Next Article in Journal
Synthesis and Electrochemical and Spectroscopic Characterization of 4,7-diamino-1,10-phenanthrolines and Their Precursors
Next Article in Special Issue
Deuterium-Depleted Water as Adjuvant Therapeutic Agent for Treatment of Diet-Induced Obesity in Rats
Previous Article in Journal
Dicentrine Potentiates TNF-α-Induced Apoptosis and Suppresses Invasion of A549 Lung Adenocarcinoma Cells via Modulation of NF-κB and AP-1 Activation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Possible Mechanisms of Biological Effects Observed in Living Systems during 2H/1H Isotope Fractionation and Deuterium Interactions with Other Biogenic Isotopes

by
Alexander Basov
1,2,
Liliya Fedulova
3,
Ekaterina Vasilevskaya
3 and
Stepan Dzhimak
2,3,4,*
1
Department of Fundamental and Clinical Biochemistry, Kuban State Medical University, Krasnodar 350063, Russia
2
Department of Radiophysics and Nanotechnology, Kuban State University, Krasnodar 350040, Russia
3
The V.M. Gorbatov Federal Research Center for Food Systems of Russian Academy of Sciences, Moscow 109316, Russia
4
Federal Research Center the Southern Scientific Center of the Russian Academy of Sciences, Rostov-on-Don 344006, Russia
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(22), 4101; https://doi.org/10.3390/molecules24224101
Submission received: 11 October 2019 / Revised: 1 November 2019 / Accepted: 11 November 2019 / Published: 13 November 2019
(This article belongs to the Special Issue Medicinal Biochemistry of Deuterium Discrimination)

Abstract

:
This article presents the original descriptions of some recent physics mechanisms (based on the thermodynamic, kinetic, and quantum tunnel effects) providing stable 2H/1H isotope fractionation, leading to the accumulation of particular isotopic forms in intra- or intercellular space, including the molecular effects of deuterium interaction with 18O/17O/16O, 15N/14N, 13C/12C, and other stable biogenic isotopes. These effects were observed mainly at the organelle (mitochondria) and cell levels. A new hypothesis for heavy nonradioactive isotope fractionation in living systems via neutron effect realization is discussed. The comparative analysis of some experimental studies results revealed the following observation: “Isotopic shock” is highly probable and is observed mostly when chemical bonds form between atoms with a summary odd number of neutrons (i.e., bonds with a non-compensated neutron, which correspond to the following equation: Nn − Np = 2k + 1, where k ϵ Z, k is the integer, Z is the set of non-negative integers, Nn is number of neutrons, and Np is number of protons of each individual atom, or in pair of isotopes with a chemical bond). Data on the efficacy and metabolic pathways of the therapy also considered 2H-modified drinking and diet for some diseases, such as Alzheimer’s disease, Friedreich’s ataxia, mitochondrial disorders, diabetes, cerebral hypoxia, Parkinson’s disease, and brain cancer.

Graphical Abstract

1. Introduction

It is well known that many physical and chemical processes in living systems are accompanied by isotope fractionation among the atoms of biologically significant elements, primarily H, C, O, and N [1,2]. Changes in the ratio of heavy and light isotopes of biogenic elements cause quantum (tunnel), thermodynamic, and consequently kinetic isotope effects that manifest themselves under natural conditions in acceleration, deceleration, or branching metabolic reactions, as well as in the changed speed of metabolite influx through transport channels; this leads to a local fluctuation in the pool of biologically active substances [3] in some compartments or organelles of the cells. Moreover, the change in the natural isotope ratio in some structural components of living systems is accompanied by the modification of some biochemical reaction mechanisms in animals and plants, which is caused, for example, by compartmentalization [4]. in general, this can lead, among other things, to faster adaptation under the influence of various stress factors.
Some manifestations of the isotope effects of certain elements under natural conditions have been studied and published [5]. However, it is extremely important to know the mechanisms of isotope effects to understand the specific features and reconstruction mechanisms of physiological processes in plants and animals, depending on their geographical origin and habitat [6]. This approach makes it possible to use different isotope concentration values obtained from studies to characterize biological objects [7], including clarification of the specific features of biochemical processes and reaction mechanisms, as well as the reconstruction of climatic, physiological, ecological, and environmental conditions during a certain period in the vital activities of organisms [8,9,10].
However, despite considerable amounts of experimental data on changes in the ratio of light and heavy isotopes of biogenic elements in biological objects depending on the habitat [11,12], possible mechanisms of the influence of nonradioactive isotopes of hydrogen, oxygen, carbon, and others on living objects remain unexplained in the study of their exchange, especially for combined (poly-isotopic) effects. Their explanation is necessary to clarify isotope action regularities and predict ultimate biological effects in organisms under the influence of naturally or artificially generated fluctuations of isotopic composition, and to create new methods for (poly-)isotopic correction of metabolic abnormalities observed in the body when exposed to various stress factors (chemical, physical [13], or emotional [14]).
This issue is of current interest, as there are now opportunities to correct the ratio of light and heavy isotopes of biogenic elements in the body via modification of the isotope composition of food and water [15,16,17]. It must be noted that changes in the isotope composition of food products can occur in both directions—increases and decreases in their heavy isotope concentrations [18,19]. The latter is especially relevant owing to the possible influence of differences in the ratio of light and heavy isotopes of hydrogen on the adaptive potential of an organism [20,21,22,23], which seems promising with regard to corrective measures that are taken in biology and medicine to prevent disadaptation of biological objects under unfavorable existence conditions. Prediction of changes in the magnitude and direction of the isotope gradient under particular physiological and pathological conditions is especially important because of different isotope exchange rates in the tissues of internal organs with different metabolic activities, owing to gender, age, and temperature differences in individuals of the same species [24]. Therefore, purposefully changing the heavy and light isotope ratio (e.g., deuterium/protium) in biological media and internal organs can provide an opportunity for a preventive increase in the adaptive potential of an organism in case of a developing pathological process. This is possible by modifying the intensity of the metabolic processes and structural rearrangements at the cellular level in living beings, which are both explained by potentially the most profound kinetic isotope effect, arising as hydrogen atoms are fractioned owing to differences in their mass (1H и 2H), and an expected positive biological impact on the body caused by the removal of heavy isotopes with pronounced negative kinetic isotope effects [25].
Another major discovery is proof of the presence of paramagnetic isotope effects in some metals (calcium, magnesium, and zinc) taking part in biocatalysis [26,27,28,29,30,31], which show a change in the activity of enzymes regulating energy exchange and the transfer of genetic information in cells, depending on the isotope composition of the medium [32,33].
Therefore, the purpose of this study was to substantiate possible formation mechanisms of paradoxical biochemical and biophysical effects arising from the interaction of deuterium or protium with other isotopes and combinations of various macro- and microelements in living systems.
For a more complete understanding of the effects observed when individual fractions of heavy or light isotopes accumulate, the key mechanisms for the realization of their bioeffects should be considered. This approach requires detailed analysis, as it has been shown by different scientific groups that metabolic processes in the body can be influenced (stimulated or inhibited) by both decreases and increases within the same range in the amount of heavy isotopes in comparison with their natural levels in the environment [34,35,36,37,38].

2. Molecular Effects of 2H/1H Exchange

Thermodynamic disparity of isotopic compounds can lead to uneven distributions of hydrogen isotopes when equilibrium is reached because of isotope exchange (2H/1H) reactions among cells, as well as the predominant accumulation of one of the isotopic forms in intra- or intercellular space. Isotope exchange reactions in biological systems can be accompanied by a change not only in thermodynamic characteristics (diffusion coefficient and specific charge of ions), but also primarily in kinetic ones (rate of biochemical reactions) at the molecular level. It is known that the replacement of deuterium by protium in high molecular compounds does not always affect their thermodynamic state significantly. Nevertheless, some works describe an intermolecular isotope effect caused by a more stable interaction between the deuterated solvation shell and the biomolecule itself (protein, DNA, etc.); this is accompanied by a decrease in their involvement in biochemical reactions due to the delayed desolvation of some regulatory parts of molecules during their transition into a functionally active state [39]. However, during selective changes in active and allosteric centers of enzymes associated with hydrogen isotope exchange as part of readily dissociating groups (hydroxyl (–OH), thiol (–SH), primary protonated, and secondary amino groups (–NH3+, =NH)), increasing the amount of protium can change the rate of catalytic processes through a decrease in the energy of the molecule’s transition state activation during biocatalysis reactions. It should be noted that in living organisms, intermolecular and intramolecular kinetic isotope effects will be combined, owing to a complex organization of high-molecular compounds (proteins and nucleic acids) that have high solvation ability, and consequently because of hydrogen isotope exchange between dissociating groups in the macromolecule and its hydration shell. Such changes not only lead to the modification of energy interactions between enzymes and substrates (intermolecular effects) but can also be accompanied by intramolecular conformational rearrangements at different rates; this occurs in “enzyme–substrate” or “enzyme–cofactor (coenzyme)” complexes, which can shorten some stages of biocatalytic transformations or accelerate the transition of the enzyme into an active form after the completion of a catalytic cycle, thereby increasing the activity of biochemical processes.
It should be noted that the primary kinetic isotopic effect is characterized by an increase in the energy (En) of rupture for bonds formed by heavier isotopes, for example, Enrupture(A–D) > Enrupture(A–H) (where A is atom). In some studies, attempts were made to extrapolate the data obtained in vitro to in vivo systems [40,41].
Some particularities of various isotope fractionations, including oxygen and hydrogen, observed during biochemical reactions are shown below; these account for the possible uneven local distribution of isotopes, not only within parts of the body, but also within some cells and even organelles with relatively isolated compartments (Reactions 1–3, Figure 1, Figure 2 and Figure 3).
Isotope fractionation in enzyme systems can occur in activated enzyme–substrate complexes (such as SnXE in Reaction 1, Figure 1), when the amount of energy required to replace light isotopes with heavy isotopes decreases significantly. Moreover, for a more energetically favorable decomposition of the enzyme–substrate complex with heavy isotopes, a substantial shift of the reaction toward the formation of a product containing a heavy isotope (Pn+mX) will be observed. This will be pronounced in case of reversibility of the transitional state in the [PnXE ↔ SnXE] system, when the rate of the forward reaction is the same as that of the reverse reaction; however, this rate should be much lower than that of the direct reaction in the transitional state of the [SnXE ↔ Pn+mXE] system, which will lead to a significant increase in the concentration of heavy isotopes in the reaction products. The described mechanisms were confirmed by the presence of the revealed similar intramolecular rearrangements (rapid keto–enol isomerizations) in the photodissociation of acetaldehyde enriched with deuterium (C2H3C1HO [42]):
2H2C2H–C1H=O → [2H2C=C1H–O2H ↔ 2H2C=C2H–O1H ↔ 2HC1H=C2H–O2H] · n → 2H2C1H–C2H=O, for n ≥ 20 (number of keto–enol isomerizations).
Considering that the energy required for intramolecular rearrangements in the enzyme–substrate complex is lower than the energy required for keto–enol isomerizations, the probability of heavy isotope fractionation significantly increases in proportion to the number of rearrangements in the [PnXE ↔ SnXE] system, which indicates a high probability of heavy isotope fractionation under natural conditions.
It should be noted that an additional possibility of fractionation is observed for heavy isotopes of hydrogen and oxygen that are part of an activated enzyme–substrate complex due to fixation of 2H and 18O from the solvent (water: HDO [39,43,44] or H218O), in which all biochemical processes (Reactions 2 and 3, Figure 2 and Figure 3) proceed. The latter can lead to a more significant and widespread fractionation of these isotopes in living systems [1,45,46,47,48].
Moreover, it was described that the solubility of individual active pharmaceutical ingredients (such as bendazole hydrochloride, series Р111106) differs about by 50% in natural water (δ2H = −101‰) and deuterium-depleted water (δ2H = −974‰ [49]). In some studies, the influence of deuterium-depleted water (δ2H = −679‰) on the efficiency of protein extractions (where these proteins were from 0.11 to 0.2 parts of the total protein material, including members of the family α- and β-chains of hemoglobin and actin-binding protein (profilin 1)) from the spleen, thymus, and lymph nodes of Sus scrofa was also shown [50,51,52,53], and this approach was allowed to receive more active livestock products.

