Next Article in Journal
Multimodality Imaging Findings in Waterhouse-Friderichsen Syndrome: A Case Report and a Short Literature Review
Previous Article in Journal
Investigation into the Optimal Strategy of Radium-223 Therapy for Metastatic Castration-Resistant Prostate Cancer
Previous Article in Special Issue
COVID-19 Update: The Golden Time Window for Pharmacological Treatments and Low Dose Radiation Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Physical Differences between Man-Made and Cosmic Microwave Electromagnetic Radiation and Their Exposure Limits, and Radiofrequencies as Generators of Biotoxic Free Radicals

by
Christos D. Georgiou
1,*,
Electra Kalaitzopoulou
1,
Marianna Skipitari
1,
Polyxeni Papadea
1,
Athina Varemmenou
2,
Vassilios Gavriil
3,
Evangelia Sarantopoulou
3,
Zoe Kollia
3 and
Alkiviadis-Constantinos Cefalas
3,*
1
Department of Biology, University of Patras, 26504 Patras, Greece
2
Department of Medicine, University of Patras, 26504 Patras, Greece
3
National Hellenic Research Foundation, Theoretical and Physical Chemistry Institute, 11635 Athens, Greece
*
Authors to whom correspondence should be addressed.
Radiation 2022, 2(4), 285-302; https://doi.org/10.3390/radiation2040022
Submission received: 31 July 2022 / Revised: 13 September 2022 / Accepted: 14 September 2022 / Published: 22 September 2022
(This article belongs to the Special Issue Feature Papers of Radiation 2022)

Abstract

:

Simple Summary

There is an inconsistency between the position that radio frequencies are sources of minor thermal effects in cells, tissues and living organisms and the experimental evidence indicating that non-ionising radiation is harmful, even at shallow power density radiation levels. A quantum mechanical survey of the interaction between microwaves and matter points to free radical-associated cytotoxic alterations of biomatter upon microwave irradiation.

Abstract

The critical arguments for radiofrequency radiation exposure limits are currently based on the principle that radio frequencies (RF) and electromagnetic fields (EMFs) are non-ionising, and their exposure limits are even 100-fold lower than those emitted from the Sun in the whole RF-EMF spectrum. Nonetheless, this argument has been challenged by numerous experimental and theoretical studies on the diverse biological effects of RF-EMF at much lower power density (W/m2) levels than today’s exposing limits. On the other hand, less attention has been given to counterarguments based on the differences in the physics concepts underlying man-made versus natural electromagnetic radiation (EMR) and on the fact that man’s biology has been adapted to the natural EMR levels reaching Earth’s surface at single EMF wavelengths, which are the natural limits of man’s exposure to EMFs. The article highlights the main points of interaction of natural and man-made radiation with biomatter and reveals the physical theoretical background that explains the effects of man-made microwave radiation on biological matter. Moreover, the article extends its analysis on experimental quantum effects, establishing the “ionising-like” effects of man-made microwave radiation on biological matter.

1. Introduction

Electromagnetic radiation (EMR) emits and transmits energy waves travelling through a vacuum or matter from packets of different frequency (energy) photons. Oscillating electric charges mainly generate EMR. The flow of waves in Space creates local and time-changing magnetic and electric fields. Many researchers argue that man-made EMR is potentially biologically harmful [1,2,3] because it differs from natural radiation (mainly originating from the Sun). Natural EMR is filtered mainly by Earth’s atmosphere, allowing man’s biology to adapt to shallow levels of radiation reaching the Earth’s surface. Natural EMR is electromagnetic (EM) waves covering a broad spectral range of wavelengths (mainly λ > 0.01 nm), including the UV, visible (the Sun’s spectral power density maximum emission), infrared, X- and gamma-rays. Furthermore, EMR is emitted by Earth’s electric fields (developing between the ionosphere and Earth’s surface) and from the Schumann resonances, the ultra-low frequency (7.83 Hz) window [4]. Besides the Sun, visible light is also emitted from flames and artificial lamps, such as the light-emitting diodes (LED) used by Visible Light Communications (VLC) and Optical Fibre Communications (OFC) technologies.
Earth’s magnetosphere and ionosphere are protective shields against the deadly cosmic radiation (fast-moving electrons) coming from Space. Electromagnetic fields (EMFs) are also blocked by the Earth’s atmosphere at microwave frequencies. As a result, only shallow levels, similar to 2G-5G man-made radiation sources emitted by the Sun, reach Earth’s surface. In contrast, Earth’s atmosphere transmits visible and infrared light and Schumann resonances, suggesting that human biology might be incompatible with the anthropogenic RF-EMF [5,6,7], because humans and other organisms have adapted well to the cosmic radiation background by synchronizing their biological clocks accordingly [8]. Natural evolution permits humans to tolerate and use visible photons to regulate the melatonin cycle or synthesise vitamin D from UV-A radiation as our eyes have adapted to visible light. Moreover, specific frequencies of natural EMFs are exploited by trillions of cells within the human body for intercommunications, e.g., the waves α (8–12 Hz), β (13–30 Hz), Δ (1–4 Hz) and θ (4–8 Hz) are used for heartbeat regulation, the nervous system’s neural signals, and electrical activities of the brain. Naturally, the evolution of man’s biology has adapted only to the cosmic EMFs that reach Earth’s surface.
Man’s exposure shifted from the cosmic background to the man-made EMR environment in 1882 when the first man-made EMFs started emitting from the high voltage cables of the first commercial power plant in the U.S., the Pearl Street Station plant, which powered lower Manhattan [9]. Today, the global expansion of wireless communication—a development of military applications since the 1960s—exposes the global community to RF-EMF (2G–5G) radiation many orders of magnitude above the natural exposure limit at a single wavelength. The following facts support the argument:
According to Planck’s law, the power density I (ν, T) per solid angle (Sr) and per frequency ν (Hz) emitted from the Sun’s surface, approximated as a black body, depends on the frequency of the emitted radiation and the Sun’s surface temperature T ~ 5.5 × 103 K, and is given by,
I ( v , T ) = 2 h ν 3 c 2 1 e h ν Κ Τ 1   ( Δ ν 2 )      
where h = 6.626 × 10−34 Js is Planck’s constant, K is Boltzmann’s constant equal to 1.39 × 10−23 JK−1, Δ ν 2 is the bandwidth at full width at half-maximum (FWHM) at the emitted frequency ν. For the conditions applied at the Sun’s surface, it is plausible to consider Δν to be the thermal Doppler broadening at the emitted frequency,
Δ ν = 8 K T l n 2 m c 2 ν          
where m is the helium atom mass equal to 1.67 × 10−27 Kg, and c = 3 × 108 ms−1 is the speed of light. Finally, Δν = 5.2 × 10−3 ν. From Equations (1) and (2), the cosmic EMR microwave background energy density at, e.g., 2.4 GHz, emitted from the Sun’s surface per solid angle is ~10−11 Wm−2 Sr−1. For an average Sun-to-Earth distance r = 1.5 × 1011 m and a Sun’s radius R = 6.7 × 108 m, the average solid angle Ω of the Sun viewed from the Earths distance r (Ω = πr2/R2) is ~6.26 × 10−5 Sr. From Equations (1) and (2), the radiation power density from the Sun reaching the Earth’s surface at 2.4 GHz is ~10−16 W/m−2 (~10−17 mW/cm2). The Sun’s theoretically calculated value of radiation power density at microwave frequencies reaching the Earth at a point near the Earth’s orbit in Space is seven orders of magnitude higher than NASA’s measured power density on the Earth’s surface at 2.4 GHz [10], which is equal to ~10−23 W/m2 (~10−24 mW/cm2) (Figure 1). The difference between the measured and theoretical value of power density is possibly due to radiation scattering by cosmic particles, small-size bodies and plasma in the Space area between the Sun and the Earth. The experimentally measured natural microwave levels for the non-ionising cosmic EMF spectrum associated with the 2G–5G wireless frequencies (~0.8 to 12 GHz) at the Earth’s surface lay between ~10−22 and 10−20 W/m2 (~10−23 to 10−21 mW/cm2), Figure 1. The above cosmic microwave radiation levels at the specific frequencies are biologically safe because of the adaption of organisms through natural evolution over billions of years on Earth’s surface. These natural cosmic radiation levels, being the natural biological exposure limits, can be compared to the 10 W/m2 (1 mW/cm2 or 104 mW/m2) microwave and radiofrequency exposure limit set by the International Commission on Non-Ionizing Radiation Protection (ICNIRP) and the U.S. Federal Communications Commission (FCC), at 30 min for whole body exposure and 6 min for local exposure for not developing any thermal damage [11,12]; further information can be found elsewhere [13]. Finally, the conclusion is that for the whole frequency spectrum from 2G to 5G, the ICNIRP/FCC exposure limit recommendation for man-made EMR is ~1021 to 1023-fold higher than the average radiation background of the Sun’s non-ionising radiation at Earth’s surface in the 2G–5G spectral range.
The global community’s RF-EMF radiation exposure issues can be practically understood using the EMF emission power density of a typical Global System for Mobile Communications (GSM) cellular tower base station [14,15]. Taking the least harmful exposure scenario for a citizen at a distance of 1 Km from a GSM tower antenna, the average exposure radiation power density is 10−6 W/m2 (10−7 mW/cm2) and 2 × 10−7 W/m2 (2 × 10−8 mW/cm2) when in line and not in line, respectively [14]. Even in this most favourable case, the radiation levels are 1013 to 1015-fold higher than the natural exposure EMF limit. Furthermore, microwave radiation’s thermal and non-thermal interaction with materials is still debatable. Up to now, non-thermal effects are wrongly and misleadingly viewed as being caused only by ionisation (where an atom/molecule acquires a negative/positive charge by gaining/losing electrons), ignoring that microwave frequencies generate free radicals from molecular bond braking via dissociative excited electronic states. In the case of water, experimental results indicate long-term changes in the structure of water after the microwave treatment. The mechanism of this effect today is not entirely clear, although several factors rendered, including the formation of hot spots, development, and the increased dipole moment of the reacting molecules in the transition state compared to the initial state [16,17].
In this communication, biological effects from microwave radiation are viewed under the light of non-thermal microwave interaction with the matter based on Chirikov ionisation localisation [18], microwave absorption and microwave trapping between rotational molecular levels [19], and quantum microwave trapping between coherent symmetric and antisymmetric water cluster states [20,21,22,23]. The theoretical and experimental background emphasises the discrepancy of microwave exposure levels between the cosmic power density of microwave radiation emitted mainly from the Sun and reaching the Earth’s surface and today’s existing radiation exposure limits, designed to protect from thermal effects.

