Sunteți pe pagina 1din 8

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/234859899

An Analytical Equation of State for Chain


Molecules Formed from Yukawa Segments

Article in The Journal of Chemical Physics November 1999


DOI: 10.1063/1.480205

CITATIONS READS

64 49

3 authors, including:

Alejandro gil-villegas
Universidad de Guanajuato
86 PUBLICATIONS 2,722 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Adsorption theory for classical and semiclassical systems View project

Electromagnetic wave propagation through stratified media. The effect of discontinuities in the refractive
index derivative and possible applications to photonic crystals. View project

All content following this page was uploaded by Alejandro gil-villegas on 30 May 2014.

The user has requested enhancement of the downloaded file.


JOURNAL OF CHEMICAL PHYSICS VOLUME 111, NUMBER 18 8 NOVEMBER 1999

An analytical equation of state for chain molecules formed


from Yukawa segments
Lowri A. Davies
Center for Molecular Modeling, Department of Chemistry, University of Pennsylvania, Philadelphia,
Pennsylvania 19104-6323
Alejandro Gil-Villegas
Instituto de Fsica, Universidad de Guanajuato, Leon 37150, Mexico
George Jackson
Department of Chemical Engineering and Chemical Technology, Imperial College of Science, Technology,
and Medicine, Prince Consort Road, London SW7 2BY, United Kingdom
Received 1 March 1999; accepted 12 August 1999
We present an analysis of the thermodynamic properties of chain molecules formed from Yukawa
segments using the statistical associating fluid theory with interactions of variable range SAFT-VR
and the high-temperature expansion of the mean-spherical solution MSA-HTE to the Ornstein
Zernike equation for a simple Yukawa fluid. The SAFT-VR expressions derived previously for this
system allow the MSA-HTE equation of state to be reformulated in terms of first-order perturbation
quantities, thus improving its accuracy. Furthermore, the MSA-HTE solution provides a full
theoretical derivation of the perturbation theory used in SAFT-VR, together with a completely
analytical equation of state for chain molecules composed of segments which interact via the
Yukawa potential. 1999 American Institute of Physics. S0021-96069950242-0

I. INTRODUCTION of the solvent. The expression in Eq. 2 also corresponds to


a wave equation for evanescent waves, described by Eq. 1
The Yukawa potential, used in statistical mechanics to for r , that appears, for example, in the propagation of
describe properties of a wide range of systems in condensed electromagnetic waves in an absorbent medium skin
matter, is important for two main reasons. Firstly, it appears depth phenomena28 or in the solution of the Schroedinger
naturally in a large variety of physical problems involving equation for nuclear binding.29 An evanescent wavefunction
screened interactions, such as those observed in polymer so- which gives rise to an interaction between nucleons protons
lutions, molten salts, and dilute solutions of strong and neutrons of the form of Eq. 1 was first demonstrated
electrolytes.1 Secondly, there are a number of analytical re- by Yukawa in 1935.30,1 In this case, is a parameter which
sults for the thermodynamic and structural properties of sys- is directly proportional to the mass of a meson the particle
tems interacting with a Yukawa potential.225 exchanged between a neutron and a proton and inversely
The hard-core Yukawa potential for two spherical par- proportional to the range of the nuclear force between the
ticles of diameter separated by a distance r is given by nucleons.29 The Yukawa potential in both nuclear physics

if r , and statistical mechanics historically corresponds to Eq. 1


for r , whereas the full interaction, including the infinite
r; exp r/ 1
Y
1
if r , hard-sphere-type repulsion, is more commonly known as the
r/ hard-core Yukawa.
where 1 is the parameter which characterizes the range of The analytical solution to the OrnsteinZernike OZ
the attractive forces and is the strength of the interaction equation2 within the mean-spherical approximation MSA is
energy. The potential satisfies the LaplacePoisson differen- an important theory where the hard-core Yukawa potential is
tial equation for distances r , 1 used to provide thermodynamic and structural properties of
fluids. The OZ equation links the total correlation function
2 Y r; 2 Y r; . 2 h(r) to the direct correlation function c(r) according to
This equation appears in theories describing different types
of systems with screened interactions. For example, electro- h r 12 c r 12 s dr 3 c r 12 h r 12 , 3
lyte solutions are governed by the PoissonBoltzmann equa-
tion which reduces to Eq. 2 in the low-concentration re- where s is the number density of the fluid. The MSA clo-
gime, i.e., the DebyeHuckel theory see Refs. 26,27, which sure is described by31
describe extensions to this theory and its application to c r u r; , 4
strong electrolytes. In the DebyeHuckel theory the param-
eter is the inverse Debye length, which is a function of if r , where 1/kT, together with the exact condition
density, temperature, ionic charge, and the dielectric constant for r

