Natural Antioxidants - Wiley Online Library

5 downloads 0 Views 331KB Size Report
has motivated the food industry to seek natural alternatives. ... The objective of this article is to provide an overview of natural antioxidants, their mechanisms of ...
Natural Antioxidants: Sources, Compounds, Mechanisms of Action, and Potential Applications M.S. Brewer

Abstract: While use of synthetic antioxidants (such as butylated hydroxytoluene and butylated hydroxyanisole) to maintain the quality of ready-to-eat food products has become commonplace, consumer concern regarding their safety has motivated the food industry to seek natural alternatives. Phenolic antioxidants can inhibit free radical formation and/or interrupt propagation of autoxidation. Fat-soluble vitamin E (α-tocopherol) and water-soluble vitamin C (L-ascorbic acid) are both effective in the appropriate matrix. Plant extracts, generally used for their flavoring characteristics, often have strong H-donating activity thus making them extremely effective antioxidants. This antioxidant activity is most often due to phenolic acids (gallic, protocatechuic, caffeic, and rosmarinic acids), phenolic diterpenes (carnosol, carnosic acid, rosmanol, and rosmadial), flavonoids (quercetin, catechin, naringenin, and kaempferol), and volatile oils (eugenol, carvacrol, thymol, and menthol). Some plant pigments (anthocyanin and anthocyanidin) can chelate metals and donate H to oxygen radicals thus slowing oxidation via 2 mechanisms. Tea and extracts of grape seeds and skins contain catechins, epicatechins, phenolic acids, proanthocyanidins, and resveratrol, all of which contribute to their antioxidative activity. The objective of this article is to provide an overview of natural antioxidants, their mechanisms of action, and potential applications.

Introduction Ultimately, food quality is defined in terms of consumer acceptability: taste, aroma, and appearance characteristics. The increasing demand for convenient foods has led to rapid growth in the ready-to-eat product category (Hofstrand 2008). Many of the food ingredients contain unsaturated fatty acids that are quite susceptible to quality deterioration, especially under oxidative stress. For this reason, efforts to reduce oxidation have increased. Most often, the best strategy is the addition of antioxidants. Synthetic phenolic antioxidants (butylated hydroxyanisole [BHA], butylated hydroxytoluene [BHT], and propyl gallate) effectively inhibit oxidation; chelating agents, such as ethylene diamine tetra acetic acid (EDTA), can bind metals reducing their contribution to the process. Some vitamins (ascorbic acid [AA] and α-tocopherol), many herbs and spices (rosemary, thyme, oregano, sage, basil, pepper, clove, cinnamon, and nutmeg), and plant extracts (tea and grapeseed) contain antioxidant components as well (Hinneburg and others 2006). Natural phenolic antioxidants, such as synthetics, can effectively scavenge free radicals, absorb light in the ultraviolet (UV) region (100 to 400 nm), and chelate transition metals, thus stopping progressive autoxidative damage and production of off-odors and off-tastes. MS 20101442 Submitted 10/22/2010, Accepted 3/8/2011. Author is with the Dept. of Food Science and Human Nutrition, 202 ABL, 1302 W. Pennsylvania Ave., Univ. of Illinois, Urbana, IL 61801, U.S.A. Direct inquiries to author Brewer (E-mail: [email protected]).

 c 2011 Institute of Food Technologists® doi: 10.1111/j.1541-4337.2011.00156.x

Since 1994, consumers have expressed concern about the safety of preservatives and additives in their food (Brewer and Russon 1994; Brewer and Prestat 2002; Rojas and Brewer 2008b). More than 12 y ago, Sloan (1999) reported that one of the top 10 trends for the food industry to watch included the sales of natural, organic, and vegetarian foods. There is a clear trend in consumer preference for clean labeling (Hillmann 2010), for food ingredients and additives that are organic/natural with names that are familiar, and that are perceived to be healthy (Joppen 2006). In addition, the call for sustainable sources and environmentally friendly production is forcing the food industry to move in that direction (Berger 2009). Defrancesco and Trestini (2008) estimated that consumers were willing to pay a price premium (up to 70% more) for organic fresh produce (tomatoes) for their health-promoting antioxidant content. Of an estimated $17 billion in sales in the United States, organic foods account for only 3% of total retail food sales. This category has been growing at 7 times the rate of the average food category and has maintained a growth rate of more than 15% per year. However, based on the results of a recent study, Evans and others (2010) concluded that products with physical changes, less processing, with identifiable ingredients would also be perceived to be more natural. Consumer expectations about ingredients that may suitably be labeled as “natural” do not always coincide with current guidelines (Williams and others 2009). A clear definition can guide manufacturers. However, the lack of consumer consensus makes it difficult for them to understand the implications. For consumers, “natural”

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 221

Natural antioxidants . . . and “clean label” are related to what they perceive an ingredient to be. This article provides an overview of naturally occurring antioxidant compounds, their sources, and mechanisms of action. Various different mechanisms may contribute to oxidative processes in complex systems, such as foods. These include reactions that generate reactive oxygen species that target different structures (lipids, proteins, and carbohydrates), and Fenton reactions, where transition metal ions play a vital role. It should be noted that antioxidant activity of food extracts can be determined using a variety of tests (stable free radical scavengers: galvinoxyl, diphenylb-picrylhydrazyl [DPPH]; lipid oxidation: peroxide oxygen, conjugated dienes, Rancimat [measurements of oxygen consumption of a linoleic acid emulsion and oxidation induction period in lard at 100 ◦ C], oxygen radical absorbance capacity [ORAC] values ), active oxygen method, iodine value (measure of the change in number of double bonds that bind I), anisidine value (reaction of acetic acid p-anisidine and aldehydes to produce a yellow color that absorbs at 350 nm), measurement of absorbance at 234 nm (conjugated dienes) and 268 nm (conjugated trienes) to assess oxidation in the early stages, and chromatographic methods; however, extraction procedures strongly influence the composition of the extracts and, therefore, also influence the antioxidant activity results (Halliwell 1997; Schwarz and others 2001; Trojakova and others 2001). In addition, the effect of the antioxidant compound in a food matrix may be significantly different than the activity of a purified extract.

Food Lipids The fatty acids in the lipids of food tissues may be saturated or unsaturated and may be part of the neutral triglyceride fraction (triacylglycerol) or part of the phospholipid fraction. Free fatty acids are electron-deficient at the oxygen atom of the carbonyl group (C=O); unsaturated fatty acids are also electron-deficient at points of carbon–carbon unsaturation (C=C). These electrondeficient regions make fatty acids susceptible to attack by a variety of oxidizing and high-energy agents generating free radicals (Nawar 1996). Triglycerides contain straight chains of primarily 16- to 18-carbon fatty acids and minimal amounts of unsaturated fatty acids. Phospholipids in tissue membranes contain up to 15 times the amount of unsaturated fatty acids (C18:4, C20:4, C20:5, C22:5, and C22:6) found in triglycerides. They are much more susceptible to oxidation because of the increase in the number of points of carbon–carbon unsaturation (C=C) (Elmore and others 1999).

Oxidation

r

When a hydrogen atom (H ) is abstracted rfrom an unsaturated fatty acid (R:H) forming an alkyl radical (R ), lipid oxidation is initiated (see nr 1 below). Generation of this lipid radical is thermodynamically unfavorable and ris usually initiated by the presence of other radical compounds (R ), singletstate oxygen (1 O2 ), decomposition of hydroperoxides (ROOH), or pigments that r act as photoensitizers. In order to stabilize, the alkyl radical (R ) usually undergoes a shift in the position of the double bond (cis r to trans) and production of a conjugated diene system. The R r can react with O2 to form a high-energy peroxyl radical (ROO ; see nr 2 below). The peroxyl radical can then abstract a hydrogen atom r (H ) from another unsaturated fatty acid (see nr 3 below) forming r a hydroperoxide (ROOH) and a new, free alkyl radical (R ). This process then propagates to another fatty acid (see nr 4 below; Srinivasan and others 2008). Lipid hydroperoxides (ROOH) are the

Table 1–Flavors and aromas associated with volatile compounds resulting from lipid oxidation. Compound Butanoic acid Propanoic acid Pentanal Hexanal Heptanal Octanal Nonanal Decanal Decanal 2-Nonenal 2-Hexenal Nona-2(E)-enal E-2-Hexenal E-2-Octenal E-2-Decenal E,E-Deca-2, 4-dienal E,E-Nona-2, 4-dienal 2-Propanone 3-Hydroxy-2-butanone 2,3-Octanedione Nonenone 2-Heptanone Dimethyl disulfide Dimethyl trisulfide Methanethiol

Flavors and aromas Rancid Pungent, rancid, soy Pungent, malt Green, grassy, tallowy Green, oily, rancid Sweet, fatty, soapy Tallowy, waxy Rancid, burnt Soapy, tallowy Paper Rancid Tallowy, fatty Green, fat, rancid Fatty, nutty, green Green, pungent, Fatty, fried potato, green Green, grassy, fatty Livery Rancid, beany, grassy Oxidized fat or oil Pungent Soapy Onion, cabbage, putrid Sulfur, fish, cabbage Garlic, sulfur

Kerler and Grosch (1996), Yong and others (2000), Ahn and others (2002, 2007), Jensen and others (2002), Jo and others (2003), Mahrour and others (2003), Nam and Ahn (2003), Nam and others (2003), Acree and Arn (2004), Rochat and Chaintreau (2005), Obana and others (2006), and Yancey and others (2006).

primary products of lipid oxidation. They are tasteless and odorless; however, in the presence of heat, metal ions, and/or light, they can decompose to compounds that contribute roff-odors and offr tastes. Alkoxy radicals (RO ) can also abstract H from unsaturated fatty acids continuing the chain reaction (see nr 5 below; r Decker 2002; Srinivasan and others 2008). Hydroxyl radicals ( OH) can react with conjugated systems continuing the oxidation process. This chain reaction terminates when 2 radical species combine to form a nonradical species (see nr 6 below). Antioxidants (A:H) r inhibit the chain reaction by donating hydrogen atoms (H ) to radicals (see nr 7 below). The antioxidant free radical may then form a stable peroxy-antioxidant compound (see nr 8 below). (1) (2) (3) (4) (5) (6) (7) (8) (9) (10)

r

r

R:H r + O::O + Initiator r → R + HOO R + rO::O → ROO r ROO + R:H → r ROOH r +R RO:OHr→ RO + HO r R::R +r OH → R:R-O r Rr + R →r R:R R + rROO → rROOR ROOr + ROO → ROOR +r O2 ROOr + AH r → ROOH + A ROO + A → ROOA.

Ultimately, oxidation depends on the addition of oxygen to a compound; however, the energy level of the oxygen has a significant impact on the ease of the oxidation reaction. Singlet state oxygen (1 O2 ) has spin-coupled electrons and is a nonradical, high-energy species (Decker 2002; Min and Boff 2002). It is electrophilic and can react with other electron-rich, nonradical, singlet state compounds containing double bonds (C=C, C=O). However, oxygen in its lowest energy state, triplet state oxygen (3 O2 ), has 2 unpaired, parallel spin electrons. It is very reactive (primarily with radical species). Most food components, such as carbohydrates and proteins, are nonradical (singlet state) and are relatively unreactive with triplet state oxygen (3 O2 ); however, they

222 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . . are reactive with singlet state oxygen (1 O2 ) that can be generated in response to temperature change, reduction of activation energy (presence of transition metals), exposure to UV light, and physical damage to tissues. Singlet oxygen and free radicals can cause biological damage to macromolecules and membrane constituents. The presence of natural antioxidants may help control these degradative reactions. Oxidation of unsaturated fatty acids can produce a variety of aldehydes, alkanals, alkenes, and alkanes; many of which contribute off-odors that are perceptible at very low concentrations. Odor detection thresholds for pentanal, hexanal, and heptanal, compounds typically generated from the breakdown of oxidized linoleic acid have been reported to be kaempferol > galangin. Authors interpret these findings as indicating that hydrogen bond force plays an important role. Flavonoids with multiple hydroxyl groups are more effective antioxidants than those with only one. The presence of the ortho3,4-dihydroxy structure increases the antioxidative activity (Geldof and Engeseth 2002). Flavonoids can dampen transition metal

224 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . .

CH3 H 3C

Figure 2–Antioxidative phenolic diterpenes (carnosol, carnosic acid, rosmanol, and rosmarinic acid) found in plants.

CH 3

CH3 OH

CH 3

Diterpene structure

OH

H O

HO

O

COOH

OH

OH

OH

Carnosol

Carnosic acid OH OH

O

HO O OH

O

CH 3

HO CH3 O

CH

O

HO

OCH 2CH 3 H 3C

CH3

OH

Rosmanol

Rosmarinic acid r

enhancement of oxidation by donating a H to them, rendering them less proxidative. In addition, flavones and some flavanones (naringenin) can preferentially bind metals at the 5-hydroxyl and 4-oxo groups (Fernandez and others 2002). Brown and Kelly (2007) evaluated the antioxidative activity of structurally related (poly)phenols, anthocyan(id)ins, and phenolic acids at physiologically relevant concentrations (100 to 1000 nM) using a Cu2+ -mediated low-density lipoprotein oxidation model. (Poly)phenols with an ortho-dihydroxy substituted arrangement (cyanidin-3-glucoside, cyanidin, and protocatechuic acid) were the most effective, while trihydroxy-substituted compounds (gallic acid) had only intermediate efficacy. This was explained, in part, by their ability to chelate Cu2+ ions. It seems likely that the steric relationship of these −OH groups and their arrangement on the ring(s) both play a role in the ability of the substance to chelate metal ions. However, differences in lipid/hydrophilic phase parti c 2011 Institute of Food Technologists®

tioning and in H-donating abilities were also hypothesized to have contributed to the structure-activity relationships. Alamed and others (2009) reported that the order of free radicalscavenging activity of a group of polar compounds was ferulic acid > coumaric acid > propyl gallate > gallic acid > AA; the free radical-scavenging activity of a group of nonpolar compounds was rosmarinic acid > BHT, tert-butylhydroquinone (TBHQ) > α-tocopherol. Only propyl gallate, TBHQ, gallic acid, and rosmarinic acid inhibited lipid oxidation in an oil-in-water emulsion that may reflect the ability of these compounds to orient at the interface of the oil droplet in the emulsion. Evaluating the antioxidative activity of hydroxycinnamic acids with similar structures (caffeic, chlorogenic, o-coumaric, and ferulic acids) in a fish muscle system, Medina and others (2007) found that the capacity of these compounds to donate electrons (bond dissociation energies) appeared to play the most significant

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 225

Natural antioxidants . . .

OH OH

OH OH

HO

O

HO

O

OH

OH

OH

OH

Epicatechin

O

Quercetin (flavanol)

OH

OH

HO

OH

O

OH

HO

HO

O OH

O

HO

HO

O

OH

OH

O

O

HO

OH OH

OH

Epicatechin gallate

Epigallocatechin gallate OH

HO

O

O

O

OH

O

O

OH

O HO O

HO OH

HO

Rutin

OH OH

Figure 3–Antioxidative flavonoids (epicatechin, quercetin, epicatechin gallate, epigallocatechin gallate, and rutin) found in plant extracts.

226 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . .

OH 3C

O HO

O

HO

Eugenol

Carvacrol

Safrole

H 3C O

O

OH

OH O

H2 C

Thymol

Menthol

Myristicin H 3C

OH

CH 3

O

CH3 H3 C O

CH 3

1,8-Cineol

CH3

- Terpineol

p-Cymene

Cinnamaldehyde

O

O

N

O

Piperine Figure 4–Antioxidant volatile oils (eugenol, carvacrol, safrole, thymol, menthol, 1,8-cineole, α-terpineol, p-cymene, cinnamaldehyde, myristicin, and piperine) found in plant extracts.

role in delaying rancidity, while the ability to chelate metals and the distribution between oily and aqueous phases were not correlated with inhibitory activities. The latter finding may reflect the type of matrix, fish muscle, in which the oxidative activity was studied. Caffeic acid was the most effective of this antioxidant group (similar to propyl gallate).  c 2011 Institute of Food Technologists®

Potapovich and Kostyuk (2003) reported that, of a variety of flavonoids (rutin, dihydroquercetin, quercetin, epigallocatechin gallate, and epicatechin gallate), the catechins were most effective in inhibiting microsomal lipid peroxidation. All rwere able to chelate Fe2+ , Fe3+ , and Cu2+ and were effective O2 − scavengers to varying degrees. Authors speculate that the

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 227

Natural antioxidants . . .

B 8 7

9

6

4

A 5

2 3

1

Basic C6, C3, C6 structure

Chalcone

B

B HO

O

O

C

A

C

A

B

O

C

A OH

OH

O

O

O

Flavone

Flavonol

Flavanone HO

R1 O

OH

R2

3' 4'

+

O

O

R7

O

5'

7

OH

R3

HO

O O

3

6 R6

5

HO

R4

HO

OH

OH OH

R5

OH

Anthocyanin (R2= -OH, R4= -OH or H, R5 and R7= -OH or –OCH3

Anthocyanidin-3,5-glucoside

Figure 5–Flavonoid structure and flavonoids (flavanones: chalcone, flavone, flavanol, and flavanones; and flavans: anthocyanin and anthocyanidin-3,5-glycoside) found in plant extracts.

r

relative ability to scavenge O2 − may be responsible for the relative antioxidative difference among these compounds. Many of the antioxidative flavonoid compounds are naturally occurring pigments. It appears that chloroplast-located flavonoids r perform a photo-protective role against O2 − in plants (Agati and others 2007). Anthocyanins are the glycosides of polyhydroxy or polymethoxy derivatives of the flavylium cation. Hydrolysis of the sugar moiety yields an aglycone, anthocyanindin (Figure 5). Anthocyanins and anthocyanindins exhibit visual color because of the extreme mobility of the electrons within the molecular structure (double bonds) in response to light in the visible spectrum (approximately 400 to 700 nm). The pigments are quite water-

soluble and 4 −OH groups are bound to the aromatic rings. pH has a significant effect on anthocyanin pigments. r These −OH groups can give up H+ (in a basic solution) or H to an oxidizing r lipid (ROO ). Proanthocyanidins also contain multiple −OH groups that can r donate hydrogen, quench O2 − , and chelate metals (Shahidi and Wanasundara 1992; Fukumoto and Mazza 2000). Free radicalscavenging ability increases as the number of phenolic −OH groups increases (Kondo and others 2001). Some phenols can polymerize into polyphenols that can bind minerals. Proanthocyanidins often occur as oligomers or polymers of monomeric flavonoids, polyhydroxy flavan-3-ols such as

228 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . .

H 3C

H 3C

CH3 H

H

CH3

CH3

CH2

H 2C

CH2

H 2C

HO

H

H

HO

O

CH3

H3 C

O

CH 3

CH 3

CH3

CH3

Alpha tocopherol

Gamma tocopherol O

OH CH3

CH OH

CH 2O

(CH2) 14

C

CH 3

HCOH

O

H C

O

O

C

O

C C OH

OH

OH

OH

Ascorbic acid

Ascorbyl palmitate OH

OH HO

OH

O

O

HO

Propyl gallate

OH

Resveratrol

Figure 6–Natural antioxidants (alpha- and gamma-tocopherol, ascorbic acid, ascorbyl palmitate, propyl gallate, and resveratrol).

[+]-catechin and [−]-epicatechin (Dixon and others 2005; Figure 3 and 5). The polymeric procyanidins are better antioxidants than the corresponding monomers, catechin, and epicatechin (Ursini and others 2001). Catechin and epicatechin can combine to form esters, such as catechin/epicatechin gallate, or bond with sugars and proteins to yield glycosides and polyphenolic proteins. Glycosylation of flavonoids at the 3 −OH group usually decreases the antioxidative activity due to the reduction of the number of phenolic groups (quercetin/rutin; Figure 3). Proanthocyanidins with demonstrated antioxidant activity and potential biologically therapeutic effects occur in fruits (apples and  c 2011 Institute of Food Technologists®

cherries), some berries (rosehips, raspberries, blackberries, and strawberries), as well as in the leaves (tea), seeds (grape, sorghum, soy, and cocoa bean), and bark of many plants (Dixon and others 2005; Buricova and Reblova 2008; Bak and others 2010).

α-Tocopherol α-Tocopherol (vitamin E) is a fat-soluble carotenoid whose antioxidative capacity has been studied extensively (Figure 6). α-Tocopherol is the major vitamin E compound in plant leaves where it is located in the chloroplast envelope and thylakoid membranes in proximity to phospholipids (Onibi and others 2000). It

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 229

Natural antioxidants . . . deactivates r photosynthesis-derived reactive oxygen species (especially O2 − ) and prevents the propagation of lipid peroxidation by scavenging lipid peroxyl radicals in thylakoid membranes (Munn´e-Bosch 2005). Trolox is a water-soluble derivative of vitamin E. Structurally related lipid-soluble antioxidants that differ in the number of methyl groups (δ-tocopherol compared with α-tocopherol) have different free radical-scavenging activities and different surface activities (Figure 6; Chaiyasit and others 2005). Giuffrida and others (2007) evaluated the ability of α-tocopherol, δ-tocopherol, ascorbyl palmitate, and propyl gallate (300 mg/kg; Figure 6) to prevent oxidation in sunflower oil and high-oleic sunflower oil, both rich in di-unsaturated fatty acids, and in partially hydrogenated palm oil containing monounsaturated fatty acids. δ-Tocopherol was the most effective antioxidant in sunflower oil, and propyl gallate was the most effective in the more saturated oils. Yeum and others (2009) reported synergistic effects between AA and α-tocopherol in protecting an in vitro biological model system. It may be that r AA regenerates α-tocopherol after α-tocopherol donates a H to an oxidizing lipid. α-Tocopherol can also inhibit oxidation of protein. Est´evez and Heinonen (2010) demonstrated that α-tocopherol reduced formation of α-aminoadipic and γ -glutamic semialdehydes from oxidized myofibrillar proteins. In general, vitamin E added to water-based food systems using an oil carrier targets the neutral lipid fraction (triacylglycerols) rather than the polar lipid fraction (phospholipids) and is not an effective antioxidant. However, δ-tocopherol added using a polar carrier can be incorporated into the phospholipid fraction and is an effective antioxidant (Wills and others 2007). In a lard model system, the antioxidative activity of the tocopherols is temperature dependent (Reblova 2006). At 80 ◦ C, the antioxidative activity of δ-tocopherol is about twice that of α-tocopherol; however, it decreases as temperature increases. Antioxidative activity of α-tocopherol decreases above 110 ◦ C, and both lose their activity above 150 ◦ C. Dietary supplementation of α-tocopherol increases incorporation of the antioxidant into the phospholipid membrane region where the polyunsaturated fatty acids are located. Including α-tocopherol in livestock diets has been shown to have significant effects on the antioxidative activities of their tissues and the stability of meat derived from them (Formanek and others 2001; Swigert and others 2004; Guo and others 2006; Boler and others 2009; Lahucky and others 2010).

Ascorbic acid AA has 4 −OH groups that can donate hydrogen to an oxidizing system (Figure 6). Because the −OH groups (2 pairs of 2) are on adjacent carbon atoms, AA is able to chelate metal ions r (Fe++ ). It also scavenges free radicals, quenches O2 − , and acts as a reducing agent. At high levels (>1000 mg/kg), AA shifts the balance between ferrous (Fe2+ ) and ferric iron (Fe3+ ), acts as an oxygen scavenger, and inhibits oxidation. However, at low levels ( carvacrol > γ -terpinene > myrcene > linalool > p- antibacterial activity against foodborne pathogens. Gram-positive cymene > limonene > 1,8-cineole > α-pinene (Youdim and bacteria were generally more sensitive than Gram-negative bacothers 2002). Carvacrol and thymol each have 1 aromatic ring and teria. Staphylococcus aureus was the most sensitive, while Echerichia 1 −OH group, 1-terpineol has 1 −OH group, while p-cymene coli was the most resistant. The antibacterial activity of the extracts has 1 aromatic group. The presence of aromatic groups and the was closely associated with their phenolic content. 232 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

 c 2011 Institute of Food Technologists®

Nakatani and Inatani (1981), Arouma and others (1992), Ahn and others (2002, 2007), Dorman and others (2003), Shan and others (2005), Carrillo and Tena (2006), Jirovetz and others (2006), Jukic and others (2006), Schmidt and others (2006), Yanishieva and others (2006), Chaieb and others (2007), Hatzidimitrioua and others (2007), Hernandez-Hernandez and others (2009), and El-Ghorab and others (2010).

X X ∗ Flavonoids: kaempferol glycosides, rutin, apigenin, kaempferol, hesperetin, dihydroquercetin, quercetin, and catechins (catechin, epicatechin epigallocatechin gallate, and epicatechin gallate).

Clove The primary components of clove (Eugenia caryophyllus) essential oil are phenylpropanoids such as eugenol, carvacrol, thymol, and cinnamaldehyde (Figure 4; Chaieb and others 2007). Clove also contains a variety of nonvolatile compounds (tannins, sterols, flavonoids, and triterpenes). Jirovetz and others (2006) identified 23 compounds in clove oil including eugenol (76.8%), βcaryophyllene (17.4%), α-humulene (2.1%), and eugenyl acetate (1.2%). A variety of the antioxidative compounds are shown in Table 5. Clove essential oil is inhibitory toward hydroxyl radicals and can chelate iron. Comparing 16 spices, Khatun and others (2006) found that clove had the highest radical-scavenging activity followed by allspice and cinnamon. Eugenol has been reported to have an antioxidative activity equivalent to Trolox, carvacrol (oregano), and thymol (thyme; Dorman and others 2000). The essential oil scavenges free radicals at concentrations lower than those of eugenol, BHT, and BHA alone. Using peroxide values and formation of conjugated dienes, Marinova and others (2008) established that in sunflower oil at 100 ◦ C, myricetin is a more effective and stronger antioxidant than α-tocopherol. Mixtures of

Table 5–Selected antioxidant compounds identified in selected spices.

Cinnamon Cinnamon (Cinnamonum zeylanicum) contains a number of antioxidative components including vanillic, caffeic, gallic, protochatechuic, p-hydroxybenzoic, p-coumaricd, and ferulic acids and p-hydroxybenzaldehyde (Table 5, Figure 1, 2, 4, and 5; (Muchuweti and others 2007). Of a number of herbs and spices (bay leaves, rosemary, sage, marjoram, oregano, cinnamon, parsley, sweet basil, and mint) evaluated, cinnamon has been reported to have the highest polyphenolic compound concentration (13.7 mg GAE/g; Muchuweti and others 2007). Of 42 commonly used essential oils, cinnamon bark, oregano, and thyme have been reported to have the strongest free radical-scavenging abilities (Wen and others 2009). At 5 mg/mL, cinnamon a radical-scavenging activity of 92% (Muchuweti and others 2007). The major components responsible for this activity were identified as eugenol, carvacrol, and thymol. Jayaprakasha and others (2003) identified 27 compounds in the volatile oil of cinnamon stalks. The volatile oil was 44.7% hydrocarbons and 52.6% oxygenated compounds. Using a β-carotene-linoleate model system, the volatile oil inhibited 55.9% and 66.9% of the oxidation at 100 and 200 ppm, respectively, compared to the control. The antioxidant capability of cinnamon essential oil is stronger than its free radical-scavenging capacity (Chen 2008). However, it is a better superoxide radical scavenger than propyl gallate, mint, anise, BHA, licorice, vanilla, ginger, nutmeg, or BHT (Murcia and others 2004). Cinnamon (bark and leaf) oleoresin can significantly inhibit formation of primary and secondary oxidation products. Singh and others (2007) identified 13 components, which accounted for 100% of the total amount, in cinnamon bark volatile oil. The bark oleoresin contained 17 components that accounted for 92.3% of the total amount. The major component in cinnamon bark oleoresin was (E)-cinnamaldehyde (49.9%). Schmidt and others (2006) identified small amounts of β-caryophyllene, benzyl benzoate, linalool, eugenyl acetate, and cinnamyl acetate in cinnamon leaf essential oil. The major component identified in cinnamon leaf oil was eugenol (87.2%). Cinnamon leaf oil has a significant inhibitory effect on hydroxyl radicals and acts as an iron chelator efficiently inhibiting formation of conjugated dienes and generation of secondary products from lipid peroxidation at a concentration equivalent to BHT.

Phenolic acids Volatiles Phenylpropanoids p-Hy p-Hy α, β Van-illic Caffeic Gallic droxy droxy p- Coum α, βCaro Cur Argen Cam- Lina Gin Piper Cy- Carv CuminCinn-amacid acid acid benzoic acid benz aldehye aric acid Pinene phyllene cumin taene phene lool gerol ine Limonene mene acrol alde hyde alde hyde Cinn-amon X X X X X X X X X X X X X Clove X X X X X X X X X X Nut-meg X X X X X X X X X X Ginger X X X X X Turm-eric X X X X X X X Black pepper X X X X X X X X X

Eug enol X X X

Natural antioxidants . . .

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 233

Natural antioxidants . . . the 2 exhibited a synergistic effect that was optimized in an equal molar ratio of the 2. The antioxidant activity of glycosidically bound volatile compounds in clove essential oil has been reported to be significantly greater than that of the volatile aglycones (Politeo and others 2010). The glycosides can undergo enzymatic hydrolysis releasing their aglycones, therefore, could be considered as potential antioxidant precursors. Heating at 100 ◦ C for up to 6 h increases the peroxy radical-scavenging activity of clove (Khatun and others 2006).

out Asia for centuries. Ground turmeric consists primarily of curcumin, dimethoxycurcumin and bis-dimethoxycurcumin, and 2,5-xylenol (Zhang and others 2009). Curcumin is an unsaturated diketone that exhibits keto-enol tautomerism (Anand and others 2008).r It is a classical phenolic chain-breaking antioxidant, donating H from the phenolic groups rather than from the CH2 group (Ross and others 2000). Jayaprakashaa and others (2006) reported that the antioxidant activity of the curcuminoids is curcumin > BHT, dimethoxycurcumin > bisdemithoxycurcumin. All of these polyphenolic molecules have limited water solubility. Curcumin is highly effective in neutralizing free radicals (Yu and others 2008). At the same concentration, curcumin has about twice the antioxidative activity of the polyphenol resveratrol (Aftab and Vieira 2009). Turmeric oil has a free radical-scavenging ability comparable to vitamin E and BHT (Yu and others 2008). The major components of turmeric oil responsible for this antioxidant activity are α- and β-turmerone, curlone, and α-terpineol (Carolina and others 2003). Heating dry ginger and turmeric and their essential oils at 120 ◦ C results in different degrees of retention of antioxidant activity (Tiwari and others 2006). Antioxidant activity of turmeric oil is higher after heating while that of ginger oil is lower. This may be due to the difference in monoterpene content or to the release of bound antioxidants caused by the heat treatment. A variety of the antioxidant compounds found in ginger and nutmeg are shown in Table 5. Cumin is derived from Cuminum cyminum. The major components in cumin volatile oil are cuminal, γ -terpinene, and pinocarveol (El-Ghorab and others 2010). Cumin essential oil is better at reducing Fe3+ ions than dried or fresh ginger or cumin.

