<<

University of Nebraska - Lincoln DigitalCommons@University of Nebraska - Lincoln

Xiao Cheng Zeng Publications Published Research - Department of Chemistry

10-1-2006

Homogeneous nucleation at high supersaturation and heterogeneous nucleation on microscopic wettable particles: A hybrid thermodynamic/density-functional theory

T. V. Bykov McMurry University

Xiao Cheng Zeng University of Nebraska-Lincoln, [email protected]

Follow this and additional works at: https://digitalcommons.unl.edu/chemzeng

Part of the Chemistry Commons

Bykov, T. V. and Zeng, Xiao Cheng, "Homogeneous nucleation at high supersaturation and heterogeneous nucleation on microscopic wettable particles: A hybrid thermodynamic/density-functional theory" (2006). Xiao Cheng Zeng Publications. 3. https://digitalcommons.unl.edu/chemzeng/3

This Article is brought to you for free and open access by the Published Research - Department of Chemistry at DigitalCommons@University of Nebraska - Lincoln. It has been accepted for inclusion in Xiao Cheng Zeng Publications by an authorized administrator of DigitalCommons@University of Nebraska - Lincoln. THE JOURNAL OF CHEMICAL PHYSICS 125, 144515 ͑2006͒

Homogeneous nucleation at high supersaturation and heterogeneous nucleation on microscopic wettable particles: A hybrid thermodynamic/density-functional theory T. V. Bykov and X. C. Zeng Department of Physics, McMurry University, Abilene, Texas 79697 and Department of Chemistry, University of Nebraska-Lincoln, Lincoln, Nebraska 68588 ͑Received 7 July 2006; accepted 1 September 2006; published online 11 October 2006͒

Homogeneous nucleation at high supersaturation of vapor and heterogeneous nucleation on microscopic wettable particles are studied on the basis of Lennard-Jones model system. A hybrid classical thermodynamics and density-functional theory ͑DFT͒ approach is undertaken to treat the nucleation problems. Local-density approximation and weighted-density approximation are employed within the framework of DFT. Special attention is given to the disjoining of small liquid droplets, which is dependent on the thickness of wetting film and radius of the wettable particle. Different contributions to the disjoining pressure are examined using both analytical estimations and numerical DFT calculation. It is shown that van der Waals interaction results in negative contribution to the disjoining pressure. The presence of wettable particles results in positive contribution to the disjoining pressure, which plays the key role in the heterogeneous nucleation. Several definitions of the of liquid droplets are discussed. Curvature dependence of the surface tension of small liquid droplets is computed. The important characteristics of nucleation, including the formation free energy of the droplet and nucleation barrier height, are obtained. © 2006 American Institute of Physics. ͓DOI: 10.1063/1.2357937͔

I. INTRODUCTION rierless if the chemical potential of the supersaturated vapor 1 exceeds the spinodal value. However, the classical nucleation This paper is closely related to an earlier work by us, in 2 which we considered heterogeneous nucleation on mesos- theory cannot predict the threshold chemical potential in the copic wettable particles. The theoretical method developed case of homogeneous nucleation. Several attempts have been there was a hybrid thermodynamics/density-functional made recently to modify this theory in order to account for theory approach within the weighted-density approximation. the existence of the threshold. For instance, the scaling rela- tions for critical nucleus were introduced by McGraw and It is well known that heterogeneous nucleation is the most 3 4 common mechanism to initiate a first-order Laaksonen and further tested by Talanquer. The McGraw- from metastable supersaturated vapor to bulk liquid. With Laaksonen phenomenological theory gives significant im- wettable particles the nucleation proceeds through formation provement over the classical nucleation theory even though of a liquidlike film on the surface of the particles. Although a there is still a tendency to overestimate the nucleation barrier variety of particles can act as heterogeneous nucleation cen- at high supersaturation and underestimate the height of ters, in Ref. 1 we only considered mesoscopic wettable par- nucleation barrier at low supersaturation. The scaling law ticles as a possible source to initiate heterogeneous nucle- found further support in thermodynamically and kinetically ation. In this work we extend the approach proposed in the consistent extended modified liquid drop ͑EMLD͒ model, 5 earlier study to deal with microscopic wettable particles. In proposed by Reguera and Reiss. This theory gives even bet- addition, we studied homogeneous nucleation as a limiting ter agreement with simulation/experimental results since case of heterogeneous nucleation as the size of wettable par- EMLD takes into account fluctuations existing in extremely ticles approaches to zero. small clusters. Thermodynamically this means that small ho- In Ref. 1 we showed that the behavior of disjoining pres- mogeneous systems are already heterogeneous and addi- sure and chemical potential of liquid films formed on the tional contributions have to be introduced into thermody- surface of mesoscopic wettable particles is responsible for namic potential to study these systems. Another purely the heterogeneous nucleation mechanism. One of the key thermodynamic approach has been pursued by Kashchiev,6 characteristics of heterogeneous nucleation is the existence in which a phenomenological approximation has been made of a threshold supersaturation. If the particle was placed in a to account for the existence of the spinodal. We shall use the supersaturated vapor with supersaturation higher than this idea of nonhomogeneity of small droplets and explore some threshold value then the barrierless nucleation would occur. physical reasons behind aforementioned phenomenological This is why the initiation of homogeneous nucleation always approximations. requires higher value of supersaturation than that for hetero- The similarity in heterogeneous and homogeneous geneous nucleation. nucleation mechanisms becomes closer as the size of the Note that even homogeneous nucleation can become bar- impurity particle becomes smaller. When the impurity par-

0021-9606/2006/125͑14͒/144515/18/$23.00125, 144515-1 © 2006 American Institute of Physics

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-2 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒ ticle and the molecules of the nucleating substance are about scribe highly nonuniform solid-liquid interfacial systems. the same size, one could view the nucleation as a binary Here we used the same DFT/WDA approach but for a spheri- nucleation rather than heterogeneous nucleation. Here we cal droplet system rather than for the planar system. The only consider very dilute concentration of impurity particles local-density approximation ͑LDA͒, although cannot qualita- and that the size of the particle is a few times larger than that tively describe the density profile of fluid near the solid sur- of molecules of nucleating substance. As such, we still can face, is very effective to deal with homogeneous treat the nucleation process as heterogeneous nucleation. We nucleation.12,14,23 However, to our knowledge, WDA has not also used the assumption made in Ref. 1, that is, the droplet been applied to deal with homogeneous nucleation. Using formation proceeds via forming a thin fluid film on the sur- the DFT/WDA approach in spherical coordinates, not only face of the impurity particle ͑“complete wetting”͒. The thin we can treat the problem of heterogeneous nucleation on film then grows into a larger spherical droplet with the im- microscopic particles but also we can compare DFT/WDA purity particle located at the center. with DFT/LDA in calculating quantities relevant to the ho- Thermodynamic theory of heterogeneous nucleation7–9 mogeneous nucleation. For both types of nucleation, as in as well as our previous theoretical work1 are all based on the Ref. 1, we calculated dependence of the chemical potential perturbation approach with an expansion over small curva- of the liquid condensate on the droplet size, the threshold ture of the liquid-vapor interface. The classical theory7–9 value of the chemical potential ͑or the threshold value of used a zero-order approximation which is only valid for mac- supersaturation͒ at which barrierless nucleation takes place, roscopic nucleation centers. The theory in Ref. 1 involves the the height of the activation barrier, the location of minimum first-order approximation which includes the Tolman curva- and maximum on the curve of free energy of droplet forma- ture correction and compressibility correction. The first-order tion versus the droplet size, and a few other characteristics approximation allows us to deal with mesoscopic wettable necessary to describe kinetics of nucleation. particles. However, for microscopic wettable particles the This paper is organized as follows. In Sec. II thermody- perturbation approach is no longer effective. The reasons are namics of heterogeneous and homogeneous nucleations is given below. discussed. The disjoining pressure is described in Sec. III. In First, the surface tension of a very small droplet is very Sec. IV the DFT/WDA approach is detailed to treat a spheri- different from that in the planar limit. For large droplets the cal system. Characteristics of homogeneous nucleation at Tolman correction to the surface tension can be used in the high supersaturation of vapor are obtained in Sec. V. Key planar limit.1 In fact, the value of the Tolman length is ex- properties of heterogeneous nucleation on microscopic par- tremely small, possibly even close to zero.1,10,11 However, ticles are presented and discussed in Sec. VI. Discussions the first-order curvature correction cannot be used for micro- and conclusions are given in Sec. VII. scopic droplets, since the Tolman length itself is highly de- pendent on the curvature12 while the surface tension of small II. THERMODYNAMICS OF NUCLEATION droplets ͑in conventional thermodynamic sense͒ is close to A. Heterogeneous nucleation 13–15 zero. 1 Second, the fluid density of small droplet is no longer In the previous work we showed several thermody- close to bulk liquid density ͑in coexistence with saturated namic equations such as the Laplace equation and Gibbs- vapor͒. Moreover, this density cannot be approximated using Duhem equation to describe heterogeneous nucleation on compressibility correction. While the compressibility route mesoscopic wettable particles. The general forms of these predicts the bulk density to be larger than that in the planar equations remain valid to describe heterogeneous nucleation limit, the fluid density of the small droplet approaches to on microscopic wettable particles. vapor value14–16 as supersaturation increases. The EMLD We summarize some main results of Ref. 1 in this sec- model5 also supports that at high supersaturations the critical tion. We consider an open system which consists of a wet- cluster is a diffusive physical entity containing a significant table solid particle and a metastable vapor at given tempera- number of vaporlike molecules. This is analog to the behav- ture and volume. Upon vapor condensation, a fluid film with ior of fluid density of a thin liquid film on the solid substrate, a uniform thickness is formed on the surface of the particle. where surface layers overlap.1,7 To describe this behavior, the Assuming spherical symmetry, the density profile of the sys- ␳͑ ͒ disjoining pressure of a thin liquid film has to be considered. tem is given by r , where r is measured from the center of Similarly, the disjoining pressure should be introduced when the solid particle, which is also the center of the droplet. We considering very small droplets, due to the small thickness of denote metastable vapor as phase 1, the fluid film as phase 2, the film on the surface of the impurity particles, as well as and the solid particle as phase 3. To describe thermodynamic the small size of the droplet. We expect similar effect in properties of this nonuniform system, we employed the homogeneous nucleation as well.14,15,17 Gibbs method of dividing surfaces to replace the real density ␳͑ ͒ We employed classical density-functional theory ͑DFT͒ profile r by a step-function density profile. This step- ͑ ͒ as a theoretical tool to calculate thermodynamic characteris- function density profile consists of several regions. 1 A 1 ␳ Ͻ˜ tics of liquid condensate as in previous study. It is known uniform region of solid particle with density 3 for r Rn, ˜ ͑ ͒ that DFT can be used to investigate behavior of fluid near where Rn is the physical radius of the solid particle. 2 A 18–22 ␳ planar solid surface. We also used the weighted-density uniform fluid film with density 2 which is equal to the bulk ͑ ͒ 18 Ͻ Ͻ approximation WDA developed by Tarazona, and Tara- density of liquid at the same supersaturation for Rn r R, 19 zona et al., which has been proven to be effective to de- where Rn is the radius of solid-liquid dividing surface. It

