Sie sind auf Seite 1von 8

Phase Equilibrium Behavior in Water (1) + n-Alkane (2) Mixtures

Christian Bidart,

Hugo Segura,*
,
and Jaime Wisniak

Departamento de Ingenier a Qu mica, Facultad de Ingenier a, UniVersidad de Concepcion POB 160-C, Correo
3, Concepcion, Chile, and Department of Chemical Engineering, Ben-Gurion UniVersity of the NegeV,
Beer-SheVa 84105, Israel
The natural water content in hydrocarbon mixtures in a wide range of concentrations, temperature, and pressure
conditions leads to important technological changes in the processes involved in the petroleum industry. For
this reason, the physical understanding and mathematical modeling of these aqueous-organic mixtures
constitutes a challenging task, both for scientists and for applied engineers. The present work focuses on the
description of the phase behavior of binary mixtures of water (1) + n-alkane (2) [n) 1-36], considering the
topological approach introduced by van Konynenburg and Scott. The phase behavior of the mixtures under
analysis is predicted from a Carnahan Starling van der Waals equation of state with quadratic mixing rules,
which is in qualitative agreement with experimental results. In terms of global transitional mechanisms, we
describe the topological changes that systematically occur in the phase diagrams of water (1) + n-alkane (2)
mixtures as the carbon number n of the aliphatic chain increases. These transitional mechanisms give coherence
to the barotropic phenomena (phase density inversion) that is observed along the three-phase equilibrium line
of water (1) + n-alkane (2) mixtures for n ) 28-36.
Introduction
Water (1) + n-alkane (2) mixtures constitute a typical case
of liquid-phase immiscibility, where the equilibrium behavior
is not only of scientific interest, but also important in engineer-
ing
1
and geological
2
applications. Certainly, mixtures of natural
hydrocarbons are inevitably in contact with water, influencing,
for example, petroleum extraction processes from oil wells,
3
ordinary hydrocarbon processing,
4
or groundwater remediation
procedures during oil spill events. At room temperature, the
phase equilibrium behavior of water (1) +n-alkane (2) mixtures
may be treated with fair accuracy using the ideal simplification
of absolute liquid-phase immiscibility. However, this is not the
case when an accurate prediction of the solubility is needed, or
when we consider the equilibrium behavior in a wide range of
temperatures and/or pressures, especially in the vicinity of the
critical point.
5
Tsonopoulos
6,7
has presented a critical collection of experi-
mental data for the mutual solubility of water + hydrocarbons
in the low-temperature range. Additionally, in an authoritative
experimental research, Brunner
7
reported gas-liquid and gas-
liquid-liquid equilibrium data for water (1) + n-alkane (2)
mixtures [n ) 1-36] from low temperature up to the critical
range, including the critical lines of the pertinent binary
mixtures. From his experimental results, Brunner observed the
following:
According to the van Konynenburg and Scott
8
classification,
aqueous-organic mixtures with n g 2 classify as Type III
systems, where a branch of the critical line ends in an upper
critical end point (UCEP) of a three-phase line (3PL or
heteroazeotropic line), while the other one diverges to the high-
pressure range.
In all of the investigated binaries, the pressure of the 3PL is
larger than the vapor pressure of the pure components, thus
exhibiting heteroazeotropic behavior of the first kind
9
(i.e., the
vapor-phase mole fraction lies between the mole fractions of
the immiscible liquids).
The temperature and pressure of the UCEP increase mono-
tonically with the carbon number n of the aliphatic chain.
When 2 e n e 24, and as shown in Figure 1, the critical line
connecting the critical point of the hydrocarbon [2] ends in an
UCEP characterized by a gas-liquid critical point and an
aqueous-rich liquid phase. In the work of Brunner, such a
behavior was classified as subtype III
a
.
For the case of 26 e n e 32, and as shown in Figure 2, the
critical line connecting the critical point of pure water [1] ends
in an UCEP characterized by a gas-liquid critical point and an
organic-rich liquid phase. Brunner classified this latter behavior
as subtype III
b
.
Besides these previous experimental facts, Brunner also found
clear evidence of barotropic inversion during the transition from
Type III
a
to III
b
. The barotropic phenomenon, which was
discovered experimentally by Kamerlingh Onnes,
10
is character-
ized by mass density inversion, a condition in which the relative
position of immiscible phases changes in a gravity field.
Barotropy may be frequently observed in mixtures that exhibit
gas-gas immiscibility (for example, helium + hydrogen
11
or
carbon dioxide + linear hydrocarbons
12
), although, as Rowlin-
son
13
pointed out, gas-gas equilibrium is not a necessary
condition for barotropy.
For the case of the barotropic behavior of water (1) +
n-alkane (2) mixtures, phase density inversion between im-
miscible liquids is observed for 26 < n e 36. In addition, the
temperature of the barotropic point decreases monotonically as
n increases, exhibiting an interesting coordination with the
thermal dependence of the UCEP. In fact, as can be seen in
Figure 3, the temperature of the UCEP reported by Brunner
shows a singular break point for a fictitious value 26 e n e
28, where, according to the available experimental evidence,
barotropy seems to begin and to persist with increasing carbon
number n. Brunner pointed out that the singular point where
the temperatures of the UCEP and the barotropic point intersect
corresponds to an equilibrium state where three phases of equal
mass density become critical at the UCEP, thus reaching a
tricritical end point (TCEP).
TCEPs are of theoretical interest because, in such a condition,
* To whom correspondence should be addressed. E-mail: hsegura@
diq.udec.cl.