Nutmeg Jukic and others (2006) isolated glycosidically bound volatiles from nutmeg and identified free aglycones in the essential oil. The glycosidically bound and aglycone fractions had only 2 compounds in common, eugenol and terpinen-4-ol. The aglycone fraction had stronger antioxidant properties than did the free volatiles from the oil. Nutmeg (Myristica fragans and M. argentea) contains argenteane, a flavanol diglycoside, which appears to be the primary antioxidative compound (Calliste and others 2010). Bis-erythro argenteane is a di-lignan that has been isolated from nutmeg mace (the lace-like seed membrane of nutmeg). The argenteane central moiety (3,3 -dimethoxy-1,1 -biphenyl-4,4 -diol) appears to ower its free radical-scavenging ability to its ability to release 1 or 2 H (Chatterjee and others 2007; Calliste and others 2010). Nutmeg also contains significant amounts of myristicin and safrole that are responsible for the characteristic aroma of nutmeg (Fisher 1992). Myristicin and safrole have similar structures: a 6-membered aromatic ring bound to an oxygenated 5-carbon ring on one side and a hydrocarbon side chain on the other (Figure 4). After heating (180 ◦ C, 10 min), nutmeg oil has a significantly higher free radical-scavenging activity, compared to basil, cinnamon, clove, oregano, and thyme (Tomaino and others 2005). A variety of the Black pepper antioxidant compounds found in nutmeg are shown in Table 5. Black pepper (Piper nigrum) is a highly valued spice for its distinct biting quality that occurs at 1.35 ppm. It has a pungency 150 times Ginger, turmeric, and cumin that of capsaisan (United States Consumer Product Safety ComGinger is derived from the root of Zinger officinale. Fresh and mission 1992) due to the alkaloid piperine (Figure 4; Srinivasan dried ginger contain relatively large amounts of the volatile oils 2007). The flavor quality is measured by the volatile oil and by the camphene, p-cineole, alpha-terpineol, zingiberene, and pentade- nonvolatile methylene chloride extract, piperine. Piperine stimcanoic acid (Figure 4; Tiwari and others 2006; El-Ghorab and ulates the digestive enzymes of the pancreas, enhances digestive others 2010). The maximum total phenolic contents were ex- capacity, and reduces gastrointestinal food transit time. Piperine tractable with methanol from fresh ginger (95.2 mg/g dry extract) can also quench free radicals and reactive oxygen species. It can followed by hexane extraction of fresh ginger (87.5 mg/g dry ex- protect against oxidative damage in vitro. Piperine acts as a hydroxyl tract). Hydrodistillation produced 23 mg GAE/g (Hinneman and radical scavenger at low concentrations (Mittal and Gupta 2000). others 2006). Ginger extract has been shown to have antioxidant Kapoor and others (2009) reported that black pepper (P. nigrum) activity almost equal to that of synthetic antioxidants (BHA and volatile oil contains 54 components that represent about 97% of BHT; Rehman and others 2003). the total weight. β-Caryophylline (30%) is the major compoKikuzaki and Nakatani (2006) reported that 12 of the 5 nent along with limonene (13%), β-pinene (7.9%), and sabinene gingerol-related compounds and 8 diarylheptanoids isolated (5.9%). Pepper essential oils also contain α- and β-pinene, from ginger rhizomes exhibit higher antioxidative activity than cyclohexene, 1-methyl-4-(1-methylethylidene)-2,3-cyclohexenα-tocopherol. Authors suggest that this is likely dependent upon 1-ol, limonen-6-ol, (E)-3(10)-caren-4-ol, and t-caryophyllene side chain structures in addition to substitution patterns on the (Liang and others 2010). The major component of both ethanolbenzene ring. Hinneburg and others (2006) have suggested that, and ethyl acetate-extracted oleoresins is piperine (63.9% and for ginger, it may advisable to use extraction media that are able 39.0%, respectively, Liang and others 2010). Using peroxide, pto extract the lipophilic antioxidant compounds. anisidine, and thiobarbituric acid tests, the oil and oleoresins Turmeric is a spice derived from the rhizomes the Cur- have been shown to have stronger antioxidant activity than cuma longa plant, which is a member of the ginger family BHA and BHT (Kapoor and others 2009) but less than that of (Zingiberaceae). Rhizomes are horizontal underground stems propyl gallate. Gurdip and others (2004) reported that, while exthat send out shoots as well as roots. The bright yellow color tracts were predominantly piperine, piperolein B, piperamide, and of turmeric is primarily due to fat-soluble, polyphenolic pig- guineensine, the predominant compounds in essential oils were βments known as curcuminoids, primarily curcumin (diferuloyl caryophyllene, limonene, sabinene, β-bisabolene, and α-coapene. methane; Priyadarsini and others 2003; Anand and others 2008). Some of the antioxidant compounds found in black pepper are Turmeric has been used as a spice and medicinal herb through- shown in Table 5. 234 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . .

Figure 7–Antioxidative sulfur-containing compounds (S-allyl—cysteine sulfoxide, diallyl sulfide, allyl trisulfide, and allyl-cysteine) in allium plant extracts.

O

S

OH

O

S

N

S-allyl L-cysteine sulfoxide

Diallyl sulfide O N OH

S

S

S S

Allyl trisulfide

Allyl-cysteine

The antioxidative effects of shallots are related primarily to their phenol content (Leelarungrayub and others 2006). According to Nuutila and others (2003), methanol extracts of onions have significantly higher radical-scavenging activities than garlic and red onion has higher activity than yellow onion. Quercetin content is highest in red onions (Gorinstein and others 2008). The radicalscavenging activities are positively correlated with the total pheGarlic and related herbs Garlic (Allium sativum L.) has been widely used as a foodstuff nolics in these extracts. since antiquity. It has acquired a reputation as a therapeutic agent and herbal remedy in many cultures to prevent and treat heart and Tea extracts metabolic diseases, such as atherosclerosis, thrombosis, hypertenThe 3 primary types of tea, green, black, and oolong, are prosion, dementia, cancer, and diabetes (Tyler 1993). duced by different processing procedures. Of these types, green tea Garlic and shallots (Allium ascalonicum) have antioxidant and extracts have the highest total phenolics content, 94% of which free radical-scavenging characteristics and identifiable odors at are flavonoids (catechins; Duh and others 2004). Oolong tea conlow concentrations. They contain 2 main classes of antioxi- tains about 18% total phenolics and 4.4% flavonoids. Theaflavins dant compounds: flavonoids (flavones and quercetins; Figure 3) and thearubigins predominate in black tea. Black tea also conand sulfur-containing compounds (allyl-cysteine, diallyl sulfide, tains chlorogenic, caffeic, p-coumaric, and quinic acids (Figure 1; and allyl trisulfide; Figure 7). The sulfur-containing amino acid Kiehne and Engelhardt 1996). Some of the antioxidant comderivative, alliin (S-allyl-L-cystein sulfoxide), can be converted pounds found in tea are shown in Table 6. into allicin (diallyldisulfide-S-oxide), the compound commonly Much of the antioxidative activity of green tea (C. sinensis) associated with garlic odor, by the enzyme alliinase. Thiosulfi- appears to be due to natural flavonoids, tannins, and some vitamins nates, such as allicin, give garlic its characteristic odor; however, (Abdullin and others 2001). The antioxidant activity is linearly they are not necessarily responsible for all of the various antiox- related to the phenol content (Apak and others 2006) that has idative and health benefits attributed to it (Amagase 2006). Okada been reported to be about 450 mg/g (Peschel and others 2007). and others (2005) have suggested that a combination of the allyl Catechins in green tea consist primarily of gallic acid derivatives group (−CH2 CH=CH2 ) and the −S(O)S− group is necessary (Chen and others 2007). Catechin flavanols appear to account for for the antioxidant action of thiosulfinates in garlic extracts. S- more than 80% of the total antioxidant activity of green tea but allylcysteine, S-allyl mercaptocysteine, and nonsulfur compounds, less than 60% of that of black tea (Gardner and others 1998). The such as saponins, may contribute to the health benefits (hy- radical-quenching ability of green tea has been shown to be more polipidemic, antiplatelet, procirculatory, immune enhancement, than 20% more effective than that of black tea in both aqueous anticancer, and chemopreventive activities) associated with garlic. and lipophilic systems. In tea extracts, the strongest antioxidant and H2 O2 -scavenging Gorinstein and others (2008) reported that trans-hydroxycinnamic acid (caffeic, p-coumaric, ferulic, and sinapic acids) concentrations activity is due to phenols, with 3 −OH groups bonded to the in garlic were twice that in onions. Some of the antioxidant com- aromatic ring, adjacent to each other (Sroka and Cisowski 2003). pounds found in garlic are shown in Table 6, and sulfur-containing Epigallocatechin, which has 3 adjacent −OH substitutions on compounds are shown in Figure 7. the B ring, has the highest antioxidant activity (Figure 3). In Given that piperine is one of the most effective antioxidative components and also one of the primary aroma compounds in pepper, using extracts of this herb would likely impart unwanted flavors to foods to which they are added unless other antioxidative but nonaromatic components can be separated from the extract.

 c 2011 Institute of Food Technologists®

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 235

X X

Others Myre cetin

X

Res veratrol

Hydrocinnamic acids pCaffeic Coumaric Ferulic X X X X X X X X X

X X

X

Sulfur-containing compounds Allyl Diallyl Tri cysteine sulfide sulfide X X X Epi gallo catechin gallate Flavonoids Epi catechin gallate Epi catechin

Cate quin X X X Garlic and shallots Tea extract Grape seed extract

Quer cetin X X X

Table 6–Selected antioxidant compounds identified in garlic, tea, and grapeseed extract.

Nakatani and Inatani (1981), Arouma and others (1992), Cuvelier and others (1994), Kiehne and Engelhardt (1996), Dorman and others (2000, 2003), Ahn and others (2002, 2007), Okada and others (2005), Carrillo and Tena (2006), Yanishieva and others (2006), Chen and others (2007), Hatzidimitrioua and others (2007), Gorinstein and others (2008), Iacopini and others (2008), Hernandez-Hernandez and others (2009), and Calliste and others (2010).

Rutin

Natural antioxidants . . . addition to a flavonoid ring, a 3 ,4 ,5 -trihydroxy (galloyl) group are required for Fe++ binding in catechins (Khokhar and OwusuApenten 2003). These polyphenolic flavonoids are particularly effective free radical scavengers (Figure 3; Lien and others 2008). The primary catechin polyphenol [(−)-epigallocatechin-3-gallate] is also the primary peroxyl-radical-scavenging compound in tea extracts (Caldwell 2001; Cabrera and others 2003). In terms of free radical-scavenging ability, epicatechin gallate > epigallocatechin > epicatechin (Guo and others 1996). The first 2 compounds have 3 ,4 ,5 -trihydroxy (galloyl) groups while the last does not. Both their iron-chelating and free radical-scavenging activities appear to be responsible for the ability of these compounds to protect membranes from Fe2+ /Fe3+ -initiated lipid oxidation. In an ironmediated reaction, Grey and Adlercreutz (2006) demonstrated that catechin inhibited oxidation better than AA. They concluded that catechin’s chelating ability, rather than its radical-scavenging mechanism alone, is responsible for the observed antioxidative activity. In more complex food systems, tea catechins can have varying effects. Mitsumoto and others (2005) found that adding tea catechins to raw beef (200 or 400 mg/kg) inhibited (P < 0.05) lipid oxidation to a greater extent than vitamin C (200 or 400 mg/kg); however, they increased discoloration in cooked beef and chicken meat. Chen and others (1998) reported that green tea catechin extract, consisting primarily of 4 epicatechin isomers, was much more antioxidative than rosemary extract when added to canola oil, pork lard, and chicken fat. In maize (corn) oil triglycerides, Huang and Frankel (1997) found that epigallocatechin (140 M), epigallocatechin gallate, and epicatechin gallate were better antioxidants than either epicatechin or catechin. Both gallic acid and propyl gallate were more effective than epicatechin and catechin. However, in a maize oil-in-water emulsion, all tea catechins, gallic acid, and propyl gallate were prooxidative (5 and 20 M) accelerating hydroperoxide and hexanal formation. The improved antioxidant activity of tea catechins in liposomes, compared with emulsions, may be due to the greater affinity of the polar catechins toward the polar surface of the lecithin bilayers, thus affording better protection (Hatzidimitrioua and others 2007). In a model system mixture of the flavanols, in the same concentrations as they occur in the tea extracts, the antioxidant potential has been shown to be a simple summation of the activity of the individual components with no apparent synergism or antagonism occurring (Gardner and others 1998). Alkyl compounds with double bond(s), such as 3,7-dimethyl1,6-octadien-3-ol in green tea extracts and heterocyclic compounds (furfural) in roasted green tea extracts, are major volatile constituents that also exhibit some antioxidative activity (Yanagimoto and others 2003).

