Next Article in Journal
Effects of Long-Term Blue Light Exposure on Body Fat Synthesis and Body Weight Gain in Mice and the Inhibitory Effect of Tranexamic Acid
Previous Article in Journal
RIPK1 in Diffuse Glioma Pathology: From Prognosis Marker to Potential Therapeutic Target
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Diversity of Effects of Mechanical Influences on Living Systems and Aqueous Solutions

by
Sergey V. Gudkov
,
Vladimir I. Pustovoy
,
Ruslan M. Sarimov
,
Dmitriy A. Serov
,
Alexander V. Simakin
and
Ivan A. Shcherbakov
*
Prokhorov General Physics Institute of the Russian Academy of Sciences, Vavilove St. 38, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2025, 26(12), 5556; https://doi.org/10.3390/ijms26125556
Submission received: 22 April 2025 / Revised: 31 May 2025 / Accepted: 4 June 2025 / Published: 10 June 2025
(This article belongs to the Collection Latest Review Papers in Molecular Biophysics)

Abstract

Water is the basis of life. Any factors acting on water will also affect the functioning of living organisms, including humans. Mechanical effects are as ubiquitous as temperature or magnetic fields. Numerous works have been devoted to the action of mechanical impacts on living systems, aqueous solutions, and water. However, no unified theory that would allow predicting the consequences of mechanical effects on living organisms based on their characteristics. In this review, we have attempted to systematize the available quantitative data on the effects of mechanical impacts on living organisms, cells, aqueous solutions, and purified water. In addition, in this review, we provide a basic overview of the variety of mechanical effects and the mechanisms of their realization. The responses of living systems and aqueous solutions depend quantitatively on different sets of characteristics of the vibration action. The magnitude of responses of living systems (cells and organisms) to mechanical action correlates with frequency, acceleration, and force. Mechanical action changes the characteristics of water and aqueous solutions as a function of frequency, acceleration, and duration. The data obtained may find application in a wide range of fields: from analytical chemistry and pharmacology to environmental protection.

1. Introduction

Water and aqueous solutions are subjected to mechanical action on a daily basis: from bars to pharmaceutical production, from water purification systems to the propellers of supertankers, from air humidifiers to watering plants in agriculture [1,2,3,4]. Mechanical action on aqueous solutions is used for various tasks in human activities. The most common application of mechanical action to aqueous solutions is to dissolve and/or mix and/or change temperature and/or transfer torque/force and/or wet something [5]. Mechanical effects are even used in nanotechnology, e.g., to concentrate particle separation molecules from the solvent by flotation [6] or to control nanoparticle aggregation [7]. More than a hundred years ago, it was found that mechanical impacts of high intensities are observed to alter/damage organic molecules in water. For example, denaturation of proteins in aqueous colloids is observed when they are mechanically impacted [8]. This discovery was later proposed to be used as a stress test to check the quality of antibodies [9]. Today, a whole section of chemistry (mechanochemistry) studies changes in the properties of substances and their mixtures, as well as physicochemical transformations under mechanical effects [10]. The negative effects of prolonged mechanical impact on the level of living tissues and organisms have been well described. There is a separate group of so-called vibration diseases, which belong to occupational diseases and affect various body systems. The complex of vibration diseases affects both pathologies of the nervous system (peripheral neuropathy) and musculoskeletal system (aseptic necrosis, fatigue fractures, degenerative joint disease, carpal tunnel syndrome), as well as the cardiovascular system (Raynaud’s syndrome) [11]. Even though modern working conditions are designed to provide a high level of safety in chronic vibration exposure due to modernization of impact tools and new technological elements of vibration protection [12], the incidence of vibration diseases in workers of construction, mining, metalworking industries and operators of heavy machinery is still about 10–40%, depending on the length of service and occupation [13]. It is known that the highest frequency of vibration disease is observed among workers in the mining industry and construction workers. The risk of vibration diseases increases significantly after 5–7 years of work in these industries. Unfortunately, there are no unified working recommendations on the prevention and ideal treatment of vibration diseases. Unfortunately, there are no unified working recommendations on prevention and ideal therapy of vibration diseases today [14].
In order to create more effective ways to treat aqueous solutions, to create new biotechnological approaches, to improve the quality and safety of life, it is necessary to understand in detail the regularities of mechanical (especially weak) effects on water, aqueous solutions and living organisms, which are 70–90% composed of water [15].
Almost 300,000 works have been published on the mechanobiology issues (https://pubmed.ncbi.nlm.nih.gov/?term=mechanobiology accessed on 26 March 2025). However, there is no common unified theory linking the characteristics of heterogeneous mechanical effects (vibration, shear stress, pressure, and sound action) to date. In this review, we have attempted to find general regularities that determine the magnitude of the effect of mechanical action on aqueous solutions and biological objects depending on the characteristics of mechanical action (duration, frequency, acceleration, force, and other features).
In addition, this review will describe the main effects of mechanical action on purified water, aqueous solutions of organic compounds, cell cultures, and whole organisms. The main mechanisms of the realization of mechanical impact effects will be discussed.

2. Diversity of Effects of Mechanical Influences on Living Systems and Aqueous Solutions

2.1. Impact Types

Mechanical effects on aqueous solutions are ubiquitous around us in all spheres of life: household, technology, agriculture, and the pharmacological industry [16,17,18,19]. However, the exact dependence of the effects of mechanical actions on water and aqueous solutions on their characteristics has not yet been determined. The mechanical effects that can be exerted on water, aqueous solutions, and living organisms are quite diverse. They include the following: sinusoidal vibrations [20,21], different mixing modes (turbulent, laminar, etc.) [22,23,24], centrifugation [25], shear stress in constant flow [26], dropping from a height in the form of drops or vials [24,27,28], formation of nanobubbles under pressure, and others [29]. Separately, ultrasound exposure is the subject of research as a potential therapeutic agent [30,31]. In addition, mechanical effects are often used simultaneously with other weak effects, such as weak permanent and alternating magnetic fields, electric fields, and others [32,33,34,35]. Vibration effects have an impact on the physicochemical properties of solutions. Their effects have already found application in practice, in particular, it is used in the technology of preparing vibratory iterations [16,36] and the technology of preparing highly dilute solutions [37,38].
In the present work, we attempted to systematize the available literature data and to identify the general regularities that determine the magnitude of mechanical effects on water, aqueous solutions, and living organisms based on the characteristics of these effects.
Analyzed works can be conditionally divided into works on the study of mechanical effects such as shaking, impact, ultrasonic effects, etc. (Appendix A). In order to evaluate quantitative dependencies of the magnitude of the effects of impacts, we further divided them into two groups: mechanical (vibration) and ultrasonic.

2.2. Effects of Mechanical Influences

According to the objects of study, the analyzed works can be divided into three large groups:
  • The whole organism in vivo;
  • Cell cultures;
  • Aqueous solutions or deionized water.
For this reason, we will use this classification when describing the effects.

2.2.1. Mechanical Impact on Living Systems

Most of the work on in vivo mechanical effects focuses on the cardiovascular, immune, and musculoskeletal systems. The circulatory system can be a target of weak mechanical effects. At the systemic level, whole-body vibration causes an increase in HR and blood pressure [39]. In rodents, it was shown that sinusoidal mechanical vibration reduced the number of capillaries (lectin-positive vessels), venules, and arterioles (α-actin-positive blood vessels) in muscle fibers compared to controls [20]. Another work showed a decrease in vascular lumen under the effect of vibration by increasing the smooth muscle work of microvessels [40]. Micro-damage of endothelial cells (increase in the number of vacuoles, membrane ruptures), as well as damage to the integrity of the inner walls of blood vessels, were found at the cellular-tissue level [41]. Intermittent mechanical impacts, even on isolated body parts, can induce a proinflammatory response in the body, expressed in the activation of NFAT-dependent signaling pathways, increased IL-6 synthesis, and oxidative stress [40,41]. Intermittent jade vibration of isolated limbs increases the activity of total plasma creatine phosphokinase (t-CPK), the excess of which is a marker of tissue damage [42,43]. However, more moderate loads with low acceleration and high total duration can have a positive effect on bone mineralization in humans by reducing TRAP-dependent osteoclast activity [44,45].
In humans, whole-body vibration after 8 days increased plasma concentrations of the osteoporosis markers P1NP and CTx [46,47]. An increase in leg skin temperature has also been reported after vibration impact [46]. In another study on rats with hip implants, an improvement in bone-implant contact was shown [48]. In a rat osteoporosis model, whole-body vibration promoted protection against obesity and increased bone mineralization, and worsened their mechanical characteristics (stiffness, maximum load in vertebra compression test) [28]. The mechanism of the vibration effect involves a decrease in the viability of osteoblasts, a reduction in alkaline phosphatase (ALP) activity, and a decrease in the expression of osteogenesis regulators, including MMP-2 and OSX [28,49,50]. A positive effect of intermittent vibration on muscle strength assessed by apparent mass and mechanical resistance was observed in studies on volunteers [51].
A positive effect of mechanical vibration on human respiratory system functions (peak inspiratory flow rate, inspired volume, and expired volume) has been described [52].
Weak mechanical vibration can also act on the nervous system, in particular, it increases the diameter of myelin fibers in muscles in rabbits [53]. Similar results were obtained in rats, demonstrating an acceleration of regeneration of the injured nerve due to an increase in IGF-I expression [54]. In part of the in vivo studies, the quantitative acceleration-effect or frequency-effect relationship was bell-shaped or reached a plateau after certain values [48].
Shaking can also act on plants in addition to animals. In particular, it has been shown that shaking at an angle of 45° slows seed germination by reducing seedling stem length, number of leaves, basal diameter, and total shoot biomass [55].
The effects of mechanical action (centrifugation, shear stress, vibration) depend largely on the cell lineage and may be opposite for different cell types. Periodic application of centrifugal force (100–200 g centrifugation) to culture stem cells from human exfoliated deciduous teeth inhibited their proliferation [25]. Vibrations altered the proliferation rate of Chinese hamster ovary cells and T cells [56]. The direction of the effect depended on the frequency: at 30 Hz, an increase in proliferation was observed, and at >50 Hz, an inhibition was observed [56]. Sinusoidal vibration causes a dramatic decrease in the viability of fibroblasts [55]. Erythrocytes become more susceptible to lysis under shock shaking [57]. Vibration caused an acceleration of proliferation and differentiation in MC3T3-E1 osteoblast culture [26]. The degree of differentiation was assessed by alkaline phosphatase activity, osteocalcin levels, Runx2, and Osx gene expression [26,58]. Notably, the application of constant shear stress (ductal culturing) did not produce similar results as vibration [26]. Cyclic shear stress (agitation of the medium) also enhanced the differentiation of primary fibrochondrocytes: a number of BrdU-positive cells, ACAN, and SOX9 expression [59]. Vibration increased the proliferation of embryonic cell line GD25 cells and human osteoblast-like cells [60,61]. Vibration also decreased alkaline phosphatase activity in the primary culture of human osteoblast-like cells [60]. Shear stress increases the metabolism of human periodontal ligament cells (hPDLSC), increases the expression and activity of indoleamine 2,3-dioxygenase, cyclooxygenase-2, kynurenine synthesis, TGF-β1 growth factor expression, and decreases the synthesis of the pro-inflammatory cytokine IFN-β [62]. The combination of these latter effects suggests an anti-inflammatory effect of the applied exposure [63]. Vibration stimulated the synthesis of osteoarthritis markers collagen II and aggrecan in pig chondrocytes [64,65]. On the other hand, vibration increased proliferation and protected the MLO-Y4 osteocyte lineage from apoptosis [66]. The molecular mechanisms of the observed effects are decreased expression of RANKL and TNF-α [66].
However, constant shear exposure causes an increase in collagen content in mesenchymal stem cells MSCs [67]. The attenuation of differentiation in pluripotent iPSC stem cells under the influence of vibration has been described [68]. For another type of stem cell (human adipose-derived stem cells, hASCs), an increase in A cytochrome P450 monooxygenase CYP1B1 expression and metabolic activity has been described [69]. The combination of mechanical shaking and weak alternating magnetic fields causes an enhancement of fMLF- and PMA-induced ROS production by mouse granulocytes or ‘neutrophil respiratory burst’ [32,33]. Continuous shear stress can affect the calcium homeostasis of endotheliocytes and induce periodical changes in resting calcium concentration, and alter the characteristics of calcium oscillations on the background of AChR ligand application [70,71].
The application of constant pressure can be considered as a special case of mechanical action [72]. For example, ultrasonic influence can be expressed not only in intensity and frequency but also in the pressure exerted by the ultrasonic wave on the sample [73]. The application of constant pressure of different magnitudes, in the case of cells, can be modeled using substrates with different densities. Growing differentiated dendritic cells on substrates of different densities can alter cell metabolism by reducing the efficiency of glycolysis, lactase activity, and oxygen consumption [72]. In addition, a softer substrate (calculated pressure of 2000 Pa) increased oxygen consumption and gene expression of glycolysis enzymes Hexokinase II (Hk2) and glucose transporter GLUT1 (Slc2a1) compared to the harder standard substrate (calculated pressure of 50,000 Pa) [72]. In addition, cells grown on the harder substrate showed more pronounced immune activity: their immune memory and antitumor activity were improved when injected in vivo into mice with solid tumors [72]. Thus, it can be concluded that the density of the extracellular matrix is one of the potential mechanisms of regulation of immune cell functioning. Water-swelling polymers such as poly(ethylene glycol) (PEG), poly(vinyl alcohol) (PVS), poly(2-hydroxyethyl methacrylate) (PHEMA), and poly(acrylamide) (PAAM) can form elastic hydrogels that can reproduce the basic mechanical characteristics of soft tissues and be used for mechano-mediated regulation of cell function in culture [74,75,76].
The wastewater treatment problem has been relevant since ancient times and is still unsolved. The search for new, cheaper, faster, and more effective methods of wastewater treatment continues [77,78]. Vibration can be applied to increase the efficiency of bioelectrolytic wastewater treatment systems. In particular, constant agitation with nitrogen purging of water has been shown to increase nitrate uptake and sodium removal from wastewater by T. denitrificans bacteria immobilized on the electrode. At the same time, the voltage on the electrodes increased almost 2-fold compared to non-stirred water [79].
A summary of the mechanical impact effects is shown in Table 1.

2.2.2. Mechanical Impact on Water Solution

Mechanical effects on aqueous solutions include shaking, tumbling, and high-pressure blowing (formation of microbubbles). Mechanical shaking of pure water at 30 Hz also changed the pH to 0.6 pH units [80]. The effect may seem small, but it must be recalled that physiological pH in a living organism is actively maintained within narrow limits. For example, a deviation of blood pH by a fraction of a unit from the normal range of 7.35–7.45 can lead to dangerous conditions such as acidosis or alkalosis [81].
Water stirring also enhances the absorption of water in the terahertz region 140–230 cm−1 [82] and changes the characteristics of water fluorescence in the UV region: the absorption maximum shifts from 250 nm to 340 nm and the emission maximum from 270 to 330 nm [83]. The fluorescence intensity increases ~5 times [83]. Prolonged (several hours) shaking changes the spectral parameters of water in the UV region. Fluorescence increases at a wavelength of 255 nm, and light scattering increases at a wavelength of 355 nm [83].
Vertical mechanical shaking or agitation causes the generation of hydrogen peroxide H2O2, singlet oxygen, and OH-radicals in pure water [80,84,85]. It is noteworthy that the amount of ROS generated is positively dependent on the frequency, amplitude, and duration of vibration impact [86]. The type of agitation affects the efficiency of ROS generation: turbulent agitation causes the generation of ~10 times more ROS than laminar agitation [85]. However, it is worth noting that under the conditions studied by the authors (frequency 10–30 Hz, amplitude 1 cm), the concentration of ROS is very small (tens of nM), which is less than the described concentrations exerting biological effects [87,88]. The falling of deionized water drops on a solid surface also causes the generation of H2O2, and OH-radicals in water, the concentration of which grows in proportion to the height from which the drop fell [27]. High-frequency mechanical vibration of water coated with a layer of oil increased its redox potential by 2.5 times compared to untreated samples [89].
Mechanical impact changes not only the properties of water, but aqueous solutions too (both one-substance solutions and complex mixtures). Circular agitation of an aqueous solution of interferon-gamma (IFNγ) increases its electrical conductivity compared to an unagitated solution [34]. Vertical shaking at frequencies of 10–50 Hz of deionized water causes an increase in its spontaneous chemiluminescence for at least the next 15 min [21]. The addition of 1 mg/mL of BSA protein to the water cancels the effects of mechanical action on the water [21]. Circular agitation of an aqueous solution of polyclonal antibody to IFNγ enhances the luminescence of the solution [16]. A series of impacts also causes luminescence enhancement of aqueous solutions of carbohydrates, using lactose as an example [17]. Vertical shaking of an aqueous solution of human immunoglobulin causes the generation of ~200 nm nanobubbles [23,90]. At the same time, the concentration of molecular oxygen in water and water temperature (by a fraction of a percent) increased [80]. The magnitudes of the effects were proportional to the frequency of shaking.
Dropping of vials with aqueous buffer solutions of anti-antistreptavidin IgG1 enhances protein aggregation as assessed by the number of suspended particles and separation of proteins from the walls after treatment with urea and guanidine hydrochloride [24,91]. Shaking even with small amplitudes (less than 0.1 mm) and frequencies (10 Hz) retarded protein crystallization using chicken egg lysozyme in a buffer solution, for example [92]. The change in flow rate affects the precipitation of lysozyme after treatment with a high concentration of NaCl [93], indicating possible conformational changes of the protein molecule, and could potentially be applied in protein production technology.
The method of solution stirring can change the geometric characteristics of crystals formed during the evaporation of aqueous extracts of medicinal plants [22]. Turbulent mixing produces less ordered and more heterogeneous crystals, while circular mixing produces more ordered structures (having less lacunarity and entropy) with more pronounced fractal properties [22]. Serial impacts alter the shape of crystals formed from plant and inorganic components during the evaporation of aqueous extracts of echinacea, baptisia, and luffa [23].
Microfluidic agitation can also be considered a variant of mechanical action. It has been shown that microfluidic mixing of a solution of acetate of europium (AcEu) causes a change in its Raman spectra in the region of 1360–1770 cm−1, in particular increasing the intensity of the OH groups’ stretching vibrations peak [94]. The stretching of a column of an aqueous solution of hydrogel precursors between two sites affects the spatial structure of the synthesized hydrogel. By varying the tension alone, a controlled (including mathematically predicted) change in the net structure of the final polymer can be achieved [95].
Passing deionized water through a nozzle at high speed causes the creation of water nanodroplets that contain hydrogen peroxide. The size of the microdroplets determined the concentration of hydrogen peroxide that was generated in them. The smaller the microdroplet, the greater the concentration of hydrogen peroxide. It is noteworthy that degassing water by nitrogen purging did not cause a significant decrease in hydrogen peroxide generation in the microdroplets compared to the control. Hence, the process may occur without the involvement of oxygen [90]. The creation of microdroplets can be one of the ways to catalyze chemical reactions. In particular, when microdroplets of aqueous solution of substrates of the Pomeranz-Fritsch reaction are squeezed together under accelerated nitrogen purging, the reaction rate can be increased by 5–7 times compared to the reaction rate in the control solution [29]. The combination of mechanical and electrical effects can weakly change the volt-ampere characteristics of solutions of organic and inorganic mercury compounds [35]. Stirring of the solution can also be attributed to mechanical effects. Stirring of water creates nanobubbles in it, the number of which is proportional to the duration of stirring [96].

2.3. Ultrasound Impact

Works on the ultrasound (US) impact on living systems are usually represented by studies on cell lines and aqueous solutions [95]. The influence of ultrasound on living organisms is investigated by the new field of sonobiology [96]. We will combine in vivo and in vitro results in one section when discussing the effects.

2.3.1. US Impact on Living Systems

Periodic US exposure promoted more effective tissue regeneration in rats with a model of muscle injury (transection), which was estimated by total protein mass and muscle fiber diameter due to increased myocyte proliferation [97,98]. The US may also accelerate drug delivery in eye tissues [99]. It was demonstrated that exposure to ultrasound at a frequency of 1 MHz and an intensity greater than 8 W/cm2 disrupted the ER structure. Immediately after exposure, disruption to the ER microstructure was observed in all the cells studied, and after three and 24 h, recovery of the ER was observed in 16% and 26% of cells, respectively [100]. The effect of ultrasound depends on the organism’s taxonomic classification. For example, ultrasound did not affect Vicia faba root meristem cells at a frequency of 1.1 MHz and 8 mW/cm2 [101], whereas animal cells showed effects at lower frequencies and intensities.
The works on the influence of the US on cell proliferation and differentiation are quite numerous. There are data on the effect on the primary culture of human fibroblasts, human amnion-derived mesenchymal stem cells (hAD-MSC), and murine myoblasts C2C12 [102,103,104]. Ultrasound causes the acceleration of human fibroblasts’ proliferation, increased expression of collagen, and secretion of proinflammatory cytokines IL-1β and IL-8 [102]. The regulation of proliferation occurs with the participation of EGFb regulatory factors [102,105]. Ultrasound enhances the secretion of angiogenesis stimulator VEGF by fibroblasts, which may indicate the ability of the US to enhance angiogenesis in tissues [102,106]. The effect is slightly dependent on the intensity and, to a lesser extent, on the US frequency [102]. Ultrasound enhances the differentiation and division rate of hAD-MSC stem cells by increasing protein synthesis of the cell cycle regulators Cyclin D1/β, Cyclin B1/β, Cyclin E1/β, and Cyclin A1/β [103]. Since the concentrations of all four forms of Cyclins were increased, it can be concluded that US affects beyond the G1, G2, and M stages of the cell cycle [107]. In terms of blood cells, the following were recorded in the US: induction of Ca2+ influx and platelet aggregation; induction of apoptosis in the HL-60 and U937 leukocyte cell lines; hemolysis of erythrocytes; and an increase in the number of chromosomal aberrations in whole blood cells [108,109,110,111]. Ultrasound with a frequency of over 5 MHz can alter the spatial orientation of platelets [112].
Ultrasound almost doubled the rate of division of mouse myoblast culture and bovine aortic endothelial cells, and promoted differentiation, expressed as an increase in the length of myotubes and the number of cells fused into one myotube, and regeneration of myofibrils [104,113,114]. At the molecular level, the expression of myogenin, Pax7, COX2, and the inhibition of pro-inflammatory reactions were observed [113,115]. The action of the US on neuronal cells accelerates the differentiation, assessed by the expression of markers Alk, Cenpf, and Pcdh17, as well as the expression of actin involved in the formation of outgrowths [116].
The US destroys bacterial spores and cells, and induces DNA release from them [117,118]. The US also accelerates DNA plasmid transfer in bacterial cultures [119]. The US inhibits plant photosynthesis [120]. A summary of US impacts on living systems is shown in Table 2.

2.3.2. US Impact on Water Solutions

The ultrasonic treatment of deionized and artesian water changes its pH by 0.3–0.6 and 0.5–1.05 pH units, respectively [121]. At the same time, the US treatment causes acidification of deionized water and alkalization of artesian water, indicating the role of dissolved salts in the effect of ultrasonic treatment on water pH [121]. US treatment promotes accelerated crystallization of HEWL in acetate buffer [93]. Currently, approaches are being developed for the practical application of the US in accelerating the precipitation of soya proteins in biotechnological processes [122,123]. In addition to proteins, ultrasonics accelerate the precipitation of clay particles in aqueous colloids [124]. Increasing the duration of the US treatment favors faster precipitation. Methods for accelerating precipitation using ultrasonics have also found applications in other fields, such as nanoparticle synthesis [4,125]. The US can catalyze H2O2 generation reactions in water in the presence of solid catalysts (Bi3TiNbO9) [126]. Currently, there is a problem of finding a cheaper and safer method of hydrogen peroxide synthesis than the currently used anthraquinone process [127], so the use of US and solid catalysts for this purpose seems reasonable.

