Next Article in Journal
Natural Products Counteracting Cardiotoxicity during Cancer Chemotherapy: The Special Case of Doxorubicin, a Comprehensive Review
Previous Article in Journal
Synergistic Effect of Simultaneous versus Sequential Combined Treatment of Histone Deacetylase Inhibitor Valproic Acid with Etoposide on Melanoma Cells
Previous Article in Special Issue
The Protective Effects of EMF-LTE against DNA Double-Strand Break Damage In Vitro and In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Oxidative Stress and NADPH Oxidase: Connecting Electromagnetic Fields, Cation Channels and Biological Effects

by
Christos D. Georgiou
1 and
Lukas H. Margaritis
2,*
1
Department of Biology, Section of Genetics, Cell & Developmental Biology, University of Patras, 10679 Patras, Greece
2
Section of Cell Biology and Biophysics, Department of Biology, National and Kapodistrian University of Athens, 26504 Athens, Greece
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(18), 10041; https://doi.org/10.3390/ijms221810041
Submission received: 10 July 2021 / Revised: 1 September 2021 / Accepted: 15 September 2021 / Published: 17 September 2021
(This article belongs to the Special Issue Biological Effects of Non-ionizing Radiation)

Abstract

:
Electromagnetic fields (EMFs) disrupt the electrochemical balance of biological membranes, thereby causing abnormal cation movement and deterioration of the function of membrane voltage-gated ion channels. These can trigger an increase of oxidative stress (OS) and the impairment of all cellular functions, including DNA damage and subsequent carcinogenesis. In this review we focus on the main mechanisms of OS generation by EMF-sensitized NADPH oxidase (NOX), the involved OS biochemistry, and the associated key biological effects.

1. Introduction

Oxidative stress (OS) generation, expressed by the production of reactive oxygen/nitrogen species (ROS/RNS), has been linked to the exposure of animals and cells to man-made EMF. The related key experimental studies of the last decade have been presented elsewhere [1]. However, the biochemical specifics of the biological mechanisms involving the generation of OS are unclear, with the most well studied mechanisms being those involving ion channels. The main hypothesis underlining these mechanisms portrays EMF (RF 900, 1800, 1900, 2450 MHz, ELF 0–3000 Hz) exerting their oscillatory forces on every free ion on both sides of any biological membrane, thereby causing the ions, cations, in particular, to pass through at abnormal rates. Such abnormal cation movement can alter the biochemical properties of membranes and can cause deterioration of cation channel functions, especially those of the voltage-gated (VGC) ones [2,3,4,5,6,7,8]. These, in turn, can trigger an increase of OS, leading to the impairment of most cellular functions and DNA damage [9,10] as well as to numerous associated diseases including carcinogenesis [11,12,13]. In this review, we examine (i) the main mechanisms of OS generation by EMF-induced VGCs and their association with the ROS-generating NADPH oxidase (NOX) and the voltage-gated calcium channels (VGCC), (ii) the involved OS biochemistry, and (iii) the VGC-OS-originating key biological effects. We conclude with the need for research to unequivocally prove that non-ionizing EMF can directly generate ROS.

