Sie sind auf Seite 1von 118

1

I. Thermodynamics of mixtures

I. 1. Introduction
The chemical industry typically deals with mixtures in the process of making pure components or
even mixtures as a product. Dealing with pure substances is the exception, and in some cases
streams can be approximated by assuming that they are pure substances. It is then important to
know when this assumption will fail. It is often the task of a chemical engineer to ensure the
separation of a relatively pure target product from a mixture.

The task of a chemical engineer is to design, optimize, and operate chemical processes. The starting
point for these activities are the mass and energy balance around a system. On a plant, we typically
measure volumetric flow rates (e.g. via an orifice plate, Venturi meter, etc.), although more mass
flow meters are coming into practice (even taken the old bucket and stopwatch method).


Hence, to understand and being able to model an existing process, the volumetric flow rate must be
converted into a mass flow rate. This can be done using the density of the particular mixture
(
mixture
):


with w
i
: the mass fraction of component i in the mixture

i
: the density of the pure component i
The underlying assumption here is, that the volume of the various components added to the mixture
is additive. This assumption needs to be investigated.

Knowing the temperature and pressure (easily measured quantities) and the composition (not that
easily determined), the system is completely described. This means that all other thermodynamic
quantities are now fixed. For the energy balance around a system, it is important to know what the
enthalpy of the streams are (in relation to a pre-defined reference state), so that the heat required
to be added or removed can be predicted. We need therefore a tool to calculate the enthalpy of the
streams relative to the reference state knowing the temperature, pressure and composition of a
stream.

For a chemical engineer it is not only important to know how much of a product can be produced (as
given by the mass balance, thermodynamic constraints, and the size of the equipment), but also
whether the production is as efficient as possible. Thermodynamics can set targets with respect to
the amount of product formed (chemical equilibrium considerations), but also with respect to the
efficiency of the process. The thermodynamic efficiency of a process can be defined as
SYSTEM
V
in,1
T
in,1
,p
in,1
,x
in,1
State of aggregation
V
in,2
T
in,2
,p
in,2
,x
in,2
State of aggregation
V
out
T
out
,p
out
,x
out
State of aggregation
.
. .
2


(for inlet and outlet streams at 298 K). Knowing the maximum theoretical efficiency of a process sets
a target for the optimization of the process and can be used to analyze the part of the process with
the largest thermodynamic losses.

It is thus important to understand and possibly predict the behavior of mixtures in order to be able
to determine the volumetric, enthalpic and entropic properties. Mixtures differ from pure
substances. Mixtures differ from pure components upon phase change. Pure substances will
melt/evaporate at a particular temperature (for a given pressure) or at a particular pressure (for a
given temperature). This implies that there is only one single state of aggregation possible for a pure
substance (i.e. either gas, liquid or solid). In general, mixtures of a particular composition do not
melt/evaporate at a single temperature (for a given pressure) or at a single pressure (for a given
temperature). They typically show a range where two or more phases can be present (depending on
the number of components present in the system).

Mixing of pure components does yield some interesting properties, such as change in volume and
change in temperature upon adiabatic mixing of streams.


Example: Mixing equal amount of water and ethanol
Mixing 50 ml of water with 50 ml of ethanol yields a mixture of containing 50 vol.-% ethanol. This
process is associated with a volume contraction (see Fig. 1.1). What is the volume of this mixture?
Data:
H
2
O
= 0.99708 g/cm
3
; M
H
2
O
= 18.02 g/mol;
ethanol
= 0.78506 g/cm
3
; M
ethanol
= 46.07 g/mol


Figure I.1.1: Change in the volume of the mixture per mole of mixture as a function of the mole
fraction of ethanol in the mixture (redrawn from J.-P.E. Grolier and E. Wilhelm, Fluid
Phase Equilibria 6 (1981), 283-287)

50 ml of ethanol contains 39.465 g ethanol 0.8567 mol ethanol

50 ml of water contains 49.850 g water 2.7664 mol water

A mixture of 50 ml ethanol and 50 ml of water contains the same number of moles of water and
ethanol as for the pure substances (mole balance!). Hence, the mole fraction of ethanol in the
mixture is given by:



-1.2
-0.9
-0.6
-0.3
0
0 0.2 0.4 0.6 0.8 1
C
h
a
n
g
e

i
n

v
o
l
u
m
e

u
p
o
n

m
i
x
i
n
g
,

c
m
3
/
m
o
l

o
f

m
i
x
t
u
r
e
Mole fraction of ethanol in mixture, x
EtOH
AV (x
EtOH
= 0.2364) =
-0.9552 cm
3
/mol
3

Hence, the change in the volume upon mixing is -0.9552 cm
3
/mol of mixture (see Fig. 1.1). This
mixture contains 3.6230 moles (0.8567 mol of ethanol and 2.7664 mol of water). Thus, the volume
reduction is now


And the total volume of the mixture is


(AND NOT 100 ml!).
The reduction in the volume is due to the molecular interaction between ethanol and water. The
resulting interaction (due to H-bonding) will result in a closer packing of water and ethanol
molecules resulting in a volume reduction in comparison to the pure liquids.

The molar volume of the mixture can thus be described in terms of the molar volume of the pure
components and the change in the volume upon mixing to form 1 mole of mixture:


The change in the volume upon mixing to form 1 mole of mixture is dependent on temperature,
pressure and the composition of the mixture:

( ) (

( ); (

( ))
(the dependency on pressure is typically neglected when dealing with liquids, since liquids can be
considered to be incompressible).

Furthermore, the mixing process may involve heat effects. The mixing process can be exothermic,
upon adding two pure components together, i.e. the mixture has a higher temperature upon
adiabatic mixing than the original components or otherwise stated heat has to be removed to keep
the temperature constant (e.g. in the case of mixing sulfuric acid with water or to a lesser extent
mixing water and ethanol), or endothermic, i.e. the mixture has a lower temperature upon
adiabatic mixing than the original components or otherwise stated heat has to be added to keep the
temperature constant (e.g. in the case of mixing ethanol with benzene). It may even have a more
complex dependency as given in the example below.

The molar enthalpy of a mixture can be defined in a similar manner as the molar volume of a
mixture, i.e. as a function of the molar enthalpy of the pure components and the change in the
enthalpy upon mixing to form 1 mole of a mixture:


The change in the volume upon mixing to form 1 mole of mixture is dependent on temperature,
pressure and the composition of the mixture:

( ) (

( ); (

( ))
(the dependency on pressure is typically neglected when dealing with liquids, since molar volume of
liquids is rather small).


Example: Isothermal mixing of water and acetone
For a process it is necessary to separate isothermally at 25
o
C a liquid stream (2 m
3
/s) containing 20
mol-% acetone in water into its pure components. How much heat has to be added/removed for the
separation process? What is the volumetric flow rate of each of the pure component streams?
Data:
H
2
O
= 0.9970 g/cm
3
; M
H
2
O
= 18.02 g/mol;
acetone
= 0.79042 g/cm
3
; M
acetone
= 58.08 g/mol
4


Figure I.1.2: Change in the volume (left) and change in the enthalpy (right) for mixtures of acetone
and water at 298.15 K (data: change in volume H.K. Bae and H.-C. Song, Korean
Journal of Chemical Engineering 15(6) (1998), 615-618; change in enthalpy B. Lwen
and S. Schulz, Thermochimica Acta 262 (1995). 69-82)



A mole balance on the process:
Acetone




Water






The main problem is now to find the molar flow rate of the mixture entering the process. Taking a
basis of 1 mol of a mixture containing 0.8 mol of water and 0.2 mol of acetone, the mass of water
and acetone can be determined:



The change in the volume for a mixture containing 20 mol-% acetone is -1.1929 cm
3
/mol.
The volume of 1 mole of the mixture is given by:



Thus the molar volume of the mixture is:
-1.6
-1.2
-0.8
-0.4
0
0 0.2 0.4 0.6 0.8 1
C
h
a
n
g
e

i
n

v
o
l
u
m
e

u
p
o
n

m
i
x
i
n
g
,

c
m
3
/
m
o
l

o
f

m
i
x
t
u
r
e
Mole fraction of acetone in mixture, x
acetone
AV (x
acetone
= 0.2) =
-1.1929 cm
3
/mol
-800
-400
0
400
0 0.2 0.4 0.6 0.8 1
C
h
a
n
g
e

i
n

e
n
t
h
a
p
l
y

u
p
o
n

m
i
x
i
n
g
,

J
/
m
o
l

o
f

m
i
x
t
u
r
e
Mole fraction of acetone in mixture, x
acetone
AH (x
acetone
= 0.2) =
-651.768 J/mol
S
e
p
a
r
a
t
i
o
n

p
r
o
c
e
s
s Mixture (2 m
3
/s)
x
acetone
= 0.2
x
water
= 0.8
T = 298.15 K
Acetone (? m
3
/s)
x
acetone
= 1.0
x
water
= 0.0
T = 298.15 K
water (? m
3
/s)
x
acetone
= 0.0
x
water
= 1.0
T = 298.15 K
5



Hence, a flow of the mixture of 2m
3
/s corresponds to a flow of 70.07 kmol/s (i.e. 56.06 kmol/s of
water and 14.01 kmol/s of acetone). The flow rates of the pure streams leaving the process are thus
given by


Similarly for acetone



The heat added/removed can now be determined using an energy balance. For this, the definition of
the change in the enthalpy as given in Fig. 1.2 needs to be defined
()

()

()

()
The enthalpy is a function of temperature. It is important to realize that all enthalpies must be taken
all at the same temperature (298.15 K). The energy balance on the process:



with

((

))

)

and

)

Substituting the mole balance and realizing that T
inlet
=T
oulet
:

)

Substituting into the energy balance

)
Thus, the heat to be added to the system amounts to 45.67 MW.


The change in the volumetric properties and enthalpy upon mixing can be understood in terms of
molecular interactions between the various molecules present in the mixture. The molecular
interactions are associated with energetic effects (internal energy, enthalpy). Furthermore, mixing
results in an increase in the entropy of the mixture.

6

For pure substances, the interaction between the molecules was taken into account using an
appropriate equation of state (and it was shown that cubic equations of state were useful to predict
properties around the vapour-liquid equilibrium). The thermodynamic functions were then
developed around the idea that the property can be described as a deviation of the thermodynamic
function from its value, when the compound at a particular temperature and pressure could be
considered to be an ideal gas.
Enthalpy
Temperature change


pressure change (

) ( (



Entropy
Temperature change


pressure change (

) ((



Similarly, the departure of the Gibbs free energy from ideality was formulated in terms of the
fugacity:

(



Hence, the appropriate equation of state relating the molar volume to temperature and pressure
yields the information on the thermodynamic functions at a given temperature and pressure for a
pure substance. The specific molecular interactions between the molecules were introduced through
the critical properties of the pure components.



7

I.2. Ideal gas mixtures non-interacting mixtures
Mixtures differ from pure components due to the interaction of the different components in the
mixture with each other. An ideal gas is a state, in which molecules are point-like objects, which do
not interact with each other, and an ideal gas mixture is thus a mixture where the molecules do not
interact with each other. The difference between an ideal gas and an ideal gas mixture lies in the
fact that the ideal gas contains indistinguishable molecules, whereas the ideal gas mixture contains
molecules which are distinguishable.

In an ideal gas mixture (IGM), the pressure is given by the ideal gas law:


The partial pressure of each component in the mixture is defined in terms of the total pressure of
the system, and the mole fraction of each component:


or otherwise stated the partial pressure of a component i in an ideal gas mixture (p
i
IGM
)


Thus, the partial pressure of a component in an ideal gas mixture is the pressure exerted by the
same number of molecules as present in the mixture, at the same temperature of the mixture (and
thus average velocity), and in the same volume as the volume of the mixture.

It should be noted that the volume of the system does not change upon mixing to form an ideal gas
mixture. In an ideal gas mixture molecules do not interact. The volume change, which is often
observed upon mixing, is due to the interactions between molecules. Hence, in an ideal gas mixture
there will be no volume change upon mixing.

In an ideal gas mixture, molecules do not interact. Hence, the internal energy associated with each
component is not changed due to the presence of other molecules. Hence, interactions, such as van
der Waals interactions (dipole-dipole, dipole-induced dipole, induced dipole-induced dipole
interactions), can be neglected. The only contributing factor to the internal energy is the energy
associated with a single molecule. Hence, the internal energy of an ideal gas mixture can be
expressed in terms of the internal energy of the individual components present in the mixture:

( )

()


(Note:

( (

)

This immediately leads to the conclusion that the change in the internal energy upon mixing ideal
gases to form an ideal gas mixture (A
mix
U
IGM
) equals zero:

()




The enthalpy is defined as:

Hence, the change in enthalpy upon mixing two ideal gases forming an ideal gas mixture:

( )







8

The entropy will change due to mixing. A mixing process at constant temperature and pressure
(p
total
) is equivalent to reducing the pressure of each component from the total pressure p
total
to its
partial pressure, p
i
, in the mixture. The change in entropy upon isothermal reduction of the pressure
for an ideal gas is given by
(

) (

)
Thus, the entropy of a component in an ideal gas mixture can be deduced from the entropy of the
pure component as an ideal gas at the same temperature and pressure of the system taking into
account the reduction of the pressure from p
system
to its partial pressure p
i
in the ideal gas mixture.

) (

)
Hence the change in entropy upon mixing ideal gases to yield an ideal gas mixture is given by


or in terms of the mole fractions in the ideal gas mixture



The mole fraction of each component is less than 1, and hence ln(x
i
) is less than zero. This means
that the entropy of the system increases upon mixing.

The Gibbs free energy is defined as:

Hence, the change in the Gibbs free energy upon mixing ideal gases to form an ideal gas mixture is
given by:


Hence, the Gibbs free energy of the ideal gas mixture will be lower than the Gibbs free energy of its
contributing pure components.

The consideration of an ideal gas mixture leads to the conclusion that the change in the internal
energy and the enthalpy equals zero since molecules do not interact, i.e. no energy associated with
the interaction. However, the change in the entropy and the Gibbs free energy is not equal to zero
despite the absence of any interaction between the molecules. This originates from an effective
reduction of the pressure of each component from the system pressure to its partial pressure in the
mixture (similarly, it can be shown that the Helmholtz free energy changes upon mixing, since it is
defined as the difference between the internal energy and T
.
S).


Example: Separation of an ideal gas mixture
Air at 298.15K, which at low pressure can be considered to be an ideal gas mixture, is to be separated
into its components oxygen and nitrogen in a continuous, reversible, isothermal and isobaric process.
Determine the amount of work required per mole of air fed to the process, and the amount of heat
released/required.

Mole balance


9

Energy balance


For air as an ideal gas mixture

( )

( )

( )
Hence, for the isothermal and adiabatic separation of air:



Entropy balance


For a reversible process entropy generation is zero

) (

))



Hence, for the isothermal and isobaric separation process:
(

))

))






10

I.3 Describing mixtures with interactions partial molar properties
Molecules can be polarized resulting in a charge distribution within the molecule (possibly leading to
dipole, quadrupole or multipoles) resulting an electro-static interaction between the molecules.
These interactions will be minimal, when the molecules are far from each other, e.g. in a gas/vapour
at low pressure, but will affect the energy of the system severely in dense systems, such as liquids or
even gases at high pressure. In systems containing pure compounds, this was taken into account via
the critical properties of the pure component involved in the equation of state.

In mixtures, different types of molecules are interacting with each other. The interaction between
the various molecules manifests itself in various forms. One of the forms is the molar volume of the
mixture. The molar volume of a mixture cannot be represented by a linear relationship between the
molar volume of the components making up the mixture (see Figure I.3.1).




Figure I.3.1: Molar volume of methanol-water mixture (in cm
3
per mol of mixture) at 293.15 K as a
function of mole fraction of water (dashed line represents the linear relationship
between the molar volume of the mixture and the molar volumes of pure methanol and
water; Data from Handbook of Physics and Chemistry, D-238, 67
th
ed. (R.C. Weast, M.J.
Astle, W.H. Beyer, Eds.), CRC Press, Boca Raton, 1986)

A small difference exists between the molar volume of the mixture and the molar volume of the
pure components. The difference originates from the structure of the liquid, which is a consequence
of the inter-molecular interactions. This small difference is better illustrated by looking at the change
in the molar volume upon mixing:


or in the case of the methanol-water mixture:


which shows a contraction in the volume upon mixing the two components together (see Fig. 3.2).
The highest contraction in the volume (1 cm
3
/mol of mixture) is obtained at a mole fraction of water
of 0.52. The change in volume upon mixing is thus not symmetric with respect to the mole fraction
of its components!

0
10
20
30
40
0 0.2 0.4 0.6 0.8 1
M
o
l
a
r

v
o
l
u
m
e

o
f

m
i
x
t
u
r
e
,

c
m
3
/
m
o
l

o
f

m
i
x
t
u
r
e
Mole fraction of water, x
w
11


Figure I.3.2: Change in the molar volume upon mixing pure methanol and pure water to form a
mixture as a function of the mole fraction at 293.15 K (solid curve represents fit to the
Redlich-Kister equation - see text; Data of mixture from Handbook of Physics and
Chemistry, D-238, 67
th
ed. (R.C. Weast, M.J. Astle, W.H. Beyer, Eds.), CRC Press, Boca
Raton, 1986)

The change in the molar volume as a function of the mole fraction of water can be represented by
the empirical Redlich-Kister equation:

in its general form


The change in the molar volume upon mixing water and methanol can be reasonably fitted with a 3
rd

order Redlich-Kister equation (i.e. i goes up to 3)

)
(with the coefficients a
0
= -4.0012 cm
3
/mol; a
1
= -0.3961 cm
3
/mol, a
2
= 0.3827 cm
3
/mol; a
3
= 1.2836
cm
3
/mol)

The molecular interaction between the different molecules in the mixture results in a change in the
molar volume upon mixing two pure components, as well as other thermodynamic properties (such
as A
mix
U, A
mix
H, A
mix
S, A
mix
G). This implies that the molar properties of the mixture cannot be
represented in terms of the molar properties of the pure components (V
i
, U
i
, S
i
,. H
i
, G
i
). Mixtures are
described in terms of the partial molar properties.

A partial molar property is the change in the property of the mixture upon the addition of a small
amount of compound i to the mixture keeping all other variables (T,p and the number of moles of all
other components in the mixture constant)

( ) (
(

in its general form


partial molar volume

( ) (
(


partial molar enthalpy

( ) (
(


-1.5
-1
-0.5
0
0 0.2 0.4 0.6 0.8 1
A
m
i
x
V
,

c
m
3
/
m
o
l

o
f

m
i
x
t
u
r
e
Mole fraction of water, x
w
12

The partial molar property is compound specific and is a function of temperature, pressure and the
molar composition of the mixture. It describes the change in the molar property of the mixture upon
addition of an infinite small number of moles of component i to the mixture.

The thermodynamic property of the mixture can now be expressed in terms of the partial molar
properties of the components in the mixture:

( )

in its general form



Molar volume of mixture

( )


Molar enthalpy of mixture

( )



The partial molar property can be obtained from the experimentally observed change in the
property upon mixing. For instance, the change in the molar volume upon mixing methanol and
water was modelled as:

)
The change in the molar volume upon mixing is defined as:

( )

( )

( )

( )

( ) (


The partial molar volume of water is defined as:

( ) (
(

( ) (
(

( )
(
(

))

(
(

(
(



Evaluating each of the three terms in the equation:
(
(

, since the number of methanol is kept constant


(

, since the molar volume of pure methanol is only a function of


temperature and pressure


(
(

, since the molar volume of pure water is only a function


temperature and pressure

, since (

)


(
(

))

(
(

) (


13

(

) (

) (
(

(
(

(
(

(
(

(
(

(
(

(
(



Substituting it all back:

( ) (

(
(



Knowing the fitting function f(x
w
), the partial molar volume of water can be found:
Realize:
(

( )

( )

)
((

( )





The partial molar volume can be determined without previous knowledge on the fit of the change in
the volume upon mixing as a function of the mole fraction in the mixture. For instance, the change in
the molar volume upon mixing methanol and water is defined as:

( )

( )

( )

( )

( )

( )

)

This equation has two variables which are a function of the mole fraction, viz.