3. Mechanisms of the Mechanical Impact Effects

3.1. Receptor Response of Living Systems. Mechanotransduction

Mechanotransduction is a group of processes of signal conversion from receptor structures of cells into molecular and biochemical responses of cells [128,129,130]. Mechanotransduction involves millisecond processes of the response of sensitive molecules and structures (usually receptors) and triggering of signaling pathways (changes in transcriptional activity, auto- and paracrine regulation), which further determine the fate of cells and the organism as a whole [31]. Mechanotransduction events can be conditionally divided into three main stages: mechanosensing, mechanotransmission, and mechanoresponse. Classically, these processes appear to be sequential; however, parallel and cyclic processes are possible [131,132,133].
The first stage includes receptor perception of a mechanical stimulus due to changes in the conformation of receptor proteins. Mechanosensing can occur passively in either an outside-in or an inside-out way. Outside-in scheme realized through events: “external force → change of protein structure → signal transduction”.
When implementing the inside-out mechanism, two modes are possible:
  • “elements of cytoskeleton → creation of internal force on the mechanoreceptor from the intracellular matrix → activation of mechanoreceptor → detection of changes in the extracellular matrix under external mechanical stimulation” (inside out).
  • “absence of internal force on the mechanoreceptor from the intracellular matrix → inhibition of mechanoreceptor → inhibition of detection of changes in the extracellular matrix under external mechanical stimulation”
The advantage of the first pathway is the ability of the cell to respond to mechanical stimuli at any time (excluding refractoriness). The advantage of the second method is the possibility for selective “on/off” mechanosensitivity by the cell [128].
Examples of outside-in receptors are TRP receptors (TRPV1, TRPV2, TRPV4, etc.), which can open directly upon mechanical membrane deformation or open indirectly via a phospholipase A2 (PLA2)-dependent pathway [134,135]. Nuclear proteins associated with cytoskeleton proteins transmit mechanical stress to the nuclear apparatus and alter the transcriptional activity of the cell. For example, LINC (Linker of Nucleoskeleton and Cytoskeleton) complexes include a group of protein INM on the inner nuclear membrane, SUN1 and SUN2 proteins in the nuclear envelope lumen and outer nuclear membrane, and KASH proteins on the outer side of the outer nuclear membrane [136]. KASH protein mediates interactions with actin cables through Nesprin 2. A-type and B-type lamins are also involved in mechanoreception through the spatial reorganization of chromosomes [137,138,139]. Some subtypes of K+ channels, K2P and ASIC families, can also respond to mechanical stress [140,141,142,143]. Different ASIC subtypes can act as primary mechanosensing receptors (ASIC1a, ASIC1b) or modulate and regulate the pain pathway and nociception as a special case of mechanoreception (ASIC2 and ASIC3) [14,144,145,146].
The inside-out system consists of actin, talin, myosin II, vinculin, and integrin proteins bound in the extracellular matrix from the fibrillar proteins surrounding the cell [147,148].
It is worth noting that many receptor structures and ensembles are capable of operating in both modes. In particular, FA (focal adhesions)-proteins, membrane G-protein coupled receptors (GPCRs), dense intercellular AJs contacts (adherens junctions), and most of the stress-activated channels (SACs) [21,74,75,76,128,131,132,133,134,135,137,138,139,147,148,149,150,151,152,153,154,155,156,157,158]. Therefore, the division into inside-out and outside-in is more valid for how a particular response of the same structure is regulated than for a specific structure.
SACs include numerous ion channels of varying degrees of specificity belonging to the K2P potassium channel (TREK-1, TREK-2), DEG-ENa-ASIC, TRP, and piezo channel families [142,143,159,160,161]. There are two models of regulation of ion channel opening under the action of mechanical load: “force-from-lipid” and “force-from-filament”. According to the first “force-from-lipid” model, the response to mechanical stimulation follows the scheme: external force → reorganization of the lipid environment of receptors → conformational changes of the receptor → opening/closing of the channel or another way of signal transduction [149,154,157]. According to the second model, ion channels, such as part of SACs, are physically bound to proteins of the cytoskeleton and/or extracellular matrix. In this case, the channel opening scheme would be as follows: external force → tension changes of the extracellular or intracellular matrix → conformational change of the channel. It is assumed that changes in the tension of the cell matrix cause reorganization of membrane lipids. The latter further facilitates signal transmission from an open channel to a still closed channel [151,152]. Thus, signal transduction from an open mechanosensitive channel to long distances along the membrane can be achieved.
Further signal transduction (mechanotransmission and mechanoresponse steps) can occur through several signaling pathways. It is noteworthy that the type of receptors involved and the further signal transduction pathway are determined by the type of acting mechanical action [31]. The whole variety of mechanical impacts on the cellular level is finally reduced to compression, tension, shear, and their combinations.
Cell compression is perceived by receptors from the integrin families, piezo channels, and TRP- and FA-receptors that trigger Wnt/β- and YAP/TAZ-dependent signaling pathways [162,163,164,165]. Tension is perceived by a similar set of receptors (integrins, TRPs, FAs) with additional cytoskeletal involvement. However, the signaling pathways triggered are different and include Rho-ROCK, MAPK, and ERK kinase cascades [166,167]. Shift (shear stress) is recognized by a wider range of receptors than compression and tension. A shift-sensing system including GPCRs, tyrosine kinase receptors, lipid rafts, gap junctions, and cell contacts, triggering both Wnt/β-dependent and kinase signaling pathways (Rho-ROCK, MAPK ERK) [168,169,170,171,172,173,174,175,176].
Mechanotransduction during US treatment should be separately mentioned. The main acting factor in this case is the shock wave formed by the cavitation of microbubbles [168]. Acoustic stimulation activates TRPs, piezo channels, and cell contact areas. Signaling pathways include cascades involving Rho-ROCK, MAPK, ERK, Wnt, ESCRT, Epac-Rap, YAP, and PI3K/Akt [177]. The reader can learn more about each of these signaling pathways in the cited reviews. Another mechanism involves action on cell membrane lipids [178].

3.2. ROS Generation in Water Solutions

Water is subjected to mechanical action, not only by human will, but also by nature. For example, when water droplets fall on a solid surface during a rainstorm or when water falls from a waterfall. In this case, physicochemical parameters such as concentration of dissolved gases, hydrogen index, temperature, electrical conductivity [179], redox potential [89], and light scattering spectra in the near-ultraviolet region [83] change in water. Changes in some physicochemical properties of water after mechanical impact can persist for up to several days [180]. It has been found that mechanical impact leads to the formation of reactive oxygen species (ROS) [27]. ROS are extremely reactive compounds that, on the one hand, can damage biological macromolecules [181] and, on the other hand, play an important signaling and regulatory role in biological systems [182]. Of all ROSs, only hydrogen peroxide is a stable molecule. The other ROS species have short lifetimes [183]. This allows hydrogen peroxide to accumulate in the medium under prolonged exposure, which facilitates detection and allows us to work with large concentrations. In this connection, most of the works most often investigate the effect of a particular low-intensity physical factor on hydrogen peroxide generation. The generation of hydrogen peroxide under mechanical action seems to occur according to two main scenarios:
  • Activation of oxygen molecules dissolved in water;
  • Dissociation of water molecules and hydroxyl anions.
The first scenario is related to the activation of molecular oxygen (Figure 1). It has been shown that under mechanical action, a part of the molecules of dissolved oxygen in water change from a triplet to a singlet state [184]. It is assumed that the transition of oxygen molecules from the triplet state to the singlet state occurs due to the impacts of dipoles [185]; it is also assumed that a local magnetic field can be generated under mechanical action, affecting the triplet-singlet transition [186]. More detailed views on the processes occurring with oxygen molecules dissolved in water under mechanical action can be found in studies [86,187,188]. Experimental validation of scenario 1 and its detailed discussions in studies [80,179]. The mechanical-induced generation of singlet oxygen in water was proofed by inhibitory analysis with sodium azide, NaN3 (singlet oxygen quencher). NaN3 inhibits mechanically induced ROS production (measured by luminol-dependent chemiluminescence) by 30%. These data indicate that singlet oxygen is involved in the generation of hydrogen peroxide during mechanical stress (30 Hz, 5 mm amplitude, 5 min) [80].
Singlet oxygen is further reduced to superoxide anion radical, which is protonated to hydroperoxide radical. The dismutation of hydroperoxide radicals produces hydrogen peroxide and molecular oxygen, in some cases in the singlet state; the cycle is closed. A similar scenario is observed in the generation of hydrogen peroxide upon exposure to visible radiation [189], infrared radiation [190], and heat [191]. The chemical transformations of ROS characteristic of this scenario are described in detail [192].
The second scenario is related to the oxidation of hydroxyl anions and/or the dissociation of water molecules. The possibility of the existence of such a process was theoretically predicted almost half a century ago [193]. This prediction has been criticized in terms of the energetics of the process [194]. It is believed that a quantum with an energy of about 5 eV is required to break the O-H bond in a water molecule [195]. The formation of hydroxyl radicals from water molecules has been observed under the action of ionizing radiation [196], plasma [197], and other high-energy effects, for example, during the optical breakdown of water [198]. In general, the scientific community was rather skeptical about the information concerning the generation of hydroxyl radicals from hydroxyl anion or water molecules under ‘weak’ non-ionizing effects, including mechanical ones, until the publication of theoretical works by Prof. Nowakowka’s group [199] and others. Five years ago, the phenomenon of hydroxyl radical generation under mechanical action was brilliantly experimentally confirmed by the example of the generation of micron-sized water droplets [90]. Today, it is assumed that the oxidation of hydroxyl anions is based on the mechanism of asymmetric charge of water droplets. When water is atomized, charged microdroplets are formed carrying excessive amounts of their own OH or H+ ions, which repel each other towards the droplet surface. The formation of ROS is attributed to limited hydration at the water-air phase interface [200]. It is assumed that both of the described scenarios occur in parallel during mechanical action in water. The efficiency of ROS generation in the first or second scenario mainly depends on such parameters as the frequency and amplitude of impact, conditions for the generation of micron-sized droplets during impact, and the concentration of molecular oxygen dissolved in water.

4. Dependence of the Magnitude of the Effects of Mechanical Impacts on Their Characteristics

4.1. Mechanical Influence

4.1.1. Principal Conditions of Realisation of Effects

In the first stage, we evaluated the principal limits of the characteristics of the described mechanical effects. m/s2), and Characteristics we chose: frequency of vibration (Hz), acceleration that the sample acquires (sum duration of impact (s). Impact times, if expressed in hours or minutes, were converted to seconds. If periodic incubation was described (e.g., 5 times 100 s), the sum of the times was calculated and expressed in seconds (e.g., 500 s). Acceleration values were taken as initial values given in publications, or acceleration was calculated as the second derivative of the coordinates using the formula:
x ( t ) = A sin ( 2 π f t ) ,
x ( t ) = 4 A π 2 f 2 sin ( 2 π f t )
x max = 4 A π 2 f 2
where x(t) is the sample coordinate, f is the frequency, A is the oscillation amplitude, and x’’max is the maximum acceleration.
Frequency in cases of expression in rpm, etc., was converted into Hz. In this case, to determine the specified limits, we constructed three-dimensional distributions. It can be seen that the boundaries of applied frequencies, accelerations, and durations are quite strictly localized (Figure 2). Most effects were registered at frequencies from 0.01 to 500 Hz and at accelerations from 0.01 to 1000 m/s2. The duration of exposure required to achieve the effects increased as both frequency and acceleration decreased. Consequently, the effects of mechanical influence are realized when a certain “threshold” of transferred kinetic energy is overcome. This threshold is determined simultaneously by all three components: frequency, acceleration, and duration. At the same time, the values are partially interchangeable: low values of frequency and acceleration can be compensated by high duration and vice versa. Please note that Figure 2 shows not the magnitudes of the effects, but the threshold values of frequency, amplitude, and duration at which any effects were observed (taken as an “all-or-nothing” response). The quantitative values of the effects are already presented in the following figures.

4.1.2. Dependences of Mechanical Effects on the Combination of Factors

The articles analyzed assessed a wide variety of effects of mechanical exposure: from the concentration of ROS and inorganic ions to the fraction of differentiated cells and apparent human muscle strength [26,51,79,86,104]. All values are expected to have different units of measurement. Thus, it is not possible to directly compare the data obtained in different articles. For this reason, we have converted all data into dimensionless quantities: effects were expressed as a percentage of the control values. The effect of mechanical influence can be expressed both as an increase in the value of any arbitrarily chosen parameter (e.g., an increase in cell proliferation) and a decrease (a slowdown in cell proliferation). Since we are not interested in the direction but numerically in the magnitude of the treatment facilitation effect, we decided to take all calculated percentages modulo the magnitude. We will hereafter refer to the resulting values as “relative effect moduli” or simply “effects”.
To estimate quantitative dependencies, we applied the approach we used and described earlier [201]. To assess the influence of the combination of conditions, we constructed dot plots in 3 axes. Further, we will denote the axes under consideration in the following order: “ordinate axis—abscissa axis—z (colirbar) axis”.
We evaluated the dependence of the magnitude of the effects of mechanical influence. The magnitude of the moduli of the relative effects of mechanical influence was weakly dependent on the frequency, acceleration, and duration of the influence. The most obvious dependence was found for the “acceleration-duration-effect” dependence (Figure 3c,d). We can see a clear enough boundary of acceleration and duration values, below which the effects of mechanical influence cease to be realized, which confirms the observed “threshold” effect described above. For the set of “frequency-acceleration-effect” characteristics, the dependence takes an unobvious form (Figure 3a,b). The separation into regions of 0.001–1 Hz and above 1 Hz is visible. In the first region, the effects are registered relatively rarely. In the second region, most of the effects are registered. There is a general tendency for the magnitudes of mechanical effects to increase with increasing acceleration and/or frequency. However, the magnitudes of the effects depend on the values of acceleration and frequency, not monotonically. The region between frequencies of 10–100 Hz and accelerations of 10–50 m/s2, at which the magnitudes of the effect moduli are the largest, can be seen. In the investigated works, there are often additional factors that are difficult to take into account (sample mass, laboratory temperature, and other microclimate features). In order to minimize the contribution of extraneous factors, we have attempted to calculate the force of the mechanical effect using the standard formula:
F = m a ,
where F is the force acting on the sample, m and a are the mass and the acceleration of the sample, respectively (taken from published papers or calculated from published data).
In the case of studies where the mass of the sample was clearly described, the values given by the authors were used. If the volumes of aqueous solutions studied were described, the assumption was made that mV, since the density of water under standard conditions is 0.99707 g/cm3. If “whole body mass” was specified in the in vivo experiments, the average mass of the animals or volunteer subjects specified in the paper was taken as mass. If the study was performed on an isolated limb or tail of the animal, the literature data on the mass of the indicated body parts were used [202,203].
We found that the shape of the dependence of the magnitude of the effect of mechanical impact on the combination of force and duration (Figure 3d,e) is similar to the shape of the acceleration-duration-effect relationship, but the point cloud is more widely distributed.
The increase in the “spread” of the dots indicates the dependence of the magnitude of the mechanical action on the mass of the sample, which determines the acceleration threshold required to achieve a certain value of acceleration. It is noteworthy that we found inhomogeneously distributed maxima of the effect moduli values within the force ranges from 0.0001 to 1000 N and from 1 s to several days. The presence of well-defined regions of frequency and acceleration values at which the effects are observed is consistent with the data of mathematical modeling and the theory of potential energy surfaces [204]. The form of the double force-duration or acceleration-duration dependence is similar to the previously described dependence obtained during the action of shear stress on non-neurons and myoblasts [153,155].
At the next stage, we attempted to answer the question: are there any fundamental differences between the magnitudes of the effects of mechanical influence from its characteristics, estimated earlier, for living and nonliving systems? For this purpose, we divided the data we analyzed into two groups. For the division, we used the classical paradigm of biology that the cell is the basic structural and functional unit of all life forms, and the manifestation of all the properties of living things begins at the level of the whole cell [205]. In the first group, we referred to the literature data on non-living systems: purified water, buffer solutions, isolated proteins in aqueous solutions, solutions of monomers in water, etc. The results on living systems included data obtained on cell lines, animals, and humans (Figure 4).
The threshold limits for non-living systems are broader than for living systems (Figure 4a) and are shifted to lower frequencies and accelerations. This is partly due to methodological limitations that are imposed on biological studies. However, this may explain the use of higher frequencies, accelerations, and durations. Reducing the above parameters should not cause methodological difficulties and/or confirmation of living subjects. In addition, mechanical action needs a much longer exposition to achieve an effect when acting on living systems than non-living systems. The difference in action time is on average 10–15 times. This phenomenon can be explained by fundamental differences in the way systems respond. “Response” of non-living systems to mechanical impact is provided exclusively by physical and chemical processes (energy dissipation, free radical reactions, etc.). Whereas, the responses of living systems are often mediated by specific receptors that trigger intracellular signaling cascades of events with some delay (see above). Receptor-mediated responses can take from fractions of seconds to minutes, hours, or even longer. Responses of living systems are also characterized by significant time delays due to the speed of the processes regulated by the cell or organism. Protein synthesis requires time delays of more than an hour, cell division—days, and tissue regeneration and remodeling—days or weeks.
Such a difference in “sensitivity” can be explained by the presence of homeostasis in living systems: up to a certain threshold, living systems can compensate for the effects of mechanical influence. According to our evaluation, the limits of metabolic compensation of mechanical impact lie in the areas of frequencies up to 10 Hz and accelerations up to 1 m/s2 with durations not exceeding 100 s.
In the next step, we assessed the dependence of effect sizes on the combination of force and duration (Figure 4b). We also found a significant difference in the limits of force and duration values at which the effects are registered. For non-living systems, a force of 10−10 N or a duration of 0.01 s is sufficient for the development of effects. Whereas, for living systems, forces of at least 0.001 N and a duration of ~100 s are required. This difference can be explained by the ability of living systems to actively self-regulate and maintain homeostasis. With increasing force and/or duration, the effects of mechanical impact on nonliving systems grow relatively monotonically. The largest effect sizes (up to 2–4 orders of magnitude) for aqueous solutions are observed in the range of forces >10 N and durations of 100–1000 s. For living systems, the dependence of the magnitude of effects on force and duration is not obvious. We can detect a region of increased effect values between 0.001 and 10 N and 300 and 10,000 s, where the largest number of effects with the largest moduli are observed. However, the presence of this region may be due more to the convenience of studying cells and animals in the laboratory under these conditions. Thus, the effects of mechanical impact on non-living systems depend on a combination of factors, the leading ones being acceleration, force, and duration. In the case of non-living systems, there is an increase in the magnitude of the effect from the post force and/or duration. For living systems, a threshold mechanism of mechanical action close to the “all or nothing” principle and/or a complex character of dependence of the efficiency of mechanical action on its characteristics is most likely.

4.1.3. Correlations of Effects from Individual Parameters

At the next stage of the analysis, we tried to answer the question: which of the parameters of mechanical impact determines the magnitude of their effects to a greater extent? For this purpose, we evaluated the correlations between the magnitude of the effect of mechanical impact and its frequency, acceleration, force, and duration. Previously, we found that living and non-living systems respond differently to mechanical forcing, so in evaluating the correlations, we divided all evaluated responses into those of non-living (Figure 5) and living (Figure 6) systems. The strength, direction, and statistical significance of correlations were assessed using the Spearman criterion (Table 3).
We found that the magnitude of the effects of mechanical influence in the case of living systems depended on frequency, acceleration, and force. Significant correlations of the effect size with frequency (correlation strength 0.173 at p < 0.05), acceleration (correlation strength 0.253 at p < 0.001), and force (correlation 0.195, p < 0.05) were found. For non-living systems, the situation was somewhat different. Significant correlations were found between the magnitude of the effect size of mechanical impact and the frequency (p < 0.01), acceleration (p < 0.01), and duration (p < 0.001) of mechanical impact. However, the correlations found were of low strength: ~0.18 for the frequency-effect pair, ~0.22 for the acceleration-effect, and ~−0.26 for the duration-effect. Thus, mechanical effects with higher values of frequency and acceleration and the shortest durations should be most effective for water and aqueous solutions. It is noteworthy that living systems are more sensitive to acceleration and force than non-living systems. This can be explained by the ability of living systems to actively receive force (compression, stretching, shear stress) due to specialized receptor structures [128,129,130]. At the same time, nonliving systems are more sensitive to the duration of impact, which may indicate the contribution of mechanisms for maintaining homeostasis in living systems.

4.2. Dependence of the Magnitude of Ultrasound Effects on Its Characteristics

To assess the dependence of ultrasound (US) effects on its characteristics, we applied an approach similar to that described in Section 4.1.2 and Section 4.1.3. We chose the frequency, intensity, and duration of exposure as the main characteristics of US treatment (Figure 7).
The characteristics of ultrasound exposure used in most of the analysed works lie in the frequency ranges from 104 to 107 Hz, and intensity from 0.1 mW/cm2 to 100 W/cm2. The durations are from 10 s to several hours. The bulk of the effects with high values are concentrated in the frequency range from 0.1 to 1.0 MHz (Figure 7). Increasing both frequency and intensity enhances the effects of ultrasound (US) on living and non-living systems (Figure 7). The contribution of frequency appears to be more significant than that of power. Similarly, the contribution of duration appears to be more significant than that of intensity. A threshold duration of 1000 s is observed, after which the magnitude of the effects increases significantly, regardless of power.

Correlations of Ultrasound Effects on Individual Parameters

When assessing the correlations between the magnitudes of the effects of the US on living and non-living systems, we found no differences, so we calculated Spearman correlation coefficients for the whole sample of analyzed data (Figure 8 and Table 4).
We found statistically significant correlations between the strength of the effect of ultrasound exposure and the intensity (correlation strength 0.38, p = 3.60925 × 10−9) and duration of exposure (correlation strength −0.27719, p = 2.8004 × 10−5). The positive correlation between the magnitude of US effects and its power seems more logical and does not require further explanation. The absence of a significant correlation between the magnitude of the effects of ultrasound (US) and its frequency indicates that frequency plays a less pronounced role in regulating the magnitude of US effects. Additionally, the frequency range used is narrower (3.5 orders of magnitude) than the duration and intensity ranges (5 and 7 orders of magnitude, respectively). The observed negative relationship between US exposure duration and effects can be explained by the fact that shorter exposure times were investigated at higher intensities, and vice versa (Figure 7a).

5. Limitations and Prospects

As can be seen from the Appendix A, there are more than a dozen (vibrations, shocks, stirring in different modes, falling, etc.) fundamentally different protocols of mechanical influence on aqueous solutions and living systems. For each of the effects, there may be critical conditions for the realization of effects that are not important for the other effects. For example, frequency is not applicable to falling, and height above the floor is not important in vibration. The difference in methods causes a difference in the information provided by the authors (see Appendix A), so the task of combining and comparing data becomes more difficult.
We have attempted to “arrive at a common denominator” when evaluating heterogeneous mechanical effects in the form of acceleration or force calculations. This approach allowed us to discover several interesting regularities, but it is not without disadvantages. In particular, when calculating the acceleration, we made the assumption that the oscillations are sinusoidal. However, we cannot guarantee that in all the analyzed articles the oscillations were sinusoidal. When calculating the force, we also used approximate mass values, since the authors’ exact mass values may simply not have been measured.
However, the approach we used was sufficient to find a number of generalizable acceleration-duration-effect and force-duration-effect patterns, and we also found differences between the responses to mechanical actions of living and non-living systems.
Two approaches can be suggested as future directions for research on this topic. The first is to deepen the analysis of already available data with the involvement of new methods of data analysis, for example, the principal component method, and the calculation of multiple correlation coefficients.
When planning future experiments, we would recommend limiting the number of measured effects in favor of expanding the number of exposure options. It is obligatory to perform adequate sham controls at all stages of the experiment [206]. In addition, when planning experiments (including those in other fields, not only mechanobiology) and interpreting the obtained data, it is necessary to take into account the possible effects of vibration and other influences on experimental samples, especially in vivo experiments.
The processes of cellular mechanotransduction are fundamentally different from those of solution mechanochemistry. In particular, actin-dependent transduction of mechanical signals in the cell is 40 to 50 times faster than diffusion of chemical reaction products [207,208]. However, as follows from the cited articles, mechanical action can affect the conformational states of proteins and enzymatic action in aqueous solution [91,92,209]. We proceed from the position that protein mechanochemistry, albeit altered, also occurs in the cytosol as in aqueous solutions, independent of signaling systems. Consequently, in a broad sense, the response of living systems to mechanical action will be the sum of aqueous solution mechanochemistry and mechanotransduction at the cellular level. Therefore, these processes are discussed together in this review.
We consider the use of mechanical effects for stem cell differentiation, more detailed in vivo studies, catalysis of chemical reactions in solutions, and nanobubbles as promising directions for future research.

