Next Article in Journal
Inflammation Confers Healing Advantage to Corneal Epithelium Following Subsequent Injury
Next Article in Special Issue
The Role of NQO1 in Ovarian Cancer
Previous Article in Journal
Chronic Voluntary Alcohol Consumption Alters Promoter Methylation and Expression of Fgf-2 and Fgfr1
Previous Article in Special Issue
Combination Treatment of a Phytochemical and a Histone Demethylase Inhibitor—A Novel Approach towards Targeting TGFβ-Induced EMT, Invasion, and Migration in Prostate Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Thermodynamic Approach to the Metaboloepigenetics of Cancer

by
Umberto Lucia
1,*,†,
Thomas S. Deisboeck
2,
Antonio Ponzetto
3 and
Giulia Grisolia
1,4,*,†
1
Dipartimento Energia “Galileo Ferraris”, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
2
Department of Radiology, Harvard-MIT Martinos Center for Biomedical Imaging, Massachusetts General Hospital and Harvard Medical School, Charlestown, Boston, MA 02129, USA
3
Department of Medical Sciences, University of Torino, Corso A.M. Dogliotti 14, 10126 Torino, Italy
4
Dipartimento Scienza Applicata e Tecnologia, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2023, 24(4), 3337; https://doi.org/10.3390/ijms24043337
Submission received: 18 January 2023 / Revised: 3 February 2023 / Accepted: 3 February 2023 / Published: 7 February 2023

Abstract

:
We present a novel thermodynamic approach to the epigenomics of cancer metabolism. Here, any change in a cancer cell’s membrane electric potential is completely irreversible, and as such, cells must consume metabolites to reverse the potential whenever required to maintain cell activity, a process driven by ion fluxes. Moreover, the link between cell proliferation and the membrane’s electric potential is for the first time analytically proven using a thermodynamic approach, highlighting how its control is related to inflow and outflow of ions; consequently, a close interaction between environment and cell activity emerges. Lastly, we illustrate the concept by evaluating the Fe 2 + -flux in the presence of carcinogenesis-promoting mutations of the TET1/2/3 gene family.

1. Introduction

The main aim of cancer appears to be continuous cell replication as well as spatial expansion of the tumour system itself, a process involving local invasion and distant metastasis [1]. As the scientific knowledge of complex systems increases, cancer emerges as a disease of the breakdown of the natural biological order within the body, a consequence of the malfunctioning of the controls of cell replication. In this context, an epigenetic approach could be useful in aiding our understanding of the processes involved. ‘Epigenetics’ is the biological science that studies phenotypic changes that do not involve alterations in the DNA sequence. Over the last few decades, evidence supporting the importance of gene-environment interactions for the regulation of gene expression and phenotypic outcome has mounted [2]. Furthermore, metabolism has been shown to affect gene expression: a review of the interaction between some metabolites and gene expression is summarised in Ref. [2]. Consequently, the link between energy metabolism and epigenetic control of gene expression can be analysed from the perspective of ‘metaboloepigenetics’.
We note that in 1956, cancer cells were first shown to be electrically different from normal cells [3]. Then, in 1969, in relation to the cell cycle phases, Cone Jr. pointed out that hyperpolarization represents a characteristic of the start of the M phase, introducing the hypothesis of a possible link between the membrane’s electric potential and cell cycle progression [4]. In addition, in 1970, he showed that membrane hyperpolarization was able to reversibly block the synthesis of DNA and mitosis [5], and in 1971, he reported that an increase in cancer cell proliferation could be ascribed to a lower than normal membrane potential [6]. These results have repeatedly been experimentally confirmed [7,8,9,10].
The fundamental role of the cellular membrane potential and its link to metaboloepigenetics represents an interesting topic of investigation. This link can be shown through the ion fluxes across the membrane, and its effect on gene activity can be studied in relation to the use of energy carried by the ions themselves. Indeed, living cell systems can use and transform energy in different forms:
  • Mechanical energy, related to cellular movement, reorganization of intracellular structures, and changes of cell shape;
  • Electrical energy, related to electron flow due to differences in voltage;
  • Electromagnetic energy, related to thermal radiation, etc.;
  • Chemical energy, related to biochemical reactions, but also to growth as an increase of molecules and biological structures;
  • Heat transfer, as outflow due to wasted energy released into the cell microenvironment;
  • Quantum energy, related to the structure of molecules and their quantum-level interactions.
Cell metabolism consists of thousands of chemical reactions which occur in organized sequences or metabolic pathways, such that they can lead from a high to low oxidative state in anabolism or vice versa in catabolism [11]. Thus, life is a complex biological phenomenon represented by numerous chemical, physical and biological processes, with the same physical laws governing processes in both animate and inanimate matter [12].
Thermodynamics is the discipline of physical science that allows us to study and interpret the evolution of any system in relation to energy use and conversion. As such, the aim of this paper is to develop an analysis of the metaboloepigenetic relationship from a thermodynamic viewpoint, in an effort to obtain a useful tool for cancer researchers and physicians to support the interpretation of experimental results in oncology.

