Next Article in Journal
A New Human Cancer Cell Proliferation Inhibition Sesquiterpene, Dryofraterpene A, from Medicinal Plant Dryopteris fragrans (L.) Schott
Next Article in Special Issue
Biologic Stress, Oxidative Stress, and Resistance to Drugs: What Is Hidden Behind
Previous Article in Journal
Protective Effect of Strawberry Extract against Inflammatory Stress Induced in Human Dermal Fibroblasts
Previous Article in Special Issue
Novel Amino-Pyridine Functionalized Chitosan Quaternary Ammonium Derivatives: Design, Synthesis, and Antioxidant Activity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Detection of Reactive Oxygen and Nitrogen Species by Electron Paramagnetic Resonance (EPR) Technique

1
Department of Pharmaceutical Chemistry, Faculty of Pharmacy, Ankara University, Tandogan, Ankara 06100, Turkey
2
Department of Pharmaceutical Toxicology, Faculty of Pharmacy, Ege University, 35100 Izmir, Turkey
3
Department of Physiology and Pharmacology “Vittorio Erspamer”, Sapienza University of Rome, P. le Aldo Moro 5, 00185 Rome, Italy
*
Author to whom correspondence should be addressed.
Molecules 2017, 22(1), 181; https://doi.org/10.3390/molecules22010181
Submission received: 29 November 2016 / Revised: 5 January 2017 / Accepted: 17 January 2017 / Published: 21 January 2017
(This article belongs to the Special Issue Chemistry and Pharmacology of Modulators of Oxidative Stress)

Abstract

:
During the last decade there has been growing interest in physical-chemical oxidation processes and the behavior of free radicals in living systems. Radicals are known as intermediate species in a variety of biochemical reactions. Numerous techniques, assays and biomarkers have been used to measure reactive oxygen and nitrogen species (ROS and RNS), and to examine oxidative stress. However, many of these assays are not entirely satisfactory or are used inappropriately. The purpose of this chapter is to review current EPR (Electron Paramagnetic Resonance) spectroscopy methods for measuring ROS, RNS, and their secondary products, and to discuss the strengths and limitations of specific methodological approaches.

1. Introduction

Reactive oxygen species (ROS) and reactive nitrogen species (RNS) are produced in almost all eukaryotic cells. These molecules are responsible regulating several physiological processes such as proliferation, migration, differentiation, and metabolism [1]. Excess amounts of ROS and RNS may react with vital molecules including lipids, proteins, and nucleic acids, thereby altering structural and functional properties of target molecules and leading to extensive tissue dysfunction and injury [2].
EPR (Electron Paramagnetic Resonance) spectroscopy is a very useful method for the direct detection of free radicals at concentrations as low as 1 µM. For short-lived ROS, the spin-trapping technique involves the addition of radicals to nitrone spin traps to form a spin adduct which has a relatively longer half-life to allow its detection using an EPR spectrometer. EPR enables the direct detection of free radicals, but other assays can be informative if used with proper controls [1]. EPR spin trapping has become an essential tool for the detection of ROS in biological systems. In this review the main benefits are described by the addition of some experimental considerations and applications to biological systems to measure oxidative stress (OS).

2. Free Radicals and Oxidative Stress

A free radical contains an unpaired electron (s) in its outer orbit. They are extremely reactive and easily take part in chemical reactions with vital cell components in the body. These reactions occur through a chain of oxidative reactions to cause tissue injury. ROS are the main cause of OS and are responsible for causing damage to proteins, lipids and DNA [3,4,5].
Although an excess amount of ROS may cause serious damage to cells, these reactive species play some important roles in a number of physiological processes. In mammalians, O2 is known as the most commonly produced ROS [2]. The production of ROS by mammalian mitochondria is important since during its progress oxygen accepts one electron at a time, which leads to the generation of superoxide (O2). The reduction of oxygen to form water occurs during the electron transport system of cellular respiration to lead to the generation of hydrogen peroxide (H2O2), and hydroxyl radicals (OH). Although hydrogen peroxide by itself is unreactive, it oxidizes Fe(II) to Fe(III) to generate hydroxyl radicals which is called the Fenton reaction [6]. The hydroxyl radicals generated by the Fenton reaction are extremely reactive and short-lived.
Both ROS and RNS are highly reactive, and as a result, they have short half-lives in biological environments. Therefore, these species are difficult to measure directly, and even indirect estimates of their abundance and reactivity are challenging. ROS/RNS are known to play a dual role in biological systems, since they can be either harmful or beneficial to living systems. The harmful effects of ROS are eliminated by the antioxidant achievement of non-enzymatic antioxidants in addition to antioxidant enzymes. Although many approaches have been adopted to bypass these limitations, and a wide range of methods have been developed to quantify ROS, RNS, and their secondary products, there are many considerations in the choice of assay and its application to a particular system.
OS occurs by an imbalance between the production of ROS and a biological system’s ability to detoxify the reactive intermediates. OS may cause vital damage in cells which is related with the development of cancer, neurodegenerative diseases such as Alzheimer’s and Parkinson’s diseases, atherosclerosis, diabetes mellitus and so on [7,8]. OS also has been increasingly recognized as a contributing factor in various forms of pathophysiology related with aging.
ROS have a significant role in a variety of cellular processes [9]. They are continuously produced during photosynthesis and respiration, whereas redox homeostasis in the cell is controlled by defensive mechanisms. Measuring OS in the cell requires delicate assays for ROS detection. Direct detection of ROS and RNS is quite difficult or sometimes impossible in solution at room temperature due to their very short half-life. Additionally, ROS are usually generated in subcellular compartments, which require detection procedures directed for specific localization. EPR allows researchers to detect ROS directly and it can also be used to monitor changes in the chemical forms of the oxidizable transition metal ions implicated in ROS generation [10]. EPR spectroscopy stands out from other methods because of its unique ability to detect either short- or long-lived radicals with high specificity and sensitivity. Correspondingly the technique of in vivo EPR spectroscopy can provide useful and even unique information pertinent to the study of oxygen/nitrogen radicals and related processes. Due to the low sensitivity of EPR, it is particularly challenging to measure ROS directly in vivo. The spin-trapping technique is used to overcome this problem. Using the spin-trapping method, ROS are allowed to react with specially selected trap molecules to produce less reactive and more stable species that can be readily detected by EPR [11].

