The Immunogenic Connection of Thermal and Nonthermal Molecular Effects in Modulated Electro-Hyperthermia

Abstract

Hyperthermia in oncology is an emerging complementary therapy. The clinical results depend on multiple conditional factors, like the type of cancer, the stage, the applied treatment device, and the complementary conventional therapy. The molecular effect could also be different depending on the temperature, heating dose, kind of energy transfer, and timing sequences compared to the concomitant treatment. This article examines the molecular impacts of a specific technique used in oncological hyperthermia called modulated electro-hyperthermia (mEHT). What sets mEHT apart is its emphasis on harnessing the combined effects of thermal and nonthermal factors. Nonthermal energy absorption occurs through the excitation of molecules, while the thermal component ensures the ideal conditions for this process. The applied radiofrequency current selects the malignant cells, and the modulation drives the nonthermal effects to immunogenic cell death, helping to develop tumor-specific antitumoral immune reactions. The synergy of the thermal and nonthermal components excites the lipid-assembled clusters of transmembrane proteins (membrane rafts) as the channels of transient receptor potentials (TRPs), the heat-shock proteins (HSPs), the voltage-gated channels, and the voltage-sensitive phosphatases (VSPs). All these transmembrane compartments channeling various ionic species (like calcium and proton) interact with the cytoskeleton and are involved in the apoptotic signal pathways.

Share and Cite:

Szász, A. , Lóránt, G. , Szász, A. and Szigeti, G. (2023) The Immunogenic Connection of Thermal and Nonthermal Molecular Effects in Modulated Electro-Hyperthermia. Open Journal of Biophysics, 13, 103-142. doi: 10.4236/ojbiphy.2023.134007.

1. Introduction

Electromagnetism appears as a continuous challenge in biology. The search for the use of electromagnetic effects for therapies ignited considerable research and hypotheses [1] . The ionizing radiation beam (high energy electromagnetic spectrum) shows immediate effects. The absorbed energy spectacularly destroys the biomaterial in the path of the beam. The non-ionizing radiation has less energy and more complexity. It modifies the chemical compounds and reactions and could impact enzymatic processes [2] . The complexity of living structures with physiological self-regulations challenges the therapeutic applications of non-ionizing effects. The challenges highlight the physiological importance of electric currents, deriving much intensive research, including neuroscience [3] and controlling the cellular effects [4] . A general hypothesis of “biologically closed electric circuits” (BCEC) introduced bioelectromagnetic homeostasis based on the existence of intrinsic electric currents in the body [5] [6] , modified by malignant diseases [7] [8] . The pathological disorders [9] and wounds induce intrinsic injury currents [10] , driven by the automatic biological charge transfers induced by the tissue-repair process [11] [12] . Observations have been made regarding the biological effects of low-level, non-stationary magnetic fields [13] [14] . The bioelectromagnetic effects may have resonance characters [15] [16] [17] .

All electromagnetic interactions deliver energy to the biomaterials. The energy could be realized by heat (which may increase the temperature) and electron excitation (which makes chemical changes). These effects are naturally combined. The bioelectromagnetic interactions partly modify the chemical bonds and structure of compounds with electromagnetic forces, while the part of the energy absorption heats the target. The preferences may change the treatments. For example, radiotherapy breaks the DNA strands, modifying the chemical bonds, where the heating is an adverse effect, while the focus of hyperthermia is to heat and neglects the direct chemical effects of the electric field. Initially, hyperthermia used both the field and heat effects combined in the middle of the 18th century, but later it split by dominant electric (by French doctor Arsene d’Arsonval) and heat (by Danish doctor Kristian Overgard) effects. To produce, control, and understand the heat effects were more accessible, promoting its worldwide spread and helped by some industrial devices manufactured by Siemens in the early 20th century.

Nowadays, a novel approach tries again to combine the thermal and nonthermal factors of non-ionizing radiation using modulated radiofrequency (RF) signals [18] . The method (modulated electro-hyperthermia, mEHT [19] ) applies definite heating in the fever range [20] and bioelectromagnetic effects in the bioprocesses using the nonthermal electromagnetic activity [21] [22] in the energy range selectively exciting transmembrane proteins on the malignant cells [23] . The principal selectivity of mEHT concentrates on the physiologic specialties of malignant cells and how they differ thermally and electrically from healthy ones.

The malignant cells have a higher metabolic rate that drives the RF current by high ionic density in the tumor-cell microenvironment (TME). Moreover, healthy cells maintain homeostatic electrolyte concentrations in various regions by well-controlled electrolyte balance having body electrolytes in the right concentrations regulated by heart, kidney, and neurological function, controlling the acid-base, fluid concentration balance, oxygen delivery, carbon dioxide transport, and other processes in the complex human body. The kidneys maintain a massive sodium regulation, which balances the important Na/K balance and calcium concentration for cellular functions. Calcium is involved in the function of enzymes and serves in signal transduction pathways, acting as a second messenger, in neurotransmitter release from neurons, in contraction of all muscle cell types, and in fertilization. Cancer cells alter the electrolyte balance concentrations. Some tumors have hypercalcemia, and the dysregulated pH causes electrolyte imbalance in cancer.

The measured impedance between healthy and cancerous tissues exhibits significant differences [24] . This impedance assists in selecting appropriate radiofrequency (RF) parameters [25] and enhances current density within the cancerous tissue [26] [27] . MRI images the selection showing high RF current density in the tumor [28] and prove the self-selection of the malignant region by the current flow [29] [30] . Electrical impedance tomography pro provides further feasibility of focusing on impedance differences [31] . The preclinical experiments in various investigations show the temperature differences between the tumor and its surroundings (Figure 1).

The growing temperature on the membrane makes a particular thermal impact compared to the conventional heating (water-bath, wHT) (Figure 2).

The mEHT focuses on the complex equilibrium of the human body [39] with an appropriate technical solution [40] , synergizing thermal and nonthermal energy components [41] by strong interaction of heat production with field

Figure 1. Temperature differences between the tumor and its surroundings, B16F10 (melanoma allograft) [32] , SCCVII (oral squamous allograft) [33] , B16F10 (melanoma lung-metastases allograft) [34] , 4T1 (triple negative breast isograft) [35] , lung (squamous xenograft) [36] , lung (squamous xenograft) [37] .

Figure 2. Measurement of the membrane temperature [38] . (a) Calibration of the membrane temperature by 10 µM DIL dye (RPMI + 10% FCS + 1% L-glutamine + 0.4% gentamycin). HT29 (human CRC cell line). Dilutions were kept at discrete temperatures for 30 minutes. (b) The temperature of the medium of the cell culture. (c) Significantly higher membrane temperature was achieved with mEHT than with wHT.

effects [42] . The central concept of mEHT uses the natural homeostatic control of the human body [43] , using a low-frequency modulation to stimulate healthy homeostatic regulation [44] [45] .

2. Thermal Impact

The intensive metabolic activity of the malignant cells [46] increases electric conductivity by the ionic density in the TME. Furthermore, the tumor has a higher water content [47] , which further increases the electric conductivity of the tumor. In this way, the entire tumor conducts better than its neighbor [48] [49] [50] [51] . An additional selection factor is that the malignant processes destroy the networking orders [52] [53] [54] . The presence of the disorder leads to an increase in the dielectric permittivity (ε) of the microregion [55] [56] [57] [58] . Consequently, the electric current will naturally follow the most accessible route, which is typically the most conductive path, thereby flowing through the tumor. The water content within the tumor microenvironment (TME) interacts with the membrane [59] , forming various bonds [60] and significantly impacting the membrane’s functionality. This phenomenon results in a low specific absorption rate (SAR) but a high voltage drop [61] , facilitating the excitation of raft proteins [62] by the signal. The electrostatic charge of the membrane attracts ions from the extracellular matrix (ECM), producing a diverse effect that is sufficient to establish a transmembrane potential [63] .

Selective raft heating makes a higher cell-killing rate with apoptotic processes than conventional homogeneous water bath heating (wHT). A calibration curve by wHT describes the apoptotic rate by temperature. The mEHT heterogeneous heating has a higher impact on the membrane proteins than wHT (Figure 3).

The transient receptor potential (TRP) channels are a set of transmembrane proteins and form a family of cation control channels [67] . These channels rectify the ionic transport, mainly calcium Ca2+, through the membrane. The rectification parameters primarily depend on thermal conditions sensing the relative to homeostasis oppositely in hot and cold temperatures. The hot sensing shifts

Figure 3. Thermal calibration with wHT measured with cell-lines U937 (o) [64] , HepG2 () [65] , CT26 (u) [66] , and the mEHT with U937 (r) [64] , and HepG2 (p) [65] , cell-lines. RF homogeneous heating with a conventional capacitive device for HepG2 cells is also given (n) [65] .

the Ca2+ ion-flux to the opening direction [68] , while the cold one shifts oppositely [69] to the closing side. TRPs regulate the membrane polarization, function as primary thermal sensors of cells [70] , inducing action potential for physiological sensation, and cover chemo-, mechano-, and photosensation [71] of individual cells. The intracellular organelles and cellular compartments also have TRP channels in various vesicular processes [72] [73] . The Ca2+ intra and extracellular transmembrane ionic exchanges have a decisional role [74] , allowing the individual cells to react to all intra and extracellular stimuli signals. The intracellular TRPs actively participate in membrane fusion and fission, signal transduction, and general vesicular homeostasis [73] . The TRPV5 and TRPV6 are the only TRPV channels that are highly selective for Ca2+ [75] . Others have low or no selection on this ion. The TRPV1 is also a proton channel [76] , which lowers the pH of the cytosol [76] . However, the Ca2+ -selective ORAI channel [77] has tight interactions with non and weak Ca2+ selective TRP channels and may activate the TRPCs while that may localize the ORAI [78] .

The TRP channels have an exceptionally high temperature-coefficient Q10 [79] . It is notable that both the enthalpy and entropy components of its transition through the ion-permeability barrier are high [79] .

The vanilloid receptors (TRPVs) cover the temperature sensing in mammals from low skin temperature (~25˚C - 45˚C TRPV4) to the necrotic high up to ~50˚C - 60˚C, TRPV2 [80] [81] . The hyperthermia fits TRPV3 (~24˚C - 34˚C) and TRPV1 (~41˚C - 50˚C) ranges [81] .

The cell-membrane rafts became in focus [82] and well-studied [83] . The TRP receptors could also be a part of these clustered microdomains in the membrane and present effective thermosensors of the cell [84] . There are essential observations indicating a coherent cluster structure of a large number (~105) of voltage-gated ionic channels [85] [86] , and it could have transient receptor potential (TRP) receptors in one temperature-sensing domain [84] . Small temperature changes may affect the TRP channels with membrane lipid assistance in the raft microdomain. The membranes are inhomogeneous, which is enhanced by the mild temperature change. Notably, the activated TRPV1 channel’s ionic current may disappear at the multiple repetitions of the thermal ignition [87] . The opening of the TRPV channel for ionic current needs a relatively large enthalpy, while its closing depends less on the provided energy [88] . This asymmetry works oppositely in cold-activated channels like TRPM8 [89] .

The primary energy absorbers in the mEHT method are the cytoplasmic membrane rafts, heating them selectively. The selection is based on these microdomains’ high specific absorption rate (SAR). The thermal influence of mEHT has traditional hyperthermic functions, promoting the cell death in various ways [90] . Nevertheless, the difference is significant: the mEHT selectively heats the tumor. The selection focuses on membrane rafts, the cholesterol-stabilized microdomain cluster of transmembrane proteins [82] , participating in the membrane dynamics [83] . The rafts have exceptionally high energy absorption from the RF current [91] , allowing cellular selection of malignant cells with significantly high raft density [92] without substantially heating the healthy ones [93] . The concentrated heating of the molecular groups in the membrane rafts creates a heterogenic situation, where the rafts heat the entire cell and, in a second step, the tumor [94] . While the tumor, on average, remains in the <40˚C fever range, the rafts reach < 3˚C higher temperatures [95] .

The temperature increase of the nano parts in the target is negligible with homogeneous heating [96] . However, when the heating is heterogenic (which is the case of all nanoparticle heating and in the mEHT too), the local SAR could be extremely high due to the small particle size. This is used in nanoparticle heating when the SAR on the nanoparticles could be as much as more than 1 MW/g (1,000,000,000 W/kg), depending on the absorber’s concentration [97] .

The inhomogeneous electric field in the case of mEHT, where the dielectrophoretic force drifts the rafts forwards, gives an additional factor to increase the micro-heterogeneity of cellular heating (Figure 4). This process exhibits a significant level of selectivity because the dielectric permittivity of the transmembrane proteins is at least two orders of magnitude higher than the permittivity of the surrounding membrane through which they traverse [98] .

The TRP regulative processes are dynamic. The transmembrane protein displaces by the temperature action [99] . In malignant cells, the motility of transient receptor potential (TRP) channels is more pronounced compared to their healthy counterparts [100] . As a result, the drift movement of these channels indicates regions of higher energy density, where the specific energy absorption (SAR) is also elevated. The SAR values increase specifically at these points of the membrane known as micro-contacts (Figure 5(a)).

The telophase → cytokinesis phase forms another significant vital direct contact during the mitotic spindle, which has high importance in the proliferative

Figure 4. The electric field polarizes the membrane. Hyperpolarizing happens on one and depolarizing on the opposite side of the cell. The rafts have the electrophoretic force to drift by the electric force. The mEHT uses a 13.56 MHz carrier frequency. Consequently, the direction of all processes changes by ~0.07 μs, and so the movable proteins will be enriched in both sides of the cell.

Figure 5. Developing extra hot spots on the cell membrane. (a) The cells touch each other, which enhances the SAR value and increases the spot’s temperature. (b) The cytokinetic phase of mitosis has a “neck” between the forming daughter cells. This small area behaves like a touching point shown in Figure 5(a).

malignant cells. The neck between the just-forming daughter cells induces cataphoretic forces [101] . The small cross-section of the neck may absorb an extremely high SAR when its directional position matches to electric-field vector. The absorbed energy by the cytokinetic “neck” depends on the cellular orientation, having a maximum when the field lines are directly parallel to the cytokinetic “neck” [102] [103] . The exceptionally concentrated energy may arrest the cytokinesis and block malignant proliferation [104] (Figure 5(b)).

In the outer membrane, dominantly the TRPV1, while in the membrane of intracellular compartments, the TRPV3 is thermally activated by mEHT’s heterogeneous heating. The TRPs actively participate in membrane fusion and fission, signal transduction, and general vesicular homeostasis [75] .

The thermal load has another well-known consequence: the development of protective chaperoning heat-shock proteins (HSPs) [105] [106] . The HSPs are also part of the complex regulation of the living organization, resulting in cellular defense or promoting cell death [107] . The complexity of HSPs questions their role as a “friend” or “foe” [108] [109] [110] . This dual behavior [111] [112] appears to decide their function as inflammatory or anti-inflammatory, pro-tumor or antitumor, immune-stimulatory or immune-suppressant, etc. The intensive thermal stress secrets membrane [113] and extracellular HSPs [114] , which may reverse their cell-protecting activity [115] . The HSP expression can link to the plasma membrane processes by mild heat [116] , which may cause non-specific clustering [117] by fever-like temperatures, where TRPs are particularly sensitive.

