Next Article in Journal
Unconventional Pathways of Secretion Contribute to Inflammation
Next Article in Special Issue
Application of Chromatographic and Spectroscopic Methods towards the Quality Assessment of Ginger (Zingiber officinale) Rhizomes from Ecological Plantations
Previous Article in Journal
Time-Dependent Nerve Growth Factor Signaling Changes in the Rat Retina During Optic Nerve Crush-Induced Degeneration of Retinal Ganglion Cells
Previous Article in Special Issue
In-Depth Two-Year Study of Phenolic Profile Variability among Olive Oils from Autochthonous and Mediterranean Varieties in Morocco, as Revealed by a LC-MS Chemometric Profiling Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources

1
Guangdong Provincial Key Laboratory of Food, Nutrition and Health, School of Public Health, Sun Yat-Sen University, Guangzhou 510080, China
2
South China Sea Bioresource Exploitation and Utilization Collaborative Innovation Center, Sun Yat-Sen University, Guangzhou 510006, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2017, 18(1), 96; https://doi.org/10.3390/ijms18010096
Submission received: 21 October 2016 / Revised: 24 December 2016 / Accepted: 27 December 2016 / Published: 5 January 2017
(This article belongs to the Special Issue Analytical Techniques in Plant and Food Analysis)

Abstract

:
Natural antioxidants are widely distributed in food and medicinal plants. These natural antioxidants, especially polyphenols and carotenoids, exhibit a wide range of biological effects, including anti-inflammatory, anti-aging, anti-atherosclerosis and anticancer. The effective extraction and proper assessment of antioxidants from food and medicinal plants are crucial to explore the potential antioxidant sources and promote the application in functional foods, pharmaceuticals and food additives. The present paper provides comprehensive information on the green extraction technologies of natural antioxidants, assessment of antioxidant activity at chemical and cellular based levels and their main resources from food and medicinal plants.

Graphical Abstract

1. Introduction

In biological system, reactive oxygen species (ROS) and reactive nitrogen species (RNS), such as superoxide, hydroxyl, and nitric oxide radicals, can damage the DNA and lead to the oxidation of lipid and proteins in cells [1,2,3]. Normally, antioxidant system occurring in human body can scavenge these radicals, which would keep the balance between oxidation and anti-oxidation. Nonetheless, the exposure of cigarette smoking, alcohol, radiation, or environmental toxins induces the production of excessive ROS and RNS, which disrupt the balance between oxidation and anti-oxidation and result in some chronic and degenerative diseases [3,4,5]. The increment of intake of exogenous antioxidants would ameliorate the damage caused by oxidative stress through inhibiting the initiation or propagation of oxidative chain reaction, acting as free radical scavengers, quenchers of singlet oxygen and reducing agents [6].
The exogenous antioxidants are mainly derived from food and medicinal plants, such as fruits, vegetables, cereals, mushrooms, beverages, flowers, spices and traditional medicinal herbs [7,8,9,10,11,12,13,14,15,16,17,18]. Besides, the industries processing agricultural by-products are also potentially important sources of natural antioxidants [19]. These natural antioxidants from plant materials are mainly polyphenols (phenolic acids, flavonoids, anthocyanins, lignans and stilbenes), carotenoids (xanthophylls and carotenes) and vitamins (vitamin E and C) [6,20]. Generally, these natural antioxidants, especially polyphenols and carotenoids, exhibit a wide range of biological effects, such as anti-inflammatory, antibacterial, antiviral, anti-aging, and anticancer [2,20,21,22,23,24,25,26,27,28,29,30,31].
Considering their important health effects, the efficient extraction methods of natural antioxidants, appropriate assessment of antioxidant activity as well as their main resources from food and medicinal plants are drawing great attention in food science and nutrition. To improve the extraction efficiency of antioxidant components from plant materials, several green non-conventional methods have been developed for reducing operational time and usage of organic solvents, such as ultrasound-assisted extraction, microwave-assisted extraction, enzyme-assisted extraction, pressurized liquid extraction, supercritical fluid extraction, high hydrostatic pressure extraction, pulsed electric field extraction and high voltage electrical discharges extraction. Moreover, to further assess the antioxidant capacities of extracts from natural products, especially those frequently consumed by people, different evaluation assays have been developed, e.g., Trolox equivalence antioxidant capacity (TEAC) assay, ferric ion reducing antioxidant power (FRAP) assay, oxygen radical absorbance capacity (ORAC) assay, inhibiting the oxidation of low-density lipoprotein (LDL) assay, cellular antioxidant activity assay, etc. These assays have been used for ranking the antioxidant plants and recommending best antioxidant foods for consumption. This review is aimed at summarizing the extraction methods of natural antioxidants, assessment methods of antioxidant activity and their main resources from food and medicinal plants.

2. Extraction Methods of Antioxidants from Foods and Medicinal Plants

Extraction is the first and crucial step for studying the natural antioxidants from plants (Figure 1). Many extraction factors play important roles in the extraction efficiency, such as type and concentration of extraction solvent, extraction temperature, extraction time, and extraction pH. Among them, the solvent is one of the most influential factors. Numerous solvents have been used for the extraction of antioxidants from food and medicinal plants. The selection of solvents is based on the chemical nature and polarity of antioxidant compounds to be extracted. Most of the phenolics, flavanoids and anthocyanins are hydrosoluble antioxidants. The polar and medium polar solvents, such as water, ethanol, methanol, propanol, acetone and their aqueous mixtures, are widely used for extraction [32,33,34,35]. Carotenoids are lipid-soluble antioxidants, and common organic solvents, such as the mixtures of hexane with acetone, ethanol, methanol, or mixtures of ethyl acetate with acetone, ethanol, methanol, have been used for extraction [36,37,38].
Various extraction procedures, including conventional extraction methods and non-conventional extraction methods, can be chosen to extract antioxidants from food and medicinal plants. The conventional extraction methods are mainly hot water bath, maceration and Soxhlet extraction, which are very time-consuming and require relatively large amounts of organic solvents with low extraction yields [39,40,41]. Furthermore, the long heating process such as hot water bath and Soxhlet extraction may lead to the degradation of the thermolabile compounds. To obtain antioxidants from plants in an energy-efficient and economically sustainable way, ultrasound, microwave, pressurized liquid, enzyme hydrolysis, supercritical fluids, high hydrostatic pressure, pulsed electric field, and high voltage electrical discharges have been studied as non-conventional methods (Table 1). In this section, the current application and developments of the non-conventional methods are summarized.

2.1. Ultrasound-Assisted Extraction (UAE)

Ultrasound assisted extraction (UAE) has been applied widely in the last three decades as an efficient extraction method in the food and pharmaceutical industries [61]. The mechanism is based on the cavitation phenomenon. The spread of ultrasound in liquid systems is via a series of compression and rarefaction waves, which can induce the production of cavitation bubbles within the fluid [62,63]. The size of these bubbles grow over the period of a few cycles until reach a critical point, then these bubbles collapse and release a great quantity of energy, which would generate extreme temperatures (5000 K) and pressures (1000 atmospheres) at room temperature. During the ultrasound assisted extraction of bioactive components from plant materials, the high temperature and pressure would destroy the cell walls, facilitate the release of bioactive compounds from plant cell walls and enhance the mass transport. The heat transfer of UAE is from outside of the plant cell to the inside, which is in the opposite direction of microwave assisted extraction.
Ultrasound frequency, intensity, temperature, and time can directly affect both extraction efficiency and yields. In addition, types of solvent, solvent volume, as well as sample characteristics such as moisture content of the sample and particle size are also the important factors for effective extraction [64]. Ultrasound-assisted extraction of anthocyanins and phenolic compounds in mulberry (Morus nigra) pulp was performed by Espada-Bellido et al. [65]. Several extraction variables, including methanol concentration (50%–100%), temperature (10–70 °C), ultrasound amplitude (30%–70%), cycle (0.2–0.7 s), solvent pH (3–7) and solvent–solid ratio (10:1.5–20:1.5) were optimized to obtain the high extraction yields. It was found that extraction temperature and solvent concentration were the most influential parameters for extraction of anthocyanins and phenolic compounds. Besides, different UAE conditions were suitable for different bioactive components. The optimum UAE conditions for anthocyanins are 76% methanol concentration, 12:1.5 solvent–solid ratio, 48 °C extraction temperature, cycle of 0.7 s and 70% ultrasound amplitude at pH 3, while the optimum UAE conditions for phenolic compounds were 61% methanol concentration, 11:1.5 solvent–solid ratio, 64 °C extraction temperature, cycle of 0.7 s and 70% ultrasound amplitude at pH 7. Under the UAE optimal conditions, the maximized extraction yields of total anthocyanins and total phenolic compounds were 149.95 μg/g and 1214.03 μg/g, respectively. Compared with conventional extraction, UAE showed several merits in terms of extraction yield and extraction time. Polyphenols from apple pomace were extracted using conventional extraction and UAE methods [66]. The results showed that ultrasound-assisted extraction increased the catechin equivalents by more than 20% and the release of flavan-3-ols and procyanidins were also enhanced by more than 25% in only 45 min. In another study, Liu et al. [67] used UAE and traditional boiling-water extraction methods to extract antioxidants from ** antioxidant parameter (TRAP), and inhibiting the oxidation of low-density lipoprotein (LDL).

3.1.1. Scavenging Free Radicals Assays

The Trolox equivalent antioxidant capacity (TEAC) assay is widely applied to evaluate the antioxidant ability to scavenge the ABTS radical [148,149,151,152]. According to the type of oxidation agent, there are two versions of this assay [153]. TEAC1: metmyoglobin-H2O2 oxidize ABTS to generate its colored ABTS•+ form; then, subsequent addition of antioxidants results in loss of the green color. TEAC2: potassium persulfate oxidize ABTS to generate its colored ABTS•+ form; then, subsequent addition of antioxidants results in loss of the green color. The version of TEAC1 is inaccurate because antioxidants also can react with the original HO• radical and the metmyoglobin except for the ABTS•+, which could cause an overestimation of antioxidant capacity. Therefore, the version of TEAC2 is more preferable. ABTS•+ has a UV-vis absorption maximum at 734 nm. The decrease of absorbance can be monitored spectrophotometrically. The difference of the absorbance tested is plotted versus the antioxidants concentrations. The antioxidant capacity was expressed as Trolox equivalents. Because ABTS•+ could react rapidly with antioxidants, the assay possesses the merits of rapidity and simplicity. Additionally, ABTS•+ is not influenced by ionic strength and is solvable in both organic and aqueous solvents, so it can be applied in multiple media to detect both hydrophilic and lipophilic antioxidant activities [154]. However, for slow reactions, the TEAC values tested is inaccurate when the duration of reaction is beyond 6 min [149].
As one of the few stable organic nitrogen radicals, the 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical is used to analyze the antioxidant activity [151,155]. The DPPH• posses a deep purple color and has a UV-vis absorption maximum at 515 nm [148,149]. The test compounds (antioxidants) reduce DPPH radical to DPPH-H and the solution color fades. The reducing ability can be assessed by measuring the decrease of its absorbance. In the end, the results are shown by EC50 and TEC50, that is, the necessary amount of antioxidant to decrease the initial DPPH concentration by 50% and the time taken to reach the steady state to EC50 concentration [155]. DPPH assay is widely used in antioxidant capacity screening of fruit and vegetable juices or extracts, for it is easy, rapid and requires only a UV-vis spectrophotometer to test. Compared with ABTS assay, the DPPH radical is commercially available and does not have to be generated before assay such as ABTS•+. However, the application of DPPH assay is limited by its disadvantage. The linear reaction range of DPPH assay is narrow, only 2–3-fold. Moreover, for steric inaccessibility, antioxidants that possess strong antioxidant activities in lipid peroxidation system may react slowly or may even be inert to DPPH [148,149].

3.1.2. Reducing the Metal Ions Assays (FRAP and CUPRAC Assays)

The ferric-reducing antioxidant power (FRAP) assay measures directly the reducing capacity of antioxidants. In ferric-reducing antioxidant reactive system, the antioxidants can reduce a ferric tripyridyltriazine complex (Fe3+-TPTZ) to the ferrous complex (Fe2+-TPTZ) under pH 3.6 condition with a blank sample in parallel. The ferrous complex (Fe2+-TPTZ) is blue ferrous form and has a UV-vis absorption maximum at 593 nm. The ability of antioxidants in samples (FRAP value) is positive related to the increase in absorbance [151,156,157,158]. In general, this assay is suitable for some antioxidants that complete the reaction rapidly within 4 to 6 min, such as ascorbic acid and uric acid [159]. However, it has been demonstrated that the absorption of several dietary polyphenols in water and methanol slowly increased even after several hours, such as tannic acid, and caffeic acid [157]. In addition, FRAP cannot detect compounds that act by radical quenching (H transfer), particularly thiols and proteins. This results in a serious underestimation in serum sample [149].
The cupric reducing antioxidant capacity (CUPRAC) assay is similar to the FRAP method. CUPRAC method is conducted by mixing Cu(II)-neocuproine (Nc) chelate with antioxidant solution. The absorbance of the color Cu(I)-chelate as a result of redox reaction is measured at 450 nm after 30 min [160]. The application on this assay is less extensive than FRAP. However, this assay exhibits several merits in some ways. For example, the reagent in this assay is useful at pH 7, which is at physiological pH (as opposed to the Folin and FRAP assays, which work at pH 10 and pH 3.6, respectively). This method could be applied for the determination of both hydrophilic and lipophilic antioxidants because the Cu(II)-Nc is soluble in both aqueous and organic environments (unlike Folin and DPPH) [160,161]. In addition, CUPRAC assay can measure the reducing power of thiol-type antioxidants, such as glutathione and nonprotein thiols [150,162].