3. Electron Paramagnetic Resonance (EPR)

EPR spectrometers are available in different sizes and models, dependent on the frequency and field in which they function. A microwave radiation source is needed for any EPR spectrometer. Microwave frequencies are produced at particular small frequency gaps called bands. The optimum sensitivity of EPR is the 8–12 GHz range which is called the X-band, a mid-range frequency. Numerous detection techniques exist for EPR spectroscopy. The most preferred EPR spectrometers are reflection spectrometers which measure the amount of radiation that is reflected back out of the resonator. A resonant cavity is the most common form of resonator used in EPR spectroscopy [12,13]. It is designed with dimensions corresponding to the specific wavelength of microwave radiation that is to be used. EPR experimentation needs an electromagnet that is used to organize the strength of the applied magnetic field. Furthermore, there are optional components which are commonly used in EPR experiments such as cooling devices to cool samples down to ultra-cold temperatures and radio frequency amplifiers to excite transitions of the nuclei.
The continuous wave EPR (CW) method allows a sample to continuously illuminate with microwave radiation at a fixed frequency and detects changes in the microwave absorption. CW EPR is the most basic experiment and is widely used [14,15].
Pulse EPR varies from CW in that microwaves are applied to the sample as a series of nanosecond-long pulses. The magnetic field strength is kept constant, and the sample is pulsed with microwave energy. This technique involves the alignment of the net magnetization vector of the electron spins in a constant magnetic field [16].
Electron nuclear double resonance spectroscopy (ENDOR) is a technique that includes pulses in the radiofrequency region to the pulse patterns utilized to produce spin echoes. ENDOR spectroscopy employs both radiofrequency radiation, as in nuclear magnetic resonance (NMR) spectroscopy and microwave radiation, and as in electron spin resonance (ESR) or EPR, to study interactions between unpaired (paramagnetic) electrons and nuclei with magnetic moments [17].

4. Electron Paramagnetic Resonance (EPR) Spectroscopy Technique

EPR spectroscopy is a method for studying materials with unpaired electrons. The basic concepts of EPR are analogous to those of nuclear magnetic resonance (NMR), but it is electron spins that are excited instead of the spins of atomic nuclei. Instead of measuring the nuclear transitions in a sample, EPR detects the transitions of unpaired electrons in an applied magnetic field.
EPR spectroscopy applications cover a wide range of areas from structural biology to quantum physics. EPR spin trapping is a technique that was first discovered in 1945 [18,19] and developed in the late 1960s. In this technique, the spin-trapping reaction occurs with the addition of the free radical to the double bond of a diamagnetic “spin trap”. After the formation of a more stable free radical, this can be examined with EPR. This technique detects species that have unpaired electrons such as free radicals and many transition metal ions. It is known that free radicals have a very short life but still play crucial roles in many processes in biological systems such as oxidation, catalysis, and polymerization reactions.
EPR spectroscopy has been used to investigate nitric oxide chemistry and biology. Spin trapping is a well-known analytical technique widely used in chemistry and biology for the detection and identification of short-lived free radicals through the use of EPR. The most frequently used spin traps are alpha-phenyl N-tertiary-butyl nitrone (PBN) and 5,5-dimethyl-pyrroline N-oxide (DMPO). Secondly C-nitroso spin traps such as 3,5-Dibromo-4-nitrosobenzenesulfonic acid (DBNBS) can be preferred.
EPR spectroscopy applications cover a wide range of areas from structural biology to quantum physics. The EPR characteristics can be obtained from EPR spectra of spin adducts such as the g-value, the hyperfine coupling constant (hfcc) and the spin concentration. The concentration of a free radical can be obtained as the double integrals of the EPR spectrum of a spin adduct which eventually allows us to investigate ROS and RNS qualitatively and quantitatively [18].
Khan and Swartz [20] described that in vivo EPR spectroscopy can provide valuable information relevant to the study of oxygen/nitrogen radicals. They pointed that there are some parameters that can be measured, including oxygen-, carbon- or sulfur-centered radicals, such as nitric oxide by spin trapping, oxygen by oxygen-sensitive paramagnetic materials and thiol groups using special nitroxides. The spin-trapping method by in vivo EPR is an effective technique for providing non-invasive measurements in animals. EPR imaging in the organs of live animals can also be observed after application of nitroxyl radicals as an imaging agent using EPR-computed tomography. In vivo EPR imaging has been established as an effective method for determining the free radicals in living organs and tissues [21]. EPR is a powerful technique when combined with an appropriate spin trap to measure O2, OH, and NO in biological samples [1,22].
Rapid-scan EPR is based on continuous direct detection of the spin response as the magnetic field is scanned upfield and downfield through the resonance thousands of times per second. The method provides an improved signal-to-noise ratio for a wide range of samples, including immobilized radicals [23]. In contrast, conventional continuous wave (CW) EPR uses phase-sensitive detection at the modulation frequency. Recent developments of rapid-scan EPR provide an alternative way to observe EPR spectra. In rapid-scan EPR the signal is detected directly and the magnetic field is scanned through resonance in a time that is short relative to the electron spin relaxation times [24,25,26]. Rapid-scan EPR can also be used to extract relaxation information from a sample.