2. EMF Effect on Membrane Function and Cation Channels Can Induce OS

Disruption of the electrochemical balance and cation channel function of the cell membrane by EMFs and the generation of OS can involve the following mechanisms:
EMF disturbance of membrane function and cation channels: Abnormal function of cell membranes and ion channels by EMFs primarily take place by means of the following mechanisms:
  • EMFs can act on membrane-associated free ions by means of a forced-vibration mechanism, which acts on the voltage-gated channels using the vibrations that EMFs cause on the transported ions [2,3]. These are the Κ+ leak channels, the Νa+ leak channels, and the leak channels of other cations such as Ca2+. Briefly, the open or closed state of these channels is determined by the electrostatic interaction between the voltage sensors of the channels and the transmembrane voltage, and they change between their open and closed state when the electrostatic force on the electric charges of their voltage sensors exceeds a critical value. Alternating electric fields exert a periodic force on all of the free ions passing across voltage-gated channels and can distort the electrochemical equilibrium state of every oscillating ion due to its site-displacement from the developed electrochemical gradient. This results in a periodical displacement of the electric charge, which will exert force on every fixed charge, such as those existing on the voltage sensors of the voltage-gated channels. Specifically, voltage-gated channels contain transmembrane a-helices in parallel, forming pores through which ions pass in their dehydrated state. The voltage sensors of these channels are four positively charged symmetrically arranged transmembrane alpha-helical domains (designated S4). It is known that small voltage changes of about 30 mV in the membrane potential are able to gate this kind of channel [14,15]. Such a change can be caused by the displacement of a single ion by 10−12 m from the electric field of the EMF and in the vicinity S4 of the voltage-gated channels. Hence, EMF-induced oscillating ions can disturb the electrochemical balance of the membrane via the gating of such channels, and those ions crossing such channels can change their normal positions and can produce a false signal for the gating such channels with their charge. This mechanism can also explain the biological action of oscillating magnetic fields by replacing the force of the electric field with the force exerted by an alternating magnetic field and also by accounting for the induced electric field, which is always generated by the pulsed magnetic one. The mechanism concludes that oscillating electric or magnetic fields with frequencies lower than 1.6 × 104 Hz (ELF and VLF fields) can be bioactive, even at very low intensities [2,16]. It is also claimed that pulsed EMFs can even further amplify their biological action compared to continuous EMFs [16,17,18].
  • In the other mechanism [7,8], low intensity EMFs cause abnormal hypersensitivity on voltage-gated calcium channels (VGCC), with consequent increases in intracellular Ca2+ (and OS) acting on the structure of the aforementioned voltage sensor (and its 20 electrically charged groups that are equally distributed on its 4 alpha-helixes), which actually regulates the opening of the channel. They transverse the lipid bilayer of the membrane and their disturbance by the EMFs relies on the following very important and distinct reasons: The EMF forces on the voltage sensor are 20 times higher since they act on all 20 charges, and according to Coulomb’s Law of physics, they are inversely proportional to the dielectric constant of the fatty part of the membrane, which is about 120 times less than the forces on the charges in the aqueous parts of the cell. Thus, the forces exerted on the 20 electrically charged groups of the VGCC’s voltage sensor by the electric field of the RF EMF will also be about 120 times stronger (or 2400 times stronger, accounting for all 20 charges), which is only because of the dielectric constant of the fatty region of the cell membrane where the voltage sensor’s 20 charges are located. A third, even more important, reason is based on the fact that the cell membrane has a very high electrical resistance, which acts as amplifier of the electrical gradient (the difference in electrical charge across the cell membrane), amplifying it by about 3000 times. Combining these three distinct reasons, it is implied that the total amplification of the exerted forces by the RF EMF electric fields on the VGCC voltage sensor’s 20 electrical charges is equal to 20 × 120 times (due to the dielectric constant of the fatty inner space of the membrane) × 3000 times (due to the electrical gradient of the membrane), totaling 7,200,000 times. That is, the forces exerted on the VGCC voltage sensor by the RF EMFs are about 7.2 million times stronger than those in the electrically charged groups that are in the hydrophilic environment of our cells, which is where the safety guidelines for the RF EMF are set by ICNIRP.
EMFs act via the activation VGCC in the plasma membrane, producing excessive Ca2+, which leads to the pathophysiological effects associated with ROS, such as nitric oxide (NO), superoxide radical (O2•−), and peroxynitrite (ONOOH) [6]. Studies on the mechanisms related to VGCC and to the associated pleiotropic effects are presented elsewhere [1].
OS generation by EMF disturbance of membrane function: NADPH oxidase (NOX) is a plasma membrane enzyme and has activity that is very likely associated with the increased generation of OS, upon membrane voltage-gated channel function interaction with EMF. NOX (found in the plasma membrane the neutrophil white blood cells) normally generates OS (or respiratory burst) to eliminate invading microorganisms by means of the production of reactive oxygen species (ROS) [19]. Specifically, NOX produces intracellular O2•−, which is dismutated to H2O2, and can be used by myeloperoxidase as substrate for the two-electron oxidation of Cl to the cytotoxic and highly reactive ROS oxidant hypochlorous acid (HOCl) (Figure 1) [20]. Alternatively, H2O2 can move extracellularly or to any intracellular site, where it can be converted to the most potent biological oxidant hydroxyl radical (OH) (Table 1, reaction 1), which can oxidatively damage all biological molecules, including DNA (see below).
NOX is the only membrane bound enzyme with channel voltage-gating associated ROS production increase by EMF (3-fold in HeLa plasma membranes exposed to 875 MHz at 0.240 mW/cm2 for 10 min; i.e., ~4-fold below the ICNIRP limit). However, this EMF-induced activity increase does not correspond to a proportional increase of NADPH oxidation but is instead displayed in comparison to its inhibition by the oxidase’s specific inhibitor diphenyleneiodonium (DPI) [21]. Nonetheless, this EMF effect on NOX may be indirect and is exerted on an oxidase-associated voltage-gated channel that is blocked by DPI, given that DPI acts not only as an inhibitor of the oxidase but also as a cation channel blocker [22]. This suggests that the NADH oxidase’s direct response to particular EMF exposure may not be conclusive [23] and that its activity is indirectly associated or controlled by neighbouring cation voltage-gated channels. There are studies that support such cation voltage-gated control of NOX activity. It has been known that electron flux within NOX (from NADPH through a redox chain comprising FAD and two haems) is electrogenic and that it rapidly depolarizes the membrane potential by ~100 mV during respiratory burst, necessitating the need of proton (H+) channels for the reduction of extracellular O2 to O2•− by the oxidase (Figure 1) [24,25,26]. Indeed, it has been shown that the activities of NOX and voltage gated H+ channels present (especially the H+ channel Hv1) in many cells (immune cell types, including neutrophils, eosinophils, B-lymphocytes, T-lymphocytes, macrophages and microglia) appear to be coordinated [27]. The voltage-gated H+ channels Hv1 contribute the majority of the charge compensation required for high-level NOX-dependent O2•− [28,29]. Concomitant K+ release elimination only occurs at maximum O2•− production, while the reversed membrane potential inhibits Ca2+ entry (e.g., by VGCC), thereby preventing the cells from becoming overloaded with Ca2+. Thus, a clear correlation exists between the rate of O2•− production by NOX and the extent of plasma membrane depolarization and inhibition of Ca2+ entry. Disturbance of this limiting mechanism by EMFs may contribute to OS-associated disease [30].
Besides the voltage-gated H+ channels Hv1 coordinating/associating with NOX activity, the oxidase itself may act directly as a proton channel due to its gp91phox transmembrane component (Figure 1), which has been demonstrated to be capable of acting as an oxidase-associated H+ channel and functioning as a voltage-gated H+ conductance pathway [31].
Nonetheless, electric fields exert their oscillatory force on any subcellular membrane as well, especially if the abnormal function of their ion channels may be associated with increased OS. This is especially true for the mitochondrial ion channels/transporters since they also function as sensors and regulators of cellular redox signaling [32,33]. An abnormal OS increase can take place at two main sites in the oxidative phosphorylation electron-transport chain; at the proton transporter complex I (NADH dehydrogenase) and at complex III (cytochrome bc1), leaked single electrons reduce O2 and generate O2•− in the mitochondrial matrix, which, in turn, may activate inner-membrane anion channels and may (through O2•− dismutation to cell membrane permeable H2O2) affect neighboring mitochondria or cytosolic targets as well [32]. The inner mitochondrial membrane ion channels that have been intensely researched are the Ca2+ Uniporter (the primary route of Ca2+ into the matrix) and the Permeability Transition Pore (PTP; has a quite high conductance, ~1 nS; is a non-selective transporter of solutes up to 1500 Da), which are considered to be among the channels that contribute the most to mitochondria dysfunction under stress. An increased Ca2+ flow in the mitochondrial matrix by the Ca2+ Uniporter stimulates oxidative phosphorylation (thus, OS increase) in one or more sites of the electron transport chain as well as in sites of the Krebs Cycle. PTP is considered to be activated by depolarization, Ca2+, and OS [32]. Irrespective of their cellular sites, EMF-induced cation channels can act as ROS sources for the redox modulation of cation voltage-gated channels in general. Intracellular redox status has been known to significantly alter the gating properties of Ca2+ channels (L-, T-, P/Q-type) and Na+ and K+ channels [34], with the oxidative modulation of voltage-gated K+ channels being the most studied [35].