6. Conclusions

Despite the diversity of mechanical effects, their characteristics can be summarised into a few key ones: frequency, acceleration, force, and duration. The magnitude of the effects of mechanical actions on living and non-living systems depends differently on the type of system affected. For non-living systems, there is a significant dependence of the response on the magnitude of acceleration, frequency, and duration of the impact. Responses of living systems depend on the magnitudes of acceleration, frequency, and force, but not duration. The magnitude of the effect changes insignificantly with the growth of these parameters of mechanical influence. Ultrasonic action depends on intensity and duration, both in aqueous solutions and in living organisms. Aqueous solutions require lower values of frequency and acceleration during mechanical action than living systems to initiate a ‘response’. This is due to the ability of living systems to maintain homeostasis. Notably, living systems are more sensitive to acceleration and force than non-living systems. This can be explained by the living system’s ability to actively receive force (compression, stretching, and shear stress) due to specialized receptor structures. At the same time, non-living systems are more sensitive to the duration of impact. These differences may indicate the contribution of mechanisms of homeostasis maintenance in living systems.

Author Contributions

Conceptualization, S.V.G., D.A.S. and I.A.S.; methodology, V.I.P. and A.V.S.; software, R.M.S.; formal analysis, D.A.S., R.M.S., A.V.S. and V.I.P.; writing—original draft preparation, D.A.S.; writing—review and editing, S.V.G. and I.A.S.; visualization, D.A.S. and R.M.S.; supervision, I.A.S.; funding acquisition, A.V.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation grant number 24-22-00363.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All collected and calculated data are provided in the table Appendix A.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ACANAggrecan
AJAdherens junctions
AktRAC-alpha serine/threonine-protein kinase
ALPAlkaline phosphatase
ASICsAcid-sensing ion channels
BMDCsBone Marrow-Derived Dendritic Cell
BrdU5-bromo-2′-deoxyuridine
BSAbovine serum albumin
CDCluster of differentiation
CenpfCentromere protein F
COL2A1collagen II type gene
CPKcreatine phosphokinase
CYP1B1cytochrome P450 1B1 gene
DKK1Dickkopf-related protein 1
DLSDynamic light scattering
DMEMDulbecco’s Modified Eagle Medium
DNdouble-network
E.G7-OVAT lymphoblast cancer cell line
EGFEpidermal Growth Factor
EMRElectromagnetic radiation
EREndoplasmic reticulum
ERKExtracellular signal-regulated kinase
fMLFN-Formylmethionine-leucyl-phenylalanine
GPCRsG-protein coupled receptors
hAD-MSCHuman amnion-derived mesenchymal stem cells
hESCHuman embryonic stem cells
HEWLHen egg white lysozyme
Hk2Hexokinase II gene
hPDLSCHuman periodontal ligament stem cells
IDOIndoleamine 2,3-dioxygenase
IFNInterferon
IGF-Iinsulin-like growth factor 1
IgGImmunoglobulin G
ILInterleukin
LINCLinker of Nucleoskeleton and Cytoskeleton
MAPKMitogen-activated protein kinase
MMP2Matrix metalloproteinase-2
MSCsMesenchymal stem cells
MT-1aMetallothionein 1A
NFATNuclear factor of activated T-cells
OCNOsteocalcin
OPGOsteoprotegerin
OSXOsterix protein
PAAmpolyacrylamide
Pcdh17Protocadherin 17
PEGpoly(ethylene glycol)
PGE-2Prostaglandin E 2
PHEMApoly(2-hydroxyethyl methacrylate)
PI3KPhosphoinositide 3-kinases
PLA2phospholipase A2
PNaAMPSpoly(2-acrylamido-2-methylpropanesulfonic acid sodium salt)
PVSpoly(vinyl alcohol)
RANKLReceptor activator of nuclear factor kappa-B ligand
ROSReactive oxygen species
RPMIRoswell Park Memorial Institute
Runx2Runt-related transcription factor 2
SACsstress-activated channels
Slc2a1major glucose transporter GLUT1 gene
SOXSuperoxide dismutase
TGFTransforming growth factor
TNFTumor necrosis factor
TREKTwo-Pore Domain K+ channels
TRPTransient receptor potential channels
USUltrasound
VEGFVascular Endothelial Growth Factor

