Sie sind auf Seite 1von 31

Thermodynamics: Phase Equilibria

Jean Vidal

Introduction
It should be mentioned first of all that this chapter is in no way a treatise of
thermodynamics. We will therefore not address the first and second principles
or the classic equations that are derived from them. We will very often use as
a basis the widely accepted methodology developed by Gibbs (concept of
chemical potential) and by Lewis (fugacity, ideal solutions). We will restrict
ourselves to presenting the methods and models that are based on them and
are utilized by engineers. Additionally, certain passages, figures and tables in
this text are excerpted from a more detailed reference work (Vidal, 1997), a
more far-reaching discussion which can be consulted to supplement the
points presented here.
In order to compute the properties of material systems, the engineer gath-
ers together in a calculation “method” representations or “models” character-
istic of the various phases (liquid, vapor, etc.) that compose these systems
and of the different properties (density, heat capacity, equilibrium conditions)
of these phases. He takes into account the greater or lesser degree of com-
plexity of the mixtures he is dealing with, the conditions in which the mixtures
are found, the available data and the accuracy and time limit requirements he
must meet. Though rigorous with respect to the application of the principles
of thermodynamics, he will accept a certain degree of empiricism in working
out models when he has to represent the properties of mixtures that are
“difficult”:whether because of the nature or multiplicity of their components,
or the conditions (temperature, pressure, proximity to the critical point, etc.)
in which they are found. Accordingly, these “models” will seem perfectible to
him and he will be attentive to new proposals, thereby recognizing that
thermodynamics is a living science.
Whatever the property to be calculated, the first thing to find out is the
state of the system under study, i.e. its possible distribution into several
16 Chapter 2. THERMODYNAMICS:
PHASE EOUlLlBRlA

homogeneous parts or phases. Volume is a good example as its value is par-


ticularly sensitive to a change in state such as vaporization. This obligatory
prior step is, however, difficult. In contrast to pure substances, a mixture may
be distributed into several phases of different composition, according to the
temperature and pressure conditions. Calculating the composition of these
phases and their proportions will therefore go hand in hand. This will take on
particular importance here since the differences in composition are at the
basis of most separation processes. Calculations are based on the assumption
of thermodynamic equilibrium.
We will deal more specifically with the vapor-liquid equilibria that are
involved in distillation, absorption and stripping. This type of equilibria has
been studied and the understanding of fluid phases has become thorough
even though it could not be termed perfect. Following a brief description of the
vaporization and condensation phenomena of mixtures, equilibrium condi-
tions will be premised, allowing a simple classification of the problems that
may arise. The expression of the equilibrium condition will then lead to the
presentation of the different models applied to phases in equilibrium. In this
instance, it will be noted that these models have evolved towards unification
to the benefit of those derived from equations of state.
These concepts can be easily transposed to liquid-liquid equilibria.
Crystallization equilibria are often characterized by the existence of a pure
substance in the solid phase, and can not be dealt with in the same way as
equilibria between fluid phases. In spite of the severity of problems such as
paraffin formation or certain purification processes, these equilibria have not
been studied with the same degree of refinement as the equilibria between
fluid phases. We will, however, endeavor to present their main features.

2.1 Vapor0Liquid Equilibria


What we know about vapor-liquid equilibria is mainly based on experimental
data. There are of course predictive methods, but they are based on the
results of measurements for the determination of the parameters they involve
and the control of their validity. Accordingly, this aspect must be emphasized.
However, we will not develop here the experimental techniques that must
often be adapted to the case under study taking into account the nature of the
mixture, and the domains of pressure and temperature. The reviews by Abbott
(1986), Marsh (1989), Deiters et al. (1986) and Holste et al. (1986) can be con-
sulted in this connection.
Likewise, the bibliographical or numerical data bases that cover the results
of these measurements are also important. The largest was initiated by the
University of Dortmund (Gmehling et al. 197&1984), but mention should also
be made of the one by Kehiaian (1973-1993), which covers all deviations from
ideality and provides carefully controlled quality data. Lastly, the biblio-
graphic base of Wichterle et a]. (1973-1985, 1993) is virtually exhaustive.
Chapter 2 THERMODYNAMICS: PHASE EQUlLlBRlA 17

A good understanding of the vaporization and condensation phenomena of


mixtures is required to use them.
It should be pointed out that the way in which vapor-liquid equilibrium
data are acquired and therefore presented has evolved. The oldest data are
generally isobaric, they reflect better the operation of a distillation column,
where normally small variations in pressure correspond to pressure drop
between the trays, while the variations in temperature may be significant. This
type of data could be directly utilized by graphic methods of calculating
columns, such as the McCabe and Thiele method for example.
More recently, the trend has been to acquire isothermal data that are eas-
ier to model. The influence of pressure on liquid phase deviations from ideal-
ity is in fact generally negligible, contrary to the influence of temperature.

2.1.1 General Description


In order to describe a mixture, it is mandatory to know the pressure, the tem-
perature and the composition. In order to give a graphic illustration of changes
in phase we will first consider binary mixtures changing at constant pressure,
then at constant temperature. Figures 2.1 and 2.3 and Table 2.1 concern the
mixture of propane (1) and n-pentane (2). They show the boundaries of the
two-phase domain versus the composition.

2.1.1.1 Isobaric Vapor*LIquidEquilibrium Diagrams


Figure 2.1 shows the compositions on the abscissa, i.e. the mole fraction of one
of the components, z1 in the overall mixture, x1 in the liquid phase, and y1 in
the vapor phase. Since the mixture is binary the mole fractions of the second
component can be deduced by complement to the unit value. As a rule, the
index 1 will be assigned to the more volatile component here.
The temperature will be placed on the ordinate. In particular, on the
abscissa axes 1 and 0 we will place the boiling point of the first and the second
component at the pressure under consideration (5 bar), i.e. 1.5"C for propane
and 92°C for n-pentane.
Let us consider a mixture of an overall composition z1 (for example
z1= 0.5). At low temperature the mixture is homogeneous in the liquid state. If
heat is supplied to the system under constant pressure, it will first cause the
liquid to heat up and the "bubble point temperature" @bubblewill be reached
(i.e. 26.4"C), where the first bubble of vapor appears. At this point the liquid
phase has of course the same composition as the overall mixture, Xi,bubble = zi,
but the vapor phase that appears has a composition Yi,bubble which is different.
It is generally richer in the volatile component. For the example presented
here, y1 = 0.915. The system is said to be at the bubble point.
If the supply of heat is maintained, the temperature is usually seen to rise
(this is the case for the mixture involved here) and at the same time the vapor
phase develops at the expense of the liquid phase. The system is distributed
into two phases whose composition xi and yi differs from that of the overall
18 Chapter 2. THERMODYNAMICS: PHASE ~QUlLfBRlA

0.0 0.25 0.5 0.75 1.o


Mole fraction of propane
Figure
Isobaric vapor-liquid equilibrium of the propane/n-pentane system. P = 5 bar.
- bubble point curve ------
dew point curve.

mixture zi. The composition of these phases is related by the equilibrium con-
ditions that we will discuss later. In addition, the quantities (number of moles)
in the liquid phase, N L , vapor phase NV and for each component in the overall
mixture, N i , are related by the following equations:
Ni = NLxi + NVyi with i = 1 to 2 in the present case (2.1)
which can also be written:
Ni = ( N L + N v ) Z;

For example at 4 0 T , there is x1= 0.34, y 1 = 0.83 and the vaporized fraction
is equal to:
0.50 - 0.34
- -NV
-- - 21 -XI -
= 0.33
N L + NV y1 - X I 0.83 - 0.34
This last equation can be represented graphically by applying the "lever
rule".
Vaporization of the mixture generally continues until the "dew point". The
mixture still verifies vapor-liquid equilibrium conditions, but the liquid phase
disappears. The coordinates of this point are the dew point temperature, the
Chapter 2. THERMODYNAMICS:
PHASE ~OUlLlERlA 19

composition Xi,dew of the liquid phase, and Yi,dew of the vapor phase with
yi,dew= zi.In the case of an equimolar mixture of propane and n-pentane at a
pressure of 5 bar, the dew point temperature is 67.5"Cand x1= 0.13 (see Fig. 2.1
and Table 2.1).