2. Man-Made versus Natural EMF Radiation and Corresponding Microwave Power Density Exposure Limits

Frequency v or wavelength λ characterises EMF waves and photons, and their energy is proportional to their frequency according to Plank’s formula E = hν, where h is Planck’s constant. Εach photon has its own electric and magnetic fields, the intensities of which are at a fixed ratio, and their vectors are perpendicular to each other and the direction of the photon’s propagation at far distances from the emitting sources. In addition, each photon is polarised in the direction of the electrical field vector. Τhe physical characteristics of each photon define its dual character as a photon and an EM wave.
EM waves (photons) emitted from natural sources (Sun) are uncorrelated. They exist as independent and incoherent EM waves because their photons’ phases are not linked. The Sun’s total radiation power density reaching Earth’s surface in the whole emission spectral frequency range is ~1400 W/m2 (140 mW/cm2) (in Western, Central, and Eastern Europe around noon and a clear sky [24]), and is scaled proportionally to the number of photons. In contrast, the energy of an artificial (man-made) coherent EM beam falling on a surface is scaled as the square of the number of photons compared with the linear scaling of an incoherent photon source [25].
The Sun’s EMFs reaching the Earth’s surface penetrate human and animal skin, and individual EMR photons are partly reflected, absorbed and converted to heat. The worst-case scenario for humans from the Sun’s radiation reaching the Earth is melanoma after prolonged exposure to the UV part of the Sun’s radiation.
Man-made EMFs might also affect biological activity because their physical properties are quite different from natural EMF/EMR [26,27]. Artificial “non-ionizing” EMFs, including those emitted by mobile phone base stations, WiFi, etc., mainly cover the spectral range from 300 kHz to 30 GHz. Additionally, the low-frequency (LF) range waves from 3 to 3 × 103 Hz (105 to 10 Km) are generated by oscillating currents in metal wires (e.g., antennas, power lines). Artificial EMFs are emitted as near-field and far-field radiating modes (with a short transient transmission interval in between) according to the ratio of the distance to the wavelength.
The near-field is the EMF segment emitted from an antenna or a power line wire at a distance equal to one wavelength. The near-field itself is further divided into the reactive near-field and the radiative near-field, both fractional functions of wavelength. The reactive near-field is extended at a distance of 0.159 λ from the antenna. The radiative near-field (Fresnel region) covers the remainder of the near-field region at a distance of one wavelength. Near-field RF-EMF is a complicated space mathematical multipole structure of electric and magnetic fields, mutually independent with no fixed ratio intensities. For the 3G–4G systems, a transmission distance equal to one wavelength follows the radiative near-field length [28,29].
In contrast, far-field dipole-type EMFs produce a fixed phase relationship between electric and magnetic components [28,29]. Consequently, RF near-fields do not emit standard coherent or incoherent EMFs. However, a phase locking of individual waves and possibly non-linear photon interactions with matter make near-fields potentially harmful to man [1]. For example, in the 3G (1.8–2.5 GHz) and 4G (2–8 GHz) wireless communications RF bands, the carrier frequencies range from just under 1 GHz to just over 2 GHz (above 0.3 to below 0.15 m, respectively), and the average near-field length is 0.225 m. Therefore, when someone holds a smartphone next to its ear or a laptop/tablet at a distance less than 0.225 m, the head and some more sensitive body parts are exposed to the near-field, and, possibly, humans experience harmful biological effects [30,31,32].
In the Extremely Low Frequency (ELF) spectrum, the most predominant frequencies are those emitted by the high voltage power lines acting as antennas of alternating currents at 50 or 60 Hz, frequencies that correspond to near-field lengths of 6 × 106 or 5 × 106 m, respectively. Therefore, the power lines’ electric and magnetic field components behave independently for long distances and bear variable intensity ratios. When no power is consumed, the power lines emit constant electrical fields. However, as the current flows across the power lines at a variable strength (because of variable power consumption), it generates a variable strength magnetic field (proportional to the fluctuating current intensity). Nonetheless, in the range of near-field distances, the electric and magnetic fields act independently for possible biological effects [33,34]. ELF electric fields penetrate living tissues with a certain degree of attenuation, and magnetic fields penetrate with almost zero attenuation. Based on statistical and laboratory shreds of evidence, the International Agency for Research on Cancer (IARC) has classified both ELF magnetic fields and RF-EMFs as possibly carcinogenic to humans [33,35].
It is commonly claimed that the small electric absorption depth of the human skin prevents electric and magnetic fields from penetrating the body. However, this argument is not physically justified. The electromagnetic field is a dual entity with an electric and a magnetic component, simultaneously created by the time-variable electric charges and currents. However, the elaboration of the differential coupling equations of electric and magnetic fields propagated in dielectric, diamagnetic or paramagnetic media suggests that neither Maxwell’s equations nor their solutions indicate continuous and constant causal links between the fields. Instead, they only form a dual entity of an electric and a magnetic component, created by their time-variable electric charges and currents [36]. Along the above theoretical lines, it has also been documented that the magnetic field component of artificial EMFs is biologically active [37,38,39,40,41]. Therefore, the emphasis on electric field intensity (V/m)-based on current recommended radiation exposure limits underestimates the biological effects caused by the cumulative magnetic EMF component that primarily penetrates the human body [42,43]. Furthermore, constant or far-field magnetic field components increase carcinogenic free radicals’ cellular levels via the free radical pair spin flipping mechanism [35,44] and/or via coherent trapping of microwave radiation in rotational levels of water clusters [20,21].
Furthermore, man-made RF-EMF far-fields are produced additively by individual EMF photons as being coherent or partially coherent with each other. RF-EMF far-fields are produced by photons fully synchronised to each other (in a frequency, polarization, phase, pulse, and propagation direction). In contrast to the non-synchronised photons emitted by the Sun, the intensities of man-made individual electric and magnetic fields add up coherently or partially coherent, conditions that make them possibly biologically active since coherent photons are cumulative on a macroscopic scale; thus, they possess the potency to form cores of biological effects over exposing time [42,45,46].
The following section discusses the effect of “Chirikov localisation”, which refers to the microwave ionisation of hydrogen atoms.

3. The Physical Basis of Microwave Interaction with Matter

3.1. Microwave Adiabatic Tunnelling Ionisation

Electric and magnetic fields penetrate through the skin, even with low penetrating skin-depths. Consequently, the RF-EMF induces electrical currents, resonant interactions and interferences on charged/polar and magnetic particles, cell membrane surfaces and free radical pairs, thus leading to potential biological effects [44,47,48]. Similarly, ELF EMF induces voltage differences, currents and magnetic field effects on the membranous components of human cells [48].
The principal counterargument is that RF-EMFs are non-ionizing, and their effects are only thermal. The counterargument is based on the fact that the EMR photons are not energetic enough to break the bonds holding organic molecules together. However, many biological environmental interactions originate from non-ionising electric fields [49,50,51,52,53,54], such as cancer proliferation from radiation-polarised dielectric nanoparticles [47]. Most importantly, the non-ionisation-based argument of “no-biological” effects is weakened by the microwave excitation and ionisation of hydrogen atoms known as microwave adiabatic tunnelling, the most striking and less known effect of the destructive ionic interaction of the “non-ionising” microwave frequencies with the matter. Indeed, today it is commonly known that high-energy ionising UV radiation induces bond breaks in organic and biological matter via excited dissociative electronic molecular states, populated from the ground electronic state via one-photon absorption. For example, the binding energy of the C–C bond is 3.6 eV, and the two carbon molecules can be separated from each other (by covalent bond homolytic split) upon irradiation with one photon at 345 nm [55]. Molecular bond breaking by photons is the reason why UV radiation is biologically harmful [56,57]. At shorter photon wavelengths or higher photon energies, the excess photon energy is compensated as kinetic energy of photofragments, causing localised mechanical damages. In contrast, the common misconception about microwave photons, e.g., at 2.4 GHz, is that they are theoretically unable to break molecular bonds or ionise atoms. However, striking gas ionisation experiments with low amplitude photons at ~1 and ~10 GHz [58] demonstrated that the ionisation of the hydrogen atom exited first with laser light at a principal quantum number n = 69. The excited electronic level with n = 69 is separated from the ionisation continuum of hydrogen by 7.57 × 10−4 eV. Therefore, at 1 (4.134 × 10−6 eV) and 2.4 GHz (9.921 × 10−6 eV), 183 and 76 photons from the electronic level with n = 69 ionise the hydrogen atom, arguing against the idea that microwave interaction with matter is only thermal and, thus, against the thermal basis of the current RF-EMF exposure limits.
Given that adiabatic tunnelling breaks the electron-proton binding energy in the hydrogen atom [58] and that RF-EMF photons can provide this energy in the 2G–5G range, it can be expected that microwave adiabatic tunnelling will provide the cumulative RF photons needed to split the antiparallel spin electron pair holding the O–H bond in H2O of 1.88 eV (117.61 Kcal/mol) [59,60], and generate hydroxyl (OH) and hydrogen (H) free radicals. Therefore, photons at the 2G–5G spectral range could provide free radicals with biological and medical implications for man’s health [61].
Microwave ionisation is viewed as adiabatic quantum mechanical electron tunnelling through a time-varying potential barrier produced by the Coulomb and the external forces of the microwave field exerted on the atom. Electron tunnelling occurs when the conditions T > τ and Ε < Ε0 are satisfied, where T is the ratio of the microwave interaction with the atom, τ is the electron period around the nucleus, and E and E0 are the strengths of the microwave and the atomic Coulomb field, respectively. The adiabatic parameter γ = ν/nE describes the ionisation tunnelling for γ < 1, while the condition of γ > 1 covers any possible combination of parameters. Strong ionisation occurs at relatively weak field amplitudes, several times smaller than the static field ionisation threshold. Delone et al. described the microwave excitation and ionisation by an energy diffusion equation using random phase approximation [62], and along the same lines, Meerson et al. [63] gave a correct estimation of the ionisation threshold based on the Chirikov criterion of resonance [18]. In this approximation, the classical diffusion equation describes the photonic transitions and quantum interference effects count for the localisation length and the number of photons [64,65,66]. Additional references and results for the microwave ionisation of excited atoms can be found in past reviews [67,68,69]. Arndt et al. also provided additional experimental work on Rydberg atoms’ microwave ionisation [70]. The “Chirikov localisation” refers to the discovery and investigation of microwave ionisation by Boris Chirikov [18]. This chaotic diffusion can be localised by quantum interference effects and is similar to the Anderson localisation [71], which appears in disordered solids when a diffusive spreading in space existing for classical trajectories becomes exponentially localised due to quantum interference effects. Chirikov localisation stresses the dynamical origin of this phenomenon emerging without any disorder.