Appendix A

Table A1. Mechanical impact on water solutions and living systems.
Table A1. Mechanical impact on water solutions and living systems.
#SolutionObjectImpact TypeConcetration (If Applied)Frequency. HzAmplitude (Value)Amplitude (Units)Mass. gTime. sMesured CharacteriscticEffect. %Ref.
1Culture medium DMEMhuman fibroblastsUltrasound-45,00015mW/cm25300Cells proliferation+30%[102]
2Culture medium DMEMhuman fibroblastsUltrasound-45,00050mW/cm25300Cells proliferation+45%[102]
3Culture medium DMEMhuman fibroblastsUltrasound-1,000,0000.7mW/cm25300Cells proliferation+45%[102]
4Culture medium DMEMhuman fibroblastsUltrasound-1,000,0001mW/cm25300Cells proliferation+55%[102]
5Culture medium DMEMhuman fibroblastsUltrasound-45,00015mW/cm25300Collagen expression+45%[102]
6Culture medium DMEMhuman fibroblastsUltrasound-45,00050mW/cm25300Collagen expression+38%[102]
7Culture medium DMEMhuman osteoblastsUltrasound-45,0005mW/cm25300Cells proliferation+32%[102]
8Culture medium DMEMhuman fibroblastsUltrasound-1,000,0000.1mW/cm25300Collagen expression+49%[102]
9Culture medium DMEMhuman fibroblastsUltrasound-1,000,0000.4mW/cm25300Collagen expression+58%[102]
10Culture medium DMEMhuman fibroblastsUltrasound-1,000,0000.7mW/cm25300Collagen expression+53%[102]
11Culture medium DMEMhuman osteoblastsUltrasound-45,00030mW/cm25300Cells proliferation+36%[102]
12Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.7mW/cm25300Cells proliferation+46%[102]
13Culture medium DMEMhuman osteoblastsUltrasound-1,000,0001mW/cm25300Cells proliferation+38%[102]
14Culture medium DMEMhuman osteoblastsUltrasound 45,00015mW/cm25300Collagen expression+45%[102]
15Culture medium DMEMhuman osteoblastsUltrasound-45,00050mW/cm25300Collagen expression+37%[102]
16Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.1mW/cm25300Collagen expression+53%[102]
17Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.4mW/cm25300Collagen expression+38%[102]
18Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.4mW/cm2575Secretion of IL-1β+16[102]
19Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.7mW/cm2575Secretion of IL-1β+20%[102]
20Culture medium DMEMhuman osteoblastsUltrasound-1,000,0001mW/cm2575Secretion of IL-1β+17%[102]
21Culture medium DMEMhuman osteoblastsUltrasound-450,00030mW/cm25300Secretion of IL-8+55%[102]
22Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.4mW/cm25300Secretion of IL-8+60%[102]
23Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.7mW/cm25300Secretion of IL-8+70%[102]
24Culture medium DMEMhuman osteoblastsUltrasound-45,00005mW/cm25300Secretion of EGFb+3000%[102]
25Culture medium DMEMhuman osteoblastsUltrasound-45,000015mW/cm25300Secretion of EGFb+1500%[102]
26Culture medium DMEMhuman osteoblastsUltrasound-45,000050mW/cm25300Secretion of EGFb+750%[102]
27Culture medium DMEMhuman osteoblastsUltrasound-1,000,0000.1mW/cm25300Secretion of EGFb+300%[102]
28Culture medium DMEMhuman osteoblastsUltrasound 450,00015mW/cm25300Secretion of VEGF+25%[102]
29Culture medium DMEMhuman osteoblastsUltrasound 1,000,0000.1mW/cm25300Secretion of VEGF+40%[102]
30Culture medium RPMI 1640human monocytesUltrasound-45,00015mW/cm25300Secretion of IL-1β+25%[102]
31Culture medium RPMI 1640human monocytesUltrasound-45,00030mW/cm25300Secretion of IL-1β+103%[102]
32Culture medium RPMI 1640human monocytesUltrasound-45,00050mW/cm25300Secretion of IL-1β+80%[102]
33Culture medium RPMI 1640human monocytesUltrasound-45,00015mW/cm25300Secretion of VEGF+25%[102]
34Culture medium RPMI 1640human monocytesUltrasound-45,00030mW/cm25300Secretion of VEGF+30%[102]
35Culture medium RPMI 1640human monocytesUltrasound-45,00050mW/cm25300Secretion of VEGF+35%[102]
36Culture medium RPMI 1640human monocytesUltrasound-1,000,0000.1mW/cm25300Secretion of VEGF+35%[102]
37Culture medium RPMI 1640human monocytesUltrasound-1,000,0000.4mW/cm25300Secretion of VEGF+15%[102]
38Culture medium RPMI 1640human monocytesUltrasound-1,000,0001mW/cm25300Secretion of VEGF+10%[102]
39Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm22600Cells differentiation+10%[103]
40Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm22600Cells differentiation+15%[103]
41Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm22600Cells differentiation+15%[103]
42Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00060mW/cm221200Cells differentiation+15%[103]
43Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm221200Cells differentiation+15%[103]
44Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm221200Cells differentiation+10%[103]
45Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm221200Cells differentiation+11%[103]
46Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00030mW/cm221800Cells differentiation+15%[103]
47Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm22600Concentration of Cyclin D1/β+60%[103]
48Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm22600Concentration of Cyclin D1/β+40%[103]
49Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm22600Concentration of Cyclin D1/β+45%[103]
50Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00060mW/cm221200Concentration of Cyclin D1/β+100%[103]
51Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm221200Concentration of Cyclin D1/β+90%[103]
52Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm221200Concentration of Cyclin D1/β+80%[103]
53Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm221200Concentration of Cyclin D1/β+70%[103]
54Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00030mW/cm221800Concentration of Cyclin D1/β+110%[103]
55Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm22600Concentration of Cyclin В1/β+200%[103]
56Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm22600Concentration of Cyclin В1/β+380%[103]
57Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm22600Concentration of Cyclin В1/β+500%[103]
58Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00060mW/cm221200Concentration of Cyclin В1/β+500%[103]
59Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm221200Concentration of Cyclin В1/β+400%[103]
60Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm221200Concentration of Cyclin В1/β+530%[103]
61Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm221200Concentration of Cyclin В1/β+490%[103]
62Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00030mW/cm221800Concentration of Cyclin В1/β+490%[103]
63Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm22600Concentration of Cyclin Е1/β+80%[103]
64Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm22600Concentration of Cyclin Е1/β+110%[103]
65Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm22600Concentration of Cyclin Е1/β+150%[103]
66Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00060mW/cm221200Concentration of Cyclin Е1/β+200%[103]
67Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm221200Concentration of Cyclin Е1/β+250%[103]
68Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm221200Concentration of Cyclin Е1/β+200%[103]
69Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm221200Concentration of Cyclin Е1/β+180%[103]
70Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00030mW/cm221800Concentration of Cyclin Е1/β+190%[103]
71Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm22600Concentration of Cyclin A1/β+250%[103]
72Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm22600Concentration of Cyclin A1/β+100%[103]
73Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm22600Concentration of Cyclin A1/β+90%[103]
74Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00060mW/cm221200Concentration of Cyclin A1/β+250%[103]
75Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000100mW/cm221200Concentration of Cyclin A1/β+245%[103]
76Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000200mW/cm221200Concentration of Cyclin A1/β+240%[103]
77Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,000300mW/cm221200Concentration of Cyclin A1/β+70%[103]
78Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00030mW/cm221800Concentration of Cyclin A1/β+270%[103]
79Culture medium DMEMHuman amnion-derived mesenchymal stem cells (hAD-MSC)Ultrasound-250,00030mW/cm221800Cells proliferation+180%[103]
80Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-500,000500mW/cm2186,400Cells proliferation+90%[104]
81Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,000,000500mW/cm2186,400Cells proliferation+95%[104]
82Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-3,000,000500mW/cm2186,400Cells proliferation+150%[104]
83Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-3,000,0001000mW/cm2186,400Cells proliferation+180%[104]
84Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-3,000,000250mW/cm2186,400Cells proliferation+145%[104]
85Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,000,000500mW/cm212400myotube width+20%[104]
86Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,000,000500mW/cm212400myotube length+50%[104]
87Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-5,000,000500mW/cm212400myotube length+10%[104]
88Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,000,000500mW/cm212400fusion index+100%[104]
89Rat (trauma model)in vivoUltrasound-870,0001000mW/cm22501500Total tissue protein content−10%[97]
90Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm2107200Cells proliferation+18%[113]
91Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm2109600Cells proliferation+12%[113]
92Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm2107200Expression of myogenin+47%[113]
93Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm2109600Expression of myogenin+34%[113]
94Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21016,800Numbers of regenerating
myofibers
+90%[113]
95Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21025,200Numbers of regenerating
myofibers
+250%[113]
96Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21033,600Numbers of regenerating
myofibers
+320%[113]
97Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21025,200Restoration of fast twitch strength+150%[113]
98Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21033,600Restoration of fast twitch strength+300%[113]
99Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21025200Restoration of tetanus strength strength+90%[113]
100Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-1,500,00030mW/cm21033600Restoration of fast twitch strength+100%[113]
101in vivoWistar ratsUltrasound-1,000,0001mW/cm2220900Muscle fibre diameter+10%[98]
102in vivoWistar ratsUltrasound-1,000,0001mW/cm2220900BrdU positive cells count+30%[98]
103Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-3,000,00030mW/cm25900Pax7 positive cells count+60%[115]
104Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-3,000,00030mW/cm25900Expression of gene COX2+80%[115]
105Culture medium DMEMMouse myoblasts (C2C12)Ultrasound-3,000,00030mW/cm25900Proiflammatory cells count−40%[115]
106Culture medium DMEMBMDCsConstant pressure from below-050,00050,000Pa5950,400Efficiency of glycolysis−30%[72]
107Culture medium DMEMBMDCsConstant pressure from below-02000Pa5950,400Efficiency of glycolysis−50%[72]
108Culture medium DMEMBMDCsConstant pressure from below-050,000Pa5950,400Lactase activity−5%[72]
109Culture medium DMEMBMDCsConstant pressure from below-02000Pa5950,400Lactase activity−60%[72]
110Culture medium DMEMBMDCsConstant pressure from below-050,000Pa5950,400overall rate curves of oxygen
consumption rate
−70%[72]
111Culture medium DMEMBMDCsConstant pressure from below-02000Pa5950,400overall rate curves of oxygen
consumption rate
−80%[72]
112Culture medium DMEMBMDCsConstant pressure from below-050,000Pa5950,400Expression of gene Hk2−55%[72]
113Culture medium DMEMBMDCsConstant pressure from below-02000Pa5950,400Expression of gene Slc2a1−96%[72]
114Culture medium DMEMBMDCsConstant pressure from below-02000Pa5950,400Cells proliferation−95%[72]
115Culture medium DMEMBMDCsConstant pressure from below-050,000Pa5950,400Tumor size (11 days post E.G7-OVA injection) in WT mice injected with E.G7-OVA cells and WT BMDCs−50%[72]
116Culture medium DMEMBMDCsConstant pressure from below-050,000Pa5950,400tumor weight (14 days post E.G7-OVA injection) in WT mice injected with E.G7-OVA cells and WT
BMDCs
−85%[72]
117Culture medium DMEMCD8+ T cellsConstant pressure from below-050,000Pa5950,400Immunology memory+120%[72]
118Culture medium DMEMCD4+ T cellsConstant pressure from below-050,000Pa5950,400Immunology memory+110%[72]
119WaterWaterShaking-3012mm1060Chemiluminescence intensity−20%[21]
120WaterWaterShaking-302.3mm1060Chemiluminescence intensity+25%[21]
121WaterWaterShaking-1012mm1015Decline rate of chemiluminescence intensity−20%[21]
122WaterWaterShaking-1012mm1060Decline rate of chemiluminescence intensity−25%[21]
123WaterWaterShaking-302.3mm1060Decline rate of chemiluminescence intensity+30%[21]
124WaterWaterShaking-302.3mm1060Standard deviation of chemiluminescence intensity+10%[21]
125WaterBSAShaking1 mg/mL302.3mm1015Chemiluminescence intensity−20%[21]
126WaterBSAShaking1 mg/mL1012mm1015Chemiluminescence intensity−15%[21]
127WaterBSAShaking1 mg/mL1012mm1060Chemiluminescence intensity−10%[21]
128WaterBSAShaking1 mg/mL302.3mm1015Chemiluminescence intensity−8%[21]
129WaterBSAShaking1 mg/mL302.3mm1060Chemiluminescence intensity−20%[21]
130WaterBSAShaking1 mg/mL302.3mm1015Chemiluminescence intensity−12%[21]
131WaterBSAShaking1 mg/mL302.3mm1060Chemiluminescence intensity−20%[21]
132WaterBSAShaking1 mg/mL102.3mm1015Decline rate of chemiluminescence intensity−18%[21]
133WaterBSAShaking1 mg/mL1012mm1015Decline rate of chemiluminescence intensity−17%[21]
134WaterBSAShaking1 mg/mL1012mm1060Decline rate of chemiluminescence intensity−16%[21]
135WaterBSAShaking1 mg/mL302.3mm1015Decline rate of chemiluminescence intensity−25%[21]
136WaterBSAShaking1 mg/mL302.3mm1060Decline rate of chemiluminescence intensity−16%[21]
137WaterBSAShaking1 mg/mL302.3mm1060Standard deviation of chemiluminescence intensity−10%[21]
138WaterHuman IgGAlternating magnetic field1 mg/mL5050мкТл10300Chemiluminescence intensity+10%[21]
139WaterHuman IgGAlternating magnetic field1 mg/mL850мкТл10300Standard deviation of chemiluminescence intensity+5%[21]
140Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL12.60.01мкТл 5400fMLF-induced ROS generation+45%[32]
141Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL12.60.05мкТл 5400fMLF-induced ROS generation+45%[32]
142WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150local connected fractal dimension−27%[22]
143WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150local connected fractal dimension+3%[22]
144WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150mass fractal dimension−9%[22]
145WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150mass fractal dimension+2%[22]
146WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150lacunarity+37%[22]
147WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150lacunarity+4%[22]
148WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150ascending second moment+30%[22]
149WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150ascending second moment−15%[22]
150WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150contrast−30%[22]
151WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150contrast−12%[22]
152WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150Correlation+30%[22]
153WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150Correlation+90%[22]
154WaterViscum album Quercus L. 3. plant extractTurbulent mixing0.1% (10−3)102sm10150Entropy−3%[22]
155WaterViscum album Quercus L. 3. plant extractCircular mixing0.1% (10−3)102sm10150Entropy+2%[22]
156Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL000 3600fMLF-induced ROS production primered by PMA−21%[33]
157Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL12.60.1мкТл 3600fMLF-induced ROS production primered by PMA+67%[33]
158Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL48.50.1мкТл 3600fMLF-induced ROS production primered by PMA+635%[33]
159Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL12.60.1мкТл 2400fMLF-induced ROS generation+36%[34]
160Hanks’ balanced salts solutionMurine neutrophils (Male Balb/c mice weighing 22–25 g)Alternating magnetic field2.5 × 105 cells/mL48.50.1мкТл 2400fMLF-induced ROS generation−19%[34]
161WaterIFNγCircular mixing0.1 М501sm1050Conductivity of the solution+20%[16]
162Waterpolyclonal antibodies to human IFNγCircular mixing26 μg/mL501sm1050Solurion EMR+150%[17]
163WaterLactoseStrokes820 nM21300 (20)g (CM)241050Luminescence at 550 nm+100%[24]
164WaterLactoseStrokes820 нМ21300 (20)g (CM)24525Luminescence at 550 nm+200%[24]
16520 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL025.6sm10.025Particle counts (protein aggregation?)+150%[91]
16620 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL051.2sm10.025Particle counts+300%[91]
16720 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL076.8sm10.025Particle counts+1000%[91]
16820 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL0102.4sm10.025Particle counts+350%[91]
16920 mM histidine buffer.antistreptavidin IgG1Downfall35 mg/mL025.6sm10.025Particle counts+900%[91]
17020 mM histidine buffer.antistreptavidin IgG1Downfall35 mg/mL051.2sm10.025Particle counts+1100%[91]
17120 mM histidine buffer.antistreptavidin IgG1Downfall35 mg/mL076.8sm10.025Particle counts+1000%[91]
17220 mM histidine buffer.human growth hormoneDownfall35 mg/mL0102.4sm10.025Particle counts+800%[91]
17320 mM histidine bufferhuman growth hormoneDownfall1.75 mg/mL0102.4sm10.025Particle counts+60%[91]
17420 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL025.6sm10.025recovered from the walls of dropped vials following treatment with urea+20%[91]
17520 mM histidine bufferantistreptavidin IgG1Downfall1 mg/mL076.8sm10.025recovered from the walls of dropped vials following treatment with urea+110%[91]
17620 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL0102.4sm10.025recovered from the walls of dropped vials following treatment with urea+180%[91]
17720 mM histidine bufferantistreptavidin IgG1Downfall35 mg/mL051.2sm10.025recovered from the walls of dropped vials following treatment with urea+60%[91]
17820 mM histidine buffer.antistreptavidin IgG1Downfall35 mg/mL076.8sm10.025recovered from the walls of dropped vials following treatment with urea+120%[91]
17920 mM histidine bufferantistreptavidin IgG1Downfall35 mg/mL0102.4sm10.025recovered from the walls of dropped vials following treatment with urea+90%[91]
18020 mM histidine bufferantistreptavidin IgG1Downfall1 mg/mL025.6sm10.025recovered from the walls of dropped vials following treatment with guanidine hydrochloride+35%[91]
18120 mM histidine buffer.antistreptavidin IgG1Downfall1 mg/mL076.8sm10.025recovered from the walls of dropped vials following treatment with guanidine hydrochloride+30%[91]
18220 mM histidine bufferantistreptavidin IgG1Downfall35 mg/mL051.2sm10.025recovered from the walls of dropped vials following treatment with guanidine hydrochloride+200%[91]
183WaterEchinaceasuccussion strokes0.8 × 10−2 g/mL120sm0.05100Characteristics of crystal shape after evaporation+10%[23]
184WaterEchinaceasuccussion strokes0.8 × 10−2 g/mL120sm0.0510Characteristics of crystal shape after evaporation+8%[23]
185WaterBaptisiasuccussion strokes0.8 × 10−3 g/mL120sm0.05100Characteristics of crystal shape after evaporation−10%[23]
186WaterBaptisiasuccussion strokes0.8 × 10−3 g/mL120sm0.0510Characteristics of crystal shape after evaporation−10%[23]
187WaterBaptisiasuccussion strokes0.8 × 10−4 g/mL120sm0.05100Characteristics of crystal shape after evaporation+60%[23]
188WaterBaptisiasuccussion strokes0.8 × 10−4 g/mL120sm0.0510Characteristics of crystal shape after evaporation+50%[23]
189WaterLuffasuccussion strokes0.8 × 10−4 g/mL120sm0.05100Characteristics of crystal shape after evaporation−20%[23]
190WaterLuffasuccussion strokes0.8 × 10−4 g/mL120sm0.0510Characteristics of crystal shape after evaporation−20%[23]
191WaterSpongiasuccussion strokes0.8 × 10−6 g/mL120sm0.05100Characteristics of crystal shape after evaporation+5%[23]
192WaterSpongiasuccussion strokes0.8 × 10−6 g/mL120sm0.0510Characteristics of crystal shape after evaporation+10%[23]
193WaterIgGVertical shakingIgG molecules 3 × 1012 cm−3510mm2030DLS peak intensity ~200 nm (bubble count)+300%[6]
194Water +36.7% ethanolIgGVertical shakingIgG molecules 3 × 1012 cm−3510mm2030DLS peak intensity ~200 nm (bubble count)+20%[6]
195Water-Shaking-305mm10300Concentration of molecular oxigen−4%[80]
196Water-Shaking-305mm10600Concentration of molecular oxigen−5%[80]
197Water-Shaking-305mm10900Concentration of molecular oxigen−7%[80]
198Water-Shaking-305mm101800Concentration of molecular oxigen−8%[80]
199Water-Shaking-305mm102700Concentration of molecular oxigen−9%[80]
200Water-Shaking-305mm10600Water temperature+2%[80]
201Water-Shaking-305mm10900Water temperature+15%[80]
202Water-Shaking-305mm101800Water temperature+20%[80]
203Water-Shaking-305mm102700Water temperature+24%[80]
204Water-Shaking-155mm10300Concentration of molecular oxigen−10%[80]
205Water-Shaking-455mm10300Concentration of molecular oxigen−13%[80]
206Water-Shaking-605mm10300Concentration of molecular oxigen−15%[80]
207Water-Shaking-305mm10300light scattering intensity+50%[80]
208Water-Shaking-305mm10900pH+12%[80]
209Water-Shaking-305mm101800pH+16%[80]
210Water-Shaking-305mm102700pH+17%[80]
211Water-Shaking-455mm10300pH+6%[80]
212Water-Shaking-605mm10300pH+8%[80]
213Water-Shaking-155mm10300Concentration of H2O2+2000%[80]
214Water-Shaking-305mm10300Concentration of H2O2+5000%[80]
215Water-Shaking-455mm10300Concentration of H2O2+14000%[80]
216Water-Shaking-305mm1060Concentration of H2O2+1000%[80]
217Water-Shaking-305mm10150Concentration of H2O2+2500%[80]
218Water-Shaking-305mm10210Concentration of H2O2+3500%[80]
219Water-Shaking-305mm10300Concentration of H2O2+4500%[80]
220Water-Shaking-305mm10180Concentration of OH-radicals+1500%[80]
221Water-Shaking-305mm10300Concentration of OH-radicals+2500%[80]
222Water-Shaking-305mm10420Concentration of OH-radicals+4000%[80]
223Water-Shaking-305mm10600Concentration of OH-radicals+4300%[80]
224Water-Shaking-155mm10300Concentration of OH-radicals+900%[80]
225Water-Shaking-455mm10300Concentration of OH-radicals+5000%[80]
226Water-Shaking-605mm10300Concentration of OH-radicals+9000%[80]
227WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.01160Concentration of H2O2+50[90]
228WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.011120Concentration of H2O2+50%[90]
229WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.011180Concentration of H2O2+100%[90]
230WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.011240Concentration of H2O2+500%[90]
231WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.011300Concentration of H2O2+600%[90]
232WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.011420Concentration of H2O2+1000%[90]
233WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.0111800Concentration of H2O2−15%[90]
234WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.0113600Concentration of H2O2−35%[90]
235WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.01114,400Concentration of H2O2−45%[90]
236WaterO2 gasCreating an aerosol with N2-Constant5μL/min0.01121,600Concentration of H2O2−50%[90]
237Water-Mixing-121sm10500Absorption at 180–220 sm−1+30%[82]
238Water-Mixing-121sm1050Wavelength of maximum absorption+80% (280–340 нм)[83]
239Water-Mixing-121sm1050Wavelength of maximum absorption+18% (280–330 нм)[83]
240Water-Mixing-121sm1050Fluorescence intensity+400%[83]
240Water-Vibration 1,000,0001sm10500Redox potentials ΔE+1 mV (+100%)[89]
240Water coated by oil layes 10 mm Vibration 1,000,0001sm10500Redox potentials ΔE+2.5 mV (+250%)[89]
241WaterAcEu solutionsMicrofluidic mixing-Constant400
35.5–250.6
Reynolds number1030Position of the IR band maximum of the OH groups’ stretching vibrations on Raman spectrum+0.1%[94]
242WaterAcEu solutionsMicrofluidic mixing-Constant400
35.5–250.6
Reynolds number10250Position of the IR band maximum of the OH groups’ stretching vibrations on Raman spectrum0.3[94]
24325 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL1099мкм0.15172,800Protein crystallisation success−25%[92]
24425 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL1594.2мкм0.15172,800Protein crystallisation success−25%[92]
24525 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL3060мкм0.15172,800Protein crystallisation success−25%[92]
24625 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL5034.8мкм0.15172,800Protein crystallisation success−25%[92]
24725 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL604мкм0.15172,800Protein crystallisation success−25%[92]
24825 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL702.4мкм0.15172,800Protein crystallisation success−25%[92]
24925 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL800.85мкм0.15172,800Protein crystallisation success−25%[92]
25025 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL900.6мкм0.15172,800Protein crystallisation success−25%[92]
25125 mM 4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid
sodium buffer (pH 7.0)
HEWLShaking12.5 mg/mL1000.4мкм0.15172,800Protein crystallisation success−25%[92]
2520.1 M acetate buffer pH 4.5HEWLUltrasound25 mg/mL100,000100,000mW/cm29.440Number of protein crystals+150%[93]
2530.1 M acetate buffer pH 4.5HEWLUltrasound25 mg/mL100,000100,000mW/cm29.470Number of protein crystals+600%[93]
2540.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL1.620mm0.53600Concentration of protein after precipitation in 100 mg/mL NaCl−5%[209]
2550.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL2.520mm0.53600Concentration of protein after precipitation in 100 mg/mL NaCl−10%[209]
2560.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL3.320mm0.53600Concentration of protein after precipitation in 100 mg/mL NaCl−15%[209]
2570.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL1.620mm0.54800Concentration of protein after precipitation in 100 mg/mL NaCl−10%[209]
2580.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL2.520mm0.54800Concentration of protein after precipitation in 100 mg/mL NaCl−20%[209]
2590.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL3.320mm0.54800Concentration of protein after precipitation in 100 mg/mL NaCl−40%[209]
2600.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL1.620mm0.510,800Concentration of protein after precipitation in 100 mg/mL NaCl−35%[209]
2610.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL2.520mm0.510,800Concentration of protein after precipitation in 100 mg/mL NaCl−40%[209]
2620.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL3.320mm0.510,800Concentration of protein after precipitation in 100 mg/mL NaCl−50%[209]
2630.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL1.620mm0.51800Concentration of protein after precipitation in 100 mg/mL NaCl−5%[209]
2640.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL2.520mm0.51800Concentration of protein after precipitation in 100 mg/mL NaCl−5%[209]
2650.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL3.320mm0.51800Concentration of protein after precipitation in 100 mg/mL NaCl−60%[209]
2660.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL1.620mm0.53600Concentration of protein after precipitation in 100 mg/mL NaCl−50%[209]
2670.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL2.520mm0.53600Concentration of protein after precipitation in 100 mg/mL NaCl−70%[209]
2680.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL3.320mm0.53600Concentration of protein after precipitation in 100 mg/mL NaCl−80%[209]
2690.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL1.620mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−75%[209]
2700.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL2.520mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−78%[209]
2710.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations90 mg/mL3.320mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−80%[209]
2720.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL0.120mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−5%[209]
2730.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL0.220mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−10%[209]
2740.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL0.520mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−45%[209]
2750.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL120mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−75%[209]
2760.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL0.55mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−10%[209]
2770.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL0.510mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−20%[209]
2780.1 M acetate buffer pH 4.8HEWL 40,000 U/mgFlow rate oscillations50 mg/mL0.530mm0.55400Concentration of protein after precipitation in 100 mg/mL NaCl−80%[209]
279Water1 mmol (110 μL) benzaldehyde was with 1 mmol (145 μL) aminoacetaldehyde diethylAtomization (formations of nanobubbles in the flow N2)1 mmol030mL/min0.000000020.01Pomeranz–Fritsch reaction rate+588%[29]
280Methanol1 mmol (110 μL) benzaldehyde was with 1 mmol (145 μL) aminoacetaldehyde diethylAtomization (formations of nanobubbles in the flow N2)1 mmol030mL/min0.000000020.01Pomeranz–Fritsch reaction rate+644%[29]
281ACN/DMF (1:1; v/v)1 mmol (110 μL) benzaldehyde was with 1 mmol (145 μL) aminoacetaldehyde diethylAtomization (formations of nanobubbles in the flow N2)1 mmol030mL/min0.000000020.01Pomeranz–Fritsch reaction rate+233%[29]
2821% (v/v) m-NBA in water1 mmol (110 μL) benzaldehyde was with 1 mmol (145 μL) aminoacetaldehyde diethylAtomization (formations of nanobubbles in the flow N2)1 mmol030mL/min0.000000020.01Pomeranz–Fritsch reaction rate+736%[29]
283WaterBi3TiNbO9Ultrasoundnot applicable. solid catalyst40,00050,000mW/cmm21018,000H2O2 production+40,705%
[126]
284Water-Shaking-301sm1060H2O2 production+100%[84]
285Water-Shaking-301sm10150H2O2 production+200%[84]
286Water-Shaking-301sm10210H2O2 production+280%[84]
287Water-Shaking-301sm10300H2O2 production+300%[84]
288Water-Shaking-300.5sm1060H2O2 production+20%[86]
289Water-Shaking-300.5sm10150H2O2 production+50%[86]
290Water-Shaking-300.5sm10210H2O2 production+60%[86]
291Water-Shaking-300.5sm10300H2O2 production+80%[86]
292Water-Shaking (turbulent)-151sm10300H2O2 production+150%[85]
293Water-Shaking (turbulent)-301sm10300H2O2 production+250%[85]
294Water-Shaking (turbulent)-601sm10300H2O2 production+1500%[85]
295Water-Shaking (turbulent)-301sm10900[O2] concentration+100%[85]
296Water-Shaking (turbulent)-301sm101500[O2] concentration+100%[85]
297Water-Shaking (turbulent)-301sm101800[O2] concentration+150%[85]
298Water-Shaking (turbulent)-301sm102400[O2] concentration+160%[85]
299Water-Shaking (turbulent)-301sm102700[O2] concentration+200%[85]
300Water-Shaking (turbulent)-301sm104200[O2] concentration+200%[85]
301Water-Shaking (laminar)-301sm10900[O2] concentration+10%[85]
302Water-Shaking (laminar)-301sm101500[O2] concentration+10%[85]
303Water-Shaking (laminar)-301sm101800[O2] concentration+20%[85]
304Water-Shaking (laminar)-301sm102400[O2] concentration+20%[85]
305Water-Shaking (laminar)-301sm102700[O2] concentration+25%[85]
306Water-Shaking (laminar)-301sm104200[O2] concentration+25%[85]
307Hedysarum laeve seedlings-Shaking 45°-26.4m/sm2160Primary stem length−15%[55]