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-3 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

˜ ϵ ˜ ␴ ⌫ ␮ ͑ ͒ differs from the radius of the particle Rn by z0 Rn −Rn d 32 =− 32d 4 ͑about few molecular diameters͒ which depends on the thick- and ness of solid-liquid interface. The uniform density is set to be ␴ץ Ͻ Ͻ ˜ zero in a thin region Rn r Rn. This reflects the fact that d␴ − 21dR =−⌫ d␮, ͑5͒ R 21ץ solid-liquid interaction potential is repulsive near the solid 21 ͑ ˜ ͒ ͑ ͒ surface this potential diverges at r=Rn . 3 The uniform ⌫ ␲ 2 where 32=N32/4 Rn is the adsorption of liquid on the sur- vapor with density of ␳ for rϾR, where R is the radius of ⌫ ␲ 2 1 face of the solid particle, and 21=N21/4 R is the adsorp- liquid-vapor dividing surface. Thus, the thickness of the liq- tion at the liquid-vapor interface. The generalized Laplace uid film is given by h=R−Rn. Finally L represents the out- equation ͑3͒ and Gibbs-Duhem equations ͑4͒ and ͑5͒ are the most radius of the entire system and is chosen in such a way key thermodynamic equations required to study heteroge- that beyond L the physical density of vapor is nearly uniform neous nucleation. ͑ ␳ ͒ = 1 . For the open system with long-range interactions L →ϱ. The supersaturation can be described by the chemical potential ␮. For given this chemical potential B. Homogeneous nucleation also determines the density ␳ of bulk vapor and ␳ of bulk 1 2 To describe homogeneous nucleation, we set Rn =0 so liquid, as well as the bulk vapor pressure p1 and the bulk that Eq. ͑1͒ becomes liquid pressure p2. The grand thermodynamic potential of the ͑ ͒ 4␲ 4␲ system excluding the solid particle is given by ⍀ =−p R3 − p ͑L3 − R3͒ +4␲R2␴ ͑R͒ + ⍀˜ ͑R͒. 2 3 1 3 21 4␲ 4␲ ⍀ =−p ͑R3 − R3͒ − p ͑L3 − R3͒ +4␲R2␴ ͑R ͒ ͑6͒ 2 3 n 1 3 n 32 n The last term is now given by ␲ 2␴ ͑ ͒ ⍀˜ ͑ ͒ ͑ ͒ +4 R 21 R + h , 1 L ⍀˜ ͑R͒ ␲͵ r2␻˜ ͑r͒dr ͑ ͒ where the notations are the same as used in Ref. 1: ␴ ͑R ͒ is =4 , 7 32 n R the surface tension of the solid-liquid interface, which is de- ␴ ͑ ͒ pendent on the radius of dividing surface Rn; and 21 R is because the grand thermodynamic potential is dependent on the surface tension of the liquid-vapor interface, which is the finite size of the droplet. We can still write ␻˜ ͑R͒ dependent on the radius of the dividing surface R. Both =⌸͑0,R͒=⌸͑R͒ as a definition of disjoining pressure. As ␴ ͑ ͒ ␴ ͑ ͒ ͑ ͒ 32 Rn and 21 R are defined in such a way that they are such, the Laplace equation 3 for the critical cluster now independent of the liquid film’s thickness. The fact that ther- becomes ␴ ͑R͒ץ modynamic potential depends on the finite thickness of the 2␴ ͑R͒ ͑ ͒ ⍀˜ ͑ ͒ p − p = 21 + 21 − ⌸͑R͒. ͑8͒ Rץ film h is accounted for with the last term in Eq. 1 , h , 2 1 R which approaches to zero as h approaches infinity. This last 24 term can be expressed by According to the definition of Derjaguin et al., the disjoin- ing pressure is the difference between the normal component L ͓ ⍀˜ ͑h͒ =4␲͵ r2␻˜ ͑r͒dr, ͑2͒ of the pressure tensor inside of the liquid film i.e., at the ͑ ͒ ͔ R +h center of droplet pN 0 here and the pressure of bulk liquid n ␮ p2, given the same chemical potential , where ␻˜ ͑r͒=⌸͉͑R ,h͉͒ is Derjaguin’s disjoining pres- n h=r−Rn ⌸ = p ͑0͒ − p . ͑9͒ sure of thin fluid film. N 2 1 Based on Eq. ͑1͒ we derived the generalized Laplace Combining Eqs. ͑8͒ and ͑9͒ results in equation which is valid for both equilibrium and critical clus- ␴ ͑R͒ץ 2␴ ͑R͒ ters, p ͑0͒ − p = 21 + 21 , ͑10͒ Rץ N 1 R ␴ץ ͒ ͑ ␴ 2 21 R 21 p − p = + ͑R͒ − ␻˜ ͑R + h͒. ͑3͒ which is an alternative version of the Laplace equation com- R nץ R 1 2 pared to the conventional thermodynamic version often used Up to this point we have not made any assumptions about the in classical nucleation theory: ␴ ͑R͒ץ size of the impurity particle covered by the droplet. So the 2␴ ͑R͒ 21t 21t ͑ ͒ Laplace equation ͑3͒ is also valid for microscopic nucleation p2 − p1 = + , 11 Rץ R center as for mesoscopic nucleation center. Moreover Eq. ͑3͒ is valid for homogeneous nucleation too. But for each of where the subscript t denotes a different definition of ther- these cases we need to specify the nature of the last term modynamic surface tension. ␻͑ ͒ ␻͑ ͒ ⌸͑ ͒ ͑ ͒ ͑ ͒ ˜ Rn +h = ˜ R = Rn ,h , which is given below. The reason why Eqs. 10 and 11 are different is be- Another important thermodynamic equation is the gen- cause the density inside a very small liquid droplet differs eralized Gibbs-Duhem equation, which can be derived based from that in bulk liquid. As a result, the normal component on Eq. ͑1͒ also. In the case of heterogeneous nucleation not of pressure tensor inside of this small nonuniform droplet is one but two equations should be considered: different from the bulk liquid pressure. This effect is similar

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-4 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

to the overlap of surface layer of thin liquid films, which can neous nucleation and depends on the nature of interaction ␳ be accounted for by introducing the disjoining pressure in between the solid substrate and liquid, while term − 2ϱt3 Eq. ͑8͒. contributes the disjoining pressure even in the case of homo- 14,15 ␳ In our earlier works on homogeneous nucleation we geneous nucleation. Both u3 and t3 are positive, but 3u3 is ͑ ͒ ␳ used Eq. 11 as the Laplace equation. This conventional much greater than the term 2ϱt3, yielding a positive constant form cannot distinguish different physical contributions to B and a positive disjoining pressure. Positive and monotonic thermodynamic potential. In contrast, Eq. ͑8͒ includes two disjoining pressure versus the film thickness has profound terms. The curvature dependence of surface tension is solely effect on heterogeneous nucleation. It gives a nonmonotonic due to the Tolman effect, which reflects the fact that liquid- behavior for chemical potential of liquid condensate versus vapor interface has a finite thickness. Even without the over- liquid film thickness and is responsible for the existence of lap of the surface layer the change of the shape ͑curvature͒ of stable equilibrium liquid clusters on the surface of wettable the interface should cause change in surface tension. This is particles. a relatively weak effect due to small but finite compressibil- Without the solid particles the u3 term is zero and the ity of liquid, which is not included in classical capillary ap- main contribution to B ͓Eq. ͑13͔͒ is a negative van der Waals proximation. The last term ͑disjoining pressure͒ in Eq. ͑8͒ is term. Hence, B is negative and so is the disjoining pressure. due to interactions between molecules of liquid, which is not In the homogeneous nucleation the thickness of the fluid film directly related to the compressibility. This contribution can is the same as the radius of the droplet, as Rn =0 and R be viewed as volume contribution rather than surface contri- =Rn +h=h. As such, one cannot use the planar limit to esti- bution, as traditionally treated in theory of nucleation. We mate the disjoining pressure since it strongly depends on the will discuss a simple estimation for this term in the next droplet radius. In addition to van der Waals contribution, the section. The conventional form of the Laplace equation ͑11͒ disjoining pressure may depend on other properties, espe- attributes all curvature-related effects to the surface tension cially when the thickness of the liquid film or the size of the term. Finally, the Gibbs-Duhem equation ͑5͒ is still valid for droplet is on the same order of molecular size.1 homogeneous droplet, while Gibbs-Duhem equation ͑4͒ is Let us consider an incompressible liquid that can form a not required because of the nonexistence of the solid-liquid thin film of thickness h on the surface of the particle of interface. ͑ ˜ radius Rn here we ignore the difference between Rn and Rn ͒ for the purpose of simple estimation . The Rn value can III. ANALYTICAL ESTIMATION OF THE DISJOINING range from microscopic to mesoscopic. If one neglects all PRESSURE the interaction between the liquid and solid particles and only accounts for the van der Waals attraction in liquid, the ␻͑ ͒ ͑ ͒ The free energy density ˜ r in Eq. 2 is also the dis- intermolecular potential can be presented as ⌸͑ ͒ 1 joining pressure Rn ,h . In the previous work we consid- ered the wettable particles to be mesoscopic in size so that ␭ w͑r͒ =− , ͑14͒ we can approximate the disjoining pressure by the corre- r6 sponding planar value ⌸͑h͒=lim˜ →ϱ⌸͑˜R ,h͒, assuming Rn n that the model solid-liquid interaction potential depends only where ␭ is a positive constant. on the distance from the surface of the solid. We showed in To analyze the effect of the disjoining pressure, we re- ␳ Ref. 1 that the power-law approximation for disjoining pres- place the bulk liquid of density 2 around the solid particle of sure is valid when the liquid film is thicker than about five radius Rn by a thin liquid film of the same density and a ␳ Ͼ molecular diameters, in which the overlapping effect is not vapor of density 1 for r R. Thus, the disjoining pressure is very strong so that the center of the film can be treated as equal to the difference in interaction energy per unit volume bulklike phase. Meanwhile, small changes in liquid density between these two configurations due to compressibility are almost compensated by weak overlapping. In that case van der Waals interactions play a ␭ ⌸ ␳ ͵ dV͑␳ ␳ ͒ ͑ ͒ main role and =− 2 2 − 1 6 , 15 sur r B ⌸͑h͒Ϸ , ͑12͒ ͑z + h͒3 where the volume integration is taken over the space sur- 0 rounding the droplet. To calculate this integral we used the 25 where constant B ͑which is related to Hamaker constant ͒ cylindrical coordinate system depicted in Fig. 1. Here the can be fitted to DFT data, or can be simply estimated from distance between two molecules is r=ͱ͑z−R+h͒2 +y2,sothe 26 the equation integral becomes

B = ͑␳ ϱ − ␳ ϱ͒͑␳ u − ␳ ϱt ͒, ͑13͒ 2 1 3 3 2 3 ⌸ ␳ ͑␳ ␳ ͒ =− 2 2 − 1 where ␳ ϱ and ␳ ϱ are vapor and liquid densities at phase ϱ 1 2 −R 1 coexistence and coefficients u3 and t3 are integration charac- ϫ2␲␭ͩ͵ dz͵ ydy ͑͑ ͒2 2͒3 teristics of intermolecular potentials. Here u3 stems from at- −ϱ 0 z − R + h + y tractive contribution of the substrate potential, while t is due ϱ 3 +R 1 to van der Waals interactions between liquid molecules. This ͵ ͵ + dz ydy 2 2 3 ␳ ͱ 2 2 ͑͑z − R + h͒ + y ͒ means that the term 3u3 exists only in the case of heteroge- −R R −z

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-5 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

IV. DFT CALCULATION OF MODEL SYSTEM

Again, we used the DFT/WDA approach to study the inhomogeneous system of microscopic wettable particle sur- rounded by a fluid film. The homogeneous nucleation is also considered as a special case of heterogeneous nucleation with Rn =0, even though the DFT treatment of homogeneous nucleation does not require WDA ͑since LDA is sufficient͒. We will compare results based on both LDA and WDA to see whether WDA can bring some new insights into homoge- neous nucleation problem. For example, it is known that FIG. 1. A schematic plot of a thin liquid film formed on a spherical solid small oscillations of liquid density may occur inside of the particle. planar liquid-vapor interface even without the solid substrate.27 In Ref. 1 we discussed how this density- ϱ ϱ + 1 oscillation behavior can affect the planar surface tension and + ͵ dz͵ ydy ͪ. ͑16͒ the Tolman length. Here we will see whether similar effect ͑͑ ͒2 2͒3 +R 0 z − R + h + y can change the surface tension of small liquid droplet. First, we briefly summarize main aspects of DFT/LDA After this integration the disjoining pressure is given by and DFT/WDA ͑Refs. 18 and 19͒ theoretical formalisms. Let ͑ ͒ ͑ ͒ 4␲ R3 w r1 ,r2 =w r12 represent the pairwise intermolecular po- ⌸ =−␳ ͑␳ − ␳ ͒␭ . ͑17͒ tential for the simple nonpolar molecules. In DFT the free 2 2 1 ͑ ͒3 3 3 2R − h h energy of a nonuniform system, within random-phase ap- ͑ ͒ Note that ⌸ is a function of two variables R and h.Sowe proximation RPA , is given by consider two special cases. The first one is for heterogeneous 1 ͓␳͔ ͵ ͑␳͑ ͒͒ ͵͵ ͑ ͒␳͑ ͒␳͑ ͒ nucleation on impurity particles of a very large size, which F = drf r + dr1dr2wp r12 r1 r2 , 2 was considered in Ref. 1. In this case we only need to con- ӷ ⌸ ͑20͒ sider Rn h, because if h is comparable to Rn, is negli- ͑ ͒ gible. Hence, Eq. 17 can be expanded in terms of small where ␳͑r͒=␳͑r͒ is the density profile for spherically sym- ͑ ͒ parameter h/Rn or h/R . The zero-order contribution gives metrical system ͑the r value is measured from the center of ͒ w ͑r ͒ 4␲ 1 the droplet and p 12 is the weak attractive part of inter- ⌸ =−␳ ͑␳ − ␳ ͒␭ molecular potential. We then approximate the repulsive part 2 2 1 ͑ ͒3 3 3 2−h/R h of free energy density f͑␳͒ either with the free energy density ͑␳͑ ͒͒ 28 4␲ 1 of hard spheres fhs r in Carnahan-Starling form in LDA Ϸ − ␳ ͑␳ − ␳ ͒␭ . ͑18͒ ͑␳͑ ͒͒ ͑␳͑ ͒͒ ␳͑ ͒⌬␺ ͑␳͒ 2 2 1 3 8h3 or with f r = fid r + r hs ¯ in WDA. Here ͑␳͑ ͒͒ ⌬␺ ͑␳͒ fid r is the free energy density of ideal gas and hs ¯ is ␳ 28 This estimation is equivalent to that given by the − 2ϱt3 term the excess free energy density in Carnahan-Starling form. ⌬␺ ͑␳͒ ␳͑ ͒ in Eq. ͑13͒, that is, to replace the disjoining pressure by its This hs ¯ is a function of weighted density ¯ r . The value in the planar limit.1 This replacement is quite accurate, latter can be determined by averaging the true local density ␳ ͑ ͒ ␳͑ ͒ 18 since the van der Waals term − 2ϱt3 in Eq. 13 is much r over certain local volume. Following Tarazona and ␳ 19 smaller than the first contribution from the 3u3 term. Tarazona et al., The second special case is for homogeneous nucleation without the presence of impurity particles, so that R =0 and ␳͑ ͒ ͵ ␳͑ ͒␻͑ ␳͑ ͒͒ ͑ ͒ n ¯ r1 = dr2 r2 r12,¯ r1 , 21 R=h. Thus, Eq. ͑17͒ becomes where the weighting function ␻ is chosen according to the 4␲ 1 ⌸ ␳ ͑␳ ␳ ͒␭ ͑ ͒ 18 19 =− 2 2 − 1 3 . 19 model of Tarazona and Tarazona et al.. 3 R The open system is considered in grand canonical en- semble. In DFT the grand thermodynamic potential is given If one ignores vapor density ␳ as ␳ Ӷ␳ , it gives the ex- 1 1 2 by pression for disjoining pressure of the small droplet, obtained earlier by Tsekov et al.17 Note that the disjoining pressure in ⍀͓␳͔ ͓␳͔ ͵ ␳͑ ͒ ͑ ͒ ␮ ͵ ␳͑ ͒ ͑ ͒ this case is eight times larger compared to that of the planar = F + dr1 r1 Vext r1 − dr1 r1 , 22 thin film ͓Eq. ͑18͔͒, assuming R=h. ͑ ͒ In the case of microscopic wettable particles, for which where Vext r1 is the external potential. In the case of hetero- both R and h are comparable, the disjoining pressure can geneous nucleation this external potential arises due to the only be estimated based on Eq. ͑17͒, not mentioning other impurity particle at the center of the droplet. For homoge- complicated contributions which are not accounted for in this neous nucleation this potential is zero. The equation for equi- simple estimation of the disjoining pressure. In the following librium density profile can be derived based on the variation sections we apply DFT to obtain more quantitative values of condition ␦⍀͓␳͔/␦␳=0. If one uses LDA ͑for homogeneous ͒ the disjoining pressure. nucleation with Vext=0 , the variation gives

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-6 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

ϱ ͉ ͉ 2␲ r1+r2 thin liquid film on the surface of the particle. This is a stable ␮ ͑␳͑ ͒͒ ␮ ͵ ␳͑ ͒͵ ͑ ͒ h r1 = − r2dr2 r2 r12dr12wp r12 . resulting from the minimization of the thermody- r1 ͉ ͉ 0 r1−r2 namic potential. ͑2͒ Critical cluster of the liquid droplet with ͑23͒ solid particle at the center, which is in unstable equilibrium ͑ ͒ Using WDA one can obtain with the surrounding vapor. 3 Infinitely thick bulklike liq- ϱ uid surrounding the solid particle. To start with the numerical 1 2␲ ␳͑r ͒ = expͭ− ͫ ͵ r dr ␳͑r ͒⌬␺Ј ͑¯␳͑r ͒͒ iteration procedure, the initial step function should mimic 1 2 2 2 hs 2 kBT r1 0 two possible dropletlike as close as possible. We ͉ ͉ calculated density profiles for several given chemical poten- r1+r2 ␻͑r ,¯␳͑r ͒͒ ϫ͵ 12 2 tials. The density profiles are harder to obtain as the chemical r12dr12 ͉ ͉ 1−¯␳ ͑r ͒ −2¯␳ ͑r ͒¯␳͑r ͒ r1−r2 1 2 2 2 2 potential approaches to its threshold value, where the differ- ϱ ence between two solutions is nearly undistinguishable. 2␲ ⌬␺ ͑␳͑ ͒͒ ͵ ␳͑ ͒ Whenever the threshold supersaturation is reached, heteroge- + hs ¯ r1 + r2dr2 r2 r1 0 neous nucleation becomes barrierless. Nevertheless, it is very ͉ ͉ r1+r2 difficult to find accurate threshold value of chemical poten- ϫ͵ r dr w ͑r ͒ + V ͑r ͒ − ␮ͬͮ, ͑24͒ 12 12 p 12 ext 1 tial, since the solution is highly unstable when dealt with the ͉ ͉ r1−r2 iteration procedure. ␳ ␳ 18 As expected, the density profiles exhibit a strong oscil- where ¯1 and ¯2 are defined by Tarazona and Tarazona et al.19 latory behavior near the surface of the solid particle. Since To solve the last two equations numerically, the exact we are mainly interested in microscopic particles here, we ͑ ͒ ˜ potential function should be known. Again, we used the performed calculations for two sizes: 1 Rn =5.0d which is ͑ ͒ Lennard-Jones LJ model potential, for which the attractive somewhat close to mesoscopic size, and ͑2͒ ˜R =3.5d which ͑ ͒ n part of intermolecular potential wp r12 is approximated by is surely a microscopic size, but still much larger than the ͓ using Weeks-Chandler-Anderson perturbation scheme see molecular size. ͔ Eq. 19 in Ref. 1 . We considered a number of values of chemical potentials For homogeneous nucleation, it can be shown that Eq. ͑supersaturations͒ within the range ␮ Ͼ␮Ͼ␮ϱ. The maxi- ͑ ͒ ␳ ␳ max 23 has three solutions: two uniform solutions 1 and 2 ␮ ␮ mum value of the chemical potential max= s1 is the spin- representing the bulk vapor and liquid, and a nonuniform odal value, above which the only equilibrium solution is the ␳͑ ͒ solution r representing the spherical density profile of the uniform liquid in the case of homogeneous nucleation. The ͑ ͒ unstable critical droplet. The latter refers to the saddle point maximum value of the chemical potential in the case of het- of the grand thermodynamic potential. We followed the same erogeneous nucleation is the threshold value ␮ =␮ . 23 ͑ ͒ max th numerical iterational procedure to solve Eq. 23 as well as Nucleation becomes barrierless when ␮Ͼ␮ . The minimum Eq. ͑24͒. For Eq. ͑24͒, the numerical solution is more chal- th value of the chemical potential ␮=␮ϱ corresponds to vapor- lenging to derive. The procedure has to be adjusted, using the liquid coexistence, where vapor and liquid bulk are technique introduced by Tarazona.18 One has to choose an equal p =p =pϱ. For ␮Ͼ␮ϱ we have liquid pressure p appropriate initial density profile and to conduct iterations 1 2 2 Ͼp which leads to metastability of vapor phase. Based on very carefully to monitor changes of ⍀. This numerical pro- 1 the density profile solutions obtained by solving Eqs. ͑23͒ cedure becomes even harder to carry out for low supersatu- and ͑24͒ with the chemical potential in the range ␮ Ͼ␮ rations, at which the size of the critical cluster is large but max Ͼ␮ϱ, we can study properties of the liquid droplets and not very sensitive to the supersaturation. kinetics of liquid-vapor nucleation. These properties are dis- It is worthy to note that at low temperature T* ⑀ ͑⑀ ͒ cussed in the following sections. =kBT/ LJ=0.7 LJ is a parameter of LJ potential , the den- sity profiles obtained from Eq. ͑24͒ still exhibit weak oscil- lations on the liquid side of the interface as in the planar case, even though the surface layer is much thinner now. For heterogeneous nucleation with impurity particles of V. HOMOGENEOUS NUCLEATION any size, a particle-liquid interaction potential function V ͑r͒ is required. Here, we still used the same potential as ext We proceeded with a definition of the droplet’s radius R in Ref. 1, that is, ͑or the dividing surface͒. In fact, thermodynamic quantities u u on the right-hand side of Eq. ͑6͒ are strongly dependent on V ͑r͒ =−␳ ͩ 3 + 9 ͪ, ͑25͒ ext 3 ͑r − ˜R ͒3 ͑r − ˜R ͒9 the choice of the dividing surface, and yet the definition of n n the dividing surface is not unique for droplets. The most ␳ 3 ⑀ ␴6 ⑀ ␴12 with 3 =12/d , u3 =2.348 LJ , and u9 =−5.326 LJ as in common definition is such that the dividing surface can be ͑ ͒ ϱ Ref. 26. For numerical calculation we set Vext r = for r solely determined by the density profile. For instance, we can Ͻ˜ * Rn +0.84d. Calculations were performed for the given T use the equimolar dividing surface, which is defined such ⌫ =0.7. Although Eq. ͑24͒ also has three solutions, all of them that the adsorption 21=0. Then, according to the Gibbs are nonuniform due to the presence of the impurity particle. method of the dividing surfaces, the number of particles in a The solutions are as follows. ͑1͒ Equilibrium cluster with droplet can be given by

