleached oil during 6 d storage at 55 C in the dark (Jung and others 1989). Higher
oxidative stability of crude oil than refined oil was suggested to
be partly due to higher concentrations of tocopherols in crude
oil (1670 ppm) than refined oil (1546 ppm). The induction time
in hexane-extracted rapeseed oil oxidation at 90 C was 10.5
1.9 h compared with an induction time of 8.1 0.7 h for pressed
rapeseed oil (Krygier and Platek 1999). Oxidative stability was
significantly lower in supercritical carbon dioxide-extracted walnut oil than in pressed oil (Crowe and White 2003). Roasting of
safflower and sesame seeds before oil extraction improved the
oxidative stability of their oils (Yen and Shyu 1989; Lee and others 2004), partly due to the Maillard products produced during
roasting. Some Maillard reaction products were reported to be
antioxidants. Oxidative stability increased as the roasting and expelling temperatures of the seeds increased.
Temperature and light

Autoxidation of oils and the decomposition of hydroperoxides


increase as the temperature increases (Shahidi and Spurvey 1996;
St. Angelo 1996). The formation of autoxidation products during the induction period is slow at low temperature (Velasco
and Dobarganes 2002). The concentration of the hydroperoxides increases until the advanced stages of oxidation. The content
of polymerized compounds increases significantly at the end of
the induction period of autoxidation (Marquez-Ruiz and others
1996). The hydroperoxide decomposition rate of crude herring
oil stored at 50 C in the dark was higher than the formation rate
of hydroperoxide. The reverse phenomena were observed in the
same crude herring oil at 0 or 20 C in the dark (Aidos and others
2002).
Temperature has little effect on 1 O 2 oxidation due to the low activation energy of 0 to 6 kcal/mole (Yang and Min 1994; Rahmani

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

175

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


CH3 (CH2)4 CH CH CH2 CH CH CH2 (CH2)6 COOH
1

O2

CH3 (CH2)4 CH CH CH2 CH CH CH (CH2)6 COOH


O
OH
CH3 (CH2)4 CH CH CH2 CH
O
1
O2
CH3 (CH2)4 CH CH CH2 CH
O
O
O

CH3 (CH2)4 CH CH CH2 CH


O

+ 2 RH
CH3 (CH2)4 CH CH2 CH2 CH
O
O
OH

+ 2R

CH3 (CH2)4 CH CH2 CH2 CH


O
O

CH3 (CH2)4 C CH2 CH2 CH


O
O

CH3 (CH2)4 C CH CH CH
OH

OH
- H2O

CH3 (CH2)4
O
Figure 9 --- Formation of 2-pentylfuran from linoleic acid by singlet oxygen oxidation

and Saari Csallani 1998). Light is much more important than temperature in 1 O 2 oxidation. Light of shorter wavelengths had more
detrimental effects on the oils than longer wavelengths (Sattar
and others 1976). Reportedly, the effect of light on oil oxidation
becomes less as temperature increases (Velasco and Dobarganes
2002).
Because 1 O 2 oxidation occurs in the presence of light, the
packaging of oils is very important. Transparent plastic bottles increase oil oxidation. The incorporation of Tinuvin 234
176

(2-(2-hydroxy-3,5-di(1,1-dimethylbenzyl)phenyl) benzotriazole)
or Tinuvin 326 (2-(3-tert-butyl-2-hydroxy-5-methylphenyl)-5chlorobenzo-triazole), which is a UV absorber, into the transparent plastic bottles improved oxidative and sensory stability of
soybean oil under light (Pascall and others 1995; Azeredo and
others 2003).
Oxygen

The oxidation of oil can often take place when oil, oxygen, and
catalysts are in contact. Both concentration and type of oxygen
affect the oxidation of oils. The oxygen concentration in the oil
is dependent on the oxygen partial pressure in the headspace of
the oil (Andersson 1998). A higher amount of oxygen is dissolved
in the oil when the oxygen partial pressure in the headspace is
high. Oxidation of the oil increased with the amount of dissolved
oxygen (Min and Wen 1983). The solubility of oxygen is higher
in the oil than in water, and also in crude oil than refined oil (Aho
and Wahroos 1967). One gram of soybean oil dissolves 55 g
oxygen at room temperature (Andersson 1998). The amount of
oxygen dissolved in oil is sufficient to oxidize the oil to a peroxide
value of approximately 10 meq/kg in the dark (Przybylski and
Eskin 1988). Min and Wen (1983) reported that the rate constants
for oxygen disappearance in soybean oil containing 2.5, 4.5, 6.5,
and 8.5 ppm dissolved oxygen during storage at 55 C in the dark
were 0.049, 0.058, 0.126, and 0.162 ppm/ h, respectively.
The effect of oxygen concentration on the oxidation of oil increased at high temperature and in the presence of light and metals such as iron or copper. Higher oxygen dependence of oil oxidation at high temperature is due to low solubility of oxygen in
the oil at high temperature (Andersson 1998). The oxygen has to
be transported into the oil by diffusion when the oil is not stirred,
such as during storage at low temperature. Convection is another
important pathway for oxygen penetration into the oil from the
surface when the oil is stirred, for example, during processing at
high temperature.
Oil oxidation rate is independent of oxygen concentration at
sufficiently high oxygen concentrations, for example, above 10%
in the oxidation of methyl linoleate (Kacyn and others 1983).
When the oxygen content is low, the oxidation rate is dependent
on oxygen concentration and is independent of lipid concentration (Andersson 1998). The autoxidation rate of oil at more than
4% to 5% oxygen in the headspace was independent of oxygen concentration and was directly dependent on the lipid concentration (Labuza 1971). However, the reverse was true at low
oxygen pressure of less than 4% in the headspace (Karel 1992).
The oxidation of rapeseed oil at 50 C in the dark, measured as
oxygen consumption or peroxide values, became faster as the
oxygen concentration increased at less than 0.5% oxygen in the
headspace, whereas the oxidation rate decreased with oxygen
concentration at more than 1% oxygen (Andersson and Lingnert
1999).
Oxygen and edible oil can react more efficiently when an oil
sample size is small or an oil sample has a high ratio of surface
to volume (Tan and others 2002; Kanavouras and others 2005).
When the surface-to-volume ratio increases, the relative rate of
oxidation is less oxygen-dependent with a low oxygen content.
The container surface can act as a reduction catalyst, and its
effect was shown to be proportional to the area of the container
in contact with the oils (Brimberg and Kamal-Eldin 2003).
The interaction between temperature and oxygen concentration affects the volatile formation in rapeseed oil in the dark;
production of 2-pentenal and 1-pentene-3-one correlated positively with oxygen concentration at 50 C, but negatively at 35 C
(Andersson and Lingnert 1999).
The reaction rate between lipid and oxygen is dependent on
the type of oxygen; the reaction rate of 1 O 2 with lipid is much

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Mechanisms and factors for edible oil oxidation


O
CH3 CH2 CH CH CH2 CH CH CH2 CH CH CH2 (CH2)6 COH

Figure 10 --- Formation of 2-pentenylfuran from linolenic


acid by singlet oxygen oxidation

O2
O

CH3 CH2 CH CH CH2 CH CH CH2 CH CH CH (CH2)6 COH


O
O
H
O
CH3 CH2 CH CH CH2 CH CH CH2 CH CH CH (CH2)6 COH
O

CH3 CH2 CH CH CH2 CH CH CH2 CH


O
1

O2

CH3 CH2 CH CH CH2 CH CH CH2 CH


O
O
O
H
CH3 CH2 CH CH CH2 CH CH CH2 CH
O
O

CH3 CH2 CH CH CH2 C CH2 CH2 CH


O
O

CH3 CH2 CH CH CH2 C CH CH CH


OH
OH
- H2O

CH3 CH2 CH CH CH2


O

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

177

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


higher than that of 3 O 2 because 1 O 2 can directly react with lipids.
3
O 2 reacts with the radical state of lipids. Linoleates react with
1
O 2 at a rate 1,450 times faster than with 3 O 2 (Rawls and Van
Santen 1970).
Minor components present in oil

Edible oil consists of mostly triacylglycerols, but it also contains minor components such as free fatty acids, mono- and
diacylglycerols, metals, phospholipids, peroxides, chlorophylls,
carotenoids, phenolic compounds, and tocopherols. Some of
them accelerate the oil oxidation and others act as antioxidants.
Free fatty acid and mono- and diacylglycerols. Crude oil contains free fatty acids, and oil processing, such as refining, decreases the free fatty acid contents. Crude soybean oil contains
about 0.7% free fatty acids; however, refined oil contains 0.02%
free fatty acids (Jung and others 1989). Sesame oil extracted from
roasted sesame seeds contains 0.72% free fatty acids and bleaching with acid clay decreases free fatty acid contents of the oil to
0.56% (Kim and Choe 2005). Free fatty acids are more susceptible to autoxidation than esterified fatty acids (Kinsella and others
1978). Free fatty acids act as prooxidants in edible oil (Miyashita
and Takagi 1986; Mistry and Min 1987a). They have hydrophilic
and hydrophobic groups in the same molecule and prefer to be
concentrated on the surface of edible oils. The hydrophilic carboxy groups of the free fatty acids will not easily dissolve in the
hydrophobic edible oil and are present on the surface of edible
oil. Mistry and Min (1987a) reported that free fatty acids decrease
the surface tension of edible oil and increase the diffusion rate of
oxygen from the headspace into the oil to accelerate oil oxidation.
Mono- and diacylglycerol, usually present at 0.07% to 0.11%
and 1.05% to1.20% in soybean oil, respectively, acted as prooxidants to increase oxidation at 55 C in the dark (Mistry and Min
1987b; Mistry and Min 1988). The mono-and diacylglycerols,
which have hydrophilic hydroxy groups and hydrophobic hydrocarbons, decrease the surface tension of edible oil and increase the diffusion rate of oxygen from the headspace to the
oil to accelerate the oxidation of oil. Mono- and diacylglycerols should be removed from the oil during the oil-refining process to improve the oxidative stability of edible oils (Mistry and
Min 1988).
Metals. Crude oil contains transition metals such as iron or
copper. For example, crude soybean oil contains 13.2 ppb and
2.80 ppm of copper and iron, respectively. However, refining reduces their contents. Edible oils manufactured without refining,
such as extra virgin olive oil and sesame oil, contain relatively
high quantities of transition metals (Table 4). Metals increase the
rate of oil oxidation due to the reduction of activation energy of
the initiation step in the autoxidation down to 63104 kJ/mol
(Jadhav and others 1996). Metals react directly with lipids to
produce lipid alkyl radicals. They also produce reactive oxygen
Table 4 --- Copper and iron contents in edible oils
Metal contenta
Oils

