Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Fluid Phase Equilibria 165 Ž1999.

23–40
www.elsevier.nlrlocaterfluid

Vapor–liquid equilibria predictions of hydrogen–hydrocarbon


mixtures with the Huron–Vidal mixing rule
Socrates Ioannidis 1, Dana E. Knox )

Department of Chemical Engineering, Chemistry and EnÕironmental Science, New Jersey Institute of Technology, UniÕersity
Heights, Newark, NJ 07102-1982, USA
Received 15 March 1999; accepted 9 July 1999

Abstract

An excess Gibbs-equation of state Ž G E-EoS. framework based on the Huron–Vidal mixing rule, has been
applied to study vapor–liquid equilibria ŽVLE. of hydrogen–hydrocarbon mixtures. The mixing rule couples the
Peng–Robinson–Stryjek–Vera ŽPRSV. EoS with a local composition solution model. The solution model is
based on one-fluid theory treatment and assigns a single energy parameter to each binary pair. This energy
parameter relates to the preference of the molecules for like to unlike interactions. The allocation of a system’s
number of interactions to the individual species in a binary mixture, incorporates the use of size parameters
which gain significance only in the liquid phase. In a two parameter form, the framework has been used for the
simultaneous data reduction of a large number of binary and several ternary hydrogen–hydrocarbon mixtures.
These systems were taken over an extended range of pressures and temperatures. Results from the data
reduction are reported in both tabular and graphical forms. Correlations for the model parameters have been
identified with the acentric factor of the hydrocarbon in hydrogen–hydrocarbon binary mixtures. In a fully
predictive mode, the model has shown to describe well VLE of binary hydrogen–linear alkane systems.
Comparisons of these results with calculations from the Peng–Robinson ŽPR. EoS and the classical mixing rule
ŽvdW. are included. q 1999 Elsevier Science B.V. All rights reserved.

Keywords: Vapor–liquid equilibria; Equation of state; Activity coefficient; Mixture; Hydrogen; Hydrocarbon

1. Introduction

The phase behavior of hydrogen containing mixtures is of major importance to the chemical
industry, for applications such as development of hydrotreating processes Žcoal liquefaction area. , and

)
Corresponding author. Tel.: q1-973-596-3599; fax: q1-973-596-8436; e-mail: knoxd@admin.njit.edu
1
Presently at OLI Systems, 108 American Road, Morris Plains, NJ 07950, USA.

0378-3812r99r$ - see front matter q 1999 Elsevier Science B.V. All rights reserved.
PII: S 0 3 7 8 - 3 8 1 2 Ž 9 9 . 0 0 2 6 3 - 0
24 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