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-7 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

FIG. 2. The dependence of the reduced density ␩͑0͒=␲d3␳͑0͒/6 at the center of a homogeneous droplet on the dimensionless equimolar radius * * ⑀ Re =Re /d of the droplet at T =kBT/ LJ =0.7. The horizontal line represents the liquid density at phase coexistence.

L 4␲ 4␲ ␲͵ 2␳͑ ͒ 3␳ ͑ 3 3͒␳ ͑ ͒ 4 r r dr = Re 2 + L − Re 1, 26 0 3 3

where R=Re is the radius of equimolar surface. This equation can be rewritten as 3͐Lr2͑␳͑r͒ − ␳ ͒dr R3 = 0 1 . ͑27͒ e ␳ ␳ 2 − 1 Clearly, the equimolar surface is only dependent on the den- ␳ ␳ sity profile and bulk 1 and 2 values, thus it is unique. With the dividing surface R defined we turn our attention ͑ ͒ ͑⌸* ⌸␲ 3 ͒ FIG. 3. a The reduced disjoining pressure = d /6kBT of small to the main objective of this section—properties of droplets ͑ * ͒ * homogeneous droplets vs the equimolar dividing surface Re =Re /d at T as a function of droplet size. It is known that the higher the ⑀ ͑ ͒ ͑⌸* ⌸␲ 3 ͒ =kBT/ LJ =0.7. b The reduced disjoining pressure = d /6kBT of ͑ ␮͒ ͑ ͒ ͑ * supersaturation or , the smaller the size or radius R of small homogeneous droplets vs two different dividing surfaces RE =Re /d * ͒ * ⑀ the critical droplet, and the higher density of bulk liquid due and R =R/d at T =kBT/ LJ =0.7. to compressibility factor. The density at the center of the critical droplet, however, only behaves this way at low su- grand thermodynamic potential which is readily calculated persaturations ͑see Fig. 2͒. In contrast, for extremely small from DFT. Eventually, one can obtain the relation between droplets, the center density decreases considerably at certain size. It can only be explained by overlapping of the surface the disjoining pressure and equimolar radius of the critical ͑ ͒ layer of the droplet. This behavior can be seen from both droplet. This relation is presented in Fig. 3 a , obtained from LDA and WDA calculations of the density profile ͑Fig. 2͒,as both LDA and WDA calculations. Both calculations show well as from the gradient-expansion calculation.29 Note that that in the case of small droplets the disjoining pressure is the weighted density profile is used to illustrate density at the negative and increases in absolute value with decreasing the center of the droplet in WDA calculation ͑to exclude second- radius of the droplet. In the same figure, we also plotted the ⌸͑ ͓͒ ͑ ͔͒ ␭ ⑀ ␴ ary effect of density oscillation about the weighted value.͒ At estimated R Eq. 19 , where was taken to be 4 LJ LJ the spinodal, the critical cluster completely turns into uni- for the Lennard-Jones potential. The simple estimation ͑␳͑ ͒ ␳ ͒ ͑ ͒ form vapor r = 1s and Re =0 according to Eq. 27 . agrees reasonably well with the DFT numerical results, at least for RϾ3d, even though neither the compressibility of ⌸ A. Disjoining pressure „R… liquid nor overlapping of the surface layer was taken into ␮ ␮ ␮ account. For the smaller droplets ͑less than 3d in radius͒ the For a given in between ϱ and s1, the disjoining pressure ⌸͑R͒ can be calculated on the basis of the definition overlapping of the surface layer is much stronger and it is ͑ ͒ mainly responsible for the difference between the computed Eq. 9 in which the bulk liquid pressure p2 is known. The ͑ ͒ ⌸ ͑ ͒ normal component of the pressure tensor pN 0 at the center and the estimated one from Eq. 19 . Similar behavior was of the droplet can be calculated based on the general me- seen in Ref. 1, where we showed that power-law approxima- ͑ ͒ ͑ ͒ chanical relation pN 0 =pT 0 in the case of spherical sym- tion works well for thick liquid films, but an exponential metry ͑e.g., Ref. 16͒. Meanwhile, the tangential component approximation has to be used for thinner structural films due of the pressure tensor is equal to the negative density of the to strong overlapping of the surface layer.

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-8 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

B. Surface tension There are two ways to define the surface tension of small homogeneous droplets, depending on whether the disjoining pressure contribution is considered as part of the surface ten- sion term or not. To calculate the surface tension from DFT, we rewrite Eq. ͑6͒ as ⍀ + p 4␲L3/3 R L r2 ␴ = 1 + ͑p − p ͒ − ͵ ⌸͑r͒dr, ͑28͒ 21 ␲ 2 2 1 2 4 R 3 R R

where the grand thermodynamic potential ⍀ can be calcu- ͑ ͒ lated according to Eq. 22 , and the bulk pressures p1 and p2 are known for a given value of ␮, as well as the disjoining pressure ⌸͑r͒. An alternative way to define the surface tension is to include the disjoining pressure term. As such, one obtains the traditional thermodynamic surface tension ⍀ + p 4␲L3/3 R ␴ = 1 + ͑p − p ͒ . ͑29͒ 21t 4␲R2 2 1 3 In principle, both definitions entail arbitrary radius of divid- ing surface R. If the radius is chosen such that it is indepen- dent of the definition of surface tension, then there is a simple relationship between both surface tensions L r2 ␴ ␴ ͵ ⌸͑ ͒ ͑ ͒ 21 = 21t − 2 r dr. 30 R R

In the previous section, we have shown that the disjoining pressure is negative for small droplet. This means that the ␴ ͑ ͒ surface tension 21 calculated according to Eq. 28 is always higher than ␴ . 21t ͑ ͒ ͑␴* ␴ ␲ 2 ͒ FIG. 4. a The reduced surface tension e = 21e d /6kBT of small ho- Let us first consider a few limiting cases. When droplets ͑ * ͒ * mogeneous droplets vs the equimolar dividing surface Re =Re /d at T ⑀ ͑ ͒ ͑␴* ␴ ␲ 2 ͒ become very large, the disjoining pressure decreases rapidly =kBT/ LJ =0.7. b The reduced surface tension = 21 d /6kBT of small * ⑀ and the last term in Eqs. ͑28͒ and ͑30͒ diminishes. Thus, both homogeneous droplets vs two different dividing surfaces at T =kBT/ LJ surface tensions approach to the same limiting value of pla- =0.7. nar liquid-vapor surface tension which is independent of the dividing surface. Both LDA and WDA give similar behavior, although the When the chemical potential approaches to the value of more accurate WDA gives smaller surface tension and liquid spinodal, the equimolar size of the droplet approaches weaker nonmonotonic behavior than LDA. In contrast, the to zero. In such a case the first term in Eq. ͑29͒ vanishes, surface tension defined based on Eq. ͑28͒, which excludes ⍀ ͑ ␲ ͒ 3 since is equal to its bulk value −p1 4 /3 L . The second the disjoining pressure effect, shows a monotonic increase term also becomes zero as R=0 and thus, thermodynamic from large to small droplets due to the negative value of the surface tension calculated according Eq. ͑29͒ is zero. On the Tolman length in the planar limit. other hand, the surface tension defined via Eq. ͑28͒ diverges The existence of multiple definitions of radius of the as 1/R2. Note that this effect occurs for equimolar dividing droplet has caused debates in the theory of curved surfaces, surface since the droplet completely disappears in vapor at especially on the dependence of surface tension to the cur- the spinodal, that is, the liquid-vapor interface is effectively vature. There is another common definition of radius, in nonexistence, while the surface energy per unit of area of namely, the radius of surface of tension, which is not unique. the droplet grows infinitely in order to compensate the effect The surface of tension is defined such that the Laplace equa- of the disjoining pressure. tion ͑8͒ can be expressed in a simple form Figure 4͑a͒ shows the dependence of the surface tension on the equimolar radius of the droplet based on both WDA 2␴ 21s ͑ ͒ and LDA calculations. Thermodynamic surface tension cal- p2 − p1 = , 31 culated according to Eq. ͑29͒ is a nonmonotonic function of Rs the droplet size. It becomes slightly higher than the planar ␴ value as the size of the droplet decreases, exhibits a maxi- where 21s refers to the surface tension at the surface of mum at intermediate radius around 7–9d, and eventually tension and Rs refers to the radius of the surface of tension. approaches to zero as the size of the droplet goes to zero. Thus, one has

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-9 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

␴ cal surface of tension. This may be one of the reasons whyץ 21s = ⌸͑R ͒͑32͒ R s mechanical and thermodynamic definitions of the surfaceץ tension and the Tolman length give different results for the and small droplets. For large droplets, when disjoining pressure is negligible, all definitions of the surface of tension give the ͑p − p ͒R ␴ 2 1 s ͑ ͒ 21s = . 33 same result. 2 In summary, the radius of equimolar dividing surface Substituting Eq. ͑33͒ into Eq. ͑28͒ gives appears to be more reliable to describe the size of the droplet, since it is independent of the surface tension and the disjoin- 6͑⍀ −4␲/3L3p ͒ L ⌸͑r͒ ing pressure. We use this radius hereafter. R3 = 1 −6͵ r2 dr. ͑34͒ s 4␲͑p − p ͒ p − p 2 1 Rs 2 1 C. Chemical potential of the liquid condensate The solution of Eq. ͑34͒ is the radius of surface of tension. The equation can be solved numerically using iteration A prerequisite to calculate the barrier height to nucle- method. For very large droplets the second term on the right ation is the chemical potential of the liquid condensate as a hand is almost negligible and the first term can provide an function of the critical droplet radius R or the number of accurate estimation for the radius of tension ͑which is also particles ␯ in the critical droplet. In fact, the definition of the the classical thermodynamic definition of this radius͒. This surface tension and the disjoining pressure should not affect estimation of Rs can then be inserted into the second term the experimentally measurable thermodynamic characteris- and a new approximation for the radius can be obtained. tics such as chemical potential. We showed in Ref. 1 that the After several iterations the radius can be calculated with high presence of disjoining pressure in heterogeneous droplets re- Ͼ ␮͑␯͒͑ accuracy. For large droplets we have Rs Re because the sults in a nonmonotonic behavior of see Fig. 4 in Ref. Tolman length is a negative quantity in the limit of the large 1͒. The disjoining pressure also occurs in homogeneous droplets. For smaller droplets, the first term on the right-hand droplets. We will examine whether it will cause a nonmono- side of Eq. ͑34͒ becomes smaller and vanishes in the small- tonic behavior of the chemical potential or not. Note that the droplet limit while the second term remains finite. This classical nucleation theory does not predict such nonmono- means that the radius of tension calculated according to Eq. tonic behavior. However, the fact that the nucleation rates ͑34͒ is nonzero even at the liquid spinodal and that the sur- calculated based on the capillary approximation differ from face tension according to Eq. ͑33͒ still remains finite. Figure the measured nucleation rates at high supersaturations led 3͑b͒ shows how the disjoining pressure depends on the ra- some researchers30,31 to speculate that chemical potential of ⌸͑ ͒ ␯ dius of surface of tension Rs. Rs is still a monotonic func- the liquid condensate may be a nonmonotonic function of tion but the curve is shifted to the right in the region of the even for homogeneous nucleation. In Fig. 5͑a͒, we plot small droplet sizes. Figure 4͑b͒ shows the surface tension chemical potential as a function of particle number ␯ ͑within ͑ ͒ ͒ calculated according to Eq. 33 as a function of Rs. Both the equimolar dividing surface using both LDA and WDA. WDA and LDA curves are monotonic and close to each To calculate ␯, we used the equation other. The curves are shifted to the right compared to those 4␲ 4␲ ␯ 3␳ ␲ 2⌫ 3␳ ͑ ͒ obtained based on the equimolar dividing surface. = R 2 +4 R 21 = Re 2, 37 As mentioned above, definition of the surface of tension 3 3 is not unique. Another definition can be based on the Laplace ⌫ where the adsorption 21=0 at the equimolar dividing sur- equation in the form of Eqs. ͑10͒ and ͑11͒. Equation ͑11͒ face. We also plot the curve based on the capillary approxi- provides the first term on the right-hand side of Eq. ͑34͒, mation which only agrees with the full version of this equation at ␴ ͑ ͒ ͑ ͒ 2 21ϱ low supersaturations large droplets . If one uses Eq. 10 as ⌬␮ = . ͑38͒ ␳ the Laplace equation and sets 2ϱR 2␴ One can see that all the ␮͑␯͒ curves are monotonic. Hence p ͑0͒ − p = 21s , ͑35͒ N 1 R we conclude that the monotonic behavior of chemical poten- s tial of the critical droplets is a distinct feature of homoge- then Eq. ͑34͒ becomes neous nucleation. This feature characterizes a qualitative dif- ference between the homogeneous and heterogeneous ͑⍀ ␲ 3 ͒ L 1 6 −4 /3L p1 ͑ ͒ R3 = −6͵ r2⌸͑r͒dr. nucleations. On the other hand, from Fig. 5 a one can see a s p − p +3⌸ 4␲ 2 1 Rs large difference among predictions of capillary approxima- ͓ ͑ ͔͒ ͑36͒ tion Eq. 38 and DFT for small droplets. In fact, the higher the supersaturation ͑the chemical potential of vapor͒, the ͑ ͒ It can be seen that Eq. 36 yields negative value for Rs at larger differences between predictions of classical capillary high supersaturations, due to overlapping of the surface approximation and DFT calculation. The DFT curve is much ͉⌸͉ ͑ ͒ layer, when p2 −p1 becomes less than 3 . It seems that Eq. lower than that based on Eq. 38 . At the spinodal the clas- ͑36͒ may not be the right way to introduce the surface of sical theory fails to predict the existence of threshold chemi- tension. However, Eq. ͑35͒ is often used to calculate me- cal potential at which the barrier becomes zero. For large chanical surface tension and to define the radius of mechani- droplets, the capillary-approximation curve and the LDA

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-10 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

tional to the pressure difference between liquid and vapor phases. The pressure difference can be determined from the Laplace equation ͑8͒. In the case of heterogeneous nucleation1 the disjoining pressure term in this equation is negative, causing chemical potential to be a nonmonotonic function of the droplet size. In the case of homogeneous nucleation this term is positive. In contrast to mesoscopic droplets, for which the surface tension is only slightly af- fected by the small Tolman correction1, now the surface ten- sion is strongly dependent on the curvature. Thus, this strong curvature dependence of the surface tension in the first two terms of the Laplace equation ͑8͒ gives an explanation to the difference in the calculated chemical potential from the cap- illary approximation and DFT ͓Fig. 5͑b͔͒. The approximation of Eq. ͑40͒ slightly overestimates the threshold value if one term is considered, but underestimates it if two terms are considered. By and large, Eq. ͑40͒ is in good agreement with numerical results. Note that the values of bulk pressures p1 and p2, compressibility, and liquid density are all measurable quantities, and thus experimental data can be used in Eq. ͑40͒. When the chemical potential of the liquid condensate is close to its maximum threshold value, one can introduce the near-threshold region defined by condition of Eq. ͑57͒ in Ref. 1. In this region the nucleation rate is sufficiently dif- ferent from zero and the ␮͑␯͒ curve for the critical clusters can be approximated by parabolic one, that is, ͑␯͒Ϸ 1 ␯2 ͑ ͒ b bs1 − 2 A , 41 ͑␯͒ ⌬␮͑␯͒ where b = /kBT is the reduced chemical potential ͑ ͒ which is dependent on two parameters: 1 bmax=bs1, the spinodal value of chemical potential which can be calculated FIG. 5. ͑a͒ The dependence of the reduced chemical potential of the liquid from the bulk properties, and ͑2͒ A=͉d2b͑␯͒/d␯2͉ , which ͑␮ ␮ ͒ ␯ 0 condensate b= − ϱ /kBT on the number of particles of small homoge- * ⑀ ͑ ͒ can be fitted to DFT curves in Fig. 5. The curves fitting from neous droplets at T =kBT/ LJ =0.7. b The dependence of the reduced ͑␮ ␮ ͒ Eq. ͑41͒ are also shown in Fig. 5. At the reduced temperature chemical potential of the liquid condensate b= − ϱ /kBT on the number ␯ * ⑀ * ⑀ of particles of small homogeneous droplets at T =kBT/ LJ =0.7. T =kBT/ LJ=0.7, the fitting value A=0.0040. This parabolic approximation works well for very small clusters with less curve are located higher than the WDA curve. This is be- than 25 particles. We will use it to calculate the free energy cause WDA predicts lower planar surface tension than LDA of droplet formation in the near-threshold region. ͑at least far below critical temperature͒. More insights into the behavior of chemical potential can D. Free energy of droplet formation and nucleation be gained from the relation between the chemical potential barrier height and the pressure in liquid condensate. The pressure differ- ence p2 −p1 can be calculated via the compressibility route, The free energy of droplet formation and the barrier as shown before in Eq. ͑37͒ of Ref. 1, that is, height are key characteristics that determine the rate of nucleation.1,32,33 Following the same approach as in Ref. 1 ␳ ⌬␮ 1 ␳2 ␹ ⌬␮2 ͑ ͒ p2 − p1 = p1ϱ − p1 + 2ϱ + 2 2ϱ 2 , 39 we computed the free energy of droplet formation for droplet of any size on the basis of b͑␯͒ for critical clusters ͑derived where ␹ is the compressibility of liquid. In Eq. ͑39͒ the term 2 in the previous subsection͒. When a liquid droplet is in me- p1ϱ −p1 can be neglected compared to the next two terms as ␳ Ӷ␳ ͑ ͒ chanical and thermal equilibriums with the surrounding va- 1 2 below the critical temperature . By neglecting p1ϱ ␹ ͑ ͒ por, one has −p1 and expanding over small parameter 2 p2 −p1 ,weob- ͒ ͑ ␯ ͑␯͒ץ ץ tained W/ = b − b1, 42 ͑p − p ͒ 1 ͑p − p ͒2 where W is the reduced free energy of droplet formation ͑in ⌬␮ 2 1 ␹ 2 1 ͑ ͒ = − 2 . 40 units of k T͒ and b is the reduced chemical potential of ␳ ϱ 2 ␳ ϱ B 1 2 2 vapor surrounding liquid cluster. To derive the formation free In Eq. ͑40͒ the first term is the main term and the second energy as a function of the size of the droplet, we integrated term is the correction term due to liquid compressibility. This Eq. ͑42͒ with the boundary condition W͑0͒=0 ͑the formation equation shows that the chemical potential is largely propor- free energy is zero if the size of droplet is zero͒ and obtained