3.2. Absorption of RF-EMF by Biological Matter and Water

Besides microwave tunnelling adiabatic ionisation, a destructive molecular mechanism is the microwave interaction with the rotational levels of molecules or their clusters [20,22,48]. The quantum theory of microwave spectroscopy developed by Townes and Schawlow [19] verifies that the absorption probability of EMF between two rotational molecular states with an energy difference close to the energy of the microwave radiation is higher than the relaxation probability of the excited, energetic quantum state to the lower energy state. For a particular symmetric-top molecule, similar to a water molecule in an initial quantum state J   M , the absorption and emission probabilities γ J , M J + 1 , M ,   γ J + 1 , M J , M between the quantum states J , M J + 1 , M and J + 1 , M J , M are given by the expressions,
γ J , M J + 1 , M = 2 π h 2 N f υ 9 c ( k T ) 2 B π C h k T   μ 2 ( 4 J + 3 ) ( J + 2 ) ( J + 1 ) 2 ν 0 2 ν 2 Δ ν ( ν ν 0 ) 2 + Δ ν 2 + τ 2          
γ J + 1 , M J , M = 2 π h 2 N f υ 9 c ( k T ) 2 B π C h k T   μ 2 ( 4 J + 7 ) ( J + 3 ) ( J + 2 ) 2 ν 0 2 ν 2 Δ ν ( ν ν 0 ) 2 + Δ ν 2 + τ 2          
For an averaging over all the magnetic quantum numbers M of a particular rotational angular momentum state J. J and M are the rotational and magnetic quantum numbers, N is the number of molecules per unit volume, f υ = e ε κ Τ ( 1 e ε κ Τ ) is the fraction of molecules in a particular vibrational state of energy ε = h ω e ( υ + 1 2 ) , k = 1.38 × 10−23 JK−1 is the Boltzmann’s constant, B = h 8 π 2 Ι Β   and C = h 8 π 2 Ι C are the rotational constants, I B , I c are the molecular moments of inertia along the molecular axes x, y and z, respectively, μ 2 is the square of the dipole moment matrix element,   ν 0 2 is the resonant frequency or the central frequency of the absorption line, Δ ν is the half-width of the absorption line at FWHM, c is the velocity of light,   T is the absolute temperature, and τ 2 is the collision dephasing time during a molecular rotation. From Equations (3) and (4), the absorption probability of radiation at microwave frequencies by molecular rotational levels is higher than the emission probability between higher and lower rotational states with similar quantum numbers. This result must be valid to maintain thermal equilibrium when transitions occur, since there are 2 J + 3 states of rotational angular momentum   J + 1 . Therefore, it is evident that microwave electromagnetic energy is stored in rotational states. The storage efficiency falls with the increasing quantum number (Figure 2).
The rotational frequencies of water molecules and clusters lay in the terahertz and microwave spectral bands, respectively. For example, the rotational frequency of water molecules at 2.466454 THz stands for terahertz resonance of water molecules with quantum numbers J = 6 and M = 4. Similarly, the operation of a microwave oven in the microwave spectral region, coming as a result of preferred size (e.g., at 2.4 GHz), is based on the resonance between microwave radiation and the rotational levels of large molecules and water clusters, implying that electromagnetic energy can be stored on rotationally excited quantum states of water clusters. To specify the limits of the energy spectrum for electromagnetic interaction with the rotational levels in the first approximation, we consider that the separation between nuclei in linear rotating molecules is fixed. The possible frequencies of the end-over-end rotation of this “molecular rigid rotor” can be obtained. Using assumptions of the “old” quantum mechanics, the angular momentum of the rotating molecular system must be some integral multiple of h / 2 π , and hence the rotational frequencies expected from the system are,
ν = h 2 π 2 I J ( J + 1 )          
where h = 1.056 × 10−34 Js is Planck’s constant, ν is the frequency of rotation, I   is the molecular moment of inertia about the axes perpendicular to the internuclear axis, and J is a positive integer, giving the angular momentum in h 2 π   units. For the OH free radical, the moment of inertia (I = 1.27 × 10−47 kgm2) comes from the nuclei mass and the molecular separating distance of 9.7 × 10 −11 m. For small integral values J   between 1 and 10, the rotation frequency ν lies between 0.2 and 11 GHz, and therefore the rotational levels of the OH molecules are potentially in resonance with the 2G to 5G frequencies, and EMF energy storage is potentially possible. This strengthens the suggestion, made in sub-Section 3.1, for a plausible OH generation from splitting the O–H bond in H2O by RF-EMF photons in the 2G–5G range.
When exposed to the radiation level of, e.g., 10−3 mW/m2 (10−4 mW/cm2) at 1.8 GHz (104-fold lower than the ICNIRP/FCC limit), emitted by, e.g., a GSM tower located ~100 m far away [14], the number of fully synchronised photons per s and m2 that bombards the human body is ~8.4 × 1020 [72,73]. Therefore, because of the microwave energy-storing effect [19], the biologically harmful effects from antenna power stations can immensely increase for the billions of people exposed every day to about 10 W/m2 (1 mW/cm2), which is today’s microwave exposure limit, or to about a specific absorption rate (SAR) of radiation of 0.125 W m2kg−1 for an 80 kg heavy person. This effect is even more profound when holding a mobile phone against the ear, generating a microwave power density of 1 W/m2 (0.1 W/m2) in the cranium area that is equivalent to a higher SAR value of about 0.222 W m2kg−1 for an average human head weighing 4.5 kg. An even more significant concern is for children’s still-developing small size brains. As Dr. Vriens aptly points out: “Saying that the RF radiation from wireless communication cannot do any harm because the individual photon energies are not large enough is the same as saying that a tsunami cannot cause any harm because the individual water molecules do not have enough energy” [74].

3.3. Microwave Interaction with Symmetric and Antisymmetric Water Cluster States

A plethora of CaCO3 precipitation experiments under the application of external magnetic fields in water flow systems verifies that CaCO3 precipitates in the form of aragonite than calcite, which is the crystal form of CaCO3, at standard conditions and zero magnetic fields. CaCO3 precipitants appear in three different crystal forms, and the one with the lower ground electronic state is the rhombohedral calcite. Aragonite has a lower symmetry and belongs to the orthorhombic crystal group. A less stable form of CaCO3 is the hexagonal vaterite, which undergoes a crystal phase transition to calcite or aragonite. Because the ground electronic energy state of aragonite is ~0.8 eV above the ground electronic state of calcite, CaCO3 precipitates as calcite at room temperature. However, theoretical and experimental results also show that CaCO3 aragonite precipitates at low magnetic fields of 50 mT because the magnetic field modes can become trapped in an ensemble of coherent antisymmetric quantum rotating states of water molecules, in resonance with the magnetic field modes [22]. By applying second quantisation for the magnetic field and the rotating water molecular rotors, where both systems are described as ensembles of interacting vibrations, the interaction Hamiltonian in the dipole moment approximation between the magnetic field and the molecular rotors is given by
H ^ I = g k λ 2 ( (   i ^ j ^ ) σ ^ ω R α ^ κ λ + ( e x p i ( ω k t ) + 2 ( i ^ + i j ^ ) σ ^ ω R α ^ κ λ + ( e x p i ( ω k t )
where α ^ κ λ + , is the creation operator of the magnetic field modes for the polarization λ and mode κ, and σ ^ ω R is the annihilation operator for the molecular rotational mode R at frequency ω. The summation over λ ,   k is for all the molecular rotors and the magnetic field modes. By changing the notation, Equation (6) is simplified to
H ^ I = g λ , k ( σ ^ λ + a ^ k + σ ^ λ a ^ k + )          
At time t = 0, the quantum state of an ensemble of molecular rotors by the wavefunction | Ψ ( 0 ) > . When the magnetic field is switched on, the wavefunction of the system at time t becomes
| Ψ ( τ ) = U ^ ( t , 0 )   | Ψ ( 0 )
where U ^ ( t , 0 ) is the Dyson time ordering operator. The first-order approximation of the operator is
U ^ ( t , 0 ) ( 1 ) = 1 1 0 t d t 1 H I ( t 1 ) = 1 g t   ( σ λ + a ^ k + σ ^ λ a ^ k + )      
The wavefunction | Ψ ( 0 ) is the product of molecular rotor | Ψ ( 0 ) Μ R and magnetic field wavefunctions magnetic field | Ψ ( 0 ) Μ R . The ground and the excited states of one two-level molecular rotor, say the nth rotor, will be | b n and | c n   respectively. If the molecular rotors do not interact, the total wavefunction is
| Ψ ( 0 ) M R = | b 1 | b 2 ] .. | b n = | b 1 b 2 .. b n
For N modes of the magnetic field, and in case the jth mode is in resonance with one molecular rotor, the wavefunction of the MF | Ψ ( 0 ) Μ F is
| Ψ ( 0 ) M F = | 1 .. | j 1 | j + 1 .. | N = | 1 . 2 .. j 1 , 0 , j + 1 , .. N
Because the molecular rotor has absorbed the j mode, the total wavefunction of the system is
| Ψ ( 0 ) = | b 1 b 2 .. b n | 1 . 2 .. j 1 , 0 , j + 1 , .. N
In a first-order approximation at time t, the wavefunction of the system is
| Ψ 1 ( t ) = | Ψ ( 0 ) i g t λ , k ( σ ^ λ + a ^ k + σ ^ λ a ^ k + ) | Ψ ( 0 )        
Allowing for the action of both the creation and annihilation operators, Equation (13) becomes
| Ψ 1 ( t ) = j N A j | b 1 b 2 .. b n | 1 . 2 .. j 1 , 0 , j + 1 , .. N
Equation (14) describes a quantum state where the jth molecular rotor in the ensemble absorbs one magnetic field mode. The coefficients A j are either negative or positive, and therefore | Ψ 1 ( t ) will be either symmetric or antisymmetric. The conclusion is that the first-order perturbation of the system creates a set of symmetric and antisymmetric quantum states of the ensemble of molecular rotors. When Nb rotors are in the ground state and Nc in the excited state, then the molecular rotors will absorb Nc magnetic field modes. In this case, the wavefunction should be approximated to the N c t h order of the U ^ ( t , 0 ) . Allowing for all the existing permutations, the number of states is
| N c = ( N ! N c ! N b ! ) 1 2 p e r m A n ( 1 ) n | b 1 b 2 .. b b c b + 1 .. c N | 1 , 2 , .. b
The coefficient A n ( 1 ) n is a function of the normalisation conditions. If all the coefficients are positive, the states are symmetric. The system is in a Dicke state, and the interaction between the molecular rotors and the magnetic field is similar to the super fluorescent interactions in laser systems. For an even number of negative coefficients, the states are antisymmetric. The decay probability P a b s   of a magnetic field mode absorbed by the jth molecular rotor will be proportional to the matrix element
P a b s = N c 1 | σ ^ j | N c      
If only one of the N c 1 | ,   N c | states is antisymmetric, the transition probability is zero, and therefore the magnetic field mode is trapped between the coherent antisymmetric state and the coherent ground state of the ensemble, in agreement with Equations (3) and (4).
The quantum cyclotronic resonance effect is similar to the ion cyclotronic resonance effect in Zhadin [75]. Therefore, the aragonite precipitation can be interpreted as the magnetic vector quantum states being trapped and amplified between a pair of coherent symmetric and antisymmetric quantum states formed by individual water molecular rotors under the action of the external EMF [20]. The coherent symmetric and antisymmetric states are similar to the theory of coherent domains [21,23].