3. Isotope Exchange Effects Observed at the Organelles

The abovementioned influence of the rates of biochemical processes at different ratios of light and heavy isotopes of biogenic elements on the functioning of organelles and subcellular structures is evident in the evaluation of energy metabolism intensity in mitochondria, which makes it possible to characterize the resulting impact of isotope fluctuations on metabolism [54,55,56]. In this case, the replacement of deuterium by protium leads, according to a number of studies, is due to the acceleration of proton fluxes in mitochondria; consequently, this leads to an increase in the production of some substrates, providing, among other things, a higher energy exchange, thereby increasing the resistance of the cell to negative external or internal effects (for example, to hypoxia and intoxication). The effect of isotope exchange on catalytic complexes in some organelles (mitochondria [57], peroxisome [58,59], and lysosome) can change the intensity of metabolic processes at the cellular level as well as substantially modify the resistance or response of biological tissue.
The latter is explained by the fact that one of the key features of metabolism in eukaryotic cells is the compartmentalization of reactions in different organelles. The most critical effect of 2H/1H isotope exchanges is the mitochondria generating and using universal reducing substrates that are widely involved in redox reactions in both the organelle itself and the cytosol, supporting biosynthesis, redox homeostasis, transduction of intracellular signals, adenosine triphosphate generation, etc. Moreover, deuterium atoms transferred, for example, as part of carbohydrates, because of metabolic transformations associated with NADH2 and NADPH + H systems, can migrate from the cytosol and back; they enter amino acids, fatty acids, and the pentose–phosphate backbone of nucleic acids, thereby enriching the pool of heavy isotopes in organic compounds, which is also a special case of 2H/1H fractionation at the cellular level [60,61].
A cell in a varied functional state is characterized by a specific level of energy and biosynthetic needs that can vary within a certain range, modifying the chains of metabolic transformations with pronounced competition for the pyruvic acid fund; this is accompanied by a change in the proportions of parts of the pyruvic acid fund selectively extracted for the energy system of the cell and metabolite synthesis, causing isotope selection [62]. In addition, metabolic water can be intensively produced in the peroxisomes under particular conditions [59], which also creates auxiliary possibilities for the fractionation of non-radioactive hydrogen isotopes.

4. In Vivo 2H/1H Exchange Effects

Changes in the 2H/1H isotope composition in living systems can often be caused by different ratios of deuterium and protium in drinking water and food; these changes are observed both during the migration of living organisms and consumption of food produced artificially with reduced amounts of heavy nonradioactive isotopes of biogenic elements [1,2,5], mostly deuterium (for example, deuterium-depleted water).
Fluctuations in the isotope composition at the organism level, and probably, the formation of a new 2H/1H isotope gradient opposite to the physiological gradient (δ2H “visceral organs” > δ2H “blood plasma”), can lead to an increase in the activity of humoral and cellular protective systems; this causes a nonspecific phenomenon of increasing resistance of the organism because of preconditioning, during which protective mechanisms at the cellular level are potentiated and information transfer through secondary messengers occurs [2,23].
The described processes can lead to an increase in the rate of transcription and synthesis of heat shock proteins and antioxidant enzymes (for example, superoxide dismutase and catalase), an increase in the content of antinociceptive factors and low-molecular-weight reducing equivalents of the antioxidant system, and changes in the production of reactive oxygen species and free radicals, activity of ion channels (Na+, K+, Ca2+, and Mg2+), and ratio of energy substrate transporters in membranes [61,63]. Moreover, the consumption of deuterium-depleted drinking water by mice prevents behavioral, transcriptional, and proliferative changes typical of a depression-like state. The consumption of deuterium-depleted water also leads to electroencephalography changes observed during sleep, which resemble the effects of norepinephrine and serotonin reuptake inhibitors [64].
Several works [65,66,67,68,69,70] indicate the multidirectional effect of isotope exchange reactions on the functional activity of biological systems, their native properties, and structural organization. Nevertheless, most bioeffects associated with changes in the isotope composition of biogenic elements in the body are still not fully understood [71], which primarily concern the effect of low (compared to the natural level) concentrations of heavy nonradioactive isotopes on living systems. The latter is often associated with the traditional explanation of kinetic isotope effects, which is based on the idea of an increase in their intensity proportional to the concentration of heavy isotopes. In this case, the isotopic effects associated with a deliberate decrease in the concentration of heavy nonradioactive isotopes in comparison with their natural content, arising in complexly organized living systems during the formation of various isotope gradients, are often not fully considered [72]. According to the data available on this topic, among the studies describing changes in the ratio of the isotopes of nutrient elements, most studies pertain to the effect of various concentrations of deuterium on the body, which is explained by more pronounced isotope differences in the masses of the 1H and 2H nuclei compared to the analogous atomic weight ratio parameters of stable isotopes of oxygen, carbon, and nitrogen. However, despite the increasingly large number of studies on the effect of low δ2H values on living systems, researchers mostly consider changes in the 2H/1H isotope ratio in blood plasma and do not pay enough attention to comparative studies on δ2H values in body tissues and fluids during abundant consumption of deuterium-depleted water with modified isotope compositions [73].
In general, the effect of deuterium-depleted water on the isotope composition of tissues and morphofunctional indices in multicellular organisms is understudied; nevertheless, this topic is of particular interest, as the morphofunctional status is one of the most meaningful indicators of the development of individual organisms and their health conditions. Further, the consumption of deuterium-depleted water affects the adaptive capacity of organisms in different periods of ontogeny. It is the choice of the object of research for properly assessing the multifaceted effects of isotope exchange reactions on biological systems that is important in such scientific studies; therefore, in scientific literature, one can find studies of various unicellular and multicellular organisms [74,75,76,77,78], but often in a discrete form, without taking into account the genetic heterogeneity of individuals. However, recently the latter seems to be a relevant factor, because the different sensitivity was shown of animals of different stocks (Wistar and random bred albino male rats) in different periods of ontogeny to the effects of the drinking diet with deuterium-depleted content (δ2H = −705‰ [79]), including their morphofunctional indices and biochemical processes (e.g., hydrogen peroxide production by isolated mitochondria of rat liver). Besides, the isotope effect associated with a decrease in deuterium in the environment is not only expressed in animal cells but also in plant cells (e.g., deuterium-depleted water (δ2H = −839‰)) prevents the appearance of hyperhydricity in red beet [80].
However, the increase of 2H2O in living systems also leads to various effects [81]. Therefore, in the mice, which received deuterium (up to 10% 2H2O) as part of a drinking diet for 5 weeks and who were inoculated with 4T1 mammary carcinoma after third week consumption 2H2O, the median survival rate was 4 days longer than that in the control group. However, of interest is the fact that deuterium-depleted water (δ2H were equaled from −807‰ to −358‰) was capable of causing cell cycle arrest in the G1/S transition and reducing the number of the cells in the S phase, while the superoxide dismutase and catalase activities were increased. The populations of cells in the G1 phase in MCF-7 cells were also significantly increased [82]. Therefore, such a generally similar effect on different living systems of two opposing processes (for example, the abovementioned processes associated with both increasing and decreasing 2H in media) requires the creation of a new hypothesis that can explain why both reduction and increase of concentrations of the same isotope can lead to significantly changes in the activity of biological substances, cells, and organisms in approximately the same amount under certain conditions (such as on the border of melting ice [83]).

5. 2H/1H Exchange Effects and Other Heavy Nonradioactive Isotopic Fractionations in Living and Natural Systems via Neutron Effect Realization (Hypothesis)