308Hedysarum laeve seedlings-Shaking 45°-26.4m/sm2160Leaf numbe−15%[55]
309Hedysarum laeve seedlings-Shaking 45°-26.4m/sm2160Basal diameter−12%[55]
310Hedysarum laeve seedlings-Shaking 45°-26.4m/sm2160Total biomass−30%[55]
311Water-Downfall of drops001m0.050.1[H2O2] concentration+300%[27]
312Water-Downfall of drops002m0.050.1[H2O2] concentration+550%[27]
313Water-Downfall of drops003m0.050.1[H2O2] concentration+750%[27]
314Water-Downfall of drops004m0.050.1[H2O2] concentration+900%[27]
315Water-Downfall of drops001m0.050.1[OH] concentration+200%[27]
316Water-Downfall of drops002m0.050.1[OH] concentration+350%[27]
317Water-Downfall of drops003m0.050.1[OH] concentration+500%[27]
318Water-Downfall of drops004m0.050.1[OH] concentration+550%[27]
319WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.07Value of the first Lame parameter λ201800Normalised width w(t)/w(0)+2%[95]
321WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.14Value of the first Lame parameter λ201800Normalised width w(t)/w(0)+2%[95]
322WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.21Value of the first Lame parameter λ201800Normalised width w(t)/w(0)+1%[95]
323WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.23Value of the first Lame parameter λ201800Normalised width w(t)/w(0)+0%[95]
324WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.27Value of the first Lame parameter λ201800Normalised width w(t)/w(0)−1.5%[95]
325WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.32Value of the first Lame parameter λ201800Normalised width w(t)/w(0)−2%[95]
326WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.38Value of the first Lame parameter λ201800Normalised width w(t)/w(0)−3%[95]
327WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.07Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−6%[95]
328WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.14Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−0%[95]
329WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.21Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−0%[95]
330WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.23Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−1%[95]
331WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.27Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−3%[95]
332WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.32Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−14%[95]
333WaterPNaAMPS/PAAm double-network (DN) hydrogelsStretching1.4 М01.38Value of the first Lame parameter λ201800normalized engineering stress σ(t)/σ(0)−17%[95]
334waste water after blowing by 99.99% N2T. denitrificans. immobilised on electrodMixing99.99% N2105mm26.2259,200Concntration of Na+−20%[79]
335waste water after blowing by 99.99% N2T. denitrificans. immobilised on electrodMixing99.99% N2105mm26.2259,200Electrode voltage+80%[79]
336waste water after blowing by 99.99% N2T. denitrificans. immobilised on electrodMixing99.99% N2105mm26.2259,200Consumption of NO3-+35%[79]
337Water-Mixing-1037mm30180Concentration of bulk nanobubbles+150%[96]
338Water-Mixing-537mm30180Concentration of bulk nanobubbles+80%[96]
339Water-Mixing-1537mm30180Concentration of bulk nanobubbles+200%[96]
340Water-Mixing-1537mm30180Concentration of bulk nanobubbles+30%[96]
341Water-Mixing-1537mm3060Concentration of bulk nanobubbles+200%[96]
342Water-Mixing-1537mm30360Concentration of bulk nanobubbles+1500%[96]
343Water-Mixing-1537mm30600Concentration of bulk nanobubbles+1100%[96]
344Water-Ultrasound-22,0002000mW/cm25000300pH+6%[121]
345Water-Ultrasound-100,0001,000,0002000mW/cm25000300pH−17%[121]
346Water (artesian)total hardness. carbonate hardness; Ca+2 content; Mg+2 content; pH 7. sulfates. chloridesUltrasound16.6 mg-eq/L. 4.8 mg-eq/. 17.7 mg/L. 8.9 mg/L. 7.8 mg/L. 576 mg/L. 12.0 mg/L22,0002000mW/cm25000300pH+12%[121]
347WaterClay particlesUltrasound20 mg/mL20,0002000mW/cm2100010Precipitation rate−30%[124]
348WaterClay particlesUltrasound20 mg/mL20,0002000mW/cm2100030Precipitation rate+1%[124]
349WaterClay particlesUltrasound20 mg/mL20,0002000mW/cm2100060Precipitation rate+50%[124]
350WaterClay particlesUltrasound20 mg/mL20,0002000mW/cm21000120Precipitation rate+60%[124]
351WaterClay particlesUltrasound20 mg/mL20,0002000mW/cm21000300Precipitation rate+101%[124]
352Mousein vivoSinusoidal vibrations-450.3g2137,800Number of lectin-positive vessels−29%[20]
353Mousein vivoSinusoidal vibrations-450.3g2137,800α-actin-positive vessels−36%[20]
355Rabbitin vivoShaking-605g49002,160,000Myelin fibre diameter+80%[53]
356Humanin vivoVibration-253m/s292,90060vibration intensity transmission+90%[51]
357Humanin vivoVibration-2003m/s292,90060vibration intensity transmission+900%[51]
358Humanin vivoVibration-253m/s292,90060Apparent mass−50%[51]
359Humanin vivoVibration-2003m/s292,90060Apparent mass−80%[51]
360Humanin vivoVibration-253m/s292,90060Mechanical impedance+25%[51]
361Humanin vivoVibration-2003m/s292,90060Mechanical impedance+150%[51]
362Ratin vivoVibration-8032m/s220590,000Length of nerve regeneration by 6 days+57%[54]
363Ratin vivoVibration-8032m/s220590,000Length of nerve regeneration by 6 days+23%[54]
364Ratin vivoVibration-810.5mm24028,800Expression of IGF-I in nerve fibers+100%[210]
365Rat. tailin vivoVertical shaking-62.549m/s26144,00Expression of MT-1a+700%[40]
366Rat. tailin vivoVertical shaking-12549m/s26144,00Expression of MT-1a+660%[40]
367Rat. tailin vivoVertical shaking-25049m/s26144,00Expression of MT-1a+630%[40]
368Rat. tailin vivoVertical shaking-62.549m/s26144,00Expression of IL-6+900%[40]
369Rat. tailin vivoVertical shaking-12549m/s26144,00Expression of IL-6+500%[40]
370Rat. tailin vivoVertical shaking-25049m/s2614,400Expression of IL-6+300%[40]
371Rat. tailin vivoVertical shaking-62.549m/s26144,00Expression of IL-6+400%[40]
372Rat. tailin vivoVertical shaking-12549m/s26144,00Expression of IL-6+220%[40]
373Rat. tailin vivoVertical shaking-25049m/s2614,400Internal diameter of the caudal vein−70%[40]
374Rat. tailin vivoVertical shaking-25049m/s26144,00Vascular smooth muscle thickness+35%[40]
375Rat. tailin vivoVertical shaking-12549m/s2614,400Synthesis of nitrotyrosine+125%[40]
376Rat. tailin vivoVertical shaking-25049m/s26144,00Synthesis of nitrotyrosine+150%[40]
377Rat. tailin vivoVertical shaking-12549m/s26144,00Synthesis of IL-6+50%[40]
378Rat. tailin vivoVertical shaking-25049m/s2614,400Synthesis of IL-6+70%[40]
379Rat. tailin vivoVertical shaking-6049m/s2514,000Number of vacuoles in endothelial cells+2900%[41]
380Rat. tailin vivoVertical shaking-12049m/s2514,000Number of vacuoles in endothelial cells+1300%[41]
381Rat. tailin vivoVertical shaking-3049m/s25144,00Number of ruptures of the inner vessel wall+1600:[41]
382Rat. tailin vivoVertical shaking-6949m/s25144,00Number of ruptures of the inner vessel wall+2100%[41]
383Rat. tailin vivoVertical shaking-12049m/s2514,400Number of ruptures of the inner vessel wall+1500%[41]
384Rat. tailin vivoVertical shaking-80049m/s2514,400Number of ruptures of the inner vessel wall+1600%[41]
385Rat. tailin vivoVertical shaking-3049m/s2514,400Expression of NFATc3+100%[41]
386Rat. tailin vivoVertical shaking-6049m/s2514,400Expression of NFATc3+200%[41]
387Rat. tailin vivoVertical shaking-12049m/s2514,400Expression of NFATc3+200%[41]
388Rat. lower limbsin vivoVertical shaking-3049m/s21557,600Total plasma creatine phosphokinase (t-CPK) activity+1830%[42]
389Rat. lower limbsin vivoVertical shaking-6049m/s21557,600Total plasma creatine phosphokinase (t-CPK) activity+760%[42]
390Rat. lower limbsin vivoVertical shaking-12049m/s21557,600Total plasma creatine phosphokinase (t-CPK) activity+700%[42]
391Rat. lower limbsin vivoVertical shaking-24049m/s21557,600Total plasma creatine phosphokinase (t-CPK) activity+400%[42]
392Rat. lower limbsin vivoVertical shaking-48049m/s21557,600Total plasma creatine phosphokinase (t-CPK) activity+200%[42]
393Culture medium DMEMMouse Fibroblast CellsSinusoidal shaking-125016.5m/s25900Cells viability−59%[57]
394Culture medium DMEMMouse Fibroblast CellsShock shaking-125012.9m/s25900Cells viability−99%[57]
395Culture medium DMEMMouse red blood cellsShock shaking-125030,000m/s25900Erythrocyte lysis+100%[57]
396Culture medium DMEMMouse red blood cellsShock shaking-125015,000m/s25900Erythrocyte lysis+0.4%[57]
397Culture medium DMEMMouse red blood cellsShock shaking-12502000m/s25900Erythrocyte lysis+0.07%[57]
398Culture medium DMEMMouse red blood cellsShock shaking-12501000m/s25900Erythrocyte lysis+0.1%[57]
399Culture medium DMEMMouse red blood cellsShock shaking-1250500m/s25900Erythrocyte lysis+0.1%[57]
400Humanin vivoVibration-5.574.4N60,000600Peak inspiratory flow rate+136%[52]
401Humanin vivoVibration-5.574.4N60,000600Inspired volume+200%[52]
402Humanin vivoVibration-5.574.4N60,000600expired volume+290%[52]
403Human. healthy pre-pubertal boysin vivoVibration (platform)-402.1g34,000600Legs’ skin temperature+8%[46]
404Human. healthy pre-pubertal boysin vivoVibration (platform)-402.1g34,0003000P1NP plasma level after 8 days+18%[46]
405Human. healthy pre-pubertal boysin vivoVibration (platform)-402.1g34,0003000CTx plasma level after 8 days+10%[46]
406Rat with titanium hip implantsin vivoVibration (whole body)-800.3g353300Bone-to-implant contact+63%[48]
407Rat with titanium hip implantsin vivoVibration (whole body)-1400.3g353300Bone-to-implant contact+65%[48]
408Rat with titanium hip implantsin vivoVibration (whole body)-260.3g353300Peri-implant bone formation+20%[48]
409Rat with titanium hip implantsin vivoVibration (whole body)-800.3g353300Peri-implant bone formation+25%[48]
410Rat with titanium hip implantsin vivoVibration (whole body)-1400.043g353300Peri-implant bone formation+19%[48]
411Rat with titanium hip implantsin vivoVibration (whole body)-1400.3g353300Peri-implant bone formation+25%[48]
412Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body)-300.3g285134,400Obesity−50%[28]
413Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body) 300.3g28567,200Bone mineral densities+15%[28]
414Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body)-300.3g285134,400Vertebra compression test (maximum load)−18%[28]
415Rat (healthy)in vivoVibration (whole body)-300.3g285134,400vertebra compression test (rigidity)−45%[28]
416Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body)-300.3g285134,400Osteoblastic Cell Viability−10%[28]
417Rat (healthy)in vivoVibration (whole body)-300.3g285134,400Osteoblastic Cell Viability−12%[28]
418Rat (healthy)in vivoVibration (whole body)-300.3g285134,400Alkaline phosphatase (ALP) activity 4 day+15%[28]
419Rat (healthy)in vivoVibration (whole body)-300.3g285134,400alkaline phosphatase (ALP) activity 7 day−10%[28]
420Rat (healthy)in vivoVibration (whole body)-300.3g285134,400Expression of ALP−66%[28]
421Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body)-300.3g285134,400Expression of ALP−50%[28]
422Rat (healthy)in vivoVibration (whole body)-300.3g285134,400Expression of MMP2−40%[28]
423Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body)-300.3g285134,400Expression of MMP2−40%[28]
424Rat (healthy)in vivoVibration (whole body)-300.3g285134,400Expression of OSX−10%[28]
425Rat with removed ovaries (osteoporosis model)in vivoVibration (whole body)-300.3g285134,400Expression of OSX+20%[28]
426C57BL/6 mousein vivoVibration (whole body)-900.75m/s227.5300Arterial blood pressure+25%[39]
427C57BL/6 mousein vivoVibration (whole body)-800.75m/s227.5300Heart beat rate+19%[39]
428C57BL/6 mousein vivoVibration (whole body)-900.75m/s227.5300Heart beat rate+17%[39]
429Humanin vivoVibration (whole body)-300.3m/s26150014,400Bone mineralisation density+40%[44]
430Humanin vivoVibration (whole body)-300.3m/s26150014,400TRAP synthesis−20%[44]
431Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00024,000mW57200Expression of markers of neuronal differentiation and Alk+250%[116]
432Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,000310,000mW57200Expression of markers of neuronal differentiation and Alk+200%[116]
433Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00024,000mW57200Expression of Cenpf+40%[116]
434Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00031,000mW57200Expression of Cenpf+20%[116]
435Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00039,000mW57200Expression of Cenpf−30%[116]
436Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00024,000mW57200Expression of actin−30%[116]
437Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00031,000mW57200Expression of actin−40%[116]
438Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00039,000mW57200Expression of actin−80%[116]
439Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00024,000mW57200Expression of Pcdh17+60%[116]
440Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00031,000mW57200Expression of Pcdh17+70%[116]
441Culture medium DMEMHuman embryonic stem cells (hESC)Ultrasound-19,690,00039,000mW57200Expression of Pcdh17+20%[116]
442Culture medium DMEMSHED: stem cells from human exfoliated deciduous teethCentrifugation--100g530 min. every 24 h.. 7 днейCells proliferation−16.32%[25]
443Culture medium DMEMSHED: stem cells from human exfoliated deciduous teethCentrifugation--200g530 min. every 24 h.. 7 днейCells proliferation−18.36%[25]
444Culture medium DMEMSHED: stem cells from human exfoliated deciduous teethCentrifugation--300g530 min. every 24 h.. 7 днейCells proliferation−16.32%[25]
445Culture medium α-MEMMC3T3-E1Vibration-12.50.5G51,382,400Confluency+19.4%[26]
446Culture medium α-MEMMC3T3-E1Vibration-12.50.5G41,123,200Confluency+15.6%[26]
447Culture medium α-MEMMC3T3-E1Constant flow-00.28 ± 0.02mL/min41,382,400Confluency−22.2%[26]
448Culture medium α-MEMMC3T3-E1Constant flow-00.28 ± 0.02mL/min41,123,200Confluency+12.5%[26]
449Culture medium α-MEMhuman PDL stem cellsVibration-500.3G1030 min. every 24 h.Alkaline phosphatase activity+123.8%[211]
450Culture medium α-MEMhuman PDL stem cellsVibration-600.3G1030 min. every 24 h.Alkaline phosphatase activity+96.23%[211]
451Culture medium α-MEMhuman PDL stem cellsVibration-100.3G1030 min. every 24 h.Expression of gene Runx2−43.75%[211]
452Culture medium α-MEMhuman PDL stem cellsVibration-400.3G1030 min. every 24 h.Expression of gene Runx2+175%[211]
453Culture medium α-MEMhuman PDL stem cellsVibration-500.3G1030 min. every 24 h.Expression of gene Runx2+275%[211]
454Culture medium α-MEMhuman PDL stem cellsVibration-600.3G1030 min. every 24 h.Expression of gene Runx2+68.75%[211]
455Culture medium α-MEMhuman PDL stem cellsVibration-900.3G1030 min. every 24 h.Expression of gene Runx2+34.37%[211]
456Culture medium α-MEMhuman PDL stem cellsVibration-1800.3G1030 min. every 24 h.Expression of gene Runx2−25%[211]
457Culture medium α-MEMhuman PDL stem cellsVibration-200.3G1030 min. every 24 h.Expression of gene Osx−50%[211]
458Culture medium α-MEMhuman PDL stem cellsVibration-400.3G1030 min. every 24 h.Expression of gene Osx+62.5%[211]
459Culture medium α-MEMhuman PDL stem cellsVibration-500.3G1030 min. every 24 h.Expression of gene Osx+109.3%[211]
460Culture medium α-MEMhuman PDL stem cellsVibration-600.3G1030 min. every 24 h.Expression of gene Osx−43.75%[211]
461Culture medium α-MEMhuman PDL stem cellsVibration-1200.3G1030 min. every 24 h.Expression of gene Osx−43.75%[211]
462Culture medium α-MEMhuman PDL stem cellsVibration-1500.3G1030 min. every 24 h.Expression of gene Osx−56.25%[211]
463Culture medium α-MEMhuman PDL stem cellsVibration-1800.3G1030 min. every 24 h.Expression of gene Osx+37.5%[211]
464Culture medium α-MEMhuman PDL stem cellsVibration-400.3G1030 min. every 24 h.Osteocalcin levels (OCN)+340.9%[211]
465Culture medium α-MEMhuman PDL stem cellsVibration-500.3G1030 min. every 24 h.Osteocalcin levels (OCN)+390.9%[211]
466Culture medium α-MEMhuman PDL stem cellsVibration-600.3G1030 min. every 24 h.Osteocalcin levels (OCN)+318.18%[211]
467Culture medium α-MEMhuman PDL stem cellsVibration-900.3G1030 min. every 24 h.Osteocalcin levels (OCN)+122.72%[211]
468Culture medium α-MEMhuman PDL stem cellsVibration-1200.3G1030 min. every 24 h.Osteocalcin levels (OCN)+154.54%[211]
469Culture medium DMEM low glucosemesenchymal stem cells (MSCs)Stretching-0.035%2240 min/dayCollagen content+26.22%[67]
470StemFit AK02NiPSCVibration-1000.02mm23000Cells differentiation−28.5%[68]
471StemFit AK02NiPSCVibration-1500.04mm23000Cells differentiation−22.85%[68]
472(DMEM)/Ham’s F12human adipose-derived stem cells (hASCs)Cyclic stretching-0.55%%2864,000Expression of CYP1B1+68.18%[69]
473(DMEM)/Ham’s F12human adipose-derived stem cells (hASCs)Cyclic stretching-0.55%%2864,000Metabolic activity−30.6%[69]
474Ham’s F–12 KChinese Hamster Ovary (CHO)-adherent cellsVibration-10–500<1g10345,600Cells proliferation+79%[56]
475Freestyle CHO ExpressionChinese Hamster Ovary (CHO)- suspension cellsVibration-10–500<1g10345,600Cells proliferation−13.0%[56]
476Freestyle CHO ExpressionChinese Hamster Ovary (CHO)- suspension cellsVibration-300.7g10345,600Cells proliferation+210%[56]
477RPMIЕ-cellsVibration-300.7g10345,600Cells proliferation+20.3%[56]
478DMEMPrimary fibrochondrocytesCyclic stretching-18%228,800Expression of COL2A1+110%[59]
479DMEMPrimary fibrochondrocytesCyclic stretching-18%228,800Expression of SOX9+200%[59]
480DMEMPrimary fibrochondrocytesCyclic stretching-18%228,800Expression of ACAN+215.78%[59]
481DMEMPrimary fibrochondrocytesCyclic stretching-18%228,800Proportion of BrdU-positive cells+46.6%[59]
482DMEMHuman osteoblast-like cells primary
explant cultures
Vibration-6030µm596 hConfluency+23.07%[60]
483DMEMHuman osteoblast-like cells primary
explant cultures
Vibration-6030µm596 hAlkaline phosphatase activity−29.82%[60]
484Culture medium α-MEMGD25 cellsVibration-12.50.5G512 daysCells culture density+50%[61]
485Culture medium α-MEMGD25 cellsVibration-12.50.5G515 daysCells culture density+60%[61]
486Culture medium α-MEMGD25 cellsVibration-12.50.5G518 daysCells culture density+83.67%[61]
487Culture medium α-MEMGD25 cellsVibration-12.50.5G521 dayCells culture density+73%[61]
488Culture medium α-MEMGD25 cellsVibration-12.50.5G524 daysCells culture density+45.8%[61]
489Culture medium α-MEMGD25 cellsVibration-12.50.5G515 daysCell layer thickness+46.15%[61]
490Culture medium DMEMHuman periodontal ligament stem cell (hPDLSC)Shear stress--0.5dyn/cm20.510,800Expression of gene IDO+80%[62]
491Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Expression of gene IDO+140%[62]
492Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--0.5dyn/cm20.510,800Expression of gene COX2+83.3%[62]
493Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Expression of gene COX2+266.6%[62]
494Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--0.5dyn/cm20.510,800Activity of IDO−18.18%[62]
495Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Activity of IDO+106%[62]
496Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--10dyn/cm20.510,800Activity of IDO+78.78%[62]
497Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--0.5dyn/cm20.510,800Kynurenine syntethis+58.85%[62]
498Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Kynurenine syntethis+170.5%[62]
499Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--0.5dyn/cm20.510,800General concentration of TGF-β1 in medium+17.3%[62]
500Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800General concentration of TGF-β1 in medium−9.6%[62]
501Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--0.5dyn/cm20.510,800Concentration of active TGF-β1 in medium+2.87%[62]
502Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Concentration of active TGF-β1 in medium+125%[62]
503Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--10dyn/cm20.510,800Concentration of active TGF-β1 in medium+125%[62]
504Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Concentration of mature TGF-β1/β-ACTIN−6.66%[62]
505Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--10dyn/cm20.510,800Concentration of mature TGF-β1/β-ACTIN−43.3%[62]
506Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--0.5dyn/cm20.510,800Concentration of IFN-γ in lysate of cells−51.21%[62]
507Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--5dyn/cm20.510,800Concentration of IFN-γ in lysate of cells−51%[62]
508Culture medium DMEMHuman periodontal ligament stem cells (hPDLSC)Shear stress--10dyn/cm20.510,800Concentration of IFN-γ in lysate of cells−65.85%[62]
509Culture medium DMEMChondrocytes from pigVibration-250.5G0.586,400Type II collagen expression+308.3%[64]
510Culture medium DMEMChondrocytes from pigVibration-250.5G0.586,400Aggrecan expression+225%[64]
511Culture medium DMEMChondrocytes from pigVibration-250.5G0.586,400Type I collagen expression+372.72%[64]
512Culture medium DMEMChondrocytes from pigVibration-250.5G0.586,400Fibronectin expression+206.25%[64]
513Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Cells viability+47.36%[66]
514Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Proportion of apoptosis events−47.36%[66]
515Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Relative cell area+52.17%[66]
516Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Relative fluorescence intensity+37%[66]
517Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Cell anisotropy+111%[66]
518Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of Wnt3a+37.9%[66]
519Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of β-catenin+52%[66]
520Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of Sost−36.53%[66]
521Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of DKK1−29.5%[66]
522Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of RANKL−33.33%[66]
523Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of OPG+62.2%[66]
524Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of PGE-2+40%[66]
525Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of TNF- α−31%[66]
526Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of RANKL−20.5%[66]
527Culture medium α-MEMOsteocytic MLO-Y4 cellsVibration-450.5g0.5259,200Expression of OPG+16.6%[66]
528Holtfreter’s salineNewt Triturus alpestri notochord (tails)US-1,000,0008000mW/cm2-300Summ of cells with abnorman endoplasmatic reticulum+7000%[100]
529TC culture mediunWhole blood cellsUS-2,250,00030mW53600Cell with chromosomal aberrations+560[111]
530TC culture mediunWhole blood cellsUS-2,250,00030mW57200Cell with chromosomal aberrations+620[111]
531TC culture mediunWhole blood cellsUS-2,250,00030mW510,800Cell with chromosomal aberrations+360[111]
532TC culture mediunWhole blood cellsUS-2,250,00030mW53600Chromotid aberrations per 100 cells+375[111]
533TC culture mediunWhole blood cellsUS-2,250,00030mW57200Chromotid aberrations per 100 cells+800[111]
534TC culture mediunWhole blood cellsUS-2,250,00030mW510,800Chromotid aberrations per 100 cells+375[111]
535TC culture mediunWhole blood cellsUS-2,250,00030mW53600Chromosome aberrations per 100 cells+1100[111]
536TC culture mediunWhole blood cellsUS-2,250,00030mW57200Chromosome aberrations per 100 cells+1800[111]
537TC culture mediunWhole blood cellsUS-2,250,00030mW510,800Chromosome aberrations per 100 cells+700[111]
538TC culture mediunWhole blood cellsUS-2,250,00030mW53600Total aberrations+520[111]
539TC culture mediunWhole blood cellsUS-2,250,00030mW57200Total aberrations+1000[111]
540TC culture mediunWhole blood cellsUS-2,250,00030mW510,800Total aberrations+440[111]
541WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-500,00080mW/cm20.5100Cell death+100[212]
542WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-1,000,000200mW/cm20.5100Cell death+100[212]
543WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μm=US-400,000140mW/cm20.5100Cell death+100[212]
544WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-1,500,000400mW/cm20.5100Cell death+100[212]
545WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-200,000500mW/cm20.5100Cell death+100[212]
546WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-250,000450mW/cm20.5100Cell death+100[212]
547WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-500,0000400mW/cm20.5100Cell death+100[212]
548WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-7,500,000750mW/cm20.5100Cell death+100[212]
549WaterElodea leaves cells with gas bodies sizes 10 × 14.7 μmUS-1,000,00001500mW/cm20.5100Cell death+100[212]
550WaterElodea leaves cells with gas bodies sizes less than 5.3 × 7.4 μmUS-500,0001100mW/cm20.5100Cell death+100[212]
551WaterElodea leaves cells with gas bodies sizes less than 5.3 × 7.4 μmUS-1,000,0001300mW/cm20.5100Cell death+100[112]
552WaterElodea leaves cells with gas bodies sizes less than 5.3 × 7.4 μmUS-1,500,000400mW/cm20.5100Cell death+100[212]
553WaterElodea leaves cells with gas bodies sizes less than 5.3 × 7.4 μmUS-2,000,000300mW/cm20.5100Cell death+100[212]
554WaterElodea leaves cells with gas bodies sizes less than 5.3 × 7.4 μmUS-500,0000200mW/cm20.5100Cell death+100[212]
555WaterElodea leaves cells with gas bodies sizes less than 5.3 × 7.4 μmUS-1,000,000900mW/cm20.5100Cell death+100[212]
556Isotonic salineHealthy human latelets after incubation at 22 °C for 30 minUS-1,000,000200mW/cm21.5300Floculation during incubation+10[213]
557Isotonic salineHealthy human latelets after incubation at 22 °C for 30 minUS-1,000,000600mW/cm21.5300Floculation during incubation+100[213]
558Isotonic salineRed blood cells of healthy humanUS-1,000,0001000mW/cm20.4560Hemolysis (percent)+1900[110]
559Isotonic salineRed blood cells of healthy humanUS-1,000,0002000mW/cm20.4560Hemolysis (percent)+2300[110]
560Isotonic salineRed blood cells of healthy humanUS-1,000,0003000mW/cm20.4560Hemolysis (percent)+3200[110]
561Isotonic salineRed blood cells of healthy humanUS-1,000,0004000mW/cm20.4560Hemolysis (percent)+1600[110]
562Isotonic salineRed blood cells of healthy humanUS-1,000,0005000mW/cm20.4560Hemolysis (percent)+1400[110]
563Albunex solutionRed blood cells of healthy humanUS33 μL/mL1,000,0001000mW/cm20.4560Hemolysis (percent)+2600[110]
564Albunex solutionRed blood cells of healthy humanUS33 μL/mL1,000,0002000mW/cm20.4560Hemolysis (percent)+4500[110]
565Albunex solutionRed blood cells of healthy humanUS33 μL/mL1,000,0003000mW/cm20.4560Hemolysis (percent)+5600[110]
566Albunex solutionRed blood cells of healthy humanUS33 μL/mL1,000,0004000mW/cm20.4560Hemolysis (percent)+5100[110]
567Albunex solutionRed blood cells of healthy humanUS33 μL/mL1,000,0005000mW/cm20.4560Hemolysis (percent)+4800[110]
568photosensitive drug merocyanine 540 (MC 540)HL-60 cell lineUS15 μg/mL225,000400mW/cm2530Cell death+49.5[214]
569Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+800[117]
570Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120CFU counts growth from spores−99[117]
571Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+800[117]
572Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+12800[117]
573Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+25600[117]
574Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+51200[117]
575Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+204800[117]
576Culture mediumBacillus subtilis sporesUS-67,00050,000mW0.125120DNA releazing from spores+204800[117]
577RPMI1640leukemic cells HL-60US-750,000103,700mW/cm−21530DNA repair synthesis+60[109]
578RPMI1640leukemic cells HL-60US-750,000103,700mW/cm−21530Cell viability−45[109]
579RPMI1640leukemic cells HL-60US-750,000103,700mW/cm−21530Apoptose percantage+200[109]
580RPMI1640leukemic cells HL-60US-750,000103,700mW/cm−21530Proliferation rate−60[109]
581RPMI1640leukemic cells K562US-750,00054,600mW/cm−21530Apoptose percantage+300[109]
582RPMI1640leukemic cells U937US-750,00054,600mW/cm−21530Apoptose percantage+400[109]
583RPMI1640leukemic cells M1/2US-750,00054,600mW/cm−21530Apoptose percantage+110[109]
584RPMI1640leukemic cells HL-60US-750,000103,700mW/cm−21530Late apoptose percantage+300[109]
585RPMI1640leukemic cells HL-60US-750,00022,400mW/cm−21530Early apoptose percantage+300[109]
586RPMI1640leukemic cells HL-60US-750,000103,700mW/cm−21530Early apoptose percantage+400[109]
587RPMI1640Healthy human plateletUS-22,0001600mW510Agregation rate+15[108]
588RPMI1640Healthy human plateletUS-22,0001600mW530Agregation rate+35[108]
589RPMI1640Healthy human plateletUS-22,0001600mW560Agregation rate+55[108]
590RPMI1640Healthy human plateletUS-22,0001400mW515Ca2+ concentration is cytoplasm+100[108]
591RPMI1640Healthy human plateletUS-22,0001400mW530Ca2+ concentration is cytoplasm+180[108]
592RPMI1640Healthy human plateletUS-2200016,000mW560Ca2+ concentration is cytoplasm+900[108]
593RPMI-1640K562 cellsUS-1,800,000220mW/cm−22.51800Early apoptic cells count+410[215]
594RPMI-1640K562 cellsUS-1,800,000220mW/cm−22.57200Early apoptic cells count+740[215]
595RPMI-1640K562 cellsUS-1,800,000220mW/cm−22.518,000Early apoptic cells count+1450[215]
596RPMI-1640K562 cellsUS-1,800,000220mW/cm−22.57200bcl-2 protein expression+10000[215]
597RPMI-1640K562 cellsUS-1,800,000220mW/cm−22.57200bax protein expression+1000[215]
598YEPD mediumE. coli strain DH5-alphaUS-1,000,0005200mW/cm−20.00230Cell viability−99[118]
599YEPD mediumSaccharomyces cerevisiaeUS-1,000,0005200mW/cm−20.00230Cell viability−99[118]
600Tryptic soy brothS. epidermidisUS-70,0003000mW/cm−223600Proliferation rate+20[216]
601Tryptic soy brothS. epidermidisUS-70,0003000mW/cm−227200Proliferation rate+20[216]
602Tryptic soy brothS. epidermidisUS-70,0003000mW/cm2210,800Proliferation rate+15[216]
603Healthy rabbits’ eyesIn vivoUS-880,000190mW/cm21300Sodium fluorescein concentration in the aqueous humor+240[99]
604Healthy rabbits’ eyesIn vivoUS-880,000340mW/cm21300Sodium fluorescein concentration in the aqueous humor+380[99]
605Healthy rabbits’ eyesIn vivoUS-880,000560mW/cm21300Sodium fluorescein concentration in the aqueous humor+1060[99]
606Healthy rabbits’ eyesIn vivoUS-880,000190mW/cm21300Damaged cells+179[99]
607Healthy rabbits’ eyesIn vivoUS-880,000340mW/cm21300Damaged cells+174[99]
608Healthy rabbits’ eyesIn vivoUS-880,000560mW/cm21300Damaged cells+171[99]
609MCDB-131 mediumBovine aortic endothelial cellsUS (continuous wave)-1,000,0001200mW/cm25900Cell proliferation after 48 h+162[114]
610MCDB-131 mediumBovine aortic endothelial cellsUS (continuous wave)-3500,0001200mW/cm25900Cell proliferation after 48 h+154[114]
611MCDB-131 mediumBovine aortic endothelial cellsUS (pulse wave)-1,000,0001200mW/cm25900Cell proliferation after 48 h+135[114]
612MCDB-131 mediumBovine aortic endothelial cellsUS (pulse wave)-3500,0001200mW/cm25900Cell proliferation after 48 h+123[114]
61350 mM CaCl2Pseudomonas putida UWC1US-40,000240mW/cm−20.510Ultrasound DNA transfer+100[119]
61450 mM CaCl2Escherichia coli DH5αUS-40,000240mW/cm−20.510Ultrasound DNA transfer+100[119]
61550 mM CaCl2Pseudomonas fluorescens SBW25US-40,000240mW/cm−20.510Ultrasound DNA transfer+100[119]
616BG11 mediumMicrocystis aeruginosa cyanobacteriaUS-25,000320mW/cm−2250300Proliferation−90[120]
617BG11 mediumMicrocystis aeruginosa cyanobacteriaUS-25,000320mW/cm−2250300Chlorophyll a concentration−21[120]
618BG11 mediumMicrocystis aeruginosa cyanobacteriaUS-25,000320mW/cm−2250300PC absorbance−45[120]
619BG11 mediumMicrocystis aeruginosa cyanobacteriaUS-25,000320mW/cm−2250300Extracellular microcystins−17[120]
620BG11 mediumMicrocystis aeruginosa cyanobacteriaUS-25,000320mW/cm−2250300Oxygen evolution rate−40[120]
621F12 mediumHuman Airway Smooth Muscle (HASM) cellsUS-1,000,0001000mW/cm−20.8300Cell contractile moment−40[217]
622F12 mediumHuman Airway Smooth Muscle (HASM) cellsUS-1,000,0002000mW/cm−20.8300Cell contractile moment−65[217]