, and
we need therefore another equation to obtain the partial molar volumes of the individual
compounds. This can be obtained by differentiating the change in molar volume upon mixing with
respect to the mole fraction of water (keeping the temperature and pressure constant):
14

(
(

())

) (

(
(

) (

(
(


For a binary mixture of methanol-water:

and (


The molar volume of a pure component is independent of the composition in the mixture, i.e.
(



Thus, the differential equation simplifies to:
(
(

())

(
(



The right hand side of the equation still contains two partial differentials. We can show that the
Gibbs-Duhem equation shows that



Hence, the two equations required to obtain the partial molar volume of methanol and the partial
molar volume of water as a function of the mole fraction of water are:

( )

)
and
(
(

())

) (

)

Now working out the two unknowns:
(

( )

(
(

())

( )

(
(

())



Now, the actual experimental data or the fitted Redlich-Kister equation can be used to find the
partial molar volume of methanol and water in a water-methanol mixture.

The so deduced partial molar volume of water in a water-methanol mixture shows a monotonic
increase with increasing mole fraction of water (see Fig. 3.3). The monotonic increase can be
interpreted as a steady decrease in the number of molecules of methanol surrounding a water
molecule with increasing mole fraction of water. The partial molar volume of methanol as a function
of the mole fraction of water shows a peculiar behavior with a minimum for the mole fraction of
water of ca. 0.62. This behavior can be explained in terms of a preferential surrounding of methanol
with water due to the formation of H-bridges.
15


Figure I.3.3: Partial molar volume of water (left axis) and methanol (right axis) in a water-methanol
mixture at 293.15 K as a function of the mole fraction (Data of mixture from Handbook
of Physics and Chemistry, D-238, 67
th
ed. (R.C. Weast, M.J. Astle, W.H. Beyer, Eds.), CRC
Press, Boca Raton, 1986)


The partial molar enthalpy can be easily determined from calorimetric measurements on the heat
release/removal upon the addition of a compound to another compound. From here the partial
molar specific heat can be determined. The specific heat of a mixture is defined as:


The partial molar heat capacity is thus defined as:


Hence, the determination of the partial molar enthalpy at a variety of temperatures will yield the
partial molar heat capacity.

How to determine the partial molar Gibbs free energy?
A useful relationship between the Gibbs free energy is given by considering the change in the Gibbs
free energy relative to temperature with respect to temperature:
(
(


Knowing that (

and
(
(


(
(


And thus also for the partial molar properties:
39
41
43
45
47
12
14
16
18
20
0 0.2 0.4 0.6 0.8 1
P
a
r
t
i
a
l

m
o
l
a
r

v
o
l
u
m
e

o
f

m
e
t
h
a
n
o
l
,

V
m
e
t
h
a
n
o
l
,

c
m
3
/
m
o
l
P
a
r
t
i
a
l

m
o
l
a
r

v
o
l
u
m
e

o
f

w
a
t
e
r
,

V
w
a
t
e
r
,

c
m
3
/
m
o
l
Mole fraction of water, x
w
16

(
(


The enthalpy is in principle a function of temperature



Integration between T=0 and T=T yields
(


At 0K the entropy goes to zero (3
rd
law of thermodynamics), and the Gibbs free energy at 0K equals
the enthalpy at 0 K:
(



More often however, the partial molar Gibbs free energy is determined from VLE-data using the
excess Gibbs free energy (vide verde).

The partial molar entropy can then be determined from:




17

I.4 Gibbs-Duhem equation
Any thermodynamic property of a mixture can be represented by:

( )


or in terms of extensive property u
mixture
:

( )

(3.1.1)
The thermodynamic property is a complete function of T,p and the number of moles of the various
components in the mixture. The complete differential is thus given by:

) (
(

(
(

(
(


Substituting the definition for the partial molar property:

) (
(

(
(

(3.1.2)

However, the complete differential is also equal to

(3.1.3)

Equating (3.1.3) and (3.1.2):
(
(

(
(



or
(
(

(
(


Thus at constant temperature and pressure:



This can be expressed in terms of the mole fraction (by dividing both left and right hand side of the
equation with the total number of moles in the mixture)



This can be transformed into a partial differential equation by dividing both sides of the equation
with the same differential (e.g. dx
A
)



Thus for the mixing process of methanol and water (with i = water, methanol,; and A is water), this
equation becomes:



A consequence of the Gibbs-Duhem equation is that for a multi-component system with C
components (and thus the variables T,p, N
1
, N
2
, , N
C
) there are only C+1 independent variables!


I.4.1 Partial molar Gibbs free energy and the Gibbs-Duhem equation
The Gibbs free energy is for chemical engineers an important property, since this function is at its
minimum at equilibrium for systems at constant temperature and pressure. The Gibbs free energy of
18

a multi-component mixture is a complete function of temperature, pressure and the number of
moles present of each species.
(

)
(


(the partial molar Gibbs free energy has classically been termed the chemical potential)

We can do the same with other thermodynamic functions. For instance, the enthalpy of a multi-
component mixture is a function of pressure, entropy, and the number of components
(

)
(



NOTE: (



The partial differential of the enthalpy with respect to the number of moles of i keeping the entropy,
pressure and the number of moles of all other components constant is most conveniently evaluated
using the definition of the Gibbs free energy


Hence,



Thus, the change in enthalpy is given by


Similarly



The partial molar Gibbs free energy is related to the fugacity of a species in the mixture. The fugacity
of a species in a mixture is a departure function of the Gibbs free energy for a species from its value
in an ideal gas mixture:

()

()


The partial molar Gibbs free energy of a compound in a mixture can be obtained from:

( ) (

(
(


19

( )

(
(


(
(

(
(

(
(

(
(

) (

)
For k=i

(
(

(
(

) (


For k=i

(
(

(
(

) (

( )

)

Thus, the fugacity of a species in the mixture is given by

()

()





20

I.5 Ideal mixture and excess mixture properties
An ideal mixture is defined as a mixture, in which the partial molar volume and enthalpy of each
component in the mixture equals the molar volume/enthalpy of the pure component for all
temperatures/pressures and compositions :

( )

( )

( )

( )

The fugacity of a species in an ideal mixture is given by:

()

()



The partial molar Gibbs free energy in an ideal mixture is given by:

( )

( ) (

)

Other functions of ideal mixtures (similar to ideal gas mixtures):

( )

( ) (

( )

( )


Excess property is defined as the property of a component relative to the property in an ideal
mixture:

( )

( )



or otherwise stated:


e.g.



The excess properties are the basis of the definition of activity coefficients of compounds in a
solution. The activity coefficient of a component (
i
) in a solution is defined in terms of the fugacity
of the pure component and the fugacity of a component in a mixture:

) (

) (

*
(

( )

( )
(

( ) (
(





Example: Is a binary mixture for which the molar Gibbs free energy is given by

( )

( )

an ideal mixture?
With a a constant independent of T,p,x.
21


Condition for ideal mixture:

( )

( )

( )

( )

( ) (


(hence the first condition is fulfilled)

( )

(
(

(
(

( )

( )

()

( )

( )



For an ideal mixture

( )

( )

; hence this mixture is not ideal.






22

I.6 Properties of activity coefficient

The activity coefficient of a particular component in a liquid mixture can be derived from the excess
molar Gibbs free energy of the mixture, since
(

( ) (
(




Example: Developing an expression of the activity coefficient
Develop an expression for the activity coefficients of component 1 and component 2 in a binary
mixture for which the molar Gibbs free energy is given by

( )

( )



The excess Gibbs free energy is defined as:

( )

( )

( )

( )

) (

(
(

) (
(

))


(
(

)
(

) (

) (

) (

(
(

) (
(

))

) (



The activity coefficient is defined as:


The fugacity of a component in the mixture is given by:







23

Thermodynamic consistency of activity coefficients
Both the real mixture and the ideal mixture satisfy the Gibbs-Duhem equation:
(
(

(
(



Subtracting these two equations yields:
(

) (



Thus at constant temperature and pressure

((

))



Thus, for a binary mixture:

((

))

((

))
Hence, for a variation with respect to the mole fraction of component 1 in an isothermal and isobaric
system:

(
(

(
(




Temperature dependency of the activity coefficient
The activity coefficient is defined as: (



and the temperature dependency is given by:
(
(

*)



For small temperature variations, we may assume that the partial molar enthalpy does not vary
significantly, and hence

)



Pressure dependency of the activity coefficient
The activity coefficient is defined as: (



and the pressure dependency is given by:
(
(

*)



For liquids, the pressure dependency on the excess volume is small (liquids can be considered to be
incompressible fluids), and hence

)


24

I.7 Fugacities in mixtures

The fugacity of a compound in a mixture is defined as:

()

()


or

()

()



Equations of state are typically functions of the volume of the system rather than the pressure of the
system. Hence, we are changing the integration variable from pressure to volume. The integral can
be split into two parts:

)
The first integral is over the partial molar volume of compound i. From the triple product rule:
(


or



The second integral can be evaluated realizing that:
( )

( )


Furthermore and at constant temperature ( )
Hence,


(Note that the integration boundaries change with the change in integration variable. Each time we
integrate from the ideal gas state - Z=1, infinite large volume, zero pressure to real gas state)

Realizing that



Substituting the two evaluated integrals back into the expression for the fugacity of the mixture
yields:

( (

( (

) ()

Hence, knowing the equation of state for the mixture (in the appropriate state of aggregation)
should yield the fugacity of the specific component in a mixture.




25

Gaseous mixtures
Some gas mixtures follow Amagats law, i.e.

( )

(y
i
is the mole
fraction of component i in the gas phase). This also means that the excess volume is zero. A
consequence of this is that partial molar volume of a component i in this mixture equals molar
volume of the pure component. Hence, the expression for the fugacity of a component i in the
mixture becomes

)

and the fugacity of the species in the mixture is given by the fugacity of the pure component times
its mole fraction

Lewis-Randall rule
This expression is valid as long as Amagats law is valid. This is typically for gas mixtures at low
pressures and at extreme high pressures.

For intermediate pressures an equation of state has to be used. For instance, the truncated form of
the virial equation (neglecting the higher order terms):

()

with

( )



()

Thus

()


()


()

)


(


()

(
(


())



This can now be further developed into an expression for the fugacity of a species of a compound i
according to the truncated form of the virial equation of state to:
(



()

( (

) ()

( (



())

) ()



() ()



() ()
(with Z expressed as a function of pressure

) )

Other equations of state (such as the Peng-Robinson equation of state) can be used as well. The
Peng-Robinson equation of state is given as:

()
()()

The parameters a and b now refer to the parameters for the mixture. Mixing rules have been
identified based on statistical mechanics:

with

)
(here a
ii
represents the pure component parameter a)
26

with b
i
the pure component parameter b
These mixing rules introduce a new parameter, k
ij
, the binary interaction parameter. This parameter
is obtained by fitting experimental data to Peng-Robinson equation of state. Tabe 9.4-1 (p. 424)
gives a selection of binary interaction parameters for binary hydrocarbon mixtures. Others are given
in specialist journals (e.g. H. Nishiumi, T. Arai, K. Takeuchi, Fluid Phase Equilibria 42 (1988), 43-62).
This approach only takes into consideration binary interactions. Mixtures containing more than two
components can be built up by considering the various binary pairs present in the mixture (D.S.H.
Wong, S.I. Sandler, AIChE J. 38 (1992), 671-680).

The fugacity obtained from the Peng-Robinson equation of state is now given by:

( ) ( )

) (
()

()

*
with


The compressibility is obtained by solving the cubic equation of state:

) (

()

()





Example: Fugacity of ethane in an ethane-butane mixture
Calculate the fugacity of ethane (Et) in an ethane (Et) n-butane (Bu) mixture at 373.15 K and 1 bar,
10 bar, 50 bar, 100 bar, 500 bar as a function of the mole fraction of ethane in the mixture using the
Lewis-Randall rule, the truncated virial equation of state, and the Peng-Robinson equation of state.

Data for the virial equation of state:
B
Et-Et
= -1.15
.
10
-4
m
3
/mol; B
Et-Bu
= -2.15
.
10
-4
m
3
/mol; ; B
Bu-Bu
= -4.22
.
10
-4
m
3
/mol

Lewis-Randall rule:



The fugacity of ethane at 373.15 K can be calculated using the Sandler program Peng-Robinson
equation of state. The calculated fugacity of pure ethane at 373.15K is
p, bar 1 10 50 100 500
f
Et
, bar 0.99571 9.5784 40.2516 65.1960 208.8461
|
Et
=f
Et
/p 0.99571 0.95784 0.80503 0.65196 0.41769

Truncated virial equation of state



() ()

()

()

( )



()

( )

( )



27

It should be noted that this equation can only be used for pressures up to ca. 15 bar; higher
pressures will result in

and hence imaginary compressibilities. This is an indication


that at the higher pressures the higher order terms in the virial equation of state need to be taken
into account.

Peng-Robinson equation of state:

( ) ( )

) (
()

()

*
Hence, the fugacity of ethane in the mixture can be obtained using Sandlers program Peng-
Robinson equation of state for mixtures. The binary interaction parameter for the system ethane
n-butane is 0.010 (see p. 424)

Figure I.7.1 shows the fugacity coefficient for ethane in an ethane n-butane mixture as a function
of the mole fraction of ethane in the mixture. At low pressure the fugacity coefficient does not
deviate significantly from 1, and the choice of the model to describe the fugacity of ethane in the
mixture is immaterial. Even at 10 bar, the fugacity coefficient is still close to 1, and any one of the
models can be used. At pressures in the range between 10 and 100 bar, significant deviations
between the prediction using Lewis-Randall rule and the prediction from the Peng-Robinson
equation of state are observed. However, at very high pressures the predicted values for the fugacity
coefficient of ethane become similar. The question is always, which model is better. We cannot
decide on the quality of the models without comparison to experiment. For this the experimental
compressibility of the mixture at various compositions must be measured and compared with the
model prediction (see Fig. 9.4-1).

Figure I. 7.1: Fugacity coefficient of ethane in an ethane/n-butane mixture as predicted using the
Lewis-Randall relationship (dashed lines), the truncated virial equation of state (dotted
curves), and the Peng-Robinson equation of state (solid curves).



Liquid mixtures
The fugacity of compounds in a liquid mixture can be obtained from an appropriate equation of
state (e.g. Peng-Robinson equation of state), if the binary interaction parameter is known:

( ) ( )

) (
()

()

*
with Z the compressibility of the liquids phase (i.e. the smallest root of the cubic equation).

For mixtures in which one or more components cannot be described by an equation of state, the
fugacity of a component is described in terms of the activity coefficient.


0.95
0.96
0.97
0.98
0.99
1
1.01
1.02
0 0.2 0.4 0.6 0.8 1
F
u
g
a
c
i
t
y

c
o
e
f
f
i
c
i
e
n
t
,

|
E
t
=
f
E
t
/
(
y
E
t
.
P
)
Mol fraction ethane, y
Et
1 bar
10 bar
0.4
0.7
1
1.3
1.6
0 0.2 0.4 0.6 0.8 1
F
u
g
a
c
i
t
y

c
o
e
f
f
i
c
i
e
n
t
,

|
E
t
=
f
E
t
/
(
y
E
t
.
P
)
Mol fraction ethane, y
Et
50 bar
100 bar
500 bar
28

This then requires an appropriate model for the activity coefficient, or better for the excess Gibbs
free energy.

Solid mixtures
In most cases, solids do not form solid mixtures, but separate domains with pure crystalline phases.
Hence, a solid containing more than one component can be viewed as a collection of solid phases,
each containing a pure compound. Hence, the fugacity of the solid component in this solid phase is
the fugacity of the pure solid:



It should however be noted that this is not true for all solids. Some solids can form mixtures, e.g. in
alloys. In that case, the fugacity of a compound in this mixture (alloy) can be formulated as:


This then requires an appropriate model for the excess Gibbs free energy to obtain the activity
coefficient for the particular component in the mixture.

29

I.8 Activity coefficient models

I.8.1 Random mixtures
The excess Gibbs free energy for a binary mixture can be modeled using the Redlich-Kister equation:


The excess Gibbs free energy is zero (and thus we are then dealing with an ideal solution), when a
i
is
equal to 0 for all values of i.
The use of the Redlich-Kister equation has as an underlying assumption that the microscopic
structure of the mixture is identical to the macroscopic structure. This means that the mole fraction
for compound j is a measure for the likelihood (probability) that the molecule i is surrounded by
molecules j.

A simple mixture might be modeled with a
0
not equal to zero, but all other constants a
i
equal to zero
for I larger than 0:



The activity coefficient for component 1 is now given by:
(

(
(


and for component 2
(

(
(


These equations are called the one-constant Margules equations.

A more complex mixture might be modeled with a
0
and a
1
unequal zero and a
i
equal to zero for i
larger than 1:

))

) (

))

(

(
(

)))

) (
(

)))


(
(

)))

))

))

) (

))

) (

(
(

)))

) (






30

Van Laar equation
The van Laar equation has its origins in the van der Waals equation and takes into account the
volume occupied by each molecule. It is assumed that binary mixtures are composed of species of
similar size and similar energies of interaction, and that the van der Waals equation of state is
applicable to both the mixture and the pure fluids. The implicit assumption here is that the
molecules of each species are uniform distributed throughout the mixture (i.e. random mixture).

Since the molecules are of similar size, it can be assumed that

or


and since the molecules are randomly mixed:

)) or



Thus, the molar excess Gibbs free energy is given by:



In order to evaluate the excess molar internal energy, we start with a cycle:
Step 1: Start with two pure liquids, x
1
moles of liquid 1 and x
2
moles of liquid 2, and decrease the
pressure to evaporate each liquid to an ideal gas
Step 2: Mix the two ideal gases to form an ideal gas mixture
Step 3: Compress the ideal gas mixture to form a liquid at the starting pressure P

The change in the internal energy associated with step 1:


( (

( (



The change in internal energy associated with step 2



since an ideal gas mixture is formed!

The change in internal energy associated with step 3 can be obtained in the same manner as for step
1:




Hence, the excess molar Gibbs free energy is then given by:



In liquids, molecules are closely packed and the molar volume is (approximately) the volume taken in
by the molecule, which is represented by the factor b in the van der Waals equation of state:


Based on statistical mechanics:

with k
12
=0



31

The activity coefficient is defined as:
(

(
(

) (
(

(
(

) (



In its general form the van Laar equation looks like:
(


The parameters, a and b, can in principal be estimated from the van der Waals equation of state.
However, few fluids can be accurately predicted using this equation of state, and hence these
parameters are often estimated empirically. The coefficients o and | can be obtained experimentally
from a measured activity coefficient for each component at a particular composition
(

)
)

) (

)
)

)


Regular solution theory
Not many liquids and liquid mixtures obey the van der Waals equation of state. The basic
assumptions of the van Laar model (i.e. V
ex
and S
ex
are zero) is however a good approximation for
many mixtures. It was suggested to use the experimental change in the internal energy upon
evaporation rather than an equation of state (G. Scatchard, Chem. Rev. 8 (1931), 321). Hence, the
molar excess Gibbs free energy is given by:


(the change in the internal energy upon evaporation can easily be obtained from the enthalpy
change upon evaporation, since (for the evaporation yielding a gas at low
pressure; note that the enthalpy upon evaporation at the temperature of interest must be used!).
The change of the internal energy of the system upon evaporation can be approximated by:



32

* (

* (



Defining the volume fraction of component 1 and 2 in the solution:



and Hildebrandt solubility parameter is defined as




The activity coefficient is defined as:
(

(
((

) (

(
((

) (

((

))

)
(


(
((

))


(
(

(
(


(
(

(
(
(

)(

)
(

(
(

)
(



(
(

)
(
(

) (

))
33

(

) (



Similarly for compound 2:
(

) (





Extension of regular solution theory to multi-component mixtures:
(

) (


with



(i.e. use a volume average solubility parameter with





Note: The heat of evaporation must be taken at the temperature at which the activity coefficient
needs to be determined (which is usually not equal to the normal boiling temperature)

()





I.8.2 Non-random mixtures
The local composition in a mixture is not necessarily the same as the macroscopic composition. For
instance, a molecule might be highly solvated, i.e. preferentially surrounded by the solvent
molecules, in which case the dissolved molecules will only experience the interaction with the
solvent molecules (up to some limiting concentration). In general, non-random mixtures can be
formed due to the energetics or interaction and/or difference in the size of the molecules involved.

Wilson equation
The Wilson equation is a two-parameter model, which is based on the following expression for the
excess molar Gibbs free energy:

)
From which the activity coefficient is
(

) (

)
(

) (

)
This model can be fitted to experimental data to obtain the fitting parameters A
12
and A
21
.