4. Exposure Limits of Man-Made RF/ELF, EMF Do Not Apply to Near-Fields: Biological Consequences

ICNIRP makes the following clarifications on the RF-EMF exposure limits about near-fields: “The situation in the near-field region is rather more complicated because the maxima and minima of E (electric) and H (magnetic) fields do not occur at the same points along the direction of propagation as they do in the far-field. In the near-field, the electromagnetic field structure may be highly inhomogeneous, and there may be substantial variations from the plane wave impedance of 377 ohms; that is, there may be almost pure E fields in some regions and almost pure H fields in others. Therefore, exposures in the near-field are more difficult to specify because both E and H fields must be measured and because the field patterns are more complicated; in this situation, power density is no longer an appropriate quantity to use in expressing exposure restrictions (as in the far-field).” [76,77,78]. Additionally, ICNIRP notes that for the RF-EMF spectrum range of 5G, it is accepted that “Because the incident power density used for the reference levels above 6 GHz does not appropriately correlate with the absorbed power density used for the basic restrictions in the reactive near-field region, reference levels cannot be used to determine compliance in the reactive near-field” [11]. A culmination of the above statements is the conclusion by ICNIRP that the current exposure limits can be exceeded in RF/ELF EMF near-fields (e.g., emitted by electric appliances and cell phones): “Near-field exposure situations, localised and non-uniform field exposure, are of special interest. Typical EM sources with near-field exposure are hand-held mobile telephones, inductive or capacitive heating equipment, anti-theft devices or electric appliances in homes and workplaces. Such devices can emit localised fields above the reference levels.” [79]. Therefore, billions of individuals worldwide are overexposed daily by the RF-EMFs from their cell phones, tablets, laptops, etc.
The current man-made power density exposure limits are compared to nature’s limits, and some observed biological effects are shown in Figure 3. The biological effects are based on studies presented in the BioInitiative 2012 Report [34,80,81,82,83,84,85,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139], which have been updated in the present study. Nonetheless, the number of such studies has grown since then, listing as indicative ones that focus on oxidative stress [2,44,53,54,106,128,140,141], DNA damage [54,142] and carcinogenesis [143,144,145,146,147].
Indeed, the lowest power density of 1 × 10−9 W/m2 (1 × 10−6 mW/m2) with observed biological effects (oxidative damage, reactive oxygen species generation, DNA damage/repair failure) is ~1013-fold higher than a natural exposure limit of ~1 × 10−22 W/m2 (10−19 mW/m2) of exposure to the cosmic frequencies and 1010-fold lower than the today’s power density exposure limit of 10 W/m2 (1 × 104 mW/m2). Even the 3 to 6 × 10−6 W/m2 (3 to 6 × 10−3 mW/m2) limit proposed by the BioInitiative 2012 Report [148] is 1014- to 1018-fold higher than the natural exposure limit.

5. Conclusions

This article follows a novel approach to reveal the main principles of the microwave interaction of natural and man-made radiation in resonance with biomatter. Quantum theory is applied to elucidate the effects of man-made microwave electromagnetic radiation on biological matter, in reference to the Earth’s surface receiving cosmic radiation levels to which man’s biology has been adapted. For this, the article extends its analysis to experimentally established quantum physics effects, such as microwave adiabatic tunnelling, the coherent interaction of microwaves with the rotational levels of water clusters, and the quantum properties of microwave interactions with water clusters.
The fact that the Sun’s radiation reaching the Earth sets the actual natural exposure limits for man should bring into focus the sunlight-based VLC and OFC technologies, because VLC LEDs emit Sun-like photons that are also incoherent in terms of frequencies, waveforms, and phase difference [149,150], making them non biologically damaging. The latter is also supported by studies showing that VLC LEDs do not cause any damaging effects, even on the eye retina [150,151]. Moreover, VLCs, besides transmitting at speeds far above 5G, have been suggested as being capable of solving the significant challenges of 5G/IoT communication systems [152].

Author Contributions

C.D.G. conceived, managed, and supervised the study, and wrote and edited the draft. E.K., M.S., P.P. and A.V. contributed equally (in bibliography search and draft reviewing). A.-C.C., E.S., V.G. and Z.K. contributed to establish the quantum basis of microwave radiation interaction with matter and edited the draft. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Greek Ministry of Education. Part of the work presented was supported by the PhD Fellowships, (i) “Andreas Mentzelopoulos Foundation” fellowship of the University of Patras, (ii) HFRI PhD Fellowship grant (Fellowship Number 653) of the Hellenic Foundation for Research and Innovation (HFRI). It was also supported by Greece and the European Union (European Social Fund-ESF) through the Operational Program «Human Resources Development, Education and Lifelong Learning» in the context of the Act “Enhancing Human Resources Research Potential by undertaking a Doctoral Research” Sub-action 2: IKY Scholarship Program for PhD candidates in the Greek Universities.