3.1.3. Folin–Ciocalteu Reagent (FCR) Assay

Folin–Ciocalteu reagent (FCR) assay is a widespread method for quantitative determination of phenolic compounds. The mechanism of Folin–Ciocalteu method is electron transfer (ET) [134,135]. It involves reducibility of phenols in alkaline solution (pH = 10), which is capable of turning yellow molybdotungsto-phosphoric heteropolyanion reagent into the blue resultant molybdotungsto-phosphate [148,163,164]. These blue pigments have a maximum absorption in the 700–760 nm range. The maximum absorption depends on the qualitative and/or quantitative composition of phenolic mixtures. The total phenols assay by FCR is simple, convenient, and has produced a large body of comparable data. Thus, it has become a routine assay in studying phenolic antioxidants from fruits, vegetables and medicine plants [163,165,166,167,168]. A large number of publications found excellent linear correlations between the “total phenolic profiles” by FCR and “the antioxidant activity” by other ET-based antioxidant capacity assay (e.g., FRAP, TEAC, etc.) [148,169,170].

3.1.4. Oxygen Radical Absorbance Capacity (ORAC) Assay

Generally, these assays estimate the capacity of antioxidants to protect a target molecule exposed to a free radical source. The oxygen radical absorbance capacity (ORAC) assay has been applied widely in the field of antioxidant and oxidative stress via H atom transfer [148,149,171,172]. In the basic assay, the peroxyl radical mixes with a fluorescent probe (FL; 3′,6′-dihydroxyspiro[isobenzofuran-1[3H],9′[9H]-xanthen]-3-one), then form a nonfluorescent product, which can be quantitated easily by fluorescence [149,171,173]. When an antioxidant is added into the mixture, peroxyl radical induced oxidation is inhibited and the decay of FL is prevented. Antioxidant capacity is reflected by determining the decreased rate and amount of product formed over time. Using AUC (area under curve) to reflect the antioxidant capacity is favorable because it applies to an antioxidant that has a lag phase or one that do not [171]. It is useful for a broad range of sample types, including raw fruit and vegetable extracts, plasma, and pure phytochemicals. Furthermore, the high-throughput assay is able to test several hundred samples daily by just using one plate-reader coupled with a multichannel automatic liquid handling system [148].

3.1.5. Total Radical Trap** Antioxidant Potential (TRAP) Assay

The total radical trap** antioxidant potential (TRAP) assay measures the ability of antioxidants to suppress the oxidation progress of 2,2′-azobis-2-methyl-propanimidamide, dihydrochloride (AAPH) or 2,2′-azobis(2-amidinopropane) dihydrochloride (ABAP) [148,174]. The variation in the reaction progress is monitored fluorometrically (λex = 495 nm and λem = 575 nm). The fluorescence decay rate in the reaction slows after the addition of antioxidants compared with the rate before the antioxidants addition. The quantification is based on the lag-phase duration compared with the lag phase of Trolox [149,162]. The application of the lag phase is based on the assumption that the antioxidants show a lag phase and the length of the lag phase is positively correlated to antioxidant capacity. However, not every antioxidant component possesses an obvious lag phase and the potential of antioxidants that play a role after the lag phase is totally ignored.

3.1.6. Inhibiting the Oxidation of Low-Density Lipoprotein (LDL) Assay

Inhibition of induced lipid autoxidation has been developed as a measure of antioxidant capacity in a more physiologically relevant system [148,149,150]. Usually, the reaction solution contains free radical initiator (Cu(II) or 2,2′-azobis(2-amidinopropane) dihydrochloride (AAPH)), substrate (linoleic acid or LDL), and antioxidants. The autoxidation of linoleic acid or LDL is induced by Cu(II) or AAPH. The peroxidation of the lipid components is monitored at 234 nm by UV spectrometer for conjugated dienes. In the presence of a radical initiator, the reaction starts and the absorbance at 234 nm increases as the evidence of the accumulation of conjugated diene oxides. After the addition of antioxidants, the reaction rate slows down until the antioxidant is exhausted. In the period, the lag time is measured and used to evaluate antioxidant capacity.
Compared with other in vitro assays, the major advantage of this method is the use of a biological relevant substrate, which makes the results relevant to oxidative reactions in vivo [149]. Because LDL is isolated from blood samples, one of the major flaws of this method is the variability of the LDL samples, which can vary with different donors. Thus, this method is hard to be developed as a consistent, reproducible, high throughput antioxidant evaluation assay [148,149,150]. On the contrary, using linoleic acid or its methyl ester as an oxidation substrate would make the results more reproducible than using LDL. However, linoleic acid would form micelles in the presence of water, and the reaction progress in micelles cannot be monitored directly by UV absorbance, thus the accuracy of the method can be affected.

3.2. Cellular-Based Assays

The antioxidant capacity evaluated by chemical assays cannot completely reflect the behavior of the sample in vivo. It is necessary to estimate the effectiveness of antioxidants in more biologically relevant conditions. Animal models and human studies are more suitable for evaluation but more expensive and time-consuming [150]. As intermediate testing methods, cellular antioxidant activity (CAA) assay has been developed for evaluating the antioxidant capacities [175,176]. Dichlorofluorecin (DCFH) method is a commonly used CAA assay, which tests the capacity of antioxidants to prevent the oxidation of DCFH. In general, DCFH trapped within cells is easily oxidized to fluorescent dichlorofluorescein (DCF) by ABAP-generated peroxyl radicals in human hepatocarcinoma HepG2 cells. DCF could be monitored by fluorescence (λexc = 485 nm, λem = 538 nm). The decrease in cellular fluorescence is proportional to the antioxidant capacity of bioactive components. Except for HepG2 cells, several cell types have been applied for the CAA assay, such as human red blood cell, human endothelial EA.hy926, human colon cancer Caco-2 cell, human macrophage U937 cell and mouse macrophage RAW264.7 cell [150,176,177]. Besides, a CAA assay based on microfluidic cell chip with arrayed microchannels has been developed to assess plant antioxidants [178]. The microfluidic chip contains 288 round cell culture micro chambers and 48 independent parallel array channels. In this method, eight groups of different samples with six different concentrations could be tested simultaneously with multimode reader.
The assessment of antioxidant activity at cellular level is not limited to the test of ability of ROS/RNS scavenging but also includes tests of expression of antioxidant enzymes, inhibition of pro-oxidant enzymes, and activation vs. repression of redox transcription factors [150]. The antioxidant activities of the extracts prepared from five brown seaweeds was assessed in Caco-2 cells. Glutathione (GSH) content and antioxidant enzyme activity (catalase (CAT) and superoxide dismutase (SOD)) were evaluated. These cellular assays indicated that Pelvetia canaliculata could exert the antioxidant capacity mainly by preventing H2O2-mediated SOD depletion in Caco-2 cells [179]. Besides, antioxidant enzyme activities of glutathione peroxidase (GPx) and glutathione reductase (GR) were measured in three Argentinean red wines. Some protective effects of wine were observed in cells exposed to H2O2, which was attributed to the increased activity of antioxidant enzymes GPx and GR [180]. In addition, suppression of NF-κB activation as an anti-oxidant response has been observed in cultured cells with the treatment of phenols (e.g., curcumin) or food extracts (e.g., blueberries) [150,181,182]. In a study, it was observed that the activation of NF-κB and activator protein-1, as well as IL-8 release were suppressed in curcumin-treated alveolar epithelial cells. Additionally, in comparison with untreated cells, the levels of GSH and glutamylcysteine ligase catalytic subunit mRNA expression were increased [181].

4. Main Resources of Natural Antioxidants

Among various chemical based assays, the TEAC assay evaluates the ability to scavenge the free radical, the FRAP assay measures directly the reducing capacity of antioxidants, the total phenols assay by FCR assesses the phenolic contents from tested samples. To comprehensively study different aspects of antioxidants, the combination of TEAC, FRAP and FCR methods is frequently used to evaluate the antioxidant activity. As shown in Table 3, the antioxidant activities of many food and medicinal plants have been widely estimated, e.g., fruits, vegetables, cereal grains, edible and wild flowers, macro-fungi, medicinal plants, spices, etc., and the varieties showing strong antioxidant activities were displayed in Table 3 based on a combinative consideration of the results obtained by FRAP, TEAC and FCR assays. Overall, these results showed that different categories exhibited diverse antioxidant capacities and the variation was very large. The FRAP, TEAC and FCR values of 62 fruits varied from 0.11 ± 0.01 to 72.11 ± 2.19 μmol Fe(II)/g, 0.84 ± 0.03 to 80.68 ± 2.11 μmol Trolox/g, and 11.88 ± 0.11 to 585.52 ± 18.59 mg GAE/100 g, respectively [11]. The FRAP, TEAC and FCR values of 56 vegetables varied from 2.69 to 60.9 μmol Fe(II)/g, 6.93 to 33.63 μmol Trolox/g, and 4.99 to 23.27 mg GAE/g, respectively [14]. The FRAP, TEAC and FCR values of 24 cereal grains varied from 5.23 ± 0.23 to 126.19 ± 2.91 μmol Fe(II)/g, 0.62 ± 0.14 to 30.03 ± 1.10 μmol trolox/g, and 1.35 ± 0.15 to 9.47 ± 0.48 mg GAE/g, respectively [12]. The FRAP, TEAC and FCR values of 223 medicinal plants varied from 0.14 to 1844.85 μmol Fe(II)/g, 0.99 to 1544.38 μmol Trolox/g, and 0.19 to 101.33 mg GAE/g, respectively [15]. Obviously, among these varieties showing strong antioxidant activities, the antioxidant activities and total phenolic contents of medicinal plants were significantly higher than those of fruits, vegetables and cereals.
Additionally, the antioxidant activities of food and medicinal plants have also been evaluated by cellular antioxidant activity assays based on different cell types, and the varieties showing strong antioxidant activities were displayed in Table 3. The cellular antioxidant activities of 27 vegetables ranged from not detected (tomato) to 41.9 ± 6.2 μmol of QE/100 g (beet) [183]. The cellular antioxidant activities of 25 fruits ranged from 3.15 ± 0.21 (banana) to 292 ± 11 μmol of QE/100 g (wild blueberry) [184]. In both studies, these results showed that CAA values were significantly associated with total phenolic content. Surarit et al. [185] reported that the ethanolic bran extracts of 11 Thai pigmented (red and purple) and two non-pigmented rice varieties exerted the cellular antioxidant activities based on HL-60 cells in the following order: red > purple > non-pigmented rice, which showed the same order of phenolic and flavonoid contents in these rice extracts.

4.1. Natural Sources of Polyphenols

Polyphenols are abundant in food and medicinal plants, including phenolic acids, flavonoids, lignans, and stilbenes [20].
Phenolic acids comprise of derivatives of cinnamic acid (e.g., p-coumaric, caffeic, and ferulic) and derivatives of benzoic acid (e.g., gallic acid and hydroxybenzoic acids). Compared with the hydroxybenzoic acids, the hydroxycinnamic acids are more abundant in edible plants [20]. Fruits such as blueberries, kiwis, plums, cherries, apples are found to be rich in the hydroxycinnamic acids (0.5–2 g hydroxycinnamic acids/kg fresh wt). Caffeic acid is the most abundant phenolic acid and account for 75%–100% of the total hydroxycinnamic acid content in many fruits, whereas, ferulic acid is the most abundant phenolic acid in cereal grains and account for about 90% of total polyphenol content of wheat grain. The hydroxybenzoic acid content in edible plants is usually very low, except for certain red fruits, black radish, and onions. Due to the low content, they are not considered to be of great nutritional interest.
Flavonoids are abundant in most edible fruits and vegetables. Its subclasses include flavonols, flavanones, catechins, flavones, anthocyanidins and isoflavonoids. The concentration and type of flavonoids vary in different dietary sources [186]. Quercetin is usually the most abundant flavonol in edible plants. The richest dietary source of quercetin is onion. Tea and wine have relatively low amounts of quercetin. Other flavonols include myricetin (berries), isorhamnetin (onion), and kaempferol (broccoli). Flavanones almost only exist in citrus fruits. Hesperidin and narirutin are the main flavonoids of oranges and mandarins, whereas, naringin and narirutin are the main flavonoids of grapefruit. Catechins usually exist in the form of aglycones or esterified with gallic acid. The richest dietary sources of catechins are tea and red wine. In addition, the major flavones are apigenin and luteolin. The major dietary sources are red pepper and celery. Anthocyanins, such as pelargonidin, cyanidin, and delphinidin, contribute to the red, blue, or violet color of edible plants, such as plums, eggplant, and many berries. The isoflavonoids, such as isoflavones genistein and daidzein, mainly exist in legumes. The richest dietary source are soybean and soy products [186,187].
The richest dietary source of lignans is linseed containing secoisolariciresinol (up to 3.7 g/kg dry wt) and low quantities of matairesinol [20]. Other algae, leguminous plants (lentils), cereals (triticale and wheat), fruit (pears, prunes) and certain vegetables (garlic, asparagus, carrots) also have traces of these same lignans. Resveratrol is a stilbene, which has been extensively studied for its multiple bioactivities. Red wine is rich in resveratrol (0.3–7 mg aglycones/L and 15 mg glycosides/L).

