The liquid–vapor transition starts with the formation of a sufficiently large bubble in the metastable liquid to trigger the phase transition. Understanding this process is of fundamental and practical interest, but its study is challenging because it occurs over timescales that are too short for experiments but too long for simulations. The seeding method estimates cavitation rates by simulating a liquid in which a bubble is inserted, thus avoiding the long times needed for its formation. In one-component systems, in the NpT ensemble, the bubble grows or redissolves depending on whether its size is larger or smaller than the critical size, whereas in the NVT ensemble (i.e., at constant number of particles, volume, and temperature), the critical bubble can remain in equilibrium. Provided that a good criterion is used to determine the bubble size, this method, combined with the Classical Nucleation Theory (CNT), gives cavitation rates consistent with those obtained by methods independent of the CNT. In this work, the applicability of NVT seeding to homogeneous cavitation in mixtures is demonstrated, focusing on a partially miscible symmetrical binary Lennard-Jones (LJ) liquid at a temperature within the mixing regime. At the same stretching pressure, cavitation rates are higher in the binary mixture than in the pure liquid due to the lower interfacial free energy of the mixture. Curiously, the cost of creating a bubble is similar in the pure and binary LJ liquids at the same metastability, Δμμspin, with Δμ being the difference in chemical potential between the metastable liquid and coexistence, and Δμspin between the spinodal and coexistence.

Understanding the formation of bubbles from metastable liquids is a problem of high interest in many fields, such as sonochemistry1 or geology.2 Experiments often provide information about the cavitation rates, i.e., about the number of critical bubbles formed per unit of time and per unit of volume, but not about the microscopic mechanism that leads to their formation. With the aim of understanding this process, the Classical Nucleation Theory3,4 (CNT) was proposed a century ago, initially to rationalize the condensation of supersaturated vapors, but later it was also applied to other transitions, including the cavitation of bubbles.5 

Molecular simulations have also been extensively used to study the nucleation of bubbles.6–14 Experiments are often performed at conditions in which the nucleation rates are too low to be computed with the typical simulation times; therefore, special simulation techniques such as umbrella sampling,15–18 forward-flux sampling,6,7,19 and transition path sampling20–22 have been employed. Even though these methods are rigorous, they are often technically challenging and computationally demanding. As an alternative, during the last few years, the seeding method has been gaining popularity, mainly due to the lower computational cost and easier implementation than the preceding methods, with a focus not only on bubble cavitation9–11,13 but also on other transitions, such as crystallization or condensation.23–27 This method consists in inserting a seed of the stable phase within the metastable liquid and searching for the conditions for which the inserted seed is critical, i.e., it stands at the top of the free energy barrier that separates the metastable and stable phases. For one-component systems, this method can be implemented both in the NpT ensemble, in which the critical nucleus is in unstable equilibrium, and in the NVT ensemble, in which it is in stable equilibrium, thus enabling the study of the properties of the critical cluster over longer times.10,11,13 The information about the critical cluster size is then combined with expressions from CNT to calculate the cavitation rates. The application of NVT and NpT seeding to study the cavitation of bubbles in the truncated and force-shifted Lennard-Jones (TFS-LJ) liquid has provided rates consistent with those obtained from other simulation techniques that do not depend on CNT.9–11 

The main goal of this article is to extend the method of NVT seeding to study the nucleation of bubbles in binary systems. The applicability of NVT seeding has already been demonstrated for other transitions besides cavitation, in particular for crystal nucleation, but always for one-component systems.28NpT seeding was not used to study bubble nucleation in multi-component systems, but it has been successfully applied to study the nucleation of crystals in several multi-component systems, such as, for example, the crystallization of NaCl and ice from aqueous solutions29–31 and the nucleation of methane hydrates.32 In this work, the applicability of NVT seeding to multi-component systems will be investigated by studying bubble nucleation in a partially miscible binary symmetrical LJ mixture. This is one of the simplest mixtures one can think of, and this makes it an ideal model system in which to investigate the applicability of NVT seeding before addressing nucleation in more complex binary systems.

Simulation studies of cavitation of bubbles from binary mixtures are rather scarce, and often this phenomenon has been studied theoretically, for example, in the framework of density functional theory or in the framework of CNT.33–36 However, the extension of CNT to multi-component systems has not been straightforward.37–39 Several of the proposed formulations lead to unphysical results, especially for non-ideal mixtures, which has been ascribed to the fact that the effect of the curvature on the interfacial free energy cannot be neglected.39 The nucleation of bubbles in binary mixtures is very rich and involves many different scenarios depending on the interaction parameters. Curious phenomena have been found, such as, for example, an anomalous decrease in the nucleation rate when temperature increases at constant pressure in mixtures of species of different volatility.34 However, as our main goal is to illustrate the applicability of NVT seeding, we will focus on the case of a partially miscible symmetrical LJ mixture at a temperature at which the liquid is miscible. In these conditions, we expect that the cavitation behavior follows a similar mechanism as that in the pure system.

The model system is a binary symmetrical mixture of equal-sized spheres. The interaction between any pair of molecules is described by the truncated and force-shifted Lennard-Jones model (TSF-LJ),6 

UTSF−LJ(r)=ULJ(r)ULJrcrrcULJrc,
(1)

where r is the distance between two molecules, ULJr is the standard 12-6 LJ potential and ULJr is its first derivative, and rc = 2.5σ is the cut-off distance. We have chosen this short-range model to avoid dealing with long range corrections, which are problematic in heterogeneous systems with large variations in density, as the liquid–bubble systems considered in this work. Besides, there are numerous studies of cavitation in pure TSF-LJ that can be compared with the results for the binary symmetrical mixture.6,7,9–11 Here we consider a mixture of spheres of equal size, σAA = σBB, and equal mass, mA = mB.

The phase diagram of partially miscible symmetrical TSF-LJ mixtures (σAA = σBB = σABσ and ϵAA = ϵBBϵ, and ϵAB/ϵ less than but close to 1) has not yet been reported to the best of our knowledge. However, for the standard LJ potential, the global phase diagram (which delimits regions of parameter space with the same phase diagram topology) has been calculated40 from the equation of state proposed by Johnson et al.41 These authors found that symmetrical LJ mixtures with ϵAB/ϵ > 0.75 exhibit a type II diagram according to Bolz’s classification.42 This means that in a pressure vs temperature projection, at low temperature, there is a liquid–liquid–vapor three coexistence line that, as T increases, transforms first into a critical azeotropic line and then into a critical liquid–vapor line. According to Ref. 40, for a mixture with ϵAB/ϵ = 0.8, T* = kBT/ϵ = 0.785 (which is the T* for which cavitation was studied for the pure TSF-LJ liquid in previous studies10–12) falls within the mixing region. As our intention is to study the cavitation of bubbles from a homogeneous binary liquid mixture, we set ϵAB/ϵ = 0.9 so that the liquid does not undergo demixing at the considered T*. In the seeding simulations performed in this work, the concentration of the liquid is set to the molar fraction xA,l = 0.75, where the subscript l is added to differentiate the composition of the liquid from that of the bubble.

Molecular dynamics simulations were performed with the GROMACS 4.6.7 package (double precision).43 To implement the chosen model system, we used the argon parameters, i.e., the diameter of the particles is set to σ = 3.405 Å, the energy parameter to ϵ = 0.996 078 kJ mol−1, and the mass is m = 39.946 u. In the remainder of this article, all the quantities will be presented in LJ reduced units: the volume V* = V/σ3, the density ρ* = (N/V)σ3 (N is the number of particles in the system), the temperature T* = kBT/ϵ, the pressure p* = 3/ϵ, and the time t* = t/τ, with τ=mσ2/ϵ= 2.156 ps. In these units, p* = 1 corresponds to p = 419 bar.

The pressure was controlled using an isotropic Parrinello–Rahman barostat,44 with a relaxation time of t = 2 ps (t* ≈ 0.92) and a compressibility of 7 × 10−5 bar−1. For temperature coupling, velocity rescaling was used with a stochastic term45 and a relaxation time of t = 1 ps (t* ≈ 0.46). The temperature was fixed at T* = 0.785. The equations of motion were integrated using the velocity-Verlet algorithm with a time step of t = 0.005 ps (t* ≈ 0.002).

According to CNT, the nucleation rate, i.e., the number of nuclei from the stable phase that form in the metastable phase per unit of volume and per unit of time at constant pressure, temperature, and composition, can be calculated from3,4,37,39,46,47

J=KexpΔGckBT,
(2)

where kB is the Boltzmann constant, ΔGc is the Gibbs free energy barrier for the formation of a critical bubble, and K is the kinetic prefactor. ΔGc can be computed as

ΔGc=VΔp+γA,
(3)

where V and A are, respectively, the volume and the area of the surface of the critical bubble, γ is the interfacial free energy of the critical bubble, and Δp = pbubblepl is the difference in pressure between the critical bubble and the surrounding liquid. Assuming a spherical critical bubble of radius Rc, and using the Laplace equation

Rc=2γΔp,
(4)

the free energy barrier can also be expressed as

ΔGc=2πΔpRc33.
(5)

Note that this expression for ΔGc is formally identical to that for pure systems.9,46 The main difference with respect to mono-component systems is the procedure used to calculate Δp. In equilibrium, the chemical potential of the liquid and the bubble is the same. In previous studies of pure systems, the pressure inside the bubble was estimated by integrating the equation of state of a bulk vapor to find the pressure at which the chemical potential of the bubble is the same as that of the surrounding liquid.9–11 This calculation is a bit more complex in the mixture because we have to find the pressure and composition at which the chemical potentials of the two components are equal in both phases. It is also important to point out that the radius of the critical cluster Rc that enters in Eq. (5) is the one that satisfies the Laplace equation [Eq. (4)], i.e., the radius at the surface of tension where the value of γ is a minimum with respect to formal changes of the dividing surface.48 The kinetic factor can be calculated from34,46

K=ρlΔpRcπxA,vmA+xB,vmB,
(6)

where mA and mB are the masses of the particles of types A and B, xA,v and xB,v are their vapor molar fractions, and ρl is the number density of the metastable liquid. In our system, the mass of both particle types is the same (mA = mB = m), and the kinetic prefactor simplifies down to the same expression as that used for one-component systems,5 

K=ρlΔpRcmπ.
(7)

In the seeding method,27 a spherical seed from the stable phase is inserted into the equilibrated metastable phase. For one-component systems, simulations of this heterogeneous system can be performed either in the NpT or NVT ensembles.10,28 In pure systems, the critical cluster corresponds to a maximum in G at constant N, p, and T and to a minimum in F at constant N, T, and V (corresponding at the same p).10,49 Therefore, it is possible to have the critical bubble in stable equilibrium with the liquid, allowing one to study its properties over longer periods of time. The main goal of this paper is to show the applicability of NVT seeding for binary systems. In our simulations, the number of particles of each type is kept constant, so strictly speaking, we are working in the NANBVT ensemble or the NVTxA ensemble, with xA being the overall molar fraction of component A. However, for simplicity, we will use the notation NVT (and NpT), keeping in mind that simulations are performed at constant overall composition. The critical bubble size obtained from seeding will be used together with Eq. (5) to estimate the free energy barrier and the nucleation rate using Eq. (2). As the results of seeding are very sensitive to the size of the critical cluster, it is extremely important to choose a criterion to determine the bubble sizes that gives rates consistent with those derived from methods that do not depend on it (Refs. 9 and 10). In this work, we will use brute force simulations to test two criteria to define the bubble sizes.

