2013 Book NucleationTheory
2013 Book NucleationTheory
Volume 860
Founding Editors
W. Beiglböck
J. Ehlers
K. Hepp
H. Weidenmüller
Editorial Board
B.-G. Englert, Singapore, Singapore
U. Frisch, Nice, France
P. Hänggi, Augsburg, Germany
W. Hillebrandt, Garching, Germany
M. Hjorth-Jensen, Oslo, Norway
R. A. L. Jones, Sheffield, UK
H. von Löhneysen, Karlsruhe, Germany
M. S. Longair, Cambridge, UK
M. L. Mangano, Geneva, Switzerland
J.-F. Pinton, Lyon, France
J.-M. Raimond, Paris, France
A. Rubio, Donostia San Sebastian, Spain
M. Salmhofer, Heidelberg, Germany
D. Sornette, Zurich, Switzerland
S. Theisen, Potsdam, Germany
D. Vollhardt, Augsburg, Germany
W. Weise, Munchen, Germany
The series Lecture Notes in Physics (LNP), founded in 1969, reports new
developments in physics research and teaching—quickly and informally, but with
a high quality and the explicit aim to summarize and communicate current
knowledge in an accessible way. Books published in this series are conceived as
bridging material between advanced graduate textbooks and the forefront of
research and to serve three purposes:
• to be a compact and modern up-to-date source of reference on a well-defined
topic
• to serve as an accessible introduction to the field to postgraduate students and
nonspecialist researchers from related areas
• to be a source of advanced teaching material for specialized seminars, courses
and schools
Both monographs and multi-author volumes will be considered for publication.
Edited volumes should, however, consist of a very limited number of contributions
only. Proceedings will not be considered for LNP.
Volumes published in LNP are disseminated both in print and in electronic
formats, the electronic archive being available at springerlink.com. The series
content is indexed, abstracted and referenced by many abstracting and information
services, bibliographic networks, subscription agencies, library networks, and
consortia.
Proposals should be sent to a member of the Editorial Board, or directly to the
managing editor at Springer:
Christian Caron
Springer Heidelberg
Physics Editorial Department I
Tiergartenstrasse 17
69121 Heidelberg/Germany
christian.caron@springer.com
V. I. Kalikmanov
Nucleation Theory
123
Dr. V. I. Kalikmanov
Twister Supersonic Gas Solutions BV
Rijswijk
The Netherlands
and
Faculty of Geosciences
Delft University of Technology
Delft
The Netherlands
One of the most striking phenomena in condensed matter physics is the occurrence
of abrupt transitions in the structure of a substance at certain temperatures or
pressures. These are first-order phase transitions, and examples such as the
freezing of water and the condensation of vapors to form mist in the atmosphere
are familiar in everyday life. A fascinating aspect of these phenomena is that the
conditions at which the transformation takes place can sometimes vary. The
freezing point of water is not always 0 C: the liquid can be supercooled con-
siderably if it is pure enough and treated carefully. Similarly, it is possible to raise
the pressure of a vapor above the so-called saturation vapor pressure, at which
condensation ought to take place according to the thermodynamic properties of the
separate phases. Both these phenomena occur because of the requirement for
nucleation. In practice, the transformation takes place through the creation of small
aggregates, or clusters, of the daughter phase out of the parent phase. In spite of
the familiarity of the phenomena involved, accurate calculation of the rate of
cluster formation for given conditions of the parent phase meets serious difficul-
ties. This is because the properties of the small clusters are insufficiently well
known.
The development from the 1980s onwards of increasingly accurate experimental
measurements of the formation rate of droplets from metastable vapors has driven
renewed interest in the problems of nucleation theory. Existing models, largely
based upon versions of the classical nucleation theory developed in the 1920s–
1940s, have on the whole explained the trends in nucleation behavior correctly, but
have often failed spectacularly to account for this fresh data. The situation is more
dramatic in the case of binary- or, more generally, multi-component nucleation
where the trends predicted by the classical theory can be qualitatively in error
leading to unphysical results.
This book, starting with the classical phenomenological description of nucleation,
gives an overview of recent developments in nucleation theory. It also illustrates
application of these various approaches to experimentally relevant problems
focusing on the nonequilibrium gas–liquid transition, i.e., formation of liquid
v
vi Preface
1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
4 Nucleation Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
4.2 First Nucleation Theorem for Multi-Component Systems . . . . 44
4.3 Second Nucleation Theorem . . . . . . . . . . . . . . . . . . . . . . . . 49
4.4 Nucleation Theorems from Hill’s Thermodynamics
of Small Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 51
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .... 53
vii
viii Contents
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
Symbols
xiii
xiv Symbols
• semi-phenomenological models
The semi-phenomenological approach [16], discussed in Chap. 7, bridges the
microscopic and macroscopic description of nucleation combining statistical
mechanical treatment of clusters and empirical data.
• direct computer simulations
Simulations of nucleation on molecular level by means of Monte Carlo and Mole-
cular Dynamics methods is a technique that complements theoretical and exper-
imental studies and as such may be regarded as a virtual (computer) experiment.
Chapter 8 gives an introduction to molecular simulation methods that are relevant
for modelling of the nucleation process and presents their application to various
nucleation problems.
An important link between theory and nucleation experiment is provided by the so
called nucleation theorems discussed in Chap. 4. Chapter 9 outlines the peculiar fea-
tures of nucleation behavior at deep quenches near the upper limit of metastability.
Chapter 10 is devoted to argon nucleation because of the exceptional role argon plays
in various areas of soft condensed matter physics; here a comparison is presented of
the predictions of theoretical models (outlined in the previous chapters), computer
simulation and available experimental data on argon nucleation. Extensions of the-
oretical models to the case of binary nucleation are discussed in Chaps. 11–13; a
general approach to the multi-component nucleation is outlined in Chap. 14.
Chapters 3–14 refer to homogeneous nucleation (unary, binary, multi-component). If
the nucleation process involves the presence of pre-existing surfaces (foreign bodies,
dust particles, etc.) on which clusters of the new phase are formed, the process is
termed heterogeneous nucleation; it is discussed in Chap. 15.
Though the main aim of the book is to present various theoretical approaches to
nucleation, the general picture would be incomplete without a reference to experi-
mental methods. Chapter 16 gives a short insight into the experimental techniques
used to measure nucleation rates.
References
1. F. Bakhtar, M. Ebrahami, R. Webb, Proc. Instn. Mech. Engrs. 209(C2), 115 (1995)
2. V. Kalikmanov, J. Bruining, M. Betting, D. Smeulders, in SPE Annual Technical Conference
and Exhibition (Anaheim, California, USA, 2007), pp. 11–14. Paper No: SPE 110736
3. P.E. Wagner, G. Vali (eds) Atmospheric Aerosols and Nucleation (Springer, Berlin, 1988)
4. M. Toner, E.G. Cravalho, M. Karel, J. Appl. Phys. 67, 1582 (1990)
5. R. Becker, W. Döring. Ann. Phys. 24, 719 (1935)
6. M. Volmer, Kinetik der Phasenbildung (Steinkopf, Dresden, 1939)
7. Ya. B. Zeldovich, Acta physicochim. URSS 18, 1 (1943)
8. J.E. McDonald, Am. J. Phys. 30, 870 (1962)
9. J.E. McDonald, Am. J. Phys. 31, 31 (1963)
10. H. Reiss, A. Tabazadeh, J. Talbot, J. Chem. Phys. 92, 1266 (1990)
4 1 Introduction
11. H.M. Ellerby, C.L. Weakliem, H. Reiss, J. Chem. Phys. 95, 9209 (1991)
12. H.M. Ellerby, H. Reiss, J. Chem. Phys. 97, 5766 (1992)
13. D. Reguera et al., J. Chem. Phys. 118, 340 (2003)
14. R. Evans, Adv. Phys. 28, 143 (1979)
15. D.W. Oxtoby, R. Evans, J. Chem. Phys. 89, 7521 (1988)
16. V.I. Kalikmanov, J. Chem. Phys. 124, 124505 (2006)
Chapter 2
Some Thermodynamic Aspects
of Two-Phase Systems
Two phases (1 and 2) can coexist of they are in thermal and mechanical equilibrium.
The former implies that there is no heat flux and therefore T1 = T2 , and the latter
implies that there is no mass flux, which yields equal pressures p1 = p2 . However,
this is not sufficient. Let N be the total number of particles in the two-phase system
N = N1 + N2 . The number of particles in either phase can vary while N is kept fixed.
If the whole system is at equilibrium, its total entropy S = S1 + S2 is maximized,
which means in particular that
∂S
=0
∂ N1
∂S1 ∂S2
= (2.1)
∂ N1 ∂ N2
dU p μ
dS = + dV − dN
T T T
we find
∂S μ
=−
∂N T
From (2.1): μ1 /T1 = μ2 /T2 and since T1 = T2 , the chemical potentials of the
coexisting phases must be equal. Hence, two phases in equilibrium at a temperature
T and pressure p must satisfy the equation
μ1 ( p, T ) = μ2 ( p, T ) (2.3)
which implicitly determines the p(T )-phase equilibrium curve. Thus, T and p cannot
be fixed independently, but have to provide for equality of the chemical potentials
of the two phases. Differentiating this equation with respect to the temperature and
bearing in mind that p = p(T ), we obtain:
S dT − V d p + N dμ = 0 (2.5)
we find
dμ = −s dT + v d p (2.6)
where s = S
N and v =
V
N are entropy and volume per particle, implying that
∂μ ∂μ
= −s, =v
∂T p ∂p T
Using (2.4), we obtain the Clapeyron equation describing the shape of the ( p, T )-
equilibrium curve:
dp s 2 − s1
= (2.7)
dT v2 − v1
In the case of vapor-liquid equilibrium this curve is called a saturation line and the
pressure of the vapor in equilibrium with its liquid is called a saturation pressure,
psat . Liquid and vapor coexist along the saturation line connecting the triple point
corresponding to three-phase coexistence (solid, liquid and vapor) and the critical
point. Below the critical temperature Tc one can discriminate between liquid and
vapor by measuring their density. At Tc the difference between them disappears. The
line of liquid-solid coexistence has no critical point and goes to infinity since the
difference between the symmetric solid phase and the asymmetric liquid phase can
not disappear.
2.1 Bulk Equilibrium Properties 7
δ Q = δU + pδV (2.8)
In the theory of phase transitions the quantity δ Q is called the latent heat L. For
processes at constant pressure the latent heat is given by the change in the enthalpy:
l = T (s2 − s1 ) (2.9)
dp l
= (2.10)
dT T (v2 − v1 )
For the gas–liquid transition at temperatures far from Tc the molecular volume in the
liquid phase v1 ≡ vl is much smaller than in the vapor v2 ≡ vv . Neglecting v1 and
applying the ideal gas equation for the vapor
p v2 = kB T (2.11)
where
kB = 1.38 × 10−16 erg/K (2.13)
is the Boltzmann constant. Considering the specific latent heat to be constant, which is
usually true for a wide range of temperatures and various substances,1 and integrating
(2.12) over the temperature, we obtain:
1
psat (T ) = p∞ e−β l , β = (2.14)
kB T
where p∞ is a constant.
1 For example, for water in the temperature interval between 0 and 100 ◦ C, l changes by only 10 %.
8 2 Some Thermodynamic Aspects of Two-Phase Systems
Let us discuss the interface between the two bulk phases in equilibrium. For con-
creteness we refer to the coexistence of a liquid with its saturated vapor at the
temperature T . Equilibrium conditions are characterized by equality of temperature,
pressure, and chemical potentials in both bulk phases. The density, however, is not
constant but varies continuously along the interface between two bulk equilibrium
values ρ v (T ) and ρ l (T ). Note, that local fluctuations of density take place even in
homogeneous fluid, where, however, they are small and short-range. In the two-
phase system these fluctuations are macroscopic: for vapor–liquid systems at low
temperatures the bulk densities ρ v and ρ l can differ by 3–4 orders of magnitude.
dF = − p dV − S dT + γ d A + μ dN (2.15)
dG = V d p − S dT + γ d A + μ dN (2.16)
dΩ = − p dV − S dT + γ d A − N dμ (2.17)
(in (2.17) N is the average number of particles in the system). The coefficient γ is the
surface tension; its thermodynamic definition follows from the above expressions:
Mv
exc
M
dividing surface
bulk liquid Vl
l
M
∂F
γ = (2.18)
∂A N ,V,T
∂G
γ = (2.19)
∂A N , p,T
∂Ω
γ = (2.20)
∂A μ,V,T
Vv + Vl = V
The idea of Gibbs was that any extensive thermodynamic quantity M (the number
of particles, energy, entropy, etc.) can be written as a sum of bulk contributions M v
and M l and an excess contribution M exc that is assigned to the chosen dividing
surface:
M = M v + M l + M exc (2.21)
Equation (2.21) is in fact a definition of M exc ; its value depends on the location of
the dividing surface, and so do the values of M v and M l (as opposed to M , which
is an actual physical property and as such can not depend on the location of the Gibbs
surface). Several important examples are
N = N v + N l + N exc (2.22)
S = S v + S l + S exc
Ω = Ω v + Ω l + Ω exc
F = F v + F l + F exc
V = Vv + Vl
10 2 Some Thermodynamic Aspects of Two-Phase Systems
By definition the dividing surface has a zero width implying that V exc = 0. Since
the location of the dividing surface is arbitrary, the excess quantities accumulated on
it can be both positive or negative.
One special case that will be useful for future discussions is the equimolar surface
defined through the requirement N exc = 0. The surface density of this quantity
N exc
Γ = (2.23)
A
is called adsorption. Thus, the equimolar surface corresponds to zero adsorption. The
thermodynamic potentials, such as F , Ω, G, are homogeneous functions of the first
order with respect to their extensive variables. We can derive their expressions for
the two-phase system by integrating Eqs. (2.15)–(2.17) using Euler’s theorem for
homogeneous functions (see e.g. [1], Sect. 1.4). In particular, integration of (2.17)
results in
Ω = −p V + γ A (2.24)
Ω exc = γ A (2.25)
irrespective of the choice of the dividing surface. Independence of Ω exc on the loca-
tion of a dividing surface gives rise to the most convenient thermodynamic route
for determination of the surface tension. Equation (2.25) is used in density func-
tional theories of fluids (discussed in Chap. 5) to determine γ from the form of the
intermolecular potential.
By definition
Ω exc = Ω − Ω v − Ω l (2.26)
dΩ v = − p dV v − S v dT − N v dμ (2.27)
dΩ l = − p dV l − S l dT − N l dμ (2.28)
dΩ exc = γ d A + A dγ
2.2 Thermodynamics of the Interface 11
describing the change of the surface tension resulting from the changes in T and μ.
An important consequence of (2.29) is the expression for adsorption:
∂γ
Γ =− (2.30)
∂μ T
4π 3 4π 3
Vl = R , Vv = V − R , A = 4π R 2
3 3
A sketch of a spherical interface is shown in Fig. 2.3. The change of the Helmholtz
free energy F of the two-phase system “droplet + vapor” when its variables change
at isothermal conditions is given by [3]:
dγ
(dF )T = − p (dV )T − p (dV )T +μ(dN )T +γ (d A)T + A
l l v v
(d R)T (2.31)
dR
Rv
zs
Re
ze ~ξ
Rs
l
R
Fig. 2.3 Sketch of a spherical interface. The z axis is perpendicular to the interface pointing away
from the center of curvature. Re and Rs ≡ Rt denote, respectively, the location of the equimolar
surface and the surface of tension (see the text). The width of the transition zone between bulk vapor
and bulk liquid is of the order of the correlation length ξ (Reprinted with permission from Ref. [1],
copyright (2001), Springer-Verlag.)
dividing surface of the radius R; the term in the square brackets gives the change of
γ with respect to a mathematical displacement of the dividing surface. It is impor-
tant to stress that the physical quantities F , p v , p l , μ, N , V , do not depend on the
location of a dividing surface. So they remain unchanged when only R is changed
and from (2.31)
dγ
0 = [dF ] = −Δp 4π R 2 [d R] + 8π R γ [d R] + 4π R 2 [d R]
dR
It is clear that since Δp, as a physical property of the system, is independent of R, the
surface tension must depend on the choice of dividing surface. A particular choice
R = Rt , such that
dγ
= 0, (2.33)
d R R=Rt
corresponds to the so-called surface of tension; it converts (2.32) into the standard
Laplace equation
2γt
Δp = (2.34)
Rt
2.2 Thermodynamics of the Interface 13
where γt = γ [Rt ]. One can relate the surface tension taken at an arbitrary dividing
surface of a radius R to γt . To this end let us write (2.32) in the form
d
Δp R = 2
R 2 γ [R]
dR
and integrate it from Rt to R. Using (2.34) for Δp we obtain the Ono-Kondo equation
[4]
R 1 1 2
γ [R] = γt f ok , with f ok = 2
+ x (2.35)
Rt 3 x 3
When R differs from Rt by a small value, γ [R] remains constant to within terms of
order 1/Rt2 .
Among various dividing surfaces we distinguished two special cases—the equimolar
surface Re and the surface of tension Rt —which are related to the certain physi-
cal properties of the system. Let us introduce a quantity describing the separation
between them
δ = Re − Rt
δT = lim δ = z e − z t (2.37)
Rt →∞
is called the Tolman length. Its sign can be both positive and negative depending on
the relative location of the two dividing surfaces. By definition δT does not depend
on either radius Rt , or Re (whereas δ does) but can depend on the temperature. Both
dividing surfaces lie in the interfacial zone implying that δT is of the order of the
correlation length. Let Γt be the adsorption at the surface of tension. From the Gibbs
adsorption equation
∂γt
Γt = −
∂μ T
Using the thermodynamic relationship (2.6) in both phases we rewrite this result as
d pv d pl
dγt = −Γt dμ = −Γt = −Γt l
ρ v ρ
14 2 Some Thermodynamic Aspects of Two-Phase Systems
ρl
d pl = d pv
ρv
resulting in
d(Δp) = Δρ dμ
For a curved surface Tolman [5] showed (see also [3]) that
Γt δT 1 δT 2
= δT 1+ + (2.39)
Δρ Rt 3 Rt2
but the terms in (δT /Rt ) and (δT /Rt )2 can be omitted to the order of accuracy we
need. This means that in all derivations below we need to keep only the linear terms
in δT . With this in mind Eqs. (2.38) and (2.39) give
2γt
dγt = −δT d
Rt
γt Rt
= (2.40)
γ∞ Rt + 2δT
where γ∞ is the planar surface tension discussed in Sect. 2.2.1. Keeping the linear
term in δT we finally obtain the Tolman equation
2δT
γ t = γ∞ 1− + ... (2.41)
Rt
It is important to emphasize that the second order term in the δT can not be obtained
from (2.40) since this equation is derived to within the linear accuracy in the Tolman
length.
2.2 Thermodynamics of the Interface 15
Equation (2.41) represents the expansion of the surface tension of a curved interface
(droplet) in powers of the curvature. Its looses its validity when the radius of the
droplet becomes of the order of molecular sizes. The concept of a curvature dependent
surface tension frequently emerges in nucleation studies. It is therefore important
to estimate the minimal size of the droplet for which the Tolman equation holds.
For simple fluids (characterized by the Lennard-Jones and Yukawa intermolecular
potentials) near their triple points the density functional calculations [6] reveal that
the Tolman equation is valid for droplets containing more than 106 molecules.
References
1. V.I. Kalikmanov, in Statistical Physics of Fluids. Basic Concepts and Applications (Springer,
Berlin, 2001)
2. J.W. Gibbs, in The Scientific Papers (Ox Bow, Woodbridge, 1993)
3. J.S. Rowlinson, B. Widom, Molecular Theory of Capillarity (Clarendon Press, Oxford, 1982)
4. S. Ono, S. Kondo, in Encyclopedia of Physics, vol. 10, ed. by S. Flugge (Springer, Berlin, 1960),
p. 134
5. R.C. Tolman, J. Chem. Phys. 17(118), 333 (1949)
6. K. Koga, X.C. Zeng, A.K. Schekin, J. Chem. Phys. 109, 4063 (1998)
Chapter 3
Classical Nucleation Theory
Nucleation refers to the situation when a system (parent phase) is put into a nonequi-
librium metastable state. Experimentally it can be achieved by a number of ways (for
definiteness we refer to the vapor-liquid transition): e.g. by isothermally compress-
ing vapor up to a pressure p v exceeding the saturation vapor pressure at the given
temperature psat (T ). At this state, characterized by p v and T , the chemical potential
in the bulk liquid μl ( p v , T ) is lower than in the bulk vapor at the same conditions
μv ( p v , T ), which makes it thermodynamically favorable to perform a transformation
from the parent phase (vapor) to the daughter phase (liquid). The driving thermody-
namic force for this transformation is the chemical potential difference
Δμ = μv ( p v , T ) − μl ( p v , T ) > 0 (3.1)
Free energy
metastable state A (super-
barrier
saturated vapor). The global
minimum corresponds to the
stable state B (bulk liquid).
Transition between the states
A and B involves overcoming
the free energy barrier
metastable stable
state A state B
1
Δμ ≈ [μv ( p v , T ) − μsat (T )] − ( p v − psat ) (3.2)
ρsat
l
where ρsat
l is the liquid number density at saturation. Using the compressibility factor
we present Δμ as
pv
Δμ = kB T ln S − kB T Z sat
l
−1
psat
and using Eq. (2.6) for the vapor phase, the supersaturation can be approximated as
pv
S= (3.5)
psat
so that
Δμ = kB T ln S − kB T Z sat
l
(S − 1) (3.6)
Δμ = kB T ln S (3.7)
3.2 Thermodynamics
If the lifetime of the metastable state is much larger than the relaxation time necessary
for the system to settle in this state, we can apply the concept of quasi-equilibrium
treating the metastable state as if it were an equilibrium. Instead of a thermodynamic
probability of occurrence of a nucleus we shall discuss the “equilibrium” (in the
above sense) distribution function of n-clusters, ρeq (n), which is proportional to it.
Considerations based on the thermodynamic fluctuation theory [1] yield:
Wmin (n)
ρeq (n) = ρ1 exp − (3.8)
kB T
Uf = Ufv + U l + U exc
20 3 Classical Nucleation Theory
where the first term refers to the bulk vapor, the second term refers to the bulk liquid
and the third one gives the surface (excess) contribution. The total change in the
internal energy ΔU = Uf − U0 , caused by the change in the physical state, includes:
• the work W exerted on the system by an external source (creating pressure);
• the work performed by the heat bath to create a droplet, and
• the heat received by the system from the heat bath.
The heat bath is considered to be large enough so that its pressure pr and temperature
Tr remain constant (quantities referring to the heat bath are denoted with a subscript
“r ”). The work performed by the heat bath is pr ΔVr , and the heat given by it is
−Tr ΔSr . Thus,
ΔU = W + pr ΔVr − Tr ΔSr (3.9)
The total volume of heat bath and the system remains unchanged, ΔVr = −ΔV .
According to the second law of thermodynamics the total change of entropy (heat
bath + system) should be nonnegative:
ΔSr + ΔS ≥ 0 (3.10)
W ≥ ΔU − Tr ΔS + pr ΔV
Equality sign corresponds to the reversible process (in which the total entropy remains
unchanged) yielding the minimum work
Wmin = ΔU − Tr ΔS + pr ΔV (3.11)
If furthermore we assume that the process of transformation to the final state (i.e.
the droplet formation) takes place at a constant temperature T = Tr and pressure
p v = pr then
Wmin = ΔU − T ΔS + p v ΔV = ΔG (3.12)
U vf = T S fv − p v V fv + μv N vf (3.15)
U l = T S l − p l V l + μl N l (3.16)
U exc = T S exc + γ (R) A(R) + μexc N exc (3.17)
Since the vapor pressure and temperature are constant the vapor chemical potential
μv ( p v , T ) does not change during the transformation of vapor from the initial to the
final state. In the last expression γ (R) is a surface tension at the dividing surface R
with a surface area A(R). The change of the system volume is
ΔV = V fv + V l − V0v (3.18)
Substituting (3.13)–(3.18) into Eq. (3.12) and taking into account conservation of the
number of molecules N0v = N vf + N l + N exc , we obtain
ΔG = ( p v − p l )V l +γ A + N l μl ( p l ) − μv ( p v ) + N exc μexc − μv ( p v ) (3.19)
This is an exact general expression for the Gibbs free energy of cluster formation.
The location of a dividing surface is not specified. Choosing the equimolar surface
Re (with γ = γe , A = Ae ), the last term in (3.19) vanishes. Considering the liquid
phase to be incompressible we write
μl ( p l ) = μl ( p v ) + vl ( p l − p v ) (3.20)
ΔG = −n Δμ + γe Ae (3.21)
ΔG = −n Δμ + γ∞ A(n) (3.22)
rn = r l n 1/3
and 1/3
3 vl
r =
l
(3.23)
4π
A(n) = s1 n 2/3
where 2/3
s1 = (36π )1/3 vl (3.24)
is the “surface area of a monomer”. Using (3.7) the reduced free energy of cluster
formation reads:
βΔG(n) = −n ln S + θ∞ n 2/3 (3.25)
The function ΔG(n) is schematically shown in Fig. 3.2. At small n the positive
surface term dominates making it energetically unfavorable to create a very small
droplet in view of the large uncompensated surface energy. For large n the negative
bulk contribution prevails. The function has a maximum at
3
2 θ∞
nc = (3.27)
3 ln S
1 16π (vl )2 γ∞
3
ΔG ∗ = γ∞ A(n c ) = (3.28)
3 3 (kB T ln S)2
3.2 Thermodynamics 23
G
0
the new bulk phase. Within the
critical region, characterized
by ΔG(n c ) − ΔG(n) ≤ kB T ,
the fluctuation development
of clusters occurs
nc n
4 θ∞ 3
βΔG ∗ = (3.29)
27 (ln S)2
The critical cluster is in metastable (quasi-) equilibrium with the surrounding vapor
yielding
μv ( p v ) = μl ( p l )
This equality together with Eq. (3.20) leads to an alternative form of the nucleation
barrier:
16π γ∞ 3
ΔG ∗ = (3.30)
3 (Δp)2
where Δp = p l − p v ; note that p l refers to the bulk liquid held at the same temperature
T and the same chemical potential μ as the supersaturated vapor.
Maximum of ΔG(n) corresponds to the exponentially sharp minimum of the distri-
bution function (3.8). Therefore instead of speaking about the critical point n = n c
it would be more correct to discuss the critical region around n c , where ΔG(n) to a
good approximation has the parabolic form:
1 d
ΔG(n) ≈ ΔG ∗ + ΔG (n c )(n − n c )2 , = (3.31)
2 dn
(the term linear in (n − n c ) vanishes). This quadratic expansion yields the Gaussian
form for ρeq (n), centered at n c and having the width
−1/2
1
Δ = − βΔG (n c ) (3.32)
2
24 3 Classical Nucleation Theory
It follows from (3.31) and (3.32) that the critical region corresponds to the clusters
satisfying
ΔG(n c ) − ΔG(n) ≤ kB T
Recall that the average free energy associated with an independent fluctuation in a
fluid is of order kB T . Therefore, the above expression indicates that the clusters in
the critical region fall within the typical fluctuation range around the critical cluster.
As a result fluctuation development of nuclei in the domain of cluster sizes (n c , n c +
Δ) may with an appreciable probability bring them back to the subcritical region but
nuclei which passed the critical region will irreversibly develop into the new phase.
is limited by the stage before the actual phase transition and therefore it can not
predict the development of this process. At large n, the Gibbs formation energy is
dominated by the negative bulk contribution −nΔμ implying that for large clusters
ρeq (n) diverges
ρeq → ∞ as n→∞
The true number density ρ(n, t) (nonequilibrium cluster distribution), as any other
physical quantity, should remain finite for any n and at any moment of time t. To
determine ρ(n, t) it is necessary to discuss kinetics of nucleation. Within the CNT
the following assumptions are made:
• the elementary process which changes the size of a nucleus is the attachment to it
or loss by it of one molecule
• if a monomer collides a cluster it sticks to it with probability unity
• there is no correlation between successive events that change the number of par-
ticles in a cluster.
The last assumptions means that nucleation is a Markov process. Its schematic illus-
tration is presented in Fig. 3.3. Let f (n) be a forward rate of attachment of a molecule
to an n-cluster (condensation) as a result of which it becomes an (n + 1)-cluster, and
b(n) be a backward rate corresponding to loss of a molecule by an n-cluster (evap-
oration) as a result of which it becomes and (n − 1)-cluster. Then the kinetics of the
nucleation process is described by the set of coupled rate equations
3.3 Kinetics and Steady-State Nucleation Rate 25
f(n-1) f(n)
n-1 n n+1
b(n) b(n+1)
∂ρ(n, t)
= f (n − 1) ρ(n − 1, t) − b(n) ρ(n, t) − f (n) ρ(n, t) + b(n + 1) ρ(n + 1, t)
∂t
(3.34)
A net rate at which n-clusters become (n + 1)-clusters is defined as
implying that
∂ρ(n, t)
= J (n − 1, t) − J (n, t) (3.36)
∂t
The set of equations (3.36) for various n with J (n, t) given by (3.35) was proposed
by Becker and Döring [3] and is called the Becker-Döring equations. The expression
for the forward rate f (n) depends on the nature of the phase transition. For gas-to-
liquid transition f (n) is determined by the rate of collisions of gas monomers with
the surface of the cluster
f (n) = ν A(n) (3.37)
Here the monomer flux to the unit surface ν, called an impingement rate, is found
from the gas kinetics [1]:
pv
ν=√ , (3.38)
2π m 1 kB T
ρeq (n)
b(n + 1) = f (n) (3.39)
ρeq (n + 1)
26 3 Classical Nucleation Theory
1 ρ(n, t) ρ(n + 1, t)
J (n, t) = − (3.40)
f (n) ρeq (n) ρeq (n) ρeq (n + 1)
The kinetic process described by Eq. (3.34) rapidly reaches a steady state: a character-
istic relaxation time, τtr , is usually ∼1 μs (we briefly discuss the transient nucleation
behavior in Sect. 3.8) which is much smaller than a typical experimental time scale.
In the steady nonequilibrium state the number densities of the clusters no longer
depend on time. This implies that all fluxes are equal
The flux J , called the steady-state nucleation rate, is a number of nuclei (of any
size) formed per unit volume per unit time. Summation of both sides of Eq. (3.40)
from n = 1 to a sufficiently large N yields (due to mutual cancelation of successive
terms):
N
1 ρ(1) ρ(N + 1)
J = − (3.41)
f (n) ρeq (n) ρeq (1) ρeq (N + 1)
n=1
For small clusters the free energy barrier is dominated by the positive surface con-
tribution θ n 2/3 , implying that the number of small clusters continues to have its
equilibrium value in spite of the constant depletion by the flux J :
ρ(n)
→1 as n → 1+ (3.42)
ρeq (n)
For large n the forward rate exceeds the reverse: the system evolves into the new,
thermodynamically stable, phase. As n grows ρeq (n) increases without limit whereas
the true distribution ρ(n) remains finite. Thus,
ρ(n)
→0 as n → ∞ (3.43)
ρeq (n)
By choosing large enough N we can neglect the second term in (3.41). Extending
summation to infinity we rewrite it as
∞
−1
1
J= (3.44)
f (n) ρeq (n)
n=1
Examine the terms of this series. The cluster distribution at constrained equilibrium
is given by Eqs. (3.33) and (3.25)
In this form it was first discussed by Frenkel [8, 9] and is called the Frenkel distri-
bution. Initially (for small n) ρeq (n) decreases due to the surface contribution, then
reaches a minimum at n = n c beyond which it exponentially grows due to the bulk
contribution S n .
The major contribution to the series (3.44) comes from the terms in the vicinity of
n c . If n c is large enough (and the validity of CNT requires large n c ) the number of
these terms is large while the difference between the successive terms for n and n + 1
is small. This makes it possible to replace summation by an integral:
∞ −1
1
J= dn (3.46)
1 f (n)ρeq (n)
The integral can be calculated using the steepest descent method if one takes into
account the sharp exponential minimum of ρeq at n c . Expanding ρeq about n c we
write:
1 1
ρeq (n) ≈ ρeq (n c ) exp − ΔG (n c )(n − n c ) , ΔG (n c ) < 0
2
2 kB T
J = Z f (n c )ρeq (n c ) (3.47)
where
1 1
Z = − ΔG (n c ) (3.48)
2π kB T
is called the Zeldovich factor. Comparison of (3.48) with (3.32) shows that Z is
inversely proportional to the width of the critical region:
1
Δ= √ (3.49)
πZ
Using (3.48) and (3.25) the Zeldovich factor takes the form:
1 θ∞ −2/3
Z = nc (3.50)
3 π
or equivalently
γ∞ 1
Z = (3.51)
kB T 2πρ l rc2
1/3
where rc = r l n c is the radius of the critical cluster.
Summarizing, the main result of the CNT states that the steady state nucleation rate
is an exponential function of the energy barrier
28 3 Classical Nucleation Theory
ΔG ∗
J = J0 exp − (3.52)
kB T
16π (vl )2 γ∞
3
ΔG ∗ = (3.53)
3 |Δμ|2
J0 = Z ν A(n c ) ρ v (3.54)
It is important to realize that the steady state flux J does not depend on size, but is
observed at the critical cluster size.
Nucleation of water vapor plays an exceptionally important role in a number of
environmental processes and industrial applications. That is why water can be cho-
sen as a first example to analyze the predictions of the CNT. Two experimental
groups, Wölk et al. [10] and Labetski et al. [11] reported the results of nucleation
experiments for water in helium (as a carrier gas) in the wide temperature range:
220–260 K in [10] and 200–235 K in [11]. Though, the two groups used differ-
ent experimental setups—an expansion chamber in [10] and a shock wave tube in
[11]—the results obtained are consistent. Figure 3.4 shows the relative—experiment
to theory (CNT)—nucleation rate of water for the temperature range 200 < T <
260 K. Circles correspond to the experiment of [10], squares—to the experiment
of [11]. The thermodynamic data for water are given in Appendix A. Figure 3.4
demonstrates a clear trend: the CNT underestimates the experiment (up to 4 orders
of magnitude) for lower temperatures, and slightly overestimates it for higher tem-
peratures. The dashed line (“ideal line”) corresponds to Jexp = Jcnt ; the predictions
of CNT coincide with experiment for water at temperatures around 240 K. These
results indicate a general feature of the CNT: while its predictions of the nucleation
rate dependence on S are quantitatively correct, the dependence of J on the tempera-
ture are in many cases in error—as illustrated by the long-dashed line in Fig. 3.4. The
3.3 Kinetics and Steady-State Nucleation Rate 29
-2
200 220 240 260
T (K)
discrepancy between the CNT and experiment grows as the temperature decreases.
This is not surprising since the critical cluster at 200 < T < 220 K at experimen-
tal conditions (supersaturation) contains only ≈15–20 molecules as follows from
Eq. (3.27), implying that the purely phenomenological approach, based on the capil-
larity approximation becomes fundamentally in error. In the next chapters we discuss
alternative models of nucleation which do not invoke the capillarity approximation.
Consider in a more detail the metastable equilibrium between the liquid droplet
of a radius R and the surrounding supersaturated vapor at the pressure p v and the
temperature T . We characterize the droplet radius by its value at the surface of tension
R = Rt . Equilibrium implies that the chemical potentials of a molecule outside and
inside the droplet are equal:
μvR ( p v ) = μlR ( p l ) (3.56)
Here p l is the pressure inside the cluster. The same expression written for the bulk
equilibrium at the temperature T yields
In this case the pressure in both phases is equal to the saturation pressure: psat (T ).
Note that Eq. (3.57) can be viewed as an asymptotic form of (3.56) when the droplet
radius R → ∞. Subtracting (3.56) from (3.57) and using the thermodynamic rela-
tionship (2.6) we obtain
30 3 Classical Nucleation Theory
pv 1
dp = μlR ( p l ) − μsat (3.58)
psat ρ( p)
For temperatures not close to Tc the vapor density on the left-hand side can be written
in the ideal gas form resulting in
pv
kB T ln = μlR ( p l ) − μsat (3.59)
psat
Within the capillarity approximation using the thermodynamic relationship (2.6) for
the liquid phase, the right-hand side of Eq. (3.59) becomes
l
For low temperatures the liquid compressibility factor at saturation Z sat 1 and the
last term in (3.60) can be safely neglected yielding
2γ∞ vl
p = psat exp
v
(3.61)
k B T Rt
This result, called the Kelvin equation, was formulated by Sir William Thomson
(Lord Kelvin) in 1871 [12]. It relates the vapor pressure, p v , over a spherical liquid
drop to its radius.
In the original Kelvin’s work nucleation was not discussed. Later on the same equa-
tion naturally appeared in the formulation of the CNT since the critical cluster is
in the metastable equilibrium with the surrounding vapor which can be expressed
in the form of Eq. (3.56). At the same time in nucleation theory this equilibrium
corresponds to the maximum of the Gibbs energy of cluster formation
μvR ( p v ) = μlR ( p l ) ⇔ maxn ΔG(n) (3.62)
Thus, the alternative way to derive the Kelvin equation is to maximize the free energy
of cluster formation. Using this equivalent formulation and applying the capillarity
approximation to ΔG(n), we derive the classical Kelvin equation (3.61). The latter
has long played a very important role in nucleation theory (for a detailed discussion
see [13]). Since the radius of a spherical n-cluster is R = r l n 1/3 , we can rewrite
it in the form containing only dimensionless quantities which is more suitable for
nucleation studies:
3.4 Kelvin Equation 31
3
2 θ∞
nc = (3.63)
3 ln S
To find the evaporation rates CNT uses the concept of constrained equilibrium, which
would exist for a vapor at the same temperature T and supersaturation S > 1 as the
vapor in question. Such a fictitious state can be achieved by introducing “Maxwell
demons” which ensure that monomers are continuously replenished by artificial
dissociation of clusters which grow beyond a certain (critical) size. The necessity
of such an artificial construction clearly follows the fact that a supersaturated vapor
is not a true equilibrium state. In an alternative procedure formulated by Katz and
coworkers [14, 15] and called a “kinetic theory of nucleation”, the evaporation rate
is obtained from the detailed balance condition at the (true) stable equilibrium of
the saturated vapor at the same temperature T . Within this procedure no artificial
construction is needed.
Assuming, as in the CNT, that b(n) is independent of the gas pressure, we apply
the detailed balance condition at the saturation point, where J = 0 and the cluster
distribution ρsat (n) is independent of time. From (3.35) this results in
ρsat (n)
b(n + 1) = f sat (n) (3.64)
ρsat (n + 1)
Since the chemical potentials of liquid and vapor at saturation are equal, the
Gibbs formation energy of a cluster at saturation contains only the positive surface
contribution
sat (n) = γ∞ s1 n
ΔG CNT 2/3
(3.65)
implying that
ρsat (n) = ρsat
v
exp[−θ∞ n 2/3 ] (3.66)
Let us divide both sides of Eq. (3.35) by f (n)ρsat (n)S n . Since the forward rate is
proportional to the pressure, we have:
N −1
J (n, t) ρ(N , t)
=1− N (3.67)
f (n) ρsat (n)S n S ρsat (N )
n=1
Examine the second term on the right-hand side for large N . In the nominator ρ(N , t)
is limited for any N (as any other physical quantity). In the denominator the first
term exponentially diverges as e N ln S while the second term vanishes exponentially,
but slower than the first one—see (3.66). As a whole the last term on the right-hand
side becomes asymptotically small as N → ∞. Extending summation to infinity we
obtain for the steady state nucleation rate:
∞
−1
1
J= (3.68)
f (n) S n ρsat (n)
n=1
This result looks almost similar to the CNT expression (3.46). The difference between
the two expressions is in the prefactor of the cluster distribution function. Nucleation
rates given by the kinetic approach, Jkin,phen and by the CNT, JCNT , differ by a factor
1/S known as the Courtney correction [16]:
1
Jkin,phen = JCNT (3.69)
S
Both theories yield the same critical cluster size. Thus, if in the kinetic approach one
uses the same as in the CNT (phenomenological) model for ΔG, two approaches
become identical in all respects except for the 1/S correction in the prefactor J0 .
This conclusion, however, looses its validity if within the kinetic approach a differ-
ent expression for ΔG is chosen. It gives rise to the different form of the cluster
distribution function. The importance of the kinetic approach is, thus, in setting the
methodological framework for nucleation models with other than classical forms of
the Gibbs formation energy.
The equilibrium Frenkel distribution employed in the CNT, based on the capillarity
approximation, has the form (3.45)
pv ρ1
S= = v (3.71)
psat (T ) ρsat (T )
where ρsat
v (T ) is the monomer concentration of the vapor at saturation. The law of
mass action (see e.g. [17], Chap. 6) written for the “chemical reaction” of formation
of the n-cluster E n from n monomers E 1
n E1 En (3.72)
states that the equilibrium cluster distribution function should have the form:
where K n (T ) is the equilibrium constant for the reaction (3.72) which can depend
on n and T but can not depend on ρ1 , or equivalently on the actual pressure p v . The
Frenkel distribution does not satisfy this requirement:
n
ρeq (n) 1
= (ρ1 )1−n S n e−θ∞ n = ρ1 e−θ∞ (T ) n
2/3 2/3
(ρ1 ) n ρsat (T )
v
Katz’s kinetic approach replaces the Frenkel distribution (3.70) by the Courtney
distribution
ρeq (n) = ρsat
v
exp[n ln S − θ∞ n 2/3 ] (3.74)
(T ))1−n e−θ∞ (T ) n
2/3
K n (T ) = (ρsat
v
satisfies the law of mass action. Weakliem and Reiss [18] showed that the Courtney
distribution is not unique but represents one of the possible corrections to the Frenkel
distribution which converts it to a form compatible with the law of mass action.
Although the Courtney distribution satisfies the law of mass action, it does not satisfy
the limiting consistency requirement [13]: in the limit n → 1 it does not return the
identity
ρ1 = ρ1 , for n = 1
The same refers to the Frenkel distribution. At the same time, the limiting consistency
is not a fundamental property, to which a cluster distribution should obey, but rather a
mathematical convenience to have a single formula which could be valid for all cluster
sizes [13]. However, the fact that the CNT does not satisfy the limiting consistency can
not be considered as its “weakness”, since CNT is valid for relatively large clusters
which can be treated as macroscopic objects. However, for nucleation models which
are constructed to be valid for small clusters the requirement of limiting consistency
deserves special attention.
34 3 Classical Nucleation Theory
Nucleation and growth of clusters can be viewed as the flow in the space of cluster
sizes. This space is one-dimensional if the cluster size is determined by the number
of molecules, n, in the cluster. The flow in this space is characterized by the “density”
ρ(n, t) and the “flow rate”
v = dn/dt ≡ ṅ(n)
In the “cluster language”, ρ(n, t) is the cluster distribution function and ṅ(n) is the
cluster growth law. By definition ρ(n, t)dn is the number of clusters (in the unit
physical volume) having the sizes between n and n + dn; thus, the “mass” of the
cluster fluid inside the (one-dimensional) cluster space volume Vn is
ρ(n, t) dn
Vn
Using the analogy with hydrodynamics we can apply general hydrodynamic consid-
erations [19] to calculate the density of the “cluster fluid” ρ(n, t). In the absence of
nucleation ρ(n, t) satisfies the continuity equation
∂
ρ(n, t) dn = − ρ v d An (3.75)
∂t Vn
∂ρ ∂
+ (ρ ṅ) = 0 (3.76)
∂t ∂n
Equations (3.75)–(3.76) describe the evolution of the cluster distribution in the
absence of nucleation.
Nucleation introduces an extra, source term in (3.75), describing an additional flux in
the cluster space with the density i. Its role is analogous to diffusion in the physical
space for a real fluid. Using the standard hydrodynamic considerations (see [19],
Chap. 6), Eq. (3.75) is modified to:
∂
ρ(n, t) dn = − ρ v d An − i d An (3.77)
∂t Vn
Similar to hydrodynamics we write the flux i using Fick’s law (now in the cluster
size space):
∂ρ
i = −B
∂n
where B is the diffusion coefficient. Equation (3.78) becomes
∂ρ(n, t) ∂ J (n, t)
=− (3.79)
∂t ∂n
∂ρ(n, t)
J (n, t) = −B + ṅ ρ(n, t) (3.80)
∂n
This is the Fokker-Planck equation (FPE) [20] for diffusion in the cluster size space.