Apart from the abovementioned thermodynamic and kinetic effects, in the presence of light isotopes of biogenic elements, there is possibly a mechanism that is connected with the quantum tunneling effect of some heavy isotopes, as well as of covalently bound atoms of light and heavy isotopes. Comparative analysis of some experimental results from different scientific groups revealed the following regularity (Table 1): “Isotopic shock” is observed mostly when the formation of chemical bonds with an odd number of neutrons (with a non-compensated neutron, firstly by mass and charge of proton, and secondly by mass of another neutron) is highly probable, or when the system contains a chemical element (usually some metal), which also has one or several (odd number) non-compensated neutrons. The probability of such a regularity, hereinafter referred to as the neutron Basov–Dzhimak hypothesis (BADz phenomenon), was analyzed; possible mechanisms for its implementation may be connected with the ability of a non-compensated neutron to modify nuclear spins in atoms and consequently influence the reactivity of a chemical bond formed by isotopes containing non-compensated neutrons.
Based on the above discussions, we could specify the following conditions for the occurrence of the neutron BADz phenomenon in biological systems:
(1) The neutron BADz phenomenon in biological systems is observed when there are chemical bonds between the atoms showing complete predominance of neutrons over protons, marked by odd positive numbers (1, 3, 5, 7, etc.); Nn − Np = 2k + 1, where k ϵ Z (k is the integer, Z is the set of non-negative integers), N denotes number, n denotes neutrons, and p denotes protons;
(2) The neutron BADz phenomenon is observed in biological systems with chemical bonds between atoms having a fractional resultant nuclear spin, given as follows:
(a) 2k ≠ A-spin + A’-spin ≠ 2k + 1, where k ϵ Z, A-spin denotes atomic spin of one of the isotopes involved in the formation of the chemical bond, A’-spin denotes atomic spin of another of the isotopes involved in the formation of the same chemical bond;
(b) 2k ≠ A-spin ≠ 2k + 1, where k ϵ Z, A-spin denotes atomic spin of individual isotope without any chemical (covalent) bond with other chemical elements;
(3) The neutron BADz phenomenon is observed in biological systems where there are chemical bonds between the atoms with nuclear spins of opposite signs, as follows: R-spin: − +, where “+” and “–” show the parity of two nuclei involved in the formation of the same chemical bond;
(4) Isotopic resonance intensity slightly decreases under each isolated condition (postulates), which are presented below:
Nn − Np = 2k + 1 >> 2k ≠ A-spin + A’-spin ≠ 2k + 1 >> R-spin: − +
where k ε Z, N denotes number, n denotes neutrons, p denotes protons, “+” and “−” show the parity of two nuclei involved in the formation of the same chemical bond.
(5) A nonadditive amplification of isotopic resonance can be expected if any two conditions (postulates) among the first three conditions listed above are paired (points 1 and 2a, 1 and 3, or 2a and 3);
(6) The most favorable conditions for maximum isotopic resonance in biological systems are the first three of the abovementioned conditions (postulates 1–3), which were confirmed via experimental evaluations [84];
(7) In metal atoms (Me), in the absence of covalent or coordinate bonds, isotopic resonance is proportional to the properties of the metal atom (i.e., the difference between its neutrons and protons, nuclear spin, and parity of the nucleus).
It is possible to explain the mechanism of the BADz phenomenon by a violation of the mass balance in systems with equilibrium charge.
Charge equilibrium in an atom is achieved through the interaction of protons and electrons:
q [p+ + e] = 0 (equilibrium of charge).
Mass equilibrium in an atom is achieved by the interactions of protons with neutrons and those of paired neutrons:
a.m. [p+ + n0] ≈ 1:1 (equilibrium of atomic mass),
a.m. [ni0 + ni+10] = 1:1 (equilibrium of atomic mass).
In triads containing a proton, electron, and neutron, one can observe charge and mass equilibriums. The stronger interactions between charged particles (protons and electrons) compared to mass effects explain the absence of mass imbalance in the nuclei of atoms with a smaller number of neutrons than protons. At the same time, the presence of a neutron that is not compensated in mass can cause a mass-dependent imbalance in systems with equilibrium charge, which is characteristic of some heavy isotopes or bonds formed by them:
[p+ + e + ni0] · ni+10 = 0, which is equilibrium of charge.
However, this pattern is observed ([p+ + e + ni0] · ni+10) ≠ equilibrium of mass (such as 1proton:1neutron); that is, non-equilibrium of mass (in model with non-compensated neutron or neutrons: 1proton:1neutron + odd neutron or neutrons).
A neutron not compensated in mass (ni+10) does not affect the whole nucleus at once, but probably affects each triad (proton–electron–neutron) stochastically in time; this causes the mass effect (which is proportional to at least one-half of the equilibrium proton–neutron interactions), leading indirectly to a change in the forces of the interactions between charged particles (proton–electron interactions). The latter is confirmed by the fact that it is the hydrogen bonds that are extremely susceptible to the electron density distribution throughout the molecule [94,95]; therefore, local attenuation and amplification of the proton–electron interactions can cause tunnel effects [96].
In addition, in a system with three neutrons, the abovementioned effect can also arise because of the incapability of neutron pairs to compensate for mass fluctuations of a non-compensated neutron in an atom; however, in isotopes containing at least five more neutrons than the number of protons, partial mass equilibrium can occur because of the manner of distribution of these neutrons, which can decrease the intensity of the BADz phenomenon.
Another possible mechanism capable of significantly increasing the rate of enzyme reactions is the ability of a neutron not compensated in mass to initiate quantum tunneling by involving one of the above-described atomic triads in this process; there is a subsequent release of energy sufficient to form a new chemical bond, which can dramatically accelerate the process of substrate formation necessary for the growth of cellular structures and the overall development of organisms. This effect offers an explanation for the “isotopic shock” phenomenon observed in living systems with heavy nonradioactive isotopes of certain macro- and microelements, as well as the exponential enhancement of these manifestations when combining different isotope fractions in biological objects.
Everything stated above is highly possible, as there are certain examples of collective proton tunneling in some systems (a Zundel-like complex, [HO⋅⋅⋅D⋅⋅⋅OH]) [97]; this is because for some electron–nuclear magnetic interactions in particular electron spins, selectivity is observed with regard to the nuclear spin.
It should be noted that in living systems constituting parts of organic molecules, the BADz phenomenon will occur not in pure (isolated) isotopes, but within groups of atoms bound by covalent and noncovalent interactions. Therefore, the BADz phenomenon should be calculated at least for each atomic pair with overlapping electronic clouds. In this regard, a linear increase in the weight of isotopes that form a chemical bond will cause no linear increase in the BADz phenomenon (for example, in a 13C–1H bond, the neutron number is equal to 0 and there is no probability of isotope resonance [98]; in a 13C–2H bond, the neutron number is equal to 1, and the occurrence of isotopic resonance is highly probable (i.e., BADz phenomenon takes place); in an 18O–1H bond, where the neutron number equals 1, isotopic resonance is expected, whereas in an 18O–2H bond, the neutron number is equal to 2, and isotopic resonance is not expected). This can explain the ambiguous results of studies carried out by many researchers on the enrichment of biological systems with heavy isotopes and their mixtures [99]. The enrichment of 13C and 2H cancer cells, when the formation of isotope resonance pairs is expected (the neutron number for 13C–2H equals 1), can be considered as proving the BADz phenomenon in practice. Similar fractionations of 12C/13C and 1H/2H with the accumulation of heavy atoms are accompanied by an additional advantage in terms of the energy and metabolism in oncocytes compared to cells with natural isotope compositions (the neutron number for 12C–1H equals –1, i.e., BADz phenomenon is not observed).
Further, an increase in the bond energy and oscillation frequency of a nucleus accompanied by a decrease in the internuclear distance can occur at different rates and with different intensities in bonds [100] formed by isotopes containing neutron pairs (12C–2H, 18O–2H) and those containing a non-compensated neutron (13C–2H). After some time, the difference between the initial energy and that observed before tunneling increases. The restriction of freedom in covalently bound resonance pairs of atoms (having a non-compensated neutron) causes an increase in the internal atomic energy, which ensures bond breakage without the necessity of achieving activation energy. This can explain the occurrence of the BADz phenomenon as one of the tunnel effect mechanisms for enzyme catalysis.
Based on the above discussions, it can be concluded that isotope fractionation in biological systems is only a prerequisite for the occurrence of isotope resonance; this resonance is observed only when the BADz phenomenon is realized, being associated with the process of incorporating heavy isotopes into biological molecules, and above all, their interaction with other light and heavy isotopes. Therefore, if the medium is excessively enriched with heavy isotopes only, the BADz phenomenon, and consequently isotopic resonance, cannot be observed in the following pairs: 18O–2H, 34S–2H, 13C–15N, 12C–18О, 13C–17О, 14N–18O, and 15N–17O.
When the concentration of deuterium in the environment is low, the body requires a much higher accumulation of 13C and 15N [87,101,102] than 18O and 34S to realize the required isotopic effect and for survival (i.e., δ13C (organism) >> δ13C (environment) and δ15N (organism) >> δ15N (environment), whereas δ18O (organism) ≤ δ18O (environment) [103,104,105,106], and similarly for δ34S (organism) ≤ δ34S (environment)). In addition, a more pronounced isotopic resonance effect of 17O and 18O compared to 13C and 15N [84] can be caused by a more active participation of oxygen atoms in functionally active groups (R–17/18O–1H) in comparison with nitrogen and carbon atoms. Therefore, modification of the natural medium by enriching it with 18O will lead to an increase in catalytic activity or transcription rates, accompanied by an increase in the number of metabolically active molecules in cells (enzymes) or a qualitative increase in their resistance to adverse intracellular effects (cellular stress); however, enrichment of the natural medium through the introduction of 13C and 15N into its composition is more likely to be accompanied by a smaller isotopic resonance due to slowing down of the nonspecific decomposition of molecular structures upon isotopic weighting of some chemical groups (heterocycles, isoprenoids, condensed polycyclic and polyunsaturated structures, and noncovalent interactions in macromolecular complexes). In this regard, the possibility of isotopic shocks is high in metabolically active cells and tissues, for example, during their fission and growth; however, these shocks occur only in those cases where white noise, in terms of the amplitude of energy oscillations, accounts for approximately 50% of the energy potential of cells, thus initiating the occurrence of stochastic resonance. Therefore, the same isotope can resonate at different phases of the cell cycle, but it would only take place at different isotope concentrations and with different intensities of resonance, or vice versa; the same isotope pair would inhibit cellular effects upon exposure that was too weak (e.g., less than 40%, that significantly less above-mentioned 50% required for resonance phenomenon, i.e., <<50%) or too strong (more than 60% or >>50%). This also explains the ability of some combinations of isotopes in a specific cell cycle to cause super resonance, a phenomenon described by Zubarev [84].
Despite the fact that monitoring the growth of bacteria is limited by the range of temperatures of their vital activity, this effect can be explained by the following experimental data that explore the temperature effect on the growth rate of E. coli on the culture media with 13C enrichment. It was grown in a temperature ranging from 15 °C to 41 °C. It was noted that the maximum growth rates of E. coli were always higher in the 13C enrichment media than in the isotopically normal media [84].
This generally confirms the possibility of realization of the neutron effect (BADz phenomenon) due to electron density delocalization in the nuclei of heavy nonradioactive isotopes, even with zero (and also other integers) values of nuclear spin and the absence of magnetic moments.
Moreover, the known facts of selective isotope accumulation in biological objects in comparison with the medium, caused by the interaction of different nonradioactive isotopes with each other leading to the formation of resonance pairs, can serve as indirect proof of the BADz phenomenon. For example, at high concentrations of deuterium in the environment, in order to realize the necessary isotopic effect and to survive, the body requires much higher accumulation of 18O and 34S than 13C and 15N (i.e., δ18O (organism) > δ18O (environment) and δ34S (organism) > δ34S (environment), whereas δ13C (organism) ≈ δ13C (environment); similarly, δ15N (organism) ≈ δ15N (environment) [102]). Correspondingly, more pronounced negative correlations between δ13C and δ2H, δ15N, and δ2H are expected, while between δ18O and δ2H, as well as δ34S and δ2H, weak positive correlations or absence thereof will be observed [107]. Also, this is confirmed by some data about human infant ability to assimilate more 13C content and more 15N in fingernails than food sources contain [108], and excrete less 2H and 18O in urine [109].
The difference in the resonance effects shown by Zubarev [84], which is characterized by a 3-fold effect that is greater for 13C compared to 15N, can be explained (under the condition of equally pronounced neutron BADz phenomenon) by a much higher ability of the organism to absorb 13C from the enriched 13C diet (43.2–60.8% of the 80% enriched 13C diet [110], and also in [111]), compared to its ability to absorb 15N from the enriched 15N-diet (by 250% from 95% enriched 15N-diet [99,112]). At the same time, the co-enrichment of the medium with 13C and 15N, on the contrary, reduces the growth rate of the living organisms (for example, saprotrophic zygomycete fungus Absidia cylindrospora [113] or Poecilia reticulata guppies [86]), which is explained by the absence of a non-compensated neutron effect in these biological systems. This is because each isotope has 1 non-compensated neutron, which together form a complete pair of neutrons that do not allow the realization of the BADz phenomenon. However, in [114], it was shown that the parthenogenic rotifer Brachionus plicatilis receiving 15N resonance diet had higher fitness with a longer lifespan and similar or greater lifetime fecundity as compared to all other treatments, including the control group.
Another example is the fractionation of isotopes of various metals when they interact with individual isotopes of nonmetals. In this case, the BADz phenomenon will be realized, both in the presence of stable Me–O bonds, including coordination bonds, and in the absence of stable covalent Me–O bonds, for example in structures such as Me+...O. The reactions based on the realization of the isotope BADz phenomenon will occur due to predominant participation of a metal ion [32,115] and its own non-compensated neutron (or, more rarely, an odd number of neutrons). For example, owing to the fact that the neutron BADz phenomenon in CuCO3 compounds is realized via 63/65Cu–17/18O bonds, the accumulation in this salt of both 17O [116] and 18O isotopes is expected, and therefore the ratio [δ17O(organism)17O(environment)] can be slightly bigger or even equal (i.e., ≥) to the ratio [δ18O(organism)18O(environment)], unlike 40CaCO3, in which 17O will be predominantly accumulated (i.e., ratio [δ17O(organism)17O(environment)] can be significantly lager (>>>) than ratio [δ18O(organism)18O(environment)]). In 43CaCO3, the differences in the isotopic ratios of 17O and 18O will be substantially less pronounced due to the fractionation of both atoms; for example, the ratio [δ17O(organism)17O(environment)] can be moderately larger or even equal (> or ≥) to ratio [δ18O(organism)18O(environment)].
Thus, the created hypothesis explains the BADz phenomenon through the occurrence of isotopic resonance, both with decreasing and increasing 2H, and through other non-radioactive isotopes in living and natural systems due to the appearance of resonant pairs between chemical bound atoms, whose interaction result in an uncompensated neutron.

