Next Article in Journal
Mylar Balloon and Associated Geometro-Mechanical Moments
Previous Article in Journal
On Nearly Sasakian and Nearly Kähler Statistical Manifolds
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

From Ion Fluxes in Living Cells to Metabolic Power Considerations

1
Dipartimento Energia “Galileo Ferraris”, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
2
Istituto Nazionale di Fisica Nucleare—Sezione di Torino, 10125 Torino, Italy
3
Dipartimento di Scienza Applicata e Tecnologia, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy
*
Authors to whom correspondence should be addressed.
Mathematics 2023, 11(12), 2645; https://doi.org/10.3390/math11122645
Submission received: 16 May 2023 / Revised: 5 June 2023 / Accepted: 8 June 2023 / Published: 9 June 2023

Abstract

:
Recently, the role of thermal resonance has been highlighted in living cells. As a consequence of this approach, the electrochemical potential was obtained in a partial differential equation concerning the cell membrane depth and its external temperature surface. In this paper, this last equation is studied and its solution’s consequences are discussed concerning the cells’ ion fluxes and their related entropy variation and power generation. Moreover, the metabolic power of the whole body is evaluated by using these previous numerical results.

1. Introduction

The wave model approach to heat transfer in solids pointed out the finite speed propagation [1,2], analytically expressed by the relaxation time [3]. Experimental evidence supports the wave model approach [4,5], with a particular interest in the thermal wave speed and relaxation time for media with non-homogeneous inner structures [3]. Moreover, the wave model resulted in being very effective for temperature ripples propagating in solids under abrupt boundary heating [6] and thermal shock formation [7], for fast-propagating crack tips [8], in transient stages of material damage due to thermal cracking [9], and in solids subjected to periodic surface irradiations of heat flux [10]. In analysing the interaction between waves and systems, the resonant effect plays a fundamental role in physics. Any system has a proper oscillation frequency. Let us consider a wave (mechanical or electromagnetic) with the resonant frequency of a system, which affects the system itself. This wave forces the system to enter into vibration [11]. The resonant approach was also introduced in a biophysical study of cancer cell behaviour [12,13,14] based on the thermodynamic experimental analysis of heat flux [15,16,17]. Indeed, cells are open systems that convert their metabolic energy into mechanical and chemical processes. A thermodynamic approach models the cell as a thermodynamic engine able to convert inflow energy into work [18]. Healthy and tumour cells have two different metabolisms [19]:
  • Healthy cells use the Krebs cycle, based on the oxidation of acetyl-CoA, derived from carbohydrates, fats, and proteins;
  • Cancer cells use the Warburg cycle, which is a modified cell metabolism that favours a specialised fermentation over the aerobic respiration pathway.
Independently of the metabolic cycle used, any cell must outflow heat into its environment [18,20,21] through the cell membrane. However, different heat outflows occur for different cycles. Heat transfer through the cell membrane was studied using the thermo-kinetic lumped model [12,13,14], obtaining a characteristic resonant time. The theoretical results were experimentally confirmed [15,16,17], too.
This paper develops the analytical consequences of this approach to propose some considerations on cell membrane interactions with the cell environment, which could represent the analytical basis for experimental investigation with the aim of improving our knowledge on the role of heat and mass transfer through living cell membranes.

2. Materials and Methods

The non-equilibrium thermodynamic approach was introduced [22] to study the heat and mass transfer through the living cell membrane concerning cancer [14,23,24] and glaucoma [25], obtaining a physical–mathematical model in agreement with the experimental results reported in the literature [26,27,28,29,30,31,32,33]. To do so, the Onsager general phenomenological relations were used [21,23,34,35]:
J e = L 11 μ e T L 12 T T 2 J Q = L 21 μ e T L 22 T T 2
where J e is the net current density [A m 2 ], related to ion fluxes through the membrane; J Q denotes the heat flux [W m 2 ]; μ e = μ + Z e ϕ [18,35] is the electrochemical potential [J mol 1 ], with μ the chemical potential [J mol 1 ], z e the electric charge [A s mol 1 ], and ϕ the membrane potential [V]; T is the living cell temperature; and L i j represents the phenomenological coefficients, such that [18] L 12 ( B ) = L 21 ( B ) (Onsager–Casimir relation [36]), L 11 0 and L 22 0 , and [18] L 11 L 22 L 12 L 21 > 0 . Starting from these phenomenological relations, the following equation was deduced [22]:
μ e r = μ e T α λ T s u r f T 0
where α and λ are the convection and conduction coefficients, respectively, T s u r f is the temperature of the membrane’s external surface, and T 0 is the temperature of the cell’s environment.
In this paper, we develop the analytical solution of this last equation and the consequences on cell behaviour. To do so, we consider splitting the electrochemical potential into its spatial and thermal components:
μ e = ϑ ( r ) φ ( T )
where r = l is the internal cell membrane surface, while r = 0 is the external cell’s membrane surface with temperature T ( 0 ) = T s u r f . Consequently, if we consider the temperature gradient through the membrane, and introduce the dimensionless variable x = r / l , it follows that
1 ϑ d ϑ d x = 1 φ d φ d T l α λ T surf T 0
The integration [37] of this differential equation results in
μ e r , T = γ exp λ α l ( T s u r f T 0 ) T + r l
where γ is the integration constant.
In Equation (5), the coefficient of convection can be evaluated as [12]
α 0.023 R e 0.8 P r 0.35 λ R
where λ 0.6 W m 1 K 1 is the conductivity of water, used also for the biological tissue [14], R e 0.2 is the Reynolds number, and P r 6.9 is the Prandtl number. The mean value of the cell radius R is considered of the order of 10 5 m. Consequently, the coefficient of convection results to be around 7.49 × 10 8 W m 2 K 1 .
The surface temperature changes in relation to the metabolic and biochemical activity of the cell, but in stationary conditions it can be assumed that T s u r f T 0 0.4 °C.
The constant γ can be obtained considering that the electrochemical potential on the external surface of the cell membrane is μ e ( 0 , T s u r f ) = μ e , o u t ( T s u r f ) , so Equation (5) results in
γ = μ e , o u t ( T s u r f ) exp λ α l ( T s u r f T 0 ) T s u r f
When r = l , with l 0.004 μ m as the mean value of the cell membrane depth, then μ e l , T s u r f = μ e , i n T s u r f . Equation (5) results in
μ e r , T = μ e , o u t T s u r f exp λ ( T T s u r f ) α l ( T s u r f T 0 ) + r l
remembering that μ e 0 , T s u r f = μ e , o u t T s u r f , μ e l , T s u r f = μ e , i n T s u r f .

3. Results

These results are in agreement with the present knowledge on the physiological behaviour of cells [38]. The result obtained is interesting because it represents the analytical approach to the living cell membrane electrochemical potential gradient. Indeed, it allows us to explain some experimental evidence that represents open problems in biophysics. In this paper, we have suggested a new viewpoint in relation to these biophysical aspects of cell behaviour, as deeply discussed in the next section.
Now, we consider the second law of thermodynamics [39]:
T d s d t = · J Q i = 1 N μ i J i i = 1 N J i · μ i
where s is the specific entropy, T is the temperature, and μ is the chemical potential. J S = J Q 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 [34]. This law allows us to evaluate the effect of the ion fluxes. To do so, we consider the better condition for life ( T σ 0 ) such that Equation (9) becomes
T d s d t = · J Q i = 1 N μ i J i
Following the Prigogine approach ( d s / d t = 0 ) [39], with T constant, Equation (10) becomes
· J Q i = 1 N μ i J i = 0
Now, we consider the first law of thermodynamics for the cell membrane [23]
d u d t d V = ρ c d T d t d V = δ Q ˙ = α ( T T 0 ) d A
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 7.49 × 10 8 W m 2 K 1 , as previously evaluated, A is the area of the cell membrane, V is the cell volume, and β = α d A / d V is constant [23]. Thus, it follows that
· J Q = α d A d V ( T T 0 )
and
i = 1 N · μ i J i = δ Q ˙ d V = α d A d V ( T T 0 )
i = 1 N μ i J i l · α R ( T T 0 )
where l 0.004 μ m is the depth of the cell membrane [40] and R is the mean radius of the cell, considered, in the first approximation, as a sphere of mean radius of the order of 10 6 10 5 m, and T T 0 0.4 °C [41]. To develop the physical analysis of the mathematical results, we can evaluate the power generated by the cell by means of these fundamental fluxes for its life. Considering that the electric potential at the membrane is of the order of 10–100 mV and its electric field is of the order of 10 7 –10 8 V m 1 , we can evaluate the basal metabolism as the power generated by the fluxes of the fundamental ions (Na + , K + , Cl , Ca 2 + )
W ˙ = i l · α μ i · R ( T T 0 ) · N · Z i e · E i · 4 π R 2 = i l · 0.023 R e 0.8 P r 0.35 λ μ i · R 2 ( T T 0 ) · N · Z i e · E i · 4 π R 2 = 4 π · N e · ( T T 0 ) · l · 0.023 R e 0.8 P r 0.35 λ · i Z i · E i μ i
where N = 6.022 × 10 23 mol 1 , e = 1.6 × 10 19 . Concerning the value of the elementary electric charge, Z i = 1 for i = Na + , K + , Z i = 1 for i = Cl + , Z i = 2 for i = Ca 2 + , and E is the electric field. The metabolic power of a cell is 5.79 × 10 12 W. Considering the total number of cells in a body ( 30.0 × 10 12 [40]), the metabolic power of a human body is 174 W. Considering an efficiency of around 40% (around 60% is outflown as heat) [40], the available power for the human body is around 70 W, of which around 75% is used to perform essential body functions, while 25% is used to maintain the electrical potential of the nerve cells in agreement with the accepted value in medicine [40].

4. Discussion and Conslusions

All living cells maintain a potential difference through their membrane. This results from different concentrations and permeabilities of ions across the living cell membrane. The transport of ions, nutrients, molecules, and water (active and passive) is achieved by channels and pumps within the cell membrane. This transport changes the internal and external ionic concentrations. Any change in membrane potential allows the cell to communicate and obtain information due to the electrical signals related to any potential variation. Active and passive ion transports across the cellular membrane contribute to the change in membrane potentials [42]. So, cells must continuously balance mass transport to maintain their electric membrane potential to regulate normal cell functions [43], for example
  • Mitogen-stimulated cell proliferation, mediated by K + channel [44];
  • K + channel inhibitors can block the activation of murine B lymphocytes and murine noncytolytic T lymphocytes [45];
  • Ca 2 + inflow drives G1/S transition [46];
  • Mice teratocarcinoma cells express L-type Ca 2 + and outward channels, and Na + and inward rectifier channels during differentiation.
The transport of some molecules, such as water, can occur due to concentration gradients, while for macromolecules, such as glucose or nucleotides, channels are needed [38]. For some ions, ATPase pumps are present as ion transporters and voltage-gated channels. Moreover, transport proteins can pump ions against their concentration gradient. An example is the transport of Sodium (Na + ) and Potassium (K + ). Table 1 represents their concentration and respective electric potentials.
Concerning the Na + /K + pump, we can highlight that Na + presents a higher concentration outside of the cell while K + presents a high concentration inside the cell. Consequently, an active outflow transport of Na + and an active inflow of K + follows. Moreover, for the Ca 2 + ion inflow, concerning cancer, this induces a growth decrease.
In summary, this paper is an improvement of the previous one on thermal resonance [13,14]. In that paper, a differential equation concerning the electrochemical potential was obtained. Here, the analytical solution of that equation is obtained and its biophysical consequences are discussed. The result obtained here allows us to evaluate the power generation of the living cells obtained by ion fluxes through the cells’ membranes.

Author Contributions

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

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data are already provided in the manuscript, with quoted reference.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chester, M. Second Sound in Solids. Phys. Rev. 1963, 131, 2013. [Google Scholar] [CrossRef]
  2. Weymann, H.D. Finite Speed of Propagation in Heat conduction, Diffusion and Viscous Shear Motion. Am. J. Phys. 1967, 35, 488–496. [Google Scholar] [CrossRef]
  3. Tzou, D.Y. The resonance phenomenon in thermal waves. Int. J. Eng. Sci. 1991, 29, 1167–1177. [Google Scholar] [CrossRef]
  4. Peshkov, V. Second Sound in Helium II. J. Phys. 1944, 8, 381–389. [Google Scholar]
  5. Kaminski, W. Hyperbolic Heat Conduction Equation for Materials With Nonhomogeneous Inner Structure. J. Heat Mass Transf. 1990, 112, 555–560. [Google Scholar] [CrossRef]
  6. Özişik, M.N.; Vick, B. Propagation and Reflection of Thermal Waves in a Finite. Medium. Int. J. Heat Mass Transf. 1984, 27, 1845–1854. [Google Scholar] [CrossRef]
  7. Tzou, D.Y. On the Thermal Shock Wave Induced by a Moving Heat Source. ASME Int. J. Heat Mass Transf. 1989, 111, 232–238. [Google Scholar] [CrossRef]
  8. Tzou, D.Y. Thermal Shock Waves Induced by a Moving Crack. ASME Int. J. Heat Mass Transf. 1990, 112, 21–27. [Google Scholar] [CrossRef]
  9. Tzou, D.Y. The effects of thermal shock waves on the crack initiation around a moving heat source. Eng. Fract. Mech. 1989, 34, 1109–1118. [Google Scholar] [CrossRef]
  10. Glass, D.E.; Özişik, M.N.; Vick, B. Non-Fourier effects on transient temperature resulting from periodic on-off heat flux. Int. J. Heat Mass Transf. 1987, 30, 1623–1631. [Google Scholar] [CrossRef]
  11. Feynman, R.; Leighton, R.; Sands, M. The Feynman Lectures on Physics, Volume I; Addison Wesley: Reading, UK, 2005. [Google Scholar]
  12. Lucia, U.; Grisolia, G. Resonance in thermal fluxes through cancer membrane. AAPP Atti Della Accad. Peloritana Dei Pericolanti 2020, 98, SC11–SC16. [Google Scholar] [CrossRef]
  13. Lucia, U.; Grisolia, G. Thermal resonance and cell behavior. Entropy 2020, 22, 774. [Google Scholar] [CrossRef]
  14. Lucia, U.; Grisolia, G. Thermal resonance in living cells to control their heat exchange: Possible applications in cancer treatment. Int. Commun. Heat Mass Transf. 2022, 131, 105842. [Google Scholar] [CrossRef]
  15. Lucia, U.; Grisolia, G.; Ponzetto, A.; Bergandi, L.; Silvagno, F. Thermomagnetic resonance affects cancer growth and motility: Thermomagnetic resonance and cancer. R. Soc. Open Sci. 2020, 7, 200299. [Google Scholar] [CrossRef]
  16. 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]
  17. Bergandi, L.; Lucia, U.; Grisolia, G.; Granata, R.; Gesmundo, I.; 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 2019, 1866, 1389–1397. [Google Scholar] [CrossRef]
  18. Katchalsky, A.; Kedem, O. Thermodynamics of Flow Processes in Biological Systems. Biophys. J. 1962, 2, 53–78. [Google Scholar] [CrossRef] [Green Version]
  19. Voet, D.; Voet, J.G. Biochemistry, 3rd ed.; John Wiley & Sons: New York, NY, USA, 2004. [Google Scholar]
  20. Schrödinger, E. What’s Life? The Physical Aspect of the Living Cell; Cambridge University Press: Cambridge, UK, 1944. [Google Scholar]
  21. Callen, H.B. Thermodynamics; Wiley: New York, NY, USA, 1960. [Google Scholar]
  22. Lucia, U.; Grisolia, G. Biofuels from Micro-Organisms: Thermodynamic Considerations on the Role of Electrochemical Potential on Micro-Organisms Growth. Appl. Sci. 2021, 11, 2591. [Google Scholar] [CrossRef]
  23. Lucia, U.; Grisolia, G. Non-equilibrium thermodynamic approach to Ca2+-fluxes in cancer. Appl. Sci. 2020, 10, 6737. [Google Scholar] [CrossRef]
  24. Lucia, U. Bioengineering thermodynamics: An engineering science for thermodynamics of biosystems. Int. J. Thermodyn. 2015, 18, 254–265. [Google Scholar] [CrossRef] [Green Version]
  25. Lucia, U.; Grisolia, G. Thermal Physics and Glaucoma: From Thermodynamic to Biophysical Considerations to Designing Future Therapies. Appl. Sci. 2020, 10, 7071. [Google Scholar] [CrossRef]
  26. Rizzuto, R.; Marchi, S.; Bonora, M.; Aguiari, P.; Bononi, A.; Stefani, D.D.; Giorgi, C.; Leo, S.; Rimessi, A.; Siviero, R.; et al. Ca2+ transfer from the ER to mitochondria: When, how and why. Biochim. Biophys. Acta 2009, 1787, 1342–1351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. 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] [PubMed] [Green Version]
  28. 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]
  29. 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]
  30. Akl, H.; Bultynck, G. Altered Ca2+ signaling in cancer cells: Proto-oncogenes and tumor suppressors targeting IP3 receptors. Biochim. Biophys. Acta 2013, 1835, 180–193. [Google Scholar] [CrossRef]
  31. 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] [Green Version]
  32. Parsadaniantz, S.M.; le Goazigo, A.R.; Sapienza, A.; Habas, C.; Baudouin, C. Glaucoma: A Degenerative Optic Neuropathy Related to Neuroinflammation? Cells 2020, 9, 535. [Google Scholar] [CrossRef] [Green Version]
  33. Soto, I.; Howell, G.R. The complex role of neuroinflammation in glaucoma. Cold Spring Harb. Perspect. Med. 2014, 4, a017269. [Google Scholar] [CrossRef]
  34. Yourgrau, W.; van der Merwe, A.; Raw, G. Treatise on Irreversible and Statistical Thermophysics; Dover: New York, NY, USA, 1982. [Google Scholar]
  35. Goupil, C.; Seifert, W.; Zabrocki, K.; Müller, E.; Snyder, G.J. Thermodynamics of Thermoelectric Phenomena and Applications. Entropy 2011, 13, 1481–1517. [Google Scholar] [CrossRef] [Green Version]
  36. Degroot, S.R.; Mazur, P. Non-Equilibrium Thermodynamics; North-Holland Publishing Company: Amsterdam, The Netherlands, 1962. [Google Scholar]
  37. Apostol, T.S. Calculus. Volume 2: Multi-Variable Calculus and Linear Algebra with Applications to Differential Equations and Probability; Wiley: Hoboken, NJ, USA, 1969. [Google Scholar]
  38. Sherwood, L. Human Physiology: From Cells to Systems; Brooks Cole: Andover, MA, USA, 2009. [Google Scholar]
  39. Prigogine, I. Etude Thermodynamique des Phénoménes Irréversibles; Desoer: Liége, Belgium, 1947. [Google Scholar]
  40. Milo, R.; Phillips, R. Cell Biology by the Numbers; Garland Science: New York, NY, USA, 2003. [Google Scholar]
  41. 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, MA, USA, 1971. [Google Scholar]
  42. Tuszynski, J.A. Molecular and Cellular Biophysics; CRC: Boca Raton, FL, USA, 2019. [Google Scholar]
  43. Sundelacruz, S.; Levin, M.; Kaplan, D.L. Role of membrane potential in the regulation of cell proliferation and differentiation. Stem Cell Rev. Rep. 2009, 5, 231–246. [Google Scholar] [CrossRef]
  44. Wonderlin, W.F.; Strobl, J.S. Potassium channels, proliferation and G1 progression. J. Membr. Biol. 1996, 154, 91–107. [Google Scholar] [CrossRef]
  45. Ouadid-Ahidouch, H.; Le Bourhis, X.; Roudbaraki, M.; Toillon, R.A.; Delcourt, P.; Prevarskaya, N. Changes in the K+ current-density of MCF-7 cells during progression through the cell cycle: Possible involvement of a h-ether.a-gogo K+ channel. Recept. Channels 2001, 7, 345–356. [Google Scholar]
  46. Ouadid-Ahidouch, H.; Ahidouch, A. K+ channel expression in human breast cancer cells: Involvement in cell cycle regulation and carcinogenesis. J. Membr. Biol. 2008, 221, 1–6. [Google Scholar] [CrossRef]
Table 1. Concentration, chemical potential (in water solution), and electric membrane potential of some ions in normal cells [42].
Table 1. Concentration, chemical potential (in water solution), and electric membrane potential of some ions in normal cells [42].
IonExtracellularIntracellularChemicalMembrane
SpeciesConcentrationConcentrationPotential μ i Potential E i
[mM][mM][kJ mol 1 ][mV]
Na + 18150−261.89+56
K + 1405−283.26−89
Cl 1207−131.26−76
Ca 2 + 1.20.1−553.04+125
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.; Grisolia, G. From Ion Fluxes in Living Cells to Metabolic Power Considerations. Mathematics 2023, 11, 2645. https://doi.org/10.3390/math11122645

AMA Style

Lucia U, Grisolia G. From Ion Fluxes in Living Cells to Metabolic Power Considerations. Mathematics. 2023; 11(12):2645. https://doi.org/10.3390/math11122645

Chicago/Turabian Style

Lucia, Umberto, and Giulia Grisolia. 2023. "From Ion Fluxes in Living Cells to Metabolic Power Considerations" Mathematics 11, no. 12: 2645. https://doi.org/10.3390/math11122645

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