3. Nonthermal Impact

The electromagnetic effects differ between healthy and cancerous tissues. The essential differences appear in the conductivity and dielectric properties of the tissues. The breaking healthy cellular network in cancer better conducts the radiofrequency current, and its dielectric permittivity (polarizability) is also significantly higher. Further differences appear in the electromagnetic excitability of the signal pathways in the cells due to the expressive contrasts of the healthy and malignant cellular membranes. The low membrane potential and high number of intercellularly unconnected transmembrane proteins appear in malignancy which interacts profoundly differently with the external electric field than the healthy cell. Using the apoptotic calibration in Figure 3, the impact of mEHT for cell-killing is rather significant (Figure 6). The basic structural disruption of healthy order makes the tumor also distinguishable by its electromagnetic interactions.

Any other than thermal stress that influences the homeostatic equilibrium also activates the HSP synthesis [118] , which induces the cellular chaperoning function with HSPs. Living objects have not only thermal interactions. The dominant number of living regulation effects is not feasible with thermal effects. Enzymatic reactions and other molecular changes are mostly nonthermal, and their functions are mandatory for life. Thermal conditions are responsible for optimizing the nonthermal chemical reactions, and many biological processes

Figure 6. The heterogenic impact of mEHT (p) compared to homogeneous heating (l). The percentage of apoptosis shows much higher temperature of the process than the measured 42˚C in the cell culture medium. (a) HepG2 [65] , (b) U937 [64] .

synergize thermal and nonthermal components [119] . RF radiation induces nonthermal effects [120] together with the well-known thermal one for conventional hyperthermia treatments [121] . There is a large family devoted to voltage-gating ionic control [122] [123] , which has an also large subfamily of voltage-gated calcium channels [124] .

Even the thermally sensitive TRP channels also have nonthermal voltage dependence [125] , which is effective in different temperature ranges. The high value of the transition entropy of TRPVs [126] shows that the transition follows Eyring’s theorem [127] . This transition explains how the enzymatic reactions decrease the energy barrier with quantum-mechanical effects (e.g., tunneling [128] ) and what mechanism is behind the reaction rate changes of catalyzes. The interdisciplinary applications [129] could use the quantum-mechanical considerations [130] [131] of the transition process, making possible a first-order phase transition when only the entropy changes the overall reaction rate, and the temperature remains constant when a new phase appears. The process is nonthermal, at least in its part. In this way, the large entropy in the TRPV transition allows the nonthermal transition when the temperature is unchanged but optimizes the complete process. The high entropy points the structural changes, which have a critical role in the nonthermal activity [132] . The probability of changes has a similar expression for temperature and electric field [133] , which further supports the nonthermal activity. Notable, while the repetition of thermal stimuli decreases the signal level [90] , the repetition of the nonthermal chemical stimuli increases it [68] .

The electric field produces such excitations, which is impossible with thermal conditions [21] [22] . A notable example showing the exceptional excitation ability of electromagnetic field is that using a broad-band (0.2 - 20 MHz) signal increases the HSP70 expression [134] in such volume that to produce the same rise of HSP70 by temperature, the perturbation should have been 14 orders of magnitude greater [135] . The electric field significantly modifies the cells [136] , manipulates protein expressions, and induces extrinsic molecular pathways [137] . The modulated electric field’s vibrational effect, along with the associated electro-osmotic process, operates within the intracellular environment [138] , potentially triggering various processes within intracellular compartments. The polarization and polymerization may affect the cytoskeleton structure and the pathways that use it. The network of cytoskeleton forms and changes by the various signal transmissions in the cells. These polarizable parts of the cytoskeleton (their main components are the microfilaments, intermediate filaments, and microtubules, and their building blocks the tubulin dimers the actin subunits, and the fibrous subunits) are all interacting with external fields and are all capable of rapid growth or disassembly. The field-sensitive polymerization processes are the most basic of the mitotic rebuilding of the cell-structure of the newly born daughter cells. The membrane rafts incorporating thermosensitive TRPVs could considerably increase the Ca2+ influx to the cytoplasm by heating. The Ca2+ ionic balance has an important controlling function in homeostasis. It controls several processes in tumorigenesis [139] . It participates in gene transcription [140] , cellular motility [141] , invasion [142] , cell-cycle regulation [143] , and angiogenesis [144] . The balance between cell proliferation and apoptosis is tightly regulated by the influx of calcium ions Ca2+. The intracellular concentration of Ca2+ plays a crucial role in determining whether a cell undergoes division or apoptosis [145] . Several steps of the killing of cancer cells are Ca2+ dependent [146] . The cytosolic and mitochondrial Ca2+ overload strongly stimulates the apoptotic processes [147] [148] , but, as is usual in complex processes, the low nM Ca2+ concentration may help the survival of malignant cells by promoting proliferation by lowering the membrane potential [149] [150] and increasing the malignant differentiation [151] . On the contrary, the rise of intracellular Ca2+ concentrations to μM may support apoptosis [152] .

The regulation of various pathways by Ca2+ together with its concentration dependence, a temporal component has an important role in tumorigenesis [153] . The amplitude and duration of Ca2+ signals involved are different proinflammatory activation of B lymphocytes [154] by decoding its information of amplitude and duration. Amplitude modulation of the Ca2+ signal may produce positive or negative antigen response in gene activation of B cells [155] [156] .

Bioelectromagnetic interactions make numerous molecular excitations and chemical changes, which thermal interactions cannot achieve [157] . For example, the various intensities of electromagnetic signals make an entirely different activity of the hormone 5-hydroxytryptamine, which could even cause transepithelial potential oscillation [157] .

These nonthermal effects characterize the mEHT method [21] [22] and appear in all molecular changes made by thermal effects, even in the TRPV thermal sensors [158] and HSPs [159] [160] . The mEHT uses the gaining possibility of the cell membrane. According to in-silico models, the electric field-strength gain, which refers to the ratio of the induced field within a material compared to the externally applied field, is highest in the cell membrane [161] . Specifically, for frequencies up to a few tens of MHz, the gain remains approximately ≈ 5 × 103 and follows a power-law decrease of 1/f as the frequency increases [161] . In more realistic tissue models, the membrane gain varies depending on the cell’s position within the tissue, but it never falls below 102 in tissue arrangements [162] . In the case of cancer cells, while the intracellular gain is comparable to that of non-cancerous cells, the membrane gain in malignant cells is twice as high as that of their healthy counterparts [162] .

The TRPV investigation of activating and deactivating potentials proves their voltage dependence [163] and that thermal conditions optimize voltage action [164] . In addition to heat and electric effects (like potential change or proton influence), the chemical effects (like capsaicin) and proinflammatory cytokines may activate the TRPV1 [165] , which is upregulated in many cancer types [166] . The multiple influences make synergy [130] [167] . The multi-sensing behavior makes these channels critical for communication with the changes in TME. The voltage alone cannot wholly activate the TRPV1 channel [168] ; the thermal component works for it in synergy, in the complexity of the membrane potential, ligand binding, mechanical force, and temperature [169] .

An essential aspect of TRPs activity is its interplay with the cytoskeleton [103] , which is based on electromagnetic interactions [170] . The role of the cytoskeleton in signal transduction and its connective role between the intra and extracellular information exchange makes it especially important. The connecting structure of TRP and cytoskeleton allows Ca2+ independent signaling [171] . TRPV1 essentially regulates the dynamics of the cytoskeleton by colocalization and stable binding with microtubules when there it is resting. However, in an excited state, TRPV1 rapidly disassembles the microtubule polymers [169] .

The microtubules of the cytoskeletal network have a polymer structure [172] . The loss of the polymerization order of the cytoskeleton probably causes the high motility of cancer cells [173] because it makes the cells especially soft and detachable [174] . The increasing motility induces high metastatic potential and high deformability [175] . The extracellular matrix (ECM) plays a role in the cellular motility of cancer cells connected to its rigidity [176] . The heightened motility observed in cells is likely attributed to the loss of polymerization order within the cytoskeleton [172] , resulting in increased cell softness and mobility [173] .

The polymerization process follows a chain polymerization model known as Einstein’s polymer [177] [178] . However, this model is unable to account for multi-bonding processes where chemical bonds can form branches in tubulins, leading to the creation of various space-filling structures. The reorganization of the cytoskeleton also promotes the formation of multi-strand cases. Multi-strand structures have longer chains compared to single-strand structures due to their multiple free ends and energy centers, making them energetically less favorable. As a result, according to Boltzmann statistics, the concentration of multi-strand chains is lower than that of single-strand chains. There is a relationship between the polymer concentration [Mn] and the polymer length, expressed in terms of the number of monomers, denoted as “n” as

[ M n ] e n n 0 (1)

where n0 is constant. Consequently, the high concentration has shorter polymers.

The modulation employed in modulated electro-hyperthermia (mEHT) amplifies the effective electric field, thereby supporting the polymerization and reorganization of the cytoskeletal network. When using 1/f fractal noise modulation [15] [44] [45] , this effect becomes more potent due to the continuous spectrum of frequencies present in the non-discrete noise signal. The application of noise modulation bears a resemblance to the harmonizing method [179] , which is gaining recognition in the field of physiology [180] . In cancer cells, cytoskeletal polymerization holds particular significance, as the destabilized and incomplete polymerization of the cytoskeleton contributes to increased cell motility and facilitates metastatic spread.

The influence of the modulated electromagnetic field on the cytoskeleton can also involve voltage-sensitive phosphatase (VSP) [181] . Field-controlled phosphorous hydrolysis mediated by VSP could play crucial roles in cytoskeletal restructuring and exhibit resonant-type behavior. VSP, a macromolecule with a voltage sensor and cytoplasmic phosphatase domains [182] , regulates the influx of calcium ions Ca2+ into cells [183] . VSP is sensitive to external fields and operates within the cytoplasm, allowing the transmission of external field effects to the cell interior. This process generates biochemical signals that may contribute to intracellular organization. Through these signals, it is possible to generate biochemical cues within the cytosol that can control internal processes, most likely including cytoskeletal polymerization. The fundamental mechanism involves membrane depolarization leading to phosphoinositol hydrolysis [184] . This is a reversible decomposition reaction that the external electric field may modify.

Phosphorylation plays a crucial role in regulating the activity of microtubule-associated proteins (MAPs) within the cytoskeletal network. The activation or deactivation of phosphorous groups controls the functioning of MAPs. Specifically, the phosphorylation of MAPs destabilizes microtubules by weakening the internal bonds that contribute to their structural stability [185] . The membrane potential of proliferative cells has a lower absolute value than that of quiescent neighbors [186] . Due to this, the malignant cells present low membrane potential [187] . Consequently, the VSPs influence the cytoskeleton in the permanently depolarized cancerous cells. The low level of cytoskeletal polymerization supports the proliferation and mobility of malignant cells.

The phosphorylation of MAPs not only affects microtubule stability but also plays a crucial role in the proper functioning of various ion channels, transporters, and vesicle movement within the cell. This mechanism enables the active modulation of intercellular electrolyte levels and protein connections in response to external electric fields. The dynamic stability of the system is governed by the Le Chatelier principle: a sudden change in membrane potential triggers phosphorylation, leading to an increase in potassium transport and the simultaneous suppression of sodium transport. This intricate process aims to restore the original membrane potential and effectively interacts with the cell proliferation process. The phosphorylation of VSP is energized by ATP. This energy consumption decreases ATP concentration, which increases the depolarization of the cell membrane by suppressing the other, ATP-dependent active membrane transport of ions. The VSP has a role in anaerobic glycolysis and cancerous transformation when permanent stress conditions massively demand ATP. These conditions may activate the oncogenes, inhibiting apoptosis and producing high concentrations of stress proteins. This situation combats normal homeostatic regulation, so it is ideal for developing cancer.

The electric field polarizes the cytoskeleton’s fibers [188] , so it reorganizes the cytoskeleton in static (direct current, DC-field) [189] and dynamic (alternating current, AC-field) conditions [188] . The AC has the greatest influence at around 1 Hz [190] , while in amplitude modulation of high frequency, it is optimal around 16 Hz (Adey window) [191] . This phenomenon exhibits resonant effects that can be described by stochastic resonance [192] . It assumes a bistable two-position state of voltage-sensitive phosphatase (VSP), like voltage-gated ion channels [193] . In the presence of a DC electric field, the membrane polarizes in opposite directions at different sides of the cell relative to the field vector. One side becomes hyperpolarized while the opposite side becomes depolarized. Depolarization triggers phosphorous hydrolysis, initiating cytoskeletal formation on the hyperpolarized side. However, in the presence of an AC electric field, both sides exhibit stochastic resonance and become hyperpolarized, leading to cytoskeletal reorganization. In the belt region perpendicular to the DC field, where there is no resonance, phosphorylation proceeds normally, creating a general phosphorous gradient in this region. The reorganization of the cytoskeleton is driven by specific forces, resulting in a pattern perpendicular to the pattern induced by the DC field.

The AC electric field induces resonance-like reorganization of the cytoskeleton, with a distinct peak at a specific frequency dependent on thermal noise [193] . It has been rigorously demonstrated that amplitude-modulated carrier frequencies can generate stochastic resonance, leading to various biological effects [193] . This phenomenon selectively stimulates enzymatic reactions, activates and deactivates voltage-gated ion channels, and reorganizes cytoskeletal polymerization processes [43] [45] . Moreover, the amplitude-modulated carrier can modify complex processes [44] . The modulating signal alone may also produce resonances. However, the requested low frequency alone, due to the impedance barrier of the skin, does not penetrate deep enough into the body. The 0.1 - 15 MHz high frequency has enough penetration to the human body and, as a carrier, delivers the signal for stochastic resonances. The optimal carrier frequency, as used in mEHT, may select the cancer cells with β, δ frequency dispersions [119] [194] and cause definitive apoptotic cell destruction of malignant cells [23] .

Another approach is when the cytoskeleton polarization is optimized without modulation in high frequency (~0.2 MHz). These tumor-treating fields (TTF) target the cytokinetic “neck” with nonthermal polarization effects [195] . The electric field generated by TTF (Time-varying Tumor Treating Fields) exerts an influence on the polarizable microtubules and actin fibers within the cell. It has the potential to reorient these structures and, importantly, can impede the polymerization process of the cytoskeleton and hinder the assembly of the mitotic spindle [105] . The process does not use considerable SAR as mEHT does to arrest the proliferation by targeting the cytokinetic neck. This difference affects the treatment protocol. The TTF must be applied 18 hours/day for months, while mEHT with thermal optimizing has only 10 - 12 treatments for 60 min each.