6. Possibility of Using the Isotopic Therapy in 2H-Modified Drinking and Diet for Some Diseases

Using knowledge of isotope resonance through the BADz phenomenon will also allow more efficient poly-isotope correction of metabolic disorders in various diseases by creating conditions for targeted fractionations of isotopes in certain pathological cells and certain focuses of the disease. The latter is relevant because in current medical practice, increasing numbers of different metabolic disorders and cell pathologies are corrected by mostly monoisotopic medicines, especially 2H-isotope-modified drinking and diet therapies in healing the following diseases and certain conditions:
Alzheimer’s disease was treated by deuterium-reinforced polyunsaturated fatty acids (D-PUFAs), which improved cognition in mice fed a Western-type diet [117], and reduced the brain tissue concentrations of oxidation products and amyloid β-peptide in APP/PS1 mutant transgenic mice [118];
Parkinson’s disease was partially cured by the administration of D-PUFAs, which protected against nigrostriatal damage (in the substantia nigra) from oxidative injury in mice with a 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine model [119];
Mitochondrial disorders and oxidative stress were corrected by both deuterium-depleted water [16,55,56,65,69,120,121] and D-PUFAs, such as 11,11-D(2)-linoleic and 11,11,14,14-D(4)-α-linolenic acids and others, which preserved mitochondrial bioenergetics function under toxic effects of t-butylhydroperoxide, ethacrynic acid, or ferrous iron [122], prevented coenzyme Q (ubiquinone or CoQ)-less coq mutants and some toxic effects of lipid autoxidation products [123,124,125], and also decelerated the aging process by deuterated trilinolenin (D-TG (54:9)), resulting in extended lifespan of worms under normal and oxidative stress conditions [126];
Cerebral hypoxia was corrected by deuterium-depleted water, the long-term consumption of which resulted in activation of neuron protective systems, with higher antioxidant activity in brain tissues [68,75];
Friedreich’s ataxia was treated by deuterated linoleic and α-linolenic acids, which resulted in rescue of oxidative-stress-challenged cells by decreasing lipid peroxidation [127], and also deuterated ethyl linoleate (RT001), which recovered mitochondrial function [128];
Huntington’s disease was managed by D-PUFA diet in Q140 knock-in mice, which were able to demonstrate improved performance in novel object recognition tests after 5 months of consumption of deuterium-enriched diet [129];
Cancer adjuvant therapy showed efficacy through the administration of 1H2O [22,70,130,131], and previously-drug-processed gliomas was treated by combining deuterium-depleted drinking and a lower deuterated diet, which led to intracellular deuterium disequilibrium and formation of the 2H/1H gradient [62,132];
Inflammatory diseases were fine-tuned to potentially various inflammation stages using deuterated isotopologues of eicosanoids (for example, arachidonic acids with 2H at the bis-allylic C7, C10, and C13 positions [133]);
Pathological signaling pathways in diabetes were corrected by deuterium-depletion in drinking and diet [17,134,135];
Cardiovascular diseases were treated by improving cholesterol handling and reducing atherosclerosis development using D-PUFAs, which decreased lipid peroxidation, body weight gain, and plasma and hepatic cholesterol contents [136], as well as by using D-PUFAs to reduce the tocopherol-mediated peroxidation of lipids in human low-density lipoproteins [137].
Moreover, some effects of deuterium-reinforced amino acids (DRAAs) in the pathological pathways were recently shown (especially under the conditions of deficiency of the corresponding enzyme), such as procollagen-lysine 5-dioxygenases [138], protein-lysine 6-oxidase [139], and DNA methyltransferase 3A [140]. Also, synthesized (in vitro) hexapeptide containing asparaginyl residue with three carbon-bound atoms replaced with deuterium atoms (by reducing the speed of racemization) stabilized polypeptides, limiting the accumulation of some of the peptides and proteins in aging products [141].
The powerful effects of the organism of the selective 2H-enriched pharmaceuticals on biochemical potency are not known, which enable substantial benefits in the treatment of some diseases [142], but according to another study, livestock products were more effective in the therapy of cyclophosphamide-induced immunodeficiency after extraction of immunoactive proteins by deuterium-depleted media [52]. This points to a similar effect of heavy-isotope fluctuations, relating frequently to both their increase and decrease [37,143,144], with approximately the same number of fluctuations influencing living systems due to isotopic discrimination “up and down” in different stages of the lifespan [109]. Considering the phenomenon of isotope resonance in living systems can expand the healing possibilities of adjuvant and other types of therapies via reasonable poly-isotopic correction, this could be more effective than monoisotopic treatment.

7. Conclusions

In conclusion, it should be noted that the intensity of the isotope resonance can vary depending on the concentrations of the isotopes [144]: at low concentrations of potentially resonant isotopes (e.g., due to interaction deuterium with atoms such as 13C, 15N, and 17O), their thermodynamic and kinetic effects are predominantly realized (which are characterized by relatively insignificant differences in the fractionation rates), whereas at high concentrations of the same isotopes, the probability of the resonant pair formation with the subsequent appearance of the valence isotope BADz phenomenon, allowing for implementation of tunneling, causes anomalous (or paradoxical) isotope effects in the same biological reactions.
Although the monoisotopic-modified drinking and diet showed some efficacy in treating many pathologies (especially by changing the 2H/1H ratio in cells using deuterium depleted water, D-PUFAs, or DRAAs), such as Alzheimer’s disease, Friedreich’s ataxia, Huntington’s disease, brain cancer, cardiovascular and inflammatory diseases, and mitochondrial disorders, including Parkinson’s disease, diabetes, and cerebral hypoxia, polyisotopic resonance therapy in living systems could become a more effective type of therapy than the monoisotopic treatment.
In addition, it is important to emphasize that in living systems, the phenomenon known as “isotopic shock” can be realized by forming an isotope gradient that stimulates the work nonspecific protection system, leading to the accumulation of biologically active protective factors in the body (e.g., intracellular messengers and reactive oxygen species). These factors, together with the thermodynamic and kinetic effects described above, as well as neutron tunneling, are the driving forces causing a pronounced variety of isotopic compositions observed in biological objects, depending not only on the isotope composition of the medium but also on the functional activity of the organism itself, as well as peculiarities in the interactions of different isotopes among themselves when fractionated at different morphofunctional levels in the biological object itself.
The timeline of the reviewed papers is from 1960 to 2019.

Author Contributions

Conceptualization, methodology, writing—review and editing, A.B. and S.D.; writing—original draft preparation, L.F. and E.V.

