Open access peer-reviewed chapter

EPR Methods Applied on Food Analysis

Written By

Chryssoula Drouza, Smaragda Spanou and Anastasios D. Keramidas

Submitted: 03 June 2018 Reviewed: 29 June 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.79844

From the Edited Volume

Topics From EPR Research

Edited by Ahmed M. Maghraby

Chapter metrics overview

2,016 Chapter Downloads

View Full Metrics

Abstract

An overview of the different methodologies developed so far for the investigation of paramagnetic species in foods is presented. Electron paramagnetic resonance spectroscopy (EPR), also known as electron spin resonance spectroscopy (ESR), is the primary technique toward the development of methods for the exploration of EPR-sensitive species, such as free radicals, reactive oxygen species (ROS), nitrogen reactive species (NRS), and C-centered radicals and metal ions. These methods aim for: (a) quantification of radical species, (b) exploration of redox chemical reaction mechanisms in foods, (c) assessment of the antioxidant capacity of food, and (d) food quality, stability, and food shelf life. For these purposes, different radical initiations and detections have been used in foods depending on both the chemistry of the target system and the kind of information required, listed in: the induction of radicals by (a) microwave, UV, or γ-radiation; (b) heating; (c) addition of metals; and (d) use of oxidants.

Keywords

  • EPR
  • free radicals
  • food
  • antioxidants
  • spin traps
  • time-dependent EPR

1. Introduction

In the last few years, the applications of the magnetic resonance techniques, particularly nuclear magnetic resonance (NMR) and electron paramagnetic resonance (EPR), in food chemistry have enormously increased [1, 2, 3, 4, 5].

EPR spectroscopy is a sensitive and versatile technique for analyzing molecules that contain unpaired electrons, such as paramagnetic metal ions and organic radicals. The formation of organic radicals in foods is an indication of food degradation occurring mainly due to oxidation reactions. Metal ions present in foods are able to catalyze oxidation of the food components by activating O2 to produce reactive oxygen species (ROS). In addition to the analysis of the paramagnetic species in foods, EPR can be used for the evaluation of the food stability and shelf-life. In order to perform such studies, acceleration of the radical production and degradation in food is needed. Several methods have been applied for the production of radicals in foods, including irradiation with microwave, UV, or γ-radiation, heating, and addition of oxidants. Stable organic radicals, such as tyrosyl and semiquinone radicals, can be detected directly by EPR. However, for the detection of transient radicals, spin traps are employed in order to be measured by EPR spectroscopy. The life of the short-lived radicals can also be extended by rapid freezing of the samples after their generation. In addition, time-resolved EPR can be used for the detection of short-lived radicals. Valuable information is acquired for the mechanisms involved in these reactions by measuring the EPR signal vs time.

The main objective of this chapter is the discussion of methods for food analysis by cw X-band EPR, including the observation of endogenous unpaired electronic spin species and the initiation and detection of free radicals in foods.

Advertisement

2. Endogenous unpaired electronic spin species in foods

2.1. Metal ions in food

Foods contain metal ions originated either from the raw starting materials or from contamination with metals from metallic containers or from contamination with metals during food processing [6, 7, 8, 9]. EPR spectroscopy is particularly sensitive in detection of FeIII, MnII, and CuII metal ions, which can be found in food materials, because of their relative long relaxation times. FeIII gives at X-band EPR a singlet at ∼160 mT, MnII a six-line hyperfine pattern due to the coupling of the unpaired electrons with 55Mn nucleus (spin I = 5/2) at 300–350 mT, while CuII gives quartet hyperfine splitting after coupling with 59Cu nucleus (spin I = 3/2) for the isotropic spectra at room temperature at 250–320 mT. The axial anisotropic EPR spectra of CuII nucleus consist of four peaks for the magnetic field aligned along the z axis and one peak for the magnetic field aligned along xy plane. One example was provided by Drew et al. who employed cw X-band EPR to explore the origin of the metal ions in Scotch whiskies [7].

The EPR spectrum of a frozen whiskey, depicted in Figure 1, shows the presence of all three metal ions.

Figure 1.

Cw X-band EPR spectra of a 2008 distillate and as-bottled aged whiskies from 1960 to 1970. After the permission of Prof. SC Drew.

The EPR spectra of MnII is of particular interest because MnII is present at almost all the foods of plant origin [10]. The signal of the frozen solutions of the symmetric [MnII(H2O)6]2+ consists of six narrow lines with additional small peaks between the six main components due to forbidden transitions. However, the EPR signal of MnII is significantly different from [MnII(H2O)6]2+ when MnII is coordinated to small ligands or large biomolecules mainly because of changes in zero field splitting (ZFS) parameters [11, 12]. These EPR data can be obtained from the simulations of the experimental spectra and they can be used for investigating the coordination environment around MnII in foods. However, foods are complicated biosystems and metal ions might interact with several molecules creating around them various environments [13] of different symmetry. Thus, the MnII EPR signal is complicated and fitting of the signal by considering one MnII species is not possible in most of the cases. In order to analyze the multicomponent EPR signals, researchers combine EPR and separation techniques and analyze the EPR signals of simpler-paramagnetic fractions [14].

Trials to fit the MnII EPR signal of two Cypriot wines using Easyspin 5.2.8 [15] (Figure 2) did not result in a perfect match with the experimental spectra revealing multiple MnII species in the wines.

Figure 2.

Experimental (black continues lines) and simulated (red dashed lines) cw X-band EPR spectra of two Cypriot wines from the grapes varieties Lefkada (L) and Maratheftiko (M) at 110 K. For the simulations were used the following parameters: (L) g = 1.999, A = 258 MHz, D = 530 MHz, and E = 192 MHz; (M) g = 1.999, A = 257 MHz, D = 564 MHz, and E = 210 MHz.