5. Detection Examples of Reactive Oxygen and Nitrogen Species by EPR

EPR spectroscopy can provide exclusive data on the identity, quantity, dynamics and environment of radical species. ROS and RNS play essential roles in regulating physiological and pathophysiological pathways in cells [27]. The direct or indirect observations of O2•−/HO2 formation have been realized by spin trapping using biological systems such as mitochondria [28], nitric oxide synthase [29], endothelial cells [30] and human neutrophils [31]. The most popular cyclic nitrones can be stated as (DMPO) and 5-diethoxyphosphoryl-5-methyl pyrroline N-oxide (DEPMPO) [32] for similar studies.
The most commonly used spin traps for ROS detection are the pyrroline-based cyclic nitrones, such as DMPO and DEPMPO, which react with the OH and O2 radicals to form –OH and –OOH adducts, respectively. It was found that cyclic hydroxylamines can react with O2 to form more stable adducts compared to those with nitrone spin traps [33,34,35].
Recent studies have revealed that OS has important roles in various neuronal conditions such as stroke and traumatic brain injury. Dohi et al. [36] developed a technique for the detection of free radicals and OS using the ex vivo electron spin resonance (EPR) spin-trapping method in patients with neuroemergencies. The alkoxyl radical level was measured by ex vivo EPR spectrometry using DMPO as a spin trap in blood samples. This method was found useful and may prove valuable for clarifying the pathophysiology of neuroemergency diseases.
ROS, such as hydroxyl and superoxide radicals, have too-short lifetimes at ambient temperatures which are difficult to detect easily by EPR. When rapid-scan EPR is used, the magnetic field is scanned through resonance in a time that is short relative to the electron spin relaxation times, and data are processed to obtain the absorption spectrum. To validate the application of rapid-scan EPR to spin trapping, superoxide was generated by the reaction of xanthine oxidase and hypoxanthine and trapped with 5-tert-butoxycarbonyl-5-methyl-1-pyrroline-N-oxide (BMPO) by Mitchell et al. [37]. They observed that rapid-scan EPR can detect lower concentrations of BMPO-OOH radicals. Rapid-scan EPR will increase the chances to apply spin-trapping and stable nitroxides to explore important biological experiments in vivo and in vitro such as the detection of superoxide produced by Enterococcus faecalis at rates that are too low for detection by other methods.
Studies showed that the direct detection of some free radicals such as superoxide and hydroxyl radicals is quite problematic in solution at room temperature. The most popular spin trap is DMPO which has substantial benefits such as being most the redox-inactive. Commonly used nitrone spin traps, except DMPO, such as α-phenyl-N-tert-butylnitrone (PBN) and α-(4-pyridyl-1-oxide)-N-tert-butylnitrone (POBN), have EPR spectra of their radical adducts which show relatively little dependence on the structure of the trapped radicals. BMPO-derived adducts display a much higher signal-to-noise ratio in their EPR spectra, and this could be suitable for the detection of sulfite, hydroxyl and methyl radicals [38,39].
Recent developments have allowed short-lived ROS to be directly detected in solution using several methods. To study alternative spin configurations for important reactive intermediates such as oxenium ions, which are not very well known regarding their reactivity, lifetimes, and electronic configurations, a photo-precursor was synthesized to the m-dimethylamino phenyloxenium ion. This method achieved direct spectroscopic detection and EPR investigation of a ground-state triplet phenyl oxenium ion possible [40].
NO can be detected directly by EPR since it is a diatomic free radical. It is not possible to directly observe NO by EPR in any biological media. Several approaches have been established to trap and detect nitric oxide by EPR. One of them is diamagnetic spin trapping by using DMPO which can be used to trap superoxide, hydroxyl, and glutathionyl radicals. NO will not directly react with nitrone-based spin-traps to form spin adducts [41]. NO is considered a relatively stable radical that can conjugate with spin traps. Various studies define the use of EPR spin-trapping methods to detect NO in several in vitro and in vivo systems including in tumors [42,43]. Different iron compounds have been used to trap NO such as diethyldithiocarbamate (DETC) and N-methylglucamine dithiocarbamate (MGD). MGD was found specifically useful for the extracellular detection of NO while DETC is helpful for NO in the cellular lipid membrane [29,35].
Nitroxyl radicals display a one-electron reduction in reactions with oxidoreductases in mitochondria and microsomes, antioxidants and other free radical species in biological systems. After losing their paramagnetism with the one-electron reduction they primarily convert to the corresponding diamagnetic hydroxylamines [44]. A common OS producing such ROS is ionizing radiation. The β-ray irradiation can significantly decrease the EPR signal of a nitroxyl radical in a solution containing glutathione (GSH). In a study, the stability and reactivity of nitroxyl radicals in the reaction mixture containing an H-donor, such as GSH, NADH, or NADPH, were tested. It was found that the existence of an H-donor, such as GSH or NAD(P)H, was crucial to detect the reduction of nitroxyl radicals by low-dose irradiation. It was thought that another irreversible reaction of GSH and nitroxyl radicals was possible. The combination of 4-hydroxy-2,2,6,6-tetramethyl-1-piperidinyloxyl (4-hydroxy-TEMPO) and GSH was preferable for the quantitative detection of the free radical reaction caused by radiation [44].