the cryogenic recovery of methane from synthetic gas. EoSs are frequently used for the description of
this class of highly asymmetric mixtures. Examples have included the BWR EoS ŽSagara et al. w1,2x;
Nieto and Thodos w3x. and cubic EoSs such as the Penn–State EoS Ž Graboski and Daubert w4x., and
the RK-EoS ŽNieto and Thodos w3x; Gray et al. w5x; Grevel and Chatterjee w6x.. Recent efforts in phase
behavior modeling of highly asymmetric hydrocarbon mixtures with the cubic EoS approach have
focused on a better representation of the temperature dependence of the energy parameter of the low
boiling temperature substances. For example, Twu et al. w7x chose the RK EoS to predict the hydrogen
solubility with an improved a function for hydrogen, and Floter ¨ et al. w8x used the Peng–Robinson
ŽPR. EoS with a new a function for methane to predict solid–liquid equilibria in methane–hydro-
carbon mixtures. Simple mixing rules have been traditionally used typically requiring no more than
one adjustable parameter, in order to develop a reliable means of extending the EoS predictions to
multicomponent mixtures.
A desirable attribute of a particular model extended to asymmetric hydrogen–hydrocarbon mixtures
is the ability of the model to correlate its parameters with a molecular property of the hydrocarbon.
For example, Huang et al. w9x used the acentric factor, and Gray et al. w5x used the solubility parameter
of the hydrocarbon to correlate the model parameters in binary hydrogen–hydrocarbon mixtures.
These correlations have shown to be particularly useful for the prediction of vapor–liquid equilibria
ŽVLE. in hydrogen–linear alkane systems.
In the work of Huang et al. w9x, the Gibbs-equation of state Ž G E-EoS. formalism was used as the
basis of the thermodynamic framework for these highly nonideal systems. Recently, the G E-EoS
mixing rules have been successfully applied to a range of vapor–liquid Ž Wong and Sandler, w10x. and
liquid–liquid ŽAlvarado and Sandler, w11x. systems. The VLE mixtures studied include systems at
high pressures and temperatures typical of those found in hydrogen–hydrocarbon mixtures. Under this
framework, a solution model is used for each of the phases being at equilibrium to derive the mixture
parameters of the EoS. In the work of Huang et al. w9x, the Wong–Sandler Ž Wong and Sandler w10x.
mixing rule was chosen to combine the NRTL model Ž Renon and Prausnitz, w12x. with the
Peng–Robinson–Stryjek–Vera Ž PRSV. EoS Ž Stryjek and Vera, w13x.. For this model, three binary
parameters are assigned to each pair of molecules, two from the NRTL model and one from the
mixing rule. The authors provided correlations for the parameters of the hydrogen–hydrocarbon
mixtures studied. The mixing rule parameter was correlated with temperature, and the two NRTL
parameters were correlated with the acentric factor of the hydrocarbon. These correlations predict the
model parameters in hydrogen–hydrocarbon systems where experimental data are not available.
In the work of Huang et al. w9x, only the binary parameter introduced by the Wong–Sandler mixing
rule is temperature dependent, while a temperature-independent form is chosen for the energy
interaction parameters in the NRTL model Žcf. Eq. Ž 29.. . In contrast, in the model used by Gray et al.
w5x, the pure component covolume parameters are temperature dependent, while the binary parameters
are not. Note here, that when the solution model is not treated under the athermal solution assumption,
the Huron–Vidal model introduces a certain functional dependence of the EoS energy mixture
parameter on temperature.
In this work, a G E model derived from the work of Knox et al. w14x is coupled to the PRSV EoS
ŽStryjek and Vera, w13x., in a framework based on the Huron–Vidal mixing rule Ž Huron and Vidal,
w15x.. The Huron–Vidal mixing rule is particularly appealing for these systems, since the high
temperatures and pressures occurring in their phase equilibrium do not raise any compliance issues of
the model with the second virial coefficient constraint, which comes at the expense of an additional
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 25

parameter. We have to note here that trials to satisfy the second virial coefficient constraint as shown
in the approach followed by Tochigi et al. w16x, did not give satisfactory results. The framework used
in this work has been also successfully applied to VLE of refrigerant systems ŽIoannidis and Knox,
w17x.. In order to facilitate the application of the binary parameters to multicomponent systems, a
subject not well understood as Gray et al. w5x noted, a few ternary systems have been incorporated into
the data sets chosen for data reduction. The correlation of these binary parameters with the acentric
factor of the hydrocarbon in binary hydrogen–hydrocarbon mixtures is also investigated.

2. Solution model

The local composition concept as has been derived from Wilson w18x along with Guggenheim’s
lattice model w19x, has been the basis of a model developed by Knox et al. w14x as an alternative to
two-fluid theory models used for the description of the liquid phase properties. A modification of this
model by Ioannidis and Knox w17x, has made it suitable for multicomponent mixture predictions from
only binary interaction parameters. As shown in Ioannidis and Knox w17x, the G E value for this
modified solution model is given by:
GE fi zi x i x ii
s Ý x i ln qÝ ln Ž1.
RT i xi i 2 fi
where, x i j is the local composition of j molecules around an i molecule, and f i and z ir2 Žcf. Eqs.
Ž2. and Ž 3.. are the volume fraction and the average number of interactions assigned to species i,
respectively:
x i ri
fi s Ž2.
Ý x j rj
j

z i s Ý x j pi j Ž3.
j

with,
pii s qi Ž4.
where ri and qi are the molecular volume and area fractions, respectively. These two parameters have
been derived following the lines given by Bondi w20x, except for the qi values of hydrogen, carbon
monoxide, and ethylene which were regressed from the data sets of all of the systems in which these
molecules were involved. Values for these parameters are reported in Table 1. For the size parameters
pi j and pji of Eq. Ž3., it is assumed that:
pi j s pji . Ž5.
A single energy parameter l i j relating the local mole fractions is introduced in the model as
follows:
x i j x ji 10 4l i j
x ii x j j
s exp y
ž RT / . Ž6.
26 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

