Sei sulla pagina 1di 11

Fluid Phase Equilibria 394 (2015) 111

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

Vaporliquid equilibria for the binary systems ethylene + water,


ethylene + ethanol, and ethanol + water, and the ternary system
ethylene + water + ethanol from Gibbs-ensemble molecular simulation
Y. Mauricio Muoz-Muoz, Mario Llano-Restrepo *
School of Chemical Engineering, Universidad del Valle, Ciudad Universitaria Melendez, Building 336, Apartado 25360, Cali, Colombia

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 15 January 2015
Received in revised form 26 February 2015
Accepted 2 March 2015
Available online 5 March 2015

The conceptual design of a reactive separation process for the hydration of ethylene to ethanol requires
reliable vaporliquid equilibrium (VLE) data for the ternary system ethylene + water + ethanol. Due to the
paucity of experimental data points in the VLE phase diagrams that have been reported for that system,
molecular simulation looks appealing in order to predict such data. In this work, the Gibbs-ensemble
Monte Carlo (GEMC) method was used to calculate the VLE of the pure components (ethylene, water,
and ethanol), the binary subsystems (ethylene + water, ethylene + ethanol, and ethanol + water), and
the ternary system (ethylene + water + ethanol). A set of previously validated Lennard-Jones plus
point-charge potential models were chosen for the pure components, and the validity of these models
was corroborated from the good agreement of the GEMC simulation results for the vapor pressure and
the VLE phase diagrams of those components with respect to calculations carried out by means of the
most accurate (reference) multiparameter equations of state currently available for ethylene, water, and
ethanol. These potential models were found to be capable of predicting the available VLE phase diagrams
of the binary subsystems: ethylene + water at 200 and 250  C, ethylene + ethanol at 150, 170, 190, 200, and
220  C, and ethanol + water at 200, 250, 275, and 300  C. Molecular simulation predictions for the VLE
phase diagrams of the ternary system at 200  C and pressures of 30, 40, 50, 60, 80, and 100 atm, were
found to be in very good agreement with predictions previously made by use of a thermodynamic model
that combines the PengRobinsonStryjekVera equation of state, the WongSandler mixing rules, and
the UNIQUAC activity coefcient model. The agreement between the predictions of these two
independent approaches gives condence for the subsequent use of molecular simulation to predict the
combined phase and chemical equilibrium of the ternary system and check the validity of predictions
previously made by means of the thermodynamic model.
2015 Elsevier B.V. All rights reserved.

Keywords:
Vaporliquid equilibrium
Phase diagrams
Ternary systems
Ethylene hydration
Petrochemical ethanol

1. Introduction
The economics of the production of petrochemical ethanol by
the direct hydration of ethylene might improve by the application
of the concept of process intensication, i.e., by making the
reaction and the separation of the product (ethanol) and
the reactants (ethylene and water) to occur simultaneously in
the same piece of equipment. In the intensied process, ethanol
would be produced from ethylene by means of a vaporliquid
mixed-phase hydration, by feeding gaseous ethylene and liquid
water into a reactive separation column [1].

* Corresponding author. Tel.: +57 2 3312935; fax: +57 2 3392335.


E-mail address: mario.llano@correounivalle.edu.co (M. Llano-Restrepo).
http://dx.doi.org/10.1016/j.uid.2015.03.007
0378-3812/ 2015 Elsevier B.V. All rights reserved.

In a previous study [1], we used a thermodynamic model that


combines the PengRobinsonStryjekVera (PRSV2) equation of
state [2,3], the WongSandler (WS) mixing rules [4], and the
UNIQUAC activity coefcient model [5] in order to correlate the
available vaporliquid equilibrium (VLE) experimental data for
the binary subsystems (ethylene + water, ethylene + ethanol, and
ethanol + water) at 200  C. From the optimum values obtained
for the adjustable parameters of the PRSV2-WS-UNIQUAC
thermodynamic model, both the VLE and the combined phase
and chemical equilibrium (CPE) of the ternary system (ethylene +
water + ethanol) were predicted at 200  C and various pressures.
The thermodynamic model predicts that for many values of the
ethylene to water feed mole ratio, the vaporliquid mixed-phase
hydration of ethylene achieves equilibrium conversions much
higher than those calculated for a vapor-phase reaction that would
hypothetically occur at the same conditions of temperature,

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

Nomenclature
AAD
ARD
kB
kij
N
P
Psat
q+
q
T
U

Average absolute deviation


Average relative deviation
Boltzmann constant
Interaction parameter in the WongSandler mixing
rule
Number of molecules
Pressure
Vapor pressure
Positive electric charge
Negative electric charge
Absolute temperature
Congurational energy

Greek symbols
x
Correction factor for the Lorentz combining rule
DHv Heat of vaporization
Duij Characteristic energy parameter of the UNIQUAC model
DUT Change of total congurational energy
e
Energy parameter of the LJ potential model
r
Molar density
s
Size parameter of the LJ potential model
Subscripts
Coul Coulombic
LJ
Lennard-Jones
T
Total
Superscripts
real Real-space contribution
recip Reciprocal-space contribution
self Self-interaction contribution

pressure, and ethylene to water feed mole ratio. Another prediction


of the thermodynamic model is that the reactive phase diagram of
the ternary system exhibits a critical point at 200  C and 155 atm.
Due to the paucity of experimental data points in the VLE phase
diagrams reported by Tsiklis et al. [6] for the ternary system (in the
absence of chemical reaction), it is difcult to make a direct
assessment of the validity of the predictions made by the
thermodynamic model for both the VLE and the CPE of the
ternary system. However, an indirect assessment of those
predictions is possible by use of another method of predictive
nature and entirely independent from the thermodynamic model,
such as molecular simulation.
The Gibbs-ensemble Monte Carlo (GEMC) method of molecular
simulation [714] has been used widely and successfully by many
authors for the computation of VLE phase diagrams of binary and
ternary systems of industrial interest. In this work, the GEMC
method was carefully implemented in order to compute the VLE of
the pure components (ethylene, water, and ethanol), the binary
subsystems (ethylene + water, ethylene + ethanol, and ethanol +
water), and the ternary system (ethylene + water + ethanol). A set
of previously validated Lennard-Jones plus point-charge potential
models [1517] were chosen for ethylene, water, and ethanol. The
outline of the paper is as follows. The intermolecular potential
models for the pure components are presented in Section 2. The
simulation methods are described in Section 3. In Section 4,
simulation results for the pure components are reported,
discussed, and compared with calculations carried out by means
of the most accurate (reference) equations of state currently
available for ethylene, water, and ethanol [1820]. In Section 5,