Iron (II)-N-methyl-d-glucamine dithiocarbamate, Fe(MGD)2, is frequently used to trap NO due to making a large amount of stable adduct formation. In a study EPR spectroscopy was used to detect nitric oxide from bovine aortic endothelial cells and superoxide radical anions from human neutrophils using Fe(MGD)2 and 5,5-dimethyl-1-pyroroline-N-oxide, DMPO, respectively [45]. As a result, when the spin trapping of NO radicals was achieved using Fe2+-MGD, no EPR signal from Fe2+-MGD was found. When the DMPO spin trap was used for O2•-detection, DMPO did not show a signal, proving that the spin trap is free from paramagnetic impurities. The spectra of DMPO and PMNs (polymorphonuclear neutrophils), likewise, showed no detectable signal, suggesting that the DMPO does not cause activation of the enzyme NADPH oxidase.
New cyclic nitrones with a triphenylphosphonium or permethylated β-cyclodextrin moiety have been synthesized and their spin adducts were demonstrated to increase stability in buffer which was studied in the presence of liver subcellular fractions. The stability and kinetics of the superoxide adducts of four cyclic nitrones were detected by the EPR spectrometer. The results proved that the new spin-trapping agents CD-DEPMPO and CD-DIPPMPO are suitable for the specific detection of superoxide, especially in the presence of liver microsomes [46,47].
Nitronyl nitroxides (NNs) are the paramagnetic probes that have the potential to scavenge ROS and RNS species and are used as probes for the detection of NO and HNO by EPR spectroscopy. Investigation of the redox properties of the two most frequently used NNs was stated in a study. The results showed that the rate constants of the reaction of the NNs with HNO were found to be very close to the rate constants of scavenging superoxide and NO. The scavenging rates of the NNs towards superoxide, NO, and HNO, and their low reduction potential being thermodynamically close to the bottom of the pecking order of oxidizing radicals were thought to be significant factors contributing to their antioxidant activity [48].
Nitrone spin traps have been used as probes for the determination of radical species in chemical and biological systems using EPR spectroscopy, and they have found pharmacological activity against OS-mediated diseases. Calix(4)pyrroles have shown to exhibit high affinity to (O2•−) anions; the cyclic nitrone conjugate of calix(4)pyrroles (CalixMPO) was designed by Kim et al. [32]. Computational studies showed that a pendant-type linkage between the calix(4)pyrrole and the nitrone was the most efficient design for the spin trapping of O2. The EPR method gave larger trapping rate constant for CalixMPO and a longer half-life for its O2•− adduct compared to the commonly used nitrones. With these results the application of an anion receptor for the detection of one of the most important radical intermediates in biological and chemical systems was established.
There are complementary methods for monitoring drug candidates. For example, the distribution and the pharmacokinetics of the new potential drug NOM (1-nitrosomelatonin) were evaluated. First [O-methyl-3H]NOM was administered to and followed in mice, and then monitoring of NOM, using EPR, was done in vitro and ex vivo with the (MGD)2–Fe2+ (iron–N-methyl-d-glucamine dithiocarbamate) complex as a spin trap. According to the characterization of the EPR signal in vitro with NOM or GSNO (S-nitrosoglutathione) using the (MGD)2–Fe2+ complex, the detection of these nitroso compounds was realized ex vivo in mouse tissue extracts. Results showed that NOM was able to cross the blood-brain barrier, while GSNO was not [49]. Studies showed that nitrosation may occur on thiol and indole derivatives giving nitroso compounds, which are in turn NO donors. Regarding thiols, GSNO and S-nitroso proteins have been identified in several biological samples [50,51].
NO can coordinate with both ferric and ferrous iron to form complexes. These complexes display EPR signals and so iron has been utilized in several ways as a spin trap for NO [52,53]. The ferrous iron complex is insoluble in aqueous media but water-soluble complexes have been applied for the biological detection of NO in vitro and in vivo [54]. One of the main benefits of using iron as a spin trap is that it will mostly react with NO very quickly [55].
An EPR technique using the spin probe cyclic hydroxylamine 1-hydroxy-3-methoxycarbonyl-2,2,5,5-tetramethylpyrrolidine (CMH) was introduced for the sensitive quantification of ROS, including the superoxide radical in frozen biological samples such as cell suspensions, blood or biopsies [56]. The absolute spin concentration in wet biological samples such as biopsies, water solutions and cell cultures could be quantified with higher precision and accuracy. This technique allowed researchers to collect and store the biological samples for future incubation with a spin probe, and also to further store them up to at least six months before EPR analysis, without loss of the signal intensity.
In EPR experimentation, a common technique is to use a cell suspension to measure ROS in the cell. An in situ EPR detection methodology was developed to detect ROS produced in adherent cells. A quartz glass plate coated with poly-l-lysine (PLL) was used to improve the adhesion efficiency of cultured cells. This procedure was used for mouse fibroblasts and human malignant epithelial cells (HeLa) and significantly increased the EPR signal intensities of spin adducts related to superoxide radicals 1 h after the addition of a redox-active compound and a spin trap to the culture medium. These findings were demonstrated to be useful for measuring ROS generated under cell culture conditions [41]. Also, an EPR microinvasive method developed by Mrakic-Sposta et al. [57] provides direct evidence of the “instantaneous” presence of ROS, returning to absolute concentration levels that correlate with “a posteriori” assays of ROS-induced damage by means of biomarkers. The comparison of the results with antioxidant capacity and oxidative damage biomarker concentrations showed that all changes indicating increased OS are directly related to a ROS production increase. The method is noninvasive, sensitive, and quantitative, so it can be identified as an essential tool in the study of oxidants and OS in free radical biology and medicine.