4. Immune Effects

Immunogenic cell death is one of the main advantages of mEHT [196] , which produces damage-associated molecular pattern (DAMP) [197] by extrinsic excitation of apoptotic pathways through various channels, including the TRAIL-R2 (DR5) death receptors [198] with the complex interaction of FAD + FASS molecules [41] [199] . The concomitant immune-stimulative treatments with dendritic cells [33] [66] and another stimulator [200] , with the applied bioelectromagnetic forces by modulated RF signal [19] [44] [45] [201] , improve the immunogenic processes. The immunogenic effects represent the oncology trend, especially oncological hyperthermia oncologic hyperthermia [18] .

Due to the missing apoptosis, intensive proliferation is the assertive behavior of cancer. The mEHT, with its bioelectromagnetic excitations, promotes a massive preference for apoptosis against proliferative survival [23] [38] [202] [203] . Multiple excitable transmembrane proteins exist in the malignant cells (Figure 7(a)). The excitation could be thermal or nonthermal, but dominantly the synergy of the two, when the thermal process ensures the optimal conditions for the nonthermal excitation. The apoptosis may go through various signal pathways (Figure 7(b)).

The apoptosis uses an extrinsic pathway through Caspase-8 (Cas-8), an intrinsic pathway (Cas-9) finally, with Cas-3, to the programmed cell death (Figure 8). The caspase-independent way through apoptosis-inducing factor (AIF) is also activated, making it possible to execute the apoptosis when the caspase paths are blocked.

When one pathway is blocked by the malignant evasion of apoptosis, the other pathways may completely substitute the missing line.

The elevated levels of BAX observed in the affected cells [205] [206] further support the apoptotic effect induced by the treatment. The increased presence of BAX suggests the activation of apoptotic pathways and reinforces the notion that the treatment is triggering programmed cell death in the affected cells. The selective energy absorption of mEHT produces heterogenic membrane temperature, intensively heating the transmembrane proteins, including the TRPV channels. The TRPV promoted Ca2+ influx massively overloads the intracellular

Figure 7. The critical protein involvements in apoptosis by mEHT. (a) The caspase involvements relative to the untreated samples (HepG2 cell-line in vitro [65] ). (b) The AIF percentages in the samples from in vivo xenograft experiments for A2058 melanoma [204] and HT29 colorectal carcinoma [205] tumors.

Figure 8. Forming apoptotic signals. (a) Multiple excitable transmembrane proteins accept the thermal and nonthermal effects from mEHT. (b) Multiple signal pathways are available, ensuring that a block in one can not terminate the apoptotic process.

conditions with Ca2+ concentration. The thermal and nonthermal synergy of mEHT ensures the requested apoptotic level of Ca2+ overload.

The thermal and nonthermal stresses imposed on homeostatic control typically stimulate the production of chaperone stress proteins. The chaperoning HSPs are exhausted by mEHT and cannot meet the chaperone function to protect the cells against apoptosis [207] . A part of HSP70s is secreted on the membrane [119] , and this localization [208] promotes apoptosis [209] and has a vital role in the membrane “fluid’ to keep it functional [210] .

While cancer has strong proliferation, it is weak in its loneliness. The cellular autonomy is a weak side, which offers a correction possibility. Immune surveillance is critical in attacking the weakness of malignancy and guaranteeing homeostatic balance. The counterbalance of the evasion of immune effects by malignant cells needs local and systemic activity, which rebuilds the standard healthy conditions. The autonomy of malignant cells shows the breaking of adherent and junctional connections with the neighboring cells, substantially modifying the homeostatic electrolyte composition and concentration in the TME [26] , which offers an electric selection factor [211] [212] . The reorganization of the cytoskeleton by nonthermal electric polarization of mEHT promotes the form (β-catenin + E-cadherin) complexes [65] [213] , giving a possibility to reestablish the lost adherent connections and fix the cancer cells in their position, block the dissemination.

Immune control creates an additional possibility to block the cellular autonomy of malignant cells by destroying the freely moving cancer cells. The great and unique advantage of tumor-specific immune activity is its ability to find and eliminate distant micro and macro metastases locally and systemically. The frequently applied CAR-T method has the same purpose [214] , preparing tumor-specific T-cells from the patient’s blood sample. Immune effects integrate the TRP channels for multiple purposes [215] . All immune cells, including NK cells [216] , T-cells [217] , and dendritic cells [218] ), have TRPV1 channels with significant functions. The CD34+ hematopoietic stem cells express TRPV2, which is also expressed in granulocytes, monocytes, and CD56+ natural killer cells and orchestrates the Ca2+ signals in CD4+ and CD8+ T-cells and CD19+ B lymphocytes [219] . Moreover, TRPV1 modulates macrophage-mediated responses [220] .

The adaptive immune reaction was measured, detecting significant development of DC cells (S100) for maturation (antigen-presenting) CD3+ CD4+, CD8+ T-cells and suppressing Treg cell-population (Foxp3) (Figure 9) [33] .

The TRPV1 and TRPV4 channels have been implicated in T-cell activation and the production of effector cytokines. These channels play a role in suppressing the release of tumor necrosis factor (TNF) and interleukin-2 (IL-2), which are important immune signaling molecules involved in inflammation and immune responses [217] . Furthermore, the TRP channels have a critical role in controlling phagocytosis, the production of chemokines and cytokines, and cell survival [221] .

The HSPs are not less important in cell fate and immune activity than TRPs. The thermal and nonthermal stress combination overloads the malignant cells

Figure 9. The relative development of the critical immune-surveillance cells in SCCVII allograft after mEHT [33] . Measured with (a) flow cytometry and (b) optical density.

with chaperoning HSPs, which have much more function than only chaperoning, depending on their position in the cell [210] . To initiate the apoptotic process, the mitochondrial heat shock proteins (mHSPs) must first bind to a complex formed by tumor peptides [222] . The membrane-localized HSP70s [223] are mainly localized in the rafts [224] and activate the NK cells in the immune response [204] [225] .

A part of the HSPs may leave the cells and become extracellular [226] . Their importance in building up an appropriate immune response is crucial. These released molecules deliver tumor-specific information for antigen presentation to develop antitumoral immune reaction [115] , which process attracts much attention in studies of the overall immune reactions of the bio-systems [227] [228] . The antigen-presenting cells (APCs) develop the tumor-specific CD8+ killer and CD4+ helper T-cells, which are delivered by the bloodstream and combating with the cancer cells in the entire body (abscopal effect) [229] . The local treatment developed a whole-body effect by mEHT [230] (Figure 10). The in-situ and real-time production of tumor-specific immune activity is the advantage of mEHT [231] .

Significant development of the DAMP molecules characterizes the results of mEHT treatment. The abscopal effect of distant, untreated tumors was observed when immune stimulation was added to the protocol (Figure 11) [200] .

The thermal and nonthermal effects work in synergy to produce the in-situ immune effects with mEHT. In the thermal aspect, the relative mild tissue temperature is crucial for immune development. The enhanced temperature eases the enzymatic processes and increases the molecular reaction rate, but a higher temperature than 40˚C blocks the activity of the immune cells [232] and so does not allow the real-time processes for APC prepared by immunogenic cell death [231] . Furthermore, the > 40˚C temperature paralyzes the NK cells to attack the cells with the marks of the HSP70s on their membrane [233] . The immune cells may restore their activity with elapsing time [234] or by bloodstream replace

Figure 10. The thermal and nonthermal processes develop immunogenic cell death. The genetic information used by antigen-presenting produces tumor-specific T-cells. In this way, the adaptive immune machinery starts a systemic attack of malignant cells all over the system.

Figure 11. Development of DAMP molecules in colorectal allograft preclinical experiment of C26 tumor [22, 200]. (a) The sham treatment and untreated tumor on the same animal had treatment at a distant location. (b) Development of the DAMPs in the treated and untreated tumors on mice. (c) Development of DAMPs in treated and untreated locations when immune stimulants were added to the protocol.

them from the non-treated parts of the body, but the real-time processes with the simultaneous reactions vanish. The heterogenic heating of mEHT with the high temperature of rafts and simultaneously mild of the tumor microenvironment (TME) solves the contradictory demands of the hyperthermic temperature range and the immune requirements.

The nonthermal processes in mEHT are also essential for the abscopal effect. The excitation of the TRP channels sensitizes the immune lymphocytes [215] and promotes apoptosis [101] , helping with the ICD processes. The substantial nonthermal stress exhausts the evasion of apoptosis in cancer cells and produces transmembrane HSP70, which also supports apoptosis [209] . It activates ligands for NK cells [208] , forming an innate immune response [235] , especially in tumor cells [223] . The well-forming temporal order by mEHT of membrane secretion calreticulin, the extracellular release of HMGB1 and HSP70 developing the APCs for an active adaptive immune response [231] .

Numerous preclinical studies prove the specialties of mEHT (Figure 12). A recent review summarizes the results [236] .

Based on the regulatory conditions the mEHT method received the necessary certifications and numerous clinical trials were performed in different hospitals in various countries (Figure 13). Some protocols allow geriatric and pediatric

Figure 12. Some preclinical experiments with mEHT method. The numbers refer on the references: 1. = [42] , 2. = [237] , 3. = [238] , 4. = [94] , 5. = [202] , 6. = [198] , 7. = [239] , 8. = [65] , 9. = [207] , 10. = [203] , 11. = [240] , 12. = [64] , 13. = [207] , 14. = [206] , 15. = [204] , 16. = [32] , 17. = [205] , 18. = [66] , 19. = [241] , 20. = [200] , 21. = [33] , 23. = [205] .

considerations, too. A recent review summarizes the results up to 2019 [242] .

5. Conclusion

The modulated electro-hyperthermia focuses on the nonthermal and thermal effects synergy when the thermal component provides optimal conditions for the nonthermal electric molecular excitation. In this review, we concentrate on the role of the ionic channels as TRPs, VSPs, and voltage-gated channels in the selective antitumoral processes. These transmembrane compartments primarily

Figure 13. The clinical studies with mEHT. It contains various levels of evidence including case reports and phase II/III trials. The numbers refer to the references: 1. = [243] , 2. = [244] , 3. = [245] , 4. = [246] , 5. = [247] , 6. = [248] , 7. = [249] , 8. = [250] , 9. = [251] , 10. = [252] , 11. = [253] , 12. = [254] , 13. = [255] , 14. = [256] , 15. = [257] , 16. = [258] , 17. = [259] , 18. = [260] , 19. = [261] , 20. = [262] , 21. = [263] , 22. = [264] , 23. = [265] , 24. = [266] , 25. = [267] , 26. = [268] , 27. = [269] , 28. = [270] , 29. = [254] , 30. = [271] , 31. = [272] , 32. = [273] , 33. = [274] , 34. = [275] , 35. = [276] , 36. = [277] , 37. = [278] , 38. = [20] .

promote the Ca2+ and the H + influxes, interact with the cytoskeleton and are involved in the apoptotic signal pathways. The DAMP forming TRAIL-FAS-FADD excited extrinsic apoptotic signal combined with the Ca2+ induced apoptosis ensures a “gentle” distortion of the malignant cells, which, together with the fully functioning other DAMP molecules, uses the membrane secreted and extracellularly released HSPs with the exhaustion of their intracellular chaperoning to form APCs. The innate and adaptive tumor-specific immune activity appears by the membrane HSP70 and the APC-produced killer and helper T-cells. The bloodstream-delivered T-cells attack the cancer cells all over the body (abscopal effect), so the immunogen processes transform the local mEHT to systemic.