simulation results for the VLE phase diagrams of the binary


subsystems are reported, discussed, and compared with both the
available experimental data [2123] and the correlation results
obtained by use of the PRSV2-WS-UNIQUAC thermodynamic
model described in our previous study [1]. In Section 6, simulation
results for the VLE phase diagrams of the ternary system are
reported, discussed, and compared with the few experimental data
available [6] and, more importantly, with the predictions
previously made [1] by means of the thermodynamic model.
2. Intermolecular potential models
As explained in more detail below, the Lennard-Jones (LJ) plus
point-charge intermolecular potential models recently devised by
Weitz and Potoff [15] for ethylene, Huang et al. [16] for water, and
Schnabel et al. [17] for ethanol were chosen for the present work
due to their greater accuracy relative to other models available for
those components. These potential models are planar and rigid
(i.e., they have no internal degrees of freedom).
The potential model for ethylene devised by Weitz and Potoff
[15] comprises two LJ sites for the methylene (CH2) united-atom
groups, each site with a positive point charge q+ = 0.85e, where e is
the electron charge, and separated by a distance of 1.33 . A
negative point charge q = 1.70e is located in the middle point
(center of mass, COM) between the two LJ sites. The point charges
and the distance between the CH2 groups reproduce the gas-phase
quadrupole moment of ethylene. The local coordinates and charges
of the sites, and the size and energy parameters of the LJ potential
are given in Table 1. With this model, Weitz and Potoff used the
histogram-reweighting technique [14,24] in combination with
grand canonical ensemble Monte Carlo simulations to compute the
VLE of pure ethylene and the binary mixtures ethylene + carbon
dioxide, ethylene + xenon, and ethylene + n-butane. Saturated
liquid densities of pure ethylene were predicted in very close
agreement with experiment whereas its vapor pressure was
reproduced with a deviation of 2%. By comparing their potential
model with other two united-atom models (NERD and TraPPE-UA)
available for ethylene, Weitz and Potoff [15] found that their
proposed model yields more accurate predictions for the saturated
liquid densities, vapor pressures, critical density, and normal
boiling point of ethylene. However, no predictions for the heat of
vaporization of ethylene were reported by Weitz and Potoff.
The potential model for water devised by Huang et al. [16] is an
optimized version of the TIP4P model for water proposed 30 years
ago by Jorgensen et al. [25]. The optimized TIP4P model by

Table 1
Local coordinates (x,y) and charge q of the sites, and size (s ) and energy (e)
parameters of the LJ-site contribution to the planar (z = 0 ) intermolecular
potential models of ethylene [15], water [16], and ethanol [17].
Interaction site

x ()

Potential model of ethylene [15]


CH2
0.665
CH2
0.665
COM
0
Potential model of water [16]
O
0
H
0.7069
H
0.7069
NPC
0.20482
Potential model of ethanol [17]
CH3
0
CH2
1.9842
O
2.0127
H
2.9290

q (e)

s ()

e/kB (K)

0.85
0.85
1.70

3.72
3.72
0

83.0
83.0
0

0
0.9133
0.9133
0

0
0.41955
0.41955
0.83910

3.11831
0
0
0

208.08
0
0
0

0
0
1.7156
1.9683

0
0.25560
0.69711
0.44151

3.6072
3.4612
3.1496
0

120.15
86.291
85.053
0

y ()
0
0
0

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

Huang et al. comprises three point charges and one LJ site. One
negative point charge (NPC) q = 0.8391e is located at a distance
of 0.20482 from the oxygen (O) atom, on the line bisecting the
HOH angle, and two positive point charges, each with a value
q+ = 0.41955e, are located on each hydrogen (H) atom. Each
hydrogen atom is located at a distance of 1.1549 from the
oxygen atom. The only LJ site of the model is located on the oxygen
atom. The local coordinates and charges of the sites and the size
and energy parameters of the LJ potential are given in Table 1. With
this model, Huang et al. used the grand equilibrium method [26] in
combination with NPT-ensemble simulations and the gradual
insertion method [27,28] to compute the VLE of pure water and the
binary systems water + ethylene oxide and water + ethylene glycol.
The vapor pressure, the saturated liquid densities, and the heat of
vaporization of pure water were reproduced with deviations of
7.2%, 1.1%, and 2.8%, respectively. By comparing their potential
model with other four united-atom models (TIP4P, TIP4P-Ew,
TIP4P/2005, and SPC/E) available for water, Huang et al. [16]
found that their proposed model yields more accurate predictions
for the saturated liquid densities, vapor pressures, and heat of
vaporization of water.
The potential model for ethanol devised by Schnabel et al. [17]
comprises three LJ sites and three point charges. The LJ sites are
located on the methyl (CH3) and methylene (CH2) united-atom
groups, and on the oxygen atom of the hydroxyl (OH) group.
Positive point charges q+,1 = 0.2556e and q+,2 = 0.44151e are located
on the CH2 group and the hydrogen atom of the OH group,
respectively. A negative charge q = 0.69711e is located on the
oxygen atom of the OH group. The CH2 group is located at a
distance of 1.9842 from the CH3 group. The oxygen atom of the
OH group is located at a distance of 1.71581 from the CH2 group.
The hydrogen and oxygen atoms of the OH group are separated by a
distance of 0.95053 . The (CH3)(CH2)O and (CH2)OH angles are
90.950 and 106.368 , respectively. The local coordinates and
charges of the sites and the size and energy parameters of the LJ
potential are given in Table 1. With this model, Schnabel et al. used
the NPT + test particle method of Mller and Fischer [29,30]
in combination with the gradual insertion method [27,28] to
compute the VLE of pure ethanol, and the grand equilibrium
method [26] to compute the VLE of the binary system
ethanol + carbon dioxide. The vapor pressure, the saturated liquid
densities, and the heat of vaporization of pure ethanol were
reproduced with deviations of 3.7%, 0.3%, and 0.9%, respectively. By
comparing their potential model with other two united-atom
models (OPLS-UA and TraPPE-UA) available for ethanol, Schnabel
et al. [17] found that their proposed model yields more accurate
predictions for the saturated liquid densities, vapor pressures,
and heat of vaporization of ethanol.
3. Simulation methods
In contrast to the methods used in the works by Weitz and
Potoff [15], Huang et al. [16], and Schnabel et al. [17], the Gibbs
ensemble Monte Carlo (GEMC) method [714] was used in the
present work, in order to compute the VLE of the pure components
ethylene, water, and ethanol, the binary subsystems (ethylene +
water, ethylene + ethanol, and ethanol + water) and the ternary
mixture (ethylene + water + ethanol). The canonical version
(GEMC-NVT) of the GEMC method was applied to the pure
components and its isothermal-isobaric version (GEMC-NPT) was
applied to the binary and ternary systems. Our two previous
implementations of the GEMC method [31,32] were taken as the
starting point to develop the computational strategies explained in
Sections 3.1 and 3.2.
There are three types of random moves for the GEMC method
when dealing with rigid molecules: translational displacement

