Next Article in Journal
VB-Net: Voxel-Based Broad Learning Network for 3D Object Classification
Previous Article in Journal
Chlamydomonas reinhardtii Is a Potential Food Supplement with the Capacity to Outperform Chlorella and Spirulina
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Non-Equilibrium Thermodynamic Approach to Ca2+-Fluxes in Cancer

Dipartimento Energia “Galileo Ferraris”, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Appl. Sci. 2020, 10(19), 6737; https://doi.org/10.3390/app10196737
Submission received: 14 August 2020 / Revised: 18 September 2020 / Accepted: 23 September 2020 / Published: 26 September 2020
(This article belongs to the Section Applied Biosciences and Bioengineering)

Abstract

:
Living systems waste heat in their environment. This is the measurable effect of the irreversibility of the biophysical and biochemical processes fundamental to their life. Non-equilibrium thermodynamics allows us to analyse the ion fluxes through the cell membrane, and to relate them to the membrane electric potential, in order to link this to the biochemical and biophysical behaviour of the living cells. This is particularly interesting in relation to cancer, because it could represent a new viewpoint, in order to develop new possible anticancer therapies, based on the thermoelectric behaviour of cancer itself. Here, we use a new approach, recently introduced in thermodynamics, in order to develop the analysis of the ion fluxes, and to point out consequences related to the membrane electric potential, from a thermodynamic viewpoint. We show how any increase in the cell temperature could generate a decrease in the membrane electric potential, with a direct relation between cancer and inflammation. Moreover, a thermal threshold, for the cell membrane electric potential gradient, has been obtained, and related to the mitotic activity. Finally, we obtained the external surface growth of the cancer results related (i) to the Ca2+-fluxes, (ii) to the temperature difference between the the system and its environment, and (iii) to the chemical potential of the ion species.

1. Introduction

At present, cancer is still an problem in biophysics, medicine and pharmacology. Indeed, statistical evaluations show a continuous growth in those dead due to cancer [1,2].
In recent years, the analysis of the ion transport phenomenon in cancer has also been substantially developed [3,4,5,6,7,8,9,10,11] in relation to the consequences for the cells’ membrane potential. These experimental and theoretical results have pointed to the regulatory role of ion channels and transporters, in relation to the cell cycle phases, with relevance for neoplastic progression, resistance to apoptosis, and metastasis [12].
Indeed, since 1944, in hyperplastic mouse epidermis, the reduction in Ca2+ levels has been shown to be an important aspect of precancerous conditions [13,14]: this feature has represented a first direct correlation between Ca2+ and cancer. Today, the study of Ca2+ dynamics represents a fundamental aspect of the research on carcinogenesis and tumour evolution.
The development in the comprehension of intracellular Ca2+ signalling pathways has allowed biologists and the physicians to identify some important molecular players, with a consequent study of the activity of different cancer-related proteins, with altered functions [14]. Indeed, calcium is a fundamental second messenger, involved in a variety of cellular processes, such as proliferation and apoptosis.
Considering the experimental evidence [14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31], tumour progression has also been related to the accumulation of some alterations in the Ca2+ signal, which inhibits its cytotoxic activity [32].
Calcium signal modulation can change cells’ sensitivity to signals [15,32]. High levels of mitochondrial Ca2+ concentration for a long time, have been shown to induce the mitochondrial permeability transition pore, a pathological and physiological phenomenon, discovered over 40 years ago, and still not completely understood [15,33]. The mitochondrial permeability transition pore causes the formation of a non-specific channel within the inner mitochondrial membrane, useful for Ca2+ release and metabolite exchange, between the mitochondrial matrix and cytosol. However, a prolonged mitochondrial permeability transition pore causes changes in the inner mitochondrial membrane potential, cessation of ATP synthesis, bioenergetic crisis, and apoptotic or necrotic cell death [14,16,33].
All this experimental evidence moves the research interest towards the analysis of the Ca2+-fluxes. At present, most of the mechanisms related to intracellular Ca2+ responses have been understood by developing in vitro experiments, but comprehension of the physiological role of these processes, in relation to tumour environment, remains an problem [32].
In this paper, we wish to develop a new viewpoint in the analysis of ion fluxes, recently published in [34], based on thermodynamics, with particular regards to the non-equilibrium thermodynamics. Our aim is to suggest an approach which takes into account the ion fluxes, in relation to the membrane electric potential gradient, in order to analytically describe the link between ion fluxes and membrane potential, in relation to cancer behaviour. We will focus our analysis on Ca2+ fluxes, due to the fundamental role of this ion in the regulation of a great number of cell functions.