Acknowledgements

This work was supported by the Hungarian National Research Development and Innovation Office PIACI KFI grant: 2019-1.1.1-PIACI-KFI-2019-00011.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Valone, T.F. (2013) Bioelectromagnetic Healing, Its History and a Rationale for Its Use.
http://www.medicinacomplementar.com.br/biblioteca/pdfs/Cancer/ca-10282.pdf
[2] Liboff, A.R. (1994) Chap. 6. The Electromagnetic Field as a Biological Variable. In: Allan, H.F., Ed., On the Nature of Electromagnetic Field Interactions with Biological Systems, R.G. Landes Company, Austin, 73-82.
[3] Bear, M.F., Connors, B.W. and Paradiso, M.A. (2006) Neuroscience: Exploring the Brain. 3rd Edition, Lippincott, Philadelphia.
[4] McCaig, C.D., Rajnicek, A.M., Song, B., et al. (2005) Controlling Cell Behaviour Electrically: Current Views and Future Potential. Physiological Reviews, 85, 943-978.
https://doi.org/10.1152/physrev.00020.2004
[5] Nordenstrom, B.W.E. (1983) Biologically Closed Electric Circuits: Clinical Experimental and Theoretical Evidence for an Additional Circulatory System. Nordic Medical Publications, Stockholm.
[6] Nordenstrom, B.W.E. (1998) Exploring BCEC-Systems (Biologically Closed Electric Circuits). Nordic Medical Publications, Stockholm, Sweden.
[7] Nordenström, B.E.W. (1978) Preliminary Clinical Trials of Electrophoretic Ionization in the Treatment of Malignant Tumors. IRCS Journal of Medical Science, 6, 537-540.
[8] Nordenström, B.E.W. (1985) Electrochemical Treatment of Cancer. Annales De Radiologie, 28, 128-129.
[9] Loewenstein, W.R. (1999) The Touchstone of Life, Molecular Information, Cell Communication and the Foundations of the Life. Oxford University Press, Oxford, New York, 298-304.
[10] Zhao, M., Song, B., Pu, J., et al. (2006) Electrical Signals Control Wound Healing through Phosphatidylinositol-3-OH Kinase- and PTEN. Nature, 442, 457-460.
https://doi.org/10.1038/nature04925
[11] Becker, R.O. and Selden, G. (1985) The Body Electric, Electromagnetism and the Foundation of Life. William Morrow & Co., New York.
[12] Becker, R.O. (1990) Cross Currents. Jeremy P. Tarcher Inc., Los Angeles.
[13] Lin, J.C. (1989) Electromagnetic Interaction with Biological Systems. Pergamon Press, New York, London.
https://doi.org/10.1007/978-1-4684-8059-7
[14] Bersani, F. (1999) Electricity and Magnetism in Biology and Medicine. Kluwer Academic/Plenum Publishers, New York, Boston.
https://doi.org/10.1007/978-1-4615-4867-6
[15] Szasz, A. (2021) Therapeutic Basis of Electromagnetic Resonances and Signal-Modulation. Open Journal of Biophysics, 11, 314-350.
https://doi.org/10.4236/ojbiphy.2021.113011
[16] Szasz, A. (2022) Cancer Specific Resonances. Open Journal of Biophysics, 12, 185-222.
https://doi.org/10.4236/ojbiphy.2022.124009
[17] Foletti, A., Grimaldi, S., Lisi, A., Ledda, M. and Liboff, A.R. (2023) Bioelctromagnetic Medicine: The Role of Resonance Signaling. Electromagnetic Biology and Medicine, 116, 484-499.
[18] Lee, S.Y., Fiorentini, G., Szasz, A.M., Szigeti, G., Szasz, A. and Minnaar, C.A. (2020) Quo Vadis Oncological Hyperthermia (2020)? Frontiers in Oncology, 10, Article 1690.
https://doi.org/10.3389/fonc.2020.01690
[19] Szasz, O. (2019) Bioelectromagnetic Paradigm of Cancer Treatment—Modulated Electro-Hyperthermia (mEHT). Open Journal of Biophysics, 9, 98-109.
https://doi.org/10.4236/ojbiphy.2019.92008
[20] Lee, S.Y., Kim, J.H., Han, Y.H., et al. (2018) The Effect of Modulated Electro-Hyperthermia on Temperature and Blood Flow in Human Cervical Carcinoma. International Journal of Hyperthermia, 34, 953-960.
https://doi.org/10.1080/02656736.2018.1423709
[21] Wust, P., Kortum, B., Strauss, U., Nadobny, J., Zschaeck, S., Beck, M., et al. (2020) Non-Thermal Effects of Radiofrequency Electromagnetic Fields. Scientific Reports, 10, Article No. 13488.
https://doi.org/10.1038/s41598-020-69561-3
[22] Wust, P., Stein, U. and Ghadjar, P. (2021) Non-Thermal Membrane Effects of Electromagnetic Fields and Therapeutic Applications in Oncology. International Journal of Hyperthermia, 38, 715-731.
https://doi.org/10.1080/02656736.2021.1914354
[23] Krenacs, T., Meggyeshazi, N., et al. (2020) Modulated Electro-Hyperthermia-Induced Tumor Damage Mechanisms Revealed in Cancer Models. International Journal of Molecular Sciences, 21, Article 6270.
https://doi.org/10.3390/ijms21176270
[24] Lovelady, D.C., Richmond, T.C., Maggi, A.N., Lo, C.M. and Rabson, D.A. (2007) Distinguishing Cancerous from Noncancerous Cells through Analysis of Electrical Noise. Physical Review E, 76, Article ID: 041908.
https://doi.org/10.1103/PhysRevE.76.041908
[25] Szigeti, G.P., Szasz, O. and Hegyi, G. (2017) Connections between Warburg’s and Szentgyorgyi’s Approach about the Causes of Cancer. Journal of Neoplasm, 1, 1-13.
[26] Foster, K.R. and Schepps, J.L. (1981) Dielectric Properties of Tumor, and Normal Tissues at Radio through Microwave Frequencies. Journal of Microwave Power, 16, 107-119.
https://doi.org/10.1080/16070658.1981.11689230
[27] Blad, B. and Baldetorp, B. (1996) Impedance Spectra of Tumour Tissue in Comparison with Normal Tissue; A Possible Clinical Application for Electric Impedance Tomography. Physiological Measurement, 17, A105-A115.
https://doi.org/10.1088/0967-3334/17/4A/015
[28] Mikac, U., Demsar, F., Beravs, K. and Sersa, I. (2001) Magnetic Resonance Imaging of Alternating Electric Currents. Magnetic Resonance Imaging, 19, 845-856.
https://doi.org/10.1016/S0730-725X(01)00393-9
[29] Joy, M.L.G. (2004) MR Current Density, and Conductivity Imaging: The State of the Art, Engineering in Medicine, and Biology. IEMBS 26th Annual International Conference of the IEEE, 2, 5315-5319.
[30] Oh, S.H. (2003) Conductivity, and Current Density Image Reconstruction Using Harmonic Bz Algorithm in Magnetic Resonance Electrical Impedance Tomography. Physics in Medicine & Biology, 48, 3101-3116.
https://doi.org/10.1088/0031-9155/48/19/001
[31] Muftuler, T.L., Hamamura, M.J., Birgul, O., et al. (2006) In Vivo MRI Electrical Impedance Tomography (MREIT) of Tumors. Technology in Cancer Research, and Treatment, 5, 381-387.
[32] Besztercei, B., Vancsik, T., Benedek, A., et al. (2019) Stress-Induced, p53-Mediated Tumor Growth Inhibition of Melanoma by Modulated Electrohyperthermia in Mouse Models without Major Immunogenic Effects. International Journal of Molecular Sciences, 20, Article 4019.
https://www.mdpi.com/1422-0067/20/16/4019
https://doi.org/10.3390/ijms20164019
[33] Qin, W., Akutsu, Y., Andocs, G., et al. (2014) Modulated Electro-Hyperthermia Enhances Dendritic Cell Therapy through an Abscopal Effect in Mice. Oncology Reports, 32, 2373-2379.
https://doi.org/10.3892/or.2014.3500
[34] Thomas, M.B., Major, E., Benedek, A., Horváth, I., Máthé, D., Bergmann, R., et al. (2020) Suppression of Metastatic Melanoma Growth in Lung by Modulated Electro-Hyperthermia Monitored by a Minimally Invasive Heat Stress Testing Approach in Mice. Cancers, 12, Article 3872.
https://pubmed.ncbi.nlm.nih.gov/33371498/
https://doi.org/10.3390/cancers12123872
[35] Danics, L., Schvarcz, C.A., Viana, P., et al. (2020) Exhaustion of Protective Heat Shock Response Induces Significant Tumor Damage by Apoptosis after Modulated Electro-Hyperthermia Treatment of Triple Negative Breast Cancer Isografts in Mice. Cancers, 12, Article 2581.
https://pubmed.ncbi.nlm.nih.gov/32927720/
https://doi.org/10.3390/cancers12092581
[36] Kim, J.-K., Prasad, B. and Kim, S. (2017) Temperature Mapping and Thermal Dose Calculation in Combined Radiation Therapy and 13.56 MHz Radiofrequency Hyperthermia for Tumor Treatment. Proceedings of the SPIE, XXVI, Article 1004718.
http://spie.org/Publications/Proceedings/Paper/10.1117/12.2253163?origin_id=x4318
https://doi.org/10.1117/12.2253163
[37] Prasad, B., Kim, S., Cho, W., et al. (2018) Effect of Tumor Properties on Energy Absorption, Temperature Mapping, and Thermal Dose in 13.56-MHz Radiofrequency Hyperthermia. Journal of Thermal Biology, 74, 281-289.
https://doi.org/10.1016/j.jtherbio.2018.04.007
[38] Szasz, A. (2021) The Capacitive Coupling Modalities for Oncological Hyperthermia. Open Journal of Biophysics, 11, 252-313.
https://doi.org/10.4236/ojbiphy.2021.113010
[39] Szasz, O. and Szasz, A. (2021) Approaching Complexity: Hyperthermia Dose and Its Possible Measurement in Oncology. Open Journal of Biophysics, 11, 68-132.
https://doi.org/10.4236/ojbiphy.2021.111002
[40] Ferenczy, G.L. and Szasz, A. (2020) Chap. 3. Technical Challenges and Proposals in Oncological Hyperthermia. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 72-90.
[41] Szasz, A. (2019) Thermal and Nonthermal Effects of Radiofrequency on Living State and Applications as an Adjuvant with Radiation Therapy. Journal of Radiation and Cancer Research, 10, 1-17.
https://doi.org/10.4103/jrcr.jrcr_25_18
[42] Andocs, G., Renner, H., Balogh, L., Fonyad, L., Jakab, C. and Szasz, A. (2009) Strong Synergy of Heat and Modulated Electro-Magnetic Field in Tumor Cell Killing, Study of HT29 Xenograft Tumors in a Nude Mice Model. Strahlentherapie und Onkologie, 185, 120-126.
https://doi.org/10.1007/s00066-009-1903-1
[43] Szasz, A. (2022) Stimulation and Control of Homeostasis. Open Journal of Biophysics, 12, 89-131.
https://doi.org/10.4236/ojbiphy.2022.122004
[44] Szasz, A. and Szasz, O. (2020) Chap. 17. Time-Fractal Modulation of Modulated Electro-Hyperthermia (mEHT). In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 377-415.
[45] Szasz, A. (2022) Time-Fractal Modulation—Possible Modulation Effects in Human Therapy. Open Journal of Biophysics, 12, 38-87.
https://doi.org/10.4236/ojbiphy.2022.121003
[46] Warburg, O. (1956) On the Origin of Cancer Cells. Science, 123, 309-314.
https://doi.org/10.1126/science.123.3191.309
[47] Vaupel, P. and Piazena, H. (2022) Strong Correlation between Specific Heat Capacity and Water Content in Human Tissues Suggests Preferred Heat Deposition in Malignant Tumors upon Electromagnetic Irradiation. International Journal of Hyperthermia, 39, 987-997.
https://doi.org/10.1080/02656736.2022.2067596
[48] Sha, L., Ward, E.R. and Story, B. (2002) A Review of Dielectric Properties of Normal and Malignant Breast Tissue. Proceedings IEEE SoutheastCon 2002, Columbia, SC, 5-7 April 2002, 457-462.
[49] Scholz, B. and Anderson, R. (2000) On Electrical Impedance Scanning-Principles and Simulations. Electromedica, 68, 35-44.
[50] Haemmerich, D., Staelin, S.T., Tsai, J.Z., et al. (2003) In Vivo Electrical Conductivity of Hepatic Tumors. Physiological Measurement, 24, 251-260.
https://doi.org/10.1088/0967-3334/24/2/302
[51] Smith, S.R., Foster, K.R. and Wolf, G.L. (1986) Dielectric Properties of VX-2 Carcinoma versus Normal Liver Tissue. IEEE Transactions on Biomedical Engineering, BME-33, 522525.
https://doi.org/10.1109/TBME.1986.325740
[52] Cope, F.W. (1975) A Review of the Applications of Solid State Physics Concepts to Biological Systems. Journal of Biological Physics, 3, 1-41.
https://doi.org/10.1007/BF02308900
[53] Damadian, R. (1971) Tumor Detection by Nuclear Magnetic Resonance. Science, 171, 1151-1153.
https://doi.org/10.1126/science.171.3976.1151
[54] Hazlewood, C.F., Nichols, B.L. and Chamberlain, N.F. (1969) Evidence for the Existence of a Minimum of Two Phases of Ordered Water in Skeletal Muscle. Nature, 222, 747-750.
https://doi.org/10.1038/222747a0
[55] Durney, C.H., Massoudi, H. and Iskander, M.F. (1986) Radiofrequency Radiation Dosimetry Handbook. 4th Edition, Final Report.
https://www.osti.gov/biblio/6168126-radiofrequency-radiation-dosimetry-handbook-edition-final-report-july-december
[56] Szentgyorgyi, A. (1980) The Living State and Cancer. Physiological Chemistry and Physics, 12, 99-110.
[57] Szentgyorgyi, A. (1998) Electronic Biology and Cancer. Marcel Dekker, New York.
[58] Szentgyorgyi, A. (1968) Bioelectronics: A Study in Cellular Regulations, Defense, and Cancer. Academic Press, New York, London.
[59] Pething, R. (1979) Dielectric and Electronic Properties of Biological Materials. John Wiley and Sons, New York,.
[60] Volkov, V.V., Palmer, D.J. and Righini, R. (2007) Distinct Water Species Confined at the Interface of a Phospholipid Membrane. Physical Review Letters, 99, Article 078302.
https://doi.org/10.1103/PhysRevLett.99.078302
[61] Liu, L.M. and Cleary, S.F. (1995) Absorbed Energy Distribution from Radiofrequency Electromagnetic Radiation in a Mammalian Cell Model: Effect of Membrane-Bound Water. Bioelectromagnetics, 16, 160-171.
https://doi.org/10.1002/bem.2250160304
[62] Hendry, B. (1981) Membrane Physillogy and Membrane Excitation. Croom Helm, London.
https://doi.org/10.1007/978-1-4615-9766-7
[63] Ma, Y., Poole, K., Goyette, J., et al. (2017) Introducing Membrane Charge and Membrane Potential to T Cell Signaling. Frontiers in Immunology, 8, Article 1513.