4.2. Natural Sources of Carotenoids

Carotenoids are natural pigments, including β-carotene, lycopene, lutein, and zeaxanthin [188]. All colorful edible plants, especially dark green and yellow-orange leafy, are the good sources of carotenoids. Due to the lipid solubility of carotenoids, the absorption mainly depends on their preparation with oils or fats. Among the carotenoids, the β-carotene commonly occurs in edible plants that possesses the highest activity of provitamin A, such as acerola, mango, pumpkin, carrot, nuts, and oil palm. Lycopene is a kind of red pigment. It almost exists only in vegetables and algae tissues. Tomato products such as juices, soups, sauces, and ketchup, as well as their processing waste and peel are important sources of lycopene [189]. The lycopene (79%–91%) presented in tomatoes is mostly in the form of the trans isomer. Lutein and zeaxanthin are the most common xanthophylls in green and dark leafy vegetables such as broccoli, spinach, peas and lettuce [190]. Zeaxanthin also found in the red marine microalga P. cruentum (making up 97.4% of the total carotenoids) (Mezzomo et al., 2016; Raposo et al., 2015) [188,191].

5. Conclusions

Antioxidants derived from food and medicinal plants have been increasingly investigated for their various nutritional function and health benefits. In this review, the extraction methods of natural antioxidants, assessment methods of antioxidant activity as well as their main resources from food and medicinal plants are summarized. The non-conventional techniques described have potential to replace or enhance existing extraction techniques due to the less extraction time, energy consumption, and usage of harmful organic solvents, as well as higher extraction yields to recover antioxidant compounds from food and medicinal plants. Nevertheless, most of them are limited for industrial applications due to the high equipment costs and complicated installation procedures. Thus, establishing the balance between energetic and cost will be a key research in the future. To take advantage of the different extraction methods and to limit their drawbacks, the combinative application of multiple extraction technologies and the automated potential of these non-conventional extraction technologies would be the development tendency in future. For assessing the antioxidant activity of plant materials, several assays are suggested, such as the determination of total polyphenolic content by FCR, scavenging free radical ability by TEAC, metal-reducing activity by FRAP, and a kind of cellular-based assay. Furthermore, in order to make the comparison of different samples and studies possible, standardizing the operating conditions of the same analysis method and the expression of results are recommended.

Acknowledgments

This work was supported by the National Natural Science Foundation of China (No. 81372976), Key Project of Guangdong Provincial Science and Technology Program (No. 2014B020205002), and the Hundred-Talents Scheme of Sun Yat-Sen University.