Acknowledgments

Georgiou thanks Leendert Vriens (PhD in Physics, who has worked for 30 years at Philips Research Eindhoven) for personal communications on the differences in physical parameters of photons emitted from natural versus man-made EMFs.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Miller, A.B.; Sears, M.E.; Morgan, L.L.; Davis, D.L.; Hardell, L.; Oremus, M.; Soskolne, C.L. Risks to health and well-being from radio-frequency radiation emitted by cell phones and other wireless devices. Front. Public Health 2019, 7, 223. [Google Scholar] [CrossRef] [PubMed]
  2. 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] [PubMed]
  3. Dasdag, S.; Akdag, M.Z.; Erdal, M.E.; Erdal, N.; Ay, O.I.; Ay, M.E.; Yilmaz, S.G.; Tasdelen, B.; Yegin, K. Effects of 2.4GHz radiofrequency radiation emitted from Wi-Fi equipment on microRNA expression in brain tissue. Int. J. Radiat. Biol. 2015, 91, 555–561. [Google Scholar] [CrossRef]
  4. Price, C. ELF electromagnetic waves from lightning: The Schumann resonances. Atmosphere 2016, 7, 116. [Google Scholar] [CrossRef]
  5. Tinker, R.; Abramowicz, J.; Karabetsos, E.; Magnusson, S.; Matthes, R.; Moser, M.; Niu, S.; O’Hagan, J.; van Deventer, E. A coherent framework for non-ionising radiation protection. J. Radiol. Prot. 2022, 42, 010501. [Google Scholar] [CrossRef]
  6. Schmid, M.R.; Loughran, S.P.; Regel, S.J.; Murbach, M.; Bratic Grunauer, A.; Rusterholz, T.; Bersagliere, A.; Kuster, N.; Achermann, P. Sleep eeg alterations: Effects of different pulse-modulated radio frequency electromagnetic fields. J. Sleep. Res. 2012, 21, 50–58. [Google Scholar] [CrossRef]
  7. Singh, S.; Kapoor, N. Health implications of electromagnetic fields, mechanisms of action, and research needs. Adv. Biol. 2014, 2014, 198609. [Google Scholar] [CrossRef]
  8. König, H.L.; Krueger, A.P.; Lang, S.; Sönnig, W. Biological Effects of Environmental Electromagnetism; Springer: New York, NY, USA, 1981; p. 342. [Google Scholar]
  9. Josephson, M. Edison: A Biography; McGraw Hill Book Co: New York, NY, USA, 1959; p. 255. [Google Scholar]
  10. Raines, J.K. Electromagnetic Field Interactions with The Human Body: Observed Effects and Theories; NASA, Goddard Space Flight Center (NASA CR 166661); NASA: Washington, DC, USA, 1981. [Google Scholar]
  11. ICNIRP. Guidelines for limiting exposure to electromagnetic fields (100 KHz to 300 GHz). Health Phys. 2020, 118, 483–524. [Google Scholar] [CrossRef]
  12. FCC. Resolution of Notice of Inquiry, Second Report and Order, Notice of Proposed Rulemaking, and Memorandum Opinion and Order. U.S. Federal Communications Commission, Washington, DC, FCC19-126a1. Available online: https://docs.fcc.gov/public/attachments/FCC-19-126A1.pdf (accessed on 29 July 2022).
  13. Hardell, L.; Nilsson, M.; Koppel, T.; Carlberg, M. Aspects on the International Commission on Non-Ionizing Radiation Protection (ICNIRP) 2020 guidelines on radiofrequency radiation. J. Cancer Sci. Clin. Ther. 2021, 5, 250–283. [Google Scholar]
  14. Haumann, T.; Münzenberg, U.; Maes, W.; Sierck, P. Hf-Radiation Levels of Gsm Cellular Phone Towers in Residential Areas. In Proceedings of the 2nd International Workshop on Biological effects of EMFs, Rhodes, Greece, 7–11 October 2002; University of Ioannina: Ioannina, Greece, 2002; pp. 327–333. [Google Scholar]
  15. Marinescu, Ι.Ε.; Poparlan, C. Assessment of GSM HF-radiation impact levels within the residential area of Craiova city. Procedia Environ. Sci. 2016, 32, 177–183. [Google Scholar] [CrossRef]
  16. Yakunov, A.V.; Biliy, M.M.; Naumenko, A.P. Long-term structural modification of water under microwave irradiation: Low-frequency raman spectroscopic measurements. Adv. Opt. Technol. 2017, 2017, 5260912. [Google Scholar] [CrossRef] [Green Version]
  17. Leadbeater, N.E. Organic synthesis using microwave heating. In Comprehensive Organic Synthesis, 2nd ed.; Knochel, P., Ed.; Elsevier: Amsterdam, The Netherlands, 2014; pp. 234–286. [Google Scholar]
  18. Frahm, K.M.; Shepelyansky, D.L. Diffusion and localization for the Chirikov typical map. Phys. Rev. E 2009, 80, 016210. [Google Scholar] [CrossRef] [PubMed]
  19. Townes, C.H.; Schawlow, A.L. Microwave Spectroscopy; McGraw-Hill Book Company: New York, NY, USA, 1955. [Google Scholar]
  20. Cefalas, A.C.; Sarantopoulou, E.; Kollia, Z.; Riziotis, C.; Dražic, G.; Kobe, S.; Stražišar, J.; Meden, A. Magnetic field trapping in coherent antisymmetric states of liquid water molecular rotors. J. Comput. Theor. Nanosci. 2010, 7, 1800–1805. [Google Scholar] [CrossRef]
  21. Del Giudice, E.; Preparata, G.; Vitiello, G. Water as a free electric dipole laser. Phys. Rev. Lett. 1988, 61, 1085–1088. [Google Scholar] [CrossRef]
  22. Cefalas, A.C.; Kobe, S.; Dražic, G.; Sarantopoulou, E.; Kollia, Z.; Stražišar, J.; Meden, A. Nanocrystallization of CaCO3 at solid/liquid interfaces in magnetic field: A quantum approach. Appl. Surf. Sci. 2008, 254, 6715–6724. [Google Scholar] [CrossRef]
  23. Del Giudice, E.; Preparata, G.; Fleischmann, M. QED coherence and electrolyte solutions. J. Electroanal. Chem. 2000, 482, 110–116. [Google Scholar] [CrossRef]
  24. Sarbu, I.; Sebarchievici, C. Solar radiation. In Solar Heating and Cooling Systems: Fundamentals, Experiments and Applications; Academic Press: Cambridge, MA, USA, 2017; pp. 13–28. [Google Scholar]
  25. Nicolis, J.S. Dynamics of Hierarchical Systems: An Evolutionary Approach; Springer: Berlin/Heidelberg, Germany, 1986. [Google Scholar]
  26. Levitt, B.B.; Lai, H.C.; Manville, A.M. Effects of non-ionizing electromagnetic fields on flora and fauna, part 1. Rising ambient emf levels in the environment. Rev. Environ. Health 2021, 37, 81–122. [Google Scholar] [CrossRef]
  27. Levitt, B.B.; Lai, H.C.; Manville, A.M. Effects of non-ionizing electromagnetic fields on flora and fauna, part 2 impacts: How species interact with natural and man-made emf. Rev. Environ. Health, 2021; Epub ahead of print. [Google Scholar]
  28. Balanis, C.A. Antenna Theory: Analysis and Design, 3rd ed.; John Wiley & Sons: Hoboken, NJ, USA, 2005; p. 34. [Google Scholar]
  29. Rappaport, T.S. Wireless Communications Principles and Practice, 2nd ed.; Prentice-Hall: Hoboken, NJ, USA, 2010. [Google Scholar]
  30. Hirata, A. Temperature increase in human eyes due to near-field and far-field exposures at 900 MHz, 1.5 GHz, and 1.9 GHz. IEEE Trans. Electromagn. Compat. 2005, 47, 68–76. [Google Scholar] [CrossRef]
  31. Zhang, J.; Sumich, A.; Wang, G.Y. Acute effects of radiofrequency electromagnetic field emitted by mobile phone on brain function. Bioelectromagnetics 2017, 38, 329–338. [Google Scholar] [CrossRef]
  32. Hardell, L. World Health Organization, radiofrequency radiation and health—A hard nut to crack (review). Int. J. Oncol. 2017, 51, 405–413. [Google Scholar] [CrossRef]
  33. IARC. International agency for research on cancer (iarc) working group on the evaluation of carcinogenic risks to humans. Non-ionizing radiation, part 1: Static and extremely low-frequency (elf) electric and magnetic fields. In Iarc Monographs on the Evaluation of Carcinogenic Risks to Humans; IARC: Lyon, France, 2002; Volume 80, p. 37. Available online: https://publications.iarc.fr/Book-And-Report-Series/Iarc-Monographs-On-The-Identification-Of-Carcinogenic-Hazards-To-Humans/Non-ionizing-Radiation-Part-31-Static-And-Extremely-Low-frequency-ELF-Electric-And-Magnetic-Fields (accessed on 29 July 2022).
  34. BioInitiative. Bioinitiative Report: A Rationale for Biologically-Based Public Exposure Standards for Electromagnetic Radiation. Working Group, Cindy Sage, David O. Carpenter, 31 December 2012. Available online: https://www.bioinitiative.org (accessed on 29 July 2022).
  35. 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 29 July 2022).
  36. Jefimenko, O.D. Causality Electromagnetic Induction and Gravitation, 2nd ed.; Electret Scientific Company: Star City, WV, USA, 2000; p. 210. [Google Scholar]
  37. Verkasalo, P.K. Magnetic fields and leukemia–risk for adults living close to power lines. Scand. J. Work Environ. Health 1996, 22, 1–56. [Google Scholar] [PubMed]
  38. Crespi, C.M.; Vergara, X.P.; Hooper, C.; Oksuzyan, S.; Wu, S.; Cockburn, M.; Kheifets, L. Childhood leukaemia and distance from power lines in California: A population-based case-control study. Br. J. Cancer 2016, 115, 122–128. [Google Scholar] [CrossRef] [PubMed]
  39. Seomun, G.; Lee, J.; Park, J. Exposure to extremely low-frequency magnetic fields and childhood cancer: A systematic review and meta-analysis. PLoS ONE 2021, 16, e0251628. [Google Scholar] [CrossRef] [PubMed]
  40. Moon, J.H. Health effects of electromagnetic fields on children. Clin. Exp. Pediatr. 2020, 63, 422–428. [Google Scholar] [CrossRef]
  41. Brabant, C.; Geerinck, A.; Beaudart, C.; Tirelli, E.; Geuzaine, C.; Bruyère, O. Exposure to magnetic fields and childhood leukemia: A systematic review and meta-analysis of case-control and cohort studies. Rev. Environ. Health, 2022; Epub ahead of print. [Google Scholar]
  42. Panagopoulos, D.J.; Johansson, O.; Carlo, G. Polarization: A key difference between man-made and natural electromagnetic fields, in regard to biological activity. Sci. Rep. 2015, 5, 14914. [Google Scholar] [CrossRef]
  43. Pall, M.L. Millimeter (MM) wave and microwave frequency radiation produce deeply penetrating effects: The biology and the physics. Rev. Environ. Health 2021, 37, 247–258. [Google Scholar] [CrossRef]
  44. 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]
  45. Panagopoulos, D.J. Man-made electromagnetic radiation is not quantized. In Horizons in World Physics; Reimer, A., Ed.; Nova Science Publishers: Hauppauge, NY, USA, 2018; Volume 296. [Google Scholar]
  46. 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]
  47. Semashko, V.V.; Pudovkin, M.S.; Cefalas, A.C.; Zelenikhin, P.V.; Gavriil, V.E.; Nizamutdinov, A.S.; Kollia, Z.; Ferraro, A.; Sarantopoulou, E. Tiny rare-earth fluoride nanoparticles activate tumour cell growth via electrical polar interactions. Nanoscale Res. Lett. 2018, 13, 370. [Google Scholar] [CrossRef]