References

  1. Nasser, A.; Mingelgrin, U. Mechanochemistry: A review of surface reactions and environmental applications. Appl. Clay Sci. 2012, 67–68, 141–150. [Google Scholar] [CrossRef]
  2. Stunzhas, P.A. Mechanochemical Instability of Water and Its Applications. Phys. Wave Phenom. 2020, 28, 111–115. [Google Scholar] [CrossRef]
  3. Yin, Z.; Zhang, Q.; Li, S.; Cagnetta, G.; Huang, J.; Deng, S.; Yu, G. Mechanochemical synthesis of catalysts and reagents for water decontamination: Recent advances and perspective. Sci. Total Environ. 2022, 825, 153992. [Google Scholar] [CrossRef]
  4. Gudkov, S.V.; Sarimov, R.M.; Astashev, M.E.; Pishchalnikov, R.Y.; Yanykin, D.V.; Simakin, A.V.; Shkirin, A.V.; Serov, D.A.; Konchekov, E.M.; Gusein-zade, N.G.O.; et al. Modern physical methods and technologies in agriculture. Physics-Uspekhi 2023, 67, 194–210. [Google Scholar] [CrossRef]
  5. Males, R.M.; Clark, R.M.; Wehrman, P.J.; Gates, W.E. Algorithm for Mixing Problems in Water Systems. J. Hydraul. Eng. 1985, 111, 206–219. [Google Scholar] [CrossRef]
  6. Bunkin, N.F.; Shkirin, A.V.; Ninham, B.W.; Chirikov, S.N.; Chaikov, L.L.; Penkov, N.V.; Kozlov, V.A.; Gudkov, S.V. Shaking-Induced Aggregation and Flotation in Immunoglobulin Dispersions: Differences between Water and Water–Ethanol Mixtures. ACS Omega 2020, 5, 14689–14701. [Google Scholar] [CrossRef]
  7. Metreveli, G.; Philippe, A.; Schaumann, G.E. Disaggregation of silver nanoparticle homoaggregates in a river water matrix. Sci. Total Environ. 2015, 535, 35–44. [Google Scholar] [CrossRef]
  8. Bull, H.B. Protein Denaturation. Cold Spring Harb. Symp. Quant. Biol. 1938, 6, 140–149. [Google Scholar] [CrossRef]
  9. Kiese, S.; Papppenberger, A.; Friess, W.; Mahler, H.-C. Shaken, Not Stirred: Mechanical Stress Testing of an IgG1 Antibody. J. Pharm. Sci. 2008, 97, 4347–4366. [Google Scholar] [CrossRef]
  10. Boldyrev, V.V. Mechanochemistry and mechanical activation of solids. Russ. Chem. Rev. 2006, 75, 177–189. [Google Scholar] [CrossRef]
  11. Kákosy, T. 2 Vibration disease. Baillière’s Clin. Rheumatol. 1989, 3, 25–50. [Google Scholar] [CrossRef] [PubMed]
  12. de Souza, R.J.; Nilsson, T.; Wahlström, J.; Burström, L. Hand-arm vibration and the risk of vascular and neurological diseases—A systematic review and meta-analysis. PLoS ONE 2017, 12, e0180795. [Google Scholar] [CrossRef]
  13. Heaver, C.; Goonetilleke, K.S.; Ferguson, H.; Shiralkar, S. Hand–arm vibration syndrome: A common occupational hazard in industrialized countries. J. Hand Surg. Eur. Vol. 2011, 36, 354–363. [Google Scholar] [CrossRef] [PubMed]
  14. Lee, J.Y.P.; Saez, N.J.; Cristofori-Armstrong, B.; Anangi, R.; King, G.F.; Smith, M.T.; Rash, L.D. Inhibition of acid-sensing ion channels by diminazene and APETx2 evoke partial and highly variable antihyperalgesia in a rat model of inflammatory pain. Br. J. Pharmacol. 2018, 175, 2204–2218. [Google Scholar] [CrossRef]
  15. Giudice, E.D.; Spinetti, P.R.; Tedeschi, A. Water Dynamics at the Root of Metamorphosis in Living Organisms. Water 2010, 2, 566–586. [Google Scholar] [CrossRef]
  16. Petrova, A.; Tarasov, S.; Gorbunov, E.; Stepanov, G.; Fartushnaya, O.; Zubkov, E.; Molodtsova, I.; Boriskin, V.; Zatykina, A.; Smirnov, A.; et al. Phenomenon of Post-Vibration Interactions. Symmetry 2024, 16, 958. [Google Scholar] [CrossRef]
  17. Stepanov, G.O.; Boriskin, V.S.; Rodionova, N.N.; Fedorova, M.M.; Petrova, A.O.; Novikov, V.V.; Yablokova, E.V.; Chernikov, D.Y.; Minakov, A.V.; Tarasov, S.A. The Effect of Aqueous Solutions Processed Using Gradual Technology on the Emission of a Carbohydrate (Lactose) in the RF Range. Phys. Wave Phenom. 2024, 32, 43–47. [Google Scholar] [CrossRef]
  18. Subbotin, K.A.; Didenko, Y.S.; Titov, A.I.; Lis, D.A.; Pavlov, S.K.; Volkov, P.A.; Runina, K.I.; Voronov, V.V.; Chernova, E.V.; Lis, O.N.; et al. Influence of Ultrahigh Dilution Treatment of the Charge on the Growth and Spectroscopic Properties of Nd:MgMoO4 Potential Laser Crystal. Crystals 2024, 14, 100. [Google Scholar] [CrossRef]
  19. Don, E.S.; Stepanov, G.O.; Tarasov, S.A. The effects of physical treatment on physicochemical and biological properties of water and aqueous solutions. Fine Chem. Technol. 2023, 18, 426–445. [Google Scholar] [CrossRef]
  20. Murfee, W.L.; Hammett, L.A.; Evans, C.; Xie, L.; Squire, M.; Rubin, C.; Judex, S.; Skalak, T.C. High-frequency, low-magnitude vibrations suppress the number of blood vessels per muscle fiber in mouse soleus muscle. J. Appl. Physiol. 2005, 98, 2376–2380. [Google Scholar] [CrossRef]
  21. Astashev, M.E.; Serov, D.A.; Sarimov, R.M.; Gudkov, S.V. Influence of the Vibration Impact Mode on the Spontaneous Chemiluminescence of Aqueous Protein Solutions. Phys. Wave Phenom. 2023, 31, 189–199. [Google Scholar] [CrossRef]
  22. Kokornaczyk, M.O.; Acuña, C.; Mier y Terán, A.; Castelán, M.; Baumgartner, S. Vortex-like vs. turbulent mixing of a Viscum album preparation affects crystalline structures formed in dried droplets. Sci. Rep. 2024, 14, 12965. [Google Scholar] [CrossRef] [PubMed]
  23. Kokornaczyk, M.O.; Würtenberger, S.; Baumgartner, S. Impact of succussion on pharmaceutical preparations analyzed by means of patterns from evaporated droplets. Sci. Rep. 2020, 10, 570. [Google Scholar] [CrossRef]
  24. Tytik, D.L.; Kuzmin, V.I.; Souvorova, O.V.; Revina, A.A. Dynamics of optical properties of sequentially diluted lucigenin aqueous solutions according to luminescence data. Front. Chem. 2024, 12, 1439250. [Google Scholar] [CrossRef]
  25. Liu, Q.; Qian, H.; Yu, H.; Ren, F.; Fang, J.; Liu, F.; Liu, H.; Liang, J. Effects of mechanical force on proliferation and apoptosis of stem cells from human exfoliated deciduous teeth. Clin. Oral Investig. 2022, 26, 5205–5213. [Google Scholar] [CrossRef]
  26. Yamamoto, T.; Morishita, S. A study on the effect of medium sloshing caused by mechanical vibration on cell proliferation. In Proceedings of the Conference of Kanto Branch, Tokyo, Japan, 17–18 March 2018; Volume 2018.24. [Google Scholar] [CrossRef]
  27. Baimler, I.V.; Gudkov, S.V.; Matveeva, T.A.; Simakin, A.V.; Shcherbakov, I.A. Generation of Reactive Oxygen Species during Water Drops Fall on a Solid Surface. Phys. Wave Phenom. 2024, 32, 187–189. [Google Scholar] [CrossRef]
  28. Qing, F.; Xie, P.; Liem, Y.S.; Chen, Y.; Chen, X.; Zhu, X.; Fan, Y.; Yang, X.; Zhang, X. Administration duration influences the effects of low-magnitude, high-frequency vibration on ovariectomized rat bone. J. Orthop. Res. 2015, 34, 1147–1157. [Google Scholar] [CrossRef]
  29. Lee, J.K.; Banerjee, S.; Nam, H.G.; Zare, R.N. Acceleration of reaction in charged microdroplets. Q. Rev. Biophys. 2015, 48, 437–444. [Google Scholar] [CrossRef]
  30. Robertson, V.J.; Baker, K.G. A Review of Therapeutic Ultrasound: Effectiveness Studies. Phys. Ther. 2001, 81, 1339–1350. [Google Scholar] [CrossRef]
  31. Ambattu, L.A.; Yeo, L.Y. Sonomechanobiology: Vibrational stimulation of cells and its therapeutic implications. Biophys. Rev. 2023, 4, 021301. [Google Scholar] [CrossRef]
  32. Fesenko, E.E.; Yablokova, E.V.; Novikov, V.V. Weak Magnetic Fields Regulate the Ability of High Dilutions of Water to Enhance ROS Production by Neutrophils. Appl. Sci. 2024, 14, 3290. [Google Scholar] [CrossRef]
  33. Novikov, V.V.; Yablokova, E.V. Interaction between Highly Diluted Samples, Protein Solutions and Water in a Controlled Magnetic Field. Appl. Sci. 2022, 12, 5185. [Google Scholar] [CrossRef]
  34. Novikov, V.V.; Yablokova, E.V.; Fesenko, E.E. The Role of Water in the Effect of Weak Combined Magnetic Fields on Production of Reactive Oxygen Species (ROS) by Neutrophils. Appl. Sci. 2020, 10, 3326. [Google Scholar] [CrossRef]
  35. Petrov, S.I.; Epstein, O.I. Effect of Potentiated Solutions on Mercury(II) Signal in Inversion Voltammetry. Bull. Exp. Biol. Med. 2003, 135, 99–101. [Google Scholar] [CrossRef]
  36. Novikov, V.V. Effect of vibrational iterations of magnetized water on the physico-chemical properties of intact water. Russ. Phys. J. 2025, 67, 1718–1727. [Google Scholar] [CrossRef]
  37. Tarasov, S.A.; Gorbunov, E.A.; Don, E.S.; Emelyanova, A.G.; Kovalchuk, A.L.; Yanamala, N.; Schleker, A.S.S.; Klein-Seetharaman, J.; Groenestein, R.; Tafani, J.-P.; et al. Insights into the Mechanism of Action of Highly Diluted Biologics. J. Immunol. 2020, 205, 1345–1354. [Google Scholar] [CrossRef]
  38. Epstein, O. The Supramolecular Matrix Concept. Symmetry 2023, 15, 1914. [Google Scholar] [CrossRef]
  39. Li, Y.; Rabey, K.N.; Schmitt, D.; Norton, J.N.; Reynolds, R.P. Characteristics of Vibration that Alter Cardiovascular Parameters in Mice. J. Am. Assoc. Lab. Anim. Sci. 2015, 54, 372–377. [Google Scholar]
  40. Krajnak, K.; Miller, G.R.; Waugh, S.; Johnson, C.; Li, S.; Kashon, M.L. Characterization of Frequency-Dependent Responses of the Vascular System to Repetitive Vibration. J. Occup. Environ. Med. 2010, 52, 584–594. [Google Scholar] [CrossRef]
  41. Curry, B.D.; Govindaraju, S.R.; Bain, J.L.W.; Zhang, L.L.; Yan, J.-G.; Matloub, H.S.; Riley, D.A. Evidence for frequency-dependent arterial damage in vibrated rat tails. Anat. Rec. Part A Discov. Mol. Cell. Evol. Biol. 2005, 284A, 511–521. [Google Scholar] [CrossRef]
  42. Okada, A.; Okuda, H.; Inaba, R.; Ariizumi, M. Influence of local vibration on plasma creatine phosphokinase (CPK) activity. Occup. Environ. Med. 1985, 42, 678–681. [Google Scholar] [CrossRef] [PubMed]
  43. Morandi, L.; Angelini, C.; Prelle, A.; Pini, A.; Grassi, B.; Bernardi, G.; Politano, L.; Bruno, C.; De Grandis, D.; Cudia, P.; et al. High plasma creatine kinase: Review of the literature and proposal for a diagnostic algorithm. Neurol. Sci. 2006, 27, 303–311. [Google Scholar] [CrossRef] [PubMed]
  44. Wang, H.; Wan, Y.; Tam, K.F.; Ling, S.; Bai, Y.; Deng, Y.; Liu, Y.; Zhang, H.; Cheung, W.H.; Qin, L.; et al. Resistive vibration exercise retards bone loss in weight-bearing skeletons during 60 days bed rest. Osteoporos. Int. 2011, 23, 2169–2178. [Google Scholar] [CrossRef] [PubMed]
  45. Ballanti, P.; Minisola, S.; Pacitti, M.T.; Scarnecchia, L.; Rosso, R.; Mazzuoli, G.F.; Bonucci, E. Tartrate-resistant acid phosphate activity as osteoclastic marker: Sensitivity of cytochemical assessment and serum assay in comparison with standardized osteoclast histomorphometry. Osteoporos. Int. 1997, 7, 39–43. [Google Scholar] [CrossRef]
  46. Harrison, R.; Ward, K.; Lee, E.; Razaghi, H.; Horne, C.; Bishop, N.J. Acute bone response to whole body vibration in healthy pre-pubertal boys. J. Musculoskelet. Neuronal Interact. 2015, 15, 112–122. [Google Scholar]
  47. Szulc, P.; Naylor, K.; Hoyle, N.R.; Eastell, R.; Leary, E.T. Use of CTX-I and PINP as bone turnover markers: National Bone Health Alliance recommendations to standardize sample handling and patient preparation to reduce pre-analytical variability. Osteoporos. Int. 2017, 28, 2541–2556. [Google Scholar] [CrossRef]
  48. Ogawa, T.; Vandamme, K.; Zhang, X.; Naert, I.; Possemiers, T.; Chaudhari, A.; Sasaki, K.; Duyck, J. Stimulation of Titanium Implant Osseointegration Through High-Frequency Vibration Loading is Enhanced when Applied at High Acceleration. Calcif. Tissue Int. 2014, 95, 467–475. [Google Scholar] [CrossRef]
  49. Jiang, L.; Sheng, K.; Wang, C.; Xue, D.; Pan, Z. The Effect of MMP-2 Inhibitor 1 on Osteogenesis and Angiogenesis During Bone Regeneration. Front. Cell Dev. Biol. 2021, 8, 596783. [Google Scholar] [CrossRef]
  50. Liu, Q.; Li, M.; Wang, S.; Xiao, Z.; Xiong, Y.; Wang, G. Recent Advances of Osterix Transcription Factor in Osteoblast Differentiation and Bone Formation. Front. Cell Dev. Biol. 2020, 8, 601224. [Google Scholar] [CrossRef]
  51. Dong, R.G.; Welcome, D.E.; McDowell, T.W.; Wu, J.Z. Biodynamic Response of Human Fingers in a Power Grip Subjected to a Random Vibration. J. Biomech. Eng. 2004, 126, 447–457. [Google Scholar] [CrossRef]
  52. McCarren, B.; Alison, J.A.; Herbert, R.D. Vibration and its effect on the respiratory system. Aust. J. Physiother. 2006, 52, 39–43. [Google Scholar] [CrossRef] [PubMed]
  53. Ho, S.T.; Yu, H.S. Ultrastructural changes of the peripheral nerve induced by vibration: An experimental study. Occup. Environ. Med. 1989, 46, 157–164. [Google Scholar] [CrossRef] [PubMed]
  54. Dahlin, L.B.; Necking, L.E.; Lundström, R.; Lundborg, G. Vibration exposure and conditioning lesion effect in nerves: An experimental study in rats. J. Hand Surg. 1992, 17, 858–861. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, Y.H.; He, W.M.; Dong, M.; Yu, F.H.; Zhang, L.L.; Cui, Q.G.; Chu, Y. Effects of Shaking on the Growth and Mechanical Properties of Hedysarum leave May Be Independent of Water Regimes. Int. J. Plant Sci. 2008, 169, 503–508. [Google Scholar] [CrossRef]
  56. Chan, M.E.; Ashdown, C.; Strait, L.; Pasumarthy, S.; Hassan, A.; Crimarco, S.; Singh, C.; Patel, V.S.; Pagnotti, G.; Khan, O.; et al. Low intensity mechanical signals promote proliferation in a cell-specific manner: Tailoring a non-drug strategy to enhance biomanufacturing yields. Mechanobiol. Med. 2024, 2, 100080. [Google Scholar] [CrossRef]
  57. Lindell, H.; Peciulyte, A.; Grétarsson, S.L. Physics of Shock and Physiological Effects on Biological Systems. In Proceedings of the 15th International Conference on Hand-Arm Vibration, Nancy, France, 6–9 June 2023. [Google Scholar]
  58. Chen, S.; Gluhak-Heinrich, J.; Wang, Y.H.; Wu, Y.M.; Chuang, H.H.; Chen, L.; Yuan, G.H.; Dong, J.; Gay, I.; MacDougall, M. Runx2, Osx, and Dspp in Tooth Development. J. Dent. Res. 2009, 88, 904–909. [Google Scholar] [CrossRef]
  59. Bian, X.; Liu, X.; Zhou, M.; Tang, H.; Wang, R.; Ma, L.; He, G.; Xu, S.; Wang, Y.; Tan, J.; et al. Mechanical stimulation promotes fibrochondrocyte proliferation by activating the TRPV4 signaling pathway during tendon–bone insertion healing: CCN2 plays an important regulatory role. Burn. Trauma 2024, 12, tkae028. [Google Scholar] [CrossRef]
  60. Rosenberg, N.; Rosenberg, O.; Halevi Politch, J.; Abramovich, H. Optimal parameters for the enhancement of human osteoblast-like cell proliferation in vitro via shear stress induced by high-frequency mechanical vibration. Iberoam. J. Med. 2021, 3, 204–211. [Google Scholar] [CrossRef]
  61. Nishijima, A.; Shiraishi, T. A Study of Mechanism of Cell Proliferation Under Mechanical Vibration Considering β1 Integrin. In Proceedings of the Dynamics & Design Conference, Yamaguchi, Japan, 23–26 August 2016; Volume 2016. [Google Scholar] [CrossRef]
  62. Suwittayarak, R.; Klincumhom, N.; Ngaokrajang, U.; Namangkalakul, W.; Ferreira, J.N.; Pavasant, P.; Osathanon, T. Shear Stress Enhances the Paracrine-Mediated Immunoregulatory Function of Human Periodontal Ligament Stem Cells via the ERK Signalling Pathway. Int. J. Mol. Sci. 2022, 23, 7119. [Google Scholar] [CrossRef]
  63. Park, I.-K.; Letterio, J.J.; Gorham, J.D. TGF-β1 inhibition of IFN-γ-induced signaling and Th1 gene expression in CD4+ T cells is Smad3 independent but MAP kinase dependent. Mol. Immunol. 2007, 44, 3283–3290. [Google Scholar] [CrossRef]
  64. Watanabe, Y.; Shiraishi, T. Effects of mechanical vibrations on chondrocytes in monolayer culture for regenerative cartilage. Mech. Eng. J. 2024, 11, 24-00102. [Google Scholar] [CrossRef]
  65. Nham, G.T.H.; Zhang, X.; Asou, Y.; Shinomura, T. Expression of type II collagen and aggrecan genes is regulated through distinct epigenetic modifications of their multiple enhancer elements. Gene 2019, 704, 134–141. [Google Scholar] [CrossRef] [PubMed]
  66. Sun, T.; Yan, Z.; Cai, J.; Shao, X.; Wang, D.; Ding, Y.; Feng, Y.; Yang, J.; Luo, E.; Feng, X.; et al. Effects of mechanical vibration on cell morphology, proliferation, apoptosis, and cytokine expression/secretion in osteocyte-like MLO-Y4 cells exposed to high glucose. Cell Biol. Int. 2019, 44, 216–228. [Google Scholar] [CrossRef]
  67. Choi, K.-M.; Seo, Y.-K.; Yoon, H.-H.; Song, K.-Y.; Kwon, S.-Y.; Lee, H.-S.; Park, J.-K. Effects of mechanical stimulation on the proliferation of bone marrow-derived human mesenchymal stem cells. Biotechnol. Bioprocess Eng. 2007, 12, 601–609. [Google Scholar] [CrossRef]
  68. Kanie, K.; Sakai, T.; Imai, Y.; Yoshida, K.; Sugimoto, A.; Makino, H.; Kubo, H.; Kato, R. Effect of mechanical vibration stress in cell culture on human induced pluripotent stem cells. Regen. Ther. 2019, 12, 27–35. [Google Scholar] [CrossRef]
  69. Paul, N.E.; Denecke, B.; Kim, B.-S.; Dreser, A.; Bernhagen, J.; Pallua, N. The effect of mechanical stress on the proliferation, adipogenic differentiation and gene expression of human adipose-derived stem cells. J. Tissue Eng. Regen. Med. 2018, 12, 276–284. [Google Scholar] [CrossRef]
  70. Astashev, M.; Serov, D.; Gudkov, S. Application of Spectral Methods of Analysis for Description of Ultradian Biorhythms at the Levels of Physiological Systems, Cells and Molecules (Review). Mathematics 2023, 11, 3307. [Google Scholar] [CrossRef]
  71. Li, Y.-J.; Yu, M.; Xue, C.-D.; Zhang, H.-J.; Wang, G.-Z.; Chen, X.-M.; Qin, K.-R. Modeling of Endothelial Calcium Responses within a Microfluidic Generator of Spatio-Temporal ATP and Shear Stress Signals. Micromachines 2021, 12, 161. [Google Scholar] [CrossRef]
  72. Chakraborty, M.; Chu, K.; Shrestha, A.; Revelo, X.S.; Zhang, X.; Gold, M.J.; Khan, S.; Lee, M.; Huang, C.; Akbari, M.; et al. Mechanical Stiffness Controls Dendritic Cell Metabolism and Function. Cell Rep. 2021, 34, 108609. [Google Scholar] [CrossRef]
  73. Li, K.; Ma, H.; Li, S.; Wang, C. Numerical simulation of three physical fields in counter-current ultrasound. J. Food Process Eng. 2024, 47, e14676. [Google Scholar] [CrossRef]
  74. Annabi, N.; Tamayol, A.; Uquillas, J.A.; Akbari, M.; Bertassoni, L.E.; Cha, C.; Camci-Unal, G.; Dokmeci, M.R.; Peppas, N.A.; Khademhosseini, A. 25th Anniversary Article: Rational Design and Applications of Hydrogels in Regenerative Medicine. Adv. Mater. 2013, 26, 85–124. [Google Scholar] [CrossRef] [PubMed]
  75. Vining, K.H.; Stafford, A.; Mooney, D.J. Sequential modes of crosslinking tune viscoelasticity of cell-instructive hydrogels. Biomaterials 2019, 188, 187–197. [Google Scholar] [CrossRef] [PubMed]
  76. Charrier, E.E.; Pogoda, K.; Wells, R.G.; Janmey, P.A. Control of cell morphology and differentiation by substrates with independently tunable elasticity and viscous dissipation. Nat. Commun. 2018, 9, 449. [Google Scholar] [CrossRef]
  77. Makisha, N.; Gulshin, I. Sustainable Modernization of Wastewater Treatment Plants. Sustainability 2024, 16, 8757. [Google Scholar] [CrossRef]
  78. Abdiyev, K.; Azat, S.; Kuldeyev, E.; Ybyraiymkul, D.; Kabdrakhmanova, S.; Berndtsson, R.; Khalkhabai, B.; Kabdrakhmanova, A.; Sultakhan, S. Review of Slow Sand Filtration for Raw Water Treatment with Potential Application in Less-Developed Countries. Water 2023, 15, 2007. [Google Scholar] [CrossRef]
  79. Ye, J.; Ren, G.; Liu, L.; Zhang, D.; Zeng, R.J.; van Loosdrecht, M.C.M.; Zhou, S. Wastewater denitrification driven by mechanical energy through cellular piezo-sensitization. Nat. Water 2024, 2, 531–540. [Google Scholar] [CrossRef]
  80. Gudkov, S.V.; Penkov, N.V.; Baimler, I.V.; Lyakhov, G.A.; Pustovoy, V.I.; Simakin, A.V.; Sarimov, R.M.; Scherbakov, I.A. Effect of Mechanical Shaking on the Physicochemical Properties of Aqueous Solutions. Int. J. Mol. Sci. 2020, 21, 8033. [Google Scholar] [CrossRef]
  81. Lu, G.; Miao, X.; Liu, D. Crosstalk of physiological pH and chemical pKa under the umbrella of physiologically based pharmacokinetic modeling of drug absorption, distribution, metabolism, excretion, and toxicity. Expert Opin. Drug Metab. Toxicol. 2021, 17, 1103–1124. [Google Scholar] [CrossRef]
  82. Woods, K.N. Modeling of protein hydration dynamics is supported by THz spectroscopy of highly diluted solutions. Front. Chem. 2023, 11, 1131935. [Google Scholar] [CrossRef]
  83. Belovolova, L.V.; Glushkov, M.V.; Vinogradov, E.A.; Babintsev, V.A.; Golovanov, V.I. Ultraviolet fluorescence of water and highly diluted aqueous media. Phys. Wave Phenom. 2009, 17, 21–31. [Google Scholar] [CrossRef]
  84. Gudkov, S.V.; Lyakhov, G.A.; Pustovoy, V.I.; Shcherbakov, I.A. Influence of Mechanical Effects on the Hydrogen Peroxide Concentration in Aqueous Solutions. Phys. Wave Phenom. 2019, 27, 141–144. [Google Scholar] [CrossRef]
  85. Shcherbakov, I.A. Influence of External Impacts on the Properties of Aqueous Solutions. Phys. Wave Phenom. 2021, 29, 89–93. [Google Scholar] [CrossRef]
  86. Gudkov, S.V.; Lyakhov, G.A.; Pustovoy, V.I.; Shcherbakov, I.A. Vibration–Vortex Mechanism of Radical-Reaction Activation in an Aqueous Solution: Physical Analogies. Phys. Wave Phenom. 2021, 29, 108–113. [Google Scholar] [CrossRef]
  87. Sharapov, M.G.; Goncharov, R.G.; Parfenyuk, S.B.; Glushkova, O.V.; Novoselov, V.I. The Role of Phospholipase Activity of Peroxiredoxin 6 in Its Transmembrane Transport and Protective Properties. Int. J. Mol. Sci. 2022, 23, 15265. [Google Scholar] [CrossRef]
  88. Averill-Bates, D. Reactive oxygen species and cell signaling. Review. Biochim. Biophys. Acta BBA—Mol. Cell Res. 2024, 1871, 119573. [Google Scholar] [CrossRef]
  89. Styrkas, A.D.; Nikishina, N.G. Mechanochemical processes in water. High Energy Chem. 2007, 41, 396–402. [Google Scholar] [CrossRef]
  90. Lee, J.K.; Walker, K.L.; Han, H.S.; Kang, J.; Prinz, F.B.; Waymouth, R.M.; Nam, H.G.; Zare, R.N. Spontaneous generation of hydrogen peroxide from aqueous microdroplets. Proc. Natl. Acad. Sci. USA 2019, 116, 19294–19298. [Google Scholar] [CrossRef]
