Next Article in Journal
Cloud-Based Machine Learning Application for Predicting Energy Consumption in Automotive Spot Welding
Previous Article in Journal
A Systematic Review on the Synthesis of Silicon Carbide: An Alternative Approach to Valorisation of Residual Municipal Solid Waste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Hypomagnetic Fields and Their Multilevel Effects on Living Organisms

by
Miroslava Sinčák
and
Jana Sedlakova-Kadukova
*
Faculty of Natural Science, University of Cyril and Methodius in Trnava, Nam. J. Herdu 2, 91701 Trnava, Slovakia
*
Author to whom correspondence should be addressed.
Processes 2023, 11(1), 282; https://doi.org/10.3390/pr11010282
Submission received: 13 December 2022 / Revised: 9 January 2023 / Accepted: 12 January 2023 / Published: 16 January 2023
(This article belongs to the Topic Energy Efficiency, Environment and Health)

Abstract

:
The Earth’s magnetic field is one of the basic abiotic factors in all environments, and organisms had to adapt to it during evolution. On some occasions, organisms can be confronted with a significant reduction in a magnetic field, termed a “hypomagnetic field—HMF”, for example, in buildings with steel reinforcement or during interplanetary flight. However, the effects of HMFs on living organisms are still largely unclear. Experimental studies have mostly focused on the human and rodent models. Due to the small number of publications, the effects of HMFs are mostly random, although we detected some similarities. Likely, HMFs can modify cell signalling by affecting the contents of ions (e.g., calcium) or the ROS level, which participate in cell signal transduction. Additionally, HMFs have different effects on the growth or functions of organ systems in different organisms, but negative effects on embryonal development have been shown. Embryonal development is strictly regulated to avoid developmental abnormalities, which have often been observed when exposed to a HMF. Only a few studies have addressed the effects of HMFs on the survival of microorganisms. Studying the magnetoreception of microorganisms could be useful to understand the physical aspects of the magnetoreception of the HMF.

1. Introduction

Every living organism on Earth has adapted to the geomagnetic field during an evolutionary process lasting billions of years. The presence of a geomagnetic field (approximately 50 uT) is natural to each cell [1]. However, in a few circumstances, organisms can face the absence of magnetic fields. Understanding its effect can enhance our knowledge of magnetoreception mechanisms, with applications in space research, biotechnology or medicine. The terms “hypomagnetic”, “conditionally zero magnetic field” or “magnetic vacuum” generally refer to fields with a magnetic flux density (B) below 100 nT [2], but according to some authors, we can speak of a magnetic field weaker than 5 µT as being hypomagnetic [3].
Hypomagnetic fields (HMFs) commonly occur in the interplanetary space of the solar system and fluctuate in the range of several nanoteslas (nT). For example, the lunar magnetic field is less than 300 nT, and the magnetic field on Mars is approximately 1 µT [3]. The planetary magnetic field of Mars is extremely small, and the planetary magnetic field of Venus is practically non-existing [4] (Figure 1).
New technologies are currently being developed to enable space exploration and interplanetary flights. In the future, organisms will be exposed to a HMF during space travels, which is significantly weaker than the geomagnetic field (GMF) and expected to have diverse biological effects. During these travels, organisms will be exposed to tedious periods of a HMF that is approximately 10,000 times weaker than the Earth’s magnetic field, ranging from 0.1 to 1 nT [5]. However, attenuation of the Earth’s magnetic field is not limited to staying in space but occurs in daily life, for example, in buildings with steel walls or steel reinforcement [2]. Building walls are a natural shield against low- and high-frequency electromagnetic fields. However, a magnetic field (such as a geomagnetic field) is more difficult to shield. In contrast to radiofrequency and low-frequency electric fields, thin sheets of metal have no effect on magnetic fields [6]. However, there is evidence that buildings with steel in their construction magnetise and deform the natural geomagnetic field [5], causing an even 50-fold magnetic field attenuation according to the building size and the complexity of the steel [7].
Hypomagnetic fields can have various effects on organisms, although the underlying mechanisms remain unknown. Erdman et al. [8] suggest that the magnetoreception of the HMF differs among different organisms. The authors assume that the magnetoreception of the HMF is a nonspecific mechanism and manifests in highly different biological systems as mostly random reactions as a result of magnetic interaction with magnetic moments at a physical level. This moment, which is present in each molecule, could transfer the magnetic signal at the level of downstream biochemical events [2].
In this study, we summarise information about the observed biological effects of the HMF on eukaryotic and prokaryotic organisms and show the possible underlying mechanisms.

2. Materials and Methods

We used the scientific databases PubMed and ScienceDirect to select papers that contained the terms “hypomagnetic”, “magnetic zero”, “magnetic vacuum” or “magnetic shielding” in their titles, abstracts and keywords. This gave, after a subsequent semantic control, 65 experimental and theoretical articles that included original results suitable for further investigation. This type of search was repeated with each new relevant article iteratively until no new articles could be detected.