  48. Cefalas, A.C.; Gavriil, V.; Ferraro, A.; Kollia, Z.; Sarantopoulou, E. Dynamics and physics of integrin activation in tumor cells by nano-sized extracellular ligands and electromagnetic fields. Methods Mol. Biol. 2021, 2217, 197–233. [Google Scholar]
  49. 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]
  50. Saliev, T.; Begimbetova, D.; Masoud, A.R.; Matkarimov, B. Biological effects of non-ionizing electromagnetic fields: Two sides of a coin. Prog. Biophys. Mol. Biol. 2019, 141, 25–36. [Google Scholar] [CrossRef] [PubMed]
  51. Bertagna, F.; Lewis, R.; Silva, S.R.P.; McFadden, J.; Jeevaratnam, K. Effects of electromagnetic fields on neuronal ion channels: A systematic review. Ann. N. Y. Acad. Sci. 2021, 1499, 82–103. [Google Scholar] [CrossRef] [PubMed]
  52. Narayanan, S.N.; Jetti, R.; Kesari, K.K.; Kumar, R.S.; Nayak, S.B.; Bhat, P.G. Radiofrequency electromagnetic radiation-induced behavioral changes and their possible basis. Environ. Sci. Pollut. Res. Int. 2019, 26, 30693–30710. [Google Scholar] [CrossRef] [PubMed]
  53. 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. [Google Scholar] [CrossRef] [PubMed]
  54. Panagopoulos, D.J.; Karabarbounis, A.; Yakymenko, I.; Chrousos, G.P. Human made electromagnetic fields: Ion forced oscillation and voltage gated ion channel dysfunction, oxidative stress and DNA damage (review). Int. J. Oncol. 2021, 59, 92. [Google Scholar] [CrossRef]
  55. PSU. Coordination chemistry and crystal field theory-spectrochemical series. In Introduction to Inorganic Chemistry; Penn State University (PSU) LibreTexts: State College, PA, USA, 2020; Available online: https://chem.libretexts.org/Bookshelves/Inorganic_Chemistry/Book%3A_Introduction_to_Inorganic_Chemistry_(Wikibook)/05%3A_Coordination_Chemistry_and_Crystal_Field_Theory (accessed on 29 July 2022).
  56. Rastogi, R.P.; Richa, K.A.; Tyagi, M.B.; Sinha, R.P. Molecular mechanisms of ultraviolet radiation-induced DNA damage and repair. J. Nucleic Acids 2010, 2010, 592980. [Google Scholar] [CrossRef]
  57. Liu, C.; Zheng, Y.; Sanche, L. Damage induced to DNA and its constituents by 0–3 eV UV photoelectrons. Photochem. Photobiol. 2022, 98, 546–563. [Google Scholar] [CrossRef]
  58. Bayfield, J.E.; Koch, P.M. Multiphoton ionization of highly excited hydrogen atoms. Phys. Rev. Lett. 1974, 33, 258–261. [Google Scholar] [CrossRef]
  59. Ruscic, B.; Feller, D.; Dixon, D.A.; Peterson, K.A.; Harding, L.B.; Asher, R.L.; Wagner, A.F. Evidence for a lower enthalpy of formation of hydroxyl radical and a lower gas-phase bond dissociation energy of water. J. Phys. Chem. A 2001, 105, 1–4. [Google Scholar] [CrossRef]
  60. Estacio, S.G.; Do Couto, P.C.; Guedes, R.C.; Cabral, B.J.C.; Simoes, J.A.M. Homolytic dissociation in hydrogen-bonding liquids: Energetics of the phenol O-H bond in methanol and the water O-H bond in water. Theor. Chem. Acc. 2004, 112, 282–289. [Google Scholar] [CrossRef]
  61. Halliwell, B.; Gutteridge, J.M.C. Free Radicals in Biology and Medicine, 5th ed.; Oxford University Press: Oxford, UK, 2015; p. 904. [Google Scholar]
  62. Delone, N.B.; Zon, B.A.; Krainov, V.P. Diffusion mechanism of ionization of highly excited atoms in an alternating electromagnetic field (Zh. Eksp. Teor. Fiz. 75:445-453 (1978)). Sov. Phys. JETP Lett. 1978, 48, 223. [Google Scholar]
  63. Meerson, B.I.; Oks, E.A.; Sasorov, P.V. Stochastic instability of an oscillator and the ionization of highly-excited atoms under the action of electromagnetic radiation (Pis’ma Zh. Eksp. Teor. Fiz. 29:79 (1979)). Sov. JETP Lett. 1979, 29, 72. [Google Scholar]
  64. Casati, G.; Chirikov, B.V.; Shepelyansky, D.L.; Guarneri, I. Relevance of classical chaos in quantum mechanics: The hydrogen atom in a monochromatic field. Phys. Rep. 1987, 154, 77–123. [Google Scholar] [CrossRef]
  65. Casati, G.; Guarneri, I.; Shepelyansky, D.L. Hydrogen atom in monochromatic field: Chaos and dynamical photonic localization. IEEE J. Quantum Electron. 1988, 24, 1420–1444. [Google Scholar] [CrossRef] [Green Version]
  66. Shepelyansky, D.L. Localization of diffusive excitation in multi-level systems. Phys. D Nonlinear Phenom. 1987, 28, 103–114. [Google Scholar] [CrossRef]
  67. Jensen, R.V.; Susskind, S.M.; Sanders, M.M. Chaotic ionization of highly excited hydrogen atoms: Comparison of classical and quantum theory with experiment. Phys. Rep. 1991, 201, 1–56. [Google Scholar] [CrossRef]
  68. Koch, P.M.; van Leeuwen, K.H.A. The importance of resonances in microwave “ionization” of excited hydrogen atoms. Phys. Rep. 1995, 256, 289–403. [Google Scholar] [CrossRef]
  69. Buchleitner, A.; Delande, D.; Zakrewski, J. Non-dispersive wave packets in periodically driven quantum systems. Phys. Rep. 2002, 368, 409–547. [Google Scholar] [CrossRef]
  70. Arndt, M.; Buchleitner, A.; Mantegna, R.N.; Walther, H. Experimental study of quantum and classical limits in microwave ionization of rubidium Rydberg atoms. Phys. Rev. Lett. 1991, 67, 2435–2438. [Google Scholar] [CrossRef]
  71. Anderson, P.W. Absence of diffusion in certain random lattices. Phys. Rev. A 1958, 109, 1492–1505. [Google Scholar] [CrossRef]
  72. Du Bois, D.; Du Bois, E.F. A formula to estimate the approximate surface area if height and weight be known. Arch. Intern. Med. 1916, 17, 863–871. [Google Scholar] [CrossRef]
  73. Verbraecken, J.; Van de Heyning, P.; De Backer, W.; Van Gaal, L. Body surface area in normal-weight, overweight, and obese adults. A comparison study. Metab. Clin. Exp. 2006, 55, 515–524. [Google Scholar] [CrossRef] [PubMed]
  74. Vriens, L. On the Difference Between Man-Made and Natural Electromagnetic Fields/Radiation, in Regard to Biological Activity. 2018. Available online: https://www.jrseco.com/wp-content/uploads/Man-made-and-Natural-EMF-EMR-Dr.Leendert-Vries.pdf (accessed on 29 July 2022).
  75. Giuliani, L.; D’Emilia, E.; Grimaldi, S.; Lisi, A.; Bobkova, N.; Zhadin, M.N. Investigating the Icr effect in a Zhadin’s cell. Int. J. Biomed. Sci. 2009, 5, 181–186. [Google Scholar] [PubMed]
  76. ICNIRP. Guidelines for limiting exposure to time-varying electric, magnetic and electromagnetic fields (up to 300 GHz). Health Physics 1998, 74, 494–522. [Google Scholar]
  77. ICNIRP. Guidelines for limiting exposure to time-varying electric and magnetic fields (1 Hz to 100 KHz). Health Physics 2010, 99, 818–836. [Google Scholar] [CrossRef]
  78. Pedrazzi, G.; Galli, C.; Mattioli-Belmonte, M.; Guizzardi, S. Response to: Comment on “the use of pulsed electromagnetic fields to promote bone responses to biomaterials in vitro and in vivo”. Int. J. Biomater. 2020, 2020, 9801420. [Google Scholar] [CrossRef]
  79. ICNIRP. Use of the ICNIRP EMF Guidelines, Notes; ICNIRP Publications: Munich, Germany, 1999; pp. 297–298. Available online: https://www.icnirp.org/cms/upload/publications/ICNIRPUseEMFgdl.pdf (accessed on 13 September 2022).
  80. Belyaev, I.Y.; Alipov, Y.D.; Harms-Ringdahl, M. Effects of zero magnetic field on the conformation of chromatin in human cells. Biochim. Biophys. Acta 1997, 1336, 465–473. [Google Scholar] [CrossRef]
  81. Grundler, W.; Kaiser, F.; Keilmann, F.; Walleczek, J. Mechanisms of electromagnetic interaction with cellular systems. Naturwissenschaften 1992, 79, 551–559. [Google Scholar] [CrossRef]
  82. Behari, J.; Kesari, K.K. Effects of microwave radiations on reproductive system of male rats. Embryo Talk 2006, 1, 81–85. [Google Scholar]
  83. Velizarov, S.; Raskmark, P.; Kwee, S. The effects of radiofrequency fields on cell proliferation are non-thermal. Bioelectrochem. Bioenergetics 1999, 48, 177–180. [Google Scholar] [CrossRef]
  84. Stagg, R.B.; Thomas, W.J.; Jones, R.A.; Adey, W.R. DNA synthesis and cell proliferation in C6 glioma and primary glial cells exposed to a 836.55 MHz modulated radiofrequency field. Bioelectromagnetics 1997, 18, 230–236. [Google Scholar] [CrossRef]
  85. Mohler, E.; Frei, P.; Braun-Fahrländer, C.; Fröhlich, J.; Neubauer, G.; Röösli, M.; Team, Q. Effects of everyday radiofrequency electromagnetic-field exposure on sleep quality: A crosssectional study. Radiat. Res. 2010, 174, 347–356. [Google Scholar] [CrossRef] [PubMed]
  86. Oberfeld, G.; Enrique, N.A.; Manuel, P.; Ceferino, M.; Gomez-Perrretta, C. The microwave syndrome—further aspects of a spanish study. In Proceedings of the 3rd International Workshop on Biological Effects of Electromagnetic Fields, Kos, Greece, 4–8 October 2004. [Google Scholar]
  87. Buchner, K.; Eger, H. Changes of clinically important neurotransmitters under the influence of modulated RF fields—A long-term study under real-life conditions (original study in German). Umwelt-Medizin-Gesellschaft 2011, 24, 44–57. [Google Scholar]
  88. Heinrich, S.; Thomas, S.; Heumann, C.; von Kries, R.; Radon, K. Association between exposure to radiofrequency electromagnetic fields assessed by dosimetry and acute symptoms in children and adolescents: A population based cross-sectional study. Environ. Health 2010, 9, 75. [Google Scholar] [CrossRef]
  89. Thomas, S.; Kühnlein, A.; Heinrich, S.; Praml, G.; Nowak, D.; von Kries, R.; Radon, K. Personal exposure to mobile phone frequencies and well-being in adults: A cross-sectional study based on dosimetry. Bioelectromagnetics 2008, 29, 463–470. [Google Scholar] [CrossRef]
  90. Thomas, S.; Heinrich, S.; von Kries, R.; Radon, K. Exposure to radio-frequency electromagnetic fields and behavioural problems in bavarian children and adolescents. Eur. J. Epidemiol. 2010, 25, 135–141. [Google Scholar] [CrossRef]
  91. Hutter, H.P.; Moshammer, H.; Wallner, P.; Kundi, M. Subjective symptoms, sleeping problems, and cognitive performance in subjects living near mobile phone base stations. Occup. Environ. Med. 2006, 63, 307–313. [Google Scholar] [CrossRef]
  92. Navarro, E.A.; Sequra, J.; Portoles, M.; Gomez-Perretta de Mateo, C. The microwave syndrome: A preliminary study in Spain. Electromagn. Biol. Med. 2003, 22, 161–169. [Google Scholar] [CrossRef]
