Sie sind auf Seite 1von 11

6928

Ind. Eng. Chem. Res. 2005, 44, 6928-6938

Modeling of Polar Systems with the Perturbed-Chain SAFT


Equation of State. Investigation of the Performance of Two Polar
Terms
Aleksandra Dominik and Walter G. Chapman*
Department of Chemical and Biomolecular Engineering, Rice University, 6100 S. Main Street,
Houston, Texas 77005

Matthias Kleiner and Gabriele Sadowski


Universita t Dortmund, Lehrstuhl fu r Thermodynamik, Emil-Figge-Strasse 70, 44227 Dortmund, Germany

The Perturbed-Chain SAFT (PC-SAFT) equation of state is applied to model phase equilibria
and the thermodynamic properties of ethers and esters. A systematic study of these two
homologous series is conducted, and the performance of two different approaches for including
the dipolar interactions in the equation of state is evaluated. Although both Polar PC-SAFT
[dipolar contribution due to Jog and Chapman, Mol. Phys. 1999, 97, 307-319] and PC-SAFT
+ Fischer [dipolar contribution by Saager and Fischer, Mol. Simul. 1991, 6, 27-49] yield similar
results for the considered systems, the parameters of Polar PC-SAFT are more physically
meaningful than those of PC-SAFT + Fischer. Consequently, Polar PC-SAFT is considered to
be more suitable for extrapolations, and for application to components that have multiple polar
groups in the chain.
1. Introduction
A molecules electrostatic moment has a strong effect
on the phase behavior of the macroscopic system. In this
work, we are concerned with dipolar interactions, which
influence the phase behavior of numerous systems of
industrial importance, such as mixtures containing
ketones, ethers, and esters, as well as polar polymers
and copolymers. Mixtures containing one polar and one
nonpolar component often exhibit strong, positive deviation from ideality, resulting in azeotropic behavior. We
can cite the mixture of 2-butanone with n-heptane as
an example.1 The deviation from ideality is due to
differences in intermolecular interactions. Dipolar interactions become increasingly important at low temperatures. For instance, in the polyethylene-dimethyl
ether (PE-DME) system, the two-phase region extends
to much higher pressures at low temperatures, where
DME-DME polar interactions are favored over PEDME interactions.2 The phase behavior of polar polymer
and copolymer solutions is determined by interactions
of the multiple dipole moments present in the polymer
chain.
The effect of dipolar interactions on the phase behavior is difficult to model and predict accurately. In
conventional equations of state (EOSs), the polar interactions are not taken into account explicitly. The nonideal behavior is modeled by fitting a large binary
interaction parameter, which often is dependent on the
temperature and composition of the system. Therefore,
the predictive abilities of such models are often weak,
because of the inability to extrapolate accurately with
temperature and to multicomponent systems.
Three methods could be considered for the incorporation of the polar contribution to the Helmholtz free
* To whom correspondence should be addressed. Tel.: 713348-4900. Fax: 713-348-5478. E-mail: wgchap@rice.edu.

energy into an EOS. The first possibility would be to


obtain the polar contribution from experimental results.
An attempt to derive this contribution from experimental data has been made by Lee and Chao,3 who used an
EOS for water and incorporated their expression into
the BACK EOS.4 However, the intermolecular interactions of water are much more complex than a dipoledipole interaction, which makes the physical meaning
of the polar contribution derived from an EOS for water
unclear.
Another pathway toward including the polar contribution would be to construct it on the basis of a theory.
Several statistical mechanics-based perturbation theories have been developed for polar fluids; however, the
u-expansion is the most widely applied theory.5,6 As
stated by Gray and Gubbins,5 the u-expansion is valid
only for mixtures of dipolar spheres. Attempts to extend
the theory to nonspherical, chainlike molecules have
proved difficult. Because no rigorous theory was available, the necessity of modeling nonspherical polar
molecules for engineering applications has resulted in
the development of approximate, often empirical, models. The most widely applied model represented the
dipolar nonspherical molecule as a large sphere. The
dipole moment was positioned at the center of the
sphere, and the diameter of the sphere was chosen such
that the molecular volume was preserved.7-14 However,
the model is in poor agreement with the molecular
simulation results.5 Moreover, this approach does not
allow for modeling of molecules with multiple polar
sites, such as diethers and triethers, diketones, and,
more importantly, polar polymers and copolymers.
Consequently, its use was limited to simple fluids and
mixtures. Other approaches have used site-site perturbation theory, accounting for molecular shape through
an interaction volume parameter. For instance, Walsh
et al.15 modeled chainlike polar compounds, using the

10.1021/ie050071c CCC: $30.25 2005 American Chemical Society


Published on Web 07/14/2005

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6929

Perturbed Anisotropic Chain Theory (PACT). PACT


uses an interaction site perturbation theory with dipolar
and multipolar interactions. The advantage of PACT is
that it takes into account molecular shape through an
interaction volume parameter. Very good results are
obtained for mixtures of esters ( ) 1.8 D, on average)
and chloroalkanes with normal alkanes. Sear16 applied
a technique similar to that of Wertheim to study the
chain formation by dipolar hard spheres. This approach
neglects long-range multibody interactions; therefore,
it is applicable only at low temperature and low density.
Finally, the polar contribution to the Helmholtz free
energy can be constructed on the basis of computer
simulation results. This approach has been followed by
Fischer and co-workers17-19 for a fluid consisting of twocenter dipolar Lennard-Jones (2CLJ) molecules with
fixed elongation, with the dipole embedded along the
axis. Empirical expressions were fitted to the molecular
simulation results to obtain a dipole-dipole term and
a quadrupole-quadrupole term.
It is important to systematically test these theories
for mixtures in which the dipolar interactions (or lack
thereof, as in ketone/alkane mixtures, for example)
actually control the phase behavior and induce strong
nonideality. To illustrate this point, Kraska and Gubbins12,13 have applied their Lennard-Jones SAFT (LJSAFT), which contains explicitly the dipole-dipole
interaction contribution to the Helmholtz free energy,
to mixtures of n-alkanes with n-alkanols and water.
Hydrogen-bonding interactions are stronger than dipoledipole interactions, and they control the phase behavior
for these mixtures, thus making the performance of the
polar contribution to the EOS difficult to assess. A
revealing test for the polar contribution is the application of the EOS, including the polar term, to systems
containing one or several strongly polar components,
such as ketones, ethers, and esters.
The objective of this work is to evaluate the relative
performance of two approaches proposed for incorporating the dipolar contribution into an engineering EOS:
the dipolar theory developed by Jog and Chapman,20
and the dipolar term determined by Fischer and coworkers.17-19 The approach proposed by Jog and Chapman is based on Wertheims Thermodynamic Perturbation Theory offirst order (TPT1); the dipolar contribution
to the Helmholtz free energy is calculated on the basis
of a u-expansion. The theory has been successfully
applied to ketones and their mixtures with alkanes,21,22
as well as to polar copolymers.21,23 Saager and Fischer,
on the other hand, obtained the dipolar contribution to
the free energy by fitting empirical expressions to the
molecular simulation results, as previously mentioned.
The theory has been applied to real substances,19,24 as
well as to mixtures25,26 in the framework of the BACK
EOS.4 To recommend the best approach to be used for
complex systems including dipolar molecules, such as
polymer and copolymer solutions, a systematic study of
the homologous series of ethers and esters has been
conducted in this work. A particular emphasis was given
to molecules with multiple polar groups, to gain better
understanding of the dependence of model parameters
on the number of polar groups in the chain. Both dipolar
terms are incorporated into the PC-SAFT EOS.27 The
PC-SAFT model was chosen for this investigation,
because of its excellent predictive abilities for polymer
systems.28,29