3. Mechanisms of Magnetoreception of HMFs

Magnetoreception is the universal ability of a biological system to detect magnetic and electromagnetic fields, although it may manifest itself differently in different organisms. Any changes in magnetic field intensity may affect the organisms in many ways, including the basic metabolism of prokaryotic and eukaryotic cells [8]. Magnetoreception relates not only to geomagnetic fields and higher magnetic and electromagnetic fields, but it also explains the perception of HMFs. The hypothesis of nonspecific nonthermal magnetoreception on the physical level has not been studied since none of those has yet been identified experimentally. Typical for hypomagnetic magnetoreception experiments is a high sensitivity to the physical, chemical and physiological conditions, as well as a low reproducibility [2] and a great variety of effects in different organisms. It has not yet been possible to establish any common conditions controlling the magnetic effects in different organisms or populations rather than in their individual forms [9]. Several mechanisms have been described that could explain the mechanism of magnetoreception, such as the cyclotron resonance model, macroscopic charged vortices in the cytoplasm and the parametric resonance model, among others [10]. The most likely physical mechanisms with expected biological responses are: (i) the radical pair mechanism, (ii) the universal physical mechanism and (iii) the molecular gyroscope mechanism. However, according to Binhi and Prato [2], the radical pair mechanism is unlikely to explain all HMF effects on living organisms. The authors assume that the universal physical mechanism and the molecular gyroscope mechanism are more accurate.
These primary physical mechanisms can lead to secondary biophysical responses, which can include changes in ROS concentrations, Ca2+ ion homeostasis or influence enzymes that are involved in the electron transport chain in mitochondria or in cell cycle promotion.
1.
Radical pair mechanism
Traditionally, radicals (for example, reactive oxygen species (ROS)) are considered harmful because they can cause cell death via oxidative intracellular damage in the metabolism of sugars, fats and nucleic acids. Several studies have also shown the importance of ROS in intracellular signalling cascades such as apoptosis initiation [11,12,13]. Radicals are magnetic because an electron (along with a proton and a neutron) has a property known as spin or, more precisely, a spin momentum [14].
The radical pair consists of two radicals that have been formed simultaneously, usually by a chemical reaction. The spins of two unpaired electrons can be either parallel to each other (↑↑ which gives S = 1) or anti-parallel (↑↓, which gives S = 0, where S is the spin quantum number). The two forms of the electron pair are therefore known as triplet (S = 1) and singlet (S = 0) [15]. Influencing either singlet or triplet formations of electron pairs could be associated with the presence of an external magnetic field and leads to a longer life of the radical pairs (triplet states) [16].
This mechanism can cause a difference in the stability of radical pairs and affects the shift of the chemical reaction equilibrium. Thus, during the formation of radical pairs, external magnetic fields change the recombination rate of these radical pairs, which in turn changes the concentration of radicals such as O2 • and molecules such as H2O2 [17]. In general, the coupling between unpaired electrons and nuclei in each fragment of a radical pair can be achieved by magnetic fields in the range of 10 μT–3 mT [18]. Magnetic fields could interact with the magnetic moments of radical pairs at physical levels, which are ubiquitous in macromolecules with unpaired electrons, protons, paramagnetic ions or other magnetic nuclei in biological cells, and then transmit the magnetic signal to subsequent biochemical events such as cell oxidative stress reactions. This procedure would therefore lead to highly different biological observables and mostly random reactions [19].
This mechanism does not have frequency selectivity because the development of a magnetosensitive spin state occurs over an extremely short life of the radical pair, usually in the order of 10−9–10−7 s [20]. Many authors explain the observed results by this mechanism [21,22,23].
2.
Universal physical mechanism
The rotation of magnetic moments in a magnetic field precedes any biophysical or biochemical mechanism of magnetoreception and largely determines the spectral and nonlinear characteristics of the biological effect of the field. The mechanism is based on the external magnetic field, which influences the magnetic moment of the molecules and leads to the terminal relaxation of the magnetic moment [19]. Magnetic relaxation is known as the approach to equilibrium after a magnetic system was exposed to magnetic field change. Relaxation processes allow nuclear spins to return to equilibrium following a magnetic disturbance [24].
The biological effect is observed only when changes in the magnetic momentum dynamics go through the stages of transformation at the biochemical, physiological and biological levels of the system. A special characteristic of this mechanism is that it predicts the effects of weak magnetic fields but also those of electromagnetic fields induced by alternating electric currents (ACs) in the same biological system [2].
3.
Molecular gyroscope mechanism
The molecular gyroscope mechanism can be explained as the rotation of large fragments of macromolecules or amino acid residues with a distributed electric charge. This movement can be influenced by a magnetic field.
In some stages of protein assembly, in the final stage of their synthesis, virtual cavities without water molecules, of the order of 1 nm or less, may occur in the protein [25]. In these cavities, amino acid residues (molecular gyroscopes) rotate over milliseconds, searching for the best position. As a result of such rotation, a magnetic moment interacts with an external magnetic field [2]. The magnetic field affects these rotations, which results in possible changes in protein folding. The folding of protein chains is an evolutionarily conserved process, and improper folding can prevent a protein from performing its specific function [26]. Mostly, random changes in the proteome of the cell can explain various biological responses after HMF exposure.

4. Influences of HMFs on Organisms

In many areas, the biological effects of HMFs are contradictory, which might be explained by the length of exposure to the HMF. The authors of [27] reported that exposure to a HMF for a shorter time (1 h) could promote cell respiration, but a longer exposure time (6 h) has an inhibitory effect.
Another parameter causing conflicting results may be the method of generating the HMF. The authors used either the shielding of the present HMF or its compensation by another magnetic field, which may have caused a different result. For example, the production of free radicals caused by direct-current (DC) HMFs differs from the effect of AC HMFs. A similar difference was observed when the HMF was induced by a static field or a variable frequency-alternating magnetic field [28].
The type of organism is an important factor of the HMF effect [29]. Not only does the biological effect of HMFs vary between plant and animal cells, but according to Binhi and Prato [2], there are different targets of HMFs for different organisms or even for individuals of the same species. The observed effect of the HMF differs between eukaryotic and prokaryotic organisms, even in plant and animal cells. Few effects of HMF exposure could be similar for various organisms and are on the level of individual ions and proteins; they are generally related to cell signalling.
Regarding the spectrum of the hypomagnetic effect in various types of organisms according to their basic differences in structure and life cycle, we will separately discuss plant, animal and prokaryotic organisms.

4.1. Animals and Animal Cell Cultures

The observed effects can differ at various levels of the organization of the living organism. According to the observations described in various studies, we will discuss cell transport and respiration in a separate subsection (Section 4.1.1), and subsequently, we will discuss animals at the level of the organism or organ systems (Section 4.1.2).

4.1.1. Cell Transport and Respiration

The effects of the HMF on a single-cell level may include the effect on ion transport and concentration as well as cellular respiration (Table 1).
The cellular transport mechanisms of various nutrients can be affected by near-zero magnetic field exposure. Some studies have reported changes in the Ca2+ ion concentration in the cytosol after being subjected to hypomagnetic conditions. The effect of the HMF on Ca2+ ion concentration in tissues is the basis of the parametrical resonance theory [35], which deals with magnetoreception; it is caused by the effect of HMF on Ca2+ ions and proteins with Ca2+ binding sites. This theory agrees with the results of Kantserova et al. [32], who showed that the production of Ca2+-dependent proteases was inactivated after HMF exposure. It can be assumed that the inhibition of Ca2+-dependent enzymes under hypomagnetic conditions may negatively affect the basic calcium-mediated transduction in the cell. In eukaryotic cells, the Ca2+ ion plays a role as a primary and secondary messenger, and Ca-dependent enzymes, including calcium-dependent kinases or proteases, may participate in cell membrane fusion, cell division and apoptosis [36].
Several studies have found that a stronger magnetic field (≥100 μT) can increase the levels of reactive oxygen species (ROS) [37,38,39], whereas the HMF can significantly decrease the level of ROS in cells [21,40]. There is experimental evidence of the correlation between HMF-induced changes in cellular ROS concentration and biological effects, such as cell growth in vitro [28]. Therefore, ROS may represent a potential target for the magnetic field, which may cause the modulus of biological functions [2]. The changes in ROS concentrations in the cell are directly related to the presence of free radicals, and the authors investigating ROS lean towards the theoretical radical pair mechanism as the magnetoreception mechanism of the HMF influence.
The main source of cellular ROS are the mitochondrial electron-transport chain complexes I, II and III, which are present in the inner mitochondrial membrane. Complexes I and II are the primary sources of O2• under either physiological or pathological conditions [40]. For an individual cell, the rate of ROS generation varies depending on the availability of cellular O2, the redox state of the electron carriers, the respiration rate, the state of the electron carrier, the mitochondrial inner membrane potential and the posttranslational modifications of the respiratory protein chain [28]. Mitochondria are the organelles most sensitive to HMF exposure due to their electron-transparent matrix and lower mitochondrial membrane potential in both plant and animal cells [41,42].
Ogneva et al. [27] reported a decrease in Drosophila melanogaster sperm cell respiration as a consequence of affecting the I. mitochondrial electrical transport chain complex after 6 h in a HMF. Mitochondria may also undergo morphological rearrangements under HMF conditions. In another study, the size and relative volume of mitochondria in plant cells increased and cristae size decreased after hypomagnetic field exposure, as described in [41].
However, the mechanism by which the level of ROS is modulated by the magnetic field remains unclear. It is assumed that weak magnetic fields can alter the free radical level response and, consequently, affect specific cellular functions and inhibit or reduce cell growth [43]. In addition to metabolic changes, it is possible to consider changes at the morphological level, namely the accumulation of lipid bodies, the development of a lytic compartment (vacuoles and cytosegresomes) and the reduction of phytoferritin in plastids after HMF exposure [41].