Copper (ppb)
oilb

Cold-pressed sesame
Crude soybean oilc
Virgin olive oilb
Cold-pressed sunflower oilb
Refined olive oild
Refined soybean oilc
a Numbers in parentheses are the
b MAFF (1997).
c Sleeter (1981).
d Leonardis and Macciola (2002).

178

16 (3.038)
13.2
9.8 (1.079)
5.2 (2.28.5)
15
2.5

ranges reported.

Iron (ppm)
1.16 (0.181.52)
2.80
0.73 (nd9.79)
0.26 (0.220.31)
0.08
0.20

species such as 1 O 2 and hydroxy radical from 3 O 2 and hydrogen peroxide, respectively (Andersson 1998). Lipid alkyl radical
and reactive oxygen species accelerate oil oxidation. Copper accelerates hydrogen peroxide decomposition 50 times faster than
ferrous ion (Fe2+ ), and ferrous ion acts 100 times faster than ferric
ion (Fe3+ ):
Fe3+ + RH Fe2+ + R + H+
Fe2+ + 3 O2 Fe3+ + O
2
+

1
O
2 + O2 + 2H O2 + H2 O2
metal catalyst

H2 O2 + O
HO + OH + 1 O2
2
Fe2+ + H2 O2 Fe3+ + OH + HO
Metals also accelerate autoxidation of oil by decomposing hydroperoxides (Benjelloun and others 1991):
ROOH + Fe2+ (or Cu+ ) RO + Fe3+ (or Cu2+ ) + OH
ROOH + Fe3+ (or Cu2+ ) ROO + Fe2+ (or Cu+ ) + H+
Fe2+ is much more active with a rate of 1.5 103 M1 s1
(Halliwell and Gutteridge 2001) than Fe3+ in decomposing the
lipid hydroperoxides to catalyze autoxidation (Mei and others
1998). Fe3+ also causes decomposition of phenolic compounds
such as caffeic acid in olive oil and decreases oil oxidative stability (Keceli and Gordon 2002).
Shiota and others (2006) reported that the prooxidant activity
of iron was suppressed by lactoferrin in fish oil or soybean oil
oxidation at 50 C to 120 C, possibly due to the iron-binding
ability of lactoferrin. Depletion of iron from the surroundings by
lactoferrin can suppress the iron-catalyzed oxidation of the oil.
Phospholipids. Crude oils contain phospholipids such as
phosphatidylethanolamine, phosphatidylcholine, phosphatidylinositol, phsphatidylserine, and phosphatidic acid, but most
of them are removed by processing such as degumming.
Crude soybean oil contains phosphatidylcholine and phosphatidylethanolamine at 500.8 and 213.6 ppm, respectively.
However, refined, bleached, and deodorized (RBD) soybean
oil contains 0.86 and 0.12 ppm of phosphatidylcholine and
phosphatidylethanolamine, respectively (Yoon and others 1987).
Phospholipids act as antioxidants and prooxidants depending on
their concentration and presence of metals. Oxidation of docosahexaenoic acid at 25 C to 30 C in the dark decreased when
mixed with phosphatiylcholine at a molar ratio of 1:1 (Lyberg
and others 2005).
The mechanism of the antioxidative effects of phospholipids
has not yet been elucidated in detail, but the polar group plays an
important role and the nitrogen-containing phospholipids such as
phosphatidylcholine and phosphatidylethanolamine are efficient
antioxidants under most conditions (King and others 1992). Phospholipids decrease oil oxidation by sequestering metals, and concentration for the maximal antioxidant activity was between 3 and
60 ppm. Soybean oil oxidation decreased with addition of 5 to10
ppm phospholipids, and higher amount of phospholipids acted
as prooxidants. Yoon and Min (1987) reported that phospholipids
acted as antioxidants only in the presence of Fe2+ by chelating
iron. In purified soybean oil, which did not contain any metals, phospholipids worked as prooxidants. Phospholipids have
hydrophilic and hydrophobic groups in the same molecule. The
hydrophilic groups of the phospholipids are on the surface of oil
and hydrophobic group are in the edible oil. The phospholipids
decrease the surface tension of edible oil and increase the diffusion rate of oxygen from the headspace to the oil to accelerate oil

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Mechanisms and factors for edible oil oxidation


oxidation. Soybean oil oxidation in the dark at 60 C was the lowest with phosphatidic acid and phosphatidylethanolamine, followed by phosphatidylcholine, phosphatidylglycerol, and phosphatidylinositol (Yoon and Min 1987).
Chlorophylls. Chlorophylls are common pigments present in
edible vegetable oil. Virgin olive oil and rapeseed oil contain
chlorophyll at 10 ppm and 5 to 35 ppm, respectively (Salvador
and others 2001). Virgin olive oil also contains 4 to 15 ppm pheophytin (Psomiadou and Tsimidou 2002). The chlorophyll concentrations in crude and bleached soybean oils are 0.30 and 0.08
ppm, respectively (Jung and others 1989). Chlorophylls are generally removed during oil processing, especially the bleaching
process (Table 5).
Chlorophylls and their degradation products, pheophytins and
pheophorbides, act as sensitizers to produce 1 O 2 in the presence
of light and atmospheric 3 O 2 , and accelerate the oxidation of oil
(Fakourelis and others 1987; Whang and Peng 1988; GutierrezRosales and others 1992). Pheophytins have a higher sensitizing
activity than chlorophylls, but lower than that of pheophorbides
(Endo and others 1984; Rahmani and Csallany 1998). Soybean oil
purified by silicic acid column chromatography did not contain
any chlorophylls and did not produce volatiles in the headspace
under light at 10 C, but purified soybean oil with added chlorophyll and RBD soybean oil produced headspace volatiles under the same experimental conditions (Lee and Min 1988). Also,
appreciable amounts of volatiles were produced in soybean oil
containing chlorophyll only under light, but not in the dark. The
headspace volatile formation in soybean oil under light at 10 C
increased as the concentration of chlorophyll increased. Rahmani
and Csallani (1998) showed that the oxidation of virgin olive oil
containing pheophytin was accelerated by the illumination of
fluorescent light.
Although chlorophylls are strong prooxidants under light via
acting as a sensitizer to produce 1 O 2 , they act as antioxidants in
the dark possibly by donating hydrogen to free radicals (Endo and
others 1985; Francisca and Isabel 1992).
Thermally oxidized compounds. Processing of crude oils to refined oils is generally performed at high temperature, and it
can produce oxidized compounds such as cyclic and noncyclic
carbon-to-carbon-linked dimers and trimers, hydroxy dimers, and
dimers and trimers joined through carbon-to-oxygen linkage. The
RBD soybean oil contains 1.2% thermally oxidized compounds
(Yoon and others 1988). Thermally oxidized compounds accelerated the soybean oil autoxidation at 55 C (Yoon and others

1988). The acceleration of oil oxidation increases with the concentration of thermally oxidized compounds. Lipid hydroperoxides were also shown to act as prooxidants (Hahm 1988). The
oxidized compounds formed by hydroperoxide decomposition
can act as an emulsifier, which contain both hydrophilic and hydrophobic groups, lower surface tension in the oil, and increase
the introduction of oxygen into the oil to accelerate oil oxidation
(Min and Jung 1989).
Antioxidants. Edible oils naturally contain antioxidants such as
tocopherols, tocotrienols, carotenoids, phenolic compounds, and
sterols. Antioxidants are sometimes intentionally added to oil to
improve oxidative stability. Antioxidants are compounds that extend the induction period of oxidation or slow down the oxidation
rate. Antioxidants scavenge free radicals such as lipid alkyl radicals or lipid peroxy radicals, control transition metals, quench
singlet oxygen, and inactivate sensitizers.
Antioxidants can donate hydrogen atoms to free radicals and
convert them to more stable nonradical products (Decker 2002).
The major hydrogen-donating antioxidants are monohydroxy or
polyhydroxy phenolic compounds with various aromatic ring
substitutions. Any compound whose reduction potential is lower
than that of a free radical can donate hydrogen to that radical unless the reaction is kinetically unfavorable. Standard 1electron reduction potentials of alkoxy, peroxy, and alkyl radicals
of polyunsaturated fatty acids are 1600, 1000, and 600 mV, respectively (Buettner 1993). The standard reduction potential of
antioxidants is generally 500 mV or below. This clearly shows
that antioxidants react with lipid peroxy radicals before the peroxy radicals react with other lipid molecules to produce another
free radical. Any antioxidant radical produced from the reaction
with lipid peroxy radical has lower energy than the lipid peroxy
radical itself due to resonance structure (Figure 11).
Metal chelators such as phosphoric acid, citric acid, ascorbic
acid, and EDTA (ethylenediaminetetraacetic acid) decrease oil
oxidation in an indirect way. They can convert iron or copper
ions into insoluble complexes or can sterically hinder the formation of the complexes between metals and lipid hydroperoxides
(Halliwell and others 1995). Citric acid improved the sensory
quality of soybean oil containing 1 ppm iron during storage at
55 C (Min and Wen 1983). Citric acid is often added to oil in the
field to reduce its oxidation during storage before processing. Min
and Wen (1983) reported that the antioxidant effects of citric acid
increased as its content increased; and more than 150 ppm citric
acid was necessary to overcome the catalytic effect of 1 ppm iron.