0021-9606/99/111(18)/8659/7/$15.00 8659 1999 American Institute of Physics

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
8660 J. Chem. Phys., Vol. 111, No. 18, 8 November 1999 Davies, Gil-Villegas, and Jackson

h r 1. 5 g g HS g 1 . . ., 7
20
Henderson et al. have recently developed a completely where a HS
and g ( ) denote the corresponding hard-
HS

analytical equation of state for the Yukawa fluid using the sphere properties. In the MSA-HTE theory, the terms a n and
MSA solution in Ginozas formalism12,13,17 with a high- g n in the perturbation expansions are analytical functions of
temperature expansion HTE of the free energy and the cor- the packing fraction s 3 /6 and . The mean-attractive
relation function MSA-HTE; in essence this is a Barker and energy a 1 is obtained from31
Henderson perturbation theory32,33 obtained from the MSA
prescriptions. The thermodynamic and structural properties,
obtained in this MSA-HTE theory as explicit functions of
a 1 2 s

g HS r exp r1 rdr, 8
density, temperature, and inverse range of the potential, are
in good agreement with computer simulation results for where s N s /V is the density of a monomer Yukawa fluid.
Yukawa potentials of variable range.25,34 These expressions The corresponding MSA solution for Eq. 8 is
can hence be incorporated into global studies of the phase 0
diagram of Yukawa fluids and to applications which involve a 1 , 9
0
the modeling of real substances.
Within the framework of the statistical associating fluid where
theory SAFT,35,36 a new theory for associating chain mol-
ecules whose segments interact via potentials of variable L
0 , 10
range SAFT-VR has recently been proposed.34,37 This 1 2
2
theory provides a unified method for describing the proper-
ties of systems with different monomermonomer potentials and
such as the square-well, Sutherland, and the hard-core
exp L S
Yukawa. The family of van der Waals interaction potentials 0 , 11
which can be incorporated into the SAFT-VR equation of 3 1 2
state describe Keesom, Debye, and dispersion forces be-
with the coefficients L() and S() defined as
tween both polar and nonpolar molecules.38 The ranges of
such interaction potentials can be temperature dependent as L 12 1 /2 12 12 , 12
in Keesom interactions,38,39 or affected by the incorporation
of a specific model for dispersion interactions in the descrip- S 1 2 3 6 1 2 18 2 12 12 .
tion of phase diagrams.40 In the SAFT-VR approach the in- 13
dividual contributions to the free energy of associating chain Higher-order terms of the free energy in Eq. 6 depend
molecules are given as explicit functions of the range of the on two additional variables, 1 and , which are defined by
attractive potential; the nonconformal properties of these sys-
tems can thus be studied in the context of the phase equilib- 12 1/2
ria of, e.g., alkanes, perfluoroalkanes, and refrigerants.4144 1 , 14
2 1
In this work we consider the specific case of the Yukawa
fluid, and take advantage of the fact that SAFT-VR and MSA a1 1 1
approaches have several common features which can be . 15
combined in order to obtain compact and accurate expres- 20 0
sions to describe the thermodynamics and hence phase be- This last equation follows from the identity (1 1
havior. Furthermore, we improve the accuracy of the MSA- 0 ) 0 1 Ref. 25 and the expression for a 1 , Eq. 9.
HTE equation of state by incorporating a self-consistent The second- and third-order terms in Eq. 6 are given by
expression for the first-order perturbation term of the radial
distribution function obtained from the SAFT-VR 3
approach.34 We also illustrate how the MSA theory can be a 2 16
used to derive a completely analytical SAFT-VR equation of 40
state for chain molecules consisting of segments which inter- and
act via the Yukawa potential.
12 2 13
a 3 , 17
II. THEORY 3 60
A. Monomer properties of Yukawa fluids the subsequent terms have a similar structure
In the MSA-HTE equation of state, the Helmoltz free
n1 P n2
energy per particle aA/NkT and the radial distribution a n , 18
function at contact g( ) are given by the following high- 2n3 2n
0
temperature expansions about a hard-sphere reference
where P n2 is a polynomial of order n2 in .
system:20
The first perturbation term in the expression of the radial
aa HS a 1 2 a 2 . . . , 6 distribution function, Eq. 7, is given by