4.1.2. Animals

On the level of the whole organism, studies dealing with HMFs mostly focus on animals, humans, tissue cultures and embryos. The most common areas of study are the influences of HMFs on prenatal development as well as cardiovascular and nervous systems (Table 2).
Hypomagnetic fields can delay the development of insect eggs and nymphs, reduce the fetal size and body length, reduce female fertility in adult insects [41] and reduce the life span of daphnia [56]. Yan et al. [22] also reported negative effects on the mating ratio and developmental stages of insects (Mythimna separata) and on the foraging orientation of Nilaparvata lugens [57], but a stimulating effect on positive phototaxis and flight capacity of Sogatella furcifera [58]. Similar to insects, adverse effects on embryonal development in Xenopus laevis have been observed [59], along with the induced loss of the ability to bear offspring in pregnant mice [60]. Adverse effects of the HMF were observed even in the case of extremophilic invertebrates from the phylum Tardigrada. The obtained results showed that even partial isolation from the geomagnetic field has a negative effect on the anhydrobiotic (resting) stage of both tested species (Echiniscus testudo and Milcium inceptum). Both species exhibited lower survival rates during entering anhydrobiosis, in the anhydrobiotic state, and upon returning to the active state. The authors also observed higher mortality in E. testudo compared to M. inceptum, which suggests that different species respond to hypomagnetic conditions in different ways [29]. Developmental abnormalities caused by HMFs may be related to epigenetic modifications of embryonic stem cells, such as abnormal DNA methylation. The results suggest that a suitable electromagnetic field may be necessary for favorable epigenetic remodelling and, thus, for differentiation during the embryonic stage [63].
An effect of HMFs on the nervous system has also been observed. The results suggest that specific brain structures represent neural substrates for the orientation of the magnetic compass in certain magnetosensory animals. In several experiments, HMFs accelerated the proliferation of neuroblastoma cells and neural progenitor/stem cells [42], and this proliferative effect may be related to decreased levels of cellular reactive oxygen species (ROS). After exposure of neuroblast cells to the HMF, a Warburg effect (commonly observed in cancer metabolism) was observed, when cell metabolism is induced by the repression of oxidative stress and the up-regulation of anaerobic glycolysis. In this case, the increased activity of LDH (lactate dehydrogenase), a key member of glycolysis, could be a direct response to a HMF [49]. The other explanation for the enhanced cell proliferation, according to Mo et al. [50], is the acceleration of proliferation by a forward shift of the cell cycle in the G1 phase. In contrast to the G1 phase, G2 and M phases were not affected during the experiment. The same results could be recorded when Belyaev et al. [62] observed that the effect of the zero magnetic field on chromatin condensation is more pronounced at the beginning of the G1 phase.
A comprehensive study examining the human transcriptome after exposure to a HMF (<200 nT) for 2 days showed a change in the gene expression of 2464 genes associated with the neural system. Mentioned genes were significantly grouped into a few key processes, for example, protein transport, macromolecule localization, RNA processing and brain function. These results suggest the involvement of the MAPK pathway and cryptochrome in the early biological responses to the presence of a HMF [45].
In addition to the effects on the neural system, effects of the HMF on the cardiovascular system have been observed. Capillary blood velocity increased by 17%, cardio intervals increased by 88.7% [51], and capillary circulation rate increased by 22.4% [52] during HMF exposure. At the end of exposure, diastolic blood pressure dropped considerably relative to mid-exposure values, whereas systolic blood pressure, on the contrary, showed a significant increase [52]. One of the crucial parameters which influence the observed effects of HMFs is exposure time. Both previous studies claim to have simulated hypomagnetic conditions during interplanetary flight, but the time of HMF exposure was only 60 min. We assume that the time of exposure was not sufficient to demonstrate hypomagnetic conditions during a longer stay in space. In comparison, in two studies with a longer exposure time of 72 h [53] and up to 4 weeks [54], the authors recorded an increase in haemolysis and the weakening of the deformation and aggregation properties of human blood, along with a reduction in enzymatic activities. The reduction of these enzyme activities and the promotion of haemolysis can be related to increased protein denaturation and decreased efficiency of the proteolytic system [53].

4.2. Plants

Recent studies have shown that plants respond to near-zero magnetic fields through morphological and developmental changes, including delays in flowering time and germination [64], breath conductivity, chlorophyll content [65], photoreceptor involvement [66] and changes in auxin [67], and gibberellin concentrations [68] (Table 3).
The HMF can either have inhibitory or stimulating effects on plants, depending on the part of the growth to which the plant is exposed. Hypomagnetic fields can inhibit [41] but also promote vegetative growth, e.g., by increasing the percentage of the germination rate [69]. On the other hand, they may have a reducing effect on reproductive growth by inhibiting seed production [70]. The magnetic field, in this case, is thought to affect the activity of cryptochromes and their gene expressions [64,74]. Plant hormones are also involved in cryptochrome-mediated flowering. Exposure to HMFs reduces the gibberellin content and the expression of their biosynthetic genes in wild-type Arabidopsis thaliana but not in the cryptochrome mutant strain (cry1/cry2). Similar results have been obtained for another plant hormone, auxin [67].
As in the case of animal cells, changes in the ion concentrations of some nutrients (NH4+, K+, Ca2+ Mg2+, Cl, SO42−, NO3 and PO43−) in plant cells after exposure of the A. thaliana root system to the HMF were recorded. A few minutes of exposure to a zero magnetic field resulted in a significant reduction in the intake of all studied nutrient ions, which can be explained by the existence of a plant magnetoreceptor responding to the HMF by modulating mineral nutrient transport genes. According to Narayan et al. [75], the response to an almost zero magnetic field is rapid, suggesting that some ion channels and all transport activities may not necessarily be related to gene expression. Ion channel changes have been reported in other studies and may influence flowering time [64], photoreceptor signaling [76], and seed germination [77].
In plant cells exposed to HMFs, the functional activity of the genome declined in the early pre-replication period. The HMF can intensify protein synthesis. At the ultrastructural level, changes in condensed chromatin distribution and nuclear compaction, the accumulation of lipid bodies, the development of the lytic compartment (vacuoles, cytosegresomas and paramural bodies), and the reduction of phytoferritin in plastids in meristem cells have been observed in pea roots [39].
In contrast to the animal cell, where the HMF stimulated proliferation and accelerated the passage through the G1 phase, the observed effect on the plant cell was the opposite. The HMF had a negative effect on the speed and progress of the cell cycle. The reproductive cycle of the cells slowed down due to the expansion of the G1 and G2 phases, whereas the other phases of the cell cycle remained relatively stable. The HMF also caused a remarkable decrease in proliferating plant cells (from 68% to 95%) [41].
Tsetlin et al. [78] also recorded remarkable results when they described a synergistic inhibitory effect of HMF and ionizing radiation (α and γ) on plant germination. This experiment simulated another environmental parameter to which the plants will be exposed during interplanetary flights.