2. Results

The fundamental results of this paper are expressed in Equation (12) and can be summarised as follows:
  • Any change in the cell membrane’s electric potential generates entropy ( σ ); as such, this process is irreversible unless the cell consumes metabolites (to reverse the potential if needed). This proves the strict correlation between energy management and cell activity;
  • Cell proliferation ( d V / d t ) is related to the membrane’s electric potential, confirming Cone Jr.’s experimental results [4,5,6];
  • The change of the membrane’s electric potential can be controlled by the inflow and outflow of ions ( i μ i J ), which highlights the close interaction between environment and cell activity, as asserted by epigenetics;
  • Any change in the cell membrane potential is related to energy management and ion concentration, as represented by the Nernst equation.
Applying these thermodynamics results to cancer, we note the fundamental role played by DNA (Deoxyribonucleic acid) methylation. As an example, we discuss mutations of the TET1/2/3 gene family which have been well characterized in driving hematological carcinogenesis [13]; recently, deregulated TET1/2/3 functions via multimodal mechanisms have also been shown in solid tumors. Indeed, disrupted TET1/2/3 catalytic activity determines the reconstitution of the methylation landscape, which plays a fundamental role in carcinogenesis. Ten-eleven translocation (TET) enzymes are a family of dioxygenases that iteratively catalyse 5-methylcytosine (5mC) oxidation and promote cytosine demethylation, thereby creating a dynamic methylation landscape [2], with the promotion of a cancer phenotype. The TET enzymes use α -ketoglutarate ( α -KG) as cosubstrates to bind (Fe 2 + ) in order to activate molecular oxygen; in particular, Fe 2 + may work as a catalyst for the generation of ROS (reactive oxygen species) in pathological conditions such as carcinogenesis, inflammation, radiation and reperfusion injury. Indeed, Fe 2 + overload has been associated with carcinogenesis (with major target genes, p16(INK4A) and p15(INK4B) tumor suppressor genes, which encode cyclin-dependent kinase inhibitors) in a ferric nitrilotriacetate-induced rat renal carcinogenesis model, in which the Fenton reaction was induced in the renal proximal tubules [14].
Thus, Fe 2 + plays a fundamental role and our thermodynamic approach allows us to evaluate the iron fluxes through the cell membrane. To do so, we consider that the heat transfer from the cell to its environment occurs in a convective way, and as a consequence of (6), it follows that:
· J Q = α d A d V ( T T 0 )
where α 0.023 R e 0.8 P r 0.35 λ / R is the coefficient of convection, with λ 0.6 W m 1 K 1 conductivity, R e 0.2 the Reynolds number and P r 0.7 being the Prandtl number [15], A stands for the area of the cell membrane, V represents the cell volume.
Considering Equation (7), we can introduce the following simplifications:
  • We consider the ideal case ( σ = 0 W m 3 K 1 ): it allows us to evaluate the maximum value of the ion fluxes;
  • We consider the stationary state ( d s / d t = 0 ).
Consequently, it follows that:
J Fe + = · α μ Fe + · R ( T T 0 ) = = 0.004 × 10 6 × 0.023 × 0 . 2 0.8 × 0 . 7 0.35 × 0.6 1556.5 × 10 3 · R 2 ( T T 0 ) = = 3.45 × 10 18 [ mol s 1 ] R 2 [ m 2 ] = 1.93 × 10 19 [ kg s 1 ] R 2 [ m 2 ]
where 0.004 μ m is the depth of the cell membrane [16] and R denotes the mean radius of the cell, considered, in the first approximation, as a sphere, μ Fe 2 + = 1556.5 kJ mol 1 , and T T 0 0.4 °C [17]. The numerical result depends on the mean size of the cell. It follows that the (max) Fe 2 + -flux is of the order of 3.45 × 10 18 mol s 1 m 2 = 1.93 × 10 19 kg s 1 m 2 , beyond which the cell would suffer damage.
This value refers to a single cell. To understand whether it is meaningful and in agreement with experimental results, we consider that in a human body there are approximately 30 × 10 18 cells and also a comparable number of bacteria [16], with a radius (of mean value) of about 10 5 m. Consequently, considering a cell as a sphere, the total amount of iron mass can be evaluated per day (86,400 s) and yields a result of 20.6 mg Fe 2 + kg h b 1 , where h b means ‘for an entire human body’. This numerical result must be compared with well-known results from etiology: ingestion of less than 20 mg kg 1 of elemental iron is nontoxic, while ingestion of 20 mg kg 1 to 60 mg kg 1 results in moderate symptoms, and ingestion of more than 60 mg kg 1 can lead to significant toxicity with severe morbidity and mortality [18,19].