Table 5 --- Chlorophyll contents of canola oil during processing (ppm)a


Oil

Chlorophyll a

Pheophytin a

Pheophytin b

Pyropheophytin a

Pyropheophytin b

1.88
0.27
0.22
---

3.31
7.16
6.27
0.56

1.34
1.07
1.12
0.32

16.57
9.40
9.13
0.21

3.13
1.84
1.79
0.25

Extracted
Degummed
Refined
Bleached
a Suzuki

and Nishioka (1993).

OH
ROO

ROOH

Figure 11 --- Resonance


stabilization of an
antioxidant radical

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

179

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


When oil contains 0.1 ppm or less iron such as in RBD oil, the
common practice of adding about 150 ppm citric acid to the oil
for the improvement of oxidative stability is not necessary (Min
and Wen 1983).
Some antioxidants quench 1 O 2 or excited sensitizers. 1 O 2 is
quenched physically and chemically. In physical quenching, 1 O 2
is converted into 3 O 2 by either energy transfer or charge transfer,
and there is no oxidation of antioxidants. In chemical quenching,
antioxidants react with 1 O 2 and produce oxidized antioxidants
(Min and Boff 2002).
Tocopherols

Tocopherols are the most important antioxidants present in edible oil. Animal fats also contain tocopherols, but at lower concentrations (Table 6). Among vegetable oils, soybean, canola,
sunflower, and corn oils contain relatively high amounts of tocopherols. Palm oil does not have large amounts of tocopherols
(118 to 146 ppm); however, it contains a higher concentration
of tocotrienols, that is, -, -, and -tocotrienols at 211, 353 to
372, and 56 to 67 ppm, respectively (Al-Saqer and others 2004).
Safflower oil also contains - and -tocotrienols in addition to
tocopherols at 3.8 to 7.0 and 7.5 to 8.4 ppm, respectively (Lee
and others 2004). Tocopherol contents of edible oils are affected
by the cultivar, processing, and storage of the oil (Deiana and
others 2002). Tocopherol contents of sesame oil range between
404 and 540 ppm depending on the cultivars (Mohamed and
Awatif 1998). The tocopherol contents of rapeseed oil are 794 and
749 ppm for hexane-extracted oil and pressed oil, respectively
(Krygier and Platek 1999). The refining process, especially deodorization, reduces tocopherol contents (Jung and others 1989;
Reische and others 2002; Eidhin and others 2003). The crude,
bleached, and deodorized soybean oil contains tocopherols at
1670, 1467, and 1138 ppm, respectively (Jung and others 1989).
In virgin olive oil, -tocopherol concentration decreased with
storage time of the oil in the dark (Deiana and others 2002).
There was no tocopherol left in olive oil stored in the dark at
room temperature for 12 mo (Morello and others 2004).
Tocopherols compete with unsaturated fats and oils for lipid
peroxy radicals. Lipid peroxy radicals react with tocopherols
much faster at 104 to 109 M1 s1 than with lipids (10 to 60
M1 s1 ). One tocopherol molecule can protect about 103 to
108 polyunsaturated fatty acid molecules at low peroxide value
(Kamal-Elden and Appelqvist 1996). Tocopherols can transfer a

hydrogen atom at the 6-hydroxy group on its chroman ring to lipid


peroxy radical and scavenge the peroxy radicals. Tocopherol (T),
with a reduction potential of 500 mV, donates hydrogen to lipid
peroxy radicals (ROO) that have a reduction potential of 1000
mV, and produces lipid hydroperoxide (ROOH) and tocopheroxy
radicals (T). Tocopheroxy radicals are more stable than lipid peroxy radicals due to resonance structures. This ultimately slows
down the oil oxidation rate in the propagation stage of autoxidation. Reaction rates of peroxy radicals in stearic and oleic acids
with -tocopherol were reported to be 2.8 106 and 2.5 106
M1 s1 , respectively (Simic 1980). Tocopheroxy radicals can interact with other compounds or each other, depending on the
lipid oxidation rates. Tocopheroxy radicals react with each other
at low lipid oxidation rates and produce tocopheryl quinone and
tocopherol. At higher oxidation rates, tocopheroxy radicals may
react with lipid peroxy radicals and produce tocopherollipid peroxy complexes ([T-OOR]); these can be hydrolyzed to tocopheryl
quinone and lipid hydroperoxide (Liebler and others 1990):
T + ROO T + ROOH
T + T T + Tocopheryl quinone
T + ROO [T OOR] Tocopheryl quinone + ROOH
The effectiveness of tocopherols as antioxidants depends on
the isomers and concentration of tocopherols. Free radical scavenging activity of tocopherols is the highest in -tocopherol followed by -, -, and -tocopherol (Reische and others 2002).
-Tocopherol was more effective in reducing oil oxidation than
-tocopherol up to 200 ppm and less effective above this concentration (Yanishlieva and others 2002).
The optimal concentration of tocopherols as antioxidants is
dependent on their oxidative stability; the lower the oxidative
stability of the tocopherol, the lower the optimal concentration
of tocopherol for the maximal antioxidant activity. -Tocopherol,
which is the least stable among tocopherol isomers, showed the
maximal antioxidant activity at 100 ppm in the oxidation of soybean oil at 55 C in the dark, but the optimal concentrations of
- and -tocopherols as an antioxidant were 250 ppm and 500
ppm, respectively (Jung and Min 1990).
Tocopherols, particularly -tocopherol, act as prooxidants
when present in high concentrations in vegetable oils via
propagation of free radicals, depending on the hydroperoxide

Table 6 --- Tocopherol contents of edible oil


Tocopherol (ppm)
Oil
Soybeana

Canolaa
Sunflowera
Corna
Roasted sesameb
Rapeseedc
Safflowerd
Olivee
Beef tallowf
Lardg

a Przybylski (2001).
b Kamal-Eldin and Andersson
c Velasco and others (2003).
d Lee and others (2004)
e Salvador and others (2001);
f Choe and Lee (1998).
g Drinda and Baltes (1999).

180

Total

116.0
272.1
613.0
134.0
4
252
386520
168226
30.4
18.0

34.0
0.1
17.0
18.0

737.0
423.2
18.9
412.0
584
314
2.47.7
--3.8
---

275.0

39.0
9

1162
695.4
648.9
603
597
566
397540
168226
34.2
18.0

8.612.4
-------

(1997).
Gutierrez and others (2001).

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

---------

Mechanisms and factors for edible oil oxidation


concentration (Cillard and others 1980; Terao and Matsushita
1986; Jung and Min 1990). When the concentration of lipid peroxy radical is very low, the tocopheroxy radical abstracts hydrogen from the lipid and produces tocopherol and lipid alkyl
radical; however, the reaction rate is very low. Formation of
lipid alkyl radicals by tocopherols accelerates lipid oxidation,
which is called tocopherol-mediated peroxidation (Bowry and
Stocker 1993; Yamamoto 2001). The highest prooxidant activity
was shown in -tocopherol followed by - and -tocopherol in
soybean oil autoxidation (Jung and Min 1990). Addition of 100
ppm -tocopherol increased the oxidation of purified olive oil at
an early stage of autoxidation; however, -tocopherol added to
moderately oxidized (POV = 15) purified olive oil or lard significantly decreased the oil oxidation (Blekas and others 1995). The
threshold value for -tocopherol as a prooxidant in virgin olive
oil oxidation was 60 to 70 ppm. When less -tocopherol was initially present in the oil, the threshold value for prooxidant activity
was reached more quickly (Deiana and others 2002). Prooxidant
activity of -tocopherol decreased as the temperature increased,
even at a high concentration (Marinova and Yanishlieva 1992).
Ascorbic acid (Asc) can reduce tocopheroxy radical and prevent
tocopherol-mediated peroxidation (Yamamoto 2001):
T + RH T + R
R + O2 ROO
T + Asc T + Dehydroascorbic acid
Oxidized tocopherols increase soybean oil oxidation and
prooxidant activity was the highest in oxidized -tocopherol followed by oxidized - and -tocopherol (Jung and Min 1992).
Prevention of tocopherol oxidation and removal of oxidized tocopherols during oil processing are strongly recommended to
improve oil oxidative stability.
In addition to free radical scavenging activity, tocopherols decrease soybean oil oxidation under light by only 1 O 2 quenching
(Min and Lee 1988). 1 O 2 quenching rate by -tocopherol was
reported to be 2.7 107 M1 s1 (Jung and others 1991). 1 O 2
quenching effect in soybean oil oxidation under light is dependent on tocpherol concentration and the type; at 1 103 M,
the activity was in the decreasing order of -, -, and - tocopherol; however, there was no significant difference in 1 O 2
quenching activity among tocopherols at 4 103 M (Jung and
others 1991). Tocopherols can form a charge transfer complex
([T+ 1 O 2 ] 1 ) with 1 O 2 by electron donation to 1 O 2 . The singlet
state of tocopherol1 O 2 complex undergoes intersystem crossing
into the triplet state ([T+ 1 O 2 ] 3 ) and produces less reactive 3 O 2
and tocopherol. Since this does not involve chemical reactions
between tocopherol and 1 O 2 , it is called physical quenching:
T + 1 O2 [T+ 1 O2 ]1 [T+ 1 O2 ]3 T + 3 O2
Tocopherols react irreversibly with 1 O 2 in chemical quenching and produce tocopherol hydroperoxydienone, tocopheryl
quinone, and quinone epoxide (Figure 12). The reaction rate of
tocopherols with 1 O 2 is affected by their structures. -Tocopherol
showed the highest reaction rate of 2.1 108 M1 s1 , followed
by -tocopherol with 1.5 108 M1 s1 , -tocopherol with 1.4
108 M1 s1 , and -tocopherol with 5.3 107 M1 s1 (Mukai and
others 1991).
Carotenoids