Table 1
Pure component parameters
Component Tc Pc k 1i v r q
Hydrogen 33.2 13.0 0.0 y0.218 0.47 0.32
Carbon monoxide 132.91 34.96 0.04279 0.04830 0.77 0.64
Methane 190.56 45.95 y0.00159 0.01045 1.19 1.28
Ethane 305.43 48.80 0.02669 0.09781 1.80 1.70
Propane 369.82 42.4953 0.03136 0.15416 2.48 2.24
n-Butane 425.16 37.97 0.03443 0.20096 3.15 2.78
Ethylene 282.35 50.42 0.04191 0.08652 1.57 4.88
n-Hexane 507.30 30.12 0.05104 0.30075 4.5 3.86
n-Heptane 540.10 27.36 0.04648 0.35022 5.17 4.40
n-Decane 617.50 21.03 0.04510 0.49052 7.20 6.02
n-Hexadecane 720.6 14.19 0.02665 0.74397 11.24 9.26
Toluene 591.8 41.06 0.03849 0.26323 3.87 2.97

The parameters l i j and pi j are the two binary model parameters assigned to an i–j pair of molecules.
The activity coefficient of this solution model is:

fi fi z i fi x ii x j Ž p ji y z j . xjj
ln g i s 1 y
xi
q ln
xi
q
2 ž ui
y 1 q ln
fi / qÝ
j 2
ln
fj
Ž7.

where, u i denotes an area mole fraction defined as:


x i zi
ui s . Ž8.
Ý xj zj
j

On a molecular basis the local composition mass balance is written as:

Ý x i j s 1. Ž9.
j

Based on one-fluid theory, the assumption is made that the number of i–j interactions are equally
assigned to the i and j species:

u i x i j s u j x ji . Ž 10.

For a binary mixture we can solve Eqs. Ž 6. , Ž 9. and Ž 10. to obtain the local mole fractions, after
specifying the z i values from Eq. Ž3. as explained in Knox et al. w14x. We first define a ratio t of the
area fractions:

u2
ts Ž 11.
u1
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 27

and a parameter b as follows:

l12

bs
exp y žRT
y1
.
/ Ž 12.
l12
exp y ž
RT /
Then we can solve for x 12 and x 21 as follows:

2t
x 12 s Ž 13.
(
1 q t q Ž 1 q t . y 4b t
2

x 21 s x 12rt. Ž 14.
Also, from Eq. Ž9., we have the local conservation equations:

x 11 s 1 y x 12 Ž 15.
and

x 22 s 1 y x 21 . Ž 16.
For multicomponent mixtures, the solution scheme has to be performed numerically.

Table 2
%AAD for pressure and vapor phase composition of component 1 for binary systems and the mixture parameters
Components Temperature Temperature Number %AADP Ža. %AADy1Žb. li j pi j
Ž1–2. sets range ŽK. of points
ŽH 2 –CO. 2 83.30–100.00 18 4.67 2.95 0.88 2.44
ŽCH 4 –CO. 5 91.60–123.90 22 3.71 8.62 0.51 1.52
ŽH 2 –CH 4 . 4 103.15–173.05 26 3.39 2.85 0.72 2.42
ŽH 2 –C 2 H 6 . 4 148.15–223.15 16 5.32 0.71 3.14 1.61
ŽCH 4 –C 2 H 6 . 12 130.37–199.93 129 2.36 0.84 0.27 2.08
ŽH 2 –C 2 H 4 . 6 158.15–255.35 34 3.35 3.50 1.49 3.26
ŽC 2 H 6 –C 2 H 4 . 4 199.82–263.15 43 2.96 3.46 0.02 3.33
ŽCH 4 –C 2 H 4 . 10 148.09–248.37 144 1.32 0.62 0.40 2.75
ŽH 2 –C 3 H 8 . 7 173.15–323.15 41 5.66 1.07 1.58 2.22
ŽH 2 – n-C 4 H 10 . 5 327.65–394.25 60 1.87 2.50 y0.49 3.48
ŽH 2 – n-C 6 H 14 . 6 277.59–444.26 94 4.23 1.26 y0.64 3.74
ŽH 2 – n-C 7 H 16 . 3 424.15–498.85 32 8.01 2.49 y0.83 3.85
ŽH 2 – n-C 10 H 22 . 4 462.45–583.45 26 3.98 1.72 y1.84 4.44
ŽH 2 – n-C 16 H 34 . 4 461.65–664.05 28 14.07 0.63 y3.07 5.34
ŽH 2 –C 7 H 8 . 3 461.85–542.15 20 1.57 1.70 y0.11 3.25
Ža. N <Ž
Ž100rN .Ý is1 Ppr-Pexp .rŽ Pexp .<.
Žb. N <Ž
Ž100rN .Ý is1 ypr-yexp .rŽ yexp .<.
28 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