The first term describes diffusion in the n-space, with B being the corresponding
diffusion coefficient, while the second term describes a drift with the velocity ṅ
under the action of an external force. The necessary input parameters to solve FPE
are:
• the growth law ṅ(n), which takes into account both condensation (growth) of
supercritical clusters (positive drift) and evaporation of subcritical ones (negative
drift) and
• the model for the Gibbs free energy of the cluster formation ΔG(n)
The idea to apply the Fokker-Planck equation to describe kinetics of cluster formation
belongs to Zeldovich [5]. This formalism can be viewed as a continuous analogue
of the set of Becker-Döring equations (3.36) and leads to the alternative formulation
of the CNT, called the Zeldovich theory, which we discuss below. In (constrained)
equilibrium: J (n, t) = 0 for all clusters and ρ(n, t) = ρeq (n), with ρeq (n) given
by the general expression (3.33). This implies that the drift coefficient in (3.80) is
related to the diffusion coefficient by
∂βΔG(n)
ṅ(n) = −B(n) g1 (n), g1 ≡ (3.81)
∂n
∂
This implies that Eq. (3.80) describes diffusion in the field of force − ∂n ΔG(n). As in
the previous sections, we will be interested in the steady state solution J (n, t) = J =
const, ρ(n, t) = ρs (n) of the kinetic equation (3.79). It is convenient to introduce a
new unknown function
y = ρs (n)/ρeq (n)
∂y
− B ρeq =J (3.82)
∂n
36 3 Classical Nucleation Theory
This equation contains two unknown constants—C and J —which can be found from
the standard boundary conditions in the limit of small and large clusters (3.42)–(3.43):
y → 1, for n → 0, y → 0, for n → ∞
Exploring the exponential dependence of the integrand on n we use the second order
expansion of the Gibbs free energy g(n) ≡ βΔG(n) around the critical cluster:
1 d2 g
g(n) ≈ g(n c ) + g (n c )(n − n c ) , with g (n c ) =
2
(3.85)
2 dn 2 n c
J = ṅ ρ, n > n c + Δ (3.87)
3.7 Zeldovich Theory 37
From the physical meaning of J (n, t) as a flux in the n-space, we identify the coef-
ficient ṅ as a velocity in this space:
dn
ṅ = (3.88)
dt macro
where the subscript “macro” indicated that the growth of the supercritical nucleus
follows a certain macroscopic equation (e.g. diffusion in the real space). Then from
(3.88) and (3.81) we find
1 dn
B(n) = − (3.89)
g1 (n) dt macro
1
ṅ = (n − n c ) + O(n − n c )2
τ
where the parameter τ , introduced by Zeldovich (Zeldovich time), has a dimension-
ality of time and is defined as
−1 dṅ
τ = (3.90)
dn nc
At the critical cluster both ṅ and g1 vanish, while B(n c ) remains finite. Using (3.48)
and (3.49) we rewrite this result in terms of the Zeldovich factor
1 1
τ= = Δ2 (3.91)
B(n c ) 2π Z 2 2B(n c )
38 3 Classical Nucleation Theory
Here ϑ = 0 and ϑ = 1 for the ballistic and diffusion limited cases, respectively;
note that ϑ = −1 corresponds to cavitation [5].
In the previous section we discussed the steady state regime. The latter is preceded by
the transient non-stationary regime characterized by a characteristic relaxation time
τtr . Strictly speaking the steady regime can be reached only at infinite time when
all transient effects have disappeared. However, one can pose a question: how much
time is required for the flux to reach an appreciable fraction of the steady state value
J ? To answer this question we start with the expression for the time-dependent flux
(3.80) rewritten in the form:
∂ ρ(n, t) J (n, t)
=− (3.93)
∂n ρeq (n) B(n)ρeq (n)
We discuss times at which the flux at the point n = n c is a noticeable fraction of the
steady state value J . Due to the sharp maximum of the integrand at n c the following
expansions are plausible:
1
J (n, t) = J (n c , t) + J (n c , t)(n − n c )2
2
ρeq (n) = ρeq (n c ) exp π Z 2 (n − n c )2
3.8 Transient Nucleation 39
In the first integral one recognizes J −1 whereas the second integral can be calculated
using (3.47) and the Gaussian identity [22]
+∞ √
π
x 2 e−a
2x2
dx = , a>0
−∞ 2a 3
resulting in
1
J (n c , t) + J (n c , t) = J (3.95)
4π Z 2
Differentiation of the continuity equation (3.79) with respect to n gives:
∂ 2 J (n, t) ∂ ∂ρ(n, t)
=− (3.96)
∂n 2 ∂t ∂n
d J (n c , t)
J (n c , t) + τtr =J
dt
with
1
τtr = (3.98)
4π B(n c ) Z 2
The parameter τtr characterizes the relaxation period to a steady state, or a time-lag.
For times t > τtr one can speak about the steady regime. Comparing (3.91) and
(3.98) one can see that the time-lag is related to the Zeldovich time τ as:
1 τ
τtr = Δ2 = (3.100)
4B(n c ) 2
GC (n) = θ (n, T ) n
βΔG surf 2/3
with
θ (n, T ) = θ∞ 1 − n −2/3 (3.102)
It is straightforward to see that the resulting nucleation model, termed the Internally
Consistent Classical Theory (ICCT) [24] results in
eθ∞
JICCT = JCNT (3.104)
S
Compared to the modest Courtney (1/S) correction the ICCT correction to the clas-
sical theory, eθ∞ /S can be very large. Expression (3.101) implicitly suggests that the
CNT form of the free energy barrier is valid down to the cluster containing just one
molecule. That is why ICCT can be viewed as a rather arbitrary choice which may,
however, empirically improve the fit to experiment [30].
References 41
References
4.1 Introduction
Various nucleation models use their own set of approximations, have their own range
of validity and certain fundamental and technical limitations. Therefore it is desirable
to formulate some general, model-independent statements which would establish the
relationships between the physical quantities characterizing the nucleation behavior.
One of such statements was proposed by Kashchiev [1] later on termed the Nucleation
Theorem (NT). In 1996 Ford [2] derived another general statement which was termed
the Second Nucleation Theorem. Since then Kashchiev’s result and its generalization
is sometimes also referred to as the First Nucleation Theorem.
Following Kashchiev, consider a general form of the Gibbs free energy of n-cluster
formation
ΔG(n,Δμ) = −n Δμ + Fs (n,Δμ) (4.1)
Here Fs (n,Δμ) is the excess beyond the first (bulk) term free energy of cluster
formation. For our present purposes we do not need to specify it. The critical cluster
satisfies:
∂ΔG
=0
∂n T,Δμ
resulting in
∂ Fs
− Δμ + =0 (4.2)
∂n n c
and noticing that the expression in the round brackets vanishes in view of (4.2), we
obtain the Nucleation Theorem in the form given in Ref. [1]:
dW ∗ ∂ Fs
= −n c + (4.3)
dΔμ ∂Δμ n c
This result is particularly useful when Fs is only weakly dependent on the supersatura-
tion; for example in the CNT within the capillarity approximation Fs = γ∞ (T )A(n)
is totally independent of Δμ. In this case one can determine the size of the critical
cluster from the nucleation experiments measuring the nucleation rates and finding
the slope of ln J − ln S curves (cf. Eq. (3.52)).
V = Vv + Vl
(see Fig. 4.1). Here V l encloses the cluster of the new phase “l” and V v contains the
original phase “v”. In a mixture of q components the total number of molecules of
component i is an extensive quantity and according to (2.22) is given by
where Niv is the number of molecules of type i in the homogeneous phase “v” occu-
pying the volume V v , Nil is the number of molecules of type i in the homogeneous
phase “l” occupying the volume V l , Niexc is the excess number of molecules of type
i accumulated on the dividing surface. The Gibbs free energy of cluster formation
4.2 First Nucleation Theorem for Multi-Component Systems 45
l
V
q
ΔG = ( p v − p l )V l + Nil μil ( p l ) − μiv ( p v )
i=1
q
+ Niexc μiexc − μiv ( p v ) + φ(V l , {μiv }, T ) (4.4)
i=1
The last term, φ(V l , {μiv }, T ), is the total surface energy of the cluster. We do not
specify here its functional form (e.g. by introducing the surface tension and the
surface area of the cluster) which implies that the general form (4.4) can be applied
to small clusters (for which the physical meaning of the surface tension looses its
validity). The critical cluster (denoted by the subscript “c”) is in the mechanical and
chemical equilibrium (though metastable) with the mother phase resulting in the
extremum of the Gibbs energy with respect to {Nil }, {Nisurf } and V l :
Substituting these expressions into (4.4) we find the work of formation of the critical
cluster at the given external conditions—the temperature T and the set of the vapor
phase chemical potentials {μiv }:
Now let us study how W ∗ changes if we change the external conditions. The variation
of W ∗ with respect to the variation of the chemical potential of the component i while
keeping the temperature and the rest chemical potentials fixed, reads:
∂W∗ l ∂( p − pc )
v l ∂ Vcl ∂φc ∂φc
= Vc + ( p v − pcl ) + +
∂μiv ∂μiv ∂μiv ∂ Vcl ∂μiv
In view of the equilibrium condition (4.6) the expression in the square brackets
vanishes resulting in
∂W∗ l ∂( p − pc )
v l ∂φc
= Vc + (4.8)
∂μiv ∂μi
v ∂μiv
q
SdT − V d p + Nk dμk = 0 (4.9)
k=1
Similarly, the Gibbs adsorption equation (2.29) for a mixture at isothermal condi-
tions is:
q
dφc + k =0
Nkexc dμexc (4.11)
k=1
In view of the equality of the chemical potentials (4.5) these relations can be
written as:
⎡ ⎤
Vcl d pcl = Ni,c
l
dμiv ( p v ) + ⎣ l
Nk,c dμvk ( p v )⎦
k=i
⎡ ⎤
V v d p v = Niv dμiv ( p v ) + ⎣ Nkv dμvk ( p v )⎦ (4.12)
k=i
⎡ ⎤
dφc = −Niexc dμiv ( p v ) − ⎣ Nkexc dμvk ( p v )⎦
k=i
4.2 First Nucleation Theorem for Multi-Component Systems 47
In all of these relations the sums in the square brackets has to be set to zero since in
Eq. (4.8) all dμvk = 0 except for k = i. Substituting (4.12) into (4.8), we find:
l
∂W∗ Vc
= −N l
i,c − N i
exc
+ Niv (4.13)
∂μiv Vv
This result can be simplified if we introduce the number densities of the component
i in the both phases “v” and “l”:
l
Ni,c Niv
ρil = , ρiv =
Vcl Vv
Then
Vcl
Niv = ρiv Vcl
Vv
is the number of molecules of component i that existed in the volume Vcl before the
critical cluster was formed. Eq. (4.13) becomes
∂ W ∗ ρv
= − 1 − l
Ni,c + Niexc ≡ −Δn i,c (4.14)
∂μiv {μv }, j=i ρl
j
The quantity Δn i,c is the excess number of molecules of component i in the cluster
beyond that present in the same volume (V l ) of the mother phase before the cluster
was formed. While the quantities Ni,c l , N v , N exc depend on the location of the
i i
dividing surface the excess number Δn i,c is independent of this choice and of the
∗
cluster shape (which for small clusters can have a fractal structure). Thus, ∂ Wv is also
∂μi
independent of the location of the dividing surface. This is consistent with the fact
that the nucleation barrier W ∗ itself is invariant with respect to the dividing surface
(which is a mathematical abstraction rather than a measurable physical quantity).
Equation (4.14) represents the Nucleation Theorem for multi-component systems.
The most important feature of this result is that it is derived without any assumptions
concerning the size, shape and composition of the critical cluster and thus is of general
validity.1 Although the presented proof is based on thermodynamics, it makes no
assumptions about the size of the critical cluster and thus is valid down to atomic
size critical nuclei.
For a unary system Eq. (4.14) reads:
dW ∗ ρv
= − 1 − Ncl + N exc ≡ −Δn c (4.15)
dμv ρl
1Bowles et al. [4] showed that the Nucleation Theorem is a powerful result which is not restricted to
nucleation, as its name suggests, but refers to all equilibrium systems containing local nonuniform
density distributions stabilized by external field (not only a nucleus in nucleation theory).
48 4 Nucleation Theorems
We can rewrite this result in terms of the supersaturation ratio S, recalling that
kB T ln S = μv ( p v , T ) − μsat (T )
d(βW ∗ )
= −Δn c (4.16)
d ln S
where the prefactor J0 depends on a particular nucleation model and on the dimen-
sionality of the problem. Taking in both sides of (4.17) the derivative with respect to
ln S, we obtain
d(βW ∗ ) ∂ ln J ∂ ln J0
=− + (4.18)
d ln S ∂ ln S T ∂ ln S T
where first term on the right-hand side can be directly measured in nucleation exper-
iments. In the CNT the prefactor J0 is given by Eq. (3.55)
J0 = ψ(T ) S 2
and the Nucleation Theorem for the single-component systems takes a particularly
simple form:
∂ ln J
Δn c = −2 (4.20)
∂ ln S T
Within the kinetic approach to nucleation, discussed in Sect. 3.5, the prefactor
J0 contains the Courtney (1/S) correction
J0 = ψ(T ) S
4.2 First Nucleation Theorem for Multi-Component Systems 49
For binary nucleation of components a and b the prefactor J0 has a more com-
plex form than in the single-component case. An important feature of J0 is that it
depends on the composition of the critical cluster; a simple expression describing
this dependence is not available. Applying (4.14) to the binary case we find:
∂ ln J ∂ ln J0
Δn i,c = − , i = a, b
∂(βμiv ) T ∂(βμiv ) T
A contribution of the second term to Δn i,c is small and typically ranges from 0 to 1
[3] resulting in
∂ ln J
Δn i,c = − (0 to 1) , i = a, b (4.22)
∂(βμiv ) T
In view of the equilibrium condition (4.6) the expression in the square brackets
vanishes resulting in
∂W∗ l ∂( p − pc )
v l ∂φc
= Vc + (4.23)
∂T ∂T ∂ T Vcl ,{μv }
i
The Gibbs–Duhem equations for both of the bulk phases at the fixed chemical
potentials read
where ΔSc is the excess entropy due to the critical cluster formation. (Note that
S v/V v is the entropy per unit volume in the vapor phase). Equation (4.26) represents
the Second Nucleation Theorem.
As in the case of the First Nucleation Theorem, the application of the Second Theorem
to the analysis of experiments requires its representation in the form containing
measurable quantities. Since the nucleation barrier enters the nucleation rate in the
form of the Boltzmann factor, we combine Eq. (4.26) with the identity
∂(βW ∗ ) ∂W∗ βW ∗
=β −
∂T ∂T T
which results in
∂(βW ∗ ) βW ∗ (W ∗ + T ΔSc )
= −β ΔSc − =− (4.27)
∂T T kB T 2
According to the thermodynamic consideration of Sect. 3.2 (see Eq. (3.12)) the
expression in the round brackets is the excess internal energy of the critical clus-
ter, i.e. the change in the internal energy beyond that present in the same volume
(V l ) of the mother phase before the cluster was formed:
∂(βW ∗ ) ΔU ∗
= − (4.28)
∂ T {μi } kB T 2
The contribution of the pre-exponential term can be found easily for the one com-
ponent case. Approximating J0 by the classical expression (3.55) we find for the
leading temperature dependence:
∂ ln J0 d ln psat 2
=2 −
∂T S dT T
The first term can be worked out using the Clausius–Clapeyron equation (2.14)
d ln psat l
=
dT kB T 2
4.3 Second Nucleation Theorem 51
where l is the latent heat of evaporation per molecule. Thus, for a unary system the
Second Nucleation Theorem reads:
∂ ln J 2(l − kB T ) + ΔU ∗
= (4.30)
∂ T S kB T 2
With its help one can find the excess internal energy of the critical cluster from the
nucleation rate measurements and the known specific latent heat.
where ΔSc is the excess entropy of the critical nucleus and Δn i,c is the excess number
of molecules of component i in it. This result does not depend on the choice of a
dividing surface.
It is straightforward to see that both the First and the Second Nucleation theorems
follow from Eq. (4.31). In particular, fixing the temperature and all but one chemical
potential in the mother phase, we obtain
∂ W ∗
= −Δn i,c (4.32)
∂μiv T ; {μv }, j=i
j
which is the First Nucleation Theorem (cf. (4.14)). Varying T and keeping all chem-
ical potentials in (4.31) fixed results in the Second Nucleation Theorem:
∂ W ∗
= −ΔSc (4.33)
∂ T {μv }
j
(cf. (4.26)). There exist various other forms of NT resulting from Hill’s theory (see
[8] for the detailed discussion). Below we consider one such form which can be
particularly useful for the analysis of the effect of total pressure on multi-component
nucleation. Let us choose the following set of independent variables: the temperature
52 4 Nucleation Theorems
T , the total pressure p v and the molar fractions of components in the mother phase
{y j }, j = 1, . . . , q. All these parameters are measurable in experiments. In view of
normalization
q
yj = 1
j=1
and viv and siv are, respectively, the partial molecular volume and entropy in the
mother phase
∂ V v
viv ≡ , i = 1, 2, . . . (4.36)
∂ Niv pv ,T, N v
j, j =i
∂S v
si ≡
v
, i = 1, 2, . . . (4.37)
∂ Niv pv ,T, N v
j, j =i
Substituting (4.34) into Hill’s fundamental equation (4.31) we obtain its alternative
form in the p v , T, {y j } variables:
q
q
∗
dW = − ΔSc − siv Δn i,c dT − viv Δn i,c d pv
i=1 i=1
⎛ ⎞
q ∂μv
− ⎝ i
dy j ⎠ Δn i,c (4.38)
∂yj
i=1 j=i
(for a single-component case summation over j in the last term should be omitted).
Let us differentiate the general expression for the steady-state nucleation rate
J
ln = −β W ∗
J0
4.4 Nucleation Theorems from Hill’s Thermodynamics of Small Systems 53
with respect to ln p v :
∂ ln(J/J0 ) ∂ W ∗
= −β
∂ ln p v {yk },T ln p v {yk },T
This result, termed Pressure Nucleation Theorem for multi-component systems [9],
makes it possible to study the effect of total pressure on the nucleation rate in mixtures.
References
1 Historically the DFT in the theory of fluids originates from the quantum mechanical ideas
formulated by Hohenberg and Kohn [4] and Kohn and Sham [5]; these authors showed that the
intrinsic part of the ground state energy of an inhomogeneous electron liquid can be cast in the
form of a unique functional of the electron density ρe (r). By doing so the quantum many-body
problem—the solution of the many-electron Shrödinger equation—is replaced by a variational
one-body problem for an electron in an effective potential field.
equimolar surface
between the liquid-like value
in the center of a cluster to
the bulk vapor density far
from it. Re indicates the Gibbs
vapor
equimolar dividing surface
0 Re
r
Oxtoby and coworkers [7–11]. The general feature of all density functional models
is the assumption that the thermodynamic potential of a nonuniform system can be
approximated using the knowledge of structural and thermodynamic properties of
the corresponding uniform system. Various DFT models differ from each other in the
way this approximation is formulated [12]. Although the results of DFT calculations
can not be presented in a closed form, these calculations are much faster than the
purely microscopic computer simulations (Monte Carlo, molecular dynamics). One
can thus classify the DFT as a semi-microscopic approach.
An important feature of DFT in nucleation is that calculation of the nucleation bar-
rier does not invoke the a priori information about the surface tension. The curvature
effects of the surface free energy are incorporated into the DFT so that no ad hoc
assumptions are necessary. This is the consequence of the fact that DFT uses the
microscopic (interaction potential) rather the macroscopic input. The DFT approach
naturally recovers the CNT when the system is close to equilibrium (low supersat-
urations, large droplets). However, it considerably deviates from CNT at higher S.
In particular, the DFT predicts vanishing of the nucleation barrier at some finite S
(which signals the spinodal) while the CNT barrier remains finite even as the spinodal
is approached.
Before studying the application of DFT to nucleation we briefly formulate its fun-
damentals for the theory of nonhomogeneous fluids.
The cornerstone of DFT is the statement that the free energy of an inhomogeneous
fluid is a functional of the density profile ρ(r). On the basis of the knowledge of
5.2 Fundamentals of the Density Functional Approach in the Theory of Liquids 57
this functional one can calculate interfacial tensions, properties of confined sys-
tems, adsorption properties, determine depletion forces, study phase transitions, etc.
From these introductory remarks it is clear that DFT represents an alternative—
variational—formulation of statistical mechanics. Determination of the exact form
of the free energy functional is equivalent to calculating the partition function, which,
as it is well known, is not possible for realistic potentials. Therefore, one has to for-
mulate approximations which could lead to computationally tractable results, and
at the same time be applicable to a number of practical problems. However, even
without knowing the exact form of the functional one can formulate several rigorous
statement about it.
Let us consider the N -particle system in the volume V at the temperature T . Being
concerned with classical systems, we can always consider the momentum part of the
Hamiltonian to be described by the equilibrium Maxwellian distribution. This implies
that the arbitrariness in the distribution function refers only to the configurational
part of the Hamiltonian. We require that an arbitrary N -body distribution function,
ρ̂ N (r N ), should be positive and satisfy the normalization
ρ̂ (N ) (r N ) dr N = N !, ρ̂ (N ) (r N ) > 0 (5.1)
The “hat” indicates that the distribution does not necessarily have its equilibrium
form; the corresponding equilibrium function will be denoted without the “hat”. The
requirement (5.1) shows that functions ρ̂ (N ) are subject to the same normalization
as the equilibrium function ρ (N ) (r N ) (see e.g. [13] Chap. 2). To limit the class of
functions ρ̂ (N ) , we employ the following considerations [14]. For a given interatomic
interaction potential u(r ) every fixed external field u ext (r) gives rise to a certain
equilibrium one-particle distribution function
i.e. only one u ext (r) can determine a given ρ(r). Here the square brackets denote
that ρ(r) is a functional of u ext (r). Keeping the form of this functional and varying
u ext (r) we can generate a set of singlet density functions ρ̂(r) so that each of them
will be an equilibrium density corresponding to some other external field, û ext (r),
but is nonequilibrium with respect to the original field u ext (r). For every ρ̂(r) there
exists a unique N -particle distribution function ρ̂ (N ) (for a proof see [2]). Thus, any
functional of ρ̂ (N ) (r N ) can be equally considered a functional of the one-particle
distribution function ρ̂(r).
Let us formally define the functional of intrinsic free energy
1
Fint [ρ̂] = dr N ρ̂ (N ) [U N + kB T ln(Λ3N ρ̂ (N ) )] (5.2)
N!
58 5 Density Functional Theory
where
2π 2
Λ=
m 1 kB T
e−βU N
ρ (N ) = (5.3)
Λ3N Z N
where Z N is the canonical partition function. Substitution of (5.3) into (5.2) results in
which is the Helmholtz free energy of the system in the absence of external fields
(intrinsic free energy). In the presence of an external field the free energy functional
can be defined as a straightforward extension of Eq. (5.2):
1
F [ρ̂] = dr N ρ̂ (N ) [U N + U N ,ext + kB T ln(Λ3N ρ̂ (N ) )] (5.5)
N!
where
N
U N ,ext = u ext (ri ) (5.6)
i=1
is the total external field energy of the configuration r N . Substituting (5.6) into (5.5)
we obtain
ρ̂ (N )
N
F [ρ̂] = Fint [ρ̂] + dr N u ext (ri )
N!
i=1
implying that
F [ρ̂] = Fint [ρ̂] + dr u ext (r)ρ̂(r) (5.7)
Thus, the free energy functional in the presence of an external field is a sum of the
intrinsic free energy functional and the (average) energy of the system in an external
field. Similarly to Eq. (5.4), for equilibrium conditions F [ρ] recovers the Helmholtz
free energy of the system in an external field.
Finally, we define the grand potential functional:
Ω[ρ̂; u ext ] = F [ρ̂] − μ dr ρ̂(r) (5.8)
Ω[ρ; Uext ] = F − μN = Ω
the equilibrium profile ρ(r) minimizes the functional of the free energy. The proof
is presented elsewhere (see e.g. [13], Chap. 9). Using the Lagrange multipliers the
minimizing property for F under the condition (5.9) can be cast in the form of the
unconditional minimum of the grand potential functional:
δΩ
=0 (5.10)
δ ρ̂(r) ρ(r)
60 5 Density Functional Theory
where
δFint [ρ̂]
μint (r) ≡ (5.12)
δ ρ̂(r) ρ(r)
is the intrinsic chemical potential. The spatial dependence of μint must be exactly can-
celed by the radial dependence of u ext (r ), since the “full” chemical potential μ (which
is the Lagrange parameter in this variational problem) is constant. Equations (5.11)–
(5.12) represent the fundamental result of DFT. If we had means to determine Fint ,
then (5.11) would be an exact equation for the equilibrium density.
For realistic interactions the exact expression for the functional Fint is not available,
and one has to invoke approximations. To this end let us study the response of the
system to a small change in the pair potentials δu(ri j ) that alters the total interaction
energy (assumed to be pairwise additive)
δU N (r N ) = δu(ri j )
i< j
where
Ξ (μ, V, T ) = λN Z N
N ≥0
is the grand partition function of the system, λ = eβμ is the activity. We have
δΞ 1 N 1
δΩ = −kB T = λ dr N e−β(U N +U N ,ext ) δU N (r N )
Ξ Ξ Λ3N N !
N ≥0
Examination of the right-hand side reveals that it represents the thermal average (in
the grand canonical ensemble) of δU N
δΩ = δU N
where ρ (2) (r1 , r2 ) is the pair distribution function. Using the definition of a varia-
tional derivative, this result can be expressed as
δΩ 1
= ρ (2) (r1 , r2 ) (5.13)
δu(r12 ) 2
From the definition of Ω[ρ] it follows that the same expression is valid for Fint :
δFint [ρ] 1
= ρ (2) (r1 , r2 ) (5.14)
δu(r12 ) 2
u = u0 + u1 (5.15)
u α (r12 ) = u 0 (r ) + α u 1 (r ), 0 ≤ α ≤ 1 (5.16)
which gradually change from u 0 to u when the formal parameter α changes from
zero to unity. The functional integration of (5.14) then gives:
1
1
Fint [ρ] = Fint,0 [ρ] + dα dr1 dr2 ρα(2) (r1 , r2 )u 1 (r12 ) (5.17)
2 0
∂u α
δu α = dα = u 1 dα
∂α
The first term in (5.17) is the reference contribution—the intrinsic free energy of the
system with the interaction potential u 0 (r ). The second term refers to the perturbative
(2)
part. The pair distribution function ρα is that of a system with the density ρ and
the interaction potential u α . Equation (5.17) is the second fundamental equation of
the DFT which gives the exact (though intractable!) expression for the intrinsic free
energy. To make the theory work we imply the perturbation approach in which u 1 is
considered a small perturbation. Expanding (5.17) in u 1 to the first order we obtain
1
Fint [ρ] = Fint,0 [ρ] + dr1 dr2 ρ0(2) (r1 , r2 ) u 1 (r12 ) + O(u 21 )
2
62 5 Density Functional Theory
u 0(r)
u(r)
rm rm
r r
- u1 (r)
-
(2)
where the distribution function ρ0 is now that of the reference system with the
density ρ(r) and interaction potential u 0 (r ). One can go further and treat the reference
part in the local density approximation (LDA):
Fint,0 [ρ] ≈ dr ψ0 (ρ(r)) (5.18)
and
(2)
ρ0 (r1 , r2 ) ≈ ρ(r1 ) ρ(r2 ) g0 (ρ̄; r12 ) (5.19)
where ψ0 (ρ) is the free energy density of the uniform reference system with number
density ρ; g0 (ρ̄; r12 ) is the pair correlation function of the uniform reference system
evaluated at some mean density ρ̄, e.g. ρ̄ = [ρ(r1 ) + ρ(r2 )]/2. The LDA is valid
for weakly inhomogeneous systems, such as a liquid–vapor interface. For strongly
inhomogeneous systems, e.g. liquid at a wall, it becomes too crude and one has to
use a nonlocal approximation, such as the weighted density [15] or the modified
weighted-density approximation [12].
The most widely used decomposition of the interaction potential, (5.15) is given by
the Weeks–Chandler–Anderson theory (WCA) [16] (see Fig. 5.2):
u(r ) + ε for r < rm
u 0 (r ) = (5.20)
0 for r ≥ rm
−ε for r < rm
u 1 (r ) = (5.21)
u(r ) for r ≥ rm
where ε is the depth of the potential u(r ) and rm is the corresponding value of r :
u(rm ) = −ε. The advantage of the WCA scheme is that all strongly varying parts of
the potential are subsumed by the reference model describing the harshly repulsive
interaction, whereas u 1 (r ) varies slowly and therefore the importance of fluctuations
in the free energy expansion (represented by the second order term) is reduced.
5.2 Fundamentals of the Density Functional Approach in the Theory of Liquids 63
The free energy of the reference model can be expressed as the free energy of a hard-
sphere system with a suitably defined effective diameter d at the same temperature T
and the number density ρ as the original system. The thermodynamic properties of
a hard-sphere system are readily available from the Carnahan–Starling theory [17].
In particular the pressure and the chemical potential are (see e.g. [13]):
pd 1 + φd + φd2 − φd3
= (5.22)
ρkB T (1 − φd )3
μd 8φd − 9φd2 + 3φd3
= ln(ρΛ3 ) + (5.23)
kB T (1 − φd )3
where φd = (π/6)ρd 3 is the volume fraction of effective hard spheres. Then, from
the standard thermodynamic relationship the Helmholtz free energy density of the
hard-sphere system reads
ψd = ρμd − pd (5.24)
In the WCA theory it equals the free energy density of the reference model: ψ0 (ρ) =
ψd (ρ).
One can go further and ignore all correlations between particles in the perturbative
term which results in setting g0 = 1 in (5.19); this is the random phase approximation
(RPA), equivalent to the mean field (van der Waals-like) theory. A number of studies
showed that for systems with weak inhomogeneities, the RPA is sufficient when the
system is not too close to the critical point. Combining LDA and RPA, we obtain the
intrinsic free energy in its simplest form:
1
Fint [ρ] = dr ψd (ρ(r)) + dr1 dr2 ρ(r1 ) ρ(r2 ) u 1 (r12 ) (5.25)
2
where μd (ρ(r)) is the local chemical potential of the hard-sphere fluid. The integral
equation (5.26) can be solved iteratively for ρ(r) starting with some initial profile
satisfying the boundary conditions corresponding to bulk equilibrium:
ρ(z) → ρsat
v
in the bulk vapor
ρ(z) → ρsat
l
in the bulk liquid
where
1
a=− dr u 1 (r ) (5.29)
2
Given μsat one can perform an iteration process for the density profile ρ(z) starting
l for z < 0 and ρ(z) =
with an initial guess, say a step-function ρ(z) with ρ(z) = ρsat
ρsat for z > 0. This profile is put into the rhs of Eq. (5.26) and the latter is solved
v
δ2 Ω
>0
δρ(z)δρ(z )
Differentiation of F with respect to the volume yields the virial equation of state:
p = pd − ρ 2 a (5.32)
ρkB T
p= − ρ 2 a vdW
1 − bvdW
5.2 Fundamentals of the Density Functional Approach in the Theory of Liquids 65
where for the Lennard–Jones fluid the van der Waals parameters are:
16π 2π 3
vdW
aLJ = εσ 3 , bLJ
vdW
= σ
9 3
At the same time the DFT uses a more sophisticated approach to describe the repul-
sive part of the potential than the free volume considerations of van der Waals. Fur-
thermore, due to the different decomposition
√ schemes, the background interaction
parameters are different: aLJWCA = 2 a vdW .
LJ
Ω exc Ω[ρ] + pV
γ = = (5.35)
A A
where A is the surface area. The grand potential functional (5.8) with u ext = 0 then
reads
Ω[ρ] = − dr pd (ρ) + dr ρ μd (ρ)
1
+ dr ρ(r) dr ρ(r ) u 1 (|r − r |) − μ dr ρ(r)
2
where we used the intrinsic free energy functional in the RPA-form (5.25). Using the
DFT equation (5.26) and taking into account that dr = A dz we find
1
γ =− dz pd (ρ(z)) + ρ(z) dr ρ(z ) u 1 (|r − r |) − psat (5.36)
2
where the equilibrium vapor–liquid density profile ρ(z) satisfies Eq. (5.26).
Let us apply the DFT to a Lennard–Jones fluid [9, 18] characterized by the interaction
potential u LJ (r ). We begin by searching for the bulk equilibrium conditions at a
given temperature T < Tc . Performing the WCA decomposition, we determine the
effective hard-sphere diameter d for the reference model. Densities in the bulk phases,
together with the equilibrium chemical potential and pressure, are found from the
coupled nonlinear equations (5.28)–(5.32)
μ = μd (ρ v ) − 2ρ v a = μd (ρ l ) − 2ρ l a
p = pd (ρ v ) − (ρ v )2 a = pd (ρ l ) − (ρ l )2 a
66 5 Density Functional Theory
3
0.4
0.2
0
0 5 10 15 20
z/
1
2
0.5
0
0.6 0.8 1 1.2
kBT/
G = Ω + p v V + μv N (5.37)
where V is the volume of the system “droplet + vapor” and μv ( p v , T ) is the chemical
potential of a vapor molecule. In the absence of a droplet the Gibbs free energy and
the grand potential are
G 0 = μv N , and Ω0 = − p v V
Then from (5.37) the change in the Gibbs energy due to the droplet formation is
ΔG = Ω − Ω0 ≡ ΔΩ (5.38)
and therefore the energy barrier to nucleation can be calculated in the grand ensemble.
The grand potential functional is related to the intrinsic free energy (in the absence
of external field) via
Ω[ρ(r )] = Fint [ρ(r )] − μ drρ(r ) (5.39)
For Fint [ρ(r )] we take the mean field form (5.25) which consists of the local den-
sity approximation for the (effective) hard-sphere part of the interaction potential
and the random phase approximation for the attractive part u 1 (r ) considered as a
perturbation:
1
Fint [ρ] = drψd (ρ(r)) + dr1 dr2 ρ(r1 )ρ(r2 )u 1 (r12 ) (5.40)
2
δΩ
=0
δρ(r )
resulting in:
μd (ρ(r)) = μ − dr ρ(r )u 1 (|r − r |) (5.42)
refers to a local (rather than the global) minimum of the free energy. Still there
is a nontrivial solution of Eq. (5.42) which corresponds to a saddle point of the
functional Ω[ρ] in the functional space. This solution describes a critical nucleus.
The iteration process is now unstable. Nevertheless the solution can be found once
an appropriate initial guess is chosen. As an initial guess for a radial droplet profile
a step-function can be taken with a range parameter Rinit . If Rinit is small enough
the droplet will shrink in the process of iteration giving rise to a metastable vapor
density solution. If Rinit is large the droplet will grow into a stable liquid. There
∗ which in the process of iteration will give rise to the
exist an intermediate value Rinit
critical droplet neither growing, nor shrinking over a large number of iteration steps
n. So the function Ω(n) will exhibit a long plateau staying at a constant value Ω ∗ .
The energy barrier for nucleation is given by
ΔΩ ∗ = Ω ∗ − Ω0
since the nucleation rate is far less sensitive to J0 than to the value of the energy
barrier.
The great advantage of the DFT over the purely phenomenological models is that
in the DFT one does not have to invoke the macroscopic equilibrium properties
and equation of state. Yet, calculations of nucleation behavior are much faster than
direct computer simulations using Monte Carlo or Molecular Dynamics methods
(discussed in Chap. 8).
5.3 Density Functional Theory of Nucleation 69
5.3.2 Results
Oxtoby and Evans [6] calculated density profiles for a critical droplet with a Yukawa
attractive potential
e−κr
u 1 (r ) = −ακ 3
4π λr
According to CNT one expects that the density in the center of the droplet is equal
l (T ). However, the results of [6] show that the
to the liquid density at coexistence ρsat
density in the center of the droplet is lower than ρsat
l and is lower than the density
of the bulk liquid at the same chemical potential as the supersaturated vapor ρ l (μv ).
Another important feature of this approach is that the barrier to nucleation, ΔΩ ∗ ,
vanishes as spinodal is approached whereas in the classical theory it remains finite.
Zeng and Oxtoby [9] applied the DFT to predict nucleation rates for a Lennard-
Jones fluid. Following the preceding discussion the free energy functional is written
using the hard-sphere perturbation analysis based on the WCA decomposition of the
non-truncated Lennard-Jones potential.
Figure 5.5 based on the results of Ref. [9] shows the comparison of the nucleation rates
predicted by the DFT and the CNT. To make such a comparison consistent the macro-
scopic surface tension used in the CNT was obtained by means of Eq. (5.35). The
results in Fig. 5.5 correspond to the fixed classical nucleation rate JCNT = 1 cm−3 s−1 .
At each temperature this condition determines the chemical potential difference (and
consequently the supersaturation) for which the DFT calculation is carried out. The
results of Ref. [9] demonstrate that CNT and DFT predict the same dependence of
nucleation rate on the supersaturation but show the essentially different temperature
dependence. Due to the latter the disagreement between the two approaches is up
to 5 orders of magnitude in J . The difference between two theories becomes pro-
nounced when the nucleation temperature is away from kB T /ε ≈ 1.1. As one can see
-6
0.6 0.7 0.8 0.9 1 1.1 1.2
k BT/
70 5 Density Functional Theory
from Fig. 5.5, at lower temperatures CNT underestimates the nucleation rate (com-
pared to DFT) while at higher temperatures it overestimates it. The same trend is
demonstrated by the CNT when it is compared to experimental nucleation rates: e.g.
for water CNT underestimates experimental rates at T < 230 K, and overestimates
experimental rates at T > 230 K [23, 24].
References
The CNT studies formation of droplets in an open system. The main feature of the
modified liquid drop model of Ref. [4] is that it considers the closed system containing
N molecules confined within a small spherical volume V at a temperature T . This
small N V T system is called an EMLD-cluster. Within the volume V various sharp
n-clusters can form, n = 1, . . . , N . The important difference between the closed an
open system is that in the system with the fixed total amount of molecules N the
formation and growth of droplets is accompanied by the depletion of the vapor—the
effect neglected in CNT, which assumes the existence of an infinite source of vapor
molecules. The depletion of the vapor molecules in EMLD results in the decrease
of supersaturation so that the droplet can not become arbitrarily large. Following
n r
Ref. [6] we analyze the properties of the EMLD-cluster using purely thermodynamic
considerations.
The spherical volume V of the radius R is assumed to have impermeable hard walls.
Under certain conditions a liquid drop with n molecules, n ≤ N − 1, can be formed
inside V . The rest N − n molecules remain in the vapor phase, occupying the volume
V − n vl , where vl is the volume per molecule in the bulk liquid phase (see Fig. 6.1).
This vapor has the pressure described within the ideal gas approximation
(N − n) kB T
p1 = (6.1)
V − nvl
The sharp n-cluster is treated within the capillarity approximation, i.e. it is assumed
that it has a sharp interface, characterized by the macroscopic surface tension γ∞ ,
and the bulk liquid properties inside it. We stress that within this model the entire
N V T -system is considered as the EMLD-cluster, and not just one (sharp) n-droplet.
Since we discuss the closed system, the appropriate thermodynamic potential is the
Helmholtz free energy
F = U − TS
where U is the internal energy and S is the entropy of the EMLD-cluster. Denoting
the vapor and liquid subsystems inside the EMLD-cluster by subscripts 1 and 2,
respectively, we write the differentials of U1 and U2 as:
whereA = 4πr 2 is the surface area of the n-cluster. The free energy change associated
with the formation of a sharp n-droplet inside the EMLD-cluster is:
2γ∞
p2 − p1 = (6.6)
r
The second equality is the Laplace equation. We can rewrite these results applying the
ideal gas approximation for the vapor and considering liquid to be incompressible.
Using the thermodynamic relationship
1
(dμ)T = dp
ρ
for the vapor and liquid phases, we relate μi to the bulk vapor-liquid equilibrium
properties
p1
μ1 ( p1 ) − μsat = kB T ln (6.7)
psat
p1 2γ∞ l
kB T ln = v + vl ( p1 − psat ) (6.9)
psat r
The last term is usually very small and can be neglected. The result is the classical
Kelvin equation ( 3.61) relating the pressure inside the n-droplet to its radius
2γ∞ vl
p1 = psat exp (6.10)
r kB T
Using Eq. (6.1) for the vapor pressure inside the EMLD-cluster, we can solve (6.10)
for the size of the coexisting droplet. In CNT, dealing with the open μV T system,
the solution of the Kelvin equation determines the critical cluster radius at the given
temperature and supersaturation. In the closed system the supersaturation p1/psat is
not fixed but depends on the amount of molecules in the n-cluster since the total
74 6 Extended Modified Liquid Drop Model and Dynamic Nucleation Theory
amount N is conserved. Another difference is that while in the open system the
solution of the coexistence equations is unique and corresponds to the critical cluster,
in the closed system Eqs. (6.5)–(6.6) have two solutions: one corresponding to the
critical cluster, i.e. the cluster in the metastable equilibrium with the vapor, and the
other one—corresponding to the stable cluster.
For an arbitrary n-cluster (i.e. not necessarily a critical one), which is not in equi-
librium with the surrounding vapor, the free energy change (dF ) N V T = 0 and the
chemical potential difference μ2 ( p2 ) − μ1 ( p1 ) reads:
p1
μ2 ( p2 ) − μ1 ( p1 ) = vl ( p2 − psat ) − kB T ln (6.11)
psat
Substituting (6.11) into (6.4) and using the incompressibility of the liquid phase
(dV2 = vl dn) we obtain:
2γ∞ l p1 (n)
(dF ) N V T = v + v ( p1 (n) − psat ) − kB T ln
l
dn (6.12)
r (n) psat
is the free energy difference between the state in which EMLD-cluster contains the
n-droplet and the state of pure vapor; thus, ΔF(n) is the Helmholtz free energy
of the n-droplet formation. Taking into account that r (n) = r l n 1/3 (where r l =
(3vl /4π )1/3 ) and using (6.1) we find after the integration
p1 vl psat p1
ΔF (n) = −n kB T ln psat + γ∞ A + n k B T 1 − kB T + N kB T ln p0
(6.13)
where
N kB T
p0 =
V
is the pressure corresponding to the pure vapor. In Eq. (6.13) the first and the second
terms are, respectively, the standard bulk and surface contributions to the free energy
of cluster formation (recall, that here the supersaturation is S(n) = p1 (n)/ psat ); the
third term is the volume work (usually small and is commonly neglected); the last
term originates from depletion of the vapor molecules in the EMLD-cluster when an
n-droplet is formed. In the thermodynamic limit p1 = p0 recovering the CNT result
for the free energy of the cluster formation. Note, that in this model the state of pure
vapor corresponds to n = 0 and not to n = 1; the value n = 1 corresponds to a
hypothetical liquid cluster of size 1, which is not the same as a molecule of the vapor.
6.1 Modified Liquid Drop Model 75
The EMLD-cluster can contain sharp droplets of various sizes n. The total free energy
of the cluster ΔFtot is found by accounting of all possible fluctuations of the droplet
size. Each such fluctuation enters the configuration integral Q of the cluster with the
Boltzmann factor e−βΔF (n) resulting in
N
Q= e−βΔF (n)
n=0
Then,
N
ΔFtot = −kB T ln Q = −kB T ln e−βΔF (n) (6.14)
n=0
e−βΔF (n)
N
f (n) = , f (n) = 1
Q
n=0
The total pressure Ptot of the EMLD-cluster is the weighted sum of the vapor pressure
in the confinement sphere over all possible n. The vapor pressure consists of p1 and
the pressure exerted by the n-drop (for n = 0), modelled as a single ideal gas molecule
moving within the container
N
kB T
Ptot = f (n) p1 + Ξ (n)
Vc
n=0
Here Vc (n) = 4π(R − r (n))3 /3 is the volume accessible for the center of mass of
the n-droplet with the radius r (n), R is the radius of the EMLD-cluster
1/3
3V
R=
4π
Considerations put forward in the previous section, have not yet answered the ques-
tion: how to choose the volume V of the EMLD-cluster which would be physically
relevant for nucleation at the given external conditions? To address it we notice that
76 6 Extended Modified Liquid Drop Model and Dynamic Nucleation Theory
∂ Ptot
=0
∂V Vm
The combined (EMLD and DNT) model is called the EMLD-DNT theory.
In the open (μV T )-system the work of formation of the physical (N , Vm )-cluster at
a temperature T is given by
where Δμ0 = μ0 − μ. The role of the cluster size in this expression is played by N .
By construction the EMLD-DNT cluster represents the diffuse interface system. That
is why N is not the molecular excess quantity determined by the nucleation theorem
6.3 Nucleation Barrier 77
(discussed in Chap. 4). The nucleation barrier corresponds to the maximum of ΔG:
∂ΔG
=0 (6.16)
∂N Vm ,T
Having determined the critical cluster N ∗ from this equation, we substitute it into
Eq. (6.15) to obtain the nucleation barrier
In the thermodynamic limit p0 coincides with the actual vapor pressure p v . The
EMLD-DNT nucleation rate reads:
ΔG ∗EMLD−DNT
JEMLD−DNT = J0 exp − (6.18)
kB T
References
a small parameter implying that the Tolman expansion becomes dubious. Therefore,
it is desirable to formulate a nonperturbative semi-phenomenological approach valid
for all cluster sizes. In this chapter we discuss such a model—a Mean-field Kinetic
Nucleation Theory (MKNT) of Ref. [6].