2. Theory
The molar residual Helmholtz free energy is given in
terms of a perturbation expansion:

ares ) ahs + achain + adisp + apolar + aassoc

(1)

where ares is the Helmholtz free energy residual to an


ideal gas at the same temperature and density as the
fluid of interest. The hard-sphere contribution (ahs) is
due to Carnahan and Starling,30 and the chain term
(achain) was developed by Chapman and co-workers,31,32
on the basis of Wertheims TPT1. The species considered
in this work are nonassociating; therefore, the association contribution (aassoc) to the residual Helmholtz free
energy vanishes. The expression for the dispersion
contribution (adisp) used in this work was obtained by
Gross and Sadowski,27 who proposed the PC-SAFT
EOS. These authors also conducted a thorough parametrization of the PC-SAFT EOS for nonpolar fluids.
Research led by Jog and Chapman showed that the
dipolar contribution to the Helmholtz free energy can
be added to the sum as in eq 1 without modifying the
hard-sphere, chain, and dispersion terms.20,33 Two approaches for expressing the dipolar contribution (apolar)
are considered in this study; therefore, the basic assumptions and the fundamental equations for both
dipolar terms are presented in this section.
2.1. Jog and Chapman Dipolar Term. Thermodynamic properties obtained using the dipolar term developed by Jog and Chapman20 on the basis of Wertheims TPT1 are in excellent agreement with molecular
simulation for hard-sphere chains with a dipole moment
oriented perpendicularly to the axis of the molecule. The
model is realistic for many substances of great interest,
such as ketones, esters, and polar polymers and copolymers. The change in free energy due to polar interactions is accurately obtained by dissolving all the bonds
in a chain and then applying the u-expansion to the
resulting mixture of polar and nonpolar spherical segments. The polar contribution written in the Pade
approximate has the following form:

apolar )

a2
1 - a3/a2

(2)

where a2 and a3 are the second- and third-order terms


in the perturbation expansion. Written for mixtures,
and allowing for multiple dipolar segments, these terms
have the following form:

a2 ) -

9 (kT)2

i j

i2j2
xixjmimjxpixpj
I2,ij
dij2

(3)

a3 )
5
162

F2

i j k

(kT)3

xixjxkmimjmkxpixpjxpk

i2j2k2
I3,ijk
dijdjkdik
(4)

In the aforementioned equations, I2,ij and I3,ijk are the


angular pair and triplet correlation functions, F is the
number density of molecules, k is the Boltzmann
constant, T is the temperature, dij is the average
diameter of segments i and j, and i is the dipole
moment for component i. As shown by Jog et al.,21 they

6930

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

are related to the corresponding pure fluid integrals


by

j)
I2,ij ) I2(,m

(5)

j)
I3,ijk ) I3(,m

(6)

As mentioned previously, mi is the segment number of


component i, and xi is the mole fraction of component i
in the mixture. The universal parameters ci, ni, li, ki,
oi, as well as R and , are given in the work of Muller et
al.19 The reduced dipole moment /i in eq 9 is defined
as

where

i ximi

m
j )

(7)

and

i ximidi3

(8)

In eqs 3 and 4, xpi is the fraction of dipolar segments in


component i, whereas xi is the mole fraction of component i in the mixture, and mi is the segment number of
component i. In the case of a molecule with one dipolar
function in the chain, xp should be equal to 1/m. The
PC-SAFT model extended to polar systems, and including the dipolar term proposed by Jog and Chapman,
will hereafter be referred to as Polar PC-SAFT. The
parameters of the model are the segment number m,
the segment diameter , the segment dispersion energy
/k, and the fraction of polar segments in the chain xp.
2.2. Fischers Dipolar Term. The dipolar term of
Fischer and co-workers17-19 was obtained on the basis
of molecular simulations by fitting empirical expressions
to the simulation data. For this purpose, simulations
with two-center Lennard-Jones (2CLJ) molecules of a
fixed elongation L* ) L/ ) 0.505 were performed for
different temperatures, densities, and dipole moments.
The dipolar EOS was applied to real fluids24 and
mixtures.25,26 One restriction of the model in its original
form is the fixed elongation, which makes it nontrivial
to apply to strongly asymmetric mixtures. Another
limitation is that only one polar group per molecule is
considered. As a result, the original model does not
account for multiple functional groups with different
polarity within a molecule explicitly in the EOS. To
consider the nonspherical shape and multiple dipolar
functional groups explicitly, the polar term was modified
following a methodology similar to that proposed by Jog
and Chapman.20 As in the previously presented approach, the parameter xp is introduced in this model.
The dipolar term can then be written as

apolar )

i j

28

xixjmimjxpixpj

( ) ()
[ ( )]
T

cq 1.13T
q)1

nq/2

lq/2

F0

(/i /j )kq/8 exp -oq

F0

(9)

with

( )

R + (1 - R) T

F
) Fd 3
F0 6 ij
0.1617 T0

(10)

and

T0 )

()

ij
1.268
k

(11)

/2
i )

i2
idi3

(12)

where i is the molecular or group dipole moment, which


can be obtained from experiments or quantum chemical
calculations. By applying this modified dipolar term of
Saager and Fischer with PC-SAFT (hereafter referred
to as PC-SAFT + Fischer), the EOS has four adjustable
pure-component parameters as Polar PC-SAFT: m, ,
, xp.
The goal of this work is to choose the best methodology for modeling components with multiple dipolar
groups, and eventually to propose a methodology for
modeling polar polymers and copolymers in the framework of PC-SAFT. The relatively small number of
systems previously examined with both approaches is
not sufficient for determining which term will perform
better for polar polymers. A more-detailed study of polar
systems is conducted here to evaluate the relative
performance of both approaches.
3. Model Parameters and Regression Method
The PC-SAFT model including either of the dipolar
terms considered in this work has four adjustable purecomponent parameters: the segment number (m), the
segment diameter (), the segment dispersion energy
(/k), and the fraction of polar segments in a chain (xp).
These parameters can be regressed by fitting purecomponent vapor pressure and liquid density data. For
a molecule with a single polar function in the chain, the
experimental value of the dipole moment of the molecule
can be used. Alternatively, if experimental values are
not readily available, the value of the dipole moment
can be obtained from quantum mechanics. For molecules with multiple polar sites, the value of the
functional groups dipole moment is to be used in the
model. Ideally, the value of xp should be k/m, where k
is the number of polar functions in the chain and m is
the number of segments of the molecule. In the framework of SAFT, molecules are depicted as chains of
tangentially connected segments, which is somewhat
oversimplified. Thus, SAFT reflects averages, and not
the exact shape and structure of the molecules. Consequently, it is not realistic to expect that the value of xp
can be fixed to k/m for all molecules with k polar
functions. The value of the polar segment fraction is
dependent on the strength of the dipole moment of the
molecule and its chain length, as will be demonstrated
later in this article. Therefore, xp will be an adjustable
parameter in this model. One of the objectives of this
work, however, is to establish the relationship between
the strength of the dipole moment, the number of polar
functions, the chain length of the molecule, and the
value of the xp parameter, to obtain a predictive model
for polar polymers and copolymers.
The four parameters were obtained for the homologous series of ethers, including symmetric, nonsymmetric, and polyfunctional ethers, and the homologous
series of esters. The pure-component data used in this

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6931

study are from the Daubert and Danner pure-component data compilation.34
EOS parameters are usually regressed by fitting
vapor pressure and liquid density data, applying a
Levenberg-Maquart minimization algorithm. The objective function Fobj is defined as

nexp

obj

i)1

fexp
- fcalc
i
i
fexp
i

(13)

In eq 13, f [Pvap, Fliq], and nexp is the total number of


experimental data points. In the case of nonpolar PCSAFT, parameters obtained from this method give
satisfactory results for mixtures,27 because the minimum of the objective function is narrow and the
parameters are unique. When the polar contribution is
included, it has been observed that there exist several
sets of parameters (m, , /k, xp) that reproduce purecomponent experimental data with good accuracy. Yet,
only one set of parameters gives results in good agreement with mixture data. This behavior of the objective
function was observed for Polar PC-SAFT and PCSAFT + Fischer. Consequently, the method for obtaining pure-component parameters for both dipolar PCSAFT models has been modified. To obtain a unique set
of parameters, one set of binary mixture data must be
included in the regression. However, the choice of the
second component of the mixture should be made
carefully. It is important to choose a component for
which pure-component parameters have been determined and tested, and that does not contain any
functionality (polar group, self-associating components,
quadrupolar moment, cyclic components). Therefore,
mixtures of ethers and esters with alcohols, water,
carboxylic acids, cyclic alkanes, and chloroalkanes and
fluoroalkanes have been excluded from the parameterfitting procedure. Only mixtures of ethers and esters
with normal alkanes are considered for the regression.
The values of the parameters, as well as the average
absolute deviations (AADs) over the considered temperature range for vapor pressures and saturated liquid
densities are reported in Table 1 for Polar PC-SAFT
and in Table 2 for PC-SAFT + Fischer.
In the case of Polar PC-SAFT, it has been observed
that, after the value of the product xpm that gives good
results for both binary and pure-component data is
determined for one component, the same value of xpm
can be retained for other components with the same
molecular dipole moment and the same number of
dipolar groups in the chain. For example, all the ethers
with a single polar group have a dipole moment of 1.2
D. The product of xpm has been determined for diethyl
ether, for which good-quality pure-component data, as
well as a large selection of binary mixture data, were
available. The value of xpm has been maintained
constant for the entire series, and the results obtained
for other mixtures of monofunctional ethers are in very
good agreement with experimental data. The same
behavior has been observed for the homologous series
of ketones22 and esters (this work). This remarkable
property of Polar PC-SAFT demonstrates the fact that
the model captures, to a certain extent, the molecular
structure of fluids. Moreover, only one set of binary data
is needed for parameter regression for a homologous
series of polar components. For PC-SAFT + Fischer,
the values of the product xpm range between 2.07 and

2.19 for ethers with one polar function, as opposed to


Polar PC-SAFT, for which xpm is 1.0. This indicates
that the parameter xp is not as physically meaningful
in the Fischer and Saager approach as it is in the Jog
and Chapman approach. This variation of xpm could be
explained by the fact that, for the model of Fischer and
Saager, simulation data of nonspherical molecules were
used. If the segment appproach is applied to that model,
the dipole moment ranges over more than one (spherical) segment, which may result in higher values of the
product xpm, compared to Polar PC-SAFT.
The obtention of Polar PC-SAFT parameters is more
complex for components with multiple dipolar groups
in the chain. In this work, we focus on the series of
ethers, because no experimental data are available for
molecules with multiple dipolar groups from other series
of polar components. For polyfunctional ethers, only the
value of their molecular dipole moment is available.35,36
Yet, the value of the dipole moment of the polar segment
is needed for the calculations. This value could be
obtained from quantum calculations. In this study,
however, the value of the polar segment dipole moment
is set to 1.2 D for all polyfunctional ethers when
applying Polar PC-SAFT. For PC-SAFT + Fischer, it
turned out that the experimental dipole moments of the
molecules must be used as described in the next section.
4. Results
4.1. Monofunctional Components. A thorough
parametrization was conducted for the homologous
series of ethers and esters in the framework of both
Polar PC-SAFT and PC-SAFT + Fischer (see Table 1
for pure-component parameters from Polar PC-SAFT,
and Table 2 for parameters of PC-SAFT + Fischer).
Both approaches give an excellent representation of
the pure-component properties of dipolar components
with one dipole moment on the chain. As previously
mentioned, however, the value of xpm for Polar PCSAFT is constant for any given homologous series,
which greatly facilitates parameter regression. A comparison of results obtained for mixtures including a
normal alkane, and a component with one polar function
in the chain reveals that both Polar PC-SAFT and PCSAFT + Fischer describe the phase behavior of mixtures
very accurately, without the need to adjust a binary
interaction parameter. This is illustrated in Figure 1
for the system of 1,2-propylene oxide in n-hexane, and
in Figure 2 for the butyl acetate-cyclohexane system.
A few parameters were regressed for ethers, and an
almost-complete study for esters was conducted by
Gross and Sadowski27 with the PC-SAFT model without accounting for dipolar interactions. Consequently,
it is possible to compare the binary results obtained for
binary mixtures using the original, nonpolar PC-SAFT,
with the results from PC-SAFT including a polar
contribution. The comparison is performed here for
Polar PC-SAFT, because the performances of both
polar terms for mixtures including monofunctional polar
molecules are more or less identical. In Figure 3, the
results for a mixture of dimethyl ether and propane at
four different temperatures are presented. Polar PCSAFT represents the binary data with good accuracy
at all considered temperatures, whereas PC-SAFT does
not capture the azeotropic behavior of the mixture. The
PC-SAFT results are expected, because the EOS is
unaware of the effect of the polar groups on the phase
behavior. The nonideality is caused by a lack of attrac-

6932

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

Table 1. Pure-Component Parameters for the Polar PC-SAFT Model


AAD%
Pvap

xp

temperature
range [K]

Monoethers
215.98
220.59
236.19
235.89
236.95
217.84
230.20
237.92
240.24
231.49
221.35

0.4977
0.3474
0.2947
0.2322
0.1931
0.4007
0.3639
0.3276
0.2900
0.3248
0.2966

200-400
200-466
250-520
300-580
350-610
280-420
220-460
200-500
250-530
300-490
250-500

0.80
0.72
0.47
0.63
0.36
0.35
0.92
0.28
0.36
0.75
3.32

3.5850
3.5973
3.5556
3.5319

Diethers
234.98
239.59
221.20
223.34

0.8602
0.9451
0.5955
0.6806

250-470
350-520
300-510
300-520

4.6265
6.6346
7.8270

3.5194
3.4077
3.4184

Polyethers
225.62
214.38
225.24

0.9078
0.7401
0.7082

44.05
58.08
58.08
72.11
72.11

1.4953
2.0105
1.9227
2.1875
2.2846

methyl formate
ethyl formate
propyl formate
butyl formate
methyl acetate
ethyl acetatef
propyl acetate
butyl acetate
methyl propionate
ethyl propionate
propyl propionate

60.053
74.08
88.11
102.13
74.08
88.11
102.13
116.16
88.11
102.13
116.16

1.8742
2.2246
2.4498
2.9234
2.3967
2.7481
3.1606
3.6600
2.7813
3.1522
3.4836

3.5425
3.6793
3.8039
3.7940
3.5477
3.6511
3.6866
3.6751
3.5969
3.6463
3.7476

acetone
2-butanone
2-pentanone
3-pentanone
2-hexanone
3-hexanone
2-heptanone
4-heptanone
2-octanone
3-octanone
2-nonanone
5-nonanone
2-undecanone
6-undecanone
2-tridecanone

58.078
72.1
86.13
86.13
100.16
100.16
114.19
114.19
128.22
128.22
142.24
142.24
170.3
170.3
198.35

2.221
2.418
2.826
2.812
3.232
3.202
3.704
3.651
4.002
4.134
4.55
4.448
5.441
5.399
6.25

component

molecular weight,
Mw [g/mol]

dimethyl ether
diethyl etherf
di-n-propyl ether
di-n-butyl ether
di-n-pentyl ether
methyl ethyl ether
methyl n-propyl ether
methyl n-butyl ether
methyl n-pentyl ether
ethyl n-propyl ether
diisopropyl ether

46.069
74.123
102.177
130.23
158.28
60.1
74.123
88.15
102.177
88.15
102.177

2.0090
2.8787
3.3930
4.3064
5.1782
2.4953
2.7480
3.0529
3.4480
3.0791
3.3721

3.4343
3.5549
3.7425
3.7429
3.7474
3.4580
3.6183
3.6801
3.7199
3.7021
3.7856

dimethoxymethane
1,2-dimethoxyethane
diethoxymethanef
1,2-diethoxyethane

76.095
90.122
104.149
118.17

2.4994
2.9626
3.6103
4.1143

DEG DMEb,f
TRG DMEc
TEG DMEd

134.17
178.23
222.282

ethylene oxide
1,2-propylene oxidef
1,3-propylene oxide
1,2-butene oxide
tetrahydrofuran

a [D]

xpm

0.90
0.90
1.14
0.71
2.06
1.30
1.12
1.28
0.67
1.30
1.84

1.3
1.2
1.2
1.2
1.2
1.2
1.2
1.2
1.2
1.2
1.2

1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0
1.0

0.98
0.38
0.25
0.56

0.43
1.92
1.86
1.30

1.2
1.2
1.2
1.2

2.15
2.8
2.15
2.8

350-600
300-620
400-680

2.32
0.39
0.30

1.36
2.18
3.66

1.2
1.2
1.2

4.2
4.91
5.54

Epoxies and Cyclic Ethers


3.5860
288.93 0.5350
3.6095
258.82 0.3979
3.5990
289.54 0.4161
3.7903
275.31 0.3657
3.6241
279.07 0.33502

170-455
310-470
260-510
200-510
200-530

3.01
2.21
2.37
1.66
0.44

1.55
1.87
1.68
2.62
1.40

1.7
2.0
1.8
1.7
1.75

0.8
0.8
0.8
0.8
0.8

Esters
232.19
251.54
252.77
249.41
238.87
236.99
238.87
237.43
240.92
236.76
241.51

0.8004
0.6743
0.6123
0.5131
0.6259
0.5458
0.4746
0.4098
0.5393
0.4759
0.4306

200-470
210-495
215-550
215-550
200-480
215-485
200-530
230-560
200-515
215-545
260-560

1.66
2.06
1.91
2.25
1.29
1.13
1.41
0.78
1.41
0.32
2.09

1.36
1.46
2.75
3.69
1.66
1.64
1.41
0.53
2.70
0.71
2.78

1.75
1.96
1.91
1.90
1.7
1.84
1.86
1.86
1.7
1.75
1.79

1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5
1.5

Ketonese
3.607908253 259.99
3.689051518 270
3.710210701 264.97
3.705656195 265.83
3.730568917 263.8
3.732817265 262.1
3.731131258 259.17
3.745135035 255.72
3.752375796 257.07
3.720417659 254.12
3.743460108 256.81
3.765670418 256.25
3.742901466 253.8
3.745693011 251.65
3.762355566 253.34

0.2258
0.2074
0.177
0.1774
0.1543
0.156
0.136
0.135
0.1254
0.12
0.1101
0.1125
0.0919
0.0924
0.0799

253-463
270-374
282-353
285-347
299-332
299-324
306-416
275-384
319-420
316-360
340-422
301-356
309-427
303-352
341-427

4.726
0.647
0.235
0.137
0.025
0.036
0.556
0.864
1.523
0.327
0.099
0.053
0.047
0.183
0.022

1.338
0.498
0.033
0.197
0.013
0.01
0.638
9.703
12.33
3.671
0.155
0.013
0.144
1.305
0.177

2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7
2.7

0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5
0.5

[]

/k [K]

a Values of the segment dipole moment used in the calculations. In the case of monofunctional molecules, it is the experimental value
of the molecular dipole moment. b Diethylene glycol dimethyl ether. c Triethylene glycol dimethyl ether. d Tetraethylene glycol dimethyl
ether. e Parameters for ketones taken from ref 22. f The following binary systems were used in the fitting procedure: diethyl ether +
n-hexane (from ref 47), diethoxymethane + n-heptane (from refs 45 and 46); 1,2-propylene oxide + n-hexane (from ref 38); DEG DME +
n-decane (from ref 45); and ethyl acetate + n-heptane (from ref 42).

tion between the unlike components, as compared to the


attraction between the pure, or like, components. When
the EOS does not explicitly account for dipolar interactions, this lack of attraction cannot be predicted. A
relatively high value of the binary interaction parameter
(kij ) 0.03) is necessary for PC-SAFT to reproduce the
nonideal behavior accurately, despite the weak dipole
moment of dimethyl ether. Moreover, it is likely that
the interaction parameter is temperature-dependent

outside of this temperature range. On the other hand,


the polar term captures the temperature dependence
well within this range of temperatures and is expected
to yield good results at temperatures outside the range.
For esters, results for PC-SAFT and Polar PC-SAFT
are compared for the binary mixture of ethyl acetate
with heptane in Figure 4. The mixture exhibits azeotropic behavior, which is very accurately reproduced by
polar PC-SAFT with a binary interaction parameter

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6933
Table 2. Pure-Component Parameters for Ethers and Esters Obtained for PC-SAFT + Fischer
AAD%
Pvap

xp

temperature
range [K]

Monoethers
3.5652
223.76
3.6768
230.82
3.7797
241.81
3.4899
223.44
3.6414
211.26

0.7550
0.6145
0.5170
0.8545
0.5749

200-450
250-520
300-580
280-420
250-500

0.29
0.35
2.71
0.46
2.18

3.1970

Diethers
3.4819
225.81

0.6397

300-500

134.175
222.282

4.7104
7.9200

Polyethers
3.4839
225.81
3.3661
227.43

0.6397
0.325

1.2-propylenoxide
1.3-propylenoxide
tetrahydrofuran

58.08
58.08
72.12

2.1341
1.9754
2.2631

methyl formate
ethyl formate
propyl formate
ethyl acetate
n-propyl acetate
n-butyl acetate
methyl propionate
ethyl propionate
propyl propionate

60.053
74.08
88.11
88.11
102.13
116.16
88.11
102.13
116.16

2.3829
2.5818
2.9067
3.1878
3.2563
3.6102
3.0235
3.4309
3.7225

component
diethyl ether
di-n-propyl ether
di-n-butyl ether
methyl ethyl ether
diisopropyl ether
1,2-dimethoxyethane
DEG DMEb
TEG DMEc

molecular weight,
Mw [g/mol]

74.12
102.177
130.23
60.1
102.177

2.8311
3.5577
4.0882
2.4254
3.6883

90.122

a [D]

xpm

1.43
1.47
5.04
1.47
3.28

1.22
1.21
1.2
1.22
1.26

2.14
2.17
2.11
2.07
2.12

1.04

1.21

1.71

2.73

300-500
410-680

0.8
2.73

1.18
2.63

1.97d
2.45d

3.00
2.57

Epoxies and Cyclic Ethers


3.4739
252.95
0.916
3.5336
290.97
0.936
3.62159
282.42
0.86

310-470
280-510
250-500

3.92
2.68
0.78

1.62
0.18
0.75

2
1.8
1.75

1.95
1.85
1.95

Esters
239.57
233.64
241.65
221.85
237.44
240.05
238.65
231.56
234.61

180-480
200-500
220-520
190-510
263-548
200-560
220-520
200-540
260-560

1.13
0.67
0.38
0.43
1.86
0.8
0.84
0.67
2.5

0.99
1.85
1.61
1.72
1.44
2.65
1.96
0.97
2.52

1.75
1.96
1.91
1.84
1.86
1.86
1.7
1.75
1.79

2.01
2.32
2.44
2.68
2.78
2.95
2.59
2.85
2.93

[]

3.2219
3.4424
3.5409
3.4284
3.5692
3.6721
3.4717
3.5421
3.5877

/k [K]

0.8431
0.8988
0.8399
0.8406
0.8537
0.8169
0.8561
0.8297
0.787

a Molecular dipole moments used in the calculations. Values taken from ref 49. b Diethylene glycol dimethyl ether. c Tetraethylene
glycol dimethyl ether. d Values taken from ref 50. e The following binary systems were used in the fitting procedure: diisopropyl ether +
n-heptane (from ref 45); 1,2-dimethoxyethane + n-heptane (from ref 45); 1,2-propylene oxide + n-hexane (from ref 38); DEG DME +
n-decane (from ref 45); TEG DME + n-heptane (from ref 43); and ethyl acetate + n-heptane (from ref 42).

Figure 1. Vapor-liquid equilibrium for the 1,2-propylene oxidehexane system. The solid line represents Polar PC-SAFT, and
the dashed line represents PC-SAFT + Fischer; the binary
interaction parameter was set to zero in both cases. Symbols are
experimental data by Bougard and Jadot.38

Figure 2. Vapor-liquid equilibrium for the butyl acetatecyclohexane system. The solid line represents Polar PC-SAFT,
and the dashed line represents PC-SAFT + Fischer; the binary
interaction parameter was set to zero in both cases. Symbols are
experimental data from the Dechema compilation.39

set to zero. This is not the case for the original PCSAFT, which again does not capture the nonideal
behavior of the mixture.
The case of mixtures containing two polar components
was also examined, and good results were observed with
both polar terms. Such mixtures show almost-ideal
behavior, because of the similarity of both components;
therefore, they are not as challenging a test for the polar
terms as the strongly nonideal mixtures that contain
polar and nonpolar components.

4.2. Difunctional Components. The true challenge


of this work is to choose xp for components with multiple
polar groups, and to be able to predict values of xp for
polar polymers and copolymers. Ideally, the product xpm
should be equal to k, where k is the number of polar
groups in the chain. For monofunctional molecules,
however, this equation holds only for weakly polar
ethers in the case of Polar PC-SAFT, and it is not
applicable for PC-SAFT + Fischer for any components
considered in the study. Only one molecule with two

6934

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

Figure 3. Vapor-liquid equilibria of the dimethyl ether-propane


system. Solid lines are the results from Polar PC-SAFT, dashed
lines are the results from nonpolar PC-SAFT, and symbols are
experimental data from Horstmann et al.40 and Giles and Wilson.41
Binary interaction parameters are set to zero.

Figure 4. Vapor-liquid equilibria of the ethyl acetate-heptane


system. Dashed line represents nonpolar PC-SAFT, and the solid
line represents polar PC-SAFT. Binary interaction parameters
set to zero. Symbols are experimental data from Shealy and
Sandler.42

polar sites (2,4-pentadione) has been previously considered in the framework of Polar PC-SAFT. The results
obtained by Sauer and Chapman22 for the 2,4-pentadione + cyclohexane system were very good, but applying the model to only one system that contains a
component with more than one polar site is not sufficient for proposing a methodology for modeling systems that contain components with multiple polar
groups. In this work, parameters were obtained for two
difunctional and several polyfunctional ethers (see
Tables 1 and 2). The results obtained from both models
for systems including polyethers reveal the differences
between Polar PC-SAFT and PC-SAFT + Fischer.

Figure 5. Vapor-liquid equilibria of the di-n-propyl etherheptane and diethoxymethane-heptane systems from Polar PCSAFT. Binary interaction parameters set to zero. Symbols are
experimental data from Treszczanowicz et al.45,46

Figure 6. Vapor-liquid equilibria of the 1,2-dimethoxyethaneheptane and 1,2-diethoxyethane-heptane systems from Polar PCSAFT. Binary interaction parameters set to zero. Symbols are
experimental data from Treszczanowicz et al.45,46

In the case of Polar PC-SAFT, the segment dipole


moment is used in the calculations of thermodynamic
properties for components that have more than one
polar function in their chain. Results obtained for
mixtures of diethers with n-alkanes are in excellent
agreement with experimental data, as shown in Figures
5 and 6. The experimental molecular dipole moments
differ for the diethers considered in this study. Two of
them (1,2-dimethoxyethane and 1,2-diethoxyethane)
have a molecular dipole moment equal to 1.71 D.
Diethoxymethane and dimethoxymethane have molecular dipole moments of 1.22 and 1.0 D,36 respectively.
However, in the parameter regression, the value of 1.2
D used for all ethers has been retained for the segment
dipole moment. Therefore, to obtain satisfactory results
for binary mixtures of diethers from Polar PC-SAFT,

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6935

Figure 7. Vapor-liquid equilibrium of the 1,2-dimethoxyethaneheptane system from PC-SAFT + Fischer. Binary interaction
parameters set to zero. Dashed line: segment dipole moment of
1.2 D used in the calculations. Solid line: experimental molecular
dipole moment of 1.71 D used in the calculations. Symbols are
experimental data from Treszczanowicz and Lu.45

Figure 8. Vapor-liquid equilibria of the DEG DME-decane


system at 120 C. Solid lines are the results from Polar PC-SAFT,
dashed lines are the results from PC-SAFT + Fischer, and dotdashed lines are results from UNIFAC. Binary interaction parameters of the PC-SAFT versions were set to zero. Symbols are
experimental data from Treszczanowicz.46

different values of the product xpm are necessary for the


strongly polar molecules on one hand (1,2-dimethoxyethane and 1,2-diethoxyethane), and the weakly polar
components on the other hand (diethoxymethane and
dimethoxymethane).
The parametric study of the PC-SAFT + Fischer
model, in the case of diethers, revealed that the experimental molecular dipole moments must be used when
the PC-SAFT + Fischer model is applied to get a good
parameter set that describes binary mixtures, as shown
in Figure 7. The calculations of the binary system of
1,2-dimethoxyethane and heptane show exemplarily
that it is not possible to predict the azeotropic behavior
with the parameters fitted using the group dipole
moment of ethers of 1.2 D. In contrast, the results using
the parameters fitted with the molecular dipole moment
of 1.71 D for 1,2-dimethoxyethane are in good agreement
with the experimental data.
The results of the study of diethers with both Polar
PC-SAFT and PC-SAFT + Fischer demonstrate that
both models are able to represent the pure-component
properties, as well as the binary mixture phase equilibria, with good accuracy. When Polar PC-SAFT is
used, however, the segment dipole moment is used,
which is consistent with the theoretical derivations of
the dipolar contribution (see the reports by Jog and coworkers20,21). To get satisfactory agreement with experimental data for mixtures in the case of PC-SAFT
+ Fischer, the experimental molecular dipole moment
had to be used for diethers. This potentially could cause
problems for polar components with multiple polar
groups, for which the molecular dipole moment is not
readily available.
4.3. Polyfunctional Molecules. The results obtained with Polar PC-SAFT and with PC-SAFT +
Fischer for diethylene glycol dimethyl ether (DEG DME)
in n-decane are compared to results from modified
UNIFAC in Figure 8. The parameters for DEG DME
for both Polar PC-SAFT and PC-SAFT+Fischer were
obtained using the method described in the previous

section of this article (mixture data was also considered


in the regression). Both models accurately describe the
binary vapor-liquid equilibrium, as well as the pure
component thermodynamic properties of DEG DME.
The group contribution parameters from Gmehling et
al.37 were used for the modified universal quasichemical
functional group activity coefficient (UNIFAC) calculations. Although PC-SAFT with either polar term
describes the vapor-liquid equilibrium of the system
very accurately, the UNIFAC calculation overestimates
the azeotropic pressure. This comparison shows that,
for the considered mixtures of polyfunctional molecules
with alkanes, any of the applied PC-SAFT versions
including the dipolar contribution had superior results,
compared to modified UNIFAC calculations. A similar
result was obtained earlier for a mixture of a diketone
with an alkane. In this case, modified UNIFAC predicts
a heterogeneous azeotrope, while the system actually
exhibits a homogeneous azetorope.22 The polar parameter xp exhibits a consistent monotonic variation with
the number of polar functions in the chain in the case
of Polar PC-SAFT (see Table 1). This behavior is not
observed with the PC-SAFT + Fischer model, which
corroborates the assertion that the xp parameter is not
as physically meaningful in the PC-SAFT + Fischer
model as it is in Polar PC-SAFT.
Finally, systems that contained tetraethylene glycol
dimethyl ether (TEG DME) with various n-alkanes were
investigated. Polar PC-SAFT gives a satisfactory description of both pure component properties (see Table
1) and phase behavior of the mixture (see Figure 9). To
reproduce the liquid-liquid equilibria of TEG DME with
various alkanes, very small values of the binary interaction parameter are sufficient. The values of the product
xpm for TRG DME and TEG DME were obtained by
fitting a simple correlation to the values of xpm regressed for shorter ethers, thus taking advantage of the
monotonic variation of the polar parameter in Polar
PC-SAFT. To propose a correlation of the polar parameter with the number of polar groups in the chain that

6936

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

Figure 9. Liquid-liquid equilibria of TEG DME and various


n-alkanes. Results shown are from Polar PC-SAFT. Symbols are
experimental data from Mozo et al.43 and Treszczanowicz and
Cieslak.44

Figure 11. Excess enthalpies of the binary mixtures of polyethers


with n-dodecane. Comparison of experimental data from Burgdorf
et al.48 with Polar PC-SAFT predictions using kij ) 0.0 for
DEG DME + n-dodecane, kij ) 0.004 for TRG DME + n-dodecane,
and kij ) 0.005 for TEG DME + n-dodecane.

Figure 10. Liquid-liquid equilibria of TEG DME and various


n-alkanes. Results from PC-SAFT + Fischer. Symbols are experimental data from Mozo et al.43 and Treszczanowicz and
Cieslak.44

Figure 12. Excess enthalpies of the binary mixtures of polyethers


with n-dodecane. Comparison of experimental data from Burgdorf
et al.48 with PC-SAFT + Fischer predictions using kij ) 0.0 for
DEG DME + n-dodecane and kij ) 0.008 for TEG DME + ndodecane.

is valid for extrapolation, investigation of longer chain


molecules is necessary. We note that the number of
polar segments increases with the number of polar
functions but is not doubled or tripled, as compared to
a monofunctional molecule, as one might have expected.
This is due to the partial cancellations by one another
of the dipole moments in the chain, and illustrates the
difficulties encountered for modeling of polar systems.
Remarkably, Polar PC-SAFT is able to predict the
excess enthalpies of mixtures of polyethers with nalkanes, as shown in Figure 10. The binary interaction
parameter of the system DEG DME + n-dodecane was
set to zero for this calculation. The binary interaction
parameter of the system TEG DME + n-dodecane was
obtained by assuming a linear dependence of kij between
TEG DME and n-alkanes on the molecular weight of
n-alkanes. Recall that binary interaction parameters

were obtained for TEG DME and three n-alkanes from


phase equilibrium data, as shown in Figure 9. Finally,
no binary phase equilibrium data were considered in
this study for TRG DME. Consequently, the value of kij
for the TRG DME + n-dodecane system was interpolated from the values of kij used for DEG DME and
TEG DME in n-dodecane. A linear dependence of kij on
the molecular weight of the polyether was assumed in
this case. The good agreement with experimental data
and the ability to predict the value of the binary
interaction parameter for polyfunctional molecules illustrate the predictive abilities of Polar PC-SAFT.
PC-SAFT + Fischer is able to represent the pure
component properties of TEG DME, as well as its
liquid-liquid-phase equilibria with n-alkanes (see Figure 11). The excess enthalpies of mixtures of DEG DME

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005 6937

+ n-dodecane and TEG DME + n-dodecane are in good


agreement with experimental data (see Figure 12). The
binary parameter for the TEG DME + n-dodecane
system was obtained by assuming a linear dependence
on the molecular weight of n-alkanes, as it is done for
Polar PC-SAFT. Because no monotonic dependence of
the product xpm is observed in the case of PC-SAFT +
Fischer, the value of xpm must be adjusted to the binary
data in this case. However, the product xpm for
TEG DME is detemined to be even smaller than that
for DEG DME (see Table 2), which is not physically
meaningful.
5. Conclusions
The PC-SAFT model extended to polar systems was
applied to model phase equilibria of ethers and esters.
The results obtained from PC-SAFT, including two
different approaches for accounting for dipolar interactions, were compared, and a parametric study of both
Polar PC-SAFT and PC-SAFT + Fischer was conducted. The study revealed that both models yield very
similar results for all systems considered. However, it
was observed that the polar parameters obtained from
Polar PC-SAFT are physically more meaningful than
those regressed using PC-SAFT + Fischer. Therefore,
the extrapolative capabilities of Polar PC-SAFT, with
respect to parameter estimation of complex molecules,
are expected to be superior to those of PC-SAFT +
Fischer.
Acknowledgment
The authors gratefully acknowledge the financial
support of the Consortium of Complex Fluids.
Literature Cited
(1) Takeo, M.; Nishii, K.; Nitta, T.; Katayama, T. Fluid Phase
Equilib. 1979, 3, 123.
(2) Hasch, B.; Lee, S.-H.; McHugh, M. Strenghts and limitations
of SAFT for calculating polar copolymer-solvent phase behavior.
J. Appl. Polym. Sci. 1994, 59, 1107.
(3) Lee, M.; Chao, K. Augmented BACK EOS for polar fluids.
AIChE J. 1988, 34, 825-833.
(4) Chen, S.; Kreglewski, A. Applications of augmented van der
Waals theory of fluids. I. Pure fluids. Ber. Bunsen-Ges. 1977, 51,
1048.
(5) Gray, C.; Gubbins, K. Theory of Molecular Fluids; Clarendon
Press: Oxford, U.K., 1984; Vol. 1.
(6) Twu, C.; Gubbins, K. Thermodynamics of polyatomic fluid
mixtures. II. Polar, quadrupolar and octopolar molecules. Chem.
Eng. Sci. 1978, 33, 879.
(7) Vimalchand, P.; Donohue, M. Thermodynamics of quadrupolar molecules: the perturbed-anisotropic-chain theory. Ind. Eng.
Chem. Res. 1985, 24, 246.
(8) Gubbins, K.; Gray, C. Perturbation Theory for the Angular
Pair Correlation Function in Molecular Fluids. Ind. Eng. Chem.
Res. 1972, 23, 187-191.
(9) Cotterman, R.; Schwartz, B.; Prausnitz, J. Molecular thermodynamics of fluids at low and high densities. Part I: Pure fluids
containing small and large molecules. AIChE J. 1986, 32, 17871798.
(10) Cotterman, R.; Schwartz, B.; Prausnitz, J. Molecular
thermodynamics of fluids at low and high densities. Part II: phase
equilibria for mixtures containing components with large differences in molecular size or potential energy. AIChE J. 1986, 32,
1799-1812.
(11) Xu, K.; Li, Y.-G.; Liu, W.-B. Application of perturbation
theory to chain and polar fluids: pure alkanes, alkanols and water.
Fluid Phase Equilib. 1998, 142, 55-66.

(12) Kraska, T.; Gubbins, K. Phase Equilibria Calculations with


a Modified SAFT EOS. 1. Pure Alkanes, Alkanols and Water. Ind.
Eng. Chem. Res. 1996, 35, 4727-4737.
(13) Kraska, T.; Gubbins, K. Phase Equilibria Calculations with
a Modified SAFT EOS. 2. Binary Mixtures of n-Alkanes, nAlkanols and Water. Ind. Eng. Chem. Res. 1996, 35, 4738-4746.
(14) Tang, Y.; Wang, Z.; Lu, B.-Y. Thermodynamic calculations
of linear chain molecules using a SAFT model. Mol. Phys. 2001,
399, 65-76.
(15) Walsh, J.; Gang, J.; Donohue, M. Thermodymamics of short
chain polar compounds. Fluid Phase Equilib. 1991, 65, 209.
(16) Sear, R. Low-density fluid phase of dipolar hard spheres.
Phys. Rev. Lett. 1996, 76, 2310.
(17) Saager, B.; Fischer, J. Construction and application of
physically based EOS. Fluid Phase Equilib. 1992, 72, 67-88.
(18) Saager, B.; Fischer, J.; Neumann, M. Reaction Filed
simulations of monatomic and diatomic dipolar fluids. Mol. Simul.
1991, 6, 27-49.
(19) Muller, A.; Winkelmann, A.; Fischer, J. Simulation studies
on mixtures of dipolar with nonpolar linear molecules. Fluid Phase
Equilib. 1996, 120, 107-119.
(20) Jog, P.; Chapman, W. Application of Wertheims thermodynamic perturbation theory to dipolar hard-sphere chains. Mol.
Phys. 1999, 97, 307-319.
(21) Jog, P.; Sauer, S.; Blaesig, J.; Chapman, W. G. Application
of Dipolar Chain Theory to the Phase Behavior of Polar Fluids
and Mixtures. Ind. Eng. Chem. Res. 2001, 40, 4641-4648.
(22) Sauer, S.; Chapman, W. G. A Parametric Study of Dipolar
Chain Theory with Applications to Ketone Mixtures. Ind. Eng.
Chem. Res. 2003, 42, 5687-5696.
(23) Tumakaka, F.; Sadowski, G. Application of PC-SAFT EOS
to polar systems. Fluid Phase Equilib. 2004, 217, 234-239.
(24) Calero, S.; Weinland, M.; Fischer, J. Description of alternative refrigerants with Backone equations. Fluid Phase Equilib.
1998, 152, 1-22.
(25) Weingerl, U.; Weinland, M.; Fischer, J.; Muller, A. Backone
family of equations of state: 2. Nonpolar and polar fluid mixtures.
AIChE J. 2001, 47, 705-717.
(26) Weingerl, U.; Fischer, J. Consideration of dipole-quadrupole interactions in molecular based EOS. Fluid Phase Equilib.
2002, 202, 49-66.
(27) Gross, J.; Sadowski, G. Perturber-Chain SAFT: An Equation of State Based on a Perturbation Theory for Chain Molecules.
Ind. Eng. Chem. Res. 2001, 40, 1244-1260.
(28) Tumakaka, F.; Gross, J.; Sadowski, G. Modeling of polymer
phase equilibria using Perturbed-Chain SAFT. Fluid Phase Equilib.
2002, 194-197, 541-551.
(29) Gross, J.; Spuhl, O.; Tumakaka, F.; Sadowski, G. Modeling
of Copolymer Systems Using Perturbed-Chain SAFT Equation of
State. Ind. Eng. Chem. Res. 2003, 42, 1266-1274.
(30) Carnahan, N.; Starling, K. An equation of state for the
hard-sphere chain fluid: theory and Monte Carlo simulation. J.
Chem. Phys. 1969, 51, 635-636.
(31) Chapman, W.; Gubbins, K.; Jackson, G. Phase equilibria
of associating fluids: chain molecules with multiple bonding sites.
Mol. Phys. 1988, 65, 1057-1079.
(32) Chapman, W.; Gubbins, K.; Jackson, G.; Radosz, M. New
Reference Equation of State for Associating Liquids. Ind. Eng.
Chem. Res. 1990, 29, 1709-1721.
(33) Ghosh, A.; Blaesing, J.; Jog, P.; Chapman, W. G. Perturbed
Dipolar Chains: A Thermodynamic Model for Polar Copolymers.
Macromolecules 2005, 38, 1025-1027.
(34) Daubert, T. E.; Danner, R. P.; Sibul, H. M.; Stebbins, C.
Physical and Thermodynamic Properties of Pure Chemicals: Data
Compilation; Taylor and Francis: Washington, DC, 1989.
(35) Design Institute for Physical Property Data (DIPPR);
American Institute of Chemical Engineers (AIChE): New York,
1989.
(36) Reid, R.; Prausnitz, J.; Poling, B. Properties of Gases and
Liquids; McGraw-Hill: New York, 1987.
(37) Gmehling, J.; Li, J.; Schiller, M. A Modified UNIFAC
Model. 2. Present Parameter Matrix and Results for Different
Thermodynamic Properties. Ind. Eng. Chem. Res. 1993, 32, 178193.
(38) Gmehling, J.; Onken, U.; Rarcy, J. R. Vapor-Liquid Equilibrium Data Collection; Dechema Chemistry Data Series, Vol. 1,
Part 4a; Dechema: Flushing, NY, 1977.

6938

Ind. Eng. Chem. Res., Vol. 44, No. 17, 2005

(39) Gmehling, J.; Onken, U. Vapor-Liquid Equilibrium Data


Collection; Dechema Chemistry Data Series, Vol. 1, Part 5b;
Dechema: Flushing, NY, 1977.
(40) Horstmann, S.; Birke, G.; Fischer, K. VLE and excess
enthalpy data for the binary systems Propane + dimethyl ether
and propene + dimethyl ether at temperatures from 298 to 323
K. J. Chem. Eng. Data 2004, 49, 38-42.
(41) Giles, N.; Wilson, G. Phase equilibria on seven binary
mixtures. J. Chem. Eng. Data 2000, 45, 146-153.
(42) Shealy, G.; Sandler, S. J. Chem. Thermodyn. 1985, 17, 143.
(43) Mozo, I.; Gonzalez, J.; de la Fuente, I. G.; Cobos, J.
Thermodynamics of Mixtures Containing Ethers. Part III. LiquidLiquid Equilibria for 2,5,8,11-Tetraoxadodecane or 2,5,8,11,14Pentaoxapentadecane + Selected N-Alkanes. J. Chem. Eng. Data
2004, 49, 1091-1094.
(44) Treszczanowicz, T.; Cieslak, D. LLE in a dimethyl ether
of a poly(ethylene glycol) and an alkane. J. Chem. Thermodyn.
1993, 25, 661-665.
(45) Treszczanowicz, T.; Lu, B. C. Y. Isothermal Vapor-liquidequilibria for 11 examples of (an ether + a Hydrocarbon). J. Chem.
Thermodyn. 1986, 18, 213-220.

(46) Treszczanowicz, T. Bull. Acad. Pol. Sci., Ser. Sci. Chim.


1973, 21, 107.
(47) Klon-Placzewska, M.; Cholinski, J.; WyrzykowskaStankiewicz, D. Chem. Stosow. 1980, 24, 197.
(48) Burgdorf, R.; Zocholl, A.; Arlt, W.; Knapp, H. Thermophysical properties of binary liquid mixtures of polyether and
n-alkane at 298.15 and 323.15 K: heat of mixing, heat capacity,
viscosity, density and thermal conductivity. Fluid Phase Equilib.
1999, 164, 225-255.
(49) Dean, J. A. Langes Handbook of Chemistry; McGrawHill: New York, 1992.
(50) Kimura, K.; Fujishiro, R. Bull. Chem. Soc. Jpn. 1966, 39,
608-610.

Received for review January 18, 2005


Revised manuscript received May 17, 2005
Accepted June 7, 2005
IE050071C

Das könnte Ihnen auch gefallen