21 Obubble ec> 'bubble mar) 'dew @@


0.000 92.018 92.018 0.717 0.717
0.050 81.909 89.987 1.125 0.752
0.100 72.650 87.888 1.537 0.791
0.150 64.302 85.713 1.954 0.835
0.200 56.849 83.456 2.375 0.884
0.250 50.223 81.106 2.80 1 0.938
0.300 44.333 78.652 3.232 1.ooo
0.350 39.086 76.081 3.668 1.071
0.400 34.392 73.376 4.109 1.152
0.450 30.174 70.518 4.556 1.246
0.500 26.363 67.482 5.009 1.357
0.550 22.902 64.237 5.467 1.490
0.600 19.743 60.743 5.933 1.651
0.650 16.844 56.946 6.405 1.851
0.700 14.173 52.770 6.885 2.106
0.750 11.699 48.110 7.373 2.440
0.800 9.399 42.801 7.870 2.898
0.850 7.251 36.570 8.378 3.563
0.900 5.236 28.908 8.899 4.605
0.950 3.338 18.644 9.434 6.425
1.ooo 1.543 1.543 9.987 9.987
ITableI-
Vapor-liquidequilibrium of the mixture of propane ( 1 ) and n-pentane (Z), iso-
baric (P = 5 bar) and isothermal (8 = 26.85"C,i.e. 300 K)* z1 represents the
mole fraction of component I in the mixture.

If heat continues to be supplied to the system it then becomes homoge-


neous and is said to be "superheated" vapor.
The description of the vaporization process could be repeated for all of the
compositions of this binary mixture: the bubble points would describe the
"bubble point curve" and in the same way the "dew point curve" would be
determined. These two curves that delimit the two-phase domain come
together on the axes representing pure substances, since bubble point and
dew point temperatures coincide (monovariant system). Below the bubble
point curve is the domain representing the homogeneous liquid phase, usually
termed "subcooled", above the dew point curve is the zone of the "super-
heated vapor phase. Inside the system is in a partially vaporized state, it is
20 Chapter 2. THERMODYNAMICS: PHASE ~ O U I L S R I A

neither at its bubble point nor at its dew point and is distributed into two
phases. The liquid phase, in contrast, is at its bubble point and the vapor
phase at its dew point.
Ternary systems
It is possible to represent the composition of a ternary system by means of a
triangular diagram. The apexes represent the components of the system, the
three sides correspond to the binaries formed by two of the three components
and the inside of the triangle represents the ternary mixtures per se. In
Figure 2.2, the temperature has been plotted perpendicular to the plane of the
triangle, thereby delimiting a prism. On the faces of the prism at constant pres-
sure, the three equilibrium two phase regions of the binaries are inscribed.
Inside the prism the changing bubble point or dew point temperatures can be
plotted versus the composition, thereby defining the corresponding surfaces.

2'
Projection
plane 8 = eo
B

I Figure
2.2 Vapor-liquid equilibrium diagram For a ternary system at constant pressure.
Chapter 2. THERMODYNAMICS:
PHASE EOUlLlERlA 21

The intersection of the bubble point and dew point surfaces with an
isothermal plane determines two curves. Along the curves the representative
points of the mixtures at equilibrium correspond to each other in pairs.
A point situated between the two surfaces represents a two-phase system,
the representative points of the two phases in equilibrium are situated on the
bubble point and dew point surfaces and on the two curves corresponding to
the temperature under consideration. They are lined up with the point repre-
senting the overall mixture due to the linear equations stemming from mate-
rial balances.

2.1.1.2 Isothermal Vapor0Liquid Equilibrium Diagrams


The presentation of the isothermal vapor-liquid equilibrium diagram for a
binary mixture is similar to the one we have just discussed for an isobaric dia-
gram. Figure 2.3 concerns the mixture of propane and n-pentane at 2635°C.
Pressures are on the ordinates and on the axes representing pure substances,
the vapor pressures of the two system components are plotted, i.e. 10 bar for
propane and 0.7 bar for n-pentane. At low density, the mixture with a compo-
sition z1= 0.5 is in the vapor state and the pressure is low. A decrease in vol-
ume will cause in turn an increase in pressure, the appearance of the first drop

1
10 2 z

I
I
I
I
I
I
I
I
I

0.0 0.25 0.5 0.75

Figure
2.3 Isothermal vapor-liquid equilibrium diagram for the propaneh-pentane
- system. 8 = 26.85"C.
22 Chapter 2. THERMODYNAMICS: PHASE EOUlLlBRlA

of liquid at P = 1.36 bar (dew point pressure), then partial condensation of the
mixture which is distributed into two phases: liquid and vapor, then the dis-
appearance of the last bubble of vapor at P= 5 bar (bubble point pressure) and
finally the compression of a homogeneous liquid phase. On the isothermal dia-
gram the equilibrium two phase region can be seen, but the position of the liq-
uid and vapor domains, of the bubble point and dew point curves are reversed
in comparison with what is seen on an isobaric diagram.

2.1.1.3 mo-phase Envelope


The changes in this type of isothermal diagram versus temperature can be
examined, and this is what has been done in Figures 2.4a and 2.4b. They con-
cern the ethane/benzene system. On this diagram we will follow the changes
in state of a mixture with an overall composition z1= 0.7 for example.
At low temperature, 283 and 298 K (Fig. 2.4a), the equilibrium two phase
regions look the same as described previously. It should be noted, however,
that the dew point curve is mostly very close to one of the axes P = 0 or y , = 1.
This distorsion is due to the great difference in volatility of the two compo-
nents. For the mixture under consideration, the bubble point and dew point
pressures (respectively P = 23 and 31 bar) increase with temperature.

50

40

2
e
!? 30
3
m
!?
a
20

10 I
I
n

0.0 0.25 0.5 0.75 1.0


Mole fraction of ethane
Figure
Figure 2.4a Isothennal vapor-liquid equilibrium diagrams for the ethane/ben-
zene system at 283 and 298 K.
bubble point curves -- - - - dew point curves.
Chapter 2 THERMODYNAMICS: PHASE EQUILIBRIA 23

If the temperature is higher than the critical temperature of one of the com-
ponents (Fig. 2.4b), for example 363 K, since the critical temperature of ethane
is 305 K, then the two phase region extends only over part of the composition
domain. The mixture z1= 0.7 still has a dew point (P = 4.7 bar) and a bubble
point (P = 76.9 bar) at this temperature. The bubble point and dew point
curves no longer come together on the axis zl = 1, but rather at a point char-
acterized by the maximum value of the pressure. At this point the liquid and
vapor phases in equilibrium are identical in composition and properties (den-
sity, enthalpy, entropy, etc.). This is a critical vapor-liquidequilibrium point.
This critical point, which corresponds to an observable physical reality, must
not be confused with the pseudocritical point (Paradowski, Vol. 1, Chapter 4).
The pseudocritical point’s coordinates are used in the application of the equa-
tion of states corresponding to mixtures, are defined by a set of empirical rules
and are only calculation parameters.