NRTL model
The non-random two liquid (NRTL) model is based on a slightly different formulation of the molar
excess Gibbs free energy, which involves three parameters:

) with

and



This model requires 3 parameters to be fitted to experimental data

Flory-Huggins model
Mixtures of polymers and their solvents (e.g. n-hexane) have molecules with vastly different sizes.
Due to the size difference, the entropy change upon mixing is not approximately zero. The change in
the entropy upon mixing is controlled rather by the volume fraction of the various species in the
solution rather than the mole fraction:

)) with


34

with


(the parameter m is here the volume ratio of species 2 relative to species 1).

Hence, the excess molar entropy is given by:

))
The enthalpy of mixing can be expressed by a one-constant term in terms of the volume fraction of
the species in solution:



Thus, the molar excess Gibbs free energy is given by:

))

The activity coefficient is defined as:
(

(
((

))+


(
((

(
((


(
((

) (
(


(
(

(see regular solution theory)


(
(



(
((

)
(

)
(
((

)

(
((




(
((

))+

(
(

))


(
(

))

(
(


(
(

(
(

(
(

()


(
(

))

()



35

(
(

))

(
(



(
(

(
(

(
(


(
(

))



(
((

))+

()


(
((

))+

()

()


(
((

))+

()

)
(
((

))+

)


(

)
(with the first term describing the contribution of the excess enthalpy and the latter two terms
describing the contribution of the excess molar entropy to the activity coefficient)

Similarly,
(

()


UNIQAC model
The universal quasi-chemical (UNIQAC) model is based on statistical mechanics and allows a
different local composition based on differences of size and energy between the molecules in the
mixtures. The basic idea centers on the additive properties of energy. The contribution of the size
difference and the difference in energetics to the molar excess Gibbs free energy is separated into a
combinatorial term (mainly for size differences) and a residual term (mainly for the energy
differences):

()

()



The combinatorial part of the excess Gibbs free energy is expressed in terms of volume fraction:

()


36

The first term in this expression is familiar and originates from the entropic contribution to the Gibbs
free energy (see Flory-Huggins model). The second term expresses the direct neighbor-neighbor
interaction in the mixture.

The volume fraction in a molecule can be thought to originate from the various segments, which
each contribute to the size of a molecule (Group Contribution Method). In a mixture, each species
can be thought to be made up from various segments. The volume fraction of a species is then given
by:



(the contributing volume parameters r
i
are given in Table 9.5-2)


Example: Using the group contribution method to estimate the volume fraction
Determine the volume fraction of benzene in a mixture of benzene and 2,2,4 trimethylpentane as a
function of the mole fraction of benzene in the mixture.

The volume parameter can be computed by breaking each molecule into its constituents. Benzene is
made up by 6 aromatic CH-units (ACH) each contributing 0.5313. Hence, the volume parameter for
benzene is



2,2,4 trimethylpentane constitutes of 5 CH
3
-groups, 1 CH
2
-group, 1 CH-group and 1 C-group. Hence,
the volume parameter for 2,2,4 trimethylpentane is



Thus, the volume fraction of benzene in the mixture is given by:





The second term contributing to the combinatorial excess Gibbs free energy is determined by the
number of neighboring molecules, z (a closed packing can have between 8 and 12 nearest species; z
is typically taken to be 10) and the surface area parameter for a species. The term u
i
is the area
fraction of species i and is defined as:



The surface area parameter is also thought to be made up from the contributing factors of the
segments within a molecule.


Example: Using the group contribution method to estimate the surface area fraction
Determine the surface area fraction of benzene in a mixture of benzene and 2,2,4 trimethylpentane
as a function of the mole fraction of benzene in the mixture.

The surface area parameter can be computed by breaking each molecule into its constituents.
Benzene is made up by 6 aromatic CH-units (ACH) each contributing 0.5313. Hence, the surface area
parameter for benzene is


2,2,4 trimethylpentane constitutes of 5 CH
3
-groups, 1 CH
2
-group, 1 CH-group and 1 C-group. Hence,
the surface area parameter for 2,2,4 trimethylpentane is
37



Thus, the surface area fraction of benzene in the mixture is given by:




Thus the combinatorial contribution to the excess Gibbs free energy for the system benzene (B)-
2,2,4 trimethylpentane (TMP) is given by:

()

()

)

The residual contribution to the excess Gibbs free energy is a function of the surface area
contribution and the interaction energy between the various species (u
ij
) in the mixture:

()

with




A consequence of the splitting of the excess Gibbs free energy into two terms is that the activity
coefficient can also be viewed as having two contributions:

()

()

() (




with

) (

()

( (

*

The UNIQAC model is a (very advanced), empirical model, since the parameters t
ij
still need to be
determined from experiment.

UNIFAC model
The UNIFAC (UNIqac Functional group Activity Coefficient) model is based on the UNIQAC model.
The combinatorial contribution to the activity coefficient for the UNIFAC model is slightly different
from the UNIQAC model

() (

( (

)
with

*
The residual term is in terms of the interaction parameter (interaction energy), which is now also
evaluated in terms of a group contribution method.

COSMO-RS model
The COSMO-RS (Conductor-like Screening Model for Real Solvents-) model uses quantum-chemical
models to predict the electro-static interaction energy, hydrogen bonding and van der Waals
interactions between components in the mixture (see e.g. A. Klamt, F. Eckert, Fluid Phase Equil. 172
(2000), 43). In principle COSMO-RS is analogous to UNIFAC, in that in both models an ensemble of
pair-wise interacting surface segments is considered. The description of the interaction energy
(important for the residual contribution to the activity coefficient) is different. COSMO-RS uses the
38

actual determined screening charge density as a measure for the interaction energy, whereas in
UNIFAC the interaction of between groups is determined as an average value.


39

I.9 Fugacity of species in a non-simple mixture

The fugacity of a species in a mixture is defined in terms of the fugacity of the same species in the
same phase as a pure compound at the same temperature and pressure of the mixture.
Gas phase

( )

( ) Lewis-Randall rule
Liquid phase

( )

( )

( )

Vapour phase fugacity of pure component
A compound in a mixture at a specific temperature and pressure may be normally in the liquid
phase. For instance, water at 298.15K and 1 bar is in its liquid form. The fugacity of water as a
vapour can be obtained from

( ) (

) with (


At low pressure the fugacity of the pure component at the temperature and pressure of the mixture
can be approximated by:

( ) with p the pressure of the system



Liquid phase fugacity of pure component

Dissolution of a solid in a liquid
The normal state of aggregation of the pure compound may be the solid state. This is commonly
known as the dissolution process.
The ratio of the fugacity of a component as a liquid relative to the fugacity of a component as a solid
is given by:
(

()

()
*

()

()

()


Hence, the fugacity of the component as a liquid can be estimated knowing the fugacity of the pure
component as a solid and the change in the Gibbs free energy of fusion at the temperature and
pressure of the mixture. The fusion and melting process are hardly dependent on pressure, and only
the temperature change will be considered. At its normal fusion/melting point

) . A
three step process is used to determine the change in the Gibbs free energy upon fusion at the
temperature and pressure of the mixture.
1. Heating the pure component from the temperature of the mixture to the melting
temperature
2. Melting the pure component at its melting temperature
3. Cooling the liquid formed down to the temperature of the mixture

Enthalpy change for this three step process:



Defining the difference between the specific heat of the pure compound as a liquid and the specific
heat as a pure component as a solid as AC
p
:



At the normal melting temperature

)
or


40

Substituting it all in

) (



For many substances, the specific heat of the solid is approximately the same as the specific heat of
the liquid:

) (

)

And the fugacity coefficient of the pure component as a liquid is thus given by:

*



Dissolving a vapour in a liquid
The pure compound may exist at the temperature and pressure of the mixture as a vapour. The
fugacity of the pure component as a liquid at this temperature and pressure, if the pressure is not
too high, is then represented by the vapour pressure of the pure compound even if the pressure of
the system is lower than the vapour pressure of the pure compound:

) can be higher than p!




Dissolving a gas in a liquid
Gases can be sparingly soluble in liquids (e.g. nitrogen is to some extent soluble in water). The mole
fraction of the dissolved gas is then typically low. It has been experimentally observed that the
fugacity of the dissolved gas in the mixture depends linearly on the mole fraction of that component
in the mixture:

)
The proportionality factor is called the Henry coefficient (with the units bar/mole-fraction). The
Henry coefficient can be interpreted as the fugacity of the pure hypothetical liquid.
At higher mole fractions of the dissolved gas, a deviation from the linear behavior is typically
observed, which can be accounted for by the introduction of a new (and differently defined!) activity
coefficient:

)
The difference becomes clear by examining the limiting cases, i.e. this new activity coefficient
i
*
goes to 1 for x
i
0, whereas the usual activity coefficient
i
goes to 1 for x
i
1. They are however
linked since

)
or

)

The Henry coefficient can be found by taking x
i
0, for which
i
* becomes 1:

)
Thus,

)


This approach can also be used to describe the fugacity of other sparingly soluble species in solution
such as proteins.

41

I.10 Electrolyte solutions

Solutions containing ions pose another problem, since the fugacity of the ion as a pure liquid does
not exist. All solution containing ions are electrically neutral, i.e. the negative charges balance out
with the positive charges. For instance, in a dilute aqueous solution sulphuric acid dissociates
completely:
H
2
SO
4
2 H
+
+ SO
4
2-

The solution is electro-neutral, i.e. the charges balance:


At high concentration the dissociation of sulphuric acid, the equilibrium between the protons and
sulphate ions on the one side and the hydrogen sulphate ion on the other side becomes noticeable,
i.e.
H
2
SO
4
HSO
4
-
+ H
+
2 H
+
+ SO
4
2-

The charge balance is given now by:


(and the pH of the solution can be calculated knowing the concentration of sulphuric acid in
solution, and the equilibrium constant for the dissociation of the hydrogen sulphate ion). In general,
the charge balance of a solution containing cations with a charge z
+
and anions with a charge z
-
is
given by:



The implication of the charge balance is that the anion and cation always go together. The Gibbs free
energy of the mixture is in principle given in terms of the partial molar Gibbs free energy of its
constituents (in an ionic solution, the solvent, S, the undissociated ionic compound, A
v+
B
v-
, the
dissociated ions A
z+
and B
z-
):



Since we more mostly interested in electrolyte solutions with a low concentration, the activity
coefficient of the ionic compounds and the ions can in principle be defined as:

( ) (

)
Where M
i
is the molality of the solution (units mole of solute per kg of solvent),

( ) is the
partial molar Gibbs free energy in an 1 molal ideal solution, and

is the activity coefficient, which


approaches 1 when the molality goes to zero. According to this definition, the partial molar Gibbs
free energy of the undissociated ionic compound and the ions is given by

( ) (

( ) (

( ) (

*

However, the partial molar Gibbs free energy of the cation cannot be determined since the change
in Gibbs free energy of the mixture upon changing the number of moles of the cation keeping the
number of moles of the anion constant is physical nonsense, violating the principle of the electro-
neutrality balance. Hence, the Gibbs free energy of the mixture is described in terms of the amount
of dissociated ionic compound, N
AB,D
:



Thus,



42

Substituting the definition of the partial molar Gibbs free energy for the ions in terms of the activity
coefficient:

(
(

)

Defining
Mean ionic activity coefficient

)
(


Mean ionic molality

)
(


Partial molar Gibbs free energy of the dissociated compound as an ideal 1 molal solution


Thus,

(
(

)

The mean ionic coefficient for dilute solutions can be modeled using the Debye-Hckel limiting law,
which is based on an electro-static model of ions in a solvent:
(

) |

|
Here I is the ionic strength of a solution containing cations with a charge of z
+
and anions with a
charge of z
-
. The ionic strength is defined as:




The parameter a is solvent dependent and is for water given in Table 9.10-1. It should be noted that
the mean ionic activity coefficient is less than 1 and decreases with increasing ionic strength.

The Debye-Huckel limiting law is exact for low concentrations, but significant deviations are
observed at concentrations higher than ca. 0.01 molal. Empirical corrections have been proposed to
the limiting Debye-Huckel law:
(

)
|


with a the average radius of the hydration sphere of the ions (typically taken as 4) and the
parameter | for aqueous solutions given in Table 9.10-1.

For more concentrated solutions (ca. 0.5 molal and higher) a further empirical correction has been
proposed to account for the increase in the mean ionic activity coefficient at high ionic strength
(

)
|

with |

|
It should however be noted that nowadays robust models for the description of electrolyte solutions
at high ionic strength are available (K.S. Pitzer, J. Phys. Chem. 77 (1973), 268).






43

II Phase equilibria

II.1. Introduction
Chemical processes produce typically mixtures, which have to be separated into streams of the
required purity in order to obtain a useful product. Phase equilibria play an important role in these
separation processes, which use transfer of a particular component into another phase as a method
to separate mixtures. This often occurs in multiple steps (stages). Knowledge on the equilibrium
position between the phases is a pre-requisite for the design of these processes.

In a phase equilibrium situation, each of the compounds in a mixture can be distributed over two (or
more) phases, phase I and phase II. If the system is a closed, adiabatic, constant volume system:


i.e. the decrease of number of moles of component i in phase I corresponds to an increase in the
number of moles of component i in phase II



In a closed, adiabatic, constant volume system, the entropy strives towards a maximum. This means
that the change in the entropy is equal to zero



Since, the differential of the entropy must be equal to zero with respect to all independent variables
(i.e. dU
I
, dV
I
and dN
i
I
), we obtain 3 equilibrium conditions, viz.

temperature in each phase equal


pressure in each phase equal


partial molar Gibbs free energy of each component in each phase equal


The same criteria can be derived for a system at constant temperature and pressure. The Gibbs free
energy in such a system is at a minimum:

(dT and dp are equal to zero)


temperature in each phase equal (pre-condition of system)


pressure in each phase equal (pre-condition of system)


partial molar Gibbs free energy of each component in each phase equal


The equilibrium condition can also be expressed in terms of the fugacity of the species in the
mixture. The partial molar Gibbs free energy for each species must be equal between the phases,
and hence

partial molar Gibbs free energy of each component in each phase equal

44

II.1.1 How many variables are needed to describe a multi-component mixture containing
multiple phases? Gibbs phase rule

A single phase system containing C components, has requires C+1 variables to describe it. The last
variable is fixed due to the constraint of the Gibbs-Duhem equation. When the system contains P
phases, would require P(C+1) number of variables (a full set of equations for each phase). However,
at equilibrium
- Temperature in each phase is the same instead of P temperatures we have only one
temperature leading to a reduction in number of variables with P-1
- Pressure in each phase is the same instead of P pressures we have only one pressure
leading to a reduction in number of variables with P-1
- Partial molar Gibbs free energy for each component in each phase is the same instead of
CP partial molar Gibbs free energies we have only one partial molar Gibbs energy for each
component leading to a reduction in number of variables with C(P-1)

Hence the total number of variables needed (or degree of freedom in the system, F) to describe a
system at (phase) equilibrium:
( ) ( ) ( ) ( )


For instance, 1 variable is required to specify a system containing 1 component and 2 phases
present. The transition of a liquid to a vapour for a pure component will occur at a certain
temperature at a given pressure (the temperature cannot be freely chosen, when pressure has been
chosen). Two variables are required to specify a system containing 2 components and 2 phases
present. The description of the transition of a liquid to a vapour for a binary mixture requires two
parameters to be specified (e.g. temperature/pressure or temperature/composition).


45

II.2. Vapour-liquid equilibrium (VLE) of ideal binary mixtures at low pressure
The conditions for a vapour-liquid equilibrium of a binary mixture are in addition to the:


and the temperature and pressure in each phase are equal.

At low pressure, vapours obey the Lewis-Randall rule. Hence the fugacity of a component in the
vapour phase mixture is given by:



Furthermore, at low pressure the fugacity of the pure component is equal to the system pressure:



The fugacity of a component in the liquid phase is given by:



The activity coefficient of in an ideal liquid mixture is equal to 1.



The fugacity of the pure, liquid component at low pressure is equal to the vapour pressure of the
component:



The equilibrium conditions for an ideal binary mixture at vapour-liquid equilibrium are thus:


or



Thus, for a given mole fraction in the liquid phase in a mixture at a given temperature, the pressure
and the composition of the gas phase can be calculated. It should be noted that the pressure of the
system depends linearly on the mole fraction of compound 1 in the liquid.


Example Low pressure p-x-y VLE-diagrams for ideal mixtures
The vapour pressure of n-hexane (Hx) at 60
o
C is 0.7536 bar and the vapour pressure of n-heptane
(Hp) at 60
o
C is 0.2766 bar. Calculate the equilibrium pressure for liquid mixtures of n-hexane (H) and
n-heptane (Hp) at 60
o
C and the corresponding gas phase composition assuming that n-hexane (H)
and n-heptane (Hp) form an ideal mixture.

From the equilibrium condition

for an ideal mixture it is derived that



46


Figure II.2.1: x-y diagram (left) and P-x-y diagram (right) for the system n-hexane/n-heptane at 60
o
C
at vapour liquid equilibrium

In chemical processes, we often keep the pressure (approximately) constant and not the
temperature. Hence, we are often interested in the position of the vapour-liquid equilibrium (VLE) at
a constant pressure. The vapour pressure of the compounds is dependent on temperature, and is
typically given by the Antoine equation. Hence, the temperature of the mixture in a vapour-liquid
equilibrium at a given pressure must be solved from:

Antoine equation

with p the system pressure






Example Low pressure T-x-y VLE-diagrams for ideal mixtures
The vapour pressure of n-hexane (Hx) is given by

( )

()
and the
vapour pressure of n-heptane is given by

( )

()
. Calculate the
equilibrium temperature for liquid mixtures of n-hexane (Hx) and n-heptane (Hp) at a pressure of 0.5
bar and the corresponding gas phase composition assuming that n-hexane (Hx) and n-heptane (Hp)
form an ideal mixture.

The procedure is the same as for the calculations of the P-x-y diagram, although the calculations now
require solving a non-linear equation by iteration. From the equilibrium condition

for an
ideal mixture it is derived that


( )

()
(

()

()
(

()


and

()



0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
M
o
l
e

f
r
a
c
t
i
o
n

n
-
h
e
x
a
n
e

i
n

v
a
p
o
u
r

p
h
a
e
,

y
H
x
Mole fraction of n-hexane in liquid phase, x
Hx
T = 60
o
C
0.2
0.4
0.6
0.8
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e
,

b
a
r
Mole fraction of n-hexane
Pressure as a function
of mole fraction n-hexane
in liquid phase, x
Hx
Pressure as a function
of mole fraction n-hexane
in vapour phase, x
Hx
T = 60
o
C
Liquid mixture
vapour mixture
47


Figure II.2.2: x-y diagram (left) and T-x-y diagram (right) for the system n-hexane/n-heptane at 0.5
bar at vapour-liquid equilibrium (VLE)


Bubble point calculation
The bubble point is defined as the pressure (in the case of bubble point pressure) or temperature (in
the case of bubble point temperature), at which in a particular liquid mixture with a given
composition the first vapour bubble is formed upon decreasing the pressure/increasing the
temperature (see Figure 2.3). This means that the composition of the liquid phase is known. A
minute amount of the liquid has evaporated, and thus the composition of the liquid phase is the
same as the composition of the starting liquid mixture. At this point a large amount of liquid is in
equilibrium with a very small amount of vapour


Example Calculating bubble point pressure/temperature for ideal mixtures
Determine the bubble point pressure for a mixture containing 50 mole-% n-hexane and 50 mole-% n-
heptane at 60
o
C, and the bubble point temperature for the same mixture at a constant pressure of
0.5 bar.
Vapour pressure of n-hexane (Hx):

( )

()

Vapour pressure of n-heptane:

( )

()
.

The bubble point pressure is given by the equilibrium relationship. Operating at a constant
temperature of 60
o
C, the


The bubble point temperature is given by the same equilibrium relationship. Operating at a constant
pressure of 0.5 bar, the bubble temperature is given by
( )

()
(

()

()

()


(quite similar answers as could be expected after calculating the bubble pressure at a temperature
of 60
o
C)
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1
M
o
l
e

f
r
a
c
t
i
o
n

n
-
h
e
x
a
n
e

i
n

v
a
p
o
u
r

p
h
a
e
,

y
H
x
Mole fraction of n-hexane in liquid phase, x
Hx
p = 0.5 bar
40
50
60
70
80
0 0.2 0.4 0.6 0.8 1
T
e
m
p
e
r
a
t
u
r
e
,

o
C
Mole fraction of n-hexane
Temperature as a function
of mole fraction n-hexane
in liquid phase, x
Hx
Temperature as a function
of mole fraction n-hexane
in vapour phase, x
Hx
p = 0.5 bar
Liquid mixture
vapour mixture
48


Figure II.2.3: Bubble point pressure at 60
o
C (left) and bubble point temperature at 0.5 bar (right) for
the ideal mixture n-hexane/n-heptane containing 50 mole-% n-hexane


Dew point calculation
The dew point is defined as the pressure (in the case of dew point pressure) or temperature (in the
case of dew point temperature), at which in a particular vapour mixture with a given composition
the first liquid droplet is formed upon increasing the pressure/decreasing the temperature (see
Figure 2.4). This means that the composition of the vapour phase is known. A minute amount of the
vapour has condensed out, and thus the composition of the vapour phase is still the same as the
composition of the starting vapour mixture. At this point a large amount of vapour is in equilibrium
with a very small amount of liquid.