3. Discussion

In 1942, Conrad Waddington introduced the concept of epigenetics to describe the developmental processes between genotype and phenotype [20]. As for cancer, the packaging of the genome and its regulation were thought to be the fundamental processes involved in the preservation of cellular health [21]. We add here a novel thermodynamic viewpoint to cancer metaboloepigenetics. Our results confirm the close interaction between cell activity and cellular microenvironment, and we analytically prove the link between cell proliferation and the membrane’s electric potential, controlled by the inflow and outflow of ions.
TET genes are frequently mutated in various cancers, and they play an important role in tumorigenesis due to their role in regulating DNA methylation and transcription [22]; indeed, cancer initiation and progression are related to modifications of DNA methylation [23]. TET1,2,3 oxidize the 5-methyl group of 5-methylcytosine (5mC) in DNA by involving oxygen and 2-oxoglutarate as substrates, and Fe 2 + as a cofactor to yield 5-hydroxymethylcytosine (5-hmC), CO 2 , and succinate [24,25]. In our paper, the role of fluxes has been highlighted, and specifically, Fe 2 + was evaluated using a nonequilibrium thermodynamic approach. Indeed, Fe 2 + is a pro-oxidant agent that can produce ROS by reacting with hydrogen peroxide (H 2 O 2 ). If saturation of the antioxidant system occurs then an excess of ROS determines lipid peroxidation, amino acid oxidation, loss of protein structure, and DNA damage, with related tissue damages. Moreover, lipid peroxidation can change the plasma membrane’s composition, fluidity and permeability, modifying the activity of integral proteins, particularly Na + ,K + -ATPase and the Ca 2 + -ATPase [26], with an increase in Ca 2 + intracellular concentration. In this context, the H + -ATPase plays a fundamental role, due to its function of inflowing positive charges into the cell. Recently, in relation to iron metabolism within the tumor microenvironment, the iron promotion of the production of reactive oxygen species was emphasized. This highlights the function of triggering ferroptosis (iron-dependent cell death) or supporting malignant transformation [27].
Our result, as obtained in Equation (12), represents the quantitative evaluation of the Fe fluxes required by cells to maintain the best thermodynamic conditions for supporting cellular viability. Theoretically, this value informs physicians to supplement current anticancer therapies with a defined amount of iron which sustains TET activities, in an effort to counter the impact of tumorigenesis-promoting TET mutations [28].
We note that, while we have focused our analysis on Fe 2 + fluxes to showcase the utility of thermodynamic concepts in elucidating metaboloepigenetics, in reality, several other elements cross the cell membrane and should be of critical importance in maintaining homeostasis, each one in theory necessitating its own, specific Equation (2) description. However, while a combined analysis is beyond the scope of this paper, the fact that our numerical evaluations of Fe only come within ∼4.5% error when compared with experimental data (that implicitly include these other elements) gives credence to our stepwise approach.
In summary, as exemplified by iron and the TET loop, our approach presented here can aid in understanding the global patterns of epigenetic modifications in cancer.