150

125

-ez 100

2
3
v)

s!
a 75

50

0.0 0.25 0.5 0.75 1 .o


Mole fraction of ethane

Figure
Isothermal vapor-liquid equilibrium diagrams for the ethane/benzene system
at 333, 363, 428, 448, 483 and 508 K
bubble point curves - -
- - - dew point curves.
24 Chapter 2. THERMODYNAMICS: PHASE EQUILIBRIA

At 428 K, the critical point corresponds to the mixture z1= 0.7; the changes
in the physical state of this mixture versus pressure causes a dew point
(P = 26.4 bar) and then a two-phase mixture to appear from a homogeneous
vapor phase. N o bubble point will be seen and when the critical pressure
(P= 108 bar) is reached, the vapor-liquid interface will disappear and the phe-
nomenon termed "critical opalescence" will be seen. At higher pressures the
mixture will once again be homogeneous.
At 448 K, the ethane composition of the critical point is lower than that of
the mixture under consideration. For the latter, the increase in pressure
always means a first dew point (P = 44.5 bar), a separation into two phases.
However, the proportion of liquid phase goes through a maximum and it is by
a second dew point that the two phase region is left behind and the homoge-
neous zone is entered at high pressure (P= 106.7bar). The mixture has no bub-
ble point.
Finally at 483 and 508 K for the mixture z1= 0.7, there is no longer any dew
point or bubble point and the mixture is homogeneous whatever the pressure.
The changing bubble point and dew point pressures can be represented
versus temperature for a mixture of fixed composition. This is what is done in
Figure 2.5 for the example under discussion here. The convergence of the bub-
ble and dew point curves can be seen at the critical point (with at this point
the same slope dP/dT). There is an extremum of temperature (criconden-
therm) and pressure (cricondenbar, very close to the critical point in the pre-
sent example). Between the temperature corresponding to the critical point
and the cricondentherm, for example at 175"C, the mixture under considera-
tion has two dew point pressures. As the mixture changes by decreasing pres-
sure levels from the domain of high pressures where it is homogeneous, there
will first be a "high" dew point, then a liquid phase will appear (this phe-
nomenon is called retrograde condensation), the quantity of liquid deposited
will increase up to a maximum, a "low" dew point, and then finally a homoge-
neous vapor phase. Bubble point and dew point curves delimit the two-phase
domain, inside which this particular mixture is broken down into two phases,
and outside which it is homogeneous.
On the same diagram we have plotted the two-phase domain of another
mixture (zl = 0.2), whose cricondenbar (T = 533.9 K and P = 68.2 bar) is more
clearly distinct from the critical point (T= 540.6 K and P= 67.8 bar). In the pres-
sure interval defined as 67.8 c P (bar) < 68.2, a temperature variation causes
two bubble points to appear.
The set of critical points makes up the "locus of critical points",the enve-
lope of two-phase domains, which extends from the critical point of the light
component to the critical point of the heavy component. Outside of the
domain delimited by the two curves of vapor pressure and the locus of critical
points, the system is homogeneous whatever its composition.
According to the composition of the mixture, the cricondenbar can be
located on the bubble point or the dew point curve. For pressures between the
pressure corresponding to the critical point and the pressure of the cricon-
denbar, two dew point or bubble point temperatures can be seen. This case
Chapter 2. THERMODYNAMICS: PHASE fQUlLlERlA 25

can be termed second kind retrograde condensation; this is what is shown on


Figure 2.5 for the mixture with composition z1= 0.2.

100

75
z
h

a
v

$
v)
50
??
a
25

Figure
2.5 Two-phase domain of the ethane/benzene system.
- -bubble point curves ----- dew point curves

The qualitative description that we have just given involves a binary mix-
ture. For a complex "multicomponent" mixture, the representation in the tem-
perature-pressure diagram remains valid and the twephase envelope can be
plotted for a mixture of specified composition. Retrograde condensation phe-
nomena will be seen. This is the case in particular of natural gases for which
the liquid deposit caused by a decrease in pressure occurs when a reservoir is
being produced (pressure decrease due to production of gas), in the well
(pressure loss as produced fluids are coming up to the surface) and in sepa-
ration units at the wellhead.
It is important to note that the locus of critical points can extend over a
very wide pressure domain. This is the case for example if the mixture con-
tains hydrocarbons with very different molecular weights. The pressure cor-
responding to the critical point of the mixture that we have considered was
108 bars, and for mixtures of methane with high molecular weight paraffins,
hexadecane for example, it can reach several hundred bars.
26 Chapter 2. THERMODYNAMICS: PHASE kJUILIBFtIA

Let us point out finally that the presentation we have made corresponds to
the simplest case. Depending on the nature of the components, and the occur-
rence of liquid-liquid equilibria, the two-phase envelope and the locus of criti-
cal points may have more complex configurations. A mixture may have no
critical point or have several of them (Rowlinson and Swinton, 1982, p. 191;
Kreglewski, 1984, p. 133).
Distillation processes are of course not located in the critical zone, since
the composition of the liquid and vapor phases converges as it approaches the
critical point. Additionally, the pressure is generally high at this time. Mention
should be made of supercritical extraction processes that used a light com-
pound as a selective solvent (very often carbon dioxide) in temperature and
pressure conditions close to the critical point of the solvent. The solvent
power is then significant and also varies rapidly with operating conditions and
selectivity follows suit.

2.1.1.4 Azeotropes

Up to now we have considered that it is normal in a binary system at constant


pressure for the bubble point and dew point temperatures, or at constant tem-
perature for the bubble point and dew point pressures, to be monotonic func-
tions of the composition. This is in no way the rule. Let us assume that we have
two components with the same vapor pressure at a given temperature: either
the bubble point curve will show an extremum in isothermal representation,
or it will be isobaric, i.e. perfectly flat. The bubble point pressure will be inde-
pendent of the composition and the system will behave like a pure substance
as regards vapor-liquid equilibria at least at this temperature. This is the first
hypothesis that must be admitted: the bubble point curve will show an
extremum. The special case of components with the same vapor pressure is
not at all unlikely, and in any case there are numerous examples of mixtures
whose components have very similar vapor pressures. Examples are the cyclo-
hexane/benzene system whose isothermal equilibrium diagram shows a maxi-
mum pressure, or the hexafluorobenzene/benzene system which exhibits both
a maximum and a minimum pressure.
The appearance of an extremum of the bubble point curve is not specific
to systems whose vapor pressures are very similar. The deviations from ideal-
ity that we will examine later can also cause the same phenomenon. This is the
case for example of heptane/dimethylformamide mixtures that are rich in h e p
tane. The difference in boiling point of these two compounds is however
approximately 50°C. Mention can also be made of the carbon dioxide/ethane
system whose components have vapor pressures that differ by 50% at 10°C.
To sum up, this phenomenon can be caused by two types of distinct p r o p
erties: either similar volatility for the components of a mixture or deviations
from ideality. It has been shown that in this case, if the bubble point curve has
an extremum, then the same holds true with the same composition for the dew
point curve. At the extremum the bubble point and dew point pressures have
the same value and liquid and vapor phases have the same composition. The
Chapter 2 PHASEEOUILBRIA
THERMODYNAMICS: 27

vapor-liquid equilibrium is therefore not selective, there is said to be an


“azeotrope”.
Figures 2.6A and 2.6B concerning the n-hexane (1) and acetone (2) system
illustrate this phenomenon. It can be seen that for the compositions in com-
ponent 1 (which in this case is the less volatile) that are lower than the
azeotropic composition, the vapor phase is richer in this compound than the
liquid phase, contrary to what was seen in the “normal” case. In contrast,
above the azeotropic composition, it is the opposite that occurs. Distillation of
this type of mixture which, in the case of the example presented here, exhibits
maximum bubble point and dew point pressures, will produce the azeotropic
mixture rather than the more volatile component at the top of the column. At
the bottom one or the other of the two components will be obtained depend-
ing on the composition of the original mixture. The azeotropic phenomenon
therefore constitutes a limit for separation processes by distillation.

1.1
A B
1 65

-,.
-3 0.9
F
?. 60
g!
L

u)
E
$ 0.8
n
55
e
0.7 50 -

0.6

0.5 ,
0.0
I
0.25
I
0:5
I
0.75 1 I
45
40
1I
0.0 0.25 0.5 0.75 1
Mole fraction of mhexane Mole fraction of mhexane

!
I
Figure
2.6 Azeotropic vapor-liquid equilibrium of the n-hexane/acetone system.
bubble point curves - - - --
dew point curves.
A. Isothermal diagram (0 = .5O0C>.B. Isobaric diagram (P = 1.013 bar).