Carotenoids are a group of tetraterpenoids consisting of isoprenoid units. Double bonds in carotenoids are conjugated forms
and usually are all trans forms. -Carotene is one of the most

studied carotenoids. Edible oils, especially unrefined oils, contain -carotene. Crude palm oil and red palm olein contain 500
to 700 ppm carotenoids (Bonnie and Choo 2000). Virgin olive
oil contains 1.0 to 2.7 ppm -carotene as well as 0.9 to 2.3 ppm
lutein (Psomiadou and Tsimidou 2002).
-Carotene can slow down oil oxidation by light filtering, 1 O 2
quenching, sensitizer inactivation, and free radical scavenging.
Fakourelis and others (1987) reported that oxidation of olive oil
containing only -carotene under light at 25 C was decreased
by filtering out some light energy, mainly between 400 and 500
nm. In the co-presence of chlorophylls, -carotene decreased
the oxidation of soybean oil stored under light by 1 O 2 quenching
(Lee and Min 1988). 1 O 2 quenching by carotenoids is mainly
by energy transfer from 1 O 2 to carotenoids, without generating
oxidized products. Excited carotenoids return to the ground state
by giving off heat:
1

O2 + 1 Carotenoid 3 O2 + 3 Carotenoid
3

Carotenoid Carotenoid + Heat

One mole of -carotene can quench 250 to 1000 molecules of


O 2 at a rate of 1.3 1010 M1 s1 (Foote 1976). 1 O 2 quenching
of carotenoids is dependent on the number of conjugated double bonds (Beutner and others 2001). Carotenoids having at least
9 conjugated double bonds act as an effective 1 O 2 quencher;
-carotene, lycopene, and lutein are good 1 O 2 quenchers; however, phytoene, phytofluene, and -carotene, which are precursor
compounds of lycopene, are not. The 1 O 2 quenching activity of
carotenoids increases as the number of conjugated double bonds
increases (Min and Boff 2002; Foss and others 2004). Lycopene
and -carotene demonstrate higher 1 O 2 quenching ability than
-carotene (Miller and others 1996).
Carotenoids inactivate excited sensitizers physically by absorbing energy from sensitizers. The excited carotenoids return to the
ground state by transferring the energy to the oil (Stahl and Sies
1992). Lee and others (2003) reported that -carotene with a
high 1-electron reduction potential of 1060 mV had great difficulty in donating hydrogen to alkyl (E o = 600 mV) or peroxy
radical (E o = 770 to1440 mV) of polyunsaturated fatty acid. Nuclear magnetic resonance study showed that -carotene did not
donate a hydrogen atom to quench these free radicals (Lee and
others 2003). However, -carotene can donate hydrogen to hydroxy radical, which has a high reduction potential (2310 mV),
and produces a carotene radical (Car). A carotene radical is a
fairly stable species due to delocalization of unpaired electrons
through its conjugated polyene system. At low oxygen concentration, a carotene radical may react with other radicals such as
lipid peroxy radicals and form nonradical products (Burton and
Ingold 1984; Beutner and others 2001):
1

Car + HO Car + H2 O
Car + ROO Car OOR
A lipid peroxy radical may be added to -carotene and produce carotene peroxy radical (ROOCar), especially at higher
than 150 mm Hg of oxygen (Burton and Ingold 1984). Carotene
peroxy radical reacts with 3 O 2 and then with lipid molecules, and
produces lipid alkyl radicals, which propagate the chain reaction
of lipid oxidation (Iannone and others 1998):
ROO + Car ROO Car
ROO Car + 3 O2 ROO Car OO
ROO Car OO + R H ROO Car OOH + R

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

181

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


CH3

CH3
HO

O2

H3C

Figure 12 --- Singlet oxygen


oxidation of -tocopherol

H3C

C16H33

O
OOH

C16H33

CH3

CH3
-Tocopherol

Hydroperxydienone

O2
CH3
HO
O

H3C

C16H33

CH3
-Tocopherol endoperoxide

CH3

CH3

O
+

H3C

O
HO

C16H33

H3C

CH3
-Tocopheryl quinone

CH3
-Tocopheryl quinone epoxide

Antioxidant activity of -carotene was not shown in soybean oil


stored in the dark (Lee and others 2003). During photosensitized
oxidation of soybean or rapeseed oil in open vessels, -carotene
increased oil oxidation, but the co-presence of tocopherols decreased oxidation of the oil (Haila and Heinonen 1994). The antioxidant activities of carotenoids by hydrogen donation remain
controversial.
-Carotene may also donate electrons to free radicals and
become a -carotene cation radical (Liebler 1993; Mortensen
and others 2001). The transfer of hydrogen or electrons from
carotenoids to free radicals depends on the reduction potentials
of the free radicals and chemical structures of the carotenoids,
especially the presence of oxygen-containing functional groups
(Edge and others 1997). The -carotene cation radical has a relatively high standard 1-electron reduction potential (1060 mV),
and does not readily react with oxygen to form peroxide (Edge
and others 2000). The -carotene cation radical may react with
alkyl, alkoxy, or peroxy radicals formed in soybean oil during
oxidation. Lee and others (2003) reported that -carotene was a
prooxidant in soybean oil oxidation in the dark due to the electron transfer mechanism in their 2, 2-diphenyl-1-picryl hydrazyl
and NMR study.
182

C16H33

Carotenoids are degraded by hydroperoxides to hydroxyl- or


epoxycarotenes, and the degradation slows down at higher concentration; the rate of degradation of carotenes is lycopene >
-carotene -carotene (Anguelova and Warthesen 2000).
Other phenolic compounds

Phenolic compounds other than tocopherols in edible oils also


exert antioxidant activity. Sesame oil, which contains a high
amount of unsaturated fatty acids (iodine value = 109), has good
oxidative stability (Tan and others 2002). The autoxidation rate of
sesame oils at 60 C was much lower than that of corn oil, safflower oil, and a mixture of soybean and rapeseed oils (Fukuda
and Namiki 1988). Roasted sesame oil was more stable for the
oxidation than unroasted sesame oil (Yoshida and Takagi 1997;
Yoshida and others 2000). The remarkable oxidative stability of
sesame oil is due to the presence of lignan compounds as well as
tocopherols (Yoshida 1994; Namiki 1995). Lignan compounds in
sesame oil include sesamin, sesamol, sesamolin, sesaminol, and
sesamolinol (Figure 13). Sesamin is the predominant lignan compound in unroasted sesame oil (474 ppm) followed by sesamolin
(159 ppm). Sesamol concentration in unroasted sesame oil was
less than 7 ppm (Fukuda and others 1986; Dachtler and others

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Mechanisms and factors for edible oil oxidation


O

Figure 13 --- Phenolic


acids present in oils

O
O
OH

sesamol

sesamin

O
O
O

O
O

sesamolin

O
O
O

HO

sesaminol

OH
HO

OH

COOH

HO
HO

Caffeic acid

OH

Hydroxytyrosol

2003); however, roasted sesame oil contains a higher concentration of sesamol (36 ppm) than unroasted oil (Kim and Choe
2005). Sesamol is produced by hydrolysis of sesamolin during
oil processing such as heating and bleaching (Fukuda and others
1985; Osawa and others 1985; Kim and Choe 2005). Sesamol is
converted to sesamol dimer and then to sesamol dimer quinone
(Kurechi and others 1981; Kikugawa and others 1983). Sesamol
and sesaminol showed higher antioxidant activity than sesamin
in sunflower oil autoxidation by scavenging radicals (Dachtler
and others 2003).
Sesamol acts as an antioxidant in the oxidation under light as
well as in the dark. The antioxidation by sesamol in chlorophyllsensitized photooxidation of soybean oil was lower than by -

HO

Tyrosol

tocopherol, similar to -tocopherol, and higher than DABCO (diazabicyclo[2,2,2]octane) at the same molar concentration (Kim
and others 2003). The decrease of photooxidation of soybean oil
by sesamol was due to its 1 O 2 quenching, and the quenching rate
was 1.9 107 M1 s1 at 20 C (Kim and others 2003).
Sesame oil also contains phytosterols such as campesterol, stigmasterol, -sitosterol, and 4,5-avenasterol, with -sitosterol as
the predominant sterol (Dachtler and others 2003). Sitosterol behaves partly as a prooxidant by increasing the solubility of oxygen
in the oil (Yanishlieva and Schiller 1983) and partly as a weak
antioxidant in sunflower oil and lard by competing with lipid
molecules for oxidation at the oil surface (Maestroduran and Borjapadilla 1993; Brimberg and Kamal-Eldin 2003).