3. Biochemistry of ROS

The main cytotoxic ROS is OH and its production involves the transition metal-catalyzed oxidation of H2O2 (Fenton reaction [36]), which, once formed, can freely move intra-/extra-cellularly and cross the nuclear and sub-cellular membranes, and can therefore reach DNA. The ROS-associated main source of H2O2 is the dismutation of O2 (which can be generated either by the plasma membrane NOX or by O2 after the uptake of single electrons that have been leaked during oxidative phosphorylation). In turn, H2O2 can be converted to OH by free Fe2+ (or Cu+) via the Fenton reaction (Table 1, reaction 1), and its association with the cell’s response to EMF in relation to DNA damage has been also suggested [37]. To sustain the production of OH, Fe2+ (and Cu+) must be regenerated via reduction by O2•− or H2O2 (Table 1, reactions 2 or 3, respectively) [38,39,40] and by ascorbic acid (AH2) (Table 1, reaction 4) [41].
Nonetheless, OH is not the sole product of the Fenton reaction in the presence of bicarbonate (HCO3, even present at low concentrations, i.e., under physiological conditions), since—in the presence of the Fenton reactants Fe2+ and H2O2—the carbonate radical anion (CO3•−) is the active oxidizing product [42]. CO3•− can then migrate over long distances and can reach DNA, where as a potent one-electron oxidant (as OH is also), it can generate base radical cations, which ultimately generate 8-oxo-7,8-dihydroguanine (since guanine is the most oxidatively sensitive base in DNA [43]).
However, OH can be produced even without transition metal mediation. That is, it can be produced either via the oxidation of the ascorbate monoanion (AH) by H2O2 (Table 1, reaction 5) [44] or by a sequence of reactions involving the reactive nitrogen species (RNS) nitric oxide radical (NO). The latter, upon reaction with O2•−, produces the biotoxic RNS ONOOH (Table 1, reaction 6), the decomposition of which, besides OH, produces the RNR radical nitrogen dioxide (NO2) (Table 1, reaction 7) [45,46].
Considering O2•− and H2O2 balance in vivo, it has been suggested that cell survival is promoted by an intracellular increase of O2•−, which is a conjugated base that sustains a cytosolic pH that is mildly alkaline. On the other hand, apoptosis is promoted by intracellular H2O2 production accompanied by a decrease in O2•− levels and concomitant cytosolic acidification, while an excessive production of both O2•− and H2O2 induces necrotic cell death [47].

4. EMF-Induced OS, and Biological Effects

The primary biological effect of EMF involves increased OS, which is demonstrated by the elevated levels of ROS/RNS in exposed experimental models. EMFs have been reported to be associated with many downstream effects that can lead to cancer. These involve DNA damage (single/double strand DNA breaks, oxidized DNA bases); increased ornithine decarboxylase, NF-kappa B, tumour promotion (via degradation of gap junction proteins, and DNA breaks/gene amplification), tissue invasion/metastasis (via increased tight junction protein degradation), CaMKII (via protein oxidation), cellular oncogene transcription (by Ca2+), and CaMKII (via produced Ca2+); lowered melatonin levels; activation of Ca2+-dependent phosphatidylserine flippase and c-src (by calcium/calmodulin); and calpain activation of tumour migration, tissue invasion, and metastasis [7].
Through these events, several indirect changes may occur that can alter the physiology, e.g., of brain cells, and can cause translation–transcription interference [48] (through protein conformation changes [49]), cellular metabolism dysfunction, and membrane dyspermeability [50]. The association of the plasma membrane and mitochondrial cation channels with OS-induced damage and their potential participation in pathological liver conditions has been reviewed elsewhere [11].
It has been proposed that RF EMFs invoke a mechanism that involves the rapid activation of ERK/MAPKs (mitogen-activated protein kinase), which is mediated in the plasma membrane (of Rat1 and HeLa cells) by NADH oxidase, followed by a rapid generation of ROS [21]. In this mechanism, ROS accumulation-induced OS leads to a signal transduction pathway involving ERK kinase activation [21,51], cation channel disturbance [21,52], heat shock protein activation [21,48,53], and enzyme conformation change [54]. This, in turn, may affect gene overexpression/suppression [53], possibly through transcription factor activation/deactivation [55], which is random, given the non-targeted impact of EMF. As an end result, stress response related functions may be triggered, including the altered expression of proteins, some of which are related to neural plasticity, whereas others are involved in general metabolic processes [56].
Recent evidence indicates that ROS/RNS-induced OS is among the main intracellular signal transducers, sustaining lysosomal autophagy and nuclear DNA damage response [57,58]. In general, DNA base damage by ROS involves the formation of single lesions in the pyrimidine and purine bases, intra/inter-strand cross-links, purine 5′,8-cyclonucleosides, and DNA-protein adducts formed by the reactions of the 2-deoxyribose moiety and/or the nucleobases with ROS such as singlet oxygen (1O2), OH, and HOCl [59]. Even the oxidant role of the ROS singlet oxygen, 1O2, on DNA damage was suggested three decades ago [60], and its physiological consequences have been highlighted since 1O2 is produced in human tissues and human neutrophils [61]. However, OH is considered as the most potent oxidant of DNA, as verified by extensive studies on the oxidative modifications on DNA bases, strand breakage, and DNA structure [62,63,64]. Additionally, ROS production by mitochondria leads to mtDNA exposure to OS and subsequent damage (being higher compared to that of nucleic DNA [65]), the accumulation of which leads to further increased production of ROS/RNS and damaged mitochondria. The elimination of the latter by lysosomes, so called mitophagy, has been reported to take place both in yeasts and mammals [66]. The mechanisms of oxidative DNA damage leading to mutations and disease are also well known [67,68], and OS-induced DNA damage [69,70] has been reviewed elsewhere [71].
Experimental studies on RF EMF-induced OS effects in various cellular systems, molecules or model organisms have been reviewed by Yakymenko et al. 2016 [72]. Indicative ones are on sperm [73,74] and Drosophila [75] (body/ovaries [76]). Near-field EMF GSM (at 900 MHz, ‘‘modulated’’ via speaking vs. nonspeaking emission) decrease the reproductive capacity of Drosophila by 50–60% vs. 15–20% (after exposure for 6 min/day during only the first 2–5 days of adult life) [77], the oogenesis of which has been proposed as a biomarker for EMF sources [78]. Extending research in the chronic whole body exposure of mice Balb/c to EMF (GSM at 900 MHz for 3 h/day and by wireless DECT base for 8 h/day), proteome changes (overexpression/downregulation) were recorded in the frontal lobe, hippocampus, and cerebellum [56]. More distinct changes in all of these brain regions were those for synapsin-2 and NADH dehydrogenase, while for some others are indicative of OS increase in the nervous system [79]. In human neuroblastoma cells, low-level GSM EMFs cause alterations on Amyloid Precursor Protein processing and cellular topology, and changes in monomeric alpha-synuclein accumulation and multimerization, which can happen concurrently by means of the induction of OS and cell death, which are possibly linked to Alzheimer’s and Parkinson’s diseases [80]. Neurological abnormalities by RF EMF (GSM) are extended to effects on transient and cumulative memory impairments [81] and on short-term memory in mice (by impairing them to pass successfully the Object Recognition Task [82]), possibly due to disturbance of cation channels, particularly that of Ca2+ (as also suggested by the EMF effect on the calcium binding protein [83]), and to proteome expression changes in the mouse brain hippocampus and other memory-related brain regions [56].