and rotation of molecules inside each of the two simulation boxes,


volume changes for the boxes, and transfer of molecules between
the two boxes (i.e, simultaneous removal and insertion moves).
These moves are accepted or rejected in accordance with a
particular probability recipe that involves the calculation of the
total intermolecular potential energy change DU T UnT  U oT
between the new (trial) (UnT ) and the old (UoT ) congurations.
The probability formulas for the acceptance or rejection of the
three kinds of moves in the GEMC method have been discussed in
various papers [714] and also in the textbook by Frenkel and
Smit [33]. The total intermolecular energy was computed as the
sum of non-electrostatic and electrostatic contributions. The
non-electrostatic contribution (U LJ ) was calculated by means of
the Lennard-Jones pair potential model and the electrostatic
contribution (UCoul) was calculated by means of the Ewald
summation method [3337] for the Coulombic potential, as
follows:
recip
self
U T U LJ U Coul U trunc
U corr
U real
LJ
LJ
Coul U Coul U Coul

(1)

where U trunc
LJ

is the truncated LJ potential, U corr


LJ is the corresponding
recip
self
long-range correction [33,34,38], and U real
Coul , U Coul , and U LJ are

LJ
the real space, reciprocal space, and self-interaction terms of the
Ewald sum, respectively [3337]. Periodic boundary conditions
and the minimum image convention, as explained in detail by
Allen and Tildesley [34], were applied to the calculation of the total
intermolecular potential energy. For the rotational moves required
for water and ethanol, the orientational displacements followed
the scheme [34] that chooses random values for the Euler angles in
the rotation matrix, and employs the internal coordinates of the
sites of the molecule (see Table 1) to calculate their simulation-box
coordinates. For all simulations in this work, a spherical cutoff
distance rc = 8.5 was used to truncate the LJ potential and
calculate its long-range correction from the analytical expression
given in Refs. [33,34] and extended to mixtures by de Pablo and
Prausnitz [38]. The value of 8.5 for the cutoff distance is a
compromise between the values obtained by applying the typical
criterion rc = 2.5s to the LJ contributions to the potentials models of
ethylene, water, and ethanol (see Table 1), and was always less than
half the simulation box length so that consistency with the
minimum image convention was preserved. Since the Ewald
summation method was used in the present work to compute the
Coulombic interactions between point charges, then there was no
need to truncate those interactions and, accordingly, no need
either for such a long cutoff distance (e.g., 17.5 ) as used in other
works [16,17] for the calculation of the Coulombic interactions by
truncation and correction by means of the reaction eld method.
The time-saving strategy devised by Fartaria et al. [39] was
carefully implemented in the present work to obtain a signicant
decrease in the computer time required for the calculation of the
Ewald sum. In that strategy, two repository matrices are used to
store pairwise potential energies like the real space term of the
Ewald sum and the truncated LJ potential, which can be retrieved
when needed, thereby saving time for the calculation of the
intermolecular energy associated to trial congurations. Whenever
a translational displacement or rotation of a molecule is carried
out, the only change experienced by the repository matrices is for
the row and column associated to the molecule being moved.
Therefore, the change of total congurational energy is given by
the summation (over all of the other molecules) of the difference
between the trial and the old pair energies of each of those
molecules and the molecule being moved. Whenever a trial
conguration is accepted, the corresponding pairwise potential
energies are recorded in the repository matrices, to be retrieved
whenever needed along the simulation run. Fartaria et al. [39]
showed that time-savings of up to 52% are possible when this

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

strategy is implemented for molecular models comprising point


charges and the Ewald sum is used to compute the Coulombic
interactions. Besides the two repository matrices already
mentioned, an additional repository matrix was used in the
present work, in order to store and retrieve the complex-variable
summation [3337] that runs over molecular sites for each
reciprocal space vector, allowing an efcient calculation of changes
in the reciprocal space term of the Ewald sum.
3.1. Simulation method for pure components
For the simulation of VLE for the pure components, the
following ve-stage strategy was implemented. In the rst stage,
estimates of the saturated liquid and vapor densities of the pure
components at a specied temperature T were obtained by means
of reference multiparameter equations of state [1820]. These
equations of state were solved by means of the calculation
approach proposed by Mao et al. [40] and Shi and Mao [41] for
water and steam, and extending its application to ethylene and
ethanol. In the second stage, for each phase of the pure component,
an NVT-ensemble simulation (with N = 400 molecules), with the
corresponding density xed at the value estimated from the
equation of state, was carried out for a total number of 2  107
moves (molecular displacements and rotations), 60% of which
were typically used to equilibrate the congurational energy. In the
third stage, starting from the nal conguration obtained after the
NVT runs, an NPT-ensemble simulation, with the pressure xed at
the value of the vapor pressure estimated from the equation of
state, was carried out for each phase for a total number of 3  107
moves (using a ratio of one box-volume change to N molecular
displacements and rotations), 60% of which were typically used to
equilibrate the molecular density and the congurational energy.
In the fourth stage, starting from the nal congurations obtained
after the NPT runs, a GEMC-NVT simulation was carried out for the
set of two boxes (each of them with an initial number of
N = 400 molecules), for a total number of at least 1 105 moves
(using a ratio of one volume change to 2N molecular displacements
and rotations and 2N molecular transfers between the boxes), 60%
of which were typically used to equilibrate the properties being
averaged (molecular densities and congurational energies of the
two phases). The attainment of phase equilibrium was ascertained
from the statistical equality of the chemical potentials for the two
phases, calculated by means of the particle insertion method, and
the length of the simulation run was extended until that equality
was achieved. Since the equality of the chemical potentials is
achieved precisely by means of the transfer moves, the achievement of that equality in the simulation run indicates that the
number of successful transfer moves was sufcient. Statistical
uncertainties associated to the GEMC-NVT ensemble-averages
were calculated by means of the method by Flyvbjerg and Petersen
[42]. In the fth stage, an additional NVT-ensemble simulation
was carried out for the nal vapor-phase box obtained after the
GEMC-NVT run, in order to compute the pure-component vapor
pressure by means of the method of Harismiadis et al. [43].