Author Contributions

Dong-** Xu and Hua-Bin Li conceived this paper; Dong-** Xu and Ya Li wrote this paper; and **ao Meng, Tong Zhou, Yue Zhou, Jie Zheng, Jiao-Jiao Zhang and Hua-Bin Li revised the paper.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fang, Y.Z.; Yang, S.; Wu, G. Free radicals, antioxidants, and nutrition. Nutrition 2002, 18, 872–879. [Google Scholar] [CrossRef]
  2. Peng, C.; Wang, X.; Chen, J.; Jiao, R.; Wang, L.; Li, Y.M.; Zuo, Y.; Liu, Y.; Lei, L.; Ma, K.Y.; et al. Biology of ageing and role of dietary antioxidants. BioMed Res. Int. 2014, 2014, 831841. [Google Scholar] [CrossRef] [PubMed]
  3. Li, S.; Tan, H.Y.; Wang, N.; Zhang, Z.J.; Lao, L.; Wong, C.W.; Feng, Y. The role of oxidative stress and antioxidants in liver diseases. Int. J. Mol. Sci. 2015, 16, 26087–26124. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Wang, F.; Li, Y.; Zhang, Y.J.; Zhou, Y.; Li, S.; Li, H.B. Natural products for the prevention and treatment of hangover and alcohol use disorder. Molecules 2016, 21, 64. [Google Scholar] [CrossRef] [PubMed]
  5. Zhou, Y.; Zheng, J.; Li, S.; Zhou, T.; Zhang, P.; Li, H.B. Alcoholic beverage consumption and chronic diseases. Int. J. Environ. Res. Public Health 2016, 13, 522. [Google Scholar] [CrossRef] [PubMed]
  6. Baiano, A.; del Nobile, M.A. Antioxidant compounds from vegetable matrices: Biosynthesis, occurrence, and extraction systems. Crit. Rev. Food Sci. Nutr. 2015, 56, 2053–2068. [Google Scholar] [CrossRef] [PubMed]
  7. Cai, Y.Z.; Luo, Q.; Sun, M.; Corke, H. Antioxidant activity and phenolic compounds of 112 traditional Chinese medicinal plants associated with anticancer. Life Sci. 2004, 74, 2157–2184. [Google Scholar] [CrossRef] [PubMed]
  8. Shan, B.; Cai, Y.Z.; Sun, M.; Corke, H. Antioxidant capacity of 26 spice extracts and characterization of their phenolic constituents. J. Agric. Food Chem. 2005, 53, 7749–7759. [Google Scholar] [CrossRef] [PubMed]
  9. Fu, L.; Xu, B.T.; Gan, R.Y.; Zhang, Y.; Xu, X.R.; **a, E.Q.; Li, H.B. Total phenolic contents and antioxidant capacities of herbal and tea infusions. Int. J. Mol. Sci. 2011, 12, 2112–2124. [Google Scholar] [CrossRef] [PubMed]
  10. Fu, L.; Xu, B.T.; Xu, X.R.; Qin, X.S.; Gan, R.Y.; Li, H.B. Antioxidant capacities and total phenolic contents of 56 wild fruits from south China. Molecules 2010, 15, 8602–8617. [Google Scholar] [CrossRef] [PubMed]
  11. Fu, L.; Xu, B.T.; Xu, X.R.; Gan, R.Y.; Zhang, Y.; **a, E.Q.; Li, H.B. Antioxidant capacities and total phenolic contents of 62 fruits. Food Chem. 2011, 129, 345–350. [Google Scholar] [CrossRef]
  12. Deng, G.F.; Xu, X.R.; Guo, Y.J.; **a, E.Q.; Li, S.; Wu, S.; Chen, F.; Ling, W.H.; Li, H.B. Determination of antioxidant property and their lipophilic and hydrophilic phenolic contents in cereal grains. J. Funct. Food. 2012, 4, 906–914. [Google Scholar] [CrossRef]
  13. Guo, Y.J.; Deng, G.F.; Xu, X.R.; Wu, S.; Li, S.; **a, E.Q.; Li, F.; Chen, F.; Ling, W.H.; Li, H.B. Antioxidant capacities, phenolic compounds and polysaccharide contents of 49 edible macro-fungi. Food Funct. 2012, 3, 1195–1205. [Google Scholar] [CrossRef] [PubMed]
  14. Deng, G.F.; Lin, X.; Xu, X.R.; Gao, L.L.; **e, J.F.; Li, H.B. Antioxidant capacities and total phenolic contents of 56 vegetables. J. Funct. Food 2013, 5, 260–266. [Google Scholar] [CrossRef]
  15. Li, S.; Li, S.K.; Gan, R.Y.; Song, F.L.; Kuang, L.; Li, H.B. Antioxidant capacities and total phenolic contents of infusions from 223 medicinal plants. Ind. Crop. Prod. 2013, 51, 289–298. [Google Scholar] [CrossRef]
  16. Li, A.N.; Li, S.; Li, H.B.; Xu, D.P.; Xu, X.R.; Chen, F. Total phenolic contents and antioxidant capacities of 51 edible and wild flowers. J. Funct. Food 2014, 6, 319–330. [Google Scholar] [CrossRef]
  17. Li, Y.; Zhang, J.J.; Xu, D.P.; Zhou, T.; Zhou, Y.; Li, S.; Li, H.B. Bioactivities and health benefits of wild fruits. Int. J. Mol. Sci. 2016, 17, 1258. [Google Scholar] [CrossRef] [PubMed]
  18. Zhang, J.J.; Li, Y.; Zhou, T.; Xu, D.P.; Zhang, P.; Li, S.; Li, H.B. Bioactivities and health benefits of mushrooms mainly from China. Molecules 2016, 21, 938. [Google Scholar] [CrossRef] [PubMed]
  19. Deng, G.F.; Shen, C.; Xu, X.R.; Kuang, R.D.; Guo, Y.J.; Zeng, L.S.; Gao, L.L.; Lin, X.; **e, J.F.; **a, E.Q. Potential of fruit wastes as natural resources of bioactive compounds. Int. J. Mol. Sci. 2012, 13, 8308–8323. [Google Scholar] [CrossRef] [PubMed]
  20. Manach, C.; Scalbert, A.; Morand, C.; Remesy, C.; Jimenez, L. Polyphenols: Food sources and bioavailability. Am. J. Clin. Nutr. 2004, 79, 727–747. [Google Scholar] [PubMed]
  21. Jenab, M.; Riboli, E.; Ferrari, P.; Sabate, J.; Slimani, N.; Norat, T.; Friesen, M.; Tjonneland, A.; Olsen, A.; Overvad, K.; et al. Plasma and dietary vitamin C levels and risk of gastric cancer in the European Prospective Investigation into Cancer and Nutrition (EPIC-EURGAST). Carcinogenesis 2006, 27, 2250–2257. [Google Scholar] [CrossRef] [PubMed]
  22. Li, A.N.; Li, S.; Zhang, Y.J.; Xu, X.R.; Chen, Y.M.; Li, H.B. Resources and biological activities of natural polyphenols. Nutrients 2014, 6, 6020–6047. [Google Scholar] [CrossRef] [PubMed]
  23. Arathi, B.P.; Raghavendra-Rao Sowmya, P.; Vijay, K.; Baskaran, V.; Lakshminarayana, R. Metabolomics of carotenoids: The challenges and prospects—A review. Trends Food Sci. Technol. 2015, 45, 105–117. [Google Scholar] [CrossRef]
  24. Zhang, Y.J.; Gan, R.Y.; Li, S.; Zhou, Y.; Li, A.N.; Xu, D.P.; Li, H.B. Antioxidant phytochemicals for the prevention and treatment of chronic diseases. Molecules 2015, 20, 21138–21156. [Google Scholar] [CrossRef] [PubMed]
  25. Wojtunik-Kulesza, K.A.; Oniszczuk, A.; Oniszczuk, T.; Waksmundzka-Hajnos, M. The influence of common free radicals and antioxidants on development of Alzheimer’s disease. Biomed. Pharmacother. 2016, 78, 39–49. [Google Scholar] [CrossRef] [PubMed]
  26. Balmus, I.M.; Ciobica, A.; Trifan, A.; Stanciu, C. The implications of oxidative stress and antioxidant therapies in inflammatory bowel disease: Clinical aspects and animal models. Saudi J. Gastroenterol. 2016, 22, 3–17. [Google Scholar] [CrossRef] [PubMed]
  27. Prasad, K.N. Simultaneous activation of Nrf2 and elevation of antioxidant compounds for reducing oxidative stress and chronic inflammation in human Alzheimer’s disease. Mech. Ageing Dev. 2016, 153, 41–47. [Google Scholar] [CrossRef] [PubMed]
  28. Salomone, F.; Godos, J.; Zelber-Sagi, S. Natural antioxidants for non-alcoholic fatty liver disease: Molecular targets and clinical perspectives. Liver Int. 2016, 36, 5–20. [Google Scholar] [CrossRef] [PubMed]
  29. Zhou, Y.; Li, Y.; Zhou, T.; Zheng, J.; Li, S.; Li, H.B. Dietary natural products for prevention and treatment of liver cancer. Nutrients 2016, 8, 156. [Google Scholar] [CrossRef] [PubMed]
  30. Zhou, Y.; Zheng, J.; Li, Y.; Xu, D.P.; Li, S.; Chen, Y.M.; Li, H.B. Natural polyphenols for prevention and treatment of cancer. Nutrients 2016, 8, 515. [Google Scholar] [CrossRef] [PubMed]
  31. Zheng, J.; Zhou, Y.; Li, Y.; Xu, D.P.; Li, S.; Li, H.B. Spices for prevention and treatment of cancers. Nutrients 2016, 8, 495. [Google Scholar] [CrossRef] [PubMed]
  32. De Camargo, A.C.; Bismara Regitano-d’Arce, M.A.; Telles Biasoto, A.C.; Shahidi, F. Enzyme-assisted extraction of phenolics from winemaking by-products: Antioxidant potential and inhibition of α-glucosidase and lipase activities. Food Chem. 2016, 212, 395–402. [Google Scholar] [CrossRef] [PubMed]
  33. Belwal, T.; Dhyani, P.; Bhatt, I.D.; Rawal, R.S.; Pande, V. Optimization extraction conditions for improving phenolic content and antioxidant activity in Berberis asiatica fruits using response surface methodology (RSM). Food Chem. 2016, 207, 115–124. [Google Scholar] [CrossRef] [PubMed]
  34. Sharmila, G.; Nikitha, V.S.; Ilaiyarasi, S.; Dhivya, K.; Rajasekar, V.; Kumar, A.N.M.; Muthukumaran, K.; Muthukumaran, C. Ultrasound assisted extraction of total phenolics from Cassia auriculata leaves and evaluation of its antioxidant activities. Ind. Crop. Prod. 2016, 84, 13–21. [Google Scholar] [CrossRef]
  35. Van Tang, N.; Hong, N.T.P.; Bowyer, M.C.; van Altena, I.A.; Scarlett, C.J. Influence of solvents and novel extraction methods on bioactive compounds and antioxidant capacity of Phyllanthus amarus. Chem. Pap. 2016, 70, 556–566. [Google Scholar]
  36. Strati, I.F.; Oreopoulou, V. Effect of extraction parameters on the carotenoid recovery from tomato waste. Int. J. Food Sci. Technol. 2011, 46, 23–29. [Google Scholar] [CrossRef]
  37. Ho, K.K.H.Y.; Ferruzzi, M.G.; Liceaga, A.M.; Martin-Gonzalez, M.F.S. Microwave-assisted extraction of lycopene in tomato peels: Effect of extraction conditions on all-trans and cis-isomer yields. LWT-Food Sci. Technol. 2015, 62, 160–168. [Google Scholar] [CrossRef]
  38. Faidi, K.; Baaka, N.; Hammami, S.; El Mokni, R.; Mighri, Z.; Mhenni, M.F. Extraction of carotenoids from Lycium ferocissimum fruits for cotton dyeing: Optimization survey based on a central composite design method. Fibers Polym. 2016, 17, 36–43. [Google Scholar] [CrossRef]
  39. Azmir, J.; Zaidul, I.S.M.; Rahman, M.M.; Sharif, K.M.; Mohamed, A.; Sahena, F.; Jahurul, M.H.A.; Ghafoor, K.; Norulaini, N.A.N.; Omar, A.K.M. Techniques for extraction of bioactive compounds from plant materials: A review. J. Food Eng. 2013, 117, 426–436. [Google Scholar] [CrossRef]
  40. Barba, F.J.; Zhu, Z.; Koubaa, M.; Sant’Ana, A.S.; Orlien, V. Green alternative methods for the extraction of antioxidant bioactive compounds from winery wastes and by-products: A review. Trends Food Sci. Technol. 2016, 49, 96–109. [Google Scholar] [CrossRef]
  41. Xu, D.P.; Zhou, Y.; Zheng, J.; Li, S.; Li, A.N.; Li, H.B. Optimization of ultrasound-assisted extraction of natural antioxidants from the flower of Jatropha integerrima by response surface methodology. Molecules 2016, 21, 18. [Google Scholar] [CrossRef] [PubMed]
  42. He, B.; Zhang, L.L.; Yue, X.Y.; Liang, J.; Jiang, J.; Gao, X.L.; Yue, P.X. Optimization of ultrasound-assisted extraction of phenolic compounds and anthocyanins from blueberry (Vaccinium ashei) wine pomace. Food Chem. 2016, 204, 70–76. [Google Scholar] [CrossRef] [PubMed]
  43. Li, A.N.; Li, S.; Xu, D.P.; Xu, X.R.; Chen, Y.M.; Ling, W.H.; Chen, F.; Li, H.B. Optimization of ultrasound-assisted extraction of lycopene from papaya processing waste by response surface methodology. Food Anal. Meth. 2015, 8, 1207–1214. [Google Scholar] [CrossRef]
  44. Li, Y.; Fabiano-Tixier, A.S.; Tomao, V.; Cravotto, G.; Chemat, F. Green ultrasound-assisted extraction of carotenoids based on the bio-refinery concept using sunflower oil as an alternative solvent. Ultrason. Sonochem. 2013, 20, 12–18. [Google Scholar] [CrossRef] [PubMed]
  45. Milutinovic, M.; Radovanovic, N.; Corovic, M.; Siler-Marinkovic, S.; Rajilic-Stojanovic, M.; Dimitrijevic-Brankovic, S. Optimisation of microwave-assisted extraction parameters for antioxidants from waste Achillea millefolium dust. Ind. Crop. Prod. 2015, 77, 333–341. [Google Scholar] [CrossRef]
  46. Bouras, M.; Chadni, M.; Barba, F.J.; Grimi, N.; Bals, O.; Vorobiev, E. Optimization of microwave-assisted extraction of polyphenols from Quercus bark. Ind. Crop. Prod. 2015, 77, 590–601. [Google Scholar] [CrossRef]
  47. Tomaz, I.; Maslov, L.; Stupic, D.; Preiner, D.; Asperger, D.; Kontic, J.K. Recovery of flavonoids from grape skins by enzyme-assisted extraction. Sep. Sci. Technol. 2016, 51, 255–268. [Google Scholar] [CrossRef]
  48. Ranveer, R.C.; Patil, S.N.; Sahoo, A.K. Effect of different parameters on enzyme-assisted extraction of lycopene from tomato processing waste. Food Bioprod. Process. 2013, 91, 370–375. [Google Scholar] [CrossRef]
  49. Kamali, H.; Khodaverdi, E.; Hadizadeh, F.; Ghaziaskar, S.H. Optimization of phenolic and flavonoid content and antioxidants capacity of pressurized liquid extraction from Dracocephalum kotschyi via circumscribed central composite. J. Supercrit. Fluids 2016, 107, 307–314. [Google Scholar] [CrossRef]
  50. Golmakani, E.; Mohammadi, A.; Sani, T.A.; Kamali, H. Phenolic and flavonoid content and antioxidants capacity of pressurized liquid extraction and perculation method from roots of Scutellaria pinnatifida a. Hamilt. Subsp alpina (Bornm) Rech. F. J. Supercrit. Fluids 2014, 95, 318–324. [Google Scholar] [CrossRef]
  51. Shang, Y.F.; Kim, S.M.; Um, B. Optimisation of pressurised liquid extraction of antioxidants from black bamboo leaves. Food Chem. 2014, 154, 164–170. [Google Scholar] [CrossRef] [PubMed]
  52. Sanagi, M.M.; See, H.H.; Ibrahim, W.; Abu Naim, A. Determination of carotene, tocopherols and tocotrienols in residue oil from palm pressed fiber using pressurized liquid extraction-normal phase liquid chromatography. Anal. Chim. Acta 2005, 538, 71–76. [Google Scholar] [CrossRef] [Green Version]