The first step to study the cavitation in the symmetrical TSF-LJ mixture is to locate the liquid–vapor coexistence point at the chosen concentration, xA,l = 0.75. To facilitate the comparison with cavitation in the pure TSF-LJ liquid, simulations are performed at the same reduced temperature as in previous studies (T* = 0.785).10,11 The coexistence point was evaluated by two independent routes: by direct coexistence simulations50–52 and by calculation of the chemical potential of the two components in the two phases and imposing the condition of phase equilibrium.

1. Direct coexistence simulations

In order to locate the liquid–vapor equilibrium using direct coexistence simulations, we first equilibrate a liquid composed of 2370 particles of type A and 790 particles of type B (i.e., a liquid with a molar fraction xA = 0.75) at T* = 0.785 in the NpT ensemble. The approximate simulation box dimensions are 4.32 × 4.32 × 8.64 nm3. Once the liquid is equilibrated, we introduce an empty space next to the liquid of length about twice its longer dimension (the size of the simulation box is 4.32 × 4.32 × 30 nm3 when the empty space is inserted). This system is then allowed to evolve in the NVT ensemble.

Once the system reaches equilibrium, the coexistence pressure is estimated from the normal component of the pressure tensor, which for the conditions just described is plv*= 0.0334(3). The equilibrium composition of each phase is calculated by measuring the total and partial density profiles along the direction perpendicular to the interface. We obtained that, after equilibration, the concentration of particles of type A in the liquid phase increased up to about xA,l = 0.77(1) and that of the gas phase decreased down to about xA,v = 0.69(1). This enrichment of component A in the liquid phase can be easily understood by energetic arguments. As crossed interactions are lower than those between like particles, and interactions are much more relevant in the liquid phase than in the gas phase (in which particles are on average far away from each other), it is energetically favorable to reduce the number of crossed interactions in the liquid by releasing to the vapor a larger proportion of particles of type B.

As all the cavitation studies presented in this work were performed for a liquid with composition xA,l = 0.75, we repeated the direct coexistence simulations to try to locate the liquid–vapor transition when the liquid has a composition closer to this value. As the concentration of A in the liquid, xA,l, tends to increase when brought into contact with vacuum, we now start the direct coexistence simulations from a liquid phase containing 2323 particles of type A and 837 particles of type B, so that the initial composition of the liquid is xA = 0.735, i.e., somewhat below the target xA,l = 0.75. This system equilibrates at a slightly lower coexistence pressure than before, plv*=0.0341(3), but with the desired liquid equilibrium concentration of A, namely, xA,l = 0.75(1) and xA,v = 0.68(1) in the vapor. Finally, we also checked that the coexistence pressure coincides with that obtained using the equations of state (EOS) of the bulk liquid and vapor phases together with the densities and compositions of the liquid and vapor phases extracted from the bulk region of each phase (i.e., sufficiently away from the interface) in the direct coexistence simulations. Using this procedure, we get pl*=0.033(1) from the EOS of the bulk liquid and pv*=0.034(1) from the EOS of the bulk vapor. Both values are in very good agreement with each other and also with the estimate from the normal component of the pressure tensor.

The interfacial free energy, γ, at coexistence can be calculated from the average pressure tensor,48,53–58

γ=Lzp̄zzp̄xx+p̄yy22,
(8)

where Lz is the length of the simulation box in the direction normal to the interface, p̄zz is the average of the normal component of the pressure, and p̄xx and p̄yy are the averages of the tangential components of the pressure. The values of the average components of the pressure tensor do not depend on any particular formulation of the local pressure tensor.53,57,58 The factor 2 in the denominator of Eq. (8) takes into account that two liquid–vapor interfaces are formed by the periodicity of the simulation box.

The coexistence conditions as well as the interfacial free energy for the TSF-LJ mixture are gathered in Table I. For completeness, we also performed simulations for the pure TSF-LJ system at T* = 0.785, and these results are compared with data from previous studies6,11 in the same Table. The first observation is that the liquid–vapor coexistence and the interfacial free energy obtained by us for the pure system are fully consistent with those found in the literature.6,11 Comparing now the data for the pure and the symmetrical TSF-LJ mixtures at the same temperature, it can be seen that the liquid–vapor coexistence pressure of the symmetrical mixture is higher than that of the pure TSF-LJ system, which is in keeping with the trends observed in Ref. 40. Finally, the interfacial free energy is appreciably lower in the symmetrical TSF-LJ mixture than in the pure system. The same qualitative behavior was found in previous work,59 but a quantitative comparison cannot be made as this work considered a mixture with a different composition (xA = 0.5) and used a different truncation of the interatomic potential.

TABLE I.

Liquid–vapor equilibrium for the pure and the symmetric TSF-LJ mixture at T* = 0.785 obtained by direct coexistence (DC) simulations and by computation of the chemical potentials (CPs). ρl* and ρv* are the overall densities of the liquid and vapor at coexistence, and xA,l and xA,v are the molar fractions of component A in the liquid and in the vapor. xA is the global molar fraction in the whole system, i.e., xA = NA/(NA + NB), where NA and NB are the total number of molecules of type A and type B in the simulation box. plv* is the coexistence pressure, which in DC simulations is obtained from the normal component of the pressure tensor. γ* is the liquid–vapor interfacial free energy and is given in LJ reduced units (γ* = γσ2/ϵ). The DC simulations labeled Big were carried out in a larger system with 7110 particles of type A and 2370 particles of type B, in a box of 4.22 × 4.22 × 60 nm3.

SystemMethodT*plv*ρl*ρv*xA,lxA,vxAγ*
Pure LJ DC 0.785 0.0267(1) 0.665(2) 0.044(2) ⋯ ⋯ ⋯ 0.196(4) 
Pure LJ DC11  0.785 0.0267 ⋯ ⋯ ⋯ ⋯ ⋯ 0.193 
Pure LJ DC6  0.785 0.026 0.668 0.043 ⋯ ⋯ ⋯ 0.204(7) 
Symmetric LJ DC 0.785 0.0334(3) 0.639(4) 0.059(4) 0.77(1) 0.69(1) 0.75 0.145(4) 
Symmetric LJ (big) DC 0.785 0.0335(3) 0.639(4) 0.059(4) 0.76(1) 0.68(1) 0.75 0.145(4) 
Symmetric LJ DC 0.785 0.0336(3) 0.638(4) 0.059(4) 0.75(1) 0.67(1) 0.73 0.142(4) 
Symmetric LJ CP 0.785 0.0341(3) ⋯ ⋯ 0.75 0.67(1) ⋯ ⋯ 
SystemMethodT*plv*ρl*ρv*xA,lxA,vxAγ*
Pure LJ DC 0.785 0.0267(1) 0.665(2) 0.044(2) ⋯ ⋯ ⋯ 0.196(4) 
Pure LJ DC11  0.785 0.0267 ⋯ ⋯ ⋯ ⋯ ⋯ 0.193 
Pure LJ DC6  0.785 0.026 0.668 0.043 ⋯ ⋯ ⋯ 0.204(7) 
Symmetric LJ DC 0.785 0.0334(3) 0.639(4) 0.059(4) 0.77(1) 0.69(1) 0.75 0.145(4) 
Symmetric LJ (big) DC 0.785 0.0335(3) 0.639(4) 0.059(4) 0.76(1) 0.68(1) 0.75 0.145(4) 
Symmetric LJ DC 0.785 0.0336(3) 0.638(4) 0.059(4) 0.75(1) 0.67(1) 0.73 0.142(4) 
Symmetric LJ CP 0.785 0.0341(3) ⋯ ⋯ 0.75 0.67(1) ⋯ ⋯ 

2. Coexistence point from chemical potential calculation

An alternative method to calculate the liquid–vapor coexistence point is to find the thermodynamic conditions at which the chemical potential of the two components is equal in both phases at the same temperature T and at the same pressure p,

μA,l(p,T,xA,l)=μA,v(p,T,xA,v),μB,l(p,T,xB,l)=μB,v(p,T,xB,v),
(9)

where μi,α(p, T, xi,α) and xi,α are, respectively, the chemical potential and molar fraction of component i (i being an A or B component) in phase α (α being the liquid or vapor phase).

The excess chemical potential of components A and B in a liquid mixture with xA,l = 0.75 and in the vapor with xA,v = 0.60, 0.65, and 0.70 as a function of pressure were calculated using the particle test Widom method60,61 as implemented in GROMACS.43 For that aim, we performed NpT simulations of each phase in a cubic simulation box containing 1000 particles. Typically, the chemical potential was averaged over 2000 configurations, collected over a MD simulation run consisting of 2 × 107 steps, after reaching equilibrium. In each configuration, 10 million particle insertion tests were performed. The total chemical potential was obtained by adding to the excess chemical potential the ideal contribution, which for component i is given by μi,id/(kBT) = ln(xiρσ3), where xi is its molar fraction. The thermal de Broglie length is the same in both phases, and its value does not change the chemical potential difference between them. Here it was assigned to the arbitrary value σ.

The chemical potential of the two components in the liquid and vapor phases is shown in Fig. 1(a). As can be seen, the chemical potential of component A in the liquid with xA,l = 0.75 (black lines and symbols) crosses the chemical potential of component A in the vapor with concentrations xA,v = 0.60 (red), 0.65 (blue), and 0.70 (green) at decreasing pressures, and the opposite occurs for the minority component B (i.e., the chemical potential of the liquid crosses that of the gas phase at increasing pressures as xA,v increases). These crossing points fulfill the condition that the chemical potential of the corresponding component is equal in the liquid and vapor phases. The crossing points for each component can be represented as a function of the vapor composition, xA,v, obtaining two lines, one for each component, with opposite slopes [see Fig. 1(b)]. Only at the crossing point between these two lines does the equality of chemical potentials of each component occur at the same composition of the vapor (xa,v) and at the same pressure. From this, we obtain that, at T* = 0.785, the symmetrical TSF-LJ liquid with composition xA,l = 0.75 coexists with the vapor phase with composition xA,v = 0.67(1) at plv*=0.0341(3). As can be seen in Table I, these results are in full agreement with those obtained by the direct coexistence route.

FIG. 1.

(a) Chemical potential vs pressure for component A (circles) and B (squares) in a symmetrical TSF-LJ mixture with ϵAB* = 0.9 at T* = 0.785. Black symbols show data for the liquid with xA,l = 0.75 (dashed lines are linear fits), and red, blue, and green symbols show data for the vapor with xA,v = 0.6, xA,v = 0.65, and xA,v = 0.7, respectively. (b) Estimation of the coexistence point. The points at which the chemical potentials of each component in the liquid and vapor at different concentrations are equal are marked as diamonds (purple for A and pink for B) in (a) and presented jointly in (b). The crossing of these two lines gives the coexistence point. (c) Calculation of the pressure inside the bubble for system 10. The pressure and the chemical potential of the surrounding liquid are shown as vertical and horizontal gray lines in panel (a), respectively. Purple triangles mark the points at which the chemical potential of A in the vapor is the same as in the surrounding liquid, whereas red ones are the equivalent points for B. The crossing point of the two lines gives the equilibrium bubble–liquid conditions.

FIG. 1.

(a) Chemical potential vs pressure for component A (circles) and B (squares) in a symmetrical TSF-LJ mixture with ϵAB* = 0.9 at T* = 0.785. Black symbols show data for the liquid with xA,l = 0.75 (dashed lines are linear fits), and red, blue, and green symbols show data for the vapor with xA,v = 0.6, xA,v = 0.65, and xA,v = 0.7, respectively. (b) Estimation of the coexistence point. The points at which the chemical potentials of each component in the liquid and vapor at different concentrations are equal are marked as diamonds (purple for A and pink for B) in (a) and presented jointly in (b). The crossing of these two lines gives the coexistence point. (c) Calculation of the pressure inside the bubble for system 10. The pressure and the chemical potential of the surrounding liquid are shown as vertical and horizontal gray lines in panel (a), respectively. Purple triangles mark the points at which the chemical potential of A in the vapor is the same as in the surrounding liquid, whereas red ones are the equivalent points for B. The crossing point of the two lines gives the equilibrium bubble–liquid conditions.

Close modal

As mentioned earlier, one of the difficulties of studying nucleation by seeding is that this method depends critically on having a good criterion to estimate the size of the critical nucleus. Typically, this problem has been sorted out by comparing the nucleation rates obtained with seeding with those coming from more rigorous simulation techniques, such as umbrella sampling, forward-flux sampling, or brute force simulations.9,10,27 Given that we are not aware of previous estimates of the bubble cavitation rate for the symmetrical TSF-LJ mixture, we started by studying cavitation by brute force simulations. At high metastabilities, the free energy barrier for nucleation decreases, and spontaneous cavitation becomes accessible within the typical length and time scales of simulations.

For that aim, we performed NpT simulations of a liquid with composition xA,l = 0.75, containing 18 960 particles of type A and 6320 of type B. At these conditions, we were able to observe spontaneous cavitation only at negative pressures. For each considered pressure, we ran 10 independent simulations. Cavitation events can be easily detected by an abrupt increase in density. The cavitation rate can be estimated from these simulations as

J=1tV,
(10)

where ⟨t⟩ is the average time for observing spontaneous cavitation, which is calculated simply as the average of the time for cavitation in the 10 independent simulations performed for each pressure, and V is the average volume of the simulation box before the cavitation event. We checked that 10 simulations are enough to obtain an accurate estimate of the nucleation rate with an error significantly lower than one order of magnitude by comparing the rate obtained from 20 independent simulations with those obtained from only 10. The cavitation rates obtained by brute force simulations are given in Table II.

TABLE II.

Cavitation rate for the symmetrical TSF-LJ mixture with ϵAB* = 0.9 at xA,l = 0.75 and T* = 0.785 as obtained from brute force simulations. The rates are given in LJ reduced units (J* = 3τ). For the three lower pressures, averages were performed over 20 independent simulations, whereas 10 simulations were used for the two higher pressures.

p*t*⟩V*⟩J*
−0.0300 4.3(4) × 102 42 380(20) 5.5(5) × 10−8 
−0.0297 2.6(7) × 103 42 255(4) 9(2) × 10−9 
−0.0235 1.7(4) × 104 41 993(3) 1.4(4) × 10−9 
−0.0200 2.9(6) × 105 41 746.8(9) 8(2) × 10−11 
−0.0177 4(1) × 106 41 599(2) 6(2) × 10−12 
p*t*⟩V*⟩J*
−0.0300 4.3(4) × 102 42 380(20) 5.5(5) × 10−8 
−0.0297 2.6(7) × 103 42 255(4) 9(2) × 10−9 
−0.0235 1.7(4) × 104 41 993(3) 1.4(4) × 10−9 
−0.0200 2.9(6) × 105 41 746.8(9) 8(2) × 10−11 
−0.0177 4(1) × 106 41 599(2) 6(2) × 10−12 

As seen in Sec. II B, the estimation of the cavitation rate requires to calculate the free energy barrier ΔGc and the kinetic prefactor K. The cavitation free energy barrier can be obtained from the size of the critical cluster, Rc, and the pressure difference between the bubble and the surrounding liquid, Δp [Eq. (5)]. In the next subsections, we will see how to estimate these properties. It is important to recall that the expression of the free energy barrier ΔGc, Eq. (5), was derived under the assumption that the composition of the metastable phase remains constant during the bubble nucleation. This is true in the thermodynamic limit, where the size of the critical bubble is negligible compared to the size of the system (so that the formation of the bubble causes only an infinitesimal change in the composition of the external liquid phase). In previous studies of the nucleation of droplets and crystals from binary mixtures, simulations were performed in the isothermal-isobaric semigrand ensemble (in which p, T, N, and the chemical potential difference between the two species are kept constant).18,47,62 As will be shown later, in the liquid–bubble simulations performed in this work, the composition of the liquid remains almost unchanged. The reason is that, even though the composition of the bubble is different from that of the metastable liquid, the number of particles in the bubble is so small compared to that in the surrounding liquid that it barely changes its composition. Therefore, the use of the NVT ensemble, keeping the overall composition constant, is justified in this case.

1. Estimation of the radius of the critical cluster, Rc

The protocol used to construct initial configurations of the liquid in coexistence with bubbles is as follows: first, we equilibrate the liquid with a composition xA,l = 0.75 by performing NpT simulations at T* = 0.785. In order to be able to accommodate bubbles of different sizes, we considered two system sizes: a smaller one with N = 10 000 particles and another one with N = 130 000 that equilibrated during 1 × 106 MD steps. Bubbles of different sizes are then generated by removing all the particles from a spherical region defined in the equilibrated liquid simulation box. The sizes of the bubbles were chosen so that the ratio between the volumes of bubble and simulation box was within 5%–15% (see Table V). Previous work suggests that within these limits seeding simulations are not significantly affected by finite size effects.9 

TABLE III.

Details of the systems for which bubbles in equilibrium with the liquid phase were obtained in NVT simulations. Each system is labeled with a numerical index for further reference. Besides the total number of particles of each type (NA and NB), the pressure of the surrounding liquid pl* as calculated from the virial is given. Δp* is the difference in pressure between the bubble and the surrounding liquid. The pressure inside the bubble was estimated using the thermodynamic route, i.e., by finding the pressure and composition for which the chemical potentials of the two components are the same in the liquid and vapor phases. ρl* and ρv* are the overall densities of the liquid and the bubble, and xA,v is the composition of the vapor, as obtained from the radial overall and partial density profiles. The composition of the liquid xA,l remains very close to 0.75, with maximum deviations of 0.003 from this value.

IndexNANBpl* ± 0.0001Δp* ± 0.0002ρl* ± 0.0002ρv* ± 0.0003xA,v ± 0.001
6 706 2 239 −0.0156 0.0448 0.6110 0.0485 0.676 
18 173 6 074 −0.0083 0.0383 0.6155 0.0503 0.673 
17 636 5 852 −0.0068 0.0370 0.6163 0.0507 0.673 
17 187 5 727 −0.0022 0.0328 0.6191 0.0518 0.672 
17 918 5 973 −0.0019 0.0326 0.6192 0.0519 0.672 
17 240 5 738 −0.0004 0.0312 0.6201 0.0522 0.672 
17 605 5 875 0.0014 0.0296 0.6211 0.0527 0.671 
16 788 5 583 0.0033 0.0279 0.6222 0.0531 0.671 
17 319 5 771 0.0034 0.0278 0.6223 0.0532 0.671 
90 900 30 338 0.0031 0.0285 0.6221 0.0542 0.671 
10 90 108 30 063 0.0082 0.0235 0.6249 0.0543 0.671 
11 89 191 29 781 0.0107 0.0212 0.6263 0.0549 0.670 
12 88 274 29 438 0.0123 0.0197 0.6272 0.0553 0.670 
13 87 290 29 072 0.0136 0.0185 0.6279 0.0555 0.670 
14 86 219 28 712 0.0147 0.0175 0.6285 0.0558 0.670 
15 85 044 28 330 0.0157 0.0166 0.6290 0.0560 0.671 
16 82 545 27 468 0.0172 0.0153 0.6297 0.0563 0.671 
17 79 572 26 512 0.0184 0.0142 0.6303 0.0566 0.671 
IndexNANBpl* ± 0.0001Δp* ± 0.0002ρl* ± 0.0002ρv* ± 0.0003xA,v ± 0.001
6 706 2 239 −0.0156 0.0448 0.6110 0.0485 0.676 
18 173 6 074 −0.0083 0.0383 0.6155 0.0503 0.673 
17 636 5 852 −0.0068 0.0370 0.6163 0.0507 0.673 
17 187 5 727 −0.0022 0.0328 0.6191 0.0518 0.672 
17 918 5 973 −0.0019 0.0326 0.6192 0.0519 0.672 
17 240 5 738 −0.0004 0.0312 0.6201 0.0522 0.672 
17 605 5 875 0.0014 0.0296 0.6211 0.0527 0.671 
16 788 5 583 0.0033 0.0279 0.6222 0.0531 0.671 
17 319 5 771 0.0034 0.0278 0.6223 0.0532 0.671 
90 900 30 338 0.0031 0.0285 0.6221 0.0542 0.671 
10 90 108 30 063 0.0082 0.0235 0.6249 0.0543 0.671 
11 89 191 29 781 0.0107 0.0212 0.6263 0.0549 0.670 
12 88 274 29 438 0.0123 0.0197 0.6272 0.0553 0.670 
13 87 290 29 072 0.0136 0.0185 0.6279 0.0555 0.670 
14 86 219 28 712 0.0147 0.0175 0.6285 0.0558 0.670 
15 85 044 28 330 0.0157 0.0166 0.6290 0.0560 0.671 
16 82 545 27 468 0.0172 0.0153 0.6297 0.0563 0.671 
17 79 572 26 512 0.0184 0.0142 0.6303 0.0566 0.671 

These systems containing the artificially constructed bubbles are then allowed to evolve in the NVT ensemble for 2 × 106 MD steps. Initially, bubbles grow or shrink until an equilibrium with the surrounding liquid is established. Note that not all the bubbles will equilibrate, depending on the choice of the total number of particles, size of the simulation box, and total density.63 In those systems in which the bubble remains stable after equilibration, i.e., it does not completely dissolve in the timescale of our simulation, we estimate their average radius in simulations consisting of 100–400 × 106 MD steps.