  93. Khurana, V.G.; Hardell, L.; Everaert, J.; Bortkiewicz, A.; Carlberg, M.; Ahonen, M. Epidemiological evidence for a health risk from mobile phone base stations. Int. J. Occup. Environ. Health 2010, 16, 263–267. [Google Scholar] [CrossRef]
  94. Kundi, M.; Hutter, H.P. Mobile phone base stations—Effects on wellbeing and health. Pathophysiology 2009, 16, 123–135. [Google Scholar] [CrossRef]
  95. Otitoloju, A.A.; Obe, I.A.; Adewale, O.A.; Otubanjo, O.A.; Osunkalu, V.O. Preliminary study on the induction of sperm head abnormalities in mice, mus musculus, exposed to radiofrequency radiations from global system for mobile communication base stations. Bull. Environ. Contam. Toxicol. 2010, 84, 51–54. [Google Scholar] [CrossRef] [PubMed]
  96. Augner, C.; Florian, M.; Pauser, G.; Oberfeld, G.; Hacker, G.W. GSM base stations: Short-term effects on well-being. Bioelectromagnetics 2009, 30, 73–80. [Google Scholar] [CrossRef] [PubMed]
  97. Zwamborn, A.P.M.; Vossen, S.H.J.A.; van Leersum, B.J.A.M.; Ouwens, M.A.; Makel, W.N. Effects of Global Communication System Radio-Frequency Fields on Well Being and Cognitive Functions of Human Subjects With and Without Subjective Complaints. 2003. Netherlands Organisation for Applied Scientific Research (TNO). Fel-03-c148. Available online: http://electrosmog.grappe.be/doc/TNO_rapport_Nederland_sept_2003.pdf (accessed on 29 July 2022).
  98. Kolodynski, A.A.; Kolodynska, V.V. Motor and psychological functions of school children living in the area of the Skrunda Radio Location Station in Latvia. Sci. Total Environ. 1996, 180, 87–93. [Google Scholar] [CrossRef]
  99. Schwartz, J.L.; House, D.; Mealing, G. Exposure of frog hearts to CW or amplitude-modulated VHF fields: Selective efflux of calcium ions at 16 Hz. Bioelectromagnetics 1990, 11, 349–358. [Google Scholar] [CrossRef]
  100. Saunders, R.D.; Kowalczuk, C.I. Effects of 2.45 GHz microwave radiation and heat on mouse spermatogenic epithelium. Int. J. Radiat. Biol. Relat. Stud. Phys. Chem. Med. 1981, 40, 623–632. [Google Scholar] [CrossRef]
  101. Magras, I.N.; Zenos, T.D. RF radiation-induced changes in the prenatal development of mice. Bioelectromagnetics 1997, 18, 455–461. [Google Scholar] [CrossRef]
  102. Riddervold, I.S.; Pedersen, G.F.; Andersen, N.T.; Pedersen, A.D.; Andersen, J.B.; Zachariae, R.; Mølhave, L.; Sigsgaard, T.; Kjaergaard, S.K. Cognitive function and symptoms in adults and adolescents in relation to RF radiation from UMTS base stations. Bioelectromagnetics 2008, 29, 257–267. [Google Scholar] [CrossRef]
  103. Avendano, C.; Mata, A.; Sanchez Sarmiento, C.A.; Doncei, G.F. Use of laptop computers connected to internet through Wi-Fi decreases human sperm motility and increases sperm DNA fragmentation. Fertil. Steril. 2012, 97, 39–45. [Google Scholar] [CrossRef]
  104. Persson, B.R.; Salford, L.G.; Brun, A. Blood-brain barrier permeability in rats exposed to electromagnetic fields used in wireless communication. Wirel. Netw. 1997, 3, 455–461. [Google Scholar] [CrossRef]
  105. Fesenko, E.E.; Makar, V.R.; Novoselova, E.G.; Sadovnikov, V.B. Microwaves and cellular immunity. I. Effect of whole body microwave irradiation on tumor necrosis factor production in mouse cells. Bioelectrochem. Bioenergetics 1999, 49, 29–35. [Google Scholar] [CrossRef]
  106. Novoselova, E.G.; Fesenko, E.E.; Makar, V.R.; Sadovnikov, V.B. Microwaves and cellular immunity. II. Immunostimulating effects of microwaves and naturally occurring antioxidant nutrients. Bioelectrochem. Bioenergetics 1999, 49, 37–41. [Google Scholar] [CrossRef]
  107. Eltiti, S.; Wallace, D.; Ridgewell, A.; Zougkou, K.; Russo, R.; Sepulveda, F.; Mirshekar-Syahkal, D.; Rasor, P.; Deeble, R.; Fox, E. Does short-term exposure to mobile phone base station signals increase symptoms in individuals who report sensitivity to electromagnetic fields? A double-blind randomized provocation study. Environ. Health Perspect. 2007, 115, 1603–1608. [Google Scholar] [CrossRef] [PubMed]
  108. Pyrpasopoulou, A.; Kotoula, V.; Cheva, A.; Hytiroglou, P.; Nikolakaki, E.; Magras, I.N.; Xenos, T.D.; Tsiboukis, T.D.; Karkavelas, G. Bone morphogenetic protein expression in newborn rat kidneys after prenatal exposure to radiofrequency radiation. Bioelectromagnetics 2004, 25, 216–227. [Google Scholar] [CrossRef] [PubMed]
  109. Nittby, H.; Grafstrom, G.; Tian, D.P.; Malmgren, L.; Brun, A.; Persson, B.R.; Salford, L.G.; Eberhardt, J. Cognitive impairment in rats after long-term exposure to GSM-900 mobile phone radiation. Bioelectromagnetics 2008, 29, 219–232. [Google Scholar] [CrossRef]
  110. Kesari, K.K.; Behari, J. Microwave exposure affecting reproductive system in male rats. Appl. Biochem. Biotechnol. 2010, 31, 495–498. [Google Scholar] [CrossRef]
  111. Wolke, S.; Neibig, U.; Elsner, R.; Gollnick, F.; Meyer, R. Calcium homeostasis of isolated heart muscle cells exposed to pulsed high-frequency electromagnetic fields. Bioelectromagnetics 1996, 17, 144–153. [Google Scholar] [CrossRef]
  112. D’Inzeo, G.; Bernardi, P.; Eusebi, F.; Grassi, F.; Tamburello, C.; Zani, B.M. Microwave effects on acetylcholine-induced channels in cultured chick myotubes. Bioelectromagnetics 1988, 9, 363–372. [Google Scholar] [CrossRef]
  113. Dolk, H.; Elliot, P.; Shaddick, G.; Walls, P.; Thakrar, B. Cancer incidence near radio television and transmitters in great britain, part II. All high-power transmitters. Am. J. Epidemiol. 1997, 145, 10–17. [Google Scholar] [CrossRef]
  114. Tattersall, J.E.; Scott, I.R.; Wood, S.J.; Nettell, J.J.; Bevir, M.K.; Wang, Z.; Somasiri, N.P.; Chen, X. Effects of low intensity radiofrequency electromagnetic fields on electrical activity in rat hippocampal slices. Brain Res. 2001, 904, 43–53. [Google Scholar] [CrossRef]
  115. Hocking, B.; Gordon, I.R.; Grain, M.L.; Hatfield, G.E. Cancer incidence and mortality and proximity to TV towers. Med. J. 1996, 165, 601–605. [Google Scholar] [CrossRef]
  116. Hocking, B.; Gordon, I. Decreased survival for childhood leukemia in proximity to television towers. Arch. Environ. Health. 2003, 58, 560–564. [Google Scholar] [CrossRef] [PubMed]
  117. Boscolo, P. Effects of electromagnetic fields produced by radiotelevision broadcasting stations on the immune system of women. Sci. Total Environ. 2001, 273, 1–10. [Google Scholar] [CrossRef]
  118. Kwee, S.; Raskmark, P.; Velizarov, S. Changes in cellular protiens due to environmental non-ionizing radiation. I. Heat-shock proteins. Electro Magnetobiol. 2001, 20, 141–152. [Google Scholar] [CrossRef]
  119. Akoev, I.G.; Pashovkina, M.S.; Dolgacheva, L.P.; Semenova, T.P.; Kalmykov, V.L. Enzymatic activity of some tissues and blood serum from animals and humans exposed to microwaves and hypothesis on the possible role of free radical processes in the nonlinear effects and modification of emotional behavior of animals. Radiatsionnaia Biol. Radioecol. 2002, 42, 322–330. [Google Scholar]
  120. Phillips, J.L.; Ivaschuk, O.; Ishida-Jones, T.; Jones, R.A.; Campbell-Beachler, M.; Haggren, W. DNA damage in Molt-4 T-lymphoblastoid cells exposed to cellular telephone radiofrequency fields in vitro. Bioelectrochem. Bioenergetics 1998, 45, 103–110. [Google Scholar] [CrossRef]
  121. Dumansky, J.D.; Shandala, M.G. The biologic action and hygienic significance of electromagnetic fields of super high and ultra high frequencies in densely populated areas. In Biologic Effects and Health Hazard of Microwave Radiation: Proceedings of an International Symposium; Czerski, P., Ostrowski, K., Shore, M.L., Silverman, C., Suess, M.J., Waldeskog, B., Eds.; Polish Medical Publishers: Warsaw, Poland, 1974. [Google Scholar]
  122. Marinelli, F.; La Sala, D.; Cicciotti, G.; Cattini, L.; Trimarchi, C.; Putti, S.; Zamparelli, A.; Giuliani, L.; Tomassetti, G.; Cinti, C. Exposure to 900 MHz electromagnetic field induces an unbalance between pro-apoptotic and pro-survival signals in T-lymphoblastoid leukemia CCRF-CEM cells. J. Cell Physiol. 2004, 198, 324–332. [Google Scholar] [CrossRef]
  123. Chiang, H.; Yao, G.D.; Fang, Q.S.; Wang, K.Q.; Lu, D.Z.; Zhou, Y.K. Health effects of environmental electromagnetic fields. J. Bioelectr. 1989, 8, 127–131. [Google Scholar] [CrossRef]
  124. Navakatikian, M.A.; Tomashevskaya, L.A. Phasic behavioral and endocrine effects of microwaves of nonthermal intensity. In Biological Effects of Electric and Magnetic Fields; Carpenter, D.O., Ed.; Academic Press, Inc.: San Diego, CA, USA, 1994; Volume 1, pp. 333–342. [Google Scholar]
  125. Dutta, S.K.; Ghosh, B.; Blackman, C.F. Radiofrequency radiation-induced calcium ion efflux enhancement from human and other neuroblastoma cells in culture. Bioelectromagnetics 1989, 10, 197–202. [Google Scholar] [CrossRef]
  126. Sarimov, R.; Malmgren, L.; Markova, E.; Persson, B.; Belyaev, I.Y. Non-thermal GSM microwaves affect chromatin conformation in human lymphocytes similar to heat shock. IEEE Trans. Plasma Sci. 2004, 32, 1600–1608. [Google Scholar] [CrossRef]
  127. Mann, K.; Wagner, P.; Brunn, G.; Hassan, F.; Hiemke, C.; Röschke, J. Effects of pulsed high-frequency electromagnetic fields on the neuroendocrine system. Neuroendocrinology 1998, 67, 139–144. [Google Scholar] [CrossRef]
  128. Yurekli, A.I.; Ozkan, M.; Kalkan, T.; Saybasili, H.; Tuncel, H.; Atukeren, P.; Gumustas, K.; Seker, S. GSM base station electromagnetic radiation and oxidative stress in rats. Electromagn. Biol. Med. 2006, 25, 177–188. [Google Scholar] [CrossRef] [PubMed]
  129. Veyret, B.; Bouthet, C.; Deschaux, P.; de Seze, R.; Geffard, M.; Joussot-Dubien, J.; le Diraison, M.; Moreau, J.M.; Caristan, A. Antibody responses of mice exposed to low-power microwaves under combined, pulse-and-amplitude modulation. Bioelectromagnetics 1991, 12, 47–56. [Google Scholar] [CrossRef] [PubMed]
  130. Forgács, Z.; Somosy, Z.; Kubinyi, G.; Bakos, J.; Hudák, A.; Surján, A.; Thuróczy, G. Effect of whole-body 1800 MHz GSM-like microwave exposure on testicular steroidogenesis and histology in mice. Reprod. Toxicol. 2006, 22, 111–117. [Google Scholar] [CrossRef]