3. Mixing rule

The PRSV EoS Ž Stryjek and Vera, w13x. is implemented in our framework:
RT aŽ T .
Ps y Ž 17.
Õyb ÕŽ Õqb. qbŽ Õyb.
where Õ denotes molar volume, aŽ T . and b are the energy and size parameters of either the mixture
Ždenoted as a m , bm . or a pure species i Ž denoted as a i , bi . , T is the temperature, and R is the ideal
gas constant. The EoS energy and size parameters for a pure species i are given by Eqs. Ž 18. and
Ž19.:
R 2 Tc2i
a i Ž T . s 0.457235 ai ŽT . Ž 18 .
Pc i
RTc i
bi s 0.077796 Ž 19.
Pc i
where Tc i and Pc i are the critical temperature and pressure of species i, respectively. The alpha
function in Eq. Ž18. is given by:
2
a i Ž T . s 1 q k i Ž 1 y Tr0.5 . Ž 20.

Fig. 1. Experimental points and VLE calculations with the two Žlines. and one-parameter Žcross-haired points. models for the
system hydrogen–methane.
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 29

where Tr is the reduced temperature and k i is given by the following function:


k i s k o i q k 1i Ž 1 y Tr0.5 . Ž 1 q Tr0.5 . Ž 0.7 y Tr . Ž 21.
where k 1i and k o i are pure component parameters with k o i given from a global correlation to the
acentric factor of pure species i:
k o i s 0.378893 q 1.4897153 v y 0.17131848 v 2 q 0.0196544v 3 . Ž 22.
The form of the k function Eq. 21 , along with the correlation given in Eq. 22 , allows the
Ž Ž .. Ž .
equation of state not only to reproduce the universal function existing between the a function and the
acentric factor at a reduced temperature 0.7, but also to fit experimental vapor pressures over a range
of temperatures, as explained in Zabaloy and Vera w21x. This EoS has also been used in the work of
Huang et al. w9x. The values of the critical temperature and pressure, along with the acentric factor and
the k 1i values, for each component are included in Table 1. Values for these parameters for several
molecules are given by Stryjek and Vera w13x. Proust and Vera w22x report the parameters for ethylene
and carbon monoxide, while those of hydrogen were taken from Reid et al. w23x.
The mixture energy parameter Ž a m . for the Huron–Vidal mixing rule is given by:
ai G`E
a m s bm Ý x i q Ž 23.
bi C
where, a i and bi are the EoS energy and size parameters of the pure species i, respectively, x i is its
mole fraction in either the liquid or vapor phase, C is a constant dependent on the EoS used Žy0.62

Fig. 2. Experimental points and VLE calculations with the two Žlines. and one-parameter Žcross-haired points. models for the
system hydrogen–carbon monoxide.
30 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

for the PRSV EoS. , G`E is the Gibbs energy taken at the infinite pressure limit Žobtained from Eq.
Ž1.., and bm is the EoS mixture size parameter for which a linear mixing rule is assumed:

bm s Ý x i bi . Ž 24.
i

VLE were performed with the application of the isofugacity equilibrium condition:

f iL s f iV Ž 25.
where f i denotes the fugacity of species i in either the liquid Žsuperscript L. or vapor Ž superscript V.
phase. For the PRSV EoS and the suggested mixing rule, the fugacity coefficient of species i in the
mixture in either phase is given by:

P Ž Õm y bm . bi PÕ 1 ai lng i` Õm q bm Ž 1 y '2 .
ln w i s yln
RT
q
bm RT ž /
y1 q
2'2 bi RT
q
C
ln
Õm q bm Ž 1 q '2 .
Ž 26.
where m denotes mixture properties and ln g i` is the activity coefficient at the infinite pressure limit
Žobtained from Eq. Ž7...