4. Materials and Methods

As a consequence of the cancer system’s properties of heterogeneous clonal expansion, replicative immortality, patterns of longevity, rewired metabolic pathways, altered reactive oxygen species, evasion of death signals, and metastatic invasion [1,29,30], cancer can be modelled as an adaptive system based on natural selection that renders any single cancer cell independent of its neighbours [1]. Therefore, a thermodynamic approach emerges for modelling cells as open systems with the ability to convert metabolic energy into mechanical and chemical useful works, while heat discharges into their microenvironments. From a thermodynamic standpoint, metabolic energy represents the energetic inflow of a thermodynamic ‘cellular’ engine, with the following differences for the case of normal versus cancer cells, respectively [31]:
  • For normal cells: the Krebs cycle, which consists of chemical reactions to convert stored energy through the oxidation of acetyl-CoA, using carbohydrates, lipids, and proteins;
  • For cancer cells: the Warburg cycle, which consists of chemical reactions for specialised fermentation over the aerobic respiration pathway [32,33,34,35].
As referenced above, in 1971, Con Jr. showed that an increase in proliferation of cancer cells is caused by their lowered membrane potential, compared with the reference potential of non-cancerous, normal cells [6]. Consequently, the cytoplasmatic pH and the extracellular environment present a link to the cells’ membrane potential [36]. The electric potential difference, Δ ϕ , between the cytoplasm and the extracellular environment is evaluated with respect to the environment [37] by using the Goldman–Hodgkin–Katz equation [38,39,40]:
Δ ϕ = R T F ln P Na + [ Na + ] outside + P K + [ K + ] outside + P Cl [ Cl ] outside P Na + [ Na + ] inside + P K + [ K + ] inside + P Cl [ Cl ] inside
where [A] represents the concentration of the ion A, R = 8.314 J mol 1 K 1 is the universal constant of ideal gasses, T depicts the absolute temperature, F = 96,485 C mol 1 is the Faraday constant, and P is the relative permeability [41,42,43], such that P Na + = 0.04 , P K + = 1 and P Cl = 0.45 [41,42,43].
Cells can alter their membranes’ electric potential by changing the concentration of the different ions, i.e., by modulating ion inflow and outflow [44,45]; with particular regard to cancer cells, an increase in the Na + intracellular concentration, with a K + constant intracellular concentration [46], has been shown to lead to depolarization during malignant transformation of cells [7,47,48]. This experimental evidence demonstrates the fundamental role of the cell membrane’s electric potential for the control of malignant behaviour such as cell de-differentiation, proliferation, and migration [49,50,51].
Here, we propose a thermodynamic analysis of these processes to highlight the epigenetic and metaboloepigenetic conditions to support the treatment of cancer by conditioning ion fluxes. To do so, the Onsager phenomenological equations [52,53,54,55,56] must be considered:
J e = L 11 ϕ T L 12 T T 2 J Q = L 21 ϕ T L 22 T T 2
where J e depicts the current density [A m 2 ], J Q stands for the heat flux [W m 2 ], T is the living cell temperature, and L i j are the phenomenological coefficients, such that L 12 = L 21 in the absence of magnetic fields, and L 11 0 and L 22 0 , and L 11 L 22 L 12 2 > 0 [52,53,54,55,56,57,58]: L 11 and L 22 represent the heat conductivity and the electrical conductivity, respectively, while L 12 and L 21 are cross coefficients, independent of both L 11 and L 22 [57,58].
Now, we consider that:
  • The fluxes of an ion cause a variation in the concentration of the ion itself, balanced by the following equation [52,53]
    d c i d t = · J i
    where c i denotes the concentration of the i-th ion (Na + , K + , Ca 2 + , Cl , etc.), t is the time, and J i stands for the current density of the i-th ion;
  • The heat flux can be evaluated by the First Law of Thermodynamics as follows [52,53]:
    d u d t = · J Q
    where u represents the specific internal energy;
  • The Second Law of Thermodynamics results in [59,60]
    T d s d t = · J Q i = 1 N μ i J i i = 1 N J i · μ i
    where s represents the specific entropy, T is the temperature, J S = J Q i = 1 N μ i J i denotes the contribution of the inflows and outflows, and T σ = i = 1 N J i · μ i is the dissipation function [52], with μ being the chemical potential, defined as:
    μ i = G n i T , p , n k i G n i = g
    where G is the Gibbs energy, g represents the Gibbs molar specific energy, n is the number of moles, and p stands for the pressure. The entropy outflow σ is fundamental to generate order from disorder, as Schrödinger himself pointed out [61].
From Equation (5), we can state that an ion flux implies a variation in ion concentration on both sides of the membrane, with a related variation of the pH. Now, considering the Nernst equation [40]:
F d ϕ = d g + 2.3 R   T 0   d pH
where F = 96,485 A s mol 1 is the Faraday constant and R = 8.314 J mol 1 K 1 is the universal constant of ideal gas, it follows that a cell can control its membrane electric potential by managing ion flow through the membrane itself; but the change in the electric membrane potential is also related to the Gibbs free energy, which itself is related to the internal energy and entropy by its definition:
G = U + p V T S
Then, from Equation (6), it follows that [62]
V d u d t   d V = V ρ   c   d T d t d V = V · J Q d V = Q ˙
where ρ 10 3 kg m 3 is the cell density, c 4186 J kg 1 K 1 is the specific heat of the cell.
Finally, considering these results together with Equation (7), it follows that:
F d ϕ d t = p n d V d t · i μ i J i + T σ
which means that a change in the cell’s electric potential determines:
  • The volume of the cell ( d V / d t ), which is related to cell proliferation;
  • The fluxes of ions ( μ i J i ), which are related to the metabolic requirements, respiration, communication, molecule formation and epigenetic effects;
  • The heat discharged towards the environment ( T σ ), due to irreversibility.