https://doi.org/10.3389/fimmu.2017.01513
[64] Andocs, G., Rehman, M.U., Zhao, Q.-L., Tabuchi, Y., Kanamori, M. and Kondo, T. (2016) Comparison of Biological Effects of Modulated Electro-Hyperthermia and Conventional Heat Treatment in Human Lymphoma U937 Cell. Cell Death Discovery, 2, Article No. 16039.
http://www.nature.com/articles/cddiscovery201639
https://doi.org/10.1038/cddiscovery.2016.39
[65] Yang, K.-L., Huang, C.-C., Chi, M.-S., Chiang, H.-C., Wang, Y.-S., Andocs, G., et al. (2016) In Vitro Comparison of Conventional Hyperthermia and Modulated Electro-Hyperthermia. Oncotarget, 7, 84082-84092.
https://doi.org/10.18632/oncotarget.11444
[66] Tsang, Y.W., Huang, C.C., Yang, K.L., et al. (2015) Improving Immunological Tumor Microenvironment Using Electro-Hyperthermia Followed by Dendritic Cell Immunotherapy. BMC Cancer, 15, Article No. 708.
https://doi.org/10.1186/s12885-015-1690-2
[67] Zhu, M.X. (2011) TRP Channels, Methods in Signal Transduction. CRC Press, Boca Raton, FL.
[68] Chung, M.K., Guler, A.D. and Caterina, M.J. (2005) Biphasic Currents Evoked by Chemical or Thermal Activation of the Heat-Gated Ion Channel, TRPV3. Journal of Biological Chemistry, 280, 15928-15941.
https://doi.org/10.1074/jbc.M500596200
[69] Brauchi, S., Orio, P. and Latorre, R. (2004) Clues to Understanding Cold Sensation: Thermodynamics and Electrophysiological Analysis of the Cold Receptor TRPM8. Proceedings of the National Academy of Sciences of the United States of America, 101, 15494-15499.
https://doi.org/10.1073/pnas.0406773101
[70] Clapham, D.E. (2003) TRP Channels as Cellular Sensors. Nature, 426, 517-524.
https://doi.org/10.1038/nature02196
[71] Startek, J.B., Boonen, B., Talavera, K. and Meseguer, V. (2019) TRP Channels as Sensors of Chemically-Induced Changes in Cell Membrane Mchanical Properties. International Journal of Molecular Sciences, 20, Article 371.
https://doi.org/10.3390/ijms20020371
[72] Samie, M. and Xu, H. (2011) Chap. 19. Studying TRP Channels in Intracellular Membranes. In: Zhu, M.X., Ed., TRP Channels, CRC Press/Taylor & Francis, Boca Raton, FL.
[73] Dong, X.P., Wang, X. and Xu, H. (2010) TRP Calcium Channels in Intracellular Membranes. Journal of Neurochemistry, 113, 313-328.
https://doi.org/10.1111/j.1471-4159.2010.06626.x
[74] Sakaguchi, R. and Mori, Y. (2019) Transient Receptor Potential (TRP) Channels: Biosensors for Redox Environmental Stimuli and Cellular Status. Free Radical Biology and Medicine, 146, 36-44.
https://doi.org/10.1016/j.freeradbiomed.2019.10.415
[75] Nilius, B. and Owsianik, G. (2011) The Transient Receptor Potential Family of Ion Channels. Genome Biology, 12, Article No. 218.
http://genomebiology.com/2011/12/3/218
https://doi.org/10.1186/gb-2011-12-3-218
[76] Hellwig, N., Plant, T.D., Janson, W., et al. (2004) TRPV1 Acts as Proton Channel to Induce Acidification in Nociceptive Neurons. The Journal of Biological Chemistry, 279, 34553-34561.
https://doi.org/10.1074/jbc.M402966200
[77] Liu, X., Wu, G., You, Y., et al. (2019) Molecular Understanding of Calcium Permeation through the Open Orai Channel. PLOS Biology, 17, e3000096.
https://doi.org/10.1371/journal.pbio.3000096
[78] Saul, S., Stanisz, H., Backes, C.S., Schwarz, C., et al. (2013) How ORAI and TRP Channels Interfere with Each Other: Interaction Models and Examples from the Immune System and the Skin. European Journal of Pharmacology, 739, 49-59.
https://doi.org/10.1016/j.ejphar.2013.10.071
[79] Ito, E., Ikemoto, Y. and Yoshioka, T. (2015) Thermodynamic Implications of High Q10 of Thermo-TRPV Channels in Living Cells. Biophysics, 11, 33-38.
https://doi.org/10.2142/biophysics.11.33
[80] Lezama, G.K., Mota, R.D., Pereira, A.M.F., Martinez, B.J., et al. (2022) Transient Receptor Potential (TRP) and Thermoregulation in Animals: Structural Biology and Nurophysiological Aspects. Animals, 12, Article 106.
https://doi.org/10.3390/ani12010106
[81] Clapham, D.E. and Miller, C. (2011) A Thermodynamic Framework for Understanding Temperature Sensing by Transient Receptor Potential (TRP) Channels. Proceedings of the National Academy of Sciences of the United States of America, 108, 19492-19497.
https://doi.org/10.1073/pnas.1117485108
[82] Karnovsky, M.J., Kleinfeld, A.M., Hoover, R.L. and Klausner, R.D. (1982) The Concept of Lipid Domains in Membranes. Journal of Cell Biology, 94, 1-6.
https://doi.org/10.1083/jcb.94.1.1
[83] Rajendran, L. and Simons, K. (2005) Lipid Rafts and Membrane Dynamics. Journal of Cell Science, 118, 1099-1102.
https://doi.org/10.1242/jcs.01681
[84] Horváth, I., et al. (2008) Membrane-Associated Stress Proteins: More than Simply Chaperones. Biochimica et Biophysica Acta (BBA)—Biomembranes, 1778, 1653-1664.
https://doi.org/10.1016/j.bbamem.2008.02.012
[85] Hille, B. (1992) Ionic Channels of Excitable Membranes. Sinauer, Sunderland, Massachusetts, 57.
[86] Adair, R.K. (2003) Biophysical Limits on Athermal Effects of RF and Microwave Radiation. Bioelectromagnetics, 24, 39-48.
https://doi.org/10.1002/bem.10061
[87] Sanchez, M.A., Guevara, H.E., Contreras, C.R., et al. (2018) Irreversible Temperature Gating in trpv1 Sheds Light on Channel Activation. eLife, 7, e36372.
https://doi.org/10.7554/eLife.36372
[88] Yao, J., Liu, B. and Qin, F. (2010) Kinetic and Energetic Analysis of Thermally Activated TRPV1 Channels. Biophysical Journal, 99, 1743-1753.
https://doi.org/10.1016/j.bpj.2010.07.022
[89] Raddatz, N., Castillo, J.P., Gonzalez, C., Alvarez, O. and Latorre, R. (2014) Temperature and Voltage Coupling to Channel Opening in Transient Receptor Potential Melastatin 8 (TRPM8). Journal of Biological Chemistry, 289, 35438-35454.
https://doi.org/10.1074/jbc.M114.612713
[90] Hildebrandt, B., Wust, P., Ahlers, O., Dieing, A., et al. (2002) The Cellular and Molecular Basis of Hyperthermia. Critical Reviews in Oncology/Hematology, 43, 33-56.
https://doi.org/10.1016/S1040-8428(01)00179-2
[91] Papp, E., Vancsik, T., Kiss, E. and Szasz, O. (2017) Energy Absorption by the Membrane Rafts in the Modulated Electro-Hyperthermia (mEHT). Open Journal of Biophysics, 7, 216-229.
https://doi.org/10.4236/ojbiphy.2017.74016
[92] Staunton, J.R. and Wirtz, D. (2013) Physical Sciences-Oncology Centers Network. A Physical Sciences Network Charactrization of Non-Tumorigenic and Metastatic Cells. Scientific Reports, 3, Article No. 1449.
[93] Szasz, O., Szasz, A.M., Minnaar, C. and Szasz, A. (2017) Heating Preciosity—Trends in Modern Oncological Hyperthermia. Open Journal of Biophysics, 7, 116-144.
https://doi.org/10.4236/ojbiphy.2017.73010
[94] Andocs, G., Rehman, M.U., Zhao, Q.L., Papp, E., Kondo, T. and Szasz, A. (2015) Nanoheating without Artificial Nanoparticles Part II. Experimental Support of the Nanoheating Concept of the Modulated Electro-Hyperthermia Method, Using U937 Cell Suspension Model. Biology and Medicine, 7, 1-9.
[95] Szasz, A. (2022) Heterogeneous Heat Absorption Is Complementary to Radiotherapy. Cancers, 14, Article 901.
https://doi.org/10.3390/cancers14040901
[96] Wust, P., Ghadjar, P., Nadobny, J., Beck, M., Kaul, D., Winter, L. and Zschaeck, S. (2019) Physical Analysis of Temperature-Dependent Effects of Amplitude-Modulated Electromagnetic Hyperthermia. International Journal of Hyperthermia, 36, 1246-1254.
https://doi.org/10.1080/02656736.2019.1692376
[97] Dutz, S. and Hergt, R. (2013) Magnetic Nanoparticle Heating and Heat Transfer on a Microscale: Basic Principles, Realities and Physical Limitations of Hyperthermia for Tumour Therapy. International Journal of Hyperthermia, 29, 790-800.
https://doi.org/10.3109/02656736.2013.822993
[98] Lund, M., Jönsson, B. and Woodward, C.E. (2007) Implications of a High Dielectric Constant in Proteins. The Journal of Chemical Physics, 126, Article 225103.
https://doi.org/10.1063/1.2741543
[99] Voets, T., Droogmans, G., Wissenbach, U., Janssens, A., Flockerzi, V. and Nilius, B. (2004) The Principle of Temperature-Dependent Gating in Cold- and Heat-Sensitive TRP Channels. Nature, 430, 748-754.
https://doi.org/10.1038/nature02732
[100] Goswami, C. and Hucho, T. (2008) Sub-Membranous Microtubule Cytoskeleton: Biochemical and Functional Interplay of TRP Channels with the Cytoskeleton. FEBS Journal, 275, 4684-4699.
https://doi.org/10.1111/j.1742-4658.2008.06617.x
[101] Vincze, G., Szigeti, G., Andocs, G. and Szasz, A. (2015) Nanoheating without Artificial Nanoparticles. Biology and Medicine, 7, Article 1000249.
[102] Wong, E.T., Lok, E. and Swanson, K.D. (2015) An Evidence-Based Review of Alternating Electric Fields Therapy for Malignant Gliomas. Current Treatment Options in Oncology, 16, Article No. 40.
https://doi.org/10.1007/s11864-015-0353-5
[103] Wenger, C., MirandA, P.C., Salvador, R., et al. (2018) A Review on Tumor-Treating Fields (TTFields): Clinical Implications Inferred from Computational Modeling. IEEE Reviews in Biomedical Engineering, 11, 195-207.
https://doi.org/10.1109/RBME.2017.2765282
[104] Kirson, E.D., Dbaly, V., Tovarys, F., et al. (2007) Alternating Electric Fields Arrest Cell Proliferation in Animal Tumor Models and Human Brain Tumors. Proceedings of the National Academy of Sciences of the United States of America, 104, 10152-10157.
https://doi.org/10.1073/pnas.0702916104
[105] Ritossa, F. (1962) A New Puffing Pattern Induced by Temperature Shock and DNP in Drosophila. Experientia, 18, 571-573.
https://doi.org/10.1007/BF02172188
[106] Hu, C., Yang, J., Qi, Z., Wu, H., et al. (2020) Heat Shock Proteins: Biological Functions, Pathological Roles, and Therapeutic Opportunities. MedComm, 3, e161.
https://doi.org/10.1002/mco2.161
[107] Tang, D., Kang, R., Berghe, T.V., et al. (2019) The Molecular Machinery of Regulated Cell Death. Cell Research, 29, 347-364.
https://doi.org/10.1038/s41422-019-0164-5
[108] Giri, B., Sethi, V., Modi, S., et al. (2017) Heat Shock Protein 70 in Pancreatic Diseases: Friend or Foe. Journal of Surgical Oncology, 116, 114-122.
https://doi.org/10.1002/jso.24653
[109] Pockley, A.G. and Multhoff, G. (2008) Cell Stress Proteins in Extracellular Fluids: Friend or Foe? Novartis Foundation Symposium, 291, 86-95, Discussion 96-100, 137-140.
https://doi.org/10.1002/9780470754030.ch7
[110] Wu, T. and Tanguay, R.M. (2006) Antibodies against Heat Shock Proteins in Environmental Stresses and Diseases: Friend or Foe? Cell Stress and Chaperones, 11, 1-12.
[111] Hance, M.W., Nolan, K.D. and Isaacs, J.S. (2014) The Double-Edged Sword: Conserved Functions of Extracellular Hsp90 in Wound Healing and Cancer. Cancers, 6, 1065-1097.
https://doi.org/10.3390/cancers6021065
[112] Tittelmeier, J., Nachman, E. and Nussbaum, K.C. (2020) Molecular Chaperones: A Double-Sword in Neurodegenerative Diseases. Frontiers in Aging Neuroscience, 12, Article 581374.
https://doi.org/10.3389/fnagi.2020.581374
[113] Shevtsov, M., Balogi, Z., Khachatryan, W., et al. (2020) Membrane-Associated Heat Shock Proteins in Oncology: From Basic Research to New Theranostic Targets. Cells, 9, Article 1263.
https://doi.org/10.3390/cells9051263
[114] Calderwood, S.K., Mambula, S.S., Gay Jr., P.J. and Theriault J.R. (2007) Extracellular Heat Shock Proteins in Cell Signaling. FEBS Letters, 581, 3689-3694.
https://doi.org/10.1016/j.febslet.2007.04.044
[115] Minnaar, C.A. and Szasz, A. (2022) Forcing the Antitumor Effects of HSPs Using a Modulated Electric Field. Cells, 11, Article 1838.
https://doi.org/10.3390/cells11111838
[116] Escriba, P.V., Gonzalez, R.J.M., Goni, F.M., Kinnunen, P.K., Vigh, L., Sanchez, M.L., Fernandez, A.M., Busquets, X., Horvath, I. and Barcelo, C.G. (2008) Membranes: A Meeting Point for Lipids, Proteins and Therapies. Journal of Cellular and Molecular Medicine, 12, 829-875.
https://doi.org/10.1111/j.1582-4934.2008.00281.x
[117] Vigh, L., Escribá, P.V., Sonnleitner, A., Sonnleitner, M., Piotto, S., Maresca, B., Horvath, I. and Harwood, J.L. (2005) The Significance of Lipid Composition for Membrane Activity: New Concepts and Ways of Assessing Function. Progress in Lipid Research, 44, 303-344.
https://doi.org/10.1016/j.plipres.2005.08.001
[118] Latchman, D.S. (1998) Stress Proteins. Springer Verlag, Berlin.
https://doi.org/10.1007/978-3-642-58259-2
[119] Challis, L.J. (2005) Mechanisms for Interaction between RF Fields and Biological Tissue. Bioelectromagnetics, 26, S98-S106.
https://doi.org/10.1002/bem.20119
[120] Beyer, C. (1980) Investigation of Potential Interactions of Electromagnetic Fields with Proteins and Cells. PhD Thesis, Technical University of Dresden, Dresden.
https://www.google.com/url?sa=t&rct=j&q=&esrc=s&source=web&cd=&ved=2ahUKEwjS-dDsp_n7AhV3hP0HHXfEAcEQFnoECAkQAQ&url=https%3A%2F%2Fwww.research-collection.ethz.ch%2Fbitstream%2Fhandle%2F20.500.11850%2F88374%2F1%2Feth-14448-01.pdf&usg=AOvVaw2MbmdDCEvX22QwyBoCIkn6