the values estimated from the thermodynamic model, was carried


out for a total number of 2  107 moves (molecular displacements
and rotations), 60% of which were used to equilibrate the
congurational energy. In the third stage, starting from the nal
conguration obtained after the NVT runs, an NPT-ensemble
simulation was carried out for each phase for a total number of
3  107 moves (using a ratio of one box-volume change to
N molecular displacements and rotations), 60% of which were
used to equilibrate the molecular density and the congurational
energy. In the fourth stage, starting from the nal congurations
obtained after the NPT runs, a GEMC-NPT simulation was
carried out for the set of two boxes (each of them with an initial
number of N = 400 molecules), for a total number of moves in the
range from 1 107 to 1.5  107 (using a ratio of one volume change
for each box to 2N molecular displacements and rotations and
2N molecular transfers between the boxes). Properties of the
coexisting phases were sampled every 5  105 moves, and running
averages were recalculated until the statistical equality for the
chemical potentials of each component in the two phases was
attained. Statistical uncertainties associated to the GEMC-NPT
ensemble-averages were also calculated by means of the method
by Flyvbjerg and Petersen [42]. For all VLE simulations for the
binary and ternary mixtures, the LorentzBerthelot combining
rules were used to calculate the size and energy parameters of the
LJ potential for the unlike interactions.
4. Simulation results for the pure components
By following the ve-stage strategy explained in Section 3.1,
pure-component VLE molecular simulations were carried out by
using the potential models of ethylene, water, and ethanol
described in Section 2. The VLE phase diagrams of the pure
components were computed over the temperature intervals
(160 K, 265 K) for ethylene, (300 K, 600 K) for water, and (270 K,
493 K) for ethanol. These phase diagrams are shown in Figs. 13 as
graphs of absolute temperature vs. molar density of the two
phases, where the empty squares correspond to the molecular
simulation results and the solid lines correspond to the results
obtained by means of the most accurate (reference) multiparameter equations of state currently available: the equation of Smukala
et al. [18] for ethylene, the equation of Wagner and Pru [19] for
water, and the equation of Dillon and Penoncello [20] for ethanol.
A comprehensive account of multiparameter equations of state
as an accurate source of thermodynamic property data has been

[(Fig._1)TD$IG]

3.2. Simulation method for mixtures


For the simulation of VLE for the binary and ternary mixtures,
the following four-stage strategy was implemented. In the rst
stage, by specifying both the temperature and the pressure of the
system, estimates of the molar compositions and the densities of
the coexisting phases were obtained from a TP ash calculation
with the help of the PRSV2-WS-UNIQUAC thermodynamic model
described in our previous study [1]. In the second stage, for each
phase of the mixture, an NVT-ensemble simulation (with
N = 400 molecules), with the density and mole fractions xed at

Fig. 1. VLE phase diagram for ethylene. &, molecular simulation results of this
work; , calculated with the reference multiparameter equation of state of Smukala
et al. [18] for ethylene.

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

[(Fig._2)TD$IG]

[(Fig._4)TD$IG]

Fig. 2. VLE phase diagram for water. &, molecular simulation results of this work;
, calculated with the reference multiparameter equation of state of Wagner and
Pru [19] for water.

Fig. 4. Clausius-Clapeyron plot for the vapor pressure of ethylene as a function of


temperature. &, molecular simulation results of this work; , calculated with the
reference multiparameter equation of state of Smukala et al. [18] for ethylene.

given in the book by Span [44]. These equations are regarded as


fundamental equations of state because they are based on
the Helmholtz free energy. The reference character of these
multiparameter equations of state is supported by the very low
deviations that are associated to the thermodynamic properties
calculated from those equations. Indeed, the reference equation
of state by Smukala et al. [18] for ethylene reproduces the
experimental data for the saturated liquid and vapor densities
with deviations within 0.004% and 0.02%, respectively, and the
data for the vapor pressure with a deviation less than 0.01%.
The reference equation of state by Wagner and Pru [19] for water,
which is known as the IAPWS-95 formulation, reproduces the
experimental data for the saturated liquid and vapor densities with
deviations within 0.0025% and 0.1%, respectively, and the data for
the vapor pressure with a deviation within 0.025%. The reference
equation of state by Dillon and Penoncello [20] for ethanol
reproduces the experimental data for the saturated liquid and
vapor densities with deviations less than 0.2% and 0.5%,
respectively, and the data for the vapor pressure with a deviation
less than 0.4%. These reference equations of state were compared
by their proponents [1820] with previous equations and were
found to be much more accurate.

The average relative deviation (ARD) of the simulation results


for the saturated liquid densities with respect to the values
obtained from the corresponding reference equation of state is
2.5% for ethylene, 1.1% for water, and 1.8% for ethanol. An
assessment of the deviation for ethylene cannot be made because
no corresponding deviation value was reported by Weitz and Potoff
[15], using the same potential model for ethylene. The deviation for
water agrees perfectly with the value (1.1%) reported by Huang
et al. [16], using the same potential model for water but a different
simulation method. In contrast, the deviation for ethanol (1.8%)
turns out to be larger than the value (0.3%) reported by Schnabel
et al. [17], using the same potential model for ethanol but a
different simulation method.
As mentioned in Section 3.1, the vapor pressure of the pure
components was computed by applying the method of
Harismiadis et al. [43] to the nal vapor-phase simulation box
obtained after each GEMC-NVT simulation run. The resulting
Clausius-Clapeyron diagrams (i.e., graphs of ln Psat vs. 1/T) are
shown in Figs. 46, where the emtpy squares correspond to the
molecular simulation results and the solid line corresponds
to the results obtained by means of the respective reference
equation of state [1820]. The ARD of the simulation results for the

[(Fig._3)TD$IG]
[(Fig._5)TD$IG]

Fig. 3. VLE phase diagram for ethanol. &, molecular simulation results of this work;
, calculated with the reference multiparameter equation of state of Dillon and
Penoncello [20] for ethanol.

Fig. 5. Clausius-Clapeyron plot for the vapor pressure of water as a function of


temperature. &, molecular simulation results of this work; , calculated with the
reference multiparameter equation of state of Wagner and Pru [19] for water.

[(Fig._6)TD$IG]

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

[(Fig._8)TD$IG]

Fig. 8. Heat of vaporization for water as a function of temperature. &, molecular


simulation results of this work; , calculated with the reference multiparameter
equation of state of Wagner and Pru [19] for water.