Departamento de Ingenier a Qu mica.

Department of Chemical Engineering.


947 Ind. Eng. Chem. Res. 2007, 46, 947-954
10.1021/ie061036o CCC: $37.00 2007 American Chemical Society
Published on Web 12/24/2006
a binary system violates Gibbss phase rule as happens in every
contour of a global phase diagram
8
(GPD). Constraining our
attention to binary mixtures that exhibit heteroazeotropic
behavior from the low temperature up to the critical gas-liquid
range, in an ordinary GPD it is possible to recognize three
classes of TCEPs: van Laar points,
14
where three phases become
critical and the 3PL asymptotes the critical line in a tangent
UCEP point; any one of the vertexes of the so-called shield
region,
15
where three phases become critical and reach equi-
librium with a fourth subcritical phase; and, although not clearly
recognized, the two tricritical vertexes of the, almost unknown
and possibly unlikely, sword region,
16
which exhibit features
similar to those of the vertexes of the shield region.
No experimental systems have been found exhibiting TCEPs,
and, indeed, water + hydrocarbon mixtures are not the excep-
tion. This is so because the apparent TCEP that Brunner invokes
for explaining his experimental observations occurs for a
hypothetical non-integer n value. Anyway, the relation between
TCEPs and barotropy is interesting because, in such a case, the
density inversion phenomena could be explained in terms of
critical transitions only. Concerning this latter point, in a recent
publication by Quinones-Cisneros,
17
it has been established and
demonstrated that barotropy appears to be intrinsically linked
to weak transitions between Types II and III. However,
according to the topological features of their P-T projections,
water +n-alkane mixtures display Type III phase behavior only,
and no transition to Type II is observed. In addition, considering
that the 3PL exhibits heteroazeotropy of the first kind and that
no homogeneous azeotrope is present, water + n-alkane
mixtures should be located at the upper range of a GPD, in the
vicinity of the interesting shield region, where, with the
exception of mixtures composed by molecules of equal sizes,
little has been discussed concerning a detailed behavior of phase
diagrams.
This work is devoted to analyzing in-depth the connectivity
between barotropy and the phase behavior of water (1) +
n-alkane (2) mixtures, using the approach of global phase
diagrams. For this purpose, we introduce and describe a special
projection of the GPD that depends on the carbon number n of
the hydrocarbon that mixes with water. Such a projection of
the GPD is applied to describe the mechanisms that induce
transition from Type III
a
to III
b
. The capability of these
mechanisms for predicting barotropy is then analyzed and
contrasted with Brunners hypothesis.
Theory and Methods
Let us assume that water, n-alkanes, and their mixtures can
be, at least, qualitatively represented by the Carnahan-Starling-
Redlich-Kwong (CSRK; Carnahan and Starling
18
) equation of
state (EOS) in the high-temperature/pressure range. The CSRK
model is given by the following pressure function:
Figure 1. Schematic P-T projection for binary mixtures of water (1) +
n-alkane (2) [n ) 2-26]. Type IIIa systems. (-) Critical line, (- - )
heteroazeotropic line, ( ) vapor pressure line.
Figure 2. Schematic P-T projection for binary mixtures of water (1) +
n-alkane (2) [n ) 28-36]. Type IIIb systems. (-) Critical line, (- - )
heteroazeotropic line, ( ) vapor pressure line.
Figure 3. Thermal dependence of the UCEP and barotropic points in water
(1) + n-alkane (2) mixtures. (0-0) UCEP temperature, (O-O) barotoropic
temperature.
P )
RT
V
1 + +
2
-
3
(1 - )
3
-
a
V(V + b)
(1)
948 Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007
where P is pressure, R is the gas constant, T is temperature, V
is molar volume, and is the volume packing factor defined as
)b/(4V). In addition, a and b are the cohesion and covolume
parameters, respectively, and are defined by the following