4.3. Prokaryotes

Only a few studies have examined the effects of the HMF on microorganisms. Magnetotactic bacteria, i.e., bacteria capable of perceiving the Earth’s geomagnetic field by means of magnetosomes, have been investigated most frequently [5] (Table 4).
Studies on the magnetotactic bacterium Magnetospirillum magneticum AMB-1 have shown that after 16 h of magnetic compensation (500 nT), AMB-1 synthesises larger magnetosomes due to the up-regulation (stimulation) of genes encoding larger magnetosomes and the down-regulation (inhibition) of genes encoding smaller magnetosomes. The gene responsible for magA iron transport remained unchanged [82]. Inhibition of the growth and viability of magnetotactic bacteria (MO-1) after exposure to a 2-nT magnetic field for 2 days has also been noted [79].
In addition to magnetotactic bacteria, changes in antibiotic resistance in human pathogenic bacteria have been studied. The susceptibility of 26 strains of Escherichia coli to selected antibiotics was examined after HMF exposure. Susceptibility to antibiotics (ampicillin, ceftazidime, tetracycline, ofloxacin, and kanamycin) either increased or decreased in different strains, depending on the studied drug. The authors detected two types of E. coli strains: non-sensitive and sensitive to geomagnetic field compensation, which represents about one-third of the strains. Magneto-sensitive E. coli strains showed modified minimum inhibitory concentration (MIC) values to two of five tested antibiotics after HMF exposure [80]. According to Creanga et al. [81], half of the eight tested human pathogen strains (Pseudomonas and Enterobacter strains) were magnetosensitive and showed a change in antibiotic susceptibility (increase or decrease, depending on the tested antibiotics) from 2- to 16-fold.
Ilyin et al. [83] have recently isolated bacteria from the nasopharynx of cosmonauts after their return to Earth from a space mission. The authors observed multiple decreases in antibiotic resistance after exposure to space conditions. Although the effect of the HMF was not especially investigated in this study, its effect on bacterial life is undeniable and can be related to the observed changes in resistance.
The authors further suggest that the HMF can affect cell metabolism by changing the ion transport mechanism in cell plasma membranes in the prokaryotic cell, and this can be applicable in eucaryotic magnetoreception by influencing endoplasmic reticulum (including ribosomal membranes) and mitochondrial membranes [81].

5. Conclusions

So far, the impacts of HMFs on biological systems have been rarely investigated, and the exact mechanism of action remains unclear. The authors explain their results by several theoretical mechanisms, most often by the mechanism of radical pairs which influence the reactive oxygen species concentration. Experimental studies on HMFs yielded conflicting results on the development and functioning of the nervous and cardiovascular systems. However, HMFs are likely to have a negative effect on early developmental stages and fertility in both plants and animals. The conflicting results may have been due to the different exposure times, organism types, and methods of creating a HMF, which seem to be the key factors in the observed biological effects.
However, fewer studies have focused on the effect of the HMF on microorganisms. In our opinion, it is the research of prokaryotic models that can offer useful insight into the magnetoreception of the HMF. Based on this review, the level of magnetoreception can take place at the level of ions, protein complexes, or the cell membrane, and thus, the primary targets of the HMF could be similar for both prokaryotic and eukaryotic organisms.
Hypomagnetic fields seem to affect cell signaling on the level of ion transport and ROS, which has been demonstrated for disruptive embryonal development.

Author Contributions

M.S. wrote the paper; J.S.-K. edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The work was supported by financial aid from the Slovak Grant Agency project No. VEGA 1/0018/22.

Data Availability Statement

The manuscript has no associated data.