2. Materials and Methods

The living cell membrane is characterized by a different permeability in relation to the distinct ions (Na+, K+, Cl, Ca2+, etc.) which cause an electric potential difference, Δ ϕ , between the cytoplasm and the extracellular environment, measured in reference to the environment [35,36].
Since 1956, it has been clear that cancer cells are electrically different from normal ones [37]. Cone Jr. pointed out that hyperpolarization:
  • Characterises the start of the cell M phase [38];
  • Can reversibly block the synthesis of DNA and the mitosis [39];
  • Was found to be a characteristic of the normal cells: the lowered membrane potential was identified as a cause of an increase in proliferation of the cancer cells [40].
Consequently, in 1971, Cone Jr. conjectured a relation between the cell cycle progression and the membrane electric potential changes [40]: this hypothesis has always been experimentally confirmed [41,42,43,44,45].
Moreover, the fundamental role of the membrane electric potential has recently been highlighted in relation to the control of the critical cell functions (proliferation, migration, and differentiation) [46,47,48]. In this context, the role of the ion fluxes has also been highlighted; indeed, an increase in the Na+ intracellular concentration in tumour causes a depolarisation, during malignant transformation of normal cells [41,49,50]. On the other hand, the K+ intracellular concentration remains approximately constant [51].
The membrane electric potential can be theoretically described by the Goldman–Hodgkin–Katz equation [52,53,54]
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] is the concentration of the ion A, R = 8.314 J mol−1K−1 is the universal constant of ideal gasses, T is the absolute temperature, F = 96 , 485 C mol−1 is the Faraday constant, and P is the relative permeability [55,56,57], such that P Na + = 0.04 , P K + = 1 and P Cl = 0.45 [55,56,57].
In order to develop a non-equilibrium thermodynamic analysis of the cell membrane, we must consider the interrelationship between the fluxes through the cell membrane (heat and ion fluxes) and the potentials at the borders of the membrane itself (temperature and electric potential). To do so, we follow the Onsager approach, by introducing the phenomenological equations [34,58,59,60,61]
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 is the current density [A m−2], J Q is 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 [34,58,59,60,61,62,63]. The phenomenological coefficients in the Equations (2) are constant over the range where the linear laws hold, and they must be determined experimentally [60,64]: L 11 is named the heat conductivity, L 22 is commonly called the electrical conductivity, while L 12 and L 21 are named the cross coefficients. Moreover, the cross coefficients are independent of both L 11 and L 22 [62,63].
When ion fluxes occur J e 0 , it follows that [58,59]
d c i d t = · J i
where c i is the concentration of the i-th ion (Na+, K+, Ca2+, Cl, etc.), t is the time, and J i is the current density of the i-th ion. In this condition, considering the Equation (2), it follows that [34,58,59]
d ϕ d T = L 21 L 11 1 T
which highlights that a Peltier-like effect occurs [58], and a related heat flux is also generated [58,59]
d u d t = · J u
where u is the specific internal energy. Living cells exchange heat power towards their environment by convection, therefore, following the First Law of Thermodynamics, we can write [65]
d u d t d V = ρ c d T d t d V = δ Q ˙ = α ( T T 0 ) d A · J u = α d A d V ( T T 0 ) = β ( T T 0 )
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, α 0.023 R e 0.8 P r 0.35 λ / R is the coefficient of convection, with λ 0.6 W m−1K−1 conductivity, R e 0.2 the Reynolds number and P r 0.7 the Prandtl number [66], A area of the cell membrane, V is the cell volume, and β = α d A / d V is constant. Therefore, considering Equation (2), we can obtain [34]
d ϕ d = α L 22 L 11 L 12 L 12 T ( T T 0 ) = α k T ( T T 0 )
which links the membrane electric potential to the temperature of the cell, with being the length of the cell membrane. Moreover, considering the Schrödinger approach to living systems [67], we can point out that life always requires T T 0 > 0 , and, consequently, d ϕ / d r < 0 . This last inequality explains hyperpolarization in cells [41,42,43,44].
The model obtained allows us to describe life as [34,67]
  • A continuous metabolic generation, characterised by ion and metabolite fluxes, for which a Peltier-like effect occurs, and d ϕ / d T = L 21 / L 11 T
  • A continuous heat exchange, towards the environment, for which a Seebeck-like effect occurs, and d ϕ / d = α T ( T T 0 ) / k
Consequently, a specific entropy rate is generated [68]
T d s d t = · J u i = 1 N μ i J i i = 1 N J i · μ i
where s is the specific entropy, T is the temperature, J S = J u i = 1 N μ i J i is the contribution of the inflows and outflows, and T σ = i = 1 N J i · μ i is the dissipation function [58], and μ is the chemical potential, defined as
μ i = G n i T , p , n k i
where G is the Gibbs energy, n is the number of moles, and p is the pressure. The entropy outflow σ is fundamental to generate order from disorder, as Schrödinger himself pointed out [67].
In relation to Ca2+ fluxes, we rewrite Equation (8) as follows
T d s d t = · J u μ Ca J Ca
which, considering T constant, and following Prigogine ( d s / d t = 0 ) [69], becomes
· J u μ Ca J Ca = 0
Now, considering that · J u = β ( T T 0 ) , we can write
β ( T T 0 ) + · ( μ Ca J Ca ) = 0 · ( μ Ca J Ca ) = β ( T T 0 ) = δ Q ˙ d V