As mentioned earlier, in seeding, nucleation rates can change significantly depending on the criterion used to define the cluster size.9,29 The radius that enters a CNT is that given by the surface of tension, i.e., the surface on which the surface tension acts.48,63–65 However, the location of that interface is not known.66,67 For that reason, in previous implementations of the seeding method, the radius of the critical nucleus is estimated using a criterion that yields nucleation rates consistent with those obtained from rigorous simulations that do not depend on the critical cluster size, such as umbrella sampling or brute force simulations.9,11,31

Here we employed two criteria used in previous studies to define the bubble size.9,11 Later, we will compare the cavitation rates derived from these two alternative criteria to see which one gives results in closer agreement with those from brute force simulations. In the first criterion, the radius of the bubble is taken as the equidensity radius,9 which is defined as the distance at which the density is equal to the average of the liquid and gas densities measured sufficiently away from the interface (where radial densities remain constant). The equidensity radius is estimated from the radial overall density profile measured from the center of the bubble, as shown in Fig. 2(a). Following previous work,10 the center of the bubble in each instantaneous configuration is located by searching for the positions of the minima in the density profiles calculated along the x, y, and z directions using a bin width of about 0.7σ. The radial density profile is calculated using a bin of width 0.15σ in the radial direction and is typically averaged over 1000 independent configurations stored every 20 000 steps, so that the typical length of the simulations used for these calculations is 20 × 106 MD steps. The radial overall density profile is then fitted to the following sigmoid function:

ρ(r)=ρv,dp+ρl,dp2+ρl,dpρv,dp2tanhrRed/α,
(11)

that has four fitting parameters: ρv,dp and ρl,dp provide estimates of the densities of the vapor and liquid phases (the subscript dp stands for density profile), α is a measure of the width of the interfacial region, and Red is the distance at which the density adopts the average liquid and vapor densities. The values for ρv,dp and ρl,dp obtained from this fit are coherent (within their statistical uncertainty) with the values inferred from averaging the density profiles of the vapor and the liquid regions, respectively, in Fig. 2. Notice that the error in the estimation of the vapor density ρv,dp is higher than that of the liquid density ρv,dp since there are only a few particles in the bubble, resulting in large fluctuations in the density.

FIG. 2.

(a) Radial overall density profile and (b) radial concentration profile for the system labeled with the index 10 in Table III. R* = R/σ is the radial distance to the bubble center.

FIG. 2.

(a) Radial overall density profile and (b) radial concentration profile for the system labeled with the index 10 in Table III. R* = R/σ is the radial distance to the bubble center.

Close modal

Besides the radial density profile, we also calculated the radial concentration profile by dividing the density of particles of type A by the total density for each radial distance. A sample profile is shown in Fig. 2(b). The concentration of particles of type A is higher in the liquid phase than in the vapor phase, similarly to what was observed in the direct coexistence simulations. In this case, as the volume of the gas phase is relatively small, the composition of the liquid phase remains fairly close to the starting composition, namely, xA,l ≃ 0.75. However, the molar fraction of A in the vapor phase is again slightly lower than that in the surrounding liquid (xA,v ≈ 0.66). As explained earlier, this is due to the fact that the strength of the crossed interactions is 10% lower than that of like-interactions. In general, in all the NVT seeding simulations performed in this work, the concentrations of the bubbles are within xa,v ≈ 0.670–0.673, and that of the surrounding liquid is typically within xA,l ≈ 0.750–0.753.

Alternatively, one can also resort to the equimolar Gibbs dividing surface as a criterion to define the bubble radius.9,66 The Gibbs dividing surface is obtained by assuming that the densities of the liquid and gas phases are constant up to the dividing surface with radius RG. In multi-component systems, a Gibbs dividing surface can be defined for each component. Here we define the Gibbs radius using the majority component A,

NA=VbubbleρA,v+(VVbubble)ρA,l,
(12)

where Vbubble and V are, respectively, the volume of the bubble and of the simulation box. Substituting the volume of the bubble by

Vbubble=43πRG3,
(13)

and reorganizing we obtain that the Gibbs radius can be estimated as

RG=34πNAL3ρA,lρA,vρA,l1/3,
(14)

where L is the edge of the simulation box. As mentioned earlier, we could have also defined the Gibbs radii using the minoritary component, B. These two Gibbs radii are only identical when the composition in the liquid and vapor phases are the same. As can be seen in Tables I and III, in our system, this is neither true at coexistence nor in states where nucleation is studied. As discussed earlier, as a consequence of the reduced crossed interactions, the minoritary component tends to go to the vapor phase, where interactions are less relevant. From Eq. (12), it is easy to see that this leads to RG,B > RG,A. However, as can be seen in Table IV, this effect is mild in our system. RG,B is only about 2% bigger than RG,A. Consequently, nucleation rates calculated with RG,B are 0–6 orders of magnitude smaller than those calculated with RG,A, but, as will be shown later, this difference is within the uncertainty of our calculations. Since it seems more sensible to use the majority component, we have chosen to use RG,A in our calculations. In what follows, for simplicity in the notation, this radius will be denoted RG.

TABLE IV.

Gibbs radii calculated using component A (RG,A), component B (RG,B), and the average (RG,AB), as well as the corresponding nucleation rates (JG,A*,JG,B*,JG,AB*) obtained with these radii. In the remainder of the article, RG,A is designed as RG and JG,A* as JG* for simplicity in the notation.

IndexRG,A*RG,B*RG,AB*log(JG,A*)log(JG,B*)log(JG,AB*)
6.21 6.22 6.21 −13 −13 −13 
7.34 7.25 7.30 −18 −18 −18 
7.37 7.73 7.55 −18 −21 −19 
8.36 8.48 8.42 −23 −24 −23 
8.43 8.54 8.48 −23 −24 −24 
8.80 8.97 8.89 −25 −27 −26 
9.30 9.38 9.34 −28 −29 −29 
9.82 9.73 9.77 −32 −31 −32 
9.94 10.08 10.01 −32 −34 −33 
9.88 10.09 9.98 −32 −34 −33 
10 11.72 11.77 11.75 −45 −45 −45 
11 13.06 13.07 13.06 −55 −56 −55 
12 14.10 14.26 14.18 −65 −67 −66 
13 15.03 15.30 15.17 −74 −78 −76 
14 15.92 16.20 16.06 −82 −87 −85 
15 16.77 17.04 16.91 −92 −96 −94 
16 18.32 18.66 18.49 −110 −116 −113 
17 19.86 20.16 20.01 −129 −135 −132 
IndexRG,A*RG,B*RG,AB*log(JG,A*)log(JG,B*)log(JG,AB*)
6.21 6.22 6.21 −13 −13 −13 
7.34 7.25 7.30 −18 −18 −18 
7.37 7.73 7.55 −18 −21 −19 
8.36 8.48 8.42 −23 −24 −23 
8.43 8.54 8.48 −23 −24 −24 
8.80 8.97 8.89 −25 −27 −26 
9.30 9.38 9.34 −28 −29 −29 
9.82 9.73 9.77 −32 −31 −32 
9.94 10.08 10.01 −32 −34 −33 
9.88 10.09 9.98 −32 −34 −33 
10 11.72 11.77 11.75 −45 −45 −45 
11 13.06 13.07 13.06 −55 −56 −55 
12 14.10 14.26 14.18 −65 −67 −66 
13 15.03 15.30 15.17 −74 −78 −76 
14 15.92 16.20 16.06 −82 −87 −85 
15 16.77 17.04 16.91 −92 −96 −94 
16 18.32 18.66 18.49 −110 −116 −113 
17 19.86 20.16 20.01 −129 −135 −132 

As can be seen in Fig. 3(a) and Tables V and VI, the Gibbs radius is systematically larger than the equidensity radius. Taking into account Eqs. (2) and (5), this means that the Gibbs radius will lead to a higher cavitation free energy barrier and, consequently, to a lower cavitation rate than the equidensity radius. The same trend was also observed in previous seeding studies of cavitation9–11 [see Fig. 3(a)] and droplet nucleation11 in the pure TSF-LJ system. Gibbs radii are not provided in Refs. 10 and 11, but they can be estimated from the data reported in Ref. 11. The results for the pure TSF-LJ system using the Gibbs radii are also included in Figs. 3 and 5.

FIG. 3.

(a) Radius of the critical bubble as a function of the stretching pressure (measured as pl*plv*, with plv*=0.0341(3)). Solid black squares were obtained in this work with the equidensity criterion and solid red squares with the Gibbs criterion. Open symbols show the radii of bubbles for the pure TSF-LJ system reported in the literature10,11 (the Gibbs radii were calculated in this work with the data provided in those references). (b) Bubble–liquid pressure difference as a function of pl*plv*. Black dots were obtained by measuring the density inside the bubble and using an EOS to get the pressure of the bubble, whereas red dots were obtained by looking for the conditions at which the chemical potentials of components A and B in the bubble and surrounding liquid are equal. For comparison, data from the literature on the pressure difference between the bubble and the surrounding liquid for the pure TSF-LJ are also shown with blue symbols (calculated by imposing that the chemical potential in the bubble and surrounding liquid is equal).10,11 Δp* shows a fairly linear behavior, and is fitted to Δp*=4.17871050.90275(pl*plv*) (c) Density of the surrounding liquid, ρl*, as a function of pl*plv*. Black crosses were obtained by performing NpT simulations of a liquid with xA,l = 0.75. This EOS was fitted to a second degree polynomial (ρl*=0.63835+0.46246(pl*plv*)1.7164(pl*plv*)2). For comparison, light blue symbols show the density of the liquid resulting from the fit of the radial density profile to a sigmoid function vs the virial pressure in the seeding simulations.

FIG. 3.

(a) Radius of the critical bubble as a function of the stretching pressure (measured as pl*plv*, with plv*=0.0341(3)). Solid black squares were obtained in this work with the equidensity criterion and solid red squares with the Gibbs criterion. Open symbols show the radii of bubbles for the pure TSF-LJ system reported in the literature10,11 (the Gibbs radii were calculated in this work with the data provided in those references). (b) Bubble–liquid pressure difference as a function of pl*plv*. Black dots were obtained by measuring the density inside the bubble and using an EOS to get the pressure of the bubble, whereas red dots were obtained by looking for the conditions at which the chemical potentials of components A and B in the bubble and surrounding liquid are equal. For comparison, data from the literature on the pressure difference between the bubble and the surrounding liquid for the pure TSF-LJ are also shown with blue symbols (calculated by imposing that the chemical potential in the bubble and surrounding liquid is equal).10,11 Δp* shows a fairly linear behavior, and is fitted to Δp*=4.17871050.90275(pl*plv*) (c) Density of the surrounding liquid, ρl*, as a function of pl*plv*. Black crosses were obtained by performing NpT simulations of a liquid with xA,l = 0.75. This EOS was fitted to a second degree polynomial (ρl*=0.63835+0.46246(pl*plv*)1.7164(pl*plv*)2). For comparison, light blue symbols show the density of the liquid resulting from the fit of the radial density profile to a sigmoid function vs the virial pressure in the seeding simulations.

Close modal
TABLE V.

Cavitation rates, Jed*, obtained with seeding use the equidensity criterion to estimate the critical radii along with the interfacial free energy γed*, free energy barrier, ΔGed/(kBT), and kinetic factor, Ked*. The fraction of system volume occupied by the bubble, Vbubble/Vbox, is also given.

IndexRed*γed*ΔGed/(kBT)Ked*log(Jed*)Vbubble/Vbox
5.2 0.116 16.4 0.165 −8 0.04 
6.5 0.125 28.3 0.174 −13 0.03 
6.7 0.124 29.8 0.173 −14 0.03 
7.8 0.128 41.1 0.177 −19 0.05 
7.8 0.128 41.8 0.177 −19 0.05 
8.2 0.128 46.4 0.178 −21 0.06 
8.8 0.130 53.2 0.179 −24 0.07 
9.4 0.131 61.6 0.180 −27 0.09 
9.5 0.131 62.5 0.180 −28 0.09 
9.4 0.134 63.1 0.182 −28 0.02 
10 11.3 0.133 91.4 0.182 −40 0.03 
11 12.7 0.135 115.5 0.184 −51 0.04 
12 13.9 0.136 139.5 0.185 −61 0.06 
13 14.8 0.137 158.7 0.186 −70 0.07 
14 15.7 0.137 181.2 0.186 −79 0.08 
15 16.6 0.138 201.9 0.187 −88 0.10 
16 18.1 0.138 241.6 0.188 −106 0.13 
17 19.7 0.139 288.1 0.189 −126 0.16 
IndexRed*γed*ΔGed/(kBT)Ked*log(Jed*)Vbubble/Vbox
5.2 0.116 16.4 0.165 −8 0.04 
6.5 0.125 28.3 0.174 −13 0.03 
6.7 0.124 29.8 0.173 −14 0.03 
7.8 0.128 41.1 0.177 −19 0.05 
7.8 0.128 41.8 0.177 −19 0.05 
8.2 0.128 46.4 0.178 −21 0.06 
8.8 0.130 53.2 0.179 −24 0.07 
9.4 0.131 61.6 0.180 −27 0.09 
9.5 0.131 62.5 0.180 −28 0.09 
9.4 0.134 63.1 0.182 −28 0.02 
10 11.3 0.133 91.4 0.182 −40 0.03 
11 12.7 0.135 115.5 0.184 −51 0.04 
12 13.9 0.136 139.5 0.185 −61 0.06 
13 14.8 0.137 158.7 0.186 −70 0.07 
14 15.7 0.137 181.2 0.186 −79 0.08 
15 16.6 0.138 201.9 0.187 −88 0.10 
16 18.1 0.138 241.6 0.188 −106 0.13 
17 19.7 0.139 288.1 0.189 −126 0.16 
TABLE VI.

Cavitation rates, JG*, obtained with seeding using the Gibbs criterion to estimate the critical radii, along with the interfacial free energy γG*, free energy barrier, ΔGG/(kBT), and kinetic factor, KG*.

IndexRG*γG*ΔGG/(kBT)KG*log(JG*)
6.2 0.139 28.8 0.181 −13 
7.3 0.140 39.0 0.183 −18 
7.4 0.136 45.5 0.186 −21 
8.4 0.137 53.4 0.185 −24 
8.4 0.137 54.1 0.185 −24 
8.8 0.137 60.2 0.186 −27 
9.3 0.138 65.1 0.185 −29 
9.9 0.138 76.3 0.187 −34 
9.9 0.138 75.9 0.186 −34 
9.8 0.140 69.9 0.185 −31 
10 11.7 0.138 102.1 0.186 −45 
11 13.1 0.138 126.3 0.186 −56 
12 14.1 0.139 152.2 0.188 −67 
13 15.0 0.139 177.1 0.189 −78 
14 15.9 0.139 198.5 0.189 −87 
15 16.8 0.139 219.6 0.189 −96 
16 18.3 0.140 264.9 0.191 −116 
17 19.9 0.141 309.7 0.191 −135 
IndexRG*γG*ΔGG/(kBT)KG*log(JG*)
6.2 0.139 28.8 0.181 −13 
7.3 0.140 39.0 0.183 −18 
7.4 0.136 45.5 0.186 −21 
8.4 0.137 53.4 0.185 −24 
8.4 0.137 54.1 0.185 −24 
8.8 0.137 60.2 0.186 −27 
9.3 0.138 65.1 0.185 −29 
9.9 0.138 76.3 0.187 −34 
9.9 0.138 75.9 0.186 −34 
9.8 0.140 69.9 0.185 −31 
10 11.7 0.138 102.1 0.186 −45 
11 13.1 0.138 126.3 0.186 −56 
12 14.1 0.139 152.2 0.188 −67 
13 15.0 0.139 177.1 0.189 −78 
14 15.9 0.139 198.5 0.189 −87 
15 16.8 0.139 219.6 0.189 −96 
16 18.3 0.140 264.9 0.191 −116 
17 19.9 0.141 309.7 0.191 −135 

Following previous work on one-component systems,10,11,28 we checked what happens to the stabilized bubbles in the NVT ensemble when simulated in the NpT ensemble. For that aim, we launched 36 simulations in the NpT ensemble starting from bubble–liquid configurations taken along the trajectory of the system with index 9 in Table III at the corresponding pl*(pl*=0.0031). The outcome of those simulations is shown in Fig. 4, where the evolution of the total density of the system in all the NpT simulations is monitored. As can be seen, in about half of the trajectories, the bubble dissolves, and in the other half, the bubble grows irreversibly to the vapor phase, indicating that the cluster is indeed critical.

FIG. 4.

Evolution of the density of the system with time for 36 NpT simulations starting from different configurations collected along the NVT trajectory of the bubble–liquid system with index 9 in Table III. As can be seen, in about half of the trajectories, the bubble dissolves, and in the other half, it leads to the vapor phase.

FIG. 4.

Evolution of the density of the system with time for 36 NpT simulations starting from different configurations collected along the NVT trajectory of the bubble–liquid system with index 9 in Table III. As can be seen, in about half of the trajectories, the bubble dissolves, and in the other half, it leads to the vapor phase.

Close modal

One might intuitively expect that by virtue of the Gibbs phase rule, for binary systems, it would be possible to have a critical cluster in stable equilibrium at constant p and T. Indeed, this has been exploited to not only have a planar interface of a binary system at equilibrium in the NpT ensemble68,69 but also to stabilize spherical crystal seeds of NaCl68 in coexistence with a salt solution and of methane hydrate in coexistence with water with some methane dissolved.70 However, as we have just seen, the critical bubbles in the binary symmetrical TSF-LJ mixture cannot be stabilized in the NpT ensemble. Therefore, why is it possible to stabilize the critical seed in the NpT ensemble for some multi-component systems but not for others? Further work is needed to clarify this issue, but it is important to highlight that in all cases in which stable clusters in the NpT ensemble were achieved, they had a constant composition, given by the stoichiometry of the crystals (the NaCl rock salt and the methane hydrate), whereas in our case, the composition of the bubbles is not fixed. Small fluctuations in the composition of the clusters change the free energy barrier for bubble formation, making the cluster precritical or postcritical at the specified p and T and, therefore, leading to their dissolution or growth. As for the one-component systems, we found that the bubbles only remain stable in simulations at constant volume.

2. Pressure difference between the bubble and the surrounding liquid, Δp*

Once we have estimated the bubble size, we proceed to calculate the difference in pressure between the bubble and the surrounding liquid, Δp*=pbubble*pl*, which is the other quantity needed to estimate ΔGc [Eq. (5)]. In this work, we followed two different routes to achieve this aim. In both approaches, the pressure of the liquid that surrounds the bubble, pl*, was taken as the virial pressure in the NVT simulations, as provided by GROMACS.43 Note that as we are dealing with a heterogeneous system in which the bubble is in coexistence with the liquid, the pressure is not uniform in the system. However, the average of the pressure tensor in the whole system coincides with that of the external phase, i.e., with that of the liquid.71 The pressure of the liquid in all the seeding simulations is reported in Table III. The difference between the two methods relies on how to estimate the pressure inside the bubble, pbubble*.

In the first approach, the pressure inside the bubble is taken as the mechanical pressure, which can be estimated from the density and composition inside the bubble together with an EOS for the bulk vapor phase obtained from simulations [pbubble*=p*(ρv*,xA,v)]. We checked that the EOS of the bulk vapor shows a mild dependence on composition within the range xA,v = 0.670–0.673. This is rather convenient, because even though the composition of the bubbles is not exactly the same in all the seeding simulations, it remains within these boundaries, and the use of the same bulk EOS for all the bubbles is justified. Here, the EOS was calculated by performing NpT MD simulations of a vapor with composition xA,v = 0.672 and sweeping pressures within p* = 0.002–0.033.

In the second approach, we use the thermodynamic route to estimate pbubble*. The equilibrium established between the bubble and the surrounding liquid implies that the chemical potential of each component is the same in the two phases. The pressure inside the bubble is the one that fulfills this condition, which can be obtained from the chemical potentials calculated in Sec. III A 2 and shown in Fig. 1(a). The procedure is as follows: first, we estimate the chemical potentials of components A and B in a liquid with xA,l = 0.75 at pressure pl*, namely, μA,l(pl*,xA,l) and μB,l(pl*,xA,l). The gray vertical line in Fig. 1(a) marks the pressure of the surrounding liquid, pl*, for the system with index 10 (Table III), and the two horizontal gray lines mark the chemical potentials of components A and B in such a liquid. Second, we look for the crossing points of these horizontal lines with the chemical potential lines of the corresponding component in the vapor phase at several vapor compositions (xA,v = 0.60, 0.65, and 0.70). At these crossing points, the chemical potential of each component is the same in the liquid and in the vapor at concentration xA,v, i.e., they fulfill μA,l(pl*,xA,l)=μA,v(pv*,xA,v) (purple symbols) and μB,l(pl*,xA,l)=μB,v(pv*,xA,v) (red symbols). These crossing points are represented and fitted to straight lines in Fig. 1(c). Only at the crossing point between these two lines does the equality of chemical potentials of the two components in both phases occur at the same composition of the vapor phase, xA,v, and at the same vapor pressure pv*, giving the equilibrium pressure inside the bubble pbubble* and its composition. For the system labeled with index 10 and shown in Fig. 1(c), the pressure and composition of the bubble in equilibrium with the liquid are pbubble*=0.0316 and xA,v = 0.671, which compare well with the values pbubble*=0.0309 and xA,v = 0.656 obtained with the first described method, i.e., using estimates of the density and composition inside the bubble together with the simulated EOS of the bulk vapor.

The results of Δp* obtained from the mechanical and thermodynamic routes are plotted jointly in Fig. 3(b) for comparison. As can be seen, both methods give results that are consistent with each other, but the chemical potential route provides less noisy data than the EOS route. Consistent with previous findings in the pure system, this means that the mechanical pressure is equivalent to the thermodynamic pressure.11,72 Note that this does not always hold, as has been shown in a recent study of hard-sphere crystal nucleation.28 The higher statistical uncertainty in the mechanical pressure is expected, given the difficulty of accurately estimating the density inside the bubble. The number of vapor particles is very small, and fluctuations in the number of particles lead to relatively large changes in the density. Besides, only the results obtained from the thermodynamic route guarantee that the chemical potentials are the same in the bubble and the surrounding liquid. For those reasons, the values of Δp* obtained from the chemical potential route are the ones that will be used in the remainder of the article (these data are tabulated in Table III). Curiously, as can be seen in Fig. 3(b), the difference in pressure between the bubble and surrounding liquid is very similar in the pure and the symmetrical binary TSF-LJ mixture when compared at the same stretching pl*plv*, despite the bubbles being much bigger in the pure system.