Conflicts of Interest

The authors declare no conflict of interest or interpretation of data. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Monteil, C.L.; Lefevre, C.T. Magnetoreception in Microorganisms. Trends Microbiol. 2019, 28, 266–275. [Google Scholar] [CrossRef] [PubMed]
  2. Binhi, V.N.; Prato, F.S. Biological effects of the hypomagnetic field: An analytical review of experiments and theories. PLoS ONE 2017, 12, e0179340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Mo, W.; Liu, Y.; He, R. Hypomagnetic field, an ignorable environmental factor in space? Sci. China Life Sci. 2014, 57, 726–728. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kivelson, M.G.; Bagenal, F. Planetary magnetospheres. In Encyclopedia of the Solar System, 3rd ed.; Elsevier: Amsterdam, The Netherlands, 2014; Chapter 7; pp. 137–157. [Google Scholar]
  5. Zhang, Z.; Xue, Y.; Yang, J.; Shang, P.; Yuan, X. Biological Effects of Hypomagnetic Field: Ground-Based Data for Space Exploration. Bioelectromagnetics 2021, 42, 516–531. [Google Scholar] [CrossRef]
  6. Pavlík, M. Compare of shielding effectiveness for building materials. Prz. Elektrotechniczny 2019, 95, 137–140. [Google Scholar] [CrossRef]
  7. Guo, C.; Liu, D. Quantitative Analyses of Magnetic Field Distributions for Buildings of Steel Structure. In Proceedings of the 2012 Sixth International Conference on Electromagnetic Field Problems and Applications, Dalian, China, 19–21 June 2012. [Google Scholar]
  8. Erdmann, W.; Kmita, H.; Kosicki, J.Z.; Kaczmarek, Ł. How the Geomagnetic Field Influences Life on Earth—An Integrated Approach to Geomagnetobiology. Space Life Sci. 2021, 51, 231–257. [Google Scholar] [CrossRef]
  9. Wajnberg, E.; Acosta-Avalos, D.; Alves, O.C.; de Oliveira, J.F.; Srygley, R.B.; Esquivel, D.M.S. Magnetoreception in eusocial insects: An update. J. R. Soc. Interface 2010, 7, S207–S225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Binhi, V.N.; Savin, A.V. Molecular gyroscopes and biological effects of weak extremely low-frequency magnetic fields. Phys. Rev. E 2002, 65, 051912. [Google Scholar] [CrossRef] [Green Version]
  11. Dröge, W. Free Radicals in the Physiological Control of Cell Function. Physiol. Rev. 2002, 82, 47–95. [Google Scholar] [CrossRef] [Green Version]
  12. Gauron, C.; Rampon, C.; Bouzaffour, M.; Ipendey, E.; Teillon, J.; Volovitch, M.; Vriz, S. Sustained production of ROS triggers compensatory proliferation and is required for regeneration to proceed. Sci. Rep. 2013, 3, srep02084. [Google Scholar] [CrossRef]
  13. Van Huizen, A.V.; Morton, J.M.; Kinsey, L.J.; Von Kannon, D.G.; Saad, M.A.; Birkholz, T.R.; Czajka, J.M.; Cyrus, J.; Barnes, F.S.; Beane, W.S. Weak magnetic fields alter stem cell–mediated growth. Sci. Adv. 2019, 5, eaau7201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Adams, B.; Sinayskiy, I.; Petruccione, F. An open quantum system approach to the radical pair mechanism. Sci. Rep. 2018, 8, 15719. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Hore, P.J.; Mouritsen, H. The Radical-Pair Mechanism of Magnetoreception. Annu. Rev. Biophys. 2016, 45, 299–344. [Google Scholar] [CrossRef] [PubMed]
  16. Ruiz-Gómez, M.J.; Sendra-Portero, F.; Martínez-Morillo, M. Effect of 2.45 mT sinusoidal 50 Hz magnetic field on Saccharomyces cerevisiae strains deficient in DNA strand breaks repair. Int. J. Radiat. Biol. 2010, 86, 602–611. [Google Scholar] [CrossRef]
  17. Barnes, F.; Greenenbaum, B. Some Effects of Weak Magnetic Fields on Biological Systems: RF fields can change radical concentrations and cancer cell growth rates. IEEE Power Electron. Mag. 2016, 3, 60–68. [Google Scholar] [CrossRef]
  18. Brocklehurst, B.; Mclauchlan, K.A. Free radical mechanism for the effects of environmental electromagnetic fields on biological systems. Int. J. Radiat. Biol. 1996, 69, 3–24. [Google Scholar] [CrossRef]
  19. Binhi, V.N.; Prato, F.S. A physical mechanism of magnetoreception: Extension and analysis. Bioelectromagnetics 2016, 38, 41–52. [Google Scholar] [CrossRef] [PubMed]
  20. Otsuka, H.; Mitsui, H.; Miura, K.; Okano, K.; Imamoto, Y.; Okano, T. Rapid Oxidation Following Photoreduction in the Avian Cryptochrome4 Photocycle. Biochemistry 2020, 59, 3615–3625. [Google Scholar] [CrossRef] [PubMed]
  21. Novikov, V.V.; Yablokova, E.V.; Fesenko, E.E. The Effect of a “Zero” Magnetic Field on the Production of Reactive Oxygen Species in Neutrophils. Biophysics 2018, 63, 365–368. [Google Scholar] [CrossRef]
  22. Yan, M.-M.; Zhang, L.; Cheng, Y.-X.; Sappington, T.W.; Pan, W.-D.; Jiang, X.-F. Effect of a near-zero magnetic field on development and flight of oriental armyworm (Mythimna separata). J. Integr. Agric. 2021, 20, 1336–1345. [Google Scholar] [CrossRef]
  23. Zhang, B.; Wang, L.; Zhan, A.; Wang, M.; Tian, L.; Guo, W.; Pan, Y. Long-term exposure to a hypomagnetic field attenuates adult hippocampal neurogenesis and cognition. Nat. Commun. 2021, 12, 1174. [Google Scholar] [CrossRef] [PubMed]
  24. Gupta, A.; Stait-Gardner, T.; Price, W.S. Is It Time to Forgo the Use of the Terms “Spin–Lattice” and “Spin–Spin” Relaxation in NMR and MRI? J. Phys. Chem. Lett. 2021, 12, 6305–6312. [Google Scholar] [CrossRef] [PubMed]
  25. Zangi, R.; Hagen, M.; Berne, B.J. Effect of Ions on the Hydrophobic Interaction between Two Plates. J. Am. Chem. Soc. 2007, 129, 4678–4686. [Google Scholar] [CrossRef] [PubMed]
  26. Zhao, V.; Jacobs, W.M.; Shakhnovich, E.I. Effect of Protein Structure on Evolution of Cotranslational Folding. Biophys. J. 2020, 119, 1123–1134. [Google Scholar] [CrossRef]
  27. Ogneva, I.V.; Usik, M.A.; Burtseva, M.V.; Biryukov, N.S.; Zhdankina, Y.S.; Sychev, V.N.; Orlov, O.I. Drosophila melanogaster Sperm under Simulated Microgravity and a Hypomagnetic Field: Motility and Cell Respiration. Int. J. Mol. Sci. 2020, 21, 5985. [Google Scholar] [CrossRef]
  28. Zhang, B.; Tian, L. Reactive Oxygen Species: Potential Regulatory Molecules in Response to Hypomagnetic Field Exposure. Bioelectromagnetics 2020, 41, 573–580. [Google Scholar] [CrossRef]
  29. Erdmann, W.; Idzikowski, B.; Kowalski, W.; Kosicki, J.Z.; Kaczmarek, Ł. Tolerance of two anhydrobiotic tardigrades Echiniscus testudo and Milnesium inceptum to hypomagnetic conditions. PeerJ 2021, 9, e10630. [Google Scholar] [CrossRef]