  53. Pereira, P.; Cebola, M.J.; Oliveira, M.C.; Bernardo-Gil, M.G. Supercritical fluid extraction vs. conventional extraction of myrtle leaves and berries: Comparison of antioxidant activity and identification of bioactive compounds. J. Supercrit. Fluids 2016, 113, 1–9. [Google Scholar] [CrossRef]
  54. Kazan, A.; Koyu, H.; Turu, I.C.; Yesil-Celiktas, O. Supercritical fluid extraction of Prunus persica leaves and utilization possibilities as a source of phenolic compounds. J. Supercrit. Fluids 2014, 92, 55–59. [Google Scholar] [CrossRef]
  55. Jimenez-Aguilar, D.M.; Escobedo-Avellaneda, Z.; Martin-Belloso, O.; Gutierrez-Uribe, J.; Valdez-Fragoso, A.; Garcia-Garcia, R.; Torres, J.A.; Welti-Chanes, J. Effect of high hydrostatic pressure on the content of phytochemical compounds and antioxidant activity of prickly pears (Opuntia ficus-indica) beverages. Food Eng. Rev. 2015, 7, 198–208. [Google Scholar] [CrossRef]
  56. Lee, D.; Ghafoor, K.; Moon, S.; Kim, S.H.; Kim, S.; Chun, H.; Park, J. Phenolic compounds and antioxidant properties of high hydrostatic pressure and conventionally treated ginseng (Panax ginseng) products. Qual. Assur. Saf. Crops Foods 2015, 7, 493–500. [Google Scholar] [CrossRef]
  57. Teh, S.; Niven, B.E.; Bekhit, A.E.A.; Carne, A.; Birch, E.J. Microwave and pulsed electric field assisted extractions of polyphenols from defatted canola seed cake. Int. J. Food Sci. Technol. 2015, 50, 1109–1115. [Google Scholar] [CrossRef]
  58. Segovia, F.J.; Luengo, E.; Corral-Perez, J.J.; Raso, J.; Pilar Almajano, M. Improvements in the aqueous extraction of polyphenols from borage (Borago officinalis L.) leaves by pulsed electric fields: Pulsed electric fields (PEF) applications. Ind. Crop. Prod. 2015, 65, 390–396. [Google Scholar] [CrossRef]
  59. Luengo, E.; Alvarez, I.; Raso, J. Improving the pressing extraction of polyphenols of orange peel by pulsed electric fields. Innov. Food Sci. Emerg. Technol. 2013, 17, 79–84. [Google Scholar] [CrossRef]
  60. Rosello-Soto, E.; Barba, F.J.; Parniakov, O.; Galanakis, C.M.; Lebovka, N.; Grimi, N.; Vorobiev, E. High voltage electrical discharges, pulsed electric field, and ultrasound assisted extraction of protein and phenolic compounds from olive kernel. Food Bioprocess Technol. 2015, 8, 885–894. [Google Scholar] [CrossRef]
  61. Esclapez, M.D.; Garcia-Perez, J.V.; Mulet, A.; Carcel, J.A. Ultrasound-assisted extraction of natural products. Food Eng. Rev. 2011, 3, 108–120. [Google Scholar] [CrossRef]
  62. Soria, A.C.; Villamiel, M. Effect of ultrasound on the technological properties and bioactivity of food: A review. Trends Food Sci. Technol. 2010, 21, 323–331. [Google Scholar] [CrossRef]
  63. Chemat, F.; Zill-e-Huma; Khan, M.K. Applications of ultrasound in food technology: Processing, preservation and extraction. Ultrason. Sonochem. 2011, 18, 813–835. [Google Scholar] [CrossRef] [PubMed]
  64. Talmaciu, A.I.; Volf, I.; Popa, V.I. A comparative analysis of the “green” techniques applied for polyphenols extraction from bioresources. Chem. Biodivers. 2015, 12, 1635–1651. [Google Scholar] [CrossRef] [PubMed]
  65. Espada-Bellido, E.; Ferreiro-Gonzalez, M.; Carrera, C.; Palma, M.; Barroso, C.G.; Barbero, G.F. Optimization of the ultrasound-assisted extraction of anthocyanins and total phenolic compounds in mulberry (Morus nigra) pulp. Food Chem. 2017, 219, 23–32. [Google Scholar] [CrossRef] [PubMed]
  66. Virot, M.; Tomao, V.; Le Bourvellec, C.; Renard, C.M.C.G.; Chemat, F. Towards the industrial production of antioxidants from food processing by-products with ultrasound-assisted extraction. Ultrason. Sonochem. 2010, 17, 1066–1074. [Google Scholar] [CrossRef] [PubMed]
  67. Liu, C.T.; Wu, C.Y.; Weng, Y.M.; Tseng, C.Y. Ultrasound-assisted extraction methodology as a tool to improve the antioxidant properties of herbal drug **ao-chia-hu-tang. J. Ethnopharmacol. 2005, 99, 293–300. [Google Scholar] [CrossRef] [PubMed]
  68. Kazemi, M.; Karim, R.; Mirhosseini, H.; Hamid, A.A. Optimization of pulsed ultrasound-assisted technique for extraction of phenolics from pomegranate peel of Malas variety: Punicalagin and hydroxybenzoic acids. Food Chem. 2016, 206, 156–166. [Google Scholar] [CrossRef] [PubMed]
  69. Pan, Z.; Qu, W.; Ma, H.; Atungulu, G.G.; McHugh, T.H. Continuous and pulsed ultrasound-assisted extractions of antioxidants from pomegranate peel. Ultrason. Sonochem. 2011, 18, 1249–1257. [Google Scholar] [CrossRef] [PubMed]
  70. Xu, W.J.; Zhai, J.W.; Cui, Q.; Liu, J.Z.; Luo, M.; Fu, Y.J.; Zu, Y.G. Ultra-turrax based ultrasound-assisted extraction of five organic acids from honeysuckle (Lonicera japonica Thunb.) and optimization of extraction process. Sep. Purif. Technol. 2016, 166, 73–82. [Google Scholar] [CrossRef]
  71. Tan, Z.J.; Yi, Y.J.; Wang, H.Y.; Zhou, W.L.; Wang, C.Y. Extraction, preconcentration and isolation of flavonoids from Apocynum venetum L. Leaves using ionic liquid-based ultrasonic-assisted extraction coupled with an aqueous biphasic system. Molecules 2016, 21, 262. [Google Scholar] [CrossRef] [PubMed]
  72. Zhang, H.F.; Yang, X.H.; Wang, Y. Microwave assisted extraction of secondary metabolites from plants: Current status and future directions. Trends Food Sci. Technol. 2011, 22, 672–688. [Google Scholar] [CrossRef]
  73. Florez, N.; Conde, E.; Dominguez, H. Microwave assisted water extraction of plant compounds. J. Chem. Technol. Biotechnol. 2015, 90, 590–607. [Google Scholar] [CrossRef]
  74. Pap, N.; Beszedes, S.; Pongracz, E.; Myllykoski, L.; Gabor, M.; Gyimes, E.; Hodur, C.; Keiski, R.L. Microwave-assisted extraction of anthocyanins from black currant marc. Food Bioprocess Technol. 2013, 6, 2666–2674. [Google Scholar] [CrossRef]
  75. Dahmoune, F.; Spigno, G.; Moussi, K.; Remini, H.; Cherbal, A.; Madani, K. Pistacia lentiscus leaves as a source of phenolic compounds: Microwave-assisted extraction optimized and compared with ultrasound-assisted and conventional solvent extraction. Ind. Crop. Prod. 2014, 61, 31–40. [Google Scholar] [CrossRef]
  76. Dahmoune, F.; Nayak, B.; Moussi, K.; Remini, H.; Madani, K. Optimization of microwave-assisted extraction of polyphenols from Myrtus communis L. Leaves. Food Chem. 2015, 166, 585–595. [Google Scholar] [CrossRef] [PubMed]
  77. Tatke, P.; Jaiswal, Y. An overview of microwave assisted extraction and its applications in herbal drug research. Res. J. Med. Plant 2011, 5, 21–31. [Google Scholar] [CrossRef]
  78. Pasrija, D.; Anandharamakrishnan, C. Techniques for extraction of green tea polyphenols: A review. Food Bioprocess Technol. 2015, 8, 935–950. [Google Scholar] [CrossRef]
  79. Routray, W.; Orsat, V. Microwave-assisted extraction of flavonoids: A review. Food Bioprocess Technol. 2012, 5, 409–424. [Google Scholar] [CrossRef]
  80. Li, Y.; Fabiano-Tixier, A.S.; Vian, M.A.; Chemat, F. Solvent-free microwave extraction of bioactive compounds provides a tool for green analytical chemistry. TrAC-Trends Anal. Chem. 2013, 47, 1–11. [Google Scholar] [CrossRef]
  81. Wang, Z.M.; Ding, L.; Li, T.C.; Zhou, X.; Wang, L.; Zhang, H.Q.; Liu, L.; Li, Y.; Liu, Z.H.; Wang, H.J.; et al. Improved solvent-free microwave extraction of essential oil from dried Cuminum cyminum L. and Zanthoxylum bungeanum Maxim. J. Chromatogr. A 2006, 1102, 11–17. [Google Scholar] [CrossRef] [PubMed]
  82. Sahin, S. A novel technology for extraction of phenolic antioxidants from mandarin (Citrus deliciosa Tenore) leaves: Solvent-free microwave extraction. Korean J. Chem. Eng. 2015, 32, 950–957. [Google Scholar] [CrossRef]
  83. Bendjersi, F.Z.; Tazerouti, F.; Belkhelfa-Slimani, R.; Djerdjouri, B.; Meklati, B.Y. Phytochemical composition of the Algerian Laurus nobilis L. Leaves extracts obtained by solvent-free microwave extraction and investigation of their antioxidant activity. J. Essent. Oil Res. 2016, 28, 202–210. [Google Scholar] [CrossRef]
  84. Zill-e-Huma, M.A.V.; Maingonnat, J.F.; Chemat, F. Clean recovery of antioxidant flavonoids from onions: Optimising solvent free microwave extraction method. J. Chromatogr. A 2009, 1216, 7700–7707. [Google Scholar] [CrossRef] [PubMed]
  85. Perino-Issartier, S.; Zill-e-Huma; Abert-Vian, M.; Chemat, F. Solvent free microwave-assisted extraction of antioxidants from sea buckthorn (Hippophae rhamnoides) food by-products. Food Bioprocess Technol. 2011, 4, 1020–1028. [Google Scholar] [CrossRef]
  86. Michel, T.; Destandau, E.; Elfakir, C. Evaluation of a simple and promising method for extraction of antioxidants from sea buckthorn (Hippophae rhamnoides L.) berries: Pressurised solvent-free microwave assisted extraction. Food Chem. 2011, 126, 1380–1386. [Google Scholar] [CrossRef]
  87. Gu, H.; Chen, F.; Zhang, Q.; Zang, J. Application of ionic liquids in vacuum microwave-assisted extraction followed by macroporous resin isolation of three flavonoids rutin, hyperoside and hesperidin from Sorbus tianschanica leaves. J. Chromatogr. B Anal. Technol. Biomed. Life Sci. 2016, 1014, 45–55. [Google Scholar] [CrossRef] [PubMed]
  88. Cheng, X.; Bi, L.W.; Zhao, Z.D.; Chen, Y.X. Advances in enzyme assisted extraction of natural products. In AER-Advances in Engineering Research, 3nd ed.; Yarlagadda, P., Ed.; Atlantis Press: Paris, France, 2015; Volume 27, pp. 371–375. [Google Scholar]
  89. Liu, X.; Hu, Y.; Wei, D. Optimization of enzyme-based ultrasonic/microwave-assisted extraction and evaluation of antioxidant activity of orcinol glucoside from the rhizomes of Curculigo orchioides Gaertn. Med. Chem. Res. 2014, 23, 2360–2367. [Google Scholar] [CrossRef]
  90. Dinkova, R.; Heffels, P.; Shikov, V.; Weber, F.; Schieber, A.; Mihalev, K. Effect of enzyme-assisted extraction on the chilled storage stability of bilberry (Vaccinium myrtillus L.) anthocyanins in skin extracts and freshly pressed juices. Food Res. Int. 2014, 65, 35–41. [Google Scholar] [CrossRef]
  91. Nguyen, T.H.; Smagghe, G.; Gonzales, G.B.; van Camp, J.; Raes, K. Enzyme-assisted extraction enhancing the phenolic release from cauliflower (Brassica oleracea L. Var. Botrytis) outer Leaves. J. Agric. Food Chem. 2014, 62, 7468–7476. [Google Scholar]
  92. Mushtaq, M.; Sultana, B.; Bhatti, H.N.; Asghar, M. RSM based optimized enzyme-assisted extraction of antioxidant phenolics from underutilized watermelon (Citrullus lanatus Thunb.) rind. J. Food Sci. Technol.-Mysore 2015, 52, 5048–5056. [Google Scholar] [CrossRef] [PubMed]
  93. Liu, J.J.; Gasmalla, M.A.A.; Li, P.; Yang, R. Enzyme-assisted extraction processing from oilseeds: Principle, processing and application. Innov. Food Sci. Emerg. Technol. 2016, 35, 184–193. [Google Scholar] [CrossRef]
  94. Puri, M.; Sharma, D.; Barrow, C.J. Enzyme-assisted extraction of bioactives from plants. Trends Biotechnol. 2012, 30, 37–44. [Google Scholar] [CrossRef] [PubMed]
  95. Park, C.; Kim, S.; Lee, D.; Park, D.J.; Imm, J. Enzyme and high pressure assisted extraction of tricin from rice hull and biological activities of rice hull extract. Food Sci. Biotechnol. 2016, 25, 159–164. [Google Scholar] [CrossRef]
  96. Mustafa, A.; Turner, C. Pressurized liquid extraction as a green approach in food and herbal plants extraction: A review. Anal. Chim. Acta 2011, 703, 8–18. [Google Scholar] [CrossRef] [PubMed]
  97. Herrero, M.; del Pilar Sanchez-Camargo, A.; Cifuentes, A.; Ibanez, E. Plants, seaweeds, microalgae and food by-products as natural sources of functional ingredients obtained using pressurized liquid extraction and supercritical fluid extraction. TrAC-Trends Anal. Chem. 2015, 71, 26–38. [Google Scholar] [CrossRef]
  98. Setyaningsih, W.; Saputro, I.E.; Palma, M.; Barroso, C.G. Pressurized liquid extraction of phenolic compounds from rice (Oryza sativa) grains. Food Chem. 2016, 192, 452–459. [Google Scholar] [CrossRef] [PubMed]
  99. Breithaupt, D.E. Simultaneous HPLC determination of carotenoids used as food coloring additives: Applicability of accelerated solvent extraction. Food Chem. 2004, 86, 449–456. [Google Scholar] [CrossRef]
  100. Bozan, B.; Altinay, R.C. Accelerated solvent extraction of flavan-3-ol derivatives from grape seeds. Food Sci. Technol. Res. 2014, 20, 409–414. [Google Scholar] [CrossRef]
  101. Cai, Z.; Qu, Z.; Lan, Y.; Zhao, S.; Ma, X.; Wan, Q.; **g, P.; Li, P. Conventional, ultrasound-assisted, and accelerated-solvent extractions of anthocyanins from purple sweet potatoes. Food Chem. 2016, 197, 266–272. [Google Scholar] [CrossRef] [PubMed]