  131. Salford, L.G.; Brun, A.E.; Eberhardt, J.L.; Malmgren, L.; Persson, B.R. Nerve cell damage in mammalian brain after exposure to microwaves from GSM mobile phones. Environ. Health Perspect. 2003, 111, 881–883, discussion A408. [Google Scholar] [CrossRef] [PubMed]
  132. Mann, K.; Röschke, J. Effects of pulsed high-frequency electromagnetic fields on human sleep. Neuropsychobiology 1996, 33, 41–47. [Google Scholar] [CrossRef]
  133. Richter, E.; Berman, T.; Ben-Michael, E.; Laster, R.; Westin, J.B. Cancer in radar technicians exposed to radiofrequency/microwave radiation: Sentinel episodes. Int. J. Occup. Environ. Health 2000, 6, 187–193. [Google Scholar] [CrossRef]
  134. Somozy, Z. Radiation response of cell organelles. Micron 2000, 31, 165–181. [Google Scholar] [CrossRef]
  135. Stankiewicz, W.; Dąbrowski, M.P.; Kubacki, R.; Sobiczewska, E.; Szmigielski, S. Immunotropic influence of 900 MHz microwave GSM signal on human blood immune cells activated in vitro. Electromagn. Biol. Med. 2006, 25, 45–51. [Google Scholar] [CrossRef]
  136. Lebedeva, N.N.; Sulimov, A.V.; Sulimova, O.P.; Kotrovskaya, T.I.; Gailus, T. Cellular phone electromagnetic field effects on bioelectric activity of human brain. Crit. Rev. Biomed. Eng. 2000, 28, 323–337. [Google Scholar] [CrossRef]
  137. Ivaschuk, O.I.; Jones, R.A.; Ishida-Jones, T.; Haggren, W.; Adey, W.R.; Phillips, J.L. Exposure of nerve growth factor-treated PC12 rat pheochromocytoma cells to a modulated radiofrequency field at 836.55 MHz: Effects on c-jun and c-fos expression. Bioelectromagnetics 1997, 18, 223–229. [Google Scholar] [CrossRef]
  138. Elekes, E. Effect on the immune system of mice exposed chronically to 50 Hz amplitude modulated 2.45 GHz microwaves. Bioelectromagnetics 1996, 17, 246–248. [Google Scholar] [CrossRef]
  139. Salford, L.G.; Brun, A.; Sturesson, K.; Eberhardt, J.L.; Persson, B.R. Permeability of the blood-brain barrier induced by 915 MHz electromagnetic radiation, continuous wave and modulated at 8, 16, 50, and 200 Hz. Microsc. Res. Tech. 1994, 27, 535–542. [Google Scholar] [CrossRef] [PubMed]
  140. 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] [PubMed]
  141. 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] [PubMed]
  142. Smith-Roe, S.L.; Wyde, M.E.; Stout, M.D.; Winters, J.W.; Hobbs, C.A.; Shepard, K.G.; Green, A.S.; Kissling, G.E.; Shockley, K.R.; Tice, R.R.; et al. Evaluation of the genotoxicity of cell phone radiofrequency radiation in male and female rats and mice following subchronic exposure. Environ. Mol. Mutagen. 2020, 61, 276–290. [Google Scholar] [CrossRef] [PubMed]
  143. NTP. National Toxicology Program (NTP) Technical Report on the Toxicology and Carcinogenesis Studies in b6c3f1/n Mice Exposed to Whole-Body Radio Frequency Radiation at a Frequency (1,900 MHz) and Modulations (GSM and CDMA) used by Cell Phones. NTP TR 596, 26–28 March 2018. Available online: https://ntp.Niehs.Nih.Gov/ntp/htdocs/lt_rpts/tr596_508.Pdf?Utm_source=direct&utm_medium=prod&utm_campaign=ntpgolinks&utm_term=tr596 (accessed on 29 July 2022).
  144. NTP. Toxicology and Carcinogenesis Studies in Sprague Dawley (Hsd:Sprague Dawley SD) Rats Exposed to Whole-Body Radio Frequency Radiation at a Frequency (900 MHz) and Modulations (GSM and CDMA) Used by Cell Phones. Natl. Toxicol Program Tech. Rep. Ser. 2018, 595, NTP-TR-595. Available online: https://pubmed.Ncbi.Nlm.Nih.Gov/33562898 (accessed on 29 July 2022). [CrossRef]
  145. Falcioni, L.; Bua, L.; Tibaldi, E.; Lauriola, M.; De Angelis, L.; Gnudi, F.; Mandrioli, D.; Manservigi, M.; Manservisi, F.; Manzoli, I.; et al. Report of final results regarding brain and heart tumors in Sprague-Dawley rats exposed from prenatal life until natural death to mobile phone radiofrequency field representative of a 1.8 GHz GSM base station environmental emission. Environ. Res. 2018, 165, 496–503. [Google Scholar] [CrossRef]
  146. Tillmann, T.; Ernst, H.; Streckert, J.; Zhou, Y.; Taugner, F.; Hansen, V.; Dasenbrock, C. Indication of cocarcinogenic potential of chronic UMTS-modulated radiofrequency exposure in an ethylnitrosourea mouse model. Int. J. Radiat. Biol. 2010, 86, 529–541. [Google Scholar] [CrossRef]
  147. Lerchl, A.; Klose, M.; Grote, K.; Wilhelm, A.F.X.; Spathmann, O.; Fiedler, T.; Streckert, J.; Hansen, V.; Clemens, M. Tumor promotion by exposure to radiofrequency electromagnetic fields below exposure limits for humans. Biochem. Biophys. Res. Commun. 2015, 459, 585–590. [Google Scholar] [CrossRef]
  148. Powell, R.M. Biological Effects from RF Radiation at Low-Intensity Exposure, Based on the Bioinitiative 2012 Report, and the Implications for Smart Meters and Smart Appliances. 2013. Available online: https://stopsmartmetersau.files.wordpress.com/2014/09/biological_effects_from_rf_radiation_and-implications_for_smart_meters_june_11_2013_b.pdf (accessed on 29 July 2022).
  149. Wolf, E. Introduction to the Theory of Coherence and Polarization of Light; Cambridge University Press: Cambridge, UK, 2007. [Google Scholar]
  150. DiLaura, D.L.; Houser, K.W.; Mistrick, R.G.; Steffy, G.R. The Lighting Handbook: Reference and Application, 10th ed.; Illuminating Engineering Society of North America (ILESNA): New York, NY, USA, 2011; p. 1087. [Google Scholar]
  151. Ahfayd, M.H.; Farhat, Z.A.; Sibley, M.J.N.; Mather, P.J.; Lazaridis, P.I. Selection of High Power Leds for Li-Fi Applications. In Proceedings of the 25th International Conference on Telecommunications, Saint-Malo, France, 26–28 June 2018; pp. 170–174. [Google Scholar]
  152. Oyewobi, S.S.; Djouani, K.; Kurien, A.M. Visible light communications for internet of things: Prospects and approaches, challenges, solutions and future directions. Technologies 2022, 10, 28. [Google Scholar] [CrossRef]
Figure 1. Day and night natural limit of power density (mW/cm2) of EMR emitted by the Sun and reaching Earth’s surface in the microwave spectral regions, from 0.1 to 3 × 104 MHz. The associated power density natural limit of ~10−23 to 10−21 mW/cm2 is compared with the ICNIRP exposure limit of ~1 mW/cm2 for man-made microwave sources. The ICNIRP/FCC thermal exposure limit for man-made EMR is ~1021- to 1023-fold higher than the natural microwave radiation limit at Earth’s surface of single microwave frequencies from the Sun (the plot is an extensive modification of that in NASA’s report CR 166661 [10]).
Figure 1. Day and night natural limit of power density (mW/cm2) of EMR emitted by the Sun and reaching Earth’s surface in the microwave spectral regions, from 0.1 to 3 × 104 MHz. The associated power density natural limit of ~10−23 to 10−21 mW/cm2 is compared with the ICNIRP exposure limit of ~1 mW/cm2 for man-made microwave sources. The ICNIRP/FCC thermal exposure limit for man-made EMR is ~1021- to 1023-fold higher than the natural microwave radiation limit at Earth’s surface of single microwave frequencies from the Sun (the plot is an extensive modification of that in NASA’s report CR 166661 [10]).
Radiation 02 00022 g001
Figure 2. The ratio of absorption to emission probability of microwave radiation from molecular rotational states as a function of the rotational quantum number J. Resonance of microwave radiation with the rotational levels of water clusters or resonances via microwave adiabatic tunnelling is responsible for microwave energy trapping between the rotational energy levels of water clusters, causing biological stress during the interaction of EMR with biological matter.
Figure 2. The ratio of absorption to emission probability of microwave radiation from molecular rotational states as a function of the rotational quantum number J. Resonance of microwave radiation with the rotational levels of water clusters or resonances via microwave adiabatic tunnelling is responsible for microwave energy trapping between the rotational energy levels of water clusters, causing biological stress during the interaction of EMR with biological matter.
Radiation 02 00022 g002
Figure 3. The biological effects of RF-EMF radiation with radiation power intensities lower of the International Commission on Non-Ionizing Radiation Protection/U.S. Federal Communications Commission (ICNIRP/FCC) exposure limit. The chart is a reference-updated and slightly modified from the original version [148].
Figure 3. The biological effects of RF-EMF radiation with radiation power intensities lower of the International Commission on Non-Ionizing Radiation Protection/U.S. Federal Communications Commission (ICNIRP/FCC) exposure limit. The chart is a reference-updated and slightly modified from the original version [148].
Radiation 02 00022 g003
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.; Kalaitzopoulou, E.; Skipitari, M.; Papadea, P.; Varemmenou, A.; Gavriil, V.; Sarantopoulou, E.; Kollia, Z.; Cefalas, A.-C. Physical Differences between Man-Made and Cosmic Microwave Electromagnetic Radiation and Their Exposure Limits, and Radiofrequencies as Generators of Biotoxic Free Radicals. Radiation 2022, 2, 285-302. https://doi.org/10.3390/radiation2040022

AMA Style

Georgiou CD, Kalaitzopoulou E, Skipitari M, Papadea P, Varemmenou A, Gavriil V, Sarantopoulou E, Kollia Z, Cefalas A-C. Physical Differences between Man-Made and Cosmic Microwave Electromagnetic Radiation and Their Exposure Limits, and Radiofrequencies as Generators of Biotoxic Free Radicals. Radiation. 2022; 2(4):285-302. https://doi.org/10.3390/radiation2040022

Chicago/Turabian Style

Georgiou, Christos D., Electra Kalaitzopoulou, Marianna Skipitari, Polyxeni Papadea, Athina Varemmenou, Vassilios Gavriil, Evangelia Sarantopoulou, Zoe Kollia, and Alkiviadis-Constantinos Cefalas. 2022. "Physical Differences between Man-Made and Cosmic Microwave Electromagnetic Radiation and Their Exposure Limits, and Radiofrequencies as Generators of Biotoxic Free Radicals" Radiation 2, no. 4: 285-302. https://doi.org/10.3390/radiation2040022

Article Metrics

Back to TopTop