Author Contributions

Conceptualization, U.L. and T.S.D.; methodology, U.L. and G.G.; software, G.G.; validation, U.L., T.S.D. and A.P.; formal analysis, U.L.; investigation, G.G.; resources, U.L.; data curation, G.G.; writing—original draft preparation, U.L., T.S.D. and G.G.; writing—review and editing, U.L., T.S.D. and G.G.; visualization, G.G.; supervision, U.L.; project administration, U.L.; funding acquisition, U.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data used have been here evaluated and the reference to literature is quoted in the References.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hanahan, D.; Weinberg, R.A. Hallmarks of Cancer: The Next Generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef]
  2. Donohoe, D.R.; Bultman, S.J. Metaboloepigenertics: Interrelationships between energy metabolism and epigenetic control of gene expression. J. Cell Physiol. 2012, 227, 3169–3177. [Google Scholar] [CrossRef]
  3. Ambrose, E.J.; James, A.M.; Lowick, J.H. Differences between the electrical charge carried by normal and homologous tumour cells. Nature 1956, 177, 576–577. [Google Scholar] [CrossRef]
  4. Cone, C.D. Electroosmotic interactions accompanying mitosis initiation in sarcoma cells in vitro. Trans. N. Y. Acad. Sci. 1969, 31, 404–427. [Google Scholar] [CrossRef] [PubMed]
  5. Cone, C.D. Variation of the transmembrane potential level as a basic mechanism of mitosis control. Oncology 1970, 24, 438–470. [Google Scholar] [CrossRef] [PubMed]
  6. Cone, C.D. Unified theory on the basic mechanism of normal mitotic control and oncogenesis. J. Theor. Biol. 1971, 30, 151–181. [Google Scholar] [CrossRef]
  7. Tokuoka, S.; Marioka, H. The membrane potential of the human cancer and related cells (I). Gann 1957, 48, 353–354. [Google Scholar] [PubMed]
  8. Altman, P.L.; Katz, D. Biological Handbook Vol. 1: Cell Biology; Federation of American Society for Experimental Biology: Bethesda, MD, USA, 1976. [Google Scholar]
  9. Balitsky, K.P.; Shuba, E.P. Resting potential of malignant tumour cells. Acta-Unio Int. Contra Cancrum 1964, 20, 1391–1393. [Google Scholar]
  10. Jamakosmanovic, A.; Loewenstein, W. Intracellular communication and tissue growth. III. Thyroid cancer. J. Cell Biol. 1968, 38, 556–561. [Google Scholar] [CrossRef]
  11. Doelle, H.W.; Rokem, S.; Berovic, M. (Eds.) Biotechnology. Fundamentals in Biotechnology; Eolss Publisher: Oxford, UK, 2009; Volume I. [Google Scholar]
  12. Popovic, M. Thermodynamic properties of microorganisms: Determination and analysis of enthalpy, entropy, and Gibbs free energy of biomass, cells and colonies of 32 microorganism species. Heliyon 2019, 5, e01950. [Google Scholar] [CrossRef]
  13. Bray, J.K.; Dawlaty, M.M.; Verma, A.; Maitra, A. Roles and Regulations of TET Enzymes in Solid Tumors. Trends Cancer 2021, 7, 635–646. [Google Scholar] [CrossRef] [PubMed]