A mixture with the composition of the azeotrope therefore vaporizes at a


given temperature in the same way as a pure substance: the pressure remains
constant during vaporization. However, if the same diagram is examined at
another temperature, it can be seen that the azeotropic composition has
changed. This change therefore differentiates the azeotropic mixture from
the pure substance and allows the separation limit mentioned above to be
overcome. The two components of a system exhibiting an azeotrope can be
28 Chapter 2. THERMODYNAMICS:
PHASE EOUILIBRIA

separated by using two distillations operating at different pressures (see


Chapter 4). Table 2.2 gives the data regarding the ethanol/water system for
pressures higher than atmospheric pressure.
The azeotrope may exist only in a limited pressure domain. In other cases,
the carbon dioxide/ethane mixture for example, it continues to be present
right up until the critical zone (see Fig. 2.7).

Pressure mar) Temperature @) x (ethanol)

1 35 1.50 0.894
3.44 385.75 0.882
6.89 408.85 0.874
13.78 437.35 0.862
20.68 455.75 0.852

Changing composition of the ethanol/water azeotrope versus


pressure (Otsuki and Williams, 1953).

Figure
2.7 Isothermal vapor-liquid equilibrium curves for the ethaneC0, system. The
curves are computed by means of the SRK method, with kij = 0.13. The points
Chapter 2. THERMOOYNAM~CS:
PHASE fQUiLi6R/A 29

The extremum of bubble point and dew point pressures characterizing an


azeotrope is usually a maximum and corresponds to "positive" deviations
from ideality from the standpoint of excess Gibbs energy and activity coeffi-
cients (see Section 2.3.2). This is what can be observed for mixtures of hydro-
carbons or for mixtures of hydrocarbons with polar solvents. "Negative"
deviations from ideality can cause azeotropes to occur at a pressure minimum,
which are often attributed to liquid phase complex formation (by hydrogen
bonding or hydration).
The correspondence between isothermal and isobaric azeotropic diagrams
is simple for a given system: a pressure maximum observed on the isothermal
equilibrium diagram corresponds to a temperature minimum on the isobaric
diagram for the azeotropic composition, as shown in Figures 2.6a and 2.6b.
The n-hexane (l)/acetone (2) azeotrope has the same coordinates:
x1= y , = 0.35, P = 1.013 bar, 8 = 50°C.

2.1.2 Equilibrium Condition and Generalized Variance


2.1.2.1 General Expression of the Equilibrium Condition;
Chemical Potential and Fugacity
It is on the basis of the second principle of thermodynamics that the equilib-
rium condition between two phases can be stated: for any transformation
occurring at constant pressure and temperature, Gibbs energy will decrease,
in a stable equilibrium state Gibbs energy is minimal. This condition is gen-
erally expressed by the following equation:
dC, = 0 (2.2)
which is in fact an extremum condition.
Given the constraint on temperature and pressure, the quantities alone of
each component in the liquid phase, Nk, or in the vapor phase, NY, can vary by
transfer of mass from one phase to the other. It is therefore important to char-
acterize the influence of these variables on the free enthalpy of each phase.
The chemical potential of a component in a homogeneous phase is defined
as the partial derivative of the Gibbs energy of the phase in relation to the
quantity of matter (number of moles) of the component under consideration,
all the other variables being kept constant. Accordingly, for the liquid and
vapor phases:

and for the heterogeneous system at equilibrium:


dGTp = dCip + dC&. = C@\ dN\ + pydNy) = 0
Additionally, with the material balance imposing the condition:
dNk + dNY=O
the equation below can be deduced:
p; = f l y
30 ChaDter 2. THERMODYNAMICS: PHASE EQUILIBRIA

As a result, at equilibrium, the chemical potential of any component has


the Same value in the liquid and vapor phase. This can be extended to the
case of a system distributed into any number of phases (for example liquid-liq-
uid-vapor) and the chemical potential of any component can be said to be the
same in all the phases comprising the system. It can therefore be considered
that this factor is the potential governing exchanges of matter, in the same way
as temperature determines thermal exchanges and pressure determines
mechanical equilibrium.
The problem of calculating phase equilibria is thus reduced to evaluating
chemical potentials. These potentials are, however, rather inconvenient to
use. First of all the concepts are fairly abstract, additionally, chemical poten-
tials can be calculated only from an arbitrary origin in the same way as inter-
nal energy and enthalpy. Lastly, they approach --tc, when the pressure or the
mole fraction approach 0. As a result, by a very simple change in variable they
have been replaced by fugacity 4, which is the same as partial pressure at low
densities. It can be defined by the following equations:
(dj& = RT d In t;. (2.5)
- 4+ l ifw 0
(2.6)
PY;
Its most common use is that the equilibrium condition (2.4) can be
expressed by means of fugacities:
f; = fy (2.7)
As mentioned earlier, in the vapor phase at low pressure fugacities are the
same as pressure itself for a pure substance, to partial pressures for the com-
ponents in a mixture, and to vapor pressure (if it is low) for a pure substance
in the liquid phase. They are of course expressed in pressure units and
approach zero if the pressure or the mole fraction approaches zero.
In the same way as chemical potentials, fugacities are evaluated by means
of liquid and vapor phase models in relation to temperature, pressure and
composition conditions.
Equilibrium conditions are frequently expressed by the “equilibrium
constant“, Ki,whose value is of course evaluated from fugacities and depends
generally on all of the intensive variables that characterize the system:

Intensive and Extensive Variance


Before addressing the calculation of the K value, we would like to briefly
review the concept of “variance”. This is the number of data required to cal-
culate the properties of a system, or the number of degrees of freedom the sys-
tem has.
A simple balance among the intensive variables that characterize a system
(temperature, pressure, composition of each phase) on the one hand, and on
the other the equations of equality that chemical potentials must verify leads
Chapter 2. PHASE EOUILIBR~A 31
THERMODYNAMICS:

to Gibbs’s rule of phases. If cp stands for the number of phases, n the number
of components, then the variance V is equal to:
V=n+2-(p (2.8)
Another case can be considered: the system is described by the quantity of
each of the components (number of moles, Ni), and not only the intensive vari-
ables (temperature, pressure, composition of each phase) but also the quanti-
ties of matter present in the phases (NL,N v ) and the extensive properties of
the system (volume, enthalpy, etc.) are to be determined.
The properties are calculated in relation to the number of moles present in
the liquid and vapor phases, N L and NV and the mole properties of each phase.
For volume, for example:
V=VL + Vv= N L u L+ NVuV
where u L and u v stand for the mole volume in the liquid and in the vapor phase
respectively. These last properties can be calculated in relation to tempera-
ture, pressure and the composition of the phases by correlations, models and
equations of state for example.
Accordingly, in order to characterize the system and calculate its proper-
ties we must know or determine the pressure, temperature, quantity of each of
the phases present (NL,N v ) , their composition (xi,yi), i.e. a total of 2n + 4 vari-
ables. Based on these variables, the existing models can be used to verify that
equilibrium conditions are complied with (calculation of chemical potentials
or fugacities) and to evaluate the volume occupied, the thermodynamic prop-
erties (enthalpy, entropy, etc.).
We have material balances among these variables involving each compo-
nent:
Ni = NLxi + NVyi with i = 1,n (2.9
i.e. n equations, equilibrium conditions involving chemical potentials or fugac-
ities:
p; = p y (2.4)
or:
f; =fy (2.7)
i.e. a further n equations. Finally there are equations among the mole fractions:
C x i = l and C y i = 1 (2.9)
i.e. at total of 2n + 2 equations.
Two variables therefore have to be known in order to determine the prob-
lem, and they can be chosen from among those we have mentioned or from
among those that depend on them (total volume, enthalpy, etc.).
As a result, it can be said that the total variance of a system is equal to
two,and this can be extended to a multiphase equilibrium (Duhem’s rule of
phases). There is no contradiction here with Gibbs’s rule of phases, since we
have considered on the one hand that the total quantity Ni of each component
32 Chaoter 2, THERMODYNAMICS: PHASE fOUlL/BRlA

was known while on the other hand, by including the quantity of each phase
among the unknowns, we have extended variance to the extensive variables.
For example, we know that for water at atmospheric pressure, the vapor-liquid
equilibrium temperature is fixed (variance equal to one) and equal to 100°C.
The volume occupied, however, can not be specified; according to the vapor-
ization rate it can range from 20 cm3 to around 30 liters.
In the choice of the two data in a vapor-liquid equilibrium problem, inten-
sive variance must of course be taken into account. For a pure substance,
there can only be one intensive variable out of the two, and the same is true
for a binary mixture at three-phase liquid-liquid-vapor equilibrium, etc.