Fig. 3. Experimental points and VLE calculations with the two Žlines. and one-parameter Žcross-haired points. models for the
system hydrogen–ethylene.
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 31

4. Results

The experimental VLE data for the chosen systems were mostly obtained from the Dechema Data
Series ŽKnapp et al. w24x.. The system hydrogen–decane was taken from Sebastian et al. w25x, the
system hydrogen–n-hexadecane was taken from Lin et al. w26x, and the ternary systems hydrogen–
ethylene–methane and hydrogen–ethylene–ethane were taken from Sagara et al. w1x.
The regression of the model parameters was performed with bubble point pressure calculations and
N Ž
the use of the summation Ý is1 Ppr y Pexp . 2 as the objective function. A large number of data sets
spanning a range of temperatures and pressures were used for each system. In Table 2, we report the
results of the percentage absolute average deviations in pressure Ž %AADP. and vapor phase
composition of the first component Ž %AADy1 . . Results for several of the binary systems are
presented in graphical form in Figs. 1–6. These systems are reported at low temperatures and they
cover a range of low ŽCO–CH 4 . to high ŽH 2 –C 2 H 4 . pressures. In Table 2, we also report the values
of the two parameters for all the binary systems studied. By examining the values of these parameters,
we observe that the energy parameters decrease as the acentric factor of the hydrocarbon increases.
On the basis of the solution model used, this fact can be physically interpreted as an increase in the
number of unlike interactions as the asymmetricity of the mixture increases. Also, along these lines of
interpretation, we should expect a higher local mole fraction of hydrogen molecules around a central
hydrocarbon molecule, than of a hydrocarbon molecule around a central hydrogen molecule, mainly
due to steric effects. Indeed, this is the case from our solution model. For example, for the

Fig. 4. Experimental points and VLE calculations Žlines. for the system methane–ethylene.
32 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

Fig. 5. Experimental points and VLE calculations Žlines. for the system methane–ethane.

Fig. 6. Experimental points and VLE calculations Žlines. for the system carbon monoxide–methane.
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 33

Table 3
Results for two hydrogen–hydrocarbon binaries with the three-parameter model
Components Ž1–2. li j pi j Pji %AADP %AADy1
ŽH 2 – n-C 7 H 16 . 1.24 2.94 5.21 4.48 3.04
ŽH 2 – n-C 16 H 34 . y0.75 4.01 6.09 3.09 0.82

hydrogen–n-hexadecane equimolar system at 664 K, the model predicts a hydrogen local composition
around a hydrocarbon molecule of 0.72, and a hydrocarbon local composition around a hydrogen
molecule of 0.41.
We note in Table 2 that the pressure deviations for the heptane and hexadecane systems are greater
than for the rest of the systems. For these two systems, we relaxed the assumption of Eq. Ž 5. and
fitted the energy and the two size parameters anew. These results are shown in Table 3. As Table 3
shows, the %AADP values have been significantly reduced. We note in Table 3 that the energy
parameter decreases as the acentric factor of the hydrocarbon increases, as was observed for the
systems in Table 2. Also, the values of the size parameters for each system bracket the corresponding
value shown in Table 2. It is interesting to note that the values of the size parameters in Table 3 will
result in a higher number of interactions assigned to the hydrocarbon molecule than will those in
Table 2. At the same time, the energy parameters in Table 3 point towards an increase of the like
interactions, as compared to the two-parameter model. This implies that the three-parameter model
shares the same characteristics as the two-parameter model. This is probably so, due to their common
functional form for the energy parameter. Although it is difficult to assign a physical significance to
the molecular parameters that resulted from the data reduction procedure, we can imagine that a
central hydrogen molecule ‘‘observes’’ less of a hydrocarbon molecule, than a central hydrocarbon
molecule does a hydrogen molecule, which is smaller in size. This correctly translates, by Eq. Ž 3. , to a

Fig. 7. Experimental points and VLE calculations Žlines. for the ternary system hydrogen–methane–ethane at 144.26 K and
68.95 bars.
34 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

Fig. 8. Experimental points and VLE calculations Žlines. for the ternary system hydrogen–carbon monoxide–methane at
163.17 K and 68.95 bars, and 120.00 K and 50.00 bars.

higher number of interactions assigned to the hydrocarbon molecule. The values of the energy
parameter also draw from this analogy, by considering that the substitution of C 16 H 34 in place of
C 7 H 16 increases the unlike interactions in the system, and, hence, the energy parameter decreases.
In Figs. 7–10, we show calculations for a few ternary systems with the binary parameters reported
in Table 2. In Fig. 7, the system hydrogen–methane–ethane is shown at 144.26 K and 69 bars. In Fig.
8, the system hydrogen–methane–carbon monoxide, a key ternary in the cryogenic recovery of
methane ŽGray et al. w5x., is shown at two conditions. In Figs. 9 and 10, we plot the ternary systems

Fig. 9. Experimental points and VLE calculations Žlines. for the ternary system hydrogen–ethylene–methane at 123.15 K
and 20.26 bars, and 198.15 K and 60.78 bars.
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 35