  102. Nieto, A.; Borrull, F.; Pocurull, E.; Marce, R.M. Pressurized liquid extraction: A useful technique to extract pharmaceuticals and personal-care products from sewage sludge. TrAC Trends Anal. Chem. 2010, 29, 752–764. [Google Scholar] [CrossRef]
  103. Franquin-Trinquier, S.; Maury, C.; Baron, A.; Le Meurlay, D.; Mehinagic, E. Optimization of the extraction of apple monomeric phenolics based on response surface methodology: Comparison of pressurized liquid-solid extraction and manual-liquid extraction. J. Food Compos. Anal. 2014, 34, 56–67. [Google Scholar] [CrossRef]
  104. Pineiro, Z.; Palma, M.; Barroso, C.G. Determination of trans-resveratrol in grapes by pressurised liquid extraction and fast high-performance liquid chromatography. J. Chromatogr. A 2006, 1110, 61–65. [Google Scholar] [CrossRef] [PubMed]
  105. Wianowska, D.; Typek, R.; Dawidowicz, A.L. Chlorogenic acid stability in pressurized liquid extraction conditions. J. AOAC Int. 2015, 98, 415–421. [Google Scholar] [CrossRef] [PubMed]
  106. Yen, H.; Yang, S.; Chen, C.; Jesisca; Chang, J. Supercritical fluid extraction of valuable compounds from microalgal biomass. Bioresour. Technol. 2015, 184, 291–296. [Google Scholar] [CrossRef] [PubMed]
  107. Radzali, S.A.; Baharin, B.S.; Othman, R.; Markom, M.; Rahman, R.A. Co-solvent selection for supercritical fluid extraction of astaxanthin and other carotenoids from Penaeus monodon Waste. J. Oleo Sci. 2014, 63, 769–777. [Google Scholar] [CrossRef] [PubMed]
  108. Capuzzo, A.; Maffei, M.E.; Occhipinti, A. Supercritical fluid extraction of plant flavors and fragrances. Molecules 2013, 18, 7194–7238. [Google Scholar] [CrossRef] [PubMed]
  109. Da Silva, R.P.F.F.; Rocha-Santos, T.A.P.; Duarte, A.C. Supercritical fluid extraction of bioactive compounds. TrAC-Trends Anal. Chem. 2016, 76, 40–51. [Google Scholar] [CrossRef]
  110. Sharif, K.M.; Rahman, M.M.; Azmir, J.; Mohamed, A.; Jahurul, M.H.A.; Sahena, F.; Zaidul, I.S.M. Experimental design of supercritical fluid extraction—A review. J. Food Eng. 2014, 124, 105–116. [Google Scholar] [CrossRef]
  111. Konar, N.; Dalabasmaz, S.; Poyrazoglu, E.S.; Artik, N.; Colak, A. The determination of the caffeic acid derivatives of Echinacea purpurea aerial parts under various extraction conditions by supercritical fluid extraction (SFE). J. Supercrit. Fluids 2014, 89, 128–136. [Google Scholar] [CrossRef]
  112. Maran, J.P.; Priya, B.; Manikandan, S. Modeling and optimization of supercritical fluid extraction of anthocyanin and phenolic compounds from Syzygium cumini fruit pulp. J. Food Sci. Technol. Mysore 2014, 51, 1938–1946. [Google Scholar] [CrossRef] [PubMed]
  113. Da Silva, G.F.; Konat Gandolfi, P.H.; Almeida, R.N.; Lucas, A.M.; Cassel, E.; Figueiro Vargas, R.M. Analysis of supercritical fluid extraction of lycopodine using response surface methodology and process mathematical modeling. Chem. Eng. Res. Des. 2015, 100, 353–361. [Google Scholar] [CrossRef]
  114. Haghayegh, M.; Zabihi, F.; Eikani, M.H.; Moghadas, B.K.; Yazdi, S.A.V. Supercritical fluid extraction of flavonoids and terpenoids from herbal compounds: Experiments and mathematical modeling. J. Essent. Oil Bear. Plants. 2015, 18, 1253–1265. [Google Scholar] [CrossRef]
  115. Zaghdoudi, K.; Framboisier, X.; Frochot, C.; Vanderesse, R.; Barth, D.; Kalthoum-Cherif, J.; Blanchard, F.; Guiavarc’h, Y. Response surface methodology applied to supercritical fluid extraction (SFE) of carotenoids from Persimmon (Diospyros kaki L.). Food Chem. 2016, 208, 209–219. [Google Scholar] [CrossRef] [PubMed]
  116. Liu, X.; Yang, D.; Liu, J.; Xu, K.; Wu, G. Modeling of supercritical fluid extraction of flavonoids from Calycopteris floribunda leaves. Chem. Pap. 2014, 68, 316–323. [Google Scholar] [CrossRef]
  117. Talmaciu, A.I.; Volf, I.; Popa, V.I. Supercritical fluids and ultrasound assisted extractions applied to spruce bark conversion. Environ. Eng. Manag. J. 2015, 14, 615–623. [Google Scholar]
  118. Solana, M.; Mirofci, S.; Bertucco, A. Production of phenolic and glucosinolate extracts from rocket salad by supercritical fluid extraction: Process design and cost benefits analysis. J. Food Eng. 2016, 168, 35–41. [Google Scholar] [CrossRef]
  119. Pasquel Reategui, J.L.; Da Fonseca Machado, A.P.; Barbero, G.F.; Rezende, C.A.; Martinez, J. Extraction of antioxidant compounds from blackberry (Rubus sp.) bagasse using supercritical CO2 assisted by ultrasound. J. Supercrit. Fluids 2014, 94, 223–233. [Google Scholar] [CrossRef]
  120. Mushtaq, M.; Sultana, B.; Anwar, F.; Adnan, A.; Rizvi, S.S.H. Enzyme-assisted supercritical fluid extraction of phenolic antioxidants from pomegranate peel. J. Supercrit. Fluids 2015, 104, 122–131. [Google Scholar] [CrossRef]
  121. Corrales, M.; Garcia, A.F.; Butz, P.; Tauscher, B. Extraction of anthocyanins from grape skins assisted by high hydrostatic pressure. J. Food Eng. 2009, 90, 415–421. [Google Scholar] [CrossRef]
  122. Altuner, E.M.; Tokusoglu, O. The effect of high hydrostatic pressure processing on the extraction, retention and stability of anthocyanins and flavonols contents of berry fruits and berry juices. Int. J. Food Sci. Technol. 2013, 48, 1991–1997. [Google Scholar] [CrossRef]
  123. Toepfl, S.; Mathys, A.; Heinz, V.; Knorr, D. Review: Potential of high hydrostatic pressure and pulsed electric fields for energy efficient and environmentally friendly food processing. Food Rev. Int. 2006, 22, 405–423. [Google Scholar] [CrossRef]
  124. **, J.; Shen, D.; Zhao, S.; Lu, B.; Li, Y.; Zhang, R. Characterization of polyphenols from green tea leaves using a high hydrostatic pressure extraction. Int. J. Pharm. 2009, 382, 139–143. [Google Scholar] [CrossRef] [PubMed]
  125. Uzelac, V.D.; Putnik, P.; Zoric, Z.; Jezek, D.; Karlovic, S.; Kovacevic, D.B. Winery by-products: Anthocyanins recovery from red grape skin by high hydrostatic pressure extraction (HHPE). Ann. Nutr. Metab. 2015, 671, 522–523. [Google Scholar]
  126. Kovacevic, D.B.; Putnik, P.; Pedisic, S.; Jazek, D.; Karlovic, S.; Uzelac, V.D. High hydrostatic pressure extraction of flavonoids from freeze-dried red grape skin as winemaking by-product. Ann. Nutr. Metab. 2015, 671, 521–522. [Google Scholar]
  127. **, J.; Luo, S. The mechanism for enhancing, extraction of ferulic acid from Radix Angelica sinensis by high hydrostatic pressure. Sep. Purif. Technol. 2016, 165, 208–213. [Google Scholar] [CrossRef]
  128. Lopez, N.; Puertolas, E.; Condon, S.; Raso, J.; Alvarez, I. Enhancement of the extraction of betanine from red beetroot by pulsed electric fields. J. Food Eng. 2009, 90, 60–66. [Google Scholar] [CrossRef]
  129. Zderic, A.; Zondervan, E. Polyphenol extraction from fresh tea leaves by pulsed electric field: A study of mechanisms. Chem. Eng. Res. Des. 2016, 109, 586–592. [Google Scholar] [CrossRef]
  130. Bouras, M.; Grimi, N.; Bals, O.; Vorobiev, E. Impact of pulsed electric fields on polyphenols extraction from Norway spruce bark. Ind. Crop. Prod. 2016, 80, 50–58. [Google Scholar] [CrossRef]
  131. Zhou, Y.; Zhao, X.; Huang, H. Effects of pulsed electric fields on anthocyanin extraction yield of blueberry processing by-products. J. Food Process Preserv. 2015, 39, 1898–1904. [Google Scholar] [CrossRef]
  132. Bansal, V.; Sharma, A.; Ghanshyam, C.; Singla, M.L. Optimization and characterization of pulsed electric field parameters for extraction of quercetin and ellagic acid in emblica officinalis juice. J. Food Meas. Charact. 2014, 8, 225–233. [Google Scholar] [CrossRef]
  133. Gachovska, T.; Cassada, D.; Subbiah, J.; Hanna, M.; Thippareddi, H.; Snow, D. Enhanced anthocyanin extraction from red cabbage using pulsed electric field processing. J. Food Sci. 2010, 75, E323–E329. [Google Scholar] [CrossRef] [PubMed]
  134. Luengo, E.; Martinez, J.M.; Bordetas, A.; Alvarez, I.; Raso, J. Influence of the treatment medium temperature on lutein extraction assisted by pulsed electric fields from Chlorella vulgaris. Innov. Food Sci. Emerg. Technol. 2015, 29, 15–22. [Google Scholar] [CrossRef]
  135. Roohinejad, S.; Oey, I.; Everett, D.W.; Niven, B.E. Evaluating the effectiveness of beta-carotene extraction from pulsed electric field-treated carrot pomace using oil-in-water microemulsion. Food Bioprocess Technol. 2014, 7, 3336–3348. [Google Scholar] [CrossRef]
  136. Lopez, N.; Puertolas, E.; Condon, S.; Alvarez, I.; Raso, J. Effects of pulsed electric fields on the extraction of phenolic compounds during the fermentation of must of Tempranillo grapes. Innov. Food Sci. Emerg. Technol. 2008, 9, 477–482. [Google Scholar] [CrossRef]
  137. Boussetta, N.; Vorobiev, E. Extraction of valuable biocompounds assisted by high voltage electrical discharges: A review. Comptes Rendus Chim. 2014, 17, 197–203. [Google Scholar] [CrossRef]
  138. Boussetta, N.; Lanoiselle, J.; Bedel-Cloutour, C.; Vorobiev, E. Extraction of soluble matter from grape pomace by high voltage electrical discharges for polyphenol recovery: Effect of sulphur dioxide and thermal treatments. J. Food Eng. 2009, 95, 192–198. [Google Scholar] [CrossRef]
  139. Rajha, H.N.; Boussetta, N.; Louka, N.; Maroun, R.G.; Vorobiev, E. Electrical, mechanical, and chemical effects of high-voltage electrical discharges on the polyphenol extraction from vine shoots. Innov. Food Sci. Emerg. Technol. 2015, 31, 60–66. [Google Scholar] [CrossRef]
  140. Sarkis, J.R.; Boussetta, N.; Tessaro, I.C.; Ferreira Marczak, L.D.; Vorobiev, E. Application of pulsed electric fields and high voltage electrical discharges for oil extraction from sesame seeds. J. Food Eng. 2015, 153, 20–27. [Google Scholar] [CrossRef]
  141. Le, H.V.; Le, V.V.M. Comparison of enzyme-assisted and ultrasound-assisted extraction of vitamin C and phenolic compounds from acerola (Malpighia emarginata DC.) fruit. Int. J. Food Sci. Technol. 2012, 47, 1206–1214. [Google Scholar] [CrossRef]
  142. Rodriguez-Perez, C.; Gilbert-Lopez, B.; Mendiola, J.A.; Quirantes-Pine, R.; Segura-Carretero, A.; Ibanez, E. Optimization of microwave-assisted extraction and pressurized liquid extraction of phenolic compounds from Moringa oleifera leaves by multiresponse surface methodology. Electrophoresis 2016, 37, 1938–1946. [Google Scholar] [CrossRef] [PubMed]
  143. Das, A.K.; Mandal, V.; Mandal, S.C. A brief understanding of process optimisation in microwave-assisted extraction of botanical materials: Options and opportunities with chemometric tools. Phytochem. Anal. 2014, 25, 1–12. [Google Scholar] [CrossRef] [PubMed]
  144. Chemat, F.; Rombaut, N.; Sicaire, A.G.; Meullemiestre, A.; Fabiano-Tixier, A.S.; Abert-Vian, M. Ultrasound assisted extraction of food and natural products. Mechanisms, techniques, combinations, protocols and applications. A review. Ultrason. Sonochem. 2017, 34, 540–560. [Google Scholar] [CrossRef] [PubMed]
  145. Eskilsson, C.S.; Bjorklund, E. Analytical-scale microwave-assisted extraction. J. Chromatogr. A 2000, 902, 227–250. [Google Scholar] [CrossRef]
  146. Senorans, F.J.; Ibanez, E.; Cifuentes, A. New trends in food processing. Crit. Rev. Food Sci. Nutr. 2003, 43, 507–526. [Google Scholar] [CrossRef] [PubMed]
  147. Parniakov, O.; Barba, F.J.; Grimi, N.; Lebovka, N.; Vorobiev, E. Impact of pulsed electric fields and high voltage electrical discharges on extraction of high-added value compounds from papaya peels. Food Res. Int. 2014, 65, 337–343. [Google Scholar] [CrossRef]
  148. Huang, D.J.; Ou, B.X.; Prior, R.L. The chemistry behind antioxidant capacity assays. J. Agric. Food Chem. 2005, 53, 1841–1856. [Google Scholar] [CrossRef] [PubMed]
  149. Prior, R.L.; Wu, X.L.; Schaich, K. Standardized methods for the determination of antioxidant capacity and phenolics in foods and dietary supplements. J. Agric. Food Chem. 2005, 53, 4290–4302. [Google Scholar] [CrossRef] [PubMed]
  150. Lopez-Alarcon, C.; Denicola, A. Evaluating the antioxidant capacity of natural products: A review on chemical and cellular-based assays. Anal. Chim. Acta 2013, 763, 1–10. [Google Scholar] [CrossRef] [PubMed]
  151. Frankel, E.N.; Meyer, A.S. The problems of using one-dimensional methods to evaluate multifunctional food and biological antioxidants. J. Sci. Food Agric. 2000, 80, 1925–1941. [Google Scholar] [CrossRef]
  152. Arts, M.; Dallinga, J.S.; Voss, H.P.; Haenen, G.; Bast, A. A new approach to assess the total antioxidant capacity using the TEAC assay. Food Chem. 2004, 88, 567–570. [Google Scholar] [CrossRef]
  153. Schaich, K.M.; Tian, X.; **e, J. Hurdles and pitfalls in measuring antioxidant efficacy: A critical evaluation of ABTS, DPPH, and ORAC assays (Reprinted). J. Funct. Food. 2015, 18, 782–796. [Google Scholar] [CrossRef]