2.1.2.2 Classification of Vapor=LiquidEquilibrium Problems


According to the TVpe of Data Available
The fact that total variance (intensive + extensive) is reduced to two and inde-
pendent of the number of components will enable us to present the main
vapor-liquid equilibrium problems briefly according to the type of data avail-
able.
a. The Data Available to Calculate the Equilibrium Are
Temperature (or Pressure) and the Vaporized Fraction
Since the total number of moles is known, the quantities of each phase, NLand
N V , can be deduced directly from the vaporized fraction. The other unknowns
are pressure (or temperature) and the composition of each phase.
Two important special cases must be mentioned. If the vaporized fraction
approaches zero or one, then the composition of the phase that is disappear-
ing approaches a definite limit that should be determined by equilibrium equa-
tions, while the composition of the remaining phase becomes identical to that
of the overall mixture. These borderline cases are precisely the bubble and
dew points of the mixture.
Finally, let us point out that the problem may not have a solution when the
mixture contains one or more supercritical components, or it may have sev-
eral solutions in the case of retrograde condensation.
b. The Data Available to Calculate the Equilibrium Are
Temperature and Pressure
The unknowns here are the quantities of each phase, NL and NV, and their
composition. As discussed earlier, we use material balances and equilibrium
equations as a basis for solving the problem. It should be noted of course that
the problem has a "physical" solution only if the data are located inside the
two-phase envelope. The problem is undetermined if the data coincide exactly
with the coordinates of an azeotrope. Finally it is necessary to verify that the
intensive variance is at least equal to two.
c. The Data Available to Calculate the Equilibrium Are Temperature
(or Pressure) and an Extensive Property of the System
Problems whose data are temperature or pressure and in addition a property
of the system can be envisaged quite well in practice.
Chapter 2. THERMODYNAMICS:
PHASE EQUILIBRIA 33

A case in point is the calculation of the pressure prevailing in a bottle of


known volume, containing a known quantity of a liquefied gas of known com-
position at a specified temperature.
Mention should also be made of expansion with partial liquefaction of the
mixture. Depending on the case, expansion is considered as isenthalpic or
isentropic, the problem data being the pressure at the end of expansion and
the enthalpy or entropy value.
Lastly and most important, it should be noted that in a distillation column
the phase counterflow is isenthalpic since it is assumed that there is no heat
transfer by conduction either from stage to stage or from the outside to the
separation unit. Calculation of a column will therefore combine around each
stage:
material balances for each component,
equilibrium equations,
the enthalpy balance,
with pressure given elsewhere.
As we have mentioned earlier, among the equations that are used to solve
vapor-liquid equilibrium problems there are first and foremost the equilibrium
conditions, i.e. equalities of chemical potentials.
We will now address the different methods used to calculate these chemi-
cal potentials, or in practice, fugacities and the Kvalue. The methods naturally
depend on the chemical nature of the mixture under consideration as well as
the temperature and pressure conditions. They involve experimental data to a
greater or lesser degree and the availability of these data must be taken into
account when a model is chosen.
We will distinguish between two main categories of methods:
“heterogeneous” methods that apply different models to the two phases
comprising the system, thereby accounting for the radical differences in
the liquid and vapor properties, far from the critical point;
“homogeneous” methods which in contrast apply the same model (an
equation of state) to the two existing phases, thereby respecting the con-
tinuity of fluid states of matter.
Each approach has its advantages that we will endeavor to underscore.

2.1.3 “Heterogeneous” Methods Applied to


Calculate the K Value
2.1.3.1 The K Value
Heterogeneous methods are usually applied at low pressure. The behavior of
the vapor phase is then similar to that of an ideal gas. The methods were par-
ticularly developed to represent the liquid phase and the phase equilibria of
mixtures containing polar components.
34 Chapter 2. THERMODYNAMICS: PHASE EQUlLlBRlA

Fugacity in the vapor phase is expressed based on the partial pressure:


fy = Py;(py (2.10)
where represents the “fugacity coefficient”, which is close to one at low
pressure. If the desired precision so demands or if the pressure is higher than
one or two bars, then the coefficient can be calculated by means of an equa-
tion of state such as the Redlich-Kwong equation (1949) (see Section 2.1.4.1)
or the virial equation (see Vidal, 1997, Chapters 4 and 8).
There are several steps used to calculate fugacity in the liquid phase.
First the fugacity of the pure substance is evaluated at vapor-liquid equi-
librium (state of “saturation”, represented here by the exponent o), i.e. at a
pressure equal to its vapor pressure, Pp, at the temperature under considera-
tion:
fp = ppqp
This fugacity is the same in the liquid or vapor phase since there is equi-
librium between the two phases. The fugacity coefficient at saturation, cpp, is
close to one if the vapor pressure is low. Otherwise, it is like (py evaluated with
an equation of state.
Then the value obtained is corrected to take the influence of the pressure,
P, into account. The corrective term introduced is called “Poynting correc-
tion”, 9;:
ffSL(T,P) = f p 9 ; = Ppq~pP,
the value of 9;
generally remains close to one, as can be seen below:

9;=exp [ vi*.L ( P - Pp)


RT ]
where u:, represents the molar volume in the liquid phase, with the exponent
* standing for the properties of pure components in the same temperature and
pressure conditions as the mixture.
Then the composition is taken into account. In an “ideal” solution by defi-
nition, fugacity is proportional to the mole fraction. However, this simplifica-
tion is possible only if the mixture is made up solely of compounds that are
close to one another in terms of their chemical structure and their molecular
volume: for example mixtures of C&, paraffins or C,-C, aromatics.
Generally speaking, the liquid phase activity coefficient, y;, must finally be
introduced:
fk(T, P, X) = ff.‘(T, P)x;Y;
The following is the final result:
fk(T, P, X) = ff.L(T,P) xiy; = PP(ppY;x,y) (2.1 1)
and the equilibrium condition is expressed as follows:
Py;(py = Pp(pp9;xiy;
Chapter 2 THERMODYNAMICS: PHASE EOWLIBR/A 35

The K value expression helps to distinguish the main elements of the equa-
tion better:
(2.12)

The quotient Pp/P represents the K value that is obtained by applying


Raoult’s law. It is appropriate if the pressure and the vapor pressure of the
component under consideration are low and if the solution is ideal. The sec-
ond quotient stands for the corrective terms which are generally close to one
and basically depend on the deviations from ideal gas laws in the vapor phase.
Lastly the activity coefficient plays an essential part since it represents the sin-
gularities that are specific to the liquid mixture. It depends on the nature of the
components, the temperature and the composition, but the influence of the
pressure is usually negligible.
It should be pointed out that even though the preceding expression is gen-
eral, it does not apply if the temperature is higher than the critical tempera-
ture of the component under consideration, since in this case the vapor
pressure Pp can not be evaluated. In addition, it should be used very carefully
if the vapor pressure or the total pressure is high, since here the calculation of
corrective terms would take on particular importance.
In conclusion, it can be seen that close to atmospheric pressure the fol-
lowing equation can be written as a first approach:

(2.13)

This simplified form concentrates the two main elements of the K value at
low pressure:
the vapor pressure: it is easy to understand that the precision of vapor-
liquid equilibrium calculations for mixtures is limited by the precision of
the vapor pressure values for the components of the mixture:
the deviations from ideality in the liquid phase, represented by the sym-
bol yk, that we plan to develop.