Fig. 6. Clausius-Clapeyron plot for the vapor pressure of ethanol as a function of


temperature. &, molecular simulation results of this work; , calculated with the
reference multiparameter equation of state of Dillon and Penoncello [20] for
ethanol.

where U T

and U T are the total congurational energies of the

vapor pressure with respect to the values obtained from the


corresponding reference equation of state is 5.0% for ethylene, 9.6%
for water, and 4.1% for ethanol. The deviation for ethylene (5.0%)
turns out to be much larger than the value (2%) reported by
Weitz and Potoff [15], using the same potential model for ethylene
but a different simulation method. The deviation for water (9.6%)
is 33% larger than the value (7.2%) reported by Huang et al. [16],
using the same potential model for water but a different
simulation method. In contrast, the deviation for ethanol (4.1%)
is just 11% larger than the value (3.7%) reported by Schnabel et al.
[17], using the same potential model for ethanol but a different
simulation method.
Simulation results for the vapor pressure and the congurational energies and densities of the coexisting phases at a given
temperature were used to compute the heat of vaporization (DHv)
of the pure components from the following thermodynamic
relationship:


1
1
L
sat
DHv UV
 L
(2)
T  UT P
V

vapor and liquid phases, respectively, and rV and rL are their


corresponding densities. The resulting graphs of DHv vs. T are
shown in Figs. 79, where the empty squares correspond to the
molecular simulation results and the solid line corresponds to the
results obtained by means of the respective reference equation of
state [1820]. The ARD of the simulation results for the enthalpy of
vaporization with respect to the values obtained from the
corresponding reference equation of state is 1.7% for water, 4.1%
for ethanol, and 12.4% for ethylene. The deviation for water (1.7%)
turns out to be 39% smaller than the value (2.8%) reported by
Huang et al. [16], using the same potential model for water but a
different simulation method. In contrast, the deviation for ethanol
(4.1%) turns out to be much larger than the value (0.9%) reported by
Schnabel et al. [17], using the same potential model for ethanol but
a different simulation method. For ethylene, simulation results
exhibit a signicant offset (by overestimation) with respect to the
equation of state results (see Fig. 7). Since the heat of vaporization
was not reported by Weitz and Potoff [15] (proponents of the
potential model of ethylene), a direct assessment of the seemingly
large deviation (12.4%) obtained in the present work cannot be

[(Fig._7)TD$IG]

[(Fig._9)TD$IG]

Fig. 7. Heat of vaporization for ethylene as a function of temperature. &, molecular


simulation results of this work; , calculated with the reference multiparameter
equation of state of Smukala et al. [18] for ethylene.

Fig. 9. Heat of vaporization for ethanol as a function of temperature. &, molecular


simulation results of this work; , calculated with the reference multiparameter
equation of state of Dillon and Penoncello [20] for ethanol.

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

[(Fig._10)TD$IG]

[(Fig._12)TD$IG]

Fig. 10. VLE phase diagram of ethylene (1) + water (2) at 200  C. &, molecular
simulation results of this work; , calculated with the PRSV2-WS-UNIQUAC
thermodynamic model; +, experimental data of Tsiklis et al. [21].

made. However, similarly large deviations for the heat of


vaporization predicted by other LJ plus point-charge potential
models have been reported, even when the vapor pressure and the
saturated liquid densities are reproduced accurately. For instance,
Huang et al. [16] reported an ARD (with respect to the
experimental data) of 13.4% for the heat of vaporization predicted
by their potential model of ethylene glycol (see Fig. 4 in Ref. [16], in
which an offset by overestimation is also shown), even though the
saturated liquid densities and vapor pressure of that substance
were reproduced with acceptable accuracy by the same potential
model (see Figs. 2 and 3 in Ref. [16]).
As shown in Figs. 19, the statistical uncertainties (error bars) of
our GEMC simulation results are small (most of them are within or
smaller than the symbol sizes). Therefore, the resulting discrepancy for the average deviations between the present and the
previous works is more likely due to the difference between the
simulation techniques used.
The good agreement of the GEMC simulation results for the
vapor pressure and the VLE phase diagrams of the pure
components, with respect to calculations carried out by means
of the most accurate (reference) equations of state currently
available for those components, corroborates the validity of
the potential models proposed by Weitz and Potoff [15] for
ethylene, Huang et al. [16] for water, and Schnabel et al. [17]
for ethanol.

Fig. 12. VLE phase diagram of ethylene (1) + ethanol (2) at 200  C. &, molecular
simulation results of this work; , calculated with the PRSV2-WS-UNIQUAC
thermodynamic model; +, experimental data of Tsiklis and Kofman [22].

5. Simulation results for the binary mixtures


By following the four-stage strategy explained in Section 3.2,
binary-mixture VLE molecular simulations were carried out by
using the potential models of ethylene, water, and ethanol
described in Section 2. Isothermal VLE phase diagrams of the
binary subsystems were computed at the following temperatures: 200, 250, and 300  C for ethylene + water, 150, 170, 190,
200, and 220  C for ethylene + ethanol, and 200, 250, 275, and
300  C for ethanol + water. The resulting phase diagrams at
200  C are shown in Figs. 1013 as graphs of pressure vs.
molar composition of the two phases (P-xy diagrams), where the
empty squares (with their respective error bars) correspond to
the molecular simulation results, the crosses correspond to the
available experimental data [2123], and the solid lines
correspond to the results obtained by means of the PRSV2WS-UNIQUAC thermodynamic model described in our previous
study [1]. The resulting phase diagrams at the other temperatures
listed above are shown as Figs. S1S9 in the Supporting
information.

[(Fig._13)TD$IG]

[(Fig._1)TD$IG]

Fig. 11. VLE phase diagram of ethylene (1) + water (2) at 200  C. &, molecular
simulation results of this work using a correction factor x = 0.9 for the Lorentz
combining rule; , calculated with the PRSV2-WS-UNIQUAC thermodynamic
model; +, experimental data of Tsiklis et al. [21].

Fig. 13. VLE phase diagram of ethanol (1) + water (2) at 200  C. &, molecular
simulation results of this work; , calculated with the PRSV2-WS-UNIQUAC
thermodynamic model; +, experimental data of Barr-David and Dodge [23].