Grape seed extract Because red wines are produced from red grapes, the antioxidant capacity and chemical composition are related to the grapes. The antioxidant activity of red wines is associated with the content of polyphenols such as flavonoids, phenolic acids, stilbenes, coummarines, and lignoids (Radovanovic and others 2009). The phenolic composition varies greatly due to grape variety, environmental and climate conditions, soil type, degree of ripeness, and winemaking process (Table 6; Jayaprakasha and others 2001; Hatzidimitrioua and others 2007; Lachman and others 2007; Iacopini and others 2008; Rababah and others 2008; Xu and others 2010;). Touns and others (2009) reported wide variations in the contents of total phenols (122 to 441 mg GAE/g), flavonoids

236 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . . 17 to 48 mc epicatechin [EC]/g), and tannins (15 to 37 mc EC/g) in the methanolic extracts from seeds of 3 varieties of grapes. Phenolic compounds in grape seeds and skins include catechins, epicatechins, epicatechin-3-O-gallate, phenolic acids, caffeic acid, quercetin, myricetin, proanthocyanidins, and resveratrol (Figure 1, 3, and 4; Jayaprakasha and others 2001; Hatzidimitrioua and others 2007). Many have strong antiradical activity. Most of the phenolic compounds found in red wines are derived from the condensation of flavan-2-ol into oligomers (proanthocyanidins) and polymers (condensed tannins). Resveratrol, quercetin, and rutin are generally found in grape skin extracts, while catechin and epicatechin are found in the seeds (Figure 3 and 6). The phenolic content of grape seeds defatted with hexane then extracted with methanol and dried under vacuum has been reported to be about 5 mg/100 g, while the anthocyanin content is between 0.14 and 0.68 g/100 g (Rababah and others 2008). Iacopini and others (2008) assessed the antioxidant activity of the extracts and pure compounds using 2 different in vitro tests: scavenging of the stable DPPH radical and of authentic peroxynitrite (ONOO−). Antioxidant activities of grape seed extract ranged from 66.4% to 81.4%, compared to vitamin E that ranges from 90.3% to 94.7%. Monophenols, quercetin, rutin, and resveratrol may act either synergistically or antagonistically depending on their concentrations and the reaction temperature. Grape seed extract has been shown to inhibit both lipid hydroperoxide and propanal formation in an emulsion system (Hu and Skibsted 2002). Oligomeric procyanidins may be better antioxidants than their monomeric counterparts due to their ability to concentrate where the oxidative reaction is likely to occur. Resveratrol (trans-3,4 ,5-trihydroxystilbene), produced primarily in the grapevine, is present in various parts of the grape, including the skin. It has strong antioxidant activity exceeding that of propyl gallate, vanillin, phenol, BHT, and α-tocopherol (Murcia and Martinez-Tome 2001). This may be because it has more phenolic rings (2 compared with 1) than propyl gallate, phenol, and BHT, and because it has more −OH groups than α-tocopherol (3 compared with 1). Resveratrol inhibits peroxidation in a concentration-dependent manner. However, it does not scavenge hydroxyl radicals or does it react with H2 O2 , making it an inefficient catalyst of subsequent oxidation (Murcia and MartinezTome 2001). Some of the antioxidant compounds found in grape seed extract are shown in Table 6. Soares and others (2003) demonstrated that resveratrol, vitamins C and E, BHT, and propyl gallate were all able to significantly inhibit the oxidation of β-carotene by hydroxyl free radicals. Polyphenolic fractions from grape pomace can repair α-tocopherol by reducing the α-tocopheroxyl radical (Pazos and others 2009). Most of the phenolic compounds in fresh wine are derived from condensation of flavan-3-ol into oligomers (proanthocyanidins) and polymers (tannins; Granato and others 2011). Granato and others (2011) reported that the primary phenolics exerting antioxidant effects (DPPH and ORAC assays) in Brazilian red wines were nonanthocyanin flavonoids. Anthocyanins present in these wines were present solely in their monomeric form and ranged from about 9 to 237 mg/mL. Flavonoid content varied from 520 to 1795 mg catechin equivalents [CTE]/L. However, after evaluating 80 Spanish red wines, Rivero Perez and others (2008) found that the free anthocyanin fraction is the primary fraction responsible for antioxidant capacity and is correlated with electron transfer processes. Pazos and others (2006) evaluated the effectiveness of a grape phenol fraction, isolated grape procyanidins, hydroxytyrosol (from  c 2011 Institute of Food Technologists®

olive oil), and propyl gallate in inhibiting lipid oxidation in a fish (hake) microsomal model system. Oxidation was initiated by hemoglobin, enzymatic NADH iron and nonenzymatic ascorbate iron. The relative antioxidant efficiency was independent of the prooxidant system and was isolated grape procyanidin > propyl gallate > grape phenolic extract > hydroxytyrosol. Antioxidative effectiveness was positively correlated with incorporation of the substance into microsomes. However, polarity appeared to play less of a role in inhibition of hemoglobin oxidation by phenolics underscoring the fact that an exogenous antioxidant must be incorporated into membranes where unsaturated fatty acids and iron-reducing enzymes are located in order to be effective. Poiana and others (2008) demonstrated that during the ageing of red wine, polymeric anthocyanins increased from about 9% to over 75% after 6 mo, while monomeric anthocyanins decreased from over 75% to less than 24%. Total antioxidant capacity decreased and was highly correlated with the monomeric anthocyanin fraction (r > 0.98); however, free radical-scavenging ability increased and was highly correlated with the polymeric anthocyanin fraction. Granato and others (2010) also evaluated the antioxidant activity and the phenolic content of red wines and verified that ORAC values correlated well to flavonoid content (r = 0.47; P = 0.01), total phenolics (r = 0.44), and DPPH (r = 0.67). DPPH values also correlated well to the content of flavonoids (r = 0.69), total phenolic compounds (r = 0.60), and nonflavonoid compounds (r = 0.46) (in beers; Granato and others 2011).

The Stereochemistry of Flavanones Enantiomers are molecules that are mirror images of each another but cannot be superimposed onto one another. Molecules exhibit stereoisomerism (enantiomers) because they have one or more chiral centers. A chiral center results from the presence of an assymetrical carbon atom, that is, one that is attached to 4 different atoms or 4 different groups of atoms (making its mirror image nonsuperimposable). Enantiomers rotate the plane of polarized light in opposite directions. Enantiomer names use the R/S system. This system involves no reference but labels each chiral center R or S using a system in which its substituents are each assigned a priority, according to the Cahn-Ingold-Prelog priority rules (Cahn and others 1966), based on atomic number. If the center is oriented so that the lowest priority of the 4 substituents is pointed away from a viewer, the viewer will then see 2 possibilities: if the priority of the remaining 3 substituents decreases in a clockwise direction, it is labeled R (rectus), and if it decreases in a counterclockwise direction, it is S (sinister). Flavanones can have chiral carbon atoms; therefore, they can exist as S- and R-enantiomers. These enantiomers can be produced in different quantities in different plant materials under different growing conditions (Yanez and others 2005, 2008, 2007). They can have different effects in both biological and inorganic systems. For these reasons, separating and quantifying them has been of interest to the medical, biological, and agricultural industries in the recent past. The flavanone glycosides naringin and neohesperidin found in some citrus species have a chiral center in the C-2 position of the flavanone moiety (Uchiyama and others 2008; Figure 8). The flavanone hesperetin, the aglycone of hesperidin and major flavonoid in oranges, contains a chiral C-atom, so it can also exist as an S- and R-enantiomer. The 2S-herperidin and its Shesperitin aglycone predominate in nature (Uchiyama and others 2008).

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 237

Natural antioxidants . . .

OH HO

OH HO

OH

HO O

O

OH

O

O

OH

OH O O

O

OH

O

O

OH

Hesperitin

OH

Naringen OH OH

HO

O HO OH

O O

O

O

OH

O

HO

OH

OH

OH

Neohesperidin

OH OH

O

O

HO OH

O

O

O

O

HO OH

OH OH

OH OH

Hesperidin Figure 8–Natural antioxidants that exist as stereoisomers (hesperitin, naringin, neohesperidin, and hesperidin).

Enantiomers can react with other compounds or other enantiomers in different ways or at different rates. Brand and others (2010) have demonstrated small, but significant, differences in the metabolism and transport characteristics, and bioactivity between S- and R-hesperetin. Naringin, the major flavanone7-O-glycoside of sour orange, is responsible for the bitter taste of the fruit (Caccamese and others (2010). The relative ratios of naringin and neohesperidin to their C-2 epimers varies depend-

ing on species, maturity, and processing. Separation of naringin from neohesperidin is complicated by the presence of stereoisomers (Belboukhari and others 2010). Takemoto and others (2008) developed a high-performance liquid chromatography (HPLC) method using UV detection for the stereospecific analysis of the flavan, sakuranetin, found in grapefruit and oranges. Stereospecific HPLC methods have been developed for separation of epimers in tea, grapes, orange juice, and C-2 epimers from other sources

238 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . . and others 2001). Three primary extraction techniques are used for polyphenols: solvents, solid-phase extraction, and supercritical extraction. Using a Soxhlet apparatus combines percolation and immersion that increases extraction efficiency. Several extractions can be accomplished with solvents having different polarities (petrol ether, toluene, acetone, ethanol, methanol, ethyl acetate, and water). Methanol/water/HCl (70:29:1, v/v/v) has been shown to be the best among several solvents evaluated for extracting phenolics from grape seed (Xu and others 2010). Grinding in a mortar in liquid nitrogen provides uniform particle size allowing for a more consistent extraction. Ultrasound can be used to assist liquid solvent extraction. Xu and others (2010) reported that sequential sonication was a preferred to mechanical agitation as an extraction method for assessing phenolic content in grapeseed. Supercritical CO2 extraction can also be used (Schwarz and others 2001). Hydrodistillation of plant materials has several advantages. The essential oils that carry the intrinsic flavor of a spice can be removed and polyphenols, primary antioxidant compounds, are concentrated. In addition, the hydrodistilled compounds are generally more soluble in aqueous media than are those extracted using organic solvents. They are often more soluble than synthetic antioxidants as well. Hydrodistillation also avoids potential residues from organic solvents. Hydrodistilled extracts have also been reported to have a variety of functional effects in foods and in human health (Hinneburg and others 2006). Optimizing the extraction process could lead to even better results. The distillation process can also concentrate antioxidant components. Naz and others (2011) found that deodorizer distillates from sunflower oil processing were richer in tocopherols than the Effects of Processing Endogenous antioxidant systems (enzymatic) can be damaged deodorized oil itself. The implication is that, while the distillation during food processing (particle size reduction and heating), by process removes unwanted materials from the oil, it may, in fact, certain ingredients (salts and organic acids), and by storage con- concentrate some of the antioxidants. ditions (presence of oxygen) such that they are ineffective (Chen and others 1998). NaCl, in particular, reduces the activity of the Heat treatment When we think of processing, heat treatment is often the first antioxidant enzymes catalase, glutathione peroxidase, and superoxide dismutase that reduces their capacity to perform antioxidative process that comes to mind. Antioxidative activity of a given compound may increase, defunctions (Lee and others 1997). Ingredients, such as AA and citric crease, or remain unchanged as a function of temperature. Stability acid, can work synergistically with flavonoid antioxidants. Spices and herbs can be added to foods in various forms: whole, of an antioxidant to heat is advantageous in the food industry, since ground, or as isolates from their extracts. Extracting antioxidant many fat- and oil-containing foods are heated during processing ◦ components from a complex matrix depends on the solubility of and since heat is often the initiator of lipid oxidation. At 80 C, the extractant, the solvent, and the presence of other substances the antioxidative activity of δ-tocopherol is about twice that of αthat may compete with the extraction process, and the extraction tocopherol; however, it decreases as temperature increases from 80 ◦ process itself (vacuum, distillation, pressure, and so on). Because to 150 C. Antioxidative activity of α-tocopherol remains fairly ◦ these substances are aromatic, pungent food ingredients, they may constant between 80 and 110 C, decreasing only at tempera◦ or may not be desirable in a nonflavoring (antioxidant or other) tures above 110 C. Neither retains their antioxidative activity at ◦ application (Ruberto and others 2000; Teissedre and Waterhouse 150 C. Ginger extract has good thermal stability and inhibits more 2000). For example, even at low concentrations, some components ◦ of rosemary essential oil (verbenone, borneol, and camphor) can than 85% of linoleic acid peroxidation when heated at 185 C for ◦ impart a rosemary odor to foods (Carrillo and Tena 2006). Solid 120 min (Rehman and others 2003). Heating (120 C) dry ginger rosemary extract can contain >356 μg/g verbenone, 190 μg/g and turmeric essential oils results in different degrees of antioxidant activity retention. The antioxidant activity of turmeric oil is higher borneol, and >135 μg/g camphor (Carrillo and Tena 2006). after heating (120 ◦ C), unlike ginger oil that loses antioxidant activity (Tiwari and others 2006). Turmeric oil contains a higher Extraction Because many antioxidants are unstable to oxygen and endoge- concentration of monoterpenes than does ginger oil; however, nous enzymes, most are extracted from freeze-dried plant materi- release of bound antioxidants by the heat treatment should not be als. Selecting an appropriate extraction procedure can increase the ruled out. concentration of the antioxidant compound. Extraction using edible oil or fat is relatively simple. Herbs and spices can be mixed with Adding antioxidants to livestock diets fats, oils, or medium-chain triglycerides, allowed to extract under Including herb distillates into livestock diets can have positive defined time/temperature control, then filtered for use (Pokorny effects. Moclino and others (2008) found that feeding a steam (Uchiyama and others 2008; Caccamese and others 2010; VegaVillas and others 2008; Kim and others 2009; Belboukhari and others 2010). Si-Ahmed and others (2010) reported that different mobile phases in different ratios are required to accomplish enantiomeric and diastereomeric separation of a variety of flavanones (flavanone, 2 -hydroxyflavanone, 4 -hydroxyflavanone, 6-hydroxyflavanone, 7-hydroxyflavanone, 4 -methoxyflavanone, 6-methoxyflavanone, 7-methoxyflavanone, hesperetin, hesperidin, naringenin, and naringin). Others have reported similar differences in chiral discrimination ability (toward flavanones) depending on the buffer and alcohol modifier enantioselectivity (Cirilli and others 2008). Abbate and others (2009) described a method for assessing configurational and conformational properties (of naringenin) using vibrational circular dichroism. The stereoselectivity of chiral flavanones and epimers has significant biological effects in terms of their pharmacological activity and disposition in humans and livestock. Gardana and others (2009) reported that some human intestinal bacteria can transform diadzein to equol, O-desmethylangolensin, or dihydrodaidzein. Diet has a clear effect. A diet lower in fiber, vegetables, and cereals and higher in lipids from animal sources increases production of equol. These stereoselective differences in the chiral forms of flavonone antioxidants may result in differences in antioxidative effects of the various epimers depending on the matrix and oxidizing group. For these reasons, a concerted effort is being made to separate these chiral compounds and to evaluate their specific characteristics under defined conditions.

 c 2011 Institute of Food Technologists®

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 239

Natural antioxidants . . . distilled rosemary by-product to ewes increased rosmarinic acid, carnosol, and carnosic acid content in the meat. Fresh meat from these animals had higher total ferric reducing antioxidant power and lower DPPH values than controls indicating that the rosemary distillate partitioned into the meat tissues and reduced susceptibility to oxidation. McCarthy and others (2001) have shown similar results with pigs. Boler and others (2009) found that feeding vitamin E to pigs increased pork stability during storage. Simitzis and others (2008) found that meat from lambs fed a feed that had been sprayed with oregano essential oil (1 mL/kg) was much more stable to lipid oxidation during both refrigerated and frozen storage than that from controls. Gobert and others (2010) found that adding antioxidants to diets of cattle fed a polyunsaturated fatty acid (PUFA)-rich diet improved lipid stability in steaks; the combination of vitamin E and plant extracts rich in polyphenols was more efficient than vitamin E alone indicating some synergism between the 2.