These EPR spectra features of the metal ions, which are originated from the various environments occurring for metal ions in foods, might be used for the food classification such as geographical or botanical discrimination. An example of the use of MnII X-band EPR spectroscopy for the discrimination of Cypriot wines from various grape varieties is shown in Figure 3 (unpublished results). In addition to the characteristic shape of the spectrum, the quantity of MnII in each wine can be measured from the double-integrated spectra in the presence of standard [14, 16] information that can be additionally used as a variable for the wine discrimination.

Figure 3.

Cw X-band EPR spectra of various Cypriot wines from the grape varieties Xynisteri (X), Lefkada (L), Shiraz (S), and Maratheftiko (M) at 110 K.

The MnII cw X-band EPR spectra are also useful for analyzing the degradation of the food [10, 17]. An example of the alternation of the MnII signal in the wines up to exposure to air is shown in Figure 4. After the exposure, a new signal is appeared at g = 2.000 and A ∼ 185 MHz. Such signals have been assigned to multinuclear manganese clusters of higher oxidation states than MnII as previously reported for studies in solutions of model MnII compounds after their exposure to O2 [18, 19]; therefore, similar clusters might be formed also in wines.

Figure 4.

Cw X-band EPR spectra of two fresh samples and one sample exposed to atmospheric oxygen for 1 day of the Cypriot wine from the grape variety Maratheftiko (M) at 110 K.

The presence of free ions, such as FeIII and CuII, might accelerate degradation of foods, through Fenton reactions, leading to undesirable taste, color, or food spoilage [20, 21, 22, 23, 24, 25, 26]. Sometimes the removal of excessive free ions from foods is required in order to preserve their quality [8]. Metal chelators have found to inhibit the oxidation and increase the stability of model wines [27]. On the other hand, addition of metal ions in foods emerges reactive radical species that can be detected by EPR and used further for food characterization.

2.2. Organic radicals

In addition to metallic radicals, foods might contain persistent organic radicals formed by the exposure of food in atmospheric oxygen or the food preparation processes. Metal ions might play an important catalytic role in the formation of organic radicals. For example, although X-band EPR spectrum of fresh tea leaves gives at g = 2.000 only the sextet of MnII, the ground tea from tea bags gives a sharp peak due to the stable semiquinone radical, in addition to the MnII peak (Figure 5).

Figure 5.

Cw X-band EPR spectra of ground tea from tea bags at room temperature.

An extensive EPR study of dry tea leaves from various origins has shown that except the semiquinone radicals, stable carbohydrate radical can also be detected [28]. The same study showed that the type of radical is depended on the content of flavan-3-ols in tea. The teas owned the highest content of flavan-3-ols (unfermented teas) form carbohydrate radicals, whereas fermented teas have high quantities of semiquinone radicals.

Troup et al. have investigated the organic radicals formed in roasted coffee beans and the brewed coffee solutions by EPR spectroscopy [14]. They have assigned the radicals to high-molecular-weight phenolic compounds present in the coffee brew and melanoidin compounds generated in the course of the Maillard reaction from reducing sugars and amino acids.

Phenolics are also the compounds which form radicals in red wines [29]. In addition, stable radicals were detected directly in the extracts of carrot root, celery stalk, cress shoots, cucumber, parsley, and cabbage leaf appeared upon maceration. The EPR signal is a double peak in the EPR spectrum, attributed to the monodehydroascorbyl radical formed in the aqueous solution. A wide single peak overlays the above signals in some samples and is attributed to the stressed biotic or abiotic conditions [30].

In general, fresh foods, protected from the oxidation, do not form organic radicals. However, such radicals might be induced and used for the characterization of food shelf-stability.

Advertisement

3. Induction and monitoring of radicals in foods

3.1. Methods for induction of radicals

Several methods have been used for the induction of free radicals in foods, including irradiation with UV, microwaves, or γ-radiation, heating, addition of ozone, metal ions, or other oxidants. The EPR signal of stable radicals formed in food could be monitored directly, whereas unstable radicals can be measured indirectly with the addition of spin traps.

The use of EPR spectroscopy to monitor radicals in γ-radiated foods is a common practice which is very well documented in the literature [31, 32, 33, 34, 35, 36, 37, 38, 39]. The most of the studies were focused on consumer safety due to the use of this method in some countries for food product sterilization.

Microwave irradiation also causes formation of radicals in foods which can be monitored by EPR spectroscopy [40]. X-band EPR studies of the effect of microwave radiation on rice flour and rice starch [41, 42, 43] have shown the formation of tyrosyl and semiquinone radicals, after food irradiation, localized in the starch and the protein fraction of rice flour. These radicals exist in the native rice flour; however, their intensity increases exponentially by increasing microwave power and radiation time. The authors have proposed that transition metal redox process might be associated with the formation of the radicals [42, 43]. On the other hand, the rate of radical generation in flour starch is not related to the microwave power and irradiation time but increases rapidly at about 100°C [41].

UV-irradiation is a very popular technique for the generation of radicals measured by EPR [44, 45, 46]. Foods are directly irradiated with UV-light [47, 48, 49] or after the addition of a photosensitive radical initiator in foods [50, 51]. The radicals, produced from UV-irradiation, usually are trapped by spin traps before being measured by EPR. However, there are examples of direct measurement of stable radicals formed in food. For example, UV-irradiation of grains resulted in the formation of reactive oxygen species and stable semiquinone and phenoxyl radicals [49]. In addition to the formation of organic radicals, the MnII and the FeIII EPR signals alternate, pointing to a disturbance of the biomolecules’ structures.

