You are on page 1of 21

Fluid Phase Equilibria 194–197 (2002) 619–639

Applications of the group-contribution,


lattice-fluid equation of state
Ronald P. Danner∗ , Mourad Hamedi, B.-C. Lee1
Department of Chemical Engineering, Center for the Study of Polymer-Solvent Systems, The Pennsylvania State University,
University Park, PA 16802, USA
Received 20 May 2001; accepted 20 August 2001

Abstract
The group-contribution, lattice-fluid (GCLF) equation of state (EoS) is now more than 10 years old. The primary
objective of our effort has been to develop a predictive model for the phase equilibria in polymer-solvent systems.
The model has been developed by introducing group-contribution to the Panayiotou-Vera EoS. This paper reviews
and updates the application of this approach along with its shortcomings. This EoS has been applied to prediction
of the density of pure polymers, VLE of vapors, supercritical gas solubility, LLE, solubility parameters, and glass
transition temperatures. Homopolymers and random copolymers can be successfully treated. The key features of
this model are that it is purely predictive based only on the molecular structure of the molecules, and it is able to
describe the pressure as well as temperature and composition effects because it is based on an EoS. © 2002 Elsevier
Science B.V. All rights reserved.
Keywords: Equation of state; Data; Vapor–liquid equilibria; Liquid–liquid equilibria; Activity coefficient; Density; Polymer

1. Introduction

Models for treating the phase equilibria of solutions can be classified into two general categories—
van der Waals or lattice-fluid models. Activity coefficient relations or equations of state can be derived
from these approaches. Activity coefficient models are not functions of volume and therefore are not
dependent on the pressure. Equations of state, on the other hand, are functions of volume, and pressure
does influence the results. In either approach, the characteristics of the molecules are described in terms of
two contributions. The first is the combinatorial (athermal, entropic, or free volume) term and depends only
upon the size and shape of the molecules. The second term is the enthalpic (attractive energy, residual,
or potential energy) term which results from the interactions between molecules. Since we wished to

Corresponding author. Tel.: +1-814-863-4814; fax: +1-814-865-7846.
E-mail address: rpd@psu.edu (R.P. Danner).
1
Present address: Department of Chemical Engineering, Hannam University, Taejon 306-791, South Korea.

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 6 9 8 - 7
620 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

develop a model which could be applied at elevated pressures, we elected to use an equation of state
(EoS) approach.
In the lattice model, each molecule is assumed to occupy a number of three-dimensional lattice cells.
The arrangement of the molecules is calculated from the number of statistical possibilities. In the van
der Waals model the volume in which the molecules can translate is determined by the total volume of
the system less the volume occupied by the molecules, i.e. the free volume. In these descriptions, the
molecular interactions are not “turned on” and thus the combinatorial term is obtained. In more advanced
models the concept of local composition is used to account for preferential arrangements rather than
the random arrangements generated from the simpler approach. The most widely used approach is the
quasichemical theory of Guggenheim [1].
The enthalpic contribution for the lattice model is calculated from the product of a characteristic
energy and the number of contacts in the system. Van der Waals models on the other hand use a simple
proportionality with density.
The lattice theories of Flory [2] and Huggins [3–6] have been the foundation of most of the successful
descriptions of polymer solutions. In this activity coefficient approach the combinatorial term was based
on a completely random statistical distribution and the enthalpic term was expressed as a constant (the
well-known interaction parameter, chi) times the square of the solvent volume fraction. Sanchez and
Lacombe [7] and Lacombe and Sanchez [8] proposed a lattice-fluid EoS which incorporated holes in the
lattice and assumes a random mixing of the molecules in the system. The presence of holes in the lattice
provides the key to the density variation with pressure. As the fraction of holes is changed the density
changes proportionately.
Panayiotou and Vera [9,10] developed an EoS also based on lattice theory, but eliminated the assumption
of random distribution of the molecules in the case of mixtures. They introduced a non-random term
which leads to local composition effects for the molecules. The holes in the lattice, on the other hand,
are randomly distributed. This equation has been shown to be quite successful in correlating phase
equilibria in polymer-solvent systems and is versatile in theory. Thus, it was selected as the basis for the
development of the group-contribution, lattice-fluid (GCLF) EoS (GCLF-EoS). The objective, of course,
was to develop a model which is purely predictive, requiring only the structure of the molecules in the
system. This approach will necessarily lead to some significant errors, particularly as the systems become
more complex as found in hydrogen bonding or ionic systems. It can, however, provide a powerful tool
for the screening properties for the innumerable possible combinations of solvents and polymers, if it is
reasonably accurate and shows the correct trends. In this respect the GCLF-EoS has been a success.
There are, of course, other models available for the prediction of the thermodynamic properties of
polymer-solvent systems. Among these are the UNIFAC-free volume model [11], the revised group-
contribution Flory EoS [12] and the entropic free volume model [13]. Lee and Danner [14] made an
extensive comparison of these models in the case of infinitely dilute solvent in a polymer and found the
GCLF-EoS and the UNFAC-FV model to be the best. The UNIFAC-free volume model is more versatile
in that it was built upon the previous extensive UNIFAC work for small molecules and thus, has more
groups available for it. On the other hand, it requires accurate density data for the components and cannot
predict pressure effects.
The original GCLF-EoS was developed by High and Danner [15]. This initial effort did not include a
binary interaction parameter. Nevertheless, it was successful in treating many systems that are not strongly
polar. Lee and Danner [16] extended the model to include a molecular binary interaction parameter which
is obtained from group binary interaction values. This preserved the predictive nature of the model. Of
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 621

course, if one has data for a system, the parameters can be fitted to these data and the model can be used
as a correlation with improved accuracy for extrapolation to different conditions of pressure, temperature,
and composition.