  30. Jia, W.; Fan, Z.; Du, A.; Shi, L. Molecular mechanism of Mare Nectaris and magnetic field on the formation of ethyl carbamate during 19 years aging of Feng-flavor Baijiu. Food Chem. 2022, 382, 132357. [Google Scholar] [CrossRef]
  31. Tombarkiewicz, B. Effect of long-term geomagnetic field deprivation on the concentration of some elements in the hair of laboratory rats. Environ. Toxicol. Pharmacol. 2008, 26, 75–79. [Google Scholar] [CrossRef]
  32. Kantserova, N.P.; Krylov, V.V.; Lysenko, L.A.; Ushakova, N.V.; Nemova, N.N. Effects of Hypomagnetic Conditions and Reversed Geomagnetic Field on Calcium-Dependent Proteases of Invertebrates and Fish. Izv. Atmos. Ocean. Phys. 2017, 53, 719–723. [Google Scholar] [CrossRef]
  33. Fu, J.-P.; Mo, W.-C.; Liu, Y.; Bartlett, P.F.; He, R.-Q. Elimination of the geomagnetic field stimulates the proliferation of mouse neural progenitor and stem cells. Protein Cell 2016, 7, 624–637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Hu, P.D.; Mo, W.C.; Fu, J.P.; Liu, Y.; He, R.Q. Long-term Hypogeomagnetic field exposure reduces muscular mitochondrial function and exercise capacity in adult male mice. Prog. Biochem. Biophys. 2020, 47, 426–438. [Google Scholar]
  35. Lednev, V.V. Bioeffects of weak static and alternating magnetic fields. Biofizika 1996, 41, 224–232. [Google Scholar] [PubMed]
  36. Krebs, J. Structure, Function and Regulation of the Plasma Membrane Calcium Pump in Health and Disease. Int. J. Mol. Sci. 2022, 23, 1027. [Google Scholar] [CrossRef] [PubMed]
  37. Poniedziałek, B.; Rzymski, P.; Karczewski, J.; Jaroszyk, F.; Wiktorowicz, K. Reactive oxygen species (ROS) production in human peripheral blood neutrophils exposed in vitro to static magnetic field. Electromagn. Biol. Med. 2013, 32, 560–568. [Google Scholar]
  38. Vergallo, C.; Ahmadi, M.; Mobasheri, H.; Dini, L. Impact of Inhomogeneous Static Magnetic Field (31.7–232.0 mT) Exposure on Human Neuroblastoma SH-SY5Y Cells during Cisplatin Administration. PLoS ONE 2014, 9, e113530. [Google Scholar] [CrossRef]
  39. Tang, R.; Xu, Y.; Ma, F.; Ren, J.; Shen, S.; Du, Y.; Hou, Y.; Wang, T. Extremely low frequency magnetic fields regulate differentiation of regulatory T cells: Potential role for ROS-mediated inhibition on AKT. Bioelectromagnetics 2016, 37, 89–98. [Google Scholar] [CrossRef]
  40. Angelova, P.R.; Dinkova-Kostova, A.T.; Abramov, A.Y. Assessment of ROS Production in the Mitochondria of Live Cells. In Reactive Oxygen Species; Humana: New York, NY, USA, 2021; pp. 33–42. [Google Scholar] [CrossRef]
  41. Belyavskaya, N. Biological effects due to weak magnetic field on plants. Adv. Space Res. 2004, 34, 1566–1574. [Google Scholar] [CrossRef]
  42. Fu, J.-P.; Mo, W.-C.; Liu, Y.; He, R.-Q. Decline of cell viability and mitochondrial activity in mouse skeletal muscle cell in a hypomagnetic field. Bioelectromagnetics 2016, 37, 212–222. [Google Scholar] [CrossRef]
  43. Montoya, R.D. Magnetic fields, radicals and cellular activity. Electromagn. Biol. Med. 2017, 36, 102–113. [Google Scholar] [CrossRef]
  44. Zhang, H.-T.; Zhang, Z.-J.; Mo, W.-C.; Hu, P.-D.; Ding, H.-M.; Liu, Y.; Hua, Q.; He, R.-Q. Shielding of the geomagnetic field reduces hydrogen peroxide production in human neuroblastoma cell and inhibits the activity of CuZn superoxide dismutase. Protein Cell 2017, 8, 527–537. [Google Scholar] [CrossRef] [PubMed]
  45. Mo, W.; Liu, Y.; Bartlett, P.F.; He, R. Transcriptome profile of human neuroblastoma cells in the hypomagnetic field. Sci. China Life Sci. 2014, 57, 448–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Mo, W.C.; Zhang, Z.J.; Wang, D.L.; Liu, Y.; Bartlett, P.F.; He, R.Q. Shielding of the geomagnetic field alters actin assembly and inhibits cell motility in human neuroblastoma cells. Sci. Rep. 2016, 6, 22624. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Zhang, B.; Lu, H.; Xi, W.; Zhou, X.; Xu, S.; Zhang, K.; Jiang, J.; Li, Y.; Guo, A. Exposure to hypomagnetic field space for multiple generations causes amnesia in Drosophila melanogaster. Neurosci. Lett. 2004, 371, 190–195. [Google Scholar] [CrossRef] [PubMed]
  48. Sarimov, R.M.; Binhi, V.N.; Milyaev, V.A. The influence of geomagnetic field compensation on human cognitive processes. Biophysics 2008, 53, 433–441. [Google Scholar] [CrossRef]
  49. Wang, G.-M.; Fu, J.-P.; Mo, W.-C.; Zhang, H.-T.; Liu, Y.; He, R.-Q. Shielded geomagnetic field accelerates glucose consumption in human neuroblastoma cells by promoting anaerobic glycolysis. Biochem. Biophys. Res. Commun. 2022, 601, 101–108. [Google Scholar] [CrossRef]
  50. Mo, W.-C.; Zhang, Z.-J.; Liu, Y.; Bartlett, P.F.; He, R.-Q. Magnetic Shielding Accelerates the Proliferation of Human Neuroblastoma Cell by Promoting G1-Phase Progression. PLoS ONE 2013, 8, e54775. [Google Scholar] [CrossRef]
  51. Gurfinkel, Y.; At’Kov, O.; Vasin, A.; Breus, T.; Sasonko, M.; Pishchalnikov, R. Effect of zero magnetic field on cardiovascular system and microcirculation. Life Sci. Space Res. 2016, 8, 1–7. [Google Scholar] [CrossRef]
  52. IuI, G.; Vasin, A.L.; Matveeva, T.A.; Sasonko, M.L. Evaluation of the hypomagnetic environment effects on capillary blood circulation, blood pressure and heart rate. Aviakosmicheskaia I Ekol. Meditsina= Aerosp. Environ. Med. 2014, 48, 24–30. [Google Scholar]
  53. Ciorba, D.; Morariu, V.V. Life in zero magnetic field. III. Activity of aspartate aminotransferase and alanine aminotransferase during in vitro aging of human blood. Electro- Magn. 2001, 20, 313–321. [Google Scholar] [CrossRef]
  54. Katiukhin, L.N. Rheological Properties of the Erythrocytes in Weakened Static Magnetic Field of the Earth In vitro Study. J. Sci. Res. Rep. 2019, 22, 1–12. [Google Scholar] [CrossRef]
  55. Krylov, V.V.; Bolotovskaya, I.V.; Osipova, E.A. The response of European Daphnia magna Straus and Australian Daphnia carinata King to changes in geomagnetic field. Electromagn. Biol. Med. 2013, 32, 30–39. [Google Scholar] [CrossRef] [PubMed]
  56. Wan, G.-J.; Yuan, R.; Wang, W.-J.; Fu, K.-Y.; Zhao, J.-Y.; Jiang, S.-L.; Pan, W.-D.; Sword, G.A.; Chen, F.-J. Reduced geomagnetic field may affect positive phototaxis and flight capacity of a migratory rice planthopper. Anim. Behav. 2016, 121, 107–116. [Google Scholar] [CrossRef]