7.2 Kinetics
where S is the supersaturation, f (n) is the forward rate of n-cluster formation in the
supersaturated vapor, and ρsat (n) is the equilibrium cluster distribution at saturation
(corresponding to S = 1). The purely phenomenological considerations adopted
in the CNT and in Katz’s kinetic version of the CNT are valid when the major
contribution to the nucleation rate comes from big clusters (the notion of a “big
cluster” will be specified below). In this case the free energy of cluster formation
reads
sat (n) = γ∞ s1 n
ΔG CNT 2/3
The formal extension of these expressions to all cluster sizes does not pose the
problem since the contribution of small clusters to J is negligible.
On the other hand, if nucleation behavior is primarily driven by the formation of
small clusters, one has to apply microscopic considerations. In this regime the very
notion of a surface tension of a small cluster looses its physical meaning. In the next
section we consider the model for ρsat (n) which is valid for arbitrary n.
7.3 Statistical Thermodynamics of Clusters 81
A typical interaction between gas molecules consists of a harshly repulsive core and
a short-range attraction. The most probable configurations of the gas at low densi-
ties and temperatures will be isolated clusters of n = 1, 2, 3, . . . molecules. Hence,
to a reasonable approximation one can describe a real gas as a system of nonin-
teracting clusters.1 At the same time intracluster interactions are important—they
are responsible for the formation of a cluster All clusters are in statistical equilib-
rium, associating and dissociating. Even large clusters have a certain probability of
appearing. The partition function of an n-cluster at a temperature T is:
1
Zn = qn (7.2)
Λ3n
where Λ is the thermal de Broglie wavelength of a particle (being an atom or a
molecule); qn is the configuration integral of the n-cluster in a physical domain of
volume V :
1
qn (T ) = drn e−βUn , (7.3)
n! cl
1
Un is the potential energy of the n-particle configuration in the cluster; the factor n!
takes into account the indistinguishability of particles inside the cluster. The sym-
bol cl indicates that integration is performed only over those atomic configurations
that belong to the cluster. At this point it is important to emphasize the difference
between the n-particle configuration integral Q n and the n-cluster configuration inte-
gral qn . The latter includes only those configurations in the volume V that form the n-
cluster, while Q n contains all different configurations of n particles in the volume V ;
therefore Q n ≥ qn . The cluster as a whole can move through the entire volume V of
the system, while the particles inside the cluster are restricted to the configurations
about cluster’s center of mass that are consistent with a chosen definition of the clus-
ter. For that one can adopt, e.g., Stillinger cluster [7]: an atom belongs to a cluster
if there exists at least one atom of the same cluster separated from the given one by
a distance r < rb , where rb is some characteristic distance describing the range of
interparticle interactions. In other words, an atom belongs to the cluster if inside a
sphere of radius rb there is at least one atom belonging to the same cluster. Early
Monte Carlo studies of Lee et al. [8] showed that a cluster’s free energy is almost
independent of a cluster definition provided that the definition is reasonable and the
temperature is sufficiently low. For the present model a particular type of a cluster
definition is not important. What matters is that a cluster is a compact object around
its center of mass.
1 Note that at high temperatures interactions between clusters can not be neglected.
82 7 Mean-Field Kinetic Nucleation Theory
The partition function Z (n) of the gas of Nn noninteracting n-clusters in the volume
V at the temperature T is factorized:
1
Z (n) = Z Nn (7.4)
Nn ! n
where the prefactor 1/Nn ! takes into account the indistinguishability of clusters (recall
that indistinguishability of atoms inside the cluster is taken into account in qn ). The
Helmholtz free energy of the gas of n-clusters is: F (n) = −kB T ln Z (n) , which
using Stirling’s formula becomes:
(n) Nn
F = Nn kB T ln
Zn e
Introducing the number density of n-clusters ρ(n) = Nn/V and substituting (7.2)
into (7.5), we obtain
V Λ3n
μn = kB T ln ρ(n) (7.6)
qn
μn = nμv (7.7)
where μv is the chemical potential of a molecule in the vapor phase. Combining (7.6)
and (7.7) we find:
qn
ρ(n) = zn (7.8)
V
where
z = eβμ /Λ3
v
(7.9)
is the fugacity of a vapor molecule. From the definition of qn it is clear that the
quantity qn /V involves only the degrees of freedom relative to the center of mass
of the cluster and remains finite in thermodynamic limit (V → ∞). The pressure
equation of state for the vapor is given by Dalton’s law
∞
p v (μv , T )
= ρ(n)μv ,T (7.10)
kB T
n=1
7.3 Statistical Thermodynamics of Clusters 83
Equation (7.8) holds for every point of the gaseous isotherm. In particular for the
saturation point it reads
qn n eβμsat
ρsat (n) = z sat , z sat = (7.12)
V Λ3
where the chemical potential at saturation μsat (T ) can be found from the suitable
equation of state. The problem of finding the equilibrium cluster distribution is
reduced to the determination of the cluster configuration integral. Up to this point
all results were exact. To proceed with calculation of qn it is necessary to introduce
approximations.
n = n core + n s (7.13)
The physical idea behind this distinction is that the core of the cluster, if present,
should possess the liquid-like structure which can be characterized by a certain
property typical for the liquid phase, e.g. by the liquid coordination number N1 . The
surface molecules can then be viewed as an adsorption layer covering the core.2 By
definition the integer numbers n core and n s satisfy
n core ≥ 0, ns ≥ 1
2 This decomposition should not be confused with the Gibbs construction involving a dividing
surface discussed in Sect. 2.2.
84 7 Mean-Field Kinetic Nucleation Theory
so that δn core + δn s = 0. For the present purposes it is not necessary to specify the
form of these quantities; we postpone this discussion till Sect. 7.5.
Similar to the seminal Fisher droplet model of condensation [5], we write the internal
potential energy of an n-cluster as a sum of the bulk and surface contributions
Un = −n E 0 + Wn (7.14)
Here −E 0 (E 0 > 0) is the binding energy per particle related to the depth of inter-
particle attraction; Wn is the surface energy of the n-cluster. In the Fisher model Wn
has the form Wn = w A(n), with w being the energy per unit surface and A(n) is the
cluster surface area. In view of the previous discussion, we present Wn in a different
way making use of the concept of surface particles:
W n = w1 n s , w1 > 0 (7.15)
where w1 is the surface energy per surface particle. Both E 0 and w1 are material con-
stants independent of temperature. The difference between the two models becomes
increasingly important for small clusters, for which the surface area of a cluster can
not be properly defined.
Let us place the origin of the coordinate system in Eq. (7.3) into the center of mass
of the cluster. Then the configuration integral can be written as
where
1
drn−1 e−βw1 n
s (rn−1 )
G n (β) = (7.17)
n! cl
describes the “surface part” of qn and has the dimensionality of (volume)n−1 . The
notation n s (rn−1 ) indicates that n s depends on a particular configuration of cluster
particles. The configurational Helmholtz free energy of the cluster reads
Fnconf = −kB T ln qn
∂Fnconf ∂ ln qn
Snconf = − = kB ln qn − β (7.18)
∂T ∂β
7.4 Configuration Integral of a Cluster: Mean-Field Approximation 85
From (7.16)
ln qn = ln V + nβ E 0 + ln G n (β) (7.19)
yielding
∂ ln qn 1 ∂G n
= n E0 +
∂β G n ∂β
where n s (β) is the thermal average of n s . Substituting (7.19) and (7.20) into (7.18)
we find
The bulk entropy per molecule can be identified with the entropy per molecule in the
bulk liquid, or equivalently, in the infinitely large cluster
Snconf
S0 (β) = lim
n→∞ n
When n → ∞ most of the particles belong to the core while the relative number of
surface molecules vanishes:
ns
→ 0 as n → ∞
n
(the rigorous proof of this statement is presented in Sect. 7.5). Then
1
S0 (β) = kB lim ln G n (β) (7.21)
n→∞ n
By virtue of the cluster definition, mutual distances between molecules in the cluster
can not exceed some maximum value, therefore the integral (7.17) remains finite.
Let us introduce a temperature dependent parameter v0 with the dimensionality of
volume which can be understood as an average volume per molecule in the cluster.
1/3
Scaling all distances with v0 , we rewrite G n as
1
rn−1 e−βw1 n
s (
rn−1 )
G n (β) = v0n−1 d
n! cl
whereri are the dimensionless positions of the cluster molecules; the integral in the
square brackets is now dimensionless (and finite). The number of surface molecules
86 7 Mean-Field Kinetic Nucleation Theory
in the n-cluster depends on a particular configuration, but can have the values in
the range 1 ≤ n s ≤ n. Replacing integration over 3(n − 1) configuration space by
summation over all possible values of n s , we obtain
g(n, n s ) e−βw1 n
s
G n (β) = v0n−1 (7.22)
1≤n s ≤n
where the degeneracy factor g(n, n s ) gives the number of different molecular con-
figurations in the n-cluster having the same number n s of the surface molecules.
We calculate the positive definite series G n (β) using the mean-field approximation.
We assume that the sum in Eq. (7.22) is dominated by its largest term to the extent that
it is possible replace the entire sum by this largest term while completely neglecting
the others:
s −βw1 n s
G n (β) ≈ v0n−1 max
s
g(n, n ) e (7.23)
n
The similar approximation is used in the theory of phase transitions giving rise to
the Landau theory [10]. The maximum in (7.23) is attained at the most probable
number of surface particle in the n-cluster, n s (n; β), corresponding to the parti-
cle configuration with the maximum statistical weight. Thus, within the mean-field
approximation
G n (β) = v0n−1 g(n, n s ) e−βw1 n
s
(7.24)
Recalling the physical meaning of g(n, n s ), one can expect that ln g(n, n s ) is related
to the configurational entropy of the n-cluster. To verify this conjecture we use Eqs.
(7.21) and (7.24) to obtain:
1
S0 = kB ln v0 + lim ln g(n, n s ) (7.25)
n→∞ n
3 Note that as a thermodynamic quantity S conf depends on the average number of surface molecules
n
n s , whereas Un as a microscopic quantity depends on n s itself.
7.4 Configuration Integral of a Cluster: Mean-Field Approximation 87
yielding:
n S0 ν1 n s
g(n, n s ) = v0−n exp + (7.28)
kB kB
where
C = v0−1 (7.30)
Substituting (7.29) into (7.8), the number density of n-clusters takes the form
s
ρ(n) = C y n x n (7.31)
S0
y ≡ z exp β E 0 + (7.32)
kB
w1 − ν1 T
x ≡ exp − (7.33)
kB T
The quantity x > 0 measures the temperature; at low temperatures x is small. The
quantity y measures the fugacity, or, equivalently, the chemical potential of the vapor.
A big cluster can be associated with a liquid droplet in the vapor. The growth of a
macroscopic droplet corresponds in this picture to condensation. Following [5], let
us discuss the probability of finding an n-cluster in the vapor at the temperature T
and the chemical potential μv . This probability is proportional to ρ(n).
If in (7.31) y < 1, which corresponds to a small z, or equivalently to a large and
negative μv , then ρ(n) exponentially decays as exp[−const × n]. As y approaches
unity this decrease becomes slower. When y = 1, ρ(n) still decays but only as
exp[−const × n s (n)]. Finally, if y slightly exceeds unity, then ρ(n) first decreases,
reaching a minimum at some n = n 0 , and then increases without bounds (see
Fig. 7.1). The large (divergent) probability of finding a very large cluster signals
the condensation. Thus, we identify
ysat = 1 (7.34)
with the saturation point. Applying Eq. (7.32) to saturation and using (7.34) we find
S0 1
exp β E 0 + = (7.35)
kB z sat
88 7 Mean-Field Kinetic Nucleation Theory
(n)
y >1
diverges
n0 n
where z sat (T ) is the fugacity at saturation. From (7.31) and (7.34) the cluster distri-
bution at saturation reads:
(w1 − ν1 T ) n s
ρsat (n) = C exp −
kB T
The quantity
γmicro = w1 − ν1 T (7.36)
is the Helmholtz free energy per surface particle of the cluster, it includes both
energy and entropy contributions and depends on the temperature but not on the
cluster size. The size-dependence of the surface energy is contained in n s (n). By
analogy with the fluctuation theory γmicro can be termed a microscopic surface tension
per particle (the combination of the terms “microscopic” and “surface tension” is
purely terminological and should not cause confusion). It is convenient to introduce
the dimensionless quantity
γmicro
θmicro = (7.37)
kB T
which we term the “reduced microscopic surface tension”, being the Helmholtz
free energy per surface particle in kB T units. From (7.29), (7.35) and (7.37) the
configuration integral of the n-cluster takes the form
qn −n −θmicro n s (n)
= C z sat e (7.38)
V
This is the central result of the model. Substitution of (7.38) into (7.8) yields the
cluster distribution function in supersaturated vapor
ρ(n) = C enβ (μ
v −μ )
e−θmicro n
s (n)
sat , 1≤n<∞ (7.39)
7.4 Configuration Integral of a Cluster: Mean-Field Approximation 89
ρsat
v
C = ∞ , where h ≡ eθmicro (7.41)
−n s (n)
n=1 n h
Without going into details of the behavior of n s (n) for arbitrary n we make use of
the physically obvious fact that small clusters (with n ≤ N1 ) do not have the liquid
core implying that:
n s (n) = n, for n ≤ N1 (7.42)
h 1 (7.43)
or equivalently:
θmicro (T ) > 2 (7.44)
The above constraint is the domain of validity of the model. From (7.41) to the
leading order in 1/ h
v θmicro
C = ρsat e (7.45)
v0 = vsat
v
e−θmicro (7.47)
1
v
Z sat =1− (7.48)
h
On the other hand,
v
Z sat = 1 + Z exc,sat
v
(7.49)
v
where the first term is the ideal gas part and Z exc,sat is the excess (over ideal) contri-
bution. Comparing (7.48) and (7.49) we find
1
h=− v (7.50)
Z exc,sat
This result shows that the microscopic surface tension originates from the nonideality
of the vapor and can be determined from a suitable equation of state; its simplest
form is the second order virial expansion [11]:
B2 psat
v
Z sat =1+ (7.51)
kB T
v
Z exc,sat
(−B2 ) psat
θmicro = − ln (7.52)
kB T
This feature of the model makes it especially attractive for applications: in order to
find θmicro one does not need to solve the two-phase equilibrium equations but can
7.4 Configuration Integral of a Cluster: Mean-Field Approximation 91
Zsat
compressibility factor Z sat
increases with T . Both curves
meet at the critical point Tc
Zc
Z lsat
0 1
T/Tc
use the experimental or tabulated data on B2 (T ) and psat (T ) which are available for
a large amount of substances in a broad temperature range (see [11]). From (7.43)
and (7.52) the range of validity of the theory is given by:
B2 (T ) psat (T )
1 (7.53)
kB T
−8 ÷ 10−5 . At higher
exc,sat ≈ 10
v is close to unity, so that Z v
At low temperatures Z sat
v
temperatures Z sat decreases (see Fig. 7.2). If we formally apply (7.50) at the critical
point we would find
1
h(Tc ) =
1 − Zc
The critical compressibility factor lies in the limits Z c ≈ 0.2 ÷ 0.4 [11] indicating
that the constraint (7.43) is violated. For example, for van der Waals fluids Z c = 3/8
[12] yielding
8
h vdW
c =
5
The failure of the model at high temperatures manifests the fact that in this domain one
has to take into account intercluster interactions which are completely neglected in the
present model. Besides, in the close vicinity of Tc fluctuations become increasingly
important and the mean-field approach can be in error.
n = n core + n s (7.54)
92 7 Mean-Field Kinetic Nucleation Theory
core
n core core
R
l
R
l s
n
By definition the core possesses the liquid-like structure characterized by the coor-
dination number N1 in the liquid phase; N1 gives the average number of nearest
neighbors for a molecule in the bulk liquid. This quantity influences a number of
physical properties: density, viscosity, diffusivity, etc. The determination of N1 is a
nontrivial problem in its own right (it is discussed in Sect. 7.6). In the present section
we assume N1 to be known. By construction the clusters with n ≤ N1 do not have
the core—all their particles belong to the surface:
core surface
v
0 Rcore Re
r
ρCNT (r ) = ρ l (1 − Ξ (r − Re )) + ρ v Ξ (r − Re )
where Ξ (x) is the Heaviside unit step function. Clearly, for large clusters the
relative width of the adsorption layer λr l /R core → 0 and both approaches become
asymptotically identical.
Within the two-step approximation Eq. (7.54) reads:
4π 4π
n = ρl (R core )3 + ξρ l [(R core + λr l )3 − (R core )3 ] (7.56)
3 3
n core ns
R core
X=
rl
From (3.23)
X 3 = n core (7.57)
This cubic equation for X (n) has the unique real positive root if n − ωλ2 > 0.
For n ≥ N1 + 1: n core ≥ 1, implying that X ≥ 1. The minimum value X = 1 is
94 7 Mean-Field Kinetic Nucleation Theory
achieved when
n = 1 + N1
In this case the core contains just one particle while the rest N1 particles belong to
the surface. Eq. (7.58) then results in the relation between ω and λ:
N1
ω=
3 + 3λ + λ2
0 < ω < N1 /7
n s (n) = n − [X (n)]3
where X (n) is the solution of Eq. (7.58). It is convenient to introduce two dimen-
sionless quantities
ns X3
α= =1− and (7.60)
n n
ζ = n −1/3 (7.61)
Then
X = ζ −1 (1 − α)1/3
Equation (7.62) can be used to determine the parameter ω. To this end let us consider
the behavior of the model for large clusters: n → ∞. In this case almost all cluster
particles belong to the core:
n core ns
→ 1, → 0 for n → ∞
n n
The problem has two small parameters:
0<α 1, 0<ζ 1
7.5 Structure of a Cluster: Core and Surface Particles 95
α = 3ωζ
This result comes as no surprise, since at large n the droplet is a compact spherical
object, its radius scales as n 1/3 and the number of surface atoms scales as the surface
area ∼n 2/3 . From (7.46) to the same order in n
On the other hand, for big clusters the CNT description is valid:
1 θ∞
ω= (7.66)
3 θmicro
π d3
η≡ (7.67)
6va
96 7 Mean-Field Kinetic Nucleation Theory
For the case of a liquid the atomic volume va should be replaced by 1/ρ l and the
atomic diameter d should be replaced by the position dg of the first peak of the pair
correlation function g(r ; ρ l , T ). Thus,
π l 3
η= ρ dg (7.69)
6
If the diffraction data is not available one can find dg using the ideas of perturbation
approach in the theory of liquids applying the Weeks–Chandler–Anderson decom-
position scheme of the interaction potential given by Eqs. (5.20), (5.21) and shown in
Fig. 5.2. Within the WCA dg is approximated by the effective hard-sphere diameter
rm
3
dhs =3 1 − e−βu 0 (r ) r 2 dr (7.70)
0
Combining the results of the previous sections the steady state nucleation rate is
∞
−1
−H (n)
J = K0 e (7.71)
n=1
4 Equation (7.70) is the mean-field approximation to the original WCA expression, where the cavity
function of the hard sphere system is set to unity (for details see e.g. [12] Chap. 5).
7.7 Steady State Nucleation Rate 97
where psat s1
K 0 = ρsat
v
f 1,sat S, f 1,sat = √ (7.72)
2π m 1 kB T
and
2
H (n) = ln n + n ln S − θmicro n s (n) − 1 (7.73)
3
X (n) is the real positive root of Eq. (7.58) in which the parameters λ and ω are found
from
1 θ∞ N1 3 3
ω= , λ= − − (7.75)
3 θmicro ω 4 2
(where dhs (T ) is the effective hard sphere diameter in the theory of liquids) by means
of Eq. (7.68). We refer to this model as a Mean-field Kinetic Nucleation Theory
(MKNT) [6].
It is easy to see that −H (n) is the free energy of the cluster formation in kB T units
−H (n) = βΔG(n)
Apart from the small logarithmic corrections (which can safely be set to a constant
2
3 ln n c , n c being the critical cluster) the free energy reads
βΔG(n) = −n ln S + θmicro n s (n) − 1 (7.77)
Expression in the round brackets is positive: the supersaturation can not exceed some
maximum value given by the pseudospinodal corresponding to the nucleation barrier
≈ kB T . From the pseudospinodal condition, discussed in Chap. 9, it follows that
ln S < θmicro
Hence, ΔG(N1 − 1) < ΔG(N1 ). Due to the model construction when the clus-
ter size is increased from N1 to N1 + 1, the number of surface particles does not
change
n s (N1 ) = n s (N1 + 1) = N1
leading to ΔG(N1 ) > ΔG(N1 + 1). Thus, ΔG(n) has a maximum at n = N1 which
is an artifact of the model and has to be ignored. The second maximum corresponds
to the critical cluster n c :
dΔG dn s
ΔG ≡ = − ln S + θmicro =0 (7.79)
dn dn
1
ΔG(n) ≈ ΔG ∗ + ΔG (n c ) (n − n c )2 , ΔG ∗ ≡ ΔG(n c )
2
we have:
∞
∞ ∗
βΔG(n) βΔG ∗ 1 eβΔG
e ≈e dx exp βΔG (n c ) x 2 =
−∞ 2 Z
n=1
where
−βΔG (n c )
Z =
2π
where
J0 = Z ρsat
v
f (n c ) (7.81)
2/3
is the kinetic prefactor, f (n c ) = f 1,sat S n c , and
7.8.1 Water
with the closed symbols referring to Jth = JMKNT and open symbols referring
to Jth = JCNT . Circles (open and closed) correspond to the experiment of Wölk
Water
4
C
N
T
log 10(Jexp/Jth)
MKNT
-2
200 220 240 260
T(K)
Fig. 7.5 Relative nucleation rate Jrel = Jexp /Jth for water; Log(Jrel ) ≡ log10 Jrel . Closed symbols:
MKNT, open symbols: CNT. Circles: experiment of Wölk et al. [15]; squares experiment of Labetski
et al. [17]. The lines labelled ‘CNT’ and ‘MKNT’, shown to guide the eye, illustrate the temperature
dependence of the relative nucleation rate for the CNT and MKNT, respectively. Also shown is the
“ideal line” (dashed): Jexp = Jth
100 7 Mean-Field Kinetic Nucleation Theory
log10J (cm-3s-1)
symbols: experiment [18].
Labels: nucleation tempera-
ture in K; horizontal labels 1
refer to the theory, inclined
(italicized) labels refer to
experiment. The CNT line
0
320
for 300 K is almost coincid-
310
290
300
ing with the MKNT line for
310 K; and the CNT line for H2O/He
310 K is almost coinciding -1
with the MKNT line for 320 K 3 3.5 4 4.5
S
et al. [15]; squares (open and closed)—to the experiment of Peeters et al. [16] and
Labetski et al. [17]. The thermodynamic data for water used in both models are
given in Appendix A. In the whole temperature range the MKNT predictions are
1–2 orders of magnitude off the experimental data, while the CNT demonstrates
much larger deviation. The most important observation, however, is that MKNT
predicts the temperature dependence of the nucleation rate correctly. Meanwhile,
the discrepancy between CNT and experiment depends on the temperature: at low T
CNT underestimates the experimental data, reaching about 4 orders of magnitude at
the lowest temperature T = 201 K, while at high T CNT slightly overestimates the
experiment. At 200 < T < 220 K critical clusters, corresponding to experimental
conditions, contain ≈15–20 molecules; for such small objects the dominant role in
the cluster formation is played by the microscopic (rather than the macroscopic)
surface tension which explains the success of MKNT.
Brus et al. [18] measured water nucleation in helium in the thermal diffusion cloud
chamber. It is important to note that the measurements were performed for temper-
atures T = 290, 300, 310, 320 K which are beyond the freezing point implying
that all macroscopic properties of water are well known from experiment. Figure 7.6
shows the experimental J − S curves together with MKNT and CNT predictions.
At these relatively high temperatures CNT systematically overestimates experiment
(in qualitative agreement with the previously shown results) by 2–4 orders of magni-
tude. MKNT predictions are in perfect agreement with experiment (within one order
of magnitude). There is a regular temperature shift of CNT curves with respect to
experiment by about 10◦ K; as a result they practically overlap the MKNT curves
related to the nucleation temperatures which are 10◦ K higher. In particular, the 290 K
CNT line overlaps the 300 K MKNT line; the 300 K CNT line overlaps the 310 K
MKNT line, etc.
7.8 Comparison with Experiment 101
log (J rel)
(dashed): Jexp = Jth
10
0
Nitrogen
7.8.2 Nitrogen
Figure 7.7 shows the comparison of experimental nucleation rate Jexp of Ref. [19]
with predictions of the CNT JCNT and MKNT JMKNT . Thermodynamic data used
in the analysis is given in Appendix A. As in Fig. 7.5 the relative nucleation rate is:
Jrel = Jexp /Jth with Jth = JCNT or JMKNT .
The dashed line corresponds to the “ideal case”: Jexp = Jth . As one can see, MKNT
predictions for J deviate on average from the experimental data by 2–7 orders of
magnitude while the CNT predictions are 10–20 orders of magnitude lower then the
experimental values.
7.8.3 Mercury
ln S
(2006), American Institute of
Physics.) 12 MK
NT
Expt
10
6
260 280 300 320
T [K]
according [23]. Recently the cluster size dependent interaction potential of mercury
was proposed by Moyano et al. [24].
Experimental study of mercury nucleation in helium as a carrier gas was carried out
by Martens et al. [25]. The measurements were made in an upward diffusion cloud
chamber. This technique makes it possible to detect the onset of nucleation rather
than to measure directly the nucleation rates (as e.g. in shock wave tube experiments).
The onset corresponds to the nucleation rate
The supersaturation giving rise to the onset of nucleation is referred to as the critical
supersaturation. Figure 7.8 shows the experimental results (closed circles) for the
critical supersaturation S as a function of the nucleation temperature. Also shown
are predictions of the MKNT (solid line), CNT (dashed line). The thermodynamic
data for mercury is presented in Appendix A. As stated in [25] the measured values
of S are about 3 orders of magnitude lower than the CNT predictions; the MKNT
results are in good agreement with experiment.
Recalling how sensitive the nucleation rate is to the value of S it is instructive to
illustrate the difference between the two models in terms of J . For that purpose we
choose an experimental point T = 284 K, ln S = 9.35 and compare experimental
and theoretical results corresponding to these conditions. Experimental rate is given
by the onset condition (7.83) while theoretical predictions are:
The CNT predictions deviate from experiment by about 67 orders of magnitude (!)
while MKNT predictions are within 3 orders of magnitude.5
7.9 Discussion
The general MKNT result for the nucleation rate can be written as
⎡ ⎤−1
⎢N ⎥
⎢ 1 Nclass ∞ ⎥
⎢ ⎥
J = K0 ⎢ e−H (n) + e−H (n) + e−H (n) ⎥ (7.84)
⎢ ⎥
⎣n=1 n=N1 n=Nclass ⎦
small intermediate large
If the critical cluster falls into this domain the contribution of the other two groups
is negligible and we recover the kinetic CNT result (3.69). This case can be called
a macroscopic nucleation regime; here the surface part of the free energy is entirely
determined by the macroscopic surface tension θ∞ .
5 Note, that the nucleation rate is very sensitive to the surface tension: if γ∞ is measured within
the 10 % relative accuracy (common for most of the liquid metals), the accuracy of the predicted
nucleation rate for mercury lies within 4 orders of magnitude.
104 7 Mean-Field Kinetic Nucleation Theory
N1
N1
1
e−H (n) = (7.86)
n 2/3 S n e−θmicro (n−1)
n=1 n=1
small
The free energy is determined solely by the microscopic surface tension while the
macroscopic surface tension does not play a role.
In majority of nucleation experiments the critical cluster contains ∼10 − 102 mole-
cules, falling into the domain of intermediate clusters where the influence of both
micro- and macroscopic surface tension is important. This intermediate nucleation
regime represents a challenging problem for a nucleation theory. MKNT solves it
by providing a smooth interpolation between the two limits—of large and small
clusters—for which it becomes “exact” by construction. Such interpolation is pos-
sible because MKNT is not based on a perturbation in the cluster curvature making
it possible to describe all nucleation regimes within one model.
r −12 r −6
u LJ (r ) = 4ε −
σ σ
Scaling the distance with σ and the energy with ε we introduce the dimensionless
quantities
kB T pσ 3
tLJ = , ρLJ = ρσ 3 , pLJ =
ε ε
The compressibility factor then reads
p p
Z= = LJ
ρkB T ρLJ tLJ
7.9 Discussion 105
At coexistence v
pLJ,
v
Z sat = sat
ρLJ,
v t
sat LJ
The expression in the square brackets is also a universal function of tLJ . Introducing
the reduced temperature t = T /Tc we present tLJ as
where the critical temperature for Lennard-Jones fluids tLJ,c can be determined from
Monte Carlo simulations (see e.g. [26]): tLJ,c = kB Tc /ε ≈ 1.34. Dividing both sides
of Eq. (7.87) by kB Tc , we find
γmicro
= Ψ (t)
kB Tc
where Ψ (t) is again a universal function. From Eq. (7.36) we expect that at low
temperatures this function is linear
γmicro w1 ν1
= − t (7.89)
kB Tc kB Tc kB
implying that for Lennard-Jones fluids at low temperatures the dimensionless surface
entropy per particle, ν/kB , and the surface energy per particle, w1 /kB Tc , are universal
parameters.
This conjecture can be verified using the mean-field density functional calculations
described in Chap. 5. Bulk equilibrium follows Eqs. (5.30)–(5.31) with
Here the hard-sphere pressure, pd , and chemical potential, μd , follow the Carnahan-
Starling theory [27] and the background interaction parameter for a Lennard-Jones
fluids is given by Eq. (5.34): √
16π 2 3
a= εσ (7.92)
9
Figure 7.9 shows the results of the DFT calculations (the dotted line). Indeed, to the
high degree of accuracy the temperature dependence of γmicro turns out to be linear,
106 7 Mean-Field Kinetic Nucleation Theory
4
DFT
micro = w1 - 1T
3.5
3 Ar
/kBTc
N2
CH
2.5 4
micro
2
DFT for LJ fluids:
DF
T
1.5 w1/kBTc=3.62
1/kB=2.90
1
0.2 0.3 0.4 0.5 0.6 0.7 0.8
t=T/Tc
Fig. 7.9 Microscopic surface tension γmicro for Lennard-Jones systems. Dotted line: DFT cal-
culation for a Lennard-Jones system; solid lines: MKNT results for argon (Ar), nitrogen (N2 ) and
methane (C H4 ) obtained using Eq. (7.52) and empirical fit for the second virial coefficient [11]. The
MKNT curves for argon and methane are practically indistinguishable (Reprinted with permission
from Ref. [6], copyright (2006), American Institute of Physics.)
In Fig. 7.9 we compare the DFT result with predictions of MKNT for various simple
fluids—argon, nitrogen, methane— using Eq. (7.52) and empirical fit to the second
virial coefficient for nonpolar substances [11] (see Eq. (F.2)). All MKNT curves are
close to the DFT line (the curves for argon and methane are practically indistin-
guishable) confirming the validity of the MKNT conjecture for γmicro (T ) and the
numerical values of its parameters (7.93)–(7.94).
In order to treat small clusters within the classical approach, some nucleation models
replace γ∞ in the CNT free energy of cluster formation by the curvature dependent
form, representing the expansion of the surface tension of a cluster in powers of its
curvature. To discuss the range of validity of such an expansion we study the general
Eq. (7.62), valid for all cluster sizes.
7.9 Discussion 107
For large clusters we solve the cubic equation (7.62) for α(ζ ) keeping the terms up
to the 3rd order in ζ , which results in
2
λ λ
α = 3ωζ 1 − 2 ω − ζ+ − 3λω + 3ω ζ 2 , ζ → 0
2
(7.95)
2 3
To the same order of accuracy the cluster distribution function at saturation reads:
γn A(n)
ρsat (n) = ρsat
v
exp −
kB T
Recall that for sufficiently large droplets the surface tension at the surface of tension,
γt = γ [Rt ], can be expressed using the Tolman formula (2.41)
2δT
γ t = γ∞ 1− + ... (7.99)
Rt
Since both ω and λ are of order 1, δT is of the order of molecular size. It is tempting
to apply the Tolman equation (7.100) truncated at the first order term for relatively
small n-clusters once the Tolman correction is small
2 ω − λ n −1/3 1 (7.102)
2
This condition, however, is not sufficient. By definition (see Eq. (2.37)) the Tolman
length is independent of the radius; meanwhile, for sufficiently small droplets
δ = Re − Rt is a strong function of Rt [28, 29]. Koga et al. [30] obtained δ(Rt )
for simple fluids (Lennard-Jones and Yukawa) using the DFT. They showed that
for these systems the Tolman equation becomes valid for clusters containing more
than 105 ÷ 106 molecules. This result implies that in contrast with conventional
expectation one can not apply the Tolman equation down to droplets containing few
tens or hundreds of molecules. For the nucleation theory it means that even in the
macroscopic nucleation regime one can use Tolman’s correction only for extremely
large clusters, containing n ∼ 105 ÷ 106 Nclass molecules which are practically
irrelevant for experimental conditions (critical clusters contain usually ∼101 ÷ 102
molecules).
MKNT allows to find an independent criterion of the applicability of the Tolman
equation to nucleation problems by analyzing the next-to-Tolman term in the curva-
ture expansion (7.97). The following condition should be satisfied:
λ2
− 3λω + 3ω2
3 −1/3
n 1 (7.103)
2 ω − λ2
Combining (7.102) and (7.103) we conclude that the Tolman correction is applicable
for the clusters satisfying
⎧ ⎫
⎨
λ
λ2
3 − 3λω + 3ω2 ⎬
n 1/3 max 2 ω − , (7.104)
⎩ 2 2 ω − λ2 ⎭
For typical experimental conditions the second requirement (Eq. (7.103)) is much
stronger than the first one (Eq. (7.102)). For the cases studied in Sect. 7.8 the Tolman
Eq. (7.100) is valid for clusters containing ∼(4 ÷ 5) × 104 ÷ 105 particles which is
close to the result of Ref. [30] for simple fluids. Note, that although δT is not used
7.9 Discussion 109
15 -0.2
Water
-0.4
, m icro 10 m icro
(A)
T
T
m icro/kBTc,
-0.6
m icro/kBTc
-0.8
0
0.3 0.4 0.5 0.6 0.7 0.8
t=T/Tc
Fig. 7.10 Equilibrium properties of water used in MKNT: θ∞ , θmicro , γmicro /kB Tc (left y-axis)
and the Tolman length δT [Å] according to Eq. (7.101) (right y-axis) (Reprinted with permission
from Ref. [6], copyright (2006), American Institute of Physics.)
20
T Nitrogen
0
15
m icro
-0.2
,
(A)
10
T
m icro/kBTc,
m icro
-0.4
5
LJ m icro/kBTc
-0.6
0
0.3 0.4 0.5 0.6 0.7 0.8
t=T/Tc
Fig. 7.11 MKNT: equilibrium properties of nitrogen.: θ∞ , θmicro , γmicro /kB Tc (left y-axis) and
the Tolman length δT [Å] according to Eq. (7.101) (right y-axis). The dashed line labelled “LJ”
shows γmicro /kB Tc for Lennard-Jones fluids according to Eq. (7.89) with the universal parameters
given by (7.93)–(7.93) (for details see the text)
In the limit of small (nano-sized) clusters the behavior of the system (the free energy
barrier and the distribution function) as a function of the cluster size n will be essen-
tially different from that given by the phenomenological CNT. It is instructive to
consider the physical picture emerging from the MKNT in the case of small n. For
n ≤ N1 the equilibrium cluster distribution is
v −θmicro (n−1)
ρsat (n) = ρsat e , 1 ≤ n ≤ N1 (7.105)
Using the relation (7.12) we find for the configuration integral of an n-cluster:
Λ3 e−βμsat n−1
qn = ρsat
v
V e−θmicro (n−1) e−nβμsat Λ3n = ρsat
v
Λ3 e−βμsat V
h
(7.106)
μsat = kB T ln(ρsat
v
Λ3 )
7.9 Discussion 111
qn = V K n−1 , 1 ≤ n ≤ N1 (7.107)
where
1
K = v eθmicro (7.108)
ρsat
The factorized form (7.107) of the configuration integral suggests that the n-cluster
in this case is characterized by short range nearest neighbor interactions between
particles and contains n − 1 bonds, each bond contributing the same quantity K
to qn . This is the minimum possible number of bonds for an n-cluster (a spherical
droplet represents the opposite limit of maximum number of bonds). The latter means
that a small cluster is not a compact object but rather reminds a polymer chain of
atoms with nearest neighbor interactions [31]. Each particle of such cluster is bonded
to the two neighboring particles belonging to the same chain with exception of the
end-point particles having one neighbor. The simplest example of such a cluster is a
linear chain of molecules. A somewhat more complex cluster structure satisfying Eq.
(7.107) can have branch points where a particle of the given chain has contacts with
particles belonging to another chain; however, loops are prohibited (see Fig. 7.12).
This structure was studied in Ref. [32] where it was termed a “system of virtual
chains” indicating that the sequence of atoms in the chains is not fixed: they are
associating and dissociating.
The value of K depends on the interatomic potential u(r ) and temperature. For
the nearest neighbor pairwise additive interaction we obtain from the definition
of qn :
q2 n−1
qn = V , 1 ≤ n ≤ N1 (7.109)
V
Comparing (7.107) and (7.109) we identify
q2 1
K = = dr e−βu(r ) (7.110)
V 2 cl
112 7 Mean-Field Kinetic Nucleation Theory
From (7.109) and (7.12) it is seen that K represents the dimer association constant:
ρsat (2)
K = (7.111)
[ρsat (1)]2
Hence the microscopic surface tension can be related to the dimer association
constant: v
θmicro = − ln ρsat K
References
8.1 Introduction
N
p2i
Ekin = . (8.2)
2 mi
i=1
8.2 Molecular Dynamics Simulation 115
where mi is the mass of a particle i and pi is its momentum. The generalized momenta
pk are given by
∂L
pk = , k = 1, 2, . . . , s
∂ q̇k
The potential energy may be divided into terms which depend on external potentials,
particle interactions of pairs, triplets etc.:
N
N
N
N
N
N
U= u1 (ri ) + u2 (ri , rj ) + u3 (ri , rj , rk ) + . . . (8.3)
i=1 i=1 j>i i=1 j>i k>j>i
The first term in Eq. (8.3) contains the external potential u1 acting on the individual
molecules. The second term contains the potential u2 of the interaction between any
pair of particles in the system. We could continue to include the interactions between
triplets, quadruplets etc. The second term, however, is typically the most important
one. Neglecting any external fields, the first term in Eq. (8.3) equals zero. In addition,
the contributions from triplet or higher order interactions are typically much smaller
and are often neglected, such as for the widely used Lennard-Jones potential. Thus,
the total potential of the system reduces to a good accuracy to the sum of all pairwise
interactions:
N
N
U= u2 (ri , rj ) (8.4)
i=1 j>i
The second summation j > i provides the exclusion of double counting. If we insert
the expressions for potential and kinetic energy into Eq. (8.1), the Lagrange equation
reduces to
mi r̈i = fi , (8.5)
which is the Newton second law in terms of the force fi acting on particle i. For time
and velocity-independent interaction potentials, the force is given by
1 2
L= mv − U(x)
2
116 8 Computer Simulation of Nucleation
where v is the velocity of the particle. The Lagrange equation (8.1) now reads:
d ∂L ∂L
=
dt ∂ ẋ ∂x
mv − dU
dx
On the left-hand side the derivative of the momentum mv with respect to the time
t yields ma, where a is the acceleration of the particle, while the right-hand side is
equal to the force f . Lagrangian dynamics does not only recover the Newton law
but can also be applied to more complicated problems. For example it is useful for
the proper derivation of the equations of motion for a system including a thermostat.
In this case the Lagrange equation has to be extended by a term representing the
energy of the thermostat. The application of the Lagrange differential equation to
this approach gives the desired proper equation of motion for the NVT ensemble.
Besides Lagrangian dynamics one can also describe the evolution of the system using
the Hamiltonian dynamics. Let us introduce the Hamiltonian of the system [1]
H (p, q) = q̇k pk − L (q, q̇)
k
With its help the equations of motion can be written in the Hamiltonian form:
∂H
q̇k = , (8.7)
∂pk
∂H
ṗk = − . (8.8)
∂qk
Finally, we can write down the Hamiltonian equations (8.7)–(8.8) in Cartesian form:
Computing of the trajectories in the phase space involves solving either a system of
3N second-order differential equation (8.5) or an equivalent set of 6N first-order dif-
ferential equations (8.9)–(8.10). The classical equations of motion are deterministic
and invariant to time reversal. This means that if we change the sign of the velocities,
the particles will trace back on exactly the same trajectories. In computer simula-
tions, exact reversibility usually is not observed because of the limited accuracy of
the numerical calculations and the chaotic behavior of the dynamics of a many-body
system.
8.2 Molecular Dynamics Simulation 117
dx 1 d2 x 1 d3 x 3
x(t + Δt) = x(t) + Δt + 2
Δt 2
+ 3
Δt + O(Δt 4 )
dt
2 dt
6 dt
v(t) a(t)
From Newton’s law we can calculate the acceleration a at the moment t for the given
force f (t):
f (t)
a(t) =
m
yielding
f (t) 2 1 d3 x 3
x(t + Δt) = x(t) + v(t) Δt + Δt + Δt + O(Δt 4 ) (8.11)
2m 6 dt 3
In order to solve the above equation, we need to know the velocity at time t. However,
it can be eliminated by expanding the position of the molecules backward in time.
Replacing in (8.11) Δt by −Δt, we obtain
f (t) 2 1 d3 x 3
x(t − Δt) = x(t) − v(t) Δt + Δt − Δt + O(Δt 4 ) (8.12)
2m 6 dt 3
Adding expansions (8.11) and (8.12), the odd terms cancel resulting in
f (t) 2
x(t + Δt) = 2x(t) − x(t − Δt) + Δt + O(Δt 4 ) (8.13)
2m
This way of numerically integrating the equations of motion is called the Verlet
algorithm [2]. Though it is a rather simple approach, its error is just of the order
of O(Δt 4 ). For calculation of positions of the particles the velocity is not needed.
Meanwhile, it is required for calculation of the kinetic energy of the system, instan-
taneous temperature, transport properties, to mention just a few. One can calculate v
by subtracting the expansions (8.11) and (8.12):
Thus, velocity at time t is calculated after the positions at time (t + Δt) and (t − Δt)
have been already found. The positions in Eq. (8.13) are exact up to errors of the
order of O(Δt 4 ), while the velocities are exact up to an error of the order of O(Δt 3 ).
The Verlet algorithm is properly centered on the respective positions at (t ± Δt) and,
thus, is time reversible. However, the difference between the accuracies in positions
and velocities may result in deviations from the classical trajectories. Consequently,
most Verlet-type algorithms exhibit a drift in the energy of the system on short and
long time-scales, which is strongly influenced by the length of the time-step Δt.
Since calculation of the velocities in the original Verlet method contains a compar-
atively large error, a more accurate form, known as the velocity-Verlet algorithm,
was proposed [3]:
1
x(t + Δt) = x(t) + Δt v(t) + f (t) Δt 2 (8.15)
2m
Δt
v(t + Δt) = v(t) + f (t) + f (t + Δt) (8.16)
2m
Calculation using the velocity-Verlet algorithm typically proceeds through the two
steps: first, the new positions x(t + Δt) and the velocities are calculated at an inter-
mediate time interval:
1 1
v t + Δt = v(t) + f (t) Δt
2 2m
The original Verlet algorithm may be recovered by eliminating the velocity from
Eq. (8.16). In the velocity-Verlet method, the new velocities are obtained with the
same accuracy and at the same time as the positions and accelerations of the mole-
cules. Therefore, the kinetic and potential energy are known at each time. The
velocity-Verlet algorithm requires a minimum storage memory and its numerical
stability, time reversibility, and simplicity in both form and implementation make it
by far the most preferred method in MD simulations to date.
A number of other algorithms for solving the equations of motion have been devel-
oped. For example, the leap-frog algorithm improves the accuracy by using half-steps
between the time steps. Other methods are based on higher order terms of the Taylor
expansions. In general, the more accurate the algorithm—the larger the time step
can be used. For a complete survey of various algorithms the reader is referred to the
literature [4, 5].
8.2 Molecular Dynamics Simulation 119
Finite size effects, i.e. the influence of the system size on the simulated properties,
are also important for calculation of the surface tension [7]. Surface fluctuations of
a vapor–liquid interface lead to the so-called capillary waves with the wavelength
related to the surface tension. Besides the wavelength, also the amplitude of the
capillary waves varies with the system size leading to a widening of the interface
with increasing system size.
1
N
Eat,kin = mi vi2
2N
i=1
The equipartition theorem of statistical mechanics states that each degree of freedom
contributes kB T /2 to the total kinetic energy of the system (see e.g. [8]) yielding
Eat,kin = Nf kB T /2 (8.17)
where Nf is the number of degrees of freedom of a particle; for a particle with trans-
lational motion in 3D Nf = 3. In general, after one (or several) steps of the dynamics,
the instantaneous temperature T will be different from the desired temperature Tset .
By rescaling the velocities of the molecules one can set the system to the temperature
Tset :
Tset
vi,new = vi
T
While such velocity scaling approach is suitable in some cases and for some prop-
erties, especially in equilibrium, in the context of nucleation this method of ther-
mostating will give wrong results. Rescaling the velocities is actually a method to
keep the kinetic energy of a particle constant while all temperature fluctuations are
completely eliminated. At the same time if one wants to simulate the system in the
canonical NVT ensemble, one has to realize that this ensemble exhibits thermal fluc-
tuations. These fluctuations are the origin of the heat capacity. Thus, the velocity
scaling method is by construction unable to predict the heat capacity. Furthermore,
in nucleating systems the velocity scaling can lead to an artificial cooling down of
the remaining monomers in the vapor phase [9].
Several thermostating techniques were proposed which are able to correctly realize
the canonical ensemble. One example is the stochastic Andersen thermostat [10].
Instead of rescaling the velocities of all molecules in every time step, one or a few
molecules are randomly chosen from the vapor and their new velocities are calculated
from the Maxwell distribution function at the desired temperature. In this way, the
Andersen thermostat mimics a collision of a molecule with a carrier gas particle. The
frequency at which a molecule is picked from the vapor, determines the effectiveness
of this method. In the limit of very high frequencies one recovers a procedure similar
to velocity scaling but with a certain temperature distribution. On the other hand, a
low frequency may not be sufficient to keep the temperature constant.
122 8 Computer Simulation of Nucleation
Another widely used method is the Nosé-Hoover thermostat [11, 12]. It includes
the thermostat in the equations of motion. Adding an energy term, which describes
the thermostat, to the Lagrange (or Hamilton) function and applying the Lagrangian
(or Hamiltonian) dynamics leads to equations of motion including a friction term,
that affects the acceleration of the molecules. The Nosé-Hoover thermostat correctly
realizes the canonical ensemble allowing for fluctuations in the system temperature.
If the thermostat itself is coupled to another thermostat, one obtains a so-called Nosé-
Hoover chain thermostat [13], which represents an improved thermostat realizing
the canonical NVT ensemble.
In physical experiments the exchange of energy between the system and the environ-
ment is usually provided by a carrier gas. Such carrier gas usually does not affect the
phase transition beyond its function as a latent heat transfer agent. It is also possible
to include a carrier gas in MD simulations. To keep the computational effort low, it
is useful to employ for this purpose a mono-atomic gas, such as the Lennard-Jones
argon. It is added into the simulation box, typically in abundance. When a cluster
is formed by nucleation, the latent heat is removed from the cluster by collisions of
the cluster particles with the carrier gas atoms. In real physical experiment the heat
is removed from the carrier gas either by expanding the system or by collisions with
the container walls. In MD simulations it is possible to simulate the expansion [14]
directly or model the heat removal from the carrier gas by coupling it to one of the
regular MD thermostats as described above. This is possible because the carrier gas
remains in the gas phase and does not condense. Special care should be taken in the
cases in which the carrier gas is present in the interior of the cluster or is adsorbed at
the cluster surface. It should be checked whether the application of a MD-thermostat
to the carrier affects the simulation results.
Westergren et al. [15] analyzed several effects on the heat exchange between a cluster
and the carrier gas in MD simulation. They found an increasing energy transfer with
rising the atomic mass of the noble gas (acting as a carrier gas). By using different
forms of cluster-gas interaction potentials they also found that soft interactions are
more efficient in the heat transfer from the cluster to the gas. While Westergren et
al. employed “real” noble gas atoms, each having a distinct set of parameters of the
interaction potential and the atomic mass, one can also use a pseudo-noble gas, having
the same interaction parameters but different atomic masses [16]. This allows one to
separate the effect of the atomic mass from the effect caused by different interaction
parameters of the different noble gases. It is useful for the fundamental analysis of the
heat exchange to optimize the heat transfer in a simulation, however one should be
careful to directly draw conclusions for experimental systems. Figure 8.3 shows the
effect of the atomic mass, given in the diagrams in atomic units, on the cooling of the
largest cluster in the simulation system during a nucleation simulation. Temperature
8.2 Molecular Dynamics Simulation 123
jumps in these graphs are related to cluster-cluster collisions. In principle one can
identify two effects:
• The heavy atoms are slower than the light ones which means that the heavy atoms
move a smaller distance and, hence, less likely collide with a cluster than the light
ones. Hence, the heat exchange should become less efficient.
• On the other hand a collision with a heavy atoms itself is more efficient, i.e. more
heat is transferred.
As Fig. 8.3 shows, the effect of the mass on the velocity dominates, because the lighter
the carrier gas the more heat is removed from the clusters. In order to steer the
heat transfer during particle formation processes, one can also vary the amount of
the carrier gas. The more carrier gas is present, the more heat can be removed from the
forming clusters. Figure 8.4 shows the effect of the amount of carrier gas (Ar) on the
temperature of the zinc subsystem, which includes all zinc clusters in the simulation
box [17]. The horizontal line gives the temperature of the argon subsystem. The labels
are the ratio of Zn to Ar atoms in the system. The higher this ratio, the faster is the
convergence of the zinc temperature to the carrier gas temperature.
124 8 Computer Simulation of Nucleation
(a) (b)
Fig. 8.5 Comparison of the path during an expansion simulation for pure CO2 with the Span-
Wagner EoS (reference equation)[19]. In addition the expansion path of the CO2 /naphthalene solu-
tion is plotted. (Reprinted with permission from Ref. [14], copyright (2009) American Chemical
Society.)
8.2 Molecular Dynamics Simulation 125
simulation system agrees very well with the adiabatic expansion calculated with the
Span-Wagner EoS in all three coordinates. In both figures also the expansion path of
a dilute naphthalene solution in carbon dioxide is plotted. There is no such accurate
reference equation for this mixture but the expansion path is very close to that of
pure carbon dioxide. The small shift is related to the small amount of naphthalene
in the solution that affects the properties through its molecular interactions.
While in MD the actual equations of motion are numerically integrated, Monte Carlo
simulations sample the configurational space of the system. To be more specific, let
us consider a canonical (NVT ) ensemble of interacting particles (molecules). The
potential energy of a given configuration rN ≡ (r1 , . . . , rN ) is U(rN ). The average
value of an arbitrary function of coordinates X(rN ) is given by the integral
X(r ) =
N
X(rN ) w(rN ) drN (8.18)
1 −β U(rN )
w(rN ) = e (8.19)
QN
and
e−β U(r
N)
QN = drN (8.20)
Eq. (8.18) defines the mathematical expectation of X(rN ). Imagine that we have a
digital camera that can instantaneously take photos of the system so that we can
use this camera to scan and memorize the 3D coordinates of all N molecules in the
volume V . We can repeat these actions M times per second. Then, the computer
memory will contain the set of coordinates (rN )1 , (rN )2 , . . . , (rN )M , where M is a
number of configurations. The average observed value of X
M
AVRG(X) = (1/M) X[(rN )k ] (8.21)
k=1
gives an estimate of the true (exact) value X(rN ) given by (8.18), which we are
actually not able to calculate. The mean-square deviation
126 8 Computer Simulation of Nucleation
M
σ 2 = (1/M) X 2 [(rN )k ] − [AVRG(X)]2 (8.22)
k=1
Note, that σ becomes independent of the number of observations for large M, imply-
ing that the error of approximation (8.23) is inversely proportional to the square root
of the number of observations, which is typical for mathematical statistics.
MC simulation are based on the random generation of atom coordinates in a
simulation box. In view of the large number 3N of space coordinates of the
N-particle system it is clear, that calculations employing solely random choice of
these coordinates is a hopeless task. The introduction of the so-called importance
sampling by Metropolis et al. [21] allowed the sampling of the system with sufficient
accuracy within reasonable simulation time. As indicated by the term “importance
sampling”, not all states of the system are sampled with equal probability, but pre-
dominantly those that bring significant contribution to the configuration integral. The
criterion indicating the importance (statistical weight) of a state is its Boltzmann fac-
tor e−β U(r ) . Clearly, a very large positive U results is a very low statistical weight
N
of the state. Generating such term is hence a waist of time and should be avoided. In
molecular simulation the energy of the system can become very high if two atoms
overlap sufficiently. The repulsive part of the interaction potential is usually very
steep yielding high energies resulting in low probabilities of such states. Based on
these observations, instead of generating all coordinates of all atoms each time at
random, one can modify only those configurations that are already very likely. To do
so, one picks an atom and moves it at random throughout the simulation box. Then
the change of the configurational energy ΔU due to this (virtual) move is computed.
If the move leads to overlapping of atoms, it is clear that the energy would become
very high and the move should be rejected with high probability. On the other hand,
if the move lowers the energy of the system, it should be accepted. The resulting
Metropolis algorithm [21] is hence given by the following set of instructions:
if ΔU < 0
accept move
else
q = exp(−ΔU/kB T )
x = Random[0,1]
if x < q
accept move
else
reject move
8.3 Molecular Monte Carlo Simulation 127
endif
endif
Here Random [0,1] means a random number at the interval [0,1].
This is the core of the Metropolis scheme but, of course, a lot of additional numerical
procedures are needed to set up such an MC simulation. For example, the parameters
of the random atom move should be optimized during the simulation in order to
increase the efficiency. Periodic boundary conditions are required as in MD simula-
tions. A possible cutoff of the potential has to be corrected for. A difference between
MD and MC is that MC does not require calculation of the forces. Furthermore, the
“natural” ensemble for MC is the canonical NVT ensemble, (while in MD it is the
micro-canonical NVE ensemble). In order to switch to another ensemble in MC, one
has to modify the simulation procedure in connection with a modified Boltzmann
factor. For example, simulation of the NpT ensemble requires modification of the
simulation box at random and an additional term pV in the Boltzmann factor.
An advantage of MC compared to MD simulation is that in MC it is not necessary
to follow a real trajectory of an atom. The movement of an atom is in principle
random and hence it is possible to overcome high energy barriers that would be
impossible to overcome in MD simulations. For example, in MD an atom would not
be able to pass through two atoms which are close to each other, because that would
require a significant overlap. In MC a jump on the other side of the two atoms is
possible. However, even Metropolis sampling is in a number of cases not sufficient
to overcome high energy barriers most efficiently. A further improvement of MC
for such cases is the so-called umbrella sampling introduced by Torrie and Valleau
[22]. In this method the Boltzmann factor in the acceptance criterion is extended by
a weighting function w1 (rN ). The resulting transition probability is then
w1 (rN )e−βU(r
N)
π(rN ) = (8.24)
w1 (sN )e−βU(s ) dsN
N
The weighting function w1 (rN ) > 0, normalized to unity, is chosen in a way provid-
ing the reduction of the energy barrier which the original system has to overcome.
In order to calculate the desired properties in the canonical ensemble, it is necessary
to eliminate afterwards the effect of the weighting function. The thermodynamic
average X of a quantity X is then found from the relationship
wX1 w1
X =
w11 w1
(a) (b)
Fig. 8.7 Detection of clusters using the tWF- and Stillinger definitions. a Argon: there are 4
Stillinger clusters, within each of them dark spheres comprise a tWF-cluster; for example, the 5/22
cluster contains 22 particles according to Stillinger criterion from which only 5 comprise a tWF-
cluster. (Reprinted with permission from Ref. [26], copyright (2007), American Institute of Physics).
b Zinc: there are 2 Stillinger clusters, within each of them dark spheres comprise a tWF-cluster
(Reprinted with permission from Ref. [17], copyright (2007), American Institute of Physics.)
One of the most important properties, which can be obtained from molecular sim-
ulation of nucleation, is the nucleation rate. It is defined as the number of clusters
formed per unit time and unit volume, which continue to grow to the stable bulk
phase. In experiments the droplets or particles are usually counted by optical or scat-
tering methods. These methods require droplets which after the nucleation stage have
grown to a relatively large size, comparable to the wavelength of a laser beam. To
obtain a reliable nucleation rate estimate one has to perform the experiment under
conditions at which coagulation can be avoided because that would change the num-
ber of droplets. This can for example be accomplished with a low droplet density. In
molecular simulation the determination of the nucleation rate depends strongly on
the chosen simulation method.
The key quantity determining the nucleation behavior of a substance is the free energy
of an n-cluster formation, ΔG(n). Various theoretical models, discussed in this book,
invoke various approximations to derive this quantity; the most widely used one—
is the capillarity approximation of the classical nucleation theory. Calculation of
the free energy of cluster formation in Monte Carlo simulations, pioneered by Lee
et al. [28], is based on the analysis of cluster statistics, emerging in simulations,
without referring to a particular model for ΔG(n). Below we follow the procedure
outlined by Reiss and Bowles [29]. N molecules of the NVT -system can be grouped
in various clusters. We will consider the system configuration containing exactly Nn
clusters with n particles. Each n-cluster generates and exclusion volume vn which
is unaccessible for other N − n molecules of the system. Using the assumption of
non-interacting clusters (which is usually a good approximation for vapor–liquid
nucleation), we present the partition function of the NVT -system as
8.5 Evaluation of the Nucleation Rate 131
Here Zn (n, V , T ) is the partition function of one n-cluster in the volume V (see
Eq.(7.2)). Note, that Zn depends on the size of the system through the translational
degree of freedom of the center of mass of the cluster: Zn (n, V , T ) ∼ V /Λ3 . From
(8.25)–(8.26) using Stirling’s formula we have
For each cluster size n a variety of Nn is possible; we will be interested in the most
probable value. The latter maximizes ln Z with respect to Nn :
∂ ln Z Zn ∂ ln Z
= ln + =0 (8.27)
∂Nn N,V Nn ∂Nn
N,V
where μv and pv are the chemical potential and pressure of the vapor of volume
V − vn Nn containing N − nNn molecules. In general, μv = μv and pv = pv (μv is
the chemical potential of a molecule in the supersaturated vapor and pv is the vapor
pressure), however for rare clusters (recall the assumption of noninteracting clusters)
the difference between the barred and non-barred quantities is negligible. Therefore,
Eqs. (8.27) and (8.28) yield:
The ratio Zn (n, V , T )/V does not depend on V and remains constant in thermody-
namic limit. This means, that if we chose another volume of the system V , we would
have
132 8 Computer Simulation of Nucleation
Introducing the Helmholtz free energy of the n-cluster, confined to the volume vv
Nn
P(n) ≡ = exp −β (Fn (n, vv , T ) + pv vn ) − n μv (8.31)
N
The expression in the square brackets is
where G(n)bulk = n μv is the Gibbs free energy of n molecules in the bulk super-
saturated vapor prior to the formation of the cluster; G(n) is the same quantity after
the n-cluster was formed. Hence, ΔG(n) = G(n) − G(n)bulk is the “intensive Gibbs
free energy” of n-cluster formation and
P(n) = e−βΔG(n)
From these considerations we can schematically interpret the process of cluster for-
mation as consisting of two steps [29]:
1. n molecules are picked up anywhere in volume V of the system and gathered in
the volume equal to the molecular volume in the vapor phase vv ;
2. within the volume vv the cluster is formed with the volume vn < vv .
Thus, measuring the cluster-size probability distribution P(n) in MC simulation, we
can determine the free energy of cluster formation ΔG(n) from the relationship:
Its maximum gives the anticipated nucleation barrier ΔG∗ . This barrier can be com-
pared to the corresponding quantity resulting from nucleation theory. From Eq. (8.32)
one can determine the nucleation barrier, but not the kinetic prefactor, which deter-
mines the flux over this barrier and can be obtained from MD simulations.
In principle, Eq. (8.32) opens a possibility of calculating the nucleation barrier by
simulating the metastable vapor and counting clusters of various sizes. The total
number of particles used in modern MC simulations is of the order of N ∼ 105 −106 .
8.5 Evaluation of the Nucleation Rate 133
With this number of particles one can detect clusters from reliable statistics when
ΔG(n) < 10 kB T . Those are small n-mers, having the energy of formation of several
kB T . Only for extremely high supersaturations, close to pseudo-spinodal (discussed
in Chap. 9) will the height of the nucleation barrier be in this range (such high S
and therefore high nucleation rates are realized, e.g., in the supersonic Laval nozzle,
where J is in the range of 1016 − 1018 cm−3 s−1 [30]). For moderate supersaturations
nucleation barriers are in the range of ∼40 − 60 kB T and thus the clusters formed
in simulations are much smaller than the critical cluster. E.g. for ΔG∗ = 53 kB T the
chance to find a critical cluster is P = e−53 ≈ 10−23 , which means that the simulated
system should contain > 1023 particles which is equal to the Avogadro number. The
previously discussed umbrella sampling technique makes it possible to overcome
this difficulty. For a given n-cluster we introduce a weighting function w1 (rn ) which
according to Eq. (8.24) replaces the internal energy of the cluster U(rn ) by
1
W1 = kn (n − n0 )2 , kn > 0 (8.33)
2
It ensures that the formation of n-clusters with the sizes outside a certain range (char-
acterized by kn ) around n0 becomes highly improbable: the probability of finding the
n-cluster becomes proportional to
1
w1 = exp −β kn (n − n0 )2
2
This function forms a Gaussian umbrella in the space of cluster sizes, centered at n0 .
Only those clusters, which find themselves under this umbrella, will be sampled.
Thus, introduction of the biasing potential opens a “window” of the cluster sizes,
located at n0 , with a width of kn , which are sampled in simulations. By changing n0
one “opens consecutive windows” performing simulation runs within the windows,
thereby consecutively scanning the cluster size space.
of clusters one can draw a time dependent cluster statistics. This cluster statistics can
then be analyzed yielding the nucleation rate (besides other properties). For such an
analysis of the cluster statistics, several methods have been proposed in the literature.
Here we focus on two approaches which are commonly used in simulation studies
on nucleation.
The method of Yasuoka and Matsumoto [32], which also can be called the threshold
method, requires MD simulation system large enough to generate a significant amount
of clusters. This method can be obtained from the continuity equation in the space
of cluster sizes:
∂ ∂
N(n, t) = − j(n, t) (8.34)
∂t ∂n
Here N(n, t) is the number of clusters of size n in the simulation box at time t and
j(n, t) is the rate of the formation of clusters of size n in the box. In the steady state
∂
∂t N(n, t) = 0 yielding
∂
j(n, t) = 0
∂n
showing that j(n, t) is constant. If V is the volume of the simulation box, then the
steady-state nucleation rate is given by J = j(n, t)/V . In order to determine J from
the cluster statistics let us choose a certain, threshold, cluster size nthres and integrate
(8.34) over n from nthres to ∞:
∞ ∞
∂ ∂
N(n , t)dn = − j(n , t)dn (8.35)
∂t nthres nthres ∂n
The integral on the left-hand side is the total number of clusters with sizes larger
than nthres : ∞
N (nthres , t) = N(n , t) dn
nthres
where we took into account that the rate of formation of infinitely large clusters is
zero. Thus, Eq. (8.35) becomes
∂
N (nthres , t) = j(nthres , t)
∂t
8.5 Evaluation of the Nucleation Rate 135
(a) (b)
Fig. 8.8 Derivation of nucleation rate by means of the threshold method. a Cluster size dis-
tribution at time t: N(n, t); the shaded area under the curve gives the total number of clusters
N (nthres , t) larger than nthres at time t. b Time evolution of N (nthres , t). The slope of the domain
II ∂t∂ N (nthres , t) is proportional to the nucleation rate
1 ∂
J= N (nthres , t) (8.36)
V ∂t
In the threshold method the number of clusters larger than a threshold value,
N (nthres , t), is plotted as a function of the simulation time for different values
of nthres . As a result one obtains curves with four domains schematically depicted
in Fig. 8.8.
The slope of the linear domain II is ∂t∂ N (nthres ,t). Hence, the nucleation rate is
this value, divided by the simulation box volume V . The plateau-like domain III
and also the following descending part IV of the curve result from the finite size of
the system. The steady state situation is only possible as long as sufficient number
of monomers is present in the box to deliver the clusters of size nthres . At some
point, due to depletion of the vapor, the monomer concentration becomes too small
to provide further nucleation and at the same time clusters grow by collision. This
leads to stagnation and then decreasing of the number of clusters. Therefore, from the
analysis of the plateau domain of the data one can not derive the nucleation properties.
In practice one finds that linear parts of the curves are not necessarily all parallel.
One may use this fact for a rough estimate of the critical cluster size by calculating
the curve for each single threshold value and detect the threshold value of the cluster
size beyond which the slope of domain II does not change any more. Application of
Yasuoka-Matsumoto method to vapor-liquid nucleation of zinc is shown in Fig. 8.9.
Though the critical cluster is not known a priori, the threshold method can safely
be applied. By choosing various threshold values it is possible to detect the linear
domain of steady-state nucleation. If nthres > n∗ , each simulation curve exhibits a
linear domain where all N (t) lines are parallel (cf. Fig. 8.9). The average of the
slopes in the linear domain can be used to calculate the nucleation rate.
136 8 Computer Simulation of Nucleation
Fig. 8.9 Number of clusters larger than a certain size, indicated above each curve, for vapor-liquid
nucleation of zinc (for explanation see Fig. 8.8). The plot is used to derive the nucleation rate by
means of the threshold method. Zinc vapor density is 0.0315 mol/dm3 , temperature T = 400 K.
The resulting nucleation rate is J = 25.5 × 1028 dm−3 s−1 . (Reprinted with permission from Ref.
[17], copyright (2007), American Institute of Physics.)
The Mean First-Passage Time method (MFPT) provides an instruction to analyze the
stochastic dynamics of the nucleation process. In contrast to the threshold method, in
the MFPT a relatively small simulation system is sufficient to obtain the nucleation
rate. However, to get good statistics a large number of simulations is required. The
stochastic dynamics of a system with an activation barrier is governed by the Fokker-
Planck equation, describing the evolution of the cluster distribution function ρ(n, t)
caused by diffusion and drift in the space of cluster sizes, discussed in Sect. 3.7 (see
Eqs. (3.79)–(3.81)):
∂ρ(n, t) ∂ ∂ρ B(n) ∂ΔG(n)
= B(n) +ρ (8.37)
∂t ∂n ∂n kB T ∂n
Here B(n) is a diffusion coefficient in the space of cluster sizes. In Fig. 8.10 the work
of cluster formation ΔG(n) is plotted versus the cluster size. The solution of the
Fokker-Planck equation requires boundary conditions. In case of nucleation the left
boundary na is the monomer na = 1. It is called a reflecting boundary, since there
are no clusters smaller than a monomer. The right boundary nb is a size large enough
so that the cluster > nb has a negligibly small chance to evaporate. In view of this
feature, nb is called an absorbing boundary.
Let us fix na = 1 and an initial value n0 and follow the time evolution of the system
for various values of the absorbing boundary nb . For each nb we can identify the
mean first passage time τ (nb ) which is the average time necessary for the system,
starting at n0 , to leave the domain of cluster sizes (na , nb ) for the first time. Clearly,
8.5 Evaluation of the Nucleation Rate 137
na n0 n nb n
τ (nb = n∗ ) = τ ∗ is the average time necessary to reach the critical size. Since the
nucleation rate is the flux through the critical cluster, we may write
1 1
J= (8.38)
2 τ∗ V
where V is the volume of the simulation box; the factor 1/2 stands for the fact that
the critical cluster, corresponding to the maximum of ΔG(n), has a 50 % chance
of either growing to the new bulk phase or decaying, i.e. evaporating. Solving the
Fokker-Planck equation (8.37), Wedekind et al. [33] showed that for reasonably high
nucleation barriers the behavior of τ (nb ) in the vicinity of the critical size can be
approximated by the function
τJ
τ (nb ) = 1 + erf c nb − n∗ (8.39)
2
shown schematically in Fig. 8.11. Here
x
2
e−x dx
2
erf(x) =
π 0
and c is the inverse width of the critical region given by Eq. (3.49):
√
c = 1/Δ = πZ
na n0 n nb
providing the 3 fit parameters: τJ , n∗ and c. The nucleation rate being the reciprocal
of τJ reads:
1
J=
τJ V
An example of the MFPT curve obtained from simulation of zinc vapor–liquid nucle-
ation is shown in Fig. 8.12. Figures 8.10 and 8.11 demonstrate an advantage of the
MFPT method: one can recognize whether or not the nucleation process is coupled
to growth. If only the nucleation process takes place in the system, the simulation
data reach a plateau prescribed by the error function. Deviation from this shape is
related to the influence of particle growth on nucleation. In case of argon modelled
by the Lennard-Jones potential and small system sizes of approximately 300 atoms
it is possible to perform hundreds of simulations for averaging [33]. For larger sys-
tems with more complex interaction potentials the results have to be obtained from
a small number of simulation runs: e.g., for zinc nucleation (see Fig. 8.12) MFPT
results were obtained from 10 simulation runs.
While there are other methods to determine the nucleation rate from MD simula-
tions, the two methods described in this chapter are most frequently employed. Both
methods have advantages and drawbacks. The threshold method requires a large
simulation system in order to provide a significant amount of clusters to obtain good
statistics. If the systems are large enough possible depletion effects can be minimized.
The MFPT method requires a large number of simulation runs. In order to reach such
a large number each run should be sufficiently fast. This can be accomplished by ter-
minating a simulation at the point when the chosen cluster size is passed for the first
time. Another method for optimization is to choose a relatively small system since
only clusters not larger than few times n∗ are required for MFPT. On the other hand,
the smaller a simulation system is, the more important become finite sized effects.
8.6 Comparison of Simulation with Experiment 139
Fig. 8.12 MFPT analysis of a series of simulation runs for zinc at T = 800 K and log10 S = 2.79.
The critical cluster size, indicated by the dashed vertical line, is n∗ = 9. Because of the large system
size only 10 simulations were performed (Reprinted with permission from Ref. [17], copyright
(2007), American Institute of Physics.)
1 It must be noted that performing extensive simulations with up-to-date computers allows to
approach the region of supersonic nozzle experiments.
140 8 Computer Simulation of Nucleation
Fig. 8.13 Comparison of MD simulations, experimental data and CNT calculations for zinc vapor-
liquid nucleation. Filled symbols: calculations with the Yasuoka-Matsumoto method (“Yas”), open
symbols: calculations with MFPT method [17]. The dashed curve is the CNT prediction. An exper-
imental point from Ref. [34] is indicated by “exp”
Most molecular simulations of nucleation have been performed for pure substances;
much fewer computational studies have been devoted to binary systems.
MD simulation of nucleation in the binary vapor of iron and platinum performed
in Ref. [35] revealed that besides the mixed clusters, in which both components
are present, there are also single-component clusters containing pure iron and pure
platinum. Due to the difference in attraction strength between the two substances,
one observes a large amount of big platinum clusters compared to a relatively small
number of small iron clusters. The iron atoms have a weaker attraction compared
to platinum and either do not condense on hot platinum cluster or rapidly evaporate
after condensation.
Compared to metals, the binary mixture of n-nonane and methane is characterized
by much weaker van der Waals interactions. Nucleation in this mixture was studied
in MD simulations of Braun [36]. The peculiar feature of this system is a very low
vapor molar fraction of nonane: ynonane ≈ 10−4 ; meanwhile it is nonane that ensures
nucleation in the system. In order to tackle this problem, a simulation box containing
around 105 methane molecules was chosen and expansion simulation method was
used. From the bulk phase behavior one would expect a mole fraction of methane
in the liquid-like clusters to be around 0.2–0.4 at the given nucleation conditions
(high pressure and ambient temperature). At the same time simulation results show
that even for weakly interacting van der Waals systems the critical clusters are very
different from what one would expect from the bulk equilibrium. Clusters have
a peculiar structure resulting from minimization of the surface energy. The system
tends to lower its energy by phase separation and moving the more volatile component
(methane) towards the shell region of the cluster.
8.7 Simulation of Binary Nucleation 141
where Nna nb is the number of (na , nb )-clusters, N is the number of molecules in the
system. This expression is a generalization of the corresponding result (8.32) for the
unary case [37].
Kusaka et al. [38] performed simulations of nucleation in water-sulfuric acid sys-
tem which plays an important role in atmospheric processes. Simulations employed
water, hydronium ion, sulfuric acid and the bisulfate ion as species modeled by force
field combining the Lennard-Jones potential with electrostatic potential by means
of partial charges. From MC simulations the free energy of cluster formation was
analyzed along with the cluster structures. It was found that the shape of most of the
clusters differs from the spherical one. Different conformations of the clusters turn
out to be very close in energy and have a fairly long lifetime. This observation leads
to a conclusion that various clusters contribute to the nucleation rate.
Chen et al. [39] performed MC simulation of the nucleation in binary water/ethanol
systems. They found that ethanol is enriched in the cluster surface leading to a lower
surface tension than that of the ethanol/water mixture of the given mole fraction.
They argue that the shortcomings of the CNT in such cases might be related to this
surface enrichment.
did not contain inhomogeneities. Kimura and Maruyama [41] instead used a sur-
face composed by harmonically vibrating molecules coupled to a heat bath. They
investigated the nucleation of argon vapor for various vapor phase temperatures and
pressures. The nucleation rate was analyzed using the Yasuoka-Matsumoto threshold
method. Simulation results showed good agreement with the classical heterogenous
nucleation theory—the Fletcher model, discussed in Sect. 15.1.
In Ref. [42] the nucleation of argon on a polyethylene substrate was investigated. In
this work the polyethylene film in the center of the simulation box was coupled to a
thermostat. All latent heat of condensation was hence withdrawn from the system via
the substrate. In this way the process of heterogeneous nucleation can be modelled
realistically. In the transient stage of nucleation a temperature gradient develops,
which after condensation is complete vanishes again. Depending on the supersatu-
ration of the argon vapor, different types of growth take place. Comparison with the
classical heterogeneous nucleation theory exhibits good agreement.
The simplest possible model of interacting particles is the Ising model. It consists of a
lattice of spins that can have two values: either s = +1 or s = −1. Each spin interacts
with its nearest neighbors. Their number N1 depends on the type of the lattice. In the
simplest case of a 2D square lattice N1 = 4, while for a 3D cubic lattice N1 = 6.
Interaction between spins is given by the coupling constant K. If K is positive the
model describes ferromagnetic behavior favoring the alignment of neighboring spins
parallel to each other, while a negative K mimics antiferromagnetic behavior with
the preference for the anti-parallel alignment of neighboring spins. The Hamiltonian
of the Ising model reads:
H = −K si sj − H sk (8.40)
(i,j)nn k
Within this model molecules are only allowed to occupy the sites of a regular lattice
instead of continuous distribution in space. This requirement mimics a short-range
repulsion in a real fluid: molecules can not be closer than the lattice spacing. Attractive
interactions, i.e. the attractive well, is modelled by a nearest-neighbor potential , so
that the potential energy of a certain configuration takes the form
U = −ε ρi ρj
(i,j)nn
where ρk = 1 if the site k is occupied and ρk = 0 in the opposite case. Let us set
si = 2 ρi − 1
Then, in the Ising model si = −1 if the site i in the lattice gas model is free and
si = +1 if it is occupied. This transformation thus maps the fully occupied lattice
of spins si = ±1 on the partially occupied lattice of fluid molecules. The two
systems—lattice gas and Ising model—become thermodynamically equivalent, i.e.
their partition functions are the same, if we set [43]:
Thus, the Ising model can be used for simulation of a vapor–liquid system and
hence also for vapor–liquid nucleation. A metastable state of the spin system can be
achieved by varying the external field H which from (8.41) is equivalent to varying
of the chemical potential of a fluid molecule resulting in a supersaturation.
Glauber [44] was the first to study the kinetics of the 1D Ising model which is solvable
analytically, but does not exhibit a phase separation. Stoll et al. [45] performed
Monte Carlo simulations of the 2D spin-flip Ising model and analyzed its relaxation
towards equilibrium. Simulations demonstrated consistency with the dynamic scaling
hypothesis. Stauffer et al. [46] analyzed nucleation in 3D Ising lattice gas by Monte
Carlo simulations. They observed that the results obtained from the lattice model are
roughly in agreement with CNT.
The Ising model is very useful for the investigation of fundamental concepts of
nucleation. It can be employed to the analysis of the scaling behavior expressed in
the form of power laws. Since the pioneering works [45, 46] the model was employed
for various other systems—e.g. for heterogeneous nucleation [47], to mention just
one example. Meanwhile, it must be noted that simulation of real substances goes
beyond the scaling behavior and requires explicit force fields acting between the
molecules.
144 8 Computer Simulation of Nucleation
References
9.1 Introduction
At high supersaturations (deep quenches) the system from being metastable becomes
unstable; in the theory of phase transitions the boundary between the metastable
and unstable regions is given by a thermodynamic spinodal being a locus of points
corresponding to a divergent compressibility. Rigorously speaking the transition from
metastable to unstable states does not reduce to a sharp line but rather represents a
region of a certain width which depends on the range of interparticle interactions [1].
Within the spinodal region the fluid becomes unstable giving rise to the phenomenon
of spinodal decomposition [2], characterized by vanishing of the free energy barrier
of cluster formation at some finite value of the supersaturation. The classical theory
does not signal the spinodal: the nucleation barrier decreases with S but remains
finite for all values of S (see Eq. (3.28)). Therefore, nucleation in the spinodal region
can not be described by CNT and a more general formalism is needed.
Such a formalism, the field theoretical approach, was pioneered by Cahn and Hilliard
[3] and developed by Langer [4, 5], Klein and Unger [6, 7]. It is based on the
mean-field Ginzburg–Landau theory of phase transitions. Cahn-Hilliard’s approach
(usually termed a “gradient theory of nucleation”) leads to the existence of a well-
defined mean-field spinodal characterized by a supersaturation Ssp . A mean-field
theory becomes asymptotically accurate in the limit of infinite-range intermolecular
interactions, hence a spinodal line exists in the same limit. At the spinodal the barrier
vanishes which means that the capillary forces can no longer sustain the compact form
of a droplet, clusters in the vicinity of a spinodal are ramified fractal objects [6, 7].
In this chapter we formulate the mean-field (Cahn-Hilliard) gradient theory consid-
ering nucleation at high supersaturations. Special attention in this domain should be
paid to the role of fluctuations giving rise to the concept of pseudospinodal. Analysis
of nucleation near the pseudospinodal results in a generalized form of the classical
Kelvin equation (3.61) relating the size of the critical cluster to the supersaturation.
The starting point for the mean-field analysis of nucleation in the vicinity of the
thermodynamic spinodal is the Landau expansion of the free energy density in powers
of the order parameter m [8]
a 2 b 4
g = g0 + m + m −mh (9.1)
2 4
where
a = a0 t, t ≡ (T − Tc )/Tc , a0 > 0, b > 0
h is the external field conjugate to m. For the gas-liquid transition the order parameter
can be defined as
m = ρ − ρc
h = Δμ = μv ( p v ) − μl ( p v ) (9.2)
is the external field. At the spinodal h = h sp , while at the binodal the chemical poten-
tials of the phases are equal yielding h = 0. Below the critical point a < 0 and the
free energy density has a double-well structure shown in Fig. 9.1. In thermodynamic
equilibrium one should have
∂g ∂2g
= 0, >0
∂m ∂m 2
yielding
a m + b m3 − h = 0 (9.3)
a + 3b m 2 > 0 (9.4)
At h = 0, g(m) has two equal minima corresponding to the two coexisting phases.
For h = 0 the cubic equation (9.3) has a single real root if h 2 > h 2sp [9], where
4 a3
h 2sp = − (9.5)
27 b
This root refers to the single, stable, phase (liquid). If h 2 ≤ h 2sp , there are three real
roots; for h 2 = h 2sp two of them are equal. The left local minimum of the free energy
at m = m ∗ corresponds to the metastable state (supersaturated vapor), while the
9.2 Mean-Field Theory 147
h sp
h
h=0
metastable
g
stable
T< Tc
m * m sp 0 mglob
Fig. 9.1 Schematic plot of Landau free energy density g for T < Tc . At h = 0 (long dashed line)
there are two equal minima corresponding to the coexisting states. At 0 < h 2 < h 2sp (solid line)
the left, local, minimum at m ∗ corresponds to a metastable state (supersaturated vapor), while the
right, global, minimum refers to a stable state (liquid); the two states are separated by the energy
barrier. At h = h sp (short dashed line) the local minimum disappears—this is the case of spinodal
decomposition
right—global—minimum corresponds to the stable state (liquid); the two states are
separated by the energy barrier (these features are illustrated in Fig. 9.1). Expres-
sion (9.5) gives the maximum supersaturation corresponding to h = h sp , where
the local minimum of g becomes an inflection point—this is the case of spinodal
decomposition [2].
Nucleation takes place for 0 < h 2 < h 2sp . The discussion below refers to this case.
Since a < 0, it is convenient to introduce s = −a. Then the solutions of the cubic
equation (9.3) are [9, 10]:
s 2π
m(h) = 2 cos α + k , k = 0, 1, 2 (9.6)
3b 3
where √
3 3 b1/2 π
cos 3α = 3/2
h, 0 ≤ α ≤ (9.7)
2 s 6
Substituting (9.6) into the minimization condition (9.4) we obtain
cos α + 2π k > 1
3 2
which is satisfied for k = 0, 1 and is not satisfied for k = 2. The latter case cor-
responds to a maximum of g while the other two solutions correspond to the two
minima.
148 9 Nucleation at High Supersaturations
In order to determine which of these two roots refers to a local minimum (a super-
saturated state) we substitute (9.6) with k = 0 and k = 1 into (9.1) and after some
algebra obtain:
s2
g|k=0 = g0 + (1 + cos 2α)(1 − 3 cos 2α)
6b
s2 2π 2π
g|k=1 = g0 + 1 + cos 2 α + 1 − 3 cos 2 α +
6b 3 3
π
It is easy to check that for 0 ≤ α ≤ 6 : g|k=0 ≤ g|k=1 . Thus, the solution with k = 1,
s 2π
m ∗ (h) = 2 cos α + (9.8)
3b 3
corresponds to a metastable state of the system in an external field h, the free energy of
this state being g∗ = g|k=1 . The solution with k = 0 gives the global minimum cor-
responding to the thermodynamically stable liquid state to which the system evolves.
For the states close to m ∗ the free energy density can be expanded in powers of
φ = (m − m ∗ )/m ∗ :
b2 2 b3 3
g = g∗ + φ − φ + O(φ)4 (9.9)
2 3
where
∂ 2 g 1 ∂ 3 g
b2 (h) = m 2∗ , b3 (h) = −m 3∗ (9.10)
∂m 2 m=m ∗ 2 ∂m 3 m=m ∗
b2 (h = h sp ) = 0, b3 (h = h sp ) > 0 (9.11)
∂p ∂μ
=ρ
∂ρ ∂ρ
9.2 Mean-Field Theory 149
we have
∂ p
b2 = ρ (9.14)
∂ρ ρ=ρ v
showing that b2 is the inverse isothermal compressibility of the vapor at the given
metastable state.
Consider the behavior of the system near the mean-field spinodal. For this purpose
we construct an appropriate Ginzburg–Landau free energy functional which should
describe the state of the system undergoing a first order phase transition, characterized
by a scalar order parameter [11]
Now we allow for its spatial variations. Using (9.9), this functional reads:
c0 b2 2 b3 3
F [φ(r )] = F∗ + dr |∇φ|2 + φ − φ (9.15)
2 2 3
where F∗ is the free energy of the metastable state m = m ∗ (the local minimum of
the free energy), out of which nucleation starts. The square-gradient term in (9.15)
is an energy cost to create an interface between the phases; c0 > 0 is related to the
correlation length in the system [12] and can be well approximated by [13]
c0 ∼
1/3
= kB Tρc (9.16)
Following Unger [7], we associate the critical cluster with the saddle point of the
functional F [φ(r )]. If the saddle point is found, its substitution into (9.15) yields
the nucleation barrier
c0 b2 2 b3 3
W = F − F∗ = dr |∇φ|2 + φ − φ (9.17)
2 2 3
and denoting
ε = b2 (b32 c0 )−1/3 (9.18)
150 9 Nucleation at High Supersaturations
we rewrite (9.17) as
2/3
1 c0 ε 2 1 3
W = c0 dr |∇φ1 |2 + φ − φ
2 b3 2 1 3 1
yielding
c02 1 ε 1
W = dr1 |∇1 φ1 |2 + φ12 − φ13
b3 2 2 3
∂
where ∇1 = ∂r1 . And finally, further rescaling is useful:
∂
where ∇ = ∂r .
∇ 2φ = φ − φ2 (9.21)
The critical cluster is the nontrivial solution of (9.21) vanishing at infinity. The
existence of such solutions was proved for sufficiently large bounded domains [14].
Without presenting its full form it is instructive to study its behavior at large r , i.e.
far from the center of mass of the cluster. In this domain the amplitude of the droplet
is small and we can neglect the second term in (9.21) which leads to the equation
∇ 2 φ = φ, large r (9.22)
characterizes the size of the critical cluster. Equation (9.23) shows that within the
mean-field analysis the critical cluster size diverges as the spinodal is approached.
In the same limit the nucleation barrier (9.20) vanishes as
Finally, we must relate ε to the physical parameters of the system. To be more precise
we will determine scaling of ε near the spinodal to the leading order in (h − h sp ).
Obviously
ε(h = h sp ) = 0
From (9.10) and (9.18) it follows that ε is proportional to the curvature of the Landau
free energy at the metastable state:
∂ 2 g
ε ∼ b2 ∼ g2 (h) ≡ = −s + 3b m 2∗
∂m 2 m=m ∗ (h)
ε ∼ g2 = 2(3bs)1/4 h sp − h (9.25)
ln S
η= , 0≤η≤1 (9.26)
ln Ssp
Ssp (T ) is the upper boundary of S for nucleation at the temperature T ; its value
depends on the equation of state. For van der Waals fluids calculation of Ssp is
presented in Appendix C.
From (9.25)
ε ∼ (1 − η)1/2 (9.27)
Substituting (9.27) into (9.24) and (9.23), we find that in the vicinity of the spinodal
the nucleation barrier vanishes as
152 9 Nucleation at High Supersaturations
The excess number of molecules in the critical cluster is found from (9.28) using the
nucleation theorem (4.15):
Δn c ∼ (1 − η)−1/4 , η → 1− (9.30)
Δn c ∼ R∗ (9.31)
Previous discussions avoided an important conceptual question: how deep the quench
can be so that the concept of quasi-equilibrium (of the metastable state) can be
considered valid? In other words: what is the limit of validity of the mean-field
gradient theory, which completely neglects the effect of fluctuations? The answer to
this question can be found using the Ginzburg criterion for the breakdown of Landau
theory of phase transitions [12]. Very close to the spinodal fluctuations become
increasingly important and the mean-field theory of Sect. 9.2.2 breaks down. The
Ginzburg criterion determines the width of the domain near the spinodal, inside
which the mean-field considerations are violated. Such a thermodynamic analysis
was carried out by Wilemski and Li [15], who showed that for real fluids the Ginzburg
criterion is violated in the entire spinodal region, where the Landau expansion is
used. Having stated this, Wilemski and Li suggested that the concept of the mean-
field spinodal should be replaced by the concept of a pseudospinodal, introduced
earlier by Wang [16] in the study of polymer phase separation, which is associated
with a nucleation barrier ∼kB T .
The applicability of the mean-field approach can also be considered on the basis
of kinetic considerations. To do this let us compare two characteristic times: (i) the
time tM necessary to form a critical cluster which is a lifetime of the metastable state,
and (ii) the relaxation time tR during which the system settles in this state. The first
quantity can be related to the nucleation rate by using its definition: tM = 1/(J V ).
To find tR one must study the dynamics of the metastable state. Since the order
9.3 Role of Fluctuations 153
∂φ δF
= Γ0 ∇ 2 +ζ (9.32)
∂t δφ
where Γ0 is a transport coefficient and ζ (r, t) is a noise source (which models thermal
fluctuations) satisfying
to ensure that the equilibrium distribution associated with (9.32) is given by the
Boltzmann statistics. From the solution of (9.32) and (9.15), obtained by Patashinskii
and Shumilo [18, 19] (see also [13]), it follows that
16c0
tR =
Γ0 b22
implying that when the system approaches the thermodynamic spinodal (b2 → 0) its
relaxation time diverges. The relation between t M and t R established in [18, 19] is:
4π χ χW
t M = tR exp (9.33)
λ0 kB T
where
(b2 c0 )3/2
χ= (9.34)
kB T b32
χW ∼
=1 (9.35)
kB T
Beyond the kinetic spinodal the phase separation proceeds not via nucleation but via
the mechanism of spinodal nucleation [21], which differs both from nucleation and
spinodal decomposition. As one can see, the kinetic considerations are in agreement
with the thermodynamic analysis of [15]. Hence, in terms of the nucleation barrier
the pseudospinodal is similar to the kinetic spinodal.
The preceding discussion shows that the spinodal limit is hard to achieve in practice:
gradual quenching of the supersaturated vapor results in the barrier becoming equal
to the characteristic value of natural thermal fluctuations of the free energy, which
in a fluid is of the order of kB T ; at these conditions the time necessary to form the
154 9 Nucleation at High Supersaturations
critical cluster becomes comparable to the relaxation time during which the system
settles in the metastable state.