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

183

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


Table 7 --- Phenolic compounds in virgin olive oila
Phenolic alcohols

3,4-Dihydroxyphenyl ethanol (3,4-DHPEA, hydroxytyrosol), p-hydroxyphenylethanol (p-HPEA, tyrosol),


3,4-dihydroxyphenylethanol-glucoside
Phenolic acid and its derivatives Vanillic acid, syringic acid, p-coumaric acid, o-coumaric acid, gallic acid, caffeic acid, protocatechuic acid,
p-hydroxybenzoic acid, ferulic acid, cinnamic acid, 4-(acetoxyethyl)-1,2-dihydroxybenzene, benzoic acid
Secoiridoids
Dialdehydic form of elenolic acid linked to 3,4-DHPEA, dialdehydic form of elenolic acid linked to p-HPEA,
oleuropein aglycon, ligstroside aglycon, oleuropein, p-HPEA derivative
Lignans
1-Acetoxypinoresinol, pinoresinol, 1-hydroxypinoresinol
Flavones
Apigenin, luteolin
a Servili

and Montedoro (2002).

Olive oil, which is very stable to autoxidation (Guillen


and Cabo 2002), contains phenolic compounds and tocopherols. Phenolic compounds in olive oil include tyrosol
(4-hydroxyphenylethanol),
hydroxytyrosol
(3,4-dihydroxyphenylethanol), hydroxybenzoic acids, oleuropein, caffeic acid,
vanillic acid, p-coumaric acid, and derivatives of tyrosol and
hydroxytyrosol (Papadopoulos and Boskou 1991; Tsimidou
and others 1992), as shown in Table 7. Contents of tyrosol,
hydroxytyrosol, and phenolic acids in olive oil were 34.9, 37.8,
and 36.3 ppm, respectively (Keceli and Gordon 2002). The
phenolic compounds in olive oil acted as antioxidants mainly
at the initial stage of autoxidation (Deiana and others 2002) by
scavenging free radicals and chelating metals (Chimi and others
1991). Hydroxytyrosol was the most effective antioxidant in
olive oil oxidation (Papadopoulos and Boskou 1991; Tsimidou
and others 1992; Baldioli and others 1996). Among phenolic
compounds, o-diphenols such as caffeic acid are oxidized to
quinones by ferric ions and become ineffective in inhibiting
iron-dependent free radical chain reactions in oil (Keceli and
Gordon 2002). However, hydroxytyrosol, tyrosol, vanillic acids,
and p-coumaric acid were not oxidized by the ferric ions. Fki
and others (2005) reported high radical scavenging effects of
3,4-dihydroxyphenyl acetic acid present in olive mill wastewater
on the oxidation of olive and husk refined oils. Chlorogenic acid
and caffeic acid are the principal phenols found in sunflower oil
(Leonardis and others 2003). Osborne and Akoh (2003) reported
prooxidant effects of polar phenolic acids such as quercetin and
gallic acid on a canola oil and caprylic acid structured lipid in
the presence of iron at pH 3.0, due to an increased ability to
reduce iron.

ducing sugar of phosphatidylinositol, donates hydrogen or electron to tocopherols and delays the oxidation of tocopherols to tocopheryl quinone (Yoon and Min 1987). Sesamol and sesaminol
showed synergistic antioxidant activities with -tocopherol in the
autoxidation of sunflower oil (Dachtler and others 2003).
A combination of 11 ppm -carotene, 30 ppm citric acid, 3
to 4 ppm Tinuvin P, and 150 to 200 ppm TBHQ decreased the
oxidation rate of soybean oil during storage at 25 C under light
for 6 mo (Azeredo and others 2004).
Synergism between antioxidants is affected by hydroperoxide concentration. Olive oil autoxidation was lower when tocopherol and phenolic compounds were used together than
when used separately (Velasco and Dobarganes 2002); however,
-tocopherol exerted an adverse effect on the antioxidation of
olive oil by phenolic compounds such as tyrosol or oleuropein
during the period of low hydroperoxide formation, for example,
less than 20 meq (Blekas and others 1995).

Conclusions
Edible oil undergoes autoxidation and photosensitized oxidation during processing and storage. The oxidation of edible
oil produces off-flavor compounds and decreases oil quality.
Free fatty acids, mono- and diacylglycerols, metals, chlorophylls,
carotenoids, tocopherols, phospholipids, temperature, light, oxygen, oil processing methods, and fatty acid composition affect
the oxidative stability of edible oil. To minimize the oxidation of
edible oil during processing and storage, it is recommended to
decrease temperature, exclude light and oxygen, remove metals
and oxidized compounds, and use appropriate concentrations of
antioxidants such as tocopherols and phenolic compounds.

Antioxidant interactions

Edible oil often contains multicomponent antioxidants and REFERENCES


demonstrates interactions among them. Metal chelator and a free Aho L, Wahlroos O. 1967. A comparison between determinations of the solubility of oxygen
in oils by exponential dilution and chemical methods. J Am Oil Chem Soc 66:4649.
radical scavenger demonstrate better antioxidant activities in oil Aidos
I, Lourenco S, van der Padt A, Luten JB, Boom RM. 2002. Stability of crude herring
autoxidation when present together than when used separately. oil produced from fresh byproducts: influence of temperature during storage. J Food Sci
67:331420.
The improved antioxidation of oil by a combination of metal
JM, Sidhu JS, Al-Hooti SN, Al-Amiri HA, Al-Othman A, Al-Haji L, Ahmed N, Mansour
chelator and free radical-scavenging antioxidant is caused mainly Al-Saqer
IB, Minal J. 2004. Developing functional foods using red palm olein. IV. Tocopherols and
by the action of the chelator at the initiation step of oxidation and tocotrienols. Food Chem 85:57983.
Andersson K. 1998. Influence of reduced oxygen concentrations on lipid oxidation in food
that of the free radical scavenger at the propagation step.
during storage [Ph.D thesis]. Chalmers Reproservice, Sweden: Chalmers University of TechBecause tocopherols are the most common antioxidants in ed- nology and the Swedish Institute for Food and Biotechnology.
Andersson
K, Lingnert H. 1999. Kinetic studies of oxygen dependence during initial lipid
ible oil, they have been used in research on interactions among
oxidation in rapeseed oil. J Food Sci 64:2626.
antioxidants. -Tocopherol showed synergistic effects with - Anguelova
T, Warthesen J. 2000. Degradation of lycopene, -carotene, and -carotene during
carotene to decrease the autoxidation (Palozza and Krinsky 1992) lipid peroxidation. J Food Sci 65:715.
Azeredo
HMC,
Faria JAF, Silva MAAP. 2003. The efficiency of TBHQ, -carotene, citiric
and photosensitized oxidation of soybean oil (Choe and Min
and tinuvin 234 on the sensory stability of soybean oil packaged in PET bottles. J
1992). The synergistic effects of tocopherol and -carotene were acid,
Food Sci 68:3026.
suggested to be due to the protection of -carotene from degra- Azeredo HMC, Faria JAF, Silva MAAP. 2004. Minimization of peroxide formation rate in
dation by tocopherols (Shibasaki-Kitakawa and others 2004). - soybean oil by antioxidant combinations. Food Res Int 37:68994.
M, Servili M, Perretti G, Montedoro GF. 1996. Antioxidant activity of tocopherols
Tocopherol shows synergism with ascorbic acid and phospho- Baldioli
and phenolic compounds of virgin olive oil. J Am Oil Chem Soc 73:158993.
lipids. Ascorbic acid gives hydrogen to the tocopheroxy radical Benjelloun B, Talou T, Delmas M, Gaset A. 1991. Oxidation of rapeseed oil: effect of metal
produced from the reaction between the lipid peroxy radical and traces. J Am Oil Chem Soc 68:2101.
S, Bloedorn B, Frixel S, Blanco IH, Hoffmann T, Martin H-D, Mayer B, Noack P,
tocopherol, and then regenerates tocopherol. Nitrogen moiety of Beutner
Ruck C, Schmidt M, Schulke I, Sell S, Ernst H, Haremza S, Seybold G, Sies H, Stahl W,
phosphatidylethanolamine and phosphatidylcholine, or the re- Walsh R. 2001. Quantitative assessment of antioxidant properties of natural colorants and
184