Funding

This work was supported by The Russian Science Foundation (project №15-16-00008).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Schmidt, H.L.; Robins, R.J.; Werner, R.A. Multi-factorial in vivo stable isotope fractionation: Causes, correlations, consequences and applications. Isot. Environ. Health Stud. 2015, 51, 155–199. [Google Scholar] [CrossRef] [PubMed]
  2. Basov, A.A.; Fedulova, L.V.; Baryshev, M.G.; Dzhimak, S.S. Deuterium-depleted water influence on the isotope 2H/1H regulation in body and individual adaptation. Nutrients 2019, 11, 1903. [Google Scholar] [CrossRef] [PubMed]
  3. Galimov, E.M. On an erroneous “experience” of isotope fractionation in enzymatic reaction. Russ. J. Phys. Chem. A 2007, 81, 831–835. [Google Scholar] [CrossRef]
  4. Brenna, J.T. Natural intramolecular isotope measurements in physiology: Elements of the case for an effort toward high-precision position-specific isotope analysis. Rapid Commun. Mass Spectrom. 2001, 15, 1252–1262. [Google Scholar] [CrossRef] [PubMed]
  5. Hayes, J.M. Isotopic order, biogeochemical processes, and earth history. Geochim. Cosmochim. Acta 2004, 68, 1691–1700. [Google Scholar] [CrossRef]
  6. Walter, W.D.; Kurle, C.M.; Hopkins, J.B. Applications of stable isotope analysis in mammalian ecology. Isot. Environ. Health Stud. 2014, 50, 287–290. [Google Scholar] [CrossRef]
  7. Kosenkov, A.V.; Gulyaev, M.V.; Anisimov, N.V.; Lobyshev, V.I.; Pirogov, Y.A. Investigation of the distribution of heavy nuclei in laboratory animals using multinuclear magnetic resonance imaging. Phys. Wave Phenom. 2015, 23, 311–315. [Google Scholar] [CrossRef]
  8. Pedersen, L.G.; Bartolotti, L.; Li, L. Deuterium and its role in the machinery of evolution. J. Theor. Biol. 2006, 238, 914–918. [Google Scholar] [CrossRef]
  9. Podlesak, D.W.; Torregrossa, A.-M.; Ehleringer, J.R.; Dearing, M.D.; Passey, B.H.; Cerling, T.E. Turnover of oxygen and hydrogen isotopes in the body water, CO2, hair, and enamel of a small mammal. Geochim. Cosmochim. Acta 2008, 72, 19–35. [Google Scholar] [CrossRef]
  10. Podlesak, D.W.; Bowen, G.J.; O’Grady, S.; Cerling, T.E.; Ehleringer, J.R. δ2H and δ18O of human body water: A GIS model to distinguish residents from non-residents in the contiguous USA. Isot. Environ. Health Stud. 2012, 48, 259–279. [Google Scholar] [CrossRef]
  11. Kelly, S.; Heaton, K.; Hoogewerff, J. Tracing the geographical origin of food: The application of multi-element and multi-isotope analysis. Trends Food Sci. Technol. 2005, 16, 555–567. [Google Scholar] [CrossRef]
  12. Hobson, K.A.; Bond, A.L. Extending an indicator: Year-round information on seabird trophic ecology from multiple-tissue stable-isotope analyses. Mar. Ecol. Prog. Ser. 2012, 461, 233–243. [Google Scholar] [CrossRef]
  13. Galimov, E.M. Life is a product of molecular ordering “machine”. Geochem. Int. 2014, 52, 1190–1196. [Google Scholar] [CrossRef]
  14. Ivlev, A.A.; Knyazev, Y.A.; Logachev, M.F. Short-period variations in carbon isotope composition of CO2 of expired air under different physiological conditions in man. Biofizika 1996, 41, 508–516. [Google Scholar] [PubMed]
  15. Frolov, V.Y.; Baryshev, M.G.; Bolotin, S.N.; Dzhimak, S.S. Method of Producing Biologically Active Drinking Water with Low Deuterium Content. Patent RF 2438765, 10 January 2012. [Google Scholar]
  16. Rasooli, A.; Fatemi, F.; Akbarzadeh, K.; Dini, S.; Bahremand, S. Synergistic protective activity of deuterium depleted water (DDW) and saturejarechingeri essential oil on hepatic oxidative injuries induced by acetaminophen in rats. J. Essent. Oil Bear. Plants 2016, 19, 1086–1101. [Google Scholar] [CrossRef]
  17. Lisitsy, A.B.; Bogatyrev, A.N.; Dydykin, A.S.; Derevitskaya, O.K.; Soldatova, N.E.; Fedulova, L.V. Influence of meat semiprepared foods produced with an addition of water having a reduced deuterium content on the indicators of the laboratory animals with the model of alloxan diabetes. Vopr. Pitan. 2017, 86, 64–71. [Google Scholar]
  18. Chesson, L.A.; Valenzuela, L.O.; O’Grady, S.P.; Cerling, T.E.; Ehleringer, J.R. Hydrogen and oxygen stable isotope ratios of milk in the United States. J. Agric. Food Chem. 2010, 58, 2358–2363. [Google Scholar] [CrossRef]
  19. Bowen, G.J.; Winter, D.A.; Spero, H.J.; Zierenberg, R.A.; Reeder, M.D.; Cerling, T.E.; Ehleringer, J.R. Stable hydrogen and oxygen isotope ratios of bottled waters of the world. Rapid Commun. Mass Spectrom. 2005, 19, 3442–3450. [Google Scholar] [CrossRef]
  20. Mladin, C.; Ciobica, A.; Lefter, R.; Popescu, A.; Bild, W. Deuterium depletion induces anxiolytic-like effects in rats. Arch. Biol. Sci. 2014, 66, 947–953. [Google Scholar] [CrossRef]
  21. Stefanescu, I.; Manolescu, N.; Titescu, G.; Balint, E.; Tamaian, R.; Feurdean, V. Assessment of deuterium depletion in plymouth rock chickens’ bodies. Environ. Eng. Manag. J. 2010, 9, 1477–1480. [Google Scholar]
  22. Gyöngyi, Z.; Budán, F.; Szabó, I.; Ember, I.; Kiss, I.; Krempels, K.; Somlyai, I.; Somlyai, G. Deuterium depleted water effects on survival of lung cancer patients and expression of Kras, Bcl2, and Myc genes in mouse lung. Nutr. Cancer 2013, 65, 240–246. [Google Scholar] [CrossRef] [PubMed]
  23. Dzhimak, S.S.; Basov, A.A.; Elkina, A.A.; Fedulova, L.V.; Kotenkova, E.A.; Vasilevskaya, E.R.; Lyasota, O.M.; Baryshev, M.G. Influence of deuterium-depleted water on hepatorenal toxicity. Jundishapur. J. Nat. Pharm. Prod. 2018, 13, e69557. [Google Scholar] [CrossRef]
  24. Storm-Suke, A.; Wassenaar, L.I.; Nol, E.; Norris, D.R. The influence of metabolic rate on the contribution of stable-hydrogen and oxygen isotopes in drinking water to quail blood plasma and feathers. Funct. Ecol. 2012, 26, 1111–1119. [Google Scholar] [CrossRef]
  25. Lobyshev, V.I.; Kalinichenko, L.P. Isotopic Effects of D2O in Biological Systems; Nauka: Moscow, Russia, 1978; p. 215. [Google Scholar]
  26. Buchachenko, A.L. Photooxidation of water by molecular oxygen: Isotope exchange. Theor. Exp. Chem. 2011, 46, 359–362. [Google Scholar] [CrossRef]
  27. Buchachenko, A.L.; Kuznetsov, D.A.; Breslavskaya, N.N.; Shchegoleva, L.N.; Arkhangelsky, S.E. Calcium induced ATP synthesis: Isotope effect, magnetic parameters and mechanism. Chem. Phys. Lett 2011, 505, 130–134. [Google Scholar] [CrossRef]
  28. Buchachenko, A.L.; Kuznetsov, D.A.; Breslavskaya, N.N. Chemistry of enzymatic ATP synthesis: An insight through the isotope window. Chem. Rev. 2012, 112, 2042–2058. [Google Scholar] [CrossRef]
  29. Buchachenko, A.L.; Kuznetsov, D.A. Efficiency of ATP synthase as a molecular. Biophysics 2008, 53, 219–222. [Google Scholar] [CrossRef]
  30. Buchachenko, A.L.; Orlov, A.P.; Kuznetsov, D.A.; Breslavskaya, N.N. Magnetic control of the DNA synthesis. Chem. Phys. Lett 2013, 586, 138–142. [Google Scholar] [CrossRef]
  31. Buchachenko, A.L. Magnetic Isotope Effect in Chemistry and Biochemistry; Nova Science Publ.: New York, NY, USA, 2009; p. 105. [Google Scholar]
  32. Koltover, V.K.; Labyntseva, R.D.; Kosterin, S.O.; Karandashev, V.K. Magnetic isotope of magnesium accelerates ATP hydrolysis catalyzed by myosin. Biophysics 2016, 61, 200–206. [Google Scholar] [CrossRef]
  33. Koltover, V.K.; Avdeeva, L.V.; Kudryashova, E.A.; Shevchenko, U.G.; Royba, E.A.; Berdinsky, V.L. Magnetic-isotope effect of magnesium in the living cell. Dokl. Biochem. Biophys. 2012, 442, 12–14. [Google Scholar] [CrossRef]
  34. Xie, X.; Zubarev, R.A. Effects of low-level deuterium enrichment on bacterial growth. PLoS ONE 2014, 9, e102071. [Google Scholar] [CrossRef] [PubMed]
  35. Nikitin, D.I.; Oranskaya, M.N.; Lobyshev, V.I. Specificity of bacterial response to variations of the isotopic composition of water. Biophysics 2003, 48, 636–640. [Google Scholar]
  36. Stefanescu, I.; Manolescu, N.; Titescu, G.; Comisel, V.; Tamaian, R.; Feurdean, V. Studies concerning DDW use for deuterium depletion in the organism of laboratory animals. Environ. Eng. Manag. J. 2010, 9, 1473–1476. [Google Scholar] [CrossRef]
  37. Rodin, S.; Rebellato, P.; Lundin, A.; Zubarev, R.A. Isotopic resonance at 370 ppm deuterium negatively affects kinetics of luciferin oxidation by luciferase. Sci. Rep. 2018, 8, 16249. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Ek, C.; Garbaras, A.; Yu, Z.; Oskarsson, H.; Wiklund, A.-K.E.; Kumblad, L.; Gorokhova, E. Increase in stable isotope ratios driven by metabolic alterations in amphipods exposed to the beta-blocker propranolol. PLoS ONE 2019, 14, e0211304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Buchachenko, A.L.; Pliss, E.M. Hydrogen isotope effects and atom tunneling. Russ. Chem. Rev. 2016, 85, 557–564. [Google Scholar] [CrossRef]
  40. Shchepinov, M.S. Reactive oxygen species, isotope effect, essential nutrients, and enhanced longevity. Rejuvenation Res. 2007, 10, 47–59. [Google Scholar] [CrossRef]
  41. Li, X.; Snyder, M.P. Yeast longevity promoted by reversing aging-associated decline in heavy isotope content. Npj Aging Mech. Dis. 2016, 2, 16004. [Google Scholar] [CrossRef]
  42. Heazlewood, B.R.; Maccarone, A.T.; Andrews, D.U.; Osborn, D.L.; Harding, L.B.; Klippenstein, S.J.; Jordan, M.J.; Kable, S.H. Near-threshold H/D exchange in CD3CHO photodissociation. Nat. Chem. 2011, 3, 443–448. [Google Scholar] [CrossRef]
  43. Andrianova, Z.S.; Buchachenko, A.L.; Breslavskaya, N.N.; Pliss, E.M. Bond energies in polyunsaturated acids and kinetics of co-oxidation of protiated and deuterated acids. Russ. J. Phys. Chem. A 2016, 90, 1936–1941. [Google Scholar] [CrossRef]
  44. Martin, G.J.; Zhang, B.L.; Naulet, N.; Martin, M.L. Deuterium transfer in the bioconversion of glucose to ethanol studied by specific isotope labeling at the natural abundance level. J. Am. Chem. Soc. 1986, 108, 5116–5122. [Google Scholar] [CrossRef]
  45. Zhang, X.; Gillespie, A.L.; Sessions, A.L. Large D/H variations in bacterial lipids reflect central metabolic pathways. Proc. Natl. Acad. Sci. USA 2009, 106, 12580–12586. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Fischer, C.R.; Bowen, B.P.; Pan, C.; Northen, T.R.; Banfield, J.F. Stable-isotope probing reveals that hydrogen isotope fractionation in proteins and lipids in a microbial community are different and species-specific. ACS Chem. Biol. 2013, 8, 1755–1763. [Google Scholar] [CrossRef] [PubMed]