3. Results

In this paper, we have developed a non-equilibrium thermodynamic analysis of the cell membrane electric potential, in order to obtain an analytical model for the comprehension of the role of the ion fluxes in relation to cancer behaviour, with particular interest in Ca2+ fluxes.
Some general statements can be introduced; indeed, Equation (7) points out that:
  • Any increase in cell temperature generates a decrease in the membrane electric potential; in the case of cancer, it is caused by inflammation;
  • The possible existence of this, due to the thermal threshold ( T > T 0 ) for the cell membrane electric potential gradient, is related to the mitotic activity [40].
In relation to Ca2+, these results link the external surface growth of the cancer to Ca2+ fluxes, to the temperature difference between the internal of the system and its environment, and to the chemical potential of the ion species. Indeed, Ca2+ outflow is a flux against the gradient, so, it is negative, and the heat exchange decreases. In this case, the cell must use the energy stored in other ways (proteins formation, etc.). If the Ca2+ inflows into the cell, the sign changes, and the cell can outflow heat, decreasing its energy value. Consequently, Ca2+ inflow should allow the cell to prevent cancer development, because the cell can decrease the chemicals that are useful for proliferation, in accordance with the experimental evidence [13,15,70,71,72].

4. Discussion and Conclusions

Hyperpolarization determines the activation of the Ca2+-K+ channel, which increases the Ca2+ intracellular concentration [45]. Consequently, the Ca2+-K+ channel results in a fundamental controller of the membrane electric potential.
Proteins play a fundamental role in ion transport. Proteins in the cytosolic can be modified in their functions by phosphorylation or dephosphorylation. In this context, the H+-ATPase plays a fundamental role, because, it generates inflows of positive charges into the cell [73,74,75,76]. Consequently, protein phosphorylation results in an important cellular regulatory mechanism, because many enzymes and receptors [77,78] are activated or deactivated by phosphorylation, by involving kinase and phosphatase.
Cancer and normal cells have different metabolic pathways; indeed, cancer cells must increase their metabolism in order to support their growth [79]. Consequently, we can consider that glycolysis is the cytoplasmic catabolism of glucose (C6H12O6) and it finishes with the inflow of pyruvate (CH3COCO2) into the Krebs cycle and the mitochondrion in the presence of oxygen, but, when oxygen supply is scarce, the pyruvate is converted into lactate (CH3CHOHCO2), and pumped out of the cell. As a consequence, the production of Hydrogen ions (H+) causes acidification, with the consequence of the stabilization of the Warburg metabolic cycle. In cancer cells, there is a net conversion of serine to glycine, catalysed by the cytosolic (SHMT1) or mitochondrial (SHMT2) serine hydroxymethyl transferase, and correlated with the cell proliferation rate and the DNA synthesis rate [79,80]. Protein synthesis, proportional to the inflow rate of amino acids, requires energy: in cancer, around the 70% of glucose is converted to lactate during aerobic glycolysis [79,81]. Therefore, our results point to the fundamental role played by ion transfer in any protein cycle. Indeed, ion channels transduce surface events to the cytosolic protein machineries. They couple the sensitivity of cooperative allosteric proteins to chemical and electrical signals: to do so, they use the energy released during the passive ions flows. Here, we have developed the study of the Ca2+ flows and the related membrane potential variations, because changes in membrane potential can regulate Ca2+ influx, which can impact T cell activation. This process is triggered by an elevation of the cytosolic free calcium concentration, which activates the Ca2+/PKC-dependent pathways that regulate progress from G0 into mitosis, with a related lymphocyte proliferation, as an effective immune response to cancer. The Ca2+ inflow is obtained by means of hyperpolarization, which is induced through K+ channel activation [5].
Here, a theoretical model to analyse the ion fluxes was developed using non-equilibrium thermodynamics. It represents a useful tool for future analysis, in order to develop a new approach to anticancer therapies, based on ion fluxes.
Recently, the key role for Ca2+ was shown to be in regulating cancer, in relation to oncogenes protecting against cell death, and perturbing intracellular Ca2+ homeostasis. Indeed, oncoprotein B cell lymphoma 2 over-expression has been shown to be able to reduce steady-state Ca2+ levels within the endoplasmic reticulum, reducing Ca2+ transfer to the mitochondria, during apoptotic stimulation, and inhibiting apoptosis initiation [20,21,56]. Moreover, the protein mitogenic kinase Akt has been linked to Ca2+ homeostasis control, pointing out its modulation function on the phosphorylation state of IP3R, by inhibiting its Ca2+-channel activity, and reducing the transfer of Ca2+ from the endoplasmic reticulum to the mitochondria [25].
In conclusion, our results agree with the experimental evidence in the literature [32,79,82,83,84], and could represent their biophysical explanation based on non-equilibrium thermodynamics. Moreover, this approach could support the new frontier in cancer therapies [14,32,85,86,87,88,89].
Last, we can evaluate the Ca2+-fluxes in Equation (12) as follows:
J Ca = · α μ Ca · R ( T T 0 ) = 0.004 × 10 6 × 0.023 × 0.2 0.8 × 0.7 0.35 × 0.6 552.79 × 10 3 · R 2 ( T T 0 ) = 0.97 × 10 17 [ mol s 1 ] R 2 [ m 2 ]
where 0.004 μm is the depth of the cell membrane [90] and R is the mean radius of the cell, considered, in the first approximation, as a sphere, μ Ca = 552.79 kJ mol−1, and T T 0 0.4 °C [91]. The numerical result depends on the mean size of the cell. Considering that the mean radius for human cell is of the order of 10 6 10 5 m, it follows that the Ca2+-flux is of the order of 21–450 mmol s−1m−2, which can be expressed as ∼0.010 mol s−1kg−1, in agreement with the experimental results obtained in [92].