2. The GCLF-EoS

The general form of the canonical partition function as used by Panayiotou and Vera [9,10] for
N1 , N2 , . . . , Np semi-flexible molecules having a configurational energy E, was expressed as given
below.
p    
 ξi Ni E
Q= gc gnr exp − (1)
i=1
σi RT

Here gc is the combinatorial term assuming a random distribution of the molecules, gnr is the non-random
correction for the combinatorial term, and ξ i /σ i accounts for the flexibility and the symmetry of a
molecule. The lattice-fluid theory as used in this model approximates the molecule by its equivalent—r
identical segments occupying consecutive lattice sites. The symmetry parameter is thus equal to two.
The flexibility parameter accounts for the number of internal configurations available to a semi-flexible
molecule i occupying ri sites. Each lattice site is assumed to have a constant volume (vh = 9.75 ×
10−6 m3 /mol) and a fixed coordination number (z = 10). The lattice cells may be occupied (by segments
or molecules) or vacant (holes).
Based on this partition function the reduced form of the EoS is
   
P̃ ṽ z ṽ + (q/r) − 1 θ2
= ln + ln − (2)
T̃ ṽ − 1 2 ṽ T̃
P 2Pvh T 2RT v vh (Nh + rN)
P̃ = ∗
= , T̃ = = ∗, ṽ = = (3)
P zε∗ T ∗ zε v ∗ v∗
here ε∗ and v ∗ are the molecular interaction energy and reference volume of a binary mixture, respectively:
ε∗ = θ̄1 ε11 + θ̄2 ε22 − θ̄1 θ̄2 Γ˙12 ε, ε = ε11 + ε22 − 2ε12 (4)

v∗ = xi vi∗ (5)

εii is the molecular interaction energy between like molecules of type i, ε12 is the cross-interaction energy
between unlike molecules 1 and 2, and vi∗ is the molecular reference volume of a pure component i. The
other parameters in Eq. (1) are calculated from the following simple combining rules
  
r= xi ri , q= xi qi , θ= θi (6)

vi∗
ri = , zqi = (z − 2)ri + 2 (7)
vh
zqi Ni qi Ni
θi =  = (8)
z(Nh + j qj Nj ) Nh + qN
622 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

here ri is the number of lattice sites occupied by a molecule i, zqi is the number of contact sites by the
molecule i, θ i is the molecular surface fraction of component i including holes, and Nh is the number of
holes. ε12 in Eq. (4) is calculated from the following mixing rule
ε12 = (ε11 ε12 )1/2 (1 − k12 ) (9)
where k12 is the binary interaction parameter. In Eq. (4), θ̄i is the molecular surface fraction of com-
ponent i on a hole-free basis, and Γ˙12 is the non-randomness parameter between unlike molecules 1
and 2:
zqi Ni qi Ni xi qi
θ̄i =  = = (10)
z j qj Nj qN q

θ̄1 Γ˙11 + θ̄2 Γ˙12 = θ̄2 Γ˙22 + θ̄1 Γ˙12 = 1 (11)


The quasichemical approach gives the following relationship among the non-randomness parameters
[17].
 
Γ˙11 Γ˙22 ε
= exp θ (12)
Γ˙122 RT
The weight fraction activity coefficient (WFAC) of component i in the mixture is expressed as [16]
   
ai ṽi ṽ ṽi − 1 2θi,P − θ θ zqi
ln Ωi = ln = ln ϕi − ln wi + ln + qi ln + qi − + ln Γ˙i
wi ṽ ṽ − 1 ṽi T̃i T̃ 2
(13)
xi v ∗ xi ri
ϕi =  i ∗ =  (14)
j xj vj j xj rj

here the subscript i in T̃i and ṽi represents a pure component i, wi and ϕ i are the weight and volume
fractions of component i in the mixture, respectively, and θ i ,P is the surface area fraction of the pure
component i at the same temperature and pressure as the mixture.
For a pure component the EoS contains two adjustable parameters: molecular interaction energy, εii ;
and molecular reference volume, vi∗ . For a binary mixture, it has one more parameter the binary interaction
parameter, k12 . To make the model predictive, the following group-contribution combining rules for these
three parameters have been developed
 (i)
εii = Θk Θm(i) (ekk emm )1/2 (15)
k m
 (i)
vi∗ = nk Rk (16)
k

k12 = Θm(M) Θn(M) αmn (MR-1) (17)
m n

here ekk is the group interaction energy between like groups k, Rk is the group reference volume, and α mn
is the group binary interaction parameter between groups m and n. Θk(i) and Θk(M) are the surface area
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 623

fraction of group k in the pure component i and in the mixture, respectively:


 (i)
n(i) Qk i nk Qk
Θk(i) =  k (i) , Θk(M) =   (i)
(18)
p np Qp p i np Qp

here n(i)
k is the number of group k in component i and Qk is the surface area of group k, as used in the
UNIFAC method [18].
The characteristic parameters are also functions of temperature. A quadratic form with respect to
temperature was found to be adequate:
   2     2 
T T 1 T T
ekk = e0,k + e1,k + e2,k , Rk = 3 R0,k + R1,k + R2,k (19)
T0 T0 10 T0 T0

here ei ,k and Ri ,k are constants, T is the system temperature in Kelvins and T0 was arbitrarily set to
273.15 K.
The group parameters (ekk , Rk , α mn ) were estimated from properties of low molecular weight com-
pounds only. First of all, the molecular parameters, ε ii and vi∗ , at various temperatures for each pure
compound were simultaneously determined by fitting the experimental saturated vapor pressure and
liquid density data [19] to the EOS. Fig. 1 shows an example of the results of these regressions for a
series of aromatic compounds. The binary interaction parameters, k12 , were estimated from the VLE
data [20–25] of low molecular weight binary mixtures. Secondly, the group parameters (ekk , Rk , α mn ) at
each temperature were independently estimated from the molecular parameters (εii , vi∗ , k12 ) by non-linear
regression using the mixing rules, Eqs. (15)–(17). Finally, the parameter constants (ei ,k and Ri ,k ) were
determined by fitting the group parameters obtained at the various temperatures to Eq. (19) assuming a
quadratic dependency on temperature. Tables of the group parameters for a variety of functional groups
are available from Lee and Danner [16]. New parameters have been developed to deal primarily with
refrigerants. Values of these parameters are given in Tables 1–3.
Once the molecular structures of the components in a mixture are known with respect to their functional
groups, one can calculate the molecular parameters from the group parameters and then predict the
equilibrium behavior of polymer solutions as well as of low molecular weight compounds.

Fig. 1. Prediction of (a) vapor pressures and (b) saturated liquid densities of aromatic hydrocarbons by the GCLF-EoS (data:
[19]).
624 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

Table 1
Group interaction energy and reference volume parameters for refrigerants
Gas ek ,0 ek ,1 ek ,2 Rk ,0 Rk ,1 Rk ,2 Qk

CH2 ClF 979.749 117.708 −132.947 67.674 −39.502 22.715 1.728


CHCl2 F 927.501 −33.834 −8.92 68.212 −8.449 6.162 2.116
CCl3 F 874.072 −79.185 39.437 74.285 6.401 −0.484 2.644
CHClF2 829.428 0.437 −63.43 66.565 −27.265 18.156 1.828
CCl2 F2 690.846 49.712 −26.581 82.37 −23.72 13.848 2.376
CClF3 608.824 18.688 −46.243 68.057 −21.846 18.559 2.1
CH3 F 1075.386 −273.648 −55.579 45.836 −31.933 27.59 1.308
CH2 F2 1333.228 −588.76 101.413 41.316 −13.455 15.552 1.46
CHF3 961.509 −373.243 51.028 48.026 −17.708 20.101 1.608
CF4 507.696 53.27 −144.68 59.106 −35.376 37.128 1.84

Table 2
Group binary interaction parameters, α mn , for refrigerant and organic groups
Main groups CH2 aCH aCCH2 cyCH2 C=C CH2 O CH2 CO COO OH CH3 OH
CHClF2 0.1672 −0.0438 – 0.1530 0.0660 −0.4091 −0.0670 −0.2850 −0.2791 0.0109
CCl2 F2 0.0474 – – – 0.0087 – – – –
CClF3 0.0159 – – – 0.1213 – – – – –
CH2 F2 −0.0366 – – – 0.3973 2.0342 – – – –
CF4 0.2017 −1.0955 −0.1114 0.0882 −0.0825 −1.9318 −0.0779 −0.3312 0.2962 0.0127

Table 3
Molecular interaction parameter, kij , between refrigerants

CHClF2 CCl3 F CCl2 F2 CClF3 CH2 F2 CHF3 CF4


CHClF2 0.0000 0.0693 0.0962 0.0054 0.0141 0.0271 0.1104
CCl3 F 0.0000 0.0166 0.0204 – – −0.0846
CCl2 F2 0.0000 0.0250 0.1840 0.1797 0.2163
CClF3 0.0000 0.1660 0.1401 0.0024
CH2 F2 0.0000 −0.0288 −0.0829
CHF3 0.0000 0.1087
CF4 0.0000

3. Applications of the GCLF-EoS

Based on the above model, extensive applications of the GCLF-EoS have been developed over the
past decade. The results of these efforts have been quite positive and have provided useful tools for the
analysis of many types of systems and prediction of a number of polymer properties.

3.1. Density of pure polymers

The initial work on predicting polymer PVT behavior from the GCLF-EoS was reported by Parekh
and Danner [26]. Additional examples were reported by Lee [27] using the enhanced form of the model.
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 625

Fig. 2. GCLF-EoS prediction of specific volumes of (a) poly(dimethyl siloxane) and (b) poly(methyl methacrylate),
poly(oxymethylene), and poly(vinyl chloride).

In Fig. 2 examples of the prediction of the specific volume of a number of homopolymers are shown.
In Fig. 3 the density of a random copolymer of 21 wt.% styrene with ethylene is shown. The model
does a good job of predicting the PVT of these polymers up to high pressures. At these conditions of
temperature and pressure, the copolymer is amorphous. At lower temperatures, however, the copolymer
is about 26.1 wt.% crystalline. The model does not take into account crystallinity and significant errors
are to be expected for semi-crystalline polymers.

3.2. Low molecular weight VLE

Although the primary objective of the GCLF-EoS is to treat polymer–solvent systems, it cannot be
expected to be successful unless it can adequately describe the behavior of low molecular weight systems.
The group parameters were derived based on the density and vapor pressure behavior of low molecular
weight compounds and the interaction group values were obtained from fitting low molecular weight

Fig. 3. GCLF-EoS prediction of the density of the random copolymer of 79 wt.% ethylene −21 wt.% styrene (data: [31]).
626 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

Fig. 4. GCLF-EoS prediction of phase equilibria of alcohols with R22 (data: [38]).

binary VLE. Thus, it is not surprising that the model does a good job of fitting low molecular weight
VLE data. Examples of this have been shown by Lee and Danner [16] and Lee [27].
More recently new group values have been developed to treat fluorine-type refrigerants. In Fig. 4 the
predictions of the equilibrium between R22 (CHClF2 ) and ethanol or propanol are shown. Fig. 5 shows
the azeotropic system R13 (CClF3 )-R23 (CHF3 ). The kij values in Table 3 can be particularly useful in
the foam industry to study mixtures of blowing agents. While the quantitative results are not exceptional,
the model is successful in predicting the qualitative behavior of these systems.