  47. Bila, W.C.; Mariano, R.M.S.; Silva, V.R.; dos Santos, M.E.S.M.; Lamounier, J.A.; Ferriolli, E.; Galdino, A.S. Applications of deuterium oxide in human health. Isot. Environ. Health Stud. 2017, 53, 327–343. [Google Scholar] [CrossRef] [PubMed]
  48. Lewis, E.R.; Johansen, E.; Holman, T.R. Large competitive kinetic isotope effects in human 15-lipoxygenase catalysis measured by a novel HPLC method. J. Am. Chem. Soc. 1999, 121, 1395–1396. [Google Scholar] [CrossRef]
  49. Uspenskaya, E.V.; Anfimova, E.V.; Syroeshkin, A.V.; Pleteneva, T.V. Kinetics of active pharmaceutical ingredient solubility in water with different hydrogen isotopic content. Indian J. Pharm. Sci. 2018, 80, 318–324. [Google Scholar] [CrossRef]
  50. Vasilevskaya, E.R.; Akhremko, A.G. Proteomic study of pig’s spleen. Potravin. Slovak J. Food Sci. 2019, 13, 314–317. [Google Scholar] [CrossRef]
  51. Fedulova, L.V.; Vasilevskaya, E.R.; Kotenkova, E.A.; Elkina, A.A.; Baryshev, M.G.; Lisitsyn, A.B. Influence of different polypeptides fractions derived from Sus Scrofa immune organs on the rats immunological reactivity. J. Pharm. Nutr. Sci. 2017, 7, 35–40. [Google Scholar] [CrossRef] [Green Version]
  52. Fedulova, L.V.; Basov, A.A.; Vasilevskaya, E.R.; Dzhimak, S.S. Gender difference response of male and female immunodeficiency rats treated with tissue-specific biomolecules. Curr. Pharm. Biotechnol. 2019, 20, 245–253. [Google Scholar] [CrossRef]
  53. Fedulova, L.; Elkina, A.; Vasilevskaya, E.; Barysheva, E. Identification of tissue-specific proteins of immunocompetent organs of Sus scrofa isolated in deuterium depleted medium. Med. Sci. 2018, 22, 509–513. [Google Scholar]
  54. Darad, R.; Aiyar, A.S. Effects of heavy water on mitochondrial respiration and oxidative phosphorylation. J. Biosci. 1982, 4, 159–166. [Google Scholar] [CrossRef]
  55. Kolesova, O.E.; Pomytkin, I.A. Relationship between natural concentration of heavy water isotopologs and rate of H2O2 generation by mitochondria. Bull. Exp. Biol. Med. 2006, 142, 570–572. [Google Scholar]
  56. Dzhimak, S.S.; Basov, A.A.; Volchenko, N.N.; Samkov, A.A.; Fedulova, L.V.; Baryshev, M.G. Changes in the functional activity of mitochondria isolated from the liver of rat that passed the preadaptation to ultra-low deuterium concentration. Dokl. Biochem. Biophys. 2017, 476, 323–325. [Google Scholar] [CrossRef] [PubMed]
  57. Olgun, A. Biological effects of deuteronation: ATP synthase as an example. Theor. Biol. Med. Model. 2007, 4, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Wanders, R.J.A.; Waterham, H.R.; Ferdinandusse, S. Metabolic interplay between peroxisomes and other subcellular organelles including mitochondria and the endoplasmic reticulum. Front. Cell Dev. Biol. 2015, 3, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Kotkina, I.T.; Titov, V.N.; Parkhimovich, R.M. The different notions about beta-oxidation of fatty acids in peroxisomes, peroxisomes and ketonic bodies. The diabetic, acidotic coma as an acute deficiency of acetyl-CoA and ATP. Klin. Lab. Diagn 2014, 3, 14–23. [Google Scholar]
  60. Pollak, N.; Dolle, C.; Ziegler, M. The power to reduce: Pyridine nucleotides—small molecules with a multitude of functions. BioChem. J. 2007, 402, 205–218. [Google Scholar] [CrossRef]
  61. Lewis, C.A.; Parker, S.J.; Fiske, B.P.; McCloskey, D.; Gui, D.Y.; Green, C.R.; Vokes, N.I.; Feist, A.M.; Heiden, M.G.V.; Metallo, C.M. Tracing compartmentalized NADPH metabolism in the cytosol and mitochondria of mammalian cells. Mol. Cell 2014, 55, 253–263. [Google Scholar] [CrossRef] [Green Version]
  62. Boros, L.G.; D’Agostino, D.P.; Katz, H.E.; Roth, J.P.; Meuillet, E.J.; Somlyai, G. Submolecular regulation of cell transformation by deuterium depleting water exchange reactions in the tricarboxylic acid substrate cycle. Med. Hypotheses 2016, 87, 69–74. [Google Scholar] [CrossRef]
  63. Sormendi, S.; Wielockx, B. Hypoxia pathway proteins as central mediators of metabolism in the tumor cells and their microenvironment. Front. Immunol. 2018, 9, 40. [Google Scholar] [CrossRef]
  64. Strekalova, T.; Evans, M.; Chernopiatko, A.; Couch, Y.; Costa-Nunes, J.; Cespuglio, R.; Chesson, L.; Vignisse, J.; Steinbusch, H.W.; Anthony, D.C.; et al. Deuterium content of water increases depression susceptibility: The potential role of a serotonin-related mechanism. Behav. Brain Res. 2015, 277, 237–244. [Google Scholar] [CrossRef] [PubMed]
  65. Avila, D.S.; Somlyai, G.; Somlyai, I.; Aschner, M. Anti-aging effects of deuterium depletion on Mn-induced toxicity in a C. elegans model. Toxicol. Lett. 2012, 211, 319–324. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Letuta, U.G.; Berdinskiy, V.L. Magnetosensitivity of bacteria E. coli: Magnetic isotope and magnetic field effects. Bioelectromagnetics 2017, 38, 581–591. [Google Scholar] [CrossRef] [PubMed]
  67. Cong, F.; Zhang, Y.; Sheng, H.; Ao, Z.; Zhang, S.; Wang, J. Deuterium-depleted water inhibits human lung carcinoma cell growth by apoptosis. Exp. Ther. Med. 2010, 1, 277–283. [Google Scholar] [CrossRef] [Green Version]
  68. Kozin, S.V.; Kravtsov, A.A.; Elkina, A.A.; Zlishcheva, E.I.; Barysheva, E.V.; Shurygina, L.V.; Moiseev, A.V.; Baryshev, M.G. Isotope exchange of deuterium for protium in rat brain tissues changes brain tolerance to hypoxia. Biophysics 2019, 64, 272–278. [Google Scholar] [CrossRef]
  69. Rasooli, A.; Fatemi, F.; Hajihosseini, R.; Vaziri, A.; Akbarzadeh, K.; Mohammadi Malayeri, M.R.; Dini, S.; Foroutanrad, M. Synergistic effects of deuterium depleted water and Mentha longifolia L. essential oils on sepsis-induced liver injuries through regulation of cyclooxygenase-2. Pharm. Biol. 2019, 57, 125–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Zhang, X.; Gaetani, M.; Chernobrovkin, A.; Zubarev, R.A. Anticancer effect of deuterium depleted water—Redox disbalance leads to oxidative stress. Mol. Cell Proteom. 2019. [Google Scholar] [CrossRef] [Green Version]
  71. Shchepinov, M.S. Do “heavy” eaters live longer? BioEssays 2007, 29, 1247–1256. [Google Scholar] [CrossRef]
  72. Dzhimak, S.S.; Basov, A.A.; Baryshev, M.G. Content of deuterium in biological fluids and organs: Influence of deuterium depleted water on D/H gradient and the process of adaptation. Dokl. Biochem. Biophys. 2015, 465, 370–373. [Google Scholar] [CrossRef]
  73. O’Grady, S.P.; Wende, A.R.; Remien, C.H.; Valenzuela, L.O.; Enright, L.E.; Chesson, L.A.; Dale Abel, E.; Cerling, T.E.; Ehleringer, J.R. Aberrant water homeostasis detected by stable isotope analysis. PLoS ONE 2010, 5, e11699. [Google Scholar] [CrossRef]
  74. Somlyai, G.; Jancsó, G.; Jákli, G.; Vass, K.; Barna, B.; Lakics, V.; Gaál, T. Naturally occurring deuterium is essential for the normal growth rate of cells. FEBS Lett. 1993, 317, 1–4. [Google Scholar] [CrossRef] [Green Version]
  75. Kravtsov, A.A.; Kozin, S.V.; Vasilevskaya, E.R.; Elkina, A.A.; Fedulova, L.V.; Popov, K.A.; Malyshko, V.V.; Moiseev, A.V.; Shashkov, D.I.; Baryshev, M.G. Effect of drinking ration with reduced deuterium content on brain tissue prooxidant-antioxidant balance in rats with acute hypoxia model. J. Pharm. Nutr. Sci. 2018, 8, 42–51. [Google Scholar] [CrossRef]
  76. Samkov, A.A.; Dzhimak, S.S.; Barishev, M.G.; Volchenko, N.N.; Khudokormov, A.A.; Samkova, S.M.; Karaseva, E.V. The effect of water isotopic composition on Rhodococcus erythropolis biomass production. Biophysics 2015, 60, 107–112. [Google Scholar] [CrossRef]
  77. Letuta, U.G.; Letuta, S.N.; Berdinskiy, V.L. The influence of low magnetic fields and magnesium isotopes on E. coli bacteria. Biophysics 2017, 62, 935–941. [Google Scholar] [CrossRef]
  78. Bild, W.; Năstasă, V.; Haulică, I. In vivo and in vitro research on the biological effects of deuterium-depleted water: 1. Influence of deuterium-depleted water on cultured cell growth. Rom. J. Physiol. 2004, 41, 53–67. [Google Scholar]
  79. Basov, A.A.; Elkina, A.A.; Samkov, A.A.; Volchenko, N.N.; Moiseev, A.V.; Fedulova, L.V.; Baryshev, M.G.; Dzhimak, S.S. Influence of deuterium depleted water on the isotope D/H composition of liver tissue and morphological development of rats at different periods of ontogenesis. Iran. Biomed. J. 2019, 23, 129–141. [Google Scholar] [CrossRef]
  80. Petrus-Vancea, A. Cell ultrastructure and chlorophyll pigments in hyperhydric and non-hyperhydric Beta vulgaris var. Conditiva plantlets, treated with deuterium depleted water. Plant Cell Tissue Organ. Cult. 2018, 135, 13–21. [Google Scholar] [CrossRef]
  81. Kosenkov, A.V.; Gulyaev, M.V.; Lobyshev, V.I.; Yusubalieva, G.M.; Baklaushev, V.P. The reversible effect of deuteration on tissue fluid and biopolymers in normal and tumor tissues of mice. Biophysics 2018, 63, 820–824. [Google Scholar] [CrossRef]
  82. Yavari, K.; Kooshesh, L. Deuterium depleted water inhibits the proliferation of human MCF7 breast cancer cell lines by inducing cell cycle arrest. Nutr. Cancer 2019. [Google Scholar] [CrossRef]
  83. Lobyshev, V.I. Biphasic response of biological objects on variation of low deuterium concentration in water. Int. J. High Dilution Res. 2018, 17, 12–13. [Google Scholar]
  84. Xie, X.; Zubarev, R.A. Isotopic resonance hypothesis: Experimental verification by Escherichia coli growth measurements. Sci. Rep. 2015, 5, 9215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Crotty, F.V.; Blackshaw, R.P.; Murray, P.J. Differential growth of the fungus Absidia cylindrospora on 13C/15N-labelled media. Rapid Comm. Mass Spectrom. 2011, 25, 1479–1484. [Google Scholar] [CrossRef] [PubMed]
  86. Dennis, C.A.; MacNeil, M.A.; Rosati, J.Y.; Pitcher, T.E.; Fisk, A.T. Diet discrimination factors are inversely related to δ15N and δ13C values of food for fish under controlled conditions. Rapid Commun. Mass. Spectrom. 2010, 24, 3515–3520. [Google Scholar] [CrossRef] [PubMed]