6. Concluding Remarks

ROS are by-products of the aerobic metabolism of various pathological conditions. The EPR spectroscopy technique is able to deliver valuable information on the presence of ROS and RNS in biological systems. It is possible to derive the ultimate information on the identity of the species as well as information on their quantity, structure and possible interactions.
There are some limitations for EPR spin trapping during ROS determination in vivo/in vitro, such as detection limit of 1 nM for O2−. The technique requires spin traps that reduce specificity and stability for the measurement [34]. EPR measurements usually require low temperatures and long recording times for ROS detection in physiological conditions [35,58]. The short lifetime of free radicals and the low rates of formation expected in in vivo detection by EPR are challenges. The new rapid-scan EPR method offers improved sensitivity for these types of samples. Making interpretations in complex and non-homogenous solutions is a key advantage of EPR compare to other spectroscopic techniques [59]. EPR techniques are capable of directly measuring ROS in vivo. Recent developments in analytical techniques such as EPR offer more accurate and quantitative approaches to measure ROS in cells.

Acknowledgments

Authors acknowledge the support of TÜBİTAK Project 112S599.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Griendling, K.K.; Touyz, R.M.; Zweier, J.L.; Dikalov, S.; Chilian, W.; Chen, Y.R.; Harrison, D.G.; Bhatnagar, A. Measurement of reactive oxygen species, reactive nitrogen species, and redox-dependent signaling in the cardiovascular system: A scientific statement from the American Heart Association. Circ. Res. 2016, 119, e39–e75. [Google Scholar] [CrossRef] [PubMed]
  2. Aprioku, J.S. Pharmacology of Free Radicals and the Impact of Reactive Oxygen Species on the Testis. J. Rep. Infert. 2013, 14, 158–172. [Google Scholar]
  3. Suzen, S. Antioxidant activities of synthetic indole derivatives and possible activity mechanisms. In Topics in Heterocyclic Chemistry, Bioactive Heterocycles; Khan, M.T.H., Ed.; Spinger: Berlin/Heidelberg, Germany, 2007; Volume 11, pp. 145–178. [Google Scholar]
  4. Suzen, S.; Saso, L. Antioxidant heterocyclic compounds in drug discovery and medicinal chemistry. Mini Rev. Med. Chem. 2013, 13, 317–318. [Google Scholar] [PubMed]
  5. Gurer-Orhan, H.; Karaaslan, C.; Ozcan, S.; Firuzi, O.; Tavakkoli, M.; Saso, L.; Suzen, S. Novel indole-based melatonin analogues: Evaluation of antioxidant activity and protective effect against amyloid β-induced damage. Bioorg. Med. Chem. 2016, 24, 1658–1664. [Google Scholar] [CrossRef] [PubMed]
  6. Shaeib, F.; Banerjee, J.; Maitra, D.; Diamond, M.P.; Abu-Soud, H.M. Impact of hydrogen peroxide driven Fenton reaction on mouse oocyte quality. Free Rad. Biol. Med. 2013, 58, 154–159. [Google Scholar] [CrossRef] [PubMed]
  7. Suzen, S. Evaluation of synthetic melatonin analog antioxidant compounds. In Melatonin: Therapeutic Value and Neuroprotection; Srinivasan, V., Gobbi, G., Shillcutt, S.D., Suzen, S., Eds.; Taylor & Francis Group: Boca Raton, FL, USA, 2015; Chapter 21; pp. 259–269. [Google Scholar]
  8. Hybertson, B.M.; Gao, B.; Bose, K.S.; McCord, J.M. Oxidative stress in health and disease: The therapeutic potential of Nrf2 activation. Mol. Asp Med. 2011, 32, 234–246. [Google Scholar] [CrossRef] [PubMed]
  9. Foyer, C.H.; Noctor, G. Redox homeostasis and antioxidant signaling: A metabolic interface between stress perception and physiological responses. Plant Cell 2005, 17, 1866–1875. [Google Scholar] [CrossRef] [PubMed]
  10. Jackson, H.L.; Cardounel, A.J.; Zweier, J.L.; Lockwood, S.F. Synthesis, characterization, and direct aqueous superoxide anion scavenging of a highly water-dispersible astaxanthin-amino acid conjugate. Bioorg. Med. Chem. Lett. 2004, 14, 3985–3991. [Google Scholar] [CrossRef] [PubMed]