2.1.3.2 Deviations from Ideality in the Liquid Phase


Activity coefficients depend on the composition and it is ordinarily in a dilute
medium that a component’s behavior differs most from its properties in an
ideal solution. Infinite dilution activity coefficients therefore are a sort of
“scale of non-ideality”. Table 2.3 concerns mixtures of hydrocarbons and it
should be noted that mixtures containing paraffins and aromatics are highly
non-ideal. For example in a dilute solution in benzene, the volatility of heptane
(expressed by the product Ppyk which, in an ideal solution, is the same as
the vapor pressure) has increased by some 70%. As a result, despite the dif-
ference in boiling point (SOT for benzene, 100°C for heptane), separation by
simple distillation is impossible. In any case, this type of deviation can not be
disregarded. If the molar volumes are not very different, mixtures of hydro-
carbons of the same family can, however, be considered as ideal. This is shown
36 Chapter 2. THERMODYNAMICS:
PHASE EQUfLfBRlA

for example by the linear variation of the bubble point pressure with compo-
sition (Raoult’s law) illustrated by Figure 2.3 for the mixture of propane and
n-pentane.

Hexane Benzene 68
YPL
1 diluted in 2

1.7
2 diluted
y2”’Lin 1 I
Heptane Benzene 60 1.7
Heptane Benzene 80 1.6
Hexane Toluene 24.8 1.8
Heptane Toluene 98.5 1.4 1.3
Cyclohexane Benzene 60 1.5 1.4
Cyclohexane Toluene 81 1.35 1.35

Activity coefficientsat infinite dilution (represented by the exponent W)


in mixtures of hydrocarbons.

Mixtures containing both apolar compounds (hydrocarbons) and polar


compounds exhibit deviations from ideality of a totally different order of mag-
nitude. Table 2.4 gives the activity coefficients at infinite dilution of heptane
and benzene in several polar compounds as an example.

Acetone 40 6.4 31 1.6


Methanol 40 34 30 7.2
Ethanol 40 13 45 5.1
Dirnethylformamide 25 21 25 1.4
Dirnethylsulfoxide 25 121 25 3.5
Ethylene glycol 25 750 25 32
Table
Activity coefficientsat infinite dilution of heptane ( I ) and benzene (2)
in polar compounds.

It can be seen that in a dilute medium, the volatility of a hydrocarbon is


considerably modified and that the polar compound acts selectively since the
activity coefficient is very sensitive to the type of hydrocarbon. For example
in dimethylformamide, the relative volatility of heptane and of benzene is mul-
tiplied by the ratio of activity coefficients, i.e. 21/1.4. In extractive distillation
carried out in the presence of this solvent, heptane will be eliminated at the
top of the column even though its boiling point is 20°C higher than that of ben-
zene.
High activity coefficient values often induce partial miscibility in the liquid
phase. This is the case for heptane in methanol, dimethylformamide, dimethyl-
sulfoxide and ethylene glycol. The miscibility of aromatic hydrocarbons in the
Chapter 2 THERMODYNAMICS: PHASE EQUILIBRIA 37

same solvents is usually total, or in any case much greater, and purification
processes by liquid-liquid extraction are based on this selectivity.
The examples of deviation from ideality that we have just discussed exhibit
"positive" deviations: activity coefficients are greater than one and the excess
Gibbs energy is positive (see Section 2.1.3.3). If the azeotrope phenomenon
exists, it will be perceived by a pressure maximum on isothermal vapor-liquid
equilibrium diagrams (or a temperature minimum on isobaric diagrams). This
is the most frequent case, at least as far as "nonelectrolytic" mixtures are con-
cerned. The differences in molecular volume lead to negative deviations,
which can be considerable in the case of solutions of polymers.
On the molecular scale, the mixture of two polar compounds gives rise to
a break in dipolar interaction between identical molecules and a recombina-
tion of this type of interaction between different molecules. Generally speak-
ing, it is difficult to predict the sign and the amplitude of the deviations from
ideality that come as a result.

2.1.3.3 Prediction and Calculation of Activity Coefficients


The models that we will present sometimes involve activity coefficients them-
selves, but usually involve the excess Gibbs energy related to activity coeffi-
cients, 1.4, and to partial molar excess Gibbs energy or excess chemical
potential, p:, by the equations below:

(2.14)

GEvL= C NiRT In f (2.15)


We will not go into detail on how these models are worked out since we pre-
fer to emphasize their qualities and their area of application. It is in particular
very important to discern the "predictive" qualities, i.e. the nature and acces-
sibility of the data required to use them.
a. Methods Using the Regular Solutions Model
This model (Hildebrand and Scott, 1924; Hildebrand et al., 1970), offers the
advantage of being expressed by an equation that includes, in addition to the
temperature and composition variables, only parameters that can be readily
determined from the properties of the components. If is used to represent
the molar volumes of pure substances in the liquid phase, then the equation
below is applied:
(2.16)

where 6; represents the solubility parameter of component i, which can be cal-


culated from the heat of vaporization Ah?; 6 , is obtained by volume fraction
linear combination Qiof the values of 6;:
38 Chapter 2. THERMODYNAMICS: PHASE EQUILIBRIA

(2.18)

(2.19)

Activity coefficients can therefore be ca.culated provided that the molar


volume and solubility parameters are known for each component of the mix-
ture, or if they are not available, the heat of vaporization can be used.
However, this model can be applied only to non-polar mixtures. Accuracy
is usually satisfactory when the model is applied to hydrocarbon solutions.
The Chao and Seader Method (1 96 1)
In theory nothing prevents application of a "heterogeneous" method of calcu-
lating vapor-liquid equilibria under pressure, as long as the deviations of the
vapor phase from the ideal gas laws and the Poynting correction are taken into
account of course. This is what was done by Chao and Seader: Eq. 2.12 is writ-
ten as follows:

The term vT is independent of the composition of the mixture, it is


expressed in function of the reduced temperature, the reduced pressure and
the acentric factor by application of the law of the corresponding states
(Paradowski, Vol. 1, Chapter 4). The fugacity coefficient in the vapor phase, (PY,
is calculated by means of the Redlich-Kwong equation of state (see
Section 2.1.4.1) and the activity coefficient in the liquid phase is obtained by
application of the regular solutions model. It should be noted that this method
can be applied to mixtures containing compounds in the supercritical condi-
tion by extending the calculation of the term vT and by defining solubility
parameters inherent to the compounds. Table 2.5 provides examples of the
values of molar volumes and solubility parameters that have been used by the
above-mentioned authors.

Compound

Ethane 68 12.38 Propene 79 13.15


Propane 84 13.09 Butenel 95.3 13.83
n-Butane 101.4 13.77 Pentene-l 110.4 14.42
n-Pentane 116.1 14.36 Cyclohexane 108.7 16.77
n-Hexane 131.6 14.87 Benzene 89.4 18.74
n-Heptane 147.5 15.20 Toluene 106.8 18.25
Ethylene 61 12.44 pXylene 124 17.94
1 Table

w Molar volumes and solubility parameters (Chao and Seader, 1961).