  87. Andriukonis, E.; Gorokhova, E. Kinetic 15N-isotope effects on algal growth. Sci. Rep. 2017, 7, 44181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  88. Letuta, U.G.; Tikhonova, T.A. Magnetic fields and magnetic isotope 25Mg effects on biofilms formation by bacteria E. coli. Dokl. Biochem. Biophys. 2019, 484, 85–87. [Google Scholar] [CrossRef] [PubMed]
  89. Waki, M.; Miyamoto, K. Isotope Enrichment Method. U.S. Patent 7,943,107, 17 May 2011. [Google Scholar]
  90. Letuta, U.G.; Shailina, D.M. Magnetosensitivity of E. coli bacteria in the presence of zinc isotopes. Dokl. Biochem. Biophys. 2018, 479, 111–113. [Google Scholar] [CrossRef]
  91. Letuta, U.G.; Berdinskiy, V.L. Effects of the magnetic field and zinc isotopes on the colony forming ability and elemental composition of E. coli bacterial cells. Russ. Chem. Bull. 2018, 67, 1732–1737. [Google Scholar] [CrossRef]
  92. Letuta, U.G.; Berdinskiy, V.L. Biological effects of static magnetic fields and zinc isotopes on E. coli bacteria. Bioelectromagnetics 2019, 40, 62–73. [Google Scholar] [CrossRef] [Green Version]
  93. Lysenko, O.B.; Demikhov, Y.N.; Skul’skii, N.A.; Sobotovich, E.V. The role of a magnetic effect in uranium isotope fractionation. Russ. J. Phys. Chem. B 2014, 8, 870–873. [Google Scholar] [CrossRef]
  94. Sobczyk, L.; Obrzud, M.; Filarowski, A. H/D isotope effects in hydrogen bonded systems. Molecules 2013, 18, 4467–4476. [Google Scholar] [CrossRef] [Green Version]
  95. Ishikita, H.; Saito, K. Proton transfer reactions and hydrogen-bond networks in protein environments. J. R. Soc. Interface 2014, 11, 20130518. [Google Scholar] [CrossRef] [PubMed]
  96. Xin, Y.-B.; Hu, Q.; Niu, D.-H.; Zheng, X.-H.; Shi, H.-L.; Wang, M.; Xiao, Z.-S.; Huang, A.-P.; Zhang, Z.-B. Research progress of hydrogen tunneling in two-dimensional materials. Acta Phys. Sin. 2017, 66, 056601. [Google Scholar] [CrossRef]
  97. Drechsel-Grau, C.; Marx, D. Exceptional isotopic-substitution effect: Breakdown of collective proton tunneling in hexagonal ice due to partial deuteration. Angew. Chem. Int. Ed. Engl. 2014, 53, 10937–10940. [Google Scholar] [CrossRef] [PubMed]
  98. Northrop, D.B. Unusual origins of isotope effects in enzyme-catalysed reactions. Philos. Trans. R. Soc. B Biol. Sci. 2006, 361, 1341–1349. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  99. Uphaus, R.A.; Flaumenhaft, E.; Katz, J.J. A living organism of unusual isotopic composition. Sequential and cumulative replacement of stable isotopes in Chlorella vulgaris. Biochim. Biophys. Acta 1967, 141, 625–632. [Google Scholar] [CrossRef]
  100. Sarsa, A.; Alcaraz-Pelegrina, J.M.; Le Sech, C. Isotopic effects on covalent bond confined in a penetrable sphere. J. Phys. Chem. B 2015, 119, 14364–14372. [Google Scholar] [CrossRef]
  101. Karlson, A.M.L.; Reutgard, M.; Garbaras, A.; Gorokhova, E. Isotopic niche reflects stress-induced variability in physiological status. R. Soc. Open Sci. 2018, 5, 171398. [Google Scholar] [CrossRef] [Green Version]
  102. Symes, C.; Skhosana, F.; Butler, M.; Gardner, B.; Woodborne, S. Isotope (δ13C, δ15N, δ2H) diet–tissue discrimination in African grey parrot Psittacus erithacus: Implications for forensic studies. Isot. Environ. Health Stud. 2017, 53, 580–596. [Google Scholar] [CrossRef]
  103. Borek, E.; Rittenberg, D. Anomalous growth of microorganisms produced by changes in isotopes in their environment. Proc. Natl. Acad. Sci. USA 1960, 46, 777–782. [Google Scholar] [CrossRef] [Green Version]
  104. Abeni, F.; Petrera, F.; Capelletti, M.; Dal Prà, A.; Bontempo, L.; Tonon, A.; Camin, F. Hydrogen and oxygen stable isotope fractionation in body fluid compartments of dairy cattle according to season, farm, breed, and reproductive stage. PLoS ONE 2015, 10, e0127391. [Google Scholar] [CrossRef]
  105. Reynard, L.M.; Burt, N.; Koon, H.E.C.; Tuross, N. Limits and possibilities in the geolocation of humans using multiple isotope ratios (H, O, N, C) of hair from east coast cities of the USA. Isot. Environ. Health Stud. 2016, 52, 498–512. [Google Scholar] [CrossRef] [PubMed]
  106. Kirsanow, K.; Tuross, N. Oxygen and hydrogen isotopes in rodent tissues: Impact of diet, water and ontogeny. Palaeogeogr. Palaeoclimatol. Palaeoecol. 2011, 310, 9–16. [Google Scholar] [CrossRef]
  107. Lysenko, O.B.; Skulskiy, N.A.; Shatilo, V.B.; Koshliakova, T.A.; Lysenko, M.O.; Sobotovich, E.V. Problems and prospects of the main biogenic elements intramolecular isotope correlations using as a new resource of different etiology pathologies diagnostics. Zhivye i Biokosnye Sistemy 2014, 8, 1–23. [Google Scholar]
  108. Fuller, B.T.; Fuller, J.L.; Harris, D.A.; Hedges, R.E. Detection of breastfeeding and weaning in modern human infants with carbon and nitrogen stable isotope ratios. Am. J. Phys. Anthropol. 2006, 129, 279–293. [Google Scholar] [CrossRef] [PubMed]
  109. Li, X.; Snyder, M.P. Can heavy isotopes increase lifespan? Studies of relative abundance in various organisms reveal chemical perspectives on aging. BioEssays 2016, 38, 1093–1101. [Google Scholar] [CrossRef] [Green Version]
  110. Gregg, C.T.; Hutson, J.Y.; Prine, J.R.; Ott, D.G.; Furchuer, J.E. Substantial replacement of mammalian body carbon with carbon-13. Life Sci. 1973, 13, 775–782. [Google Scholar] [CrossRef]
  111. Strable, M.S.; Tschanz, C.L.; Varamini, B.; Chikaraishi, Y.; Ohkouchi, N.; Brenna, J.T. Mammalian DNA δ15N exhibits 40‰ intramolecular variation and is unresponsive to dietary protein level. Rapid Commun. Mass Spectrom. 2011, 25, 3555–3562. [Google Scholar] [CrossRef]
  112. Felsing, M.; Telfer, T.; Glencross, B. 15N-enrichment of an aquaculture diet and tracing of cage culture waste in an estuarine environment. J. Appl. Ichthyol. 2006, 22, 419–426. [Google Scholar] [CrossRef]
  113. Crotty, F.V.; Blackshaw, R.P.; Murray, P.J. Tracking the flow of bacterially derived 13C and 15N through soil faunal feeding channels. Rapid Commun. Mass Spectrom. 2011, 25, 1503–1513. [Google Scholar] [CrossRef]
  114. Gorokhova, E. Shifts in rotifer life history in response to stable isotope enrichment: Testing theories of isotope effects on organismal growth. R. Soc. Open Sci. 2017, 4, 160810. [Google Scholar] [CrossRef] [Green Version]
  115. Avdeeva, L.V.; Koltover, V.K. Nuclear spin catalysis in living nature. Mosc. Univ. Chem. Bull. 2016, 71, 160–166. [Google Scholar] [CrossRef]
  116. Buchachenko, A.L. Mass-independent isotope effects. J. Phys. Chem. B 2013, 117, 2231–2238. [Google Scholar] [CrossRef] [PubMed]
  117. Elharram, A.; Czegledy, N.M.; Golod, M.; Milne, G.L.; Pollock, E.; Bennett, B.M.; Shchepinov, M.S. Deuterium-reinforced polyunsaturated fatty acids improve cognition in a mouse model of sporadic Alzheimer’s disease. FEBS J. 2017, 284, 4083–4095. [Google Scholar] [CrossRef] [PubMed]
  118. Raefsky, S.M.; Furman, R.; Milne, G.; Pollock, E.; Axelsen, P.; Mattson, M.P.; Shchepinov, M.S. Deuterated polyunsaturated fatty acids reduce brain lipid peroxidation and hippocampal amyloid β-peptide levels, without discernable behavioral effects in an APP/PS1 mutant transgenic mouse model of Alzheimer’s disease. Neurobiol. Aging 2018, 66, 165–176. [Google Scholar] [CrossRef] [PubMed]
  119. Shchepinov, M.S.; Chou, V.P.; Pollock, E.; Langston, J.W.; Cantor, C.R.; Molinari, R.J.; Manning-Bog, A.B. Isotopic reinforcement of essential polyunsaturated fatty acids diminishes nigrostriatal degeneration in a mouse model of Parkinson’s disease. Toxicol. Lett. 2011, 207, 97–103. [Google Scholar] [CrossRef]
  120. Zlatska, A.; Gordiienko, I.; Vasyliev, R.; Zubov, D.; Gubar, O.; Rodnichenko, A.; Syroeshkin, A.; Zlatskiy, I. In vitro study of deuterium effect on biological properties of human cultured adipose-derived stem cells. Sci. World J. 2018, 2018, e5454367. [Google Scholar] [CrossRef] [Green Version]
  121. Dzhimak, S.S.; Basov, A.A.; Fedulova, L.V.; Didikin, A.S.; Bikov, I.M.; Arcybasheva, O.M.; Naumov, G.N.; Baryshev, M.G. Correction of metabolic processes in rats during chronic endotoxicosis using isotope (D/H) exchange reactions. Biol. Bull. 2015, 42, 440–448. [Google Scholar] [CrossRef]
  122. Andreyev, A.Y.; Tsui, H.S.; Milne, G.L.; Shmanai, V.V.; Bekish, A.V.; Fomich, M.A.; Pham, M.N.; Nong, Y.; Murphy, A.N.; Clarke, C.F.; et al. Isotope-reinforced polyunsaturated fatty acids protect mitochondria from oxidative stress. Free Radic. Biol. Med. 2015, 82, 63–72. [Google Scholar] [CrossRef]
  123. Hill, S.; Hirano, K.; Shmanai, V.V.; Marbois, B.N.; Vidovic, D.; Bekish, A.V.; Kay, B.; Tse, V.; Fine, J.; Clarke, C.F.; et al. Isotope-reinforced polyunsaturated fatty acids protect yeast cells from oxidative stress. Free Radic. Biol. Med. 2011, 50, 130–138. [Google Scholar] [CrossRef]
  124. Hill, S.; Lamberson, C.R.; Xu, L.; To, R.; Tsui, H.S.; Shmanai, V.V.; Bekish, A.V.; Awad, A.M.; Marbois, B.N.; Cantor, C.R.; et al. Small amounts of isotope-reinforced polyunsaturated fatty acids suppress lipid autoxidation. Free Radic. Biol. Med. 2012, 53, 893–906. [Google Scholar] [CrossRef] [Green Version]
  125. Firsov, A.M.; Fomich, M.A.; Bekish, A.V.; Sharko, O.L.; Kotova, E.A.; Saal, H.J.; Vidovic, D.; Shmanai, V.V.; Pratt, D.A.; Antonenko, Y.N.; et al. Threshold protective effect of deuterated polyunsaturated fatty acids on peroxidation of lipid bilayers. FEBS J. 2019, 286, 2099–2117. [Google Scholar] [CrossRef] [PubMed]
  126. Beaudoin-Chabot, C.; Wang, L.; Smarun, A.V.; Vidovic, D.; Shchepinov, M.S.; Thibault, G. Deuterated polyunsaturated fatty acids reduce oxidative stress and extend the lifespan of C. Elegans. Front. Physiol. 2019, 10, 641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Cotticelli, M.G.; Crabbe, A.M.; Wilson, R.B.; Shchepinov, M.S. Insights into the role of oxidative stress in the pathology of Friedreich ataxia using peroxidation resistant polyunsaturated fatty acids. Redox. Biol. 2013, 1, 398–404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Zesiewicz, T.; Heerinckx, F.; De Jager, R.; Omidvar, O.; Kilpatrick, M.; Shaw, J.; Shchepinov, M.S. Randomized, clinical trial of RT001: Early signals of efficacy in Friedreich’s ataxia. Mov. Disord. 2018, 33, 1000–1005. [Google Scholar] [CrossRef] [PubMed]