Example Calculating dew point pressure/temperature for ideal mixtures
Determine the dew point pressure for a mixture containing 50 mole-% n-hexane and 50 mole-% n-
heptane at 60
o
C, and the dew point temperature for the same mixture at a constant pressure of 0.5
bar.
Vapour pressure of n-hexane (Hx):

( )

()

Vapour pressure of n-heptane:

( )

()
.

The starting point for the dew point pressure/temperature calculation is knowing that the mole
fraction in the vapour phase is known, but the mole fraction in the liquid phase is unknown. From
the equilibrium relationship, it is deduced that

and


or

and



The sum of the mole fractions of n-heptane and n-hexane equal 1, thus
or


and


Thus, the dew point pressure is given for a mixture containing 50 mole-% n-hexane at 60
o
C by



The dew point temperature for the mixture containing 50 mole-% n-hexane at 0.5 bar can be
estimated from
0.2
0.4
0.6
0.8
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e
,

b
a
r
Mole fraction of n-hexane
T = 60
o
C
Liquid
vapour
p
bubble
y
vapour
40
50
60
70
80
0 0.2 0.4 0.6 0.8 1
T
e
m
p
e
r
a
t
u
r
e
,

o
C
Mole fraction of n-hexane
p = 0.5 bar
Liquid
vapour
T
bubble
y
vapour
49

()

()




The corresponding composition of the liquid phase can be estimated by back substitution into



(Note the large difference between the bubble point and the dew point!)


Figure II.2.4: Dew point pressure at 60
o
C (left) and the dew point temperature at 0.5 bar (right) for
the ideal mixture n-hexane/n-heptane containing 50 mole-% n-hexane



Flash evaporation
The sudden decrease in pressure (or, but less common, the sudden increase in temperature) may
result in the splitting of a liquid mixture into a liquid stream and a vapour stream. The liquid and
vapour stream can be considered to be at equilibrium, and the composition of the liquid and the
vapour is then given by the temperature and pressure at the exit point of the flash evaporator
(Please note that flash evaporation is typically associated with a temperature decrease of the
mixture, if this process is carried out adiabatically, since the high enthalpy of the vapour stream
heat of evaporation - has to be taken into account). The amount of liquid and the amount of vapour
leaving the flash evaporator can be determined using a simple mass balance, since the composition
of the vapour phase and liquid phase are determined by equilibrium. For a binary mixture at a given
temperature and pressure, the composition of the liquid and vapour phase in equilibrium with each
other are given, since the degree of freedom equals zero.

If 1 mole of a mixture, containing x
A,mix
moles of A and (1-x
A,mix
) moles of B, splits into a L moles of
liquid, with a composition of x
A
, and V moles of vapour, with a composition of y
A
, in a flash
evaporation, the mole balances yield:
Overall mole balance
Mole balance for A



Substitute the overall mole balance into the mole balance for A:
Mole balance for A

( )

)

Thus, the fraction of the mixture, which appears as a liquid is given by:


0.2
0.4
0.6
0.8
0 0.2 0.4 0.6 0.8 1
P
r
e
s
s
u
r
e
,

b
a
r
Mole fraction of n-hexane
T = 60
o
C
Liquid
vapour
p
dew
x
liquid
40
50
60
70
80
0 0.2 0.4 0.6 0.8 1
T
e
m
p
e
r
a
t
u
r
e
,

o
C
Mole fraction of n-hexane
p = 0.5 bar
Liquid
vapour
T
dew
x
liquid
50

and the mole fraction appearing as a vapour is given by:




Example Flash calculation for ideal mixtures
The pressure of a mixture containing 50 mole-% n-hexane and 50 mole-% n-heptane initially at 5 bar
and 85
o
C is suddenly reduced to 1 bar. How much vapour and how much liquid are being formed, if
the decompression takes place isothermally?
Vapour pressure of n-hexane (Hx):

( )

()

Vapour pressure of n-heptane:

( )

()
.

At 90
o
C, the vapour pressure of the compounds is:
Vapour pressure of n-hexane (Hx):

( )

()

Vapour pressure of n-heptane:

( )

()
.

From the vapour pressure of the pure compounds, it is possible that we are in the VLE-region. To be
certain, we estimate the bubble point pressure and dew point pressure at 85
o
C:


The pressure is between the bubble and dew point we are in the VLE-region

At the given temperature and pressure the composition of the liquid phase is given by the
equilibrium relationship:



Hence, the fraction of the mixture that remains as a liquid is:



In reality this type of processes will be occur rather under adiabatic than isothermal conditions,
implying that the transformation of a liquid mixture into a vapour-liquid mixture will be associated
with a decrease in temperature due to the heat of evaporation. The heat of evaporation of n-hexane
is 30.1 kJ/mol and the heat of evaporation of n-heptane is 34.2 kJ/mol. Furthermore, the specific
heat of n-hexane is 26.3 J/(mol
.
K) and of n-heptane is 22.4 J/(mol
.
K)

The final temperature of the streams leaving the flash evaporator, can be obtained from an energy
balance for this adiabatic system:

( )

( )

( )
(Remember that the resulting vapour and liquid phase are in equilibrium. Hence, the temperature of
the vapour and liquid phase leaving are equal.)
51

( )

( )

( )
Hence, per mole of mixture coming into the system:

( )

( )

( )

We are dealing with ideal mixtures (thus, the change in the enthalpy upon mixing equals zero!).
Thus, the enthalpy of the mixtures can be split into the enthalpy of the individual components with
the excess enthalpy being equal to zero:

( ) (

( )
(


()
( ) (

)
()
( )) (

()
( ) (

)
()
( ))

The enthalpy of a component in the gas phase equals the enthalpy of the same component in the
liquid phase plus the enthalpy change upon evaporation:

()
()
()
()

()

()
()
()
()

()

( ) (

( )

() (

() (

()
( ) ( (

) (

))
()
( )

The mole balance for n-hexane shows

( )
()
( )) (

)
(

( )
()
( ))

()
(

()

The change in the enthalpy of the liquid is mainly due to the change in temperature, since the
enthalpy of the liquid phase is hardly dependent on temperature (incompressible fluid assumption):


()
( ()) (

)
()
( ())

() (

()

Hence, we have a set of equations which needs to be solved simultaneously:
Equilibrium condition 1:

() (

()

()

()

()

Equilibrium condition 2:

()



Mole balance


Energy balance: ( ) ( ()) (

)
()

( ())



Solving these equations simultaneously:


Thus, under these conditions hardly any vapour is formed.

52

II.3. Vapour-liquid equilibrium (VLE) of non-ideal binary mixtures at low pressure
The conditions for a vapour-liquid equilibrium of a binary mixture are in addition to the :


and the temperature and pressure in each phase are equal.

At low pressure, vapours obey the Lewis-Randall rule. Hence the fugacity of a component in the
vapour phase mixture is given by:



Furthermore, at low pressure the fugacity of the pure component is equal to the system pressure:



The fugacity of a component in the liquid phase is given by:



Now the activity coefficient for the liquid phase is no longer equal to 1, but given by an activity
coefficient model. The fugacity of a component in the mixture is now given by:



Thus, the equilibrium relationships are:



and hence the equilibrium pressure for a binary mixture is now given by



If the activity coefficients of at least one of the species in the mixture is larger than 1


(and similar for mixtures for which at least one of the species activity coefficient is less than 1,



Of special interest are mixtures where the equilibrium pressure of the system passes a maximum or
a minimum as a function of the composition:
(

(
(

(
(

*
According to the Gibbs-Duhem equation:

(
(

(
(

(
(

(
(



Substitute it back:
(

(
(

(
(

*
(

) (

(
(

*

53

Since, there is no reason why the second term should be equal to zero (i.e. (

(
(

*
), it means that at the extreme point:
(



From the equilibrium relationships:


Thus, at the extreme point



Furthermore, the sum of the mole fractions equals 1:


And thus also



Thus, at the extreme point, the composition of the liquid phase is identical to the composition of the
vapour phase (for each component!). Thus at a fixed temperature, the extreme pressure (either
maximum or minimum) results in the composition of the vapour phase and the liquid phase to be
identical. This is called an azeotropic mixture. A similar phenomenon takes place for the vapour-
liquid equilibrium of non-ideal mixtures at constant pressure. If the temperature is mixture at the
azeotropic point is larger than the boiling temperature of each component at the given pressure, it is
called a maximum boiling azeotrope. A similarly, it is called a minimum boiling azeotrope, if the
equilibrium temperature is lower than the boiling temperature of each component at the given
pressure.


Example: Constructing p-x-y and T-x-y diagrams for VLE involving non-ideal liquid mixtures
Ethyl acetate (EA) and benzene (B) form a non-ideal mixture. The activity coefficient for this system
can be described by the van Laar model with o = 1.15 and | = 0.92. The vapour pressure of
Ethyl acetate (

()

Benzene (

()

Determine the p-x-y diagram at 75
o
C and the T-x-y diagram at 1 bar.

The equilibrium condition for vapour-liquid-equilibrium (VLE) is:


At low pressure, this translates into:



With (



or


54

()

()

()


At 75
o
C, the vapour pressure of ethyl acetate (EA) is 0.946 bar and of benzene is 0.862 bar.
()


()



The corresponding composition of the vapour phase is given by:




Figure II.3.1: x-y diagram (left) and P-x-y diagram for the non-ideal mixture ethyl acetate/benzene
at 75
o
C

The position of the azeotrope can easily be defined, since this occurs at the point where the
equilibrium pressure of the system is maximum and


Solving this equation results in the mole fraction of ethyl acetate in the liquid phase (and thus the
mole fraction of ethyl acetate in the vapour phase) at the azeotropic point of 0.5180, and the
corresponding pressure of 1.166 bar.

The construction of the T-x-y diagram is analogue to the construction of the P-x-y diagram, but now
the pressure is kept constant at 1 bar:
()

()

()

()

()


(It is assumed here that the coefficients in the van Laar equation do not change significantly over the
temperature range considered. This is typically a good approximation, since the activity coefficient is
not strongly dependent on temperature.)

0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1 m
o
l
e

f
r
a
c
t
i
o
n

o
f

e
t
h
y
l

a
e
t
a
t
e

i
n

v
a
p
o
u
r

p
h
a
s
e
,

y
E
A
Mole fraction of ethyl acetate in liquid
phase, x
EA
T = 75
o
C
azeotropic point
0.8
0.9
1
1.1
1.2
0 0.2 0.4 0.6 0.8 1
p
r
e
s
s
u
r
e
,

b
a
r
Mole fraction of ethyl acetate, x
EA
/y
EA
T = 75
o
C
Liquid
Vapour
azeotropic point
55

This equation can be solved by assuming a mole fraction of ethyl acetate (and thus the mole fraction
of benzene is known), and estimate the temperature for which the total pressure corresponds to 1
bar. This can be done using e.g. Goalseek in Excel.


The corresponding composition of the vapour phase is given by:

()


The azeotropic point is given by:

()

()

This is one equation with two unknowns. However, there is a constraint on the system, viz. the total
pressure in the system must be equal to 1 bar:
()

()

()


Hence, the composition at the azeotropic point and the temperature of the azotropic point can now
be estimated (e.g. by using Solver in Excel). I typically solve for the ratio of the mole fraction in the
vapour phase relative to the mole fraction in the liquid phase (which should be 1) ensuring that the
pressure is the set pressure and varying the temperature and the mole fraction in the liquid phase
(i.e. for a given T and x
EA
, the activity coefficients and the vapour pressure of the pure components
can be calculated and thus the total pressure; knowing the total pressure in conjunction with the
mole fraction in the liquid phase, the activity coefficient and the vapour pressure of a component,
the mole fraction in the vapour phase can be calculated).
T, K x
EA

EA

B
p
vap
EA
p
vap
B
p
calc
y
EA
y
EA
/x
EA
343.5 0.515 1.237 1.348 0.808 0.742 1 0.515 1
Variables, varied during
the optimization
Constrained during
optimization
Target cell, containing the
parameter to be solved for

The azeotrope for the constant pressure system is at a mole fraction of ethyl acetate of 0.515 and a
temperature of 70.38
o
C. The position of the azeotrope depends on the applied conditions!

The system ethyl acetate/benzene shows a minimum boiling azeotrope, since this is the lowest
temperature at which any mixture of ethyl acetate and benzene will evaporate.
56


Figure II.3.2: x-y diagram (left) and T-x-y diagram for the non-ideal mixture ethyl acetate/benzene
at 1 bar

The presence of azeotropes causes difficulty in separation processes, since an azeotropic point
cannot be passed, i.e. the end point of a mixture with an azeotrope will be the azeotropic mixture
and one of the pure components. However, the position of the azeotrope is dependent on the
applied pressure for a system operating at a constant pressure or on the applied temperature for a
system operating at a constant temperature. The condition for the azeotropic point is given by:
(


and in the case of the system ethyl acetate (EA) and benzene (B)


or (

) (

) (

) (

()

()


Figure II.3.3 shows the shift in the position of the azeotrope as a function of the (constant)
temperature. As a consequence, the pressure in the system has to change as well. The extent of the
shift in the azeotropic point depends on the relative heat of evaporation of the two components
(and the excess enthalpy for the mixture)


Figure II.3.3: Shift in the azeotropic point in the mixture ethyl acetate (EA) and benzene (B) as a
function of temperature (left) and the associated shift in the equilibrium pressure
(right)

The mode of operation for the separation of ethyl acetate and benzene into its pure components
depends now on the initial composition. A mixture, which is rich in ethyl acetate (i.e. the mole
fraction of ethyl acetate is higher than the mole fraction of ethyl acetate at the azeotropic point),
will in a first distillation column operating at 1 bar separate into the azeotropic mixture
corresponding to this pressure and pure ethyl acetate. The azeotropic mixture is fed into a column
0
0.2
0.4
0.6
0.8
1
0 0.2 0.4 0.6 0.8 1 m
o
l
e

f
r
a
c
t
i
o
n

o
f

e
t
h
y
l

a
e
t
a
t
e

i
n

v
a
p
o
u
r

p
h
a
s
e
,

y
E
A
Mole fraction of ethyl acetate in liquid
phase, x
EA
p = 1 bar
azeotropic point
70
72
74
76
78
80
0 0.2 0.4 0.6 0.8 1
p
r
e
s
s
u
r
e
,

b
a
r
Mole fraction of ethyl acetate, x
EA
/y
EA
p = 1 bar
Liquid
Vapour
azeotropic point
0.4
0.5
0.6
0 50 100 150 200 250
M
o
l
e

f
r
a
c
t
i
o
n

e
t
h
y
l

a
c
e
t
a
t
e

a
t

a
z
e
o
t
r
o
p
e
,

x
E
A
Temperature,
o
C
0.01
0.1
1
10
100
0 50 100 150 200 250
p
r
e
s
s
u
r
e
,

b
a
r
Temperature,
o
C
57

operating at a higher pressure (e.g. ca. 10 bar), where it will be separated into pure benzene and an
azeotropic mixture corresponding to an equilibrium pressure of 10 bar (which contains more ethyl
acetate than the mixture fed to this column). This azeotropic mixture is then fed back to the first
column operating at 1 bar.


Determining activity coefficients from VLE
Measuring the equilibrium pressure at a given temperature for a mixture at its vapour-liquid
equilibrium point allows the determination of the activity coefficient of the components at each
point, if the composition of the gas phase and the composition of the liquid phase are measured as
well.

For a mixture at its vapour-liquid equilibrium point (at low pressure):


The partial molar Gibbs free energy for each component in the mixture can now be determined from
the activity coefficient as a function of the mole fraction in the liquid phase

)
and thus


Measuring the vapour-liquid equilibrium position at various temperatures lead to the excess
enthalpy, since

)

Example: Determining excess Gibbs free energy in the system n-pentane/propionaldehyde
The equilibrium pressure and the composition of the liquid and the vapour in equilibrium have been
measured for the system n-pentane/propionaldehyde at 40
o
C. Model the molar excess Gibbs free
energy as a function of the mole fraction
x
n-pentane
y
n-pentane
P (bar) x
n-pentane
y
n-pentane
P (bar)
0 0 0.7609 0.4463 0.5877 1.3354
0.0503 0.2121 0.9398 0.5031 0.6146 1.3494
0.1014 0.3452 1.0643 0.561 0.6311 1.3568
0.1647 0.4288 1.1622 0.6812 0.6827 1.3636
0.2212 0.4685 1.2173 0.7597 0.7293 1.3567
0.3019 0.5281 1.2756 0.8333 0.7669 1.3353
0.3476 0.5539 1.2949 0.918 0.8452 1.2814
0.4082 0.5686 1.3197 1 1 1.1541

The activity coefficient as a function of the mole fraction can be determined from:

)
(

)

Knowing the activity coefficient, the partial molar excess Gibbs free energy for each component can
be calculated:

) ( ) (

)
and thus
58




x
n-pentane

n-pentane

propionaldehyde
G
ex
n-pentane
G
ex
propionaldehyde
G
ex
, J/mol
0.0000

1.0000

0.0 0.0
0.0503 3.4337 1.0247 3211.8 63.5 221.9
0.1014 3.1394 1.0192 2978.6 49.6 346.6
0.1647 2.6218 1.0445 2509.4 113.3 507.9
0.2212 2.2340 1.0918 2092.7 228.7 641.0
0.3019 1.9334 1.1332 1716.5 325.6 745.5
0.3476 1.7879 1.1637 1512.8 394.6 783.3
0.4082 1.5928 1.2643 1212.0 610.6 856.1
0.4463 1.5237 1.3068 1096.4 696.7 875.1
0.5031 1.4284 1.3755 928.2 830.0 879.4
0.5610 1.3225 1.4984 727.8 1052.9 870.5
0.6812 1.1841 1.7837 440.0 1506.6 780.0
0.7597 1.1285 2.0086 314.8 1815.8 675.5
0.8333 1.0648 2.4539 163.5 2337.1 525.8
0.9180 1.0223 3.1792 57.3 3011.3 299.5
1.0000 1.0000

0.0

0.0

The molar excess Gibbs free energy can be modeled using a Redlich-Kister expression (see Figure
3.4) with a
0
=3543.7 J/mol (taking more terms into account would reduce the error in the fit only
marginally).


Figure II.3.4: Modeling the excess Gibbs free energy of the system n-pentane/propionaldehyde at
40
o
C with a one-constant Redlich-Kister equation:

)


The activity coefficient model belonging to a one-constant Redlich-Kister expression is the one-
constant Margules expression. Thus, the activity coefficients are given by:
(


0
200
400
600
800
1000
0 0.2 0.4 0.6 0.8 1
E
x
c
e
s
s

m
o
l
a
r

G
i
b
b
s

f
r
e
e

e
n
e
r
g
y
,

G
e
x
,

J
/
m
o
l
Mole fraction n-pentane in liquid phase,
x
n-pentane
59

(



With the model, the position of the azeotropic point at 40
o
C can be estimated. At the azeotropic
point:

) (

) (

*
(

) (

*
(

*

At the azeotropic point


and the equilibrium pressure is given by:



()




Figure II.3.5: Measured P-x-y diagram for the system n-pentane/propionaldehyde at 40
o
C and
model prediction using one-constant Margules expression (solid line)






0.7
0.9
1.1
1.3
0 0.2 0.4 0.6 0.8 1
E
q
u
i
l
i
b
r
i
u
m

p
r
e
s
s
u
r
e
,

b
a
r
Mole fraction n-pentane in liquid phase or
vapour phase, x
n-pentane
/y
n-pentane
60

II.4. High pressure vapour-liquid equilibrium (VLE)
The phase equilibrium conditions for a binary mixture at its point of vapour-liquid equilibrium are:



At high pressure, the Lewis-Randall rule is no longer valid (unless we consider very high pressures)
and the fugacity in the vapour phase cannot be expressed in terms of the total pressure and its mole
fraction. Similarly, the fugacity of the pure component in the liquid phase cannot be set equal to the
vapour pressure of that component. Now, an equation of state (e.g. the Peng-Robinson equation of
state) can be used to determine the fugacity of each species in the respective phases. The fugacity
obtained from the Peng-Robinson equation of state is now given by:

) (

) (

()

()

) (

) (

()

()

*
with


The compressibility is obtained by solving the cubic equation of state:

) (

()

()

with

)
(here a
ii
represents the pure component parameter a)

with b
i
the pure component parameter b


The corresponding phase diagram can be computed (if the binary interaction parameter is known)
by calculating the bubble point pressure for mixtures of a known composition (bubble line
calculation; see Figure 10.3-5), and the dew point pressure for mixtures of a known composition
(dew line calculation; see Figure 10.3-6). The critical point of the mixtures is defined as the point
where the dew point line and the bubble point line meet. The calculation near the critical point is
not easy to obtain convergence.