  11. Khan, N.; Wilmot, C.M.; Rosen, G.M.; Demidenko, E.; Sun, J.; Joseph, J.; O’Hara, J.; Kalyanaraman, B.; Swartz, H.M. In vitro toxicity and stability of radical adducts. Free Radic. Biol. Med. 2003, 34, 1473–1481. [Google Scholar] [CrossRef]
  12. Schweiger, A.; Jeschke, G. Principles of Pulse Paramagnetic Resonance; Oxford University Press: Oxford, UK, 2001. [Google Scholar]
  13. Weil, J.A.; Bolton, J.R. Electron Paramagnetic Resonance; Wiley-Interscience: Hoboken, NJ, USA, 2007. [Google Scholar]
  14. Cooke, J.A.; Brown, L.J. Distance measurements by continuous wave EPR spectroscopy to monitor protein folding. Methods Mol. Biol. 2011, 752, 73–96. [Google Scholar] [PubMed]
  15. Kalin, M.; Gromov, I.; Schweiger, A. The continuous wave electron paramagnetic resonance experiment revisited. J. Mag. Res. 2003, 160, 166–182. [Google Scholar] [CrossRef]
  16. Prisner, T.; Rohrer, M.; MacMillan, F. Pulsed EPR Spectroscopy: Biological Applications. Ann. Rev. Phys. Chem. 2001, 52, 279–313. [Google Scholar] [CrossRef] [PubMed]
  17. Rigby, S.E.J.; Evans, M.C.W.; Heathcote, P. Electron nuclear double resonance (ENDOR) spectroscopy of radicals in photosystem I and related Type 1 photosynthetic reaction centres. Biochim. Biophys. Acta Bioenerg. 2001, 1507, 247–259. [Google Scholar] [CrossRef]
  18. Kohno, M. Applications of Electron Spin Resonance Spectrometry for Reactive Oxygen Species and Reactive Nitrogen Species Research. J. Clin. Biochem. Nutr. 2010, 47, 1–11. [Google Scholar] [CrossRef] [PubMed]
  19. Zavoisky, E. Spin-magnetic resonance in paramagnetics. Fizicheskiĭ Zhurnal. 1945, 9, 211–245. [Google Scholar]
  20. Khan, N.; Swartz, H. Measurements in vivo of parameters pertinent to ROS/RNS using EPR spectroscopy. Mol. Cell. Biochem. 2002, 234–235, 341–357. [Google Scholar] [CrossRef] [PubMed]
  21. Egashira, T.; Takayama, F. Free radicals and oxidative stress: Targeted ESR measurement of free radicals. Nihon Yakurigaku Zasshi. 2002, 120, 229–236. [Google Scholar] [CrossRef]
  22. Reszka, K.J.; Bilski, P.; Chignell, C.F. EPR and spin-trapping investigation of nitric oxide (NO) from UV irradiated nitrite anions in alkaline aqueous solutions. J. Am. Chem. Soc. 1996, 118, 8719–8720. [Google Scholar] [CrossRef]
  23. Eaton, S.S.; Eaton, G.R. Chapter One—Rapid-Scan EPR of Nitroxide Spin Labels and Semiquinones. Methods Enzym. 2015, 563, 3–21. [Google Scholar] [PubMed]
  24. Quine, R.W.; Czechowski, T.; Eaton, G.R. A Linear Magnetic Field Scan Driver. Concept. Magn. Reson. Part B Magn. Reson. Eng. 2009, 35B, 44–58. [Google Scholar] [CrossRef] [PubMed]
  25. Quine, R.W.; Rinard, G.A.; Eaton, S.S.; Eaton, G.R. Quantitative rapid scan EPR spectroscopy at 258 MHz. J. Magn. Reson. 2010, 205, 23–27. [Google Scholar] [CrossRef] [PubMed]
  26. Tseitlin, M.; Czechowski, T.; Quine, R.W.; Eaton, S.S.; Eaton, G.R. Background removal procedure for rapid scan EPR. J. Magn. Reson. 2009, 196, 48–53. [Google Scholar] [CrossRef] [PubMed]
  27. Abdel-Rahman, E.; Mahmoud, A.M.; Khalifa, A.M.; Ali, S.S. Physiological and pathophysiological reactive oxygen species as probed by EPR spectroscopy: The underutilized research window on muscle ageing. J. Physiol. 2016, 594, 4591–4613. [Google Scholar] [CrossRef] [PubMed]
  28. Chen, Y.R.; Chen, C.L.; Yeh, A.; Liu, X.; Zweier, J.L. Direct and Indirect Roles of Cytochrome bin the Mediation of Superoxide Generation and NO Catabolism by Mitochondrial Succinate-Cytochrome c Reductase. J. Biol. Chem. 2006, 281, 13159–13168. [Google Scholar] [CrossRef] [PubMed]
  29. Weaver, J.; Porasuphatana, S.; Tsai, P.; Budzichowski, T.; Rosen, G.M. Spin trapping nitric oxide from neuronal nitric oxide synthase: A look at several iron-dithiocarbamate complexes. Free Radic. Res. 2005, 39, 1027–1033. [Google Scholar] [CrossRef] [PubMed]
  30. Varadharaj, S.; Watkins, T.; Cardounel, A.J.; Garcia, J.G.N.; Zweier, J.L.; Kuppusamy, P.; Natarajan, V.; Parinandi, N.L. Vitamin C-Induced Loss of Redox-Dependent Viability in Lung Microvascular Endothelial Cells. Antioxid. Redox. Signal. 2005, 7, 287–300. [Google Scholar] [CrossRef] [PubMed]