  129. Hatami, A.; Zhu, C.; Relaño-Gines, A.; Elias, C.; Galstyan, A.; Jun, M.; Milne, G.; Cantor, C.R.; Chesselet, M.F.; Shchepinov, M.S. Deuterium-reinforced linoleic acid lowers lipid peroxidation and mitigates cognitive impairment in the Q140 knock in mouse model of Huntington’s disease. FEBS J. 2018, 285, 3002–3012. [Google Scholar] [CrossRef] [PubMed]
  130. Somlyai, G.; Javaheri, B.; Davari, H.; Gyöngyi, Z.; Somlyai, I.; Tamaddon, K.A.; Boros, L.G. Pre-clinical and clinical data confirm the anticancer effect of deuterium depletion. Biomacromol. J. 2016, 2, 1–7. [Google Scholar] [CrossRef]
  131. Shikhlyarova, A.I.; Zhukova, G.V.; Dzhimak, S.S.; Baryshev, M.G.; Kit, O.I.; Goncharova, A.S.; Shirnina, E.A.; Protasova, T.P.; Zhadobina, A.I.; Bragina, M.I.; et al. Influence of deuterium-depleted water as a nutritive factor on the effectiveness of experimental chemotherapy in aging rats. J. Clin. Oncol. 2019, 37, e14683. [Google Scholar] [CrossRef]
  132. Boros, L.G.; Que Collins, T.; Somlyai, G. What to eat or what not to eat—that is still the question. Neuro Oncol. 2017, 19, 595–596. [Google Scholar] [CrossRef] [Green Version]
  133. Chistyakov, D.V.; Filimonov, I.S.; Azbukina, N.V.; Goriainov, S.V.; Chistyakov, V.V.; Fomich, M.A.; Bekish, A.V.; Shmanai, V.V.; Sergeeva, M.G.; Shchepinov, M.S. Deuterated arachidonic acids library for regulation of inflammation and controlled synthesis of eicosanoids: An in vitro study. Molecules 2018, 23, 3331. [Google Scholar] [CrossRef] [Green Version]
  134. Yavari, K.; Gholamali, M.; Yazdian, F. The biological effects of deuterium depleted water: A possible tool in diabetes therapy. Electron. J. Biol. 2017, 13, 314–319. [Google Scholar]
  135. Rehakova, R.; Klimentova, J.; Cebova, M.; Barta, A.; Matuskova, Z.; Labas, P.; Pechanova, O. Effect of deuterium-depleted water on selected cardiometabolic parameters in fructose-treated rats. Physiol. Res. 2016, 65, S401–S407. [Google Scholar] [PubMed]
  136. Berbee, J.F.P.; Mol, I.M.; Milne, G.L.; Pollock, E.; Hoeke, G.; Lütjohann, D.; Monaco, C.; Rensen, P.C.N.; van der Ploeg, L.H.T.; Shchepinov, M.S. Deuterium-reinforced polyunsaturated fatty acids protect against atherosclerosis by lowering lipid peroxidation and hypercholesterolemia. Atherosclerosis 2017, 264, 100–107. [Google Scholar] [CrossRef] [PubMed]
  137. Lamberson, C.R.; Xu, L.; Muchalski, H.; Montenegro-Burke, J.R.; Shmanai, V.V.; Bekish, A.V.; McLean, J.A.; Clarke, C.F.; Shchepinov, M.S.; Porter, N.A. Unusual kinetic isotope effects of deuterium reinforced polyunsaturated fatty acids in tocopherol-mediated free radical chain oxidations. J. Am. Chem. Soc. 2014, 136, 838–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Korneenko, T.V.; Pestov, N.B.; Hurski, A.L.; Fedarkevich, A.M.; Shmanai, V.V.; Brenna, J.T.; Shchepinov, M.S. A strong developmental isotope effect in Caenorhabditis elegans induced by 5,5-deuterated lysine. Amino Acids 2017, 49, 887–894. [Google Scholar] [CrossRef] [PubMed]
  139. Pestov, N.B.; Okkelman, I.A.; Shmanai, V.V.; Hurski, A.L.; Giaccia, A.J.; Shchepinov, M.S. Control of lysyl oxidase activity through site-specific deuteration of lysine. Bioorg. Med. Chem. Lett. 2011, 21, 255–258. [Google Scholar] [CrossRef] [PubMed]
  140. Woodcock, C.B.; Ulashchik, E.A.; Poopeiko, N.E.; Shmanai, V.V.; Reich, N.O.; Shchepinov, M.S. Rational manipulation of DNA methylation by using isotopically reinforced cytosine. ChembioChem 2016, 17, 2018–2021. [Google Scholar] [CrossRef]
  141. Lowenson, J.D.; Shmanai, V.V.; Shklyaruck, D.; Clarke, S.G.; Shchepinov, M.S. Deuteration protects asparagine residues against racemization. Amino Acids 2016, 48, 2189–2196. [Google Scholar] [CrossRef] [Green Version]
  142. Tung, R. The development of deuterium-containing drugs. Innov. Pharm. Technol. 2010, 32, 24–28. [Google Scholar]
  143. Dzhimak, S.S.; Drobotenko, M.I.; Basov, A.A.; Svidlov, A.A.; Fedulova, L.V.; Lyasota, O.M.; Baryshev, M.G. Mathematical modeling of open states in DNA molecule depending on the deuterium concentration in the surrounding liquid media at different values of hydrogen bond disruption energy. Dokl. Biochem. Biophys. 2018, 483, 359–362. [Google Scholar] [CrossRef]
  144. Zubarev, R.A. Role of stable isotopes in life—testing isotopic resonance hypothesis. Genom. Proteom. Bioinform. 2011, 9, 15–20. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Reaction 1: Fractionation of isotopes (atomic mass n+mX > atomic mass nX) is observed at k1 ≈ k2; k3 >> k4; k3 >> k1. Note: S—(bio-)chemical substrate; P—product of (bio-)chemical reaction; E—enzyme (catalyst).
Figure 1. Reaction 1: Fractionation of isotopes (atomic mass n+mX > atomic mass nX) is observed at k1 ≈ k2; k3 >> k4; k3 >> k1. Note: S—(bio-)chemical substrate; P—product of (bio-)chemical reaction; E—enzyme (catalyst).
Molecules 24 04101 g001
Figure 2. Reaction 2: Fractionation of deuterium is observed when k1 ≈ k2; k3 >> k4; k3 >> k1. Note: S—(bio-)chemical substrate; P—product of (bio-)chemical reaction; E—enzyme (catalyst).
Figure 2. Reaction 2: Fractionation of deuterium is observed when k1 ≈ k2; k3 >> k4; k3 >> k1. Note: S—(bio-)chemical substrate; P—product of (bio-)chemical reaction; E—enzyme (catalyst).
Molecules 24 04101 g002
Figure 3. Reaction 3: Fractionation of oxygen isotope 18O is observed when k1 ≈ k2; k3 >> k4; k3 >> k1. Note: S—(bio-)chemical substrate; P—product of (bio-)chemical reaction; E—enzyme (catalyst).
Figure 3. Reaction 3: Fractionation of oxygen isotope 18O is observed when k1 ≈ k2; k3 >> k4; k3 >> k1. Note: S—(bio-)chemical substrate; P—product of (bio-)chemical reaction; E—enzyme (catalyst).
Molecules 24 04101 g003
Table 1. Isotopic resonance regularity in biological and natural systems occurring because of the neutron Basov–Dzhimak hypothesis in individual chemical bonds and some individual isotopes.
Table 1. Isotopic resonance regularity in biological and natural systems occurring because of the neutron Basov–Dzhimak hypothesis in individual chemical bonds and some individual isotopes.
IIIIIIIVVVIVIIVIIIIXX
12C-1H 12C-2H
6n − 7p−17n − 7p0
R-spin0+ & ½+ is ½+ +R-spin0+ & 1+ is 1+ +
13C-1H [84] 13C-2H [41]
7n − 7p08n − 7p1
R-spin½ & ½+ is 1− +R-spin½ & 1+ is 1½− +
14N–1H 14N-2H
7n − 8p−18n − 8p0
R-spin1+ & ½+ is 1½+ +R-spin1+ & 1+ is 2+ +
15N-1H [84] 15N-2H
8n − 8p09n − 8p1
R-spin½ & ½+ is 1− +R-spin½ & 1+ is 1½− +
16O-1H 17O-1H 18O-1H [84]
8n − 9p−19n − 9p010n − 9p1
R-spin0+ & ½+ is ½+ +R-spin5/2+ & ½+ is 3+ +R-spin0+ & ½+ is ½+ +
16O-2H 17O-2H 18O-2H
9n − 9p010n − 9p111n − 9p2
R-spin0+ & 1+ is 1+ +R-spin5/2+ & 1+ is 3½+ +R-spin0+ & 1+ is 1+ +
32S-1H 32S-2H 34S-1H 34S-2H
16n − 17p−117n − 17p018n − 17p119n − 17p2
R-spin0+ & ½+ is ½+ +R-spin0+ & 1+ is 1+ +R-spin0+ & ½+ is ½+ +R-spin0+ & 1+ is 1+ +
12C-12C 12C-13C
12n − 12p013n − 12p1
R-spin0+ & 0+ is 0+ +R-spin0+ & ½ is ½+ −
12C-14N 12C-15N
13n − 13p014n − 13p1
R-spin0+ & 1+ is 1+ +R-spin0+ & ½ is ½ + −
13C-14N 13C-15N [84,85,86]
14n − 13p115n − 13p2
R-spin½ & 1+ is 1½ + −R-spin½ & ½ is 1 − −
12C-16О 12C-17О 12C-18О
14n − 14p015n − 14p116n − 14p2
R-spin0+ & 0+ is 0+ +R-spin0+ & 5/2+ is 5/2+ +R-spin0+ & 0+ is 0+ +
13C-16О 13C-17О 13C-18О [84]
15n − 14p116n − 14p217n − 14p3
R-spin½ & 0+ is ½− +R-spin½ & 5/2+ is 3− +R-spin½ & 0+ is ½− +
14N-16O 14N-17O 14N-18O
15n − 15p016n − 15p117n − 15p2
R-spin1+ & 0+ is 1+ +R-spin1+ & 5/2+ is 3½+ +R-spin1+ & 0+ is 1+ +
15N-16O [87] 15N-17O 15N-18O [84]
16n − 15p117n − 15p218n − 15p3
R-spin½ & 0+ is ½ − +R-spin½ & 5/2+ is 3 − +R-spin½ & 0+ is ½− +
40Ca [27] 43Ca [27]
20n − 20p023n − 20p3
R-spin0+R-spin7/2
24Mg [29] 25Mg [29,88] 26Mg [29]
12n − 12p013n − 12p114n − 12p2
R-spin0+R-spin5/2+R-spin0+
63Cu-18O
44n − 37p7
R-spin3/2 & 0+ is 3/2− +
63Cu
34n − 29p5
R-spin3/2
65Cu
36n − 29p7
R-spin3/2
28Si [89] 29Si [89] 30Si [89]
14n − 14p015n − 14p116n − 14p2
R-spin0+R-spin1/2+R-spin0+
64Zn [90,91] 67Zn [91,92]
34n − 30p437n − 30p7
R-spin0+R-spin5/2
66Zn [90,91]
36n − 30p6
R-spin0+
238U [93] 235U [93]
146n − 92p54143n − 92p51
R-spin0+R-spin7/2
Note: n—neutron; p—proton; D—deuterium; R-spin—resultant spin of individual isotopes or pair of isotopes with chemical bonds (as a sum of individual spins); k—integer. In the table, all references [27,29,41,84,85,86,87,88,89,90,91,92,93] confirm the new hypothesis and arising resonance in living and natural systems only under certain isotopes. Note: I, III, and VII—pairs of isotopes with chemical bonds and individual isotopes that must have no isotopic resonance; II—there is no isotopic resonance for: Nn − Np = −1; IV—there is no isotopic resonance for: Nn − Np = 0; V and IX—pairs of isotopes with chemical bonds and individual isotopes that have some isotopic resonance; VI—there is some isotopic resonance for: Nn − Np = 1; VIII—there is no isotopic resonance for: Nn − Np = 2, and Nn − Np = 2k; X—there is some isotopic resonance for: Nn − Np = 2k + 1; N—number.

Share and Cite

MDPI and ACS Style

Basov, A.; Fedulova, L.; Vasilevskaya, E.; Dzhimak, S. Possible Mechanisms of Biological Effects Observed in Living Systems during 2H/1H Isotope Fractionation and Deuterium Interactions with Other Biogenic Isotopes. Molecules 2019, 24, 4101. https://doi.org/10.3390/molecules24224101

AMA Style

Basov A, Fedulova L, Vasilevskaya E, Dzhimak S. Possible Mechanisms of Biological Effects Observed in Living Systems during 2H/1H Isotope Fractionation and Deuterium Interactions with Other Biogenic Isotopes. Molecules. 2019; 24(22):4101. https://doi.org/10.3390/molecules24224101

Chicago/Turabian Style

Basov, Alexander, Liliya Fedulova, Ekaterina Vasilevskaya, and Stepan Dzhimak. 2019. "Possible Mechanisms of Biological Effects Observed in Living Systems during 2H/1H Isotope Fractionation and Deuterium Interactions with Other Biogenic Isotopes" Molecules 24, no. 22: 4101. https://doi.org/10.3390/molecules24224101

Article Metrics

Back to TopTop