[121] Storm, F.K., Morton, D.L., Kaiser, L.R., Harrison, W.H., Elliott, R.S., Weisenburger, T.H., Parker, R.G. and Haskell, C.M. (1982) Clinical Radiofrequency Hyperthermia: A Review. National Cancer Institute Monograph, 61, 343-350.
[122] Bezanilla, F. (2005) Voltage-Gated Ion Channels. IEEE Transactions on NanoBioscience, 4, 34-48.
https://doi.org/10.1109/TNB.2004.842463
[123] Loussouarn, G. and Tarek, M. (2021) Editorial: Molecular Mechanisms of voltage-Gating in Ion Channels. Frontiers in Pharmacology, 12, Article 768153.
https://doi.org/10.3389/fphar.2021.768153
[124] Catterall, W.A. (2014) Voltage-Gated Calcium Channels. Cold Spring Harbor Perspectives in Biology, 3, a003947.
https://doi.org/10.1101/cshperspect.a003947
[125] Voets, T., Taavera, K., Owsianik, G. and Nilius, B. (2005) Sensing with TRP Channels. Nature Chemical Biology, 1, 85-92.
https://doi.org/10.1038/nchembio0705-85
[126] Latorre, R., Vargas, G., Orta, G., et al. (2007) Chap. 21. Voltage and Temperature Gating of ThermoTRP Channels. In: Liedtke, W.B. and Heller, S., Eds., TRP Ion Channel Function in Sensory Transduction and Cellular Signaling Cascades, CRC Press, Boca Raton, FL.
[127] Eyring, H. (1935) The Activated Complex in Chemical Reactions. The Journal of Chemical Physics, 3, 107-115.
https://doi.org/10.1063/1.1749604
[128] Miller, W.H. (1993) Beyond Transition-State Theory: A Rigorous Quantum Theory of Chemical Reaction Rates. Accounts of Chemical Research, 26, 174-181.
https://doi.org/10.1021/ar00028a007
[129] Pollak, E. and Talkner, P. (2005) Reactio Rate Theory: What It Was, Where Is It Today, and Where Is It Going? Chaos, 15, 26116-26117.
https://doi.org/10.1063/1.1858782
[130] Pelzer, H. and Wigner, E. (1932) über die Geschwindigkeitskonstante von Austauschreaktionen. Zeitschrift für Physikalische Chemie, 15B, 445-471.
https://doi.org/10.1515/zpch-1932-1539
[131] Laidler, K.J. and King, M.C. (1983) The Development of Transition-State Theory. The Journal of Physical Chemistry, 87, 2657-2664.
https://doi.org/10.1021/j100238a002
[132] Ryu, S., Liu, B., Yao, J., et al. (2007) Uncoupling Proton Activation of Vanilloid Receptor TRPV1. The Journal of Neuroscience, 27, 12797-12807.
https://doi.org/10.1523/JNEUROSCI.2324-07.2007
[133] Vincze, G. and Szasz, A. (2018) Similarities of Modulation by Temperature and by Electric Field. Open Journal of Biophysics, 8, 95-103.
https://doi.org/10.4236/ojbiphy.2018.83008
[134] Ronchi, R., Marano, L., Braidotti, P., et al. (2004) Effects of Broad Band Electromagnetic Fields on HSP70 Expression and Ischemia-Reperfusion in Rat Hearts. Life Sciences, 75, 1925-1936.
https://doi.org/10.1016/j.lfs.2003.12.033
[135] Han, L., Lin, H., Head, M., et al. (1998) Application of Magnetic Field-Induced Heat Shock Protein 70 for Presurgical Cytoprotection. Journal of Cellular Biochemistry, 71, 577-583.
https://doi.org/10.1002/(SICI)1097-4644(19981215)71:4<577::AID-JCB12>3.0.CO;2-V
[136] Taghian, T., Narmoneva, D.A. and Kogan, A.B. (2015) Modulation of Cell Function by Electric Field: A High-Resolution Analysis. Journal of the Royal Society Interface, 12, Article 20150153.
https://doi.org/10.1098/rsif.2015.0153
[137] Sheikh, A.Q., Taghian, T., Hemingway, B., Cho, H., Kogan, A.B. and Narmoneva, D.A. (2013) Regulation of Endothelial MAPK/ERK Signalling and Capillary Morphogenesis by Low-Amplitude Electric Field. Journal of the Royal Society Interface, 10, Article 20120548.
https://doi.org/10.1098/rsif.2012.0548
[138] Andreev, V.P. (2013) Cytoplasmic Electric Fields and Electroosmosis: Possible Solution for the Paradoxes of the Intracellular Transport of Biomolecules. PLOS ONE, 8, e61884.
https://doi.org/10.1371/journal.pone.0061884
[139] Monteith, G.R., McAndrew, D., Faddy, H.M. and Roberts, T.S.J. (2007) Calcium and Cancer: Targeting Ca2+ Transport. Nature Reviews Cancer, 7, 519-530.
https://doi.org/10.1038/nrc2171
[140] Carafoli, E. (2002) Calcium Signaling: A Tale for All Seasons. Proceedings of the National Academy of Sciences of the United States of America, 99, 1115-1122.
https://doi.org/10.1073/pnas.032427999
[141] Jacques, F.B., Seow, Y., Gottlieb, P.A., Sachs, F. and Gomez, T.M. (2006) Ca2+ Influx through Mechanosensitive Channels Inhibits Neurite Outgrowth in Opposition to Other Influx Pathways and Release from Intracellular Stores. Journal of Neuroscience, 26, 5656-5664.
https://doi.org/10.1523/JNEUROSCI.0675-06.2006
[142] Amuthan, G., et al. (2002) Mitochondrial Stress-Induced Calcium Signaling, Phenotypic Changes and Invasive Behavior in Human Lung Carcinoma A549 Cells. Oncogene, 21, 7839-7849.
https://doi.org/10.1038/sj.onc.1205983
[143] Cook, S.J. and Lockyer, P.J. (2006) Recent Advances in Ca2+-Dependent Ras Regulation and Cell Proliferation. Cell Calcium, 39, 101-112.
https://doi.org/10.1016/j.ceca.2005.10.014
[144] Patton, A.M., Kassis, J., Doong, H. and Kohn, E.C. (2003) Calcium as a Molecular Target in Angiogenesis. Current Pharmaceutical Design, 9, 543-551.
https://doi.org/10.2174/1381612033391559
[145] Pinton, P., Giorgi, C., Siviero, R., Zecchini, E. and Rizzuto, R. (2008) Calcium and Apoptosis: ER-Mitochondria Ca2+ Transfer in the Control of Apoptosis. Oncogene, 27, 6407-6418.
https://doi.org/10.1038/onc.2008.308
[146] Schwartz, E.C., Qu, B. and Hoth, M. (2013) Calcium, Cancer and Killing: The Role of Calcium in Killing Cancer Cells by Cytotoxic T Lymphocytes and Natural Killer Cells. Biochimica et Biophysica Acta (BBA)—Molecular Cell Research, 1833, 1603-1611.
https://doi.org/10.1016/j.bbamcr.2012.11.016
[147] Kondrantsky, A., Kondratska, K., Skryma, R. and Prevarskaya, N. (2015) Ion Channels in the Regulation of Apoptosis. Biochimica et Biophysica Acta, 1848, 2532-2546.
https://doi.org/10.1016/j.bbamem.2014.10.030
[148] Orrenius, S., Zhivotovsky, B. and Nicotera, P. (2003) Regulation of Cell Death: The Calcium-Apoptosis Link. Nature Reviews Molecular Cell Biology, 4, 552-565.
https://doi.org/10.1038/nrm1150
[149] Mimura, T. and Tazawa, T. (1983) Effect of Intracellular Ca2+ on Membrane Potential and Membrane Resistance in Tonoplast-Free Cells of Nitellopsis obtusa. Protoplasma, 118, 49-55.
https://doi.org/10.1007/BF01284746
[150] Mason, M.J., Hussain, J.F. and Mahaut, S.M.P. (2000) A Novel Role for Membrane Potential in the Modulation of Intracellular Ca2+ Oscillations in Rat Megakaryocytes. Journal of Physiology, 524, 437-446.
https://doi.org/10.1111/j.1469-7793.2000.00437.x
[151] Shapovalov, G., Lehen, K.V., Skryma, R. and Prevarskaya, N. (2011) TRP Channels in Cell Survival and Cell Death in Normal and Transformed Cells. Cell Calcium, 50, 295-302.
https://doi.org/10.1016/j.ceca.2011.05.006
[152] Qu, B., Al-Ansary, D., Kummerow, C., Hoth, M. and Schwarz, E.C. (2011) ORAI-Mediated Calcium Influx in T Cell Proliferation, Apoptosis and Tolerance. Cell Calcium, 50, 261-269.
https://doi.org/10.1016/j.ceca.2011.05.015
[153] Rizzuto, R. and Pozzan, T. (2006) Microdomains of Intracellular Ca2+: Molecular Determinants and Functional Consequences. Physiological Reviews, 86, 369-408.
https://doi.org/10.1152/physrev.00004.2005
[154] Dometsch, R., Lewis, R.S., Goodnow, C.C., et al. (1997) Differential Activation of Transcription Factors Induced by Ca2+ Response Amplitude and Duration. Nature, 386, 855-858.
https://doi.org/10.1038/386855a0
[155] Berridge, M.J. (1997) The AM and FM of Calcium Signalling. Nature, 386, 759-760.
https://doi.org/10.1038/386759a0
[156] Healy, J.I., Dolmetsch, R.E., Timmerman, L.A., et al. (1997) Different Nuclear Signals Are Activated by the B Cell Receptor during Positive versus Negative Signaling. Immunity, 6, 419-428.
https://doi.org/10.1016/S1074-7613(00)80285-X
[157] Rapp, P.E. (1981) The Control of Transepithelial Potential Oscillations in the Salivary Gland of Calliphora Erythrocephala. Journal of Experimental Biology, 93, 119-132.
https://doi.org/10.1242/jeb.93.1.119
[158] Nilius, B., Talavera, K., Owsianik, G., et al. (2005) Gating of TRP Channels: A Voltage Connection? The Journal of Physiology, 567, 35-44.
https://doi.org/10.1113/jphysiol.2005.088377
[159] Yanagida, Y., Mizuno, A., Motegi, T., et al. (2000) Electrically Stimulated Induction of hsp70 Gene Expression in Mouse Astroglia and Fibroblast Cells. Journal of Biotechnology, 79, 53-61.
https://doi.org/10.1016/S0168-1656(00)00214-5
[160] Blank, M. and Goodman, R. (2009) Electromagnetic Fields Stress Living Cells. Pathophysiology, 16, 71-78.
https://doi.org/10.1016/j.pathophys.2009.01.006
[161] Kotnik, T. and Miklavcic, D. (2000) Theoretical Evaluation of the Distributed Power Dissipation in Biological Cells Exposed to Electric Fields. Bioelectromagnetics, 21, 385-394.
https://doi.org/10.1002/1521-186X(200007)21:5<385::AID-BEM7>3.0.CO;2-F
[162] Gowrishankar, T.R. and Weaver, J.C. (2003) An Approach to Electrical Modeling of Single and Multiple Cells. Proceedings of the National Academy of Sciences of the United States of America, 100, 3203-3208.
https://doi.org/10.1073/pnas.0636434100
[163] Ahern, G.P. and Premkumar, L.S. (2002) Voltage-Dependent Priming of Rat Vanilloid Receptor: Effects of Agonist and Protein Kinase C Activation. The Journal of Physiology, 545, 441-451.
https://doi.org/10.1113/jphysiol.2002.029561
[164] Vlachova, V., Teisinger, J., Suankova, K., Lyfenko, A., Ettrich, R. and Vyklicky, L. (2003) Functional Role of C-Terminal Cytoplasmic Tail of Rat Vanilloid Receptor 1. Journal of Neuroscience, 23, 1340-1350.
https://doi.org/10.1523/JNEUROSCI.23-04-01340.2003
[165] Fernandes, E.S., Liang, L., Smillie, S.J., Kaiser, F., Purcell, R., Rivett, D.W., et al. (2021) TRPV1 Deletion Enhances Local Inflammation and Accelerates the Onset of Systemic Inflammatory Response Syndrome. The Journal of Immunology, 188, 5741-5751.
https://doi.org/10.4049/jimmunol.1102147
[166] Bujak, J.K., Kosmala, D., Szopa, I.M., et al. (2019) Inflammation, Cancer and Immunity—Implication of TRPV1 Channel. Frontiers in Oncology, 9, Article 1087.
https://doi.org/10.3389/fonc.2019.01087
[167] Voets, T., Owsianik, G., Janssens, A., Talavera, K. and Nilius, B. (2007) TRPM8 Voltage Sensor Mutants Reveal a Mechanism for Integrating Thermal and Chemical Stimuli. Nature Chemical Biology, 3, 174-182.
https://doi.org/10.1038/nchembio862
[168] Matta, J.A. and Ahern, G.P. (2007) Voltage Is a Partial Activator of Rat Thermosensitive TRP Channels. The Journal of Physiology, 585, 469-482.
https://doi.org/10.1113/jphysiol.2007.144287
[169] Brauchi, S. and Orio, P. (2017) Chap. 28. Voltage Sensing in Thermo-TRP Channels. In: Islam, M.S., Ed., Transient Receptor Potential Channels, Advances in Experimental Medicine and Biology, Vol. 704, Springer, Dordrecht, 517-530.
https://doi.org/10.1007/978-94-007-0265-3_28
[170] Brown, J.A.M. (1999) A Study of the Interactions between Electromagnetic Fields and Microtubules: Ferroelectric Effects, Signal Transduction and Electronic Conduction. PhD Thesis, University of Alberta, Edmonton.
https://ui.adsabs.harvard.edu/abs/1999PhDT........63B/abstract
[171] Goswami, C. (2012) Chap. 2. TRP-Mediated Cytoskeletal Reorganization: Implications for Disease and Drug Development. In: Szallasi, A. and Biro, T., Eds., TRP Channels in Drug Discovery. Methods in Pharmacology and Toxicology, Humana Press, Totowa, NJ, 13-39.
https://doi.org/10.1007/978-1-62703-077-9_2
[172] Friedman, E., Verderame, M., Winawer, S. and Pollack, R. (1984) Actin Cytoskeletal Organization Loss in the Benign-to-Malignant Tumor Transition in Cultured Human Colonic Epithelial Cells. Cancer Research, 44, 3040-3050.
[173] Plodinec, M., Loparic, M., Monnier, C.A., et al. (2012) The Nanomechanical Signature of Breast Cancer. Nature Nanotechnology, 7, 757-765.
http://www.nature.com/nnano/journal/v7/n11/full/nnano.2012.167.html?WT.ec_id=NNANO-201211
https://doi.org/10.1038/nnano.2012.167
[174] Wirts, D., Konstantopoulos, K. and Searson, P.C. (2011) The Physics of Cancer: The Role of Physical Interactions and Mechanical Forces in Metastasis. Nature Reviews Cancer, 11, 512-518.
https://doi.org/10.1038/nrc3080
[175] Uklrich, T.A., de Juan Pardo, E.M. and Kumar, S. (2009) The Mechanical Rigidity of the Extracellular Matrix Regulates the Structure, Motility, and Proliferation of Glioma Cells. Cancer Research, 69, 4167-4175.
https://doi.org/10.1158/0008-5472.CAN-08-4859
[176] Einstein, A. (1906) On the Theory of Brownian Movement. Annals of Physics, 19, 371-381.
https://doi.org/10.1002/andp.19063240208
[177] Condeelis, J., Hall, A., Bresnick, A., Warren, V., Hock, R., Bennett, H. and Ogihara, S. (1988) Actin Polymerization and Pseudopod Extension during Amoeboid Chemotaxis. Cell Motility and the Cytoskeleton, 10, 77-90.
https://doi.org/10.1002/cm.970100113
[178] Vincze, G. and Szasz, A. (2015) Reorganization of Actin Filaments and Microtubules by Outside Electric Field. Journal of Advances in Biology, 8, 1514-1518.