  57. Zhang, Y.; Pan, W. Removal or component reversal of local geomagnetic field affects foraging orientation preference in migratory insect brown planthopper Nilaparvata lugens. Peerj 2021, 9, e12351. [Google Scholar] [CrossRef] [PubMed]
  58. Wan, G.-J.; Jiang, S.-L.; Zhao, Z.-C.; Xu, J.-J.; Tao, X.-R.; Sword, G.A.; Gao, Y.-B.; Pan, W.-D.; Chen, F.-J. Bio-effects of near-zero magnetic fields on the growth, development and reproduction of small brown planthopper, Laodelphax striatellus and brown planthopper, Nilaparvata lugens. J. Insect Physiol. 2014, 68, 7–15. [Google Scholar] [CrossRef] [PubMed]
  59. Mo, W.; Liu, Y.; Cooper, H.M.; He, R.-Q. Altered development of Xenopus embryos in a hypogeomagnetic field. Bioelectromagnetics 2011, 33, 238–246. [Google Scholar] [CrossRef]
  60. Fesenko, E.E.; Mezhevikina, L.M.; Osipenko, M.A.; Gordon, R.Y.; Khutzian, S.S. Effect of the “zero” magnetic field on early embryogenesis in mice. Electromagn. Biol. Med. 2010, 29, 1–8. [Google Scholar] [CrossRef]
  61. Xue, X.; Ali, Y.F.; Liu, C.; Hong, Z.; Luo, W.; Nie, J.; Li, B.; Jiao, Y.; Liu, N.-A. Geomagnetic Shielding Enhances Radiation Resistance by Promoting DNA Repair Process in Human Bronchial Epithelial Cells. Int. J. Mol. Sci. 2020, 21, 9304. [Google Scholar] [CrossRef]
  62. Belyaev, I.Y.; Alipov, Y.D.; Harms-Ringdahl, M. Effects of zero magnetic field on the conformation of chromatin in human cells. Biochim. Biophys. Acta (BBA)-Gen. Subj. 1997, 1336, 465–473. [Google Scholar] [CrossRef]
  63. Baek, S.; Choi, H.; Park, H.; Cho, B.; Kim, S.; Kim, J. Effects of a hypomagnetic field on DNA methylation during the differentiation of embryonic stem cells. Sci. Rep. 2019, 9, 1333. [Google Scholar] [CrossRef]
  64. Agliassa, C.; Narayana, R.; Christie, J.M.; Maffei, M.E. Geomagnetic field impacts on cryptochrome and phytochrome signaling. J. Photochem. Photobiol. B Biol. 2018, 185, 32–40. [Google Scholar] [CrossRef]
  65. Maffei, M.E. Magnetic field effects on plant growth, development, and evolution. Front. Plant Sci. 2014, 5, 445. [Google Scholar] [CrossRef]
  66. Agliassa, C.; Mannino, G.; Molino, D.; Cavalletto, S.; Contartese, V.; Bertea, C.M.; Secchi, F. A new protein hydrolysate-based biostimulant applied by fertigation promotes relief from drought stress in Capsicum annuum L. Plant Physiol. Biochem. 2021, 166, 1076–1086. [Google Scholar] [CrossRef] [PubMed]
  67. Xu, C.; Zhang, Y.; Yu, Y.; Li, Y.; Wei, S. Suppression of Arabidopsis flowering by near-null magnetic field is mediated by auxin. Bioelectromagnetics 2017, 39, 15–24. [Google Scholar] [CrossRef] [PubMed]
  68. Xu, C.; Yu, Y.; Zhang, Y.; Li, Y.; Wei, S. Gibberellins are involved in effect of near-null magnetic field on Arabidopsis flowering. Bioelectromagnetics 2016, 38, 1–10. [Google Scholar] [CrossRef] [PubMed]
  69. Mo, W.-C.; Zhang, Z.-J.; Liu, Y.; Zhai, G.-J.; Jiang, Y.-D.; He, R.-Q. Effects of a hypogeomagnetic field on gravitropism and germination in soybean. Adv. Space Res. 2011, 47, 1616–1621. [Google Scholar] [CrossRef]
  70. Xu, C.; Wei, S.; Lu, Y.; Zhang, Y.; Chen, C.; Song, T. Removal of the local geomagnetic field affects reproductive growth inArabidopsis. Bioelectromagnetics 2013, 34, 437–442. [Google Scholar] [CrossRef]
  71. Islam, M.; Maffei, M.; Vigani, G. The Geomagnetic Field Is a Contributing Factor for an Efficient Iron Uptake in Arabidopsis thaliana. Front. Plant Sci. 2020, 11, 325. [Google Scholar] [CrossRef] [Green Version]
  72. Negishi, Y.; Hashimoto, A.; Tsushima, M.; Dobrota, C.; Yamashita, M.; Nakamura, T. Growth of pea epicotyl in low magnetic field implication for space research. Adv. Space Res. 1999, 23, 2029–2032. [Google Scholar] [CrossRef]
  73. Xu, C.; Yin, X.; Lv, Y.; Wu, C.; Zhang, Y.; Song, T. A near-null magnetic field affects cryptochrome-related hypocotyl growth and flowering in Arabidopsis. Adv. Space Res. 2012, 49, 834–840. [Google Scholar] [CrossRef]
  74. Xu, C.; Li, Y.; Yu, Y.; Zhang, Y.; Wei, S. Suppression of Arabidopsis flowering by near-null magnetic field is affected by light. Bioelectromagnetics 2015, 36, 476–479. [Google Scholar] [CrossRef] [PubMed]
  75. Narayana, R.; Fliegmann, J.; Paponov, I.; Maffei, M.E. Reduction of geomagnetic field (GMF) to near null magnetic field (NNMF) affects Arabidopsis thaliana root mineral nutrition. Life Sci. Space Res. 2018, 19, 43–50. [Google Scholar] [CrossRef] [PubMed]
  76. Vanderstraeten, J.; Gailly, P.; Malkemper, E.P. Low-light dependence of the magnetic field effect on cryptochromes: Possible relevance to plant ecology. Front. Plant Sci. 2018, 9, 121. [Google Scholar] [CrossRef] [Green Version]
  77. Soltani, F.; Kashi, A.; Arghavani, M. Effect of magnetic field on Asparagus officinalis L. seed germination and seedling growth. Seed Sci. Technol. 2006, 34, 349–353. [Google Scholar] [CrossRef]
  78. Tsetlin, V.; Moisa, S.; Levinskikh, M.; Nefedova, E. EFFECT OF VERY SMALL DOSES OF IONIZING RADIATION AND HYPOMAGNETIC FIELD CHANGE PHYSIOLOGICAL CHARACTERISTICS OF HIGHER PLANT SEEDS. Aviakosmicheskaia I Ekol. Meditsina= Aerosp. Environ. Med. 2016, 50, 51–58. [Google Scholar] [CrossRef]
  79. Zhang, S.-D.; Petersen, N.; Zhang, W.-J.; Cargou, S.; Ruan, J.; Murat, D.; Santini, C.-L.; Song, T.; Kato, T.; Notareschi, P.; et al. Swimming behaviour and magnetotaxis function of the marine bacterium strain MO-1. Environ. Microbiol. Rep. 2013, 6, 14–20. [Google Scholar] [CrossRef] [PubMed]
  80. Poiata, A.; Creanga, D.E.; Morariu, V.V. Life in zero magnetic field. VE coli resistance to antibiotics. Electromagn. Biol. Med. 2003, 22, 171–182. [Google Scholar] [CrossRef]
  81. Creanga, D.; Poiata, A.; Morariu, V.; Tupu, P. Zero-magnetic field effect in pathogen bacteria. J. Magn. Magn. Mater. 2004, 272–276, 2442–2444. [Google Scholar] [CrossRef]
  82. Wang, X.K.; Ma, Q.F.; Jiang, W.; Lv, J.; Pan, W.D.; Song, T.; Wu, L.-F. Effects of Hypomagnetic Field on Magnetosome Formation of Magnetospirillum Magneticum AMB-1. Geomicrobiol. J. 2008, 25, 296–303. [Google Scholar] [CrossRef]
  83. Ilyin, V.K.; Orlov, O.I.; Morozova, Y.A.; Skedina, M.A.; Vladimirov, S.K.; Plotnikov, E.V.; Artamonov, A.A. Prognostic model for bacterial drug resistance genes horizontal spread in space-crews. Acta Astronaut. 2022, 190, 388–394. [Google Scholar] [CrossRef]