relations:
Equation 1 may be applied to binary mixtures considering the
following quadratic mixing rules:
where the cross parameters are calculated as
In eq 4, k
12
is an interaction parameter for the cohesion term. It
should be pointed out that the CSRK-EOS and the mixing rules
presented in eq 3 may not be the most accurate selection for
representing the mixtures considered in this work. However,
as mixtures approximate high-temperature gas-liquid critical
points, polarity and rotational degrees of freedom do not affect
significantly the phase behavior of binary systems. In addition,
in a recent paper, Wang and Wong
19
demonstrated the qualita-
tive superiority of quadratic mixing rules and simple EOS
models for the simultaneous prediction of liquid-liquid im-
miscibility and the entire P-T projection for water + n-alkane
mixtures.
The theoretical basis of GPDs has been discussed in-depth
in previous references
20,21
so that our discussion is limited to
present a brief summary, as required to understand the scope
of this work. A GPD is a parametric plot whose coordinates
are the parameters of the EOS model used in its construction.
It maps entire regions inside of which mixtures, as predicted
by the EOS model, exhibit common phase behavior (for
example, presence or absence of azeotropy, presence or absence
of partial miscibility, etc.). In this work, the parameters of the
GPD for the CRSK model have been defined in agreement with
those used in the work of van Konynenburg and Scott:
As follows from eq 2, the cohesion parameter a
i
depends on
T
r
. Consequently, we have selected T
c1
as the reference
temperature for reducing the critical temperature of component
[2] at its critical temperature. In addition, the transitional
mechanisms required for calculating the GPD contours as in
the cases of the tricritical line (TC), double critical end points
(DCEP), critical pressure step points (CPSP), critical azeotropic
end points (CAzEP), zero temperature (ZT) states, and additional
transition boundaries have been calculated according to the
procedures reported by Boshkov,
22
Kraska,
20,21
and Deiters.
23
Figure 4 depicts the complete version of the GPD of CSRK
binary fluids composed by molecules of equal size. Our
calculations are in excellent agreement with the partial results
reported by Kraska and Deiters
20
for the same EOS model,
suggesting that our procedures for calculating transitional
boundaries are reliable. However, contrary to the claim of these
authors, we have verified that CSRK binary fluids exhibit van
Laar points (vLP) at ) (0.4056, ) 0.0835, and, conse-
quently, CSRK mixtures also exhibit Type IV* behavior.
Constraining our attention to the traditional transitions
between Types II, III, IV, IV*, and V, the GPD in Figure 4
exhibits common characteristics with the GPD of the van der
Waals equation. However, in contrast to this latter EOS, the
CPSP and DCEP contours of the CSRK model present
mathematical singularities that allow the presence of mecha-
nisms able to predict closed loops (CLP) of immiscibility (the
pertinent range is indicated by the CLP area in Figure 4. For
further details concerning phase diagrams and transitional
mechanisms that give place to CLP behavior in Figure 4, see
the paper of Yelash and Kraska
24
).
As indicated previously, and as is shown in Figures 1 and 2,
water (1) + n-alkane (2) mixtures present Type III behavior
and exhibit heteroazeotropy of the first kind. In Figure 4, the
topology of such an equilibrium behavior may be found inside
the III-H domain, above the shield region.
It should be pointed out that the GPD projection in Figure 4
is useful for comparing EOS models. However, its usefulness
for analyzing water + hydrocarbon series has the drawback that
the molecular size of the hydrocarbon that mixes with water
changes with the carbon number n. Inspection of eqs 2, 4, and
5 reveals that the global coordinates (, , ) depend on the
a ) 0.4618
(RT
c
)
2
P
c
T
r
-1/2
b ) 0.1050
RT
c
P
c
(2)
a )