  14. Toyokuni, S. Iron and carcinogenesis: From Fenton reaction to target genes. Redox Rep. 2002, 7, 189–197. [Google Scholar] [CrossRef]
  15. Lucia, U.; Grisolia, G. Resonance in Thermal Fluxes Through Cancer Membrane. Atti Accad. Peloritana Pericolanti 2020, 98, SC1–SC6. [Google Scholar] [CrossRef]
  16. Milo, R.; Phillips, R. Cell Biology by the Numbers; Garland Science: New York, NY, USA, 2015. [Google Scholar]
  17. Mercer, W.B. Technical Manuscript 640—The Living Cell as an Open Thermodynamic System: Bacteria and Irreversible Thermodynamica; Department of the U.S. Army-Fort Detrick: Frederic, MD, USA, 1971.
  18. Yuen, H.W.; Becker, W. Iron Toxicity; StatPearls Publishing: Treasure Island, FL, USA, 2022. [Google Scholar]
  19. He, W.L.; Feng, Y.; Li, X.L.; Wei, Y.Y.; Yang, X.E. Availability and toxicity of Fe(II) and Fe(III) in Caco-2 cells. J. Zhejiang Univ. Sci. B 2008, 9, 707–712. [Google Scholar] [CrossRef]
  20. Deichmann, U. Epigenetics: The origins and evolution of a fashionable topic. Dev. Biol. 2016, 416, 249–254. [Google Scholar] [CrossRef] [PubMed]
  21. Sharma, S.; Kelly, T.K.; Jones, P.A. Epigenetics in cancer. Carcinogenesis 2010, 31, 27–36. [Google Scholar] [CrossRef]
  22. Rasmussen, K.D.; Helin, K. Role of TET enzymes in DNA methylation, development, and cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef] [PubMed]
  23. Feinberg, A.; Vogelstein, B. Hypomethylation distinguishes genes of some human cancers from their normal counterparts. Nature 1983, 301, 89–92. [Google Scholar] [CrossRef]
  24. Tahiliani, M.; Koh, K.P.; Shen, Y.; Pastor, W.A.; Bandukwala, H.; Brudno, Y.; Agarwal, S.; Iyer, L.M.; Liu, D.R.; Aravind, L.; et al. Conversion of 5-methylcytosine to 5-hydroxymethylcytosine in mammalian DNA by MLL partner TET1. Science 2009, 324, 930–935. [Google Scholar] [CrossRef]
  25. Ito, T.; Ando, H.; Suzuki, T.; Ogura, T.; Hotta, K.; Imamura, Y.; Yamaguchi, Y.; Handa, H. Identification of a primary target of thalidomide teratogenicity. Science 2010, 327, 1345–1350. [Google Scholar] [CrossRef] [PubMed]
  26. Sousa, L.; Oliveira, M.M.; Pessô, M.T.C.; Barbosa, L.A. Iron overload: Effects on cellular biochemistry. Clin. Chim. Acta 2020, 504, 180–189. [Google Scholar] [CrossRef]
  27. Sacco, A.; Battaglia, A.M.; Botta, C.; Aversa, I.; Mancuso, S.; Costanzo, F.; Biamonte, F. IronMetabolism in the TumorMicroenvironment—Implications for Anti-Cancer Immune Response. Cells 2021, 10, 303. [Google Scholar] [CrossRef] [PubMed]
  28. Jiang, S. Tet2 at the interface between cancer and immunity. Commun. Biol. 2020, 3, 667. [Google Scholar] [CrossRef] [PubMed]
  29. Uthamacumaran, A. A Review of Complex Systems Approaches to Cancer Networks. Complex Syst. 2020, 20, 779–835. [Google Scholar] [CrossRef]