Author Contributions

Conceptualization, U.L.; methodology, U.L. and G.G.; software, G.G.; validation, U.L. and G.G.; formal analysis, U.L.; investigation, G.G.; resources, U.L.; data curation, G.G.; writing—original draft preparation, U.L. and G.G.; writing—review and editing, U.L. 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.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. WHO. Fact Sheet; World Health Organization: Geneva, Switzerland, 2020; Available online: https://www.who.int/news-room/fact-sheets/detail/cancer (accessed on 24 September 2020).
  2. Verginadis, I.; Velalopoulou, A.; Karagounis, I.; Simos, Y.; Peschos, D.; Karkabounas, S.; Evangelou, A. Beneficial effects of electromagnetic radiation in cancer. In Electromagnetic Radiation; Bashir, S.O., Ed.; InTech: Shanghai, China, 2012. [Google Scholar]
  3. Tozzi, M.; Sørensen, C.E.; Magni, L.; Christensen, N.M.; Bouazzi, R.; Buch, C.M.; Stefanini, M.; Duranti, C.; Arcangeli, A.; Novak, I. Proton Pump Inhibitors Reduce Pancreatic Adenocarcinoma Progression by Selectively Targeting H+, K+-ATPases in Pancreatic Cancer and Stellate Cells. Cancers 2020, 12, 640. [Google Scholar] [CrossRef] [Green Version]
  4. Becchetti, A.; Crescioli, S.; Zanieri, F.; Petroni, G.; Mercatelli, R.; Coppola, S.; Gasparoli, L.; D’Amico, M.; Pillozzi, S.; Crociani, O.; et al. The conformational state of hERG1 channels determines integrin association, downstream signaling, and cancer progression. Sci. Signal. 2020, 10, 473. [Google Scholar] [CrossRef] [Green Version]
  5. Becchetti, A. Ion channels and transporters in cancer. 1. Ion channels and cell proliferation in cancer. Am. J. Physiol. Cell Physiol. 2011, 301, C255–C265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Lucia, U.; Deisboeck, T.S. The importance of ion fluxes for cancer proliferation and metastasis: A thermodynamic analysis. J. Theor. Biol. 2018, 445, 1–8. [Google Scholar] [CrossRef] [PubMed]
  7. Lucia, U. Thermodynamic approach to nano-properties of cell membrane. Physica A 2014, 407, 185–191. [Google Scholar] [CrossRef]
  8. Lucia, U. Transport processes and irreversible thermodynamics analysis in tumoral systems. Physica A 2014, 410, 380–390. [Google Scholar] [CrossRef]
  9. Cardone, R.A.; Casavola, V.; Reshkin, S.J. The role of distrurbed pH dynamics and the Na+/H+ exchanger in metastasis. Nat. Rev. Cancer 2005, 5, 786–795. [Google Scholar] [CrossRef] [PubMed]
  10. Demirel, Y. Nonequilibrium Thermodynamics: Transport and Rate Processes in Physical, Chemical and Biological Systems; Elsevier: Amsterdam, The Netherlands, 2007. [Google Scholar]
  11. Lucia, U. A link between nano- and classical thermodynamics: Dissipation analysis (The entropy generation approach in nano-thermodynamics). Entropy 2015, 17, 1309–1328. [Google Scholar] [CrossRef]
  12. Djamgoz, M.B.A.; Arcangeli, A. Bioelectricity of Cancer. Bioelectricity 2020, 1, 113. [Google Scholar] [CrossRef] [Green Version]
  13. Marchi, S.; Pinton, P. Alterations of calcium homeostasis in cancer cells. Curr. Opin. Pharmacol. 2016, 29, 1–6. [Google Scholar] [CrossRef]
  14. Bonora, M.; Bononi, A.; Marchi, E.D.; Carlotta, G.; Lebiedzinska, M.; Marchi, S.; Patergnani, S.; Rimessi, A.; Suski, J.M.; Wojtala, A.; et al. Role of the c subunit of the FO ATP synthase in mitochondrial permeability transition. Cell Cycle 2013, 12, 674–683. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Rimessi, A.; Patergnani, S.; Bonora, M.; Wieckowski, M.R.; Pinton, P. Mitochondrial Ca2+ remodeling is a prime factor in oncogenic behavior. Front. Oncol. 2015, 5, 143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Bonora, M.; Wieckowski, M.R.; Chinopoulos, C.; Kepp, O.; Kroemer, G.; Galluzzi, L.; Pinton, P. Molecular mechanisms of cell death: Central implication of ATP synthase in mitochondrial permeability transition. Oncogene 2015, 34, 1475–1486. [Google Scholar] [CrossRef] [PubMed]