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

Mechanisms and factors for edible oil oxidation


phytochemicals: carotenoids, flavonoids, phenols and indigoids. The role of -carotene in
antioxidant functions. J Sci Food Agr 81:55968.
Blekas G, Tsimidou M, Boskou D. 1995. Contribution of -tocopherol to olive oil stability.
Food Chem 52:28994.
Bonnie TYP, Choo YM. 2000. Valuable minor constituents of commercial red palm olein:
carotenoids, vitamin E, ubiquinone and sterols. J Oil Palm Res 12:1424.
Bowry VW. Stocker R. 1993. Tocopherol-mediated peroxidation. The prooxidant effect of
vitamin E on the radical-initiated oxidation on human low-density lipoprotein. J Am Chem
Soc 115:602944.
Brimberg UI, Kamal-Eldin A. 2003. On the kinetics of the autoxidation of fats: influence of
pro-oxidants, antioxidants and synergists. Eur J Lipid Sci Technol 105:8391.
Buettner GR. 1993. The pecking order of free radicals and antioxidants: lipid peroxidation,
-tocopherol and ascorbate. Arch Biochem Biophys 300:53543.
Burton GW, Ingold KU. 1984. -Carotene: an unusual type of lipid antioxidant. Science
224:56973.
Chang SS, Smouse YH, Krishnamurthy RG, Mookherjee BD, Reddy RB. 1966. Isolation and
identification of 2-pentyl-furan as contributing to the reversion flavour of soybean oil. Chem
Ind 11:19267.
Chang SS, Shen GH, Tang J, Jin QZ, Shi H, Carlin JT, Ho CT. 1983. Isolation and identification
of 2-pentenylfurans in the reversion flavor of soybean oil. J Am Oil Chem Soc 60:5537.
Chen Y, Xu S, Li L, Zhang M, Shen J, Shen T. 2001. Active oxygen generation and photooxygenation involving temporin (m-THPC). Dyes Pigments 51:639.
Chimi H, Cillard J, Cillard P, Rahmani M. 1991. Peroxyl and hydroxyl radical scavenging
activity of some natural phenolic antioxidants. J Am Oil Chem Soc 68:30712.
Choe E. 1997. Effects of heating time and storage temperature on the oxidative stability of
heated palm oil. Korean J Food Sci Technol 29:40711.
Choe E, Lee J. 1998. Thermooxidative stability of soybean oil, beef tallow and palm oil during
frying of steamed noodles. Korean J Food Sci Technol 30:28892.
Choe E, Min DB. 1992. Interaction effects of chlorophyll, -carotene and tocopherol on the
photooxidative stabilities of soybean oil. Foods and Biotechnol 1:10410.
Cillard J, Cillard P, Cormier M. 1980. Effect of experimental factors on the prooxidant behavior
of -tocopherol. J Am Oil Chem Soc 57:25561.
Crowe TD, White PJ. 2003. Oxidative stability of walnut oils extracted with supercritical
carbon dioxide. J Am Oil Chem Soc 80:5758.
Dachtler M, FVan dePut HM, Stijn FV, Beindorff CM, Fritsche J. 2003. On-line LC-NMR-MS
characterization of sesame oil extracts and assessment of their antioxidant activity. Eur J
Lipid Sci Technol 105:48896.
Decker EA. 2002. Antioxidant mechanisms. In: Akoh CC, Min DB, editors. Food lipids. New
York: Marcel Dekker. p 51742.
Deiana M, Rosa A, Cao CF, Pirisi FM, Bandino G, Dessi MA. 2002. Novel approach to study
oxidative stability of extra virgin olive oils: importance of -tocopherol concentration. J
Agric Food Chem 50:43426.
Drinda H, Baltes W. 1999. Antioxidant properties of lipoic and dihydrolipoic acid in vegetable
oils and lard. Z Lebensm Unters For 4:2706.
Edge R, Garvey MC, Truscott TG. 1997. The carotenoids as antioxidants a review. J Photochem Photobiol B: Biol 41:189200.
Edge R, Land EJ, McGarvey DJ, Burke M, Truscott TG. 2000. The reduction potential of the
-carotene+ /-carotene couple in an aqueous micro-heterogeneous environment. FEBS
Letters 471:1257.
Eidhin DN, Burke J, OBeirne. 2003. Oxidative stability of 3-rich camelina oil and camelina
oil-based spread compared with plant and fish oils and sunflower spread. J Food Sci 68:345
53.
Endo Y, Usuki R, Kaneda T. 1984. Prooxidant activities of chlorophylls and their decomposition products on the photooxidation of methyl linoleate. J Am Oil Chem Soc 61:7814.
Endo Y, Usuki R, Kaneda T. 1985. Antioxidant effects of chlorophyll and pheophytin on the
autoxidation of oils in the dark. I. Comparison of the inhibitory effects. J Am Oil Chem Soc
62:13758.
Evans CD, List GR, Moser HA, Cowan JC. 1973. Long-term storage of soybean and cottonseed
salad oils. J Am Oil Chem Soc 50:21822.
Fakourelis N, Lee EC, Min DB. 1987. Effects of chlorophyll and -carotene on the oxidation
stability of olive oil. J Food Sci 52:2346.
Fki I, Allouche N, Sayadi S. 2005. The use of polyphenolic extract, purified hydroxytyrosol
and 3,4-dihydroxyphenyl acetic acid from olive mill wastewater for the stabilization of
refined oils: a potential alternative to synthetic antioxidants. Food Chem 93:197204.
Foote C. 1976. Photosensitized oxidation and singlet oxygen: consequences in biological
systems. In: Pryor WA, editor. Free radicals in biology. New York: Academic Press. p 85
133.
Foss BJ, Silwka H-R, Partali V, Caedounel AJ, Zweier JL, Lockwood SF. 2004. Direct superoxide anion scavenging by a highly water-dispersible carotenoid phospholipids evaluated
by electron paramagnetic resonance (EPR) spectroscopy. Bioorg Med Chem Lett 14:2807
12.
Francisca G, Isabel M. 1992. Action of chlorophylls on the stability of virgin olive oil. J Am
Oil Chem Soc 69:86671.
Frankel EN. 1985. Chemistry of autoxidation: mechanism, products and flavor significance.
In: Min DB, Smouse TH, editors. Flavor chemistry of fats and oils. Champaign, Ill.: American
Oil Chemists Society. p 134.
Fukuda Y, Namiki M. 1988. Recent studies on sesame seed and oil. Nippon Shokuhin Kogyo
Gakkaishi 35:55260.
Fukuda Y, Nagata M, Osawa T, Namiki M. 1986. Contribution of lignan analogues to antioxidative activity of refined unroasted sesame seed oil. J Am Oil Chem Soc 63:102731.
Fukuda Y, Osawa T, Namiki M, Ozaki T. 1985. Studies on antioxidative substances in sesame
seed. Agric Biol Chem 49:3016.
Girotti AW. 1998. Lipid hydroperoxide generation, turnover, and effector action in biological
systems. J Lipid Res 39:152942.
Gollnick K. 1978. Mechanism and kinetics of chemical reactions of singlet oxygen with
organic compounds. In: Ranby B, Rabek JF, editors. Singlet oxygen. New York: John Wiley
& Sons. p 11134.
Guillen MD, Cabo N. 2002. Fourier transform infrared spectra data versus peroxide and
anisidine values to determine oxidative stability of edible oils. Food Chem 77:50310.

Gutierrez F, Arnaud T, Garrido A. 2001. Contribution of polyphenols to the oxidative stability