Chapter 2 THERMODYNAMICS: PHASE ~OUlLlBRlA 39

The advantage of the Chao and Seader method is that it is predictive in the
sense that application requires only the relevant, readily accessible data for
the components of the system such as critical coordinates, acentric factor, sol-
ubility parameter and molar volume in the liquid phase. However, the limits of
application detailed by the authors must be complied with:
It is applied only to mixtures of hydrocarbons, with hydrogen also
included.
For hydrocarbons, except for methane, the reduced temperature of each
component must range between 0.5 and 1.3, and the pressure must be
lower than 130 bar, without however going over 80% of the critical pres-
sure of the mixture.
In the presence of hydrogen or methane, the temperature must range
between -75°C and 93% of the pseudocritical temperature of the mixture
(expressed in K), without however going over 260°C. The pressure must
be lower than 540 bar and the concentration of the components must not
exceed 20% molar.
According to the authors, under these conditions the mean relative error
on K values is approximately 8%.
We will see later on (see Section 2.1.4) that “homogeneous”methods can be
used in vapor-liquid equilibrium calculations over a wider pressure range and
usually with better accuracy. However, the presence of binary interaction
terms makes them less predictive. The Chao and Seader method must there-
fore not be disregarded on the grounds that it is older.
Grayson and Streed (1963) made some minor modifications to the method
which, according to the authors, enhance the results obtained for mixtures
containing hydrogen or methane.
b. Models Using the Concept of Local Composition
These models, whose considerable advantage is that they can be applied to
mixtures containing polar or non-polar components, are based on the concept
of local composition. For a molecule belonging to a mixture, the composition
of its immediate environment, which determines the molecular interactions it
is involved in, can differ considerably from that of the mixture. For example in
a mixture of hydrocarbons and alcohol, the alcohol molecules tend to be sur-
rounded preferentially by other alcohol molecules. The tendency - a com-
monly known fact -can go to the point of the mixture separating out into two
liquid phases. Stated and implemented by Wilson (1964), this concept gave
rise to several models, among which first and foremost “Wilson’s equation”.
We will discuss the more recent NRTL and UNIQUAC models here.
The NRTL Model (Non-Random Two Liquids)
This model (Renon et al., 1971) is expressed by the equation below:
40 Chapter 2. THERMODYNAMICS: PHASE ~OUlLlBRlA

with:
T..=
q,i
- (2.21)
J~' RT
It therefore depends on three parameters: q,;,
Ci,j,.and aj,;= au, which must
be determined by binary experimental data correlation.
It has also been proposed to take the variation of these parameters with
temperature into account by using:
J , l + C(T)
C.1.I. = C(OJ J.1 (T - To) and aj,;=a$J+ al? (T - To) (2.22)
The model then contains six parameters. In fact the number and nature of
parameters that can be determined depend largely on the number and nature
of data that are available and the amplitude of deviations from ideality as
shown in Table 2.6 (Renon et al., 1971, p. 30). The variation of parameters with
temperature can only be taken into account if the vapor-liquid equilibrium
data extend over a wide temperature range, or if heat of mixing data for the
mixture can be added. Even if the limit is fixed at three parameters, their appli-
cation to systems that are moderately non-ideal, or not very polar, will show
that these parameters are correlated and it is then recommended to set the
parameter a,whose stand values are equal to 0.2 for systems that are not very
polar and 0.3 for polar systems.

Available data
g E ,hM g E ,hM gE
in a wide at only one at only one
temperature temperature temperature
range
Order of mag-
NRTL parameters to be determined
nitude of gE
6 parameters 4 parameters 2 parameters
ap2= 0.2 ap2= 0.2
Weak aT2= 0 CT1 = cT2= a;* = 0
45, < 0,3 RT Example: Example: Example:
benzene - n-heptane water - acetic acid benzene -
dimethylsulfoxide
6 parameters 6 parameters 3 parameters
Strong
cT -cT
21-
-
12-'YT2=0
> 0.3 RT Example: Example: Example:
acetone - water ethanol - n-hexane - ethanol
cyclohexane

Determination o fNRTL model parameters depending on the nature ofthe data


available and the degree o f non-ideality
gE: molar excess Gibbs energy Cjoules/mole o f mixture)
hM:heat of mixing (joules/mole o f mixture).
Chapter 2. THERMODYNAMICS: PHASE ECJUlLlBRlA 41

The UNIQUAC Model (Universal Quasi Chemical)


Proposed shortly after the NRTL model, the UNIQUAC model (Abrams and
Prausnitz, 1975; Maurer and Prausnitz, 1978; Prausnitz et al., 1980) is also
based on the concept of local composition. However, it expresses the energy
balance of the mixture operation by taking the size and shape of molecules
into account with what is called a “combinatorial” term. This term involves
molecular volume and area parameters in the expression of the activity coeffi-
cient. Likewise, in the calculation of molecular interactions the area of the
molecules and surface fractions are used to define what is called a ”residual”
term. The result is the following equations:

2
1; = - (ri- q> - (r;- 1) with z = 10
2

where the parameters ri and qi stand for the molecular areas and volumes of
the components respectively. They can be calculated, as we will see for the
UNIFAC method, by contributing groups. The terms @; and 8, are surface and
volume fractions:
rixi (7;xi
@;= 7 ei. = -
n (2.26)
c
‘ixj
j=1
C
qjxj
j=1

The binary parameters depend on the temperature:

so that the UNIQUAC method characterizes each binary by means of two coef-
ficients, aU and ai,;,which must be determined from experimental data in the
same way as for the NRTL model.
The NRTL and UNIQUAC models are comparable in their area of application
(polar or non-polar compounds) and their predictive qualities. Their binary
parameters must be determined by experimental data correlation and the
NRTL model, which contains three binary parameters, is less advantageous
from this standpoint.
They usually result in an excellent correlation of binary vapor-liquid equi-
libria and satisfactory prediction of vapor-liquid equilibria for ternary or
higher order mixtures. They can also be used to calculate heat of mixing, par-
ticularly if the variation in parameters with temperature is considered. Finally
they can be applied to the calculation of liquid-liquid equilibria. It should, how-
ever, be noted that in this case the presence of three parameters in the NRTL
model is somewhat advantageous.
42 Chapter 2. THERMODYNAMICS:
PHASE ~CJUlLl8RlA

In any case, the application of these models to a complex mixture entails con-
siderable correlation work, since a system made up of ten components, which
is not exceptional in practice, contains 45 binary systems. However, there are
data bases which, for a large number of systems, propose the parameter val-
ues for the models we have presented alongside experimental values. The best
known and the most complete is the one initiated by the University of
Dortmund (Gmehling et al., 1978-1984), an example of its contents is given in
Table 2.7.
Wilson's equation, and the NRTL and UNIQUAC models are based on the con-
cept of local composition. However, they can not be considered as particularly
rigorous and their derivation involves necessary but questionable assump
tions and sometimes even incoherence and lastly a large dose of empiricism
(Renon, 1985). They are tried and true though, even if they should be applied
only by those with long thorough experience in the field.
What may happen is that no experimental data related to the system under
study are available. If the system is one of the "key components" in a separa-
tion then we have to "go back to the laboratory" or if not try to apply a "group
contribution" method.

c. Models Using Group Contribution: ASOG and UNIFAC Models


Even if ionic solutions are excluded, the number of compounds that can be
involved in a mixture is so high that there can be no hope of having all the
numerical values of the parameters used in the models that we have just dis-
cussed. However, the molecular interactions that the models are based on
lend themselves to a more refined, and mainly simplifying, analysis. If excep
tion is made of the simplest compounds, molecular interactions are in fact the
result of interactions existing between the groups that make up the molecules.
For example in a mixture of n-alkanes, the following interactions will be distin-
guished: CH,-CH,, CH3-CH2 and CH2-CH2, etc. This analysis provides consider-
able simplification since the number of atomic groups is much smaller than
the number of chemical individuals. A parallel can be established between the
structures of groups, compounds and mixtures on the one hand, and on the
other between letters, words and sentences.
Nevertheless, there are some difficulties.
The first concerns the definition of groups: can the terminal methyl group
of a paraffin be regarded as identical to the group present in ethanol, acetone
or even in toluene? Do the three methyl groups in isopentane have identical
properties? Each group should be defined not only by its nature but also by
the nature of its immediate environment in the molecule that contains it.
Accordingly, a distinction would be made for the preceding example between:
CH, (CHd, CH, (CH), CH, (CH,OH), CH,(CO), and CH, (C arom.). This has been
done in certain cases in calculating the thermodynamic properties in the ideal
gas state for example (Benson et al., 1969). However, the result is a multiplica-
tion of distinct groups, and therefore of parameters representing their inter-
actions for the calculation of properties in the dense phase. Since these
parameters will have to be determined from experimental measurements, the
(2)

(1)
(2)
-
(1) ACEXINE

ANTOINE CONSTANTS TEMPERATURERANGE


7.11714 1210.595 229.664 from-13to ~ 5 5 ° C
6.91058 1189.640 226.280 from-30to +17O"C
Method 1
Method 2
C3H60
'BH14
c0"cy
+
Temperature = 45.00 degree C
Lit : Schaefer K., Rall W., 2. EleMrochem. 62,1090 (1958).