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 111, No. 18, 8 November 1999 Analytical equation of state 8661

1 a 1 a V
1 g 1; eff ,
DW HS
24
g 1 , 19
0 2
where
and the higher-order terms are given by the general form
aV
1
DW
12 1 2 , 25
n1 Q n2
g n , 20 is the mean-field term van der Waals free energy for a
2n3 2n
0 Yukawa fluid, and g HS(1; eff) is the contact value of the
where Q n2 is a polynomial of order n2 in . These radial distribution function of the hard-sphere fluid obtained
expressions for a n and g n can be rewritten in terms of g 1 by from the CarnahanStarling45 equation, evaluated at an ef-
using Eq. 19: fective packing fraction eff . The quantity eff is a function
of the actual packing fraction of the system
n1 P n2 g n1
a n , 21 eff , c 1 c 2 2 , 26
2n3
where the coefficients c 1 and c 2 are given by
n1 Q n2 g n1
g n . 22
2n3 c 1 0.9006781.50051 1 0.776577 2 , 27
These can be simplified further if we introduce the extra c 2 0.3143000.257101 1 0.0431566 2 . 28
relationship between g 1 and a 1 for a Yukawa fluid, obtained
previously:34 Values for c 1 and c 2 were obtained using a fitting procedure
1 a1 a 1 1 which reproduces the values of a 1 by incorporating a very
g 1 a . 23 accurate radial distribution function of the hard-sphere refer-
4 12 12 1
ence system in the integral of Eq. 8, as is explained in
Substitution of this expression for g 1 into Eqs. 21 and 22 Ref. 34.
results in a simplification of the MSA-HTE theory, since all These coefficients can be obtained more formally as ana-
the higher-order terms of the free energy and the radial dis- lytical functions of the range parameter using the MSA
tribution function are thus directly related to the first-order theory. By expanding both the MSA and SAFT-VR expres-
term a 1 , and its derivatives with respect to and . Further- sions for a 1 Eq. 9 and Eq. 24, respectively in powers of
more, the accuracy of the MSA-HTE theory is improved by , we obtain virial coefficients as predicted by each theory.
the use of this relationship between g 1 and a 1 . Equating these two expansions, using only the first three
In the SAFT-VR formalism the mean-attractive energy virial coefficients, gives a prediction for the coefficients c 1
a 1 of the monomer Yukawa fluid is given by34 and c 2 as

242412 2 4 3 5 4 24 exp 12 2
c 1 , 29
5 3 1

and fitting procedure. More importantly, the procedure outlined


here can be applied to other model potentials, such as the
0 exp 1 exp 2 2
c 2 , 30 square-well or Lennard-Jones interactions, since there are
125 6 1 2 MSA high-temperature expansions for these systems4648
and the SAFT-VR approach can also be applied to soft-core
where
models.49 This complementary use of the MSA and
0 25 8 140 7 364 6 1176 5 672 4 SAFT-VR theories in systems with different interaction po-
tentials provides a route to a purely analytical equation of
2688 3 40322016, 31
state for associating chain molecules, as we will outline in
1 840 6 1008 5 2520 4 672 3 the following section for the Yukawa model. In general, the
advantage of using the SAFT-VR expressions for the mean-
10080 2 120964032, 32 attractive energy a 1 Eq. 24 instead of the MSA prescrip-
tions Eq. 9 is a result of their explicit dependence on the
2 2016 4 8064 3 12096 2 80642016. 33
contact value of the radial distribution function for a hard-
Although these prescriptions for c 1 and c 2 are more al- sphere system g HS( ). Since the mean-attractive energy is
gebraically complex than the expressions initially proposed the basic quantity which is involved in the thermodynamic
in the SAFT-VR approach Eqs. 27and 28, they are theo- properties of associating chain molecules, as is illustrated in
retically derived functions rather than being obtained from a the next section, a dependence on the reference system struc-

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
8662 J. Chem. Phys., Vol. 111, No. 18, 8 November 1999 Davies, Gil-Villegas, and Jackson

ture is favorable. Additionally, since g HS( ) is known for 3. Chain contribution


mixtures, an extension of the SAFT-VR formalism to incor- The contribution to the free energy due to the bonding
porate the description of mixture properties is between the m s monomers forming a chain is given by50
straightforward.37
A
m s 1 ln y M , 37
NkT
B. Associating chain molecules composed of Yukawa where y M( ) is the monomermonomer background corre-
segments lation function evaluated at hard-core contact which is re-
In the SAFT-VR approach molecules are modeled as lated to the radial distribution function g M( ),
chains of m hard-core segments of equal diameter tangen- y M g M exp /kT . 38
tially bonded together. The monomermonomer interactions
are described by the hard-core Yukawa potential, Eq. 1; g ( ) is obtained from the high-temperature expansion
M

additional bonding sites are included to deal with molecular given by Eq. 7. The contact value for the hard-sphere radial
association. The Helmholtz free energy A can be written in distribution function g HS( ) is obtained form the Carnahan
terms of separate contributions as and Starling equation45 and the first-order perturbation term
g 1 is given by Eq. 23. As in the case of the monomer free
A A IDEAL A MONO. A CHAIN A ASSOC. energy, the expression for g 1 obtained within the MSA-HTE
, 34
NkT NkT NkT NkT NkT theory Eq. 19 can also be used.
where N is the total number of molecules, A IDEAL is the ideal
free energy, A MONO. is the residual free energy due to the 4. Association contribution
monomer segments, A CHAIN is the contribution due to chain
formation, and A ASSOC. is the term that describes intermo- The contribution to the free energy due to the association
lecular association. Each of the contributions has been de- of s sites on chain molecules is obtained from the theory of
scribed in detail in Ref. 34; here we briefly describe the Wertheim5156 as:57,58
general expressions for a pure component fluid, and focus on
the new features introduced by the use of the MSA.
A ASSOC.
NkT
s

a1
ln X a
Xa
2
s
,
2
39

1. Ideal contribution
where the sum is over all s sites a on a molecule, and Xa is
the fraction of molecules not bonded at site a. Xa is given by
The free energy of an ideal fluid is given by26
1
A IDEAL X a s , 40
ln 3 1, 35
NkT 1
b1
X b K a,b f a,b g
M

where N/V is the molecular number density and is the


thermal de Broglie wavelength. where K a,b is the volume available for bonding between as-
sociating sites a and b on different molecules, and can be
determined from the parameters of the bonding site such as
2. Monomer contribution its position and range. The other terms appearing in Eq. 40
are the Mayer function f a,b exp(a,b /kT)1 of the ab
In the general case of a fluid of chain molecules, the sitesite bonding interaction a,b , and the monomer
monomer free energy is written in terms of the m s spherical monomer radial distribution function at contact g M( ). As
segments in the chain as
A MONO. AM
m s m s a M, 36 TABLE I. The mean-attractive energy a 1 and the first fluctuation term a 2
NkT N s kT for the Yukawa fluid with 1.8 obtained using the MSA Ref. 6 and
SAFT-VR Ref. 34 approaches for a series of reduced densities s*
where N s is the total number of spherical segments and a M is s 3.
obtained from a high-temperature expansion Eq. 6. In the
original SAFT-VR formalism34 the free energy of the hard- a1 a2
sphere fluid a HS is obtained from the Carnahan and Starling s* MSA SAFT-VR MSA SAFT-VR
expression,45 the mean-attractive energy a 1 is given by Eq.
24, and the first fluctuation term a 2 is obtained from the 0.1 0.566 0.565 0.054 0.067
0.2 1.175 1.172 0.065 0.104
local compressibility approximation.32,33 Alternatively, the
0.3 1.824 1.822 0.057 0.119
MSA-HTE expressions for a 1 Eq. 9 and a 2 Eq. 16 0.4 2.513 2.511 0.044 0.120
could be used in the perturbation expansion. However, a 0.5 3.237 3.237 0.030 0.112
more accurate description of the monomer properties is ob- 0.6 3.995 3.996 0.019 0.099
tained using the general expression for the perturbation terms 0.7 4.784 4.785 0.011 0.083
0.8 5.602 5.599 0.006 0.066
a n in terms of g 1 Eq. 21 and the self-consistent relation
0.9 6.446 6.432 0.003 0.051
between g 1 and a 1 Eq. 23, which results in an expression 1.0 7.314 7.278 0.001 0.036
for the monomer-free energy solely in terms of a 1 .

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 111, No. 18, 8 November 1999 Analytical equation of state 8663

TABLE II. The Helmholtz free energy A/N s kT for the Yukawa fluid with
1.8 obtained using the MSA Ref. 6 and SAFT-VR Ref. 34 ap-
proaches to first- and second-order, for a series of reduced densities s*
s 3 and temperatures T * kT/ .

First-order Second-order

s* T* MSA SAFT-VR MSA SAFT-VR

0.4 2.0 0.137 0.125 0.139 0.155


1.5 0.565 0.549 0.569 0.597
1.0 1.426 1.381 1.443 1.501

0.6 2.0 0.040 0.044 0.039 0.019


1.5 0.630 0.622 0.631 0.666
1.0 1.972 1.954 1.976 2.053

0.8 2.0 0.601 0.603 0.600 0.587


1.5 0.334 0.330 0.334 0.359
1.0 2.205 2.200 2.206 2.262 FIG. 1. The compressibility factor Z PV/N s kT for the Yukawa fluid with
0.7 4.611 4.595 4.164 4.731 1.8 obtained from the SAFT-VR equation of state Ref. 34 curves
compared with values obtained using the MSA approach crosses and from
MC simulation Ref. 20 filled circles. The reduced density is defined s*
s 3 , and the curves are labeled with the corresponding values of the
reduced temperatures T * kT/ .
for the chain contribution, g M( ) can be obtained using ei-
ther the SAFT-VR or MSA-HTE prescriptions for g 1 Eqs.
23or 19, respectively.
Values of the mean-attractive energy a 1 and the first
In summary, the free energy of associating chain mol-
fluctuation term of the free energy a 2 at different densities
ecules composed of Yukawa segments can thus be obtained
obtained with the SAFT-VR approach and the numerical so-
from the SAFT-VR perturbation theory,34 from the MSA-
lution of the MSA are compared in Table I. No significant
HTE theory,20 or by the combination of expressions from
difference is found between the values for a 1 obtained by
both theories.
either method, but the first fluctuation term a 2 from the
SAFT-VR approach is larger for all densities. The same
III. RESULTS trend is observed when the SAFT-VR values for a 2 are com-
pared with the MC data,6 indicating that the local compress-
In this section we compare the predictions for several ibility approximation overestimates the first fluctuation term.
properties of the hard-core Yukawa fluid obtained using dif- The SAFT-VR and MSA-HTE Helmholtz free energies
ferent approaches: the SAFT-VR formalism,34 the numerical are shown in Table II, and the corresponding predictions for
MSA results,6 the MSA-HTE theory,20 and Monte Carlo the compressibility factor Z PV/(N s kT) in Table III and
MC simulation results.6 Since all the properties of associ- Fig. 1. These results illustrate that there is good agreement
ating chain molecules interacting via the Yukawa potential between the two approaches. The negative values of Z at low
can be obtained from those for the monomer fluid, a com-
parison of the monomer properties gives the best indication
of the effectiveness of the different approaches. We also pro-
vide a description of the vaporliquid phase equilibria of
chain molecules composed of Yukawa segments.

TABLE III. The compressibility factor Z PV/(N s kT) for the Yukawa
fluid with 1.8 obtained from MC simulation Ref. 6, and the MSA Ref.
6 and SAFT-VR Ref. 34 approaches for a series of reduced densities
s* s 3 and temperatures T * kT/ .

s* T* MC MSA SAFT-VR

0.4 2.0 1.08 1.119 1.106


1.5 0.69 0.659 0.637
1.0 0.21 0.251 0.298

0.6 2.0 2.04 1.976 1.981


1.5 1.21 1.213 1.223
1.0 0.27 0.303 0.277 FIG. 2. The mean-attractive energy a 1 for the Yukawa fluid as a function of
reduced density s* s 3 for several inverse ranges obtained from the
0.8 2.0 4.27 4.432 4.486 MSA-HTE theory Ref. 20 dashed curve, the SAFT-VR expression evalu-
1.5 3.31 3.330 3.412 ated with the coefficients given by Eqs. 27 and 28 Ref. 34 dotted
1.0 1.29 1.131 1.287 curve, and the SAFT-VR expression evaluated with the MSA-HTE coeffi-
0.7 1.63 1.687 1.401 cients obtained in this work, Eqs. 29 and 30 continuous curve.The
curves are labeled with the values of .

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
8664 J. Chem. Phys., Vol. 111, No. 18, 8 November 1999 Davies, Gil-Villegas, and Jackson

TABLE IV. The contact value of the radial distribution function g( ) for the Yukawa fluid with 1.8
obtained from MC simulation Ref. 6, the MSA numerical solution Ref. 6, the MSA-HTE prediction to first
order Ref. 20, SAFT-VR Ref. 34, and a combination of the MSA-HTE and SAFT-VR approaches for a
series of reduced densities s* s 3 and temperatures T * kT/ .

s* T* MC MSA MSA-HTE SAFT-VR SAFT-VRMSA-HTE

0.4 2.0 2.128 1.963 1.944 2.106 2.124


1.5 2.378 2.040 2.003 2.204 2.228
1.0 2.943 2.222 2.121 2.401 2.436

0.6 2.0 2.921 2.561 2.555 2.821 2.846


1.5 2.966 2.598 2.586 2.891 2.923
1.0 3.205 2.681 2.649 3.029 3.079

0.8 2.0 4.109 3.629 3.628 4.177 4.194


1.5 4.257 3.646 3.643 4.226 4.249
1.0 4.490 3.681 3.674 4.325 4.359
0.7 4.622 3.729 3.713 4.452 4.500

temperatures correspond to metastable states within the and higher-order terms can be included using Eqs. 21 and
vaporliquid coexistence region of the system. 22.
Values of the mean-attractive energy a 1 for different The vaporliquid phase equilibria of chain molecules of
ranges obtained using the MSA-HTE theory, the SAFT-VR length m s 1, 2, 4, and 16 interacting via the Yukawa poten-
approach where the coefficients c 1 and c 2 are given by Eqs. tial obtained using the SAFT-VR equation of state are shown
27 and 28 and the MSA-HTE/SAFT-VR approach in Figs. 3 and 4 for the inverse ranges 1.8 and 4.0, re-
where c 1 and c 2 are obtained analytically from the MSA, spectively. The region of vaporliquid coexistence is seen to
Eqs. 29 and 30, are presented in Fig. 2. Good agreement move to higher temperatures as the chain length increases,
exists between the results obtained from the different ap- and also becomes more extensive as the inverse range of the
proaches. Use of the MSA-HTE coefficients in the potential increases. In the case of the monomer fluid the
SAFT-VR expression has the important advantage over the SAFT-VR prediction of the phase envelope is seen to com-
fitted expressions as they are not restricted to a specific den- pare favorably with the Gibbs ensemble simulation results of
sity range; the only limitation comes from the accuracy of Refs. 59,60 for both values of . No simulation data cur-
the MSA itself. rently exists for chain molecules composed of Yukawa seg-
A combination of both the SAFT-VR and the MSA ap- ments.
proaches can be used to give an expression for the contact
value of the radial distribution function, since the MSA IV. CONCLUSIONS
mean-attractive energy a 1 given by Eq. 9 can be used in the
SAFT-VR expression for the radial distribution function, Eq. In this work we have incorporated certain features of the
23. Predictions for the contact value of the radial distribu- MSA theory into the SAFT-VR approach for chain mol-
tion function are compared in Table IV. The values for
g( ) obtained using the SAFT-VR approach are in better
agreement with the simulation results than those obtained
using the MSA theory both from the numerical solution6
and the MSA-HTE approach20. However, the contact value
of the radial distribution function obtained by substituting
the values of a 1 obtained with the MSA-HTE approach Eq.
9 into the SAFT-VR expression for g 1 Eq. 23 is in
closest agreement with the simulation data.
Hence, the most accurate equation of state for chain mol-
ecules composed of Yukawa segments is found to be one
where a combination of the MSA-HTE and SAFT-VR ap-
proaches is used. The procedure can be summarized as fol-
lows: the hard sphere properties are obtained with the Car-
nahan and Starling expressions45 and the mean-attractive
energy from the SAFT-VR expression Eq. 24, but with the
coefficients c 1 and c 2 obtained from the analytical MSA re- FIG. 3. The vaporliquid coexistence densities for chain molecules of m s
sults Eqs. 29 and 30. The first fluctuation term a 2 is 1, 2, 4, and 16 segments interacting via the Yukawa potential of inverse
better predicted by the MSA relation of Eq. 16, which can range 1.8. The curves are labeled with the values of the chain length m s .
The reduced parameters used are defined as T * kT/ and s* s 3 . The
be rewritten in terms of g 1 as a 2 3 g 21 /. Finally, the
data points correspond to the Gibbs ensemble simulation data for Yukawa
contact value of the radial distribution function is obtained monomers of Lomba and Almarza Ref. 59 squares and of Smit and
from the self-consistent SAFT-VR relation for g 1 Eq. 23 Frenkel Ref. 60 circles.

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
J. Chem. Phys., Vol. 111, No. 18, 8 November 1999 Analytical equation of state 8665

14
G. Giunta, M. C. Abramo, and C. Caccamo, Mol. Phys. 56, 319 1985.
15
M. C. Abramo, C. Caccamo, and G. Giunta, Phys. Rev. A 34, 3279
1986.
16
E. Arrieta, C. Jedrzejek, and K. N. Marsh, J. Chem. Phys. 86, 3607
1987.
17
M. Ginoza, Mol. Phys. 71, 145 1990.
18
C. Rey, L. J. Gallego, and L. E. Gonzalez, J. Chem. Phys. 96, 6984
1992.
19
L. Blum, F. Vericat, and J. N. Herrera, J. Stat. Phys. 66, 249 1992.
20
D. Henderson, L. Blum, and J. P. Noworyta, J. Chem. Phys. 102, 4973
1995.
21
J. N. Herrera, H. Ruiz-Estrada, and L. Blum, J. Chem. Phys. 104, 6327
1996.
22
J. N. Herrera, L. Blum, and E. Garcia-Llanos, J. Chem. Phys. 105, 9288
1996.
23
J. N. Herrera, L. Blum, and P. T. Cummings, Mol. Phys. 93, 73 1998.
24
R. J. F. Leote de Carvalho and R. Evans, Mol. Phys. 92, 211 1997.
25
D-M. Duh and L. Mier-y-Teran, Mol. Phys. 90, 373 1997.
26
L. L. Lee, Molecular Thermodynamics of Nonideal Fluids Butterworth,
FIG. 4. The vaporliquid coexistence densities for chain molecules of m s Washington, DC, 1988.
1, 2, 4, and 16 segments interacting via the Yukawa potential of inverse 27
K. S. Pitzer, J. Phys. Chem. 77, 268 1973.
range 4.0. The curves are labeled with the values of the chain length m s 28
M. Born and E. Wolf, Principles of Optics Pergamon, New York, 1975.
and the reduced parameters used are defined in Fig. 3. The data points 29
A. I. Miller, Early Quantum Electrodynamics Cambridge University
correspond to the Gibbs ensemble simulation data for Yukawa monomers of
Press, Cambridge, England, 1994.
Lomba and Almarza Ref. 59. 30
H. Yukawa, Proc. Phys. Math. Soc. Jpn. 17, 48 1935.
31
J. A. Barker and D. Henderson, Rev. Mod. Phys. 48, 587 1975.
32
J. A. Barker and D. Henderson, J. Chem. Phys. 47, 2856 1967.
ecules composed of Yukawa segments, thus finding addi- 33
J. A. Barker and D. Henderson, J. Chem. Phys. 47, 4714 1967.
34
tional relationships which improve the accuracy of the pre- A. Gil-Villegas, A. Galindo, P. J. Whitehead, S. J. Mills, G. Jackson, and
diction of the thermodynamic properties. This description A. N. Burgess, J. Chem. Phys. 106, 4168 1997.
35
W. G. Chapman, K. E. Gubbins, G. Jackson, and M. Radosz, Fluid Phase
also provides a theoretical basis for the coefficients used in Equilibria 52, 31 1989.
the SAFT-VR expression for the mean-attractive energy a 1 . 36
W. G. Chapman, K. E. Gubbins, G. Jackson, and M. Radosz, Ind. Eng.
From the options available to describe the monomer fluid Chem. Res. 29, 1709 1990.
37
properties, essential to the description of the free energy of A. Galindo, L. A. Davies, A. Gil-Villegas, and G. Jackson, Mol. Phys. 93,
241 1998.
chain molecules, the most accurate approach appears to be a 38
J. N. Israelachvili, Intermolecular and Surface Forces Academic, New
combination of the MSA-HTE and SAFT-VR expressions. York, 1992.
Due to the theoretical basis of all the terms which appear in 39
S. J. Miklavcic, Phys. Rev. E 56, 1142 1997.
the equation of state, such an approach can be used to pro-
40
A. Gil-Villegas, F. Del Rio, and C. Vega, Phys. Rev. E 53, 2326 1996.
41
C. M. McCabe, A. Gil-Villegas, and G. Jackson, J. Phys. Chem. B 102,
vide a detailed description of the phase behavior of noncon-
4183 1998.
formal associating chain molecules. 42
A. Galindo, A. Gil-Villegas, P. J. Whitehead, G. Jackson, and A. N. Bur-
gess, J. Phys. Chem. B 102, 7632 1998.
43
ACKNOWLEDGMENTS C. M. McCabe, A. Galindo, A. Gil-Villegas, and G. Jackson, J. Phys.
Chem. B 102, 8060 1998.
A.G.V. acknowledges CONACYT-Mexico for repatria- 44
C. M. McCabe and G. Jackson, Phys. Chem. Chem. Phys. 1, 2057 1999.
tion Grant No. 970215. We also acknowledge support from
45
N. F. Carnahan and K. E. Starling, J. Chem. Phys. 51, 635 1969.
46
Y. P. Tang and B. C. Y. Lu, J. Chem. Phys. 99, 9828 1993.
the Royal Society, the EPSRC and the European Union. 47
Y. P. Tang and B. C. Y. Lu, J. Chem. Phys. 100, 3079 1994.
48
Y. P. Tang and B. C. Y. Lu, J. Chem. Phys. 100, 6665 1994.
49
1
J. S. Rowlinson, Physica A 156, 15 1989. L. A. Davies, A. Gil-Villegas, and G. Jackson, Int. J. Thermophys. 19, 675
2
E. Waisman, Mol. Phys. 25, 45 1973. 1998.
3
D. Henderson, G. Stell, and E. Waisman, J. Chem. Phys. 62, 4247 1975.
50
W. G. Chapman, J. Chem. Phys. 93, 4299 1990.
4
J. S. Ho ye and G. Stell, Mol. Phys. 32, 195 1976.
51
M. S. Wertheim, J. Stat. Phys. 35, 19 1984.
5
J. S. Ho ye and G. Stell, J. Stat. Phys. 16, 399 1977.
52
M. S. Wertheim, J. Stat. Phys. 35, 35 1984.
6
D, Henderson, E. Waisman, J. L. Lebowitz, and L. Blum, Mol. Phys. 35,
53
M. S. Wertheim, J. Stat. Phys. 42, 459 1986.
241 1978.
54
M. S. Wertheim, J. Stat. Phys. 42, 477 1986.
7
L. Blum and J. S. Ho ye, J. Stat. Phys. 19, 317 1978. 55
M. S. Wertheim, J. Chem. Phys. 85, 2929 1986.
8
P. T. Cummings and E. R. Smith, Chem. Phys. 42, 241 1979. 56
M. S. Wertheim, J. Chem. Phys. 87, 7323 1987.
9
P. T. Cummings and E. R. Smith, Mol. Phys. 38, 997 1979. 57
G. Jackson, W. G. Chapman, and K. E. Gubbins, Mol. Phys. 65, 1 1988.
10
L. Blum, J. Stat. Phys. 22, 661 1980. 58
W. G. Chapman, G. Jackson, and K. E. Gubbins, Mol. Phys. 65, 1057
11
P. T. Cummings and G. Stell, J. Chem. Phys. 78, 1917 1983. 1988.
12
M. Ginoza, J. Phys. Soc. Jpn. 54, 2783 1985. 59
E. Lomba and N. G. Almarza, J. Chem. Phys. 100, 8367 1994.
13
M. Ginoza, J. Phys. Soc. Jpn. 55, 95 1986. 60
B. Smit and D. Frenkel, Mol. Phys. 74, 35 1991.

Downloaded 02 Aug 2003 to 200.78.116.102. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
View publication stats

S-ar putea să vă placă și