of virgin olive oil. J Sci Food Agric 81:146370.
Gutierrez-Rosales F, Garrido-Femandez J, Gallardo-Guerrero L, Gandul-Rojas B, MinguezMosquera MI. 1992. Action of chlorophylls and the stability of virgin olive oil. J Am Oil
Chem Soc 69:86671.
Hahm TS. 1988. Effects of initial peroxide contents on the oxidative stability of soybean oil
to prevent environmental pollution. J Environ Research 1:11220.
Haila K, Heinonen M. 1994. Action of -carotene on purified rapeseed oil during light storage.
Food Sci Technol 27:5737.
Halliwell B, Gutteridge JMC. 2001. Free radicals in biology and medicine. 3rd ed. New York:
Oxford University Press. p 297.
Halliwell B, Murcia MA, Chirico S, Aruoma OI. 1995. Free radicals and antioxidants in food
and in vivo: What they do and how they work. Crit Rev Food Sci Nutr 35:720.
Hamilton RJ. 1994. The chemistry of rancidity in foods. In: Allen JC, Hamilton RJ, editors.
Rancidity in foods. 3rd ed. London: Blackie Academic & Professional. p 121.
Heinonen M, Haila K, Lampi AM, Piironen V. 1997. Inhibition of oxidation in 10% oil-inwater emulsions by carotene with and tocopherols. J Am Oil Chem Soc 74:1047
52.
Hiatt R, Mill T, Irwin KC, Mayo TR, Gould CW, Castleman JK. 1968. Homolytic decomposition
of hydroperoxides. J Org Chem 33:141641.
Ho CT, Smagula MS, Chang SS. 1978. The synthesis of 2-(1-pentenyl) furan and its relationship
to the reversion flavor of soybean oil. J Am Oil Chem Soc 55:2337.
Iannone A, Rota C, Bergamini S, Tomasi A, Canfield LM. 1998. Antioxidant activity of
carotenoids: an electron-spin resonance study on -carotene and lutein interaction with
free radicals generated in a chemical system. J Biochem Mol Toxicol 12:299304.
Jadhav SJ, Nimbalkar SS, Kulkami AD, Madhavi DL. 1996. Lipid oxidation in biological and
food systems. In: Madhavi DL, Deshpande SS, Salunkhe DK, editors. Food antioxidants.
New York: Marcel Dekker. p 564.
Jung MY, Min DB. 1990. Effects of -, -, and -tocopherols on the oxidative stability of
soybean oil. J Food Sci 55:14645.
Jung MY, Min DB. 1992. Effects of oxidized -, - and -tocopherols on the oxidative stability
of purified soybean oil. Food Chem 45:1837.
Jung MY, Yoon SH, Min DB. 1989. Effects of processing steps on the contents of minor
compounds, and oxidation stability of soybean oil. J Am Oil Chem Soc 66:11820.
Jung MY, Choe E, Min DB. 1991. Effects of -, -, -, and -tocopherols on the chlorophyll
photosensitized oxidation of soybean oil. J Food Sci 56:80715.
Kacyn LJ, Saguy I, Karel M. 1983. Kinetics of oxidation of dehydrated food at low oxygen
pressures. J Food Proc Preserv 7:16178.
Kamal-Eldin A, Andersson R. 1997. A multivariate study of the correlation between tocopherol
content and fatty acid composition in vegetable oils. J Am Oil Chem Soc 74:37580.
Kamal-Eldin A, Appelqvist L-A. 1996. The chemistry and antioxidant properties of tocopherols
and tocotrienols. Lipids 31:671701.
Kanavouras A, Cert A, Hernandez RJ. 2005. Oxidation of olive oil under still air. Food Sci
Technol Inter 11:1839.
Karel M. 1992. Kinetics of lipid oxidation. In: Schwarzberg HG, Hartel RW, editors. Physical
chemistry of foods. New York: Marcel Dekker. p 65168.
Keceli T, Gordon MH. 2002. Ferric ions reduce the antioxidant activity of the phenolic fraction
of virgin olive oil. J Food Sci 67:9437.
Kellogg EW, Fridovich I. 1975. Superoxide, hydrogen peroxide, and singlet oxygen in lipid
peroxidation by a xanthine oxidase system. J Biol Chem 250:88127.
Kikugawa K, Arai M, Kurechi T. 1983. Participation of sesamol in stability of sesame oil. J
Am Oil Chem Soc 60:152833.
Kim I, Choe E. 2005. Effects of bleaching on the properties of roasted sesame oil. J Food Sci
70:C4852.
Kim JY, Choi DS, Jung MY. 2003. Antiphoto-oxidative activity of sesamol in methylene blueand chlorophyll-sensitized photo-oxidation of oil. J Agric Food Chem 51:34605.
King MF, Boyd LC, Sheldon BW. 1992. Effects of phospholipids on lipid oxidation of a salmon
oil model system. J Am Oil Chem Soc 69:23742.
Kinsella JE, Shimp JL, Mai J. 1978. The proximate composition of several species of freshwater
fishes. Food Life Sci Bull 69:120.
Kochevar IE, Redmond RW. 2000. Photosensitized production of singlet oxygen. Methods
Enzymol 319:208.
Korycka-Dahl MB, Richardson T. 1978. Activated oxygen species and oxidation of food
constituents. Crit Rev Food Sci Nutr 10:20940.
Krygier K, Platek T. 1999. Comparison of pressed and extracted rapeseed oils characteristics. Proceedings of the 10th International Rapeseed Congress. Canberra. Australia.
http://www.regional.org.au/au/gcirc/6/641.htm
Kurechi T, Kikugawa K, Aoshima S. 1981. Studies on the antioxidants. XIV. Reaction of
sesamol with hydrogen peroxide-peroxidase. Chem Pharm Bull 29:23518.
Labuza TP. 1971. Kinetics of lipid oxidation in foods. CRC Crit Rev Food Technol 2:355
405.
Lee EC, Min DB. 1988. Quenching mechanism of beta-carotene on the chlorophyll-sensitized
photooxidation of soybean oil. J Food Sci 53:18945.
Lee JH, Ozcelik B, Min DB. 2003. Electron donation mechanisms of -carotene as a free
radical scavenger. J Food Sci 68:8615.
Lee SH, Min DB. 1991. Effects, quenching mechanisms, and kinetics of nickel chelates in
singlet oxygen oxidation of soybean oil. J Agric Food Chem 39:6426.
Lee YC, Oh SW, Chang J, Kim IH. 2004. Chemical composition and oxidative stability of
safflower oil prepared from safflower seed roasted with different temperatures. Food Chem
84:16.
Leonardis AD, Macciola V. 2002. Catalytic effect of the Cu(II)- and Fe(III)cyclohexanebutyrates on olive oil oxidation measured by Rancimat. Eur J Lipid Sci Technol
104:15660.
Leonardis AD, Macciola V, Di Rocco A. 2003. Oxidative stabilization of cold-pressed sunflower oil using phenolic compounds of the same seeds. J Sci Food Agric 83:5238.
Liebler DC. 1993. Antioxidant reactions of carotenoids. Ann NY Acad Sci 691:2031.
Liebler DC, Baker PF, Kaysen KL. 1990. Oxidation of vitamin E: evidence for competing
autoxidation and peroxyl radical trapping reaction of the tocopheroxyl radical. J Am Chem
Soc 112:69957000.

Vol. 5, 2006COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETY

185

CRFSFS: Comprehensive Reviews in Food Science and Food Safety


Liu Q, Singh S, Green A. 2002. High-oleic and high-stearic cottonseed oils: nutritionally
improved cooking oils developed using gene silencing. J Am Coll Nutr 21:205S11S.
Long CA, Kearns DR. 1975. Radiationless decay of singlet molecular oxygen in solution. II.
Temperature dependence and solvent effects. J Am Chem Soc 97:201820.
Lyberg A-M, Fasoli E, Adlercreutz. 2005. Monitoring the oxidation of docosahexaenoic acid
in lipids. Lipids 40:96979.
Maestroduran R, Borjapadilla R. 1993. Antioxidant activity of natural sterols and organic
acids. Grasas Aceites 44:20812.
Marinova EM, Yanishlieva NV. 1992. Effect of temperature on the antioxidative action of
inhibitors in lipid autoxidation. J Sci Food Agric 60:3138.
Marquez-Ruiz G, Martin-Polvillo M, Dobarganes MC. 1996. Quantitation of oxidized triglyceride monomers and dimers as a useful measurement for early and advanced stages of
oxidation. Grasaa Aceites 47:4853.
Martin-Polvillo M, Marquez-Ruiz G, Dobarganes MC. 2004. Oxidative stability of sunflower
oils differing in unsaturation degree during long-term storage at room temperature. J Am
Oil Chem Soc 81:57783.
Mei L, McClements DJ, Wu J, Decker EA. 1998. Iron-catalyzed lipid oxidation in emulsions
as affected by surfactant, pH, and NaCl. Food Chem 61:30712.
Merkel PB, Kearns DR. 1972. Radiationless decay of singlet molecular oxygen in solution. An
experimental and theoretical study of electronic to vibrational energy transfer. J Am Chem
Soc 94:724453.
Miller NJ, Sampson J, Candeias LP, Bramley PM, Rice-Evans CA. 1996. Antioxidant activities
of carotenes and xanthophylls. FEBS Lett 384:2402.
Min DB, Boff JM. 2002. Lipid oxidation of edible oil. In: Akoh CC, Min DB, editors. Food
lipids. New York: Marcel Dekker. p 33563.
Min DB, Bradley GD. 1992. Fats and oils: flavors. In: Hui YH, editor. Wiley encyclopedia of
food science and technology. New York: John Wiley & Sons. p 82832.
Min DB, Jung MY. 1989. Effects of minor components on the flavor stability of vegetable oils.
In: Min DB, Smouse TH, editors., Flavor chemistry of lipid foods. Champaign, Ill.: AOCS
Press. p 24264.
Min DB, Lee EC. 1988. Factors affecting singlet oxygen oxidation of soybean oil. In: Charalambous G, editor. Frontiers of flavor. New York: Elsevier. p 47398.
Min DB, Wen J. 1983. Effects of citric acid and iron levels on the flavor quality of oil. J Food
Sci 48:7913.
Min DB, Wen J. 1983. Effects of dissolved free oxygen on the volatile compounds of oil. J
Food Sci 48:142930.
Min DB, Callison AL, Lee HO. 2003. Singlet oxygen oxidation for 2-pentylfuran and 2pentenylfuran formation in soybean oil. J Food Sci 68:11758.
MAFF. 1997. Ministry of Agriculture, Fisheries and Food of United Kingdom. Food surveillance information sheet 138. Metals in cold-pressed oils. http://archive.food.gov.uk/maff
Mistry BS, Min DB. 1987a. Effects of fatty acids on the oxidative stability of soybean oil. J
Food Sci 52:8312.
Mistry BS, Min DB. 1987b. Isolation of sn-a-monolinolein from soybean oil and its effects on
oil oxidative stability. J Food Sci 52:78690.
Mistry BS, Min DB. 1988. Prooxidant effects of monoglycerides and diglycerides in soybean
oil. J Food Sci 53:18967.
Miyashitau K, Takagi T. 1986. Study on the oxidative rate and prooxidant activity of free fatty
acids. J Am Oil Chem Soc 63:13804.
Mohamed HMA, Awatif II. 1998. The use of sesame oil unsaponifiable matter as a natural
antioxidants. Food Chem 62:26976.
Morello JR, Motilva MJ, Tovar MJ, Romero MP. 2004. Changes in commercial virgin olive oil
(cv Arbequina) during storage, with special emphasis on the phenolic fraction. Food Chem
85:35764.
Mortensen A, Skibsted LH, Truscott TG. 2001. The interaction of dietary carotenoids with
radical species. Arch Biochem Biophys 385:139.
Mukai K, Daifuku K, Okabe K, Tanigaki T, Inoue K. 1991. Structure-activity relationship in
the quenching reaction of singlet oxygen by tocopherol (vitamin E) derivatives and related
phenols. Finding of linear correlation between the rates of quenching of singlet oxygen and
scavenging of peroxyl and phenoxyl radicals in solution. J Org Chem 56:418892.
Namiki N. 1995. The chemistry and physiological functions of sesame. Food Rev Int 11:281
329.
Osawa T, Nagata M, Namiki M, Fukuda Y. 1985. Sesamolinol, a novel antioxidant isolated
from sesame seeds. Agric Biol Chem 49:33512.
Osborne HT, Akoh CC. 2003. Effects of natural antioxidants on iron-catalyzed lipid oxidation
of structured lipid-based emulsions. J Am Oil Chem Soc 80:84752.
Palozza P, Krinsky NI. 1992. -Carotene and -tocopherol are synergistic antioxidants. Arch
Biochem Biophys 297:1847.
Papadopoulos G, Boskou D. 1991. Antioxidant effect of natural phenols on olive oil. J Am
Oil Chem Soc 68:66971.
Parker TD, Adams DA, Zhou K, Harris M, Yu L. 2003. Fatty acid composition and oxidative
stability of cold-pressed edible seed oils. J Food Sci 68:12403.
Pascall MA, Harte BR, Giacin JR, Gray JI. 1995. Decreasing lipid oxidation in soybean oil by
a UV-absorber in the packaging material. J Food Sci 60:11169.
Przybylski R. 2001. Canola oil: physical and chemical properties. Canola Council of Canada.
pp 112.
Przybylski T, Eskin NAM. 1988. A comparative study on the effectiveness of nitrogen or
carbon dioxide flushing in preventing oxidation during the heating of oil. J Am Oil Chem
Soc 65:62933.
Przybylski R, Eskin NAM. 1995. Methods to measure volatile compounds and the flavor
significance of volatile compounds. In: Warner K, Eskin NAM, editors. Methods to assess quality and stability of oils and fat-containing foods. Champaign, Ill.: American Oil
Chemists Society. p 10733.
Psomiadou E, Tsimidou M. 2002. Stability of virgin olive oil. 2. Photo-oxidation studies. J
Agric Food Chem 50:7227.
Rahmani M, Saari Csallany A. 1998. Role of minor constituents in the photooxidation of
virgin olive oil. J Am Oil Chem Soc 75:83743.

