Skip to main content
Log in

Phase Growth with Heat Diffusion in a Stochastic Lattice Model

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

When a stable phase is adjacent to a metastable phase with a planar interface, the stable phase grows. We propose a stochastic lattice model describing the phase growth accompanying heat diffusion. The model is based on an energy-conserving Potts model with a kinetic energy term defined on a two-dimensional lattice, where each site is sparse-randomly connected in one direction and local in the other direction. For this model, we calculate the stable and metastable phases exactly using statistical mechanics. Performing numerical simulations, we measure the displacement of the interface R(t). We observe the scaling relation \(R(t)=L_x \bar{\mathcal {R}} (Dt/L_x^2)\), where D is the thermal diffusion constant and \(L_x\) is the system size between the two heat baths. The scaling function \(\bar{\mathcal {R}}(z)\) shows \(\bar{\mathcal {R}}(z) \simeq z^{0.5}\) for \(z \ll z_c\) and \(\bar{\mathcal {R}}(z) \simeq z^{\alpha }\) for \(z \gg z_c\), where the cross-over value \(z_c\) and exponent \(\alpha \) depend on the temperatures of the baths, and \(0.5\le \alpha \le 1\). We then confirm that a deterministic phase-field model exhibits the same scaling relation. Moreover, numerical simulations of the phase-field model show that the cross-over value \(\bar{\mathcal {R}}(z_c)\) approaches zero when the stable phase becomes neutral.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Data Availibility

The datasets generated during and/or analysed during the current study are available from the corresponding author on reasonable request.

References

  1. Callen, H.B.: Thermodynamics and an Introduction to Thermostatistics, 2nd edn. Wiley, New York (1985)

    MATH  Google Scholar 

  2. Pomeau, Y.: Front motion, metastability and subcritical bifurcations in hydrodynamics. Physica D 23, 3 (1986)

    Article  ADS  Google Scholar 

  3. Stefan, J.: über die theorie der eisbildung, insbesondere über die eisbildung im polarmeere, Sitzungsberichte der Österreichischen Akademie der Wissenschaften Mathematisch- Naturwissenschaftliche Klasse. Abteilung 2, Mathematik, Astronomie, Physik, Meteorologie und Technik 98, 965 (1889)

  4. Stefan, J.: über die theorie der eisbildung, insbesondere über die eisbildung im polarmeere. Ann. Phys. (Leipzig) 42, 269 (1891)

  5. Fix, G.J.: Phase field methods for free boundary problems. In: Fasano, A., Primicerio, M. (eds.) Free Boundary Problems: Theory and Applications, vol. 2. Piman, Boston (1983)

    MATH  Google Scholar 

  6. Caginalp, G.: Surface tension and supercooling in solidification theory. In: Springer Lecture Notes in Physics, Applications of Field Theory to Statistical Mechanics. Springer, Berlin (1984)

  7. Langer, J.S.: Models of pattern formation in first-order phase transitions. In: Grinstein, G., Mazenko, G. (eds.) Directions in Condensed Matter Physics. World Scientific, Philadelphia (1986)

    Google Scholar 

  8. Collins, J.B., Levine, H.: Diffuse interface model of diffusion-limited crystal growth. Phys. Rev. B 33, 2020 (1986)

    Article  ADS  Google Scholar 

  9. Penrose, O., Fife, P.C.: Thermodynamically consistent models of phase-field type for the kinetics of phase transition. Physica D 43, 44 (1990)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Kobayashi, R.: Modeling and numerical simulations of dendritic crystal growth. Physica D 63, 410 (1993)

    Article  ADS  MATH  Google Scholar 

  11. Zener, C.: Theory of growth of spherical precipitates from solid solution. J. Appl. Phys. 20, 950 (1949)

    Article  ADS  Google Scholar 

  12. Dewynne, J.N., Howison, S.D., Ockendon, J.R., Xie, W.: Asymptotic behavior of solutions to the Stefan problem with a kinetic condition at the free boundary. J. Austral. Math. Soc. Ser. B 31, 81 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  13. Löwen, H., Bechhoefer, J., Tuckerman, L.S.: Crystal growth at long times: critical behavior at the crossover from diffusion to kinetics-limited regimes. Phys. Rev. A 45, 2399 (1992)

    Article  ADS  Google Scholar 

  14. Hiraizumi, M., Sasa, S.-i.: Perturbative solution of a propagating interface in the phase field model. J. Stat. Mech. 103203 (2021)

  15. Halperin, B.I., Hohenberg, P.C., Ma, S.K.: Renormalization group methods for critical dynamics. I. Recursion relations and effects of energy conservation. Phys. Rev. B 10, 139 (1974)

    Article  ADS  Google Scholar 

  16. Anderson, D.M., McFadden, G.B., Wheeler, A.A.: Diffuse-interface methods in fluid mechanics. Annu. Rev. Fluid Mech. 30, 139 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Hohenberg, P.C., Halperin, B.I.: Theory of dynamic critical phenomena. Rev. Mod. Phys. 49, 435 (1977)

    Article  ADS  Google Scholar 

  18. Townsend, R.M., Rice, S.A.: Molecular dynamics studies of the liquid-vapor interface of water. J. Chem. Phys. 94, 2207 (1991)

    Article  ADS  Google Scholar 

  19. Burton, W.K., Cabrera, N., Frank, F.C.: The growth of crystals and the equilibrium structure of their surfaces. Phil. Trans. R. Soc. Lond. Ser. A 234, 299 (1951)

    ADS  MathSciNet  MATH  Google Scholar 

  20. Gilmer, G.H., Bennema, P.: Simulation of Crystal Growth with Surface Diffusion. J. Appl. Phys. 43, 1347 (1972)

    Article  ADS  Google Scholar 

  21. Müller-Krumbhaar, H.: Master-equation approach to stochastic models of crystal growth. Phys. Rev. B 10, 1308 (1974)

    Article  ADS  Google Scholar 

  22. Weeks, J.D., Gilmer, G.H., Jackson, K.A.: Analytical theory of crystal growth. J. Chem. Phys. 65, 712 (1976)

    Article  ADS  Google Scholar 

  23. Saito, Y., Müller-Krumbhaar, H.: Diffusion and relaxation kinetics in stochastic models for crystal growth. J. Chem. Phys. 70, 1078 (1979)

    Article  ADS  Google Scholar 

  24. Gilmer, G.H.: Computer models of crystal growth. Science 208, 355 (1980)

    Article  ADS  Google Scholar 

  25. Witten, T.A., Sander, L.M.: Diffusion-limited aggregation, a kinetic critical phenomenon. Phys. Rev. B 27, 5686 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  26. Moss, R., Harrowell, P.: Dynamic Monte Carlo simulations of freezing and melting at the 100 and 111 surfaces of the simple cubic phase in the face-centered-cubic lattice gas. J. Chem. Phys. 100, 7630 (1994)

    Article  ADS  Google Scholar 

  27. Novotny, M.A., et al.: Simulations of metastable decay in two- and three-dimensional models with microscopic dynamics. J. Non-Cryst. Solids 274, 356 (2000)

    Article  ADS  Google Scholar 

  28. Wu, F.Y.: The Potts model. Rev. Mod. Phys. 54, 235 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  29. Kadanoff, L., Swift, J.: Transport coefficients near the liquid-gas critical point. Phys. Rev. 165, 310 (1968)

    Article  ADS  Google Scholar 

  30. Creutz, M.: Microcanonical Monte Carlo simulation. Phys. Rev. Lett. 50, 1411 (1983)

    Article  ADS  MathSciNet  Google Scholar 

  31. Casartelli, M., Macellari, N., Vezzani, A.: Heat conduction in a two-dimensional Ising model. Eur. Phys. J. B 56, 149 (2007)

    Article  ADS  Google Scholar 

  32. Sahni, P.S., Grest, G.S., Anderson, M.P., Srolovitz, D.J.: Kinetics of the \(Q-\)state Potts model in two dimensions. Phys. Rev. Lett. 50, 263 (1983)

    Article  ADS  Google Scholar 

  33. Kaski, K., Nieminen, J., Gunton, J.D.: Domain growth and scaling in the Q-state Potts model. Phys. Rev. B 31, 2998 (1985)

    Article  ADS  Google Scholar 

  34. Clément, Sire., Satya, N., Majumdar: Coarsening in the q-state Potts model and the Ising model with globally conserved magnetization. Phys. Rev. E 52, 244 (1995)

  35. Ohta, H., Rosinberg, M.L., Tarjus, G.: Morphology transition at depinning in a solvable model of interface growth in a random medium. Europhys. Lett. 104, 16003 (2013)

    Article  ADS  Google Scholar 

  36. Mézard, M., Montanari, A.: Information, Physics, and Computation. Oxford University Press, Oxford (2009)

    Book  MATH  Google Scholar 

  37. Dembo, A., Montanari, A.: Gibbs measures and phase transitions on sparse random graphs. Braz. J. Probab. Stat. 24, 137 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  38. Karma, A., Rappel, W.J.: Phase-field model of dendritic sidebranching with thermal noise. Phys. Rev. E 60, 3614 (1999)

    Article  ADS  Google Scholar 

  39. Hiraizumi, M., Ohta, H., Sasa, S.-I.: unpublished

  40. Mullins, W.W., Sekerka, R.F.: Stability of a planar interface during solidification of a dilute binary alloy. J. Appl. Phys. 34, 323 (1963)

    Article  ADS  Google Scholar 

  41. Kupfermann, R., Shochet, O., Ben-Jacob, E., Schuss, Z.: Phase-field model: boundary layer, velocity of propagation, and the stability spectrum. Phys. Rev. B 46, 16045 (1992)

    Article  ADS  Google Scholar 

  42. Braun, R.J., McFadden, G.B., Coriell, S.R.: Morphological instability in phase-field models of solidification. Phys. Rev. E 49, 4336 (1994)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

This work was supported by KAKENHI (Grant Nos. 17H01148, 19H05795, and 20K20425).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shin-ichi Sasa.

Ethics declarations

Conflict of interest

The authors have no financial or proprietary interests in any material discussed in this article.

Additional information

Communicated by Hal Tasaki.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Derivation of the Formulas in Sect. 3

In this section, we derive formulas (16), (17), (18), and (22) in Sect. 3.

1.1 A.1 Derivation of (16)

We study a Cayley tree with a root site connected with four sites in the first generation. Each site in the n-th generation \((n \ge 1)\) is connected with three sites in the \(n+1\)-th generation. See Fig. 13 for the illustration of the Cayley tree.

Fig. 13
figure 13

Illustration of the Cayley tree

Let Z be the partition function of the Potts model on the lattice. We consider the partition function of a system in which a root site is replaced by the cavity and the state of a site in the first generation is fixed as \(\sigma ' \in \{1,\cdots , q\}\), which is denoted by \(\tilde{Z}_1(\sigma ')\). Z is then the partition function of the model expressed as

$$\begin{aligned} Z= \sum _{\sigma } \left( \sum _{\sigma '} e^{\beta \delta (\sigma ,\sigma ')} \tilde{Z}_1(\sigma ')\right) ^4. \end{aligned}$$
(A.1)

A graphical representation is displayed in Fig. 14.

Fig. 14
figure 14

Graphical representation of (A.1)

By setting

$$\begin{aligned} \gamma&\equiv e^\beta -1,\end{aligned}$$
(A.2)
$$\begin{aligned} G_1&\equiv \sum _\sigma \tilde{Z_1}(\sigma ) , \end{aligned}$$
(A.3)

we rewrite (A.1) as

$$\begin{aligned} Z= \sum _{\sigma } \left( \gamma \tilde{Z}_1(\sigma )+G_1 \right) ^4. \end{aligned}$$
(A.4)

Defining \(\tilde{Z}_n(\sigma )\) and \(G_n\) similarly, we have the iterative equation

$$\begin{aligned} \tilde{Z}_{n}(\sigma )= \left( \gamma \tilde{Z}_{n+1}(\sigma )+G_n \right) ^3, \end{aligned}$$
(A.5)

whose graphical representation is shown in Fig. 15.

Fig. 15
figure 15

Graphical representation of (A.5)

We define \(u_n(\sigma )\) as

$$\begin{aligned} u_n(\sigma )\equiv \frac{\tilde{Z_n}(\sigma )}{G_n}, \end{aligned}$$
(A.6)

which corresponds to the probability of the state \(\sigma \) of the cavity-connected site in the n-th generation. By substituting (A.6) into (A.5), we obtain

$$\begin{aligned} G_n u_n(\sigma )&=G_{n+1}^3 \left[ \gamma u_{n+1}(\sigma ) +1\right] ^3. \end{aligned}$$
(A.7)

We also have

$$\begin{aligned} G_n=G_{n+1}^3\sum _\sigma \left[ \gamma u_{n+1}(\sigma )+ 1\right] ^3 \end{aligned}$$
(A.8)

using \(\sum _{\sigma } u_n(\sigma )=1\). From (A.7) and (A.8), we derive the iterative equation for \(u_n(\sigma )\),

$$\begin{aligned} u_n(\sigma )=\frac{\left[ \gamma u_{n+1}(\sigma )+1\right] ^3}{\sum _\sigma \left[ \gamma u_{n+1}(\sigma )+1\right] ^3 }. \end{aligned}$$
(A.9)

Assuming homogeneity in the equilibrium state, \(u_n(\sigma )\) is independent of n in the large-size limit. This provides (16).

1.2 A.2 Derivation of (17)

The order parameter m for the model is calculated by the expectation value of \(\delta (\sigma ,1)\) at the root site. That is,

$$\begin{aligned} m=\frac{1}{Z}\sum _{\sigma }\delta (\sigma ,1) \left[ \gamma \tilde{Z}_1(\sigma ) + G_1 \right] ^4. \end{aligned}$$
(A.10)

Using the expression given in (A.4), we have

$$\begin{aligned} m=\frac{[\gamma u_1(1)+1]^4}{\sum _\sigma [\gamma u_1(\sigma )+1]^4}. \end{aligned}$$
(A.11)

By replacing \(u_1\) with the solution of the self-consistent equation (16), we obtain (17).

1.3 A.3 Derivation of (18)

To derive the free energy density, we use a tactical method manipulating graphs. We first remove one edge connected to the root site. The partition function of this system with \(\sigma \) at the root site and \(\sigma '\) at the other site connected by the removed edge is \(\tilde{Z}_0(\sigma )\tilde{Z}_1(\sigma ')\). See Fig. 16. We thus express the partition function Z as

$$\begin{aligned} Z&=\sum _{\sigma ,\sigma ^\prime } e^{\beta \delta (\sigma ,\sigma ^\prime )} \tilde{Z_0}(\sigma )\tilde{Z_1}(\sigma ^\prime )\end{aligned}$$
(A.12)
$$\begin{aligned}&=G_0G_1\left[ \gamma \sum _\sigma u_0(\sigma )u_1(\sigma ) +1 \right] , \end{aligned}$$
(A.13)

where \(G_0\equiv \sum _{\sigma }\tilde{Z}_0(\sigma )\) and \(u_0(\sigma )\equiv \tilde{Z}_0(\sigma )/G_0\). Note that \(u_0(\sigma )\) also satisfies (A.9).

Fig. 16
figure 16

By removing one edge, we get two rooted graphs. \(\times \) represents the cavity

Next, we prepare four independent systems. The partition function of the total system is \(Z^4\). We remove one edge connected to the root site for each graph. Then, we combine four graphs with the root site by adding one site. See Fig. 17. The partition function of this system, \(\check{Z_0}\), is expressed as

$$\begin{aligned} \check{Z_0}=G_0^4\sum _\sigma \left[ \gamma u_0(\sigma )+1 \right] ^4. \end{aligned}$$
(A.14)
Fig. 17
figure 17

By adding one site, we combine four graphs with the root site

Similarly, another Cayley tree is obtained by combining the other graphs with another added site, and the partition function is written as

$$\begin{aligned} \check{Z_1}=G_1^4\sum _\sigma \left[ \gamma u_1(\sigma )+1 \right] ^4 . \end{aligned}$$
(A.15)

The free energies of the original system and the new system are \(-T\log Z^4\) and \(-T\log \check{Z_0}\check{Z_1}\), respectively. The difference in free energy is equal to 2f, where f is the free energy density, because the two systems have the same free energy density in the large-size limit and the new system is the original system with two sites added. That is,

$$\begin{aligned} - T\log \check{Z_0}\check{Z_1} + T\log Z^4 = 2f. \end{aligned}$$
(A.16)

This is rewritten as

$$\begin{aligned}&e^{-\beta f}=\left( \frac{\check{Z}_0\check{Z}_1}{Z^4}\right) ^{1/2}\end{aligned}$$
(A.17)
$$\begin{aligned}&=\left( \frac{\sum _{\sigma '}[\gamma u_0(\sigma ')+1]^4 \sum _{\sigma ''}[\gamma u_1(\sigma '')+1]^4}{\left[ \gamma \sum _\sigma u_0(\sigma )u_1(\sigma )+1 \right] ^4} \right) ^{1/2}. \end{aligned}$$
(A.18)

By replacing \(u_0\) and \(u_1\) with the solution of the self-consistent equation (16), we obtain (18).

1.4 A.4 Derivation of (22)

For later convenience, we set

$$\begin{aligned} \tilde{c} \equiv \frac{1-c}{q-1}. \end{aligned}$$
(A.19)

From the definition of \(\tilde{f}\) given in (20), the left-hand side of (22) is calculated as

$$\begin{aligned}&\frac{\partial }{\partial c} e^{-\beta \tilde{f}(\beta ,c)} \nonumber \\&\quad =4\gamma \frac{ (\gamma c+1)^3-(\gamma \tilde{c}+1)^3}{(\gamma c^2+\gamma (q-1)\tilde{c}^2+1)^2} \nonumber \\&\qquad -4\gamma (c-\tilde{c} ) \frac{(\gamma c+1)^4+ (q-1)(\gamma \tilde{c}+1)^4 }{ (\gamma c^2+\gamma (q-1)\tilde{c}^2+1)^3 }. \end{aligned}$$
(A.20)

The self-consistent equation (16) is expressed as

$$\begin{aligned} (\gamma \tilde{c}+1)^3 =\frac{\tilde{c}}{c}(\gamma c+1)^3. \end{aligned}$$
(A.21)

Thus, the right-hand side of (A.20) for the special values of c satisfying (A.21) is calculated as

$$\begin{aligned}&4\gamma \frac{(\gamma c+1)^3 (c-\tilde{c} ) }{c (\gamma c^2+\gamma (q-1)\tilde{c}+1)^2 } \nonumber \\&\quad \times \left( 1-\frac{\gamma c^2+\gamma (q-1)\tilde{c}^2 +c+(q-1)\tilde{c} }{ \gamma c^2+\gamma (q-1)\tilde{c}^2 +1} \right) , \end{aligned}$$
(A.22)

which turns out to be zero from (A.19).

Appendix B: Estimation of \(\varDelta \)

In this section, we estimate the value of \(\varDelta \) defined by (1) for the model we study.

We first calculate the energy density h defined as

$$\begin{aligned} h \equiv \lim _{|\varLambda | \rightarrow \infty } \frac{1}{|\varLambda |} \sum _{\sigma ,p} P_\mathrm{can}(\sigma ,p) H(\sigma ,p), \end{aligned}$$
(B.23)

where \(P_\mathrm{can}(\sigma ,p)\) is given in (12). Using the free energy density f calculated in Sect. 3, we express the energy density h as

$$\begin{aligned} h=T+g, \end{aligned}$$
(B.24)

where g is the potential energy density given by

$$\begin{aligned} g(\beta )\equiv \frac{\partial }{\partial \beta } \left( \beta f(\beta ) \right) . \end{aligned}$$
(B.25)

As with the free energy density, \(g_0(\beta )\) and \(g_*(\beta )\) denote the potential energy densities corresponding to the trivial solution \(u_0\) and the nontrivial solution \(u_*\) of (16), respectively. In Fig. 18a, \(g_0(\beta )\) and \(g_*(\beta )\) are displayed. Then, the latent heat per unit volume \(T_c\delta s\) at the equilibrium transition temperature is determined by the entropy jump defined as

$$\begin{aligned} \delta s\equiv \beta _c(g_0(\beta _c)-g_*(\beta _c)). \end{aligned}$$
(B.26)

For the model with \(q=10\), we obtain \(T_c\delta s\simeq 1.07\). Next, we consider the heat capacity per unit volume C expressed as

$$\begin{aligned} C(\beta )\equiv \frac{\partial h}{\partial T} =1+\frac{\partial g}{\partial T}. \end{aligned}$$
(B.27)

Using similar notations, we obtain \(C_0(\beta )\) and \(C_*(\beta )\) from \(g_0(\beta )\) and \(g_*(\beta )\). These are shown at the bottom of Fig. 18b. For the cases \(q=10\) and \(T_L=1.01T_c\), we obtain \(C_*(T_L)\simeq 6.95\). Therefore, for the model we numerically study, we have

$$\begin{aligned} \varDelta \simeq 0.05, \end{aligned}$$
(B.28)

which is less than unity. Note that in the stochastic model studied in this paper, C corresponds to \(c_p\) in the phase-field model.

Fig. 18
figure 18

a Potential energy density g as a function of \(\beta \). The solid (purple) line represents \(g=g_*\) in the range \(\beta > \beta _\mathrm{sp}\). The dashed (green) line represents \(g=g_0\) in the range \(\beta < \beta _\mathrm{un}\). The black line indicates \(\beta =\beta _c\). b Heat capacity per unit volume C as a function of \(\beta \). The styles (colors) of lines correspond to the graphs in (a) (Color figure online)

Appendix C: Estimation of D

In this section, we estimate the value of the thermal diffusion constant D by measuring the relaxation property of the temperature profile T(xt), where

$$\begin{aligned} T(x,t)\equiv \frac{1}{L_y}\sum _{y=1}^{L_y} p_{x,y}(t). \end{aligned}$$
(C.29)

For simplicity, we study the case \(T_R=T_L=1.2T_c\) with the initial condition

$$\begin{aligned} T(x,0)\equiv 1.2T_c+\sin \left( \frac{\pi (x-1)}{L_x-1}\right) . \end{aligned}$$
(C.30)

To realize the initial condition T(x, 0), \(\sigma _i\) is randomly chosen with equal probability and \(p_i=T(i_x,0)\) for any i. We then define the spatial average of the local temperature as

$$\begin{aligned} \bar{T}(t)\equiv \left\langle \frac{1}{L_x}\sum _{x=1}^{L_x}T(x,t) \right\rangle , \end{aligned}$$
(C.31)

where \(\langle \cdot \rangle \) denotes the average over ten independent samples. Assuming the diffusion equation for T(xt), we have

$$\begin{aligned} \frac{\bar{T}(t)}{T_c}=1.2+Be^{-D\pi ^2 t/L_x^2}, \end{aligned}$$
(C.32)

where D is the thermal diffusion constant and B is a parameter associated with the initial condition. As shown in Fig. 19, we find that the fitting of (C.31) with (C.32) works well for various system sizes with \(B = 1\). From this fitting, we estimate \(D = 1.9\times 10^{-2}\).

Fig. 19
figure 19

\(\bar{T}/T_c\) as a function of \(\pi ^2t/L_x^2\). The solid line represents the fitting curve (C.32) (Color figure online)

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hiraizumi, M., Ohta, H. & Sasa, Si. Phase Growth with Heat Diffusion in a Stochastic Lattice Model. J Stat Phys 189, 28 (2022). https://doi.org/10.1007/s10955-022-02990-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-022-02990-8

Keywords

Navigation