  17. Rizzuto, R.; Marchi, S.; Bonora, M.; Aguiari, P.; Bononi, A.; Stefani, D.D.; Giorgi, C.; Leo, S.; Rimessi, A.; Siviero, R.; et al. Ca(2+) transfer from the ER to mitochondria: When, how and why. Biochim. Biophys. Acta 2009, 1787, 1342–1351. [Google Scholar] [CrossRef] [Green Version]
  18. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [Green Version]
  19. Giorgi, C.; Missiroli, S.; Patergnani, S.; Duszynski, J.; Wieckowski, M.R.; Pinton, P. Mitochondria-associated membranes: Composition, molecular mechanisms, and physiopathological implications. Antioxid. Redox Signal. 2015, 22, 995–1019. [Google Scholar] [CrossRef]
  20. Pinton, P.; Ferrari, D.; Rapizzi, E.; Virgilio, F.D.; Pozzan, T.; Rizzuto, R. The Ca2+ concentration of the endoplasmic reticulum is a key determinant of ceramide-induced apoptosis: Significance for the molecular mechanism of Bcl-2 action. EMBO 2001, 20, 2690–2701. [Google Scholar] [CrossRef]
  21. Pinton, P.; Ferrari, D.; Magalhaes, P.; Schulze-Osthoff, K.; Virgilio, F.D.; Pozzan, T.; Rizzuto, R. Reduced loading of intracellular Ca(2+) stores and downregulationof capacitative Ca(2+) influx in Bcl-2-overexpressing cells. J. Cell Biol. 2000, 148, 857–862. [Google Scholar] [CrossRef] [Green Version]
  22. Foyouzi-Youssefi, R.; Arnaudeau, S.; Borner, C.; Kelley, W.L.; Tschopp, J.; Lew, D.P.; Demaurex, N.; Krause, K.H. Bcl-2 decreases the free Ca2+ concentration within the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 2000, 97, 5723–5728. [Google Scholar] [CrossRef] [Green Version]
  23. Akl, H.; Vervloessem, T.; Kiviluoto, S.; Bittremieux, M.; Parys, J.B.; Smedt, H.D.; Bultynck, G. A dual role for the anti-apoptotic Bcl-2 protein in cancer: Mitochondria versus endoplasmic reticulum. Biochim. Biophys. Acta 2014, 1843, 2240–2252. [Google Scholar] [CrossRef] [Green Version]
  24. Akl, H.; Bultynck, G. Altered Ca(2+) signaling in cancer cells: Proto-oncogenes and tumor suppressors targeting IP3 receptors. Biochim. Biophys. Acta 2013, 1835, 180–193. [Google Scholar] [PubMed]
  25. Marchi, S.; Marinello, M.; Bononi, A.; Bonora, M.; Giorgi, C.; Rimessi, A.; Pinton, P. Selective modulation of subtype III IP(3)R by Akt regulates ER Ca2+ releaseand apoptosis. Cell Death Dis. 2012, 3, e304. [Google Scholar] [CrossRef] [Green Version]
  26. Giorgi, C.; Ito, K.; Lin, H.K.; Santangelo, C.; Wieckowski, M.R.; Lebiedzinska, M.; Bononi, A.; Bonora, M.; Duszynski, J.; Bernardi, R.; et al. PML regulates apoptosis at endoplasmic reticulum by modulating calcium release. Science 2019, 330, 1247–1251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Stewart, T.A.; Yapa, K.T.; Monteith, G.R. Altered calcium signaling in cancer cells. Biochim. Biophys. Acta 2015, 1848, 2502–2511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Bononi, A.; Bonora, M.; Marchi, S.; Missiroli, S.; Poletti, F.; Giorgi, C.; Pandolfi, P.P.; Pinton, P. Identification of PTEN at the ER and MAMs and its regulation of Ca2+ signaling and apoptosis in a protein phosphatase-dependent manner. Cell Death Differ. 2013, 20, 1631–1643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. Giorgi, C.; Bonora, M.; Sorrentino, G.; Missiroli, S.; Poletti, F.; Suski, J.M.; Galindo Ramirez, F.; Rizzuto, R.; Di Virgilio, F.; Zito, E.; et al. p53 at the endoplasmic reticulum regulates apoptosis in a Ca2+-dependent manner. Proc. Natl. Acad. Sci. USA 2015, 112, 1779–1784. [Google Scholar] [CrossRef] [Green Version]