The thermal stability of foods, in particular, edible oils, is a property associated with the storage life of food staff explored through various spectroscopic methods and rancimat analysis [52, 53, 54, 55, 56, 57]. The thermal process of foods generates radicals that can be detected by EPR spectroscopy. An example of heating-induced radical formation is the coffee beans roasting with formed radicals to be monitored in real time [14, 58, 59]. Goodman et al. have shown that the organic radicals produced from the heating of coffee beans are dependent on the variety of the bean, but the experimental data were not enough to support an explanation. In addition, they noticed that the quantity of radicals is higher at the presence of O2, and the oxidation rate of beans is considerably higher during the cooling process [58]. The radicals produced from the heating of edible oils are trapped with radical traps such as N-tert-butyl-α-phenylnitrone (PBN). Monitoring the signal of the PBN spin adducts by EPR consists a promising method for the determination of the lipid oxidation lag phase but not suitable for the lag phase of hydroperoxides and thus oil shelf-life [60]. The formation of free radicals in edible oils is catalyzed by unsaturated lipids, and in this autoxidation mechanism, there is a direct involvement of β-carotene and chlorophyll [61]. The EPR spectra of the heated oils showed also the formation of α-tocopheryl radical, suggesting that the α-tocopheryl radical might be used as an alternative marker for studying the oxidation state of edible oils [61, 62]. The EPR spectra of edible oils heated at 180 °C in contact with metals suggested that iron and aluminum do not significantly affect the oils. On the other hand, heating the oil with copper resulted in the dissolution of large quantities of CuII in the oil promoting the decomposition of primary oxidation products, while increasing the buildup of secondary oxidation products [63].

Ozone is a nonthermal technology with promising application in food processing. It is primarily used as a disinfectant and antimicrobial agent for food safety applications and for food preservation [64, 65, 66]. However, processing of foods with ozone results in the formation of radicals that can be detected with EPR [67, 68]. The ozonation of grains was found to be safe for the consumers; however, the application of ozone directly on food products containing crushed grains, for instance, meal, might pose a threat to consumers.

The initiation of radicals with addition of metal ions or with the addition of metal ions with H2O2 (Fenton-like reagents) is also a usual strategy for the characterization of foods. The formation of radicals with the Fenton reagents is based in the reactions (1) and (2).

Fe 2 + + H 2 O 2 Fe 3 + + HO + HO E1
Fe 2 + + H 2 O 2 Fe VI O 2 + + H 2 O E2

However, in addition to Fenton reagent, other reagents [69], reacting like the Fenton reagent, such as CoII/H2O2, CuI/H2O2 [70], and K2S2O8 [71], might be used. Usually, the radicals formed from the reaction with the Fenton reagents are trapped by spin traps and monitored by various spectroscopies including EPR. This methodology has been applied on several types of foods including plant extracts [72], strawberry fruit [73], sugar and other molecules found in foods [74], edible oils [48, 75], tea [76], wines [27], etc. Investigation of the reactivity of FeII complexes with quinolinic acid as Fenton reagent has shown that Fe(II)-Quin produces more hydroxyl radicals and is more stable than Fe(II) alone [72]. In addition, metal ions being in the form of salts are insoluble in lipids; thus, in order to be used as radical initiators in lipids, they require their solubility to be increased by the addition of emulsifiers [77, 78, 79]. Recently, Drouza et al. have synthesized lipophilic metal complexes soluble in oils that initiate radicals in the presence of oxygen [3], whereas α-tocopherol is used as a marker for the investigation of the olive oils’ stability.

3.2. Addition of radicals

A common use of EPR spectroscopy is the addition of reactive organic radicals, usually DPPH, galvanoxyl radical, ABTS+•, TEMPO, TEMPOL, or Fremy’s salt for the determination of the antioxidant activity of foods [80, 81, 82, 83, 84]. The EPR signal is reduced after the addition of radicals in oil because of the reduction of the radicals from the antioxidant food components, and the antioxidant activity can be calculated from Eq. (3) or more complicate mathematical equations [85, 86, 87, 88, 89].

Inhibition activity % = A 0 A / A 0 × 100 E3

where A0 and A are the double integrals of the signal of the control and the sample after the addition of the antioxidant, respectively.

Stable radicals can also be added as probes. The EPR signal of the radical is dependent on the environment around the radical, thus structural information can be acquired. The radical probes could be organic [90, 91, 92, 93, 94] or inorganic [13]. The X-band EPR spectra of aqueous solutions containing extracts of green or black tea and CuII showed the formation of six complexes, probably of CuII with amino acids. The interactions of CuII with teas are pH dependent. At high pH, the CuII ions form complexes with polyphenols [13].

3.3. Lipophilic metal initiators

Although metal ions have been used as insoluble salts to induce free radicals in edible oil samples, a novel approach has been presented by the utilization of lipophilic metal complexes as radical initiators for the oxidation of lipids in olive oils, targeting the activation of α-tocopheryl radical naturally contained in edible oils [3].

The new metal initiators consist the VV and VIV complexes, 1 and 2 (Figure 6), containing a lipophilic tail enabling them to perfectly dissolve in the oil matrix. This has been presented as an advantage of the new method because it allows the retaining of the chemical environment neighboring the polar phenols as it is in the bulk pure oil. Thus, phenols are allowed to participate in the free radical interplay between the redox species unaffected by any phase change discontinuation as it occurred in the case of the emulsions. In this method, the evolution of the phenol scavenging activity is recorded versus time revealing information for all the time framework of the food exposure to radicals (Figures 7 and 8).

Figure 6.

Vanadium (IV/V) complexes 1 and 2.

Figure 7.

X-band EPR spectrum of virgin olive oil (0.500 g) vs time after addition of 1 (100 μL, 7.00 mM) at RT. The time period between two adjacent spectra is 6.5 min.

Figure 8.

First integral X-band EPR spectrum of virgin olive oil (0.500 g) vs time after addition of 1 (100 μL, 7.00 mM) at RT. The time period between two adjacent spectra is 6.5 min.

The particular metal ion, vanadium, was selected because it participates in redox reactions, producing radicals and stabilizing semiquinone radicals [95, 96, 97], and activate molecular dioxygen [98, 99]. Cw X-band variable temperature (VT)-EPR spectroscopy reveals strong interactions between complex 2 and phenols suggesting that such interactions in the presence of O2 might promote the initiation of the radicals.