  30. Schwab, E.D.; Pienta, K.J. Cancer as a complex adaptive system. Med. Hypothesis 1996, 47, 235–241. [Google Scholar] [CrossRef] [PubMed]
  31. Voet, D.; Voet, J.G. Biochemistry, 3rd ed.; John Wiley & Sons: New York, NY, USA, 2004. [Google Scholar]
  32. Warburg, O.; Wind, F.; Negelein, E. The metabolism of tumors in the body. J. Gen. Physiol. 1927, 8, 519. [Google Scholar] [CrossRef]
  33. Weinhouse, S. On respiratory impairment in cancer cells. Science 1956, 124, 267–269. [Google Scholar] [CrossRef]
  34. Warburg, O. On respiratory impairment in cancer cells. Science 1956, 124, 269–270. [Google Scholar] [CrossRef] [PubMed]
  35. Burk, D.; Schade, A.L. On respiratory impairment in cancer cells. Science 1956, 124, 270–272. [Google Scholar] [CrossRef]
  36. Binggeli, R.; Cameron, I.L. Cellular potentials of normal and cancerous fibroblasts and hepatocytes. Cancer Res. 1980, 40, 1830. [Google Scholar]
  37. Yang, M.; Brackenbury, W.J. Membrane potential and cancer progression. Front. Physiol. 2013, 4, 185. [Google Scholar] [CrossRef]
  38. Goldman, D.E. Potential impedance, and rectification in membranes. J. Gen. Physiol. 1943, 27, 37–60. [Google Scholar] [CrossRef] [Green Version]
  39. Hodgkin, A.L.; Katz, B. The effect of sodium ions on the electrical activity of giant axon of the squid. J. Physiol. 1949, 108, 37–77. [Google Scholar] [CrossRef] [PubMed]
  40. Grabe, M.; Wang, H.; Oster, G. The mechanochemistry of V-ATPase proton pumps. Biophs. J. 2000, 78, 2798–2813. [Google Scholar] [CrossRef] [PubMed]
  41. Junge, D. Nerve and Muscle Excitation, 2nd ed.; Sunderland: New York, NY, USA, 1981. [Google Scholar]
  42. Adams, D.J.; Dwyer, T.M.; Hille, B. The permeability of endplate channels to monovalent and divalent metal cations. J. Gen. Physiol. 1980, 75, 493–510. [Google Scholar] [CrossRef] [PubMed]
  43. Wright, S.H. Generation of resting membrane potential. Adv. Physiol. Educ. 2004, 28, 139–142. [Google Scholar] [CrossRef]
  44. Lucia, U.; Grisolia, G.; Astori, M.R. Constructal law analysis of Cl transport in eyes aqueous humor. Sci. Rep. 2017, 7, 6856. [Google Scholar] [CrossRef] [PubMed]
  45. Lucia, U. Thermodynamic approach to nano-properties of cell membrane. Physica A 2014, 407, 185–191. [Google Scholar] [CrossRef]
  46. Cameron, I.L.; Smith, N.K.; Pool, T.B.; Sparks, R.L. Intracellular concentration of sodium and other elements as related to mitogenesis and oncogenesis in vivo. Cancer Res. 1980, 40, 1493–1500. [Google Scholar]
  47. Johnstone, R.M. Microelectrode penetration of ascites tumour cells. Nature 1959, 183, 411. [Google Scholar] [CrossRef]
  48. Marino, A.A.; Morris, D.M.; Schwalke, M.A.; Iliev, I.G.; Rogers, S. Electrical potential measurements in huma breast cancer and benign lesions. Tumour Biol. 1994, 15, 147–152. [Google Scholar] [CrossRef]