  91. Randolph, T.W.; Schiltz, E.; Sederstrom, D.; Steinmann, D.; Mozziconacci, O.; Schöneich, C.; Freund, E.; Ricci, M.S.; Carpenter, J.F.; Lengsfeld, C.S. Do Not Drop: Mechanical Shock in Vials Causes Cavitation, Protein Aggregation, and Particle Formation. J. Pharm. Sci. 2015, 104, 602–611. [Google Scholar] [CrossRef]
  92. Lu, Q.-Q.; Yin, D.-C.; Liu, Y.-M.; Wang, X.-K.; Yang, P.-F.; Liu, Z.-T.; Shang, P. Effect of mechanical vibration on protein crystallization. J. Appl. Crystallogr. 2010, 43, 473–482. [Google Scholar] [CrossRef]
  93. Kakinouchi, K.; Adachi, H.; Matsumura, H.; Inoue, T.; Murakami, S.; Mori, Y.; Koga, Y.; Takano, K.; Kanaya, S. Effect of ultrasonic irradiation on protein crystallization. J. Cryst. Growth 2006, 292, 437–440. [Google Scholar] [CrossRef]
  94. Slatinskaya, O.V.; Pyrkov, Y.N.; Filatova, S.A.; Guryev, D.A.; Penkov, N.V. Study of the Effect of Europium Acetate on the Intermolecular Properties of Water. Front. Phys. 2021, 9, 641110. [Google Scholar] [CrossRef]
  95. Imaoka, C.; Nakajima, T.; Indei, T.; Iwata, M.; Hong, W.; Marcellan, A.; Gong, J.P. Inverse mechanical-swelling coupling of a highly deformed double-network gel. Sci. Adv. 2023, 9, eabp8351. [Google Scholar] [CrossRef] [PubMed]
  96. Fang, Z.; Wang, X.; Zhou, L.; Zhang, L.; Hu, J. Formation and Stability of Bulk Nanobubbles by Vibration. Langmuir 2020, 36, 2264–2270. [Google Scholar] [CrossRef] [PubMed]
  97. Fisher, B.D.; Hiller, C.M.; Rennie, S.G.A. A Comparison of Continuous Ultrasound and Pulsed Ultrasound on Soft Tissue Injury Markers in the Rat. J. Phys. Ther. Sci. 2003, 15, 65–70. [Google Scholar] [CrossRef]
  98. Matsumoto, Y.; Nakano, J.; Oga, S.; Kataoka, H.; Honda, Y.; Sakamoto, J.; Okita, M. The Non-Thermal Effects of Pulsed Ultrasound Irradiation on the Development of Disuse Muscle Atrophy in Rat Gastrocnemius Muscle. Ultrasound Med. Biol. 2014, 40, 1578–1586. [Google Scholar] [CrossRef]
  99. Zderic, V.; Clark, J.I.; Vaezy, S. Drug Delivery Into the Eye With the Use of Ultrasound. J. Ultrasound Med. 2004, 23, 1349–1359. [Google Scholar] [CrossRef]
  100. Selman, G.G.; Jurand, A. An Electron Microscope Study of the Endoplasmic Reticulum in Newt Notochord Cells after Disturbance with Ultrasonic Treatment and Subsequent Regeneration. J. Cell Biol. 1964, 20, 175–183. [Google Scholar] [CrossRef]
  101. Miller, M.W. Comparison of micronuclei induction for X-ray and ultrasound exposures of Vicia faba root meristem cells. Ultrasound Med. Biol. 1978, 4, 263–267. [Google Scholar] [CrossRef]
  102. Doan, N.; Reher, P.; Meghji, S.; Harris, M. In vitro effects of therapeutic ultrasound on cell proliferation, protein synthesis, and cytokine production by human fibroblasts, osteoblasts, and monocytes. J. Oral Maxillofac. Surg. 1999, 57, 409–419. [Google Scholar] [CrossRef]
  103. Ling, L.; Wei, T.; He, L.; Wang, Y.; Wang, Y.; Feng, X.; Zhang, W.; Xiong, Z. Low-intensity pulsed ultrasound activates ERK1/2 and PI3K-Akt signalling pathways and promotes the proliferation of human amnion-derived mesenchymal stem cells. Cell Prolif. 2017, 50, e12383. [Google Scholar] [CrossRef]
  104. Salgarella, A.R.; Cafarelli, A.; Ricotti, L.; Capineri, L.; Dario, P.; Menciassi, A. Optimal Ultrasound Exposure Conditions for Maximizing C2C12 Muscle Cell Proliferation and Differentiation. Ultrasound Med. Biol. 2017, 43, 1452–1465. [Google Scholar] [CrossRef] [PubMed]
  105. Wee, P.; Wang, Z. Epidermal Growth Factor Receptor Cell Proliferation Signaling Pathways. Cancers 2017, 9, 52. [Google Scholar] [CrossRef] [PubMed]
  106. Ito, T.K.; Ishii, G.; Chiba, H.; Ochiai, A. The VEGF angiogenic switch of fibroblasts is regulated by MMP-7 from cancer cells. Oncogene 2007, 26, 7194–7203. [Google Scholar] [CrossRef] [PubMed]
  107. Uzbekov, R.; Prigent, C. A Journey through Time on the Discovery of Cell Cycle Regulation. Cells 2022, 11, 704. [Google Scholar] [CrossRef]
  108. Samal, A.B.; Adzerikho, I.D.; Mrochek, A.G.; Loiko, E.N. Platelet aggregation and change in intracellular Ca2+ induced by low frequency ultrasound in vitro. Eur. J. Ultrasound 2000, 11, 53–59. [Google Scholar] [CrossRef]
  109. Ashush, H.; Rozenszajn, L.A.; Blass, M.; Barda-Saad, M.; Azimov, D.; Radnay, J.; Zipori, D.; Rosenschein, U. Apoptosis induction of human myeloid leukemic cells by ultrasound exposure. Cancer Res. 2000, 60, 1014–1020. [Google Scholar]
  110. Miller, M.W.; Azadniv, M.; Doida, Y.; Brayman, A.A. Effect of a Stabilized Microbubble Contrast Agent on CW Ultrasound Induced Red Blood Cell Lysis In Vitro. Echocardiography 2007, 12, 1–11. [Google Scholar] [CrossRef]
  111. Macintosh, I.J.C.; Davey, D.A. Chromosome Aberrations Induced by an Ultrasonic Fetal Pulse Detector. BMJ 1970, 4, 92–93. [Google Scholar] [CrossRef]
  112. Trenchard, P.M. Ultrasound-induced orientation of discoid platelets and simultaneous changes in light transmission: Preliminary characterisation of the phenomenon. Ultrasound Med. Biol. 1987, 13, 183–195. [Google Scholar] [CrossRef]
  113. Chan, Y.-S.; Hsu, K.-Y.; Kuo, C.-H.; Lee, S.-D.; Chen, S.-C.; Chen, W.-J.; Ueng, S.W.-N. Using Low-Intensity Pulsed Ultrasound to Improve Muscle Healing After Laceration Injury: An in vitro and in vivo Study. Ultrasound Med. Biol. 2010, 36, 743–751. [Google Scholar] [CrossRef]
  114. Raz, D.; Zaretsky, U.; Einav, S.; Elad, D. Cellular Alterations in Cultured Endothelial Cells Exposed to Therapeutic Ultrasound Irradiation. Endothelium 2009, 12, 201–213. [Google Scholar] [CrossRef] [PubMed]
  115. Nagata, K.; Nakamura, T.; Fujihara, S.; Tanaka, E. Ultrasound Modulates the Inflammatory Response and Promotes Muscle Regeneration in Injured Muscles. Ann. Biomed. Eng. 2013, 41, 1095–1105. [Google Scholar] [CrossRef] [PubMed]
  116. Sun, C.; Dong, Y.; Wei, J.; Cai, M.; Liang, D.; Fu, Y.; Zhou, Y.; Sui, Y.; Wu, F.; Mikhaylov, R.; et al. Acoustically accelerated neural differentiation of human embryonic stem cells. Acta Biomater. 2022, 151, 333–345. [Google Scholar] [CrossRef]
  117. Belgrader, P.; Hansford, D.; Kovacs, G.T.A.; Venkateswaran, K.; Mariella, R.; Milanovich, F.; Nasarabadi, S.; Okuzumi, M.; Pourahmadi, F.; Northrup, M.A. A Minisonicator to Rapidly Disrupt Bacterial Spores for DNA Analysis. Anal. Chem. 1999, 71, 4232–4236. [Google Scholar] [CrossRef]
  118. Li, G.; Xiao, H.; Glo, M.; Cheg, J. Miniaturized cell lysis device using spherically focused ultrasound. Tsinghua Sci. Technol. 2003, 8, 487–492. [Google Scholar]
  119. Song, Y.; Hahn, T.; Thompson, I.P.; Mason, T.J.; Preston, G.M.; Li, G.; Paniwnyk, L.; Huang, W.E. Ultrasound-mediated DNA transfer for bacteria. Nucleic Acids Res. 2007, 35, e129. [Google Scholar] [CrossRef]
  120. Zhang, G.; Zhang, P.; Liu, H.; Wang, B. Ultrasonic damages on cyanobacterial photosynthesis. Ultrason. Sonochem. 2006, 13, 501–505. [Google Scholar] [CrossRef]
  121. Vikulin, P.; Vikulina, V. Influence of ultrasound on the concentration of hydrogen ions in water. IOP Conf. Ser. Mater. Sci. Eng. 2020, 869, 042007. [Google Scholar] [CrossRef]
  122. Villalón-López, D.N.; Martínez-Padilla, L.P. Ultrasound-assisted process to improve proteins recovery from industrial canola and soybean byproducts. J. Food Sci. Technol. 2024, 61, 1–12. [Google Scholar] [CrossRef]
  123. Yao, S.; Li, W.; Wu, Y.; Martin, G.; Ashokkumar, M. The Impact of High-Intensity Ultrasound-Assisted Extraction on the Structural and Functional Properties of Hempseed Protein Isolate (HPI). Foods 2023, 12, 348. [Google Scholar] [CrossRef]
  124. Vikulina, V.; Vikulin, P. Effect of ultrasound on sedimentation of suspended solids in water. IOP Conf. Ser. Mater. Sci. Eng. 2018, 365, 032001. [Google Scholar] [CrossRef]
  125. Serov, D.A.; Gritsaeva, A.V.; Yanbaev, F.M.; Simakin, A.V.; Gudkov, S.V. Review of Antimicrobial Properties of Titanium Dioxide Nanoparticles. Int. J. Mol. Sci. 2024, 25, 10519. [Google Scholar] [CrossRef]
  126. Cui, Y.; Wang, F.; Yuan, P.; Liu, W.; Fang, B.; Wang, Z.; Pu, Y. Harvesting Vibration Energy to Produce Hydrogen Peroxide with Bi3TiNbO9 Nanosheets through a Water Oxidation Dominated Dual-Channel Pathway. ACS Sustain. Chem. Eng. 2024, 12, 3595–3607. [Google Scholar] [CrossRef]
  127. Wen, Y.; Zhang, T.; Wang, J.; Pan, Z.; Wang, T.; Yamashita, H.; Qian, X.; Zhao, Y. Electrochemical Reactors for Continuous Decentralized H2O2 Production. Angew. Chem. Int. Ed. 2022, 61, e202205972. [Google Scholar] [CrossRef]
  128. Martino, F.; Perestrelo, A.R.; Vinarský, V.; Pagliari, S.; Forte, G. Cellular Mechanotransduction: From Tension to Function. Front. Physiol. 2018, 9, 824. [Google Scholar] [CrossRef]
  129. Burridge, K.; Monaghan-Benson, E.; Graham, D.M. Mechanotransduction: From the cell surface to the nucleus via RhoA. Philos. Trans. R. Soc. B Biol. Sci. 2019, 374, 20180229. [Google Scholar] [CrossRef]
  130. Gillespie, P.G.; Walker, R.G. Molecular basis of mechanosensory transduction. Nature 2001, 413, 194–202. [Google Scholar] [CrossRef]
  131. Janota, C.S.; Calero-Cuenca, F.J.; Gomes, E.R. The role of the cell nucleus in mechanotransduction. Curr. Opin. Cell Biol. 2020, 63, 204–211. [Google Scholar] [CrossRef]
  132. Maurer, M.; Lammerding, J. The Driving Force: Nuclear Mechanotransduction in Cellular Function, Fate, and Disease. Annu. Rev. Biomed. Eng. 2019, 21, 443–468. [Google Scholar] [CrossRef]
  133. Na, S.; Collin, O.; Chowdhury, F.; Tay, B.; Ouyang, M.; Wang, Y.; Wang, N. Rapid signal transduction in living cells is a unique feature of mechanotransduction. Proc. Natl. Acad. Sci. USA 2008, 105, 6626–6631. [Google Scholar] [CrossRef]
  134. Lin, S.-Y.; Corey, D.P. TRP channels in mechanosensation. Curr. Opin. Neurobiol. 2005, 15, 350–357. [Google Scholar] [CrossRef] [PubMed]
  135. Zhang, M.; Ma, Y.; Ye, X.; Zhang, N.; Pan, L.; Wang, B. TRP (transient receptor potential) ion channel family: Structures, biological functions and therapeutic interventions for diseases. Signal Transduct. Target. Ther. 2023, 8, 261. [Google Scholar] [CrossRef] [PubMed]
  136. King, M.C. Dynamic regulation of LINC complex composition and function across tissues and contexts. FEBS Lett. 2023, 597, 2823–2832. [Google Scholar] [CrossRef] [PubMed]
  137. Swift, J.; Discher, D.E. The nuclear lamina is mechano-responsive to ECM elasticity in mature tissue. J. Cell Sci. 2014, 127, 3005–3015. [Google Scholar] [CrossRef]
  138. Uhler, C.; Shivashankar, G.V. Regulation of genome organization and gene expression by nuclear mechanotransduction. Nat. Rev. Mol. Cell Biol. 2017, 18, 717–727. [Google Scholar] [CrossRef]
  139. Khilan, A.A.; Al-Maslamani, N.A.; Horn, H.F. Cell stretchers and the LINC complex in mechanotransduction. Arch. Biochem. Biophys. 2021, 702, 108829. [Google Scholar] [CrossRef]
  140. Plant, L.D. A Role for K2P Channels in the Operation of Somatosensory Nociceptors. Front. Mol. Neurosci. 2012, 5, 21. [Google Scholar] [CrossRef]
  141. Salzer, I.; Ray, S.; Schicker, K.; Boehm, S. Nociceptor Signalling through ion Channel Regulation via GPCRs. Int. J. Mol. Sci. 2019, 20, 2488. [Google Scholar] [CrossRef]
  142. Lesage, F.; Terrenoire, C.; Romey, G.; Lazdunski, M. Human TREK2, a 2P Domain Mechano-sensitive K+ Channel with Multiple Regulations by Polyunsaturated Fatty Acids, Lysophospholipids, and Gs, Gi, and Gq Protein-coupled Receptors. J. Biol. Chem. 2000, 275, 28398–28405. [Google Scholar] [CrossRef]
  143. Lauritzen, I.; Chemin, J.; Honoré, E.; Jodar, M.; Guy, N.; Lazdunski, M.; Jane Patel, A. Cross-talk between the mechano-gated K2P channel TREK-1 and the actin cytoskeleton. EMBO Rep. 2005, 6, 642–648. [Google Scholar] [CrossRef]
  144. Ruan, N.; Tribble, J.; Peterson, A.M.; Jiang, Q.; Wang, J.Q.; Chu, X.-P. Acid-Sensing Ion Channels and Mechanosensation. Int. J. Mol. Sci. 2021, 22, 4810. [Google Scholar] [CrossRef] [PubMed]
  145. Bornstein, B.; Watkins, B.; Passini, F.S.; Blecher, R.; Assaraf, E.; Sui, X.M.; Brumfeld, V.; Tsoory, M.; Kröger, S.; Zelzer, E. The mechanosensitive ion channel ASIC2 mediates both proprioceptive sensing and spinal alignment. Exp. Physiol. 2023, 109, 135–147. [Google Scholar] [CrossRef] [PubMed]
  146. Stephan, G.; Huang, L.; Tang, Y.; Vilotti, S.; Fabbretti, E.; Yu, Y.; Nörenberg, W.; Franke, H.; Gölöncsér, F.; Sperlágh, B.; et al. The ASIC3/P2X3 cognate receptor is a pain-relevant and ligand-gated cationic channel. Nat. Commun. 2018, 9, 1354. [Google Scholar] [CrossRef] [PubMed]
  147. Jaalouk, D.E.; Lammerding, J. Mechanotransduction gone awry. Nat. Rev. Mol. Cell Biol. 2009, 10, 63–73. [Google Scholar] [CrossRef]
  148. Springer, T.A.; Dustin, M.L. Integrin inside-out signaling and the immunological synapse. Curr. Opin. Cell Biol. 2012, 24, 107–115. [Google Scholar] [CrossRef]
  149. Anishkin, A.; Loukin, S.H.; Teng, J.; Kung, C. Feeling the hidden mechanical forces in lipid bilayer is an original sense. Proc. Natl. Acad. Sci. USA 2014, 111, 7898–7905. [Google Scholar] [CrossRef]
  150. Chen, W.; Lou, J.; Evans, E.A.; Zhu, C. Observing force-regulated conformational changes and ligand dissociation from a single integrin on cells. J. Cell Biol. 2012, 199, 497–512. [Google Scholar] [CrossRef]
  151. Chuang, Y.-C.; Chen, C.-C. Force From Filaments: The Role of the Cytoskeleton and Extracellular Matrix in the Gating of Mechanosensitive Channels. Front. Cell Dev. Biol. 2022, 10, 886048. [Google Scholar] [CrossRef]
  152. Cox, C.D.; Bavi, N.; Martinac, B. Biophysical Principles of Ion-Channel-Mediated Mechanosensory Transduction. Cell Rep. 2019, 29, 1–12. [Google Scholar] [CrossRef]
  153. Grevesse, T.; Dabiri, B.E.; Parker, K.K.; Gabriele, S. Opposite rheological properties of neuronal microcompartments predict axonal vulnerability in brain injury. Sci. Rep. 2015, 5, 9475. [Google Scholar] [CrossRef]
  154. Martinac, B.; Adler, J.; Kung, C. Mechanosensitive ion channels of E. coli activated by amphipaths. Nature 1990, 348, 261–263. [Google Scholar] [CrossRef] [PubMed]
  155. Mohammed, D.; Versaevel, M.; Bruyère, C.; Alaimo, L.; Luciano, M.; Vercruysse, E.; Procès, A.; Gabriele, S. Innovative Tools for Mechanobiology: Unraveling Outside-In and Inside-Out Mechanotransduction. Front. Bioeng. Biotechnol. 2019, 7, 162. [Google Scholar] [CrossRef] [PubMed]
  156. Sachs, F. Stretch-Activated Ion Channels: What Are They? Physiology 2010, 25, 50–56. [Google Scholar] [CrossRef]
  157. Shah, N.; Morsi, Y.; Manasseh, R. From mechanical stimulation to biological pathways in the regulation of stem cell fate. Cell Biochem. Funct. 2014, 32, 309–325. [Google Scholar] [CrossRef]
  158. Wilde, C.; Mitgau, J.; Suchý, T.; Schöneberg, T.; Liebscher, I. Translating the force—Mechano-sensing GPCRs. Am. J. Physiol.-Cell Physiol. 2022, 322, C1047–C1060. [Google Scholar] [CrossRef]
  159. Murthy, S.E.; Dubin, A.E.; Patapoutian, A. Piezos thrive under pressure: Mechanically activated ion channels in health and disease. Nat. Rev. Mol. Cell Biol. 2017, 18, 771–783. [Google Scholar] [CrossRef]
  160. Enyedi, P.; Czirják, G. Molecular Background of Leak K+ Currents: Two-Pore Domain Potassium Channels. Physiol. Rev. 2010, 90, 559–605. [Google Scholar] [CrossRef]
  161. Venkatachalam, K.; Montell, C. TRP Channels. Annu. Rev. Biochem. 2007, 76, 387–417. [Google Scholar] [CrossRef]
  162. Millward-Sadler, S.J.; Salter, D.M. Integrin-Dependent Signal Cascades in Chondrocyte Mechanotransduction. Ann. Biomed. Eng. 2004, 32, 435–446. [Google Scholar] [CrossRef]
  163. Chen, Q.; Yang, D.; Zong, H.; Zhu, L.; Wang, L.; Wang, X.; Zhu, X.; Song, X.; Wang, J. Growth-induced stress enhances epithelial-mesenchymal transition induced by IL-6 in clear cell renal cell carcinoma via the Akt/GSK-3β/β-catenin signaling pathway. Oncogenesis 2017, 6, e375. [Google Scholar] [CrossRef]
  164. Quiroga, X.; Walani, N.; Disanza, A.; Chavero, A.; Mittens, A.; Tebar, F.; Trepat, X.; Parton, R.G.; Geli, M.I.; Scita, G.; et al. A mechanosensing mechanism controls plasma membrane shape homeostasis at the nanoscale. eLife 2023, 12, 72316. [Google Scholar] [CrossRef] [PubMed]
  165. Roberts, S.R.; Knight, M.M.; Lee, D.A.; Bader, D.L. Mechanical compression influences intracellular Ca2+ signaling in chondrocytes seeded in agarose constructs. J. Appl. Physiol. 2001, 90, 1385–1391. [Google Scholar] [CrossRef] [PubMed]
  166. Wang, J.G.; Miyazu, M.; Matsushita, E.; Sokabe, M.; Naruse, K. Uniaxial Cyclic Stretch Induces Focal Adhesion Kinase (FAK) Tyrosine Phosphorylation Followed by Mitogen-Activated Protein Kinase (MAPK) Activation. Biochem. Biophys. Res. Commun. 2001, 288, 356–361. [Google Scholar] [CrossRef]
  167. Xu, B.; Song, G.; Ju, Y.; Li, X.; Song, Y.; Watanabe, S. RhoA/ROCK, cytoskeletal dynamics, and focal adhesion kinase are required for mechanical stretch-induced tenogenic differentiation of human mesenchymal stem cells. J. Cell. Physiol. 2012, 227, 2722–2729. [Google Scholar] [CrossRef]
  168. Abuammah, A.; Maimari, N.; Towhidi, L.; Frueh, J.; Chooi, K.Y.; Warboys, C.; Krams, R. New developments in mechanotransduction: Cross talk of the Wnt, TGF-β and Notch signalling pathways in reaction to shear stress. Curr. Opin. Biomed. Eng. 2018, 5, 96–104. [Google Scholar] [CrossRef]
  169. Chachisvilis, M.; Zhang, Y.-L.; Frangos, J.A. G protein-coupled receptors sense fluid shear stress in endothelial cells. Proc. Natl. Acad. Sci. USA 2006, 103, 15463–15468. [Google Scholar] [CrossRef]
  170. Choi, H.Y.; Yang, G.-M.; Dayem, A.A.; Saha, S.K.; Kim, K.; Yoo, Y.; Hong, K.; Kim, J.-H.; Yee, C.; Lee, K.-M.; et al. Hydrodynamic shear stress promotes epithelial-mesenchymal transition by downregulating ERK and GSK3β activities. Breast Cancer Res. 2019, 21, 6. [Google Scholar] [CrossRef]
  171. Das, J.; Maji, S.; Agarwal, T.; Chakraborty, S.; Maiti, T.K. Hemodynamic shear stress induces protective autophagy in HeLa cells through lipid raft-mediated mechanotransduction. Clin. Exp. Metastasis 2018, 35, 135–148. [Google Scholar] [CrossRef]
  172. Kuipers, A.J.; Middelbeek, J.; van Leeuwen, F.N. Mechanoregulation of cytoskeletal dynamics by TRP channels. Eur. J. Cell Biol. 2012, 91, 834–846. [Google Scholar] [CrossRef]
  173. Li, H.; Zhou, W.-Y.; Xia, Y.-Y.; Zhang, J.-X. Endothelial Mechanosensors for Atheroprone and Atheroprotective Shear Stress Signals. J. Inflamm. Res. 2022, 15, 1771–1783. [Google Scholar] [CrossRef]
  174. Park, M.G.; Jang, H.; Lee, S.-H.; Lee, C.J. Flow Shear Stress Enhances the Proliferative Potential of Cultured Radial Glial Cells Possibly Via an Activation of Mechanosensitive Calcium Channel. Exp. Neurobiol. 2017, 26, 71–81. [Google Scholar] [CrossRef] [PubMed]
  175. Simfia, I.; Schiavi, J.; McNamara, L.M. Alterations in osteocyte mediated osteoclastogenesis during estrogen deficiency and under ROCK-II inhibition: An in vitro study using a novel postmenopausal multicellular niche model. Exp. Cell Res. 2020, 392, 112005. [Google Scholar] [CrossRef] [PubMed]
  176. Yuan, L.; Sakamoto, N.; Song, G.; Sato, M. Low-Level Shear Stress Induces Human Mesenchymal Stem Cell Migration Through the SDF-1/CXCR4 Axis Via MAPK Signaling Pathways. Stem Cells Dev. 2013, 22, 2384–2393. [Google Scholar] [CrossRef]
  177. Belosludtseva, N.V.; Uryupina, T.A.; Pavlik, L.L.; Mikheeva, I.B.; Talanov, E.Y.; Venediktova, N.I.; Serov, D.A.; Stepanov, M.R.; Ananyan, M.A.; Mironova, G.D. Pathological Alterations in Heart Mitochondria in a Rat Model of Isoprenaline-Induced Myocardial Injury and Their Correction with Water-Soluble Taxifolin. Int. J. Mol. Sci. 2024, 25, 11596. [Google Scholar] [CrossRef] [PubMed]
  178. Rubin, D.M.; Anderton, N.; Smalberger, C.; Polliack, J.; Nathan, M.; Postema, M. On the Behaviour of Living Cells under the Influence of Ultrasound. Fluids 2018, 3, 82. [Google Scholar] [CrossRef]
  179. Gudkov, S.V.; Ivanov, V.E.; Matveeva, T.A.; Sarimov, R.M.; Lomonosov, A.M.; Astashev, M.E.; Baimler, I.V.; Simakin, A.V.; Bunkin, N.F.; Pustovoy, V.I.; et al. Physical and chemical changes that occur in water when drops fall on a hard surface (rain, waterfall, shower). J. Mol. Liq. 2024, 401, 124686. [Google Scholar] [CrossRef]
  180. Penkov, N.V. Temporal Dynamics of the Scattering Properties of Deionized Water. Phys. Wave Phenom. 2020, 28, 135–139. [Google Scholar] [CrossRef]
  181. Chernikov, A.V.; Gudkov, S.V.; Shtarkman, I.N.; Bruskov, V.I. Oxygen effect in heat-mediated damage to DNA. Biofizika 2007, 52, 244–251. [Google Scholar]
  182. Sharapov, M.G.; Gudkov, S.V.; Lankin, V.Z.; Novoselov, V.I. Role of Glutathione Peroxidases and Peroxiredoxins in Free Radical-Induced Pathologies. Biochemistry 2021, 86, 1418–1433. [Google Scholar] [CrossRef]
  183. Sharapov, M.G.; Gudkov, S.V.; Lankin, V.Z. Hydroperoxide-Reducing Enzymes in the Regulation of Free-Radical Processes. Biochemistry 2021, 86, 1256–1274. [Google Scholar] [CrossRef]
  184. Lyakhov, G.A.; Man’ko, V.I.; Suyazov, N.V.; Shcherbakov, I.A.; Shermeneva, M.A. Physical Mechanisms of Activation of Radical Reactions in Aqueous Solutions under Mechanical and Magnetic Effect: Problem of Singlet Oxygen. Phys. Wave Phenom. 2022, 30, 174–181. [Google Scholar] [CrossRef]
  185. Minaev, B.F. Electronic mechanisms of activation of molecular oxygen. Russ. Chem. Rev. 2007, 76, 1059–1083. [Google Scholar] [CrossRef]
  186. Lyakhov, G.A.; Man’ko, V.I.; Shcherbakov, I.A.; Suyazov, N.V. Impact of Classical Vibrations and Magnetic Fields on Quantum Objects. Phys. Wave Phenom. 2024, 32, 67–72. [Google Scholar] [CrossRef]