  30. Giorgi, C.; Bonora, M.; Missiroli, S.; Poletti, F.; Ramirez, F.G.; Morciano, G.; Morganti, C.; Pandolfi, P.P.; Mammano, F.; Pinton, P. Intravital imaging reveals p53-dependent cancer cell death induced by phototherapy via calcium signaling. Oncotarget 2015, 6, 1435–1445. [Google Scholar] [CrossRef] [Green Version]
  31. Rimessi, A.; Marchi, S.; Patergnani, S.; Pinton, P. H-Ras-driven tumoral main-tenance is sustained through caveolin-1-dependent alterations in calcium signaling. Oncogene 2014, 33, 2329–2340. [Google Scholar] [CrossRef]
  32. Bonora, M.; Giorgi, C.; Pinton, P. Novel frontiers in calcium signaling: A possible target for chemotherapy. Pharmacol. Res. 2015, 99, 82–85. [Google Scholar] [CrossRef]
  33. Šileikytė, J.; Forte, M. The Mitochondrial Permeability Transition in Mitochondrial Disorders. Oxid. Med. Cell. Longev. 2019, 2019, 3403075. [Google Scholar] [CrossRef] [Green Version]
  34. Lucia, U.; Grisolia, G. How Life Works—A Continuous Seebeck-Peltier Transition in Cell Membrane? Entropy 2020, 22, 960. [Google Scholar] [CrossRef]
  35. Yang, M.; Brackenbury, W.J. Membrane potential and cancer progression. Front. Physiol. 2013, 4, 185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. 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] [Green Version]
  37. 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]
  38. 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]
  39. Cone, C.D. Variation of the transmembrane potential level as a basic mechanism of mitosis control. Oncology 1970, 24, 438–470. [Google Scholar] [CrossRef]
  40. 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]
  41. Tokuoka, S.; Marioka, H. The membrane potential of the human cancer and related cells (I). Gann 1957, 48, 353–354. [Google Scholar]
  42. Altman, P.L.; Katz, D. Biological Handbook Vol. 1: Cell Biology; Federation of American Society for Experimental Biology: Bethesda, MD, USA, 1976. [Google Scholar]
  43. Balitsky, K.P.; Shuba, E.P. Resting potential of malignant tumour cells. Acta-Unio Int. Contra Cancrum 1964, 20, 1391–1393. [Google Scholar]
  44. Jamakosmanovic, A.; Loewenstein, W. Intracellular communication and tissue growth. III. Thyroid cancer. J. Cell Biol. 1968, 38, 556–561. [Google Scholar] [CrossRef] [Green Version]
  45. Binggelli, R.; Cameron, I.L. Cellular Potential of Normal and Cancerous Fibroblasts and Hepatocytes. Cancer Res. 1980, 40, 1830–1835. [Google Scholar]
  46. Sundelacruz, S.; Levin, M.; Kaplan, D.L. Role of the membrane potential in the regulation of cell proliferation and differentiation. Stem Cell Rev. 2009, 5, 231–246. [Google Scholar] [CrossRef] [PubMed]
  47. 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]
  48. 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]
  49. Johnstone, R.M. Microelectrode penetration of ascites tumour cells. Nature 1959, 183, 411. [Google Scholar] [CrossRef]
  50. Marino, A.A.; Morris, D.M.; Schwalke, M.A.; Iliev, I.G.; Rogers, S. Electrical potential measurements in huma breast cancer and benign lesions. Tumor Biol. 1994, 15, 147–152. [Google Scholar] [CrossRef]
  51. 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]
  52. Goldman, D.E. Potential impedance, and rectification in membranes. J. Gen. Physiol. 1943, 27, 37–60. [Google Scholar] [CrossRef] [Green Version]
  53. 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]
  54. Grabe, M.; Wang, H.; Oster, G. The mechanochemistry of V-ATPase proton pumps. Biophys. J. 2000, 78, 2798–2813. [Google Scholar] [CrossRef] [Green Version]
  55. Junge, D. Nerve and Muscle Excitation, 2nd ed.; Sunderland: New York, NY, USA, 1981. [Google Scholar]
  56. 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] [Green Version]