Figure 1. Presence of hypomagnetic field in the solar system [3].
Figure 1. Presence of hypomagnetic field in the solar system [3].
Processes 11 00282 g001
Table 1. Impact of hypomagnetic field on cell transport and metabolism (B—magnetic flux density in Tesla (T)).
Table 1. Impact of hypomagnetic field on cell transport and metabolism (B—magnetic flux density in Tesla (T)).
Impact onEffect Hypomagnetic Field Properties
OrganismMechanismB (nT)DurationReferences
Mineral density of bonesReductionSprague-Dawley ratsShielding<3003 days[30]
The concentration of Fe, Mn, Cu, CrReductionFur of laboratory rats WistarShielding<207 months[31]
Ca2+ dependent proteasesInactivationEnzymes from fish and invertebratesCompensation 1 h[32]
The concentration of Co, NiNo effectFur of laboratory rats WistarShielding<208 months[31]
Mitochondrial activityReductionSkeletal muscle cellsCompensation<2007 days[33]
Mitochondrial activityReductionMouse (C57BL/6)Compensation0–50030 days[34]
ATP levelsReductionSkeletal muscle cellsCompensation<30,0003 days[35]
Cell respirationReductionDrosophila melanogasterCompensation16 h[27]
Cell respirationPromotionDrosophila melanogasterCompensation11 h[27]
Table 2. Impact of hypomagnetic field on animal neural systems (B—magnetic flux density in Tesla (T)).
Table 2. Impact of hypomagnetic field on animal neural systems (B—magnetic flux density in Tesla (T)).
Impact onEffectOrganismHypomagnetic Field Properties
Mechanism of GenerationB (nT)DurationReferences
Neural systemROS levelsReductionMouse (C57BL/6 J), malesShielding170Every 3 days/150 days[23]
ROS levelsReductionPeritoneal mice neutrophilsShielding201.5 h[21]
GrowthPromotionPrimary neural progenitor/mouse stem cellsShielding0–2007 days[42]
ROS levelsReductionHuman cells of neuroblastShielding0–20016 h[44]
ROS genes expressionReductionMouse (C57BL/6 J), malesShielding1703 day/150 days[23]
Gene expressionReduction (down-regulation)Human neuroblast cellsCompensation<2002 days[45]
Migratory propertiesReductionHuman cells of neuroblastShielding0–20048 h[46]
ProliferationPromotionHuman cells neuroblast (SH-SY5Y)Shielding0–2003 days[46]
MemoryReductionDrosophila melanogasterCompensation100–68010–19 generations[47]
ProliferationPromotionHuman neuroblastoma cellsShielding--[48]
Cognitive abilitiesReductionHuman (volunteers)Compensation40045 min[49]
ProliferationPromotionHuman neuroblastoma (SHSY5Y) cellsShielding<2003 days[50]
Hippocampal neurogenesisInhibitionMouse (C57BL/6 J), malesShielding170every 3 day/150 days[23]
Cardiovascular systemBlood pressurePromotionHuman (volunteers)Compensation±10 60 min[51]
Blood circulationPromotionHuman (volunteers)Compensation±10 60 min[52]
HaemolysisPromotionHuman bloodCompensation100 72 h[53]
HaemolysisPromotionBlood of ratsCompensation1926 h to 4 weeks[54]
Life cycle and survivalSurvivalReductionMilnesium inceptumShielding-21 days[29]
SurvivalReductionTardigada (Echiniscus testudo and Milnesium inceptum)Shielding-21 days[29]
Life expectancyReductionDaphninia magneCompensation15Generational period[55]
Larval developmentInhibitionMythimna separataCompensation<50012 h[56]
Development of eggs and nymphsDelayedNilaparvata lugensCompensation0–1060Generational period[57]
FertilityReductionNilaparvata lugensCompensation0–1060Generational period[58]
Production of abnormal embryoysPromotionXenopus larvaeShielding104 ± 12.64 days[59]
FertilityReduction (sterility)NMRI mouse zygotesShielding20012 days[60]
AbortionPromotionPregnant NMRI miceShielding2003–12 days[60]
Survival of cells exposed to X-raysPromotionImmortalised human bronchial epithelial cellsShielding<5024 h[61]
Chromatic condensationChangesHuman fibroblasts and lymphocytesCompensation180020–70 min[62]
Table 3. Impact of hypomagnetic field on plants (B—magnetic flux density in Tesla (T)).
Table 3. Impact of hypomagnetic field on plants (B—magnetic flux density in Tesla (T)).
Impact onEffectOrganismHypomagnetic Field Properties
Mechanism of GenerationB (nT)DurationReferences
GrowthReductionGlycine maxShielding111 ± 1524 h[69]
GrowthReductionArabidopsis thalianaCompensation0–133035 days[70]
GrowthReductionArabidopsis thalianaCompensation40–4496 h[71]
Epicotyl elongationPromotionPisum sativumShielding-24 h[72]
Gene expansionReductionArabidopsis thalianaCompensation0–133033 days[68]
Activity of photoreceptors phyAReductionArabidopsis thalianaCompensation403 h[64]
Activity of phyB photoreceptorsPromotionArabidopsis thalianaCompensation403 h[64]
The content of auxin in flowerReductionArabidopsis thalianaCompensation0–133033 days[67]
Gene expression (associated with flowering)PromotionArabidopsis thalianaCompensation5033 days[73]
Auxin content in rootsPromotionArabidopsis thalianaCompensation1–133033 days[67]
Iron intake by rootsReductionArabidopsis thalianaCompensation40–4496 h[71]
Concentration of Ca2+ ionsPromotionPisum sativum (root system)Shielding0.5–23 days[41]
Table 4. Impact of hypomagnetic field on procaryotic microorganisms (B—magnetic flux density in Tesla (T)).
Table 4. Impact of hypomagnetic field on procaryotic microorganisms (B—magnetic flux density in Tesla (T)).
Impact onEffectOrganismField Properties
MechanismB (nT)DurationReferences
Growth and number of cellsReductionMagnetotactic bacteria (MO-1)Shielding22 days[79]
Tolerance to antibioticsBoth reduction and promotionEscherichia coliCompensation-6 days[80]
Tolerance to antibioticsBoth reduction and promotionPseudomonas and Enterobacter strainsField compensation-6 days[81]
Magnetosome sizePromotionMagnetospirillum magneticum AMB-1Compensation50016 h[82]
Gene expressionModificationMagnetospirillum magneticum AMB-1Compensation50016 h[82]
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

Sinčák, M.; Sedlakova-Kadukova, J. Hypomagnetic Fields and Their Multilevel Effects on Living Organisms. Processes 2023, 11, 282. https://doi.org/10.3390/pr11010282

AMA Style

Sinčák M, Sedlakova-Kadukova J. Hypomagnetic Fields and Their Multilevel Effects on Living Organisms. Processes. 2023; 11(1):282. https://doi.org/10.3390/pr11010282

Chicago/Turabian Style

Sinčák, Miroslava, and Jana Sedlakova-Kadukova. 2023. "Hypomagnetic Fields and Their Multilevel Effects on Living Organisms" Processes 11, no. 1: 282. https://doi.org/10.3390/pr11010282

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