  31. Jackson, S.K.; Thomas, M.P.; Smith, S.; Madhani, M.; Rogers, S.C.; James, P.E. In Vivo EPR Spectroscopy: Biomedical and Potential diaGnostic Applications. Faraday Discuss. 2004, 126, 103–117, discussion 169–183. [Google Scholar] [CrossRef] [PubMed]
  32. Kim, S.U.; Liu, Y.; Nash, K.M.; Zweier, J.L.; Rockenbauer, A.; Villamena, F.A. Fast Reactivity of Cyclic Nitrone-Calix(4)Pyrrole Conjugate with Superoxide Radical Anion: Theoretical and Experimental Studies. J. Am. Chem. Soc. 2010, 132, 17157–17173. [Google Scholar] [CrossRef] [PubMed]
  33. Münzel, T.; Afanas’ev, I.B.; Kleschyov, A.L.; Harrison, D.G. Detection of superoxide in vascular tissue. Arterioscler. Thromb. Vasc. Biol. 2002, 22, 1761–1768. [Google Scholar] [CrossRef]
  34. Cai, H.; Dikalov, S.; Griendling, K.K.; Harrison, D.G. Detection of reactive oxygen species and nitric oxide in vascular cells and tissues: Comparison of sensitivity and specificity. Methods Mol. Med. 2007, 139, 293–311. [Google Scholar]
  35. Rodriguez-Rodriguez, R.; Simonsen, U. Science Publishers Measurement of Nitric Oxide and Reactive Oxygen Species in the Vascular Wall. Curr. Anal. Chem. 2012, 8, 485–494. [Google Scholar] [CrossRef]
  36. Dohi, K.; Satoh, K.; Nakamachi, T.; Ohtaki, H.; Yofu, S.; Nakamura, S.; Shioda, S.; Aruga, T. Novel free radical monitoring in patients with neurological emergency diseases. Acta Neurochir. Suppl. 2010, 106, 315–319. [Google Scholar] [PubMed]
  37. Mitchell, D.G.; Rosen, G.M.; Tseitlin, M.; Symmes, B.; Eaton, S.S.; Eaton, G.R. Use of Rapid-Scan EPR to Improve Detection Sensitivity for Spin-Trapped Radicals. Biophys. J. 2013, 105, 338–342. [Google Scholar] [CrossRef] [PubMed]
  38. Zhao, H.T.; Joseph, J.; Zhang, H.; Karoui, H.; Kalyanaraman, B. Synthesis and biochemical applications of a solid cyclic nitrone spin trap: A relatively superior trap for detecting superoxide anions and glutathiyl radicals. Free Radical Biol. Med. 2001, 31, 599–606. [Google Scholar] [CrossRef]
  39. Ranguelova, K.; Mason, R.P. The fidelity of spin trapping with DMPO in biological systems. Magn. Reson. Chem. 2011, 49, 152–158. [Google Scholar] [CrossRef] [PubMed]
  40. Li, M.D.; Albright, T.R.; Hanway, P.J.; Liu, M.; Lan, X.; Li, S.; Peterson, J.; Winter, A.H.; Phillips, D.L. Direct Spectroscopic Detection and EPR Investigation of a Ground State Triplet Phenyl Oxenium Ion. J. Am. Chem. Soc. 2015, 137, 10391–10398. [Google Scholar] [CrossRef] [PubMed]
  41. Ando, T.; Yonamoto, Y. In Situ EPR Detection of Reactive Oxygen Species in Adherent Cells Using Polylysine-Coated Glass Plate. Appl. Magn. Reson. 2015, 46, 977–986. [Google Scholar] [CrossRef]
  42. Loibl, S.; von Minckwitz, G.; Weber, S.; Sinn, H.P.; Schini-Kerth, V.B.; Lobysheva, I.; Nepveu, F.; Wolf, G.; Strebhardt, K.; Kaufmann, M. Expression of endothelial and inducible nitric oxide synthase in benign and malignant lesions of the breast and measurement of nitric oxide using electron paramagnetic resonance spectroscopy. Cancer 2002, 95, 1191–1198. [Google Scholar] [CrossRef] [PubMed]
  43. Alzawahra, W.F.; Talukder, M.A.; Liu, X.; Samouilov, A.; Zweier, J.L. Heme proteins mediate the conversion of nitrite to nitric oxide in the vascular wall. Am. J. Phys. Heart Circ. Phys. 2008, 295, H499–H508. [Google Scholar] [CrossRef] [PubMed]
  44. Matsumoto, K.I.; Okajo, A.; Nagata, K.; Degraff, W.G.; Nyui, M.; Ueno, M.; Nakanishi, I.; Ozawa, T.; Mitchell, J.B.; Krishna, M.C.; et al. Detection of Free Radical Reactions in an Aqueous Sample Induced by Low Linear-Energy-Transfer Irradiation. Biol. Pharm. Bull. 2009, 32, 542–547. [Google Scholar] [CrossRef] [PubMed]
  45. Gopalakrishnan, B.; Nash, K.M.; Velayutham, M.; Villamena, F.A. Detection of Nitric Oxide and Superoxide Radical Anion by Electron Paramagnetic Resonance Spectroscopy from Cells using Spin Traps. J. Vis. Exp. 2012, 66, 2810–2821. [Google Scholar] [CrossRef] [PubMed]