Example: Determine the P-x-y diagram for the mixture ethane/propane at 298.15 K and 350 K.
At 298.15 K both propane (T
c
= 369.8 K) and ethane (T
c
= 305.4 K) are below their respective critical
temperature. However 350K is higher than the critical temperature for ethane. Hence, it is expected
that at 298.15 K, the P-x-y diagram corresponds to a normal (quite ideal) P-x-y-diagram. At 350K no
condensation of pure ethane is possible.
61


Figure II.4.1: P-x-y diagram for the system propane/ethane at 298.15 K and 350 K.

Interesting behavior can be observed close to the critical point. Figure II.4.2 shows the p-T-diagram
for a mixture containing 40 mole-% n-heptane and 60 mole-% ethane. The critical point is not given
by the maximum pressure or by the maximum temperature, but by the point at which the dew line
and the bubble line meet. Interesting behavior can be observed when compressing this mixture from
10 bar to 80 bar at 478 K. Initially, the mixture is in the vapour phase. Increasing the pressure to 39
bar, and the first liquid start forming, since the dew line is crossed. The amount of liquid increases
upon increasing the pressure further, but at some point the amount of liquid starts to decrease. At a
pressure of ca. 68 bar the liquid disappears and the mixture is again in the vapour phase (since we
passed the dew line again). This double crossing of the dew line at different conditions means that
the mixture is initially present as a vapour, then as a vapour-liquid mixture, and then again as a
vapour. This behavior is called retrograde condensation of the 1
st
kind.


Figure II.4.2: P-T diagram for a mixture containing 40 mole-% n-heptane and 60 mole-% ethane.
Left: Critical point determination for the mixture (i.e. the meeting point of the bubble
line (red) and the dew line (blue)
Right: Retrograde condensation



0
20
40
60
0 0.2 0.4 0.6 0.8 1
E
q
u
i
l
i
b
r
i
u
m

p
r
e
s
s
u
r
e
,

b
a
r
Mole fraction propane
T = 298.15 K
T = 350 K
0
20
40
60
80
100
300 340 380 420 460 500
E
q
u
i
l
i
b
r
i
u
m

p
r
e
s
s
u
r
e
,

b
a
r
Temperature, K
Mixture,
40% n-heptane/60% ethane
critical point
p
max
>p
c,mix
T
max
>T
c,mix
0
20
40
60
80
100
300 340 380 420 460 500
E
q
u
i
l
i
b
r
i
u
m

p
r
e
s
s
u
r
e
,

b
a
r
Temperature, K
Mixture,
40% n-heptane/60% ethane
62

II.5. Solubility of a gas in a liquid
The solubility of a gas in a liquid is given by the equilibrium conditions:


The fugacity of the dissolved gas in the liquid can be described using Henrys law, which is based on
the experimentally observed linear increase in the mole fraction of the dissolved gas with increasing
pressure:

)
(with
i
* describing the deviation from the linear increase in the mole fraction with increasing
pressure). Hence the equilibrium condition for the dissolved gas is given by:



The activity coefficient
i
* can be related to the usual activity coefficients via:

)


The activity coefficients of gases dissolved in a liquid can typically be estimated using e.g. the regular
solution theory
(

) (

or (

) (


(dissolution in pure solvent) (dissolution in a mixture)

which require pre-knowledge about the molar volume of the dissolved gas as a pure liquid and the
solubility parameters. Table 11.1-1 lists these values for some common gases. It should be noted
that these values represent typically a hypothetical liquid since these substances are typically non-
condensable at the temperatures of interest.

For very dilute systems (as typically given for gases dissolved in liquids), the activity coefficient
i
*
approaches 1 (it is typically larger than 0.98, but less than 1, if the volume fraction of the dissolved
gas is less than 0.01). Hence, the Henry coefficient can typically be estimated from the measured
mole fraction at atmospheric pressure (see Table 11.1-2 for the solubility of gases in water):


If p is 1 atm (as in Table 11.1-2):



The Henry coefficient is dependent on pressure. The Henry coefficient is defined as:





Example: Determine the solubility of oxygen in water
The equilibrium mole fraction of oxygen dissolved in water in equilibrium with oxygen at 1.013 bar at
298.15 K is 2.3
.
10
-5
(see Table 11.1-2). Estimate the mole fraction of oxygen in water in equilibrium
with oxygen at 10 bar.

Data:

Handbook of Chemistry and Physics


63

Table 11.1-1

Handbook of Chemistry and Physics

Table 11.1-1

The Henry coefficient for oxygen in water at 298.15 K and 1.013 bar is:

)


The activity coefficient of oxygen in water, *, is given by:

)


Suppose that the activity coefficient can be described using the regular solution theory:
(

) (



Thus,
(

)) (

( )

)

(

)) (

( )

()




For some gas-solvent systems the Henry coefficient is not available. The solubility can be estimated
from:


Now the fugacity of the dissolved gas as a pure liquid at the temperature and pressure of the
mixture needs to be estimated. This can be obtained from the Prausnitz-Shair correlation (see Fig.
11.1-1), which relates the fugacity of a hypothetical liquid at atmospheric pressure relative to the
critical pressure as a function of the relative temperature (i.e. the temperature relative to the critical
temperature). Knowing the fugacity of a component as a liquid at 1.013 bar, the fugacity of this
compound can be estimated at any pressure since



Since liquids are (relatively) incompressible

()


64

II.5.1 Solubility of gases as a function of temperature
The solubility of a gas in a liquid is given by:


In order to evaluate the temperature dependency of the solubility (given by the mole fraction xi), we
will consider the derivative of this condition at constant pressure and constant gas phase
composition:
(
(

))

(
(

divide by



The temperature dependency of the activity coefficient is given by:
(

()


The temperature dependency of the fugacity of the gas as a pure liquid is given by:
(



Thus, the temperature dependency of the solubility of a gas at a given pressure and a given gas
phase composition is given by:
(

()


(Note,

( ) (

) can be interpreted as the heat of evaporation from a


mixture).
The heat of evaporation is typically much larger than the partial molar excess enthalpy. Hence,
(


Hence, the solubility of the gas will decrease if

. It should be noted that beyond the


critical point the (extrapolated) heat of evaporation becomes less than zero. Thus, the solubility of
gases can increase with increasing temperature if the temperature is much larger than the critical
temperature of the gas.



65

II.6. Liquid-liquid equilibrium
The appearance of two or more liquid phases in equilibrium with each other implies that a single
liquid phase is not stable. The equilibrium condition for a system at equilibrium (at constant
temperature and pressure) is that the Gibbs free energy is at a minimum. This yields two conditions,
i.e.
and

at constant N,T,p
The total Gibbs free energy for a binary liquid mixture is given by:

))


or

))

))

))

)

For an ideal mixture (G
ex
=0), the Gibbs free energy of the mixture is always less than the sum of the
Gibbs free energy of the pure components since ln(x
i
) is less than zero. Hence, the addition of two
pure compounds forming an ideal mixture will always result in a lowering of the Gibbs free energy
and hence yielding a mixture. However, the excess molar Gibbs free energy can be larger than zero
resulting in a Gibbs free energy of the mixture larger than the Gibbs free energy of the pure
components.

Suppose the excess Gibbs free energy s given by the one-constant Margules equation:


Then depending on the value of A, the Gibbs free energy of the mixture can be larger than the Gibbs
free energy of the pure components (see Figure II.6.1).


Figure II.6.1: Difference between the molar Gibbs free energy of the mixture and the molar Gibbs
free energy of the pure components as a function of the mole fraction for binary
mixtures for which the excess Gibbs free energy can be described by the Margules
expression (A/RT = 0 represents an ideal mixture)

Two liquid phases can form, if the difference between the Gibbs free energy of the mixture and the
Gibbs free energy of its components as a function of the composition passes a minimum:

(


since in that situation the formation of two liquid phases will result in a lower Gibbs free energy than
the formation of a single liquid phase. The composition of the phases cannot be determined from
-0.8
-0.6
-0.4
-0.2
0
0.2
0 0.2 0.4 0.6 0.8 1
(
G
m
i
x
t
u
r
e
-
x
1
.
G
1
-
x
2
.
G
2
)
/
R
T
Mole fraction x
1
A/RT = 0
A/RT = 1
A/RT = 2
A/RT = 3
66

this consideration. The equilibrium conditions have to be considered. For a binary system containing
two liquids to be at equilibrium, the following equilibrium conditions must be valid:


The fugacity of each species in each of the liquid phases must be equal, and thus
(


The vapour pressure of the pure component is independent of the phase:
(


Hence, the composition of the two liquid phases in equilibrium can be found using:
(


With the boundary conditions (

and (




Example: Phase splitting in the liquid furfural/iso-butane mixture
The mixture furfural/iso-butane splits into two phases at 37.8
o
C and 5 bar. Determine the
composition of the phases, if the activity coefficients in the mixtures can be represented by the van
Laar expression:
(



The equilibrium conditions are (besides constant T and p):



With the boundary conditions (

and (



These 4 equations have to be solved simultaneously:


Solving these equations (e.g. by trial and error) yields:

and




As shown in Figure II.6.1 the likelihood for the formation of two liquid phases in equilibrium with
each other diminishes with increasing temperature (with increasing temperature the term A/RT will
decrease). The highest temperature at which a mixture will split into two liquid mixtures is called the
upper consolute temperature. The upper consolute temperature can be determined from the
stability criterion of a single liquid. A phase is stable if:

at constant N,T,p
The change in the stability occurs, when
67

(


since if (

a single phase will be stable and if (

a single phase will be


unstable. If the mixture can be described with a one constant Margules expression and

))


The 1
st
derivative is given by:
(

((

((

) (

))


And the 2
nd
derivative is thus given by:
(

)
Thus, the highest temperature at which two liquid phases co-exist for a mixture of a given
composition is given by:



At temperatures below this maximum temperature, two liquid phases will co-exist and above the
upper consolute temperature, a single liquid phase will exist for a Margules mixture. The highest
possible temperature at which two liquid phases co-exists is given when the product of the mole
fractions is maximum. This occurs for a Margules mixture at x
1
= x
2
= 0.5. This temperature is the
upper consolute temperature, which for a Margules mixture is given by:




II.6.1 Distribution of compounds over two immiscible liquids
The distribution of species over two in essential immiscible liquids is used industrially for the
extraction of material from a process stream. For instance methyl bromide dissolved in water can be
extracted by using a hydrocarbon, e.g. n-decane. The hydrocarbon and water are essentially
immiscible. The distribution of the compound, which is to be extracted, over the two phases is given
by the equilibrium relationship:


(only the compound to be extracted needs to be considered, since the other components in this
system remain in essence in the same phase). The fugacity of the compound to be extracted in each
of the liquid phases must be equal, and thus
(


The distribution of the compound over the two phases is thus given by:


In literature, you will find often a concentration partition or distribution coefficient, which is defined
as


68

The partition coefficient is related to the thermodynamic distribution coefficient. If the mole fraction
of the species involved in the solution is small, then


Hence, the concentration of this species is given by:




Example: Distribution of bromine between an organic and aqueous phase
The concentration distribution coefficient of bromine (Br
2
) between carbon tetrachloride (CCl
4
) and
water at 25
o
C is
[Br
2
]
in CCl
4
, kmol/m
3
0.04 0.1 0.5 1.0 1.5 2.0 2.5



26.8 27.2 29.0 30.4 31.4 33.4 35.0
Calculate the ratio of the activity coefficient of Br2 in water to that in carbon tetrachloride.

Data
Species M, g/mol , kg/m
3

CCl
4
153.84 1.595
.
10
3

Water 18.0 1.0
.
10
3

Br
2
159.83 3.119
.
10
3


The concentration in the water phase can now easily be calculated since


[Br
2
]
in CCl
4
, kmol/m
3
0.04 0.1 0.5 1.0 1.5 2.0 2.5
[Br
2
]
in water
, kmol/m
3
1.5
.
10
-3
3.7
.
10
-3
1.7
.
10
-2
3.3
.
10
-2
4.8
.
10
-2
6.0
.
10
-2
7.1
.
10
-2


The concentration can be re-calculated in terms of the mole fraction. The molar volume of the
mixture containing the solvent and bromine assuming that the excess volume can be neglected is
given by:



The number of moles of bromine in 1 mole of the mixture:

] ((



Thus, the mole fraction of bromine is given by

](

)

(Be careful with the units!)
69



x
Br
2
in CCl
4

3.85
.
10
-3
9.60
.
10
-3
4.72
.
10
-2
9.23
.
10
-1
1.35
.
10
-1
1.77
.
10
-1
2.17
.
10
-1

x
Br
2
in water

2.69
.
10
-5
6.62
.
10
-5
3.11
.
10
-4
5.93
.
10
-4
8.61
.
10
-4
1.08
.
10
-3
1.29
.
10
-3


The relative activity coefficient can now be calculated from the equilibrium condition:
(


[Br
2
]
in CCl
4
, kmol/m
3
0.04 0.1 0.5 1.0 1.5 2.0 2.5


143.3 145.1 151.9 155.7 157.3 163.8 168.1




II.6.2 Ternary phase diagrams
A system with three components (ternary systems) have F = C-P+2 = 5-P degrees of freedom. If
systems at a given temperature and pressure are considered (i.e. taking away 2 degrees of freedom),
the degrees of freedom is F= 3-P. Hence, 2 variables are required to describe a single phase ternary
system (i.e. x
1
and x
2
, with x
3
= 1-x
1
-x
2
). To depict the dependency of the variables, ternary diagrams
are being used (see Figure 6.1). The corners in this triangle graph represent the pure components.
The lines opposite the corner representing the pure component are lines of constant composition.
For instance, the top corner in Figure II.6.2 represents the pure component A. The horizontal lines
represent mixtures with a constant mole fraction of A. A mixture containing 40% A and 30% B can be
found from the intersection between the line representing mixtures containing 40% A and the line
representing mixtures containing 30% B. The shape of the ternary diagram enforces the sum of the
mole fractions being equal to 1.


Figure II.6.2: Depicting mixtures in a ternary diagram

The ternary phase diagram is useful to assess whether the mixture will split into two or more liquid
phases at the temperature considered. Figure II.6.3 shows the ternary diagram for the system water-
methanol-acrylonitrile at 288.15K and atmospheric pressure. It is clear that in the methanol-poor
0 20 40 60 80 100
A
C B
x
B
100
80
60
40
20
0
x
c
100
80
60
40
20
0
x
A
70

region between mole fractions for acetonitrile of 2% to 91% the appearance of two liquid phases can
be expected. It should be noted that a ternary system at a constant pressure and temperature may
contain up to 3 different liquid phases (see Figure11.2-11e in Sandler, p. 620 and further examples
on p. 621).

The presence of two liquid phases does not yield any information on the composition of these
phases in equilibrium with each other. The phases in equilibrium with each other are connected with
the tie-line (see Figure II.6.4).


Example: Liquid-liquid equilibrium in the system water/methanol/acetonitrile
A stream contains 40 mole-% acetonitrile and 50 mole-% water with the balance methanol at 288.15
K and atmospheric pressure. Determine the phase composition of each of the liquid phases and the
relative amount of each phase present.

The composition of the two liquid phases can be read off from the phase diagram. After locating the
point representing the composition of the mixture on the ternary diagram, draw a tie line (almost
parallel to the nearest given tie-line). At the crossing point of the tie-line with the LLE-curve, the
composition of the two liquid phases can be read off.

Figure II.6.3: Ternary phase diagram for the system water-methanol-acrylonitrile at 288.15 K and 1
atm showing LLE-region, i.e. region where two liquids are in equilibrium with each
other (data: X. Zhang and C. Jian, J. Chem. Eng. Data 52 (2012), 142)

0 20 40 60 80 100
methanol
water acetonitrile
x
acetonitrile
100
80
60
40
20
0
x
water
100
80
60
40
20
0
x
methanol
Two liquid region
single liquid region
71


Figure II.6.4: Ternary phase diagram for the system water-methanol-acrylonitrile at 288.15 K and 1
atm showing the composition of the phases in equilibrium with each other (data: X.
Zhang and C. Jian, J. Chem. Eng. Data 52 (2012), 142)



Figure II.6.5: Ternary phase diagram for the system water-methanol-acrylonitrile at 288.15 K and 1
atm showing the splitting of a mixture containing 40 mole-% acetonitrile and 50 mole-
% water into two liquid phases (data: X. Zhang and C. Jian, J. Chem. Eng. Data 52
(2012), 142)

Composition of L
I
: 85 mole-% water, 5 mole-% acetonitrile 10 mole-% methanol
Composition of L
II
: 10 mole-% methanol, 77 mole-% acetonitrile 13 mole-% water

The relative amounts of L
I
and L
II
can be obtained from a mole balance:

Mole balance for water: (


Mole balance for ACN (acetonitrile): (


0 20 40 60 80 100
methanol
water acetonitrile
x
acetonitrile
100
80
60
40
20
0
x
water
100
80
60
40
20
0
x
methanol
tie-lines
0 20 40 60 80 100
methanol
water acetonitrile
x
acetonitrile
100
80
60
40
20
0
x
water
100
80
60
40
20
0
x
methanol
tie-lines
72


Define

and


Or



And substitute the obtained mole fractions:




73

II.6.3 Vapour-liquid-liquid equilibria
The composition of co-existing liquids in a liquid-liquid equilibrium is (almost) independent of
pressure, although dependent on temperature (see the section on the upper consolute
temperature). Hence, decreasing the pressure of a system with co-existing liquids will yield a system
where vapour-liquid-liquid equilibrium can be expected.

Determining the bubble point pressure for a system with co-existing liquids
At the bubble point the overall composition of the liquid mixture is known, and the composition of
the co-existing liquids is given by the condition(s):
(

for each component in the mixture


Furthermore, the co-existing liquids are in equilibrium with the vapour phase. For a system at low
pressure, the condition is thus given by:
(

for each component in the mixture


The bubble point pressure of the mixture containing co-existing liquids is given by:
(


The consequence is that the bubble point pressure is constant in the range of the co-existing liquids
(as represented by the horizontal line in Figure 6.6), i.e. it does not vary with the overall composition
of the mixture. In the region outside the liquid-liquid equilibrium, the bubble point pressure varies
with the composition and is given by:




Example: Bubble point determination in the presence of LLE
The binary liquid mixture of nitromethane (NM) and 2,2,5 trimethylhexane (TMH) exhibits liquid-
liquid equilibrium. The activity coefficient of nitromethane (NM) and 2,2,5 trimethylhexane (TMH) at
100
o
C can be described with the following van Laar equations:
(


The vapor pressure of the pure components at 100
o
C are 1.0 bar and 0.5 bar for nitromethane and
2,2,5 trimethyl hexane, respectively. Determine the bubble point pressure as a function of the overall
composition in the mixture.

The composition of the co-existing liquids at 100oC can be determined from:
(



This can be developed into iteration formulae taking into account the sum of the mole fractions in a
phase adds up to 1:

*
,


74


Starting values for the iterations would be

with the corresponding activity coefficients


(note that you cannot take the pure substances as the starting point)



Hence, at 100
o
C mixtures with a mole fraction of nitro-methane between 0.4 and 0.91 are expected
to exhibit phase splitting, i.e. co-existence of two liquids.

The bubble point pressure for mixtures with a mole fraction of nitro-methane between 0.4 and 0.9 is
given by:
(


( )

( )


For mixtures with a mole fraction of nitro-methane of less than 0.4 and more than 0.91, the bubble
pressure is given by :



The corresponding dew line can be calculated from:


At a given composition of the liquid, the activity coefficient and the equilibrium pressure can be
calculated. Then, the corresponding mole fraction in the vapour phase can be determined.



75

Determining the position of the azeotrope
The systems exhibiting vapour-liquid-liquid equilibria are highly non-ideal and may thus exhibit an
azeotrope. At the azeotrope:




Example: Azeotrope and liquid-liquid equilibrium in the system nitro-methane/2,2,5 trimethyl
hexane
The binary liquid mixture of nitro-methane (NM) and 2,2,5 trimethyl hexane (TMH) exhibits liquid-
liquid equilibrium. The activity coefficient of nitro-methane (NM) and 2,2,5 trimethyl hexane (TMH)
at 100
o
C can be described with the following van Laar equations:
(


The vapor pressure of the pure components at 100
o
C are 1.0 bar and 0.5 bar for nitro-methane and
2,2,5 trimethyl hexane, respectively. Determine the azeotropic point for this mixture at 100
o
C.

At the azeotropic point, the mole fraction in the vapour and liquid phase are identical, and thus


The composition of the azeotropic point is given by:

) (

) (


Solving this equations, yields a mole fraction of nitro-methane of 0.716, which lies within the region
where the system exhibits a splitting of the liquid phase into two liquid phases. Hence, the pressure
at the azeotropic point is given by
( )

( )


The overall phase diagram for the system nitro-methane/2,2,5 trimethyl hexane at 100
o
C is given in
Figure 6.6. The vapour is either in equilibrium with the liquid L
I
or with the liquid L
II
. Only mixtures
with the overall composition equal to the azeotropic composition show three phases in equilibrium
with each other. This could have been expected, since for a two component system, the degrees of
freedom is given by F = C-P+2=4-P. The choice of temperature reduces the degrees of freedom with
1. Hence, F=3-P. When 3 phases are in equilibrium (at a given temperature), the system is fully
defined and hence the position at which three phases are in equilibrium with each other are defined
by a point in the phase diagram.

Figure II.6.6: Phase diagram for the system nitro-methane/2,2,5 trimethyl hexane at 100
o
C showing
the point of the heterogeneous azeotrope

0
0.5
1
1.5
2
0 0.2 0.4 0.6 0.8 1
E
q
u
i
l
i
b
r
i
u
m

p
r
e
s
s
u
r
e
,

b
a
r
Mole fraction of nitromethane
V
L
I
L
II
LI+L
II
VI+L
II
VI+L
I heterogeneous
azeotrope
bubble line
dew line
76

II.7 Equilibrium involving solids
The discussion on the equilibrium of solids is here limited to solids, which do not form solid mixtures
and thus the fugacity of the solid species is that of the pure component


Furthermore, only phase changes, in which the molecular structure of the component remains
unaltered will be considered. Hence, the dissolution of ionic compounds, which (partially) dissociate
upon dissolution, will not be considered her.

II.7.1 Dissolution of a solid component in a liquid
The equilibrium solubility of a solid is given by the mole fraction of that component in the liquid
phase in equilibrium with the solid phase:

)

The product of the solubility (given by its mole fraction in the liquid phase) and the activity
coefficient is thus given by the ratio of the fugacity of the pure component as a solid relative to the
fugacity of the same component as a liquid.



Hence, the product of the solubility and the activity coefficient can be found, when the Gibbs free of
fusion at the temperature and pressure of the mixture is known. At the normal melting point, the
Gibbs free energy of fusion equals zero (

) ). The fusion and melting process are


hardly dependent on pressure, and only the temperature change will be considered. A three step
process is used to determine the change in the Gibbs free energy upon fusion at the temperature
and pressure of the mixture.
4. Heating the pure component from the temperature of the mixture to the melting
temperature
5. Melting the pure component at its melting temperature
6. Cooling the liquid formed down to the temperature of the mixture

Enthalpy change for this three step process:


Entropy change for this three step process:



Defining the difference between the specific heat of the pure compound as a liquid and the specific
heat as a pure component as a solid as AC
p
:



At the normal melting temperature

)
or


Substituting it all in
77

) (



For many substances, the specific heat of the solid is approximately the same as the specific heat of
the liquid:

) (

)

And the fugacity coefficient of the pure component as a liquid is thus given by:

*



Hence, the solubility of a solid in a liquid is given by:

*

The activity coefficient itself is dependent on the composition and thus an iterative process is
needed to determine the solubility of the solid in a liquid. A good starting point is often by using the
ideal solubility (i.e.
i
= 1).


Example: Solubility of naphthalene in benzene
Determine the solubility of naphthalene in benzene at 25
o
C for an ideal solution, and when the
activity coefficient can be described using the regular solution theory. The difference in the specific
heat between liquid naphthalene and solid naphthalene can be neglected.
Data for naphthalene
1
:
M (g/mol) (g/cm
3
) T
melt
(K) A
fus
H (kJ/mol) A
vap
H (kJ/mol)
128.17 1.0253 353.4 19.01 43.2
1
http://webbook.nist.gov/chemistry/

The Gibbs free energy of fusion at 298.15 K is given by:

( ) (



Hence, the product of the solubility and the activity coefficient is given by:



In an ideal solution, the activity coefficient for naphthalene is equal to 1. Thus, the solubility in an
ideal solution is 0.302

For a nono-ideal solution the activity coefficient must be considered. According to the regular
solution theory, the activity coefficient of naphthalene in benzene is given by:
(

) (


The solubility parameter of benzene is given in Table 9.6-1 as 9.2 (cal/cm
3
)
0.5
(or 18.8 (J/cm
3
)
0.5
) and
the molar volume of 89 cm
3
/mol. The molar volume for naphthalene can be estimated from its
density and molar mass:


78

The solubility parameter for naphthalene can then be estimated from the change in the enthalpy
upon evaporation assuming that the change in enthalpy upon evaporation is not dependent on
temperature (it should actually be evaluated at the temperature of the mixture, i.e. 298.15 K!):



Thus, the activity coefficient of naphthalene in the naphthalene/benzene mixture is given by:
(

)
()

(
(

)
(


(
(

)
(



and the solubility of naphthalene can be obtained from:

(
(

)
(



Iteration:
x
naphthalene
u
benzene

naphthalene

0.302461 0.621451 1.011655
0.298977 0.625342 1.011803
0.298933 0.62539 1.011804
0.298933 0.625391 1.011804

The solution of naphthalene in benzene is rather ideal resulting in a high solubility of this compound
in benzene as a solvent (note: the regular solution theory overpredicts the solubility of naphthalene
in benzene by ca. 5%.



II.7.2 Freezing point depression
The melting point of a component is depressed in the presence of another compound. The point at
which a solid starts forming is given by the equilibrium relationship:

*

(

)



Example: Freezing point depression in the system naphthalene/benzene
Pure naphthalene melts at 353.4 K. Determine the melting point in a mixture containing 10%
benzene if the solution is ideal and when the activity coefficient can be described using the regular
solution theory. The difference in the specific heat between liquid naphthalene and solid naphthalene
can be neglected.
79

Data for naphthalene
1,2
:
M (g/mol) (g/cm
3
) T
melt
(K) A
fus
H (kJ/mol) A
vap
H (kJ/mol) o (J/cm
3
)
0.5
V (cm
3
/mol)
128.17 1.0253 353.4 19.01 43.2 18.05 125
1
http://webbook.nist.gov/chemistry/
2
determined in example in section 7.1

The melting point is given by:

)

At the melting point, the composition of the mixture is 90% naphthalene and 10% benzene.


(
(

)
(


Thus, the mixture is rather ideal!!

For an ideal mixture:

()


Thus, a mixture containing 90 mole-% naphthalene melts at a temperature of 6.7 K lower than pure
naphthalene.



II.7.3 Solid-liquid equilibrium phase diagram
The extent of the freezing point depression increases with increasing content of the second
component in the mixture. However, also the freezing point of the second component decreases
with increasing content of the first component in the mixture. The crossing point of these curves
represents the lowest temperature at which this mixture can exist as a liquid. This point is also called
the eutectic point.


Example: Melting point for mixtures of naphthalene and benzene
Pure naphthalene melts at 353.4 K and pure benzene at 278.57 K. Determine the melting point of
mixtures of benzene/naphthalene as a function of the mole fraction of naphthalene. The heat of
fusion of naphthalene is 19.01 kJ/mol and the heat of fusion of benzene is 9.9 kJ/mol. The differences
in the specific heat of the solid and the liquid can be neglected and thus the heat of fusion can be
considered to be independent of temperature.The napthlene/benzene mixture can be assumed to be
ideal

The freezing point of naphthalene can be described by:


Similarly, the freezing point of benzene can be described by:


Hence, the curves for the freezing point of naphthalene and benzene from an indeal
naphthalene/benzene mixture is given in Figure II.7.1. The lowest temperature for a mixture of
naphthalene and benzene to be present as a liquid is at 269.5 K for a mixture containing 13.4 mole-%
naphthalene.
80


Figure II.7.1: Freezing point of naphthalene and benzene in an ideal naphthalene/benzene mixture

Now the phase diagram for the solid-liquid transition in the system naphthalene/benzene can be
drawn. A decrease in temperature for mixtures rich in naphthalene will precipitate out pure
naphthalene, but solid benzene will precipitate out first from mixtures with a content of less than
13.4 mole-% naphthalene (the composition of the eutectic point). The separation of the compounds
in a liquid and a pure solid can be used for purification of compounds (just like distillation, but is
here called fractionated crystallization).

Figure II.7.2: Phase diagram for an ideal naphthalene/benzene mixture




200
250
300
350
0 0.2 0.4 0.6 0.8 1
T
,

K
Mole fraction naphthalene
250
275
300
325
350
0 0.2 0.4 0.6 0.8 1
T
,

K
Mole fraction naphthalene
L
S
benzene
+S
naphthalene
L+S
naphthalene
L+S
benzene
eutectic point
81

III Chemical equilibria

III.1. Introduction
The principle of chemical equilibria at a given temperature and pressure is the minimization of the
Gibbs free energy of system due to the chemical reaction. The Gibbs free energy of a system is given
by the sum of the number of moles of each component present in the system time the partial molar
Gibbs free energy of these components



The number of moles of a component is changing due to the reaction and can be can be expressed in
terms of the extent of reaction or conversion (see notes on Reactor Design I). For instance, for the
reaction
A + b B c C + d D
the number of moles can be expressed as:


or in general



Thus, the Gibbs free energy of the system is given by:


The condition for the system to be at equilibrium requires that the change in the Gibbs free energy
with changing extent of reaction is zero:
(

) (


It can be derived from the Gibbs-Duhem relationship, that
(

) (



Hence, the equilibrium condition for a chemical reaction is:
(



This can be easily extended to systems with multiple reactions occurring (note: only independent
reactions need to be taken into consideration!)

The analysis of the equilibrium position requires therefore knowledge on the partial molar Gibbs
free energy of the components in the system at the equilibrium pressure, temperature and
composition. A standard reference state can be defined of the species at a defined composition (x
i
0
)
and 1 bar. Hence,

( )

) (

( )

)*

( )

) (

()

)
*
The ratio of the fugacity of the species in a mixture relative to the fugacity of the species at the
standard state is called the activity, a
i

()

)

Thus,

( )

) (

)

82

The position of the equilibrium is thus given by:
(



Hence, the thermodynamic equilibrium constant (K
a
) is defined as:



It should be noted that the definition of the activity of a component is strictly linked to the definition
of the fugacity of the species at the reference state, and thus the Gibbs free energy of reaction!



83

III.2. Influence of pressure and temperature on the activity based equilibrium constant
The activity based equilibrium constant is defined as:
(



Hence, the pressure dependency of the equilibrium constant is given by:
(
((

))

(
(

(
(


(
((

))

(
(

(
(

)*


Since the partial molar Gibbs free energy of the components is defined at 1 bar, it is therefore not
dependent on pressure:

(
(

)*


And hence the activity based equilibrium constant does not change with changing pressure at
constant temperature.

The temperature dependency of the equilibrium constant is given by:
(
((

))

(
(


Hence, the change in the activity based equilibrium constant with changing temperature is given by
(

)) (

))

)
)



It should be noted that the enthalpy (and thus also the change in the enthalpy upon reaction) is
dependent on temperature:
(

or ( ) (



The ideal gas specific heat can often be described as a polynomial as a function temperature


Defining the change in the specific heat upon reaction as:



And hence the enthalpy change upon reaction as a function of temperature is given by:

( )

) (

( ) (



Thus, the temperature dependency of the activity based equilibrium constant is given by:
(

)
)


84

(

)
) (
(

)


(

)
)
(

)

85

III.3 Single gas phase reactions
For gas phase reactions it is convenient to define the reference state as the pure components at 1
bar. Hence, the activity of the components in the mixture is given by:

()

( )

)


The activity based equilibrium constant is thus related to the K
p
, since

)

)

)

)

( )



The activity based equilibrium constant has the same numerical value as K
p
(when the units bar are
used!), if the Lewis-Randall rule is applicable and the fugacity coefficient is equal to 1, but does not
have the same units. The activity based equilibrium constant is dimensionless, whereas K
p
has the
units of pressure.


Example: Single gas phase reaction with a constant number of moles water gas shift reaction
The water gas shift reaction H
2
O(g) + CO(g) CO
2
(g) + H
2
(g) is an important chemical reaction. This
reaction is typically performed at temperatures of ca. 400
o
C and elevated pressure.
The following data can be found for the various components:
Compound A
f
H (298.15K) A
f
G (298.15K)

, (J/mol
.
K)
kJ/mol kJ/mol a b c d
H
2
O(g) -241.8 -228.6 32.218 0.192
.
10
-2
1.055
.
10
-5
-3.593
.
10
-9

CO(g) -110.5 -137.2 28.142 0.167
.
10
-2
0.537
.
10
-5
-2.221
.
10
-9

CO
2
(g) -393.5 -394.4 22.243 5.977
.
10
-2
-3.499
.
10
-5
7.464
.
10
-9

H
2
(g) 0 0 29.088 -0.192
.
10
-2
0.400
.
10
-5
-0.870
.
10
-9

1. Determine the change in the Gibbs free energy upon reaction at 298.15 K.
2. Determine the change in the Gibbs free energy upon reaction at 673.15 K.
3. Determine the change in the Gibbs free energy of reaction relative to the Gibbs free energy of
the starting mixture as a function of the conversion of CO for an equi-molar feed of water
and CO at a temperature of 673.15 K and 1 bar
4. Determine the equilibrium conversion of CO for an equi-molar feed at 673.15 K and 1 bar
5. Determine the equilibrium conversion of CO for an equi-molar feed at 673.15 K and 10 bar
6. Determine the equilibrium conversion of CO for an equi-molar feed at 673.15 K and 10 bar
for a feed containing 20 mole-% CO and the balance water.

1 Determine the change in the Gibbs free energy upon reaction at 298.15 K.
The change in the Gibbs energy upon reaction is given by:

)

Taking the reference states as the pure components at a pressure of 1 bar:

( )

( )

( )

( )

We can relate the Gibbs free energy of the various components (although the Gibbs free energy of
the pure components is not known), by using the elements in their natural state at 298.15 K and 1
bar as the reference state.

( )

( )

( )

( )

( )
86

( ) () () () ()


2 Determine the change in the Gibbs free energy upon reaction at 673.15 K.
The change in the Gibbs free energy with temperature (at a constant pressure is given by):
(
(


This differential equation can be solved by separation of variables:
(




In order to evaluate the change in the Gibbs free energy upon reaction, the function of the reaction
enthalpy as a function of temperature must be known.

()

( ) () () () ()



The change in enthalpy upon reaction at 298.15 K is given by:

( )

( )

( )

(
)

( ) () () () ()


with

()

) (

()

()

( )



Thus, the change in the Gibbs free energy upon reaction at 673.15 K is given by:
87

(

)

(

)

(

)

(

( ) ()



3 Determine the change in the Gibbs free energy of reaction relative to the Gibbs free energy of
the starting mixture as a function of the conversion of CO for an equi-molar feed of water and
CO at a temperature of 673.15 K and 1 bar

The number of moles in the system as function of the conversion of CO can be found from the
stoichiometric table. Define X
CO
as the conversion of CO:
Compound Initial Final y
i

H
2
O(g)

( ) ( )
CO(g)

( ) ( )
CO
2
(g) 0


H
2
(g) 0



Thus the Gibbs free energy of the system can now be defined in terms of a single variable X

( )

( )


or per mole of CO entering the system:

( )

( )



The partial molar Gibbs free energy can be related to the Gibbs free energy of the pure component
at 1 bar:

( ) (

( ))

)

Thus, the Gibbs free energy of the system

) (( ) (

) ( )
(

) (

) (

))

and the change in the Gibbs free energy of the system as a function of conversion is given by:

(( ) (

) ( ) (

) (

)
(

)*

The change in the Gibbs free energy of the system contains a linear term in terms of the conversion
(proportional to the change in the Gibbs free energy upon reaction, A
rxn
G) and a non-linear term,
due to the mixing of the reactants and the products (see Fig. 3.1). The Gibbs free energy of the
system is minimal at full conversion, if the products and reactants do not mix. A resulting partial
equilibrium conversion is caused by the contribution of the mixing term to the Gibbs free energy of
the system.
88


Figure III.3.1: The change in the Gibbs free energy of the system relative to the Gibbs free energy of
the starting mixture at 673.15 K for an equi-molar feed of CO and water at 1 bar.

4 Determine the equilibrium conversion of CO for a an equi-molar feed at 673.15 K and 1 bar
The equilibrium constant for the water-gas shift reaction at 673.15 K is given by:

()


()



The equilibrium composition in the gas in the gas phase is related to the equilibrium constant K
a



The activity of hydrogen is given as:

()

( )


at low pressure
Thus, the activity based equilibrium constant:


The mole fractions are defined in terms of the conversion of CO:

()()

()



Solving for the conversion of CO X
CO
= 0.7735


5 Determine the equilibrium conversion of CO for a an equi-molar feed at 673.15 K and 10 bar
The Gibbs free energy of reaction and thus the activity based equilibrium constant are not a function
of pressure. Hence

()


()



The composition in the gas phase is also not dependent on the pressure. since

()


both the left-had side of the equation and the right hand side of the equation are independent of
pressure. Hence, the equilibrium conversion remains unchanged upon changing the pressure at
77.35%.



-20
-15
-10
-5
0
0 0.2 0.4 0.6 0.8 1
G
s
y
s
t
e
m
/
N
C
O
,
0
-
G
C
O
-
G
H
2
O
-
2
R
T
.
l
n
(
2
)
CO-conversion, X
CO
A
mix
G
(mixing of reactants and poducts)
X
.
A
rxn
G
89

6 Determine the equilibrium conversion of CO for a an equi-molar feed at 673.15 K and 10 bar
for a feed containing 20 mole-% CO and the balance water.
The equilibrium constant is not a function of the inlet composition. However, the resulting
equilibrium composition of the gas phase is dependent on the starting composition of the mixture.
Setting up the stoichiometric table for a feed containing 20 mole-% CO and the balance water:

Compound Initial Final y
i

H
2
O(g)

( ) ( )
CO(g)

( ) ( )
CO
2
(g) 0


H
2
(g) 0



The gas phase composition is given by:

()()

()()

Solving for the conversion of CO X
CO
= 0.9639

Hence, the equilibrium conversion can be improved by increasing the excess of water fed into the
system (but not by changing the pressure).



Example: Single gas phase reaction with changing number of moles ammonia synthesis
The ammonia synthesis 3 H
2
(g) + N
2
(g) 2NH
2
(g) is an important chemical reaction. This reaction is
typically performed at temperatures of ca. 450
o
C and pressures as high as 300 bar.
The following data can be found for the various components:
Compound A
f
H (298.15K) A
f
G (298.15K)

, (J/mol
.
K)
kJ/mol kJ/mol a b c d
H
2
(g) 0 0 29.088 -0.192
.
10
-2
0.400
.
10
-5
-0.870
.
10
-9

N
2
(g) 0 0 28.883 -0.157
.
10
-2
0.808
.
10
-5
-2.871
.
10
-9

NH
3
(g) -46.1 -16.5 24.619 3.75
.
10
-2
-0.138
.
10
-5

1. Determine the change in the Gibbs free energy upon reaction at 723.15 K.
2. Determine the equilibrium conversion of N
2
for stoichiometric feed of H
2
and N
2
at 723.15 K
and 1 bar
3. Determine the equilibrium conversion of N
2
for stoichiometric feed of H
2
and N
2
at 723.15 K
and 10 bar
4. Determine the equilibrium conversion of N
2
for a feed containing 90% Ar, 7.5 % H
2
and 2.5%
N
2
at 723.15 K and 10 bar
5. Determine the equilibrium conversion of N
2
for stoichiometric feed of H
2
and N
2
at 723.15 K
and 100 bar

1 Determine the change in the Gibbs free energy upon reaction at 723.15 K.
The change in the Gibbs energy upon reaction at 298.15K is given by:

( )

( )

( )

( )
(taking the reference states as the pure components at a pressure of 1 bar and relating the Gibbs
free energy of the various components (although the Gibbs free energy of the pure components is
not known), by using the elements in their natural state at 298.15 K and 1 bar as the reference state)

( ) () () () ()


90


The change in the Gibbs free energy with temperature (at a constant pressure of 1 bar is given by):
(



In order to evaluate the change in the Gibbs free energy upon reaction, the function of the reaction
enthalpy as a function of temperature must be known.