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

5.1. Results for ethylene + water


The isothermal VLE (P-xy) phase diagrams for the binary system
ethylene + water are shown in Figs. 10 and 11 at 200  C, and in
Figs. S1 and S2 (in the Supporting information) at 250 and 300  C,
respectively. The parameter values reported in Table 2 were used
for the calculations with the PRSV2-WS-UNIQUAC thermodynamic
model. The solid line and data points at the left side of the diagrams
correspond to the liquid phase and the solid line and data points at
the right side of the diagrams correspond to the vapor phase. The
three sets of data (molecular simulation, thermodynamic model,
and experimental measurements) consistently follow the same
increasing trend for the mole fraction of ethylene as the pressure
increases. At 200 and 250  C, simulation results for the vapor phase
exhibit smaller error bars and are closer to both the experimental
data and the results obtained from the thermodynamic model, as
compared to the simulation results for the liquid phase. At 300  C,
as the pressure increases, simulation results for the vapor phase
exhibit a monotonic deviation to the left of the corresponding
experimental data (see Fig. S2), and the simulation result for the
liquid phase at 160 atm is shifted to the right, signicantly. A
possible interpretation of this behavior is that simulation would be
predicting a much narrower phase envelope for ethylene + water at
300  C, which is a surprising result considering the trend of the
experimental data.
At 200  C, the average absolute deviation (AAD) of the
simulation results (ensemble-averages) for the mole fraction of
ethylene with respect to the experimental values, is 0.0030 for the
liquid phase and 0.0345 for the vapor phase. At 250  C, the AAD is
0.0015 for the liquid phase and 0.0240 for the vapor phase. At
300  C, the AAD is 0.0050 for the liquid phase and 0.0919 for the
vapor phase.
Additional simulations were carried out in order to assess the
effect of including a correction factor x in the Lorentz combining
rule for the energy parameter eij of the LJ potential for unlike
interactions, so that ei;j xei ej 1=2 . By means of a simple random
search, several values of the correction factor x were tried for the
simulations at 200  C and the best results were obtained with
x 0:9, for which the resulting VLE phase diagram is shown in
Fig. 11. The AAD of the simulation results for the mole fraction of
ethylene at 200  C decreased from 0.0030 to 0.0017 for the liquid
phase and from 0.0345 to 0.0189 for the vapor phase, showing an
improvement with respect to the use of the default value x 1:0.
This correction factor was not applied to the simulations at 250
and 300  C.
From an overall examination of the VLE phase diagrams shown
in Figs. 10, 11, and S1, it follows that the intermolecular potential
models for ethylene and water described in Section 2 are capable of
qualitatively predicting the VLE behavior of this binary system at
200 and 250  C.
5.2. Results for ethylene + ethanol
The VLE phase diagrams for the binary system ethylene +
ethanol are shown in Fig. 12 at 200  C and in Figs. S3S6 (in the
Supporting information) for temperatures of 150, 170, 190, and
220  C. The parameter values reported in Table 2 were used for the
calculations with the PRSV2-WS-UNIQUAC thermodynamic model.
Once again, the solid line and data points at the left side of the
diagrams correspond to the liquid phase and the solid line and data
points at the right side of the diagrams correspond to the vapor
phase. The three sets of data (molecular simulation, thermodynamic model, and experimental measurements) follow the
same trend for the mole fraction of ethylene as the pressure
increases. At 150  C and in the middle range of pressures,

Table 2
Optimum values of the adjustable parameters for the t of the PRSV2-WS-UNIQUAC
thermodynamic model [1] to the VLE experimental data of Tsiklis et al. [21] for
ethylene (1) + water (2), Tsiklis and Kofmann [22] for ethylene (1) + ethanol (3), and
Barr-David and Dodge [23] for ethanol (3) + water (2).
System (ij)

T ( C)

Duij/R (K)

Duji/R (K)

kij

12
12
12
13
13
13
13
13
32
32
32
32

200
250
300
150
170
190
200
220
200
250
275
300

808.94
603.75
936.45
394.7
306.1
350.0
420.0
155.0
23.67
51.92
22.60
293.78

385.66
2007.55
1685.95
94.2
100.0
100.0
100.0
65.0
369.45
348.74
395.02
770.99

1.0181
1.1301
2.7033
0.115
0.145
0.235
0.260
0.350
0.0085
0.0003
0.0157
0.0737

simulation results for the vapor phase exhibit larger deviations


with respect to both the experimental data and the results
obtained from the thermodynamic model.
At 150  C, the average absolute deviation (AAD) of the
simulation results (ensemble-averages) for the mole fraction of
ethylene with respect to the experimental values, is 0.0332 for the
liquid phase and 0.0781 for the vapor phase. At 170  C, the AAD is
0.0275 for the liquid phase and 0.0476 for the vapor phase. At
190  C, the AAD is 0.0174 for the liquid phase and 0.0196 for the
vapor phase. At 200  C, the AAD is 0.0143 for the liquid phase and
0.0189 for the vapor phase. At 220  C, the AAD is 0.0085 for the
liquid phase and 0.0156 for the vapor phase.
Even though both molecular simulations and the TP ash
calculations made with the help of the thermodynamic model
become increasingly difcult in the region near the critical point of
the binary mixture, the results associated to these two methods are
in very good agreement in that region for some of the phase
diagrams, e.g., at temperatures of 170 and 220  C (Figs. S4 and S6).
The experimental critical points of this binary system were
reported by Tsiklis and Kofman [22] for each of the ve values of
temperature corresponding to the VLE phase diagrams shown in
Figs. 12 and S3S6. In each of these diagrams the experimental
critical point is indicated by the highest pressure cross. At 150 and
170  C (see Figs. S3 and S4), the critical pressures predicted from
both molecular simulation and the thermodynamic model are very
close to the experimental critical pressures of 127 and 118 atm,
respectively [22]. At 190 and 200  C (see Figs. S5 and 12), the
critical pressures predicted by extrapolation from the molecular
simulation results are very close to the experimental critical
pressures of 105 and 98 atm, respectively [22], whereas the critical
pressures predicted by means of the thermodynamic model turn
out to be slightly higher than the experimental ones. At 220  C (see
Fig. S6), both molecular simulation and the thermodynamic model
predict a critical pressure about 8% higher than the experimental
value of 81 atm [22].
From an overall examination of the VLE phase diagrams shown
in Figs. 12 and S3S6, it follows that the intermolecular potential
models for ethylene and ethanol described in Section 2 are capable
of qualitatively predicting the VLE behavior of this binary system.
5.3. Results for ethanol + water
The VLE phase diagrams for the binary system ethanol + water
are shown in Fig. 13 at 200  C and in Figs. S7S9 (in the Supporting
information) for temperatures of 250, 275, and 300  C. The
parameter values reported in Table 2 were used for the calculations
with the PRSV2-WS-UNIQUAC thermodynamic model. Once again,
the solid line and data points at the left side of the diagrams

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

[(Fig._14)TD$IG]

[(Fig._16)TD$IG]

Fig. 14. VLE phase diagram of ethylene + water + ethanol at 200  C and 30 atm. *,
molecular simulation results of this work; , calculated with the PRSV2-WSUNIQUAC thermodynamic model; ~, experimental data of Tsiklis et al. [6].

Fig. 16. VLE phase diagram of ethylene + water + ethanol at 200  C and 50 atm. *,
molecular simulation results of this work; , calculated with the PRSV2-WSUNIQUAC thermodynamic model; ~, experimental data of Tsiklis et al. [6].

correspond to the liquid phase and the solid line and data points at
the right side of the diagrams correspond to the vapor phase. The
three sets of data (molecular simulation, thermodynamic model,
and experimental measurements) follow the same trend for the
mole fraction of ethanol as the pressure increases. Even though
simulation results for the vapor phase in the middle range of
pressures at 200 and 250  C (Figs. 13 and S7) deviate from both
the experimental data and the results obtained from the
thermodynamic model, molecular simulation shows to be capable
of making an accurate prediction of the azeotropic point of this
binary system.
At 200  C, the average absolute deviation (AAD) of the
simulation results (ensemble-averages) for the mole fraction of
ethanol with respect to the experimental values, is 0.0160 for the
liquid phase and 0.0557 for the vapor phase. At 250  C, the AAD is
0.0215 for the liquid phase and 0.0254 for the vapor phase. At

275  C, the AAD is 0.0096 for the liquid phase and 0.0122 for the
vapor phase. At 300  C, the AAD is 0.0134 for the liquid phase and
0.0180 for the vapor phase.
From an overall examination of the VLE phase diagrams shown
in Figs. 13 and S7S9, it follows that the intermolecular potential
models for ethanol and water described in Section 2 are capable of
qualitatively predicting the VLE behavior of this binary system.