Effects of the food matrix and ingredients Natural plant antioxidants can protect food components from oxidation under the stress of heating and storage. However, the inherent characteristics (ionic strength and pH) of the food, the food matrix (emulsion, foam, aqueous, and protein), and ingredients can influence antioxidant effectiveness. Vitamin E added to water-based food systems in an oil carrier concentrates in the neutral lipid fraction rather than the polar lipid fraction and is not an effective antioxidant. However, δ-tocopherol added using a polar carrier can be incorporated into the phospholipid fraction and is an effective antioxidant (Wills and others 2007). In a lard model system, the antioxidative activity of the tocopherols is temperature dependent (Reblova 2006). Wanatabe and others (2010) demonstrated that, in a methyl lineoleate/water emulsion, the effectiveness of AA and acyl ascorbates depended on whether the oxidation process was initiated by an oil-soluble prooxidant or a water-soluble prooxidant. The AA concentrated in the aqueous phase and suppressed oxidation to a greater degree at the oil/water interface when the prooxidant was water soluble. Docecanoyl and hexadecanoyl ascorbates dissolved in the oil phase and suppressed oxidation the oil phase (droplets) rather than at the interface. Increasing the pH appeared to enhance the electron-donating ability of AA in the water phase ultimately affecting oxidation. Hexadecanoyl ascorbate in the oil phase was not susceptible to these pH effects. Authors suggest that another explanation may be destabilization of the emulsion through flocculation and coalescence of the oil droplets at low pH. Thymol can prevent loss of α-tocopherol (in oil) following heating at 180 ◦ C for 10 min (Tomaino and others 2005). Using a lipophilic model system, Lee and Shibamoto (2002) demonstrated that volatile extracts of thyme (and basil) inhibited the oxidation of hexanal for 40 d. These extracts also inhibited methyl linoleate deterioration at 40 ◦ C. In sunflower oil, aroma detection thresholds of carvacrol, thymol, and p-cymene 2,3-diol have been reported to be 30, 124, and 794 ppm, respectively (Bitar and others 2008). pCymene 2,3-diol at 335 ppm imparted no negative flavor changes and reduced oxidation by more than 46%. Estevez and others (2008) evaluated several phenols (gallic acid, cyanidin-3-glucoside, (+)-epicatechin, chlorogenic acid, genistein, and rutin) and α-tocopherol in terms of anti- or prooxidative effects of oil-in-water emulsions containing myofibrillar proteins (1%). Gallic acid, cyanidin-3-glucoside, and genistein were the most efficient inhibitors of lipid and protein oxidation. They con-

cluded that the nature and conformation of the proteins as well as the chemical structure of the phenols influenced the overall effect. Antioxidant content of raw materials can change over time and are likely related to storage conditions. Hatzidimitrioua and others (2007) reported that total phenol content of grape seeds decreases during storage. Changes were minor for samples stored at less than 55% relative humidity; however, high humidity (75%) accelerated degradation resulting in a 50% reduction of total phenol content. Based on the continuous gallic acid release, authors suggested that this degradation was related to hydrolytic reactions. Modifications of the storage process would be expected to enhance retention of antioxidative compounds in grape seeds. Ingredients, such as salt, can act as prooxidants in food systems; however, antioxidants can help reduce it. Brannan (2008) found that grape seed extract helps to mitigate the prooxidative effects of NaCl in stored ground chicken without affecting moisture content or pH. The author suggests that grapeseed extract may alter the effect of NaCl on protein solubility in salted chicken patties. Whether it affects physicochemical interactions in cooked meat quality remains to be assessed. Akarpat and others (2008) demonstrated that adding a hot water extract of rosemary (10%) to ground beef containing salt (1.5%) protected color and preserved oxidative quality during frozen storage (120 d). Fasseas and others (2008) found that essential oils from oregano and sage added to ground beef and pork (3% w/w) reduced oxidation. The effect was even more dramatic in cooked meat than in raw meat. The antioxidant components of rosemary, sage, basil, black pepper, garlic, and onion appear to be relatively stable. Microwave treatment of these herbs has no effect on reducing power or ironchelating capacity (Bertelli and others 2004). However, the effects on other components, such as flavor components and pigments, are unknown. Marinating and cooking (chicken) significantly reduces the antioxidant activities of marinating sauces and consequently reduces the amounts of antioxidant available (Thomas and others 2010). Marinating chicken (in herb and spice-based marinades) prior to cooking reduced the total antioxidant activity (45% to 70%) originally present in the sauce. This may be due to the ionic effects of various salts typically included in marinades, the effects of reduced pH on the phenolic components of the marinades, and/or to the interactions between antioxidants or between antioxidants and protein. Loss of antioxidant activity due to cooking may reflect the protective action of antioxidants on proteins (which are denatured by heating) or their protective action toward other components (vitamins). In addition to reducing lipid oxidation, antioxidants may have other benefits in food systems. Adding rosemary essential oil and/or citrus fiber washing water to bologna has been shown to lower the levels of residual nitrite (Viuda-Martos and others 2010). Flavonoids, hesperidin, and narirutin were identified in the bologna with hesperidin concentrations being higher than narirutin concentrations. The preferred (sensory) sample was that which contained 50 g/kg citrus fiber water and 200 mg/kg rosemary essential oil. There are many types of food matrices to which these antioxidant compounds might be added and many types of processing that the product might then undergo. There are currently no general guidelines as to what/when to use plant extracts in food matrices. More studies are necessary to elucidate that substances are effective in what systems and under what condition.

240 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . .

Synergism Combining antioxidants may increase their effectiveness. Smet and others (2008) found that dietary synthetic antioxidants combined with α-tocopherol were more effective than rosemary, green tea, grape seed, or tomato extracts (100 to 200 ppm) alone or in combination in sparing tocopherols oxidation and in preventing oxidation of fresh frozen chicken patties. It has been proposed that the mixed free radical acceptors involve 2 antioxidants: one that reacts with the peroxy radical (and is consumed) and a 2nd that regenerates the 1st, effectively sparing. Phenolic antioxidants and AA appear to work synergistically in this way (Uri 1961). r

(1) ROO + A:H = ROO:H +A r r (2) A + B:H = A:H + B .

r

Some acidic compounds, such as AA and citric acid, can exert a synergistic effect when added along with polyphenolic antioxidants. These acidic compounds chelate metals. These synergists form an antioxidant radical synergist complex (A:S) such thatr neir ther the antioxidant radical (A ) nor the synergist radical (S ) can catalyze oxidation reactions. This chemical association suppresses the antioxidant radical’s ability to assist in the breakdown of lipid peroxides (Aurand and Woods 1979). Addition of anthocyanin can prevent oxidation of AA by metal ions such as copper (Sarma and others 1997). Anthocyanin not only chelates metal ions, but also forms an AA (copigment-metalanthocyanin) complex that may be the basis for its antioxidative activity. Because of the number of −OH groups on the aromatic rings, and because of their water solubility, anthocyanins are pH-sensitive. In a basic solution, the −OH groups can give up r H+; in a more neutral environment, they can donater H to an r oxidizing lipid (ROO ). For this reason, the antioxidative capacity of an anthocyanin is dependent on the anthocyanin itself (number and location of −OH groups), the pH of the surrounding environment, and the other components of the system (metals, continuous phase). Lee and others (2005) found that combinations of chelators (sodium tripolyphosphate or sodium citrate) with reductants (erythorbate), and/or free radical scavengers (BHA and rosemary extract) were effective antioxidants. The combination of rosemary and erythorbate was most effective in delaying lipid oxidation in ground beef. The rosemary/citrate/erythorbate combination was most effective in stabilizing color and delaying lipid oxidation. These findings indicate that combining a reductant with a free radical scavenger is more effective at preventing lipid oxidation than either alone. In a mixture of 3 monophenols (catechin, resveratrol, and/or quercetin) derived from grapeseed, Pinelo and others (2004) found an initial increase in antioxidative activity followed by a subsequent decrease for all solution combinations. They also reported a possible synergy between quercetin, rutin, and resveratrol toward ONOO−. The effect was additive for catechin and epicatechin. These compounds may be acting independently, while other combinations may react with each other. Granato and others (2010) found that (in brown ales) flavonoids, total phenolics, and nonflavonoid phenolics (hydroxycinnamates and hydroxybenzoates), derived from both the malt and the hops, are strongly correlated with antioxidant activity (ORAC and DPPH). Ghiselli and others (2000) have shown that beer increases serum antioxidant capacity. Ethanol increases absorption of phenolic acids. However, the increase in antioxidant capacity is not due to either ethanol or phenolic acids alone, but rather because of a synergistic effect between the 2.  c 2011 Institute of Food Technologists®

Regulatory Status of Extracts, Concentrates, and Resins Synthetic antioxidants (BHA, BHT, and EDTA) are regulated by the Food and Drug Administration (FDA) as direct food additives. They may be used alone or in combination not to exceed 0.02% (2 ppm) of the final product in specified food products (21CFR172.110). These antioxidants are considered to be safe and suitable ingredients for use in meat, poultry, and egg products, alone or in combination, not to exceed 0.02% of the fat content (FSIS Directive 7120.1. revision 5). Some herb and spice extracts and oleoresins are Generally Recognized As Safe (GRAS). Some are considered to be indirect additives (21 CFR Vol. 3. Part 101); as such, solvents permitted for the extraction process and solvent residues allowed are specified. Some extracts, concentrates, and resins are regulated by the FDA “Dietary Supplement Health and Education Act of 1994” and are considered to be one (or more) of several defined dietary ingredients (a vitamin, a mineral, an herb or other botanical, amino acid, a dietary substance for use by man to supplement the diet by increasing the total dietary intake, or a concentrate, metabolite, constituent, extract, or combination of any ingredient described in clause (A), (B), (C), (D), or (E) and is excluded from regulation as a food additive. Extracts, concentrates, and resins are also regulated under the Food Labeling Regulation, Amendments; Food Regulation Uniform Compliance Date; and New Dietary Ingredient Premarket Notification Final Rule (1997). If they are added to cause flavor or color changes, they are regulated as such and specific quantities allowable for use in various foods are set forth. Based on the number of various classifications under which an extract, concentrate or resin could be covered, allowable use levels vary widely.

Summary Plant and animal tissues contain unsaturated fatty acids, primarily in the phospholipid fraction of cell membranes. These lipids are especially susceptible to oxidation because of their electrondeficient double bonds. The breakdown products of oxidation can produce off-odors, new flavors, loss of nutrient content, and color deterioration. To manufacture high-quality, stable food products, the most effective solution is often the addition of antioxidants, either synthetic or natural, which can serve as “chain breakers,” by intercepting the free radicals generated during various stages of oxidation or to chelate metals. Chain-breaking antioxidants are generally the most effective. A common feature of these compounds is that they have one or more aromatic rings (often phenolic) with one or more −OH groups capable of donating H· to the oxidizing lipid. Synthetic antioxidants, such as BHA, BHT, and propyl gallate, have one aromatic ring. The natural antioxidants AA and α-tocopherol each have 1 aromatic ring as well. However, many of the natural antioxidants (flavonoids and anthocyanins) have more than 1 aromatic ring. The effectiveness of these aromatic antioxidants is generally proportional to the number of −OH groups present on the aromatic ring(s). Depending on the arrangement of the −OH groups, these compounds may also chelate prooxidative metals. The facts that they are natural, and have antioxidative activity that is as good or better than the synthetic antioxidants, makes them particularly attractive for commercial food processors because of consumer demand for natural ingredients. References Abbate S, Burgi LF, Castiglioni E, Lebon F, Longhi G, Toscano E, Caccamese S. 2009. Assessment of configurational and conformational

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 241

Natural antioxidants . . . properties of naringenin by vibrational circular dichroism. Chirality 21(4):436–41. Abdullin IF, Turova EN, Budnikov GK. 2001. Coulometric determination of the antioxidant capacity of tea extracts using electrogenerated bromine. Zhurnal Analiticheskoi Khimi 56(6):627–9. Acree T, Arn H. 2004. Flavornet. Available from: http://www.flavornet.org,  c Datu Inc. Accessed Jun 2008. Aftab N, Vieira A. 2009. Antioxidant activities of curcumin and combinations of this curcuminoid with other phytochemicals. Phytother Res 24(4): 500–2. Agati G, Matteini P, Goti A, Tattini M. 2007. Chloroplast-located flavonoids can scavenge singlet oxygen. New Phytol 174(1):77–81. Ahn J, Gr¨un IU, Fernando LN 2002. Antioxidant properties of natural plant extracts containing polyphenolic compounds in cooked ground beef. J Food Sci 67(4):1364–9. Ahn J, Gr¨un IU, Mustapha A. 2007. Effects of plant extracts on microbial growth, color change, and lipid oxidation in cooked beef. Food Micro 24:7–14. Akarpat A, Turhan S, Ustun NN. 2008. Effects of hot water extracts from myrtle, rosemary, nettle and lemon balm leaves on lipid oxidation and color of beef patties during frozen storage. J Food Process Preserv 32(1):117–32. Alamed J, Chaiyasit W, McClements DJ, Decker EA. 2009. Relationships between free radical scavenging and antioxidant activity in foods. J Agric Food Chem 57(7):2969–76. Allam SSM, Mohamed HMA. 2002. Thermal stability of some commercial natural and synthetic antioxidants and their mixtures. J Food Lipids 9(4):277–93. Amagase H. 2006. Clarifying the real bioactive constituents of garlic. J Nut 136(Suppl 3):716S–25S. Anand P, Thomas SG, Kunnumakkara AB, Sundaram C, Harikumar KB, Sung B, Tharakan ST, Misra K, Priyadarsini IK, Rajasekharan KN, Aggarwal BB. 2008. Biological activities of curcumin and its analogues (congeners) made by man and mother nature. Biochem Pharm 7(6):1590–611. Aoki T, Wada S. 2003. Phenolics composition and antioxidant activity of sweet basil (Ocimum basilicum L. J Agric Food Chem 51(15):4442–9. Apak R, Guclu K, Ozyurek M, Karademir SE, Ercag E. 2006. The cupric ion reducing antioxidant capacity and polyphenolic content of some herbal teas. Internat J Food Sci Nut 57(5–6):292–304. Arouma OI, Halliwell B, Aeschbach R, Loligers J. 1992. Antioxidant and pro-oxidant properties of active rosemary constituents: carnosol and carnosic acid. Xenobiotica 22:257–68. Aurand LW, Woods AE. 1979. Lipids. Chapter 5. In: Aurand LW, Woods AE, editors. Food chemistry. Westport, Conn.: The AVI Publishing Company, Inc. p 125–6. Bak I, Lekli I, Juhasz B, Varga E, Varga B, Gesztelyi R, Szendrei L, Tosaki A. 2010. Isolation and analysis of bioactive constituents of sour cherry (Prunus cerasus) seed kernel: an emerging functional food. J Med Food 13(4): 905–10. Belboukhari N, Cheriti A, Roussel C, Vanthuyne N. 2010. Chiral separation of hesperidin and naringin and its analysis in a butanol extract of Launeae arborescens. Nat Prod Res 24(7):669–81. Bendini A, Toschi TG, Lercker G. 2002. Antioxidant activity of oregano (Origanum vulgare L.) leaves. Italian J Food Sci 14(1):17–24. Berger RG. 2009. Biotechnology of flavours-the next generation. Biotech Let 31(11):1651–9. Bertelli D, Plessi M, Miglietta F. 2004. Effect of industrial microwave treatment on the antioxidant activity of herbs and spices. Italian J Food Sci 16(1):97–103. Bitar A, Ghaddar T, Malek A, Haddad T, Toufeili I. 2008. Sensory thresholds of selected phenolic constituents from thyme and their antioxidant potential in sunflower oil. J Am Oil Chem Soc 85(7): 641–6. Boler DD, Gabriel SR, Yang H, Balsbaugh R, Mahan DC, Brewer MS, McKeith FK, Killefer J, 2009. Effect of different dietary levels of natural-source vitamin E in grow-finish pigs on pork quality and shelf life. Meat Sci 83(4):723–30. Bozin B, Mimica-Dukic N, Simin N, Anackov G. 2006. Characterization of the volatile composition of essential oils of some Lamiaceae spices and the antimicrobial and antioxidant activities of the entire oils. J Agric Food Chem 54(5):1822–8. Brand W, Shao J, Hoek-van den Hil EF, van Elk KN, Spenkelink B, de Haan LH, Rein MJ, Dionisi F, Williamson G, van Bladeren PJ, Rietjens