5. Conclusions

On the basis of the above findings, an EMF mechanism can involve ROS formation due to membrane and voltage-gated cation channel function deterioration [2,3,7,8] followed by stress activation and heat-shock protein overexpression [56], which may be associated with behavioural and physiological effects such as blood–brain barrier disruption, memory malfunction, changes in gene expression [53], autophagy, apoptosis [53,84] (especially due to modulation [85]), lifespan reduction, DNA damage, and cancer [18].

6. Theory and Research Perspectives for a Conclusive Linking of EMFs with ROS/RNS

EMF induction of OS via increased concentration of free radicals, has been challenged (by ICNIRP) mainly due to (i) the claimed non-ionizing nature of EMF (ELF/RF), where no covalent bonds are broken at non-thermal intensities, or so the argument goes, and because (ii) the measurement of OS is performed by non-specific methods. Indeed, OS is measured, either by methods that are not specific to the identification of generated free radicals, or indirectly by certain oxidative modifications they cause on key biological molecules (e.g., DNA damage, lipid/protein peroxidation, etc.).
Man-made EMFs do not possess high enough energy to generate free radicals, e.g., on freely moving single H2O molecules by a single photon. However, the individual EMF of such photons are fully synchronized (in terms of frequency, polarization, phase, and propagation direction), thereby producing cumulative macroscopic electric and magnetic fields and electromagnetic radiation (EMR). Nonetheless, these may be additively high enough to break covalent bonds and may directly generate free radicals. Secondly, concentrations of naturally occurring free radicals can increase by the prevention of either (i) reactions between them (e.g., appearing as the aforementioned reactants NO + O2•− and products OH + NO2), or (ii) the reassociation of free radical pairs generated enzymically as transition states. Such prevention can be assisted by the EMF-induced free radical pair mechanism [86,87]. Here, EMFs can prevent the reassociation of free radical pairs by reversing the spin direction of the single electron in one of these free radicals by flipping the direction of its magnetic field component. Thus, the magnetically affected free radical pair ends up consisting of two free radicals, the electron spins of which have become parallel, thereby preventing their re-binding and indirectly increasing their concentration. The free radical pair mechanism has been accounted for by the International Agency for Research on Cancer for the classification of the RF EMFs in the Group 2B category of “possibly carcinogenic to humans” [88]. These two mechanisms of free radical-concentration increase corroborate with the preliminary finding that ELF EMFs increase the concentration of O2•− by many fold in various organs of mice exposed to the ICNIRP limit of 100 µT at 50 Hz (pending publication by Dr. Georgiou’s lab).
Therefore, methods for the in vivo specific detection of the key biological free radicals OH and O2•− ([89,90]) are needed in order to unequivocally prove the generation of carcinogenic OS by EMFs.

Author Contributions

Conceptualization, C.D.G.; validation, C.D.G., L.H.M.; writing—original draft preparation, C.D.G.; writing—review and editing, C.D.G., L.H.M.; visualization, C.D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