3. Density of the metastable liquid ρl*

The density of the metastable liquid, ρl*, needed to calculate the kinetic factor K, Eq. (7), is shown in Fig. 3(b). Again, we tested two possible ways to estimate this density. In the first one, we used the EOS of the bulk liquid with xA,l = 0.75 to obtain the density at the corresponding liquid pressure pl*, obtained from the virial pressure in seeding simulations. In the second approach, the density is directly measured from the seeding simulations using the sigmoid fit. As can be seen in Fig. 3(c), the agreement between both sets of data is fairly good, except for the point at the lower pressure and some of the points at the highest pressures. It is also worth pointing out that the seeding points that exhibit the largest departures from the EOS are precisely those with a higher Vbubble/Vbox ratio (see Table V), which might be indicative that these small deviations arise from finite size effects. In any case, differences are always lower than 0.3%, which can be taken as an indication that finite size effects are not very significant in any of our seeding simulations.

Due to the reduced cross interactions in the mixture, the liquid density is lower than in the pure system. Therefore, the kinetic factor K is about 15% lower in the mixture11 both because of the smaller bubble size and the reduced density of the liquid as compared to those of the pure system at the same stretching.

4. Cavitation free energy barrier and rate

Now that we have estimated the radius of the critical cluster (Rc), the difference of pressure inside and outside the bubble (Δp*), and the density of the metastable liquid, we can proceed to calculate the free energy barrier for cavitation using Eq. (5). The results as a function of stretching relative to the coexistence (pl*plv*) are shown in Fig. 5(a). As the Gibbs radii are larger than the equidensity radii, the free energy barrier calculated with the Gibbs dividing surface is larger, with the difference increasing with stretching as the critical bubble gets smaller. However, both sets of data support the conclusion that the free energy barrier is significantly lower in the mixture than in the pure TSF-LJ system when compared at the same stretching relative to coexistence. Interestingly, a similar trend was observed in a simulation study on the effect of composition on the nucleation of droplets from a LJ binary mixture.18 

FIG. 5.

(a) Cavitation free energy barrier and (b) decimal logarithm of the cavitation rate as a function of pl*plv*. (c) Interfacial free energy as a function of the inverse of the radius of the critical bubble and (d) as a function of pl*plv*. In all panels, black squares show results obtained with the equidensity radius, and red squares with the Gibbs radius. Solid symbols show data for the mixture studied in this work, and open symbols display the results for the pure TSF-LJ system reported in Ref. 11. The blue triangles in (b) show rates obtained from brute force simulations. In (c) and (d), the interfacial free energy at coexistence for the pure and mixture systems is represented with a pink diamond. In (d), γ* was fitted to a straight line γ*=0.14089+0.079549(pl*plv*), which is used to obtain the CNT-like fit shown by the red dotted–dashed line in panel (b). The red dotted lines in panel (b) delimit the uncertainty in the rate.

FIG. 5.

(a) Cavitation free energy barrier and (b) decimal logarithm of the cavitation rate as a function of pl*plv*. (c) Interfacial free energy as a function of the inverse of the radius of the critical bubble and (d) as a function of pl*plv*. In all panels, black squares show results obtained with the equidensity radius, and red squares with the Gibbs radius. Solid symbols show data for the mixture studied in this work, and open symbols display the results for the pure TSF-LJ system reported in Ref. 11. The blue triangles in (b) show rates obtained from brute force simulations. In (c) and (d), the interfacial free energy at coexistence for the pure and mixture systems is represented with a pink diamond. In (d), γ* was fitted to a straight line γ*=0.14089+0.079549(pl*plv*), which is used to obtain the CNT-like fit shown by the red dotted–dashed line in panel (b). The red dotted lines in panel (b) delimit the uncertainty in the rate.

Close modal

Next, we present the cavitation rates obtained from Eq. (2) using the data summarized in Tables V and VI. As can be seen in Fig. 5(b), the small differences in the free energy barrier obtained with the two defined radii are also visible in the nucleation rates. Rates are lower using the Gibbs criteria. Both sets of data give consistent results with brute force simulations, although those derived from the Gibbs radii show better agreement. However, this should be taken with care, because for the pure TSF-LJ system, RG also coincides better with brute-force simulations,9 but at lower metastabilities Red provides rates consistent with umbrella sampling.11 

5. Interfacial free energy

The interfacial free energy can be obtained from the Laplace equation [Eq. (4)]. The values calculated from the equidensity and Gibbs criteria are shown as a function of the inverse radius and as a function of pl*plv* in Figs. 5(c) and 5(d), respectively. As can be seen, the extrapolation of both sets of data to an infinite radius gives coherent results with our calculation of the interfacial free energy at coexistence, but the values of γ obtained from RG exhibit a considerably smaller slope than those derived from Red, in this latter case predicting that the cost of creating a bubble–liquid interface is rather insensitive to the curvature of the bubble, adopting values close to those of the planar interface from very small sizes.

The values of γ are compared with those of the pure TSF-LJ system11 in Fig. 5(d). For the pure TSF-LJ system, γ was estimated using Red,11 because it gave rates coherent with those from umbrella sampling.9 Consistent with the results in the thermodynamic limit, the interfacial free energy is systematically larger for the pure system than for the symmetrical mixture, with a slope intermediate between those derived from Red and RG in the mixture. The lower interfacial free energy of the mixture as compared to that of the pure system explains why cavitation rates are significantly larger in the mixture.

The cavitation rate was fitted using CNT expressions as described in Ref. 10. ΔGc is a function of RG and Δp [Eq. (5)], and the kinetic factor K depends on RG, Δp, and ρl. As can be seen in Fig. 3, the dependence of Δp and ρl with plplv follows a smooth behavior, but that of RG is more complex. Therefore, it is convenient to rewrite ΔG and A in terms of γ, which shows a nearly linear behavior with plplv [Fig. 5(d)]. Using the fits of Δp, γ, and ρl as a function of plplv provided in the captions to Figs. 3 and 5, we obtain J as a function of plplv, which is plotted as a dashed red line in Fig. 5(b). The error in the rate comes from the uncertainty in the estimation of RG, Δp and ρl, with the larger error coming from the critical radius. Assuming that the error in the critical radius is of about 0.3σ and neglecting the uncertainty in the other quantities, we can estimate the error in the rate, which is delimited by the red dotted lines in Fig. 5(b). As can be seen, this fit and its error capture well the variation of the rates with the stretching. If instead of using the fit of γ we use the capillarity approximation (i.e., we assume that adopts the same value as the planar interface), the rate is very similar to that obtained using the rates calculated from RG, which is expected as we have seen that γ was rather insensitive to the bubble curvature. Differences are much larger for the data derived from Red because γ exhibits a larger dependence on curvature.

A theoretical work73 has shown that in the pure LJ system, the nucleation free energy barrier for the formation of bubbles is the same for all temperatures if compared at the same scaled metastability, defined as

Xm=μnuclμcoexμspinμcoex,
(15)

where μnucl, μcoex, and μspin are the chemical potentials of the liquid at nucleation, coexistence, and spinodal. With this definition, Xm varies between 0 at coexistence and 1 at the spinodal. Interestingly, in subsequent work,11 it was shown that the free energy barrier for droplet and bubble formation as a function of Xm also collapses into a single curve. In that regard, it seems that the free energy barrier for nucleation is the same for bubbles and droplets if compared at the same degree of scaled metastability.

In view of this, it is interesting to compare the free energy barrier for the LJ mixture as a function of Xm with that of the pure system. For a two component system, the driving force for nucleation depends on the chemical potential of the two components. Therefore, we define

μ=xAμA+xBμB,
(16)

and this is the quantity that will be evaluated at coexistence, at the nucleation conditions, and at the mechanical spinodal to estimate Xm. For a pure component system, the spinodal is unique and is defined by the condition (∂p/∂ρ)T = 0. For a binary mixture, one can define two spinodals, the mechanical one given by (∂p/∂ρ)T,x = 0 and the diffusional one defined as (∂μ/∂x)p,T = 0.74,75 The location of these two spinodals can be quite different.74 Since for pure components we found a correlation between the free energy barrier and Xm via the mechanical definition of the spinodal, we shall explore the possibility of a similar correlation for mixtures when also using the mechanical definition of the spinodal. We estimated its location using an approximate method, as described in Ref.11. In particular, we performed brute force simulations as described in Sec. III B at decreasing pressures. The spinodal is located at the highest pressure at which cavitation occurs immediately at the beginning of the simulations, without an induction period. We estimate that the spinodal is located at p* ≈ −0.05. The free energy barriers for bubble formation obtained either with Red or RG are plotted as a function of Xm in Fig. 6, together with the results for the pure TSF-LJ system taken from previous work.11 As can be seen, the free energy barrier for the pure and the binary systems adopt quite similar values when compared at the same scaled metastability. This means that the cost of creating a critical bubble does not depend much on the interactions (at least for symmetrical mixtures with low deviations of the reduced crossed interactions from unity as the ones considered), and it is mostly determined by the scaled metastability.

FIG. 6.

Cavitation free energy barrier for the TSF-LJ symmetrical mixture with xA,l = 0.75 as a function of the scaled metastability. Free energy barriers obtained with the equidensity radius (Red) are shown as black squares, and those obtained with the Gibbs radius (RG) are shown as red squares. The solid symbols show the results obtained for the mixture studied in this work, and the open symbols show the results for the pure TSF-LJ system at the same temperature (T* = 0.785) taken from Ref. 11.

FIG. 6.

Cavitation free energy barrier for the TSF-LJ symmetrical mixture with xA,l = 0.75 as a function of the scaled metastability. Free energy barriers obtained with the equidensity radius (Red) are shown as black squares, and those obtained with the Gibbs radius (RG) are shown as red squares. The solid symbols show the results obtained for the mixture studied in this work, and the open symbols show the results for the pure TSF-LJ system at the same temperature (T* = 0.785) taken from Ref. 11.

Close modal

In this work, we have applied NVT seeding to calculate bubble nucleation rates in a partially miscible binary symmetrical TSF-LJ mixture. The study was conducted at a temperature that falls within the mixing region of the system, for which there are previously available data for the pure TSF-LJ model. One aspect that requires special care when implementing seeding is the choice of an appropriate criterion to determine the bubble size. Here, we used both the equidensity radius and the equimolar (or Gibbs dividing surface) radius. Consistently with previous work,9,11 the equidensity radii are systematically smaller than the Gibbs radii, leading to a lower free energy barrier and a higher nucleation rate. However, what is the correct value of the critical bubble size? As mentioned before, the correct radius is the one corresponding to the surface of tension, and it is the one that enters into CNT expressions. Therefore, we compared the nucleation rates derived from the equidensity and the Gibbs surface radii with those obtained from brute force simulations. Gibbs radii give rates closer to those obtained by brute force simulations. The same trend was found for bubble nucleation in pure TSF-LJ, but at lower metastabilities, rates calculated with the equidensity radii are closer to those obtained by forward-flux sampling.9 For droplet nucleation in the same system, equidensity nucleation rates are also in closer agreement with umbrella sampling results.11 However, a recent study using enhancing sampling methods and transition state theory found results that are in agreement with seeding but obtained from Gibbs radii and concluded that forward-flux sampling gives systematically higher rates.6,7,12 Therefore, we can conclude that for the LJ model at high metastabilities, close to the region where brute force simulations are feasible, the Gibbs dividing surface radii give the correct nucleation rates. However, at low metastabilities, where special simulation techniques are needed, the situation is not yet clear, as there are conflicting results using rigorous techniques.12 Further work is needed to clarify why different methods provide different rates at low metastabilities.

In any case, the differences in the cavitation behavior between the pure and symmetrical mixtures are larger than the differences between seeding rates using different criteria for determining the bubble size. Therefore, conclusions about the difference in the cavitation behavior in the two systems can be drawn. At the same stretching, measured as the pressure difference for coexistence, the critical bubble is smaller in the binary system than in the pure system. The reason is that the interfacial free energy is about 25% lower in the mixture than in the pure system, i.e., the free energy cost of creating a bubble is lower in the mixture. This is, at least in part, related to the fact that the critical temperature is lower in the mixture than in the pure system.40,59 Curiously, when comparing both systems at the same degree of scaled metastability, X = (μliqμcoex)/(μspinμcoex), the free energy barrier is similar in both systems. This is an interesting result, as it was found in previous work that free energy barriers at different temperatures for bubble and droplet nucleation merged into a single curve when represented as a function of the scaled metastability.11,73 Our results suggest that this conclusion can also be extended to mixtures, at least for small deviations of the crossed interactions from those between like interactions, as the one considered here.

In the future work, it would be interesting to extend the NVT method to other systems and transitions, for example, to study droplet formation in strongly non ideal mixtures or crystallization from a binary mixture. In principle, NVT should also be applicable in this case; however, some difficulties might arise. For example, if important fractionation effects occur during nucleation (i.e., the composition of the nucleated and the parent phase are too different), if one wants to keep the composition constant, large simulation boxes are needed or that other ensembles, such as the semigrand canonical ensemble, need to be employed. In NVT seeding, the critical cluster is in equilibrium with the metastable phase, so that the system evolves until the chemical potentials of the two components are the same in the two phases. But if one of the phases is solid, diffusion might be too slow to reach the equilibrium.

It would also be interesting to explore alternative routes to calculate the free energy cost of creating a bubble in the metastable liquid. A possibly viable method is to introduce an external potential that creates a bubble in the liquid and design a reversible path that allows the calculation of the free energy by thermodynamic integration.

This project has been funded by Grant Nos. PID2019-105898GB-C21 and PID2020-115722GB-C21 from the Agencia Estatal de Investigación. C. P. L. thanks the Ministerio de Educación y Formación Profesional (MEFP) for a predoctoral Formación Profesorado Universitario under Grant No. FPU18/03326 and also the Ayuntamiento de Madrid for a Residencia de Estudiantes grant. The authors acknowledge the computer resources and technical assistance provided by the Red Española de Supercomputación (RES).

The authors have no conflicts to disclose.

Cintia P. Lamas: Data curation (equal); Formal analysis (equal); Methodology (equal); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal). Eduardo Sanz: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal). Carlos Vega: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (supporting); Writing – review & editing (equal). Eva G. Noya: Conceptualization (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (lead); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
K. S.
Suslick
, “
Sonochemistry
,”
Science
247
,
1439
1445
(
1990
).
2.
C. E.
Brenen
,
Cavitation and Bubble Dynamics
(
Oxford University Press
,
New York
,
1995
).
3.
M.
Volmer
and
A.
Weber
, “
Keimbildung inübersättigten gebilden
,”
Z. Phys. Chem.
119
(
1
),
277
301
(
1926
).
4.
R.
Becker
and
W.
Döring
, “
The kinetic treatment of nuclear formation in supersaturated vapors
,”
Ann. Phys.
24
,
719
752
(
1935
).
5.
M.
Blander
and
J. L.
Katz
, “
Bubble nucleation in liquids
,”
AlChE J.
21
,
833
(
1975
).
6.
Z.-J.
Wang
,
C.
Valeriani
, and
D.
Frenkel
, “
Homogeneous bubble nucleation driven by local hot spots: A molecular dynamics study
,”
J. Phys. Chem. B
113
(
12
),
3776
3784
(
2009
).
7.
S. L.
Meadley
and
F. A.
Escobedo
, “
Thermodynamics and kinetics of bubble nucleation: Simulation methodology
,”
J. Chem. Phys.
137
,
074109
(
2012
).
8.
J.
Diemand
,
R.
Angélil
,
K. K.
Tanaka
, and
H.
Tanaka
, “
Direct simulation of bubble nucleation: Agreement with classical nucleation theory and no local hot spots
,”
Phys. Rev. E
90
,
052407
(
2014
).
9.
P.
Rosales-Pelaez
,
M. I.
Garcia-Cid
,
C.
Valeriani
,
C.
Vega
, and
E.
Sanz
, “
Seeding approach to bubble nucleation in superheated Lennard-Jones fluids
,”
Phys. Rev. E
100
(
5
),
052609
(
2019
).
10.
P.
Rosales-Pelaez
,
I.
Sanchez-Burgos
,
C.
Valeriani
,
C.
Vega
, and
E.
Sanz
, “
Seeding approach to nucleation in the NVT ensemble: The case of bubble cavitation in overstretched Lennard Jones fluids
,”
Phys. Rev. E
101
(
2
),
022611
(
2020
).
11.
I.
Sanchez-Burgos
,
P. M.
de Hijes
,
P.
Rosales-Pelaez
,
C.
Vega
, and
E.
Sanz
, “
Equivalence between condensation and boiling in a Lennard-Jones fluid
,”
Phys. Rev. E
102
(
6
),
062609
(
2020
).
12.
K. M.
Bal
and
E. C.
Neyts
, “
Extending and validating bubble nucleation rate predictions in a Lennard-Jones fluid with enhanced sampling methods and transition state theory
,”
J. Chem. Phys.
157
,
184113
(
2022
).
13.
C. P.
Lamas
,
E. G.
Noya
,
C.
Vega
, and
E.
Sanz
, “
Estimation of bubble cavitation rates in a symmetrical Lennard-Jones mixture by NVT Seeding simulations
,”
J. Chem. Phys.
(
2023
) (submitted).
14.
K. R.
Protsenko
and
V. G.
Baidakov
, “
Classical nucleation theory and molecular dynamics simulation: Cavitation
,”
Phys. Fluids
35
(
1
),
014111
(
2023
).
15.
G. M.
Torrie
and
J. P.
Valleau
, “
Nonphysical sampling distributions in Monte Carlo free-energy estimation: Umbrella sampling
,”
J. Comput. Phys.
23
,
187
(
1977
).
16.
L.
Filion
,
M.
Hermes
,
R.
Ni
, and
M.
Dijkstra
, “
Crystal nucleation of hard spheres using molecular dynamics, umbrella sampling, and forward flux sampling: A comparison of simulation techniques
,”
J. Chem. Phys.
133
(
24
),
244115
(
2010
).
17.
S.
Auer
and
D.
Frenkel
, “
Prediction of absolute crystal nucleation rate in hard-sphere colloids
,”
Nature
409
,
1020
(
2001
).
18.
P. R.
ten Wolde
and
D.
Frenkel
, “
Numerical study of gas–liquid nucleation in partially miscible binary mixtures
,”
J. Chem. Phys.
109
,
9919
,
1998
.
19.
R. J.
Allen
,
C.
Valeriani
, and
P.
Rein ten Wolde
, “
Forward flux sampling for rare events simulations
,”
J. Phys.: Condens. Matter
21
,
463102
(
2009
).
20.
C.
Dellago
,
P. G.
Bolhuis
,
F. S.
Csajka
, and
D.
Chandler
, “
Transition path sampling and the calculation of rate constants
,”
J. Chem. Phys.
108
,
1964
(
1998
).
21.
P. G.
Bolhuis
,
D.
Chandler
,
C.
Dellago
, and
P. L.
Geissler
, “
Transition path sampling: Throwing ropes over rough mountain passes, in the dark
,”
Annu. Rev. Phys. Chem.
53
,
291
(
2002
).
22.
E.
Schöll-Paschinger
and
C.
Dellago
, “
Demixing of a binary symmetric mixture studied by transition path sampling
,”
J. Chem. Phys.
133
,
104505
(
2010
).
23.
X.-M.
Bai
and
M.
Li
, “
Test of classical nucleation theory via molecular dynamics simulations
,”
J. Chem. Phys.
122
,
224510
(
2005
).
24.
X.-M.
Bai
and
M.
Li
, “
Calculation of solid-liquid interfacial free energy: A classical nucleation theory based approach
,”
J. Chem. Phys.
124
,
124707
(
2006
).
25.
R. G.
Pereyra
,
I.
Szleifer
, and
M. A.
Carignano
, “
Temperature dependence of ice critical nucleus size
,”
J. Chem. Phys.
135
,
034508
(
2011
).
26.
B. C.
Knott
,
V.
Molinero
,
M. F.
Doherty
, and
B.
Peters
, “
Homogeneous nucleation of methane hydrates: Unrealistic under realistic conditions
,”
J. Am. Chem. Soc.
134
,
19544
19547
(
2012
).
27.
J. R.
Espinosa
,
C.
Vega
,
C.
Valeriani
, and
E.
Sanz
, “
Seeding approach to crystal nucleation
,”
J. Chem. Phys.
144
(
3
),
034501
(
2016
).
28.
P.
Montero de Hijes
,
J. R.
Espinosa
,
V.
Bianco
,
E.
Sanz
, and
C.
Vega
, “
Interfacial free energy and Tolman length of curved liquid-solid interfaces from equilibrium studies
,”
J. Phys. Chem. C
124
,
8795
8805
(
2020
).
29.
N. E. R.
Zimmermann
,
B.
Vorselaars
,
J. R.
Espinosa
,
D.
Quigley
,
W. R.
Smith
,
E.
Sanz
,
C.
Vega
, and
B.
Peters
, “
NaCl nucleation from brine in seeded simulations: Sources of uncertainty in rate estimates
,”
J. Chem. Phys.
148
(
22
),
222838
(
2018
).
30.
G. D.
Soria
,
J. R.
Espinosa
,
J.
Ramirez
,
C.
Valeriani
,
C.
Vega
, and
E.
Sanz
, “
A simulation study of homogeneous ice nucleation in supercooled salty water
,”
J. Chem. Phys.
148
(
22
),
222811
(
2018
).
31.
C. P.
Lamas
,
J. R.
Espinosa
,
M. M.
Conde
,
J.
Ramírez
,
P. M.
de Hijes
,
E.
Noya
,
C.
Vega
, and
E.
Sanz
, “
Homogeneous nucleation of NaCl in supersaturated solutions
,”
Phys. Chem. Chem. Phys.
23
,
26843
26852
(
2021
).
32.
J.
Grabowska
,
S.
Blazquez
,
E.
Sanz
,
E. G.
Noya
,
I. M.
Zeron
,
J.
Algaba
,
J.
Míguez
,
F. J.
Blas
, and
C.
Vega
, “
Homogeneous nucleation rate of methane hydrate formation under experimental conditions from seeding simulations
,”
J. Chem. Phys.
158
,
114505
(
2023
).
33.
S. P.
Protsenko
,
V. G.
Baidakov
,
A. S.
Teterin
, and
E. R.
Zhdanov
, “
Computer simulation of nucleation in a gas-saturated liquid
,”
J. Chem. Phys.
126
,
094502
(
2007
).
34.
V.
Talanquer
and
D. W.
Oxtoby
, “
Nucleation of bubbles in binary mixtures
,”
J. Chem. Phys.
102
,
2156
(
1995
).
35.
V.
Talanquer
,
C.
Cunningham
, and
D. W.
Oxtoby
, “
Bubble nucleation in binary mixtures: A semiempirical approach
,”
J. Chem. Phys.
114
,
6759
(
2001
).
36.
V. G.
Baidakov
,
S. P.
Protsenko
, and
V. M.
Bryukhanov
, “
Bubble nucleation in a Lennard-Jones binary liquid mixture
,”
Chem. Phys. Lett.
663
,
57
60
(
2016
).
37.
P. G.
Debenedetti
,
Metastable Liquids: Concepts and Principles
(
Princeton University Press
,
1996
).
38.
D. W.
Oxtoby
and
D.
Kashchiev
, “
A general relation between the nucleation work and the size of the nucleus in multicomponent nucleation
,”
J. Chem. Phys.
100
,
7665
(
1994
).
39.
A.
Laaksonen
,
R.
McGraw
, and
H.
Vehkamäki
, “
Liquid-drop formalism and free energy surfaces in binary homogeneous nucleation theory
,”
J. Chem. Phys.
111
,
2019
(
1999
).
40.
J. M.
Garrido
,
M. M.
Piñeiro
,
A.
Mejía
, and
F. J.
Blas
, “
Understanding the interfacial behavior in isopycnic Lennard-Jones mixtures by computer simulations
,”
Phys. Chem. Chem. Phys.
18
(
2
),
1114
1124
(
2016
).
41.
J. K.
Johnson
,
J. A.
Zollweg
, and
K. E.
Gubbins
, “
The Lennard-Jones equation of state revisited
,”
Mol. Phys.
78
,
591
618
(
1993
).
42.
A.
Bolz
,
U. K.
Deiters
,
C. J.
Peters
, and
T. W.
de Loos
, “
Nomenclature for phase diagrams with particular reference to vapour–liquid and liquid–liquid equilibria
,”
Pure Appl. Chem.
70
,
2233
(
1998
).
43.
B.
Hess
,
C.
Kutzner
,
D.
Van Der Spoel
, and
E.
Lindahl
, “
GROMACS 4: Algorithms for highly efficient, load-balanced, and scalable molecular simulation
,”
J. Chem. Theory Comput.
4
(
3
),
435
447
(
2008
).
44.
M.
Parrinello
and
A.
Rahman
, “
Polymorphic transitions in single crystals: A new molecular dynamics method
,”
J. Appl. Phys.
52
(
12
),
7182
7190
(
1981
).
45.
G.
Bussi
,
T.
Zykova-Timan
, and
M.
Parrinello
, “
Isothermal-isobaric molecular dynamics using stochastic velocity rescaling
,”
J. Chem. Phys.
130
(
7
),
074101
(
2009
).
46.
B. S.
Holden
and
J. L.
Katz
, “
The homogeneous nucleation of bubbles in superheated binary liquid mixtures
,”
AlChE J.
24
,
260
(
1978
).
47.
R.
Ni
,
F.
Smallenburg
,
L.
Filion
, and
M.
Dijkstra
, “
Crystal nucleation in binary hard-sphere mixtures: The effect of the order parameter on the cluster composition
,”
Mol. Phys.
109
,
1213
1227
(
2011
).
48.
J.
Rowlinson
and
B.
Widom
,
Molecular Theory of Capillarity
(
Clarendon
,
Oxford
,
1982
).
49.
J. F.
Lutsko
, “
How crystals form: A theory of nucleation pathways
,”
Sci. Adv.
5
,
eaav7399
(
2019
).
50.
A. J. C.
Ladd
and
L. V.
Woodcock
, “
Triple-point coexistence properties of the Lennard-Jones system
,”
Chem. Phys. Lett.
51
(
1
),
155
159
(
1977
).
51.
A. J. C.
Ladd
and
L. V.
Woodcock
, “
Interfacial and co-existence properties of the Lennard-Jones system at the triple point
,”
Mol. Phys.
36
(
2
),
619611
(
1978
).
52.
E. G.
Noya
,
C.
Vega
, and
E.
de Miguel
, “
Determination of the melting point of hard spheres by direct coexistence simulations
,”
J. Chem. Phys.
128
,
154507
(
2008
).
53.
J. P. R. B.
Walton
,
D. J.
Tildesley
,
J. S.
Rowlinson
, and
J. R.
Henderson
, “
The pressure tensor at the planar surface of a liquid
,”
Mol. Phys.
48
(
6
),
1357
1368
(
1983
).
54.
J. K.
Lee
,
J. A.
Barker
, and
G. M.
Pound
, “
Surface structure and surface tension: Perturbation theory and Monte Carlo simulation
,”
J. Chem. Phys.
60
,
1976
(
1974
).
55.
G. J.
Gloor
,
G.
Jackson
,
F. J.
Blas
, and
E.
de Miguel
, “
Test-area simulation method for the direct determination of the interfacial tension of systems with continuous or discontinuous potentials
,”
J. Chem. Phys.
123
,
134703
(
2005
).
56.
C.
Vega
and
E.
de Miguel
, “
Surface tension of the most popular models of water by using the test-area simulation method
,”
J. Chem. Phys.
126
,
154707
(
2005
).
57.
C.
Braga
,
E. R.
Smith
,
A.
Nold
,
D. N.
Sibley
, and
S.
Kalliadasis
, “
The pressure tensor across a liquid-vapour interface
,”
J. Chem. Phys.
149
(
4
),
044705
(
2018
).
58.
K.
Shi
,
E. R.
Smith
,
E. E.
Santiso
, and
K. E.
Gubbins
, “
A perspective on the microscopic pressure (stress) tensor: History, current understanding, and future challenges
,”
J. Chem. Phys.
158
(
4
),
040901
(
2023
).
59.
E.
Diaz-Herrera
,
G.
Ramirez-Santiago
, and
J. A.
Moreno-Razo
, “
Phase and interfacial behaviour of partially miscible symmetric Lennard-Jones binary mixture
,”
J. Chem. Phys.
123
,
184507
(
2005
).
60.
B.
Widom
, “
Some topics in the theory of fluids
,”
J. Chem. Phys.
39
,
2802
2812
(
1963
).
61.
D.
Frenkel
and
B.
Smit
,
Understanding Molecular Simulation
(
Academic Press
,
London
,
1996
).
62.
S.
Punnathanam
and
P. A.
Monson
, “
Crystal nucleation in binary hard sphere mixtures: A Monte Carlo simulation study
,”
J. Chem. Phys.
125
,
024508
(
2006
).
63.
P.
Montero de Hijes
and
C.
Vega
, “
On the thermodynamics of curved interfaces and the nucleation of hard spheres in a finite system
,”
J. Chem. Phys.
156
,
014505
(
2022
).
64.
J. W.
Gibbs
, “
On the equilibrium of heterogeneous substances
,”
Trans. Connect. Acad. Sci.
3
,
108
248
(
1876
).
65.
R. C.
Tolman
, “
The effect of droplet size on surface tension
,”
J. Chem. Phys.
17
(
3
),
333
337
(
1949
).
66.
B. J.
Block
,
S. K.
Das
,
M.
Oettel
,
P.
Virnau
, and
K.
Binder
, “
Curvature dependence of surface free energy of liquid drops and bubbles: A simulation study
,”
J. Chem. Phys.
133
,
154702
(
2020
).
67.
K.
Binder
,
B. J.
Block
,
P.
Virnau
, and
A.
Tröster
, “
Beyond the van der Waals loop: What can be learned from simulating Lennard-Jones fluids inside the region of phase coexistence
,”
Am. J. Phys.
80
,
1099
(
2012
).
68.
J. R.
Espinosa
,
J. M.
Young
,
H.
Jiang
,
D.
Gupta
,
C.
Vega
,
E.
Sanz
,
P. G.
Debenedetti
, and
A. Z.
Panagiotopoulos
, “
On the calculation of solubilities via direct coexistence simulations: Investigation of NaCl aqueous solutions and Lennard-Jones binary mixtures
,”
J. Chem. Phys.
145
(
15
),
154111
(
2016
).
69.
C. P.
Lamas
,
C.
Vega
, and
E. G.
Noya
, “
Freezing point depression of salt aqueous solutions using the Madrid-2019 model
,”
J. Chem. Phys.
156
,
134503
(
2022
).
70.
J.
Grabowska
,
S.
Blazquez
,
E.
Sanz
,
I. M.
Zerón
,
J.
Algaba
,
J. M.
Míguez
,
F. J.
Blas
, and
C.
Vega
, “
Solubility of methane in water: Some useful results for hydrate nucleation
,”
J. Phys. Chem. B
126
,
8553
8570
(
2022
).
71.
P.
Montero de Hijes
,
K.
Shi
,
E. G.
Noya
,
E. E.
Santiso
,
K. E.
Gubbins
,
E.
Sanz
, and
C.
Vega
, “
The Young–Laplace equation for a solid–liquid interface
,”
J. Chem. Phys.
153
,
191102
(
2020
).
72.
P. R.
ten Wolde
and
D.
Frenkel
, “
Computer simulation study of gas–liquid nucleation in a Lennard-Jones system
,”
J. Chem. Phys.
109
,
9901
(
1998
).
73.
V. K.
Shen
and
P. G.
Debenedetti
, “
Density-functional study of homogeneous bubble nucleation in the stretched Lennard-Jones fluid
,”
J. Chem. Phys.
114
,
4149
(
2000
).
74.
V. G.
Baidakov
,
S. P.
Protsenko
, and
V. M.
Bryukhanov
, “
Nucleation and relaxation processes in weak solutions: Molecular dynamics simulation
,”
Mol. Simul.
48
,
1051
1061
(
2022
).
75.
H.
Jiang
,
P. G.
Debenedetti
, and
A. Z.
Panagiotopoulos
, “
Nucleation in aqueous NaCl solutions shifts from 1-step to 2-step mechanism on crossing the spinodal
,”
J. Chem. Phys.
150
(
12
),
124502
(
2019
).