  154. Awika, J.M.; Rooney, L.W.; Wu, X.L.; Prior, R.L.; Cisneros-Zevallos, L. Screening methods to measure antioxidant activity of sorghum (Sorghum bicolor) and sorghum products. J. Agric. Food Chem. 2003, 51, 6657–6662. [Google Scholar] [CrossRef] [PubMed]
  155. Antolovich, M.; Prenzler, P.D.; Patsalides, E.; McDonald, S.; Robards, K. Methods for testing antioxidant activity. Analyst 2002, 127, 183–198. [Google Scholar] [CrossRef] [PubMed]
  156. Pulido, R.; Bravo, L.; Saura-Calixto, F. Antioxidant activity of dietary polyphenols as determined by a modified ferric reducing/antioxidant power assay. J. Agric. Food Chem. 2000, 48, 3396–3402. [Google Scholar] [CrossRef] [PubMed]
  157. Ou, B.X.; Huang, D.J.; Hampsch-Woodill, M.; Flanagan, J.A.; Deemer, E.K. Analysis of antioxidant activities of common vegetables employing oxygen radical absorbance capacity (ORAC) and ferric reducing antioxidant power (FRAP) assays: A comparative study. J. Agric. Food Chem. 2002, 50, 3122–3128. [Google Scholar] [CrossRef] [PubMed]
  158. Benzie, I.F.F.; Strain, J.J. The ferric reducing ability of plasma (FRAP) as a measure of “Antioxidant power”: The FRAP assay. Anal. Biochem. 1996, 239, 70–76. [Google Scholar] [CrossRef] [PubMed]
  159. Amorati, R.; Valgimigli, L. Advantages and limitations of common testing methods for antioxidants. Free Radic. Res. 2015, 49, 633–649. [Google Scholar] [CrossRef] [PubMed]
  160. Ozyurek, M.; Guclu, K.; Apak, R. The main and modified CUPRAC methods of antioxidant measurement. TrAC-Trends Anal. Chem. 2011, 30, 652–664. [Google Scholar] [CrossRef]
  161. Apak, R.; Guclu, K.; Demirata, B.; Ozyurek, M.; Celik, S.E.; Bektasoglu, B.; Berker, K.I.; Ozyurt, D. Comparative evaluation of various total antioxidant capacity assays applied to phenolic compounds with the CUPRAC assay. Molecules 2007, 12, 1496–1547. [Google Scholar] [CrossRef] [PubMed]
  162. Gulcin, I. Antioxidant activity of food constituents: An overview. Arch. Toxicol. 2012, 86, 345–391. [Google Scholar] [CrossRef] [PubMed]
  163. Singleton, V.L.; Orthofer, R.; Lamuela-Raventos, R.M. Analysis of total phenols and other oxidation substrates and antioxidants by means of Folin-Ciocalteu reagent. In Methods in Enzymology; Packer, L., Ed.; Elsevier Academic Press: Sandiego, Chile, 1999; Volume 299, pp. 152–178. [Google Scholar]
  164. Cicco, N.; Lanorte, M.T.; Paraggio, M.; Viggiano, M.; Lattanzio, V. A reproducible, rapid and inexpensive Folin-Ciocalteu micro-method in determining phenolics of plant methanol extracts. Microchem. J. 2009, 91, 107–110. [Google Scholar] [CrossRef]
  165. Magalhaes, L.M.; Segundo, M.A.; Reis, S.; Lima, J.L.F.C.; Toth, I.V.; Rangel, A.O.S.S. Automatic flow system for sequential determination of ABTS(center dot+) scavenging capacity and Folin-Ciocalteu index: A comparative study in food products. Anal. Chim. Acta 2007, 592, 193–201. [Google Scholar] [CrossRef] [PubMed]
  166. Shaghaghia, M.; Manzoori, J.L.; Jouyban, A. Determination of total phenols in tea infusions, tomato and apple juice by terbium sensitized fluorescence method as an alternative approach to the Folin-Ciocalteu spectrophotometric method. Food Chem. 2008, 108, 695–701. [Google Scholar] [CrossRef] [PubMed]
  167. Wootton-Beard, P.C.; Moran, A.; Ryan, L. Stability of the total antioxidant capacity and total polyphenol content of 23 commercially available vegetable juices before and after in vitro digestion measured by FRAP, DPPH, ABTS and Folin-Ciocalteu methods. Food Res. Int. 2011, 44, 217–224. [Google Scholar] [CrossRef]
  168. Yoo, K.S.; Lee, E.J.; Leskovar, D.; Patil, B.S. Development of an automated method for Folin-Ciocalteu total phenolic assay in artichoke extracts. J. Food Sci. 2012, 77, C1278–C1283. [Google Scholar] [CrossRef] [PubMed]
  169. Everette, J.D.; Bryant, Q.M.; Green, A.M.; Abbey, Y.A.; Wangila, G.W.; Walker, R.B. Thorough study of reactivity of various compound classes toward the Folin-Ciocalteu reagent. J. Agric. Food Chem. 2010, 58, 8139–8144. [Google Scholar] [CrossRef] [PubMed]
  170. Berker, K.I.; Olgun, F.A.O.; Ozyurt, D.; Demirata, B.; Apak, R. Modified Folin-Ciocalteu antioxidant capacity assay for measuring lipophilic antioxidants. J. Agric. Food Chem. 2013, 61, 4783–4791. [Google Scholar] [CrossRef] [PubMed]
  171. Sanchez-Moreno, C. Review: Methods used to evaluate the free radical scavenging activity in foods and biological systems. Food Sci. Technol. Int. 2002, 8, 121–137. [Google Scholar] [CrossRef]
  172. Nkhili, E.; Brat, P. Reexamination of the ORAC assay: Effect of metal ions. Anal. Bioanal. Chem. 2011, 400, 1451–1458. [Google Scholar] [CrossRef] [PubMed]
  173. Ou, B.X.; Hampsch-Woodill, M.; Prior, R.L. Development and validation of an improved oxygen radical absorbance capacity assay using fluorescein as the fluorescent probe. J. Agric. Food Chem. 2001, 49, 4619–4626. [Google Scholar] [CrossRef] [PubMed]
  174. Wood, L.G.; Gibson, P.G.; Garg, M.L. A review of the methodology for assessing in vivo antioxidant capacity. J. Sci. Food Agric. 2006, 86, 2057–2066. [Google Scholar] [CrossRef]
  175. Wolfe, K.L.; Liu, R.H. Cellular antioxidant activity (CAA) assay for assessing antioxidants, foods, and dietary supplements. J. Agric. Food Chem. 2007, 55, 8896–8907. [Google Scholar] [CrossRef] [PubMed]
  176. **ng, J.; Wang, G.; Zhang, Q.; Liu, X.; Yin, B.; Fang, D.; Zhao, J.; Zhang, H.; Chen, Y.Q.; Chen, W. Determining antioxidant activities of lactobacilli by cellular antioxidant assay in mammal cells. J. Funct. Food 2015, 19, 554–562. [Google Scholar] [CrossRef]
  177. Blasa, M.; Angelino, D.; Gennari, L.; Ninfali, P. The cellular antioxidant activity in red blood cells (CAA-RBC): A new approach to bioavailability and synergy of phytochemicals and botanical extracts. Food Chem. 2011, 125, 685–691. [Google Scholar] [CrossRef]
  178. Zhang, X.D.; Xu, Y.; Zhang, T.; Lu, J.J. Assessing plant antioxidants by cellular antioxidant activity assay based on microfluidic cell chip with arrayed microchannels. Chin. J. Anal. Chem. 2016, 44, 604–609. [Google Scholar] [CrossRef]
  179. O’sullivan, A.M.; O’Callaghan, Y.C.; O’Grady, M.N.; Queguineur, B.; Hanniffy, D.; Troy, D.J.; Kerry, J.P.; O’Brien, N.M. In vitro and cellular antioxidant activities of seaweed extracts prepared from five brown seaweeds harvested in spring from the west coast of Ireland. Food Chem. 2011, 126, 1064–1070. [Google Scholar] [CrossRef]
  180. Baroni, M.V.; Di Paola Naranjo, R.D.; Garcia-Ferreyra, C.; Otaiza, S.; Wunderlin, D.A. How good antioxidant is the red wine? Comparison of some in vitro and in vivo methods to assess the antioxidant capacity of Argentinean red wines. LWT-Food Sci. Technol. 2012, 47, 1–7. [Google Scholar] [CrossRef]
  181. Biswas, S.K.; McClure, D.; Jimenez, L.A.; Megson, I.L.; Rahman, I. Curcumin induces glutathione biosynthesis and inhibits NF-κB activation and interleukin-8 release in alveolar epithelial cells: Mechanism of free radical scavenging activity. Antioxid. Redox Signal. 2005, 7, 32–41. [Google Scholar] [CrossRef] [PubMed]
  182. **e, C.; Kang, J.; Ferguson, M.E.; Nagarajan, S.; Badger, T.M.; Wu, X. Blueberries reduce pro-inflammatory cytokine TNF-alpha and IL-6 production in mouse macrophages by inhibiting NF-κB activation and the MAPK pathway. Mol. Nutr. Food Res. 2011, 55, 1587–1591. [Google Scholar] [CrossRef] [PubMed]
  183. Song, W.; Derito, C.M.; Liu, M.K.; He, X.; Dong, M.; Liu, R.H. Cellular antioxidant activity of common vegetables. J. Agric. Food Chem. 2010, 58, 6621–6629. [Google Scholar] [CrossRef] [PubMed]
  184. Wolfe, K.L.; Kang, X.; He, X.; Dong, M.; Zhang, Q.; Liu, R.H. Cellular antioxidant activity of common fruits. J. Agric. Food Chem. 2008, 56, 8418–8426. [Google Scholar] [CrossRef] [PubMed]
  185. Surarit, W.; Jansom, C.; Lerdvuthisopon, N.; Kongkham, S.; Hansakul, P. Evaluation of antioxidant activities and phenolic subtype contents of ethanolic bran extracts of Thai pigmented rice varieties through chemical and cellular assays. Int. J. Food Sci. Technol. 2015, 50, 990–998. [Google Scholar] [CrossRef]
  186. Erlund, I. Review of the flavonoids quercetin, hesperetin, and naringenin. Dietary sources, bioactivities, bioavailability, and epidemiology. Nutr. Res. 2004, 24, 851–874. [Google Scholar] [CrossRef]
  187. Liggins, J.; Bluck, L.J.C.; Runswick, S.; Atkinson, C.; Coward, W.A.; Bingham, S.A. Daidzein and genistein contents of vegetables. Br. J. Nutr. 2000, 84, 717–725. [Google Scholar] [PubMed]
  188. Mezzomo, N.; Ferreira, S.R. Carotenoids functionality, sources, and processing by supercritical technology: A review. J. Chem. 2016, 2016, 3164312. [Google Scholar] [CrossRef]
  189. Borguini, R.G.; Ferraz Da Silva Torres, E.A. Tomatoes and tomato products as dietary sources of antioxidants. Food Rev. Int. 2009, 25, 313–325. [Google Scholar] [CrossRef]
  190. Abdel-Aal, E.S.M.; Akhtar, H.; Zaheer, K.; Ali, R. Dietary sources of lutein and zeaxanthin carotenoids and their role in eye health. Nutrients 2013, 5, 1169–1185. [Google Scholar] [CrossRef] [PubMed]
  191. Raposo, M.F.D.J.; de Morais, A.M.M.B.; de Morais, R.M.S.C. Carotenoids from marine microalgae: A valuable natural source for the prevention of chronic diseases. Mar. Drugs 2015, 13, 5128–5155. [Google Scholar] [CrossRef] [PubMed]
  192. Xu, B.; Chang, S.K.C. Comparative study on antiproliferation properties and cellular antioxidant activities of commonly consumed food legumes against nine human cancer cell lines. Food Chem. 2012, 134, 1287–1296. [Google Scholar] [CrossRef] [PubMed]
  193. Zhou, Q.; Lu, W.; Niu, Y.; Liu, J.; Zhang, X.; Gao, B.; Akoh, C.C.; Shi, H.; Yu, L.L. Identification and quantification of phytochemical composition and anti-inflammatory, cellular antioxidant, and radical scavenging activities of 12 plantago species. J. Agric. Food Chem. 2013, 61, 6693–6702. [Google Scholar] [CrossRef] [PubMed]
  194. Lai, Q.; Wang, H.; Guo, X.; Abbasi, A.M.; Wang, T.; Li, T.; Fu, X.; Li, J.; Liu, R.H. Comparison of phytochemical profiles, antioxidant and cellular antioxidant activities of seven cultivars of Aloe. Int. J. Food Sci. Technol. 2016, 51, 1489–1494. [Google Scholar] [CrossRef]
  195. Li, Y.; Sun, H.Y.; Yu, X.Y.; Liu, D.; Wan, H.X. Evaluation of cellular antioxidant and antiproliferative activities of five main Phyllanthus emblica L. cultivars in China. Indian J. Pharm. Sci. 2015, 77, 274–282. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The outline of extraction of antioxidants from foods and medicinal plants.
Figure 1. The outline of extraction of antioxidants from foods and medicinal plants.
Ijms 18 00096 g001
Table 1. The application of non-conventional techniques in the extraction of antioxidants from some foods and medicinal plants.
Table 1. The application of non-conventional techniques in the extraction of antioxidants from some foods and medicinal plants.
SourceCompounds ExtractedExtraction ParametersExtraction ImprovementReference
Non-Conventional MethodConventional Methods
ultrasound-assisted extraction (UAE)
blueberry wine pomaceanthocyanins and phenolicssolvents: 70% ethanol and 0.01% hydrochloric acid; conditions: 400 w, 61.03 °C, 23.67 min 70% ethanol and 0.01% hydrochloric acid; 61 °C, 35 min without ultrasound treatmentincreased total anthocyanins from 1.72 to 4.27 mg C3G/g (2.5-fold) and total phenolics from 5.08 to 16.41 mg gallic acid equivalent (GAE)/g (about 3.2-fold)[42]
papayalycopenesolvents: 42.28% ethanol in ethyl acetate conditions: 40 kHz, 800 W, 26.09 min, 50.12 °C40% ethanol in ethyl acetate (300 mL) 95 °C in a Soxhlet extractorRecovery of lycopene increased from 68.3 ± 4.1 to 189.8 ± 4.5 μg/g[43]
carrotcarotenoidssolvents: sunflower oil conditions: 22.5 W/cm2, 40 °C, 20 minhexane at room temperature for one hourobtained the β-carotene yield of 334.75 mg/L just in 20 min while the CSE method using hexane as solvent obtained the β-carotene yield of 321.36 mg/L after one-hour extraction[44]
microwave-assisted extraction (MAE)
Achillea millefolium dustantioxidantssolvents: 70% ethanol conditions: 170 W, 40 mL/g, 33 s40% ethanol at room temperature (1:10, v/v) for 48 hincreased total polyphenol content from 135.26 ± 1.72 to 237.74 ± 2.08 mg GAE/g, total flavonoid content from 30.82 ± 2.35 to 42.95 ± 1.32 mg quercetin equivalents (QE)/g, 2,2-diphenyl-1-picrylhydrazyl (DPPH) radical scavenging activity from 21.58% ± 0.88% to 71.72% ± 2.12%[45]
Quercus barkpolyphenolssolvents: ethanol content 33%, methanol content 0.38% conditions: 50 Hz, 45 W, 60 min, pH 10.75, room temperaturethe same extraction condition without microwave treatmentincreased by 3 times and 2 times respectively for total phenolic content and antioxidant recoveries[46]
enzyme-assisted extraction (EAE)
