You are on page 1of 18

A Comparison of the Peng-Robinson and Soave-Redlich-Kwong

Equations of State Using a New Zero-Pressure-Based Mixing Rule for


the Prediction of High Pressure and High Temperature Phase Equilibria

*
Chorng H. Twu , John E. Coon, and David Bluck
Simulation Sciences Inc., 601 Valencia Avenue, Brea, CA 92823 (USA)

Keywords: mixing rule, excess free energy, equation of state, cubic, PR, SRK,
liquid activity coefficient, NRTL equation

Abstract

The Peng-Robinson (PR) and Soave-Redlich-Kwong (SRK) equations of


state are probably the most widely used cubic equations of state in the refinery
and gas processing industries for the prediction of vapor-liquid equilibria for
systems containing non-polar components. The new mixing rules which have
recently been developed that combine liquid activity models with the equations
of state, however, have extended the application of such equations to highly
non-ideal systems. A new zero-pressure-based mixing rule is presented here
that reproduces, with extremely high accuracy, the excess Gibbs free energy as
well as the liquid activity coefficients of any activity model without requiring any
additional binary interaction parameters. We examine the performance of the
Peng-Robinson and Soave-Redlich-Kwong equations of state using the NRTL
liquid activity model with binary parameters determined at low temperature in
this new mixing rule, MHV1, and Wong-Sandler for the prediction of high
pressure and high temperature phase equilibria.

Introduction
Among many equations of state proposed for predicting phase behavior of
non-polar systems, cubic equations of state have been widely used because of
their simplicity and accuracy. Among cubic equations of state, two equations
which have enjoyed widespread acceptance in the refinery and gas processing
industries are the Soave (1972) and the Peng-Robinson (1976) equations of

*
Corresponding author. e-mail: ctwu@simsci.com, Tel: (714) 985-5298, Fax: (714) 579-0113
state. This paper therefore will focus on these two equations of state. One of
the most frequent questions asked by engineers is which equation of state, PR
or SRK, should be selected for phase equilibrium prediction. To answer this
question, we will compare the performance of extended PR and SRK equations
on the prediction of phase equilibria for non-ideal mixtures.

The application of the PR or SRK equation of state to systems containing


highly non-ideal components requires an appropriate mixing rule for the equation
of state parameter a. Huron and Vidal (1979) pioneered linking the equation of
state parameter a to the excess Gibbs free energy at infinite pressure. However,
their mixing rule has not become widely used because the available excess
Gibbs energy parameters at low pressure cannot be used in their mixing rule.

In view of that, several authors have proposed different approaches to


directly use the existing liquid activity model parameters in equations of state.
Among them, two models have been quite successful. One is a zero-pressure
model by Michelsen (1990a, 1990b), Dahl and Michelsen (1990), and
Heidemann and Kokal (1990) to equate the excess Helmholtz free energy (AE) at
zero pressure from an equation of state to that from an activity model and the
other is an infinite-pressure model by Wong and Sandler (1992) to force the AE
at infinite pressure derived from a cubic equation of state to be equal to that
from a liquid activity coefficient mode. Both models can directly use available
activity coefficient model parameters from low-pressure data in their mixing
rules for predicting phase equilibria at high temperatures and pressures quite
successfully. However, neither the zero pressure models such as MHV1 or
MHV2 nor infinite pressure approaches such as Huron-Vidal or Wong-Sandler
can accurately reproduce the GE model with which it is combined (Coutsikos et
al., 1995; Kalospiro et al., 1995).

A methodology has been developed to extend the infinite-pressure Twu-Coon


mixing rule (1996) to correctly reproduce the incorporated GE model without
introducing any additional binary interaction parameters. With this capability,
the available activity coefficient models at low pressure can be used directly in
this new mixing rule. We will insert the NRTL excess Gibbs free energy model
into the PR or SRK equations. The comparison of the PR or SRK equations
will be made on the same exactly basis (i.e. the same GE model and the same
binary interaction parameters reported in the DECHEMA Chemistry Data Series

2
for the GE model, without using any other regressed binary interaction
parameters, are used in both PR or SRK equations of state). The mixing rules
included in this comparison are our new mixing rule, MHV1 and the Wong-
Sandler mixing rules.

Extended Twu-Coon Mixing Rule


A two-parameter cubic equation of state is considered here:
RT a
P= − (1)
v − b ( v + ub )( v + wb )
where P is the pressure, T is the absolute temperature, and v is the molar
volume. The constants u and w are equation of state dependent (for the PR
equation: u = -0.4142, w = 2.4141 ; for the SRK equation: u=0, w=1). The
parameter a in eqn.(1) is a function of temperature and the parameter b is
assumed to be a constant for pure components. The value of a(T) at
temperatures other than the critical temperature, ac, can be calculated from

a( T ) = α( T ) ac (2)
The alpha function, α(T) in eqn.(2), is a function only of reduced
temperature, Tr=T/Tc. Since the prediction of pure component vapor pressure
must be of high accuracy for accurate vapor-liquid calculations, we have chosen
to use the alpha correlation of Twu et al. (1991):

α =Tr N ( M − 1 ) e L ( 1−T r
NM
)
(3)

Eqn.(3) has three parameters, L, M, and N which are unique to each


component and are determined from the regression of pure component vapor
pressure data. The values for the components used in this study are given in
Tables 1 and 2 for the PR and SRK equations, respectively. Consequently, the
PR and SRK equations will not suffer in accuracy by comparison if the alpha
correlation can reproduce the same accuracy for pure component vapor
pressure from the triple point to the critical point.

Twu and Coon (1996) have related the excess Helmholtz free energy, AE, with
respect to a van der Waals fluid to the Helmholtz free energy departure function,
∆A, by the following:

3
AE − AEvdw = ∆A − ∆Avdw (4)

Eqn.(4) was used by Twu and Coon (1996) to derive the following mixing
rule for the cubic equation of state mixture a and b parameters at infinite
pressure:

A∞E A∞Evdw  a* a* 
− = C1 * − vdw  (5)
RT RT b *
bvdw 

with the C1, avdw and bvdw being:

1  1+ w 
C1 = − ln  (6)
( w − u )  1 + u 

avdw = ∑ ∑ xi xj ai aj (7)
i j

1 
bvdw = ∑∑ xi xj  (bi + bj ) (8)
i j 2 
AE∞ and AE∞vdw in eqn.(5) are the excess Helmholtz energy at infinite pressure
evaluated from a cubic equation of state using the complete mixing rules for its a
and b parameters and using the van der Waals mixing rules for its a and b
parameters (avdw and bvdw), respectively.

If eqn.(4) is applied at zero pressure, instead of infinite pressure, an equation


containing liquid volume is obtained:

A0 − A0vdw = ln  v0*vdw −1  bvdw   − 1  a* ln v0* +w  − avdw  v0*vdw +w  


E E *

 *    *  *  * ln  
RT RT  v0 − 1  b   ( w − u )  b  v0 +u  bvdw  v0*vdw +u  
(9)

A0E and v0*=v0/b are the excess Helmholtz energy and reduced liquid volume
at zero pressure. As mentioned, the subscript vdw denotes that the properties
are evaluated from the cubic equation of state using the van der Waals mixing
rule for its a and b parameters. The zero pressure volume is obtained from
eqn.(1) by setting pressure equal to zero and selecting the smallest root:

4
1
2
a* 2 
v0* =  * − u − w −  u + w − *  − 4 uw + *  
1  a *  a*
(10)
2  b   b   b  

Eqn.(10) has a root as long as


a*
≥ (2 + u + w) + 2 (u + 1)( w +1) (11)
b*

Eqns.(9) and (10) represent an exact model for a new mixing rule. However,
because the equation of state parameter a* /b* and the zero pressure liquid
volume are interrelated by eqns.(9) and (10), the exact model does not permit
explicit solution of eqn.(9) for a* /b* and an iterative technique is required for the
solution. If eqns.(9) and (10) are used to solve for a* /b* , the resulting new
mixing rule will give an exact match between the excess Helmholtz free energy of
the equation of state at zero pressure and that of the incorporated excess Gibbs
free energy model. Nevertheless, the non-explicit nature of the expression for
the mixing rule becomes cumbersome in the evaluation of thermodynamic
properties such as fugacity coefficients from the equation of state. This paper
presents a methodology to overcome this obstacle to obtain an explicit
expression for this new mixing rule.

A variety of alternatives have been proposed to simplify the exact model so


that the equation of state parameters, a and b, can be explicitly expressed
(Michelsen, 1990b; Dahl and Michelsen, 1990). However, these modifications
of the exact model sacrifice to some extent the quality of the match between the
equation of state and the GE model. For example, the MHV1 model developed
by Michelsen (1990b) can be alternately derived from our new mixing rule by
assuming v0* is a constant (1.23547 for SRK and 1.22756 for PR), instead of
solving for it from eqn.(10). This means that the MHV1 model assumes that the
ratio of the zero pressure liquid volume to the close packing parameter b is the
same for the mixture and for all pure components. This is the main reason why
the MHV1 model performs poorly in reproducing the behavior of the GE model
when applied to systems either with components that are different in size or
where the value of v0* of the system is not close to the fixed values given above.

In this paper, v0* will not be assumed to be constant. We propose instead


that the ratio of the zero pressure liquid volume to the close packing parameter

5
b of the system, v0*, be assumed to be the same as that of the van der Waals
fluid, v0*vdw. Using eqns.(7) and (8) for the parameters a and b in eqn.(10),
v0*vdw can be readily calculated from the equation. Eqn.(10) is used to calculate
v0*vdw for both the mixture and the pure components. Eqn.(9) can then be
simplified to
E
A0 AE0vdw b  a* a* 
− = ln  vdw + Cv  * − vdw  (12)
RT RT  b  b 
* 0
bvdw

with the density function Cv0 being:

1  v0 * +w 
Cv = − ln  *  (13)
( w − u)  v0 +u  vdw
0

Since the equation of state parameters, a and b, are pressure independent,


these two parameters can be canceled out from eqns.(5) and (12) to give the
inter-relationship of AE between infinite pressure and zero pressure as:
E
A∞E A∞ vdw C1  A0E A0Evdw  b vdw  
− =  − − ln  (14)
RT RT Cv  RT RT 0
 b 

Substituting eqn.(14) into the mixing rule proposed by Twu and Coon (1996)
results in a new and explicit mixing rule in terms of AE0 at zero pressure:

− avdw
*
bvdw *

b* = (15)
a * 1  A0
E
AE0vdw  b vdw   
1−  vdw +  − − ln    
 bvdw
*
Cv  RT0 RT  b   

 avdw 1  AE0  bvdw   


E
*
A0vdw
a =b  * +
* *
 − − ln    (16)
 bvdw Cv0  RT RT  b   

As mentioned above, A0Evdw in eqns.(15) and (16) is the excess Helmholtz


energy at zero pressure evaluated from a cubic equation of state using the mixing
rules for its avdw and bvdw parameters, as given in eqns.(7) and (8). The zero
pressure volume v0*vdw=v0vdw/b is obtained from eqn.(10) by substituting avdw
and bvdw for the a and b parameters.

6
There are some nice features of this new mixing rule. The new mixing rule
reduces to the van der Waals mixing rule when A0E is equal to A0Evdw. The
mixing rule satisfies the quadratic composition dependence of the second virial
coefficient boundary condition. The most important aspect is that the mixing
rule is density dependent in an explicit form which allows the mixing rule to
reproduce accurately the incorporated GE model.

Incorporation of the NRTL Activity Model into the New Mixing Rule
Since A0E in eqns.(15) and (16) is at zero pressure, its value is identical to the
excess Gibbs free energy at zero pressure GE. Therefore, any activity model
such as the NRTL equation can be used directly for the excess Helmholtz free
energy expression A0E in these two equations.

For a solution of n components, NRTL equation is:


n

G
E n ∑ xjτjiGji
= ∑ xi n
j
(17)
RT
∑ xkGki
i

with τij and Gij defined as:


Aji
τ ji = (18)
T
Gji = exp ( −αji τji ) (19)
The NRTL parameters, Aij, Aji and αij, obtained from the lowest temperature
of each binary reported in the DECHEMA Chemistry Data Series, are used
directly in the mixing rule models. These values of Aij, Aji, and αij are then used
in the mixing rules at all temperatures, where temperature T is in Kelvin. The
values of the NRTL binary interaction parameters for the binaries used in this
study are given in Table 3.

In this work, we have considered ten highly non-ideal binary mixtures which
are traditionally described by liquid activity models. They are listed in Table 3.
In order to obtain liquid-like values for v0* at zero pressure from eqn.(10), we
have limited our analysis to systems with components and temperatures such
that a* /b* is larger than the limiting value of 5.82843 for SRK and 6.82843 for

7
the PR equation of state. By doing this, we have eliminated the need to include
an extrapolation methodology into our comparisons.

The mixing rule for the parameter b, as given by eqn.(15), forces the mixing
rule to satisfy the quadratic composition dependence of the second virial
coefficient. Alternatively, the conventional linear mixing rule could be chosen
for the b parameter (i.e. ignoring the second virial coefficient boundary
condition):

1 
b = ∑∑ xi xj  (bi + bj ) (20)
i j 2 
We will examine the capability of this mixing rule for phase equilibrium
prediction with and without the second virial coefficient constraint on the b
parameter. We will also compare our new mixing rule with two of the most
successful and widely used mixing rules, MHV1 (Michelsen, 1990b) and the
mixing rule proposed by Wong and Sandler (1992). Recently, Orbey and
Sandler (1997) have made a comparison of a number of Huron-Vidal type
mixing rules, including MHV1 and MHV2, for mixtures of molecules with large
differences in size. They conclude while there are differences between the
models, none was clearly superior to the others for all mixtures, though the
MHV2 was the least accurate. Since the MHV2 is the least accurate model, the
MHV1, instead of the MHV2, is selected for the comparison in this work.

Wong and Sandler assumed that the excess Helmholtz free energy at infinite
pressure can be approximated by the excess Gibbs free energy at low pressure:

AE(T, x, P=∞)= AE(T, x, P=low)= GE(T, x, P=low) (21)

The Wong-Sandler approximation will be tested in this comparison to see


how well the assumption in eqn.(21) stands. Another objective in this paper is
to test the ability of different mixing rules to reproduce the incorporated GE
model. In this work, we have performed rigorous tests of the capability of
reproducing the GE model using the PR and SRK equations of state combined
with our new mixing rules. We use ‘WS’ to refer to the Wong-Sandler mixing
rules. ‘TCB’ is used to represent the mixing rule developed by us in this work
(eqns.15 and 16), and ‘TCB(0)’ to eqns.(16) and (20). The zero in the TCB
parenthesis means no second virial coefficient constraint. The accuracy of

8
reproducing the activity coefficients of component i, γi in terms of average
absolute deviation percentage (AAD%), from the incorporated GE model using
these different mixing rules in the SRK and PR equations is given in Table 3.
Similarly, the accuracy of the VLE prediction from the different mixing rules,
expressed in terms of AAD% in bubble point pressure and K-values of
component 1 and 2, is also presented in Table 3 for the PR and SRK equations.

The activity coefficient of component i in a mixture can be derived from the


fugacity coefficients of the equation of state:
φi
γi =
φ io (22)

where γi is the activity coefficient of component i at the equilibrium temperature


and pressure, φi is the fugacity coefficient of component i in the mixture, and φio
is the fugacity coefficient of pure component i at the same temperature and
pressure.

Examining the accuracy of reproducing the activity coefficients, as given in


Table 3, the PR and SRK equations of state produce almost identical results for
all four different mixing rules. Table 3 also compares the VLE predictions of
the cubic equations of state. The comparisons show that there is little difference
in the accuracy of the predictions with these two methods. Based on these
results, it seems to indicate that both equations of state are equivalent to each
other and neither one has an advantage over the other in phase equilibrium
calculations as long as the alpha correlation of Twu et al. (1991) is used.

For the mixing rule comparison, the Wong-Sandler mixing rule gives the
largest deviation for all the systems in this comparison. The inability to match
the GE derived from the equation of state with that from the incorporated GE
model invalidates the basic assumption behind the Wong-Sandler mixing rule.
The predictions from the Wong-Sandler mixing rule without using any additional
binary interaction parameters is unacceptable. Table 3 contains the results for
the systems acetone-benzene and acetone-methanol. The zero pressure model,
MHV1, reproduces closely the GE model it is combined with for systems where
the liquid volume of the components is close to the fixed value chosen by
Michelsen (1990b). It is not surprising that if the liquid volume of the system is
not close to the fixed constant, the GE model is not reproduced by the equation

9
of state using the MHV1 mixing rule. These results are in agreement with a
statement by Kalospiro et al. (1995), although we have a different explanation for
it. The results shown in Table 3 illustrate that our new mixing rule reproduces
the GE model almost exactly. The errors in the reproduction of the activity
coefficients for these systems are minimal from our mixing rule.

For the VLE predictions, our new mixing rule gives consistent results and, in
general, provides good agreement between the experimental data and the
predictions over a wide range of temperatures and pressures using only the
information in the GE model. It was somewhat surprising that good agreement
was also obtained from MHV1, although it cannot reproduce accurately the
incorporated liquid activity coefficients. This unexpected result from MHV1
might come from some mutual cancellation of errors, but we do not know at this
time. Again, the worst predictions are obtained from the Wong-Sandler mixing
rule because of its inability to match the GE model.

As we mentioned, we want to investigate the impact on phase equilibrium


prediction of the mixing rule with and without the second virial coefficient
condition constraint. Reviewing the results shown in Table 3, they show that
our mixing rule yields almost identical results either with or without the second
virial coefficient condition constraint. This indicates that the second virial
coefficient constraint has no effect on the phase equilibrium prediction.
Theoretically, it would be nice to have the mixing rule satisfy the quadratic
composition dependence of the second virial coefficient boundary condition.
Practically, it is simpler just to use the conventional linear mixing rule for the b
parameter. The same quality of phase behavior will be predicted from both
cases.

Conclusion
We have successfully extended the Twu-Coon Mixing Rule from infinite
pressure to zero pressure. Incorporating a CEOS/AE mixing rule into either the
PR or SRK equation of state results in similar accuracy in the phase equilibrium
prediction. Therefore, as long as the same CEOS/AE mixing rule is used,
engineers can use either PR or SRK in design calculations without worrying
about the discrepancies between the equation of state methods. We also
demonstrate that CEOS/AE models such as the Wong-Sandler mixing rule do

10
not reproduce the GE models with which they are associated. We show why the
zero pressure models do not reproduce exactly the GE models at low pressure
and reveal that approximate reproduction is feasible for MHV1 only for systems
with liquid volumes close to the assumed constant volume. On the other hand,
the new model we developed in this work accurately reproduces the activity
coefficients of the GE model.

List of Symbols
a, b cubic equation of state parameters
* *
a,b reduced parameters of a and b
A Helmholtz free energy
Aij, Aji NRTL binary interaction parameters in degree kelvin
C1 infinite pressure function defined in eqn.(6)
Cv0 zero pressure function defined in eqn.(13)
G Gibbs free energy
Ki K value of component i defined as yi / xi
L, M, N parameters in the α function defined in eqn.(3)
P pressure
R gas constant
T temperature
u, w cubic equation of state constants
v molar volume
*
v0 reduced zero pressure liquid volume
xi mole fraction of component i
Z compressibility factor

Greek letters
α cubic equation of state alpha function defined in eqn.(2)
φi fugacity coefficient of component i in the mixture
γi activity coefficient of component i
∆ departure function

Subscripts
0 zero pressure
∞ infinite pressure
c critical property
i, j property of component i, j

11
ij interaction property between components i and j
vdw van der Waals

Superscripts
o pure component
* reduced property
E excess property

References
Coutsikos, P., Kalospiros, N.S. and Tassios, D.P. Capabilities and Limitations
of the Wong-Sandler Mixing Rules, Fluid Phase Equilibria, 1995, 108, 59-78.
Dahl, S. and Michelsen, M.L. High-pressure Vapor-liquid Equilibrium with a
UNIFAC-based Equation of State, AIChE J., 1990, 36, 1829-1836.
Heidemann, R.A. and Kokal, S.L. Combined Excess Gibbs Energy Models and
Equations of State, Fluid Phase Equilibria, 1990a, 56, 17-37
Huron, M.J. and Vidal, J. New Mixing Rules in Simple Equations of State for
Representing Vapor-Liquid Equilibria of Strongly Nonideal Mixtures, Fluid
Phase Equilibria, 1979, 3, 255-271.
Kalospiros, N.S., Tzouvaras, N., Coutsikos, P. and Tassios, D.P. Analysis of
Zero-reference-pressure EOS/G E Models, AIChE J., 1995, 41, 928-937.
Michelsen, M.L. A Method for Incorporating Excess Gibbs Energy Models in
Equation of State, Fluid Phase Equilibria, 1990a, 60, 47-58
Michelsen, M.L. A Modified Huron-Vidal Mixing Rule for Cubic Equations of
State, Fluid Phase Equilibria, 1990b 60, 213-219
Orbey, H. and Sandler, S.L. A Comparison of Huron-Vidal type Mixing Rules
of Mixtures of Compounds with Large Size Differences, and A New Mixing
Rule, Fluid Phase Equilibria, 1997, 132, 1-14.
Peng, D-Y. and Robinson, D.B. A New Two-constant Equation of State, Ind.
Eng. Chem. Fundam., 1976, 15, 59-64.
Soave, G. Equilibrium Constants from a Modified Redlich-Kwong Equation of
State, Chem. Eng. Sci., 1972, 27, 1197-1203.
Twu, C.H., Bluck, D., Cunningham, J.R., and Coon, J.E. A Cubic Equation of
State with a New Alpha Function and a New Mixing Rule, Fluid Phase
Equilibria, 1991, 69, 33-50.

12
Twu, C.H. and Coon, J.E. CEOS/AE Mixing Rules Constrained by the vdW
Mixing Rule and the Second Virial Coefficient, AIChE J., 1996, 42, 3212-
3222.
Wong, S.H. and Sandler, S.I. A Theoretically Correct Mixing Rule for Cubic
Equations of State, AIChE J., 1992, 38, 671-680.

13
Table 1
L, M, and N parameters of the temperature-dependent α function given by
equation 3 for pure components used with the PR cubic equation of state
Component Tc (K) Pc (bar) L M N
n-pentane 469.70 33.70 0.331853 0.829974 1.65879
n-hexane 507.85 30.31 0.132963 0.863991 3.37934
n-heptane 540.16 27.36 0.507729 0.823369 1.45872
cyclohexane 553.58 40.73 0.140885 0.844513 2.48153
benzene 562.16 48.98 0.0810842 0.877078 3.59357
acetone 508.20 47.01 0.598248 0.882439 1.27592
methanol 512.64 80.97 0.516917 0.873183 2.14538
ethanol 513.92 61.48 2.66593 4.82235 0.138949
water 647.13 220.55 0.386676 0.870460 1.96014

14
Table 2
L, M, and N parameters of the temperature-dependent α function given by
equation 3 for pure components used with the SRK cubic equation of state
Component Tc (K) Pc (bar) L M N
n-pentane 469.70 33.70 0.379229 0.841706 1.82331
n-hexane 507.85 30.31 0.158080 0.872819 3.84418
n-heptane 540.16 27.36 0.340339 0.844963 2.38332
cyclohexane 553.58 40.73 0.245880 0.845046 2.25895
benzene 562.16 48.98 0.163664 0.860016 2.98498
acetone 508.20 47.01 0.479844 0.870627 1.79010
methanol 512.64 80.97 0.690551 0.911298 1.96941
ethanol 513.92 61.48 1.07646 0.964661 1.35369
water 647.13 220.55 0.413297 0.874988 2.19435

15
Table 3
NRTL interaction parameters and results of the prediction in terms of average absolute deviation
percentage (AAD%) in activity coefficients, bubble point pressure and K-values with respect to PR
(1st value) and SRK (2nd value) equations of state, respectively.
a
ethanol(1)/n-heptane(2) from 30.12 to 70.02 C; I/2e/377, 379; I/2c/457, 458
A12=521.746, A21=727.003, α12=0.4598 at 30.12 C
mixing rule γ 1(%) γ 2(%) P(%) K1(%) K2(%)
WS 24.72 / 24.87 24.64 / 24.77 16.93 / 16.88 13.21 / 13.36 26.11 / 26.25
MHV1 4.25 / 3.46 4.52 / 3.78 2.21 / 2.02 1.78 / 1.49 2.89 / 2.62
TCB 0.10 / 0.11 0.06 / 0.06 1.23 / 1.27 1.37 / 1.27 2.33 / 2.13
TCB(0) 0.08 / 0.08 0.04 / 0.05 1.21 / 1.25 1.39 / 1.28 2.35 / 2.15
a
ethanol(1)/water(2) from 24.99 to 120 C; I/1b/93, 106, 107, 108
A12=13.3878, A21=437.683, α12=0.2945 at 24.99 C
mixing rule γ 1(%) γ 2(%) P(%) K1(%) K2(%)
WS 18.13 / 18.04 16.37 / 16.29 13.24 / 13.26 15.99 / 15.87 14.91 / 14.84
MHV1 6.24 / 5.70 4.55 / 4.14 4.01 / 4.15 6.02 / 6.17 4.17 / 4.23
TCB 0.26 / 0.25 0.17 / 0.17 4.19 / 4.33 7.17 / 7.19 5.30 / 5.36
TCB(0) 0.21 / 0.20 0.14 / 0.13 4.05 / 4.19 7.04 / 7.06 5.16 / 5.22
a
methanol(1)/cyclohexane(2) from 25 to 55 C; I/2a/242; I/2c/208, 209
A12=644.886, A21=784.966, α12=0.4231 at 25 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 25.16 / 25.34 25.02 / 25.19 19.00 / 19.27 15.84 / 16.01 24.28 / 24.53
MHV1 4.07 / 3.25 4.49 / 3.68 2.18 / 1.56 2.62 / 2.43 4.09 / 3.76
TCB 0.20 / 0.21 0.15 / 0.16 1.29 / 1.15 1.86 / 1.86 2.85 / 2.87
TCB(0) 0.15 / 0.16 0.12 / 0.12 1.29 / 1.15 1.84 / 1.85 2.82 / 2.84
a
methanol(1)/benzene(2) from 25 to 90 C; I/2c/188; I/2a/207,210,216,217,228
A12=441.228, A21=738.702, α12=0.5139 at 25 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 20.69 / 20.87 19.77 / 19.93 13.63 / 14.01 13.87 / 14.10 15.44 / 15.64
MHV1 3.71 / 3.03 3.45 / 2.86 4.17 / 3.49 7.02 / 6.62 4.87 / 4.57
TCB 0.17 / 0.17 0.13 / 0.13 3.20 / 3.00 5.89 / 5.88 3.99 / 4.00
TCB(0) 0.13 / 0.13 0.10 / 0.10 3.16 / 2.96 5.89 / 5.87 3.96 / 3.97
a
data taken from DECHEMA Chemistry Data Series by Gmehling, Onken, and Arlt; numbers
corresponding to volume/part/page.

16
Table 3 (continued)
NRTL interaction parameters and results of the prediction in terms of average absolute deviation
percentage (AAD%) in activity coefficients, bubble point pressure and K-values with respect to PR
(1st value) and SRK (2nd value) equations of state, respectively.
a
acetone(1)/benzene(2) from 25 to 45 C; I/3+4/194, 203, 208
A12=-35.4443, A21=193.289, α12=0.3029 at 25 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 4.25 / 4.32 4.08 / 4.15 2.39 / 2.55 2.15 / 2.18 4.34 / 4.37
MHV1 0.85 / 0.73 0.69 / 0.58 0.92 / 0.84 1.11 / 1.08 1.79 / 1.74
TCB 0.01 / 0.01 0.01 / 0.01 0.82 / 0.79 1.03 / 1.03 1.74 / 1.66
TCB(0) 0.01 / 0.01 0.01 / 0.01 0.81 / 0.79 1.03 / 1.03 1.75 / 1.66
a
acetone(1)/ethanol(2) from 32 to 48 C; I/2a/323, 324, 325
A12=24.3880, A21=224.395, α12=0.3007 at 32 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 4.98 / 5.02 4.93 / 4.98 3.36 / 3.49 3.04 / 3.11 2.58 / 2.63
MHV1 1.95 / 1.75 1.55 / 1.37 1.46 / 1.39 2.14 / 1.98 2.92 / 2.79
TCB 0.01 / 0.01 0.01 / 0.01 1.20 / 1.22 0.98 / 0.97 1.83 / 1.84
TCB(0) 0.01 / 0.01 0.01 / 0.01 1.20 / 1.22 0.98 / 0.97 1.81 / 1.82
a
acetone(1)/methanol(2) from 45 to 55 C; I/2a/75, 80, 81
A12=31.5237, A21=180.554, α12=0.3004 at 45 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 7.95 / 7.93 7.71 / 7.71 4.20 / 4.47 5.21 / 5.13 6.13 / 6.20
MHV1 0.92 / 0.74 0.62 / 0.45 0.51 / 0.59 1.00 / 0.97 0.82 / 0.86
TCB 0.00 / 0.00 0.00 / 0.45 0.49 / 0.62 0.92 / 0.87 0.84 / 0.86
TCB(0) 0.00 / 0.00 0.00 / 0.00 0.49 / 0.63 0.92 / 0.87 0.84 / 0.86
a
ethanol(1)/benzene(2) from 25 to 55 C; I/2a/398, 407, 415, 417, 418, 421, 422
A12=115.954, A21=584.473, α12=0.2904 at 25 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 14.84 / 14.80 12.92 / 13.06 10.20 / 10.41 12.17 / 12.32 9.99 / 10.13
MHV1 3.25 / 2.73 3.13 / 2.69 1.39 / 1.24 3.34 / 3.29 2.41 / 2.38
TCB 0.15 / 0.15 0.12 / 0.13 1.05 / 1.11 3.18 / 3.19 2.51 / 2.53
TCB(0) 0.12 / 0.12 0.10 / 0.10 1.06 / 1.12 3.20 / 3.21 2.53 / 2.54
a
data taken from DECHEMA Chemistry Data Series by Gmehling, Onken, and Arlt; numbers
corresponding to volume/part/page.

17
Table 3 (continued)
NRTL interaction parameters and results of the prediction in terms of average absolute deviation
percentage (AAD%) in activity coefficients, bubble point pressure and K-values with respect to PR
(1st value) and SRK (2nd value) equations of state, respectively.
a
methanol(1)/water(2) from 24.99 to 100 C; I/1b/29; I/1/41, 49, 72, 73
A12=-23.1150, A21=188.147, α12=0.3022 at 24.99 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 6.16 / 6.00 5.51 / 5.37 5.13 / 5.21 7.06 / 6.94 8.75 / 8.66
MHV1 3.58 / 3.37 3.03 / 2.85 2.23 / 2.32 3.56 / 3.67 3.56 / 3.66
TCB 0.16 / 0.16 0.13 / 0.13 2.81 / 2.97 4.77 / 4.78 5.02 / 5.05
TCB(0) 0.13 / 0.13 0.11 / 0.10 2.78 / 2.97 4.72 / 4.73 4.91 / 4.93
a
methanol(1)/n-hexane(2) from 25 to 45 C; I/2c/219; I/2a/252
A12=823.172, A21=848.519, α12=0.4388 at 25 C
mixing rule γ 1(%) γ 2(%) P(%) K 1(%) K 2(%)
WS 27.63 / 27.73 27.69 / 27.77 25.54 / 25.52 33.88 / 33.94 31.51 / 31.54
MHV1 3.33 / 2.47 5.06 / 4.12 3.62 / 3.13 2.41 / 1.99 2.19 / 1.85
TCB 0.12 / 0.12 0.05 / 0.05 1.73 / 1.82 1.26 / 1.23 1.13 / 1.13
TCB(0) 0.09 / 0.10 0.04 / 0.04 1.67 / 1.76 1.28 / 1.23 1.16 / 1.13
a
data taken from DECHEMA Chemistry Data Series by Gmehling, Onken, and Arlt; numbers
corresponding to volume/part/page.

18

You might also like