The effect of the polar phenols naturally contained in the edible oils on the dioxygen activation and the free radical production was explored by a key experiment based on the monitoring of the intensity of the EPR α-tocopheryl signal in the presence and/or the absence of the polar phenols. The subtraction of the polar phenols resulted in (i) the reduction of maximum intensity of the EPR signal of α-tocopheryl radical and (ii) the decrease of the time needed for the occurrence of maximum intensity, tm, for the same edible oil. This new method has been applied for evaluating the age of olive oil or the storage period associated with the amounts of the polar phenols, which are decomposed by the increase of the storage time, using the abovementioned two spectral characteristics as evaluating parameters. The mechanism of the radical initiation by 1 and 2 complexes was further investigated by spin trap experiments.

3.4. Radical traps

The life time of organic free radicals is usually very short because they undergo bimolecular self-reaction. Spin trap technique has been developed since 1968 for the detection and identification of the transient free radicals. Spin traps are diamagnetic molecules exerting a particular high affinity for reactive radicals, to which reactive radicals rapidly add to form persistent spin adducts, detectable in the EPR spectroscopy. Typically, there are two types of molecules serving as spin traps, the C-nitroso compounds and the nitrones; some of them are shown in Table 1.

Table 1.

Spin traps commonly used for detection and identification of free radicals.

The first one, the C-nitroso compounds are organic nitroxides which upon reaction form the spin adduct through addition of organic part of the radical directly on the nitrogen atom [100, 101]. This proximity to the unpaired electron occupying the p* orbital of N atom of the functional group generates additional hyperfine coupling because of the presence of the neighboring magnetic nuclei of the added free radical. These hyperfine coupling parameters can provide structural information for the identification of added radical. The spin adducts of C-nitroso compounds in general have longer life times but bound less types of radicals, usually the C-centered ones, than nitrones [102]. The second type of spin traps, nitrones are organic molecules reacting with free radicals very fast, close to the diffusion-controlled limit, forming spin adducts by the bound of the added radical to the unsaturated C atom next to the N atom of the functional group [101, 102, 103]. It appears that this type of traps is widely used because they can form spin adducts with a wide range of radical species, such as peroxy (HOO), alkoperoxy (ROO), alkoxy (RO), hydroxy (HO), acyloxy radicals, as well as with other heteroatom-centered radical, including halogen atoms. The prime drawback for this type of traps is the poor information provided by their EPR spectra: the unpaired electron gives hyperfine coupling in the very best cases only from nitrogen nuclei of the function group and the β-proton, but not from the added radical. Thus, identification of the free radical goes through comparison of the under examination EPR spectra with undoubtfully characterized spectra obtained from the spin adducts of the prototype radicals.

An example of the use of DMPO for the detection of the alkoperoxyl and the alkoxyl lipid radicals is shown in Figure 9. The spectrum was acquired 5 min after the addition of DMPO, and the vanadium complex 1 in olive oil. Deconvolution of the spectra fits to the alkoperoxyl lipid radical adduct of DMPO (DMPO-OOR) (AN = 1.37 and AH = 1.06 mT) in 33%, and the alkoxyl lipid radical adduct of DMPO (DMPO-OR) of (AN = 1.31, A = 0.65, and A = 0.17 mT) in 77%, and a minor unknown carbon adduct of DMPO (DMPO-CRR′R″).

Figure 9.

(A) X-band EPR spectra of a solution of 200 μL 1 (7.0 mM, in CH2Cl2) 0.5 g pomace olive oil and 100 μL DMPO (30.0 mM DMPO in CH3OH) at 5 min, (B, C) simulated spectra of the two components of the experimental spectra (AN = 1.37 and AH = 1.06 mT (DMPO-OOR) and with AN = 1.31, A = 0.65, and A = 0.17 mT (DMPO-OR)).

Advertisement

4. Conclusions

In this chapter, we have reviewed the main cw X-band EPR methodologies used for the study of foods, by observing endogenous unpaired electronic spin species and by the initiation and detection of radicals in foods. The use of EPR for analysis of foods is growing up rapidly. New methodologies in initiation and detection of radicals have resulted in the better understanding of the mechanisms involved in food oxidation processes. The high sensitivity and versatility of EPR makes this technique a valuable tool in food science, and further applications are expected to emerge in the future.

The cw EPR methods used for the characterization of foods are based on the recording of endogenous metal ion or organic radical preexisting in food or the initiation of radicals that can be detected directly or by the addition of radical traps. This chapter is an overview of these methods focusing to the research of the last 15 years.

Advertisement

Acknowledgments

Supported by Research Promotional Foundation of Cyprus and the European Structural Funds ΑΝΑΒΑΘΜΙΣΗ/ΠΑΓΙΟ/0308/32.

Advertisement

Notes/Thanks/Other declarations

The cw X-band EPR spectra in this review were acquired on an ELEXSYS E500 Bruker spectrometer at resonance frequency ∼9.5 GHz and modulation frequency 100 MHz. Figures were produced by the software MultiSpecEPR (the software has been developed by Prof. AD Keramidas).