  49. Sundelacruz, S.; Levin, M.; Kaplan, D.L. Role of the membrane potential in the regulation of cell proliferation and differentiation. Stemm Cell Rev. 2009, 5, 231–246. [Google Scholar] [CrossRef]
  50. Lobikin, M.; Chernet, B.; Lobo, D.; Levin, M. Resting potential, oncogene-induced tumorigenesis, and metastasis: The bioelectric basis of cancer in vivo. Phys. Biol. 2012, 9, 065002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Schwab, A.; Fabian, A.; Hanley, P.J.; Stock, C. Role of the ion channels and transporters in cell migration. Physiol. Rev. 2012, 92, 1865–1913. [Google Scholar] [CrossRef] [PubMed]
  52. Yourgrau, W.; van der Merwe, A.; Raw, G. Treatise on Irreversible and Statistical Thermophysics; Dover: New York, NY, USA, 1982. [Google Scholar]
  53. Callen, H.B. Thermodynamics; Wiley: New York, NY, USA, 1960. [Google Scholar]
  54. Katchalsky, A.; Currant, P.F. Nonequilibrium Thermodynamics in Biophysics; Harvard University Press: Boston, MA, USA, 1965. [Google Scholar]
  55. Demirel, Y.; Sandler, S.I. Nonequilibrium Thermodynamics in Engineering and Science. J. Phys. Chem. B 2004, 108, 31–43. [Google Scholar] [CrossRef]
  56. Lucia, U.; Grisolia, G. How Life Works—A Continuous Seebeck-Peltier Transition in Cell Membrane? Entropy 2020, 22, 960. [Google Scholar] [CrossRef]
  57. Onsager, L. Reciprocal relations in irreversible processes: I. Phys. Rev. 1931, 37, 405–426. [Google Scholar] [CrossRef]
  58. Onsager, L. Reciprocal relations in irreversible processes: II. Phys. Rev. 1931, 38, 2265–2279. [Google Scholar] [CrossRef]
  59. Lucia, U.; Grisolia, G. Second law efficiency for living cells. Front. Biosci. 2017, 9, 270–275. [Google Scholar] [CrossRef]
  60. Lucia, U. Entropy generation: From outside to inside! Chem. Phys. Lett. 2013, 583, 209–212. [Google Scholar] [CrossRef]
  61. Schrödinger, E. What’s Life? The Physical Aspect of the Living Cell; Cambridge University Press: Cambridge, UK, 1944. [Google Scholar]
  62. Lucia, U.; Grisolia, G. Thermal Resonance and Cell Behavior. Entropy 2020, 22, 774. [Google Scholar] [CrossRef] [PubMed]
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

Lucia, U.; Deisboeck, T.S.; Ponzetto, A.; Grisolia, G. A Thermodynamic Approach to the Metaboloepigenetics of Cancer. Int. J. Mol. Sci. 2023, 24, 3337. https://doi.org/10.3390/ijms24043337

AMA Style

Lucia U, Deisboeck TS, Ponzetto A, Grisolia G. A Thermodynamic Approach to the Metaboloepigenetics of Cancer. International Journal of Molecular Sciences. 2023; 24(4):3337. https://doi.org/10.3390/ijms24043337

Chicago/Turabian Style

Lucia, Umberto, Thomas S. Deisboeck, Antonio Ponzetto, and Giulia Grisolia. 2023. "A Thermodynamic Approach to the Metaboloepigenetics of Cancer" International Journal of Molecular Sciences 24, no. 4: 3337. https://doi.org/10.3390/ijms24043337

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