[(Fig._15)TD$IG]

Fig. 15. VLE phase diagram of ethylene + water + ethanol at 200  C and 40 atm. *,
molecular simulation results of this work; , calculated with the PRSV2-WSUNIQUAC thermodynamic model; ~, experimental data of Tsiklis et al. [6].

6. Simulation results for the ternary mixture


By following the four-stage strategy explained in Section 3.2,
ternary-mixture VLE molecular simulations were carried out by
using the potential models of ethylene, water, and ethanol
described in Section 2. Due to the improved results obtained for
the VLE phase diagram of ethylene + water at 200  C by the use of a
correction factor x 0:9 in the Lorentz combining rule for the LJ

[(Fig._17)TD$IG]

Fig. 17. VLE phase diagram of ethylene + water + ethanol at 200  C and 60 atm. *,
molecular simulation results of this work; , calculated with the PRSV2-WSUNIQUAC thermodynamic model; ~, experimental data of Tsiklis et al. [6].

10

[(Fig._18)TD$IG]

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

Fig. 18. VLE phase diagram of ethylene + water + ethanol at 200  C and 80 atm. *,
molecular simulation results of this work; , calculated with the PRSV2-WSUNIQUAC thermodynamic model; ~, experimental data of Tsiklis et al. [6].

energy parameter of unlike interactions (see Section 5.1), such


correction factor was also used in the VLE simulations of the
ternary mixture. In Section 5, VLE phase diagrams for the binary
systems were computed at various temperatures, 200  C being a
common value for the three systems with regard to the availability
of experimental data [2123] and simulation results. Due to the
availability of experimental data [6] for the ternary system at
200  C as well, the isothermal-isobaric VLE phase diagrams of the

[(Fig._19)TD$IG]

ternary system were also computed at 200  C and at the following


pressures: 30, 40, 50, 60, 80, and 100 atm. The resulting phase
diagrams are shown in Figs. 1419, where the lled circles
correspond to the molecular simulation results, the solid lines
correspond to the results obtained from the PRSV2-WS-UNIQUAC
thermodynamic model, and the triangles correspond to the few
experimental data available [6].
In the ternary diagrams, the upper locus corresponds to the
vapor phase at its dew point and the lower locus corresponds to the
liquid phase at its bubble-point. The two loci dene the saturation
envelope for the ternary system, the two phases of which coexist in
the region within the envelope. In contrast to the few experimental
data points available and included in the phase diagrams, both the
simulation and the model predictions extend over larger ranges
and follow the same trend for the dew-point and bubble-point
lines of the ternary system. At pressures of 40 and 50 atm
(Figs. 15 and 16), simulation results accurately match the results
obtained by means of the thermodynamic model. At 100 atm
(Fig. 19), molecular simulation predicts a slightly narrower
two-phase region, indicating that the critical point of the binary
mixture ethylene + ethanol at 200  C would occur at a pressure
slightly below 100 atm. At that pressure, the bubble-point and
dew-point lines of the ternary system would intersect the
ethyleneethanol side of the phase diagram at the composition
of that critical point. In contrast, the thermodynamic model
predicts that this critical point would occur at a pressure slightly
above 100 atm. From the experimental VLE measurements
reported by Tsiklis and Kofman [22] for the binary system
ethylene + ethanol, the critical point occurs at a pressure of
98 atm at 200  C; therefore, the molecular simulation prediction
of a value slightly below 100 atm for the critical pressure at 200  C
turns out to be more accurate than the prediction made by means
of the thermodynamic model (see Fig. 12).

Fig. 19. VLE phase diagram of ethylene + water + ethanol at 200  C and 100 atm. *, molecular simulation results of this work; , calculated with the PRSV2-WS-UNIQUAC
thermodynamic model; ~, experimental data of Tsiklis et al. [6].

Y.M. Muoz-Muoz, M. Llano-Restrepo / Fluid Phase Equilibria 394 (2015) 111

7. Conclusions
From the good agreement of the Gibbs-ensemble Monte Carlo
simulation results for the vapor pressure and the VLE phase
diagrams of ethylene, water, and ethanol with respect to
calculations carried out by means of the most accurate (reference)
multiparameter equations of state currently available for those
components [1820], we were able to corroborate the validity of a
set of previously published intermolecular potential models for
ethylene [15], water [16], and ethanol [17], which had been
validated by their proponents from results obtained by means of
other simulation methods.
These potential models are capable of predicting the available
VLE phase diagrams of the binary subsystems ethylene + water [21]
(at 200 and 250  C), ethylene + ethanol [22] (at 150, 170, 190, 200,
and 220  C) and ethanol + water [23] (at 200, 250, 275, and 300  C).
For the binary system ethylene + water, it was shown that the use of
a correction factor of 0.9 for the Lorentz combining rule, which is
close to the default value of 1.0, improves the calculated phase
diagram at 200  C by decreasing by 45% the corresponding
deviations for the mole fractions of ethylene in the vapor and
liquid phases. Since for most of the simulated binary VLE phase
diagrams (11 diagrams out of a total of 12) a correction factor for
the Lorentz combining rule was not needed, and the value of the
correction factor used for ethylene + water at 200  C is not far
from the default value of 1.0, then it can be said that the
LorentzBerthelot combining rules appear to be appropriate for
the three binary systems considered in the present work. This
suitability of the LorentzBerthelot combining rules for VLE
calculations from molecular simulation is in agreement with the
ndings of previous works (e.g., [31,32,45,46]).
Molecular simulation predictions for the VLE phase diagrams of
the ternary system at 200  C and pressures of 30, 40, 50, 60, 80, and
100 atm are in very good agreement with predictions that we had
previously made [1] by use of a thermodynamic model that
combines the PengRobinsonStryjekVera equation of state, the
WongSandler mixing rules, and the UNIQUAC activity coefcient
model.
The latter agreement is encouraging for the subsequent use of
molecular simulation to predict the combined phase and chemical
equilibrium of the ternary system and check the validity of
predictions that we previously made [1] by means of the
thermodynamic model.