This work was supported by the Greek Ministry of Education.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schuermann, D.; Mevissen, M. Manmade electromagnetic fields and oxidative stress—Biological effects and consequences for health. Int. J. Mol. Sci. 2021, 22, 3772. [Google Scholar] [CrossRef]
  2. Panagopoulos, D.J.; Karabarbounis, A.; Margaritis, L.H. Mechanism for action of electromagnetic fields on cells. Biochem. Biophys. Res. Commun. 2002, 298, 95–102. [Google Scholar] [CrossRef] [Green Version]
  3. Panagopoulos, D.J.; Margaritis, L.H. Theoretical considerations for the biological effects of electromagnetic fields. In Biological Effects of Electromagnetic Fields; Stavroulakis, P., Ed.; Springer Science & Business Media: Heidelberg, Germany; London, UK, 2003. [Google Scholar]
  4. Kıvrak, E.G.; Yurt, K.K.; Kaplan, A.A.; Alkan, I.; Altun, G. Effects of electromagnetic fields exposure on the antioxidant defense system. J. Microsc. Ultrastruct. 2017, 5, 167–176. [Google Scholar] [CrossRef]
  5. Pall, M.L. Scientific evidence contradicts findings and assumptions of Canadian Safety Panel 6: Microwaves act through voltage-gated calcium channel activation to induce biological impacts at non-thermal levels, supporting a paradigm shift for microwave/lower frequency electromagnetic field action. Rev. Environ. Health 2015, 30, 99–116. [Google Scholar]
  6. Pall, M.L. Electromagnetic fields act similarly in plants as in animals: Probable activation of calcium channels via their voltage sensor. Curr. Chem. Biol. 2016, 10, 74–82. [Google Scholar] [CrossRef]
  7. Pall, M.L. How cancer can be caused by microwave frequency electromagnetic field (EMF) exposures: EMF activation of voltage-gated calcium channels (VGCCs) can cause cancer including tumor promotion, tissue invasion and metastasis via 15 mechanisms. In Mobile Communications and Public Health; Markov, M., Ed.; CRC Press: Boca Raton, FL, USA, 2018. [Google Scholar]
  8. Pall, M.L. 5G Risk: The Scientific Perspective; The 5G Summit: Washington, DC, USA, 2020. [Google Scholar]
  9. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Membrane Transport of Small Molecules and the Electrical Properties of Membranes-Molecular Biology of the Cell, 4th ed.; Garland Science: New York, NY, USA, 2002. [Google Scholar]
  10. Kiselyov, K.; Muallem, S. ROS and intracellular ion channels. Cell Calcium 2016, 60, 108–114. [Google Scholar] [CrossRef] [Green Version]
  11. Ramírez, A.; Vázquez-Sánchez, A.Y.; Carrión-Robalino, H.; Camacho, J. Ion channels and oxidative stress as a potential link for the diagnosis or treatment of liver diseases. Oxid. Med. Cel. Longev. 2016, 2016, 3928714. [Google Scholar] [CrossRef] [Green Version]
  12. Patel, R.; Sesti, F. Oxidation of ion channels in the aging nervous system. Brain Res. 2016, 1639, 174–185. [Google Scholar] [CrossRef] [Green Version]
  13. Bachmann, M.; Li, W.; Edwards, M.J.; Ahmad, S.A.; Patel, S.; Szabo, I.E.G. Voltage-gated potassium channels as regulators of cell death. Front. Cell Dev. Biol. 2020, 8, 611853. [Google Scholar] [CrossRef]
  14. Bezanilla, F.; White, M.M.; Taylor, R.E. Gating currents associated with potassium channel activation. Nature 1982, 296, 657–659. [Google Scholar] [CrossRef]
  15. Liman, E.R.; Hess, P.; Weaver, F.; Koren, G. Voltage-sensing residues in the S4 region of a mammalian K+ channel. Nature 1991, 353, 752–756. [Google Scholar] [CrossRef]
  16. Panagopoulos, D.J.; Messini, N.; Karabarbounis, A.; Philippetis, A.L.; Margaritis, L.H. A mechanism for action of oscillating electric fields on cells. Biochem. Biophys. Res. Commun. 2000, 272, 634–640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Panagopoulos, D.J.; Johansson, O.; Carlo, G.L. Polarization: A key difference between man-made and natural electromagnetic fields, in regard to biological activity. Sci. Rep. 2015, 5, 14914. [Google Scholar] [CrossRef] [Green Version]
  18. Panagopoulos, D.J. Comparing DNA damage induced by mobile telephony and other types of man-made electromagnetic fields. Mutat. Res. 2019, 781, 53–62. [Google Scholar] [CrossRef] [PubMed]
  19. Panday, A.; Sahoo, M.K.; Osorio, D.; Batra, S. NADPH oxidases: An overview from structure to innate immunity-associated pathologies. Cell. Mol. Immunol. 2015, 12, 5–23. [Google Scholar] [CrossRef] [Green Version]
  20. Hurst, J.K. What really happens in the neutrophil phagosome? Free Radic. Biol. Med. 2012, 53, 508–520. [Google Scholar] [CrossRef] [Green Version]
  21. Friedman, J.; Kraus, S.; Hauptman, Y.; Schiff, Y.; Seger, R. Mechanism of short-term ERK activation by electromagnetic fields at mobile phone frequencies. Biochem. J. 2007, 405, 559–568. [Google Scholar] [CrossRef] [PubMed]
  22. Wyatt, C.N.; Weir, E.K.; Peers, C. Diphenylamine iodonium blocks K+ and Ca2+ currents in type I cells isolated from the rat carotid body. Neurosci. Lett. 1994, 172, 63–66. [Google Scholar] [CrossRef]
  23. Pall, M.L. Electromagnetic fields act via activation of voltage-gated calcium channels to produce beneficial or adverse effects. J. Cell Mol. Med. 2013, 17, 958–965. [Google Scholar] [CrossRef]
  24. DeCoursey, T.E. Interactions between NADPH oxidase and voltage-gated proton channels: Why electron transport depends on proton transport. FEBS Lett. 2003, 555, 57–61. [Google Scholar] [CrossRef] [Green Version]
  25. DeCoursey, T.; Morgan, D.; Cherny, V. The voltage dependence of NADPH oxidase reveals why phagocytes need proton channels. Nature 2003, 422, 531–534. [Google Scholar] [CrossRef] [PubMed]