3.3. Polymer–vapor equilibria

The most extensive application of the GCLF-EoS has been for the prediction of the equilibrium between
a polymer and a solvent which is well below the critical region. For applications such as devolatilization
and drying the phase equilibria in the region of infinitely dilute solvent is of primary importance. Thus,
considerable evaluation of the model has been done using the infinitely dilute, WFAC. Lee and Danner
[14] have presented comparisons of the GCLF-EoS, the UNIFAC-FV [11], the entropic free volume [13],
and the revised GC-Flory EoS [12] for this application using 342 systems. They found that the GCLF-EoS

Fig. 5. GCLF-EoS prediction of phase equilibria of R13 and R23 (data: [39]).
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 627

Fig. 6. Prediction of infinite dilution weight fraction activity coefficients of methyl ethyl ketone in poly(ethylene oxide).

gave the best predictions for the systems with non-polar and weakly polar solvents. For the systems with
strongly polar solvents, the modified GCLF-EOS under-predicted experimental values, but was as accurate
as or better than the other models. An example of this application is shown in Fig. 6 for the GCLF-EoS
and UNIFAC-FV. These results illustrate that the data in the literature for the infinitely dilute WFACs are
generally quite scattered. In fact, the predictions are typically more reliable than a lot of experimental
data. In the case shown one cannot know which model is preferable. Finite concentration measurements
are more accurate, particularly those collected using gravimetric sorption. There are numerous data of
this type in the literature that can be used as a basis to discriminate between models.
Finite solvent concentration is the important region for such things as foam production and film and coat-
ing application. Fig. 7 demonstrates some typical performances of the GCLF-EoS and the UNIFAC-FV
model in this region. With any group-contribution model there are likely to be some systems where it
does significantly worse than average. Fig. 8a shows such a case for the UNIFAC-FV model. In the case

Fig. 7. Prediction of weight fraction activity coefficients of (a) n-propyl acetate in polystyrene (data: [40]) and (b) benzene in
polyisobutylene (data: [41]).
628 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

Fig. 8. Prediction of weight fraction activity coefficients of (a) toluene in poly(methyl methacrylate) (data: [42]) and (b) benzene
in isoprene rubber (data: [43]).

of the poly(methyl methacrylate) PMMA/toluene system, Oishi and Prausnitz [11] treated PMMA as if it
contained the CH3 COO group instead of the COOCH3 group, in contradiction to the UNIFAC definition
as discussed by Macedo et al. [28]. While this change significantly improves the UNIFAC-FV prediction,
there appears to be no justification for the use of the CH3 COO group in the definition of PMMA. In Fig. 8b
a failing of the GCLF-EoS in the case of the isoprene rubber/benzene system is shown. This is a case
where there is significant steric hindrance. The group definitions are not sufficiently specific to account
for this type of effect. UNIFAC-FV also fails for water-containing systems. This is because the molar
volume of water is so low that the free volume contribution leads to a logarithm of a negative number.
Lee and Danner [29] demonstrated the ability of the GCLF-EoS to predict the change in the finite
concentration WFAC as a function of the composition of random copolymers. Fig. 9 shows an application

Fig. 9. GCLF-EoS prediction of toluene solubility in a non-crystalline, ethylene–styrene interpolymer containing 75 wt.% styrene
(data: [31]).
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 629

Fig. 10. GCLF-EoS prediction of toluene solubility in a crystalline, ethylene–styrene interpolymer containing 21 wt.% styrene
(data: [31]).

to another copolymer the non-crystalline, ethylene–styrene interpolymer (ESI) containing 21 wt.% styrene
at 150 ◦ C. The predictions are excellent.
At 50 ◦ C the copolymer forms crystals. The crystals do not absorb solvent [30] and the solvent weight
fraction should be expressed based only on the amorphous phase. Fig. 10 shows the solubility of toluene at
50 ◦ C where the ESI has about 26.1 wt.% crystals. The GCLF-EoS model over predicts the experimental
weight fraction by about 22% even after this correction is made. From studies with other ESI materials
Hamedi et al. [31] showed that the error in the GCLF-EoS predictions varied as the percent crystallinity
changed. The presence of crystals reduces the solubility of the solvent even in the amorphous portion, prob-
ably due to the restricted mobility of the molecules, some segments of which are anchored in the crystals.

3.4. Polymer–gas equilibria

A primary advantage of an EoS approach over an activity coefficient approach is that it can be applied at
high pressures. This is important in applications such as high pressure polymer extrusion and supercritical
extraction of monomers, solvents, and contaminants from polymers. The mixing rule for the binary
interaction parameter in Eq. (17) is expressed in such a way that the interaction parameter, kij , is obtained
as a correction for the interaction of molecules of the same and different species. This form is referred to
as mixing rule MR-1 hereafter. This approach has been quite successful for treating components below
their critical conditions (vapors) but has been found to give significant errors for supercritical fluids.
Hamedi et al. [32] proposed an alternative formulation in which the interaction parameter is calculated
only as a contribution of groups from different species. In this case, the interaction parameter is given by
the following mixing rule (MR-2):

k12 = Θm(1) Θm(2) αmn (MR-2) (20)
m n

where Θk(1) and Θk(2) are expressed as in Eq. (18) and represent the surface area fractions of group
k in species 1 and 2. Table 4 lists the pure component parameters for carbon dioxide and ethylene.
630 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

Table 4
Group interaction energy and reference volume parameters for carbon dioxide and ethylene
Gas ek ,0 ek ,1 ek ,2 Rk ,0 Rk ,1 Rk ,2 Qk Range, T

Below Tc
CO2 995.246 −94.447 −183.601 59.742 −66.458 44.614 1.112 220–300
C2 H 4 834.377 −453.702 204.473 38.370 11.665 0.368 1.488 110–280
Above Tc
CO2 570.152 149.875 −49.995 8.808 33.2832 −9.353 1.112 310–430
C 2 H4 500.325 80.274 1.465 50.658 −6.314 3.942 1.488 290–500

Table 5
Group binary interaction parameters, α ij , for carbon dioxide and ethylene

Mixing rule Main groups CH2 aCH aCCH2 cyCH2 C=C CH2 O COO CH3 OH
MR-1 CO2 0.1262 0.1130 0.0027 0.0089 0.0749 −0.1018 −0.1850 0.0486
C 2 H4 0.0907 0.0029 – – – – –
MR-2 CO2 0.2994 0.2483 0.0278 0.0219 0.1374 −0.0786 −0.4739 0.0974
C 2 H4 0.0380 0.0014 – – – – – –

Parameters below the critical temperature were obtained from density and vapor pressure data. Above the
critical temperature, P–V–T data were used. Different parameters were used above and below the criti-
cal temperature in order to improve the correlations. These values are applicable to either mixing rule.
Table 5 gives the group interaction parameters between these two gases and some groups commonly
found in bulk polymers. In this case there are two sets—one to be used with MR-1 and one to be used
with MR-2. The GCLF-EoS with the two different mixing rules was used to predict the solubility of
carbon dioxide in polypropylene and a copolymer (85 wt.% ethylene—15 wt.% vinyl acetate) as shown
in Fig. 11. Clearly the new mixing rule (MR-2) is superior in these cases. In Fig. 12 the prediction

Fig. 11. GCLF-EoS predictions (using the two different mixing rules) of carbon dioxide solubility in polypropylene and a
copolymer of 85 wt.% ethylene and 15 wt.% vinyl acetate) (data: [44]).
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 631

Fig. 12. GCLF-EoS predictions (using the two different mixing rules) of ethylene solubility in polypropylene (data: [45]).

of the solubility of ethylene in polypropylene using MR-2 is shown. The results are quite good for
both cases.

3.5. Liquid–liquid equilibria

The GCLF-EoS has been applied with surprising success in the liquid–liquid regime of polymer–solvent
systems [33]. This was done using the same group values for the hard-core volume, the interaction energy,
and the binary interaction parameter. Fig. 13 shows predictions of the spinodal and binodal curves for
the poly(ethylene glycol)/tert-butyl acetate. The predicted spinodal and binodal curves are shown along
with the experimental data. The results are quite good in both cases.
In Fig. 14 an example of the prediction of polymer–polymer miscibility is shown for the poly(n-butyl
acrylate)-poly(vinyl chloride) system. In polymer–polymer systems the phase behavior is extremely
sensitive to the value of the binary interaction parameter. Thus, the kij was regressed from the data. This
and the binary refrigerant correlation (Fig. 5) are the only cases shown in this paper which are correlative
and not purely predictive.

Fig. 13. GCLF-EoS prediction of spinodal and binodal curves for poly(ethylene glycol) in tert-butyl acetate.
632 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

Fig. 14. Correlation of poly(n-butyl acrylate)-poly(vinly chloride) miscibility using the GCLF-EoS (data: [46]).

3.6. Polymer glass transition temperature

Hamedi et al. [34] has shown that the glass transition temperature, Tg , of polymers and polymer solutions
can be predicted through a combination of the GCLF EoS and the Gibbs and DiMarzio [35] criterion. The
criterion of the glass transition temperature as defined by Gibbs and DiMarzio is a value of zero for the
system entropy, which is function only of the number of degeneracies. The system entropy is expressed as.
 
δi Ni
S = k ln gc (21)
σi
Thus, we see that the Tg depends upon the symmetry and flexibility of the molecules while the phase
equilibria do not. The ri -mer bonds are assumed to have two energetic states. The flexibility parameter
is expressed as a function of fi , the fraction of ri -mer bonds in the higher energy state:
 (ri −2)fi  (ri −2)(1−fi )
2 1
δi = 4 (22)
fi 1 − fi
The equilibrium number of flexed bonds, fi , is obtained by finding the minimum of the canonical
partition function with respect to fi :
2 exp(−εi /kT)
fi = (23)
1 + 2 exp(−εi /kT)
For the case of a macromolecule where r approaches infinity, the final expression for the pure component
entropy is [34]
       
S ṽi zṽi ṽi − 1 + qi /ri 1 qi
= (ṽi − 1)ln + − 1 ln + ln ṽi − 1 +
kri Ni ṽi − 1 2 ṽi ri ri
εi
+fi − ln(1 − fi ) (24)
kT
here ṽi is the reduced specific volume calculated from the EoS for pure component i. By setting this equa-
tion equal to zero and using the GCLF-EoS parameters, estimates of the glass transition temperature are
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 633

Fig. 15. GCLF-EoS prediction of the glass transition temperature of poly(methyl methacrylate) (data: [47]) and poly(vinyl
acetate) (data: [48]).

obtained. For binaries, the non-randomness term must be taken into account and the equations become a bit
more complicated. For the case of alternating copolymers, Hamedi et al. used the concept of dyads which
account for the flex energy contribution from the number of adjacent monomer pairs of different type.
Fig. 15 compares the experimental Tg of PMMA and poly(vinyl acetate) with the predictions from the
GCLF-EoS up to relatively high pressures. The predictions are quite good. The Tg of the styrene–methyl
methacrylate copolymer is treated in Fig. 16 as a function of the styrene content. Once again the GCLF-EoS
does a good job of characterizing the Tg behavior. For polymer–solvent solutions, the model predictions
are semi-quantitative depending on the system. The interaction parameter required for binary systems
was found to have little effect on the glass transition temperature predictions.

3.7. Solubility parameter prediction

Solubility parameters are widely entrenched in the polymer industry at least as a qualitative indication
of the compatibility of solvents and polymers. Thus, a method of predicting these values is of considerable
practical value. The solubility parameter, δ, is equal to the square root of the cohesive energy density, i.e.

Fig. 16. GCLF-EoS prediction of the glass transition temperature as a function of the composition of styrene–methyl methacrylate
copolymers (data: [49]).
634 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

Table 6
Average errors in the GCLF-EoS predictions of the solubility parametersa
Family No. of compounds Error (%)

Alkanes 16 5.9
Alkenes 15 5.4
Ketones 9 2.8
Acetates 7 3.2
Alcohols 15 8.2
Ethers 8 4.6
Cyclic and aromatic 13 3.5
Chlorinated 14 1.7
a
Compared to values from Daubert et al. [19].

the square root of cohesive energy per unit of volume V. The cohesive energy is explicitly expressed by
the energy of the lattice, E.

−E
δ= (25)
V
In the lattice model, the cohesive energy can be calculated from product of the number of external
energetic contacts and the energies in the different interactions. For a binary mixture, the energy of the
lattice is expressed as
−E = N11 ε11 + N12 ε12 + N22 ε22 (26)
θzqN
−E = (θ̄1 θ̄1 Γ11 ε11 + θ̄i θ̄2 Γ12 ε12 + 2θ̄2 θ̄2 Γ22 ε22 ) (27)
2
here Nij and εij are the number of external energetic interactions and the potential energy of interaction
between species i and j, respectively. The other parameters of Eq. (27) have been given in Eqs. (6), (9)
and (10). Using Eqs. (4) and (10).
zqN ∗
−E = θε (28)
2
substituting into the latter equation, the solubility parameter expression becomes [37].
 ∗ 1/2
qP θ
δ= (29)
r ṽ
The same equation is obtained for pure components.
The solubility parameters of 97 pure solvents were calculated using the GCLF-EoS using only the
chemical structure as an input and compared against the values reported by Daubert et al. [19]. The
results are reported in Table 6. The model is found to give satisfactory results for polar, non-polar,
and slightly hydrogen bonding systems. A systematic over-prediction of approximately 5–6% is found
for alkanes and alkenes. This information could be used for practical engineering purposes. Molecular
structures with steric exclusion, however, were found to cause strong discrepancies in some cases.
The predictions for solvents involving strong hydrogen bonding is poor, particularly for short chains.
As the molecular weight increases a better agreement is obtained due to the decrease of the hydrogen
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 635

Table 7
Comparison of solubility parameters predicted from GCLF-EoS with the range of values reported by Barton [36]
Polymer δ Range [36] [MPa1/2 ] GCLF-EoS prediction

Poly(vinyl chloride) 19.22–22.09 20.52


Polyisobutylene 15.95–16.57 16.95
Poly(methyl methacrylate) 18.61–26.18 19.39
Polystyrene 17.39–19.02 20.48
Poly (vinyl acetate) 19.1–22.6 20.41
Poly(dimethyl siloxane) 14.93–15.54 14.31
Polybutadiene 16.57–17.59 15.09
Polyethylene 15.75–17.8 17.97

bonding contribution to the overall interactions. The GCLF-EOS consistently under estimates the sol-
ubility parameter of alcohols. For practical purposes, a good estimate is obtained by adding 8% to the
predicted value.
The primary usefulness of the GCLF-EoS in terms of solubility parameters lies in its application to
polymers and copolymers. The model is a good alternative to the different experimental methods that
give large variations in the solubility parameters. In Table 7 a comparison of the solubility parameters
predicted from the GCLF-EoS and the range of values collected from the literature by Barton [36] is
given. The predicted values fall right in the expected range or just outside it. Again larger discrepancies
should be expected for polymers involving hydrogen bonding. Moreover, the model does not take into
account any crystallinity which may introduce an additional source of error.
Liquid mixtures can be treated easily by the GCLF-EoS model. The mixture affects are taken into
account by the mixing rule in Eq. (8). Predictions were compared to the 12 mixtures reported by Barton
[36] and an average error of only 1.3% was found. In this case the use of the interaction parameter k12 is
not crucial. A zero value increased the average error to only 1.7%.
Since supercritical carbon dioxide is a poor solvent for polymers, a cosolvent can be added to enhance
the solvent power. Fig. 17 shows the predicted variation of the solubility parameter of propanol/carbon
dioxide mixtures as a function of pressure. The propanol significantly increases the solubility parameter

Fig. 17. GCLF-EoS prediction of the change in the solubility parameter of mixtures of propanol and carbon dioxide as a function
of pressure.
636 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

towards the value typical of a polymer, which in practice allows lower operating pressures and higher
solubilities. The solubility parameters obtained for pure carbon dioxide are typically below that of all
polymers except PTFE (teflon). This agrees with the observed high solubility of carbon dioxide in PTFE. A
remedy for the incomparability of carbon dixoide with other polymers is the addition of such a cosolvent.

4. Conclusion

The GCLF-EoS has been successfully applied to the prediction of numerous forms of phase equilibria
as well as to prediction of solubility parameters and glass transition temperatures. The primary advantages
of this model are that it is purely predictive using only the molecular structure, and it can be used for
high pressure applications. Thus, it is unlike most other available models. The equation which is the
foundation of the model, the Panayiotou-Vera EoS, is soundly derived from lattice-fluid theory and has
proven to be a reliable tool for correlation of polymer–solvent properties. The group-contribution form of
this equation necessarily gives up some accuracy, but gains the significant advantage of being applicable
to a wide range of systems even if no data are available. Thus, the GCLF-EoS is a valuable tool for
screening potential systems and for determining relative behavior of components.

List of symbols
ai activity of component i in the mixture
ekk group interaction energy parameter between like groups k, J/mol
e0,k , e1,k , e2,k constants in group interaction energy parameter
E total lattice energy in the lattice-fluid theory, J/kmol
fi fraction of ri -mer bonds in the higher flex energy state
gc random (athermal) combinatorial factor
gnr non-random combinatorial factor
k12 binary interaction parameter (molecular basis)
MW molecular weight (g/mole)
n(i)
k number of groups of type k in molecule i
N total number of molecules in the mixture; number of data points
Nh total number of holes in the lattice
Ni number of molecules of type i
Nij total number of interactions (external contacts) between molecules i and j in the
non-random distribution of molecules
P pressure, Pa
Pi∗ , P ∗ characteristic pressures of pure component i and mixture, respectively
P̃1 , P̃ reduced pressures of pure component 1 and mixture, respectively
qi surface area parameter of component i
Q canonical partition function
Qk group surface area parameter for group k
r volume parameter for component i; number of lattice sites occupied by the segments
of a molecule i
R gas constant = 8.314 m3 Pa/mol K
Rk group reference volume parameter for group k, m3 /kmol
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 637

R0,k , R1,k , R2,k constants in group reference volume parameter


S entropy, J/K
T temperature, K
T0 reference temperature, K
Tg glass transition temperature, K
Ti∗ , T ∗ characteristic temperatures of pure component i and mixture, respectively
T̃i , T̃ reduced temperatures of pure component 1 and mixture, respectively
vh molar volume of lattice site (= 9.75 × 10−3 m3 /kmol)
vi molar volume of component i, m3 /kmol
vi∗ molecular reference (hard core) volume of component i, m3 /kmol
ṽ, ṽm reduced volume of the mixture
ṽi reduced volume of component i
wi weight fraction of component i
xi mole fraction of component i
z coordination number

Greek Symbols
α mn group binary interaction parameter between main groups m and n
Γ˙ij non-randomness factor for molecules of type i around molecules of type j
δ solubility parameter (MPa)1/2
ξi flexibility parameter of component i
ε∗ characteristic interaction energy of the mixture, J/kmol
ε ii characteristic interaction energy between like molecules i, J/kmol
ε ij characteristic interaction energy between unlike molecules i and j, J/kmol
ε interaction energy change, J/kmol
θi molecular surface area fraction of component i, or fraction of contacts which involve
molecules of type i
θi molecular surface area fraction of component i on a hole-free basis, or fraction
of contacts which involve molecules of type i on a hole-free basis
θ i ,P molecular surface area fraction of the pure component i at the same temperature and
pressure as the mixture
Θk(i) surface area fraction of group k in component i
Θk(M) surface area fraction of group k in the mixture
σi symmetry parameter of component i
φi volume fraction of component i
Ωi weight fraction activity coefficient (WFAC) of component i

Acknowledgements

This work has been sponsored over the years by the Design Institute of Physical Property Data of the
American Institute of Chemical Engineers and by the industrial sponsors of the Center for the Study of
Polymer–Solvent Systems. Their support and encouragement is gratefully acknowledged.
638 R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639

References

[1] E.A. Guggenheim, Mixtures, Oxford University Press, London, 1952.


[2] P.J. Flory, J. Chem. Phys. 9 (1941) 660.
[3] M.L. Huggins, J. Chem. Phys. 9 (1941) 440.
[4] M.L. Huggins, J. Phys. Chem. 46 (1942) 151.
[5] M.L. Huggins, J. Am. Chem. Soc. 64 (1942) 1712.
[6] M.L. Huggins, J. Ann. N. Y. Acad. Sci. 43 (1942) 1.
[7] I.C. Sanchez, R.H. Lacombe, J. Phys. Chem. 80 (1976) 2352.
[8] R.H. Lacombe, I.C. Sanchez, J. Phys. Chem. 80 (1976) 2568.
[9] C. Panayiotou, J.H. Vera, Polym. Eng. Sci. 22 (1982) 345.
[10] C. Panayiotou, J.H. Vera, Polym. J. 14 (1982) 681.
[11] T. Oishi, J.M. Prausnitz, Ind. Eng. Chem. Res. Dev. 17 (1978) 333.
[12] G. Bogdanic, A. Fredenslund, Ind. Eng. Chem. Res. 33 (1994) 1331.
[13] G.M. Kontogeorgis, A. Fredenslund, D.P. Tassios, Ind. Eng. Chem. Res. 32 (1993) 362.
[14] B.-C. Lee, R.P. Danner, Fluid Phase Equilib. 128 (1997) 97.
[15] M.S. High, R.P. Danner, Fluid Phase Equilib. 53 (1989) 323.
[16] B.-C. Lee, R.P. Danner, AIChE J. 42 (1996) 837.
[17] C. Panayiotou, J.H. Vera, Fluid Phase Equilib. 5 (1980) 55.
[18] A. Fredenslund, J. Gmehling, P. Rasmussen, Vapor–liquid Equilibria Using UNIFAC, Elsevier, New York, 1977.
[19] T.E. Daubert, R.P. Danner H.M. Sibul, C. Stebbins, Physical and Thermodynamic Properties of Pure Chemicals: Data
Compilation, Taylor & Francis, Bristol, PA, 1997.
[20] J. Gmehling, U. Onken, Vapor–liquid Equilibrium Data Collection, DECHEMA Chemistry Data Series, DECHEMA,
Frankfurt, Vol. I, Parts 1, 2a (1977).
[21] J. Gmehling, U. Onken, W. Arlt, Vapor–liquid Equilibrium Data Collection, DECHEMA Chemistry Data Series,
DECHEMA, Frankfurt, Vol. I, Parts 1a (1981), 2b (1978), 2c (1982), 3/4 (1979), 6a, 6b, 7 (1980), 8 (1984).
[22] J. Gmehling, U. Onken, P. Grenzheuser, Vapor–liquid Equilibrium Data Collection, DECHEMA Chemistry Data Series,
DECHEMA, Frankfurt, Vol. I, Part 5 (1982).
[23] J. Gmehling, U. Onken, B. Kolbe, Vapor–liquid Equilibrium Data Collection, DECHEMA Chemistry Data Series,
DECHEMA, Frankfurt, Vol. I, Part 6c (1983).
[24] J. Gmehling, U. Onken, J.R. Rarey-Nies, Vapor–liquid Equilibrium Data Collection, DECHEMA Chemistry Data Series,
DECHEMA, Frankfurt, Vol. I, Parts 1b (1988), 2e (1988), 2f (1990), 3b (1993).
[25] J. Gmehling, U. Onken, U. Weidlich, Vapor–liquid Equilibrium Data Collection, DECHEMA Chemistry Data Series,
DECHEMA, Frankfurt, Vol. I, Parts 2d (1982).
[26] V.S. Parekh, R.P. Danner, J. Polym. Sci.: Polym. Phys. 33 (1995) 395.
[27] B.-C. Lee, Prediction of Phase Equilibria in Polymer Solutions, Ph.D. Thesis, Pennsylvania State University, 1995.
[28] E.A. Macedo, U. Weidlich, J. Gmehling, P. Rasmussen, Ind. Eng. Chem. Process Des. Dev. 22 (1983) 676.
[29] B.-C. Lee, R.P. Danner, Fluid Phase Equilib. 117 (1996) 33.
[30] H. Chen, M.J. Guest, S.P. Chum, A. Hiltner, E. Baer, J. Appl. Pol. Sci. 70 (1998) 109.
[31] M. Hamedi, N. Lützow, H.S. Betz, J.L. Duda, R.P. Danner, Ind. Eng. Chem. Res. 40 (2001) 3002.
[32] M. Hamedi, V. Muralidharan, B.-C. Lee, Danner, R.P., Prediction of Carbon Dioxide Solubility in Polymers Based on a
Group-Contribution Equation of State, Fluid Phase Equilib. 2001, submitted.
[33] B.-C. Lee, R.P. Danner, AIChE J. 42 (1996c) 3223.
[34] M. Hamedi, R.P. Danner, J.L. Duda, A lattice-fluid, group-contribution treatment of the glass transition of homopolymers,
copolymers, and polymer solutions, J. Appl. Polymer Sci., 2001, submitted.
[35] J.H. Gibbs, E.A. DiMarzio, J. Chem. Phys. 28 (1958) 373.
[36] A.F.M. Barton, Handbook of Solubility Parameters and Other Cohesion Parameters, CRC Press, Boca Roton, FL, 1983.
[37] M. Hamedi, R.P. Danner, J. Appl. Polym. Sci. 80 (2001) 197.
[38] N. Kleiber, J. Xu, Y. Yao, J. Wang, J. Shi, Fluid Phase Equilib. 69 (1991) 261.
[39] P.F. Stein, P.C. Proust, J. Chem. Eng. Data 16 (1971) 389.
[40] C.E.H. Bawn, M.A. Wajid, Trans. Faraday Soc. 52 (1956) 1658.
[41] N.H. Wang, S. Takishima, H. Masuoka, Kagaku Kogaku Ronbunshu 15 (1989) 313.
R.P. Danner et al. / Fluid Phase Equilibria 194–197 (2002) 619–639 639

[42] P.J.T. Tait, A.M. Abushihada, Polymer 18 (1977) 810.


[43] P. Eichinger, P.J. Flory, Trans. Faraday Soc. 64 (1968) 2035.
[44] Y. Kamiya, Y. Naito, D. Bourbon, J. Pol Sci. Pol. Phys. Ed. 32 (1994) 281.
[45] J.C. Dougherty, Diffusivity and Solubility Measurements Using the Pressure Decay Technique, M.S. Thesis, Pennsylvania
State University, 1999.
[46] C.K. Sham, D.J. Walsh, Polymer 28 (1987) 804.
[47] O. Olabisi, R. Simha, Macromolecules 8 (1975) 204.
[48] J. Parry, D. Tabor, Polymer 14 (1973) 628.
[49] M. Hirooka, T. Kato, J. Pol. Sci.: Pol. Lett. Ed. 12 (1974) 31.

You might also like