()

( ) () () ()


with

()

) (

()

()

)


Thus, the change in the Gibbs free energy upon reaction at 723.15 K is given by:
(

)

(

( )




2 Determine the equilibrium conversion of N
2
for stoichiometric feed of H
2
and N
2
at 723.15 K
and 1 bar
The equilibrium composition in the gas phase is given by:



with the equilibrium constant given by

()


()




At low pressure,


*(


91


The mole fractions of the individual components can be obtained from a stoichiometric table:

Compound Initial Final y
i

H
2
(g)

( )

()
()

N
2
(g)

( )

()
()

NH
3
(g) 0

()

( )


Thus, the equilibrium gas phase composition at 723.15 K and 1 bar is given by:

()
)

(
()
()
)(
()
()
)

(


)

()

()


Solving for the conversion of nitrogen yields X = 0.004597 (i.e. only 0.45% of nitrogen will be
converted at atmospheric pressure).

3 Determine the equilibrium conversion of N
2
for stoichiometric feed of H
2
and N
2
at 723.15 K
and 10 bar
The equilibrium constant is not dependent on pressure and thus

()


()


However, the equilibrium gas phase composition is dependent on pressure. At low pressure, the
equilibrium gas phase composition is given by:


*(

()
)

(
()
()
)(
()
()
)

()
)

(
()
()
)(
()
()
)


Thus, at 10 bar

(

()
)

(
()
()
)(
()
()
)



Solving for the conversion of nitrogen X = 0.04413

Thus, increasing the pressure affects the equilibrium gas phase composition for reactions with a
change in the number of moles.

4 Determine the equilibrium conversion of N
2
for a feed containing 90% Ar, 7.5 % H
2
and 2.5% N
2

at 723.15 K and 10 bar
At a first glance, the equilibrium conversion might be similar to one obtained for a total pressure of 1
bar, since the partial pressures of the reactants (N
2
and H
2
) in this system add up to 1 bar. However,
the partial pressure of the inert (Ar) does not remain at 9 bar during the reaction due to the
contraction of the reaction. Setting up a stoichiometric table (including the inert component):





92


Compound Initial Final y
i

H
2
(g)

( )

()
()

N
2
(g)

( )

()
()

NH
3
(g) 0

()

Ar (g)

()

( )

The equilibrium conversion is given by:


*(

()
)

(
()
()
)(
()
()
)

()
)

(
()
()
)(
()
()
)


Thus, at 10 bar

(

()
)

(
()
()
)(
()
()
)


Solving for the conversion of nitrogen yields X = 0.004589 (i.e. only 0.45% of nitrogen will be
converted at atmospheric pressure). The contraction of the chemical reaction (reduction in the
number of moles due to the reaction) results in a (in this case slight) increase in the partial pressure
of the inert component (Ar), resulting in a lower conversion than when the reaction pressure was 1
bar.

5 Determine the equilibrium conversion of N
2
for stoichiometric feed of H
2
and N
2
at 723.15 K
and 100 bar
The equilibrium constant is independent of pressure:

()


()


and the equilibrium gas phase composition is given through the equilibrium constant K
a
:



The activities of the components are given as:

()

( )

()

( )

()

( )

)



The mole fractions are given from the stoichiometric table.


*(

)(

)(


93

()

(())

()

)(



The problem here is that the fugacity coefficients (

* themselves depend on the composition in


the gas phase as well. Hence, we need an iterative procedure to compute the equilibrium
conversion. A reasonable starting value would be (

* . The fugacity coefficients coefficient can


then be predicted using the Peng-Robinson equation of state (here the binary interaction
coefficients have been set equal to zero; application of the binary interaction coefficient as defined
by J.O. Valderama and E.A. Molina, Fluid Phase Equilibria 31 (1986), 209 does not seem to change
the fugacity coefficients substantially.
X
N2
y
NH3
y
N2
y
H2
f
H2
/(y
H2
.
p) f
N2
/(y
N2
.
p) f
NH3
/(y
NH3
.
p)
0.279005 0.162118 0.20947 0.628411 1.0228 1.0356 0.99457
0.288888 0.168831 0.207792 0.623377 1.0229 1.0357 0.99419
0.288888 0.168831 0.207792 0.623377 1.0229 1.0357 0.99419




94

III.4 Multiple gas phase reactions taking place simultaneously
Often, more than one reaction can take place in a system. In this case, various approaches are
possible. The most robust approach is the use of Lagrange multipliers. The Gibbs free energy of the
system must be minimal at equilibrium:

or

or

Furthermore, the elements are conserved in a system with chemical reactions occurring. This results
in constraints for the system. The elemental balance for element C can be written as


and for each element:



A new function can be generated:


A new set of variables is introduced here (
i
, the Lagrange multipliers). This new function must be
minimal at equilibrium as well. Hence, the variation of this of function with respect to N
i
and the
Lagrange multipliers,
i
, must be equal to zero.
(



Realizing that

according to the Gibbs-Duhem equation, this set of C+k


equations simplifies to:
(


The partial molar Gibbs free energy of a compound in the mixture can be expressed in term of the
Gibbs free energy of the pure compound and the activity of the compound
(



The activity of compound j can at low pressure be set equal to


(at higher pressure the fugacity coefficient has to be evaluated during the minimization process)

Thus,
(



Hence, the following non-linear equations have to be solved simultaneously:


95




Alternatively, the Gibbs free energy of the system can be minimized subject to the constraints of the
elemental balances:

))


With the constraints


Finding the minimum in the Gibbs free energy can be obtained using e.g. Solver in Excel.


Example: Multiple gas phase reactions methanol/dimethylether synthesis
The methanol synthesis from a feed containing 28% CO, 4% CO
2
and 68% H
2
at 250
o
C is equilibrium
limited. The following reactions may occur in this system
CO + 2 H
2
CH
3
OH
CO
2
+ 3 H
2
CH
3
OH +H
2
O
CO + H
2
O CO
2
+H
2

To obtain higher conversions per pass, a side reaction is allowed to occur, viz.
2 CH
3
OH CH
3
OCH
3
+ H
2
O
What is the equilibrium conversion of CO with and without the formation of dimethylether for the
reaction at 50 bar and 250
o
C?

The following data can be found for the various components:
Compound A
f
H (298.15K) A
f
G (298.15K)

, (J/mol
.
K)
kJ/mol kJ/mol a b c d
H
2
O(g) -241.8 -228.6 32.218 0.192
.
10
-2
1.055
.
10
-5
-3.593
.
10
-9

CO(g) -110.5 -137.2 28.142 0.167
.
10
-2
0.537
.
10
-5
-2.221
.
10
-9

CO
2
(g) -393.5 -394.4 22.243 5.977
.
10
-2
-3.499
.
10
-5
7.464
.
10
-9

H
2
(g) 0 0 29.088 -0.192
.
10
-2
0.400
.
10
-5
-0.870
.
10
-9

CH
3
OH (g) -200.7 -162.0 19.038 9.146
.
10
-2
-1.218
.
10
-5
-8.034
.
10
-9

CH
3
OCH
3
(g)
1
-184.1 -112.8 15.511 18.690
.
10
-2
-6.379
.
10
-5
-3.174
.
10
-9

1
: Data from Perrys Chemical Engineers Handbook, 7
th
ed.

Step 1: Calculate the Gibbs free energy of the species at 250
o
C and 1 bar with respect to the
elements at their natural state at 298.15 K and 1 bar
The Gibbs free energy of a species with respect to the elements at their natural state at 298.15 K is
given by:
(

)

(


with the enthalpy of a species with respect to the elements at their natural state at 298.15 K:

()

() (

()

()

)
(

) (

( )

)
96


Putting it all together
Compound A
f
H (298.15K) A
f
G (298.15K) G
i
(523.15K)
1

kJ/mol kJ/mol kJ/mol
H
2
O(g) -241.8 -228.6 -210.4
CO(g) -110.5 -137.2 -151.8
CO
2
(g) -393.5 -394.4 -378.8
H
2
(g) 0 0 -1.9
CH
3
OH (g) -200.7 -162.0 -129.8
CH
3
OCH
3
(g) -184.1 -112.8 -61.1
1
Gibbs free energy of the compounds with respect to their elements at 298.15 K and 1 bar in their
natural state

Step 2: Gibbs free energy of the system
The Gibbs free energy of the system is given by

))



Thus for the system excluding dimethylether

))

))

))

))

))
(in this equation the number of moles of H
2
O, CO, CO
2
, H
2
, and CH
3
OH are unknown)


Step 3: Elemental balances
The following elements need to be considered: C, O and H
C-balance


O-balance


H-balance



Taking a basis of 1 mol of CO fed to the system:



Step 3: Setting up the spreadsheet
Finding the minimum Gibbs free energy of the system is a somewhat brute force method fraught
with difficulties. Special attention is required for the initial values of the variable (i.e. the starting
values for the number of moles of water, CO, CO
2
, H
2
, and methanol) and typically a scan of the
possible initial values is required to obtain physically feasible solutions. Furthermore, this method of
solution may result in the location of a local minimum rather than the global minimum on the Gibbs
free energy surface.
97

Feed T 523.15 K
28 1 1.142857
4 0.142857 1.285714
68 2.428571 4.857143
Objective function (to be minimized)
Variables G
system
-143685
N
H2O
0.01 -2164.920751
N
CO
0.4 -56708.25511
N
CO2
0.1 -37485.18161
N
H2
1 12076.60939
N
CH3OH
0.5 -59403.04594
N
CH3OCH3
0
EN
i
2.01 Constraints g
C
-0.142857143
g
O
-0.175714286
g
H
-0.837142857
Results
X
CO
60.0%
X
CO2
30.0%
Y
CH3OH
43.8%



After solving it using Solver:
Results

X
CO
62.4%
X
CO2
4.3%
Y
CH3OH
55.2%


Step 4: Including non-idealities
At 50 bar, the fugacity coefficient is not expected to be equal to 1. The fugacity coefficients (as
calculated using the Peng-Robinson equation of state using a binary interaction coefficient of zero
for all components) can be taken into the calculation of the Gibbs free energy of the system:

))


Feed T 523.15 K
28 1 1.142857
4 0.142857 1.285714
68 2.428571 4.857143
Objective function (to be minimized)
Variables G
system
-1.67E+05
N
H2O
0.006715 -1.47E+03
N
CO
0.343901 -4.91E+04
N
CO2
0.136143 -5.09E+04
N
H2
1.09623 1.33E+04
N
CH3OH
0.662813 -7.86E+04
N
CH3OCH3
0
EN
i
2.245802 Constraints g
C
2.69919E-09
g
O
4.07979E-09
g
H
4.38408E-09
Results
X
CO
65.6%
X
CO2
4.7%
Y
CH3OH
58.0%
-------------------------------------------------------------------
Fugacity calculation using the Peng-Robinson equation of state
Temperature = 523.15 K
Pressure = 50.0000 bar
Component Feed Fugacity (bar) (f/xP)
y
H2O
0.00299 water 0.003 1.33E-01 8.87E-01
y
CO
0.153131 rbon monoxide 0.1531 7.89E+00 1.03E+00
y
CO2
0.060621 arbon dioxide0.0606 2.94E+00 9.71E-01
y
H2
0.488124 ogen (normal)0.4881 2.52E+01 1.03E+00
y
CH3OH
0.295134 methanol 0.2952 1.30E+01 8.80E-01
y
CH3OCH3
0
Compressibility 0.9865 of vapor phase

98

X
CO
65.6%
X
CO2
4.7%
Y
CH3OH
58.0%

The inclusion of the non-ideality through the fugacity coefficients results in an increase in the
methanol yield and a decrease in the CO
2
-conversion. The non-ideality results in a virtual decrease in
the mole fraction of methanol and thus shifting the equilibrium composition in the direction of
methanol.

Step 5: Including the formation of dimethylether
The spreadsheet for the minimization of the Gibbs free energy of the system is given on the
following page. The inclusion of the formation of dimethylether into the system does result in an
increase in the overall CO conversion:
Results

X
CO
97.1%

X
CO2
-128.1%

Y
CH3OH
5.7% Carbon yield
Y
CH3OCH3
63.2% Carbon yield

The yield of methanol plus dimethylether on a carbon basis is ca. 10% higher than the methanol
yield in the case when no dimethylether was formed. This is accompanied by a (strong) negative
conversion of CO
2
(i.e. CO
2
is being formed in the system). The formation of dimethylether in the
system yields water as a co-product, which may react with CO in the water-gas shift reaction yielding
CO
2
. This is undesirable since the valuable CO is converted in the less reactive CO
2
. Hence, the
strategy to increase the methanol yield per pass by co-producing dimethylether can only work, if a
catalyst is found with a relative low activity for the water-gas shift reaction in comparison to its
activity for the methanol synthesis and the dehydration of methanol yielding dimethylether. In that
case, the yield of methanol plus dimethylether on a carbon basis can be as high as 90.6 C-% (this was
calculated by considering CO
2
as an inert, i.e. the number of moles of CO
2
was not optimized during
the minimization of the Gibbs free energy of the system).
99

Feed T 523.15 K
28 1 1.142857
4 0.142857 1.285714
68 2.428571 4.857143
Objective function (to be minimized)
Variables G
system
-1.76E+05
N
H2O
0.178308 -3.65E+04
N
CO
0.029377 -4.49E+03
N
CO2
0.325816 -1.20E+05
N
H2
1.036202 1.29E+04
N
CH3OH
0.065129 -8.35E+03
N
CH3OCH3
0.361268 -1.88E+04
EN
i
1.9961 Constraints g
C
1.88932E-09
g
O
-1.51749E-09
g
H
-3.83498E-09
Results
X
CO
97.1%
X
CO2
-128.1%
Y
CH3OH
5.7% Carbon yield
Y
CH3OCH3
63.2% Carbon yield -------------------------------------------------------------------
Fugacity calculation using the Peng-Robinson equation of state
Temperature = 523.15 K
Pressure = 50.0000 bar
ComponentFeed Fugacity (bar) (f/xP)
y
H2O
0.089328 water 0.0893 3.97E+00 8.88E-01
y
CO
0.014717 rbon monoxide0.0147 7.58E-01 1.03E+00
y
CO2
0.163226 arbon dioxide 0.1632 7.93E+00 9.71E-01
y
H2
0.519113 ogen (normal) 0.5191 2.68E+01 1.03E+00
y
CH3OH
0.032628 methanol 0.0326 1.44E+00 8.81E-01
y
CH3OCH3
0.180987 imethyl ether 0.181 8.27E+00 9.13E-01
Compressibility 0.9835 of vapor phase


Step 6: Confirming the location of the global minimum using Lagrange multipliers
To find the minimum Gibbs free energy, we can also use Lagrange multipliers, and we need to solve
the following set of non-linear equations:



The first step is to find (

for each component and each element


(


Thus, the system in which the dimethylether formation is included, results in 9 non-linear equations:


100



This set of equation can be solved using a non-linear equation solver such as Polymath.
POLYMATH Results
04-24-2012, Rev5.1.225

NLES Solution

Variable Value f(x) Ini Guess
gC -14.627517 5.107E-14 1
gO 49.871504 5.573E-14 1
gH -1.4416895 -6.217E-15 1
NH2O 0.1863727 -4.315E-12 0.1
NCO 0.0286006 -1.866E-09 0.05
NCO2 0.316985 -4.832E-12 0.25
NH2 1.1113426 -1.017E-13 1
NCH3OH 0.0676989 -2.277E-12 0.1
NCH3OCH3 0.3633578 1.643E-12 0.25

NLES Report (safenewt)

Nonlinear equations
[1] f(gC) = NCO+NCO2 +NCH3OH+2*NCH3OCH3 -1.14 = 0
[2] f(gO) = NH2O+NCO+2*NCO2+NCH3OH+NCH3OCH3 -1.28 = 0
[3] f(gH) = NH2O+2*NH2+4*NCH3OH+6*NCH3OCH3 -4.86 = 0
[4] f(NH2O) = GH2O/RT+ln(NH2O/SUMNi*P*phiH2O)+0*gC+1*gO+2*gH = 0
[5] f(NCO) = GCO/RT+ln(NCO/SUMNi*P*phiCO)+1*gC+1*gO+0*gH = 0
[6] f(NCO2) = GCO2/RT+ln(NCO2/SUMNi*P)*phiCO2+1*gC+2*gO+0*gH = 0
[7] f(NH2) = GH2/RT+ln(NH2/SUMNi*P*phiH2)+0*gC+0*gO+2*gH = 0
[8] f(NCH3OH) = GCH3OH/RT+ln(NCH3OH/SUMNi*P*phiCH3OH)+1*gC+1*gO+4*gH = 0
[9] f(NCH3OCH3) = GCH3OCH3/RT+ln(NCH3OCH3/SUMNi*P*phiCH3OCH3)+2*gC+1*gO+6*gH = 0

Explicit equations
[1] RT = 523.15*8.314/1000 [8] GCH3OH = -129.8
[2] P = 50 [9] GCH3OCH3 = -61.1
[3] SUMNi = NH2O+NCO+NCO2+NH2+NCH3OH+NCH3OCH3 [10] phiH2O = 8.8979e-01
[4] GH2O = -210.4 [11] phiCO = 1.0292e+00
[5] GCO = -151.8 [12] phiCO2 = 9.7154e-01
[6] GCO2 = -378.8 [13] phiH2 = 1.0328e+00
[7] GH2 = -1.9 [14] phiCH3OH = 8.8321e-01
[15] phiCH3OCH3 = 9.1517e-01

The resulting equilibrium composition of the gas phase is in agreement with those found from
searching for the minimum in the Gibbs free energy of the system confirming the location of the
minimum (and thus the conclusions made above).





101

III.5 Reactions involving solids
For any reactive system, the Gibbs free energy of reaction strives towards a minimum:

is at a global minimum
The partial molar Gibbs free energy of a compound can be expressed in terms of the molar Gibbs
free energy of the pure component at the same temperature and a pressure of 1 bar and the activity
of that compound:

( )

( ) (

)
The activity is defined as:

()

( )

The fugacity of a component as a solid equals the fugacity of that pure component as a solid:

()

( )

()

( )

The fugacity of a solid is hardly dependent on pressure (unless extremely high pressures are
applied):

()

( )

and thus the partial molar Gibbs free energy is the molar Gibbs free energy of the pure component:

( )

( )


Example: Phase transition in a solid transformation of red PbO into yellow PbO
Lead oxide (PbO) can exist in two allotropic forms, also named red and yellow PbO. The following
data can be found for the various components:
Compound A
f
H (298.15K) A
f
G (298.15K)
kJ/mol kJ/mol
PbO-yellow (s) -217.3 -187.9
PbO-red (s) -219.0 -188.9
Determine at which temperature red-PbO transforms into yellow-PbO.

For a solid-solid reaction, the Gibbs free energy of the system can be expressed as:


For a system containing N moles of Pb with x
yellow
the fraction of yellow-PbO and (1-x
yellow
) the
fraction of red-PbO, the Gibbs free energy of the system is:


Thus, the Gibbs free energy for a reaction involving only solids is minimal if the mole fraction of a
single phase is 0 or 1. In the case of PbO, the Gibbs free energy is minimal at 298.15 K, if all PbO
exists as red-PbO (see Figure 5.1). The difference between the behavior of reacting solids and
reacting gases (and liquids) is their mixing behavior. In the case of gases and liquids, the mixing
contribution towards the Gibbs free energy of the system results in a mixture being favored as the
end product. Solids do (in general) not mix, and hence reactions involving only solids go to
completion or do not proceed at all.