  46. Beziere, N.; Hardy, M.; Poulhes, F.; Karoui, H.; Tordo, P.; Ouari, O.; Frapart, Y.M.; Rockenbauer, A.; Boucher, J.L.; Mansuy, D.; et al. Metabolic stability of superoxide adducts derived from newly developed cyclic nitrone spin traps. Free Radic. Biol. Med. 2014, 67, 150–158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Abbas, K.; Hardy, M.; Poulhes, F.; Karoui, H.; Tordo, P.; Ouari, O.; Peyrot, F. Detection of superoxide production in stimulated and unstimulated living cells using new cyclic nitrone spin traps. Free Rad. Biol. Med. 2014, 71, 281–290. [Google Scholar] [CrossRef] [PubMed]
  48. Bobko, A.A.; Khramtsov, V.V. Redox properties of the nitronyl nitroxide antioxidants studied via their reactions with nitroxyl and ferrocyanide. Free Radic. Res. 2015, 49, 919–926. [Google Scholar] [CrossRef] [PubMed]
  49. Peyrot, F.; Grillon, C.; Vergely, C.; Rochette, L.; Ducrocq, C. Pharmacokinetics of 1-nitrosomelatonin and detection by EPR using iron dithiocarbamate complex in mice. Biochem. J. 2005, 387, 473–478. [Google Scholar] [CrossRef] [PubMed]
  50. Blanchard, B.; Pompon, D.; Ducrocq, C. Nitrosation of melatonin by nitric oxide and peroxynitrite. J. Pineal Res. 2000, 29, 184–192. [Google Scholar] [CrossRef] [PubMed]
  51. Feelisch, M.; Rassaf, T.; Mnaimneh, S.; Singh, N.; Bryan, N.S.; Jourd’heuil, D.; Kelm, M. Concomitant S-, N-, and heme-nitros(yl)ation in biological tissues and fluids: Implications for the fate of NO in vivo. FASEB J. 2002, 16, 1775–1785. [Google Scholar] [CrossRef] [PubMed]
  52. Mordvintcev, P.; Mülsch, A.; Busse, R.; Vanin, A. On-line detection of nitric oxide formation in liquid aqueous phase by electron paramagnetic resonance spectroscopy. Anal. Biochem. 1991, 199, 142–146. [Google Scholar] [CrossRef]
  53. Hogg, N. Detection of Nitric Oxide by Electron Paramagnetic Resonance Spectroscopy. Free Radic. Biol. Med. 2010, 49, 122–129. [Google Scholar] [CrossRef] [PubMed]
  54. Komarov, A.; Mattson, D.; Jones, M.M.; Singh, P.K.; Lai, C.S. In vivo spin trapping of nitric oxide in mice. Biochem. Biophys. Res. Commun. 1993, 195, 1191–1198. [Google Scholar] [CrossRef] [PubMed]
  55. Kleschyov, A.L.; Wenzel, P.; Munzel, T. Electron paramagnetic resonance (EPR) spin trapping of biological nitric oxide. J. Chromatogr. B Analyt. Technol. Biomed. Life Sci. 2007, 851, 12–20. [Google Scholar] [CrossRef] [PubMed]
  56. Berg, K.; Ericsson, M.; Lindgren, M.; Gustafsson, H. A High Precision Method for Quantitative Measurements of Reactive Oxygen Species in Frozen Biopsies. PLoS ONE 2014, 9, e90964. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Mrakic-Sposta, S.; Gussoni, M.; Montorsi, M.; Porcelli, S.; Vezzoli, A. A Quantitative Method to Monitor Reactive Oxygen Species Production by Electron Paramagnetic Resonance in Physiological and Pathological Conditions. Oxid. Med. Cell Longev. 2014, 2014, 306179. [Google Scholar] [CrossRef] [PubMed]
  58. Villamena, F.A.; Zweier, J.L. Detection of reactive oxygen and nitrogen species by EPR spin trapping. Antioxid. Redox. Signal. 2004, 6, 619–629. [Google Scholar] [CrossRef] [PubMed]
  59. Shulaev, V.; Oliver, D.J. Metabolic and Proteomic Markers for Oxidative Stress. New Tools for Reactive Oxygen Species Research. Plant Phys. 2006, 141, 367–372. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Suzen, S.; Gurer-Orhan, H.; Saso, L. Detection of Reactive Oxygen and Nitrogen Species by Electron Paramagnetic Resonance (EPR) Technique. Molecules 2017, 22, 181. https://doi.org/10.3390/molecules22010181

AMA Style

Suzen S, Gurer-Orhan H, Saso L. Detection of Reactive Oxygen and Nitrogen Species by Electron Paramagnetic Resonance (EPR) Technique. Molecules. 2017; 22(1):181. https://doi.org/10.3390/molecules22010181

Chicago/Turabian Style

Suzen, Sibel, Hande Gurer-Orhan, and Luciano Saso. 2017. "Detection of Reactive Oxygen and Nitrogen Species by Electron Paramagnetic Resonance (EPR) Technique" Molecules 22, no. 1: 181. https://doi.org/10.3390/molecules22010181

Article Metrics

Back to TopTop