  187. Lyakhov, G.A.; Man’ko, V.I.; Shcherbakov, I.A. Action of Classical Fields on Quantum Systems within the Schrödinger–Robertson Uncertainty Relation. Phys. Wave Phenom. 2022, 30, 169–173. [Google Scholar] [CrossRef]
  188. Shcherbakov, I.A. Current Trends in the Studies of Aqueous Solutions. Phys. Wave Phenom. 2022, 30, 129–134. [Google Scholar] [CrossRef]
  189. Gudkov, S.V.; Karp, O.E.; Garmash, S.A.; Ivanov, V.E.; Chernikov, A.V.; Manokhin, A.A.; Astashev, M.E.; Yaguzhinsky, L.S.; Bruskov, V.I. Generation of reactive oxygen species in water under exposure to visible or infrared irradiation at absorption bands of molecular oxygen. Biophysics 2012, 57, 1–8. [Google Scholar] [CrossRef]
  190. Gudkov, S.V.; Bruskov, V.I.; Astashev, M.E.; Chernikov, A.V.; Yaguzhinsky, L.S.; Zakharov, S.D. Oxygen-Dependent Auto-Oscillations of Water Luminescence Triggered by the 1264 nm Radiation. J. Phys. Chem. B 2011, 115, 7693–7698. [Google Scholar] [CrossRef]
  191. Bruskov, V.I. Heat-induced formation of reactive oxygen species and 8-oxoguanine, a biomarker of damage to DNA. Nucleic Acids Res. 2002, 30, 1354–1363. [Google Scholar] [CrossRef]
  192. Bruskov, V.I.; Chernikov, A.V.; Ivanov, V.E.; Karmanova, E.E.; Gudkov, S.V. Formation of the Reactive Species of Oxygen, Nitrogen, and Carbon Dioxide in Aqueous Solutions under Physical Impacts. Phys. Wave Phenom. 2020, 28, 103–106. [Google Scholar] [CrossRef]
  193. Kloss, A. Electron-radical dissociation and water activation mechanism. Dokl. Akad. Nauk SSSR 1988, 303, 1403–1407. [Google Scholar]
  194. Lavrik, N.L.; Bazhin, N.M. On the impossibility of OH radical formation from the hydroxyl ion in water. Biophysics 2011, 56, 535–536. [Google Scholar] [CrossRef]
  195. Ward, J.F. DNA Damage Produced by Ionizing Radiation in Mammalian Cells: Identities, Mechanisms of Formation, and Reparability. Prog. Nucleic Acid Res. Mol. Biol. 1988, 35, 95–125. [Google Scholar] [PubMed]
  196. Draganić, I.G. Radiolysis of water: A look at its origin and occurrence in the nature. Radiat. Phys. Chem. 2005, 72, 181–186. [Google Scholar] [CrossRef]
  197. Sergeichev, K.F.; Lukina, N.A.; Sarimov, R.M.; Smirnov, I.G.; Simakin, A.V.; Dorokhov, A.S.; Gudkov, S.V. Physicochemical Properties of Pure Water Treated by Pure Argon Plasma Jet Generated by Microwave Discharge in Opened Atmosphere. Front. Phys. 2021, 8, 614684. [Google Scholar] [CrossRef]
  198. Barmina, E.V.; Gudkov, S.V.; Simakin, A.V.; Shafeev, G.A. Stable Products of Laser-Induced Breakdown of Aqueous Colloidal Solutions of Nanoparticles. J. Laser Micro/Nanoeng. 2017, 12, 254–257. [Google Scholar] [CrossRef]
  199. Novakovskaya, Y.V. Hydrogen-bond Network of Water and Irradiation Effects. Phys. Wave Phenom. 2020, 28, 161–167. [Google Scholar] [CrossRef]
  200. Colussi, A.J. Mechanism of Hydrogen Peroxide Formation on Sprayed Water Microdroplets. J. Am. Chem. Soc. 2023, 145, 16315–16317. [Google Scholar] [CrossRef]
  201. Sarimov, R.M.; Nagaev, E.I.; Matveyeva, T.A.; Binhi, V.N.; Burmistrov, D.E.; Serov, D.A.; Astashev, M.E.; Simakin, A.V.; Uvarov, O.V.; Khabatova, V.V.; et al. Investigation of Aggregation and Disaggregation of Self-Assembling Nano-Sized Clusters Consisting of Individual Iron Oxide Nanoparticles upon Interaction with HEWL Protein Molecules. Nanomaterials 2022, 12, 3960. [Google Scholar] [CrossRef]
  202. Moore, K.D.; Wu, J.Z.; Krajnak, K.; Warren, C.; Dong, R.G. Quantification of mechanical behavior of rat tail under compression. Bio-Med. Mater. Eng. 2024, 35, 337–349. [Google Scholar] [CrossRef]
  203. Yarasheski, K.E.; Lemon, P.W.; Gilloteaux, J. Effect of heavy-resistance exercise training on muscle fiber composition in young rats. J. Appl. Physiol. 1990, 69, 434–437. [Google Scholar] [CrossRef]
  204. Odintsov, S.D.; Oikonomou, V.K. Towards Modelling Mechanical Shaking Using Potential Energy Surfaces: A Toy Model Analysis. Symmetry 2024, 16, 572. [Google Scholar] [CrossRef]
  205. Orgel, L.E. The origin of life—A review of facts and speculations. Trends Biochem. Sci. 1998, 23, 491–495. [Google Scholar] [CrossRef] [PubMed]
  206. Reynolds, R.P.; Norton, J.N. Evaluation of Vibration as an Extrinsic Variable in In Vivo Research. J. Am. Assoc. Lab. Anim. Sci. 2024, 63, 107–115. [Google Scholar] [CrossRef]
  207. Na, S.; Wang, N. Application of Fluorescence Resonance Energy Transfer and Magnetic Twisting Cytometry to Quantify Mechanochemical Signaling Activities in a Living Cell. Sci. Signal. 2008, 1, pl1. [Google Scholar] [CrossRef]
  208. Cross, R.A.; McAinsh, A.D.; Straube, A. Mechanochemical cell biology. Semin. Cell Dev. Biol. 2011, 22, 913–915. [Google Scholar] [CrossRef]
  209. Yang, H.; Chen, W.; Peczulis, P.; Heng, J.Y.Y. Development and Workflow of a Continuous Protein Crystallization Process: A Case of Lysozyme. Cryst. Growth Des. 2019, 19, 983–991. [Google Scholar] [CrossRef]
  210. Hansson, H.A.; Dahlin, L.B.; LÖWenadler, B.; Lundborg, G.; Paleus, S.; Skottneru, A. Transient increase in insulin-like growth factor I immunoreactivity in rat peripheral nerves exposed to vibrations. Acta Physiol. Scand. 2008, 132, 35–41. [Google Scholar] [CrossRef]
  211. Zhang, C.; Li, J.; Zhang, L.; Zhou, Y.; Hou, W.; Quan, H.; Li, X.; Chen, Y.; Yu, H. Effects of mechanical vibration on proliferation and osteogenic differentiation of human periodontal ligament stem cells. Arch. Oral Biol. 2012, 57, 1395–1407. [Google Scholar] [CrossRef]
  212. Miller, D.L. Cell death thresholds in Elodea for 0.45–10 MHz ultrasound compared to gas-body resonance theory. Ultrasound Med. Biol. 1979, 5, 351–357. [Google Scholar] [CrossRef]
  213. Williams, A.R.; Sykes, S.M.; O’Brien, W.D. Ultrasonic exposure modifies platelet morphology and function in vitro. Ultrasound Med. Biol. 1977, 2, 311–317. [Google Scholar] [CrossRef]
  214. Tachibana, K.; Uchida, T.; Ogawa, K.; Yamashita, N.; Tamura, K. Induction of cell-membrane porosity by ultrasound. Lancet 1999, 353, 1409. [Google Scholar] [CrossRef] [PubMed]
  215. Lagneaux, L.; de Meulenaer, E.C.; Delforge, A.; Dejeneffe, M.; Massy, M.; Moerman, C.; Hannecart, B.; Canivet, Y.; Lepeltier, M.-F.; Bron, D. Ultrasonic low-energy treatment. Exp. Hematol. 2002, 30, 1293–1301. [Google Scholar] [CrossRef] [PubMed]
  216. Pitt, W.G.; Ross, S.A. Ultrasound Increases the Rate of Bacterial Cell Growth. Biotechnol. Prog. 2003, 19, 1038–1044. [Google Scholar] [CrossRef] [PubMed]
  217. Mizrahi, N.; Zhou, E.H.; Lenormand, G.; Krishnan, R.; Weihs, D.; Butler, J.P.; Weitz, D.A.; Fredberg, J.J.; Kimmel, E. Low intensity ultrasound perturbs cytoskeleton dynamics. Soft Matter 2012, 8, 2438–2443. [Google Scholar] [CrossRef]
Figure 1. Scheme of reactions leading to the formation of hydrogen peroxide at mechanical impact on water, aqueous solutions, and aqueous colloids. MI is mechanical impact. Scenario 1: activation of molecular oxygen. Scenario 2: oxidation of anions.
Figure 1. Scheme of reactions leading to the formation of hydrogen peroxide at mechanical impact on water, aqueous solutions, and aqueous colloids. MI is mechanical impact. Scenario 1: activation of molecular oxygen. Scenario 2: oxidation of anions.
Ijms 26 05556 g001
Figure 2. Dependences of the thresholds of the mechanical influence characteristics on all analysed systems: Dot plots (a) and 3D-map (b) of the characteristics ‘frequency-acceleration-duration’ at which effects were observed (irrespective of their magnitude). Each point corresponds to one set of published data (one row in Appendix A Table A1). A set of 339 published data sets was analyzed.
Figure 2. Dependences of the thresholds of the mechanical influence characteristics on all analysed systems: Dot plots (a) and 3D-map (b) of the characteristics ‘frequency-acceleration-duration’ at which effects were observed (irrespective of their magnitude). Each point corresponds to one set of published data (one row in Appendix A Table A1). A set of 339 published data sets was analyzed.
Ijms 26 05556 g002
Figure 3. Dependences of the magnitude of the effects of mechanical influence on all analyzed systems: Dot plots of the characteristics ‘frequency-acceleration-effect’ (a), ‘acceleration-duration-effect’ (c), and ‘force-duration-effect’ (e). Each point corresponds to an individual parameter value in published papers. The color of the dot corresponds to the magnitude of the modulus of the effect relative to the control taken modulo. 3D maps of the distribution of the magnitude of the averaged effect sizes for the characteristics ‘frequency-acceleration-effect’ (b), ‘acceleration-duration-effect’ (d), and ‘force-duration-effect’ (f). Each point corresponds to one set of published data (row in Appendix A Table A1). A set of 339 published data sets was analyzed.
Figure 3. Dependences of the magnitude of the effects of mechanical influence on all analyzed systems: Dot plots of the characteristics ‘frequency-acceleration-effect’ (a), ‘acceleration-duration-effect’ (c), and ‘force-duration-effect’ (e). Each point corresponds to an individual parameter value in published papers. The color of the dot corresponds to the magnitude of the modulus of the effect relative to the control taken modulo. 3D maps of the distribution of the magnitude of the averaged effect sizes for the characteristics ‘frequency-acceleration-effect’ (b), ‘acceleration-duration-effect’ (d), and ‘force-duration-effect’ (f). Each point corresponds to one set of published data (row in Appendix A Table A1). A set of 339 published data sets was analyzed.
Ijms 26 05556 g003aIjms 26 05556 g003b
Figure 4. Dependences of the effects of mechanical impact on living (red circles) and non-living (blue hexagons) systems: (a) thresholds of the values of frequency, acceleration, and duration, at which the effects of mechanical impact were observed, and dependence of ‘frequency-acceleration-effect’ (b), ‘acceleration-duration-effect’ (c) and ‘force-duration-effect’ (d) (see Table 1) on living (red circles) and non-living (blue hexagons) systems. The legend shows color scales with the maximum and minimum values marked for each group compared. Each point corresponds to one set of published data (one row in Appendix A Table A1). A set of 339 published data sets was analyzed.
Figure 4. Dependences of the effects of mechanical impact on living (red circles) and non-living (blue hexagons) systems: (a) thresholds of the values of frequency, acceleration, and duration, at which the effects of mechanical impact were observed, and dependence of ‘frequency-acceleration-effect’ (b), ‘acceleration-duration-effect’ (c) and ‘force-duration-effect’ (d) (see Table 1) on living (red circles) and non-living (blue hexagons) systems. The legend shows color scales with the maximum and minimum values marked for each group compared. Each point corresponds to one set of published data (one row in Appendix A Table A1). A set of 339 published data sets was analyzed.
Ijms 26 05556 g004
Figure 5. Dependences of the moduli of the effects of mechanical impact on water and aqueous solutions on the magnitude of frequency (orange) (a), acceleration (brown) (b), duration (green) (c), and force (cyan) (d). Each point corresponds to one set of published data (one row in Appendix A Table A1). The dash lines indicate fitting curves. A set of 192 published data sets was analyzed.
Figure 5. Dependences of the moduli of the effects of mechanical impact on water and aqueous solutions on the magnitude of frequency (orange) (a), acceleration (brown) (b), duration (green) (c), and force (cyan) (d). Each point corresponds to one set of published data (one row in Appendix A Table A1). The dash lines indicate fitting curves. A set of 192 published data sets was analyzed.
Ijms 26 05556 g005
Figure 6. Dependences of the moduli of the effects of mechanical impact on living systems (cells and organisms) on the magnitude of frequency (frequency) (a), acceleration (brown) (b), duration (green) (c), and force (cyan) (d). The dash lines indicate fitting curves. Each point corresponds to one set of published data (one row in Appendix A Table A1). A set of 174 published data sets was analyzed.
Figure 6. Dependences of the moduli of the effects of mechanical impact on living systems (cells and organisms) on the magnitude of frequency (frequency) (a), acceleration (brown) (b), duration (green) (c), and force (cyan) (d). The dash lines indicate fitting curves. Each point corresponds to one set of published data (one row in Appendix A Table A1). A set of 174 published data sets was analyzed.
Ijms 26 05556 g006
Figure 7. Dependences of the magnitude of ultrasonic influence effects on all analyzed systems: Dot plots of the characteristics “intensity-duration-effect” (a), “frequency-intensity-effect” (c). Each dot plot corresponds to an individual parameter value in published papers. The color of the point corresponds to the magnitude of the modulus of the effect relative to the control taken modulo. 3D maps of the distribution of the magnitude of the averaged effect sizes for the characteristics “intensity-duration-effect” (b), “frequency-intensity-effect” (d). In each variant, 225 results described in the literature were analysed.
Figure 7. Dependences of the magnitude of ultrasonic influence effects on all analyzed systems: Dot plots of the characteristics “intensity-duration-effect” (a), “frequency-intensity-effect” (c). Each dot plot corresponds to an individual parameter value in published papers. The color of the point corresponds to the magnitude of the modulus of the effect relative to the control taken modulo. 3D maps of the distribution of the magnitude of the averaged effect sizes for the characteristics “intensity-duration-effect” (b), “frequency-intensity-effect” (d). In each variant, 225 results described in the literature were analysed.
Ijms 26 05556 g007
Figure 8. Correlations of frequency (orange) (a), intensity (cyan) (b), and duration (green) (c) of ultrasonic exposure and magnitude of effects on all analyzed systems. In each variant, 225 results published in the literature were analyzed. The dash lines indicate fitting curves.
Figure 8. Correlations of frequency (orange) (a), intensity (cyan) (b), and duration (green) (c) of ultrasonic exposure and magnitude of effects on all analyzed systems. In each variant, 225 results published in the literature were analyzed. The dash lines indicate fitting curves.
Ijms 26 05556 g008
Table 1. Effect of mechanical impacts on different levels of living systems.
Table 1. Effect of mechanical impacts on different levels of living systems.
SystemLevelReactionBenefitReferences
CardiovascularMolecularChange Ca2+ responses to AChR ligands↓ or ↑ depend on conditions[70,71]
CellsMicro-damage of endothelial cells damage[41]
Calcium homeostasis of endotheliocytes changes[70,71]
TissuesNumber of capillaries, venules, and arterioles reduction[20]
Smooth muscle work of microvessels[40]
OrganismsHR and blood pressure, Raynaud’s syndrome, etc.[11,39]
Immune and blood cellsMolecularNFAT activity, IL-6 synthesis, ROS production, and t-CPK activity increasing[40,41,42,43]
Decreasing the synthesis of pro-inflammatory IFN-β, RANKL, and TNF-α, increased oxygen consumption, and gene expression of glycolysis enzymes and transporters (Hk2, GLUT1)[62,66,72]
CellsErythrocytes hemolysis rate increasing[57]
T-cell proliferation altered, ‘neutrophil respiratory burst’ (ROS productions), reduction of differentiated dendritic cell proliferation↓ or ↑ depend on frequency and conditions[32,33,56,72]
TissuesImmune memory and antitumor activity of dendric cells[72]
OrganismsAnti-inflammatory effect[63]
Proinflammatory effect↓ or ↑ depend on conditions[32,33]
MusculoskeletalMolecularTRAP-dependent pathway activation, increasing of ACAN and SOX9 expression, increasing of collagen content[44,45,59,67]
P1NP and CTx concentration increase, osteogenesis regulators MMP 2 and OSX expression decreasing. Decreasing of osteocalcin levels, Runx2 and Osx gene expression, synthesis of osteoarthritis markers collagen II and aggrecan[26,28,46,47,49,50,58,64,65]
CellsOsteoclast proliferation, differentiations and activity increasing, increasing of differentiation of primary fibro chondrocytes[26,44,45,59,60,61]
The viability of osteoblasts decreases; decrease of osteoblast differentiation[26,58]
TissuesImproving bone mechanical characteristics and bone structure[28,44,45]
Bone tissue damage[42,43]
OrganismsOsteoporosis markers P1NP and CTx levels[46,47]
Improvement bone-implant contact, obesity decreasing, and muscle strength increasing[28,48,51]
RespiratoryOrganismsInspiratory flow rate, inspired volume, and expired volume[52]
NervousMolecularIncreasing the expression of the IGF-I factor[54]
CellsImprove of myelinization[53]
OtherMolecularIncreasing in A cytochrome P450 monooxygenase CYP1B1 expression and metabolic activity changes[69]
CellsStem cell proliferation inhibition[25]
↑ and ↓ indicate possible positive or negative effects on organisms, respectively.
Table 2. Effect of US impacts on different levels of living systems.
Table 2. Effect of US impacts on different levels of living systems.
SystemLevelReactionBenefitReferences
CardiovascularMolecularEnhancing VEGF secretion[102]
TissuesEnhance angiogenesis [102,106]
Immune and blood cellsMolecularSecretion of proinflammatory cytokines IL-1β, IL-8[102]
Increasing the number of chromosomal aberrations[111]
CellsInduction of Ca2+ influx and platelet aggregation; induction of apoptosis of leucocyte cells[108,109,110,111,112]
Organismsinhibition of pro-inflammatory reactions↑ or ↓ depend on situation[113,115]
MusculoskeletalMolecularIncreased expression of collagen, expression of myogenin, Pax7, COX2[102,113,115]
CellsIncreasing of myocyte proliferation, myoblast differentiation[97,98,104]
TissuesIncreasing of length of myotubes and the number of cells fused into one myotube, and regeneration of myofibrils[97,98,104,113,114]
NervousMolecularEnhanced expression of markers Alk, Cenpf, Pcdh17 and actin[116]
CellsPromotions of proliferation and differentiation of neurons, formation of outgrowths[116]
OtherMolecularAcceleration of drug delivery, EGFb regulatory factor expression increasing, change Cyclin D1/β, Cyclin B1/β, Cyclin E1/β, and Cyclin A1/β levels[99,102,103,105]
Accelerate DNA plasmid transfer in bacterial cells[119]
Plants photosynthesis inhibition[120]
CellsER microstructure damage[100]
Stem cell differentiation and proliferation increasing[102,103,104,107]
Destruction of bacterial spores and cells[117,118]
↑ and ↓ indicate possible positive or negative effects on organisms, respectively.
Table 3. Spearman correlation coefficients calculated for mechanical effects.
Table 3. Spearman correlation coefficients calculated for mechanical effects.
CoefficientsMechanical Impact ParametersEffect log10
In VivoIn Vitro
Spearman’s correlationfrequency, Hz (log10)0.1730.196
acceleration, m/s2 (log10)0.2530.263
duration, s (log10)0.0520.11407
force, N (log10)0.195−0.06187
p-valuefrequency, Hz (log10)0.023440.01622
acceleration, m/s2 (log10)7.76 × 10−40.00113
duration, s (log10)0.496270.166
force, N (log10)0.013410.45196
The color scale in Spearman’s correlation corresponds to smaller negative (red) and larger positive (blue) values of Spearman’s Corr. The color scale in the p-value section shows significant (green, p < 0.05) and insignificant (orange, p ≥ 0.05) correlations.
Table 4. Spearman correlation coefficients between effect size and KM characteristics.
Table 4. Spearman correlation coefficients between effect size and KM characteristics.
CoefficientsUS Impact ParametersEffect log10
Spearman’s correlationfrequency, Hz (log10)0.1086
intensity, mW/cm2 (log10)_0.38302
duration, s (log10)−0.27719
p-valuefrequency, Hz (log10)0.10657
intensity, mW/cm2 (log10)_3.60925 × 10−9
duration, s (log10)2.8004 × 10−5
The color scale in Spearman’s correlation corresponds to negative (red) and larger positive (blue) values of Spearman’s Corr. The color scale in the p-value section shows significant (green, p < 0.05) and insignificant (orange, p ≥ 0.05) correlations.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gudkov, S.V.; Pustovoy, V.I.; Sarimov, R.M.; Serov, D.A.; Simakin, A.V.; Shcherbakov, I.A. Diversity of Effects of Mechanical Influences on Living Systems and Aqueous Solutions. Int. J. Mol. Sci. 2025, 26, 5556. https://doi.org/10.3390/ijms26125556

AMA Style

Gudkov SV, Pustovoy VI, Sarimov RM, Serov DA, Simakin AV, Shcherbakov IA. Diversity of Effects of Mechanical Influences on Living Systems and Aqueous Solutions. International Journal of Molecular Sciences. 2025; 26(12):5556. https://doi.org/10.3390/ijms26125556

Chicago/Turabian Style

Gudkov, Sergey V., Vladimir I. Pustovoy, Ruslan M. Sarimov, Dmitriy A. Serov, Alexander V. Simakin, and Ivan A. Shcherbakov. 2025. "Diversity of Effects of Mechanical Influences on Living Systems and Aqueous Solutions" International Journal of Molecular Sciences 26, no. 12: 5556. https://doi.org/10.3390/ijms26125556

APA Style

Gudkov, S. V., Pustovoy, V. I., Sarimov, R. M., Serov, D. A., Simakin, A. V., & Shcherbakov, I. A. (2025). Diversity of Effects of Mechanical Influences on Living Systems and Aqueous Solutions. International Journal of Molecular Sciences, 26(12), 5556. https://doi.org/10.3390/ijms26125556

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop