You are on page 1of 10

Fluid Phase Equilibria 194–197 (2002) 521–530

A SAFT–DFT approach for the vapour–liquid


interface of associating fluids
Guy J. Gloor a , Felipe J. Blas a,b,∗ , Elvira Martı́n del Rı́o b ,
Enrique de Miguel b , George Jackson a
a
Department of Chemical Engineering and Chemical Technology, Imperial College of Science,
Technology and Medicine, Prince Consort Road, SW7 2BY London, UK
b
Departamento de Fı́sica Aplicada, Escuela Politécnica Superior, Universidad de Huelva,
21819 La Rábida, Huelva, Spain
Received 30 March 2001; accepted 15 October 2001

Abstract
We present a density functional theory (DFT) based on the statistical associating fluid theory (SAFT) bulk
free energy to describe the behaviour of inhomogeneous associating molecular fluids, with a special emphasis on
the vapour–liquid interface. The molecules are described in terms of hard-sphere (HS) segments which interact
through an arbitrary intermolecular potential. Off-centre square-well bonding sites are incorporated to mediate the
association between molecules. The approach presented here is based on a perturbation theory about a HS reference
fluid under the local density approximation (LDA); the contributions due to chain formation and association are
also considered at the local density level. The Helmholtz free energy functional due to the dispersive attractive
interactions is treated at the mean-field level by ignoring the correlations. This description is essentially a ‘van
der Waals’ theory of non-uniform fluids. The incorporation of a SAFT free energy of the bulk associating fluid
represents the simplest extension of the theory to deal with inhomogeneous associating systems (SAFT–DFT).
The resulting functional is used to investigate the effect of the range of the attractive interactions on the interfacial
properties, such as the density profile and surface tension, for two different intermolecular potential models, namely
the Yukawa and square-well. This simple SAFT–DFT approach is also used to make quantitative comparisons with
the experimental surface tension for water. © 2002 Elsevier Science B.V. All rights reserved.
Keywords: Interfacial tension; Statistical mechanics; Equation of state; Vapour–liquid equilibria; SAFT

1. Introduction

A description of interfacial phenomena of complex fluids at the molecular level is a challenging problem.
An understanding of surfaces is central to many processes of practical importance, including detergency

Corresponding author. Tel.: +34-959-017-584; fax: +34-959-017-304/350-311.
E-mail address: felipe.jimenez@dfaie.uhu.es (F.J. Blas).

0378-3812/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 8 - 3 8 1 2 ( 0 1 ) 0 0 7 7 4 - 9
522 G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530

and the behaviour of surface active agents (surfactants), colloid stability (emulsions such as milk or
aerosols such as smoke), the formation of micellar and lyotropic liquid crystalline phases (foods and
cosmetics), and the behaviour of cell membranes. There is clearly a great desirability in developing
accurate predictive theories of interfacial phenomena. Density functional theory (DFT), in which the free
energy of the system is expressed as a functional of the spatially varying single particle density, is one
of the most powerful tools to describe inhomogeneous systems [1]. This method, which in principle can
be applied to treat any inhomogeneous problem, is particularly well suited to describe vapour–liquid
interfaces. The vapour–liquid interface can be described using a perturbation theory about a hard-sphere
(HS) reference fluid, where the reference term is treated within the local density approximation (LDA) [2]
and the correlations are ignored in the attractive perturbation term. This constitutes the so-called ‘van der
Waals theory’ of non-uniform fluids [3]. The statistical associating fluid theory (SAFT) [4,5] is proving
to be one of the most sophisticated, successful, and versatile equations of state (EoS) to describe the bulk
thermodynamic properties of complex fluid mixtures (see [6] for a recent overview). This approach is
based on the theory of Wertheim [7] where the association is treated in the context of both a perturbation
theory and an integral equation. In the early development of the theory, the perturbation expressions
of Wertheim were incorporated into an EoS for mixtures of associating chain molecules formed from
HS segments, with the dispersion forces treated at the mean-field level of van der Waals [8,9]. We will
also describe the bulk system with this simplified SAFT free energy which we refer to as the SAFT-HS
approach [10].
In this work, we examine the vapour–liquid interface of associating fluids by combining the SAFT and
DFT approaches. The use of a Wertheim-SAFT description within a DFT approach was first suggested
by Chapman [11]. Numerous studies of inhomogeneous associating fluids have now appeared in the
literature, but this has been principally for confined systems (see [12] for details). As in recent work [12],
we use the relatively simple SAFT-HS representation because it incorporates all of the essential physics
of associating fluids and provides a good representation of the vapour pressure and coexisting densities
of strongly associating fluids such as water [10]. In order to keep the representation of the bulk fluid and
interface at the same level of approximation we use the van der Waals theory for non-uniform fluids,
which is a DFT at the LDA level. There have been other studies using the SAFT free energy to describe
the surface tension of aqueous systems e.g. [13,14] involving a more phenomenological approach (see
[12] for details). However, a DFT approach offers a more promising molecular based predictive platform
for interfacial properties.

2. Model and theory

The associating molecules are formed from spherical segments (monomers) each of diameter σ .
The interaction between monomers has two contributions, a short-range repulsive part, given by a
HS potential, and a long-range attractive term. We consider two different intermolecular attractive
models, the Yukawa potential, φ att (r) = −ε exp[−λ(r/σ − 1)]/(r/σ ) where λ−1 is the range of the
potential, and the square-well potential where φ att (r) = −ε when the separation is within the range
of the spherical cutoff σ < r < λσ . Association between molecules is mediated by one (or more)
embedded off-centre bonding sites placed at a distance rd from the centre of a given mono-
mer; the site–site interaction is represented by a square-well potential of range rc and
depth εhb .
G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530 523

For convenience, we consider an open system at temperature T and chemical potential µ in a volume
V. In the absence of external fields, the grand potential Ω of the inhomogeneous system is given by [1]:

Ω[ρ(r )] = F [ρ(r )] − µ dr ρ(r ) (1)

where ρ(r ) is the number density at positions r, and F [ρ(r )] is the intrinsic Helmholtz free energy of the
system, which is a functional of ρ(r ). Following a perturbative scheme, it will be assumed that F [ρ(r )]
can be represented as [1]:

F [ρ(r )] = dr f ref (r ) + F att [ρ(r )] (2)

where the first term on the right-hand side (reference term) accounts for the relevant contributions of the
intermolecular interactions (repulsive forces, short-range associations and chain contribution) to the free
energy; Fatt is the contribution to the free energy due to the dispersive attractive interactions. It will be
assumed that the reference term can be treated in a LDA such that f ref (r ) = f ref (ρ(r )) is the Helmholtz
free energy density of a (spatially) uniform system with constant density ρ(r ) = ρ. Following the usual
perturbation DFT
 in
 the mean-field description,
  correlations between molecules at r and r are neglected,
i.e. F att = 2 dr dr  ρ(r̄)ρ(r̄  )φ att (r − r). In the spirit of the SAFT-HS EoS for the bulk fluid, we
1

propose the following expression for f ref [ρ(r )]:

f ref [ρ(r )] = f ideal [ρ(r )] + f hs [ρ(r )] + f chain [ρ(r )] + f assoc [ρ(r )] (3)

In Eq. (3), fideal is the ideal gas contribution, fhs the excess free energy density of the repulsive (HS) part, for
which the Carnahan–Starling expression will be used here [15], fchain the Helmholtz free energy density
due to the formation of the chain molecule [9], with the monomer–monomer HS pair distribution function
given by the expression of Carnahan and Starling evaluated locally. The association contribution, fassoc , is
obtained directly from Wertheim’s first-order thermodynamic perturbation theory in terms of XA (r ), the
fraction of molecules not bonded at a given site A at r [8,9]. As Chapman [11] notes, the original theory
of Wertheim is formulated in the context of inhomogeneous fluids. At the first-order level, the association
contribution is exact even in a more general non-local approach.
At equilibrium, the functional Ω[ρ(r )] is a minimum, and the corresponding density profile, ρ(r ) is
obtained from the solution of the Euler–Lagrange equation resulting from the functional differentiation
of Eq. (1):

δΩ[ρ(r )]
=0 (4)
δρ(r )

An iterative numerical scheme is employed to determine ρ(r ). Once the equilibrium density profile is
obtained, the surface tension is calculated from the usual thermodynamic definition γ = (Ω + PV)/A
where A is the area of the interface [1]. Calculation of the interfacial properties at a given temperature
requires accurate values of bulk properties of the coexisting phases. The latter are obtained from the
Helmholtz free energy functional given by Eq. (3) which, in the limit of a uniform system ρ(r ) = ρ bulk ,
reduces to the SAFT-HS free energy [10].
524 G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530

3. Results and discussion

The SAFT–DFT approach outlined in the previous section is applied to study the vapour–liquid interface
for two types of attractive intermolecular potentials of variable range, the Yukawa and square-well models.
In the following discussion, the integrated mean-field energy εmf = a/b is chosen as the unit of energy
and the HS diameter σ as the unit of length; here a = −2π φ att (r)r 2 dr and b = πσ 3 /6 are van
der Waals-like constants which are natural to any mean-field theory. Accordingly, the following reduced
quantities are defined: temperature, T ∗ = kB T /εmf ; density, ρ ∗ = ρσ 3 ; pressure, P ∗ = P σ 3 /εmf ; surface
tension, γ ∗ = γ σ 2 /εmf ; length, z∗ = z/σ .
At first, we consider systems of spherical molecules (f chain = 0) with no association (f assoc = 0), and
investigate the effect of the range of the attractive interaction on interfacial properties, such as the density
profile and surface tension. As will be shown later, the range of the potential turns out to be an important
parameter in obtaining a quantitative description of associating molecules such as water; in this case the
full SAFT functional has to be used to describe the association.
The dependence of the density profile on the range of the potential for molecules that interact through the
Yukawa and square-well models is depicted in Fig. 1; a fixed reduced temperature of Tr = T ∗ /Tc∗ = 0.8,
where the critical temperature Tc∗ , is examined for each model. For both the Yukawa and square-well
models, the density profiles become broader as the range of the potential is increased. It is important
to note that in this mean-field approach the integrated energy ε mf (van der Waals attractive energy) is
constant; this implies that an increase in the range of the potential must be accompanied by a corresponding
decrease in the strength of the interaction (well depth ε). The overall result is a broadening of the interface
as seen in Fig. 1.
In Fig. 2 we present the temperature dependence of the surface tension of the Yukawa and square-well
models for different values of the potential range. For a given reduced temperature the reduced surface
tension is seen to increase with increasing range of the attractive interaction. This indicates that the
non-local attractive contribution to the grand potential increases when the range is increased due to the
broadening of the interface (larger interfacial thickness); one should also note that in this mean-field
treatment the coexisting densities do not depend on the range of the potential. At low temperatures there
is a near linear dependence of the surface tension with temperature, with a more marked curvature close
to the critical point.
Our attention now turns to the interfacial properties of the most ubiquitous associating fluid, namely
water. As in earlier work with the SAFT-HS mean-field approach [10], the water molecule is modelled
as a hard-spherical core (f chain = 0) with four embedded off-centre square-well bonding sites. The four
bonding sites model the hydrogen bonding between the two electron lone pairs and the two hydrogen
atoms of the water molecule. The attractive interactions between the water molecules are represented by
a square-well potential, treated at the mean-field level; in the bulk fluid the precise form of the attractive
potential is unimportant. The intermolecular potential parameters for this SAFT-HS model of water were
determined to provide the optimum description of the vapour pressure and saturated liquid densities for the
bulk fluid and then rescaled to the critical point [10]. The values of the parameters for the molecular size,
integrated mean-field attractive energy, hydrogen-bond well-depth, and bonding volume are, respectively:
σ = 3.589 Å, εmf /kB = 4384 K, εhb /kB = 1534 K, and K hb = 1.35 Å3 .
A preliminary comparison of the experimental surface tension [16] with SAFT–DFT description is
shown in Fig. 3. The calculations are performed for both the square-well and Yukawa description of
the dispersive interactions. The optimum representation of the experimental surface tension for water is
G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530 525

Fig. 1. Density profiles for the vapour–liquid interface of non-associating spherical molecules interacting through the (a) Yukawa
and (b) square-well models with different ranges: (a) λ−1 = 0.333 (continuous curve), 0.4 (dashed curve), 0.5 (dotted curve),
0.667 (dot-dashed curve) and 1.0 (double dot-dashed curve); (b) λ = 1.2 (continuous curve), 1.4 (dashed curve), 1.6 (dotted
curve), 1.8 (dot-dashed curve) and 2.0 (double dot-dashed curve).
526 G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530

Fig. 2. Temperature dependence of the surface tension for the vapour–liquid interface of non-associating spherical molecules
interacting through the Yukawa (a) and square-well (b) models with different potential ranges: (a) λ−1 = 0.333 (continuous curve),
0.4 (dashed curve), 0.5 (dotted curve), 0.667 (dot-dashed curve) and 1.0 (double dot-dashed curve); (b) λ = 1.2 (continuous
curve), 1.4 (dashed curve), 1.6 (dotted curve), 1.8 (dot-dashed curve) and 2.0 (double dot-dashed curve).
G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530 527

Fig. 3. Surface tension at the vapour–liquid interface of water. The circles correspond to experimental data [16] and curves
represent the theoretical predictions using the SAFT–DFT approach with a set of parameters taken from [10]. The dashed curve
corresponds to the Yukawa potential λ−1 = 0.40, and the continuous curve to the square-well λ = 2.0.

obtained with a range of the square-well of λ = 2 and a Yukawa range λ−1 = 0.4; it is gratifying to
find that a large range provides the best description of the surface tension for water. For a system with
infinite range (an infinitesimal well-depth) the van der Waals description is exact. The bulk vapour–liquid
equilibria for our SAFT-HS model of water is obtained by treating the dispersive interactions at the
van der Waals level. In order to be consistent with this mean-field treatment, it is important that a good
description of the interfacial properties is obtained for the larger ranges of the potential. This is indeed
the case: a square-well range of λ = 2 provides the best description of the surface tension. The agreement
between the experimental data and theoretical results is excellent over the entire liquid range. It is also
interesting to note that this simple SAFT–DFT approach is able to reproduce the weak s-shape of the
surface tension–temperature curve found experimentally for water; there is a point of inflexion in the
curve as the system approaches the critical point. This behaviour is not exhibited by non-associating
fluids and can be ascribed to hydrogen bonding [17].

4. Conclusion

We have presented a DFT for the free vapour–liquid interface of associating molecular fluids. The
Helmholtz free energy functional of the system is based on SAFT under the local density approximation,
with the dispersive attractive interactions treated at the mean-field level.
This simple SAFT–DFT approach is first used to investigate the effect of the range of the intermolec-
ular dispersive potential on the interfacial properties of non-associating spherical molecules that interact
528 G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530

through the Yukawa and square-well models. The theory predicts the expected broadening of the density
profiles as the range of the potential is increased. The increase in the reduced surface tension which
accompanies this broadening is less obvious but corresponds to a corresponding increase in the integrated
dispersive energy when the range is increased. As expected, the integrated non-local dispersive contribu-
tion of the inhomogeneous system (which is the main contribution to the surface tension) increases with
the range of the potential. In the case of the square-well potential one can show analytically for a step
function representation of the density profile that our reduced surface tension is essentially a linear func-
tion of the range. The broadening of the profile is a consequence of the increase of the spatial range over
which the density varies from the vapour to the liquid values. It is important to note that the temperature
and surface tension are reduced with respect to the van der Waals size and energy parameters. In order
to make fruitful comparisons for systems of various ranges it is more appropriate to use units reduced in
the conventional way (T ∗ = k B T/ε and γ ∗ = γ σ 2 /ε).
In order to establish the adequacy of this SAFT–DFT approach in describing the interfacial properties
of associating fluids, we examine a simple model for water. The description of the surface tension of
water is in excellent agreement with the experimental values, and the theory even reproduces the change
in curvature of the surface tension with temperature. In this paper, we have only made preliminary com-
parisons with experiment by using the mean-field SAFT-HS free energy. A more accurate description
of the vapour–liquid equilibria of bulk fluids can be obtained with a more sophisticated treatment of the
dispersive interactions, e.g. using the SAFT–VR free energy for potentials of variable range [18]. In future
work, we will incorporate the SAFT–VR description into a DFT to provide a quantitative description
of the interfacial properties for a wide variety of real systems ranging from the non-polar alkanes to
polar molecules, such as ammonia and refrigerants. Local and non-local versions of the DFT will also be
examined, as well as an appropriate treatment of the correlations in the attractive term.

List of symbols
A interfacial area
f local Helmholtz free energy density
F Helmholtz free energy
Fatt Helmholtz free energy due to the attractive interactions
kB Boltzmann constant
Khb bonding volume
P pressure
T temperature
V volume
XA (r ) fraction of molecules not bonded at a given site A at position r̄
z position along the direction perpendicular to the interface

Greek letters
ε dispersive interaction energy
ε hb hydrogen-bond well-depth
ε mf
integrated mean-field energy
φ att (r) intermolecular attractive potential
γ surface tension
λ potential range
G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530 529

µ chemical potential
ρ(r ) density profile at postion r
ρ bulk bulk density
σ diameter of HS segments
Ω grand potential

Subscripts
A site A
c critical property
r reduced with respect to the critical property

Superscripts
assoc association contribution
att attractive contribution
bulk bulk property
chain contribution due to the formation of the chain
hb hydrogen bond
hs hard-sphere contribution
ideal ideal contribution
mf mean-field contribution
ref reference fluid
∗ reduced unit

Acknowledgements

G.J. G. would like to thank BP Amoco Exploration for the award of a studentship. We also acknowledge
the support from the Oil Extraction programme of the Engineering and Physical Sciences Research
Council (EPSRC) for a research fellowship (GR/N20317), the Joint Research Equipment Initiative (JREI)
of the EPSRC for computer hardware (GR/M94427), and the Royal Society/Wolfson Foundation for the
award of a refurbishment grant. Financial support from the Ministerio de Ciencia y Tecnologı́a under
project BFM2001-1420-C02-02 is also acknowledged.

References

[1] D. Henderson (Ed.), Fundamentals of Inhomogeneous Fluids, Marcel Dekker, New York, 1992.
[2] S. Toxvaerd, J. Chem. Phys. 55 (1971) 3116.
[3] D.E. Sullivan, Phys. Rev. A 25 (1982) 1669–1682.
[4] W.G. Chapman, K.E. Gubbins, G. Jackson, M. Radosz, Fluid Phase Equilib. 52 (1989) 31–38.
[5] W.G. Chapman, K.E. Gubbins, G. Jackson, M. Radosz, Ind. Eng. Chem. Res. 29 (1990) 1709–1721.
[6] E.A. Müller, K.E. Gubbins, in: J.V. Sengers, R.F. Kayser, C.J. Peters, H.J. White (Eds.), Equations of State for Fluids and
Fluid Mixtures, Elsevier, Amsterdam, 2000.
[7] (a) M.S. Wertheim, J. Stat. Phys. 35 (1984) 19–34;
(b) M.S. Wertheim, J. Stat. Phys. 35 (1984) 35–47;
(c) M.S. Wertheim, J. Stat. Phys. 42 (1986) 459–476;
(d) M.S. Wertheim, J. Stat. Phys. 42 (1986) 477–492.
530 G.J. Gloor et al. / Fluid Phase Equilibria 194–197 (2002) 521–530

[8] G. Jackson, W.G. Chapman, K.E. Gubbins, Mol. Phys. 65 (1988) 1–31.
[9] W.G. Chapman, G. Jackson, K.E. Gubbins, Mol. Phys. 65 (1988) 1057–1079.
[10] A. Galindo, P.J. Whitehead, G. Jackson, A.N. Burgess, J. Phys. Chem. 100 (1997) 6781–6792.
[11] W.G. Chapman, Ph.D. dissertation, Cornell University, Ithaca, NY, 1988.
[12] F.J. Blas, E.M. del Rı́o, E. de Miguel, G. Jackson, Mol. Phys. 99 (2001) 1851–1865.
[13] H. Kahl, S. Enders, Fluid Phase Equilib. 172 (2000) 27–42.
[14] D. Lu, J.F. Lu, T.Z. Bao, Y.G. Li, Ind. Eng. Chem. Res. 39 (2000) 320–327.
[15] N.F. Carnahan, K.E. Starling, J. Chem. Phys. 51 (1969) 635.
[16] W. Wagner, A. Kruse, Properties of Water and Steam, Springer-Verlag, Berlin, 1998.
[17] J. Rowlinson, B. Widom, Molecular theory of Capillanty Oxford Science Publications, Oxford, 1982.
[18] A. Gil-Villegas, A. Galindo, P.J. Whitehead, S.J. Mills, G. Jackson, A. Burgess, J. Chem. Phys. 106 (1997) 4168–4186.

You might also like