j
x
i
x
j
a
ij
b )

j
x
i
x
j
b
ij
(3)
a
12
)

a
1
a
2
(1 - k
12
)
b
12
)
(b
1
+ b
2
)
2
(4)
)
b
2
- b
1
b
2
+ b
1
)
a
2
/b
2
2
- a
1
/b
1
2
a
2
/b
2
2
+ a
1
/b
1
2
)
a
2
/b
2
2
- 2a
12
/(b
1
b
2
) + a
1
/b
1
2
a
2
/b
2
2
+ a
1
/b
1
2
(5)
Figure 4. GPD of a binary CSRK fluid with molecules of equal size ( )
0). (-) TC line, (- - ) CPSP line, ( ) CAzEP line, (- -) DCEP
line, (- -) ZT line, (bold - -) III-A to III-H boundary.
Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007 949
critical properties of the pure component and on the interaction
parameter k
12
. Furthermore, as we can see in Figure 5, the critical
properties of linear hydrocarbons are well correlated in terms
of the carbon number n (36 g n g 2) using the following
correlations:
where the parameters c
i
(T)
, c
i
(P)
are reported in Table 1.
Consequently, and provided that T
c1
, P
c1
are the critical
properties of water, we can deduce that water (1) + n-alkane
(2) binary mixtures may be globally parametrized by the
following functions of global coordinates:
and, therefore, its global behavior may be mapped to a
bidimensional n-k
12
plane, a projection that we have called
serial-prediction-domain GPD (or spd-GPD).
Results and Discussion
Figure 6 depicts the spd-GPD projection for water (1) +
n-alkane (2) mixtures, as predicted by the CSRK model. Here,
we can observe that the TC and DCEP transitions affect the
global phase behavior inside the range of the experimental
measurements made by Brunner (20 e n e 36). The schematic
P-T projections that may be drawn from points A-C are shown
in detail in Figure 7.
From Figures 6 and 7 we can observe that, as a consequence
of the tricritical transition (between points A and B in Figure
6), Type III
a
systems originate a discontinuity along the critical
line that emerges from pure water [1]. In this case and as shown
in Figure 7b, the 3PL presents two stable branches. The first
branch appears in the low-temperature range (-UCEP
1
), while
the additional one (LCEP
2
- UCEP
2
) meets the high-temperature
range. As shown in Figure 7c, at the right-hand side of the DCEP
line (between points B and C in Figure 6), the upper and lower
branches of the 3PLs shown in Figure 7b connect, yielding thus
a single heteroazeotropic line. In addition, an exchange of critical
lines is observed, and the specific branch that diverges in
pressure emerges now from the hydrocarbon [2]. From these
results, we can conclude the following:
The CSRK model predicts Type III
a
behavior below the TC
line (see Figure 7a) and III
b
above the DCEP line (see Figure
7c), in good qualitative agreement with the experimental data
(see Figures 1 and 2).
The CSRK model predicts a nondirect the transition between
Types III
a
and III
b
. Between these types of behavior, the hybrid
system shown in Figure 7b occurs.
In direct correspondence with the experimental behavior, a
maximum of three phases is predicted for water (1) + n-alkane
(2) mixtures. Consequently, the most probable way to generate
the TCEP that Brunner suggests is by intersecting the DCEP
and TC mechanisms in a common critical point. In such an
intersection, formally called the van Laar point, the three phases
of the 3PL become critical (as required by the TC mechanism),
Figure 5. Dependence of critical pressure and temperature of pure
hydrocarbons on the carbon number n. Critical properties for water and
n-alkanes have been taken from DIPPR.
25
Table 1. Parameters for Calculating Critical Properties from
Correlation in Eq 6
i ci
(T)
ci
(P)
1 -440.3571540 -11.9446194
2 1169.1586170 -0.9582342
3 -848.0502740 50.5748318
4 409.6516068 0.7079303
5 -117.2186560 -4.1484593
6 19.3538313
7 -1.7080074
8 0.0624194
T
c,2
/K )

i)1
8
c
i
(T)
n
(i-1)/2
P
c,2
/bar )

i)1
3
c
2i-1
(P)
n
(i-1)/2
1 +

i)2
3
c
2i-2
(P)
n
(i-1)/2
(6)
) (n); ) (n); ) (n,k
12
)
Figure 6. spd-GPD projection for water (1) + n-alkane (2) mixtures, as
predicted from the CSRK-EOS. (-) TC line, (- -) DCEP line.
950 Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007
and, in addition, the 3PL becomes tangent to the critical line
(as required by the DCEP mechanism). However, according to
Figure 6, a van Laar point is unlikely for reasonable n-k
12
values. Certainly, no TCEP mechanism is observed in the
transition presented in Figure 7.
Previous predictions of the CSRK model suggest that
Brunners explanation concerning the existence of a TCEP does
not constitute an unequivocal approach of the phase equilibrium
behavior of systems under consideration. Let us consider now
the barotropic behavior as predicted by the CSRK model for
water (1) + n-alkane (2) mixtures. As can be seen in Figure 8
and in excellent agreement with experimental results, we
conclude the following:
Mass density inversion is predicted by the CSRK model for
Type III
b
systems, while ordinary behavior is observed for
Type III
a
systems. As a consequence of the barotropic inversion,
the aqueous phase becomes lighter than the organic one (see
Figure 8b).
The critical point of the UCEP in Figure 8a occurs between
the gas and the organic phase for Type III
a
systems, while Figure
8b shows that the gas and aqueous phases become critical in
the case of Type III
b
systems.
Inspection of eq 6 reveals that mixtures with n ) 26 classify
as Type III
a
behavior, while mixtures with n ) 28 classify as
Type III
b
, if we consider CSRK mixtures with the following
linear dependency for the interaction parameter:
Figure 9 depicts the critical end points (CEPs) and barotropic
temperatures predicted by CSRK for different n values. Interac-
tion parameters have been calculated according to eq 7. Figure
9 should be directly compared with Figure 3, where the
experimental data for water (1) + n-alkane (2) are presented.
In Figure 9, where the shaded area corresponds to a
temperature range of heteroazeotropy, we can observe that Type
Figure 7. P-T projections for the binary systems shown in Figure 6. (-) Critical line, (- -) heteroazeotropic line, ( ) vapor pressure line. (a) System
in point A (Type IIIa), (b) system in point B, (c) system in point C (type IIIb).
k
12
) -0.12 + 0.01n (7)
Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007 951
III
a
and III
b
systems exhibit single CEPs, which, in turn,
correspond to the UCEPs observed in Figure 7a and c. Between
the TC and DCEP boundaries observed in Figure 9, multiple
CEPs can be predicted in direct relation with the two branches
of the 3PL shown in Figure 7b. Finally, from Figure 9 it is
clear that the temperature of barotropy reaches the critical range
in point A for 26 <n <28. Afterward, barotropic phase reversal
is observed from point A to subcritical states along the curve
A-B (for n > 27). It should be noted that no special
coordination occurs between the TC or DCEP mechanisms and
the critical temperature of barotropy. In fact, point A, where
barotropy begins, persisting for larger n values, occurs at the
ordinary LCEP
2
that we can see in Figure 10.
Figure 11 shows the mass density of each one of the phases
that appear along the two 3PLs in Figure 10. Below the
temperature of UCEP
1
, where the organic liquid and the gas-
phase collapse in a critical point, the subcritical aqueous-rich
phase is the heaviest one. Above the temperature of LCEP
2
,
where the organic liquid and the gas phase become critical once
again, the non-critical aqueous-rich phase becomes lighter than
the organic liquid. At the temperature of LCEP
2
, it is seen that,
although these three phases have the same mass density, only
two of them (the organic-rich liquid phase and the gas phase)
are characterized by a critical condition. Consequently, and
contrary to the hypothesis of a TCEP suggested by Brunner,
the CSRK model predicts that the point where barotropy begins
is not a tricritical point.
In Figure 11, it is possible to observe an additional barotropic
point B between the aqueous and the gas phase. This behavior
is predicted by the CSRK model for the range 27 < n < 28
and the interaction parameter in eq 7. Because we do not have
experimental evidence of such a barotropic behavior for the
Figure 8. Mass density along the heteroazetropic line, as predicted by the
CSRK model. (-) Aqueous phase, (- -) organic phase, ( ) gas phase,
(O) upper critical end point. (a) Type IIIa system, (b) Type IIIb system.
Figure 9. Critical end points and barotropic temperatures for different
n-values. k12 as given in eq 7. (-) CEPS, (- -) temperature of barotropy.
Figure 10. P-T projection for the binary system indicated in point A,
Figure 9. (-) Critical line, (- -) heteroazeotropic line, ( ) vapor
pressure line.
952 Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007
mixtures under consideration, it will not be discussed further
in this work.
Concluding Remarks
The approach presented in this work constitutes a theoretical
way for demonstrating that the topology observed in the phase
equilibrium of water (1) + n-alkane (2) mixtures depends
directly on the carbon number n of the aliphatic chain. In fact,
from the spd-GPD of the CSRK model, and using n as the
fundamental parameter, we have demonstrated the following:
The EOS model predicts qualitatively the high-pressure
behavior of the mixtures under consideration.
The transition from Type III
a
to Type III
b
can be completely
explained in terms of the sequential action of the TC and DCEP
mechanisms.
The sequence of the TC and DCEP mechanisms predicts
consistently the barotropic transition observed in the experi-
mental data.
Certainly, the transitional mechanisms that we present here
contrast with the hypothesis of a tricritical end point suggested
by Brunner. On the one hand, and in complete agreement with
Brunners hypothesis, at the point where barotropy begins three
phases should have the same mass density. However, three
phases with equivalent mass densities meet only a sufficient
but not a necessary condition of a tricritical end point.
Consequently, the origin of barotropy in water + n-alkane
mixtures could also be explained in terms of a more realistic
scenario, considering a mechanism which does not violate
Gibbss phase rule.
Acknowledgment
This research work has been financed by FONDECYT, Chile,
project No. 1050157. We would like to thank Dr. A. Mej a for
several helpful discussions.
Nomenclature
3PL ) three-phase line or heteroazeotropic line
CAzEP ) critical azeotropic end point
CEP ) critical end point
CLP ) closed loop of immiscibility
CPSP ) critical pressure step point
CSRK ) Carnahan-Starling-Redlich-Kwong
DCEP ) double critical end point
EOS ) equation of state
GPD ) global phase diagram
LCEP ) lower critical end point
UCEP ) upper critical end point
spd ) serial prediction domain
TC ) tricritical point
TCEP ) tricritical end point
vLP ) van Laar point
ZT ) zero temperature point
Symbols
a ) cohesion parameter
b ) covolume
c
i
(T)
, c
i
(P)
) correlation parameters, defined in eq 6 and Table 1
k
12
) interaction parameter
n ) carbon number of the aliphatic chain
P ) pressure
R ) universal gas constant
T ) temperature
V ) volume
x ) mole fraction
Greek Letters
, , ) global phase diagram coordinates, defined in eq 5
) volume packing factor [)b/(4V)]
Subscripts
c ) critical property
i, j ) component index
r ) reduced property
Literature Cited
(1) Montel, F. Phase equilibria needs for petroleum exploration and
production industry. Fluid Phase Equilib. 1993, 84, 343-367.
(2) Dutkiewicz, A.; Ridley, J.; Buick, R. Oil-bearing CO2-CH4-H2O
fluid inclusions: oil survival since the palaeoproterozoic after high
temerature entrapment. Chem. Geol. 2003, 194, 51-79.
(3) de Hemptinne, J. C.; Peumery, R.; Ruffier-Meray, V.; Moracchini,
G.; Naiglin, J.; Carpentier, B.; Oudin, J. L.; Connan, J. Compositional
changes resulting from the water-washing of a petroleum fluid. J. Pet. Sci.
Eng. 2001, 29, 39-51.