  57. Wright, S.H. Generation of resting membrane potential. Adv. Physiol. Educ. 2004, 28, 139–142. [Google Scholar] [CrossRef] [PubMed]
  58. Yourgrau, W.; van der Merwe, A.; Raw, G. Treatise on Irreversible and Statistical Thermophysics; Dover: New York, NY, USA, 1982. [Google Scholar]
  59. Callen, H.B. Thermodynamics; Wiley: New York, NY, USA, 1960. [Google Scholar]
  60. Katchalsky, A.; Currant, P.F. Nonequilibrium Thermodynamics in Biophysics; Harvard University Press: Boston, MA, USA, 1965. [Google Scholar]
  61. Demirel, Y.; Sandler, S.I. Nonequilibrium Thermodynamics in Engineering and Science. J. Phys. Chem. B 2004, 108, 31–43. [Google Scholar] [CrossRef]
  62. Onsager, L. Reciprocal relations in irreversible processes: I. Phys. Rev. 1931, 37, 405–426. [Google Scholar] [CrossRef]
  63. Onsager, L. Reciprocal relations in irreversible processes: II. Phys. Rev. 1931, 38, 2265–2279. [Google Scholar] [CrossRef] [Green Version]
  64. de Groot, S.R. Thermodynamics of Irreversible Processes; North-Holland Publishing Co.: Amsterdam, The Netherlands, 1952. [Google Scholar]
  65. Lucia, U.; Grisolia, G. Thermal Resonance and Cell Behavior. Entropy 2020, 22, 774. [Google Scholar] [CrossRef]
  66. Lucia, U.; Grisolia, G. Resonance in Thermal Fluxes Through Cancer Membrane. Atti Dell’Accademia Peloritana Dei Pericolanti 2020, 98, SC1–SC6. [Google Scholar] [CrossRef]
  67. Schrödinger, E. What’s life? The Physical Aspect of the Living Cell; Cambridge University Press: Cambridge, UK, 1944. [Google Scholar]
  68. Lucia, U.; Grisolia, G. Second law efficiency for living cells. Front. Biosci. 2017, 9, 270–275. [Google Scholar] [CrossRef] [Green Version]
  69. Prigogine, I. Structure, dissipation and life. In Theoretical Physics and Biology; Marois, M., Ed.; North Holland Pub. Co.: Amsterdam, The Netherlands, 1969. [Google Scholar]
  70. Tombes, R.M.; Grant, S.; Westin, E.H.; Krystal, G. G1 cell cycle arrest and apoptosis are induced in NIH 3T3 cells by KN-93, an inhibitor of CaMK-II (the multifunctional Ca2+/CaM kinase). Cell Growth Differ. 1995, 6, 1063–1979. [Google Scholar]
  71. Wulff, H.; Miller, M.J.; Hansel, W.; Grissmer, S.; Cahalan, M.D.; Chandy, K.G. Design of a potent and selective inhibitor of the intermediate conductance Ca2+-activated K+ channel, IKCa1: A potential immunosuppressant. Proc. Natl. Acad. Sci. USA 2000, 97, 8151–8156. [Google Scholar] [CrossRef] [Green Version]
  72. Amuthan, G.; Biswas, G.; Ananadatheerthavarada, H.K.; Vijayasarathy, C.; Shephard, H.M.; Avadhani, N.G. Mitochondrial stress-induced calcium signaling, phenotypic changes and invasive behavior in human lung carcinoma A549 cells. Oncogene 2002, 21, 7839–7849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Nakanishi-Matsui, M.; Sekiya, M.; Futai, R.K.N.M. The mechanism of rotating proton pumping ATPases. Biochim. Biophys. Acta 2010, 1797, 1343–1352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Stevens, T.H.; Forgac, M. Structure, function and regulation of the vacuolar (H+)-ATPase. Annu. Rev. Cell Dev. Biol. 1997, 13, 779–808. [Google Scholar] [CrossRef]
  75. Tuszynski, J.A.; Kurzynski, M. Introduction to Molecular Biophysics; CRC Press: Boca Raton, FL, USA, 2003; pp. 383–392. [Google Scholar]
  76. Lucia, U.; Ponzetto, A.; Deisboeck, T.S. A thermo-physical analysis of the proton pump vacuolar-ATPase: The constructal approach. Sci. Rep. 2014, 4, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Rudolph, M.G.; Stanfield, R.L.; Wilson, I.A. How TCRs bind MHCs, peptides, and coreceptors. Annu. Rev. Immunol. 2006, 24, 419–466. [Google Scholar] [CrossRef] [PubMed]
  78. Strong, R.K. Asymmetric ligand recognition by the activating natural killer cell receptor NKG2D, a symmetric homodimer. Mol. Immunol. 2002, 38, 1029–1037. [Google Scholar] [CrossRef]
  79. Dolfi, S.C.; Chang, L.L.Y.; Qiu, J.; Tedeschi, P.M.; Bertino, J.R.; Hirshfield, K.M.; Oltvai, Z.N.; Vazquez, A. The metabolic demand of cancer cells are coupled to their size and protein synthesis rates. Cancer Metab. 2013, 1, 20–32. [Google Scholar] [CrossRef] [Green Version]
  80. Racker, E. History of the Pasteur effect and its pathobiology. Mol. Cell Biochem. 1974, 54, 17–23. [Google Scholar] [CrossRef]
  81. Jain, M.; Nilsson, R.; Sharma, S.; Madhusudhan, N.; Kitami, T.; Souza, A.L.; Kafri, R.; Kirshner, M.W.; Clish, C.B.; Mootha, V.K. Metabolite profiling identifies a key role for glycine in rapid cancer cell proliferation. Science 2012, 336, 1040–1044. [Google Scholar] [CrossRef] [Green Version]
  82. Munaron, L.; Antoniotti, S.; Lovisolo, D. Intracellular calcium signals and control of cell proliferation: How many mechanisms? J. Cell. Mol. Med. 2004, 8, 161–168. [Google Scholar] [CrossRef]
  83. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion channels and the hallmarks of cancer. Trends Mol. Med. 2010, 16, 107–121. [Google Scholar] [CrossRef] [PubMed]
  84. Roderick, H.L.; Cook, S.J. Ca2+ signalling checkpoints in cancer: Remodelling Ca2+ for cancer cell proliferation and survival. Nat. Rev. Cancer 2008, 8, 361–375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Arcangeli, A.; Crociani, O.; Lastroioli, E.; Masi, A.; Pillozzi, S.; Becchetti, A. Targeting ion channels in cancer: A novel frontiers in antineoplastic therapy. Curr. Med. Chem. 2009, 16, 66–93. [Google Scholar] [CrossRef] [PubMed]
  86. Lucia, U.; Ponzetto, A.; Deisboeck, T.S. A thermodynamic approach to the ‘mitosis/apoptosis’ ratio in cancer. Physica A 2015, 436, 246–255. [Google Scholar] [CrossRef]
  87. Lucia, U.; Ponzetto, A.; Deisboeck, T.S. Constructal approach to cell membranes transport: Amending the ‘Norton-Simon’ hypothesis for cancer treatment. Sci. Rep. 2016, 6, 19451. [Google Scholar] [CrossRef] [PubMed]
  88. Bergandi, L.; Lucia, U.; Grisolia, G.; Granata, R.; Gesmundo, T.; Ponzetto, A.; Paolucci, E.; Borchiellini, R.; Ghigo, E.; Silvagno, F. The extremely low frequency electromagnetic stimulation selective for cancer cells elicits growth arrest through a metabolic shift. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 1389–1397. [Google Scholar] [CrossRef]
  89. Lucia, U.; Grisolia, G.; Ponzetto, A.; Silvagno, F. An engineering thermodynamic approach to select the electromagnetic wave effective on cell growth. J. Theor. Biol. 2017, 429, 181–189. [Google Scholar] [CrossRef]
  90. Milo, R.; Phillips, R. Cell Biology by the Numbers; Garland Science: New York, NY, USA, 2015. [Google Scholar]
  91. 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.
  92. Borle, A.B. An Overview of Techniques for the Measurement of Calcium Distribution, Calcium Fluxes, and Cytosolic Free Calcium in Mammalian Cells. Environ. Health Perspect. 1990, 84, 45–56. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Lucia, U.; Grisolia, G. Non-Equilibrium Thermodynamic Approach to Ca2+-Fluxes in Cancer. Appl. Sci. 2020, 10, 6737. https://doi.org/10.3390/app10196737

AMA Style

Lucia U, Grisolia G. Non-Equilibrium Thermodynamic Approach to Ca2+-Fluxes in Cancer. Applied Sciences. 2020; 10(19):6737. https://doi.org/10.3390/app10196737

Chicago/Turabian Style

Lucia, Umberto, and Giulia Grisolia. 2020. "Non-Equilibrium Thermodynamic Approach to Ca2+-Fluxes in Cancer" Applied Sciences 10, no. 19: 6737. https://doi.org/10.3390/app10196737

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