  26. DeCoursey, T.E. Voltage-gated proton channels find their dream job managing the respiratory burst in phagocytes. Physiology 2010, 25, 27–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Musset, B.; Cherny, V.V.; Morgan, D.; DeCoursey, T.E. The intimate and mysterious relationship between proton channels and NADPH oxidase. FEBS Lett. 2009, 583, 7–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Ramsey, I.S.; Ruchti, E.; Kaczmarek, J.S.; Clapham, D.E. Hv1 proton channels are required for high-level NADPH oxidase-dependent superoxide production during the phagocyte respiratory burst. Proc. Natl. Acad. Sci. USA 2009, 106, 7642–7647. [Google Scholar] [CrossRef] [Green Version]
  29. Seredenina, T.; Demaurex, N.; Krause, K.H. Voltage-gated proton channels as novel drug targets: From NADPH oxidase regulation to sperm biology. Antioxid. Redox Signal. 2015, 23, 490–513. [Google Scholar] [CrossRef] [Green Version]
  30. Rada, B.K.; Geiszt, M.; Hably, C.; Ligeti, E. Consequences of the electrogenic function of the phagocytic NADPH oxidase. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2005, 360, 2293–2300. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  31. Henderson, L.M. NADPH oxidase subunit gp91phox: A proton pathway. Protoplasma 2001, 217, 37–42. [Google Scholar] [CrossRef]
  32. O’Rourke, B.; Cortassa, S.; Aon, M.A. Mitochondrial ion channels: Gatekeepers of life and death. Physiology 2005, 20, 303–315. [Google Scholar] [CrossRef] [Green Version]
  33. O-Uchi, J.; Ryu, S.Y.; Jhun, B.S.; Hurst, S.; Sheu, S.S. Mitochondrial ion channels/transporters as sensors and regulators of cellular redox signaling. Antioxid. Redox Signal. 2014, 21, 987–1006. [Google Scholar] [CrossRef] [Green Version]
  34. Akbarali, H.I. Oxidative stress and ion channels. In Systems Biology of Free Radicals and Antioxidants; Laher, I., Ed.; Springer: Berlin/Heidelberg, Germany, 2014. [Google Scholar]
  35. Sahoo, N.; Hoshi, T.; Heinemann, S.H. Oxidative modulation of voltage-gated potassium channels. Antioxid. Redox Signal. 2014, 21, 933–952. [Google Scholar] [CrossRef] [Green Version]
  36. Fenton, H.J.H. Oxidation of tartaric acid in presence of iron. J. Chem. Soc. Trans. 1894, 65, 899–911. [Google Scholar] [CrossRef] [Green Version]
  37. Phillips, J.L.; Singh, N.P.; Lai, H. Electromagnetic fields and DNA damage. Pathophysiology 2009, 16, 79–88. [Google Scholar] [CrossRef]
  38. Haber, F.; Weiss, J. The catalytic decomposition of hydrogen peroxide by iron salts. Proc. Roy. Soc. A 1934, 147, 332–351. [Google Scholar]
  39. Barb, W.G.; Baxendale, J.H.; George, P.; Hargrave, K.R. Reactions of ferrous and ferric ions with hydrogen peroxide. Part I.—The ferrous ion reaction. Trans. Faraday Soc. 1951, 47, 462–500. [Google Scholar] [CrossRef]
  40. Barb, W.G.; Baxendale, J.H.; George, P.; Hargrave, K.R. Reactions of ferrous and ferric ions with hydrogen peroxide. Part II.—The ferric ion reaction. Trans. Faraday Soc. 1951, 47, 591–616. [Google Scholar] [CrossRef]
  41. Hsieh, Y.H.; Hsieh, Y.P. Kinetics of Fe(III) reduction by ascorbic acid in aqueous solutions. J. Agric. Food Chem. 2000, 48, 1569–1573. [Google Scholar] [CrossRef] [PubMed]
  42. Illés, E.; Mizrahi, A.; Marks, V.; Meyerstein, D. Carbonate-radical-anions, and not hydroxyl radicals, are the products of the Fenton reaction in neutral solutions containing bicarbonate. Free Radic. Biol. Med. 2019, 131, 1–6. [Google Scholar] [CrossRef] [PubMed]
  43. Fleming, A.M.; Burrows, C.J. On the irrelevancy of hydroxyl radical to DNA damage from oxidative stress and implications for epigenetics. Chem. Soc. Rev. 2020, 2020. [Google Scholar] [CrossRef]
  44. Nappi, A.J.; Vass, E. Hydroxyl radical production by ascorbate and hydrogen peroxide. Neurotox. Res. 2000, 2, 1029–8428. [Google Scholar] [CrossRef]
  45. Blough, N.V.; Zafiriou, O.C. Reaction of superoxide with nitric oxide to form peroxonitrite in alkaline aqueous solution. Inorg. Chem. 1985, 24, 3502–3504. [Google Scholar] [CrossRef]
  46. Gunaydin, H.; Houk, K.N. Molecular dynamics simulation of the HOONO decomposition and the HO*/NO2* caged radical pair in water. J. Am. Chem. Soc. 2008, 130, 10036–10037. [Google Scholar] [CrossRef]
  47. Clément, M.-V.; Pervaiz, S. Intracellular superoxide and hydrogen peroxide concentrations: A critical balance that determines survival or death. Redox Rep. 2001, 6, 211–214. [Google Scholar] [CrossRef] [PubMed]
  48. Blank, M.; Goodman, R. A mechanism for stimulation of biosynthesis by electromagnetic fields: Charge transfer in DNA and base pair separation. J. Cell Physiol. 2008, 214, 20–26. [Google Scholar] [CrossRef] [PubMed]
  49. Challis, L.J. Mechanisms for interaction between RF fields and biological tissue. Bioelectromagnetics 2005, 26 (Suppl. 7), 98–106. [Google Scholar] [CrossRef]
  50. McNamee, J.P.; Chauhan, V. Radiofrequency radiation and gene/protein expression: A review. Radiat. Res. 2009, 172, 265–287. [Google Scholar] [CrossRef]
  51. Lee, K.S.; Choi, J.S.; Hong, S.Y.; Son, T.H.; Yu, K. Mobile phone electromagnetic radiation activates MAPK signaling and regulates viability in Drosophila. Bioelectromagnetics 2008, 29, 371–379. [Google Scholar] [CrossRef] [PubMed]
  52. Minelli, T.A.; Balduzzo, M.; Milone, F.F.; Nofrate, V. Modeling cell dynamics under mobile phone radiation. Nonlin. Dyn. Psychol. Life Sci. 2007, 11, 197–218. [Google Scholar]
  53. Manta, A.K.; Papadopoulou, D.; Polyzos, A.P.; Fragopoulou, A.F.; Skouroliakou, A.S.; Thanos, D.; Stravopodis, D.J.; Margaritis, L.H. Mobile-phone radiation-induced perturbation of gene-expression profiling, redox equilibrium and sporadic-apoptosis control in the ovary of Drosophila melanogaster. Fly 2017, 11, 75–95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Barteri, M.; Pala, A.; Rotella, S. Structural and kinetic effects of mobile phone microwaves on acetylcholinesterase activity. Biophys. Chem. 2005, 113, 245–253. [Google Scholar] [CrossRef] [PubMed]