References

  1. 1. Gómez-Caravaca AM, Maggio RM, Cerretani L. Chemometric applications to assess quality and critical parameters of virgin and extra-virgin olive oil. A review. Analytica Chimica Acta. 2016;913:1-21
  2. 2. Siddiqui AJ, Musharraf SG, Choudhary MI, Rahman AU. Application of analytical methods in authentication and adulteration of honey. Food Chemistry. 2017;217:687-698
  3. 3. Drouza C, Dieronitou A, Hadjiadamou I, Stylianou M. Investigation of phenols activity in early stage oxidation of edible oils by electron paramagnetic resonance and 19F NMR spectroscopies using novel lipid vanadium complexes as radical initiators. Journal of Agricultural and Food Chemistry. 2017;65:4942-4951
  4. 4. Vlasiou M, Drouza C. 19F NMR for the speciation and quantification of the OH-molecules in complex matrices. Analytical Methods. 2015;7:3680-3684
  5. 5. Lund A, Shiotani M. Applications of EPR in Radiation Research. Spinger. Switzerland. 2014
  6. 6. Capece A, Romaniello R, Scrano L, Siesto G, Romano P. Yeast starter as a biotechnological tool for reducing copper content in wine. Frontiers in Microbiology. 2018;8:2632
  7. 7. Drew SC, Robertsa B, Troupb GJ. In Scotch whisky, from where are the Fe3+ and Cu2+ ions sourced? In: Proceedings—37th Annual Condensed Matter and Materials Meeting; Wagga Wagga, NSW, Australia; 2013
  8. 8. Carreon-Alvarez A, Herrera-Gonzalez A, Casillas N, Prado-Ramirez R, Estarron-Espinosa M, Soto V, de la Cruz W, Barcena-Soto M, Gomez-Salazar S. Cu (II) removal from tequila using an ion-exchange resin. Food Chemistry. 2011;127:1503-1509
  9. 9. Rousseva M, Kontoudakis N, Schmidtke LM, Scollary GR, Clark AC. Impact of wine production on the fractionation of copper and iron in Chardonnay wine: Implications for oxygen consumption. Food Chemistry. 2016;203:440-447
  10. 10. Morsy MA, Khaled MM. Novel EPR characterization of the antioxidant activity of tea leaves. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy. 2002;58:1271-1277
  11. 11. Hunsicker-Wang L, Vogt M, Derose VJ. EPR methods to study specific metal-ion binding sites in RNA. Methods in Enzymology. 2009;468:335-367
  12. 12. Morrissey SR, Horton TE, DeRose VJ. Mn2+ sites in the hammerhead ribozyme investigated by EPR and continuous-wave Q-band ENDOR spectroscopies. Journal of the American Chemical Society. 2000;122:3473-3481
  13. 13. Goodman BA, Severino JF, Pirker KF. Reactions of green and black teas with Cu(ii). Food & Function. 2012;3:399-409
  14. 14. Troup GJ, Navarini L, Liverani FS, Drew SC. Stable radical content and anti-radical activity of roasted Arabica coffee: From in-tact bean to coffee brew. PLoS One. 2015;10:e0122834
  15. 15. Stoll S, Schweiger A. EasySpin, a comprehensive software package for spectral simulation and analysis in EPR. Journal of Magnetic Resonance. 2006;178:42-55
  16. 16. Klencsár Z, Köntös Z. EPR analysis of Fe3+ and Mn2+ complexation sites in fulvic acid extracted from lignite. The Journal of Physical Chemistry. A. 2018;122:3190-3203
  17. 17. Morsy MA. Teas: Direct test on quality and antioxidant activity using electron paramagnetic resonance spectroscopy. Spectroscopy. 2002;16:371-378
  18. 18. Ishigure S, Mitsui T, Ito S, Kondo Y, Kawabe S, Kondo M, Dewa T, Mino H, Itoh S, Nango M. Peroxide decoloration of CI acid orange 7 catalyzed by manganese chlorophyll derivatives at the surfaces of micelles and lipid bilayers. Langmuir. 2010;26:7774-7782
  19. 19. Bryliakov KP, Kholdeeva OA, Vanina MP, Talsi EP. Role of MnIV species in Mn(salen) catalyzed enantioselective aerobic epoxidations of alkenes: An EPR study. Journal of Molecular Catalysis A: Chemical. 2002;178:47-53
  20. 20. Kaneda H, Kano Y, Koshino S, Ohyanishiguchi H. Behavior and role of iron ions in beer deterioration. Journal of Agricultural and Food Chemistry. 1992;40:2102-2107
  21. 21. Jaklová Dytrtová J, Straka M, Bělonožníková K, Jakl M, Ryšlavá H. Does resveratrol retain its antioxidative properties in wine? Redox behaviour of resveratrol in the presence of Cu(II) and tebuconazole. Food Chemistry. 2018;262:221-225
  22. 22. Monforte AR, Martins SIFS, Silva Ferreira AC. Strecker aldehyde formation in wine: New insights into the role of gallic acid, glucose, and metals in phenylacetaldehyde formation. Journal of Agricultural and Food Chemistry. 2018;66:2459-2466
  23. 23. Pazos M, Da Rocha AP, Roepstorff P, Rogowska-Wrzesinska A. Fish proteins as targets of ferrous-catalyzed oxidation: Identification of protein carbonyls by fluorescent labeling on two-dimensional gels and MALDI-TOF/TOF mass spectrometry. Journal of Agricultural and Food Chemistry. 2011;59:7962-7977
  24. 24. Guo A, Kontoudakis N, Scollary GR, Clark AC. Production and isomeric distribution of xanthylium cation pigments and their precursors in wine-like conditions: Impact of Cu(II), Fe(II), Fe(III), Mn(II), Zn(II), and Al(III). Journal of Agricultural and Food Chemistry. 2017;65:2414-2425
  25. 25. Clark AC, Wilkes EN, Scollary GR. Chemistry of copper in white wine: A review. Australian Journal of Grape and Wine Research. 2015;21:339-350
  26. 26. Danilewicz JC. Chemistry of manganese and interaction with iron and copper in wine. American Journal of Enology and Viticulture. 2016;67:377-384
  27. 27. Kreitman GY, Cantu A, Waterhouse AL, Elias RJ. Effect of metal chelators on the oxidative stability of model wine. Journal of Agricultural and Food Chemistry. 2013;61:9480-9487