[179] Springer, M. and Paulsson, J. (2006) Harmonies from Noise. Nature, 439, 27-28.
https://doi.org/10.1038/439027a
[180] West, J.B. (2013) Fractal Physiology and Chaos in Medicine. World Scientific, Singapore.
https://doi.org/10.1142/8577
[181] Okamura, Y., Murata, Y. and Iwasaki, H. (2009) Voltage-Sensing Phosphatase: Actions and Potentials. The Journal of Physiology, 587, 513-520.
https://doi.org/10.1113/jphysiol.2008.163097
[182] Okamura, Y. and Dixon, J.E. (2011) Voltage-Sensing Phosphatase: Its Molecular Relationship with PTEN. Physiology, 26, 6-13.
https://doi.org/10.1152/physiol.00035.2010
[183] van Petegem, F. and Minor, D.L. (2006) The Structural Biology of Voltage-Gated Calcium Channel Function and Regulation. Biochemical Society Transactionss, 34, 887-893.
https://doi.org/10.1042/BST0340887
[184] Yasushi, O. (2009) Voltage-Sensing Phosphatase: Actions and Potentials. The Journal of Physiology, 587, 513-520.
https://doi.org/10.1113/jphysiol.2008.163097
[185] Drewes, G., et al. (1998) MAPs, MARKs and Microtubule Dynamics. Trends in Biochemical Sciences, 23, 307-311.
https://doi.org/10.1016/S0968-0004(98)01245-6
[186] Levin, M. (2007) Large-Scale Biophysics: Ion Flows and Regeneration. Trends in Cell Biology, 17, 261-270.
https://doi.org/10.1016/j.tcb.2007.04.007
[187] Yang, M. and Brackenbury, W.J. (2013) Membrane Potential and Cancer Progression. Frontiers in Physiology, 4, Article 185.
https://doi.org/10.3389/fphys.2013.00185
[188] Titushkin, I. and Cho, M. (2009) Regulation of Cell Cytoskeleton and Membrane Mechanics by Electric Field: Role of Linker Proteins. Biophysical Journal, 96, 717-728.
https://doi.org/10.1016/j.bpj.2008.09.035
[189] Mycielska, M.E. and Djamgoz, M.B.A. (2004) Cellular Mechanisms of Direct-Current Electric Field Effects: Galvanotaxis and Metastatic Disease. Journal of Cell Science, 117, 1631-1639.
https://doi.org/10.1242/jcs.01125
[190] Cho, H., Tharfe, H.S., Lee, R.C. and Goland, D.E. (1996) Reorganization of Microfilament Structure Induced by ac Electric Fields. The FASEB Journal, 10, 1552-1558.
https://doi.org/10.1096/fasebj.10.13.8940302
[191] Adey, W.R. (1984) Nonlinear, Nonequilibrium Aspects of Electromagnetic Field Interactions at Cell Membranes. In: Adey, W.R. and Lawrence, A.F., Eds., Nonlinear Electrodynamics in Biological Systems, Plenum Press, New York, London, 3-22.
https://doi.org/10.1007/978-1-4613-2789-9_1
[192] Anishchenko, V.S., Neiman, A.B., Moss, F. and Schimansky, G.L. (1999) Stochastic Resonance: Noise-Enhanced Order. Physics-Uspekhi, 42, 7-36.
https://doi.org/10.1070/PU1999v042n01ABEH000444
[193] Vincze, G., Sziget, G.P. and Szasz, A. (2016) Reorganization of the Cytoskeleton. Journal of Advances in Biology, 9, 1872-1882.
[194] Martinsen, O.G., Grimnes, S. and Schwan, H.P. (2002) Interface Phenomena and Dielectric Properties of Biological Tissue. In: Encyclopedia of Surface and Colloid Science, Marcel Dekker, Inc., New York, 2643-2652.
[195] Kirson, E.D., Gurvich, Z., Schneiderman, R., Dekel, E., et al. (2004) Disruption of Cancer Cell Replication by Alternating Electric Fields. Cancer Research, 64, 3288-3295.
https://doi.org/10.1158/0008-5472.CAN-04-0083
[196] Szasz, A. (2020) Towards the Immunogenic Hyperthermic Action: Modulated Electro-Hyperthermia. Clinical Oncology and Research, Science Repository, 3, 1-6.
https://www.sciencerepository.org/towards-the-immunogenic-hyperthermic-action-modulated-electro-hyperthermia_COR-2020-9-107
https://doi.org/10.31487/j.COR.2020.09.07
[197] Andocs, G., Meggyeshazi, N., Balogh, L., et al. (2014) Upregulation of Heat Shock Proteins and the Promotion of Damage-Associated Molecular Pattern Signals in a Colorectal Cancer Model by Modulated Electrohyperthermia. Cell Stress and Chaperones, 20, 37-46.
https://doi.org/10.1007/s12192-014-0523-6
[198] Meggyeshazi, N., Andocs, G., et al. (2013) Early Changes in mRNA and Protein Expression Related to Cancer Treatment by Modulated Electro-Hyperthermia. Conference of the International Clinical Hyperthermia Society 2012, 2013, Article ID: 249563.
http://www.hindawi.com/archive/2013/249563/
https://doi.org/10.1155/2013/249563
[199] Szasz, A. (2013) Challenges and Solutions in Oncological Hyperthermia. Thermal Medicine, 29, 1-23.
https://doi.org/10.3191/thermalmed.29.1
[200] Vancsik, T., Kovago, C., Kiss, E., et al. (2018) Modulated Electro-Hyperthermia Induced Loco-Regional and Systemic Tumor Destruction in Colorectal Cancer Allografts. Journal of Cancer, 9, 41-53.
https://doi.org/10.7150/jca.21520
[201] Szasz, A. (2015) Bioelectromagnetic Paradigm of Cancer Treatment Oncothermia. In: Rosch, P.J., Ed., Bioelectromagnetic and Subtle Energy Medicine, CRC Press, Boca Raton, 323-336.
[202] Meggyeshazi, N., Andocs, G. and Krenacs, T. (2013) Programmed Cell Death Induced by Modulated Electro-Hyperthermia. Conference of the International Clinical Hyperthermia Society 2012, 2013, Article ID: 187835.
http://www.hindawi.com/archive/2013/187835/
[203] Forika, G., Balogh, A., Vancsik, T., Zalatnai, A., et al. (2020) Modulated Electro-Hyperthermia Resolves Radioresistance of Panc1 Pancreas Adenocarcinoma and Promotes DNA Damage and Apoptosis in Vitro. International Journal of Molecular Sciences, 21, Article 5100.
https://doi.org/10.3390/ijms21145100
[204] Vancsik, T., Mathe, D., Horvath, I., et al. (2021) Modulated Electro-Hyperthermia Facilitates NK-Cell Infiltration and Growth Arrest of Human A2058 Melanoma in a Xenograft Model. Frontiers in Oncology, 11, Article 590764.
https://www.frontiersin.org/articles/10.3389/fonc.2021.590764/full
https://doi.org/10.3389/fonc.2021.590764
[205] Meggyeshazi, N., Andocs, G., Balogh, L., Balla, P., Kiszner, G., Teleki, I., Jeney, A. and Krenacs, T. (2014) DNA Fragmentation and Caspase-Independent Programmed Cell Death by Modulated Electrohyperthermia. Strahlentherapie und Onkologie, 190, 815-822.
https://doi.org/10.1007/s00066-014-0617-1
[206] Meggyeshazi, N. (2015) Studies on Modulated Electrohyperthermia Induced Tumor Cell Death in a Colorectal Carcinoma Model. Doctoral Thesis, Semmelweis University, Budapest.
[207] Danics, L., Schvarcz, C., Viana, P., et al. (2020) Exhaustion of Protective Heat Shock Response Induces Significant Tumor Damage by Apoptosis after Modulated Electro-Hyperthermia Treatment of Triple Negative Breast Cancer Isografts in Mice. Cancers, 12, Article 2581.
https://doi.org/10.3390/cancers12092581
[208] Gross, C.C. (2004) Membrane-Bound Hsp70 an Activating Ligand for NK Cells. PhD Thesis, Ludwig-Maximilians-University, Munich.
[209] Sapozhnikov, A.M., Ponomarev, E.D., Tarasenko, T.N., et al. (1999) Spontaneous Apoptosis and Expression of Cell-Surface Het-Shock Proteins in Cultured EL-4 Lymphoma Cells. Cell Proliferation, 32, 363-378.
https://doi.org/10.1111/j.1365-2184.1999.tb01354.x
[210] Horvath, I., Multhoff, G., Sonnleitner, A., et al. (2008) Membrane-Associated Stress Proteins: More than Simply Chaperones. Biochimica et Biophysica Acta, 1778, 1653-1664.
https://doi.org/10.1016/j.bbamem.2008.02.012
[211] Kim, J.-K., Prasad, B. and Kim, S. (2017) Temperature Mapping and Thermal Dose Calculation in Combined Radiation Therapy and 13.56 MHz Radiofrequency Hyperthermia for Tumor Treatment. Proceedings of the SPIE, XXVI, Article 1004718.
http://spie.org/Publications/Proceedings/Paper/10.1117/12.2253163?origin_id=x4318
https://doi.org/10.1117/12.2253163
[212] Hossain, M.T., Prasad, B., Park, K.S., et al. (2016) Simulation and Experimental Evaluation of Selective Heating Characteristics of 13,56 MHz Radiofrequency Hyperthermia in Phantom Models. International Journal of Precision Engineering and Manufacturing, 17, 253-256.
https://doi.org/10.1007/s12541-016-0033-9
[213] Andocs, G., Szasz, O. and Szasz, A. (2009) Oncothermia Treatment of Cancer: From the Laboratory to Clinic. Electromagnetic Biology and Medicine, 28, 148-165.
https://doi.org/10.1080/15368370902724633
[214] Sterner, R.C. and Sterner, R.M. (2021) CAR-T Cell Therapy—Current Limitations and Potential Strategies. Blood Cancer Journal, 11, Article No. 69.
https://doi.org/10.1038/s41408-021-00459-7
[215] Khalil, M., Alliger, K., Weidinger, C., et al. (2017) Functional Role of Transient Receptor Potential Channels in Immune Cells and Epithelia. Frontiers in Immunology, 9, Article 174.
https://doi.org/10.3389/fimmu.2018.00174
[216] Kim, H.S., Kwon, H.J., Kim, G.E., Cho, M.H., Yoon, S.Y., Davies, A.J., et al. (2014) Attenuation of Natural Killer Cell Functions by Capsaicin through a Direct and TRPV1-Independent Mechanism. Carcinogenesis, 35, 1652-1660.
https://doi.org/10.1093/carcin/bgu091
[217] Majhi, R.K., Sahoo, S.S., Yadav, M., Pratheek, B.M., Chattopadhyay, S. and Goswami, C. (2015) Functional Expression of TRPV Channels in T Cells and Their Implications in Immune Regulation. The FEBS Journal, 282, 2661-2681.
https://doi.org/10.1111/febs.13306
[218] Assas, B.M., Wakid, M.H., Zakai, H.A., Miyan, J.A. and Pennock, J.L. (2016) Transient Receptor Potential Vanilloid 1 Expression and Function in Splenic Dendritic Cells: A Potential Role in Immune Homeostasis. Immunology, 147, 292-304.
https://doi.org/10.1111/imm.12562
[219] Santoni, G., Farfariello, V., Liberati, S., et al. (2013) The Role of Transient Receptor Potential Vanilloid Type-2 Ion Channels in Innate and Adaptive Immune Responses. Frontiers in Immunology, 4, Article 31.
https://doi.org/10.3389/fimmu.2013.00034
[220] Fernandes, E.S., Brito, C.X.L., Teixeira, S.A., et al. (2014) TRPV1 Antagonism by Capsazepine Modulates Innate Immune Response in Mice Infected with Plasmodium berghei ANKA. Immunology and Infection by Protozoan Parasites, 2014, Article ID: 506450.
https://doi.org/10.1155/2014/506450
[221] Santoni, G., Morelli, M.B., Amantini, C., et al. (2018) “Immuno-Transient Receptor Potential Ion Channels”: The Role in Monocyte- and Macrophage-Mediated Inflammatory Responses. Frontiers in Immunology, 9, Article 1273.
https://doi.org/10.3389/fimmu.2018.01273
[222] Multhoff, G. and Hightower, L.E. (2011) Distinguishing Integral and Receptor-Bound Heat Shock Protein 70 (Hsp70) on the Cell Surface by Hsp70-Specific Antibodies. Cell Stress and Chaperones, 16, 251-255.
https://doi.org/10.1007/s12192-010-0247-1
[223] Multhoff, G., Botzler, C., Wiesnet, M., et al. (1995) A Stress-Inducible 72-kDa Heat-Shock Protein (HSP72) Is Expressed on the Surface of Human Tumor Cells, but Not on Normal Cells. International Journal of Cancer, 61, 272-279.
https://doi.org/10.1002/ijc.2910610222
[224] Gehrmann, M., Liebisch, G., Schmitz, G., et al. (2008) Tumor-Specific Hsp70 Plasma Membrane Localization Is Enabled by the Glycosphingolipid Gb3. PLOS ONE, 3, e1925.
https://doi.org/10.1371/journal.pone.0001925
[225] Multhoff, G., Botzler, C., Jennen, L., et al. (1997) Heat Shock Protein 72 on Tumor Cells: A Recognition Structure for Natural Killer Cells. The Journal of Immunology, 158, 4341-4350.
https://doi.org/10.4049/jimmunol.158.9.4341
[226] Multhoff, G. (2007) Heat Shock Protein 70 (Hsp70): Membrane Location, Export and Immunological Relevance. Methods, 43, 229-237.
https://doi.org/10.1016/j.ymeth.2007.06.006
[227] Fleishner, M. and Johnson, J.D. (2005) Endogenous Extra-Cellular Heat Shock Protein 72: Releasing Signal(s) and Functions. International Journal of Hyperthermia, 21, 457-471.
https://doi.org/10.1080/02656730500088211
[228] Tytell, M. (2005) Release of Heat Shock Proteins (HSPs) and the Effects of Extracellular HSPs on Neural Cells and Tissues. International Journal of Hyperthermia, 21, 445-455.
https://doi.org/10.1080/02656730500041921
[229] Szasz, O. (2020) Chap. 11. Local Treatment with Systemic Effect: Abscopal Outcome. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 192-205.
[230] Minnaar, C.A., Kotzen, J.A., Ayeni, O.A., et al. (2020) Potentiation of the Abscopal Effect by Modulated Electro-Hyperthermia in Locally Advanced Cervical Cancer Patients. Frontiers in Oncology, 10, Article 376.
https://doi.org/10.3389/fonc.2020.00376
[231] Lee, S.Y. and Szasz, A. (2022) Immunogenic Effect of Modulated Electro-Hyper-thermia (mEHT) in Solid Tumors. In: Interdisciplinary Cancer Research, Springer, Cham, 1-28.
https://doi.org/10.1007/16833_2022_74
[232] Beachy, S.H. and Repasky, E.A. (2011) Toward Establishment of Temperature Thresholds for Immunological Impact of Heat Exposure in Humans. International Journal of Hyperthermia, 27, 344-352.
https://doi.org/10.3109/02656736.2011.562873
[233] Shen, R.N., Lu, L., Young, P., Shidnia, H., Hornback, N.B. and Broxmeyer, H.E. (1994) Influence of Elevated Temperature on Natural Killer Cell Activity, Lymphokine-Activated Killer Cell Activity and Lectin-Dependent Cytotoxicity of Human Umbilical Cord Blood and Adult Blood Cells. International Journal of Radiation Oncology-Biology-Physics, 29, 821-826.