  55. Kar, P.; Nelson, C.; Parekh, A.B. Selective activation of the transcription factor NFAT1 by calcium microdomains near Ca2+ release-activated Ca2+ (CRAC) channels. J. Biol. Chem. 2011, 286, 14795–14803. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Fragopoulou, A.F.; Samara, A.; Antonelou, M.H.; Xanthopoulou, A.; Papadopoulou, A.; Vougas, K.; Koutsogiannopoulou, E.; Anastasiadou, E.; Stravopodis, D.J.; Tsangaris, G.T.; et al. Brain proteome response following whole body exposure of mice to mobile phone or wireless DECT base radiation. Electromagn. Biol. Med. 2012, 31, 250–274. [Google Scholar] [CrossRef] [PubMed]
  57. Filomeni, G.; De Zio, D.; Cecconi, F. Oxidative stress and autophagy: The clash between damage and metabolic needs. Cell Death Differ. 2015, 22, 377–388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Galati, S.; Boni, C.; Gerra, M.C.; Lazzaretti, M.; Buschini, A. Autophagy: A player in response to oxidative stress and DNA damage. Oxid. Med. Cell. Longev. 2019, 2019, 5692958. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Cadet, J.; Wagner, J.R. DNA base damage by reactive oxygen species, oxidizing agents, and UV radiation. Cold Spring Harb. Perspect. Biol. 2013, 5, a012559. [Google Scholar] [CrossRef] [PubMed]
  60. Kanofsky, J.R. Singlet oxygen production by biological systems. Chem. Biol. Interact. 1989, 70, 1–28. [Google Scholar] [CrossRef]
  61. Onyango, A.N. Endogenous generation of singlet oxygen and ozone in human and animal tissues: Mechanisms, biological significance, and influence of dietary components. Oxid. Med. Cell Longev. 2016, 2016, 2398573. [Google Scholar] [CrossRef] [Green Version]
  62. Balasubramanian, B.; Pogozelski, W.K.; Tullius, T.D. DNA strand breaking by hydroxyl radical is governed by the accessible surface area of the hydrogen atom of the DNA backbone. Proc. Natl. Acad. Sci. USA 1998, 95, 9738–9743. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cadet, J.; Delatour, T.; Douki, T.; Gasparutto, D.; Pouget, J.P.; Ravanat, J.L.; Sauvaigo, S. Hydroxyl radicals and DNA base damage. Mutat. Res. 1999, 1424, 9–21. [Google Scholar] [CrossRef]
  64. Tsunoda, M.; Sakaue, T.; Naito, S.; Sunami, T.; Abe, N.; Ueno, Y.; Matsuda, A.; Takénaka, A. Insights into the structures of DNA damaged by hydroxyl radical: Crystal structures of DNA duplexes containing 5-formyluracil. J. Nucleic Acids 2010, 2010, 107289. [Google Scholar] [CrossRef] [Green Version]
  65. Yakes, F.M.; Van Houten, B. Mitochondrial DNA damage is more extensive and persists longer than nuclear DNA damage in human cells following oxidative stress. Proc. Natl. Acad. Sci. USA 1997, 94, 514–519. [Google Scholar] [CrossRef] [Green Version]
  66. Fujiwara, Y.; Wada, K.; Kabuta, T. Lysosomal degradation of intracellular nucleic acids—multiple autophagic pathways. J. Biochem. 2017, 161, 145–154. [Google Scholar] [CrossRef] [Green Version]
  67. Cooke, M.S.; Evans, M.D.; Dizdaroglu, M.; Lunec, J. Oxidative DNA damage: Mechanisms, mutation, and disease. FASEB J. 2003, 17, 1195–1214. [Google Scholar] [CrossRef] [Green Version]
  68. Barzilai, A.; Yamamoto, K. DNA damage responses to oxidative stress. DNA Repair 2004, 3, 1109–1115. [Google Scholar] [CrossRef]
  69. Lai, H.; Singh, N.P. Single- and double-strand DNA breaks in rat brain cells after acute exposure to radiofrequency electromagnetic radiation. Int. J. Radiat. Biol. 1996, 69, 513–521. [Google Scholar] [CrossRef] [PubMed]
  70. Guler, G.; Tomruk, A.; Ozgur, E.; Seyhan, N. The effect of radiofrequency radiation on DNA and lipid damage in non-pregnant and pregnant rabbits and their newborns. Gen. Physiol. Biophys. 2010, 29, 59–66. [Google Scholar] [CrossRef] [Green Version]
  71. Ruediger, H.W. Genotoxic effects of radiofrequency electromagnetic fields. Pathophysiology 2009, 16, 89–102. [Google Scholar] [CrossRef] [PubMed]
  72. Yakymenko, I.; Tsybulin, O.; Sidorik, E.; Henshel, D.; Kyrylenko, O.; Kyrylenko, S. Oxidative mechanisms of biological activity of low-intensity radiofrequency radiation. Electromagn. Biol. Med. 2016, 35, 186–202. [Google Scholar] [CrossRef]
  73. Agarwal, A.; Desai, N.R.; Makker, K.; Varghese, A.; Mouradi, R.; Sabanegh, E.; Sharma, R. Effects of radiofrequency electromagnetic waves (RF-EMW) from cellular phones on human ejaculated semen: An in vitro pilot study. Fertil. Steril. 2009, 92, 1318–1325. [Google Scholar] [CrossRef] [PubMed]
  74. De Luliis, G.N.; Newey, R.J.; King, B.V.; Aitken, R.J. Mobile phone radiation induces reactive oxygen species production and DNA damage in human spermatozoa in vitro. PLoS ONE 2009, 4, e6446. [Google Scholar] [CrossRef]
  75. Irmak, M.K.; Fadillioğlu, E.; Güleç, M.; Erdoğan, H.; Yağmurca, M.; Akyol, O. Effects of electromagnetic radiation from a cellular telephone on the oxidant and antioxidant levels in rabbits. Cell Biochem. Funct. 2002, 20, 279–283. [Google Scholar] [CrossRef]
  76. Manta, A.K.; Stravopodis, D.J.; Papassideri, I.S.; Margaritis, L.H. Reactive oxygen species elevation and recovery in Drosophila bodies and ovaries following short-term and long-term exposure to DECT base EMF. Electromagn. Biol. Med. 2014, 33, 118–131. [Google Scholar] [CrossRef] [PubMed]
  77. Panagopoulos, D.J.; Karabarbounis, A.; Margaritis, L.H. Effect of GSM 900-MHz mobile phone radiation on the reproductive capacity of Drosophila melanogaster. Electromagn. Biol. Med. 2004, 23, 29–43. [Google Scholar] [CrossRef] [Green Version]