  28. 28. Socha R, Baczkowicz M, Fortuna T, Kura A, Łabanowska M, Kurdziel M. Determination of free radicals and flavan-3-ols content in fermented and unfermented teas and properties of their infusions. European Food Research and Technology. 2013;237:167-177
  29. 29. Troup GJ, Hutton DR, Hewitt DG, Hunter CR. Free radicals in red wine, but not in white? Free Radical Research. 1994;20:63-68
  30. 30. Goodman BA, Glidewell SM, Arbuckle CM, Bernardin S, Cook TR, Hillman JR. An EPR study of free radical generation during maceration of uncooked vegetables. Journal of the Science of Food and Agriculture. 2002;82:1208-1215
  31. 31. Alberti A, Chiaravalle E, Corda U, Fuochi P, Macciantelli D, Mangiacotti M, Marchesani G, Plescia E. Treating meats with ionising radiations. An EPR approach to the reconstruction of the administered dose and its reliability. Applied Radiation and Isotopes. 2011;69:112-117
  32. 32. Alberti A, Chiaravalle E, Fuochi P, Macciantelli D, Mangiacotti M, Marchesani G, Plescia E. Irradiated bivalve mollusks: Use of EPR spectroscopy for identification and dosimetry. Radiation Physics and Chemistry. 2011;80:1363-1370
  33. 33. Aleksieva KI, Yordanov ND. Various approaches in EPR identification of gamma-irradiated plant foodstuffs: A review. Food Research International. 2018;105:1019-1028
  34. 34. Bercu V, Negut CD, Duliu OG. Irradiation free radicals in freshwater crayfish Astacus leptodactylus Esch investigated by EPR spectroscopy. Radiation Physics and Chemistry. 2017;133:45-51
  35. 35. Beshir WB. Identification and dose assessment of irradiated cardamom and cloves by EPR spectrometry. Radiation Physics and Chemistry. 2014;96:190-194
  36. 36. Bhatti IA, Akram K, Kwon JH. An investigation into gamma-ray treatment of shellfish using electron paramagnetic resonance spectroscopy. Journal of the Science of Food and Agriculture. 2012;92:759-763
  37. 37. De Oliveira MRR, Mandarino JMG, Del Mastro NL. Radiation-induced electron paramagnetic resonance signal and soybean isoflavones content. Radiation Physics and Chemistry. 2012;81:1516-1519
  38. 38. Della Monaca S, Fattibene P, Boniglia C, Gargiulo R, Bortolin E. Identification of irradiated oysters by EPR measurements on shells. Radiation Measurements. 2011;46:816-821
  39. 39. Duliu OG, Bercu V. ESR investigation of the free radicals in irradiated foods. In: Electron Spin Resonance in Food Science. Academic Press. Massachusetts. 2017. pp. 17-32
  40. 40. Yakhin RG, Samigullina NA, Yagund EM, Yakhin RR. Investigation of the influence of microwave radiation on the properties of vegetable food products by methods of EPR and IR spectroscopy. Khimiya Rastitel’nogo Syr’ya. 2017:151-157
  41. 41. Fan D, Liu Y, Hu B, Lin L, Huang L, Wang L, Zhao J, Zhang H, Chen W. Influence of microwave parameters and water activity on radical generation in rice starch. Food Chemistry. 2016;196:34-41
  42. 42. Fan DM, Lin LF, Wang LY, Huang LL, Hu B, Gu XH, Zhao JX, Zhang H. The influence of metal ions on the dielectric enhancement and radical generation of rice starch during microwave processing. International Journal of Biological Macromolecules. 2017;94:266-270
  43. 43. Lin L, Huang L, Fan D, Hu B, Gao Y, Lian H, Zhao J, Zhang H, Chen W. Effects of the components in rice flour on thermal radical generation under microwave irradiation. International Journal of Biological Macromolecules. 2016;93:1226-1230
  44. 44. Walton JC. Functionalised oximes: Emergent precursors for carbon-, nitrogen- and oxygen-centred radicals. Molecules. 2016;21:21010063
  45. 45. Hricovíni M, Dvoranová D, Barbieriková Z, Jantová S, Bella M, Šoral M, Brezová V. 6-Nitroquinolones in dimethylsulfoxide: Spectroscopic characterization and photoactivation of molecular oxygen. Journal of Photochemistry and Photobiology A: Chemistry. 2017;332:112-121
  46. 46. Hayes EC, Jian Y, Li L, Stoll S. EPR study of UV-irradiated thymidine microcrystals supports radical intermediates in spore photoproduct formation. The Journal of Physical Chemistry. B. 2016;120:10923-10931
  47. 47. Ottaviani MF, Spallaci M, Cangiotti M, Bacchiocca M, Ninfali P. Electron paramagnetic resonance investigations of free radicals in extra virgin olive oils. Journal of Agricultural centred radicals. Molecules. 2016;21:21010063
  48. 48. Skoutas D, Haralabopoulos D, Avramiotis S, Sotiroudis TG, Xenakis A. Virgin olive oil: Free radical production studied with spin-trapping electron paramagnetic resonance spectroscopy. Journal of the American Oil Chemists’ Society. 2001;78:1121-1125
  49. 49. Kurdziel M, Filek M, Łabanowska M. The impact of short-term UV irradiation on grains of sensitive and tolerant cereal genotypes studied by EPR. Journal of the Science of Food and Agriculture. 2018;98:2607-2616
  50. 50. Kameya H, Watanabe J, Takano-Ishikawa Y, Todoriki S. Comparison of scavenging capacities of vegetables by ORAC and EPR. Food Chemistry. 2014;145:866-873
  51. 51. Kohri S, Fujii H. 2,2′-Azobis(isobutyronitrile)-derived alkylperoxyl radical scavenging activity assay of hydrophilic antioxidants by employing EPR spin trap method. Journal of Clinical Biochemistry and Nutrition. 2013;53:134-138
  52. 52. Alvarenga BR, Xavier FAN, Soares FLF, Carneiro RL. Thermal stability assessment of vegetable oils by Raman spectroscopy and chemometrics. Food Analytical Methods. 2018;11:1969-1976
  53. 53. Forero-Doria O, García MF, Vergara CE, Guzman L. Thermal analysis and antioxidant activity of oil extracted from pulp of ripe avocados. Journal of Thermal Analysis and Calorimetry. 2017;130:959-966