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-11 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

with the more accurate WDA. Since this surface tension is used for classical calculations of formation free energy, the predicted difference in nucleation rate can be very large with the LDA. In this aspect, it is more desirable to use the more accurate WDA for homogeneous nucleation. To further de- scribe kinetics of nucleation at some particular nucleation conditions, we need to analyze the shape of the curve in Fig. 6͑a͒, including not just the height of nucleation barrier ͓maximum on the curve W͑␯͔͒ but also the location of the maximum ͑size of the critical cluster͒ and the half-width of the maximum. As shown in Fig. 6͑a͒, WDA predicts a smaller critical size compared to LDA. The critical size pre- dicted from the classical nucleation theory is much larger than that predicted from DFT. At low supersaturations both classical and DFT results are much closer. This is in part due to the fact that1,11 the Tolman length in the limit of the large droplets is very close to zero. As such, the classical nucle- ation theory can provide quite accurate description of the nucleation at low supersaturation. We now analyze formation free energy in the near- threshold region, where the chemical potential of vapor is ␮ very close to the spinodal value s1. In this region, we can define an ⑀ parameter via ͑ ⑀͒ ͑ ͒ b1 = bth 1− , 44 where the barrierless nucleation occurs at ⑀=0 and the near- threshold region refers to ⑀Ӷ1. Parabolic approximation of Eq. ͑41͒ can be used for the chemical potential in this region. Integration of Eq. ͑41͒ gives ␯ 1 1 W = ͵ ͩb − A˜␯2 − b + ⑀b ͪd˜␯ = ⑀b ␯ − A␯3. s1 2 s1 s1 s1 6 ͑ ͒ ͑ ͒ 0 FIG. 6. a Formation free energy in units of kBT of small homogeneous droplets as a function of the number of particles ␯ of the droplets at T* ͑45͒ ⑀ ͑ ͒ =kBT/ LJ =0.7 and different values of supersaturation. b Formation free ͑ ͒ energy in units of kBT of small homogeneous droplets in the near-threshold Critical droplets in the near-threshold region have a size of region as a function of the number of particles ␯ of the droplets at T* =k T/⑀ =0.7. 2⑀b 1/2 B LJ ␯ = ͩ s1 ͪ , ͑46͒ c A ␯ ͑␯͒ ͵ ␯ ͑␯͒ ␯ ͑ ͒ and half-width of the nucleation barrier is given by W = d˜b ˜ − b1 . 43 0 2 1/4 ⌬␯ = ͩ ͪ . ͑47͒ c ⑀ Ͻ bs1A In homogeneous nucleation with b1 bth, the formation free energy is a nonmonotonic function of the droplet size and Substituting Eq. ͑46͒ into Eq. ͑45͒ gives exhibits a maximum. This maximum corresponds to the ⑀3/2 3/2 nucleation barrier and the critical cluster. 2 bs1 ⌬W = ͱ2 , ͑48͒ Figure 6͑a͒ depicts formation free energy curves in the 3 A1/2 case of high supersaturations where the differences between the classical theory and DFT are large. In general, the DFT that is, the height of nucleation barrier. To satisfy parabolic approximation and to be able to use the classical theory for predicts a lower height of nucleation barrier than the classi- 32 cal theory. Secondly DFT/WDA gives a lower barrier height kinetics of nucleation, the condition than LDA. It is known that classical nucleation theory gen- ⌬␯ A 1/4 c = ͩ ͪ Ӷ 1 ͑49͒ erally overestimates the nucleation rate by several orders of ␯ 2⑀3b3 magnitude at high supersaturations compared to experi- c s1 34 ⑀ ments. Similar situation could occur when comparing clas- should be met. For given A and bs1 this means that should sical theory with our WDA results. Meanwhile the LDA re- be larger than 0.15. The condition of Eq. ͑49͒ also means that sults presented in Fig. 6͑a͒ would cause almost 20 orders of the dimensionless height of nucleation barrier described by magnitude difference in nucleation when compared with Eq. ͑48͒ should be more than 2/3. Another condition that classical theory. As pointed out in the previous work1 LDA allows treating ␯ as continues variable in kinetics of nucle- tends to overestimate the planar surface tensions compared ation is that32

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-12 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

⌬␯ ӷ ͑ ͒ c 1. 50 Once the radius of the dividing surface and the adsorp- tion are known, we can then calculate the equimolar radius ⑀ In our case it requires to be at least less than 0.20. So we based on the density profiles for heterogeneous droplets as can identify the relatively small region of 0.15Ͻ⑀Ͻ0.20 ͑for * L 3 given temperature T =k T/⑀ =0.7͒ as the one where the 3 Rn B LJ R3 ͩ͵ r2͑␳͑r͒ ␳ ͒dr ␳ R2⌫ ͪ ͑ ͒ e = − 1 + 2 − n 32 . 55 simplified near-threshold description of homogeneous nucle- ␳ − ␳ ˜ 3 2 1 Rn ation according to Eqs. ͑45͒–͑48͒ becomes possible. Figure Ͼ ␮Ͼ␮ 6͑b͒ shows formation free energy curves for several values This equimolar radius Re Rn for any s2. of supersaturation in the near-threshold region.

A. Chemical potential of the liquid condensate

VI. HETEROGENEOUS NUCLEATION As mentioned previously, knowing the chemical poten- tial of the liquid condensate as a function of droplet size is Again we use the radius of equimolar dividing surface to required to determine all necessary characteristics of nucle- describe the size of droplets, including some modification for ation. For heterogeneous nucleation on mesoscopic particles the case of heterogeneous nucleation. Equation ͑26͒ now be- we obtained Eq. ͑39͒ in Ref. 1. However, the approximation comes used in Ref. 1, such as the use of first-order curvature cor- rection and the planar term of disjoining pressure, is only L 4␲ ␲͵ 2␳͑ ͒ ␲ 2⌫ ͑ 3 3͒␳ valid for mesoscopic droplets. In the case of microscopic 4 r r dr =4 Rn 32 + Re − Rn 2 ˜ 3 Rn particles it is difficult to derive an analytical expression for 4␲ the dependence of the chemical potential on the radius of the ͑ 3 3͒␳ ͑ ͒ + L − Re 1. 51 droplet. But this dependence can be obtained numerically 3 based on DFT calculation. Towards this end, we first com- This equation involves three unknown quantities: radii of puted density profiles for given chemical potential and then dividing surfaces Re and Rn, and the adsorption of liquid on the radius of the droplet. ⌫ the surface of the particle 32. We used the same approach as The chemical potential of the liquid condensate as a ⌫ ͑ ͒ in Ref. 1 and defined the dividing surface Rn such that 32 function of equimolar radius is presented in Fig. 7 a for ⌫ ⌫ ˜ =const= 31s, where 31s is the adsorption of gas on the sur- Rn =5.0d, along with a plot of the analytical dependence Eq. face of the particle at the gas spinodal. Since bulk liquid ͑39͒ in Ref. 1 and chemical potential for homogeneous drop- cannot exist below the gas spinodal, the thickness of the let. It can be seen that the chemical potential is a nonmono- liquid film becomes zero at this spinodal and Eq. ͑51͒ has the tonic function of the radius for heterogeneous nucleation but form a monotonic function for homogeneous nucleation. The L 3 3 former result means that two sizes of equilibrium droplet can L 2 Rn ͵ 2␳͑ ͉͒ ͉ ␳ ⌫ ␳ ͑ ͒ form on the foreign particle for ␮ϱ Ͻ␮Ͻ␮ . The smaller r r dr ␮=␮ − 2s = Rn 32 − 1s. 52 th ˜ s2 3 3 Rn droplet represents a stable equilibrium solution, whereas the larger droplet represents an unstable critical droplet. The Another special case is the phase coexistence where liquid ␮ film becomes infinitely thick. In this case, threshold value of chemical potential th is difficult to cal- culate accurately due to the highly unstable nature of the L L3 R3 density profile near the threshold. Figure 7͑a͒ also shows that ͵ r2␳͑r͉͒dr͉ ␳ R2⌫ n ␳ ͑ ͒ ␮=␮ϱ − 2ϱ = n 32 − 2ϱ. 53 ͑ ͒ ␮ ˜R 3 3 analytical Eq. 39 in Ref. 1 overestimates th even for rela- n ˜ tively large particles with radius of Rn =5.0d, indicating that Solving the last two equations gives the solution for the two Eq. ͑39͒ in Ref. 1 is not applicable for particles of smaller ⌫ unknowns Rn and 32. So we have size. The deviation of Eq. ͑39͒ in Ref. 1 from the numerical ␳ results is due mainly to several approximations, including theץ ␳ץ L 1 R3 = ͵ r3ͩͯ ͯ − ͯ ͯ ͪdr. ͑54͒ dependence of the disjoining pressure on curvature ͑which ץ ץ n ␳ ␳ 2ϱ − 1s r ␮ ␮ r ␮ ␮ 0 = ϱ = s1 we will discuss in the next section͒, stronger effect than the Recall that z ϵ˜R −R . We found that z increases as the Tolman effect for the dependence of the surface tension on 0 n n 0 ⌫ * curvature, and dependences of Rn and 32 on the particle particle size decreases. For instance, z0 =2.35d at T ⑀ size. =kBT/ LJ=0.7 in the planar limit, while it becomes z0 ˜ Calculation of free energy of droplet formation requires =2.75d for Rn =5.0d, and z0 =3.31d for microscopic size of ˜ not only the R dependence of chemical potential but also the Rn =3.5d. In Ref. 1 we used planar limiting value of z0 to chemical potential as a function of the number of particles ␯ describe the position of the dividing surface for mesoscopic in the liquid condensate. In the case of heterogeneous nucle- particles. This seems accurate since z0 changes significantly ation this number is given by ⌫ only for microscopic particles. The adsorption 32 can then * 4␲ be calculated for given ˜R .AtT* =k T/⑀ =0.7, ⌫ ␯ ͑ 3 3͒␳ ␲ 2⌫ ͑ ͒ n B LJ 32 = Re − Rn 2 +4 Rn 32, 56 ⌫ ␲ 2 ⌫* ˜ 3 = 32 d /6=0.77 in the planar limit, but 32=0.52 for Rn ⌫* ˜ ␳ ⌫ =5d and 32=0.43 for Rn =3.5d. Thus, the adsorption de- where 2 is the bulk density of liquid and 32 is the adsorp- Ͼ creases with decreasing particle size. tion which, according to our choice, is constant for R Rn.

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-13 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

adsorbed film on the surface of the solid particle. So L 4␲ ␯͑⌫ ͒ ␲ 2⌫ ␲͵ 2͑␳͑ ͒ ␳ ͒ 3␳ 31 =4 Rn 31 =4 r r − 1 dr + Rn 1. 0 3 ͑57͒ ␮Ͻ␮ ␯ ⌫ Thus, for s2, no longer depends on R but on 31 ⌫ ͑ ͒ which changes from zero to 32. Figure 7 b shows typical dependence of chemical potential of the liquid condensate on the number of particles in the liquid film with nucleation center of microscopic size. Again, the dependence is non- monotonic. In Ref. 1 we also discussed the behavior of chemical ␮Ͻ␮ potential for s2 and its relation to the law of adsorption. ͑⌫ ͒ We showed there that the isotherm of adsorption 31 vs p1 is a stepwise monotonic function. The stepwise nature of this isotherm appears due to the oscillatory behavior of the den- sity profile for the adsorbed film near the solid surface, which is slightly different from classical Henry’s law of ad- sorption. Figure 7͑c͒ depicts the reduced chemical potential b ␯ ˜ vs for Rn =5.0d for the case of adsorbed films. The figure also shows the classical logarithmic and analytical relation for mesoscopic particles.1 One can see that all the curves are close to each other. Weak oscillations do appear on the b͑␯͒ curves, as shown before in Ref. 1. The positions of the os- cillations as well as positions of the matching point between logarithmic and structured parts of the chemical potential curve are shifted due to microscopic-particle effects. A key feature of the b͑␯͒ curves ͓Fig. 7͑b͔͒ is the exis- ␯ tence of a maximum bm =bth at certain m. Since it is difficult to determine the threshold value of the chemical potential ␯ and m with high accuracy due to strong instability of density profile, we estimated them by using a parabola relation ͓Eq. ͑56͒ in Ref. 1͔ fitting to the numeric data ͓Fig. 7͑b͔͒. With the parabolic approximation in the near-threshold region, the numbers of molecules in equilibrium and critical droplets are given by ␯ ␯ ͑ ⑀ ͒1/2 e = m − 2 bth/A , ͑58͒ ␯ ␯ ͑ ⑀ ͒1/2 c = m + 2 bth/A , ␯ ͉ 2 ͑␯͒ ␯2 ͉␯ where bth, m, and A= d b /d m are fitted to the chemi- ␯ cal potential curves. The fitting values are bth=0.2283, m ϫ −9 ˜ =10 750, and A=2.745 10 for Rn =5.0d and bth=0.2650, ␯ ϫ −9 ˜ m =7154, and A=4.492 10 for Rn =3.5d.

FIG. 7. ͑a͒ The dependence of the reduced chemical potential of the liquid ͑␮ ␮ ͒ * B. Discussion of disjoining pressure condensate b= − ϱ /kBT on the equimolar radius Re =Re /d of droplets ˜ * ⑀ ͑ ͒ formed on the foreign particle of radius Rn =5.0d at T =kBT/ LJ =0.7. b The discrepancies in predictions of analytical equation The dependence of the reduced chemical potential of the liquid condensate ͓ ͑ ͒ ͑␮ ␮ ͒ ␯ for the chemical potential of the liquid condensate Eq. 39 b= − ϱ /kBT on the number of particles of droplets with two different * ⑀ ͑ ͒ in Ref. 1͔ and numerical results of the present work ͓Fig. sizes of foreign particles at T =kBT/ LJ =0.7. c The dependence of the ͑␮ ␮ ͒ ͑ ͔͒ reduced chemical potential of the liquid condensate b= − ϱ /kBT on the 7 a are likely due to the difference in the disjoining pres- number of particles ␯ of droplets formed on the foreign particle of radius sure for the film on the spherical microscopic particle from ˜ * ⑀ Rn =5.0d at T =kBT/ LJ =0.7. the one on the planar substrate. In addition to volumelike disjoining pressure contribution, Eq. ͑39͒ in Ref. 1 also in- ⌫ ⌫ ␮ ␮ ␮ For Re =Rn, h=0 and 32= 31 at = s2. However, if cludes surface tension contributions. In our discussion of ho- Ͻ␮ ⌫ ⌫ s2, 32 is not well defined and 31 cannot be considered mogeneous droplets we showed that these surfacelike contri- ␮Ͻ␮ as constant. In fact, whenever s2 liquid condensate does butions are significantly different from the ones predicted not exist in the form of the liquid film but in the form of from pure Tolman correction as Eq. ͑39͒ of Ref. 1. This is

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-14 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

FIG. 8. ͑a͒ The reduced Laplacian pressure difference ⌬ * ⌬ ␲ 3 p = p d /6kBT as a function of the equimolar ra- * * dius Re =Re /d of the droplets of different types at T ⑀ ͑ ͒ =kBT/ LJ =0.7. b The reduced disjoining pressure ͑⌸* ⌸␲ 3 ͒ = d /6kBT of different droplets vs the equimo- ͑ * ͒ * ⑀ lar dividing surface Re =Re /d at T =kBT/ LJ =0.7.

another reason why Eq. ͑39͒ overestimates the threshold ior. Such pressure differences only include surfacelike con- value of chemical potential for microscopic droplets. tributions. The difference between the two curves is the To further explain the discrepancies we started by plot- negative disjoining pressure of homogeneous droplet. On the ting the Laplacian pressure difference for bulk phases as a other hand when the radius of the droplet is large enough function of the equimolar radius of the critical droplet. Fig- ͑larger than 15d in our graph͒ all the curves are in agreement ure 8͑a͒ depicts this pressure difference for several cases. For with the classical approximation. small homogeneous droplets, the pressure difference be- Figure 8͑a͒ also shows the bulk pressure difference in tween bulk phases differs from the one predicted by classical the case of heterogeneous nucleation on microscopic par- ͑ ␴ ͒ capillary approximation p2 −p1 =2 21ϱ /R , and the one pre- ticles, which behaves according to the general Laplace equa- ͓ dicted by first-order Tolman curvature correction theory p2 tion ͑3͒. In such a case the pressure difference is a nonmono- ␴ ͑ ␦ ͔͒ ͑ ͒ −p1 =2 21ϱ /R 1− ϱ /R . In the Laplace equation 8 both tonic function of the droplet size, resulting in the appearance the surface tension term and the disjoining pressure term are of the stable equilibrium clusters on the surface of micro- strongly dependent on the curvature. As a result the finite scopic foreign particle. The Laplace equation ͑3͒ was intro- value of the pressure difference takes place in liquid spin- duced so that the surface contribution in this equation reflects odal, while classical theory predicts this value to grow infi- the dependence of the surface tension on the external radius nitely as radius R approaches to zero. Meanwhile, we can of the droplet, but it is independent of the foreign particle plot the difference between normal components of the pres- inside of the droplet. To calculate the disjoining pressure one sure tensor inside of the droplet and outside of the droplet can subtract homogeneous surfacelike contribution from het- ͓Eq. ͑10͔͒. This curve demonstrates a nonmonotonic behav- erogeneous bulk pressure difference.

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-15 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒

The disjoining pressure as a function of the droplet Ref. 1, which is still valid with account of small compress- equimolar radius is plotted in Fig. 8͑b͒, given the size of ibility. However, since we do not have the analytical relation ˜ ˜ b͑R͒ as we had in Ref. 1 for mesoscopic foreign particles, we foreign particles Rn =3.5d and Rn =5.0d. It can be seen that the negative homogeneouslike contribution to disjoining need to perform numerical integration of the b͑␯͒ presented pressure is only significant if the external radius of the drop- in Fig. 7. The resulting free energy of droplet formation W͑␯͒ ˜ ˜ let is less than 2.5d. In the case of microscopic particles for given b1 and two radii Rn =3.5d and Rn =5.0d, respec- considered here the influence of homogeneouslike contribu- tively, are shown in Fig. 9. As expected, all the W͑␯͒ curves tion to disjoining pressure is not strong. Here, we do not exhibit an S-like feature, a signature of the heterogeneous consider microscopic particles with size less than 3d since nucleation. The minima on these curves correspond to the ␯ nucleation would be more like binary nucleation which is equilibrium clusters of size e, for which microscopic solid beyond the framework of the present study. particle is covered by a stable liquid film with finite thick- ͑␯ ͒ In the case of nucleation on mesoscopic particles consid- ness. The value We =W e is negative, indicating the spon- ered in Ref. 1 we replaced the disjoining pressure of the film taneous formation of equilibrium cluster. The absolute value on the spherical particle by the disjoining pressure of the of We depends on the accuracy of numerical integration of ͑ ͒ planar film with thickness h=R−Rn. Figure 8 b also dis- b͑␯͒ for small ␯, where less points on the b͑␯͒ curve are plays this planar disjoining pressure as function of R for both available. Note that We itself is not essential to the kinetics of ˜ ͑ particle sizes considered here. The particle size Rn =5d rep- nucleation, because it is only a reference point constant of resents a borderline case between the mesoscopic and micro- integration͒ for evaluating nucleation barrier. The minima scopic particles. It is acceptable to use the planar approxima- become less deep as the size of the foreign particle de- tion to the disjoining pressure1 for sizes larger than 5d. One creases. It disappears in the limiting case of homogeneous of the reasons is that in the present model the potential of the nucleation, namely, at ␯=0. The maxima on the W͑␯͒ curves ␯ foreign particle depends only on the distance from the par- at c correspond to the critical cluster and give the peak of ticle surface. In addition, the homogeneouslike contribution the barrier. ⌬ ϵ ͑␯ ͒ to the disjoining pressure is almost negligible for larger par- The height of the nucleation barrier W W c ͑␯ ͒ ticles. Hence, the differences between the behavior of the −W e should be less than several tens of kBT to give mea- chemical potential from Eq. ͑39͒ of Ref. 1 and the one shown surable nucleation rates. This case occurs typically at the in Fig. 7͑a͒ are not due to the disjoining pressure but rather near-threshold region as the chemical potential of vapor → ␯ ␯ the surfacelike contributions. Figure 8͑a͒ shows that these b1 bth from below. In this case both e and c are approach- ␯ ͑␯͒ contributions cause the behavior different from the Tolman- ing to the same value m for the maximum on b curve. ␯ type behavior even at R=5d. This problem should not occur The m value depends on the size of the foreign particle, ˜ approaching to zero if nucleation is homogeneous, but it for Rn larger than 5d in Ref. 1. does not depend on the supersaturation of vapor. From Fig. 8͑b͒. one can see that for ˜R =3.5d the disjoin- n It had been shown7 that in the near-threshold region, ing pressure is different from the planar one. For much where the parabolic approximation for the chemical potential thicker films it is lower than the planar value. This result is b͑␯͒ is valid, the nucleation barrier can be written as consistent with the previous study of films on the spherical core particles,35 even though a different Yukawa-type particle potential and a slightly different definition of the disjoining 4 2 1/2 pressure were used in that work. It is more difficult to com- ⌬ ⑀3/2 3/2ͩ ͪ ͑ ͒ W = bth , 59 pare planar and spherical disjoining pressures for thinner 3 A films, since not only the disjoining pressure itself but also the position of the dividing surface R depend on the particle n ⌬␯ ⌬␯ size. and the half-widths e and c of the minimum and maxi- ͑␯͒ All in all, we conclude that the analytical method devel- mum on the W curve can be given by ˜ Ͼ oped in Ref. 1 works fairly well for Rn 5d. It gives reliable results of chemical potential and disjoining pressure. Nu- 2 1/4 merical method is required for the size range within 3d ⌬␯ ⌬␯ ͩ ͪ ͑ ͒ e = c = . 60 Ͻ˜ Ͻ ⑀b A Rn 5d. th Equation ͑59͒ can be used to determine the nucleation barrier height once bth and A are obtained by fitting to the C. Free energy of droplet formation and nucleation b͑␯͒ curves. Since this equation is applicable only to the barrier height near-threshold region, we need to specify certain limits for ⑀ Once the chemical potential of the liquid condensate as parameter. By inspecting b͑␯͒ curves in Fig. 7͑b͒ we found function of the number of molecules b͑␯͒ is known, we can that ⑀ parameter should be at least 0.07 or less in order to use calculate the free energy of droplet formation based on Eqs. the parabolic approximation. The value 0.07 is about the ͑42͒ and ͑43͒. Here the boundary condition W͑0͒=0 in Eq. same as the upper limit we derived for the ⑀ in Ref. 1. That ͑43͒ means that the free energy of droplet formation is zero upper limit is also regulated by the condition similar to Eq. for a bare solid particle. ͑50͒. For the case of heterogeneous nucleation, this condition One can substitute b͑˜␯͒into Eq. ͑43͒ to obtain Eq. ͑69͒ of becomes

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-16 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

͑ ͒ ͑ ͒ FIG. 9. a Formation free energy in units of kBT of ˜ droplets formed on the foreign particle of size Rn =5.0d as function of the number of particles ␯ of the * ⑀ droplet at T =kBT/ LJ =0.7 and different supersatura- ͑ ͒ ͑ ͒ tions. b Formation free energy in units of kBT of ˜ droplets formed on the foreign particle of size Rn =3.5d as function of the number of particles ␯ of the * ⑀ droplet at T =kBT/ LJ =0.7 and different super- saturations.

⌬␯ ӷ VII. CONCLUSIONS e 1, ͑61͒ In this work we have studied two major subjects. The ⌬␯ ӷ c 1, first is the heterogeneous nucleation on microscopic wettable particles and the second is the homogeneous nucleation at high supersaturation. The theoretical approach undertaken is which allow to treat the number of particles as a continuous an extension of the hybrid thermodynamic/DFT method pro- ˜ variable in kinetic theory of nucleation. For given Rn these posed previously to deal with heterogeneous nucleation on conditions are still working very well. The low limit of ⑀ mesoscopic wettable particles. In particular, we extended the parameter is controlled by the validity condition for para- thermodynamic part of the theory in order to describe the 7,8 bolic approximation similar to the condition of Eq. ͑49͒. disjoining pressure and the surface tension of very small 7 For heterogeneous case this conditionͱ requires the height of droplets. the barrier to be larger than 2/3 3 and the ⑀ parameter to be The advantage of the present theoretical approach over ˜ at least larger than 0.005 for the value of Rn considered here. classical one is that it can account for the curvature depen- In summary, the near-threshold region predicted is dence of the surface tension as well as compressibility cor- nearly the same as in Ref. 1. The ⑀ parameter can vary from rections and effects of the disjoining pressure. We studied 0.006–0.007 to about 0.05–0.06. In the latter region one can these corrections based on DFT calculation for Lennard- use Eqs. ͑59͒ and ͑60͒ to evaluate the barrier height and Jones fluid. Even though our conclusions are based on this width. The height ranges from about 2kBT to few tens of kBT. particular model, the physical insights established from this For other values of ⑀ parameter only numerical results ͑from study are more general and can be valid for many other types formation free energy curves in Fig. 9͒ can be used. of model fluids. For both nucleation problems considered

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-17 Homogeneous and heterogeneous nucleation J. Chem. Phys. 125, 144515 ͑2006͒ here, the parameters d/R and d/Rn are no longer small numerical calculation also confirmed that the analytical ap- enough as in the case of mesoscopic nucleation center. The proximation used previously works well for mesoscopic par- absence of these small parameters renders analytical ap- ticles. proach invalid and numerical calculations are required to We obtained several key quantities relevant to kinetics of compare with previous results. Some general conclusions ob- nucleation, such as the free energy of droplet formation and tained from this study are given below. the barrier height to nucleation. We confirmed that in the We examined different contributions to the disjoining case of homogeneous nucleation both LDA and WDA give a pressure of LJ liquid film on the surface of the spherical lower barrier at high supersaturation compared to the predic- particle. We showed that van der Waals interactions provide tion of classical nucleation theory. This lowering of the bar- negative contribution to disjoining pressure. This contribu- rier is due to the overlapping of the surface layer of the small tion which depends on both thickness of the liquid film and droplet and the finite value of the threshold chemical poten- radius of the foreign particle can be accounted for by means tial. In the case of heterogeneous nucleation on microscopic of the power-law approximation if the thickness of the film particles, the formation free energy as a function of droplet exceeds several molecular diameters. The parameters in- size has a minimum which represents a stable equilibrium volved in the power-law approximation are related to the cluster formed on the surface of the foreign particle, as well parameters of intermolecular potential. The van der Waals as a maximum which represents the unstable critical cluster. contribution dominates the disjoining pressure of homoge- The height difference between these two extreme points cor- neous droplets. On the other hand, the presence of the wet- responds to the nucleation barrier. Note that the nucleation table nucleation center provides positive contribution to the rate depends exponentially on the height of nucleation bar- disjoining pressure. This contribution is predominant for me- rier and is also inversely proportional to the barrier width soscopic particles considered in Ref. 1 as well as for micro- ͓e.g., Eqs. ͑83͒–͑85͒ in Ref. 1 for heterogeneous nucleation͔. scopic particles considered here, resulting in the monotonic We also showed that certain simplifications can be made in disjoining pressure as function of the film thickness. We also calculating parameters important to the kinetics of nucleation confirmed that the planar approximation we used previously near the threshold. The results can be used as input param- 33 for mesoscopic particles, including the power-law approxi- eters to compute rates of nucleation at various conditions. mation for thick films and exponential approximation for thin films, is accurate. ACKNOWLEDGMENTS We studied curvature dependence of the surface tension. ͑ In the case of homogeneous droplets we compared the results This work is supported by grants from the DOE DE- ͒ ͑ ͒ from both LDA and WDA within the framework of DFT and FG02-04ER46164 , NSF CHE , and the Nebraska Research found that they both lead to similar predictions. The surface Initiative, and by the John Simon Guggenheim Foundation tension is strongly dependent on the curvature of the droplet, and the Research Computing Facility at University of particularly when droplets are smaller than several molecular Nebraska-Lincoln. diameters and the Tolman length cannot be treated as a con- 1 T. V. Bykov and X. C. Zeng, J. Chem. Phys. 117, 1851 ͑2002͒. stant. The importance regarding definitions of the surface 2 R. Becker and W. Doring, Ann. Phys. 24, 719 ͑1935͒. tension for different dividing surfaces is discussed. We also 3 R. McGraw and A. Laaksonen, Phys. Rev. Lett. 76, 2754 ͑1996͒. discussed the difference between the physical natures of vol- 4 V. Talanquer, J. Chem. Phys. 106,9957͑1997͒. ͑ ͒ ͑ 5 D. Reguera and H. Reiss, Phys. Rev. Lett. 93, 165701 ͑2004͒. umelike the disjoining pressure and surfacelike the surface 6 ͑ ͒ ͒ D. Kashchiev, J. Chem. Phys. 118, 1837 2003 . tension contributions to thermodynamic potentials of the 7 F. M. Kuni, A. K. Shchekin, A. I. Rusanov, and B. Widom, Adv. system. We found that the surfacelike contributions are Interface Sci. 65,71͑1996͒. mostly responsible for deviations from the predictions of 8 F. M. Kuni, A. K. Shchekin, and A. P. Grinin, Usp. Fiz. Nauk 171,345 ͑2001͒. classical nucleation theory for homogeneous nucleation and 9 7 A. K. Shchekin, D. V. Tatianenko, and F. M. Kuni, in Nucleation Theory heterogeneous nucleation on macroscopic particles. and Applications, edited by J. W. P. Schmelzer, G. Ropke, and V. B. We examined the dependence of the chemical potential Priezzhev ͑JINR, Dubna, 1999͒, p. 320. 10 of the liquid condensate on the size of the droplet. We con- T. V. Bykov and X. C. Zeng, J. Phys. Chem. B 105, 11586 ͑2001͒. 11 Y. A. Lei, T. V. Bykov, S. Yoo, and X. C. Zeng, J. Am. Chem. Soc. 127, firmed that this dependence is a monotonic function in the 15346 ͑2005͒. case of homogeneous nucleation, regardless of LDA and 12 K. Koga, X. C. Zeng, and A. K. Shchekin, J. Chem. Phys. 109,4063 WDA. However, the chemical potential dependence shows ͑1998͒. 13 significant deviations from the prediction of the classical A. I. Rusanov, Phasengleichgewichte und Grenzflachenerscheinungen ͑Akademie, Berlin, 1978͒. capillary approximation. If the size of the droplet is less than 14 T. V. Bykov and A. K. Shchekin, Colloid J. 61, 144 ͑1999͒. several molecular diameters the chemical potential is much 15 T. V. Bykov and X. C. Zeng, J. Chem. Phys. 111, 10602 ͑1999͒. lower compared to classical values. In particular, a threshold 16 I. Hadjiagapiou, J. Phys.: Condens. Matter 6, 5303 ͑1994͒. 17 value of chemical potential exists, above which the nucle- R. Tsekov, K. W. Stockelhunber, and B. V. Toshev, Langmuir 16, 3502 ͑2000͒. ation becomes barrierless. In the case of heterogeneous 18 P. Tarazona, Phys. Rev. A 31, 2672 ͑1985͒. nucleation on microscopic foreign particles the chemical po- 19 P. Tarazona, U. M. Marconi, and R. Evans, Mol. Phys. 60,573͑1987͒. tential of the liquid condensate is a nonmonotonic function 20 T. F. Meister and D. M. Kroll, Phys. Rev. A 31, 4055 ͑1985͒. 21 of the droplet size. The maximum chemical potential has also J. R. Henderson, P. Tarazona, F. van Swol, and E. Velasco, J. Chem. Phys. 96, 4633 ͑1992͒. the significance of the threshold value. Numerical calculation 22 S. Dhawan, M. E. Reimel, L. E. Scriven, and H. T. Davis, J. Chem. Phys. is required to obtain accurate value of the threshold. Our 94, 4479 ͑1991͒.

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp 144515-18 T. V. Bykov and X. C. Zeng J. Chem. Phys. 125, 144515 ͑2006͒

23 X. C. Zeng and D. W. Oxtoby, J. Chem. Phys. 94, 4472 ͑1991͒. 30 D. H. Rasmussen, J. Cryst. Growth 56,45͑1982͒. 24 B. V. Derjaguin, N. V. Churaev, and V. M. Muller, Surface forces ͑Con- 31 B. N. Hale, Phys. Rev. A 33,4156͑1986͒. sultant Bureau, New York, 1987͒. 32 F. M. Kuni and A. P. Grinin, Vestn. Leningr. Univ., Ser. 4: Fiz., Khim. 25 C. Hamaker, Physica ͑Amsterdam͒ 1058 ͑1937͒. ͑ ͒ 26 22,10 1982 . 57 ͑ ͒ 33 T. Getta and S. Dietrich, Phys. Rev. E , 655 1998 . F. M. Kuni, A. P. Grinin, and V. B. Kurasov, in Nucleation Theory and 27 R. J. F. Leote de Carvalho, R. Evans, D. C. Hoyle, and J. R. Henderson, Applications, edited by J. W. P. Schmelzer, G. Ropke, and V. B. Priez- J. Phys.: Condens. Matter 6,9275͑1994͒. ͑ ͒ 28 N. E. Carnahan and K. E. Starling, J. Chem. Phys. 51, 635 ͑1969͒. zhev JINR, Dubna, 1999 , p. 120. 29 34 ͑ ͒ A. H. Falls, L. E. Scriven, and H. T. Davis, J. Chem. Phys. 75, 3986 Y. Viisanen and R. Strey, J. Chem. Phys. 101, 7835 1994 . 35 ͑1981͒. I. Napari and A. Laaksonen, J. Chem. Phys. 119, 10363 ͑2003͒.

Downloaded 13 Apr 2007 to 129.93.16.206. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp