Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Fluid Phase Equilibria 398 (2015) 36–45

Contents lists available at ScienceDirect

Fluid Phase Equilibria


j o u r n a l h o m e p a g e : w w w . e l s e v i e r . c o m / l o c a t e / fl u i d

Regression of NRTL parameters from ternary liquid–liquid equilibria


using particle swarm optimization and discussions
Zheng Li, Kathryn H. Smith, Kathryn A. Mumford, Yong Wang, Geoffrey W. Stevens *
Particulate Fluids Processing Centre, Department of Chemical and Biomolecular Engineering, The University of Melbourne, Melbourne, Victoria 3010, Australia

A R T I C L E I N F O A B S T R A C T

Article history: The particle swarm optimization (PSO) method was successfully used to regress the nonrandom two-
Received 6 November 2014 liquid (NRTL) parameters from liquid–liquid equilibria (LLE) data and the resulting parameters had
Received in revised form 2 April 2015 smaller root-mean square deviations (RMSD) compared with literature values. Analysis of the results
Accepted 7 April 2015
reveals that multiple groups of parameters with sufficiently small RMSDs can be found for the same set of
Available online 8 April 2015
LLE data. The activities calculated using these parameters and their corresponding predicted mole
fractions can be far beyond the reasonable range of activity, clearly demonstrating that the NRTL model
Keywords:
does not represent the real activities of components with parameters regressed from LLE. In addition, five
Liquid–liquid equilibria
NRTL model
sets of random data that were normalized to resemble LLE data were correlated by the NRTL model
Particle swarm optimization equally well as correlations of regular LLE data. Therefore, the model does not have any advantage in
Activity correlating LLE data over correlating random data. These limitations of the NRTL model doubt the model’s
Random data theoretical validity and should be aware of when the model is used.
ã 2015 Elsevier B.V. All rights reserved.

1. Introduction obtaining the global minimum. To obtain global minimum of


RMSD, a robust algorithm is essential. The particle swarm
The nonrandom two-liquid (NRTL) model [1] has been used optimization method (PSO) [16] is a powerful optimization
widely to correlate and predict vapor–liquid equilibria (VLE) [2,3] method that has attracted extensive studies in recent years and
and liquid–liquid equilibria (LLE) [3–6] of non-electrolytes, which has been used across a wide range of applications [17]. Having a
have numerous applications including distillation and liquid– number of entities (particles) interacting with each other and
liquid extraction, respectively. Extensions to this model for sharing information in the search space of a function, the swarm as
predicting phase equilibria of systems containing polymers [7,8] a whole is more likely to move close to the global minimum of the
and electrolytes [9,10] have expanded the model’s versatility. function compared with the algorithms that have been used.
To predict phase equilibria of multiple component systems, it is Unlike correlation of VLE, which directly deals with the vapor
common practice to solve equations of equality of fugacities (for pressures of components that represent the absolute value of
VLE) [2,3] or activities (for LLE) [3,4,6] under constraints of mass fugacities, the LLE calculation equates the activities of each
balance with known binary interaction parameters of the NRTL component in two phases without considering the real activity of
model, which can be regressed from experimental data of either each component. As a result, the activities may not be equal to their
VLE or LLE. As a result, regressing the NRTL parameters is the real values even if good equalities are obtained. Therefore, a
precondition for predicting phase equilibria. While regression of discussion on correlation of LLE by the NRTL model is warranted.
interaction parameters of the NRTL model from VLE is straightfor- This study aims to illustrate the utilization of the PSO method to
ward, regressing parameters from LLE, which minimizes the regress the parameters of the NRTL model from ternary LLE and
difference of experimental and calculated mole fractions, is much discuss the capability of the model in representing the activities of
more complicated due to involvement of solving nonlinear components in LLE.
isoactivity equations. Although some researchers have obtained
small root-mean square deviations (RMSD) of mole fractions using 2. Parameter regression using the PSO method
various algorithms [11–15], none of these methods can guarantee
2.1. Particle swarm optimization

* Corresponding author. Tel.: +61 3 8344 6621; fax: +61 3 8344 8824. The PSO was first proposed by Kennedy and Eberhart for
E-mail address: gstevens@unimelb.edu.au (G.W. Stevens). optimization of nonlinear functions by simulating social behaviors

http://dx.doi.org/10.1016/j.fluid.2015.04.006
0378-3812/ ã 2015 Elsevier B.V. All rights reserved.
Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45 37
[(Fig._1)TD$IG]
such as bird flocks searching for corn [16]. It has since attracted
extensive studies from a wide variety of applications due to its
flexibility and ease to utilize and has undergone many variations in
algorithm [17,18]. The PSO optimizes a problem by iteratively
improving the locations of candidate solutions (particles) in the
search space. Each particle updates its location by evaluating the
objective function of its current location and the best known
location (pbest) and the best location of the entire particles (gbest).
The next iteration occurs after all the particles have been moved
and “pbest” and “gbest” have been updated. Gradually the particle
swarm is expected to move toward an optimum like a flock of birds
collectively foraging for food. The algorithm applied in this study is
as follows:

vi ¼ v  vi þ c1  R1  ðpbest  xi Þ þ c2  R2  ðgbest  xi Þ (1)

xi ¼ xi þ k  vi (2)
Fig. 1. Correlation of LLE data using Fa as objective function.
where vi is the moving velocity of the particle, xi is the current
location, v is inertia weight, c1 and c2 are acceleration factors, R1

[(Fig._2)TD$IG]

Fig. 2. Initial RMSD (a) pre-regressed particles; (b) random initial particles.

[(Fig._3)TD$IG]

Fig. 3. Decrease of RMSD with the number of iterations (a) iteration using pre-regressed particles; (b) iteration using random particles.
38 Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45

Table 1
Comparison of regression by the PSO method with literature values for ternary LLE systems.

System Method/software RMSD Reference

Literature This study Literature This study


Water(1) + phenol(2) + benzene(3) at 298.15 K Not given PSO 0.0036 0.0022 [12,p. 292]
Ethene tetrachloro(1) + 2-propanol(2) + water(3) at 303.15 K Levenberg–Marquardt PSO 0.0078 0.0045 [4,p. 148]
Heptane(1) + ethylbenzene(2) + methanol(3) at 293.15 K Simplex PSO 0.0029 0.0026 [14,p. 62]
Cyclohexane(1) + enthylbenzene(2) + sulfolane(3) at 303.15 K ASPEN Plus PSO 0.023 0.0034 [13,p. 1718]

and R2 are two random numbers distributed in the range of [0, 1], k and aji (aji = aij) are the energy parameter and non-randomness
is constriction factor. Values of the coefficients used in this study parameter, respectively.
were selected following Eberhart and Shi [19]. Two types of objective functions [11] are usually used to regress
the parameters:

2.2. Optimization method 3 X


X M  2
Fa ¼ xIik  g Iik  xIIik  g IIik (8)
i¼1 k¼1
The isoactivity equations of components in two phases are
applied for calculation of LLE of ternary systems:
3 X
X M 
2 X 2
xIi g Ii ¼ xIIi g IIi ði ¼ 1; 2and3Þ (3)
Fx ¼ xjik  ^
x jik (9)
where xi and g i are the mole fraction and activity coefficient of i¼1 j¼1 k¼1

component i, respectively, I and II represent the two phases.


where i is component, j is phase and k is the number of tie lines, xjik
Mass balance equations of components are the constraints of
isoactivity equations and can be represented as follows: and ^ x jik are the experimental and calculated mole fraction of
components, respectively.
nIi þ nIIi ¼ ni ði ¼ 1; 2and3Þ (4) While minimization of Eq. (8) is straightforward, minimization
where ni is the total moles of component i. For the prediction of of Fx (Eq. (9)) is complex as it involves solving isoactivity equations
phase equilibria, a stability test of the solution satisfying Eqs. (3) under mass balance constraints. Each iteration solves the
and (4) based on the tangent plane criterion [20] or the procedure equations for each tie line to calculate ^ xik j and then calculates
proposed by Li et al. [6] is required. This test is not necessary for the value of Fx, which in turn determines the direction of the next
regression of parameters. iteration through the algorithm of PSO. The iteration goes on until a
The activity coefficient equation of the NRTL model is given as criterion, which is the maximum number of iterations in this study,
[1] is met. The quality of regression is generally evaluated by the
0 1 RMSD.
XN XN
t ji Gji xj B xl t lj Glj C 2
X 3 X 2 X M 
3
2 1=2
j¼1 XN
xj Gij B
B
C
C j
 ^ j
lng i ¼ þ Bt  l¼1 C (5) 6 x x 7
XN X
N B ij
XN C 6 i¼1 j¼1 k¼1 ik ik 7
@ 6 7
G x j¼1
G x G x A RMSD ¼ 6 7 (10)
ki k kj k kj k 6 6M 7
k¼1 k¼1 k¼1 4 5

ðgji  gii Þ
t ji ¼ (6) For convenience of evaluating the quality of correlation, Eq. (8)
RT
is modified into Eq. (11).
3 X
X 10
Gji ¼ expðaji t ji Þ (7) Fa ¼ s ik (11)
i¼1 k¼1
where N is the number of components in the system, gji is the
energy of interaction between a j–i pair of molecules; t ji (t ji 6¼ t ij)

Table 2
Regressed parameters of the NRTL model for ternary LLE systems.

Systems i–j aij tij t ji


Water(1) 1–2 0.24962 5.5351 0.96053
Phenol(2) 1–3 0.25426 5.4491 3.1784
Benzene(3) 2–3 0.30332 6.8088 0.048392
Ethene tetrachloro(1) 1–2 0.2 7.2316 2.9827
2-propanol(2) 1–3 0.2 5.0656 5.6934
Water(3) 2–3 0.2 0.26419 9.0067
Heptane(1) 1–2 0.4 0.94695 0.5712
Ethylbenzene(2) 1–3 0.4 2.0569 2.4497
Methanol(3) 2–3 0.4 1.2377 0.0067217
Cyclohexane(1) 1–2 0.3 1.0378 3.6815
Enthylbenzene(2) 1–3 0.2 4.6788 1.6054
Sulfolane(3) 2–3 0.3 1.2566 2.7125
Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45 39

where s ik is defined as used only 30 particles because using more particles is time
8 I consuming due to the involvement of solving nonlinear equations.
>
> ðx  g Iik Þ ðxIik g ik Þ
I
> ik
<  1; if 1 The optimization method was further tested by regressing
xIIik  g IIik xIIik g II
s ik ¼ ik (12) ternary LLE systems from literature with 30 pre-regressed particles
>
> ðx  g ik Þ
II II
ðxIik g ik Þ
I
and 1000 iterations. Comparison of the results with literature is
>
: ik  1; if <1
ðxIik  g Iik Þ ðxIIik g II Þ
ik given in Table 1 and the parameters are given in Table 2. Smaller
RMSDs were obtained using the current method which demon-
strates the strength of the PSO method used in this study.
3. Results and discussions

To further illustrate the minimization, we use the ternary [(Fig._4)TD$IG]


system of “Ethene tetrachloro(1) + 2-propanol(2) + water(3)”
[4,p. 148] as an example. In this example, all the non-
randomness parameters are fixed as 0.2 and the energy
parameters are searched in the range of [–15, 15]. Firstly, Fa
was minimized using 200 particles that are generated randomly,
each particle is a group of six energy parameters of the NRTL
model. An example of calculation is given in Fig. 1. It can be seen
that Fa reaches a minimum (either local or global) after
approximately 20 iterations. A few calculations found that
60 iterations are sufficient to reach a local minimum. The
minimization procedure was then repeated for 300 times and
the resulting 300 groups of parameters (called “pre-regressed
parameters”) were evaluated over RMSD and the results are
presented in Fig. 2a. It is shown that the parameters obtained
from minimization of Fa do not necessarily result in small RMSD,
which is the target of parameter regression. As comparison,
300 groups of randomly generated parameters without any
further manipulation were also evaluated using Eq. (10) with
results presented in Fig. 2b and as shown the RMSD can be very
large.
Following optimization of Fa,Fx was optimized with either
30 pre-regressed particles or 30 randomly generated particles and
each case was repeated 10 times. The results presented in Fig. 3
clearly indicate that the pre-regressed particles are good initial
estimates for the optimization of Fx although initially they may not
have sufficiently small RMSDs. Nine out of the 10 calculations using
pre-regressed particles reached very similar RMSD within
1000 iterations, indicating that the RMSD calculated from this
group of parameters is most likely the global minimum. On the
contrary, the optimizations using 30 random particles reached
very different and much larger RMSDs. The comparison demon-
strates that the pre-regressed particles are superior over the
random ones. Therefore, it is a good strategy to minimize Fx using
pre-regressed particles obtained from minimization of Fa.
Although selection of the number of particles and iterations
appears arbitrary, there is some rationale behind their selection.
For the optimization of Fa, 200 particles were chosen to ensure a
rapid approach to a local minimum and 60 iterations aim to ensure
the minimum reached is a local one rather than the global
minimum thus to maintain diversity of the pre-regressed particles
that are to be used for optimization of Fx. The optimization of Fx

Table 3
Three groups of parameters for the same LLE data.

Parameters Group 1 Group 2 Group 3


Ethene tetrachloro(1) t12 7.2316 5.8391 483.32/303.15
2-Propanol(2) t21 2.9827 2.9387 116.92/303.15
Water(3) t13 5.0656 6.4053 2022.0/303.15
t31 5.6934 7.7491 2382.4/303.15
t23 0.26419 0.0256 291.60/303.15
t32 9.0067 7.9771 259.50/303.15
RMSD 0.0045 0.0076 0.0078
Reference This study This study [4,p. 148] Fig. 4. Representation of activities by different groups of parameters.
40 Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45

3.1. Representation of LLE by the NRTL model equilibrium data due to sensitivity of the exponential functions.
Indeed random data that are normalized to resemble experimental
3.1.1. Local and global minimum LLE data have been generated and shown to be able to be fitted
Multiple groups of parameters were obtained for the ternary equally well as correlations of experimental LLE data, refer to
system of “Ethene Tetrachloro(1) + 2 propanol(2) + water(3)” Appendix A. This reveals the NRTL model’s empiricism and weak
[4,p. 148], two of them from this study and one from literature theoretical basis, which should be considered when using the
are presented in Table 3. Each group corresponds to a local model to calculate phase equilibria.
minimum of RMSD and the smallest RMSD is most likely the global
minimum. In terms of RMSD, all three groups of parameters are
5. Conclusion
good. However, parameters in different groups show no relevance.
If two pairs of binary interaction parameters in different groups are
The parameters of the NRTL model have been successfully
exchanged, much larger RMSD will be yielded and the calculated
regressed from LLE data using the particle swarm optimization
mole fractions can even be negative which does not make sense.
(PSO) method and smaller RMSDs were obtained compared with
For example, if the energy parameters t 12 and t 21 in “Group 1” are
literature results. Further analysis reveals that multiple groups of
replaced by the values in “Group 3”, some of the predicted mole
parameters with sufficiently small RMSDs can be found for the
fractions using the new group of parameters are negative and the
same set of LLE data and the activities calculated with these
RMSD obtained is 0.5557, which is much larger than both
parameters and their corresponding predicted mole fractions can
0.0045 and 0.0078.
be far beyond the reasonable range of activity. This clearly
demonstrates that the NRTL model does not represent the real
3.1.2. Representation of activities of LLE
activities of components with parameters regressed from LLE.
Equality of fugacity of a component in VLE can be simplified to
Moreover, five sets of normalized random data have been
be [2]
correlated by the NRTL model equally well as correlations of
P  yi regular experimental LLE data. This questions the theoretical
xi  g i ¼ (13)
Pi 0 validity of the NRTL model and highlights the need to strengthen
the model’s theory. The limitations of the NRTL model revealed in
where yi and xi are the fraction of component i in the vapor and
this study should be considered when the model is used for
liquid phase, respectively, P is the total pressure and Pi0 is the vapor
correlation and prediction of phase equilibria.
pressure of the pure component i at the same temperature. It can
be seen that the activity of a component in the liquid is equal to the
Acknowledgements
ratio of its partial pressure and vapor pressure of the pure
component. The mole fraction xi is within [0, 1], the activity
Support from the Australian Research Council and the Particu-
coefficient can be larger or smaller than unity; thus, mathemati-
late Fluids Processing Centre is acknowledged. Thanks to Prof. Jian
cally we cannot derive the range of activity. In fact, it is unusual to
Chen in Tsinghua University for helpful discussions. Zheng Li is
find a component whose activity is larger than unity.
grateful for financial support from the China Scholarship Council.
Unlike correlation of VLE which directly correlates partial vapor
pressures of components and composition of liquid phase,
Appendix A.
correlation of LLE equates two sides of isoactivity equations
without knowing the real activity of each component. As a result,
representation of activities of LLE system by the NRTL model needs
1.1. Generalization of normalized random data
to be examined. The three groups of parameters in Table 3 and their
corresponding predicted mole fractions are used to calculate
Five arrays of random data in the range of 0–1 were produced by
activities and results are shown in Fig. 4. Surprisingly, activities
the mathematical software “MATLAB”. Each array comprises ten
calculated by the three groups of parameters differ significantly
rows and six columns of data with every three columns of data
and many data points are clearly incorrect as their values are far
representing a phase. The data in each phase were then normalized
beyond the range of [0, 1]. Even the data points within [0, 1] most
such that the sum of data in a row equals to 1 (Eq. (A1)), thus
likely do not represent the real activities of components as the
resembling an experimental ternary LLE data set, where each data
values of activities are not considered in the parameter regression.
point represents the mole fraction of a component in one phase.
Consequently, correlation of LLE by the NRTL model is an empirical
fitting that may not reflect the real activities of components, which X
N

are the properties we aim to correlate. This weakness of the model xjik ¼ 1 ði ¼ 1; 2; 3; j ¼ 1; 2 Þ (A1)
i¼1
limits the model’s application in predicting phase equilibria using
parameters regressed from LLE. Expansions of the NRTL model where i is component, j is phase number and k is the number of
[7–10] may also suffer from the same weakness as they are rows. The five sets of normalized data are given in Tables A1–A5
generally obtained by adding additional items to the original and the randomness of these data sets is clearly shown by the
equations of the model while maintaining the model’s theoretical phase diagrams in Fig. A1. To evaluate the correlations, five regular
concepts and core equations. ternary LLE systems that also have ten rows of data have been
taken from literatures [21, 4, 12] and are correlated as comparisons
4. Correlating normalized random data (refer to Table A6).

The activity coefficient equation of the NRTL model uses a 1. 2 Correlation and discussions
combination of exponential functions with three adjustable
parameters, providing the model with excellent flexibility in Fa in Eq. (11) of Section 2.2 was used as the objective function to
correlating data. However, it has been demonstrated by Li et al. [6] correlate all 10 sets of data using the NRTL model with all three
that the isoactivity equations of ternary LLE using the NRTL model parameters adjustable. Minimization of Fa was performed by the
represent many more data points besides the experimental PSO method as illustrated in Section 2 and the regressed
Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45 41

Table A1
Normalized random data no. 1.

Phase I Phase II

x1 x2 x3 x1 x2 x3

0.3868 0.4331 0.1801 0.0359 0.2842 0.6799

0.3971 0.5335 0.0694 0.9403 0.0410 0.0187

0.1333 0.6224 0.2443 0.1416 0.7276 0.1308

0.2848 0.4569 0.2583 0.2618 0.3405 0.3977

0.5170 0.4643 0.0187 0.1253 0.2266 0.6481

0.5626 0.3476 0.0899 0.1399 0.7893 0.0708

0.1501 0.5665 0.2834 0.2327 0.4206 0.3466

0.0332 0.7131 0.2537 0.1050 0.3462 0.5488

0.1277 0.3931 0.4791 0.4940 0.1841 0.3219

0.3246 0.3900 0.2854 0.0997 0.0985 0.8017

Table A2
Normalized random data no. 2.

Phase I Phase II

x1 x2 x3 x1 x2 x3

0.5004 0.0968 0.4028 0.4969 0.3088 0.1943

0.4737 0.5076 0.0187 0.0291 0.3491 0.6218

0.0657 0.4951 0.4392 0.1631 0.4510 0.3859

0.3915 0.2081 0.4004 0.0460 0.7920 0.1620

0.2995 0.3790 0.3215 0.2410 0.4637 0.2953

0.0978 0.1423 0.7599 0.4546 0.2704 0.2751

0.1929 0.2922 0.5149 0.3308 0.2122 0.4570

0.2948 0.4937 0.2115 0.2432 0.4957 0.2611

0.3981 0.3294 0.2725 0.4233 0.3160 0.2607

0.4604 0.4579 0.0817 0.0340 0.7450 0.2210

Table A3
Normalized random data no. 3.

Phase I Phase II

x1 x2 x3 x1 x2 x3

0.8003 0.0408 0.1589 0.5117 0.3860 0.1023

0.1278 0.6101 0.2621 0.5316 0.3167 0.1516

0.3530 0.4001 0.2470 0.3143 0.3176 0.3682


42 Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45

Table A3 (Continued)
Phase I Phase II

x1 x2 x3 x1 x2 x3

0.2882 0.1622 0.5496 0.1676 0.2852 0.5471

0.2411 0.7363 0.0226 0.3701 0.2939 0.3360

0.3294 0.5819 0.0887 0.4285 0.3980 0.1735

0.3969 0.1515 0.4516 0.0653 0.3949 0.5398

0.0476 0.8646 0.0878 0.0693 0.1473 0.7834

0.5765 0.0672 0.3563 0.3271 0.3695 0.3034

0.8353 0.0075 0.1572 0.1579 0.4580 0.3840

Table A4
Normalized random data no. 4.

Phase I Phase II

x1 x2 x3 x1 x2 x3

0.1130 0.4734 0.4136 0.1410 0.4457 0.4133

0.3687 0.3665 0.2648 0.3476 0.3987 0.2537

0.3295 0.3704 0.3001 0.5219 0.3870 0.0911

0.5622 0.2998 0.1379 0.3578 0.4432 0.1990

0.2080 0.6424 0.1496 0.2626 0.6813 0.0561

0.3816 0.3801 0.2383 0.1317 0.5413 0.3270

0.5809 0.1051 0.3141 0.1011 0.2481 0.6509

0.2738 0.0603 0.6658 0.4057 0.4611 0.1333

0.0003 0.7269 0.2727 0.4338 0.1958 0.3704

0.0451 0.7700 0.1849 0.4539 0.0965 0.4497

Table A5
Normalized random data no. 5.

Phase I Phase II

x1 x2 x3 x1 x2 x3

0.3146 0.2183 0.4671 0.3597 0.3395 0.3008

0.3517 0.2915 0.3568 0.6338 0.0725 0.2937

0.3856 0.1152 0.4992 0.1881 0.2986 0.5133

0.5042 0.0630 0.4328 0.5320 0.2341 0.2339

0.5609 0.4093 0.0297 0.2021 0.3114 0.4865

0.5180 0.3861 0.0959 0.2930 0.0603 0.6467

0.5080 0.2071 0.2849 0.4065 0.0707 0.5228

0.1920 0.3427 0.4653 0.6164 0.2453 0.1383

0.0929 0.4440 0.4631 0.0412 0.3605 0.5983

0.2152 0.3011 0.4837 0.3870 0.5936 0.0194


Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45 43
[(Fig.A_1)TD$IG]

Fig. A1. Phase diagrams of normalized random data.

Table A6
Regressed NRTL Parameters for data correlation (literature data and random data).

System i–j tij t ji aij Fa Reference


n-Hexane(1) 1–2 9.1825 3.4202 0.44505 0.594 [21,p. 111]
Benzene(2) 1–3 7.9969 5.6226 0.42441
Sulfolane(3) 2–3 3.3058 0.71576 0.49549
Benzene isopropyl(1) 1–2 15.617 3.5251 0.41036 1.578 [4,p. 64]
2-Propanone(2) 1–3 15.537 6.2466 0.4048
Formic acid amide(3) 2–3 5.7609 14.755 0.47162
Ethene tetrachloro(1) 1–2 3.4885 4.5286 0.26742 1.844 [4,p. 148]
2-Propanone(2) 1–3 13.538 10.178 0.33408
Water(3) 2–3 4.102 4.3232 0.32702
Ethene trichloro(1) 1–2 4.4067 0.87823 0.25113 0.907 [4,p. 162]
Hexanoic acid 6-amino lactam(2) 1–3 3.7488 7.9841 0.31627
Water(3) 2–3 1.0684 4.888 0.34126
Water(1) 1–2 4.2195 10.272 0.19513 4.048 [12,p. 292]
Phenol(2) 1–3 11.022 5.0282 0.46432
Benzene(3) 2–3 2.0239 2.2543 0.50000
Normalized random 1–2 3.3317 3.4104 0.42839 1.075 This study
Data no. 1 1–3 4.4288 3.2477 0.43793
2–3 2.8774 2.6955 0.43179
Normalized random 1–2 20.000 3.4461 0.34736 0.683 This study
Data no. 2 1–3 10.675 3.8366 0.40511
2–3 3.029 2.6762 0.43372
Normalized random 1–2 18.214 3.1006 0.23501 1.748 This study
Data no. 3 1–3 2.2269 2.1058 0.50000
2–3 4.2231 1.7168 0.44643
Normalized random 1–2 2.8988 19.844 0.23173 1.331 This study
Data no. 4 1–3 3.0419 4.302 0.42781
2–3 2.8322 2.4435 0.45078
Normalized random 1–2 3.011 20.000 0.30478 0.736 This study
Data no. 5 1–3 2.8082 2.8503 0.42566
2–3 16.338 3.1631 0.31597
44 Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45
[(Fig.A_2)TD$IG]

Fig. A2. Correlations of the regular LLE data and normalized random data by the NRTL model.

parameters and results are given in Table A6 and results are also [5] C.-C. Chen, Molecular thermodynamics for pharmaceutical process modeling
presented in Fig. A2. and simulation, Chemical Engineering in the Pharmaceutical Industry, John
Wiley & Sons, Inc., 2010, pp. 505–519.
Interestingly, the values of Fa for the normalized random data [6] Z. Li, K.A. Mumford, Y. Shang, et al., Analysis of the nonrandom two-liquid
and the regular LLE data from literature are comparable, indicating model for prediction of liquid–liquid equilibria, J. Chem. Eng. Data 59 (8)
that the two types of data are correlated equally well. This means (2014) 2485–2489, doi:http://dx.doi.org/10.1021/je500204v.
[7] C.-C. Chen, A segment-based local composition model for the gibbs energy of
that the NRTL model does not have any advantage in correlating polymer solutions, Fluid Phase Equilib. 83 (0) (1993) 301–312, doi:http://dx.
regular experimental LLE data over correlating normalized random doi.org/10.1016/0378-3812(93) 87033-W.
data. The normalized random data does not have any physical [8] C.-C. Chen, Y. Song, Solubility modeling with a nonrandom two-liquid segment
activity coefficient model, Ind. Eng. Chem. Res. 43 (26) (2004) 8354–8362, doi:
meaning in terms of molecular interactions, hence the regressed http://dx.doi.org/10.1021/ie049463u.
NRTL parameters are meaningless. Similarly, the parameters for [9] C.-C. Chen, H.I. Britt, J.F. Boston, L.B. Evans, Local composition model for excess
the regular LLE systems are somewhat empirical, as can be seen Gibbs energy of electrolyte systems. Part I: single solvent, single completely
dissociated electrolyte systems, AIChE J. 28 (4) (1982) 588–596, doi:http://dx.
from derivation of the model’s equations. The weak theoretical
doi.org/10.1002/aic.69028041.
basis of the model explains the inconsistency of different groups of [10] Y. Song, C.-C. Chen, Symmetric electrolyte nonrandom two-liquid activity
parameters. Therefore, the NRTL model’s weaknesses and limi- coefficient model, Ind. Eng. Chem. Res. 48 (16) (2009) 7788–7797, doi:http://
tations should be considered when using it to correlate and predict dx.doi.org/10.1021/ie9004578.
[11] J.M. Sørensen, T. Magnussen, P. Rasmussen, A. Fredenslund, Liquid–liquid
phase equilibria and further work is required to strengthen the equilibrium data: their retrieval, correlation and prediction Part II. Correlation,
model’s theoretical basis. Fluid Phase Equilib. 3 (1) (1979) 47–82, doi:http://dx.doi.org/10.1016/0378-
3812(79)80027-8.
[12] J.R. Alvarez Gonzalez, E.A. Macedo, M.E. Soares, A.G. Medina, Liquid–liquid
References equilibria for ternary systems of water–phenol and solvents: data and
representation with models, Fluid Phase Equilib. 26 (3) (1986) 289–302, doi:
[1] H. Renon, J.M. Prausnitz, Local compositions in thermodynamic excess http://dx.doi.org/10.1016/0378-3812(86)80024-3.
functions for liquid mixtures, AIChE J. 14 (1) (1968) 135–144, doi:http://dx. [13] I. Ashour, S.I. Abu-Eishah, Liquid–liquid equilibria of ternary and six-
doi.org/10.1002/aic.690140124. component systems including cyclohexane, benzene, toluene, ethylbenzene,
[2] J. Gmehling, U. Onken, Vapour–Liquid Equilibrium Data Collection Aqueous– cumene, and sulfolane at 303.15 K, J. Chem. Eng. Data 51 (5) (2006) 1717–1722,
Organic Systems, Vol. 1, DECHEMA, Frankfurt/Main, 1977. doi:http://dx.doi.org/10.1021/je060153n.
[3] J.M. Prausnitz, F.W. Tavares, Thermodynamics of fluid-phase equilibria for [14] B.E. García-Flores, J. Águila-Hernández, F. García-Sánchez, M.A. Aquino-Olivos,
standard chemical engineering operations, AIChE J. 50 (4) (2004) 739–761, doi: (Liquid–liquid) equilibria for ternary and quaternary systems of representative
http://dx.doi.org/10.1002/aic.10069. compounds of gasoline + methanol at 293.15 K: experimental data and
[4] J.M. Sørensen, W. Arlt, Liquid–liquid equilibrium data collection-ternary correlation, Fluid Phase Equilib. 348 (2013) 60–69, doi:http://dx.doi.org/
systems, DECHEMA, Frankfurt, 1980. 10.1016/j.fluid.2013.03.022.
Z. Li et al. / Fluid Phase Equilibria 398 (2015) 36–45 45

[15] Bochove Gv, Two- and Three-Liquid Phase Equilibria in Industrial Mixed- [19] R.C. Eberhart, Y. Shi, Comparing inertia weights and constriction factors in
Solvent Electrolyte Solutions, Technische Universiteit Delft, Delft, 2003. particle swarm optimization, Proceedings of the 2000 Congress on 2000 1
[16] J. Kennedy, R. Eberhart, Particle swarm optimization, Proceedings, IEEE (2000) 84–88, doi:http://dx.doi.org/10.1109/cec.2000.870279 paper
International Conference on Nov/Dec 1995 4 (1995) 1942–1948, doi:http://dx. presented at: Evolutionary Computation.
doi.org/10.1109/icnn.1995.488968 (paper presented at: Neural Networks). [20] M.L. Michelsen, The isothermal flash problem. Part I. Stability, Fluid Phase
[17] R. Poli, J. Kennedy, T. Blackwell, Particle swarm optimization, Swarm Intell. 1 Equilib. 9 (1) (1982) 1–19, doi:http://dx.doi.org/10.1016/0378-3812(82)85001-
(1) (2007) 33–57, doi:http://dx.doi.org/10.1007/s11721-007-0002-0. 2.
[18] R.C. Eberhart, S. Yuhui, Particle swarm optimization: developments, [21] J. Chen, L.-P. Duan, J.-G. Mi, W.-Y. Fei, Z.-C. Li, Liquid–liquid equilibria of multi-
applications and resources, Proceedings of the 2001 Congress on 2001 1 component systems including n-hexane, n-octane, benzene, toluene, xylene
(2001) 81–86, doi:http://dx.doi.org/10.1109/cec.2001.934374 (paper and sulfolane at 298.15 K and atmospheric pressure, Fluid Phase Equilib. 173
presented at: Evolutionary Computation). (1) (2000) 109–119, doi:http://dx.doi.org/10.1016/S0378-3812(00)00398-8.

You might also like