(4) Pedersen, K. S.; Milter, J.; Rasmussen, C. P. Mutual solubility of
water and a reservoir fluid at high temperatures and pressures. Experimental
and simulated data. Fluid Phase Equilib. 2001, 189, 85-97.
(5) Brunner, E. Fluid mixtures at high pressures. IX Phase separation
and critical phenomena in 23 (n-alkane+water) mixtures. J. Chem.
Thermodyn. 1990, 22, 335-353.
(6) Tsonopoulos, C. Thermodynamic analysis of the mutual solubilities
of normal alkanes and water. Fluid Phase Equilib. 1999, 156, 21-33.
(7) Tsonopoulos, C. Thermodynamic analysis of the mutual solubilities
of normal alkanes and water. Fluid Phase Equilib. 2001, 186, 185-206.
(8) van Konynenburg, P.; Scott, R. L. Critical lines and phase equilibria
in binary van der Waals mixtures. Philos. Trans. R. Soc. London 1980,
298A, 495-539.
(9) Van Ness, H. C.; Abbott, M. M. Classical Thermodynamics of
Nonelectrolyte Solutions; McGraw-Hill Book Co.: New York, 1982; p 390.
(10) Levelt Sengers, J. How Fluids Unmix; Royal Netherlands Academy
of Arts and Sciences: Amsterdam, 2002.
(11) Schouten, J. A.; van den Bergh, L. C. Density inversion between
fluid and solid phases in the system He-H
2 at high pressures. Fluid Phase
Equilib. 1986, 32, 1-7.
(12) van der Steen, J.; de Loos, Th. W. The volumetric analysis and
prediction of liquid-liquid-vapor equilibria in certain carbon dioxide +
n-alkane systems. Fluid Phase Equilib. 1989, 51, 353-367.
Figure 11. Mass density along the heteroazetropic lines observed in Figure
10. (-) Aqueous phase, (- -) organic phase, ( ) gas phase, (O) critical
end point.
Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007 953
(13) Rowlinson, J. S.; Swinton, F. L. Liquids and Liquid Mixtures, 3rd
ed.; Butterworths: London, 1982.
(14) Van Pelt, A.; De Loos, Th. W. Connectivity of critical lines around
the van Laar point in T, x projections. J. Chem. Phys. 1992, 97, 1271-
1281.
(15) Furman, D.; Griffiths, R. B. Global phase diagram for a Van der
Waals model of a binary mixture. Phys. ReV. A 1978, 17, 1139-1148.
(16) Deiters, U. K.; Boshkov, L. Z.; Yelash, L. V.; Mazur, V. A. On a
new mechanism of four-phase equilibria generation in two-component fluids.
Dokl. Akad. Nauk 1998, 359, 343-347.
(17) Quinones-Cisneros, S. E. Barotropic phenomena in complex phase
behavior. Phys. Chem. Chem. Phys. 2004, 6, 2307-2313.
(18) Carnahan, N. F.; Starling, K. E. Intermolecular repulsions and the
equation of state for fluids. AIChE J. 1972, 18, 1184-1189.
(19) Wang, M.; Wong, D. Calculations of critical lines of hydrocarbon/
water systems by extrapolating mixing rules fitted to subcritical equilibrium
data. Fluid Phase Equilib. 2005, 227, 183-196.
(20) Kraska, T.; Deiters, U. K. Systematic investigation of phase behavior
in binary fluid mixtures. II. Calculations based on the Carnahan-Starling-
Redlich-Kwong equation of state. J. Chem. Phys. 1992, 96, 539-547.
(21) Kraska, T. Systematic investigation of the global phase behavior
of associating binary fluid mixtures. I. mixtures containing one self-
associating substance. Ber. Bunsen-Ges. Phys. Chem. 1996, 100, 1318-
1327.
(22) Boshkov, L. Z. Bifurcations - A possibility to generalize the
thermodynamic description of phase diagrams of two component fluid. Ber.
Bunsen-Ges. Phys. Chem. 1992, 96, 940-943.
(23) Deiters, U. K.; Pegg, I. L. Systematic investigation of phase behavior
in binary fluid mixtures. I. Calculations based on the Redlich-Kwong
equation of state. J. Chem. Phys. 1989, 90, 6632-6641.
(24) Yelash, L. V.; Kraska, T. Closed-loops of liquid-liquid immiscibility
in binary mixtures of equal sized molecules predicted with a simple
theoretical equation of state. Ber. Bunsen-Ges. Phys. Chem. 1998, 102, 213-
223.
(25) Daubert, T. E.; Danner, R. P. Physical and Thermodynamic
Properties of Pure Chemicals. Data Compilation; Taylor & Francis: Bristol,
PA, 1989.
ReceiVed for reView August 7, 2006
ReVised manuscript receiVed October 19, 2006
Accepted October 30, 2006
IE061036O
954 Ind. Eng. Chem. Res., Vol. 46, No. 3, 2007

Das könnte Ihnen auch gefallen