https://doi.org/10.1016/0360-3016(94)90571-1
[234] Hietanen, T., Kapanaen, M. and Kellokumpu, L.P.L. (2016) Restoring Natural Killer Cell Cytotoxicity after Hyperthermia Alone or Combined with Radiotherapy. Anticancer Research, 36, 555-563.
[235] Multhoff, G. (2002) Activation of Natural Killer Cells by Heat Shock Protein 70. International Journal of Hyperthermia, 18, 576-585.
https://doi.org/10.1080/0265673021000017109
[236] Krenacs, T., Meggyeshazi, N., Forika, G., et al. (2020) Modulated Electro-Hyper-thermia-Induced Tumor Damage Mechanisms Revealed in Cancer Models. International Journal of Molecular Sciencess, 21, Article 6270.
https://doi.org/10.3390/ijms21176270
[237] Wust, P., Veltsista, P.D., Oberacker, E., Yavvari, P., Walther, W., Bengtsson, O., et al. (2022) Radiofrequency Electromagnetic Fields Cause Non-Temperature-Induced Physical and Biological Effects in Cancer Cells. Cancers, 14, Article 5349.
https://doi.org/10.3390/cancers14215349
[238] Kao, P.H.-J., Chen, C.-H., Chang, Y.-W., et al. (2020) Relationship between Energy Dosage and Apoptotic Cell Death by Modulated Electro-Hyperthermia. Scientific Reports, 10, Article No. 8936.
https://www.nature.com/articles/s41598-020-65823-2
https://doi.org/10.1038/s41598-020-65823-2
[239] Cha, J., Jeon, T.-W., Lee, C.-G., et al. (2015) Electro-Hyperthermia Inhibits Glioma Tumorigenicity through the Induction of E2F1-Mediated Apoptosis. International Journal of Hyperthermia, 31, 784-792.
https://doi.org/10.3109/02656736.2015.1069411
[240] Schvarcz, C.A., Danics, L., Krenacs, T., Viana, P., Beres, R., et al. (2021) Modulated Electro-Hyperthermia Induces a Prominent Local Stress Response and Growth Inhibition in Mouse Breast Cancer Isografts. Cancers, 13, Article 1744.
https://pubmed.ncbi.nlm.nih.gov/33917524/
https://doi.org/10.3390/cancers13071744
[241] Kuo, I.-M., Lee, J.-J., Wang, Y.-S., et al. (2020) Potential Enhancement of Host Immunity and Anti-Tumor Efficacy of Nanoscale Curcumin and Resveratrol in Colorectal Cancers by Modulated Electro-Hyperthermia. BMC Cancer, 20, Article No. 603.
https://doi.org/10.1186/s12885-020-07072-0
[242] Szasz, A.M., Minnaar, C.A., Szentmartoni, G., et al. (2019) Review of the Clinical Evidences of Modulated Electro-Hyperthermia (mEHT) Method: An Update for the Practicing Oncologist. Frontiers in Oncology, 9, Article 1012.
https://doi.org/10.3389/fonc.2019.01012
[243] Fiorentini, G., Sarti, D., Casadei, V., et al. (2020) Chap. 6. Modulated Electro-Hyperthermia for the Treatment of Relapsed Brain Gliomas. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 110-125.
https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia
[244] Roussakow, S. (2017) Clinical and Economic Evaluation of Modulated Electrohyperthermia Concurrent to Dose-Dense Temozolomide 21/28 Days Regimen in the Treatment of Recurrent Glioblastoma: A Retrospective Analysis of a Two-Centre German Cohort Trial with Systematic Comparison and Effect-to-Treatment Analysis. BMJ Open, 7, e017387.
http://bmjopen.bmj.com/content/bmjopen/7/11/e017387.full.pdf
https://doi.org/10.1136/bmjopen-2017-017387
[245] Parmar, G., Rurak, E., Elderfield, M., et al. (2020) Chap. 13. 8-Year Observational Study on Naturopathic Treatment with Modulated Electro-Hyperthermia (mEHT): A Single-Centre Experience. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 227-266.
https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia
[246] Chi, K.-H. (2020) Chat. 12. Tumour-Directed Immunotherapy: Clinical Results of Radiotherapy with Modulated Electro-Hyperthermia. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 206-226.
https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia
[247] Ou, J., Zhu, X., Chen, P., et al. (2020) A Randomized Phase II Trial of Best Supportive Care with or without Hyperthermia and vItamin C for Heavily Pretreated, Advanced, Refractory Non-Small-Cell Lung Cancer. Journal of Advanced Research, 24, 175-182.
https://doi.org/10.1016/j.jare.2020.03.004
[248] Lee, D.-Y., Park, S.-J., Jung, H.-C., et al. (2015) The Outcome of the Chemotherapy and Oncothermia for Far Advanced Adenocarcinoma of the Lung: Case Reports of Four Patients. Advances in Lung Cancer, 4, 1-7.
http://www.scirp.org/journal/PaperInformation.aspx?PaperID=54620
https://doi.org/10.4236/alc.2015.41001
[249] Szasz, A. (2014) Current Status of Oncothermia Therapy for Lung Cancer. Korean Journal of Thoracic and Cardiovascular Surgery, 47, 77-93.
https://doi.org/10.5090/kjtcs.2014.47.2.77
[250] Iyikesici, M.S. (2020) Survival Outcomes of Metabolically Supported Chemotherapy Combined with Ketogenic Diet, Hyperthermia, and Hyperbaric Oxygen Therapy in Advanced Gastric Cancer. Nigerian Journal of Clinical Practice, 23, 734-740.
https://doi.org/10.4103/njcp.njcp_509_18
[251] Chi, M.-S., Wu, J.-H., Shaw, S., et al. (2021) Marked Local and Distant Response of Heavily Treated Breast Cancer with Cardiac Metastases Treated by Combined Low Dose Radiotherapy, Low Dose Immunotherapy and Hyperthermia: A Case Report. Therapeutic Radiology and Oncology, 5, 1-7.
https://doi.org/10.21037/tro-21-16
[252] Nagata, T., Kanamori, M., Sekine, S., et al. (2021) Clinical Study of Modulated Electro-Hyperthermia for Advanced Metastatic Breast Cancer. Molecular and Clinical Oncology, 14, Article No. 103.
https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8010507/
https://doi.org/10.3892/mco.2021.2265
[253] Szasz, A.M.,, Szentmartoni, G., Garay, T., et al. (2020) Chap. 5. Breast Cancer Series Treated with Modulated Electro-Hyperthermia (mEHT)—A Single Centre Experience. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 105-109.
https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia
[254] Chi, M.-S., Mehta, M.P., Yang, K.-L., et al. (2020) Putative Abscopal Effect in Three Patients Treated by Combined Radiotherapy and Modulated Electrohyperthermia. Frontiers in Oncology, 10, Article 254.
https://www.frontiersin.org/articles/10.3389/fonc.2020.00254/full
https://doi.org/10.3389/fonc.2020.00254
[255] Herold, Z., Szasz, A.M. and Dank, M. (2021) Evidence Based Tools to Improve Efficiency of Currently Administered Oncotherapies for Tumors of the Hepatopancreatobiliary System. World Journal of Gastrointestinal Oncology, 13, 1109-1120.
https://doi.org/10.4251/wjgo.v13.i9.1109
[256] Gadaleta-Caldarola, G., Infusino, S., Galise, I., et al. (2014) Sorafenib and Locoregional Deep Electro-Hyperthermia in Advanced Hepatocellular Carcinoma. A Phase II Study. Oncology Letters, 8, 1783-1787.
https://doi.org/10.3892/ol.2014.2376
[257] Fiorentini, G., Sarti, D., Ranieri, G., et al. (2021) Modulated Electro-Hyperthermia in Stage III and IV Pancreatic Cancer: Results of an Observational Study on 158 Patients. World Journal of Clinical Oncology, 12, 1064-1071.
https://doi.org/10.5306/wjco.v12.i11.1064
[258] Petenyi, F.G., Garay, T., Muhl, D., et al. (2021) Modulated Electro-Hyperthermic (mEHT) Treatment in the Therapy of Inoperable Pancreatic Cancer Patients—A Single-Center Case-Control Study. Diseases, 9, Article 81.
https://doi.org/10.3390/diseases9040081
[259] Fiorentini, G., Sarti, D., Casadei, V., et al. (2019) Modulated Electro-Hyperthermia as Palliative Treatment for Pancreas Cancer: A Retrospective Observational Study on 106 Patients. Integrative Cancer Therapies, 18, 1-8.
https://doi.org/10.1177/1534735419878505
[260] Volovat, C., Volovat, S.R., et al. (2014) Second-Line Chemotherapy with Gemcitabine and Oxaliplatin in Combination with Loco-Regional Hyperthermia (EHY-2000) in Patients with Refractory Metastatic Pancreatic Cancer—Preliminary Results of a Prospective Trial. Romanian Reports in Physics, 66, 166-174.
[261] Pang, C.L.K., Zhang, X., et al. (2017) Local Modulated Electro-Hyperthermia in Combination with Traditional Chinese Medicine vs. Intraperitoneal Chemoinfusion for the Treatment of Peritoneal Carcinomatosis with Malignant Ascites: A Phase II Randomized Trial. Molecular and Clinical Oncology, 6, 723-732.
https://doi.org/10.3892/mco.2017.1221
[262] Deniz, G.I., Can, A. and Tansan, S. (2022) Chemotherapy and Radiofrequency Hyperthermia in Advanced Ovarian Cancer. Journal of Clinical Oncology, 40, e17550.
https://ascopubs.org/doi/pdf/10.1200/JCO.2022.40.16_suppl.e17550
https://doi.org/10.1200/JCO.2022.40.16_suppl.e17550
[263] Yoo, H.J., Lim, M.C., Seo, S.-S., et al. (2019) Phase I/II Clinical Trial of Modulated Electro-Hyperthermia Treatment in Patients with Relapsed, Refractory or Progressive Heavily Treated Ovarian Cancer. Japanese Journal of Clincal Oncology, 49, 832-838.
https://doi.org/10.1093/jjco/hyz071
[264] Maragkos, M., Danilidis, L., Vakalis, I., Kouridakis, P., Papastavrou, A. and Barich, A. (2017) Treatment of Advanced Metastatic Prostate Cancer by Oncothermia (EHY-2000). Oncothermia Journal, 20, 138-155.
[265] Douwes, F.R. and Lieberman, S. (2002) Radiofrequency Transurethral Hyperthermia and Complete Androgen Blockade. A Nonsurgical Approach to Treating Prostate Cancer. Alternative & Complementary Therapies, 8, 149-156.
https://doi.org/10.1089/107628002760090994
[266] Lee, S.Y. and Lee, N.-R. (2018) Positive Response of a Primary Leiomyosarcoma of the Breast Following Salvage Hyperthermia and Pazopanib. The Korean Journal of Internal Medicine, 33, 442-445.
http://www.ncbi.nlm.nih.gov/pubmed/27079325
https://doi.org/10.3904/kjim.2015.242
[267] Jeung, T.-S., Ma, S.-Y., Choi, J., et al. (2015) Results of Oncothermia Combined with Operation, Chemotherapy and Radiation Therapy for Primary, Recurrent and Metastatic Sarcoma. Case Reports in Clinical Medicine, 4, 157-168.
https://doi.org/10.4236/crcm.2015.45033
[268] Volovat, C., Volovat, S.R., Scripcaru, V., et al. (2014) The Results of Combination of Ifosfamid and Locoregional Hyperthermia (EHY 2000) in Patients with Advanced Abdominal Soft-Tissue Sarcoma after Relapse of First Line Chemotherapy. Romanian Reports in Physics, 66, 175-181.
[269] Garay, T., Kiss, E., Szentmartoni, G., et al. (2020) Chap. 8. Gastrointestinal Cancer Series Treated with Modulated Electro-Hyperthermia (mEHT)—A Single Centre Experience. In: Szasz, A., Ed., Challenges and Solutions of Oncological Hyperthermia, Cambridge Scholars Publishing, Cambridge, 159-162.
https://www.cambridgescholars.com/challenges-and-solutions-of-oncological-hyperthermia
[270] Reimnitz, U. (2010) Cholangiocellular Carcinomas: Survival without Symptoms with Hyperthermia—A Case Study. Oncothermia Journal, 1, 20-22.
[271] Ranieri, G., Laface, C., Laforgia, M., et al. (2020) Bevacizumab plus FOLFOX-4 Combined with Deep Electro-Hyperthermia as First-Line Therapy in Metastatic Colon Cancer: A Pilot Study. Frontiers in Oncology, 10, Article 590707.
https://doi.org/10.3389/fonc.2020.590707
[272] Ranieri, G., Ferrari, C., Di Palo, A., et al. (2017) Bevacizumab-Based Chemotherapy Combined with Regional Deep Capacitive Hyperthermia in Metastatic Cancer Patients: A Pilot Study. International Journal of Molecular Sciences, 18, Article Article 1458.
https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5535949/
https://doi.org/10.3390/ijms18071458
[273] Lee. D., Kim, S.-S., Seong, S., et al. (2016) Stage IV Wilms Tumor Treated by Korean Medicine, Hyperthermia and Thymosin-α1: A Case Report. Case Reports in Oncology, 9, 119-125.
https://doi.org/10.1159/000443724
[274] Jeung, T.S., Ma, S.Y., Yu, J., et al. (2013) Cases That Respond to Oncothermia Monotherapy. Conference of the International Clinical Hyperthermia Society 2012, 2013, Article ID: 392480.
https://www.hindawi.com/journals/cpis/2013/392480/
https://doi.org/10.1155/2013/392480
[275] Kim, S., Lee, J.H., Cha, J. and You, S.H. (2021) Beneficial Effects of Modulated Electro-Hyperthermia during Neoadjuvant Treatment for Locally Advanced Rectal Cancer. International Journal of Hyperthermia, 38, 144-151.
https://pubmed.ncbi.nlm.nih.gov/33557636/
https://doi.org/10.1080/02656736.2021.1877837
[276] You, S.H. and Kim, S. (2019) Feasibility of Modulated Electro-Hyperthermia in Preoperative Treatment for Locally-Advanced Rectal Cancer: Early Phase 2 Clinical Results. Neoplasma, 67, 677-683.
https://www.ncbi.nlm.nih.gov/pubmed/32039629
https://doi.org/10.4149/neo_2020_190623N538
[277] Minnaar, C.A., Maposa, I., Kotzen, J.A. and Baeyens, A. (2022) Effects of Modulated Electro-Hyperthermia (mEHT) on Two and Three Year Survival of Locally Advanced Cervical Cancer Patients. Cancers, 14, Article 656.
https://doi.org/10.3390/cancers14030656
[278] Minnaar, C.A., Kotzen, J.A. and Baeyens, A. (2018) Possible Potentiation of the Abscopal Effect of Ionising Radiation by Modulated Electro-Hyperthermia in Locally Advanced Cervical Cancer Patients. Oncothermia Journal, 24, 122-132.

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.