Fig. 10. Experimental points and VLE calculations Žlines. for the ternary system hydrogen–ethylene–ethane at 148.15 K and
20.26 bars, and 223.15 K and 81.04 bars.

hydrogen–ethylene–methane and hydrogen–ethylene–ethane at two conditions. These systems are


important ternaries in the purification of ethylene in demethanizers Ž Gray et al. w5x.. Comparisons of
our calculations for each ternary system and at several conditions, with the models used in the
references where the experimental data were taken from, are shown in Table 4. In particular, our
model predictions for the systems H 2 –CO–CH 4 and H 2 –CH 4 –C 2 H 6 are compared with the PR and
SRK EoSs for which the classical mixing rule was used ŽKnapp et al. w24x.. The ternary systems
H 2 –C 2 H 4 –CH 4 and H 2 –C 2 H 4 –C 2 H 6 are compared to the BWR equation calculations of Sagara et
al. w1x. Note, that these comparisons were made on exactly the same data set.
Large percentage errors are observed for the mole fractions of ethylene, methane, and ethane in the
vapor phase for the last two ternary systems in Table 4, due to the small numerical values of a few
experimental mole fraction data. Deletion of only five points for the system H 2 –CH 4 –C 2 H 4
decreases the absolute average deviation of methane by more than 4%. Overall, we note that the

Table 4
%AAD in pressure and vapor phase mole fractions for the ternary systems
Components Temperature sets Temperature span ŽK. %AADP %AADy1 %AADy2 %AADy3
Ž1–2–3. ŽNumber of points. wPressure span Žbars.x
ŽH 2 –CO–CH 4 . 12–95 120.00–173.25 w28.9–103.42x 2.85 6.29 2.65 3.64
PRa w24x 6.32 15.41 4.04 3.57
ŽH 2 –CH 4 –C 2 H 6 . 4–15 144.26–199.82 w34.47–68.95x 3.61 3.21 4.35 12.66
SRK w24x 6.03 8.55 7.55 20.41
ŽH 2 –C 2 H 4 –CH 4 . 20–98 123.15–248.15 w20.26–81.04x 4.67 5.72 9.27 17.86
BWR w1x 13.55 5.66 13.43 37.38
ŽH 2 –C 2 H 4 –C 2 H 6 . 16–80 148.15–223.15 w20.26–81.04x 4.69 0.92 16.21 11.98
BWR w1x 17.60 1.44 21.15 31.60
a
For the T – P ŽK, bars. sets of Ž120, 50., Ž140, 28., Ž140, 50., Ž163, 27., and Ž173, 55., the SRK EoS was used in Ref. w24x.
36 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

errors for both the pressure and the vapor phase mole fractions are low, although only the pressure
data were included in the objective function.

5. Binary hydrogen–hydrocarbon mixtures

The hydrogen–hydrocarbon energy parameters in Table 2 are plotted vs. the hydrocarbon’s acentric
factor in Fig. 11. We observe from Fig. 11 that the parameters for the normal alkanes, with the
exception of methane, are well correlated. Carbon monoxide and toluene seem to also correlate well,
while ethylene shows the highest deviation. A linear regression of all of the points in Fig. 11 gives:
pi j s 3.322 y 0.588 l i j . Ž 27.
The overall correlation coefficient of Eq. Ž 27. is 0.90. This value improves to 0.97 when ethylene is
not included in the data regression. In Fig. 12, we plot the binary size parameter of the systems, as in
Fig. 11, vs. the acentric factor of the hydrocarbon. As we can see, a straight line provides a
satisfactory correlation of the data points, with ethane, ethylene, and propane showing the highest
deviations. The equation of the line describing the data is:
pi j s 2.171 q 4.431 v j Ž 28.
with a correlation coefficient of 0.81, which improves to 0.90 if ethylene is excluded from the data
regression. Eqs. Ž 27. and Ž 28. were further applied for predictions of the binary hydrogen systems

Fig. 11. Regression plot of the size and energy parameters for binary hydrogen–hydrocarbon systems.
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 37

Fig. 12. Regression of the acentric factor of the hydrocarbon vs. the size parameter in binary hydrogen–hydrocarbon
systems.

reported in Table 2. The results for the methane, carbon monoxide, and ethylene were not satisfactory.
For these systems, application of a single parameter model with the l i j parameter obtained from
Table 2 and the pi j obtained from Eq. Ž28., is shown in Figs. 1–3. Results from the fully predictive
model for the rest of the binary hydrogen systems are shown in Table 5, and compared to the
calculations with the PRrvdW model. The binary parameters of the classical mixing rule were
obtained from Han et al. w27x, except for the hydrogen–ethane system, which was regressed to 0.161
from the data sets in Table 2. The %AADP and %AADY1 values for the hydrogen–methane system
with the PRrvdW model, not included in Table 5, were 13.06 and 4.47, respectively.