  54. 54. Redondo-Cuevas L, Castellano G, Raikos V. Natural antioxidants from herbs and spices improve the oxidative stability and frying performance of vegetable oils. International Journal of Food Science and Technology. 2017;52:2422-2428
  55. 55. Yalcin S, Schreiner M. Stabilities of tocopherols and phenolic compounds in virgin olive oil during thermal oxidation. Journal of Food Science and Technology. 2018;55:244-251
  56. 56. Zawadzki A, Alloo C, Grossi AB, do Nascimento ESP, Almeida LC, Bogusz Junior S, Skibsted LH, Cardoso DR. Effect of hop β-acids as dietary supplement for broiler chickens on meat composition and redox stability. Food Research International. 2018;105:210-220
  57. 57. Zawada K, Kozłowska M, Zbikowska A. Oxidative stability of the lipid fraction in cookies—The EPR study. Nukleonika. 2015;60:469-473
  58. 58. Goodman BA, Pascual EC, Yeretzian C. Real time monitoring of free radical processes during the roasting of coffee beans using electron paramagnetic resonance spectroscopy. Food Chemistry. 2011;125:248-254
  59. 59. Yeretzian C, Pascual EC, Goodman BA. Effect of roasting conditions and grinding on free radical contents of coffee beans stored in air. Food Chemistry. 2012;131:811-816
  60. 60. Cui L, Lahti PM, Decker EA. Evaluating electron paramagnetic resonance (EPR) to measure lipid oxidation lag phase for shelf-life determination of oils. Journal of the American Oil Chemists’ Society. 2017;94:89-97
  61. 61. Ricca M, Foderà V, Vetri V, Buscarino G, Montalbano M, Leone M. Oxidation processes in Sicilian olive oils investigated by a combination of optical and EPR spectroscopy. Journal of Food Science. 2012;77:C1084-C1089
  62. 62. Cheikhousman R, Zude M, Bouveresse DJR, Léger CL, Rutledge DN, Birlouez-Aragon I. Fluorescence spectroscopy for monitoring deterioration of extra virgin olive oil during heating. Analytical and Bioanalytical Chemistry. 2005;382:1438-1443
  63. 63. Silvagni A, Franco L, Bagno A, Rastrelli F. Thermo-induced lipid oxidation of a culinary oil: The effect of materials used in common food processing on the evolution of oxidised species. Food Chemistry. 2012;133:754-759
  64. 64. Fundo JF, Miller FA, Tremarin A, Garcia E, Brandão TRS, Silva CLM. Quality assessment of cantaloupe melon juice under ozone processing. Innovative Food Science & Emerging Technologies. 2018;47:461-466
  65. 65. Silveira AC, Oyarzún D, Escalona V. Oxidative enzymes and functional quality of minimally processed grape berries sanitised with ozonated water. International Journal of Food Science and Technology. 2018;53:1371-1380
  66. 66. Zhu F. Effect of ozone treatment on the quality of grain products. Food Chemistry. 2018;264:358-366
  67. 67. Łabanowska M, Kurdziel M, Filek M. Changes of paramagnetic species in cereal grains upon short-term ozone action as a marker of oxidative stress tolerance. Journal of Plant Physiology. 2016;190:54-66
  68. 68. Reichenauer TG, Goodman BA. Free radicals in wheat flour change during storage in air and are influenced by the presence of ozone during the growing season. Free Radical Research. 2003;37:523-528
  69. 69. Mellaerts R, Delvaux J, Levêque P, Wuyts B, G. Van Den Mooter, Augustijns P, Gallez B, Hermans I, Martens J. Screening protocol for identifying inorganic oxides with anti-oxidant and pro-oxidant activity for biomedical, environmental and food preservation applications. RSC Advances. 2013;3:900-909
  70. 70. Moore J, Yin J-J, Yu L. Novel fluorometric assay for hydroxyl radical scavenging capacity (HOSC) estimation. Journal of Agricultural and Food Chemistry. 2006;54:617-626
  71. 71. Staško A, Polovka M, Brezová V, Biskupič S, Malı́k F. Tokay wines as scavengers of free radicals (an EPR study). Food Chemistry. 2006;96:185-196
  72. 72. Fadda A, Barberis A, Sanna D. Influence of pH, buffers and role of quinolinic acid, a novel iron chelating agent, in the determination of hydroxyl radical scavenging activity of plant extracts by electron paramagnetic resonance (EPR). Food Chemistry. 2018;240:174-182
  73. 73. Dragišić Maksimović J, Poledica M, Mutavdžić D, Mojović M, Radivojević D, Milivojević J. Variation in nutritional quality and chemical composition of fresh strawberry fruit: Combined effect of cultivar and storage. Plant Foods for Human Nutrition. 2015;70:77-84
  74. 74. Pejin B, Savic AG, Petkovic M, Radotic K, Mojovic M. In vitro anti-hydroxyl radical activity of the fructooligosaccharides 1-kestose and nystose using spectroscopic and computational approaches. International Journal of Food Science and Technology. 2014;49:1500-1505
  75. 75. Valavanidis A, Nisiotou C, Papageorgiou Y, Kremli I, Satravelas N, Zinieris N, Zygalaki H. Comparison of the radical scavenging potential of polar and lipidic fractions of olive oil and other vegetable oils under normal conditions and after thermal treatment. Journal of Agricultural and Food Chemistry. 2004;52:2358-2365
  76. 76. Azman NAM, Peiró S, Fajarí L, Julià L, Almajano MP. Radical scavenging of white tea and its flavonoid constituents by electron paramagnetic resonance (EPR) spectroscopy. Journal of Agricultural and Food Chemistry. 2014;62:5743-5748
  77. 77. Atanassova D, Kefalas P, Psillakis E. Measuring the antioxidant activity of olive oil mill wastewater using chemiluminescence. Environment International. 2005;31:275-280
  78. 78. Bezzi S, Loupassaki S, Petrakis C, Kefalas P, Calokerinos A. Evaluation of peroxide value of olive oil and antioxidant activity by luminol chemiluminescence. Talanta. 2008;77:642-646