The mean-field gradient theory predicts the divergence of the critical cluster as S
approaches the spinodal. Meanwhile, experiments in the supersonic Laval nozzle
[22–26] with nucleation rates as high as 1017 − 1018 cm−3 s−1 do not support this
statement: the critical cluster, determined from the experimental J − S curves by
means of the nucleation theorem, continuously decreases with the supersaturation,
showing no signs of divergence up to the highest values of S. For those values
the critical cluster is a nano-sized object containing 5–10 molecules. A possible
explanation of this qualitative discrepancy was mentioned in the previous section:
the mean-field considerations fail in the entire spinodal region and the physically
relevant limit of the supersaturation is not the spinodal, but the pseudospinodal.
Therefore, we need to study nucleation in the vicinity of the pseudospinodal.
Can the CNT be useful for this study? Recall that in the CNT the critical cluster is
related to the supersaturation by the Classical Kelvin Equation (CKE) (3.63). For
every finite S it predicts a certain critical cluster yielding a certain finite nucleation
barrier (3.29). In other words the CNT does not signal either spinodal or pseudospin-
odal. This is not surprising since at high S the critical clusters become small and
obviously do not obey the capillarity approximation. For the analysis of the system
behavior in the vicinity of the pseudospinodal we need a generalization of the CKE
which extends the limit of its validity down to the clusters of molecular sizes; such
a generalization was proposed in Ref. [27].
The thermodynamic basis for the Kelvin equation is the metastable equilibrium
between the critical cluster and the surrounding supersaturated vapor and according
to (3.62) can be found from maximization of the Gibbs energy of cluster formation.
This is a general statement which holds irrespective of a particular form of the Gibbs
energy. Let us adopt the MKNT form for ΔG given by Eq. (7.77). Its maximum
leads to
dn s (n)
ln S = θmicro (9.36)
dn n c
which can be termed the Generalized Kelvin Equation (GKE). Taking into account
(7.63) and (7.66)), it is straightforward to see that in the limit of big clusters the GKE
recovers the Classical Kelvin Equation.
In order to illustrate the important features of the GKE let us consider the profiles
of ΔG(n) for various values of S at a fixed temperature T . To make the illustration
practically relevant, we refer to the experimental conditions for argon nucleation in
the supersonic Laval nozzle studied in Ref. [26].
9.4 Generalized Kelvin Equation and Pseudospinodal 155
30
Argon T=37.5 K
20
ln S=7
10
β ΔG(n) 0
8
-10
8.82
-20
0 10 20 30 40
n
Fig. 9.2 Gibbs free energy profiles ΔG(n) for argon at T = 37.5 K. Labels indicate the value of
ln S. Arrows indicate the critical cluster for each curve (corresponding to the second maximum of
ΔG(n). At ln S = 8.82 the second maximum disappears; the dashed red line gives the nucleation
barrier ΔG ∗ ≈ kB T )
Figure 9.2 shows the free energy profiles for T = 37.5 K and three different values
of ln S. Each curve ΔG(n) has two maxima (cf. Sect. 7.7): the first (left) one is
always at n = N1 (coordination number in the liquid phase) and is an artifact of the
MKNT, while the second (right) maximum corresponds to the critical cluster n c : in
Fig. 9.2 the critical cluster is indicated by the vertical error and the ‘ball’ on the top
of the curve. As S increases, both n c and the nucleation barrier ΔG ∗ = ΔG(n c )
decrease. At a certain supersaturation, ln S = 8.82, the second maximum disappears
which means that for supersaturations higher than this value the critical cluster does
not exist. In other words, there is an upper limit of S, beyond which there is no
nucleation. As one can see from the dashed red line, the nucleation barrier at this
value of S turns out to be ΔG ∗ ≈ kB T which corresponds to the pseudospinodal.
Below we will show that this is not a pure coincidence.
The free energy of formation of the critical cluster in MKNT takes the form (7.82)
which at the pseudospinodal gives
− n c ln S + θmicro [n s (n c ) − 1] = 1 (9.37)
Combination of (9.37) with the GKE Eq. (9.36) determines both n c and ln S at the
pseudospinodal. Excluding ln S, we find that n c satisfies the equation:
dn s 1
nc − n s (n c ) + 1 + υ = 0, where υ ≡ <1 (9.38)
dn n c θmicro
156 9 Nucleation at High Supersaturations
Before solving it let us discuss the domain of admissible values of n c . Suppose that
n c is large, then using the asymptotics (7.63) in Eq. (9.38) we find:
n c = [(1 + υ)/ω]3/2
where X (n) is the solution of Eq. (7.58). Then (9.38) can be rewritten as:
dX
[X (n c )] − 3 n c [X (n c )]
3 2
+1+υ =0 (9.39)
dn n c
From the previous discussion we can expect (the assumption to be verified later)
that near the pseudospinodal n c is small and close to the lower boundary of the
intermediate cluster range, i.e. it lies in the vicinity of N1 + 1. Correspondingly,
X (n) is close to unity. Presenting
X (n) = 1 + δ(n)
1
δ(n) = (n − N1 − 1) (9.40)
3q
q ≡ 1 + 2ω + ωλ (9.41)
Substituting (9.40) into (9.42), we find the critical cluster at the pseudospinodal:
√
1+ 1+τ
n c,psp = (N1 + 1) (9.43)
2
6q [(2 + υ)q − (N1 + 1)]
τ =
(N1 + 1)2
This result supports the assumption that n c is close to N1 + 1 (usually |τ | 1); the
cluster shows the liquid-like features only when it has a core, i.e. when n > N1 (T ).
9.4 Generalized Kelvin Equation and Pseudospinodal 157
Setting
n s (n) = n − [1 + 3δ(n)]
in Eq. (9.36) and using (9.40) we obtain the supersaturation at the pseudospinodal:
1
ln Spsp (T ) = θmicro 1− (9.44)
q
Figure 9.3 shows the classical (CKE) and generalized (GKE) Kelvin equation for
argon at T = 45 K and water at T = 220 K. The horizontal arrow points to
the pseudospinodal. For both substances the criterion (9.46) is satisfied up to the
pseudospinodal:
CKE and GKE become indistinguishable for clusters exceeding ∼200 molecules. In
terms of the cluster radius it corresponds to R ≈ 1.1−1.5 nm. This rather unexpected
result shows that the classical Kelvin equation may be still valid down to the clusters
containing ∼200 molecules. At large cluster sizes GKE approaches CKE: for argon—
from below, whereas for water—from above. The reason for this difference is the
sign of the Tolman length (7.101) which is negative for water at 220 K and positive
for argon at 45 K.
158 9 Nucleation at High Supersaturations
(a) 15 (b) 7
Argon Water
CK
T=45 K 6
E
T=220 K
C
ln Spsp
KE
10 5
ln S
ln S
ln S ps p 4
G
KE
5 GK 3
E
N1+1 2
N1+1
0 1
1 2 10 20 100 200 1 2 10 20 100 200
nc n
Fig. 9.3 Classical Kelvin equation (CKE) (dashed lines) and Generalized Kelvin Equation (GKE)
(solid lines). a Argon at T = 45 K; b Water at T = 220 K. The horizontal arrow indicates the value
of ln S at the pseudospinodal
MK
N
20
W ∗ = kB T characteristic of
the pseudospinodal conditions
10 nc nc, MKNT
,C
N T
βW =1
0
3 3.5 4 4.5 5 5.5 6
ln S
MKNT pseudospin.
ln Spsp =5.27
Figure 9.4 illustrates the behavior of the nucleation barrier W ∗ = ΔG ∗ and the
critical cluster size of water at T = 220 K (predicted by CNT and MKNT) as the
vapor approaches the pseudospinodal
Although CNT predicts smaller critical clusters than MKNT, the CNT barrier is
larger than the MKNT one, which is the manifestation of the fact that the formation
of small clusters is dominated by the microscopic surface tension rather than by the
macroscopic one.
References 159
References
Argon belongs to the class of so called simple fluids whose behavior on molecular
level can be adequately described by the Lennard-Jones interaction potential
σLJ 12 σLJ 6
uLJ (r) = 4εLJ − (10.1)
r r
where εLJ is the depth of the potential and σLJ is the molecular diameter; for argon
εLJ /kB = 119.8 K, σLJ = 3.40 Å [1]. Since argon plays an exceptional role in
various areas of soft condensed matter physics, its equilibrium properties have been
extensively studied experimentally [2], theoretically [3], in computer simulations—
Monte Carlo and molecular dynamics—[4, 5] and by means of the density functional
theory [6, 7].
Among various other issues, argon represents an important reference system for
non-equilibrium studies. In this context the phenomenon of nucleation is of special
significance. In the situation when no theoretical model can claim to be quantitatively
correct in describing nucleation in all substances under various external conditions,
argon can play a role of the test substance for which experimental, theoretical and
simulation efforts can be combined in order to obtain a better insight into the nucle-
ation phenomenon and abilities of various approaches to adequately describe it. This
chapter is aimed at obtaining a unified picture of argon nucleation combining theory,
simulation and experiment.
Early experimental studies of argon nucleation were carried out using various tech-
niques: cryogenic supersonic [8] and hypersonic [9] nozzles and cryogenic shock
tubes [10–12]. The data obtained in these experiments showed significant scatter
and results of various groups turned out to be inconsistent with each other.
The experimental situation was largely improved in 2006 due to the construction of
the cryogenic Nucleation Pulse Chamber (NPC) [13], and its further development
[14, 15]. This chamber uses a deep adiabatic expansion of the argon–helium mixture
which causes argon nucleation at temperatures below the triple point.
The onset nucleation data obtained in NPC for the temperature range 42–58 K are
reproducible and refer to the estimated nucleation rates 107±2 cm−3 s−1 . An impor-
tant breakthrough in nucleation measurements was achieved by construction of Laval
Supersonic Nozzle (SSN) [16], making it possible to accurately determine the onset
conditions corresponding to significantly higher nucleation rates. Argon nucleation
experiments in SSN [17], carried out in the temperature range 35–53 K which partly
overlaps the NPC range, correspond to higher supersaturations yielding the esti-
mated nucleation rates as high as 1017±1 cm−3 s−1 . In what follows we refer to the
experimental data obtained by these two techniques [14, 15] and [17].
Nucleation is an example of a rare-event process, that is why molecular dynamic
simulations at low temperatures are usually performed for very high supersaturations
in order to get a good statistics of nucleation events. Kraska [18] carried out MD
simulations in the microcanonical ensemble (MD/NVE) in the temperature range
30 K < T < 85 K with the nucleation rates JMD/NVE ∼ 1025 − 1029 cm−3 s−1 .
MD simulations of Wedekind et al. [19] in the canonical ensemble (MD/NVT) are
performed in the temperature range 45 K < T < 70 K with the nucleation rates in the
range of JMD/NVT = 1023 − 1025 cm−3 s−1 . All these simulations yield nucleation
rates which are far beyond experimental values of both NPC and SSN data.
In Chap. 5 we discussed DFT of nucleation and demonstrated its predictions for
Lennard-Jones fluids—see Fig. 5.5. In terms of argon properties these calculations
correspond to relatively high temperatures 83 K < T < 130 K. At this temperature
range the detectable nucleation rates require relatively low supersaturations. The
DFT nucleation rates lie in the range JDFT = 10−1 − 105 cm−3 s−1 . Figure 10.1
illustrates experimental, simulation and DFT studies in the T − S plane indicating
the corresponding typical values of J.
According to Chap. 9, the experimentally achievable upper limit of supersaturation
for nucleation at a temperature T is the pseudospinodal corresponding to the nucle-
ation barrier ΔG∗ ≈ kB T . The MKNT pseudospinodal given by Eq. (9.44) is shown
in Fig. 10.1 by the line labeled “psp”. As it is seen from Fig. 10.1 the SSN experi-
ments at low temperatures 37 K < T < 40 K are carried out in the pseudospinodal
region which implies that one can expect critical nuclei to be nano-sized objects
with the number of molecules close to the coordination number in the liquid phase.
In Fig. 10.2 the pseudospinodal is compared to the estimates of the thermodynamic
spinodal, corresponding to the limit of thermodynamic stability of the fluid. One way
to estimate the spinodal is to use a suitable equation of state (EoS) [24]. Dashed line
in Fig. 10.2 shows the spinodal calculated from the LJ EoS of Kolafa and Nezbeda
10.1 Temperature-Supersaturation Domain 163
MD/NVE
15 25-27
J MD/NVE ~ 10 exp. Iland (NPC)
MD/NVT
psp
exp. Sinha (SSN)
DFT
10 pseudosp. MKNT
J SSN
ln S
SS 16-1
~ 10
N
23-26
J MD/NVT ~10
8
5
J NP JD
NP
C FT ~ 10 2 +_ 3
C ~1
0 7-9
0
40 60 80 100
T (K)
Fig. 10.1 T -S domain of experiments and simulations. Nucleation Pulse Chamber (NPC) experi-
ments [14, 15] (blue squares), Supersonic Nozzle (SSN) experiments [16] (green squares), MD/NVE
simulations [18] (filled squares), MD/NVT simulations [19] (open rhombs) and DFT simulations
[20] (filled triangles). The line labelled “psp” is the MKNT pseudospinodal Eq. (9.44)
15
Argon
pseudospinodal
LJ spinodal
EoS spinodal
10 Spinodal, equilibr. MD
ln S
0
40 60 80 100
T (K)
Fig. 10.2 Pseudospinodal and estimates of the thermodynamic spinodal for argon. Solid line:
pseudospinodal Eq. (9.44); dashed line: the spinodal from the Lennard-Jones equation of state of
Ref. [21]; upper half-filled circles: the spinodal from the simulations of the supersaturated Lennard-
Jones vapor of Ref. [22]; lower half-filled circles: the spinodal from equilibrium simulations of Ref.
[23]. The vertical dashed-dotted line corresponds to the criterion (9.47)
164 10 Argon Nucleation
[21]. Extrapolations below the argon triple point Ttr = 83.8 K are limited because
EoS are usually fitted to experimental data only in the stable region.
Spinodal can be also found in computer simulations of Lennard-Jones fluids. Linhart
et al. [22] performed MD simulations of the supersaturated vapor of a LJ fluid and
obtained the spinodal pressure for the LJ temperature range 0.7 ≤ kB T/εLJ ≤ 1.2. For
argon this range corresponds to 84 K < T < 143 K. Spinodal was estimated by the
appearance of an instantaneous phase separation in the supersaturated vapor increas-
ing the argon density in a series of simulations. Simulations were performed for a
large cut-off radius 10 σLJ and the non-shifted LJ potential. A large cut-off radius
makes it feasible to apply the LJ simulation results to real argon.1 Unfortunately,
these simulations cover only partly the temperature domain of MD [18, 19] and
DFT [20]. Imre et al. [23] estimated the spinodal from the extremes of the tangential
component of the pressure tensor obtained from the simulations of the vapor-liquid
interface. This approach is based on a single equilibrium simulation without any
constraints and is applicable also below the triple point. Figure 10.2 indicates that
theoretical predictions of the pseudospinodal are consistent with the calculations of
thermodynamic spinodal performed by various methods: within the common tem-
perature range the MKNT pseudospinodal lies slightly below the instability points
of simulations and equation of state.
Let us discuss the predictions of various nucleation theories discussed in this book—
CNT (Chap. 3), EMLD-DNT (Chap. 6), MKNT (Chap. 7),—experiment (NPC and
SSN), MD simulations and DFT within the T − S domain bounded from above by
the pseudospinodal and for the temperatures corresponding to the range of validity
of MKNT (Eq. (7.53)): T < 92 K. Thermodynamic properties of argon are presented
in Appendix A.
The behavior of various model parameters is shown in Fig. 10.3. The bulk (macro-
scopic) surface tension θ∞ determines the surface part of the nucleation barrier in
the CNT and EMLD-DNT. It decreases with the temperature as well as the micro-
scopic surface tension θmicro used in MKNT. For all temperatures θ∞ > θmicro . The
difference between them can be substantial: e.g. at T = 70 K: θ∞ (T = 70 K) =
10.68, θmicro (T = 70 K) = 4.96; at higher temperatures this difference decreases.
The dashed line in Fig. 10.3 labeled “LJ” corresponds to the universal form of
γmicro /kB Tc for Lennard-Jones fluids (see Sect. 7.9.2). In view of methodological
reasons the experimental temperature-supersaturation domain does not overlap with
that of MD and DFT as clearly seen from Fig. 10.1. Therefore we perform separate
comparisons: theory versus experiment and theory versus MD and DFT [27].
1 It has been shown that the usually used cut-off radii of 5 σLJ and 6.5 σLJ are sufficient [25], while
2.5 σLJ gives significant deviation in the thermophysical properties [26].
10.2 Simulations and DFT Versus Theory 165
5
LJ micro /kB Tc
0
0.2 0.3 0.4 0.5 0.6 0.7 0.8
T/Tc
Figure 10.4 shows the ratio of nucleation rate log10 (Jsimul/Jtheor ), where Jsimul is the
nucleation rate found in MD simulations [18] and [19], and Jtheor refers to one of
the theoretical models. Open symbols correspond to Jtheor given by the CNT, and
filled symbols refer to the nonclassical models: MKNT and EMLD-DNT. The dashed
curve is the “ideal line” Jsimul = Jtheory . The agreement between simulations and
the nonclassical models in the whole temperature range is for most cases within 1–2
orders of magnitude, while the CNT rates are on average 3–5 orders of magnitude
lower than the simulation results. Figure 10.4 demonstrates that MKNT predicts a
better temperature dependence of the nucleation rates compared to EMLD-DNT.
An examination of MD results at T = 70 K shows that results obtained in NVE and
NVT simulations show a difference of one order of magnitude. There are two possible
reasons for this discrepancy. Firstly, in the NVT simulations the nucleation rate is
calculated from a mean first passage time analysis (MFPT) [28] while in the NVE
simulations the threshold method is employed. These two methods yield approxi-
mately one order of magnitude difference in the nucleation rate at given conditions
[29]. Since the nucleation rate obtained by the threshold method is larger, it is located
above the MFPT data. Secondly, in the NVE ensemble the latent heat heats up the
system allowing for the natural temperature fluctuations, while in the NVT simula-
tions velocity scaling is applied, which forces the system to stay at a fixed tempera-
ture thereby not allowing temperature fluctuations. Figure 10.5 compares theoretical
predictions with DFT of Ref. [20]. MKNT demonstrates a perfect agreement with the
DFT while both the CNT and EMLD-DNT underestimate DFT data by 3–5 orders
of magnitude. Recalling that nucleation rate is very sensitive to the intermolecular
interaction potential the agreement between MKNT and DFT is quite remarkable
166 10 Argon Nucleation
(a) 8
MD/NVE vs. MKNT (b) 8
S < S psp S < S psp MD/NVT vs. MKNT
MD/NVE vs. CNT
MD/NVT vs. CNT
6 MD/NVE vs. EMLD 6 MD/NVT vs. EMLD
4 4
2 2
0 0
-2 -2
-4 -4
70 75 80 85 60 65 70 75
T (K) T (K)
Fig. 10.4 MD simulations versus theory. a MD/NVE simulations of Ref. [18] versus theory. Open
circles: CNT, closed circles: MKNT, semi-filled squares: EMLD-DNT; b MD/NVT simulations
of Ref. [19] versus theory. Open triangles: CNT, filled upward triangles: MKNT, filled downward
triangles: EMLD-DNT
EMLD-DNT 4
-2
-4
85 90
T (K)
since DFT explicitly uses the interatomic interaction potential, while the MKNT is a
semi-phenomenological model using as an input the macroscopic empirical EoS, the
second virial coefficient, the plain layer surface tension and the coordination number
in the bulk liquid. Note, however, that the amount of available DFT data is insufficient
to formulate firm conclusions about the performance of different theoretical models.
Consider first experiments in the nucleation pulse chamber [14, 15]. The relative
nucleation rates together with the error bars of the experimental accuracy are shown
in Fig. 10.6. It turns out that for argon (being a simple fluid), predictions of the
10.3 Experiment Versus Theory 167
30 MKNT
CNT
EMLD
Log10(Jexp /J theor )
20
10
40 45 50 55 60
T (K)
Fig. 10.6 Argon nucleation experiments in Nucleation Pulse Chamber [14, 15] versus theory: CNT
(open circles), EMLD-DNT (filled squares) [19], MKNT (filled circles)
Fig. 10.7 Volmer plot for argon nucleation in nucleation pulse chamber and supersonic nozzle.
Hexagons: NPC data; diamonds: SSN data. Open symbols (hexagons and diamonds) refer CNT,
closed symbols (hexagons and diamonds) refer MKNT. The solid lines in the CNT graphs are shown
to guide the eye. (Reprinted from Ref. [17] copyright (2010), American Institute of Physics.)
168 10 Argon Nucleation
CNT fail dramatically: the discrepancy with experiment reaches 26–28 orders of
magnitude! This result looks even more surprising taking into account that experi-
mental nucleation points (see Fig. 10.1) are located far from the pseudospinodal. For
other models the results are somewhat better but remain poor: the disagreement with
experiment is 12–14 orders for EMLD-DNT and 4–8 orders for MKNT.
Nucleation experiments in the supersonic nozzle provide a possibility to reach the
vicinity of pseudospinodal thereby entering into the regime with extremely small
critical clusters. Comparison of SSN experiment to theories shows both quantitative
and qualitative differences with respect to the NPC results. Figure 10.7, taken from
Ref. [16], depicts the relative nucleation rate as a function of the inverse temperature
(the so-called Volmer plot). In this form the logarithm of the saturation pressure,
given by the Clapeyron equation (2.14), is approximately the straight line as well as
the lines of constant nucleation rate (for not too high rates). The upper curve (open
hexagons) reproduces the NPC data with respect to CNT (similar to the upper curve
of Fig. 10.6) as a function of inverse temperature. The qualitative difference between
NPC and SSN data is apparent: while the NPC data is a strongly decreasing function
of temperature, the SSN data (open diamonds) is a weakly increasing function of
temperature. Quantitative comparison of SSN data with theories reveals that all SSN
experiments are in perfect agreement with MKNT: the relative nucleation rates lie
within the “ideality domain”
Jexp
−1 < log10 <1
JMKNT
References
11.1 Introduction
11.2 Kinetics
(n a , n b ) + (1, 0) ↔ (n a + 1, n b )
(n a , n b ) + (0, 1) ↔ (n a , n b + 1)
J = (Ja (n a , n b ), Jb (n a , n b ))
(na,nb)
nb
Ja(na,nb)
na na
11.2 Kinetics 173
case, in which J is a scalar. From the assumptions made the kinetic equation describ-
ing the evolution of the cluster distribution function ρ(n a , n b , t) becomes
∂ρ(n a , n b , t)
= Ja (n a − 1, n b , t) − Ja (n a , n b , t) + Jb (n a , n b − 1, t) − Jb (n a , n b , t)
∂t
(11.1)
or in differential notations
∂ρ(n a , n b , t) ∂ Ja ∂ Jb
=− + (11.2)
∂t ∂n a ∂n b
The last expression manifests the conservation law for the number of particles and
can be rewritten as
∂ρ(n)
= − div J(n) (11.3)
∂t
In the steady state
div J = 0 (11.4)
As usual in the rate theories we write the fluxes along n a and n b axis in terms of the
forward (condensation) and backward (evaporation) rates:
Here νi , i = a, b is the impingement rate (per unit surface) of component i, i.e. the
rate of collisions of i-monomers with a unit surface of the cluster; βi , i = a, b is the
evaporation rate per unit surface of the component i, A(n a , n b ) is the surface area
of the (n a , n b )-cluster. The impingement rates for gas-liquid nucleation follow the
ideal gas kinetics (cf. (3.38))
yi p v
νi = √ (11.7)
2π m i kB T
νa ρeq (n a , n b )A(n a , n b )
βa = (11.8)
ρeq (n a + 1, n b )A(n a + 1, n b )
νb ρeq (n a , n b )A(n a , n b )
βb = (11.9)
ρeq (n a , n b + 1)A(n a , n b + 1)
Carrying out the kinetic procedure outlined in Sect. 11.2 requires considerable com-
putational effort due to the existence of a large number of nucleation paths. To proceed
with an analytical approach it is then necessary to identify the domain in the (n a , n b )
space bringing the major contribution to the nucleation rate. In its simplest form this
approach requires
1. identification of the “critical point” in the cluster space (corresponding to the
critical cluster),
2. determination of the direction of the flow in the critical point, and
3. making an assumption about the flow in the vicinity of the critical point
To cope with the problem of large number of paths contributing to the overall nucle-
ation rate, Reiss [4] showed that the nucleation rate is primarily determined by the
passage over the saddle point of the free energy surface ΔG(n a , n b ). This approx-
imation is based on the exponential dependence of ρeq on ΔG (recall that in the
single-component case the main contribution to the nucleation rate comes from the
vicinity of the maximum of ΔG(n)). This statement addresses the first question
raised above. Addressing the second one, Reiss suggested that the direction of the
flow at the saddle point is determined by the direction of the steepest descent of the
energy surface at this point, in other words this direction is determined solely by
energetic factors. The latter issue was later revisited by Stauffer [7] who showed that
the direction of the flow in the saddle point is also influenced by kinetics. Discussion
below follows Stauffer’s representation of binary nucleation kinetics [7].
11.3 “Direction of Principal Growth” Approximation 175
Let us start with presenting the kinetic equation in vector notations. Equations (11.5)–
(11.6) with the evaporation coefficients given by (11.8) and (11.9), can be written as
ρ(n)
J = −ρeq (n) F(n) ∇ , n = (n a , n b ) (11.11)
ρeq (n)
The steady-state nucleation rate is obtained by integration of Eq. (11.11) along all
possible nucleation paths subject to the boundary conditions
ρ ρ
lim = 1, lim =0 (11.12)
n a ,n b →0 ρeq n a ,n b →∞ ρeq
which are similar to the single-component case (see discussion in Sect. 3.3). Multi-
plying Eq. (11.11) from the left by (1/ρeq )F−1 and taking curl we obtain
1
curl (F−1 J) = 0 (11.13)
ρeq
for
1
a≡ , x ≡ F−1 J
ρeq
This equation determines the direction of the nucleation flux in any point (n a , n b )
of the cluster space. It implies that the direction of the nucleation flux depends not
only on the geometry of the energy surface ΔG(n a , n b ) but also on the impingement
rates of the components (through the matrix F).
The saddle point n∗ = (n a∗ , n ∗b ) of the free energy surface satisfies
∂ΔG ∂ΔG
= =0 (11.15)
∂n a n∗ ∂n b n∗
176 11 Binary Nucleation: Classical Theory
nb
y x
nb *
Jb
Ja
*
na * na
Fig. 11.2 Schematic illustration of the direction of principal growth approximation. Dashed lines:
curves of constant Gibbs free energy ΔG(n a , n b ). The origin of the x − y coordinate system corre-
sponds to the saddle point of ΔG. The angle ϕ gives the direction of principal growth determined
by Eq. (11.25); the angle ϕ ∗ is the approximation to ϕ given by Eq. (11.46)
where
1 ∂ 2 ΔG
m i = n i − n i∗ , Di j = , i, j = a, b
2 ∂n i ∂n j n∗
At the saddle point two eigenvalues of the symmetric Hessian matrix D have different
signs implying that
det D < 0
where ϕ is the (yet unknown) angle between the x and n a axis (see Fig. 11.2). The
rate components in the new coordinates are:
Jx
(y
y Jy )
=0
Then, the continuity equation (11.4) written in the rotated system yields
∂ Jx (x, y)
=0
∂x
implying that in the saddle point region the absolute value of the flux depends only
on y: Jx = Jx (y). Taking into account that in the same region ΔG is approximately
parabolic, Eq. (11.14) suggests that Jx can be cast in the form:
where Jx∗ is the flux at the saddle point and the dimensionless factor βW describes
the width of the saddle point region (Fig. 11.3). Together with the direction of the
flow, ϕ, it is found by substituting (11.19) into Eq. (11.14) which takes the form of
a linear combination
Qa ma + Qb mb = 0 (11.20)
with
Q a ≡ −w sin3 ϕ − wr sin ϕ cos2 ϕ + r cos ϕ + da sin ϕ (11.21)
Here
νb Daa Dbb
r= , da = − , db = − (11.23)
νa Dab Dab
and
W
w=− (11.24)
Dab
The solution of these equations for the two unknowns ϕ and w is:
1
tan ϕ = s + s2 + r , with s = (da − r db ) (11.25)
2
1 tan ϕ + r db
w= (11.26)
sin ϕ cos ϕ tan ϕ + tanr ϕ
Equation (11.25) states that the direction of the flux in the saddle point region is
determined from a combination of energetic and kinetic factors. The steepest descent
approximation of Ref. [4] would give cot(2ϕ) = (db − da )/2. This would agree with
Eq. (11.25) only when r = 1, i.e. when the impingement rates of the two components
are equal. Let us look at the limiting cases of large and small r . From (11.25)
1
tan ϕ = , for νb νa (11.27)
db
tan ϕ = da , for νb νa (11.28)
The remaining unknown quantity in (11.19) is the flux at the saddle point. To find it
let us write the vector equation (11.11) in the rotated system. Since Jy = 0 we are
interested only in the x-component of this equation which reads
1 ∂ ρ(x, y)
(F−1 J)x =− (11.29)
ρeq (x, y) ∂x ρeq (x, y)
where
Ja Jb
(F−1 J)a = , (F−1 J)b = (11.32)
νa A νb A
Using the standard linear algebra we express from (11.18) the “old” flux coordinates
in terms of the “new” ones, taking into account that Jy = 0:
1
(F−1 J)x = Jx (y) , A = A(x, y) (11.36)
νav A
Let us substitute (11.36) into (11.30); in view of the exponential dependence of ρeq
on x and y we can replace A = A(x, y) by its value A∗ at the saddle point and take
it out from the integral:
+∞ −1
∗ 1
Jx (y) = νav A dx (11.37)
−∞ ρeq (x, y)
Here p11 < 0, p22 > 0. The integral on the right-hand side of (11.37) reads
+∞ ∗ +∞
1 eg
dx = exp( p22 y 2 ) dx exp[ p11 x 2 + 2 p12 x y]
−∞ ρeq (x, y) C −∞
+∞ ∗
1 eg π 2 y2
p12
dx = exp + p22 y 2
−∞ ρeq (x, y) C (− p11 ) 4 (− p11 )
(− p11 ) −g∗
Jx∗ = C νav A∗ e (11.39)
π
2
p12
βW = + p22 (11.40)
4 (− p11 )
Finally, the total steady-state nucleation rate is given by Gaussian integration of Jx (y)
over y
+∞ π
J= dy Jx (y) = Jx∗ (11.41)
−∞ βW
resulting in
∗
J = K e−g (11.42)
∗
K = Z νav A(n ) C (11.43)
The prefactor K has the form analogous to the prefactor J0 in the single-component
case (cf. (3.54)) in which the impingement rate ν is replaced by νav . The Zeldovich
factor Z determines the shape of the Gibbs free energy surface in the saddle point
region:
1 (∂ 2 ΔG/∂ x 2 )n ∗
Z =− √ (11.44)
2 − det D
11.3 “Direction of Principal Growth” Approximation 181
Note, that although Eqs. (11.42)–(11.43) are similar to the single-component case,
it is not possible to recover the single-component nucleation rate from it by setting
one of the impingement rates to zero: this would lead to νav = 0. This result is a
manifestation of the general statement concerning the reduction of the dimensionality
of the physical problem. Such a reduction implies the abrupt change of symmetry
which can not be derived by smooth vanishing of one of the parameters of the system.
n i = n il + n iexc , i = a, b (11.47)
Each of the quantities in the right-hand side depend on the location of the dividing
surface while their sum can be assumed independent of this location to the relative
accuracy of O(ρ v /ρ l ), where ρ v and ρ l are the number densities in the vapor and
liquid phases. Therefore only n i ≥ 0 are observable physical properties; in this sense
the model quantities n il and n iexc can be both positive or negative.
In a unary system (see Sect. 3.2) we chose the equimolar dividing surface char-
acterized by zero adsorption n exc = 0. This choice made it possible to deal only
with the bulk numbers of cluster molecules. For a mixture, however, it is impossible
to choose a dividing surface in such a way that all excess terms n iexc vanish [9].
This is the reason for occurrence of the surface enrichment—preferential adsorption
of one of the species relative to the other. As a result the composition inside the
droplet can be different from that near its surface. For binary (and in general, multi-
component) nucleation problem introduction of the Gibbs surface is a nontrivial issue.
182 11 Binary Nucleation: Classical Theory
is the cluster volume, vil is the partial molecular volume of component i in the
liquid phase (see Appendix D), A is the surface area of the cluster calculated at
the location of the dividing surface; γ is the surface tension at the dividing surface.
Equation (11.48) presumes that formation of a cluster does not affect the surround-
ing vapor and therefore the chemical potential of a molecule in the vapor remains
unchanged. Within the capillarity approximation the bulk properties of the cluster
are those of the bulk liquid phase in which liquid is considered incompressible. This
yields:
μil ( p l ) = μil ( p v ) + vil ( p l − p v ) (11.49)
Then
n il μil ( p l ) − μiv ( p v ) = n il μil ( p v ) − μiv ( p v ) − ( p v − p l ) V l
i i
The last term in this expression cancels the first term in (11.48), leading to
ΔG = γ A − n il Δμi + n iexc μiexc − μiv ( p v ) (11.50)
i i
where
Δμi ≡ μiv ( p v ) − μil ( p v ) (11.51)
The last term in (11.50) contains the excess quantities. The chemical potentials μiexc
refer to a hypothetical (non-physical) surface phase and therefore can not be measured
in experiment or predicted theoretically. Therefore, one has to introduce an Ansatz
for them which serves as a closure of the model [12, 13]. The diffusion coefficient
11.4 Energetics of Binary Cluster Formation 183
in liquids is much higher than in gases (see e.g. [14]). This implies that diffusion
between the surface and the interior of the cluster is much faster than diffusion
between the surface and the mother vapor phase surrounding it. Hence, it is plausible
to assume equilibrium between surface and the interior (liquid) phase of the cluster,
resulting in the equality of the chemical potentials
Following [13] the Ansatz (11.53) can be termed “the equilibrium μ conjecture”.
Using (11.49) we may write
μiexc −μiv ( p v ) = μil ( p v ) − μiv ( p v ) +vil ( p l − p v ) ≡ −Δμi +vil ( p l − p v ) (11.54)
Substituting (11.54) into (11.50) and using Laplace equation we obtain an alternative
form of the Gibbs formation energy of the binary cluster
2γ (xbl )
ΔG = γ (xbl ) A − n il + n iexc Δμi + n iexc vil (11.55)
r
i i
ni
where γ (xbl ) is the surface tension of the binary solution of liquid composition xbl .
Here the second term contains the total numbers of molecules in the cluster, and not
the bulk liquid numbers n il . At the same time all the thermodynamic properties are
functions of the bulk composition xbl and not the total composition
Those two are not identical: xbl = xbtot . The same refers to the volume and the surface
area of the cluster
2/3
4π 3 l l
V =
l
r = n i vi , A = (36π ) 1/3
n il vil (11.57)
3
i i
since by definition any dividing surface has a zero thickness. An important feature
of the binary problem is the presence of the last term in the Gibbs energy (11.55).
In the previous section we derived the Gibbs energy of formation for an arbitrary
binary cluster. Consider now the critical cluster corresponding to the saddle point
of ΔG:
184 11 Binary Nucleation: Classical Theory
∂ΔG
= 0, j = a, b (11.58)
∂n lj
n∗
∂ΔG
= 0, j = a, b (11.59)
∂n exc
j n∗
Since the formation of a cluster does not affect the vapor properties, we have
dμiv = dp v = 0 (11.60)
∂μl ( p v )
∂A ∂γ exc ∂μi
exc
i
γ + A −Δμ j + n l
i + n i = 0, j = a, b (11.61)
∂n lj ∂n lj i
∂n lj i
∂n lj
∂γ ∂μiexc l ∂μil ( p v )
A −(μvj −μexc
j )+ n iexc exc + ni = 0, j = a, b (11.62)
∂n j
exc ∂n j ∂n exc
j
i i
We rewrite it as
∂γ exc ∂μi
exc
A + n = 0, α = l, exc (11.64)
∂n αj i
∂n αj
i
Gibbs-Duhem equation (2.5) for the bulk liquid phase of the mixture at constant T is
−V l d p l + n il dμil ( p l ) = 0
i
resulting in
∂μil ( p v )
n il = 0, α = l, exc (11.66)
∂n αj
i
11.5 Kelvin Equations for the Mixture 185
Substituting (11.64) and (11.66) into Eqs. (11.61)–(11.62) we obtain for the saddle
point:
∂A
γ − Δμ j = 0 (11.67)
∂n lj
μexc
j (p ) = μj(p )
l v v
(11.68)
The meaning of the second equality is transparent. As we discussed, for any cluster
the molecules belonging to the dividing surface are in equilibrium with the interior of
j ( p ) = μ j ( p ). For the critical cluster this condition is supplemented
the cluster μexc l l l
j (p ) = μj(p ) = μj(p )
μexc l l l v v
2γ val
− Δμa + =0 (11.69)
r∗
2γ v l
−Δμb + ∗ b = 0 (11.70)
r
This set of equations is known as the Kelvin equations for a mixture; they determine
the composition and the size of the critical cluster. In particular, the critical cluster
composition satisfies
Δμa Δμb
= l (11.71)
va
l vb
∗
Once the composition xbl is determined, the critical radius is given by
2γ v lj
r∗ = (11.72)
Δμ j
Here the surface tension γ refers to the dividing surface of the radius r ∗ .
Substituting the Kelvin equations into (11.55) we find the Gibbs free energy at the
saddle point (or, equivalently, the nucleation barrier):
2γ v l 2
∗ l∗
ΔG = γ A + ni − ∗i =γ A − Vl
r r
i
resulting in:
1
ΔG ∗ = γA (11.73)
3
186 11 Binary Nucleation: Classical Theory
where both quantities on the right-hand side depend on the critical cluster composi-
tion. Within the phenomenological approach droplets are considered to be relatively
∗
large, so that one can replace γ by γ∞ (xbl )—the surface tension of the plain layer
of the binary vapor- binary liquid system when the composition of the bulk liquid is
that of the critical cluster.
Note that if within the capillarity approximation we would set n aexc = n exc
b = 0,
then the terms with the excess quantities in the free energy would disappear and
the Gibbs adsorption equation (11.63) can not be invoked. The resulting equations
for the critical cluster would contain then the uncompensated term with the surface
tension derivative:
2γ∞ v lj ∂γ∞
− Δμ j + +A = 0, j = a, b (11.74)
r∗ ∂ x tot
j
11.6 K-Surface
In the general expression for the Gibbs energy (11.50) (or its equivalent form (11.55))
the bulk and excess numbers of molecules are not specified and treated as independent
variables. Their specification is related to a choice of the dividing surface for a cluster.
This is not a unique procedure. One of the appropriate options is the equimolar surface
for the mixture, termed also the K -surface [13, 15], defined through the requirement
n iexc vil = 0 (11.75)
i=a,b
This choice ensures that the macroscopic surface tension is independent of the
curvature of the drop; however it does depend on the composition of the cluster.
This can be easily seen if we present Eq. (11.55) in the form
ΔG = − n il + n iexc Δμi + γ (xbl ; r ) A
i
ni
For the K -surface the second term in the curl brackets vanishes implying that
γ (xbl ; r ) = γ (xbl ). Laaksonen et al. [15] showed that the K -surface brings together
various derivations of the free energy of cluster formation in the classical theory—
due to Wilemski [10], Debenedetti [16] and Oxtoby and Kashchiev [17]. In what
follows we adopt the K-surface formalism for binary clusters. Usually, the partial
molecular volumes of both components in the liquid phase are positive, implying
from (11.75) that the excess quantities n iexc have different signs. As we mentioned
already, a negative value of one of n iexc is not unphysical as soon as the total num-
ber n i (the quantity which does not depend on the choice of dividing surface) is
nonnegative.
Within the K -surface formalism, the volume of the cluster and its surface area can
be expressed either in terms of n il or in terms of the total numbers of molecules
n itot ≡ n i . Equation (11.57) reads:
Vl = n il vil = n itot vil (11.76)
i i
2/3 2/3
A = (36π ) 1/3
n il vil = (36π ) 1/3
n itot vil (11.77)
i i
2γ vil
μil ( p l , xbl ) = μil ( p v , xbl ) + (11.80)
r
where for brevity we used the notation μi ≡ μil ( p v , xbl ). Partial molecular volumes
can be written as (see Appendix D):
188 11 Binary Nucleation: Classical Theory
1 l
vil = η (11.82)
ρl i
with ηil given by (D.7)–(D.8). Combining (11.81) and (11.82) with the Gibbs-Duhem
equation
xil dμi = 0 (11.83)
i
we find
−1
∂γ 1 ∂μa 2γ ηal ∂ ln(ηal /ηbl )
n aexc = −A l + (11.84)
∂ xb xbl ηbl ∂ xbl r ρl ∂ xbl
−1
∂γ 1 ∂μb 2γ ηbl ∂ ln(ηbl /ηal )
n exc
b = −A l + (11.85)
∂ xb xal ηal ∂ xbl r ρl ∂ xbl
∂μal 1
= −kB T l (11.86)
∂ xb
l x a
∂μlb 1
= kB T (11.87)
∂ xbl xbl
It is easy to see that these expressions satisfy the Gibbs-Duhem relation (11.83).
A more accurate approximation can be formulated using one of the more sophisticated
models for activity coefficients—e.g. van Laar model discussed in Sect. 11.9.1.
Finally, the terms ∂ ln(ηil /ηlj )/∂ xbl in (11.84)–(11.85) are found from (D.7)–(D.9):
∂ ln(ηal /ηbl ) 1
= (τ2 − τ12 ) (11.88)
∂ xbl ηal ηbl
∂ ln(ηbl /ηal ) ∂ ln(ηal /ηbl ) 1
=− =− (τ2 − τ12 ) (11.89)
∂ xbl ∂ xbl ηalηbl
where
∂ ln ρ l ∂τ1
τ1 = , τ2 = l
∂ xbl ∂ xb
If the excess numbers, derived using this procedure, turn out to be not small compared
to the bulk numbers then the classical theory, probably, falls apart [8]. This happens in
11.6 K-Surface 189
the mixtures with strongly surface active components exhibiting pronounced adsorp-
tion on the K -surface. An example of such a system is the water/ethanol mixture
(discussed in Sect. 11.9.3).
If we choose the K -dividing surface, the Gibbs energy of cluster formation (11.55)
becomes
ΔG = γ (xbl , T ) A − n il + n iexc μiv ( p v , T ) − μil ( p v , xbl , T ) (11.90)
i
ni
Within the K -surface formalism for each pair of bulk cluster molecules (n al , n lb ) the
excess quantities are constructed
n aexc (n al , n lb ), n exc
b (n a , n b )
l l
implying that n il and n iexc are not any more the independent quantities. Therefore, the
saddle point of ΔG has to be determined in the space of independent variables—the
total number of molecules:
∂ΔG
= 0, j = a, b (11.91)
∂n j
leading again to the Kelvin equations (11.69)–(11.70). From the first sight it may
seem that to construct ΔG(n a , n b ) according to Eq. (11.90) one needs to know
only the total numbers of molecules n a and n b ; however this is not true, since the
thermodynamic properties—the surface tension γ , partial molecular volumes vil and
chemical potentials μil —depend on the bulk composition xil rather than on xitot . So,
in order to calculate ΔG(n a , n b ), one has to know also the bulk numbers n al and n lb
(and therefore the bulk composition xbl ), giving rise to these n a and n b .
Equation (11.90) formally coincides with the BCNT expression
ΔG BCNT = γ (xbtot , T ) A − n i μiv ( p v , T ) − μil ( p v , xbtot , T ) (11.92)
i
except for the argument of μil and γ . This means that the standard BCNT does not
discriminate between the bulk and excess molecules in the cluster and thus does not
account for adsorption effects: a cluster in this model is a homogeneous object.
190 11 Binary Nucleation: Classical Theory
The Gibbs free energy contains the chemical potentials of the species which are not
directly measurable quantities. Therefore, it is necessary to cast ΔG in an approxi-
mate form containing the quantities which are either measurable or can be calculated
from a suitable equation of state. Let us first recall that μiv is imposed by external
conditions and does not depend on the composition of the cluster; at the same time
μil is essentially determined by the cluster composition. For this quantity using the
incompressibility of the liquid phase we may write
where p0 is an arbitrary chosen reference pressure. Let us choose it from the condition
of bulk (xbl , T )-equilibrium. The latter is the equilibrium between the bulk binary
liquid at temperature T having the composition xbl and the binary vapor. Fixing
xbl and T , we can calculate from the EoS the corresponding coexistence pressure
p coex (xbl , T ) and coexistence vapor fractions of the components yicoex (xbl , T ). Now
we choose p0 as
p0 = p coex (xbl , T )
For the gaseous phase we assume the ideal mixture behavior, implying that each
component i behaves as if it were alone at the pressure piv = yi p v . Then the chemical
potential of component i in the binary vapor is approximately equal to its value for
the pure i-vapor at the pressure piv :
μiv ( p v , yi ) ≈ μi,pure
v
( piv ) (11.95)
where viv ( p) is the molecular volume of the pure vapor i and pi is another arbitrary
reference pressure. Let us choose it equal to the partial vapor pressure of component
i at (xbl , T )-equilibrium:
yi p v
μiv ( p v , yi ) = μi,pure
v
( picoex ) + kB T ln (11.98)
yicoex (xbl ) p coex (xbl )
The quantity in the curl brackets is proportional to the liquid compressibility factor,
which is a small number (∼10−6 − 10−2 ), implying that the second term can be
neglected in favor of the first one:
yi p v
βΔμi = ln (11.99)
yicoex (xbl ) p coex (xbl )
Substituting (11.99) into (11.90), we deduce the desired approximation for the free
energy containing now only the measurable quantities
yi p v
βΔG(n a , n b ) = − n i ln + β γ (xbl ) A (11.100)
i
yicoex (xbl ) p coex (xbl )
This result coincides with the BCNT expression [4] except for the argument xbl of
the coexistence properties:
yi p v
βΔG BCNT (n a , n b ) = − n i ln + β γ (xbtot ) A
yicoex (xbtot ) p coex (xbtot )
i
(11.101)
Equations (11.100)–(11.101) imply that for an arbitrary (n a , n b )-cluster the bulk
(logarithmic) terms can be both positive and negative depending on the cluster com-
position. Consequently, even at equilibrium conditions (say, at a fixed p v and T ) the
free energy of formation of a cluster with a composition, different from the equi-
librium bulk liquid composition at given p v and T , will contain a non-zero bulk
contribution. This situation is distinctly different from the single-component case,
for which formation of an arbitrary cluster in equilibrium (saturated) vapor is asso-
ciated only with the energy cost to build its surface and the bulk contribution to ΔG
vanishes.
192 11 Binary Nucleation: Classical Theory
where ρiv is the number density of monomers of species i in the vapor. Such a choice,
however, violates the law of mass action (recall the similar feature of the single-
component CNT discussed in Sect. 3.6). Another difficulty associated with (11.102),
is that the number density of pure a-clusters, ρeq (n a , 0), becomes proportional to
the number density of b-monomers and vice versa. Wilemski and Wyslouzil [18]
proposed an alternative form of C which is free from these inconsistencies. It was
suggested that C should depend on the cluster composition:
x tot x tot
CWW = ρav,coex (xatot ) a ρbv,coex (xatot ) b , xatot + xbtot = 1 (11.103)
where ρiv, coex (xitot ) is the equilibrium number density of monomers of species i in
the binary vapor at coexistence with the binary liquid whose composition is xitot .
Another possibility discussed by the same authors, is the self-consistent classical
(SCC) form of C based on the Girshick-Chiu ICCT model (3.103):
x tot x tot
CWW,SSC = exp xatot θ∞,a + xbtot θ∞,b ρav, coex (xatot ) a ρbv, coex (xatot ) b
(11.104)
where θ∞,i (T ) is the reduced macroscopic surface tension of pure component i.
Mention, that Eqs. (11.103), (11.104) are just two of possible choices of the
prefactor C.
in the vapor by its chemical potential μiv , the gas-phase activity of component i is
defined through
Aiv = exp β(μiv − μi,0 v
) (11.105)
The most frequent choice of the reference state is the vapor-liquid equilibrium of
pure component i at temperature T (assuming, of course, that such a state exists!).
Then
pi = psat,i (T ), μi,pure
v
( pi ) = μi,0
v
= μsat,i (11.107)
where μsat,i (T ) and psat,i (T ) are the saturation values of the chemical potential and
pressure of component i, respectively. With this choice (11.105) yields
piv
Aiv = (11.108)
psat,i (T )
AL
ln Γa = 2
A L xb
1+ B L xa
BL
ln Γb = 2
B L xa
1+ A L xb
where the van Laar constants A L and B L are determined from the equilibrium vapor
pressure measurements.
The liquid-phase activity of component i is defined as
picoex (xi , T )
Ail = = Γi xi (11.110)
psat,i (T )
This quantity describes the influence of the bulk liquid composition on the equi-
librium vapor pressure of components. For an ideal mixture Γi = 1, yielding
Ai,l ideal = xi , and for a single-component case (which is equivalent to the ideal
mixture with xi = 1): Ail = 1. In terms of activities Eqs. (11.100)–(11.101) read
Aiv
βΔG(n a , n b ) = − n i ln + β γ (xbl ) A (11.111)
i
Ail (xil )
Aiv
βΔG BCNT (n a , n b ) = − n i ln + β γ (xbtot ) A (11.112)
i
Ail (xitot )
Aiv
βΔG BCNT (n a , n b ) = − n i ln + β γ (xbtot ) A (11.113)
xitot psat,i (T )
i
11.9 Illustrative Results 195
The nucleation rate is given by Eqs. (11.42)–(11.43) with the Stauffer form of the
prefactor K , and C = CReiss . In Fig. 11.4 we compare BCNT with the experimental
results of Strey and Viisanen [19] for nucleation of ethanol–hexanol mixture in argon
as a carrier gas at T = 260 K. Thermodynamic parameters used in calculations are
taken from Table I of Ref. [19]. The rates are plotted against the mean vapor phase
activity
a= (AEv )2 + (AHv )2 (11.114)
where AEv and AHv are the ethanol and hexanol vapor phase activities, respectively.
The labels in Fig. 11.4 indicate the activity fractions
AHv
y= (11.115)
AEv + AHv
The line y = 0 corresponds to the pure ethanol nucleation and y = 1—to the pure
hexanol nucleation. As one can see, BCNT is in good agreement with experiment for
y < 0.9 but substantially underpredicts the experimental data for y → 1, i.e. at the
pure hexanol limit. It is instructive to present the same data as an activity plot. The
latter is the locus of gas phase activities of components required to produce a fixed
nucleation rate at a given temperature. Figure 11.5 is the activity plot corresponding
to the nucleation rate J = 107 cm−3 s−1 . The difference between the Stauffer from
of K and that of Reiss is quite small indicating that for this system the direction
of cluster growth at the saddle point is to a good approximation determined by the
steepest descend of the Gibbs free energy.
196 11 Binary Nucleation: Classical Theory
Fig. 11.5 Activity plot for ethanol–hexanol nucleation at T = 260 K. Activities a E = AEv and
a H = AHv correspond to the nucleation rate J = 107 cm−3 s−1 . Squares: experiment of Strey and
Viisanen [19]; solid lines—BCNT with Stauffer’s expression for K ; short dashed line—BCNT with
Reiss expression for K ; long dashed line—BCNT with SCC form of the prefactor C (11.104) and
Stauffer’s expression for K (Reprinted with permission from Ref. [18], copyright (1995), American
Institute of Physics.)
where n a,c and n b,c are the (total) numbers of molecules of components in the critical
cluster (we neglected the contribution of the prefactor K which is between 0 and 1).
Recalling the definition of gas phase activities, this expression can be written as
11.9 Illustrative Results 197
∂ ln J
Δn a,c = (11.117)
∂(ln Aav ) A v
b
∂ ln J
Δn b,c = (11.118)
∂(ln Abv ) A v
a
The first and the third term on the left-hand side of this expression can be written
using Eq. (11.116):
∂ ln Aav 1
Δn a,c = −1 (11.120)
∂ ln Abv ln J,T Δn b,c
resulting in
∂ ln Aav Δn b,c
=− (11.121)
∂ ln Abv ln J,T Δn a,c
The left-hand side of (11.121) gives the slope of the activity plot ln Aav = f (ln Abv ).
Since Δn a,c , Δn b,c > 0 this slope should be negative:
∂ ln Aav
<0 (11.122)
∂ ln Abv ln J,T
Fig. 11.6 Activity plot for the water/ethanol mixture corresponding to the nucleation rate J =
107 cm−3 s−1 and T = 260 K. Aw,g and Ae,g are the gas-phase activities of water and ethanol,
respectively. Points: experiment of Viisanen et al. [21], full line: BCNT predictions (Reprinted with
permission from Ref. [8], copyright (2006), Springer-Verlag.)
∂ ln J
<0
∂ ln Aiv A v ,T
j
Since Aiv is proportional to the vapor pressure (see Eq. (11.108)), the last inequality
would mean the decrease of the nucleation rate when the vapor density is increased.
The origin of this unphysical behavior lies in the prediction of the critical cluster
composition: BCNT predicts very water-rich critical clusters. Since the surface ten-
sion of pure water is much higher than that of the pure ethanol, the resulting surface
tension of the mixture becomes very high yielding low nucleation rates. These results
show that BCNT fails in describing the behavior of surface enriched nuclei. For this
system the adsorption effects, not taken into account by the BCNT, play an important
role: surface excess numbers n iexc turn out to be large and fluctuating.
Lf = 0 (p0,T0,y)
-
-
p -
-
coexistence -
-
- -
region (pv ,T,y) vapor
vapor + liquid
total pressure p v < p0 and temperature T < T0 ; the latter is characterized by the
equilibrium values of the thermodynamic parameters: the chemical potentials of
the species and their vapor and liquid molar fractions. The actual vapor composition
differs from the equilibrium one at the same p v and T showing that the mixture finds
itself in a nonequilibrium state. If p v is sufficiently high, the supercritical component
not only removes the latent heat (acting as a carrier gas) but also takes part in the
nucleation process due to the unlike a − b interactions becoming highly pronounced
at high pressures. These strong real gas effects attract considerable experimental
[24–28] and theoretical [29–31] attention.
As an example of a system showing retrograde nucleation behavior we consider the
n-nonane/methane mixture. The choice of the system is motivated by the availability
of data obtained in expansion wave tube experiments [32–36] carried out at the
nucleation pressures ranging from 10 to 40 bar and temperatures—from 220 to 250 K.
In this range methane is supercritical: Tc,b = 190 K [14]. In the vapor phase methane
is in abundance, yb ≈ 1, while ya ∼ 10−4 ÷ 10−3 . Let us first study the pressure
dependence of equilibrium properties influencing the nucleation behavior. For these
calculations we need an EoS. The most appropriate one for mixtures of alkanes is
the Redlich-Kwong-Soave equation [37].
As follows from Fig. 11.8, the miscibility of methane xb,eq in the bulk liquid grows
with the pressure and can be as high as ≈50 % for p v = 100 bar. The process of
methane dissolution in liquid nonane is accompanied by the decrease of the reduced
macroscopic surface tension θ∞ of the mixture. As opposed to the previously dis-
cussed examples, θ∞ can not be expressed as a sum of the corresponding individual
properties, θ∞,a and θ∞,b , since methane is supercritical. The surface tension for the
mixture is found from the Parachor method [14].
200 11 Binary Nucleation: Classical Theory
, ln ya,eq
reduced macroscopic surface 0.6
xb,eq
tension θ∞ (right y-axis) and
0
the vapor molar fraction of 0.4
nonane ya,eq (right y-axis).
Calculations are carried out xb,eq ln ya,eq
using the Redlich-Kwong- 0.2 -10
Soave equation of state (the
binary interaction parameter 0
0 20 40 60 80 100
ki j = 0.0448 [37])
pv (bar)
The equilibrium vapor molar fraction of nonane ya,eq shows nonmonotonous behav-
ior: at low pressures it decreases with p v since the increase of pressure results in the
growth of nonane fraction in the bulk liquid at the expense of its fraction in the vapor;
however, at higher pressures this process is partially blocked by penetration of the
supercritical methane into the liquid phase; ya,eq reaches minimum at p v ≈ 18 bar.
Qualitatively the presence and location of this minimum can be understood in terms
of the “compensation pressure effect” [30]. Consider the partial molecular volume
of component a in the vapor phase. By definition (D.1), vav is the change of the total
volume V v of the binary vapor when one extra a molecule is inserted into the vapor
at the fixed total pressure p v and the number of b-molecules. One can identify two
competing factors related to this process. The first factor is the tendency to increase
the volume in order to preserve p v . The opposite factor, manifested by the second
term in (11.123), is the tendency to reduce V v . The latter becomes pronounced for
mixtures of (partially) miscible components at sufficiently high pressures when the
separation between b molecules becomes of the order of the range of unlike (a − b)
attractions. As a result, a certain number of b molecules move in the direction of the
a molecule (usually relatively big compared to the b-molecule) thereby decreasing
V v . According to Eq. (D.7):
1 ∂ ln ρ v
vav = v 1 − yb (11.123)
ρ ∂ ya
ρ v = βp v (1 − b2 ), b2 ≡ β B2 p v (11.125)
is the second virial coefficient of the gas mixture; B2,aa (T ), B2,bb (T ) are the second
virial coefficients of the pure substances, and the cross term B2,ab is constructed
according to the combination rules [14]. Taking the logarithmic derivative in (11.125)
and linearizing in b2 we obtain from (11.124):
1 kB T
pcomp (T ) = (11.126)
2 [B2,bb (T ) − B2,ab (T )]
(after taking the derivative we can set yb ≈ 1 since methane is in abundance). This
result demonstrates the leading role in the compensation pressure effect played by a−
b interactions giving rise to B2,ab , which usually satisfies |B2,ab | > |B2,bb |. For the
nonane/methane mixture pcomp (T = 240 K) = 17.8 bar. This value approximately
coincides with the pressure, at which ya,eq is at minimum (see Fig. 11.8).
As usual we characterize the state of the system by the vapor-phase activities of the
components. Since methane is supercritical, the reference state should be different
from the pure component vapor–liquid coexistence at temperature T discussed in
Sect. 11.9.1. To this end we choose as a reference the ( p v , T )-equilibrium of the
v = μv ( p v , T ). Within the ideal gas approximation
mixture, so that in (11.105) μi,0 i,eq
Eq. (11.105) becomes
yi p v yi
Aiv ≈ = ≡ Si (11.127)
yi,eq ( p v , T ) p v yi,eq
log10J (cm-3s-1)
Closed circles: experiments of 40 33 25
expt 10
Luijten [33, 34]; open circles: 10
experiments of Peeters [35];
half-filled squares: exper- 5
iments of Labetski [36].
Dashed lines: BCNT. Labels: 0
total pressure in bar
-5
40
33
25
10
-10
0.5 1 1.5
log10Snonane
yi,eq p v
βΔG BCNT (n a , n b ) = − n i ln Si + β γ (xbtot ) A
yicoex (xbtot ) p coex (xbtot )
i
(11.128)
Figure 11.9 shows the BCNT predictions of nucleation rate as a function of
Sa = Snonane for temperature T = 240 K and pressures 10, 25, 33, 40 bar along
with the experimental results of [33–35] and [36]. Theoretical predictions are fairly
close to experiment for 10 and 25 bar. For higher pressures, however, BCNT largely
underestimates the experimental data. In particular, the 40 bar data show extremely
large deviation: more than 30 orders of magnitude. The analysis of the experimen-
tal data using nucleation theorem [34] reveals that the critical cluster is a small
object, containing 10–20 molecules. With this in mind it comes as no surprise that
application of a purely phenomenological BCNT approach to such clusters becomes
conceptually in error. Moreover, as opposed to the ethanol/hexanol mixture studied
in Sect. 11.9.2, surface enrichment in the nonane/methane mixture is expected to be
highly pronounced while the standard BCNT scheme does not take it into account.
References
Classical theory of binary nucleation can be drastically in error and even lead to
unphysical behavior when applied to strongly non-ideal systems with substantial
surface enrichment—a vivid example is the water/alcohol system, for which BCNT
predicts the decrease of nucleation rate with increasing partial pressures. An alter-
native to the classical treatment, based on purely phenomenological considerations,
is the density functional theory based on microscopical considerations. The basic
feature of DFT, discussed in Chap. 5, is the existence of the unique Helmholtz free
energy functional of the nonhomogeneous one-particle density ρ(r). The liquid-
vapor equilibrium corresponds to the minimum of this functional in the space of
admissible density profiles under the constraint of fixed particle number N. In a
nonequilibrium state (like supersaturated vapor) one has to search for the saddle
point of the free energy functional which corresponds to the critical nucleus in the
surrounding supersaturated vapor.
DFT of Chap. 5 can be extended to the case of nonhomogeneous binary mixtures.
Let us first discuss the two-phase equilibrium of the binary liquid and binary vapor.
The Helmholtz free energy functional F and the grand potential functional Ω for
a mixture is now written in terms of the one-body density profiles of components
1 and 2, ρ1 (r) and ρ2 (r), normalized as
ρi (ri ) dri = Ni , i = 1, 2
(1) uij (r) + εij for r < rm,ij
uij (r) = (12.1)
0 for r ≥ rm,ij
(2) −εij for r < rm,ij
uij (r) = (12.2)
uij (r) for r ≥ rm,ij
where εij is the depth of the potential uij (r) and rm,ij is the corresponding value of r:
uij (rm,ij ) = −εij . We assume that all interactions are pairwise additive.
The reference model is approximated by the hard-sphere mixture with appropriately
chosen effective diameters dij . Within the local density approximation the reference
part of the free energy is
Fd [ρ1 , ρ2 ] ≈ dr ψd (ρ1 (r), ρ2 (r)) (12.3)
where ψd (ρ1 (r), ρ2 (r)) is the free energy density of the uniform hard-sphere mixture
with the densities of components ρ1 and ρ2 . Using the standard thermodynamic
relationship (cf. (5.24)) it can be written as
2
ψd (r) = ρi μd,i (ρ1 (r), ρ2 (r)) − pd (ρ1 (r), ρ2 (r))
i=1
(2)
ρij (r, r ) ≈ ρi (r) ρj (r ) (12.4)
From (12.3) and (12.4) the Helmholtz free energy functional for the mixture takes
the form
2
1 (2)
F [ρ1 , ρ2 ] = dr ψd (ρ1 (r), ρ2 (r)) + dr dr ρi (r) ρj (r ) uij (|r − r |)
2
i,j=1
(12.5)
12.1 DFT Formalism for Binary Systems. General Considerations 207
2
The grand potential functional for the mixture Ω[ρ1 , ρ2 ] = F [ρ1 , ρ2 ] − i=1
μi Ni is
2
Ω[ρ1 , ρ2 ] = − dr pd (ρ1 (r), ρ2 (r)) + dr ρi μd,i (ρ1 (r), ρ2 (r))
i=1
2
1 (2)
+ dr ρi (r) dr ρj (r ) uij (|r − r |)
2
i,j=1
2
− μi dr ρi (r) (12.6)
i=1
δΩ
= 0, i = 1, 2 (12.7)
δρi (r)
or, equivalently
2
(2)
μd,i (ρ1 (r), ρ2 (r)) = μi − dr ρj (r )uij (|r − r |), i = 1, 2 (12.8)
j=1
The integral equations (12.8) are solved iteratively. First it is necessary to determine
the bulk equilibrium properties of the system: μi , ρiv ρil . Fixing two degrees of free-
dom (temperature and the total pressure, or temperature and bulk composition in one
of the phases), the two-phase equilibrium of the mixture is given by
2
F (ρ1 , ρ2 ) = Fd (ρ1 , ρ2 ) − V ρi ρj aij (12.12)
i,j=1
2
μi = μd,i (ρ1 , ρ2 ) − 2 ρj aij (12.13)
j=1
208 12 Binary Nucleation: Density Functional Theory
where
1 (2)
aij = − dr uij (r)
2
is the background interaction parameter. The virial equation for a mixture reads
2
p = pd (ρ1 , ρ2 ) − ρi ρj aij (12.14)
i,j=1
Consider the flat geometry with the inhomogeneity along the z-axis, directed towards
the bulk vapor. The bulk densities of the components in both phases provide asymp-
totic limits for the equilibrium density profiles in the inhomogeneous system:
The density profiles are calculated iteratively from Eq. (12.8) starting with an ini-
tial guess for each ρi (z), which can be a step-function or a continuous function that
varies between the bulk limits. When the equilibrium profiles are found, they can
be substituted back into the thermodynamic functionals which then become the cor-
responding thermodynamic potentials of the two-phase system. In particular, from
(12.6) and (12.8) the grand potential of the two-phase system in equilibrium reads:
2
1 (2)
Ω[ρ1 , ρ2 ] = − dr pd (ρ1 (r), ρ2 (r)) − dr ρi (r) dr ρj (r ) uij (|r − r |)
2
i,j=1
(12.15)
The plain layer surface tension of the binary system can be determined from the
general thermodynamic relationship (5.35):
where A is the interfacial area. For inhomogeneity along the z direction dr = A dz,
and Eqs. (12.15) and (12.16) yield
1
2
(2)
γ =− dz pd (z) + ρi (z) dr ρj (z ) uij (|r − r |) − p (12.17)
2
i=1
12.2 Non-ideal Mixtures and Surface Enrichment 209
As we know from Chap. 11, talking about a mixture we can not avoid the dis-
cussion of adsorption effects. On the phenomenological level it means that for a
binary (or, more generally, a multi-component) mixture it is impossible to choose
the Gibbs dividing surface in such a way that the excess (adsorption) terms for all
species simultaneously vanish. This feature gives rise to the surface enrichment:
a preferential adsorption of one of the species in the interfacial region between the two
bulk phases.
On the microscopic level the issue of adsorption boils down to the strength and range
of unlike interactions. They determine the degree of non-ideality of the system. The
DFT yields the density profiles of components in the inhomogeneous system. These
profiles have no rigid boundaries, their form is based on the microscopic interactions
in the system, implying that adsorption (surface enrichment) is naturally built into
the DFT scheme.
It is instructive to study the effects of non-ideality on the behavior of the mixture
considering the simplest system: a binary mixture of Lennard-Jones fluids with the
interaction potentials
σij 12 σij 6
uij (r) = 4εij − (12.18)
r r
where σii and εii are the Lennard-Jones parameters of the individual components.
For illustrative reasons (in order to have realistic numbers) we choose their values
corresponding to the argon/krypton mixture:
and
εAr /kB = ε11 /kB = 119.8 K, εKr /kB = ε22 /kB = 163.1 K
The unlike interactions u12 are defined via the mixing rules. We assume that u12 has
also the Lennard-Jones form. For conformal potentials it is common to present the
mixing rules in the form [2]:
1
σ12 = (σ11 + σ22 ) (12.19)
2
and √
ε12 = ξ12 ε11 ε22 (12.20)
where ξ12 is called the binary interaction parameter and is found from the fit to
experiment. When ξ12 = 1 one speaks about a Lorentz-Berthelot mixture. For real
mixtures ξ12 is usually significantly less than unity. It is necessary to have in mind
210 12 Binary Nucleation: Density Functional Theory
(a) 0.6
Kr T= 115.77 K (b) 0.6
T = 115.77 K
xAr=0.3 Kr
xAr=0.3
0.5 0.5
=1 = 0.88
12 12
3
0.4 0.4
3
Kr
= 13.7 mN/m = 10.1 mN/m
Kr
0.3 0.3
i (z)
i (z)
Ar Ar
0.2 0.2
0.1 0.1
0 0
0 5 10 15 20 0 5 10 15 20
z/ Kr z/ Kr
Fig. 12.1 Density profiles of argon and krypton at the flat vapor-liquid interface with the bulk liquid
molar fraction of argon xAr = 0.3 at T = 115.77 K and different values of the binary interaction
parameter ξ12 ; (a) ξ12 = 1 (Lorentz-Berthelot mixture); (b) ξ12 = 0.88. Distances and densities are
scaled with respect to σKr = σ22 . The decrease of ξ12 leads to the increase of the surface activity of
argon (surface enrichment), accompanied by the decrease of the surface tension of the mixture γ
that the laws of the ideal mixture are obtained only if all the potentials are the same;
in this sense even for ξ12 = 1 the mixture should not necessarily be ideal. Note
that for the equation of state of the mixture the mixing rule (12.20) leads to the
corresponding form of the “energy parameter” a12 :
√
a12 = ξ12 a11 a22
Consider implications of the mixing rule (12.20). The decrease of ξ12 from unity
will decrease the depth of unlike interactions thereby enhancing separation in the
solution. Since in our example ε11 < ε22 , this separation leads to the increase of the
surface activity of component 1 (argon) which is manifested by its pronounced surface
enrichment. These features are illustrated in Fig. 12.1. Equilibrium calculations are
performed for the argon/krypton mixture at T = 115.77 K and the bulk liquid molar
fraction of argon xAr = 0.3 (hence, we discuss the (x, T )-equilibrium).
The left graph (a) refers to the Lorentz-Berthelot mixture: ξ12 = 1. The surface
enrichment of argon in this case is very weak. The surface tension calculated from
Eq. (12.17) gives γ (ξ12 = 1) = 13.7 mN/m. On the right graph (b) we show the
DFT calculations for ξ12 = 0.88. The density profile of argon shows considerable
surface enrichment, meaning that the vapor-liquid interface is argon-rich leading to
the decrease of the surface tension: γ (ξ12 = 0.88) = 10.1 mN/m.
Formulation of the density functional theory for the study of equilibrium properties
of inhomogeneous binary mixtures can be extended to the nonequilibrium case of
binary nucleation. The corresponding development was carried out by Oxtoby and
12.3 Nucleation Barrier and Activity Plots: DFT Versus BCNT 211
coworkers [3–5] and represents a generalization of the similar approach for a single-
component case. In DFT of binary nucleation one studies the system “droplet in a
nonequilibrium vapor”, where both the droplet and the vapor are binary mixtures.
A droplet is associated with a density fluctuation which has no rigid boundary. The
Helmholtz free energy and grand potential functionals for this system are given as
before by Eqs. (12.5)–(12.6). However, the chemical potentials of the components in
this expressions refer now to the actual nonequilibrium state of the system (and not
to the thermodynamic equilibrium as in Sect. 12.1). This state can be characterized,
e.g. by fixing the gas-phase activities of the components.
The critical nucleus is in unstable equilibrium with the environment and corresponds
to the saddle point of Ω[ρ1 , ρ2 ] (as opposed to the minimum of Ω[ρ1 , ρ2 ] in the case
of equilibrium conditions). The density profiles in the critical nucleus are found as
before from the solution of Eq. (12.8). If the profiles are determined, then the change
in the grand potential
ΔΩ ∗ = Ω[ρ1 , ρ2 ] − Ωu (12.21)
where Ωu is the grand potential of the uniform nonequilibrium vapor (i.e. the binary
vapor prior to the appearance of the droplet) is the free energy associated with the
formation of the critical cluster. As in the single-component case it is straightforward
to see that ΔΩ ∗ is equal to the Gibbs energy of the critical cluster formation
ΔΩ ∗ = ΔG∗
where the pre-exponential factor K can be taken from the classical binary nucleation
theory (see Eq. (11.43)).
In the previous section we studied the effects of non-ideality of the mixture (expressed
in the terms of the unlike interaction potential) on the equilibrium properties. Let
212 12 Binary Nucleation: Density Functional Theory
us study their impact on the nucleation behavior. As before we consider the binary
mixture argon/krypton with the mixing rule (12.20). The role of non-ideality effects
can be clearly demonstrated by means of activity plots. Figures 12.2 and 12.3 show
the BCNT and DFT activity plots for T = 115.77 K corresponding to the nucleation
rates
Figure 12.2 refers to the Lorentz-Berthelot mixture: ξ12 = 1. As we know from the
equilibrium considerations of the previous section, surface enrichment of argon in
this case is very weak and the mixture is fairly ideal. It is therefore not surprising that
the difference in nucleation behavior between the BCNT and DFTs is of quantitative
nature; note that it increases with the krypton activity.
This situation is distinctly different from the case ξ12 = 0.88 shown in Fig. 12.3:
BCNT produces a “hump”—resembling the similar predictions for water/alcohol
systems (cf. Fig. 11.6). This hump, as discussed earlier, is unphysical since it violates
the nucleation theorem. Its occurrence is a consequence of the neglect within the
BCNT scheme of adsorption effects giving rise to surface enrichment. The DFT
approach does not have this drawback because adsorption is “built into it” on the
microscopic level. The DFT predictions therefore are in qualitative agreement with
the nucleation theorem. Figure 12.1b demonstrated the effect of surface enrichment
for the planar interface. DFT calculations reveal the same effect for the critical cluster
in binary nucleation [4].
References
1. G.A. Mansoori, N.F. Carnahan, K.E. Starling, T.W. Leland, J. Chem. Phys. 54, 1523 (1971)
2. J.S. Rowlinson, F.L. Swinton, Liquids and Liquid Mixtures (Butterworths, Boston, 1982)
3. X.C. Zeng, D.W. Oxtoby, J. Chem. Phys. 95, 5940 (1991)
4. A. Laaksonen, D.W. Oxtoby, J. Chem. Phys. 102, 5803 (1995)
5. I. Napari, A. Laaksonen, J. Chem. Phys. 111, 5485 (1999)
Chapter 13
Coarse-Grained Theory of Binary
Nucleation
13.1 Introduction
As we saw in Chap. 11, the classical theory proved to be successful for fairly ideal
mixtures. Meanwhile, for non-ideal mixtures BCNT can be sufficiently in error
[1–4] and even lead to unphysical results as in the case of water-alcohol systems.
The reasons for this failure are the neglect of adsorption effects and the inappropriate
treatment of small clusters. These issues are strongly coupled; they determine the
form of the Gibbs free energy of cluster formation and subsequently the composi-
tion of the critical cluster and the nucleation barrier. One can correct the classical
treatment by taken into account adsorption using the Gibbsian approximation (see
Sect. 11.6).
Taking into account adsorption within the phenomenological approach does not
resolve another deficiency associated with the capillarity approximation: the surface
energy of a cluster is described in terms of the planar surface tension. Obviously,
for small clusters the concept of macroscopic surface tension looses its meaning
and this assumption fails. This difficulty is not unique for the binary problem. In
the single-component mean-field kinetic nucleation theory (MKNT) of Chap. 7 this
problem was tackled by formulating an interpolative model between small clusters
treated using statistical mechanical considerations and big clusters described by the
capillarity approximation. In the present chapter we extend these considerations to
the binary case and incorporate them into a model which takes into account the
adsorption effects [5].
The statistical mechanical treatment of binary clusters, which we discuss in the
present chapter, originates from the analogy with the soft condensed matter theory,
where the description of complex fluids can be substantially simplified if one elim-
inates the degrees of freedom of small solvent molecules in the solution. By per-
forming such coarse-graining one is left with the pseudo-one-component system of
solute particles with some effective Hamiltonian. This approach opens the possibil-
ity to study the behavior of a complex fluid using the techniques developed in the
theory of simple fluids [6]. Situation in nucleation theory is somewhat similar: the
complexity of binary nucleation problem can be substantially reduced by tracing out
the degrees of freedom of the molecules of the more volatile component in favor
of the less volatile one. This pseudo-one-component system can be studied using
the approach developed in Chap. 7 making it possible to adequately treat clusters of
arbitrary size and composition.
Consider a binary mixture of component a and b in the gaseous state at the tempera-
ture T and the total pressure p v . The actual vapor mole fractions of components are
ya and yb . In the presence of a carrier gas with the mole fraction yc :
ya + yb = 1 − yc
In the present context the term “carrier gas” refers to a passive component, which
does not take part in cluster formation but serves to remove the latent heat. In some
cases, a passive carrier gas is absent (yc = 0), and one of the components of the
mixture (a or b) plays the double role: besides taking part in the nucleation process,
it removes the latent heat. In this case, this component should be in abundance in the
vapor phase.
Within the general formalism of Sect. 11.9.1 the state of component i is characterized
by the vapor phase activity
Aiv = exp β(μiv ( p v , T ; yi ) − μi,0
v
) , i = a, b (13.1)
value of μi at some reference state. Usually one chooses as the reference the saturated
v
state of pure components at the temperature T . This choice implicitly assumes the
existence of such a state for both species. In the case when one of the components
is supercritical this choice becomes inappropriate. With this in mind we choose as a
reference the true equilibrium state of the mixture at p v , T ; yc , so that μi,0
v = μv .
i,eq
Throughout this chapter the subscript “eq” refers to the true equilibrium state of the
mixture, and not a constrained equilibrium as in Chap. 11; to avoid confusion the
latter will be denoted by the superscript “cons”. Within the ideal gas approximation
yi p v yi
Aiv ≈ = ≡ Si (13.2)
yi,eq ( p , T, yc ) p
v v yi,eq
to emphasize that its value takes into account the presence of all components in the
mixture through yi,eq = yi,eq ( p v , T ; yc ).
Let us consider a two-dimensional n a − n b space of cluster sizes. Here n i is the total
number of molecules of component i, which according to Gibbs thermodynamics is
the sum of the bulk and excess terms
n i = n il + n iexc
∂ρ(n a , n b , t)
= Ja (n a − 1, n b , t) − Ja (n a , n b , t) + Jb (n a , n b − 1, t) − Jb (n a , n b , t)
∂t
(13.3)
where the fluxes along the n a and n b directions are
Impingement rates (per unit surface) of component i, vi , are given by gas kinetics:
yi p v
vi = √ (13.6)
2π m i kB T
Evaporation rates βi are obtained from the detailed balance condition. Recall that in
BCNT it is applied to the constrained equilibrium state which would exist for the
vapor at the same temperature, pressure and vapor phase activities as the vapor in
question. Instead of using this artificial state, we apply the detailed balance to the
true (full) equilibrium of the system at ( p v , T, yc ):
This procedure is a natural extension to binary mixtures of the Katz kinetic approach
discussed in Sect. 3.5. Assuming (following BCNT) that the evaporation rates do not
depend on the surrounding vapor, βi = βi,eq , we find from (13.7) and (13.8)
A(n a , n b ) ρeq (n a , n b )
βa = va,eq (13.9)
A(n a + 1, n b ) ρeq (n a + 1, n b )
A(n a , n b ) ρeq (n a , n b )
βb = vb,eq (13.10)
A(n a , n b + 1) ρeq (n a , n b + 1)
218 13 Coarse-Grained Theory of Binary Nucleation
where the second equality results from (13.2) and (13.6). The fluxes along n a and
n b read:
⎡ ⎤
Ja = −ρeq A va ⎣ Sin i ⎦ [H (n a + 1, n b ) − H (n a , n b )] (13.12)
i=a,b
⎡ ⎤
Jb = −ρeq A vb ⎣ Sin i ⎦ [H (n a , n b + 1) − H (n a , n b )] (13.13)
i=a,b
We can write down the same fluxes in terms of the constrained equilibrium quantities.
cons = v . Repeating the previous steps, in which the equilibrium
By definition vi,eq i
properties are replaced by the corresponding constrained equilibrium ones, we find
Ja = −ρeq
cons
A va [H (n a + 1, n b ) − H (n a , n b )] (13.14)
Jb = −ρeq
cons
A vb [H (n a , n b + 1) − H (n a , n b )] (13.15)
ρ(n a , n b )
H (n a , n b ) = (13.16)
ρeq
cons (n , n )
a b
(n a , n b ) = C e−βΔG eq
cons (n
a ,n b )
ρeq
cons
(13.19)
where
eq (n a , n b ) = −
βΔG cons n i ln Si + βΔG eq (n a , n b ) (13.20)
i=a,b
where ∇ ≡ ∂n∂ a , ∂n∂ b and the diagonal matrix F contains the rate of collisions of
a and b molecules with the surface of the cluster:
va A(n a , n b ) 0
F= (13.22)
0 vb A(n a , n b )
In these notations the kinetic equation (13.3) takes the form of the conservation law
for the “cluster fluid” (cf. Chap. 11)
∂ρ(n)
= − div J(n) (13.23)
∂t
with the steady state given by
div J = 0 (13.24)
β ΔG cons
eq (n a , n b ) = − n i ln Si + β ΔG eq (13.26)
i
At the saddle point the eigenvalues of the symmetric matrix D have different signs
implying that det D < 0. Following the standard procedure, we introduce a rotated
coordinate system (x, y) in the (n a , n b )-space with the origin at n∗ and the x axis
pointing along the direction of the flow at n∗ (see Fig. 11.2):
where ϕ is the yet unknown angle between the x and n a . At the saddle point Jb /Ja =
tan ϕ, which from (13.12)–(13.13) is written as:
∂H
vb ∂n b
∂H
= tan ϕ (13.28)
va ∂n a
Using the standard relationships between the derivatives in the original and the rotated
systems, we present after simple algebra Eq. (13.28) as:
∂H va tan2 ϕ + vb ∂ H
= (13.29)
∂x (va − vb ) tan ϕ ∂ y
Right at the saddle point Jx = Jx∗ (n∗ ), Jy (n∗ ) = 0. As in the BCNT, we use the
direction of principal growth approximation assuming that
13.2 Katz Kinetic Approach: Extension to Binary Mixtures 221
∂ Jx (x, y)
=0
∂x
Ja = Jx cos ϕ, Jb = Jx sin ϕ
where vav is the average impingement rate given by Eq. (11.35). Rewriting (13.31) as
1 ∂H
Jx (y) nb =−
ρeq San a Sb A vav ∂x
lim H = 1, lim H = 0
x→−∞ x→+∞
we find ∞
Jx (y) 1
dx =1 (13.32)
vav −∞ ρeq San a Sbn b A
Substituting the expansion (13.27) into Eq. (13.32) and performing Gaussian inte-
gration first over x and then over y, we find for the total nucleation rate
n∗
∗
J = vav A Z Si i ρeq (n a∗ , n ∗b ) (13.33)
i
where A∗ = A(n a∗ , n ∗b ),
1 (∂ 2 βΔG/∂ x 2 )n∗
Z =− √ (13.34)
2 − det D
is the Zeldovich factor. Equation (13.33) contains the yet undetermined direction ϕ of
the flow in the saddle point. The latter is found by maximizing the angle-dependent
part of J .
222 13 Coarse-Grained Theory of Binary Nucleation
The quantities with the ϕ dependence are: vav and Z . It is convenient to present
vav as
1 + t2 vb
vav = vb , t ≡ tan ϕ, r = (13.35)
r + t2 va
In the Zeldovich factor det D is invariant to rotation, implying that the only angle-
dependent part of Z is contained in
∂ 2 βΔG da − 2t + db t 2
= Daa cos2 ϕ +2Dab sin ϕ cos ϕ + Dbb sin2 ϕ = −Dab
∂x2 1 + t2
(13.36)
where we denoted
Daa Dbb
da = − , db = −
Dab Dab
Combining (13.35) and (13.36), the ϕ-dependent part of J is given by the function
da − 2t + db t 2
f (t) =
r + t2
In line with the kinetic approach we discuss the full thermodynamic equilibrium of
the system at the total pressure p v , temperature T and carrier gas composition yc
(if present). The partition function of an arbitrary (n a , n b )-cluster is:
1
Z na nb ≡ Z n = qn a n b (13.38)
Λa3n a Λ3n
b
b
13.3 Binary Cluster Statistics 223
is the potential interaction energy of the cluster comprised of a −a, b−b and (unlike)
a − b interactions. The prefactor n1i ! takes into account the indistinguishability of
molecules of type i inside the cluster. The symbol cl indicates that integration is
only over those molecular configurations that belong to the cluster. The cluster as
a whole can move through the entire volume V, while the molecules inside it are
restricted to the configurations about cluster’s center of mass that are consistent with
a chosen cluster definition.
We represent the equilibrium gaseous state of the a − b mixture as a system of
noninteracting (n a , n b ) = n clusters. Since the clusters do not interact, the partition
function Z (n) of the gas of Nn of such n-clusters is factorized:
1
Z (n) = Z Nn
Nn ! n
μn = n a μa,eq
v
+ n b μvb,eq (13.42)
z i,eq = , i = a, b (13.44)
Λi3
is the fugacity of component i in the equilibrium vapor. Thus, we reduced the prob-
lem of finding ρeq (n a , n b ) to the determination of the cluster configuration integral.
Even though we discuss the clusters at vapor-liquid equilibrium, it is important to
realize that the chemical potential of a molecule inside an arbitrary binary cluster
depends on the cluster composition and therefore is not the same as in the bulk vapor
surrounding it. Equation (13.43) shows that the quantity qn/V plays the key role in
determination of the cluster distribution function. From the definition of qn it is clear
that qn/V involves only the degrees of freedom relative to the center of mass of the
cluster. Note also, that qn contains the normalization constant C of the distribution
function.
one can easily see that the expression in the curl brackets
1
qb/a ({Ran a }) ≡ drn b e−β(Ubb +Uab ) (13.46)
nb ! cl
where the positions of b-particles are integrated out. By doing so we replaced the
binary cluster by the equivalent single-component one with the effective Hamiltonian
which is the sum of the Hamiltonian of the pure a-system, Uaa , and the free energy
of b-molecules in the instantaneous environment of a molecules. Equation (13.48)
is formally exact.
In order to derive a tractable representation of the free energy Fb/a we perform
the diagrammatic expansion of ln qb/a in the Mayer functions of a − b and b − b
interactions:
The zeroth order contribution F0 (n a , n b , T ), called the volume term, does not depend
on positions of molecules, but is important for thermodynamics since it depends on
cluster composition and therefore by no means can be neglected. The first-order term
U1 in (13.50) vanishes in view of translational symmetry [6]. Combining (13.49)–
(13.50) we write
qn = e−β F0 qnCG
a
(13.52)
where
1
dRn a e−βU
CG
qnCG (xbtot , T ) = (13.53)
a
na ! cl
pseudo—a molecules with the interaction energy U CG . The latter implicitly depends
on the fraction of b-molecules in the original binary cluster. The configuration
integral of this single-component cluster is qnCG
a
. Equation (13.52) is a key result
of the model.
Speaking about a binary cluster, we characterized it by the total numbers of molecules
n a and n b , not discriminating between the bulk and excess numbers of molecules of
each component
n i = n il + n iexc
n aexc (n al , n lb ), n exc
b (n a , n b )
l l
Thus, the point (n al , n lb ) in the space of bulk numbers yields the point (n a , n b ) in
the space of total numbers. As a result the dependence of various quantities on the
total composition xbtot can be also viewed as a dependence (though a different one)
on the bulk composition xbl .
Since the carrier gas is assumed to be passive, the cluster composition satisfies the
normalization:
xal + xbl = xatot + xbtot = 1 (13.54)
Let us discuss in more detail the volume term in the effective Hamiltonian H CG . It
can be written as a sum of the free energy of ideal gas of pure b-molecules in the
cluster, Fb,id , and the excess (over ideal) contribution, ΔF0 , due to b − b and a − b
interactions
F0 = Fb,id + ΔF0 (13.55)
where Vcl = n al val + n lb vbl is the volume of the cluster. Within the K -surface for-
malism we can equivalently write it in terms of total numbers:
13.4 Configuration Integral of a Cluster: A Coarse-Grained Description 227
The specific feature of Eq. (13.56) is that b-molecules are contained in the cluster
volume which itself depends on their number n b . Substituting (13.57) into (13.56),
we write
βFb,id = n b f b,id (13.58)
where ⎧ ⎡ ⎤⎫
⎪
⎨ ⎪
⎬
xbtot ⎢ Λ3b ⎥
f b,id = ln ⎣ ⎦ ≡ f b,id (xbl , T )
⎪
⎩ xatot x tot
val + xbtot vbl e ⎪⎭
a
na nb
βΔF0 = f 1 (xbl , T ) (13.59)
Vcl
where f 1 is some unknown function of xbl and T . Using (13.57), we present Eq.
(13.59) as
βΔF0 = n b f 0 (13.60)
where
f1
f0 = ≡ f 0 (xbl , T )
xbtot
val + xatot
vbl
by Eq. (7.38):
qnCG
= C Φan a e−θmicro n a
a s
(13.62)
V
Here
1
Φa = (13.63)
z a,sat
Using the basic result of the coarse-graining procedure Eq. (13.52), we express the
equilibrium distribution function (13.43) of binary clusters as
−β F0
qnCG
ρeq (n a , n b ) = e a
[z a,eq ]n a [z b,eq ]n b
V
where
g surf (n a ; xbl , T ) = θmicro (xbl ) n as (n a ; xbl ) (13.65)
The right-hand side of (13.64) contains the unknown intensive quantities Φa , Φb and
θmicro which depend on the bulk composition of the cluster and the temperature. To
determine them we consider appropriate limiting cases, for which the behavior of
ρeq (n a , n b ) can be deduced from thermodynamic considerations.
The ( p v , T )-equilibrium corresponds to the bulk liquid composition xb,eq ( p v , T ).
An arbitrary (n a , n b )-cluster at ( p v , T )-equilibrium has the bulk composition xbl
1Note in this respect, that n as (n a ) is always positive, while the Gibbs excess numbers n iexc can be
both positive and negative.
13.5 Equilibrium Distribution of Binary Clusters 229
different from xb,eq . Let us now fix xbl and consider the two-phase equilibrium at the
pressure p coex (xbl , T ), representing the total pressure above the bulk binary solution
with the composition xbl . Obviously, p coex (xbl , T ) = p v (the equality occurs only for
xbl = xb,eq ). At this “xbl -equilibrium” state the fugacities are: z i,coex = eβμi,coex /Λi3 ,
where μi,coex (xbl ) is the chemical potential at xbl -equilibrium. Thus, in the distribution
function for this state z i,eq in (13.64) should be replaced by z i,coex .