Table 5
Results for binary hydrogen systems with Eqs. Ž27. and Ž28. Ž C . and with the PRrvdW model ŽvdW.
Components Ž1–2. %AADP %AADy1
Model C vdW C VdW
ŽH 2 –C 2 H 6 . 16.04 5.16 0.98 0.57
ŽH 2 –C 3 H 8 . 13.14 17.88 1.40 0.64
ŽH 2 – n-C 4 H 10 . 10.95 5.87 2.39 1.98
ŽH 2 – n-C 6 H 14 . 8.67 21.34 0.69 1.07
ŽH 2 – n-C 7 H 16 . 8.85 42.66 2.45 5.63
ŽH 2 – n-C 10 H 22 . 2.91 7.09 2.45 2.35
ŽH 2 – n-C 16 H 34 . 15.49 13.21 0.56 1.36
ŽH 2 –C 7 H 8 . 8.08 4.78 1.16 3.74
38 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

Table 6
NRTL parameter regression and correlation for a few hydrogen–hydrocarbon systems ŽHuang et al. w9x.
Components Ž1–2. T ŽK. Correlation Prediction
t 12 t 21 t 12 t 21
ŽH 2 –ethane. 283 0.134 0.456 0.69 0.39
255 y0.150 0.500
227 0.499 0.499
199 y0.359 1.569
ŽH 2 –decane. 583 0.208 0.448 0.24 0.24
543 y0.029 0.550
503 y0.188 0.604
462 y0.275 0.600

In contrast to our solution model, the NRTL parameter correlations reported by Huang et al. w9x, do
not appear to reproduce the regressed parameters, as Table 6 shows for two representative systems.
The energy function of the NRTL model used by Huang et al. w9x is given in Eq. Ž 29. :
Gji s exp Ž ya jit ji . Ž 29.
where t i j and t ji are the two energy parameters, with the nonrandomness a i j parameter being fixed
to 0.36 for all binary systems.

6. Conclusions

In this work, we demonstrated the utility of the Huron–Vidal mixing rule and a one-fluid theory
based solution model, as an alternative to other G E-EoS models and the classical mixing rule, in
modeling hydrogen-containing mixtures. As the results in Table 4 showed, the incorporation of
ternary systems in the data reduction procedure insures an overall good representation of the VLE for
these ternaries. Despite the bias introduced by the inclusion of the ternary systems, the resulting
binary parameters have been correlated well with the acentric factor of the hydrocarbon. The fully
predictive model adequately describes the hydrogen–linear alkane systems, with the exception of
methane Žcf. Table 5..
An extra parameter can be introduced in the model by relaxation of Eq. Ž5.. This addition of an
extra size parameter in the model does not break its one-fluid character, since only the average
number of interactions per species is changed. This regression exercise has been carried out for two
systems that showed larger deviations in pressure in Table 2. The pressure deviations for the two
systems were greatly reduced Žcf. Table 3. . Clearly though, fewer parameters seem to be adequate for
most of the systems examined.
Regarding the effect of the solution model on the calculations, we note that the vapor phase of
these systems mostly contains hydrogen and so the solution model accounts mainly for the description
of non-idealities in the liquid phase, while the PRSV EoS alone carries this task for the vapor phase.
This, as shown, allows the local composition character of the solution model to capture the
asymmetricity observed in hydrogen–linear alkane systems. Furthermore, the values of the l i j
S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40 39

parameters for the CH 4 –C 2 H 6 , CH 4 –C 2 H 4 , and C 2 H 6 –C 2 H 4 are close to zero, which indicates the
near athermal character of these mixtures.