[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]

[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]

11

R. Stryjek, J.H. Vera, Can. J. Chem. Eng. 64 (1986) 323333.


D.S.H. Wong, S.I. Sandler, AIChE J. 38 (1992) 671680.
D.S. Abrams, J.M. Prausnitz, AIChE J. 21 (1975) 116128.
D.S. Tsiklis, A.I. Kulikova, L.I. Shenderei, Khim. Promst. (5) (1960) 401406.
A.Z. Panagiotopoulos, Mol. Phys. 61 (1987) 813826.
A.Z. Panagiotopoulos, N. Quirke, M. Stapleton, D.J. Tildesley, Mol. Phys. 63
(1988) 527545.
A.Z. Panagiotopoulos, M.R. Stapleton, Fluid Phase Equilibr. 53 (1989) 133141.
A.Z. Panagiotopoulos, Int. J. Thermophys. 10 (1989) 447457.
B. Smit, Ph. de Smedt, D. Frenkel, Mol. Phys. 68 (1989) 931950.
B. Smit, D. Frenkel, Mol. Phys. 68 (1989) 951958.
A.Z. Panagiotopoulos, Mol. Simul. 9 (1992) 123.
A.Z. Panagiotopoulos, J. Phys. Condens. Matter 12 (2000) 2552.
S.L. Weitz, J.J. Potoff, Fluid Phase Equilibr. 234 (2005) 144150.
Y.-L. Huang, T. Merker, M. Heilig, H. Hasse, J. Vrabec, Ind. Eng. Chem. Res. 51
(2012) 74287440.
T. Schnabel, J. Vrabec, H. Hasse, Fluid Phase Equilibr. 233 (2005) 134143.
J. Smukala, R. Span, W. Wagner, J. Phys. Chem. Ref. Data 29 (2000) 10531121.
W. Wagner, A. Pru, J. Phys. Chem. Ref. Data 31 (2002) 387535.
H.E. Dillon, S.G. Penoncello, Int. J. Thermophys. 25 (2004) 321335.
D.S. Tsiklis, E.V. Mushkina, L.I. Shenderei, Inzh. Fiz. Zh. 1 (8) (1958) 37.
D.S. Tsiklis, A.N. Kofman, Russ. J. Phys. Chem. 35 (1961) 549551.
F. Barr-David, B.F. Dodge, J. Chem. Eng. Data 4 (1959) 107121.
J.J. Potoff, J.R. Errington, A.Z. Panagiotopoulos, Mol. Phys. 97 (1999) 10731083.
W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, M.L. Klein, J. Chem.
Phys. 79 (1983) 926935.
J. Vrabec, H. Hasse, Mol. Phys. 100 (2002) 33753383.
I. Nezbeda, J. Kolafa, Mol. Simul. 5 (1991) 391403.
J. Vrabec, M. Kettler, H. Hasse, Chem. Phys. Lett. 356 (2002) 431436.
D. Mller, J. Fischer, Mol. Phys. 69 (1990) 463473.
D. Mller, J. Fischer, Fluid Phase Equilibr. 100 (1994) 3561.
J. Carrero-Mantilla, M. Llano-Restrepo, Fluid Phase Equilibr. 208 (2003)
155169.
J. Carrero-Mantilla, M. Llano-Restrepo, Mol. Simul. 29 (2003) 549554.
D. Frenkel, B. Smit, Understanding Molecular Simulation: from Algorithms to
Applications, Second Edition, Academic Press, 2002.
M.P. Allen, D.J. Tildesley, Computer Simulation of Liquids, Oxford University
Press, 1989.
M. Llano-Restrepo, Molecular modeling and Monte Carlo simulation of
concentrated aqueous alkali halide solutions at 25  C, Ph.D. Dissertation, Rice
University, Houston, 1994.
M. Llano-Restrepo, W.G. Chapman, J. Chem. Phys. 100 (1994) 83218339.
D.C. Rapaport, The Art of Molecular Dynamics Simulation, Cambridge
University Press, 1998.
J.J. de Pablo, J.M. Prausnitz, Fluid Phase Equilibr. 53 (1989) 177189.
R.P. Fartaria, R.S. Neves, P.C. Rodrigues, F.F. Freitas, F. Silva-Fernandes, Comp.
Phys. Comm. 175 (2006) 116121.
S. Mao, Z. Duan, J. Hu, Z. Zhang, L. Shi, Phys. Earth Planet. Inter. 185 (2011)
5360.
L. Shi, S. Mao, Geosci. Front. 3 (2012) 5158.
H. Flyvbjerg, H.G. Petersen, J. Chem. Phys. 91 (1989) 461466.
V. Harismiadis, J. Vorholz, A. Panagiotopoulos, J. Chem. Phys. 105 (1996) 8469.
R. Span, Multiparameter Equations of State: An Accurate Source of
Thermodynamic Property Data, Springer Verlag, Berlin, 2000.
N. Ferrando, P. Ungerer, Fluid Phase Equilibr. 254 (2007) 211223.
C.G. Pereira, L. Grandjean, S. Betoulle, N. Ferrando, C. Fjean, R. Lugo, J.C. de
Hemptinne, P. Mougin, Fluid Phase Equilibr. 382 (2014) 219234.

Acknowledgement
Financial support from the Colombian Administrative
Department of Science, Innovation and Technology (COLCIENCIAS),
through a research assistanship for doctoral students (Y.M.
Muoz-Muoz), is gratefully acknowledged.

Y. Mauricio Muoz-Muoz received his undergraduate degree (BS) in chemical


engineering from Universidad Nacional de Colombia, at the Manizales campus, in
2007, and his doctoral degree (in chemical engineering) from Universidad del Valle,
Cali, Colombia, in November 2014. He is currently working as a postdoctoral
researcher in the Thermodynamics and Energy Technology research group of
Professor Jadran Vrabec, at the Faculty of Mechanical Engineering of the University
of Paderborn, in Germany.

Appendix A. Supplementary data


Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.uid.2015.03.007.
References
[1] M. Llano-Restrepo, Y.M. Muoz-Muoz, Fluid Phase Equilibr. 307 (2011) 4557.
[2] D.-Y. Peng, D.B. Robinson, Ind. Eng. Chem. Fundam. 15 (1976) 5964.

Mario Llano-Restrepo is Professor of Chemical Engineering at Universidad del


Valle in Cali, Colombia. He received his Ph.D. degree (in chemical engineering) from
Rice University in 1994. His main research interests are modeling of phase and
chemical equilibria, modeling and simulation of separation processes, and
molecular simulation. In 2009, he earned a Distinguished Professor recognition
from Universidad del Valle for service and excellence in teaching since 1994.
Some of the courses he has taught are chemical thermodynamics, statistical
thermodynamics, molecular simulation, chemical reaction engineering, mass
transfer, separation operations, and separation process modeling and simulation.

Potrebbero piacerti anche