Now, from the entire cluster size space let us consider the clusters falling on the
xbl -equilibrium line, i.e. those whose bulk numbers of molecules satisfy:
n lb = n al (xbl /xal )
For them the chemical potential of the molecule inside the cluster is equal to its value
in the surrounding vapor at the pressure p coex . The Gibbs formation energy of such
a cluster will contain only the (positive) surface term:
1
Φi (xbl ) = (13.67)
z i,coex (xbl )
leading to
n a n b
ya,eq pv yb,eq pv
e−g
surf (n
a ;x b ,T )
l
ρeq (n a , n b ) pv ,T =C
yacoex (xbl ) p coex (xbl ) ybcoex (xbl ) p coex (xbl )
(13.68)
Here B2 (T ) is the second virial coefficient, psat (T ) is the saturation pressure. The
number of surface molecules n as depends parametrically on the coordination number
in the liquid phase N1 and the reduced plain layer surface tension θ∞ (T ).
Comparing (13.67) and (13.63), we can identify the “saturation state” of the pseudo-a
fluid: the latter is characterized by the chemical potential μa,coex (xbl ) and the pressure
Since the intermolecular potential u(r ; xbl , T ) of the pseudo—a fluid is not known,
we have to introduce an approximation for the second virial coefficient, which will
now depend on xbl . The simplest form satisfying the pure components limit, is given
by the mixing rule:
where B2,ii (T ) is the second virial coefficient of the pure component i; the cross virial
term B2,ab (T ) can be estimated using the standard methods [15]. Having identified
psat and B2 for the pseudo-a system, we find the reduced microscopic surface tension
θmicro ≡ θmicro,a and the normalization factor C from Eqs. (13.69)–(13.70):
B2 yacoex p coex
θmicro,a (xbl ) = − ln − , (13.73)
kB T
coex coex
ya p
C(xbl ) = eθmicro,a (13.74)
kB T
The cluster distribution (13.68) with g surf given by (13.65) will be fully determined
if we complete it by the model for n as (n a ; xbl ). Calculation of this quantity requires
the knowledge of the reduced planar surface tension of the pseudo-a fluid, θ∞,a and
the coordination number in the liquid phase, N1,a . The latter can be estimated from
(7.68) in which the packing fraction of pseudo-a molecules is approximated as
π l l 3
η= ρ (xb ) σa (13.75)
6
Here ρ l (xbl ) is the binary liquid number density at xbl -equilibrium and σa is the
molecular diameter of component a.
To determine θ∞,a let us consider the distribution function ρeq (n a , n b ) for big
(n a , n b )-clusters. These clusters satisfy the capillarity approximation in which the
surface part of the Gibbs energy takes the form
g surf (n a , n b ; xbl ) = β γ (xbl ) A = β γ (xbl ) (36π )1/3 (n al val + n lb vbl )2/3 (13.76)
where γ (xbl ) is the planar surface tension of the binary system at xbl -equilibrium.
13.5 Equilibrium Distribution of Binary Clusters 231
where
∂ ln ρ l
ηil = 1 + x j
∂ x j pcoex ,T, j =i
We require that the binary (n a , n b )-cluster has the same surface energy as the unary
n a -cluster of pseudo-a molecules for all sufficiently large n a . This implies that
2/3
xbtot l
θ∞,a = θ∞ (xbl ) ηa + tot ηb
l
(13.82)
xa
The construction of the model thus ensures that for sufficiently large clusters, sat-
isfying the capillarity approximation, the distribution function recovers the clas-
sical Reiss expression [16] (see also [17]) and is symmetric in both components.
This will not be true for small clusters, for which the capillarity approximation
fails and the Gibbs energy of cluster formation differs from its phenomenological
counterpart.
232 13 Coarse-Grained Theory of Binary Nucleation
Having determined all parameters, entering Eq. (13.64), we are in a position to write
down the equilibrium distribution function:
( )
ρeq (n a , n b ) pv ,T = β yacoex (xbl ) p coex (xbl ) e−geq (n a ,n b ;xb )
l
(13.83)
* +, -
C(xbl )
where
We have included the exponential factor with the microscopic surface tension in
(13.70) into the surface part of the free energy thereby redefining the prefactor C of
the distribution function:
The binary distribution function (13.83) recovers the single-component MKNT limit
when n b → 0.
An important feature of CGNT is that it eliminates ambiguity in the normalization
constant C in the nucleation rate inherent to BCNT and its modifications. This is
the direct consequence of replacing the constrained equilibrium concept by the full
thermodynamic equilibrium. Since all rapidly (exponentially) changing terms in the
distribution function are included into the free energy, we can safely set the value of
C in (13.90) to the one corresponding to the critical cluster:
∗ ∗ ∗
C(xbl ) = β yacoex (xbl ) p coex (xbl ) (13.88)
Our choice to trace out the b-molecules in the cluster in favor of a-molecules could
have been reversed: we could trace out a-molecules to be left with the effective
Hamiltonian for the b-molecules resulting in Eq. (13.52) with the single-component
cluster containing pseudo-b particles. This equation is formally exact. However,
calculation of the coarse-grained configuration integral in (13.52) invokes approxi-
mations inherent to MKNT. Its domain of validity is given by Eq. (7.53) which for
the system of pseudo-i particles (i = a or b) reads
B2 yicoex p coex
1 (13.89)
kB T
13.5 Equilibrium Distribution of Binary Clusters 233
It is clear that to obtain accurate predictions within the present approach one has
to trace out the more volatile component, i.e. the one with the largest bulk vapor
fraction. Throughout this chapter we assume this to be component b: ybcoex > yacoex ;
thus, the coarse-grained cluster contains pseudo-a particles.
Combining Eqs. (13.33) and (13.83)–(13.86), we obtain the total steady-state nucle-
ation rate for the binary mixture at the total pressure p v , temperature T and vapor
mole fractions yi :
∗ ∗ ∗ ∗
J = vav A∗ Z C(xbl ) e−g(n a ,n b ) (13.90)
* +, -
K
where star refers to the critical cluster being the saddle-point of the free energy
surface
∗ ∗
g(n a , n b ; xbl ) = − n i ln Si + geq (n a , n b ; xbl ) (13.91)
i
in the space of total numbers n i . Note, that search for the saddle point in the space of
bulk numbers can lead to an erroneous critical cluster composition and unphysical
results. Technical details of the saddle-point calculations are given in Appendix G.
The proposed model is termed the Coarse-Grained Nucleation Theory (CGNT).
It is instructive to summarize the steps leading to Eq. (13.90).
• Step 0. Determine the metastability parameters of components using as a reference
the ( p v , T )-equilibrium state of the mixture
yi
Si = , i = a, b
yi,eq ( p v , T ; yc )
Find the planar surface tension γ∞ (xbl , T ) for the x(lb , T )-equilibrium, using e.g.
the Parachor method, or some other available (semi-)empirical correlation.
234 13 Coarse-Grained Theory of Binary Nucleation
• Step 3. Calculate the properties of the pseudo-a fluid: the second virial coefficient
from Eq. (13.72) and the reduced microscopic surface tension from Eq. (13.73)
• Step 4. Determine the excess numbers of molecules n aexc , n exc
b from the K -surface
equations (11.84)–(11.85). The total numbers of molecules in the cluster are
n il = n il + n iexc , i = a, b
• Step 5. Calculate the free energy of cluster formation g(n a , n b ; xbl ) from Eq.
(13.91).
• Step 6. Repeat Steps. 1–5 for a “reasonably chosen” domain of bulk compositions,
accumulating the data:
n al , n lb , n a , n b , g(n a , n b ; xbl )
∗ ∗
K = vav A∗ Z C(xbl )
we end up with the virtual monomer approximation for Z proposed by Kulmala and
Viisanen [18]: .
∗ ∗ ∗
γ∞ (xbl ) xal val + xbl vbl
Z = (13.92)
kB T 2π (r ∗ )2
40
5
40
r
b
ba
r
33
ba
-5
25
10
-10
0.5 1 1.5
log 10 Snonane
236 13 Coarse-Grained Theory of Binary Nucleation
na , n b
indicates the compensation 30 nla
pressure
tot
na
20
tot
nb
10
l
nb
0
0 10 20 30 40 50 60
v
p (bar)
pcomp
Eq. (11.126) is
pcomp (T = 240 K) ≈ 17.8 bar
It is remarkable that the change in the nucleation behavior predicted by CGNT occurs
exactly at pcomp .
As the pressure is increased the total number of nonane molecules grows very slowly
while the total number of methane molecules increases rapidly, accumulating pre-
dominantly at the dividing surface; at the highest pressure of 60 bar shown in Fig. 13.2
there are only about 4 methane molecules in the interior of the cluster, while their
total number is ≈ 22 being close to the total number of nonane molecules n a ≈ 26.
Hence, at high pressures the critical cluster is a nano-sized object with a core-shell
structure: its interior is rich in nonane while methane is predominantly adsorbed on
the dividing surface.
References
11. M. Dijkstra, R. van Rooij, R. Evans, Phys. Rev. E 59, 5744 (1999)
12. R. van Rooij, J.-P. Hansen, Phys. Rev. Lett. 79, 3082 (1997)
13. R. van Rooij, M. Dijkstra, J.-P. Hansen, Phys. Rev. E 59, 2010 (1999)
14. V.I. Kalikmanov, Phys. Rev. E 68, 010101 (2003)
15. R.C. Reid, J.M. Prausnitz, B.E. Poling, The Properties of Gases and Liquids (McGraw-Hill,
New York, 1987)
16. H. Reiss, J. Chem. Phys. 18, 840 (1950)
17. G. Wilemski, B. Wyslouzil, J. Chem. Phys. 103, 1127 (1995)
18. M. Kulmala, Y. Viisanen, J. Aerosol Sci. 22, S97 (1991)
19. C.C.M. Luijten, Ph.D. Thesis, Eindhoven University, 1999
20. P. Peeters, Ph.D. Thesis, Eindhoven University, 2002
21. D.G. Labetski, Ph.D. Thesis, Eindhoven University, 2007
Chapter 14
Multi-Component Nucleation
n i = n il + n iexc , i = 1, . . . , N (14.1)
As discussed in Chap. 11, n il and n iexc separately depend on the location of the
dividing surface while their sum is independent of this location to the relative accu-
racy of O(ρ v /ρ l ), where ρ v and ρ l are the number densities in the vapor and liq-
uid phases. Straightforward generalization of Eq. (11.48) to N -component mixture
yields:
N N
ΔG = ( p − p )V + γ A +
v l l
n il μi ( p ) − μi ( p ) +
l l v v
n iexc μiexc − μiv ( p v )
i=1 i=1
(14.2)
Here p l is the pressure inside the cluster,
N
Vl = n il vil (14.3)
i=1
N 2/3
A = (36π ) 1/3
n il vil (14.4)
i=1
are, respectively, the cluster volume and surface area calculated at the location of
the dividing surface; vil is the partial molecular volume of component i in the liq-
uid phase, γ is the surface tension at the dividing surface. Within the capillarity
approximation
μil ( p l ) = μil ( p v ) + vil ( p l − p v ) (14.5)
N
N
ΔG = γ A − n il Δμi + n iexc μiexc − μiv ( p v ) (14.6)
i=1 i=1
where
Δμi ≡ μiv ( p v ) − μil ( p v ) (14.7)
is the driving force for nucleation. As in the binary case, the chemical potentials in
both phases are taken at the vapor pressure p v .
Recalling that diffusion between the surface and the interior of the cluster is much
faster than diffusion between the surface and the mother vapor phase, we assume that
there is always equilibrium between the cluster (dividing) surface and the interior,
resulting in
μiexc = μil ( p l ) (14.8)
14.1 Energetics of N -Component Cluster Formation 241
Substituting (14.9) into (14.6) and using Laplace equation, we obtain a generalization
of Eq. (11.55):
N
2γ (xl )
N
ΔG = γ (xl ) A − n il + n iexc Δμi + n iexc vil (14.10)
r
i=1 i=1
ni
where r is the radius of the cluster (assumed to be spherical), γ (xl ) is the surface
tension of the N -component liquid solution with composition
xl = (x1l , . . . , x Nl ),
nl
N
xil = N i , xil = 1
l
k=1 n k i=1
Note, that the second term in (14.10) contains the total numbers of molecules, while
the thermodynamic properties depend on the bulk cluster composition xl .
If we choose the dividing surface according to the K -surface recipe (cf.
Sect. 11.6):
N
n iexc vil = 0 (14.11)
i=1
N
ΔG = − n i Δμi + γ (xl ) A (14.12)
i=1
The volume and surface area of the cluster can now be written in terms of total
numbers of molecules
N
Vl = n i vil (14.13)
i=1
N 2/3
A = (36π ) 1/3
n i vil (14.14)
i=1
where y coexj (xl ) is the fraction of component j in the N -component vapor which
coexists with the N -component liquid having the composition xl of the cluster;
the corresponding coexistence pressure is p coex (xl ). The superscript “coex” empha-
sizes that the corresponding quantity refers to (x, T )-equilibrium, rather than the
( p v , T )-equilibrium. The quantities p coex (xl ) and ycoex = (y1coex (xl ), y2coex (xl ), . . .)
are found from the (x, T )-equilibrium equations:
p l (T, ρ l , xl ) = p coex
p v (T, ρ v , ycoex ) = p coex (14.15)
μil (T, ρ l , xl ) = μiv (T, ρ v , ycoex ), i = 1, . . . , N
N
yi p v
ΔG(n 1 , . . . , n N ) = − kB T n i ln + γ (xl ) A (14.16)
yicoex (xl ) p coex (xl )
i=1
where ni
xitot = N
k=1 n k
N
wi = U ji n j (14.18)
j=1
where the new set of coordinates is denoted as wi , and U is a unitary matrix with real
coefficients [5]. The latter by definition satisfies
UT = U−1 (14.19)
where UT is the transposed matrix and U−1 is the inverse matrix. From (14.19) it
follows that multiplication by U has no effect on inner products of the vectors, angles
or lengths. In particular, lengths of the vectors ||n|| are preserved
(U n)T (U n) = nT (UT U) n = nT n
The latter property actually provides rotation. Since UT U = I , where I is the unit
matrix, we have
det (UT U) = det UT det U = 1
resulting in
det U = 1
Now we are in the position to impose a certain form of the rotation matrix. Since the
critical cluster is associated with the saddle point of the free energy, we define U so
244 14 Multi-Component Nucleation
where δi j is the Kronecker delta; Q 1 < 0 (maximum of ΔG along w1 ), while the rest
eigenvalues are all positive: Q 2 , . . . , Q N > 0 (minimum of ΔG along w2 , . . . , w N );
the values of Q i are yet to be defined. Then, the matrix U should satisfy
∂ 2 ΔG ∂ 2 ΔG
0= Uui Uv j = , i = j (14.23)
u,v
∂n u ∂n v c ∂wi ∂w j c
implying U −1
ji = Ui j . Equation (14.23) determines the rotation angle of the original
coordinate system.
Combining Eqs. (14.22) and (14.23) we find
−1 ∂ 2 ΔG
Uiu Uv j = Q i δi j (14.25)
u,v
∂n u ∂n v c
We multiply both sides of (14.25) by Uki and sum over k. Since U is the unitary
matrix, we obtain using Eq. (14.24):
∂ 2 ΔG
− Q j δkv Uv j = 0 (14.26)
v
∂n k ∂n v c
This expression shows that Q j are the eigenvalues and Uv j are the eigenvectors of
the Hessian matrix 2
∂ ΔG
G2 =
∂n k ∂n v c
det (G2 − Q I ) = 0
We choose the rotational system such that the critical cluster corresponds to
w1 = w1c , w2c = . . . = w N c = 0
14.1 Energetics of N -Component Cluster Formation 245
Expansion of the Gibbs energy in the vicinity of the saddle point of ΔG reads:
1
N
1
ΔG = ΔG c + Q 1 (w1 − w1c )2 + Q i wi2 + . . . (14.27)
2 2
i=2
where ΔG c is the value of ΔG at the saddle point. Substituting (14.27) into (14.17),
we obtain the equilibrium cluster distribution in the vicinity of the saddle point:
⎧ ⎡ ⎤⎫
⎨ 1
N ⎬
ρeq (w1 , . . . , w N ) = ρeq,c exp − β ⎣ Q 1 (w1 − w1c )2 + Q i wi2 + . . .⎦ (14.28)
⎩ 2 ⎭
i=2
where
ρeq,c = C e−βΔG c (14.29)
14.2 Kinetics
k
E(n 1 , . . . , n j , . . . , n N ) + E j k j E(n 1 , . . . , n j + 1, . . . , n N ) (14.30)
j
ρ j k j = v j A(n 1 , . . . , n j , . . . , n N ) (14.31)
The form of v j depends on the physical nature of the nucleation process; for the
gas-liquid transition it is given by the gas kinetics expression
pj
vj = (14.32)
2π m j kB T
where p j is the partial pressure of component j in the mother phase, and m j is the
mass of the molecule of component j. Equation (14.30) describes a single act of the
cluster evolution. The net rate at which the clusters (n 1 , . . . , n j + 1, . . . , n N ) are
created in the unit volume of the system is
246 14 Multi-Component Nucleation
I j (n 1 , . . . , n j , . . . , n N ) = ρ j k j ρ(n 1 , . . . , n j , . . . , n N )
− k j ρ(n 1 , . . . , n j + 1, . . . , n N ) (14.33)
In equilibrium all net rates vanish (here, as in the BCNT, we consider the constrained
equilibrium) which yields the determination of the backward reaction rate
ρ j k j ρeq (n 1 , . . . , n j , . . . , n N )
k j (n 1 , . . . , n N ) = (14.34)
ρeq (n 1 , . . . , n j + 1, . . . , n N )
The evolution of the cluster distribution function is given by the kinetic equation
which includes all possible reactions with single molecules of species 1, . . . , N
∂ρ(n 1 , . . . , n j , . . . , n N )
N
= I j (n 1 , . . . , n j − 1, . . . , n N )
∂t
j=1
−I j (n 1 , . . . , n j , . . . , n N ) (14.35)
I j = ρ j k j ρeq (n 1 , . . . , n j , . . . , n N )
× [ f (n 1 , . . . , n j , . . . , n N ) − f (n 1 , . . . , n j + 1, . . . , n N )]
∂ f (n 1 , . . . , n j , . . . , n N )
I j = − ρ j k j ρeq (n 1 , . . . , n j , . . . , n N ) (14.37)
∂n j
N
∂ρ(n 1 , . . . , n j , . . . , n N ) ∂Ij
=− = −div I (14.38)
∂t ∂n j
j=1
where the vector I is defined as I = (I1 , . . . , I N ). The same equation in the rotated
system reads
⎡ ⎤
∂ ⎣ −1 ⎦
N N N N
∂ρ ∂Ij
=− Ui−1 =− Ui j I j ≡ −div J (14.39)
∂t j
∂wi ∂wi
i=1 j=1 i=1 j=1
14.2 Kinetics 247
where J = (J1 , . . . , J N ),
N
Ji = Ui−1
j I j , i = 1, . . . , N (14.40)
j=1
is the component of the nucleation flux along the wi -axis. Using (14.31) and (14.37)
Ji can be presented as
N
∂f
Ji = −A ρeq Biu , i = 1, . . . , N (14.41)
∂wu
u=1
in which
N
Biu = Ui−1
j v j U ju (14.42)
j=1
We search for the steady state solution of Eq. (14.39) ignoring the short time-lag
stage. Then, Eq. (14.39) becomes
div J = 0
Recall that we have chosen the rotated system in such a way that the flux J is directed
along the w1 axis, implying
J2 = . . . = J N = 0 (14.43)
∂f
By separating the term with ∂w1 from the rest of the sum, we get a linear set of N − 1
∂f
equations for N − 1 unknown variables ∂wu , i = 2, . . . , N . Its determinant is
B22 . . . B2N
· ·
D2 = · ·
· ·
BN 2 . . . BN N
where
B21 B22 . . . B2,u−1 B2,u+1 . . . B2N
· · · · · · ·
Lu = · · · · · · ·
· · · · · · ·
BN 1 BN 2 . . . B N ,u−1 B N ,u+1 . . . BN N
∂f ∂f
Having expressed N − 1 quantities ∂w u
in terms of ∂w 1
, we substitute the solution
(14.44) into the only remained equation from the set (14.41), namely the equation
for J1 :
N
∂f ∂f
J1 = −A ρeq B11 + B1u
∂w1 ∂wu
u=2
∂f 1
N
= −A ρeq B11 D2 + (−1) u+1
B1u L u
∂w1 D2
u=2
As can be easily seen, the expression in the square brackets is the determinant of the
N × N matrix of all Bi j ’s:
B11 . . . B1N
· ·
D1 = · · (14.45)
· ·
BN 1 . . . BN N
Now we apply the (standard) boundary conditions for the cluster distribution function
(cf. (3.42)–(3.43)): the concentration of small clusters is nearly equal to equilibrium
14.2 Kinetics 249
one; for large clusters ρeq diverges, while the actual distribution function ρ remains
finite. These requirements yield:
The first condition results in f 1 = 1, then from the second one we obtain
∞ −1
D1 A 1
J1 (w2 , . . . , w N ) = dw1
D2 w1 =w1c 1 ρeq (w1 , w2 , . . . , w N )
(14.49)
We can simplify this result by using the expansion (14.28) and extending the inte-
gration limits in (14.49) to ±∞:
! "
D1 A 1 N
J1 (w2 , . . . , w N ) = ρeq,c exp − β Q i wi2
D2 w1 =w1c 2
i=2
Q 1 (w1 −w1c ) 2 −1
∞
× dw1 e 2kB T
−∞
Taking into account that Q 1 < 0 and performing Gaussian integration we obtain
! " #
D1 A 1
N
(−Q 1 )
J1 (w2 , . . . , w N ) = ρeq,c exp − β Q i wi2
D2
w1 =w1c 22π kB T
i=2
(14.50)
The total nucleation rate is found by integrating J1 over all possible values of
w2 , . . . , w N . In view of the previously presented considerations we set the limits
of integration to ±∞:
∞ ∞
J= ... dw2 . . . dwN J1 (w2 , . . . , wN ) (14.51)
−∞ −∞
Assuming further that D1 A/D2 varies slowly with wi ’s compared to the exponential
terms exp[−Q i wi2 /kB T ], we perform N − 1 Gaussian integrations
∞ Q i wi2 1/2
− 2π kB T
dwi e 2kB T =
−∞ Qi
resulting in
#
D1 A (−Q 1 )
J = C e−βΔG c (2π kB T )(N −2)/2 (14.52)
D2 c Q2 . . . Q N
Let us consider application of the general formalism to the binary nucleation. Rotation
matrix U in this case is given by Eq. (14.20). The rotation angle φ is found from Eq.
(14.23) which reads
$ %
∂ 2 ΔG
2 ∂n 1 ∂n 2 c
tan(2φ) = (14.53)
∂ 2 ΔG ∂ 2 ΔG
∂n 21
− ∂n 22
c c
Equation (14.53) has two solutions for the angle φ. We choose the solution which
gives Q 1 < 0 and Q 2 > 0 as required by construction of the model. Having deter-
mined the rotation angle, we write down the coefficients Bi j from Eq. (14.42):
where
v1 v2
vav = (14.56)
v1 sin φ + v2 cos2 φ
2
c
14.3 Example: Binary Nucleation 251
is the average impingement rate. This result agrees with Eqs. (11.35), (11.42)–(11.44)
derived in Chap. 11.
Using similar assumptions as in the BCNT, the present model focuses on nucle-
ation in the vicinity of the saddle point of the free energy surface and completely
ignores nucleation along all paths other than the principal nucleation path (direction
of principal growth). This local approach identifies the predominant composition
of the observable nuclei—the one that corresponds to the saddle point. Although
this approach obviously has its merits, it, however, is unable to predict the rate of
formation of nuclei with arbitrary composition. The latter issue requires a global
approach which treats all cluster compositions on an equal footing. This problem
was addressed among others by Wu [6] who studied various nucleation paths (not
only the saddle-point nucleation). In [6] conditions were identified under which a
multi-component system behaves “as if it were simple” which means that the system
can be modelled by a one-dimensional Fokker-Planck equation (3.79)–(3.80).
Direction of principal growth approximation may not always be the best choice.
Depending on the form of the Gibbs free energy surface and values of the impinge-
ment rates of the components it is possible that the main nucleation flux bypasses
the saddle-point. In particular, as pointed out by Trinkaus [4], if one reaction rate is
essentially smaller than the other ones, the flux line can turn into the directions of the
fast-reacting component and pass a ridge before the saddle-point coordinate of the
slowly reacting component is reached. For binary systems these issues were studied
numerically by Wyslouzil and Wilemski [7].
References
1. M.J. Muitjens, V.I. Kalikmanov, M.E.H. van Dongen, A. Hirschberg, P. Derks, Revue de l’Institut
Français du Pétrole 49, 63 (1994)
2. V. Kalikmanov, J. Bruining, M. Betting, D. Smeulders, in 2007 SPE Annual Technical Conference
and Exhibition (Anaheim, California, USA, 2007), pp. 11–14. Paper No: SPE 110736
3. O. Hirschfelder, J. Chem. Phys. 61, 2690 (1974)
4. H. Trinkaus, Phys. Rev. B 27, 7372 (1983)
5. K.F. Riley, M.P. Hobson, S.J. Bence, Mathematical Methods for Physics and Engineering
(Cambridge University Press, Cambridge, 2007)
6. D. Wu, J. Chem. Phys. 99, 1990 (1993)
7. B. Wyslouzil, G. Wilemski, J. Chem. Phys. 103, 1137 (1995)
Chapter 15
Heterogeneous Nucleation
15.1 Introduction
where K is a kinetic prefactor and ΔG ∗ is the free energy of formation of the critical
embryo on the foreign particle (CN). As in the case of homogeneous nucleation, the
heterogeneous nucleation rate is determined largely by the energy barrier ΔG ∗ . That
is why it is sufficient to know the prefactor K (its various forms are discussed in
Sect. 15.5) to one or two orders of magnitude. We will show in the next section that
the presence of a foreign particle reduces the energy cost to build the cluster surface.
We consider a cluster (embryo) of phase 2 which has the form of a spherical liquid
cap of volume V2 and radius r and contains n molecules,
n = ρ l V2 (15.2)
where ρ l is the number density of the phase 2. The embryo is resting on the spherical
foreign particle (CN) 3 of a radius R p surrounded by the parent phase 1. The line
where all three phases meet, called a three-phase contact line, is characterized by
the contact angle θ . This configuration is schematically shown in Fig. 15.1. Within
the phenomenological approach embryos are considered to be the objects character-
ized by macroscopic properties of phase 2. Applying Gibbs thermodynamics to this
system, it is necessary to define two dividing surfaces: one for the gas-liquid (1–2)
interface and one for the solid-liquid (2–3) interface. Since we discuss the single-
component nucleation, we can choose for both of them the corresponding equimolar
surface so that the adsorption terms in the Gibbs energy vanish. Within the capillarity
approximation the Gibbs free energy of an embryo formation is
Rp r
Op O
d
2
3
1
Fig. 15.1 Embryo 2 (dashed area) on the foreign particle (condensation nucleus) 3 in the parent
phase 1. R p is the radius of the spherical CN, O p is its center; r is the radius of the sphere with the
center in O corresponding to the embryo, d = O p O, θ is the contact (wetting) angle
15.2 Energetics of Embryo Formation 255
Rp − rm r − R pm
cos φ = , cos α = − (15.8)
d d
Here
m = cos θ (15.9)
d = R 2p + r 2 − 2R p r m (15.10)
Geometrical factor f V relates the embryo volume (shaded area in Fig. 15.1) to its
homogeneous counterpart, being the full sphere of radius r . Standard algebra yields:
Rp
f V = q(cos α) − a 3 q(cos φ), a ≡ (15.11)
r
where the function q(y) is:
1
q(y) = (2 − 3 y + y 3 ) (15.12)
4
The angles φ and α depend not on r and R p individually, but on their ratio a; therefore
it is convenient to rewrite (15.8), (15.10) in dimensionless units:
a−m 1−ma
cos φ = , cos α = − (15.13)
w w
w = 1 + a2 − 2 m a (15.14)
Looking at (15.3)–(15.14), one can see that our model for ΔG is incomplete: the
surface areas A12 , A23 and the volume of the embryo depend on the yet undefined
256 15 Heterogeneous Nucleation
contact angle θ . Its value can be derived from the interfacial force balance at the
three phase boundary known as the Dupre-Young equation:
(to emphasize the bulk equilibrium nature of the Dupre-Young equation we added
the subscript “eq” to the contact angle). An alternative way to derive this result is to
consider a variational problem: formation of an embryo with a given volume V2 with
an arbitrary contact angle. The equilibrium angle θeq will be the one that minimizes
ΔG at fixed V2 :
∂ΔG
=0 (15.16)
∂m V2
Let us compare the free energies necessary to form a cluster with the same number
of molecules n at the temperature T and the supersaturation S (or equivalently, Δμ)
for the homogeneous and heterogeneous nucleation. Clearly, for both cases the bulk
contribution to ΔG will be the same: −n Δμ. Consider now the surface term. In the
homogeneous case:
ΔG surf (n)hom = γ12 s1 n 2/3
where
−2/3
s1 = (36π )1/3 ρ l (15.18)
ΔG surf (n)het
≤1 (15.19)
ΔG surf (n)hom
15.2 Energetics of Embryo Formation 257
The case of flat geometry deserves special attention. This is a limiting case when the
radius of an embryo is much smaller than the radius R p of the foreign particle, or
equivalently a→∞. Then, the embryo sees the particle as a flat wall while the particle
radius becomes irrelevant. An embryo becomes a spherical segment (cup) resting on
a plane as shown in Fig. 15.2. Taking the limit R p →∞ in Eqs. (15.5)–(15.8), we find
α = θeq
4
V2 = πr 3 q(m) (15.20)
3
A12 = 2πr 2 (1 − m) (15.21)
A23 = πr (1 − m )
2 2
(15.22)
1
q(m) = (2 − 3m + m 3 ) (15.23)
4
0.6
q(m)
0.4
0.2
0
-1 -0.5 0 0.5 1
m=cos
and the 2 − 3 interface becomes a circle bounded by the three-phase contact line of
the radius rt = r sin θeq . From (15.2) and (15.20) we find the relation between the
number of molecules in the embryo and its radius:
1/3
3 −1/3
r= ρl q −1/3 n 1/3 (15.24)
4π
Using the general form (15.17) of the heterogenous free energy barrier together with
Eqs. (15.20)–(15.22) for the case of flat geometry we find:
For homogeneous nucleation of a cluster with the same number of molecules at the
same temperature T and supersaturation S one would have:
Comparing (15.25) and (15.26), one can see that the heterogeneous barrier has the
form of the homogeneous barrier in which the surface tension γ12 is replaced by the
effective surface free energy γeff defined by
Figure 15.3 shows the behavior of the functions q(m) and q 1/3 (m) for different values
of the contact angle. They increase monotonously from 0 at θ = 0 to 1 at θ = π .
15.3 Flat Geometry 259
Thus, γeff ≤ γ12 and therefore the heterogeneous barrier is always smaller or equal
to ΔG hom for the same values of T and S. Expressing ΔG in terms of the radius of
the embryo, we have using (15.24):
Note that Eq. (15.28) is solely based on macroscopic considerations. For nonwetting
conditions (m = −1) we recover the homogeneous limit: ΔG het = ΔG hom .
Let us return to the case of arbitrary geometry. The critical embryo rc satisfies
∂ΔG
=0 (15.29)
∂r rc
Finding the critical radius using the general form (15.17) of the Gibbs formation
energy in the presence of a foreign particle requires a considerable amount of algebra.
Fortunately, calculation can be substantially simplified if we take into account that
irrespective of the presence or absence of the foreign particles, the critical cluster is
in metastable equilibrium with the surrounding vapor (phase 1) and thus the chemical
potentials of a molecule inside and outside the critical cluster are equal resulting in
the Kelvin equation (3.61):
2γ12
rc = (15.30)
ρ l Δμ
rc = rc,hom (T, S)
Rp
n c = n c,hom (T, S) f V (m, ac ), ac = (15.31)
rc
assuming the same bulk liquid density of phase 2 in the homogeneous and heteroge-
neous cases. Substituting (15.30) into (15.17) and using (15.5)–(15.14), we derive
the nucleation barrier:
m=0.5
0.1
fG
m=0.8
0.01
m=1
0.001
0.1 1 10 100
a = Rp/r
where
1 16π γ123
ΔG ∗hom = γ12 (4π rc2 ) = (15.33)
3 3 (ρ l )2 (Δμ)2
is called the Fletcher factor [6]. It varies between 0 and 1, depending on the contact
angle and the relative size of foreign particles with respect to the embryo. Thus,
heterogeneous nucleation reduces the nucleation barrier compared to the homoge-
neous case ΔG ∗hom due to the presence of foreign bodies; the Fletcher factor being
the measure of this reduction. It is important to bear in mind that the Fletcher fac-
tor f G refers exclusively to the critical embryo so that the expression (15.32) is
not true for an arbitrary embryo. The behavior of f G (m, a) as a function of a for
various values of the contact angle (parameter m) is shown in Fig. 15.4. The upper
line m = −1 corresponds to the non-wetting conditions recovering the homoge-
neous limit: f G = 1, ΔG ∗ = ΔG ∗hom for all R p . For the flat geometry the function
f G,∞ = lima→∞ f G (m, a) levels up yielding
Recall that in homogeneous nucleation the kinetic prefactor takes the form (3.54)
J0 = Z (v A∗ ) ρ1 (15.36)
where v A∗ is the rate of addition of molecules to the critical cluster of the radius r ∗ ,
Z is the Zeldovich factor (3.50):
γ12 1
Z = (15.37)
kB T 2πρ l (r ∗ )2
and ρ1 is the number of monomers (per unit volume) of the mother phase. The
latter quantity coincides with the number of nucleation sites, since homogeneous
nucleation can occur with equal probability at any part of the physical volume.
Kinetic prefactor in the heterogeneous case has the same form
K = Z (v A∗ ) ρ1,s (15.38)
with the number of nucleation sites ρ1,s being the number of molecules in contact
with the substrate; clearly, ρ1,s is sufficiently reduced compared to ρ1 .
The value of the prefactor K depends on the particular mechanism of the cluster
formation. Two main scenarios are discussed in the literature. The first one assumes
that nucleation occurs by direct deposition of vapor monomers on the surface of the
cluster [6]. Another possibility is surface diffusion [8–10]: vapor monomers collide
with the CN surface and become adhered to it; the adsorbed molecules further migrate
to the cluster by two-dimensional diffusion. Fundamental aspects of surface diffusion
are discussed in the seminal monograph of Frenkel [11]. It was found theoretically
and experimentally [12, 13] that the surface diffusion mechanism is more effective
and leads to higher nucleation rates.
The dimensionality of K coincides with the dimensionality of the nucleation rate. The
number of critical embryos in the process of heterogeneous nucleation depends on
the amount of pre-existing foreign particles acting as CN. That is why the nucleation
rate can be expressed as:
• number of embryos per unit area of the foreign particle per unit time; or
• number of embryos per foreign particle per unit time; or
• number of embryos formed per unit volume of the system per unit time
Let us discuss the “adsorption—surface diffusion” mechanism of an embryo for-
mation. Adsorption can be visualized as a process in which molecules of phase 1
strike the surface of a foreign particle, remain on that surface for a certain adsorption
time τ and then are re-evaporated. The motion of adsorbed molecules on the surface
262 15 Heterogeneous Nucleation
of a particle can not be a free one but reminds the 2D random walk which can be
associated with the 2D diffusion process [8, 11].
Let Nads be the surface concentration of adsorbed molecules, i.e. the number of
molecules of phase 1 adsorbed on the unit surface of CN. It is difficult to determine
a realistic value of this quantity for a general case. Fortunately, Nads appears in the
pre-exponential factor K and errors in evaluating it will not influence the nucleation
rate as critically as errors in determination of ΔG ∗ (in particular, the uncertainty in
the contact angle is much more significant than the uncertainty in Nads ). In view of
these considerations we can use for Nads an “educated estimate”:
Nads = v τ (15.39)
where v is the impingement rate of the monomers of the phase 1 per unit surface of
CN; if phase 1 is the supersaturated vapor, v is given by the gas kinetics expression
(3.38). The time of adsorption can be written in the Arrhenius form [14]
where E ads is the heat of adsorption (per molecule) and τ0 is the characteristic
time—the period of oscillations of a molecule on the surface of CN. The latter
can be estimated as the inverse of the characteristic absorption frequency f I of the
substance. For most of the substances the absorption frequencies lie in the ultraviolet
region [15]: f I = 3.3 × 1015 s−1 implying that
1
τ0 ≈ ≈ 3 × 10−16 s
fI
Obviously, E ads depends on the nature of the adsorbing surface and the adsorbed
molecules. One can find it from the data on diffusion coefficient D written in the
form of an Arrhenius plot:
E ads
ln D = const −
kB T
where we associate E ads with the activation energy for diffusion. Its value is given
by the slope of ln D as a function of the inverse temperature. Typical molar values
of E ads in liquids are found to be ≈10 − 30 kJ/mol.
If the heterogeneous nucleation rate is expressed as a number of embryos formed
per unit area of a foreign particle per unit time [16], the prefactor takes the form:
If nucleation rate is expressed as the number of embryos per foreign particle per unit
time, Eq. (15.41) will be modified to
(assuming that a foreign particle is a sphere of the radius R p ). The nucleation rate
ΔG ∗
J p = K p exp − [s−1 ] (15.43)
kB T
The presence of two or more bulk phases in contact with each other gives rise to
the discontinuity of their thermodynamic properties and results in the corresponding
interfacial tensions. Formation of an embryo on the surface of a foreign particle
results in the occurrence of a line of three-phase contact. This line has an associated
with it tension τt , which is the excess free energy of the system due to three-phase
contact, per unit length of the contact line. A thermodynamic definition of the line
tension τt can be given in a way analogous to the definition of the surface tension in
Sect. 2.2 by introducing three dividing surfaces—gas-liquid, solid-liquid and solid-
vapor—and using the methodology of Gibbs thermodynamics of nonhomogeneous
systems. The excess Ωt of the free energy (grand potential) associated with the
three-phase contact line reads [17]:
Ωt = τt L (15.45)
where L is the length of the contact line. This relation defines τt and is the analogue of
Eq. (2.25) for the surface tension. The line tension does not depend on the location of
dividing surfaces like each of the 2D interfacial tensions γ12 , γ23 , γ13 [17]. However,
unlike the 2D interfacial tensions, τt can be of either sign. A two-phase interfacial
tension should be necessarily positive: if this would not be the case the increase of
the interfacial area would become energetically favorable leading to the situation
when two phases become mutually dispersed in each other on molecular scale, so
264 15 Heterogeneous Nucleation
that the interface between the phases disappears. Thus, the separation between any
two phases requires a positive interfacial tension. The situation with the three-phase
line is different. A negative line tension makes an increase of the length L of the triple
line energetically favorable (Ωt < 0), however this increase inevitably changes the
surface areas of the 2D interfaces—characterized by positive tensions—in a such a
way that the total free energy of the system increases.1
The classical (Fletcher) theory, described in the previous sections, does not take into
account the line tension effect. At the same time, several authors [19, 20] indicated
that for highly curved surfaces (i.e. small critical embryos) it can have a substantial
influence on the nucleation behavior. The presence of the line tension modifies the
Gibbs free energy of an embryo formation (15.3)–(15.4):
where rt is the radius of the contact line. Inclusion of the line tension changes
the force balance at the contact line implying that the contact angle correspond-
ing to the new situation, which we denote as θt and call the intrinsic (or micro-
scopic) contact angle, will be different from its bulk value θeq given by the
Dupre-Young equation.
For simplicity consider the flat geometry of Fig. 15.2. To derive the force balance at
the contact line of radius rt let us consider a small arc of this line seen from its center
under a small angle α as shown in Fig. 15.5. The length of this arc is l = rt α. The
line energy of the arc is E t = τt l. Consider a change of the contact line radius δrt .
The corresponding change of the arc length δl = α δrt induces the change of the line
energy
δ E t = τt δl.
Then, the line force acting on the arc of length l in the radial direction (with the unit
vector −
→
er pointing outwards from the center C of the contact line) is
−
→ δ Et −
→ δl −
→
Ft = − lim er = −τt lim er = −τt α −
→
er
δrt →0 δrt δrt →0 δrt
1Note, that a possibility of a negative tension of the three-phase contact line was already mentioned
by Gibbs [18].
15.6 Line Tension Effect 265
This expression defines the contact angle in the presence of line tension. An alter-
native derivation of this result stems from the observation that θt minimizes the
Gibbs free energy (15.46) at a fixed embryo volume V2 . Comparing (15.48) with the
Dupre-Young equation (15.15) we find
τt 1
cos θt = cos θeq − (15.49)
γ12 r sin θt
This result is known as the modified Dupre-Young equation. A positive τt would mean
that the intrinsic angle of a small embryo is larger than the bulk value; a negative τt
leads to smaller contact angles compared to θeq . On a molecular level the line tension
stems from the intermolecular interactions between the three phases in the vicinity
of the contact line. This fact imposes the bulk correlation length ξ as a natural length
scale of the line tension effect. Equation (15.49) suggests that the relevant energy
scale (per unit area) is the gas-liquid surface tension γ12 . The order-of-magnitude
estimate of τt is then
τt ∼ γ12 ξ
In this form the line tension can be interpreted as the first order correction for the
bulk contact angle in the inverse radius of the cap. Equation (15.50) can be viewed
as an alternative definition of τt (θeq ). If one can measure (or simulate) the intrinsic
angle as a function of the cluster radius r and the bulk angle θeq , than τt can be found
as a slope of cos θt versus 1/(r sin θeq ).
Consider implication of the line tension on the free energy of an embryo formation.
Combining Eqs. (15.46) and (15.48), we write
A23 τt A23
ΔG = −n Δμ + γ12 A12 1 − mt − + τt 2πrt
A12 rt
where m t = cos θt . Using (15.22) for the surface area A23 , this expression reduces to
A23
ΔG = −n Δμ + γ12 A12 1 − mt + τt π r sin θt (15.51)
A12
where we took into account that for flat geometry rt = r sin θt . The expression in the
square brackets represents the Fletcher model in which the bulk angle θeq is replaced
by the intrinsic one θt :
An important feature of this expression is that taking into account the line tension
has a two-fold effect on the energy barrier:
• an additional term in ΔG appears which is proportional to the length of the contact
line, and
• in the Fletcher factor q the bulk angle is replaced by the intrinsic contact angle
which depends on τt and the radius of the embryo through the modified Dupre-
Young equation
Denoting m eq = cos θeq and performing the first order perturbation analysis in (1/r),
we write
q(m t ) = q(m eq ) + Δq
dq 3
Δq = Δm = − sin2 θeq Δm (15.53)
dm m eq 4
τt 1
Δm ≡ m t − m eq ≈ − (15.54)
γ12 r sin θeq
d
ΔG hom (r ) q(m eq ) + Δq + ΔG hom (r ) Δq + τt π sin θeq = 0, where =
dr
(15.56)
ΔG hom (r )r =rc = 0
4π
ΔG hom (rc ) = γ12 rc2
3
into Eq. (15.56) and taking into account (15.53) and (15.54), we find that Eq. (15.56)
becomes an identity. This means that the critical cluster size in the presence of line
tension is given by the classical Kelvin equation.2 Setting r = rc in (15.55), we
obtain the nucleation barrier in the presence of line tension:
4π
ΔG ∗ = γ12 rc2 q(m eq ) + 2 πrc τt sin θeq (15.57)
3
A positive τt increases the nucleation barrier given by the Fletcher theory (first term),
while a negative τt lowers it thus enhancing nucleation. Such an enhancement was
observed experimentally in Refs. [24, 25].
The modified Dupre-Young equation (15.49) has an analytical solution θt (θeq , r ) for
all values of the bulk contact angle θeq . However, the general form of the solution
2 Note, that if in (15.52) the bulk Fletcher factor q(m eq ) is used instead of q(m t ), the critical cluster
size, maximizing ΔG, would violate the Kelvin equation.
268 15 Heterogeneous Nucleation
is rather complex except for the special case of θeq = π/2, for which Eq. (15.49) is
simplified to:
2p τt
sin(2θt ) = − , p≡ (15.58)
r γ12
This constraint gives the range of validity of the modified Dupre-Young equation.
From the previous analysis (15.59) can be approximately expressed as
When the radius of the embryo approaches the molecular size Eq. (15.58) fails.
At the same time at large r the classical expression should be recovered: limr →∞ θt =
θeq , which in our case results in
θt →r →∞ π/2 (15.61)
At r →∞: θt,1 →0, θt,2 →π/2. Hence, the solution satisfying the asymptotic con-
dition (15.61) is
π 1 2p
θt = − arcsin − (15.62)
2 2 r
1 2p
m t = sin arcsin − (15.63)
2 r
τt 1
m t (r ) ≈ −
γ12 r
15.6 Line Tension Effect 269
coincides with the exact one up to the terms of order 1/r 2 . One can expect that for
angles close to π/2 the first order expansion in 1/r remains a good approximation.
To accomplish the model for the Gibbs formation energy (15.55) we need to have
information about the line tension for a given system. The extreme smallness of τt
makes its direct experimental measurement a very difficult task, that is why available
experimental data remains scarce. To obtain reliable estimates of τt it is important
that the droplets are of the same size as the deduced line tension values. Several
advanced techniques were used recently aiming to satisfy this requirement. Pompe
and Herminghaus [26] and Pompe [27] used Scanning Force Microscopy to study
the shape of the sessile droplets on solid substrates near the contact line. From the
droplet profile they deduced that τt lies in the range 10−12 −10−10 N and can be either
positive or negative depending on the system. In particular, it was found that line
tension increases with lowering contact angle; at large θeq it is negative and changes
sign at θeq ≈ 6◦ . Berg et al. [28] used Atomic Force Microscopy (AFM) to study
nanometer-size sessile fullerene (C60 ) droplets on the planar Si O2 interface and
observed the size-dependent variation of the contact angle which can be interpreted
as the line tension. From the modified Dupre-Young equation they found the negative
values
and obtained a characteristic length scale of the effect τt /γ12 ≈ 1.4 nm.
In most of the heterogeneous nucleation studies, which take into account the line
tension effect, the value of τt is found from fitting to the experimental data on nucle-
ation rates [24, 29, 30]. However, such fitting can not be considered reliable in view
of a number of reasons. Homogeneous nucleation in the bulk has to be distinguished
from heterogeneous nucleation on impurities, and small changes of parameters (e.g.
substrate heterogeneities) can lead to considerable difference in measured nucleation
rates. These and other artifacts can then be erroneously interpreted as line tension
effects. In view of these reasons it is highly desirable to determine τt from an inde-
pendent source: model/simulations/experiment.
The advantage of computer simulations is that the properties of interest are con-
trollable parameters. The simplest model of a fluid is the lattice-gas with nearest
neighbor interactions (Ising model) on the simple cubic lattice. The Ising Hamil-
tonian (discussed in Sect. 8.9) reads:
H = −K si s j (15.66)
nn
270 15 Heterogeneous Nucleation
where the “spins” sk are equal to ±1, and K is the coupling parameter (interaction
strength); summation is over nearest neighbors. The presence of the foreign substrate
(a solid wall) is described by the corresponding boundary condition which is charac-
terized by a surface field H1 acting on the first layer of fluid molecules adjacent to the
surface. Monte Carlo simulations of this system performed by Winter et al. [31] at
temperatures far from Tc result in the appearance of a spherical cap-shaped (liquid)
droplet surrounded by the vapor and resting on the adsorbing solid wall (favoring
liquid). The surface free energy of the embryo formation ΔG surf MC is found in simula-
tions using thermodynamic integration. Simulation results reveal that the difference
between ΔG surf
MC and the Fletcher model
MC − γ12 4π r q(θeq )
ΔG surf 2
increases linearly with the droplet radius r . This difference can be attributed to the
line tension contribution. In Ref. [31] this linear dependence is presented in the form
which is used as a definition of the line tension τMC . We introduced the notation τMC
to emphasize the difference between the latter and the previously defined quantity τt .
We require that the simulated surface free energy ΔG surfMC be equal to its theoretical
counterpart given by (15.55)
ΔG surf
th = γ 12 4π r 2
q(θeq ) + γ 12 4π r 2
Δq + τt π r sin θeq (15.68)
kB TMC /K = 3, (15.70)
reveal that for all contact angles studied τMC is negative and its absolute value
increases with θeq as shown in Fig. 15.6. The equilibrium contact angle is controlled
by varying the surface field H1 . It is important to note, that for a fixed tempera-
ture different contact angles (obtained by tuning the surface field H1 ) in Fig. 15.6
physically correspond to different solids.
Equation (15.66) being the simplest model of a magnet, can be also viewed as a model
of a fluid. Mapping of the Ising model to a model of a fluid is not a unique procedure,
therefore the results for τt for fluids can differ depending on the chosen procedure
but most probably will be qualitatively the same. One of the strategies, used e.g. in
the theory of polymers, is to equate the critical temperature of a substance to the
15.6 Line Tension Effect 271
a0/(kBTMC )
lattice spacing. The solid line -0.15
is a fit to the Monte Carlo
results of Ref. [31]
-0.2
MC
-0.25
-0.3
20 30 40 50 60 70 80 90
eq (grad)
kB Tc,Ising/K ≈ 4.51
Considering water as an example and setting Tc,Ising equal to the critical temperature
of water Tc,water = 647 K, we find that MC simulations of Ref. [31] correspond to
T = 430.3 K. Considering a substance which at this temperature has the bulk contact
angle θeq = 90◦ , the simulation results of Fig. 15.6, give
τMC σ
= −0.26
kB TMC
where we replaced the lattice spacing a0 in the Ising model by the water molecular
diameter σ = 2.64 Å [33]. Then, taking into account (15.69)
One of the important issues that has to be considered is the temperature dependence
of the line tension. Experimentally the latter can be deduced from the measurements
of the microscopic contact angle using the modified Dupre-Young equation: τt is
found from the slope of cos θt as a function of 1/r at different temperatures. Such
study was performed by Wang et al. [34] for n-octane and 1-octene in the temperature
interval 301 < T < 316 K. Fig. 15.7 shows the plot of the line tension for n-octane
(solid circles) and 1-octene (open circles) as a function of reduced temperature
t = (Tw − T )/Tw
272 15 Heterogeneous Nucleation
Fig. 15.7 Line tension as a function of reduced temperature t = (Tw − T )/Tw for n-octane and
1-octene on coated silicon. (Reprinted with permission from Ref. [34], copyright (2001), American
Physical Society.)
where Tw is the wetting temperature (corresponding to cos θeq = 1). The solid
substrate in both cases is Si wafer.
For both liquids as temperature increases towards the wetting temperature Tw , the
line tension changes from a negative to a positive value with an increasing slope
|dτt /dT |. This behavior qualitatively agrees with theoretical predictions [35, 36].
The wetting temperatures of n-octane and 1-octene on Si wafer were found to be
For illustration purposes let us analyze the implication of the line tension for water
nucleating on a large seed particle with the bulk contact angle θeq = 90◦ at T = 285 K
and supersaturation S = 2.93. At these conditions the critical cluster radius according
to CNT is rc = 1.03 nm. If the radius of a seed particle R p 1 nm, it can be
considered as a flat wall for the critical embryo and the Fletcher factor is given by
the function q(m). Figure 15.8 shows ΔG(r ) for 3 different models:
15.6 Line Tension Effect 273
100
homogeneous Water
T=285 K
80
S=2.93
60
G/kBT
Fletcher
cont. angle
40 90
Fletcher+line tension
20 t = -1.1*10-11 N
Fig. 15.8 Gibbs free energy of a cluster formation for water nucleating on a large seed particle
at T = 285 K and supersaturation S = 2.93. Solid line: classical homogeneous nucleation theory
(CNT). Dashed line: the classical heterogeneous (Fletcher) theory with the bulk the contact angle
θeq = 90◦ . Dashed-dotted line: the Fletcher theory corrected with the line tension effect according
to Eq. (15.55); the value of line tension is τt = −1.1 × 10−11 N
τt = −1.1 × 10−11 N
The Fletcher correction reduces the barrier to ΔG ∗Fletcher ≈ 43 kB T . The line tension
leads to further reduction ΔG ∗ ≈ 25 kB T resulting in considerable enhancement of
nucleation. Setting R p = 10 nm, τ0 = 2.55 × 10−13 s [7], E ads = 10.640 kcal/mol
[37] we find for the heterogenous nucleation prefactor (15.42)
Np
= 2 × 104 cm−3
V
Np
KV = K p ≈ 1018 cm−3 s−1
V
which is 7.5 orders of magnitude lower than the corresponding homogeneous
quantity J0 .
The Fletcher theory gives
while the incorporation of the line tension into the model yields a considerable
increase of the nucleation rate
Activation of a foreign particle occurs when the first critical embryo is formed on its
surface. This is a random event and as such can be studied using the methodology
of the theory of random processes. This approach to heterogeneous nucleation can
be particularly useful when analyzing the experimental data.
Let us choose some characteristic time t, during which heterogeneous nucleation is
observed. Typical value of t in experiments is ∼1 ms. Let Pk (t) be the probability
that exactly k activation events occurred during time t. The average number of such
events per unit time is given by the nucleation rate J p . A probability of activation
during an infinitesimally small interval Δt is J p Δt (assuming that two simultaneous
activation events during Δt are highly unlikely). Then the probability that no events
happened during the same interval is 1 − J p Δt. Consider the quantity Pk (t + Δt)
which is the probability that exactly k activation events occurred during time t + Δt.
Straightforward probabilistic considerations yield:
The first term on the right-hand side refers to the situation when all k events happened
during time t and no events occurred during time Δt. The second term gives the
probability that exactly k − 1 events took place during time t and one event happened
during time Δt. Dividing both sides by Δt and taking the limit at Δt→0 we obtain
15.7 Nucleation Probability 275
d Pk (t)
= −Pk (t) J p + Pk−1 (t) J p , k = 0, 1, 2, . . . (15.72)
dt
Consider the first equation of this set, corresponding to k = 0; P0 (t) describes the
probability that no events happened during time t. Obviously, we must set P−1 (t) = 0
which yields
d P0 (t)
= −J p P0 (t) (15.73)
dt
Integration of Eq. (15.73) gives
where P0 (0) is the probability that no events happened in zero time. Obviously,
P0 (0) = 1 resulting in
P0 (t) = e−J p t
is the probability that at least one foreign particle was activated to growth dur-
ing time t. For large number of events Phet (t), termed the nucleation probability
[37, 39, 40], describes the fraction of foreign particles activated to growth during
time t. The latter quantity is measured in heterogeneous nucleation experiments. Set-
ting Phet to 0.5 we refer to the situation when half of the foreign particles are activated
to growth. This can be viewed as the onset conditions for heterogeneous nucleation.
Since the nucleation rate is a very steep function of the supersaturation (activity), the
nucleation probability is expected to be close to the step-function centered around
the onset activity.
For illustration we use the example of water nucleation considered in Sect. 15.6.5.
Setting the characteristic experimental time to t = 1 ms [37] we find:
This result implies that S = 2.93 is the onset condition if the line tension effect is
taken into account; at the same S the Fletcher theory predicts no nucleation. From
experimental data (see e.g. [37]) it follows that the onset conditions are not much
sensitive to the choice of t.
276 15 Heterogeneous Nucleation
References
Fig. 16.1 Cutaway view of diffusion cloud chamber. (Reprinted with permission from Ref. [4],
copyright (1975), American Institute of Physics.)
The thermal diffusion cloud chamber consists of two metallic cylindrical plates sep-
arated by the optically transparent cylindrical ring. The region between the plates
forms the working volume of the chamber. The substance under study is placed as
a shallow liquid pool on the lower plate of the chamber and the working volume is
filled by the carrier gas (aiming at removal of the latent heat emerging in the process
of condensation). The lower plate is heated while the upper plate is cooled. Due to
the temperature difference ΔT between the plates, vapor1 evaporates from the top
surface of the liquid pool, diffuses through a noncondensable carrier gas (usually
helium, argon or nitrogen), and condenses on the lower surface of the top plate.
Construction of the diffusion cloud chamber is illustrated in Fig. 16.1.
Thermal diffusion gives rise to the profiles of temperature, density, pressure and
supersaturation inside the chamber. These profiles can be calculated from the one-
dimensional energy and mass transport equations using an appropriate equation of
state for the vapor/carrier gas mixture as shown in Fig. 16.2. At certain values of ΔT
the supersaturation in the chamber becomes sufficiently large to cause nucleation
of droplets which are subsequently detected by light scattering using a laser and a
photo-multiplier.
1 The term “vapor” in this chapter is used for the condensible component; while the term “gas”
refers to the carrier gas.
16.1 Thermal Diffusion Cloud Chamber 279
Fig. 16.2 Profiles of density, temperature, supersaturation and the nucleation rate inside the cham-
ber. (Reprinted with permission from Ref. [5], copyright (1989), American Institute of Physics.)
Diffusion cloud chamber can operate in the temperature range from near the triple
point (of the substance under study) up to the critical temperature and in the pressure
range from below the ambient to elevated pressures. A typical range of accessi-
ble nucleation rates is 10−3 − 103 cm−3 s−1 . The growing droplets are removed
by gravitational sedimentation or by convective flow which ensures the steady-state
self-cleaning operational conditions. Due to this feature and to the relatively low
nucleation rates, e.g. relatively small number of droplets to be counted, the quantita-
tive nucleation rate measurements are straightforward [4–8]. Measuring nucleation
rate as a function of supersaturation at a constant temperature, one can determine the
size of the critical cluster using the nucleation theorem (Chap. 4). The experimentally
determined critical cluster can then be compared to the nucleation models.
Note, however, that nonlinear temperature and pressure profiles inside the chamber
can lead to substantial nonuniformities of temperatures and supersaturations in the
working volume making it difficult to assign particular values to supersaturations
and temperatures corresponding to the observed nucleation rates.
pv
S=
psat (T )
280 16 Experimental Methods
Fig. 16.3 Time dependence of the supersaturation in the expansion cloud chamber during a single
nucleation experiment. Experiment starts when vapor is supersaturated. As a result of adiabatic
expansion vapor becomes supersaturated and nucleation occurs during the time of nucleation pulse.
After that a slight recompression terminates nucleation process; condensational growth of droplets
results in further reduction of the supersaturation due to vapor depletion. (Reprinted with permission
from Ref. [10], copyright (1994), American Chemical Society.)
increases because the decrease of its partial pressure p v during the isentropic expan-
sion is slower than the exponential decrease of the saturation pressure psat with
temperature, which is given by the Clapeyron equation (2.14). As opposed to the
diffusion chamber, nucleation of the supersaturated vapor in the working volume of
the expansion chamber takes place at uniform conditions.
Among various modifications of the expansion camber—single-piston chamber [11,
12], piston-expansion tube [13]—we will describe in a somewhat more detail the
nucleation pulse chamber (NPC) [10, 14, 15]. In order to ensure the constant con-
ditions during the nucleation period a small recompression pulse is issued in NPC
after the completion of the adiabatic expansion, which terminates the nucleation
process after a short time, of the order of 1 ms, called the nucleation pulse. Conden-
sational growth of droplets results in further reduction of the supersaturation due to
vapor depletion. Schematically evolution of the supersaturation in the NPC during
the nucleation experiment is depicted in Fig. 16.3. The cooling rate in the NPC is of
the order 104 K/s.
The values of supersaturation and temperature corresponding to the measured nucle-
ation rate are calculated from the following considerations. If p0 is the initial total
pressure of the vapor/gas mixture, T0 is the initial temperature and y is the vapor
molar fraction, then
p0v = y p0
is the partial vapor pressure at the initial conditions. After adiabatic expansion the
total pressure drops by Δpexpt becoming equal to
p = p0 − Δpexpt
16.2 Expansion Cloud Chamber 281
where κ = c p /cv is the ratio of specific heats for the vapor/gas mixture. Then, from
the saturation vapor pressure at temperature T , psat (T ), given by the Clapeyron
equation (2.14), the supersaturation is found to be
y ( p0 − Δpexpt )
Sexpt = (16.2)
psat (T )
During the short (∼1 ms) nucleation pulse only a negligible fraction of the vapor
is consumed, i.e. depletion effects are negligible which leaves the supersaturation
practically constant. After recompression supersaturation drops, nucleation is sup-
pressed so that only particle growth at constant number density occurs (while no new
droplets are formed). Thus, the nucleation pulse method realized in the expansion
tube makes it possible to decouple nucleation and growth processes.
The last step is to determine the number density ρd of droplets formed during the
nucleation pulse. In the NPC the nucleated droplets grow to the sizes ∼1 μm when
they are detected by the constant angle Mie scattering (CAMS) technique leading
to determination of ρd . CAMS, which uses a laser operating in the visual, is based
on the Mie theory of scattering of electromagnetic waves by dielectric spherical
particles [16].
The basic idea behind the technique is straightforward: (i) analyzing the time evolu-
tion of the intensity of light scattered by the droplets and comparing it with the Mie
theory, one finds the size rd of the droplet at time t; (ii) analyzing the evolution of
the intensity of the transmitted light one determines the number density of droplets
using the value of extinction coefficient corresponding to the droplet size rd .
To clarify CAMS, we briefly formulate the main results of the Mie theory (for details
the reader is referred to Refs. [16, 17]) relevant for the analysis of nucleation exper-
iments. Consider a single dielectric spherical particle of radius rd emerged in the
vacuum and having the refractive index m. The particle is illuminated by the inci-
dent light with a wavelength λ. Let us introduce the dimensionless droplet radius2
2π
α= rd = k rd
λ
2If instead of vacuum the particle is emerged in a homogeneous medium with the refractive index
m medium , the wavelength should be replaced by λvacuum /m medium .
282 16 Experimental Methods
If I0 is the intensity of the incident light (watt/m2 ), the sphere will intercept
Q ext πrd2 I0 watt from the incident beam, independently of the state of polarization of
the latter. The dimensionless quantity Q ext (m, α) is called the extinction efficiency.
In the Mie theory it is given by
∞
2
Q ext (m, α) = (2n + 1) (an + bn ) (16.3)
α2
n=1
Here the complex Mie coefficients an and bn are obtained from matching the bound-
ary conditions at the surface of the spherical droplet. They are expressed in terms of
spherical Bessel functions evaluated at α and y = mα:
where
(2)
and Jn+1/2 (z) is the half-integer-order Bessel function of the first kind, Hn+1/2 (z)
is the half-integer-order Hankel function of the second kind [18].
The intensity of the incident beam decreases with the a distance L (called the optical
path) as it proceeds through the cloud of droplets. The transmitted light intensity is
given by the Lambert-Beer law [19]
(here we assumed that all ρd dielectric spheres in the unit volume are identical).
The behavior of the extinction efficiency Q ext as a function of the size parameter α
is illustrated in Fig. 16.4 for the two substances with the values of refractive index
m = 1.33 (water) and m = 1.55 (silicone oil).
Consider now the light scattered by a single sphere. The direction of scattering
is given by the polar angle θ and the azimuth angle φ. The intensity Iscat,1 of the
scattered light in a point located at a large distance r from the center of the particle
has a form
I0
Iscat,1 = 2 2 F(θ, φ; m, α) (16.6)
k r
16.2 Expansion Cloud Chamber 283
Fig. 16.4 Mie extinction efficiency versus size parameter α for water (m = 1.33) and silicone oil
(m = 1.55). For small particles Q ext ∼ α 4 (Rayleigh limit); for big particles Q ext → 2 (limit of
geometrical optics α → ∞). The largest value of Q ext is achieved when the particle size is close
to the wavelength
where F is the dimensionless function of the direction (not of r ). For the linearly
polarized incident light
F = i 1 sin2 φ + i 2 cos2 φ (16.7)
where
1
πn (cos θ ) = P 1 (cos θ ) (16.10)
sin θ n
d 1
τn (cos θ ) = P (cos θ ) (16.11)
dθ n
we have
i1 + i2 |S1 |2 + |S2 |2
F φ = = (16.12)
2 2
If multiple scattering can be avoided, the total intensity of light scattered in the
direction θ by all spheres in the volume V is
I0 ρd V
Iscat = Iscat,1 ρd V = (|S1 |2 + |S2 |2 ) (16.13)
2 k 2r 2
At fixed m and θ the quantity in the round brackets as a function of α has a distinct
pattern of maxima and minima [16].
During the single nucleation experiment one measures the time dependence of the
expt
scattering intensity Iscat (see Fig. 16.5) which has a form of a sequence of maxima
and minima. Assuming that on the time scale of experiment droplets, nucleated
during the pulse, grow without coagulation and Ostwald ripening, one can state that
their number density ρd remains constant. That is why in order to determine ρd it
expt
is sufficient to compare the first peak of the function Iscat with the first maximum
of the scattering intensity from the Mie theory. This comparison yields the droplet
radius rd at the first peak, which after substitution into (16.4) and (16.5) yields
1
ln(I0 /I )
ρd = L
(16.14)
π rd2 Q ext (m, α)
16.2 Expansion Cloud Chamber 285
Note, that this procedure does not provide information about the droplet growth
rd (t)—the latter can be obtained from the analysis of series of peaks in the scattering
intensity.
Shock tube realizes the same idea of a short nucleation pulse, providing the separa-
tion in time of the nucleation and growth processes, which we discussed in Sect. 16.2.
In the shock tube this is achieved by means of the shock waves. The tube consists
of two sections: the driver-, or High-Pressure Section (HPS), and the driven-, or the
Low Pressure Section (LPS). The two sections are separated by the diaphragm. A
small amount of condensable vapor is added to the driver section. During the nucle-
ation experiment the diaphragm is rapidly ruptured and the high-pressure vapor/gas
mixture from the driver section sets up a nearly one-dimensional, unsteady flow and
the shock wave traveling from the diaphragm into the driven section. At the same
time the expansion wave travels back—from the diaphragm into the driver section.
Cooling of the rapidly expanding gas in the driver section imposes nucleation.
The construction of the shock tube for nucleation studies was proposed by Peters and
Paikert [21, 22] and further developed by van Dongen and co-workers [20, 23–26].
The scheme of the experimental set-up [20] is shown in Fig. 16.6. The HPS has a
length of 1.25 m, the length of the LPS is 6.42 m. The local widening in the LPS
plays an important role in creating the desired profile of pressure and, accordingly,
the supersaturation: after the rupture of the polyester diaphragm between HPS and
LPS the initial expansion wave traveling from LPS to HPS is followed by a set of
reflections of the shock wave at the widening (see Fig. 16.7). These reflections travel
286 16 Experimental Methods
Fig. 16.6 Pulse expansion wave tube set-up. (Reprinted with permission from Ref. [20], copyright
(1999), American Institute of Physics.)
back into the HPS and create the pulse-shaped expansion at the end wall of the HPS.
After the short pulse and a small recompression the pressure remains constant for a
longer period of time during which no nucleation occurs but the already nucleated
droplets are growing to macroscopic sizes to be detected by means of the scattering
technique. The temperature profile follows the adiabatic Poisson law (16.1).
Similar to the nucleation-pulse chamber, discussed in Sect. 16.2, the number density
of droplets, ρd , is obtained by means of a combination of the constant-angle Mie
scattering and the measured intensity of transmitted light—the procedure described
in Sect. 16.2.1. In the experiments of Refs. [20, 23–25] the droplet cloud in the
HPS was illuminated by the Ar-ion laser with a wavelength λ = 514.2 nm. Since
the observation section in the shock tube is located near the endwall of the HPS,
an obvious choice of the scattering polar angle is θ = 90◦ . Figure 16.8 shows the
optical set-up of the device. The laser beam passes the tube through two conical
windows. The transmitted light is focused by lens L 2 onto photodiode D2 . The
scattered intensity is recorded by the photomultiplier P M.
Because of the nature of the nucleation pulse method, the value of ρd should
be approximately constant in time. The steady-state nucleation rate is given by
Eq. (16.15)
ρd
J=
Δt
where Δt is the duration of the pulse.
As pointed out in the previous section, besides the steady-state nucleation rate one
can obtain from the same experimental data the growth law of the droplets. At each
moment of time during the nucleation experiment for which the measured scattered
16.3 Shock Tube 287
signal is at maximum, one can find the value of the droplet radius by comparison
with the corresponding maximum of the theoretical scattering intensity given by the
Mie theory of Sect. 16.2.1 as illustrated in Fig. 16.9. This gives the droplet growth
curve rd (t).
The absence of moving parts in the shock-tube (as opposed to the expansion chamber)
opens a possibility to study nucleation at sufficiently high nucleation pressures—
up to 40 bar—and reach nucleation rates in the range of 108 − 1011 cm−3 s−1
[20, 28].
288 16 Experimental Methods
Fig. 16.9 Theoretical and experimental scattering patterns for n-nonane droplets. From mutual
correspondence of extrema the time-resolved droplet radius rd (t) is found. (Copied from Ref. [27])
The supersonic nozzle (SSN) relies upon adiabatic expansion of the vapor/gas mix-
ture flowing through a nozzle of some sort. The most widely used type of these
devices contain the Laval (converging/diverging) nozzle [29–32]. The vapor/gas
mixture is undersaturated prior to and slightly after entering the nozzle region. Dur-
ing the flow in the nozzle the mixture becomes saturated and then supersaturated.
Nucleation and growth of the droplets takes place when the flow passes the throat
region of the nozzle. Rapid increase of the supersaturation results in spontaneous
onset of condensation which depletes the vapor and subsequently terminates the
supersaturation.
A typical nucleation pulse is very short ∼10 μs, and cooling rates are very high: ∼5×
105 K/s leading to characteristic nucleation rates as high as 1016 − 1018 cm−3 s−1 .
The droplets formed in SSN are extremely small ∼1 − 20 nm; critical embryos are
even smaller ∼0.1 nm, containing 10–30 molecules. Clearly, droplets of this size can
not be detected by optical devices operating in the visual—shorter wavelength is
required. Methods used for particle characterization in SSN are small-angle neutron
scattering (SANS ) and small-angle x-ray scattering (SAXS).
The schematic diagram of the experimental set-up with the supersonic nozzle and
SAXS unit is shown in Fig. 16.10. The experiment consists of the pressure trace mea-
surements during the expansion and the SAXS measurements. A movable pressure
probe measures the pressure profile of the gas p(x) along the axis of the nozzle. The
condensible vapor mole fraction y is determined from the mass flow measurements.
Using the stagnation conditions p0 , T0 of the vapor/gas mixture, one determines the
pressure profile of the vapor along the nozzle
p(x) g(x)
p v (x) = y p0 1−
p0 g∞
16.4 Supersonic Nozzle 289
Fig. 16.10 Schematic diagram of the experimental set-up with supersonic nozzle and SAXS unit.
(Copied from Ref. [33])
ṁ v
g∞ =
ṁ v + ṁ gas
where ṁ v , ṁ gas are the mass flow rates of the vapor and gas, respectively. Then, the
supersaturation profile is
p v (x)
S(x) =
psat (T (x))
where T (x) is the temperature at point x found from the Poisson equation.
Using SAXS technique, one studies elastic scattering of X-rays by a cloud of droplets.
Scattering leads to interference effects and results in a pattern, which can be analyzed
to provide information about the size of droplets and their number density. Let us
define a scattering vector (length) according to
4π
q= sin(θ/2)
λ
where θ is the scattering angle, λ is wavelength of the incident beam; for SAXS
λ ≈ 1Å. The scattering intensity Iscat is proportional to the number density of
290 16 Experimental Methods
where ρ S L D is the contrast factor, being the difference in the scattering length density
between the liquid droplet and surrounding bulk gas. Assuming Gaussian distribution
of droplet sizes with the mean rd and the width σ , Iscat can be written as
1 (rd − rd )2
Iscat (q) = ρd √ exp − P(q, rd ) drd
σ 2π 2σ 2
Fitting the measured scattering intensity to this expression, one finds the desired
quantities ρd and rd .
Figure 16.11 from Ref. [33] illustrates this procedure for SSN experiments with
n-butanol at plenum temperature T0 = 50◦ C and pressure p0 = 30.2 kPa. The solid
line gives the fit to the Gaussian distribution with the following set of parameters
Taking into account the pulse duration Δt ≈ 10 μs, this leads to the nucleation rate
References
Water
ρmass
l
= 0.08 tanh y + 0.7415 x 0.33 + 0.32 g/cm3
T
x = 1 − , y = (T − 225)/46.2
Tc
Nitrogen
ρsat
l
ln = 1.48654237 x 0.3294 − 0.280476066 x 4/6
ρc
+ 0.0894143085 x 16/6 − 0.119879866 x 35/6 ,
T
x = 1−
Tc
psat Tc
ln = −6.12445284 x + 1.2632722 x 3/2 − 0.765910082 x 5/2 − 1.77570564 x 5
pc T
Mercury
ρmass
l
[g/cm3 ] = 13.595 [1 − 10−6 (181.456 TCels + 0.009205 TCels
2
+ 0.000006608 TCels
3
+ 0.000000067320 TCels
4
)]
Appendix A: Thermodynamic Properties 295
Argon
psat Tc
ln = −5.904188529 x + 1.125495907 x 1.5 − 0.7632579126 x 3 − 1.697334376 x 6
pc T
N-nonane
ρmass
l
= 0.733503 − 7.87562 × 10−4 TCels − 9.68937 × 10−8 TCels
2
− 1.29616 × 10−9 TCels
3
g/cm3
where Tr = T /Tc .
Appendix B
Size of a Chain-Like Molecule
As one of the input parameters MKNT and CGNT use the size of the molecule.
For a chain-like molecule, like nonane, it can be characterized by the radius of
gyration Rg —the quantity used in polymer physics representing the mean square
length between all pairs of segments in the chain [11]:
Nsegm
1
Rg2 = 2
(Ri − R j )2
2Nsegm
i, j=1
Nsegm
1
Rg2 = (Ri − R0 )2
Nsegm
i=1
where R0 is the position of the center of mass of the chain. The latter expression
shows that the chain-like molecule can be appropriately represented as a sphere with
the radius Rg . The radius of gyration can be found using the Statistical Associating
Fluid Theory (SAFT) [12]. Within the SAFT a molecule of a pure n-alkane can be
modelled as a homonuclear chain with Nsegm segments of equal diameter σs and
the same dispersive energy ε, bonded tangentially to form the chain. The soft-SAFT
correlations for pure alkanes read [13]:
where M is the molecular weight (in g/mol). Thus, the number of segments and the
size of a single segment depend only on the molecular weight. Note, that within this
approach Nsegm is not necessarily an integer number. Having determined Nsegm , the
radius of gyration can be calculated using the Gaussian chain model in the theory of
polymers [11]:
Nsegm
Rg = σsegm (B.3)
6
σ = 2 Rg (B.4)
ρ∗
T∗ = (3 − ρ ∗ )2 (C.2)
4
∗ we obtain using the standard
Solving Eq. (C.2) for the spinodal vapor density ρsp
v
methods [15]:
∗ 1
(T ∗ ) = 2 − 2 cos β = arccos(1 − 2T ∗ )
v
ρsp β , (C.3)
3
Substitution of (C.3) into the van der Waals equation (C.1) yields the vapor pressure
at the spinodal:
∗ v 8 4T ∗ − 3 cos 13 β + 3 cos 23 β sin2 16 β
psp = 1 (C.4)
1 + 2 cos 3 β
∗ (T ∗ )
v
psp 1
Ssp (T ∗ ) = =
p ∗ (T ∗ ) ∗ (T ∗ )
psat
sat
8 4T ∗ − 3 cos 13 β + 3 cos 23 β sin2 16 β
(C.5)
1 + 2 cos 13 β
where the quantities with the superscript α refer to the phase α. Since
Nα
Vα =
ρα
1 Nα
dV α = dN α
− dρ α
ρα i
(ρ α )2
N αj
y αj = α (D.3)
k Nk
α α α
In view of normalization m k=1 yk = 1, ρ is a function of m − 1 variables yk
and one is free to choose a particular component to be excluded from the list of
independent variables. Discussing the partial molecular volume of component i,
it is convenient to exclude this component from the list, i.e. to set
ρ α = ρ α (y1α , . . . , yi−1
α α
, yi+1 , . . . , ym )
Then the right-hand side of (D.2) can be expressed using the chain rule:
∂ ln ρ α ∂ ln ρ α ∂ y αj
= (D.4)
∂ Niα pα ,T, N α ∂ y αj ∂ Niα
j, j =i j=i
Substituting (D.5) into ((D.2) and D.4) we obtain the general result
α
1 α α ∂ ln ρ
viα α
= α ηi , ηi = 1 + yj , α = v, l (D.6)
ρ ∂ y αj
j=i p α ,T
ρkB T
p= − am ρ 2 (D.10)
1 − bm ρ
where the van der Waals parameters am , bm for the mixture read [2]:
√ √
am = (ya aa + yb ab )2 (D.11)
bm = ya ba + yb bb (D.12)
where yi is the molar fraction of component i in vapor or liquid; ai and bi are the
van der Waals parameters for the pure fluid i. According to the definition of the
partial molecular volume consider a small perturbation in a number of molecules
of component a at a fixed pressure, temperature and the number of molecules of
component b. This perturbation results in the change of ρ, am and bm :
m = am + Δam
a (D.13)
bm = bm + Δbm (D.14)
ρ
= ρ + Δρ (D.15)
Substituting (D.13)–(D.15) into the van der Waals Eq. (D.10) and linearizing in
Δam , Δbm and Δρ we find
ρkB T Δρ kB T ρ kB T
p− + am ρ 2 = + (Δbm ρ + Δρ bm ) − ρ 2 Δam − 2am ρΔρ
1 − bm ρ 1 − bm ρ (1 − bm ρ)2
Van der Waals parameters am and bm change due to the variation in molar fractions
satisfying Δya = −Δyb :
√ √ √
Δam = 2 Δya am ( aa − ab ) (D.17)
Δbm = Δya (ba − bb ) (D.18)
vdW
√ √ √
∂ ln ρ (1 − bm ρ)2 a m ( aa − a b )
2ρ ρ(ba − bb )
= −
∂ ya mρ (1 − bm ρ)2
p,T 1 − 2akB T (1 − bm ρ)
2 kB T
(D.20)
Substituting this result into (D.7)–(D.8) we obtain the expression for ηi and hence
for the partial molecular volumes
1 ∂ ln ρ vdW
va = 1 − yb (D.21)
ρ ∂ ya p,T
1 ∂ ln ρ vdW
vb = 1 + ya (D.22)
ρ ∂ ya p,T
The reduced partial molecular volumes of the components in the liquid phase,
ηil , i = a, b play an important role for binary nucleation, especially at high pressures.
Table D.1 shows the values of these parameters for the mixture n-nonane (a)/methane
(b) for the nucleation temperature T = 240 K and various total pressures.
Appendix E
Mixtures of Hard Spheres
Ri = di /2
Ai = π di2
Vi = π di3 /6
Note that Ai can be regarded as a molecular surface area, whereas Vi denotes the
molecular volume of component i. Next, the parameters ξ (k) are defined as
ξ (0) = ρ1 + ρ2
ξ (1) = ρ1 R1 + ρ2 R2
ξ (2) = ρ1 A1 + ρ2 A2
ξ (3) = ρ1 V1 + ρ2 V2 .
Note that ξ (3) is the total volume fraction occupied by hard spheres. For notational
convenience, we also introduce
η = 1 − ξ (3) .
c(0) = − ln η
ξ (2)
c(1) =
η
(2) 2
(2) ξ (1) ξ
c = +
η 8π η2
(2) 3
(3) ξ (0) ξ (1) ξ (2) ξ
c = + + .
η η 2 12π η3
p3 = c(3) k B T (E.1)
(2) 3 (3)
ξ ξ
p2 = p3 − kB T (E.2)
12π η3
2 p3 + p2
pd = . (E.3)
3
Let us introduce for brevity of notations two additional quantities:
(1) 1 − 23 ξ (3) ln η
Y =3 + (3)
η2 ξ
2
ξ (2) η + 1 − 2ξ (3) 2 ln η
Y (2) = (3) + (3) .
6ξ η3 ξ
The chemical potentials μd,i can be derived from the virial equation using standard
thermodynamic relationships:
(3)
μi = c(0) + c(1) Ri + c(2) Ai + c(3) Vi (E.4)
2
(2) (3) Ri ξ (2) (1)
μi = μi + (3)
Y − 2Ri Y (2) (E.5)
3ξ
(3) (2)
+ μi
2μi
μiex = kB T (E.6)
3
μd,i = k B T ln ρi Λi3 + μiex (E.7)
The last expression presents the chemical potential of a species as a sum of an ideal
and excess contributions.
Appendix F
Second Virial Coefficient for Pure Substances
and Mixtures
from first principles requires the knowledge of the intermolecular potential u(r)
which is in most cases is not available. With this limited ability second virial
coefficient is calculated from appropriate corresponding states correlations. For
nonpolar substances such correlation has the form due to Tsanopoulos [2, 19]:
B2 pc
= f0 + ω P f1 (F.2)
kB Tc
where
Expressions (F.5)–(F.6) agree with (F.3)–(F.4) to within 0.01 for Tr > 0.6 and
ω P < 0.4, while for lower Tr the difference grows rapidly.
For the mixtures the second virial coefficient is written using the mixing rule
where B2,ii are the second virial coefficient of the pure components. For the cross term
B2,i j the combining rules should be devised to obtain Tc,i j , pc,i j and ω P,i j which
then are substituted into the pure component expression (F.2 ), with the coefficients
f 0 and f 1 satisfying (F.3)–(F.4) or (F.5)–(F.6). For typical applications the following
combining rules are used [2]
In Chap. 13 we search for the saddle-point of the free energy of cluster formation in
the space of total numbers of molecules of each species in the cluster. For calculation
of g(n a , n b ) we choose an arbitrary bulk composition n il and find the excess numbers
n iexc according to Eqs. (11.84)–(11.85). The total numbers of molecules are: n i =
n il + n iexc . Then, g(n a , n b ) is found from Eq. (13.91). Is easy to see that although
we span the entire space of (nonnegative) bulk numbers (n al , n lb ), the space of total
numbers (n a , n b ) contains “holes”, i.e. the points, to which no value of g(n a , n b ) is
assigned. This feature complicates the search of the saddle point of g(n a , n b ).
To overcome this difficulty we apply a smoothing procedure aimed at elimination
of the holes in (n a , n b )-space by an appropriate interpolation procedure between the
known values. The simplest procedure for the 2D space is the bilinear interpolation
which presents the function g at an arbitrary point (n a , n b ) as
g(n a , n b ) = a n a + b n b + c n a n b + d (G.1)
where coefficients a, b, c, d are defined by the known values of g around the point
(n a , n b ). However, due to randomness of the location of the “holes”, the straightfor-
ward application of bilinear interpolation is quite complicated. This difficulty can be
avoided if we notice that (G.1) is the solution of the 2D Laplace equation
∂2 ∂2
Δg(n a , n b ) = 0, Δ ≡ + (G.2)
∂n a2 ∂n 2b
∂2g
= g(n a − 1, n b ) − 2g(n a , n b ) + g(n a + 1, n b )
∂n a2
∂2g
= g(n a , n b − 1) − 2g(n a , n b ) + g(n a , n b + 1)
∂n 2b
g i+1 (n a − 1, n b ) + g i+1 (n a , n b − 1) + g i (n a + 1, n b ) + g i (n a , n b + 1)
g i+1 (n a , n b ) =
4
(G.4)
where g i is the value of g at i-th iteration step. The procedure is repeated until
g i+1 (n a , n b ) ≈ g i (n a , n b ).
The saddle point of the smoothed Gibbs function satisfies
∂g ∂g
= =0
∂n a n b ∂n b n a
Note, that computationally it is preferable to search for the saddle point by solving
the equivalent variational problem:
! "2 ! "2
∂g ∂g
+ → min (G.5)
∂n a n b ∂n b n a
References
12. W.G. Chapman, K.E. Gubbins, G. Jackson, M. Radosz, Fluid Phase Equilib. 52, 31 (1989)
13. J. Pamies, Ph.D. Thesis, Universitat Rovira i Vrgili, Tarragona, 2003
14. V.I. Kalikmanov, Statistical Physics of Fluids. Basic Concepts and Applications (Springer,
Berlin, 2001)
15. G.A. Korn, T.M. Korn, Mathematical Handbook (McGraw-Hill, New York, 1968)
16. G.A. Mansoori, N.F. Carnahan, K.E. Starling, T.W. Leland, J. Chem. Phys. 54, 1523 (1971)
17. Y. Rosenfeld, J. Chem. Phys. 89, 4272 (1988)
18. C.C.M. Luijten, Ph.D. Thesis, Eindhoven University, 1999
19. C. Tsanopoulos, AICHE J. 20, 263 (1974)
20. H.C. van Ness, M.M. Abbott, Classical Thermodynamics of Non-electrolyte Solutions
(McGraw, New York, 1982)
21. F. O’Sullivan, J. Amer. Stat. Assoc. 85, 213 (1990)
22. J. Stoer, R. Bulirsch, Introduction to Numerical Analysis (Springer, New York, 1993)
Index
D I
Detailed balance, 25, 31, 76, 173, 217 Ideal gas, 7, 18, 28, 29, 72, 73, 75, 90, 173,
Diagrammatic expansion, 225 190, 202, 216, 226, 240
Diffusion cloud chamber, 102 Ideal mixture, 190, 194, 209
Direction of principal growth approximation, Impingement rate, 25, 173, 180, 235, 243, 260
177, 220, 241, 242 average, in binary nucleation, 221, 235,
Distribution function 249
one-particle, 57 Importance sampling
Dupre-Young equation, 254, 262 in MC simulations, 126
Intrinsic
chemical potential, 60
free energy, 57, 58, 61, 63, 65, 67
E Ising model, 268
Entropy Isothermal compressibility, 148
configurational, 84
bulk per molecule, 85
Exclusion volume, 130 K
Expansion cloud chamber, 275, 277, 278 Kelvin equation
classical, 30
F
Fisher droplet model, 79 L
Fletcher theory, 251 Lagrange equation, 115
Fluctuation theory, 19, 88, 174, 218 Landau
Fokker-Planck equation, 35 expansion, 146
Frenkel distribution, 32 Laplace equation
Fugacity, 82, 87, 88, 224, 228 generalized, 12
Functional standard, 12
grand potential, 65, 67, 205, 207 Latent heat, 7, 51, 121, 122, 141, 165, 199, 216
Helmholtz free energy, 55, 57–59, 65, 69, Lattice-gas model, 142
149, 205, 206 Laval Supersonic Nozzle, 162, 275, 286
Law of mass action, 32
Lennard-Jones potential, 141
Limiting consistency, 33
G Line tension, 261–265, 267–273
Gibbs Local density approximation (LDA), 62, 206
dividing surface, 9–13, 21, 44, 47, 51, 56,
76, 83, 92, 181–183, 185–187, 189,
209, 226, 236, 238, 239, 252 M
equimolar surface, 10, 12, 13, 21, Matrix
107, 186 inverse, 241
surface of tension, 12, 13 transposed, 241
free energy, 20 unit, 241
of droplet formation, 20 unitary, 241, 242
Mean-field
approximation, 83, 86, 96
H Mean-field Kinetic Nucleation Theory, 80, 97
Hard-spheres Metastability
Carnahan-Starling theory, 63, 64, 67, 96 parameter , 200, 202, 216, 235
cavity function, 96 Metastable state, 17
effective diameter, 63 Microscopic surface tension, 79, 88, 90, 106,
Heat capacity, 121 112, 157, 164, 230, 232, 233
Heat of adsorption, 260 reduced, 88, 228, 230, 233
Hill equation, 51 Mie theroy, 279, 280, 282, 285
Index 315
S
N Saddle point, 55, 68, 149, 150, 171, 174,
Nucleation 176–180, 185, 189, 195, 205,
boundary conditions, 35, 36, 63, 119, 120, 211, 220, 221, 233, 234,
127, 136, 178, 211, 246 241–243, 249
Nucleation barrier, 22, 47, 50, 76, 77, 98, 99, Saturation
130, 132, 152–155, 158, 185, 211, pressure, 6, 168, 229
215, 235, 258, 265, 271 Scattering intensity, 282, 283, 285, 288
Nucleation pulse, 277–279, 282–284, 286 Shock tube, 275, 283–285
Nucleation pulse chamber, 278, 279 Small Angle Neutron Scattering (SANS), 286
Nucleation Theorem Small Angle X-ray Scattering
first, 44, 48 (SAXS), 286–288
pressure, 53 Spherical particle
second, 50 contrast factor of, 288
form-factor of, 288
Spinodal, 56, 69, 132, 145, 146, 150, 151–154,
O 162–163
Order parameter, 17, 146, 149, 152 decomposition, 145, 147, 153
kinetic, 153
thermodynamic, 145, 149, 162–164
P Supercritical solution
Packing fraction, 95, 97 rapid expansion of (RESS method), 124
Partial Supersaturation, 18, 151
molecular volume, 52, 182, 200, 231, 238 Surface diffusion, 259
vapor pressure, 190 Surface enrichment, 141, 181, 194, 202, 205,
Partition function 207, 209, 210, 213
canonical, 58 Surface tension
grand, 60 macroscopic, 69, 72, 100, 103, 104, 164,
Periodic boundary conditions, 120, 127 186, 192, 199, 200, 215
Phase transition
first order, 1, 7
Poisson law, 279, 284 T
Pseudospinodal, 98, 145, 152–158, 162, Thermal diffusion cloud chamber, 275, 276
163, 168 Thermodynamics
first law, 7
Threshold method
R in MD simulations, 134–136, 138,
Radius of gyration 141, 165
of a polymer molecule, 295 Time-lag, 39
Random number, 126 Tolman equation, 108
Random phase approximation (RPA), 63, 67 Tolman length, 13, 108
Random processes
theory of, 272
Rate U
forward, 24–26, 31, 80 Umbrella sampling, 127, 133
316 Index