IM. 2010. Stereoselective conjugation, transport and bioactivity of S- and R-hesperetin enantiomers in vitro. J Agric Food Chem 58(10):6119–25. Brannan RG. 2008. Effect of grape seed extract on physicochemical properties of ground, salted, chicken thigh meat during refrigerated storage at different relative humidity levels. J Food Sci 73(1):C36–C40. Brewer MS, Prestat C. 2002. Consumer attitudes towards issues in food safety. J Food Safety 22(2):67–85. Brewer MS, Russon C. 1994. Consumer attitudes towards food safety issues. J Food Saf 14:63–76. Brown JE, Kelly MF. 2007. Inhibition of lipid peroxidation by anthocyanins, anthocyanidins and their phenolic degradation products. Eur J Lipid Sci Techno 109(1):66–71. Buricova L, Reblova Z. 2008. Czech medicinal plants as possible sources of antioxidants. Czech J Food Sci 26(2):132–8. Cabrera C, Gimenez R, Lopez MC. 2003. Determination of tea components with antioxidant activity. J Agric Food Chem 51(15):4427–35. Caccamese S, Bianca S, Santo D. 2010. Racemization at C-2 of naringin in sour oranges with increasing maturity determined by chiral high-performance liquid chromatography. J Agric Food Chem 55(10):3816–22. Cahn RS, Ingold CK, Prelog V. 1966. Specification of molecular chirality. Angew Chem Intl Ed 5(4):385–415. DOI:10.1002/anie.196603851. Caldwell CR. 2001. Oxygen radical absorbance capacity of the phenolic compounds in plant extracts fractionated by high-performance liquid chromatography. Analyt Biochem 293(2):232–8. Calliste CA, Kozlowski D, Duroux JL, Champavier Y, Chulia AJ, Trouillas P. 2010. A new antioxidant from wild nutmeg. Food Chem 118(3):489–96. Carrillo JD, Tena MT. 2006. Determination of volatile compounds in antioxidant rosemary extracts by multiple headspace solid-phase microextraction and gas chromatography. Flavor Frag J 21(4):626–33. Carolina C, Manzan M, Toniolo FS, Bredow E, Povh NP. 2003. Extraction of essential oil and pigments from Curcuma longa [L.] by steam distillation and extraction with volatile solvents. J Agric Food Chem 51(23):6802–7. Carvalho RN, Moura LS, Rosa PTV, Meireles MAA. 2005. Supercritical fluid extraction from rosemary (Rosmarinus officinalis): kinetic data, extract’s global yield, composition, and antioxidant activity. J Supercrit Fluids 35:197–204. Chaieb K, Hajlaoui H, Zmantar T, Kahla-Nakbi AB, Rouabhia M, Mahdouani K, Bakhrouf A. 2007. The chemical composition and biological activity of clove essential oil, Eugenia caryophyllata (Syzigium aromaticum L. Myrtaceae): a short review. Phytother Res 21(6):501–6. Chaiyasit W, McClements DJ, Decker EA. 2005. The relationship between the physicochemical properties of antioxidants and their ability to inhibit lipid oxidation in bulk oil and oil-in-water emulsions. J Agric Food Chem 53(12):4982–8. Chatterjee S, Niaz Z, Gautam S, Adhikari S, Variyar PS, Sharma A. 2007. Antioxidant activity of some phenolic constituents from green pepper (Piper nigrum L.) and fresh nutmeg mace (Myristica fragrans). Food Chem 101(2):515–23. ChemBioOffice. 2008. ChemDraw Ultra 11. Available from: www.Cambridge.com. Accessed Nov 2010. Chen HY, Lin YC, Hsieh CL. 2007. Evaluation of antioxidant activity of aqueous extract of some selected nutraceutical herbs. Food Chem 104(4):1418–24. Chen Z. 2008. Research of antioxidative capacity in essential oils of plants. China Cond 11:40–3. Chen ZY, Wang LY, Chan PT, Zhang Z, Chung HY, Liang C. 1998. Antioxidative activity of green tea catechin extract compared with that of rosemary extract. J Am Oil Chem Soc 75(9):1141–5. Cirilli R, Ferretti R, De Santis E, Gallinella B, Zanitti L, La Torre F. 2008. High-performance liquid chromatography separation of enantiomers of flavanone and 2 -hydroxychalcone under reversed-phase conditions. J Chrom 1190(1–2):95–101. Cuvelier ME, Berset C, Richard H. 1994. Antioxidant constituents in sage (Salvia officinalis). J Agric Food Chem 42:665–9. Deans SG, Simpson EJM. 2000. Antioxidants from Salvia officinalis. In: Kinzios SE, editor. The genus salvia. Amsterdam, the Netherlands: Harwood Academic Publishers. p 189–92. Decker EA. 2002. Antioxidant mechanisms. In: Akoh D, Min DB, editors. Lipid chemistry. 2nd ed. New York: Marcel Dekker, Inc. p 530–2. Defrancesco E, Trestini S. 2008. Consumers’ willingness to pay for the health promoting properties of organic fresh tomato. Rivista di Economia Agraria 63(4):517–45.

242 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . . Dixon RA, Xie DY, Sharma SB. 2005. Proanthocyanidins—a final frontier in flavonoid research? New Phys 165:9–28. Dorman HJD, Peltoketo A, Hiltunen R, Tikkanen MJ 2003. Characterisation of the antioxidant properties of de-odourised aqueous extracts from selected Lamiaceae herbs. Food Chem 83(2):255–62. Dorman HJD, Surai P, Deans SG. 2000. In vitro antioxidant activity of a number of plant essential oils and phytoconstituents. J Essen Oil Res 12(2):241–8. Duh PD, Yen GC, Yen WJ, Wang BS, Chang LW. 2004. Effects of pu-erh tea on oxidative damage and nitric oxide scavenging. J Agric Food Chem 52:8169–76. Edris AE, Shalaby A, Fadel HM. 2003. Effect of organic agriculture practices on the volatile aroma components of some essential oil plants growing in Egypt II: sweet marjoram (Origanum marjorana L.) essential oil. Flav Frag J 18(4):345–51. El-Ghorab H, Nauman M, Anjum FM, Hussain S, Nadeem M. 2010. Comparative study on chemical composition and antioxidant activity of ginger (Zingiber officinale) and cumin (Cuminum cyminum). J Agric Food Chem 58(14):8231–7. Elmore S, Mottram DS, Enser M, Wood JD. 1999. Effect of the polyunsaturated fatty acid composition of beef muscle on the profile of aroma volatiles. J Agric Food Chem 47(4):1619–25. Estevez M, Kylli P, Puolanne E, Kivikari R, Heinonen M. 2008. Oxidation of skeletal muscle myofibrillar proteins in oil-in-water emulsions: interaction with lipids and effect of selected phenolic compounds. J Agric Food Chem 56(22):10933–40. Est´evez M, Heinonen M. 2010. Effect of phenolic compounds on the formation of alpha-aminoadipic and gamma-glutamic semialdehydes from myofibrillar proteins oxidized by copper, iron, and myoglobin. J Agric Food Chem 58(7):4448–55. Evans G, de Challemaison B, Cox DN. 2010. Consumers’ ratings of the natural and unnatural qualities of foods. Appetite 54(3):557–63. FDA. 2009. Compliance Policy Guides. CPG Sec. 525.750 Spices—Definitions. Available from: www.fda.gov/ICECI/ ComplianceManuals/CompliancePolicyGuidanceManual/ucm074468.htm.. Accessed Jan 2010. Fasseas MK, Mountzouris KC, Tarantilis PA, Polissiou M, Zervas G 2008. Antioxidant activity in meat treated with oregano and sage essential oils. Food Chem 106(3):1188–94. Federal Register Final Rule—62 FR 49826 September 23, 1997–Food Labeling Regulation, Amendments; Food Regulation Uniform Compliance Date; and New Dietary Ingredient Premarket Notification. Federal Register 62(184):49825–58. Fernandez MT, Mira ML, Florencio MH, Jennings KR. 2002. Iron and copper chelation by flavonoids: an electrospray mass spectrometry study. J Inorg Biochem 92(2):105–111. Fisher C. 1992. Phenolic compounds in spices. Chapter 9. In: Phenolic compounds in food and their effects on health. ACS Symposium Series. ACS 506:118–29. Formanek Z, Kerry JP, Higgins FM, Buckley DJ, Morrissey PA, Farkas J. 2001. Addition of synthetic and natural antioxidants to α-tocopheryl acetate supplemented beef patties; effects of antioxidants and packaging on lipid oxidation. Meat Sci 58:337–41. Frankel EN. 1991. Recent advances in lipid oxidation. J Agric Food Chem 54(4):495–511. Frankel EN, Huang SW, Aeschbach R, Prior E. 1996. Antioxidant activity of a rosemary extract and its constituents, carnosinic acid, carnosol and rosmarinic acid, in bulk oil and oil-in-water emulsion. J Agric Food Chem 44:131–5. Fukumoto L, Mazza G. 2000. Assessing antioxidant and prooxidant activities of phenolic compounds. J Agric Food Chem 48:3597–604. Gardana C, Canzi E, Simonetti P. 2009. The role of diet in the metabolism of daidzein by human faecal microbiota sampled from Italian volunteers. J Nut Biochem 20(12):940–7. Gardner PT, McPhail DB, Duthie GG. 1998. Electron spin resonance spectroscopic assessment of the antioxidant potential of teas in aqueous and organic media. J Agric Food Chem 76(2):257–62. Geldof N, Engeseth NJ. 2002 Antioxidant capacity of honeys from various floral sources based on the determination of oxygen radical absorbance capacity and inhibition of in vitro lipoprotein oxidation in human serum samples. J Agric Food Chem 50:3050–5. Ghiselli A, Natella F, Guidi A, Montanari I, Fantozzi P, Scaccini C. 2000. Beer increases plasma antioxidant capacity in humans. J Nutr Biochem 11(2):76–80.  c 2011 Institute of Food Technologists®

Giuffrida F, Destaillats F, Egart MH, Hug B, Gola PA, Skibsted LH, Dionisi F. 2007. Activity and thermal stability of antioxidants by differential scanning calorimetry and electron spin resonance spectroscopy. Food Chem 101(3):1108–14. Granato D, Castro IA, Katayama FCU. 2010. Assessing the association between phenolic compounds and the antioxidant activity of Brazilian red wines using chemometrics. LWT Food Sci Technol 43:1542–9. Granato D, Branco GF, Faria Jde A, Cruz AG. 2011. Characterization of Brazilian lager and brown ale beers based on color, phenolic compounds, and antioxidant activity using chemometrics. J Sci Food Agric 91(3): 563–71. Grey CE, Adlercreutz P. 2006. Evaluation of multiple oxidation products for monitoring effects of antioxidants in Fenton oxidation of 2 -deoxyguanosine. J Agric Food Chem 54(6):2350–8. Gobert M, Gruffat D, Habeanu M, Parafita E, Bauchart D, Durand D. 2010. Plant extracts combined with vitamin E in PUFA-rich diets of cull cows protect processed beef against lipid oxidation. Meat Sci 85(4): 676–83. Gorinstein M, Leontowicz H, Leontowicz M, Namiesnik J, Najman K, Drzewiecki J, Martincov´a O, Katrich E, Trakhtenberg S. 2008. Comparison of the main bioactive compounds and antioxidant activities in garlic and white and red onions after treatment protocols. J Agric Food Chem 56(12):4418–26. Guo Q, Richert JR, Burgess DM, Webel DE, Orr D, Blair M. 2006. Effects of dietary vitamin E and fat supplementation on pork quality. J Anim Sci 84(11):3089–99. Guo Q, Zhao B, Li M, Shen S, Xin W. 1996. Studies on protective mechanisms of four components of green tea polyphenols against lipid peroxidation in synaptosomes. Biochim et Biophys Acta 1304(3):210–22. Gurdip S, Marimuthu P, Catalan C, de Lampasona MP. 2004. Chemical, antioxidant and antifungal activities of volatile oil of black pepper and its acetone extract. J Agric Food Chem 84(14):1878–84. Halliwell B. 1997. Antioxidants in human health and disease. Annual Review of Nutrition 16:33–50. Halliwell B, Gutteridge JMC, Aruoma O. 1987. The deoxyribose method: a simple “test tube” assay for determination of rate constants for reactions of hydroxyl radicals. Anal Biochem 165:215–9. Halvorsen R, Carlsen M, Phillips K, Bohn S, Holte K, Jacobs D, Jr, Blonhoff R. 2006. Content of redox-active compounds (ie, antioxidants) in foods consumed in the United States. Am J Clin Nut 84:95–135. Hatzidimitrioua E, Nenadisa N, Tsimidou MZ. 2007. Changes in the catechin and epicatechin content of grape seeds on storage under different water activity (aw) conditions. Food Chem 105(4):1504–11 Hernandez-Hernandez E, Ponce-Alquicira E, Jaramillo-Flores ME, Guerrero-Legarreta I. 2009. Antioxidant effect of rosemary (Rosmarinus officinalis L.) and oregano (Origanum vulgare L.) extracts on TBARS and colour of model raw pork batters. Meat Sci 81(2):410–17. Hillmann J. 2010. Reformulation key for consumer appeal into the next decade. Food Rev 37(1):14, 16, 18–9. Hinneberg I, Dorman DHJ, Hiltunen R. 2006. Antioxidant activities of extracts from selected culinary herbs and spices. Food Chem 97: 122–9. Hofstrand D. 2008. Domestic perspectives on food versus fuel. Agriculture Marketing Resource Center. Available from: www.agmrc.org. Accessed Feb 2010. Hra HR, Hadolina M, Knez E, Bauman D. 2000. Comparison of antioxidative and synergistic effects of rosemary extract with α-tocopherol, ascorbyl palmitate and citric acid in sunflower oil. Food Chem 71(2):229–33. Hu M, Skibsted LH. 2002. Kinetics of reduction of ferrylmyoglobin by (-)-epigallocatechin gallate and green tea extract. J Agric Food Chem 50:2298–3003. Huang SW, Frankel EN. 1997. Antioxidant activity of tea catechins in different lipid systems. J Agric Food Chem 45(8):3033–8. Hussain I, Anwara F, Sherazi STH, Przybylski R. 2008. Chemical composition, antioxidant and antimicrobial activities of basil (Ocimum basilicum) essential oils depends on seasonal variations. Food Chem 108(3):986–95. Iacopini P, Baldi M, Storchi P, Sebastiani L. 2008. Catechin, epicatechin, quercetin, rutin, and resveratrol in red grapes: content, in vitro antioxidant activity and interactions. J Food Comp Anal 21(8):589–98. Jayaprakasha GK, Rao LJM, Sakariah KK. 2003. Volatile constituents from Cinnamomum zeylanicum fruit stalks and their antioxidant activities. J Agric Food Chem 5(15):4344–8.

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 243

Natural antioxidants . . . Jayaprakasha GK, Singh RP, Sakariah KK. 2001. Antioxidant activity of grape seed (Vitis vinifera) extracts on peroxidation models in vitro. Food Chem 73(3):285–90. Jayaprakashaa GK, Rao LJ, Sakariaha KK. 2006. Antioxidant activities of curcumin, demethoxycurcumin and bisdemethoxycurcumin. Food Chem 98(4):720–4. Jensen MT, Hansen L, Andersen HJ. 2002. Transfer of the meat aroma precursors (dimethyl sulfide, dimethyl disulfide and dimethyl trisulfide) from feed to cooked pork. LWT— Food Sci Technol 35(6):485–9. Jirovetz L, Buchbauer G, Stoilova I, Stoyanova A, Krastanov A, Schmidt E. 2006. Chemical composition and antioxidant properties of clove leaf essential oil. J Agric Food Chem 54(17):6303–7. Jo C, Ahn HJ, Son JH, Lee J, Byun MW. 2003. Packaging and irradiation effect on lipid oxidation, color, residual nitrite content, and nitrosamine formation in cooked pork sausage. Food Control 14(1):7–12. Joppen L. 2006. Taking out the chemistry. Food Eng Ingred 31(2): 38–9, 41. Jukic M, Politeo O, Milos M. 2006. Chemical composition and antioxidant effect of free volatile aglycones from nutmeg (Myristica fragrans Houtt.) compared to its essential oil. Croatica Chem Acta 79:209–14. Juliani HR, Simon JE. 2002. Antioxidant activity of basil (Ocimum spp.). In: Janick J. editor. New crops and new uses: strength in diversity. Virginia: ASHS Press. p 575–9. Jun WJ, Han BK, Yu KW, Kim MS, Chang IS, Kim HY, Cho HC. 2001. Antioxidant effects of Origanum majorana L. on superoxide anion radicals. Food Chem 75(4):439–44. Kapoor IPS, Singh B, Singh G, Heluani CS de, Lampasona MP, Catalan CAN. 2009. Chemical and in vitro antioxidant activity of volatile oil and oleoresins of black pepper (Piper nigrum). J Agric Food Chem 57(12):5358–64. Kerler J, Grosch W. 1996. Odorants contributing to warmed-over flavor (WOF) of refrigerated cooked beef. J Food Sci 61(6):1271–4, 1284. Khanduja KL. 2003. Stable free radical scavenging and antiperoxidative properties of resveratrol in vitro compared with some other bioflavonoids. Ind J Biochem Biophys 40:416–22. Khatun M, Eguchi S, Yamaguchi T, Takamura H, Matoba T. 2006. Effect of thermal treatment on radical-scavenging activity of some spices. Food Sci Technol Res 12(3):178–85. Khokhar S, Owusu-Apenten RK. 2003. Iron binding characteristics of phenolic compounds: some tentative structure-activity relations. Food Chem 81(1):133–40. Kiehne A, Engelhardt U. 1996. Thermospray LC-MS analysis of various groups of polyphenols in tea. II: chlorogenic acids, theaflavins and thearubigins. Z Leibenm untersForsch 202(4):299–302. Kikuzaki H, Nakatani N. 2006. Antioxidant effects of some ginger constituents. J Food Sci 58(6):1407–10. Kim K, Cho HY. 2001. Antioxidant effects of Origanum majorana L. on superoxide anion radicals. Food Chem 75(4):439–44. Kim H, Choi Y, Lim J, Paik SR, Jung S. 2009. Chiral separation of catechin by capillary electrophoresis using mon-, di-, tri-succinyl-beta-cyclodextrin as chiral selectors. Chirality 221(10):937–42. Kondo K, Kurihara M, Fukuhara K. 2001. Mechanism of antioxidant effect of catechins. Meth Enzymol 335:203–17. Lachman J, Sulik M, Shilla M. 2007. Comparison of the total antioxidant activity status of Bohemian wines during the wine-making process. Food Chem 103:802–7. Lahucky R, Nuernberg K, Kovac L, Bucko O, Nuernberg G. 2010. Assessment of the antioxidant potential of selected plant extracts—in vitro and in vivo experiments on pork. Meat Sci 85(4):779–84. Lee AJ, Umano K, Shibamoto T, Lee KG. 2005a. Identification of volatile components in basil (Ocimum basilicum L.) and thyme leaves (Thymus vulgaris L.) and their antioxidant properties. Food Chem 91(1):131–7. Lee S, Decker EA, Faustman C, Mancini RA 2005b. The effects of antioxidant combinations on color and lipid oxidation in n-3 oil fortified ground beef patties. Meat Sci 70(4):683–9. Lee J, Scagel CF. 2009. Chicoric acid found in basil (Ocimum basilicum L.) leaves. Food Chem 115:650–6. Lee KG, Shibamoto T. 2002. Determination of antioxidant potential of volatile extracts isolated from various herbs and spices. J Agric Food Chem 50(17):4947–52. Lee SK, Mei L, Decker EA. 1997. Influence of sodium chloride on antioxidant enzyme activity and lipid oxidation in frozen ground pork. Meat Sci 46(4):349–55.

Leelarungrayub N, Rattanapanone V, Chanarat N, Gebicki JM. 2006. Quantitative evaluation of the antioxidant properties of garlic and shallot preparations. Nutrition 22(3):266–74. Liang RJ, Shi SD, Ma YJ. 2010. Analysis of volatile oil composition of the peppers from different production areas. Med Chem Res 19(2):157–65. Lien A, He H, Pham-Huy C. 2008. Green tea and health: an overview. J Food Agric Environ 6(1):6–13. Lugasi A, Dworschak E, Hovari J 1995. Characterization of scavenging activity of natural polyphenols by chemiluminescence technique. Federation of the European Chemists’ Society. Proceedings of the European Food Chemists. VIII, Vienna, Austria, September. 3:639–43. Lupea AX, Pop M, Cacig S. 2008. Structure-radical scavenging activity relationships of flavonoids from Ziziphus and Hydrangea extracts. Rev Chim 59(3):309–13. McCarthy TL, Kerry JP, Kerry JF, Lynch PB, Buckley DJ. 2001. Evaluation of the antioxidant potential of natural food/plant extracts as compared with synthetic antioxidants and vitamin E in raw and cooked pork patties. Meat Sci 58(1):45–52. McCormick Institute Antioxidant Comparison. 2009. Curcumin: a natural standards review. Available from: www.mccormickscienceinstitute. com/content.cfm?ID=10480. Accessed Feb 2010. Madsen HL, Nielsen BR, Bertelsen G, Skibsted LH. 1996. Screening of antioxidative activity of spices. A comparison between assays based on ESR spin trapping and electrochemical measurement of oxygen consumption. Food Chem 57(2):331–7. Mahrour A, Caillet S, Nketsia-Tabiri J, Lacroix M. 2003. The antioxidant effect of natural substances on lipids during irradiation of chicken legs. J Am Oil Chem Soc 80(7):679–84. Marinova E, Toneva A, Yanishlieva N. 2008. Synergistic antioxidant effect of α-tocopherol and myricetin on the autoxidation of triacylglycerols of sunflower oil. Food Chem 106(2):628–33. Medina I, Gallardo JM, Gonzlez MJ, Lois S, Hedges N. 2007. Effect of molecular structure of phenolic families as hydroxycinnamic acids and catechins on their antioxidant effectiveness in minced fish muscle. J Agric Food Chem 55(10):3889–95. Miguel G, Simoes M, Figueiredo AC, Barroso JG, Pedro LG, Carvalho L. 2004. Composition and antioxidant activities of the essential oils of Thymus caespititius, Thymus camphoratus and Thymus mastichina. Food Chem 86(2):183–8. Min DB, Boff JM. 2002. Chemistry and reaction of singlet oxygen in foods. Comp Rev Food Sci Food Saf 1:58–61. Mitsumoto M, O’Grady MN, Kerry JP, Buckley DJ. 2005. Addition of tea catechins and vitamin C on sensory evaluation, colour and lipid stability during chilled storage in cooked or raw beef and chicken patties. Meat Sci 69(4):773–9. Mittal R, Gupta RL. 2000. In vitro antioxidant activity of piperine. Methods Exp Clin Pharm 22(5):271–4. Miura K, Kikuzaki H, Nakatani N. 2002. Antioxidant activity of chemical components from sage (Salvia officinalis L.) and thyme (Thymus vulgaris L.) measured by the oil stability index method. J Agric Food Chem 50(7):1845–51. Moclino I, Martinez C, Sotomayor JA, Lafuente A, Jordan MJ. 2008. Polyphenolic transmission to Segureno lamb meat from ewes’ diets supplemented with the distillate from rosemary (Rosmarinus officinalis) leaves. J Agric Food Chem 56(9):3363–7. Muchuweti M, Kativu E, Mupure CH, Chidewe C, Ndhlala AR, Benhura MAN. 2007. Phenolic composition and antioxidant properties of some spices. Am J Food Technol 2(5):414–20. Munn´e-Bosch S. 2005. The role of α-tocopherol in plant stress tolerance. J Plant Phys 162(7):743–8. Murcia MA, Egea I, Romojaro F, Parras P, Jimenez AM, Martinez-Tome M. 2004. Antioxidant evaluation in dessert spices compared with common food additives. Influence of irradiation procedure. J Agric Food Chem 52(7):1872–81. Murcia MA, Martinez-Tome M. 2001. Antioxidant activity of resveratrol compared with common food additives. J Food Protect 64(3):379– 84. Nakatani N. 2003. Biologically functional constituents of spices and herbs. J Jpn Soc Nut Food Sci 56(6):389–95. Nakatani N, Inatani R. 1981 Structure of rosmanol a new antioxidant from rosemary (Rosmarinus officinalis L). Agric Biol Chem 45:2385–6. Nam KC, Ahn DU. 2003. Use of antioxidants to reduce lipid oxidation and off-odor volatiles of irradiated pork homogenates and patties. Meat Sci 63(1):1–8.

244 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . . Nam KC, Min BR, Yan H, Lee EJ, Mendonca A, Wesley I, Ahn DU. 2003. Effect of dietary vitamin E and irradiation on lipid oxidation, color, and volatiles of fresh and previously frozen turkey breast patties. Meat Sci 65(1):513–21. Nawar WF. 1996. Lipids. In: Fennema O, editor. Food chemistry. 3rd ed. New York: Marcel Dekker, Inc. p 225–320. Naz S, Sherazi STH, Talpur FN. 2011. Changes of total tocopherol and tocopherol species during sunflower oil processing. J Am Oil Chem Soc 88:127–32. Nov´ak I, Z´ambori-N´emeth E, Horv´ath H, Sereg´ely Z, Kaffk K. 2003. Study of essential oil components in different origanum species by GC and sensory analysis. J Acta Aliment 32(2):141–50. Nuutila AM, Pimi¨a RP, Aarni M, Oksman-Caldentey KM. 2003. Comparison of antioxidant activities of onion and garlic extracts by inhibition of lipid peroxidation and radical scavenging activity. Food Chem 81(4):485–93. Obana H, Furuta M, Masakazu, Tanaka Y. 2006. Detection of 2-alkylcyclobutanones in irradiated meat, poultry and egg after cooking. J Health Sci 52(4):375–82. Okada Y, Tanaka K, Fujita I, Sato E, Okajima H. 2005. Antioxidant activity of thiosulfinates derived from garlic. Redox Rep 10(2):96–102. Onibi GE, Scaife JR, Murray I, Fowler VR. 2000. Supplementary α-tocopherol acetate in full-fat rapeseed-based diets for pigs: effect on performance, plasma enzymes and meat drip loss. J Agric Food Chem 80(11):1617–24. Orhan I, Kartal M, Naz Q, Ejaz A, Yilmaz G, Kan Y, Konuklugil B, Sener B, Choudhary MI. 2007. Antioxidant and anticholinesterase evaluation of selected Turkish Salvia species. Food Chem 103(4):1247–54. Ozsoy N, Candoken E, Akev N. 2009. Implications for degenerative disorders: antioxidative activity, total phenols, flavonoids, ascorbic acid, beta-carotene and beta-tocopherol in Aloe vera. Oxid Med Cell Long 2(2):99–106. Padurar I, Paduraru O, Miron A. 2008. Assessment of antioxidant activity of Basilica herba aqueous extract—in vitro studies. Farmacia 160(4):402–8. Papageorgiou V, Gardeli C, Mallouchos A, Papaioannou M, Komaitis M 2008. Variation of the chemical profile and antioxidant behavior of Rosmarinus officinalis L. and Salvia fruticosa Miller grown in Greece. J Agric Food Chem 56(16):7254–64. Pazos M, Lois S, Torres JL, Medina I. 2006. Inhibition of hemoglobin and iron promoted oxidation in fish microsomes by natural phenolics. J Agric Food Chem 54(12):4417–23. Pazos M, Torres JL, Andersen ML, Skibsted LH, Medina I. 2009. Galloylated polyphenols efficiently reduce alpha-tocopherol radicals in a phospholipid model system composed of sodium dodecyl sulfate (SDS) micelles. J Agric Food Chem 57(11):5042–8. Peschel W, Dieckmann W, Sonnenschein M, Plescher A. 2007. High antioxidant potential of pressing residues from evening primrose in comparison to other oilseed cakes and plant antioxidants. Indust Crops Prod 25(1):44–8. Pinelo M, Manzocco L, Nunez MJ, Nicoli MC. 2004. Interaction among phenols in food fortification: negative synergism on antioxidant capacity. J Agric Food Chem 52(5):1177–80. ¨ Pizzale L, Bortolomeazzi R, Vichi S, Uberegger E, Conte LS. 2002. Antioxidant activity of sage (Salvia officinalis and S. fructicosa) and oregano (Origanum onites and O. indercedens) extracts related to their phenolic compound content. J Agric Food Chem 82:1645–51. Poiana MA, Dobrei A, Stoin D, Ghita A. 2008. The influence of viticultural region and the ageing process on the color structure and antioxidant profile of Cabernet Sauvignon red wines. J Food Agric Environ 6(3 4): 104–8. Pokorny J, Yanislieva N, Gordom M. 2001. Antioxidants in food—practical applications. Boca, Raton, Boston, New York, Washington, D.C.: Woodhead Publishing, CRC Press. p 380. Politeo O, Jukic M, Milos M. 2010. Comparison of chemical composition and antioxidant activity of glycosidically bound and free volatiles from clove (Eugenia caryophyllata Thunb.). J Food Biochem 34(1):129–41. Potapovich AI, Kostyuk VA. 2003. Comparative study of antioxidant properties and cytoprotective activity of flavonoids. Biochem (Moscow) 68(5):514–9. Prior RL, Hoang H, Gu L, Wu X, Bacchiocca M, Howard L, Hampsch-Woodill M, Huang D, Ou B, Jacob R. 2003. Assays for hydrophilic and lipophilic antioxidant capacity (oxygen radical absorbance capacity [ORACFL]) of plasma and other biological and food samples. J Agric Food Chem 51:3273–9.  c 2011 Institute of Food Technologists®