7. Notation

Alphabetic
a Equation of state parameter
b Equation of state parameter
C EoS dependent parameter
G Gibbs energy, NRTL energy function
f Fugacity
N Number of data points
P Pressure
p Size parameter
q Molecular area parameter
r Molecular volume parameter
R Universal gas constant
T Temperature
t Area fraction ratio
Õ Molar volume
x Overallrlocal mole fraction
y Vapor phase mole fraction
z Characteristic number for pair interactions
Greek letters
a Energy function of the EoS a mixture parameter, NRTL nonrandomness parameter
b Combined energy parameter
g Activity coefficient
u Area fraction
k Pure component parameter in the PRSV EoS
l Energy parameter
t NRTL binary energy parameter
f Volume fraction
w Fugacity coefficient
Superscript
E Excess property
L Liquid
V Vapor
Subscript
` Infinite pressure state
c Critical property
exp Experimental
i, j Component species
40 S. Ioannidis, D.E. Knox r Fluid Phase Equilibria 165 (1999) 23–40

m Mixture
pr Predicted
r reduced property

References

w1x H. Sagara, Y. Arai, S. Saito, J. Chem. Eng. Jpn. 5 Ž1972. 339–348.


w2x H. Sagara, S. Mihara, Y. Arai, S. Saito, J. Chem. Eng. Jpn. 8 Ž1975. 98–104.
w3x M.G. Nieto, G. Thodos, AIChE J. 24 Ž1978. 672–678.
w4x M.S. Graboski, T.G. Daubert, Ind. Eng. Chem. Proc. Des. Dev. 18 Ž1979. 300–306.
w5x R.D. Gray, J.L. Heidman, S.C. Hwang, C. Tsonopoulos, Fluid Phase Equilib. 13 Ž1983. 59–76.
w6x K.D. Grevel, N.D. Chatterjee, Eur. J. Mineral. 4 Ž1992. 1303–1310.
w7x C.H. Twu, J.E. Coon, A.H. Harvey, J.R. Cunningham, Ind. Eng. Chem. Res. 35 Ž1996. 905–910.
w8x ¨ T.W. de Loos, J. de Swaan Arons, Ind. Eng. Chem. Res. 37 Ž1998. 1651–1662.
E. Floter,
w9x H. Huang, S.I. Sandler, H. Orbey, Fluid Phase Equilib. 96 Ž1994. 143–153.
w10x D.S.H. Wong, S.I. Sandler, AIChE J. 38 Ž1992. 671–680.
w11x G.N.E. Alvarado, S.I. Sandler, AIChE J. 44 Ž1998. 1178–1187.
w12x H. Renon, J.M. Prausnitz, AIChE J. 14 Ž1968. 135–144.
w13x R. Stryjek, J.H. Vera, Can. J. Chem. Eng. 64 Ž1986. 323–333.
w14x D.E. Knox, H.C. Van Ness, H.B. Hollinger, Fluid Phase Equilib. 15 Ž1984. 267–285.
w15x M.J. Huron, J. Vidal, Fluid Phase Equilib. 3 Ž1979. 255–271.
w16x K. Tochigi, P. Kolar, T. Iizumi, K. Kojima, Fluid Phase Equilib. 96 Ž1994. 215–221.
w17x S. Ioannidis, D.E. Knox, Fluid Phase Equilib. 140 Ž1997. 17–35.
w18x G.M. Wilson, J. Am. Chem. Soc. 86 Ž1964. 127–130.
w19x E.A. Guggenheim, Mixtures, Clarendon Press, Oxford, Great Britain, 1952.
w20x A. Bondi, Physical Properties of Molecular Crystals, Liquids and Glasses, Wiley, New York, 1968.
w21x M.S. Zabaloy, J.H. Vera, Ind. Eng. Chem. Res. 37 Ž1998. 1591–1597.
w22x P. Proust, J.H. Vera, Can. J. Chem. Eng. 67 Ž1989. 170–173.
w23x R.C. Reid, J.M. Prausnitz, B.E. Poling, The Properties of Gases and Liquids, 4th edn., McGraw-Hill, New York, 1987.
w24x H. Knapp, R. Doring, L. Oellrich, U. Plocker, J.M. Prausnitz, Vapor–Liquid Equilibria for Mixtures of Low Boiling
Substances. Chemistry Data Series, Vol. VI, Dechema, Frankfurt, 1982.
w25x H.M. Sebastian, J.J. Simnick, H.M. Lin, K.C. Chao, J. Chem. Eng. Data 25 Ž1980. 68–70.
w26x H.M. Lin, H.M. Sebastian, K.C. Chao, J. Chem. Eng. Data 25 Ž1980. 252–254.
w27x S.J. Han, H.M. Lin, K.C. Chao, Chem. Eng. Sci. 43 Ž1988. 2327–2367.

You might also like