Constants :
WILSON
Margules 1.5061
1.5055
962.8113 11.4214
699.6735 B8547 0.4817 ow 020 040 060 080 IW
Uniquac - 42.4860 55658
Experimental data Ma1 iles Van Laar Wilson I NRTL Un uac
P mm Hg XI Y. Diff. P Diff. Y1 Diff. P Diff. Y, Diff. P Diff. Y1 Diff. P Diff. Y, Diff. P Diff. Yl
339.40 0.0 0.0 4.21 0.0 4.21 0.0 421 0.0 4.21 0.0 4.21 0.0
444.60 0.0651 0.2828 3.10 - 0.0032 3.25 -0.0030 - 3.48 -0,0122 -2.44 - 0.0109 2.36 - 0.0042
545.80 0.1592 0.4442 10.11 - 0.0104 10.42 -0.0101 9.40 - 0.0065 9.34 - 0.0074 10.16 - 0.0098
590.20 0.2549 0.5163 3.32 -0.0163 3.63 -0.0161 7.95 - 0.0063 7.12 -O.OO80 4.07 -0.0149
617.30 0.3478 0.5560 4.90 -0.0177 5.12 -0.0177 10.83 - 0.0078 9.82 -0.0094 5.78 -0.0165
632.60 0.4429 0.5866 7.17 -0,0133 7.29 -0.0135 12.06 -0.0074 11.16 - 0.0084 7.82 -0,0128
639.60 0.5210 0.6068 8.65 - 0.0096 8.70 -0.0100 12.30 -0.0083 11.55 - 0.0087 9.07 - 0.0098
633.80 0.5907 0.6258 0.26 - 0.0049 0.28 -0.0053 3.12 - 0.0077 2.48 - 0.0075 0.55 - 0.0057
637.10 0.6202 0.6339 3.05 - 0.0034 3.06 -0.0039 5.69 - 0.0079 5.09 -0.0073 3.30 - 0.0044
631.00 0.7168 0.6662 - 1.42 0.0011 -1.45 0.0007 0.79 - 0.0069 0.24 - 0.0056 - 1.25 - 0.0002
627.80 0.7923 0.7034 2.88 0.0041 2.79 0.0038 4.26 -0.0037 3.81 - 0.0021 2.88 0.0029
623.30 0.8022 0.7292 0.04 0.0240 -0.07 0.0238 1.22 0.0165 0.80 0.0182 - 0.00 0.0230
603.40 0.8692 0.7583 - 2.33 0.0010 -2.52 0.0010 -3.16 -0,0025 -3.22 -0.0014 -2.71 0.0007 -0
583.20 0.9288 0.8255 6.82 - 0.0075 6.58 -0,0072 3.82 - 0.0052 4.30 - 0.0054 6.17 - 0.0069 3
543.30 0.9658 0.9003 - 4.81 -0.0044 -5.00 -0.0041 - 7.90 0.0001 -7.26 -0 . m - 5.38 - 0.0035 %
505.00 1.oooO 1.oooO - 7.38 0.0 - 7.38 0.0 - 7.38 0.0 - 7.38 0.0 - 7.38 0.0 nl
0

Table
Mean deviation:
Max. deviation:
4.21
10.11
0.0086
0.0240 I :::1 I 0.00%
0.0238
6.14
12.30
0.0071
0.0165
5.62
11.55
0.0072
0.0182
4.39
10.16
0.0082
0.0230
E5
P
Isothermal vapor-liquid equilibrium of the acetone/water mixture. Source: Vapor-liquidequilibrium data collection, Chemistry Data
P
Series, Dechema, Frankfurt. 0
44 Chapter 2. THERMODYNAMICS: PHASE ~QUlLlER/A

availability of data once again becomes a problem. The method will unques-
tionably be more accurate, but less "predictive".
Likewise, two contiguous groups ("in a")exert an undeniable influence on
each other, with a case in point being ethylene glycol which can not be
described as made up of two methylene groups and two hydroxyl groups. This
"proximity effect" makes itself felt less and less the longer the hydrocarbon
chain separating the two functional groups. The effect has been studied more
particularly by Kehiaian (1983).
Some structures, in particular the first terms of homologous series, can not
be broken down into groups. They are often compounds of considerable prac-
tical importance, for example methanol. Solvents like N,Ndimethylformamide
and chlorofluorinated compounds of methane and ethane, will form a group by
themselves whose structural properties (volume, area) will sometimes be an
order of magnitude greater than those of the other groups.
Despite these difficulties, for binary or higher order systems the group con-
tribution methods are commonly applied and rightly so, to predict excess
properties since the expectations of accuracy are not as high as for the mod-
els based directly on experimental data correlation.
The ASOG (Kojima and Tochigi, 1979) and UNIFAC methods (Fredenslund
et al., 1975, 1977) both comply with the first and the third of the propositions
and assumptions stated by Wilson and Deal (1962):
The partial molar excess Gibbs energy:
pBvL= RT In y) (2.14)
is the sum of two contributions: the first is termed combinatorial and cor-
responds to the differences in size and shape of the components of the
mixture, while the second is termed residual and corresponds to interac-
tions between groups. Therefore the following can be written:
In Y! = In Ykcombinatorial +In Ykresidual (2.23)
The residual term is estimated by substituting the concept of "solution of
groups" for the concept of mixture of chemical compounds. For example,
in a mixture containing one mole of n-hexane and one mole of acetone,
the hexane provides 2 moles of CH, groups, 4 moles of CH, groups, and
the acetone supplies 2 moles of CH3 groups and one mole of C=O group.
The molar composition of the "solution of groups" is therefore 4/9 for the
CH, groups, 4/9 for the CH, groups and 1/9 for the C=O group. Of course
the components themselves will be considered as "particular solutions of
groups", with the composition of hexane 2/6 for the CH, groups, 4/6 for
the CH, groups and that of acetone 2/3 for the CH, groups and 1/3 for the
C=O group.
The solutions of groups are not ideal and each group k is characterized
by its mole fraction X, and its activity coefficient r,.
The residual activity coefficients of the components themselves, Ykresidud
are related to them by the equation below:
ng
In Ykresidual = c
k=l
"k,i@ rk - In rk,i) (2.28)
Next Page

Chapter 2. THERMODYNAMICS:
PHASE ~OUlLIBRlA 45

where v k , ; stands for the number of groups k in component i, r k , i is the


activity coefficient of group k in the mixture that is made up of the com-
pound i pure, r k indicates the activity coefficient of this same group in
the mixture as such and ng is the number of distinct groups.
The ASOG and UNlFAC methods differ in the models that have been chosen
to express the combinatorial term and the residual term as well as in the defi-
nition of the groups. We will present the UNIFAC model which is more com-
monly used.
The UNIFACModel
The UNlFAC model was proposed by Fredenslund et al. (1975). It stems directly
from the formalism defined by Wilson (Eqs. 2.23 and 2.28) and from the
UNIQUAC method. Accordingly, the combinatorial term is calculated from the
equation:
in y~combinatoria,
= In 2
@, ei + li - - C xi$
+ z- qiIn - (2.24)
x; 2 @i Xi j = 1

z
1. = - (r. - q> - (ri - 1) with t = 10
I 2 I
To calculate the volume and surface fractions of the components, the cor-
responding molecular parameters, r; and q;,must be known. These will be cal-
culated from the volume and surface parameters of each group proposed by
Bondi (1968), Rk and Qk, by simple additivity rules:
ng ng
ri= vk,; Rk qi = C vk,; Qk (2.29)
k=l k=l

To obtain the residual term:


ng
In Ykresidual = C vk,i On rk - In l-k,;) (2.28)
k=l

the activity coefficients of the groups, r k , must be calculated. The UNIQUAC


method gives the equation below transposed from Eq. 2.25:

where the surface fractions of the groups, Ok, appear as deduced from the
mole fraction X k by the equations below:

(2.31)

Das könnte Ihnen auch gefallen