Priyadarsini KI, Maity DK, Naik GH, Kumar MS, Unnikrishnan MK, Satav JG, Mohan H. 2003. Role of phenolic O-H and methylene hydrogen on the free radical reactions and antioxidant activity of curcumin. Free Rad Biol Med 35:475–84. Rababah TM, Ereifej KI, Al-Mahasneh MA, Ismaeal K, Hidar AG, Yang W. 2008. Total phenolics, antioxidant activities, and anthocyanins of different grape seed cultivars grown in Jordan. Int J Food Prop 11(2):472–9. Radovanovic A, Radovanovic B, Jovancicevic B. 2009. Free radical scavenging and bacterial activities of southern Serbian red wines. Food Chem 117:326–31. Reblova Z. 2006. The effect of temperature on the antioxidant activity of tocopherols. Eur J Lipid Sci Technol 108(10):858–63. Rehman Z, Salariya AM, Habib F. 2003. Antioxidant activity of ginger extract in sunflower oil. J Agric Food Chem 83(7):624–9. Richheimer SL, Bernart MW, King GA, Kent MC, Bailey DT. 1996. Antioxidant activity of lipid-soluble phenolic diterpenes from rosemary. J Am Oil Chem Soc 73(4):507–14. Richter J, Schellenberg I. 2007. Comparison of different extraction methods for the determination of essential oils and related compounds from aromatic plants and optimization of solid-phase microextraction/gas chromatography. Anal Bioanal Chem 387(6):2207–17. Rivero Perez MD, Muniz P, Gonzalez Sanjose ML. 2008. Contribution of anthocyanin fraction to the antioxidant. Food Chem Toxico 46(8):2815–9. Rochat S, Chaintreau A. 2005. Carbonyl odorants contributing to the in-oven roast beef top note. J Agric Food Chem 53(24):9578–85. Rojas M, Brewer S. 2007. Effect of natural antioxidants on oxidative stability of cooked, refrigerated beef and pork. J Food Sci 72:S282–8. Rojas M, Brewer MS. 2008a. Effect of natural antioxidants on oxidative stability of frozen, vacuum-packaged beef and pork. J Food Qual 3(12):173–88. Rojas MC, Brewer MS. 2008b. Consumer attitudes towards issues in food safety. J Food Saf 28(1):1–22. Ross L, Barclay RC, Vinqvist MR, Mukai K, Goto H, Hashimoto Y, Tokunaga A, Uno H. 2000. On the antioxidant mechanism of curcumin: classical methods are needed to determine antioxidant mechanism and activity. Orig Let 2(18):2841–3. Ruberto G, Baratta MT, Deans SG, Dorman HJD. 2000. Antioxidant and antimicrobial activity of Foeniculum vulgare and Crithmum maritimum essential oils. Planta Med 66:687–93. Sarma AD, Sreelakshmi Y, Sharma R. 1997. Antioxidant ability of anthocyanins against ascorbic acid oxidation. Phytochem 45(4): 671–4. Sasse A, Colindres P, Brewer MS. 2009. Effect of natural and synthetic antioxidants on oxidative stability of cooked, frozen pork patties. J Food Sci 74(1):S30–5. Schmidt E, Bail S, Buchbauer G, Stoilova I, Krastanov A, Stoyanova A, Jirovetz L. 2008. Chemical composition, olfactory evaluation and antioxidant effects of the essential oil of Origanum majorana L. from Albania. Nat Prod Comm 3(7):1051–6. Schmidt E, Jirovetz L, Buchbauer G, Eller GA, Stoilova I, Krastanov A, Stoyanova A, Geissler M. 2006. Composition and antioxidant activities of the essential oil of cinnamon (Cinnamomum zeylanicum Blume) leaves from Sri Lanka. J Essen Oil-Bearing Plants 9(2):170–82. Schwarz K, Bertelsen G, Nissen LR, Gardner PT, Heinonen MI, Huynh-Ba AH, Lambelet P, McPhail D, Skibsted LH, Tijburg L. 2001. Investigation of plant extracts for the protection of processed foods against lipid oxidation. Comparison of antioxidant assays based on radical scavenging, lipid oxidation and analysis of the principal antioxidant compounds. Eur Food Res Technol 212:319–28. Shahidi F, Wanasundara PK. 1992. Phenolic antioxidants. Crit Rev Food Sci Nut 32(1):67–103. Shan B, Cai Y-Z, Brooks JD, Corke H. 2007. The in vitro antibacterial activity of dietary spice and medicinal herb extracts. Int J Food Microbiol 117(1):112–9. Shan B, Cai YZ, Sun M, Corke H. 2005. Antioxidant capacity of 26 spice extracts and characterization of their phenolic constituents. J Agric Food Chem 53(2):7749–59. Si-Ahmed K, Tazerouti F, Badjah-Hadj-Ahmed AY, Aturki Z, D’Orazio G, Rocco A, Fanali S. 2010. Analysis of hesperetin enantiomers in human urine after ingestion of blood orange juice by using nano-liquid chromatography. J Pharm Biomed Anal 51(1):225–9. Simitzis PE, Deligeorgis SG, Bizelis JA, Dardamani A, Theodosiou I, Fegeros K. 2008. Effect of dietary oregano oil supplementation on lamb meat characteristics. Meat Sci 79(2):217–23.

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 245

Natural antioxidants . . . Singh G, Maurya S, de Lampasona MP, Catalan CAN. 2007. A comparison of chemical, antioxidant and antimicrobial studies of cinnamon leaf and bark volatile oils, oleoresins and their constituents. Food Chem Tox 45(9):1650–61. Sloan AE. 1999. Top ten trends to watch and work on for the millennium. Food Technol 53(8):40–8, 51–8. Smet K, Raes K, Huyghebaert G, Haak L, Arnouts S, Smet S de. 2008. Lipid and protein oxidation of broiler meat as influenced by dietary natural antioxidant supplementation. Poultry Sci 87(8):1682–8. Soares DG, Andreazza AC, Salvador M. 2003. Sequestering ability of butylated hydroxytoluene, propyl gallate, resveratrol, and vitamins C and E against ABTS, DPPH, and hydroxyl free radicals in chemical and biological systems. J Agric Food Chem 51(4):1077–80. Srinivasan O, Parkin KL, Fennema O, editors. 2008. Fennema’s food chemistry. 4th ed. Boca Raton, Fla.: CRC Press. p 1144. Srinivasan K. 2007. Black pepper and its pungent principle-piperine: a review of diverse physiological effects. Crit Rev Food Sci Nut 47(8): 735–48. Sroka Z, Cisowski W. 2003. Hydrogen peroxide scavenging, antioxidant and anti-radical activity of some phenolic acids. Food Chem Tox 41(6): 753–8. Stashenko EE, Puertas MA, Martinez JR. 2002. SPME determination of volatile aldehydes for evaluation of in-vitro antioxidant activity. Anal Bioanal Chem 373(1/2):70–4. Suhaj M. 2006. Spice antioxidants isolation and their antiradical activity: a review. J Food Comp Anal 19(6–7):531–7. Swigert KS, McKeith FK, Carr TR, Brewer MS, Culbertson M. 2004. Effects of dietary vitamin D3 , vitamin E, and magnesium supplementation on pork quality. Meat Sci 67(1):81–6. Takemoto JK, Remsberg CM, Yanez JA JA, Vega-Villa KR, Davies NM. 2008. Stereospecific analysis of sakuranetin by high-performance liquid chromatography: pharmacokinetic and botanical applications. J Chrom Analyt Technol 75(1):136–41. Teissedre P, Waterhouse A. 2000. Inhibition of oxidation of human low-density lipoproteins by phenolic substances in different essential oils varieties. J Agric Food Chem 48:3801–5. Tepe B, Sokmen M, Akpulat HA, Sokmen A. 2006. Screening of the antioxidant potentials of six Salvia species from Turkey. Food Chem 95(2):200–4. Thomas RH, Bernards MA, Drake EE, Guglielmo CG. 2010. Changes in the antioxidant activities of seven herb- and spice-based marinating sauces after cooking. J Food Comp Anal 23(3):244–52. Thorsen MA, Hildebrandt KS. 2003. Quantitative determination of phenolic diterpenes in rosemary extracts: aspects of accurate identification. J Chrom 9(5):119–25. Tiwari V, Shankar R, Srivastrava J, Vankar PM. 2006. Change in antioxidant activity of spices—turmeric and ginger on heat treatment. J Environ Food Chem 5(2):1313–7. Tomaino A, Cimino F, Zimbalatti V, Venuti V, Sulfaro V, de Pasquale A, Saija A. 2005. Influence of heating on antioxidant activity and the chemical composition of some spice essential oils. Food Chem 89(4):549–54. Touns MS, Ouerghemmi I, Wannes WA, Ksouri R, Zemn H, Marzouk B, Kchou ME. 2009. Valorization of three varieties of grape. Industrial Crops Prod 30(2):292–6 Trojakova L, Reblova Z, Nguyen HTT, Pokorny J. 2001. Antioxidant activity of rosemary and sage extracts in rapeseed oil. J Food Lipids 8:1–13. Tyler V. 1993. The honest herbal. New York: The Haworth Press, Inc. p 139–41. Uchiyama N, Kim IH, Kikura-Hanajiri R, Kawahara N, Konishi T, Goda Y. 2008. HPLC separation of naringin, neohesperidin and their C-2 epimers in commercial samples and herbal medicines. J Pharm Biomed Anal 46(5):864–9. United States Consumer Product Safety Commission. 1992. Final report: study of aversive agents. Available from: www.cpsc.gov/library/ foia/foia99/os/aversive.pdf. Accessed Feb 2011. Uri N. 1961. Mechanism of antioxidation. In: Lundber WO, editor. Autoxidation and antioxidants. New York: Interscience. p 133–9. Ursini F, Rapuzzi I, Toniolo R, Tubaro F, Bontempelli G. 2001. Characterization of antioxidant effect of procyanidins. Meth Enzymol 335:338–50. USDA. 2010. USDA Database for the oxygen radical absorbance capacity (ORAC) of selected foods. Release 2. U.S. Department of Agriculture. ARS. Beltsville, MD. Available from: www.ars.usda.gov/

SP2UserFiles/Place/12354500/Data/.../ORAC_R2.pdf. Accessed May 2010. Vega JD, Brewer MS. 1995. Detectable odor thresholds of selected lipid oxidation compounds in a meat model system. J Food Sci 60(3): 592–5. Vega-Villa KR, Yanez JA, Remsberg CM, Ohgami Y, Davies NM. 2008. Stereospecific high-performance liquid chromatographic validation of homoeriodictyol in serum Yerba Santa (Eriodictyon glutinosum). J Pharm Biomed Anal 46(5):971–4. Vera RR, Chane-Ming J. 1999. Chemical composition of the essential oil of marjoram (Origanum majorana L.) from Reunion Island. Food Chem 66(2):143–5. Viuda-Martos M, Ruiz-Navajas Y, Fernandez-Lopez J, Perez-Alvarez JA. 2010. Effect of adding citrus fibre washing water and rosemary essential oil on the quality characteristics of a bologna sausage. LWT Food Sci Tech 43(6):958–63. Wanatabe Y, Nakanashi H, Goto N, Otsuka K, Kimura T, Adachi S. 2010. Antioxidative properties of ascorbic acid and acyl ascorbates in ML/W emulsion. J Am Oil Chem Soc 85:1475–80. Wang W, Wu N, Zu YG, Fu YJ. 2008. Antioxidative activity of Rosmarinus officinalis L. essential oil compared to its main components. Food Chem 108(3):1019–22. Wen LC, Wen YC, SungChuan W, KuangHway Y. 2009. DPPH free-radical scavenging activity, total phenolic contents and chemical composition analysis of forty-two kinds of essential oils. J Food Drug Anal 17(5): 386–95. Williams P, Markoska J, Chachay V, McMahon A. 2009. ‘Natural’ claims on foods: a review of regulations and a pilot study of the views of Australian consumers. Food Aust 61(9):383–9. Wills TM, Mireles DeWitt CA, Sigfusson H. 2007. Improved antioxidant activity of vitamin E through solubilization in ethanol: a model study with ground beef. Meat Sci 76(2):308–15. Wojdyło A, Oszmia´nski J, Czemerys R. 2007. Antioxidant activity and phenolic compounds in 32 selected herbs. Food Chem 105(3): 940–9. Xiao JB, Suzuki M, Jiang XY, Chen XQ, Yamamoto K, Ren FL, Xu M. 2008. Influence of B-ring hydroxylation on interactions of flavonols with bovine serum albumin. J Agric Food Chem 56(7):2350–6. Xu CM, Zhang YL, Wang J, Lu JA. 2010. Extraction, distribution and characterisation of phenolic compounds and oil in grapeseeds. Food Chem 122(3):688–94. Yancey EJ, Grobbel JP, Dikeman ME, Smith JS, Hachmeister KA, Chambers EC, Gadgil P, Milliken GA, Dressle E. 2006. Effects of total iron, myoglobin, hemoglobin, and lipid oxidation of uncooked muscles on livery flavor development and volatiles of cooked beef steaks. Meat Sci 73: 680–6. Yanagimoto K, Ochi H, Lee K, Shibamoto T. 2003. Antioxidative activities of volatile extracts from green tea, oolong tea, and black tea. J Agric Food Chem 51(25):7396–401. Yanez JA, Andrews PK, Davies NM. 2007. Methods of analysis and separation of chiral flavonoids. J Chrom Anal Tech Biomed Life Sci 848(2): 159–81. Yanez JA, Remsberg CM, Miranda ND, Vega-Villa KR, Andrews PK, Davies NM. 2008. Pharmacokinetics of selected chiral flavonoids: hesperetin, naringenin and eriodictyol in rats and their content in fruit juices. Biopharm Drug Dispos 29(2):63–82. Yanez JA, Teng XW, Roupe KA, Davies NM. 2005. Stereospecific high-performance liquid chromatographic analysis of hesperetin in biological matrices. J Pharm Biomed Anal 37(3):591–5. Yanishieva NV, Marinova E, Pokorny J. 2006. Natural antioxidants from herbs and spices. Eur J Lipid Sci Technol 108:776–93. Yetella RR, Min DB. 2008. Quenching mechanisms and kinetics of trolox and ascorbic acid on the riboflavin-photosensitized oxidation of tryptophan and tyrosine. J Agric Food Chem 56(22):10887–92. Yeum KJ, Beretta G, Krinsky NI, Russell RM, Aldini G. 2009. Synergistic interactions of antioxidant nutrients in a biological model system. Nut 25(7/8):839–46. Yong J, Hun K, Sung YP, W, JC, Seong SY, Young JY. 2000. Identification of irradiation-induced volatile flavor compounds in chicken. J Korean Soc Food Sci Nut 29(6):1050–6. Yoo KM, Lee CH, Lee H, Moon BK, Lee CY. 2008. Relative antioxidant and cytoprotective activities of common herbs. Food Chem 106(3): 929–36.

246 Comprehensive Reviews in Food Science and Food Safety r Vol. 10, 2011

 c 2011 Institute of Food Technologists®

Natural antioxidants . . . Youdim KA, Deans SG, Finlayson HJ 2002. The antioxidant properties of thyme (Thymus zygis L.) essential oil: an inhibitor of lipid peroxidation and a free radical scavenger. J Essen Oil Res 14(3):210–5. Yu XC, Wang X, Adelberg J, Barron FH, Chen Y, Chung HY. 2008. Evaluation of antioxidant activity of curcumin-free turmeric (Curcuma longa L.). Chapter 14. In: Oil and identification of its antioxidant constituents. Functional food and health. ACS Ser. 993:152–64.

 c 2011 Institute of Food Technologists®

Zhang J, Jinnai S, Ikeda R, Wada M, Hayashida S, Nakashima K. 2009. A simple HPLC-fluorescence method for quantitation of curcuminoids and its application to turmeric products. Analytical Sciences 25. Available from: http://naosite.lb.nagasakiu.ac.jp/dspace/bitstream/10069/22006/1/ AnalSci25_385.pdf. Accessed Feb 2010.

Vol. 10, 2011 r Comprehensive Reviews in Food Science and Food Safety 247