  78. Margaritis, L.H.; Manta, A.K.; Kokkaliaris, K.D.; Schiza, D.; Alimisis, K.; Barkas, G.; Georgiou, E.; Giannakopoulou, O.; Kollia, I.; Kontogianni, G.; et al. Drosophila oogenesis as a bio-marker responding to EMF sources. Electromagn. Biol. Med. 2014, 33, 165–189. [Google Scholar] [CrossRef]
  79. Martin-Romero, F.J.; Gutiérrez-Martín, Y.; Henao, F.; Gutiérrez-Merino, C. The NADH oxidase activity of the plasma membrane of synaptosomes is a major source of superoxide anion and is inhibited by peroxynitrite. J. Neurochem. 2002, 82, 604–614. [Google Scholar] [CrossRef] [Green Version]
  80. Stefi, A.L.; Margaritis, L.H.; Skouroliakou, A.S.; Vassilacopoulou, D. Mobile phone electromagnetic radiation affects Amyloid Precursor Protein and α-synuclein metabolism in SH-SY5Y cells. Pathophysiology 2019, 26, 203–212. [Google Scholar] [CrossRef]
  81. Ntzouni, M.P.; Skouroliakou, A.; Kostomitsopoulos, N.; Margaritis, L.H. Transient and cumulative memory impairments induced by GSM 1.8 GHz cell phone signal in a mouse model. Electromagn. Biol. Med. 2013, 32, 95–120. [Google Scholar] [CrossRef]
  82. Ntzouni, M.P.; Stamatakis, A.; Stylianopoulou, F.; Margaritis, L.H. Short-term memory in mice is affected by mobile phone radiation. Pathophysiology 2011, 18, 193–199. [Google Scholar] [CrossRef]
  83. Maskey, D.; Kim, M.; Aryal, B.; Pradhan, J.; Choi, I.-Y.; Park, K.-S.; Son, T.; Hong, S.-Y.; Kim, S.B.; Kim, H.G.; et al. Effect of 835MHz radiofrequency radiation exposure on calcium binding proteins in the hippocampus of the mouse brain. Brain Res. 2010, 1313, 232–241. [Google Scholar] [CrossRef]
  84. Santini, M.T.; Ferrante, A.; Rainaldi, G.; Indovina, P.; Indovina, P.L. Extremely low frequency (ELF) magnetic fields and apoptosis: A review. Int. J. Radiat. Biol. 2005, 81, 1–11. [Google Scholar] [CrossRef] [PubMed]
  85. Sagioglou, N.E.; Manta, A.K.; Giannarakis, I.K.; Skouroliakou, A.S.; Margaritis, L.H. Apoptotic cell death during Drosophila oogenesis is differentially increased by electromagnetic radiation depending on modulation, intensity and duration of exposure. Electromagn. Biol. Med. 2016, 35, 40–53. [Google Scholar] [CrossRef] [PubMed]
  86. Georgiou, C.D. Oxidative stress-induced biological damage by low-level EMFs: Mechanism of free radical pair electron spin-polarization and biochemical amplification. In Non-Thermal Effects and Mechanisms of Interaction between Electromagnetic Fields and Living Matter; Giuliani, L., Soffritti, M., Eds.; Collegium Ramazzini (European J. Oncology Library): Bologna, Italy, 2010; Volume 5, pp. 63–113. [Google Scholar]
  87. Adams, B.; Sinayskiy, I.; Petruccione, F. An open quantum system approach to the radical pair mechanism. Sci. Rep. 2018, 8, 15719. [Google Scholar] [CrossRef] [PubMed]
  88. IARC, International Agency for Research on Cancer (IARC) Working Group on the Evaluation of Carcinogenic Risks to Humans. Non-Ionizing Radiation, Part 2: Radiofrequency Electromagnetic Fields. In IARC Monographs on the Evaluation of Carcinogenic Risks to Humans; IARC: Lyon, France, 2013; Volume 102. Available online: https://www.ncbi.nlm.nih.gov/books/NBK304630/ (accessed on 16 September 2021).
  89. Georgiou, D.C.; Papapostolou, I.; Patsoukis, N.; Tsegenidis, T.; Sideris, T. An ultrasensitive fluorescent assay for the in vivo quantification of superoxide radical in organisms. Anal. Biochem. 2005, 347, 144–151. [Google Scholar] [CrossRef] [PubMed]
  90. Georgiou, C.D.; Papapostolou, I.; Grintzalis, K. Superoxide radical detection in cells, tissues, organisms (animals, plants, insects, microorganisms) and soils. Nat. Protoc. 2008, 3, 1679–1692. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Superoxide radical (O2) generated by NOX VGC/NOX H+-C gp91phox/NOX-coordinating H+-VGC Hv1, which is dismutated to H2O2, which is then used by myeloperoxidase (MPO) as substrate for the two-electron oxidation of Cl to cytotoxic ROS hypochlorous acid (HOCl).
Figure 1. Superoxide radical (O2) generated by NOX VGC/NOX H+-C gp91phox/NOX-coordinating H+-VGC Hv1, which is dismutated to H2O2, which is then used by myeloperoxidase (MPO) as substrate for the two-electron oxidation of Cl to cytotoxic ROS hypochlorous acid (HOCl).
Ijms 22 10041 g001
Table 1. Reactions involved in the biochemistry of ROS.
Table 1. Reactions involved in the biochemistry of ROS.
Fe2+/Cu+ + H2O2 → Fe3+/Cu2+ + OH + HO        (1)
Fe3+ + O2•− → Fe2+ + O2        (2)
Fe3+ + H2O2 → Fe2+ + H+ + HO2        (3)
2 Fe3+ + AH2 → 2 Fe2+ + A + 2 H+        (4)
AH + H2O2 → A•− + H2O + OH      (5)
NO + O2•− + H+ → ONOOH     (6)
ONOOH → OH + NO2              (7)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Georgiou, C.D.; Margaritis, L.H. Oxidative Stress and NADPH Oxidase: Connecting Electromagnetic Fields, Cation Channels and Biological Effects. Int. J. Mol. Sci. 2021, 22, 10041. https://doi.org/10.3390/ijms221810041

AMA Style

Georgiou CD, Margaritis LH. Oxidative Stress and NADPH Oxidase: Connecting Electromagnetic Fields, Cation Channels and Biological Effects. International Journal of Molecular Sciences. 2021; 22(18):10041. https://doi.org/10.3390/ijms221810041

Chicago/Turabian Style

Georgiou, Christos D., and Lukas H. Margaritis. 2021. "Oxidative Stress and NADPH Oxidase: Connecting Electromagnetic Fields, Cation Channels and Biological Effects" International Journal of Molecular Sciences 22, no. 18: 10041. https://doi.org/10.3390/ijms221810041

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