  79. 79. Tsiaka T, Christodouleas DC, Calokerinos AC. Development of a chemiluminescent method for the evaluation of total hydroperoxide content of edible oils. Food Research International. 2013;54:2069-2074
  80. 80. Kostecka-Gugała A, Ledwozyw-Smoleń I, Augustynowicz J, Wyzgolik G, Kruczek M, Kaszycki P. Antioxidant properties of fruits of raspberry and blackberry grown in Central Europe. Open Chemistry. 2015;13:1313-1325
  81. 81. Papadimitriou V, Maridakis GA, Sotiroudis TG, Xenakis A. Antioxidant activity of polar extracts from olive oil and olive mill wastewaters: An EPR and photometric study. European Journal of Lipid Science and Technology. 2005;107:512-520
  82. 82. Papadimitriou V, Sotiroudis TG, Xenakis A, Sofikiti N, Stavyiannoudaki V, Chaniotakis NA. Oxidative stability and radical scavenging activity of extra virgin olive oils: An electron paramagnetic resonance spectroscopy study. Analytica Chimica Acta. 2006;573–574:453-458
  83. 83. Stavikova L, Polovka M, Hohnová B, Karásek P, Roth M. Antioxidant activity of grape skin aqueous extracts from pressurized hot water extraction combined with electron paramagnetic resonance spectroscopy. Talanta. 2011;85:2233-2240
  84. 84. Osorio C, Carriazo JG, Almanza O. Antioxidant activity of corozo (Bactris guineensis) fruit by electron paramagnetic resonance (EPR) spectroscopy. European Food Research and Technology. 2011;233:103-108
  85. 85. Polak J, Bartoszek M, Chorążewski M. Antioxidant capacity: Experimental determination by EPR spectroscopy and mathematical modeling. Journal of Agricultural and Food Chemistry. 2015;63:6319-6324
  86. 86. Polak J, Bartoszek M, Stanimirova I. A study of the antioxidant properties of beers using electron paramagnetic resonance. Food Chemistry. 2013;141:3042-3049
  87. 87. Košťálova Z, Hromádková Z, Ebringerová A, Polovka M, Michaelsen TE, Paulsen BS. Polysaccharides from the Styrian oil-pumpkin with antioxidant and complement-fixing activity. Industrial Crops and Products. 2013;41:127-133
  88. 88. Bartoszek M, Polak J, Chorążewski M. Comparison of antioxidant capacities of different types of tea using the spectroscopy methods and semi-empirical mathematical model. European Food Research and Technology. 2018;244:595-601
  89. 89. Polak J, Bartoszek M. The study of antioxidant capacity of varieties of Nalewka, a traditional Polish fruit liqueur, using EPR, NMR and UV-vis spectroscopy. Journal of Food Composition and Analysis. 2015;40:114-119
  90. 90. Aliaga C, López de Arbina A, Rezende MC. “Cut-off” effect of antioxidants and/or probes of variable lipophilicity in microheterogeneous media. Food Chemistry. 2016;206:119-123
  91. 91. Balanč BD, Ota A, Djordjević VB, Šentjurc M, Nedović VA, Bugarski BM, Ulrih NP. Resveratrol-loaded liposomes: Interaction of resveratrol with phospholipids. European Journal of Lipid Science and Technology. 2015;117:1615-1626
  92. 92. Chatzidaki MD, Arik N, Monteil J, Papadimitriou V, Leal-Calderon F, Xenakis A. Microemulsion versus emulsion as effective carrier of hydroxytyrosol. Colloids and Surfaces B: Biointerfaces. 2016;137:146-151
  93. 93. Chatzidaki MD, Mitsou E, Yaghmur A, Xenakis A, Papadimitriou V. Formulation and characterization of food-grade microemulsions as carriers of natural phenolic antioxidants. Colloids and Surfaces A: Physicochemical and Engineering Aspects. 2015;483:130-136
  94. 94. Rübe A, Klein S, Mäder K. Monitoring of in vitro fat digestion by electron paramagnetic resonance spectroscopy. Pharmaceutical Research. 2006;23:2024-2029
  95. 95. Drouza C, Keramidas AD. Solid state and aqueous solution characterization of rectangular tetranuclear VIV/V-p-semiquinonate/hydroquinonate complexes exhibiting a proton induced electron transfer. Inorganic Chemistry. 2008;47:7211-7224
  96. 96. Drouza C, Vlasiou M, Keramidas AD. Vanadium(iv/v)-p-dioxolene temperature induced electron transfer associated with ligation/deligation of solvent molecules. Dalton Transactions. 2013;42:11831-11840
  97. 97. Kundu S, Maity S, Weyhermüller T, Ghosh P. Oxidovanadium catechol complexes: Radical versus non-radical states and redox series. Inorganic Chemistry. 2013;52:7417-7430
  98. 98. Stylianou M, Drouza C, Giapintzakis J, Athanasopoulos GI, Keramidas AD. Aerial oxidation of a VIV-iminopyridine hydroquinonate complex: A trap for the VIV-semiquinonate radical intermediate. Inorganic Chemistry. 2015;54:7218-7229
  99. 99. Adao P, Maurya MR, Kumar U, Avecilla F, Henriques RT, Kusnetsov ML, Pessoa CJ, Correia I. Vanadium-salen and -salan complexes: Characterization and application in oxygen transfer reactions. Pure and Applied Chemistry. 2009;81:1279-1296
  100. 100. McCormick ML, Gaut JP, Lin T-S, Britigan BE, Buettner GR, Heinecke JW. Electron paramagnetic resonance detection of free tyrosyl radical generated by myeloperoxidase, lactoperoxidase, and horseradish peroxidase. The Journal of Biological Chemistry. 1998;273:32030-32037
  101. 101. Hawkins CL, Davies MJ. Direct detection and identification of radicals generated during the hydroxyl radical-induced degradation of hyaluronic acid and related materials. Free Radical Biology & Medicine. 1996;21:275-290
  102. 102. Perkins MJ. Spin trapping. Advances in Physical Organic Chemistry. 1980;17:1-64
  103. 103. Venkataraman S, Schafer FQ, Buettner GR. Detection of lipid radicals using EPR. Antioxidants & Redox Signaling. 2004;6:631-638

Written By

Chryssoula Drouza, Smaragda Spanou and Anastasios D. Keramidas

Submitted: 03 June 2018 Reviewed: 29 June 2018 Published: 05 November 2018