wine making by-productsphenolicssolvents: 70% acetone enzyme treatment with 2% viscozyme solution stirred for 12 h at 37 °C or 1 mg/mL pronase solution stirred for 1 h, then extraction with 70% (v/v) acetone in a gyratory water bath shaker at 30 °C for 20 minthe same extraction protocol without enzyme treatmentpronase and viscozyme increased the content of soluble phenolics while reducing the content of insoluble-bound phenolics[32]
grape skinsflavonoidssolvents: buffer solution containing an appropriate amount of enzyme conditions: 10.52 mg/g Lallzyme EX-V, pH 2.0, extraction at 45 °C for 3 h70% aqueous ethanol containing 1% formic acid for one day in the darkimproved recovery of anthocyanin contents (from 40,496.19 ± 58.18 to 41,752.95 ± 76.10 mg/kg) and flavan-3-ol contents (from 329.32 ± 2.46 to 345.94 ± 2.88 mg/kg)[47]
tomato processing wastelycopenesolvents: hexane/acetone/ethanol (50:25:25 v/v) conditions: 1.5% cellulase/2% pectinase at 4 h of incubation periodwithout enzyme treatmentincreased the yield of lycopene from less than 200 to 847.33 μg/g (cellulase treatment) and to 1262.56 μg/g (pectinase treatment)[48]
pressurized liquid extraction (PLE)
Aerial parts of Dracoceph-alum kotschyiphenolics and flavonoidssolvents: methanol conditions: 74 °C, 34 bar pressure, 11.33 min static time, 17.45 min dynamic time, and 0.7 mL/min solvent flow ratepercolated with 1.0 L of methanol at room temperature (25 °C) according to the European Pharmacopeiaimproved recovery of total phenolic (from 22.29 ± 0.05 to 30.92 ± 0.03 GAE mg/g), total flavonoids (from 5.042 ± 0.04 to 6.13 ± 0.07 QE mg/g) and luteolin content (from 9550 ± 0.3 to 13,247 ± 0.2 μg/g)[49]
roots of Scutellaria pinnatifidaphenolics and flavonoidssolvents: methanol conditions: 65.8 °C, 39.2 bar pressure, 12.9 min static time, 18.9 min dynamic time, and 0.76 mL/min solvent flow ratepercolated with 1.0 L of methanol at room temperaturethe total phenolic content increased from 196.66 to 396.94 mg/g, and the total flavonoid content increased from 91.3 to 127.78 mg/g[50]
black bamboo leavesantioxidantssolvents: 50% ethanol for the total phenolic (TP) and 75% ethanol for total flavonoid (TF) and 25% ethanol for DPPH radical scavenging ability conditions: 1500 psi, 200 °C, 25 min static timereflux extraction method (~90 °C, 1 L solvent, 60 min)improved extraction yields from 240 to 500 mg/1 g Dry black bamboo leaves (DL), TP contents from 1510 ± 3.2 to 2682 ± 0.9 mg/100 g, TF contents from 182 ± 2.7 to 657 ± 1.7 mg/100 g[51]
palm pressed fiberβ-carotenesolvents: n-hexane conditions: 80 °C, 1500 psi, 2 × 10 min static extractions with flush volume 50%extracted with n-hexane and chloroform in a Soxhlet apparatus for 8 hobtained total β-carotene and vitamin contents comparable to Soxhlet extraction but with lower total organic solvent and rapid extraction process[52]
supercritical fluid extraction (SFE)
myrtle leaves and berriesantioxidantssolvents: carbon dioxide conditions: 23 MPa, 45 °C and a CO2 flow of 0.3 kg/h using absolute ethanol as co-solvent with a flow rate of 0.09 kg/hobtained by hydrodistillation using a Clevenger-type apparatus, for two hoursincreased antioxidant capacity (about 20–40 times), polyphenolic contents (about 2 times) and myricetin-3-O-rhamnoside content (about 110–170 times in fruit and about 130–210 times in leaves)[53]
Prunus persica leavesphenolic compoundssolvents: carbon dioxide conditions: 60 °C, 150 bar and 6% ethanol co-solvent at a flow rate of 15 g/min and for a duration of 60 minextracted 3 times with 30 mL of solvent system (acetone:methanol:water:formic acid, 40:40:20:0.1)the radical scavenging activity value increased from 32.23% to 53.25%[54]
high hydrostatic pressure extraction (HHPE)
prickly pear beverages prepared with 10% peel and 90% pulpPhyto-chemical Compounds400 or 550 MPa, room temperature, 0–16 minthermally treated at 138 ± 1 °C for 2 sincreased TP content (16%–35%) and antioxidant activity (8%–17%) for Cristal (A) and Rojo San Martin varieties as well as increased the betaxanthin contents (6%–8%) and betacyanin content (4%–7%) for Rojo San Martin variety[55]
Panax ginsengphenolic compounds600 MPa for 1 min at room temperatureconventional steamingincreased the total phenolic contents (from 1.13 to 1.37 mg maltol equivalent/g of red ginseng), especially maltol content (4.38 to 12.61 mg/100 g of red ginseng), also improved the ferrous ion chelating and superoxide dismutase activities[56]
Pulsed electric field extraction (PEFE)
defatted canola seed cakepolyphenols10% ethanol 30 V, 30 Hz and 10 smicrowave processing (5 min, liquid/solid ratio of 6.0 and 633.3 W)less solvent usage, a shorter extraction time[57]
Borago officinalis L. leavespolyphenolsacidic water (pH 1.5) 1 to 7 kV/cm, 15–150 μs, 0.04 to 61.1 kJ/kg the same extraction without PEF treatmentincreased the TPC (1.3–6.6 times) and ORAC values (2.0–13.7 times)[58]
orange peelpolyphenolsdistilled water 5 kV/cm, 60 μs, 0.06 to 3.77 kJ/kg, pressurization at 5 bars for 30 minthe same extraction without PEF treatmentincreased the naringin content from 1 to 3.1 mg/100 g and hesperidin content from 1.3 to 4.6 mg/100 g[59]
high voltage electrical discharges extraction (HVEDE)
olive kernelphenolic compounds49% ethanol, 66 kJ/kg, pH 2.5PEF with electric field strength E = 13.3 kV/cm and UAE at 400 W and 24 kHzmore effective polyphenol extraction (255 mg GAE/L for HVEDE versus 140 and 146 mg GAE/L for UAE and PEFE, respectively)[60]
Table 2. Comparison of non-conventional extraction techniques.
Table 2. Comparison of non-conventional extraction techniques.
MethodBrief DescriptionInvestment CostEnergy EfficiencyMeritsDrawbacksReference
ultrasound-assisted extractionSample is extracted with solvent in a vessel and immersed in an ultrasonication bath.lowmediumfast energy transfer; high extraction yield; low solvent use; short extraction time (5–60 min)lack of uniformity in the process; generating damages to the ear of the operator; filtration and clean-up step required.[64,65,144,145]
microwave -assisted extractionSample is extracted with a microwave absorbing solvent in a closed/open vessel and irradiated with microwave.mediummediumquick heating for bioactive compounds extraction; high extraction yield; low solvent use; moderate extraction time (1 min–40 min)Extraction solvent must be able to absorb microwaves; filtration and cleanup step required.[64,75,145]
enzyme-assisted extractionSample and enzyme solution are loaded into a vessel and placed in a water bath thermostated at the certain temperature and time.mediummediummoderate extraction conditions; eco-friendly; selectivity due to the specificity of enzymesExpensive cost of enzymes; activity of enzymes varying with the environmental factors; filtration and cleanup step required.[88,91,94]
pressurized liquid extractionSample and solvent are heated and pressurized in a vessel with elevated temperature and pressure. After finishing the extraction, the extract is automatically into a vial.highhighhigh extraction yield; low time and solvent consumption; protection for oxygen and light sensitive compounds; no filtration required; automated systemsclean-up step required; expensive equipment required.[96,145]
supercritical fluid extractionSample is extracted with super-critical fluid in a vessel with high pressure. the analytes are collected in a small volume of solvent or onto a solid-phase trap.highhighgreen solvents (e.g., CO2) used; high extraction yield; better separation of solute from solvent; possibility to on-line combining with chromato-graphic process; reduced the thermal degradation; no cleanup or filtration required; automated systemslimited ability to dissolve polar compounds; more parameters to optimize.[64,106,107,117,145]
high hydrostatic pressure extractionSample was mixed with solvent and placed in a sterile poly-ethylene bag, which is eliminated air from the inside and placed into a pressure extractor at different pressure values.highhighwaste-free process; short time (only about 5 min); performed at room temperature without any heating processHigh investment cost and cost-intensive maintenance and service, which make industrial application difficult.[123,127,146]
pulsed electric field extractionExtraction was performed between two plate electrodes with 2–3 cm distance and the sample is placed in the treatment chamber.highhighmild (non-thermal) processing method; short time (less than 1 s)Extraction must be applied to food products that can withstand high electric fields and have low electrical conductivity.[129,146,147]
high voltage electrical discharges extractionHVED treatment was performed between the stainless steel needle and the grounded plate electrodes with 1 cm distance and the sample is placed in the treatment chamber.highhighmild (non-thermal) processing method; high extraction efficiency; short extraction time High voltage electrical discharges may generate chemical products and free reactive radicals, which can react with antioxidant compounds, thus decreasing their bioactive activity.[147]
Table 3. The antioxidant capacities of some foods and medicinal plants.
Table 3. The antioxidant capacities of some foods and medicinal plants.
CategoryVarieties Showing Strong Antioxidant ActivitiesAssessment MethodReference
antioxidant activities at chemical level
26 spicesoregano, cinnamon stick, clove, cinnamon, sageTrolox equivalent antioxidant capacity (TEAC), Folin–Ciocalteu reagent (FCR)[8]
62 fruitsChinese date, pomegranate, guava, sweetsop, persimmon, Chinese wampee and plum, grape (red)TEAC, ferric-reducing antioxidant power (FRAP), FCR[11]
24 cereal grainsblack rice, red rice, purple rice, buckwheatTEAC, FRAP, FCR[12]
49 Edible macro-fungiThelephora ganbajun, Boletus edulis, Volvariella volvacea, Boletus regius, and Suillus bovinusTEAC, FRAP, FCR[13]
56 vegetablesChinese toon Bud, loosestrife, perilla leaf, cowpea, caraway, lotus root, sweet potato leaf, soy bean (green), pepper leaf, ginseng leaf, chives, and broccoliTEAC, FRAP, FCR[14]
223 medicinal plantsAcanthopanax gracilistylus, Agrimonia pilosa, Anemarrhena asphodeloides, Caesalpina sappan, Carthamus tinctorius, Dioscorea bulbifera, Fraxinus rhynchophylla, Lonicera japonica (flower), Magnolia officinalis, Mentha haplocalyx, Paeonia lactiflora (red), Polygonum multiflorum (Stem), Polygonum multiflorum (Root), Rhodiola sacra, Salvia miltiorrhiza, Tussilago farfara, Sargentodoxa cuneataTEAC, FRAP, FCR[15]
51 edible and wild flowersRosa rugosa, Limonium sinuatum, Pelargonium hortorum, Jatropha integerrima and Osmanthus fragrans, Orostachys fimbriatu, Chaenomeles sinensis, Calliandra haematocephalaTEAC, FRAP, FCR[16]
50 fruit wastesgrape seed, hawthorn peel, longan peel, longan seed, mango peel, Chinese olive peel and sweetsop peelTEAC, FRAP, FCR[19]
antioxidant activities at cellular level
27 vegetablesbeet, broccoli, and red pepper, eggplant, Brussels sprout, cabbagecellular antioxidant activity (CAA) based on HepG2 cells[183]
25 fruitspomegranate and berries (wild blueberry, blackberry, raspberry, and blueberry)CAA based on HepG2 cells[184]
11 Thai pigmented (red and purple) and 2 nonpigmented rice varietieshawm dowk mali deang (red), hawm deang sukhothai1 (red), hawm deang (red), man pu (red), red rose (red), klam moang (purple), klam chiang mai (purple)CAA based on HL-60 cells[185]
13 food legumesblack soybean, black bean, pinto bean, lentil, green pea, yellow soybeanCAA based on human gastric adenocarcinoma AGS cells[192]
12 plantago speciesP. lanceolata, P. himalaica, P. depressa, P. cornuti, P. jehohlensisCAA based on HepG2 cells[193]
seven cultivars of AloeAloe. greenii, Aloe. arborescensCAA based on HepG2 cells[194]
five main Phyllanthus emblica L. cultivarsqingyougan, binggan and boliganCAA based on HepG2 cells[195]

Share and Cite

MDPI and ACS Style

Xu, D.-P.; Li, Y.; Meng, X.; Zhou, T.; Zhou, Y.; Zheng, J.; Zhang, J.-J.; Li, H.-B. Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources. Int. J. Mol. Sci. 2017, 18, 96. https://doi.org/10.3390/ijms18010096

AMA Style

Xu D-P, Li Y, Meng X, Zhou T, Zhou Y, Zheng J, Zhang J-J, Li H-B. Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources. International Journal of Molecular Sciences. 2017; 18(1):96. https://doi.org/10.3390/ijms18010096

Chicago/Turabian Style

Xu, Dong-**, Ya Li, **ao Meng, Tong Zhou, Yue Zhou, Jie Zheng, Jiao-Jiao Zhang, and Hua-Bin Li. 2017. "Natural Antioxidants in Foods and Medicinal Plants: Extraction, Assessment and Resources" International Journal of Molecular Sciences 18, no. 1: 96. https://doi.org/10.3390/ijms18010096

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop