<<

VOLUME 78, NUMBER 18 PHYSICAL REVIEW LETTERS 5MAY 1997

Spinodal Decomposition in Binary

S. Bastea and J. L. Lebowitz Department of Physics and Mathematics, Rutgers University, Piscataway, New Jersey 08855-0849 (Received 10 December 1996) We carried out three-dimensional simulations, with about 1.4 3 106 particles, of segregation in a low density binary fluid mixture, described mesoscopically by energy and momentum conserving Boltzmann-Vlasov equations. Using a combination of direct simulation Monte Carlo for the short range collisions and a version of particle-in-cell evolution for the smooth long range interaction, we found dynamical scaling after the ratio of the interface thickness to the domain size is less than ϳ0.1. The a 1 scaling length R͑t͒ grows at late times like t , with a ෇ 1 for critical quenches and a ෇ 3 for off- critical ones. We also measured the variation of temperature, total particle density, and hydrodynamic velocity during the segregation process. [S0031-9007(97)03106-2]

PACS numbers: 64.75.+g, 47.70.Nd, 68.10.–m

The process of phase segregation through which a sys- coexistence region they will describe -gas segregation tem evolves towards equilibrium following a temperature [14] into two phases, one rich in particles of type 1 and quench from a homogeneous phase into a two phase re- the other rich in particles of type 2. (Examples of gas gion of its phase diagram has been of continuing interest mixtures that have a miscibility gap are helium-hydrogen, during the last decades [1], but many problems still re- helium-nitrogen, neon-xenon, etc. [14].) We believe main to be solved. This is particularly so for fluids,when that the model contains the essential features of phase particle, momentum, and energy densities are conserved separation in general binary fluid mixtures. locally; these are currently the focus of both numerical To simulate our system we modeled the Boltzmann studies [2–7] and microgravity experiments [8,9]. collisional part using a stochastic algorithm due to Bird In this Letter we present computer simulations of [15], known as direct simulation Monte Carlo (DSMC), spinodal decomposition in a three-dimensional mixture while for the Vlasov part we used the particle-to-grid- of two kinds of particles that we label 1 and 2 using a weighting method, well known in physics [16]. novel microscopic dynamics and computational scheme. In the DSMC method the physical space is divided into The particles interact with each other through short range cells containing typically tens of particles. The main interactions modeled here by hard spheres having the ingredients of this procedure are the alternation of free same mass m and diameter d. Particles of different kinds flow over a time interval Dt and representative collisions interact also through a long range repulsive Kac potential, 3 among pairs of particles sharing the same cell. In the V͑r͒ ෇ g U͑gr͒. The equilibrium properties of such particle-to-grid-weighting algorithm the particle densities a system are well understood; there is even a rigorous are computed on a spatial grid through some weighting proof of a at low temperatures to an depending on the particle position; then the Vlasov forces immiscible state [10], which in the limit g ! 0 [11] is are calculated on the same grid. Finally, the forces at the described by the mean-field theory. When the density n position of each particle are interpolated from the forces is low enough, nd3 ø 1, and the potential sufficiently on the grid. The coupling of these methods, which have long ranged, ng23 ¿ 1, the free energy of the system been extensively used individually, made possible our is well approximated by F k T n r៬ ln n r៬ 1 ෇ B ͓ 1͑ ͒ 1͑ ͒ simulations of phase segregation with 1.4 3 106 particles, n r៬ ln n r៬ dr 1 V jr៬ 2 r៬ j n r៬ n r៬ dr៬ dr៬ and 2͑ ͒ 2͑ ͔͒ ͑ 1 2 ͒ 1͑ 1R͒ 2͑ 2͒ 1 2 with only modest computational resources: a typical run the g ! 0 critical temperature, which should be an upper g R 0 1 took about 32 CPU hours on a 233 MHz Alpha Station. bound for Tc at g.0, is given by kBTc ෇ 2 n U͑r͒ dr៬. It appears that this method can be extended to the study In this regime the dynamical evolution of the system R of the effects of phase segregation on inhomogeneous should be well described by two coupled Boltzmann- hydrodynamical flows of practical importance [17]. Vlasov equations: Since one of our main interests was the late time ≠fi ≠fi F៬i ≠fi 1 y?៬ 1 ? ෇J͓fi,f1 1f2͔, (1) hydrodynamical regime, a delicate balance had to be struck ≠t ≠r៬ m ≠y៬ between the size of the system, the range of the potential, i ෇ 1, 2, where fi͑r៬, y៬ , t͒ are the one-particle distribu- the temperature, and the particle density, making sure that 0 0 0 tion functions, F៬ i͑r៬, t͒ ෇ 2= V͑jr៬ 2 r៬ j͒nj͑r៬ ͒ dr៬ , each of the methods is used within its range of validity and n r៬0 f r៬0, y, t dy with i, j 1, 2, i fi j, and that their combination remains computer manageable. On j͑ ͒ ෇ j͑ ៬ ͒ ៬ R ෇ J f, g is the Boltzmann collision operator for hard core the one hand the potential must be reasonably long ranged ͓ ͔ R interactions [12]. Kinetic equations of this type have been so that the Vlasov description is physically appropriate proposed in [13], and if the system is quenched inside the and numerically sound, and on the other hand it must

0031-9007͞97͞78(18)͞3499(4)$10.00 © 1997 The American Physical Society 3499 VOLUME 78, NUMBER 18 PHYSICAL REVIEW LETTERS 5MAY 1997 have a range much smaller than the size of the system. We first present results for critical quenches (equal This restriction made necessary the use of two spatial volume fractions of the two phases). Three different po- grids: a somewhat coarse one for the collisions and a finer tential ranges were used, and for each one of them 10–12 grid for the long range potential. It also imposed the quenches were performed. The domain size was probed use of quadratic spline interpolation for the calculation of using the pair correlation function, C͑r៬, t͒ ෇ V 21 3 grid quantities and a ten-point difference scheme for the ͗ x៬ f͑x៬, t͒f͑x៬ 1 r៬, t͒͘, where f ෇ ͑n1 2 n2͒͑͞n1 1 calculation of the forces [16]. n ͒ is the local order parameter, n and n are the local P2 1 2 Our results were obtained using a system with 1 382 400 particle densities, V is the volume, and the average particles in a cube with periodic boundary conditions. We is over the different runs. We determined C͑r៬, t͒ by also studied smaller systems to identify unavoidable finite- an inverse Fourier transform of the structure function size effects. The interaction potential used was Gaussian, S͑k៬, t͒ ෇ jf˜ ͑k៬, t͒j2; the order parameter f͑r៬, t͒ and its 3 2 2x2 ˜ ៬ U͑x͒ ෇ ap 2 e , a.0, but there is no reason to be- Fourier transform f͑k, t͒ were computed on a 64 3 lieve that different repulsive potentials would qualitatively 64 3 64 cubic grid. The first zero of the spherically change the results. All quenches were performed at a to- averaged correlation function, C͑r, t͒, was used as a tal particle density nd3 Ӎ 0.01 and an initial temperature measure of the typical domain size, R͑t͒. The data are 0 averages of the independent runs. T0, T0͞Tc ෇ 0.5. The initial conditions for each run were random positions for all particles and velocities distributed Assuming the existence of a single characteristic length according to a Maxwellian of constant temperature. (In the scale, the dynamical scaling prediction [1] for the late DSMC evolution, as in the Boltzmann-Vlasov equations, time spherically averaged correlation function is C͑r, t͒Ӎ the hard cores only enter in determining the collision cross C͓r͞R͑t͔͒. In Fig. 1 the correlation function for potential 21 sections.) The total energy of the system was very well range g ෇ 0.4 is plotted starting at t ෇ 160, showing conserved by the dynamics. This meant that the kinetic that the system is well within the scaling regime. energy and hence the temperature increased as the system Simple dimensional analysis of the hydrodynamical segregated, but at late times it changed very slowly on the evolution equations in the limit of large domain sizes, time scale of our simulations. We indicate the final tem- appropriate for late times, yields a linear growth law for the domain sizes [19], R͑t͒ ~ ͑s͞h͒t, when R is below perature T in Figs. 1–4. The effective number of particles 2 1 2 2 in the range of the potential was about 100–500. Rh ෇ h ͞rs and a t 3 law, R͑t͒ ~ ͑s͞r͒ 3 t 3 , for R above In the following we compare results of our simula- Rh [20,21]; s is the surface tension coefficient, h is the tions with available theoretical and experimental work and shear viscosity, and r is the density. We measured s check various assumptions made in the former, e.g., the ne- using Laplace’s law and h by studying the decay of a glect of density and temperature variations. We are also sinusoidal velocity profile. The time evolution of R͑t͒ in currently investigating both formal and rigorous Chapman- our simulations is at late times (see Fig. 2) R͑t͒ ෇ a 1 Enskog and Hilbert expansion methods for derivation of b͑s͞h͒t, with b Ӎ 0.13 6 0.2 (s ෇ 260 and h ෇ 3250 21 0 macroscopic evolution equations for this model [18]. The for g ෇ 0.4, T͞Tc ෇ 0.6). This numerical factor is units used for lengths and times are the mean-free path, similar to the one observed in experiments [8] and recent 1 2 21 l ෇ ͑2 2 pnd ͒ , and mean-free time, t ෇ l͞c, where 1 1 0 c ෇ ͑2kBT0͞m͒ 2 and T0 ෇ 2 Tc is the initial temperature.

FIG. 2. Scaled domain growth at critical quench; the g21 ෇ 0.3 and g21 ෇ 0.5 curves have been collapsed onto the g21 ෇ FIG. 1. Scaled two-point correlation function for critical 0.4 curve. A straight line fit (a ෇ 1) is drawn for late times 1 1 21 0 3 quench with g ෇ 0.4. Final temperature T ෇ 0.6Tc . and a ~ t fit (a ෇ 3 ) is drawn for early times. 3500 VOLUME 78, NUMBER 18 PHYSICAL REVIEW LETTERS 5MAY 1997

to be roughly equal to g21, far away from the critical temperature, as we are [22]. We also looked directly at the wall profiles separating different phases and found that they are approximately de- scribed by solitonic solutions [23]. These are particle den- sities ni͑z͒ depending on a single spatial coordinate, such 2 that fi͑z, y៬ ͒ ෇ ni͑z͒ exp͑2my ͞2kBT͒ are stationary solutions of Eq. (1) and n1͑6`͒ ෇ n2͑7`͒ are the mean- field equilibrium densities at temperature T. The order parameter profiles which satisfy the equation and the ones observed in the simulation (see Fig. 4) have both approxi- mately the hyperbolic tangent form [22], tanh͑z͞2j͒, with z the coordinate perpendicular to the domain wall and j a parameter which characterizes the interface thickness. We found that this thickness is about 50% bigger than g21. The total density is about 20% smaller (for g21 ෇ 0.4, 21 0 FIG. 3. Domain growth for off-critical quenches with g ෇ T͞Tc ෇ 0.6) at the interface than in the bulk, in very 0.4; straight line fits are drawn. good agreement with the solitonic solution (see Fig. 4). The solitonic profiles were calculated at an effective tem- perature Teff for which the asymptotic values of the order parameter matched the ones observed in the simulations. large scale molecular dynamics simulations [7]. Whether 2 These asymptotically matched mean-field profiles are or not a crossover to a t 3 growth occurs at later times, steeper (in units of g21) than the ones observed, with larger domain sizes cannot be decided by our present better agreement as g is decreased. We believe that this results. We estimated that we would need a system with at g 0 discrepancy is due to the fact that T͞Tc . Teff͞T ,as least 4 times as many particles as the present one to observe c the equilibrium curve for g.0is flatter around Tc than the growth of domains with sizes bigger than Rh. the mean-field curve. The linear regime starts around the time when dynami- The need for a clear separation between boundary width cal scaling begins to hold, i.e., at a domain size of about 21 and domain size may explain why in earlier molecular 12–15 times g . This is in agreement with Siggia [19] dynamics simulations [2], a smaller exponent is found: in who argued that the linear regime is due to surface tension those computations the maximum domain size observed is driven flows, so the interfaces should be well defined, i.e., only 6–8 times the potential range, so the hydrodynamic their width should be small compared to the domain size. This width is usually taken to be of order j, the correlation regime was probably never reached. In simulations using the lattice Boltzmann method the interaction range is of length of order parameter fluctuations. One expects j the order of one lattice spacing and the hydrodynamic exponent is observed when the domain size is about ten

lattice units [4], in agreement with our analysis. 1 At early times the growth is consistent with a t 3 behavior, for all potential ranges [19,24]. This exponent is not associated with a scaling regime and it may not be universal, but it is not inconsistent with the experimental results [8,9]. In fact, we were able to collapse the three curves onto each other through scaling of the lengths and times Fig. 2 [6]. Off-critical quenches were performed for a single range of the potential and three volume fractions. To our knowl- edge no published results of such simulations exist, al- though they are mentioned in [7]. Therefore, we compare our results to recent experimental work [9] and analyze them using the known coarsening mechanisms [19,25]. In Fig. 3 we present the domain growth for g21 ෇ 0.4 and volume fractions 0.16, 0.22, and 0.28. We plot R3͑t͒ vs t, and the late times regime is clearly consistent with and order parameter f ͑1͒ an exponent of 1 . In this regime the system satisfies ͒ء͑ FIG. 4. Total density n variations at the interface between the two phases; g21 ෇ 0.4 3 0 dynamical77 scaling very well. The presence of a late and T͞Tc ෇ 0.6. The full lines are the solitonic solution (see 1 text). times t 3 growth at volume fractions less or equal than

3501 VOLUME 78, NUMBER 18 PHYSICAL REVIEW LETTERS 5MAY 1997 about 0.3 was observed recently in microgravity experi- We thank Frank Alexander for useful discussions. This ments [9] and reasonably explained theoretically using a research was supported in part by AFOSR Grant No. 0159 droplet coalescence mechanism [19,25]. This regime was and NSF Grant No. 92-13424. also analyzed by Siggia [19] who predicted a prefactor 1 proportional to y 3 (y being the volume fraction of the [1] J. D. Gunton, M. San Miguel, and P. S. Sahni, in Phase minority phase). In our simulations the prefactors are in Transitions and Critical Phenomena, edited by C. Domb reasonable agreement with the above prediction. Further- and J. L. Lebowitz (Academic, New York, 1983), Vol. 8; more, we can clearly see the motion and coalescence of A. J. Bray, Adv. Phys. 43, 357 (1994). the droplets in movies of the dynamics. [2] Wen-Jong Ma, A. Maritan, J. Banavar, and J. Koplik, In writing down evolution equations for the order pa- Phys. Rev. A 45, R5347 (1992). rameter and velocity fields in a symmetric binary system [3] O. T. Valls and J. E. Farrell, Phys. Rev. E 47, R36 (1993); with momentum and energy conservation it is generally as- A. Shinozaki and Y. Oono, Phys. Rev. E 48, 2622 (1993). sumed that the density and temperature variations are small [4] F. J. Alexander, S. Chen, and D. W. Grunau, Phys. Rev. B [26]. We were able to check these quantities directly by 48, 634 (1993). dividing the system into “hydrodynamical” cells, each con- [5] Y. Wu, F. J. Alexander, T. Lookman, and S. Chen, Phys. taining about 100 particles, and computing the order pa- Rev. Lett. 74, 3852 (1995); T. Lookman, Y. Wu, F. J. 1 2 2 Alexander, and S. Chen, Phys. Rev. E 53, 5513 (1996). rameter, density, temperature [kBT ϵ 3 m͑͗y៬ ͘ 2 ͗y៬ ͘ ͒, [6] S. Bastea and J. L. Lebowitz, Phys. Rev. E 52, 3821 averages over the cell], and fluid velocity in each cell. (1995). While the statistical fluctuations of these quantities are not [7] M. Laradji, S. Toxvaerd, and O. G. Mouritsen, Phys. Rev. vanishingly small, as they should be for true hydrodynamic Lett. 77, 2253 (1996). variables, they are small enough for such a description to [8] P. Guenoun, R. Gastaud, F. Perrot, and D. Beysans, Phys. be at least reasonable. We found that at late times and Rev. A 36, 4876 (1987). away from the domain boundaries the values of the density [9] F. Perrot, P. Guenoun, T. Baumberger, and D. Beysens, and order parameter were close to their equilibrium values. Phys. Rev. Lett. 73, 688 (1994); Y. Garrabos, B. Le Nein- The temperature appeared to be uniform, with normal equi- dre, P. Guenoun, B. Khalil, and D. Beysans, Europhys. librium fluctuations everywhere in the system. This shows Lett. 19, 491 (1992). that the time scales over which heat transport takes place [10] J. L. Lebowitz and E. H. Lieb, Phys. Lett. 39A, 98 (1972). are much smaller than the time scales over which there is [11] J. L. Lebowitz and O. Penrose, J. Math. Phys. (N.Y.) 7,98 (1966). significant phase separation. [12] S. Chapman and T. G. Cowling, The Mathematical Theory If, as argued, the linear growth regime is due to sur- of Non-uniform Gases (Cambridge University Press, Cam- face tension driven flows, one would expect bigger hy- bridge, England, 1970). drodynamic velocities close to the interfaces than in the [13] Luis De Sobrino, J. Can. Phys. 45, 363 (1967); M. Grmela, bulk. We looked therefore at the distribution of hydrody- J. Stat. Phys. 3, 347 (1971). namic velocities as a function of the order parameter for [14] J. A. Schouten, Phys. Rep. 172, 33 (1989). critical quenches. As the system segregates, bigger hy- [15] G. A. Bird, Molecular Gas Dynamics (Clarendon Press, drodynamic velocities are typically observed close to the Oxford, 1976). domain walls, identified by small values of the order pa- [16] C. K. Birdsall and A. B. Langdon, Plasma Physics Via rameter. While this may be due in part to the density being Computer Simulation (McGraw-Hill Book Company, New smaller at the boundary between domains, it is also consis- York, 1985). [17] R. E. Goldstein, A. I. Pesci, and M. J. Shelley, Phys. Rev. tent with the idea that at late times the flows are generated Lett. 70, 3043 (1993). mainly by the curvature of well formed interfaces. We are [18] S. Bastea, R. Esposito, J. L. Lebowitz, and R. Marra (to planning to study this issue more carefully using bigger be published). systems. [19] E. D. Siggia, Phys. Rev. A 20, 595 (1979). A natural refinement of our model would be the use [20] H. Furukawa, Physica (Amsterdam) 204A, 237 (1994). of the Enskog correction for the collisions [12] (which [21] S. Bastea and J. L. Lebowitz, Phys. Rev. Lett. 75, 3776 would require numerically replacing the DSMC part with (1995). its recently proposed extension [27]) and of long range [22] J. S. Rowlinson and B. Widom, Molecular Theory of attractive forces between like particles. Capillarity (Clarendon Press, Oxford, 1982). We believe that we have introduced a new model which [23] G. Giacomin and J. L. Lebowitz, Phys. Rev. Lett. 76, 1094 is closer to reality than lattice gases usually simulated, (1996). but still tractable numerically by the new techniques that [24] A. Z. Akcasu and R. Klein, Macromolecules 26, 1429 (1993). we have introduced. These techniques are of independent [25] V. S. Nikolayev, D. Beysens, and P. Guenoun, Phys. Rev. interest and have enabled us to compute temperature Lett. 76, 3144 (1996). changes, density variations, and profiles not done before. [26] P. C. Hohenberg and B. I. Halperin, Rev. Mod. Phys. 49, Furthermore, we can compare our results to some exact 435 (1977). ones. In fact, we think that this model is a paradigm for [27] F. J. Alexander, A. L. Garcia, and B. J. Adler, Phys. Rev. the rigorous derivation of hydrodynamic equations. Lett. 74, 5212 (1995).

3502