186

Rawls HR, Van Santen PJ. 1970. A possible role for singlet oxidation in the initiation of fatty
acid autoxidation. J Am Oil Chem Soc 47:1215.
Reische DW, Lillard DA, Eitenmiller RR. 2002. Antioxidants. In: Akoh CC, Min DB, editors.
Food lipids. New York: Marcel Dekker. p 489516.
Salvador MD, Aranda F, Gomez-Alonso S, Fregapane G. 2001. Cornicabra virgin olive oil:
a study of five crop seasons. Composition, quality and oxidative stability. Food Chem
74:26774.
Sattar A, DeMan JM, Alexander JC. 1976. Effect of wavelength on light-induced quality deterioration of edible oils and fats. Can Inst Food Sci Technol J 9:10813.
Servili M, Montedoro GF. 2002. Contribution of phenolic compounds to virgin olive oil
quality. Eur J Lipid Sci Technol 104:60213.
Shahidi F, Spurvey SA. 1996. Oxidative stability of fresh and heated-processed dark and light
muscles of mackerel (Scomber scombrus). J Food Lipids 3:1325.
Shibasaki-Kitakawa N, Kato H, Takahashi A, Yonemoto T. 2004. Oxidation kinetics of carotene in oleic acid solvent with addition of an antioxidant, -tocopherol. J Am Oil
Chem Soc 81:38994.
Shiota M, Uchida T, Oda T, Kawakami H. 2006. Utilization of lactoferrin as an iron-stabilizer
for soybean and fish oil. J Food Sci 71:C1203.
Silva FAM, Borges F, Ferreira MA. 2001. Effects of phenolic propyl esters on the oxidative
stability of refined sunflower oil. J Agric Food Chem 49:393641.
Simic MG. 1980. Kinetic and mechanistic studies of peroxy, vitamin E, and anti-oxidant free
radicals by pulse radiolysis. In: Simic MG, Karel M, editors. Autoxidation in food biological
systems. New York: Plenum Press. p 1726.
Sleeter RT. 1981. Effect of processing on quality of soybean oil. J Am Oil Chem Soc 58:239
47.
Smagula MS, Ho CT, Chang SS. 1979. The synthesis of 2-(2-pentenyl) furans and their relationship to the reversion flavor of soybean oil. J Am Oil Chem Soc 56:5169.
Smouse TH, Chang SS. 1967. A systematic characterization of the reversion flavor of soybean
oil. J Am Oil Chem Soc 44:50914.
St. Angelo AJ. 1996. Lipid oxidation in foods. Crit Rev Food Sci Nutr. 36:175224.
Stahl W, Sies H. 1992. Physical quenching of singlet oxygen and cis-trans isomerization of
carotenoids. Ann NY Acad Sci 691:109.
Steenson DFM, Lee JH, Min DB. 2002. Solid-phase microextraction of volatile soybean oil
and corn oil compounds. J Food Sci 67:716.
Suzuki K, Nishioka A. 1993. Behavior of chlorophyll derivatives in canola oil processing. J
Am Oil Chem Soc 70:83741.
Tan CP, Che Man YB, Selamat J, Yusoff MSA. 2002. Comparative studies of oxidative stability
of edible oils by differential scanning calorimetry and oxidative stability index methods.
Food Chem 76:3859.
Terao J, Matsushita S. 1986. The peroxidizing effect of -tocopherol on autoxidation of methyl
linoleate in bulk phase. Lipids 21:25560.
Tsimidou M, Papadopoulos G, Boskou D. 1992. Phenolic compounds and stability of virgin
olive oilPart I. Food Chem 45:1414.
Vaisey-Genser M, Malcomson LJ, Przybylski R, Eskin NAM. 1999. Consumer acceptance of
stored canola oils in canola. 12th Project report research on canola, seed, oil meal. Canada
Canola Council. Canada. p 189202.
Velasco J, Dobarganes C. 2002. Oxidative stability of virgin olive oil. Eur J Lipid Sci Technol
104:66176.
Velasco J, Andersen ML, Skibsted LH. 2003. Evaluation of oxidative stability of vegetable oils
by monitoring the tendency to radical formation. A comparison of electron spin resonance
spectroscopy with the Rancimat method and differential scanning calorimetry. Food Chem
77:62332.
Vever-Bizet C, Dellinger M, Brault D, Rougee M, Bensasson RV. 1989. Singlet molecular
oxygen quenching by saturated and unsaturated fatty-acids and by cholesterol. Photochem
Photobiol 50:3215.
Warner K, Evans CD, List GR, Dupuy HP, Wadsworth JI, Goheen GE. 1978. Flavor score
correlation with pentanal and hexanal contents of vegetable oil. J Am Oil Chem Soc 55:252
6.
Whang K, Peng IC. 1988. Electron paramagnetic resonance studies of the effectiveness of
myoglobin and its derivatives as photosensitizers in singlet oxygen generation. J Food Sci
53:18635, 1893.
Yamamoto Y. 2001. Role of active oxygen species and antioxidants in photoaging. J Dermatol
Sci 27 Suppl 1:14.
Yang WTS, Min DB. 1994. Chemistry of singlet oxygen oxidation of foods. In: Ho CT, Hartmand TG, editors. Lipids in food flavors. Washington D.C.: American Chemical Society. p
1529.
Yanishlieva N, Schiller H. 1983. Effect of sitosterol on autoxidation rate and product composition in a model lipid system. J Sci Food Agric 35:21924.
Yanishlieva NV, Kamal-Eldin A, Marinova EM, Toneva AG. 2002. Kinetics of antioxidant
action of - and -tocopherols in sunflower and soybean triacylglycerols. Eur J Lipid Sci
Technol 104:26270.
Yen GC, Shyu SL. 1989. Oxidative stability of sesame oil prepared from sesame seed with
different roasting temperatures. Food Chem 31:21524.
Yoon SH, Min DB. 1987. Roles of phospholipids in the flavor stability of soybean oil. Korean
J Food Sci Tech 19:238.
Yoon SH, Min DB, Yeo YK, Horrocks LA. 1987. Analyses of phospholipids in soybean oils by
HPLC. Korean J Food Sci Tech 19:668.
Yoon SH, Jung MY, Min DB. 1988. Effects of thermally oxidized triglycerides on the oxidative
stability of soybean oil. J Am Oil Chem Soc 65:16526.
Yoshida H. 1994. Composition and quality characteristics of sesame seed (Sesamum indicum)
oil roasted at different temperatures in electric oven. J Sci Food Agric 65:3316.
Yoshida H, Takagi S. 1997. Effects of seed roasting temperature and time on the quality
characteristics of sesame (Sesamum indicum) oil. J Sci Food Agric 75:1926.
Yoshida H, Kirakawa Y, Takagi S. 2000. Roasting influences on molecular species of triacylglycerols in sesame seeds (Sesamun indicum). J Sci Food Agric 80:1495502.

COMPREHENSIVE REVIEWS IN FOOD SCIENCE AND FOOD SAFETYVol. 5, 2006

You might also like