The Gibbs free energy of the system changes with temperature. The temperature at which the Gibbs
free energy of yellow-PbO equals the Gibbs free energy of red-PbO, the phase transition will become
thermodynamically possible. The change in the Gibbs free energy of the substances can be described
with:
(



102


Figure III.5.1: Gibbs free energy of the system yellow-PbO/red-PbO at 298.15 K as a function of the
mole fraction of yellow-PbO in the mixture

Assuming that the heat of reaction is independent of temperature:
(

)
Thus, the transition temperature is given by:
(

) (

)
or



Substituting the values and solving for T T = 724 K
Thus, above 724 K red-PbO can be transformed into yellow-PbO.



Example Formation of a gas by decomposition of a solid transformation of CaCO
3

The decomposition of CaCO
3
yields a solid, CaO, and a gas, CO
2
:
CaCO
3
(s) CaO(s) + CO
2
(g)
Hence, the products and the reactants do not mix either. Thus, this reaction may go to completion (or
proceed not at all) depending on the pressure of CO2. The equilibrium pressure above which the solid
decomposes is called the decomposition pressure. Determine the decomposition pressure of CaCO
3
as
a function of temperature in the range between 298.15 K and 1400 K.
The following data can be found for the various components:
Compound A
f
H (298.15K) A
f
G (298.15K)

, (J/mol
.
K)
kJ/mol kJ/mol a b c d
CO
2
(g) -393.5 -394.4 22.243 5.977
.
10
-2
-3.499
.
10
-5
7.464
.
10
-9

CaCO
3
(s) -1206.9 -1128.8
CaO (s) -635.1 -604.0
The heat capacity for CaCO
3
and CaO (in J/(mol
.
K) are given as:



-190
-189
-188
-187
0 0.2 0.4 0.6 0.8 1
G
s
y
s
t
e
m
(
2
9
8
.
1
5

K
)
/
n
P
b
,

k
J
/
m
o
l
Mole fraction of yellow PbO, x
yellow
103

The Gibbs free energy of the system can be expressed as:


The number of moles of CaCO
3
, CaO, and CO
2
are related. Define Y as the number of moles of CaCO
3

which decomposes:
Moles of CaCO
3


Moles of CaO


Moles of CO
2


The Gibbs free energy of the system is minimum, if the variation of the Gibbs free energy of the
system with respect to the extent of reaction (Y) is zero. Thus, taking into consideration the Gibbs-
Duhem equation:



The partial molar Gibbs free energy
for the solid compounds (CaCO
3
and CaO)


for the gaseous compound CO
2

)
which at low pressure becomes


)

Hence, the minimum of the Gibbs free energy of the system is given by:


)*


)
with



Thus, the system strives to establish a partial pressure of CO
2
equal to:


by decomposing sufficient CaCO
3
to yield the decomposition pressure, or if not sufficient CaCO
3
is
present by decomposing all CaCO
3
, if the partial pressure of CO2 is initially below the decomposition
pressure, or by converting CaO into CaCO
3
if the partial pressure of CO
2
is higher than the
decomposition pressure.

The change in the Gibbs free energy upon reaction can be obtained from:
(


with

() (

) in J/mol

) (

( )

)

104






Example: Decomposition of a gas yielding a solid formation of carbon black
Carbon black is produced by the decomposition of of methane
CH
4
(g) C (s) +2 H
2
(g)
in a reactor maintained at 700
o
C and 1 bar. The equilibrium constant for this reaction at 700
o
C is
7.403 based on the pure component standard states (gaseous CH
4
and H
2
and solid C) Calculate the
equilibrium gas phase composition in the reactor.

The activity based equilibrium constant is:


The activity of solid carbon is 1, and at low pressure the activity can be expressed in terms of the
mole fraction and the total pressure (1 bar):


The fugacity coefficient will be equal to one at these low pressures. The mole fraction can be
obtained from the stoichiometric table:
Compound Initial Final y
i

Gas phase Solid phase
CH
4
(g)

( )

()
()

H
2
(g) 0

()

C (s) 0

( )

Thus, the equilibrium gas phase composition is given by:

()

()()

Solving for X X = 0.806

Thus, the equilibrium gas phase composition will contain y
CH4
= 0.108 and y
H2
=0.892. This means
that in this case a partial conversion of methane can be obtained. The mixing contribution between
T A
rxn
G p
CO2
, bar Integral
298.15 130.4 1.42E-23 0
300 130.1028 2.22E-23 -3.68774
400 114.1008 1.26E-15 -152.112
500 98.24608 5.44E-11 -240.872
600 82.54286 6.51E-08 -299.792
700 66.98582 1E-05 -341.67
800 51.57117 0.000429 -372.9
900 36.29808 0.007821 -397.033
1000 21.1683 0.078387 -416.195
1100 6.185395 0.508475 -431.741
1200 -8.64588 2.378807 -444.569
1300 -23.3201 8.650645 -455.302
1400 -37.8317 25.79707 -464.386
105

the product H
2
and the reactant CH
4
makes this possible. Furthermore, the formation of carbon can
continue in the reactor, if the flow of methane is maintained at such a level that the exit mole
fraction of methane is larger than 0.108.




106

III.6 Non-ionic reactions in the liquid phase
The condition for equilibrium is that the Gibbs free energy for the system is minimal. Hence, the
variation of the Gibbs free energy with respect to the extent of reaction must be equal to zero.
(


With

( )

( ) (

) if the reference state is chosen to be the pure


compound as a liquid. The activity of compound i is then at low pressure (fugacity coefficient equal
to 1) given by

()

( )


Hence, the equilibrium position of the system is given by



The activity based equilibrium constant can be regarded as two constants, one based on the mole
fractions and one based on the activity coefficients:



In the older literature a concentration-based equilibrium constant is often given rather than the
activity based equilibrium constant. The concentration of a component in the mixture and the mole
fraction are linked, since:

; []

([]

([])



Thus, for an ideal solution (K

= 1), the concentration based equilibrium constant can be obtained


from the activity based equilibrium constant from




Example: Liquid phase esterification of methanol and phenol yielding anisole
The esterification of phenol with methanol yielding anisole (methyl phenyl ether) may occur in the
liquid phase.
CH
3
OH(l) +C
6
H
5
OH(l) C
6
H
5
OCH
3
(l) +H
2
O(l)
A liquid mixture containing equi-molar amounts of methanol and phenol is allowed to react at 1 bar
and 298.15 K. What is the expected conversion of methanol? The following data are available:
Compound A
f
H (298.15K) A
f
G (298.15K) T
melt
A
fus
H
kJ/mol kJ/mol K kJ/mol
CH
3
OH(l) -238.7 -166.3
C
6
H
5
OH(s)
1
-165.0 -50.9 314 11.514
C
6
H
5
OCH
3
(l)
2
-114.5 -0.1
H
2
O(l) -285.8 -237.1
1
Phenol in the standard state at 298.15 K is a solid
2
Data from derived from Gibbs free energy as an ideal gas as given in Perrys Chemical Engineers
Handbook
The product mixture will contain four components, i.e. a quartenary system. However, starting from
an equi-molar mixture of methanol and phenol, the mole fraction of methanol and phenol will always
be the same and the mole fraction of water will always be identical to the mole fraction of methyl
phenyl ether. The activity coefficients of the components can then be modeled using UNIFAC.
107

The equilibrium liquid phase composition is given by:



The change in the Gibbs free energy upon reaction taking the compounds as a pure liquid state as
the reference state, is given by:
with

()

()

()

()


We have the Gibbs free energy of formation of all compounds as a liquid with the exception of
phenol. The Gibbs free energy of formation of phenol as a liquid can be calculated from the Gibbs
free energy of phenol as a solid:

()

()

( )
and thus the question reduces to finding the Gibbs free energy for fusion at 298.15 K. We have seen
that neglecting the difference in the heat capacity between solid and liquid phenol

()

) (

) (



Hence,

()

()

()

()

()

()



Thus, the equilibrium constant at 298.15



The composition of the liquid is given by

()

()

()

()


Setting up the stoichiometric table for a reaction mixture starting with an equi-molar mixture:
Compound In Out x
i

CH
3
OH(l) 1 1-X 0.5
.
(1-X)
C
6
H
5
OH(s)
1
1 1-X 0.5
.
(1-X)
C
6
H
5
OCH
3
(l)
2
0 X 0.5
.
X
H
2
O(l) 0 X 0.5
.
X

()

()

()

()

()

()



The activity coefficients are dependent on the mixture composition. Start off the calculation
assuming K

= 1 and then iterate through.




Hence, a conversion of 92.4% can be achieved. This is lower than the one obtained for the ideal
solution (which would result in a conversion of 98.4%), due to the non-ideality of the solution with
K

>1.

Methanol

phenol

water

anisole
K

K
x
X x
Methanol
x
phenol
x
water
x
anisole
1 4035.254 0.984502 0.007749 0.007749 0.492251 0.492251
1.6282 0.161 3.4295 1.9348 25.3124 159.4181 0.926611 0.036694 0.036694 0.463306 0.463306
1.3866 0.1782 3.431 1.9381 26.91151 149.9453 0.924501 0.037749 0.037749 0.462251 0.462251
1.38 0.179 3.4315 1.9378 26.91912 149.9029 0.924491 0.037754 0.037754 0.462246 0.462246
108


Example Combined chemical and phase equilibrium esterification of phenol with methanol
The esterification of phenol with methanol yielding anisole (methyl phenyl ether) may occur at 150
o
C
at a constant pressure of 1.5 bar. What is the equilibrium composition of the mixture starting from
an equi-molar mixture of methanol and phenol?
Data:
Compound p
vap
(373.15K) A
vap
H(T
boil
) T
boil

bar kJ/mol K
CH
3
OH 3.303 35.21 337.8
C
6
H
5
OH 0.055 45.91 455.0
C
6
H
5
OCH
3
0.186 38.97 426.8
H
2
O 1.0135 40.65 373.15

At a temperature of 150
o
C and 1.5 bar, liquid phase and vapour phase may co-exist. Hence, both the
chemical reaction and the vapour-liquid equilibrium need to be considered. The chemical and
vapour-liquid split will ensure that the Gibbs free energy of the system strives towards a minimum:

is at a global minimum
The partial molar Gibbs free energy of a compound can be expressed in terms of the molar Gibbs
free energy of the pure component at the same temperature and a pressure of 1 bar and the activity
of that compound:

( )

( ) (

)
Thus, the Gibbs free energy of the system can be expressed as:



with

( )

( ) (

( ) (

( )

( ) (

( ) (


)

The molar Gibbs free energy of the pure component in the liquid phase and in the vapour phase are
linked:

( )

( )

( )

The change in the Gibbs free energy upon evaporation can be linked to the change in the heat of
vaporization and the difference in the specific heat between the gas phase and the liquid phase:

( )

) (


(with

)
This can be simplified to

( )

) (

), if the
temperature range is rather small or the change in the specific heat is rather small. Thus,

( )

( )

) (

)



Step 1: Calculate the Gibbs free energy of the species in each phase at 150
o
C and 1 bar with respect
to the elements at their natural state at 298.15 K and 1 bar
Gibbs free energy of a species with respect to the elements at their natural state at 298.15 K is given
by:
109

(

)

(


with the enthalpy of a species with respect to the elements at their natural state at 298.15 K. The
contribution of the integral of the specific heat to the enthalpy of the species will be rather small,
since the temperature difference is not too large. Thus, the enthalpy of the species with respect to
the elements at their natural state at 298.15 K can be approximated by the change in the enthalpy
upon formation of the species, which is assumed to be constant:
(

)

(

)

Compound G
i
L

(kJ/mol) G
i
V

(kJ/mol)
kJ/mol kJ/mol
CH
3
OH -135.9 -144.8
C
6
H
5
OH -7.1 -3.9
C
6
H
5
OCH
3
47.9 48.2
H
2
O -216.7 -222.1

Step 2: Setting up the spreadsheet
We can now define the variables (the components in each phase a total of 8 variables; in principle
this number of variables can be reduced by considering that the total number of moles of phenol
must be equal to the number of moles of methanol and similar for the moles of water and anisole),
the function to be minimized (Gibbs free energy of the system), and the constraints (the elemental
balances and the vapour-liquid equilibrium condition). The latter is best defined as


to take into account the large variation possible in the partial pressures of the components.

Step 3: Incorporating non-ideality
The non-ideality of the solution can be taken into account iteratively, i.e. first calculate the
composition of the various phase using ideal solution, and then estimate the activity coefficient in
the liquid phase using e.g. UNIFAC.

Compound Number of moles Mole fraction Activity coefficient Partial molar Gibbs free energy
Liquid Gas Total Liquid Gas Liquid Liquid Vapour
Methanol 8.75312E-05 0.041575438 0.0417 0.0040 0.0425 1.2173 -154673 -154527 -13.5387 -6424.51
Phenol 0.00472131 0.036941659 0.0417 0.2161 0.0378 0.6733 -13846.6 -13951.9 -65.374 -515.408
Anisole 0.016320091 0.44201694 0.4583 0.7471 0.4519 0.9869 46790.33 46827.69 763.6225 20698.63
Water 0.000715628 0.457621403 0.4583 0.0328 0.4678 4.5076 -223412 -223375 -159.88 -102221
0.0218 0.9782 1.0000
Temperature: 423.15 Deg K G
system
-87937.8
-------------------------------
No Name Mole Fraction Activity Coefficient
Elemental balances Vapour-liquid equilibrium -- ------------ ------------- --------------------
C 0 methanol -4.92939E-14 1 methanol 0.0040 1.2173
H 0 phenol -5.38458E-14 2 phenol 0.2160 0.6733
O 0 anisole 1.64313E-14 3 yl phenyl ether 0.7472 0.9869
water 7.10543E-15 4 water 0.0328 4.5076
Iteration
Methanol

phenol

water

anisole
x
Methanol
x
phenol
x
anisole
x
water
y
Methanol
y
phenol
y
anisole
y
water
1 1 1 1 1 0.0046 0.1198 0.7221 0.1535 0.0400 0.0311 0.4426 0.4864
2 1.3414 0.547 1.0436 4.1894 0.0039 0.2599 0.7009 0.0354 0.0451 0.0369 0.4483 0.4697
3 1.1406 0.6993 0.9777 4.2526 0.0042 0.2064 0.7546 0.0348 0.0420 0.0375 0.4522 0.4684
4 1.2341 0.6662 0.9894 4.5483 0.0040 0.2186 0.7450 0.0325 0.0426 0.0378 0.4517 0.4678
5 1.2127 0.6751 0.9862 4.4956 0.0040 0.2155 0.7477 0.0328 0.0425 0.0377 0.4519 0.4679
6 1.2182 0.673 0.987 4.5105 0.0040 0.2162 0.7470 0.0327 0.0425 0.0378 0.4519 0.4678
7 1.2169 0.6735 0.9868 4.5065 0.0040 0.2160 0.7472 0.0328 0.0425 0.0378 0.4519 0.4678
0.0040 0.2161 0.7471 0.0328 0.0425 0.0378 0.4519 0.4678
composition liquid phase Composition vapour phase


110

III.7 Reactions involving ionic species
Quite a lot of systems of interest are aqueous systems containing ions. Chemical equilibria of
interest in these systems are e.g. dissociation of acids, dissolution of salts, etc. The treatment of the
chemical equilibrium condition (i.e. the Gibbs free energy of the system is minimal) is similar to that
for gas- or other type of liquid-phase systems. The difference originates from the definition of the
reference state for the Gibbs free energy. For ionic systems, it is common to take an ideal 1 molal
solution (i.e. activity coefficient equal to 1) as the reference. The partial molar Gibbs free energy of
an ion Z+ is then given in terms of the Gibbs free energy of formation of this ion from its elements as
an ideal 1 molal ionic solution (see Table 13.1-4)

( )

) (


)

The activity coefficient of the individual ions cannot be determined and a mean ionic activity
coefficient has been defined. Thus, if the formation of ions Z
z+
and Y
y-
in the stoichiometric amounts
v
+
and v
-
are considered:

( )

( )

) ((

*
with (

(extended Debye-Hckel model)



The minimization of the Gibbs free energy of the system, in which the dissociation of a salt Z
v+
Y
v-

takes place, yields as an equilibrium condition:

( )

( )

( )

(
(

)
With

( )

( )

( )

or otherwise stated:



(the units for K
C
are molal to the power of v
+
+v
-
-1). For very dilute solutions, the activity coefficients
go to 1 (in this definition of the activity coefficients!) and

( )



Thus,

) (

) (

) (

) (

) (

) (

)

For dilute solution

and the limited Debye-Hckel approximation for the mean ionic


activity coefficient:
(

) (

) (


111

The ionic strength depends on the concentration of ions, and hence the concentration dependent
equilibrium constant is dependent on the concentration of ions in solution!


Example: Dissociation of acetic acid determination of K
a

MacInnes and Shedlovsky (J. Amer. Chem. Soc. 54 (1932), 1429) report the following data for the
ionization of acetic acid in water at 25
o
C:
Acetic acid added, C
T
, mol/liter [CH
3
COO
-
], mol/liter
0.028
.
10
-3
0.1511
.
10
-4

0.1532
.
10
-3
0.4405
.
10
-4

1.0283
.
10
-3
1.273
.
10
-4

2.4140
.
10
-3
2.001
.
10
-4

5.9115
.
10
-3
3.193
.
10
-4

20.000
.
10
-3
5.975
.
10
-4

Calculate A
dis
G for this reaction.

The ionization of acetic acid in water can be seen as:

()

()

()

Stoichiometric table (define I as the amount of acetic acid dissociated)
Compound Initial Final
CH
3
COOH(aq) C
T
C
T
-I
CH
3
COO
-
(aq) - I
H
+
(aq) - I

Thus, the concentration based equilibrium constant is given by:

] [

]
[


but the equilibrium constant depends also on the ionic strength of the solution. The ionic strength of
the solution is given by:

(()

] ()

])
We are dealing here with (extremely) diluted solutions. Thus,
(

) (

) (

)
(

) (

)

Plotting the determined value for K
C
as a function of the square root of the ionic strength in a semi-
logarithmic plot should yield a straight line.
Acetic acid added,
C
T
, mol/liter
[CH
3
COO
-
],
mol/liter
K
C
,
mol/liter
2.80
.
10
-5
1.51
.
10
-4

1.77
.
10
-5

1.53
.
10
-4
4.41
.
10
-4

1.78
.
10
-5

1.03
.
10
-3
1.27
.
10
-4

1.8
.
10
-5

2.41
.
10
-3
2.00
.
10
-4

1.81
.
10
-5

5.91
.
10
-3
3.19
.
10
-4

1.82
.
10
-5

2.00
.
10
-2
5.98
.
10
-4

1.84
.
10
-5


112


Figure III.7.1: Dependency of the concentration-based equilibrium constant on the ionic strength of
the solution

The semi-logarithmic plot of the concentration-based equilibrium constant as a function of the
square root of the ionic strength is reasonable linear with a noticeable deviation at the high
concentration level. A possible explanation is that the limited Debye-Hckel theory starts to fall off
at this stage. In order to ascertain the origin of this deviation more data points would be required. It
can be further noted that the slope of the line is not as expected equal to 2.35, but 2.231. This may
indicate that the activity coefficient of acetic acid deviates from 1. However, the concentration at
the high ionic strength represents still a mole fraction of acetic acid in water close to infinite dilution
(i.e. x
acetic acid
less than 5
.
10
-4
) and hence the activity coefficient of acetic acid in this solution is not
expected to change much and hence can be considered to be equal to 1 taking the ideal solution
with a molality of 1 as a reference.

The intercept of the correlating line represents ln(K
C
0
), which is equal to ln(K
a
). Hence, Ka is
determined to be e
-10.952
= 1.752
.
10
-5
. Hence, the Gibbs free energy for this dissociation reaction is
given by:

) ()





Example: Influence of ionic strength on pH of aqueous acetic acid solutions
The pH of aqueous acetic acid solutions is dependent on the ionic strength of the solution. Hence, the
pH of the solution can be altered by adding NaCl to the solution. Calculate the pH of the solution as a
function of the ionic strength of the solution for solutions with containing initially 10 mmol acetic
acid per liter to which between 0 and 1 mol NaCl per liter is added.

In this ionic strength range, the extended Debye-Hckel needs to be taken. Hence, the
concentration-based equilibrium constant can be determined using:
(

) (

) (

) (

) (


(the stoichiometric factors refer to the ions involved in the dissociation reaction)

The ionic strength of the solution is given by:

(()

] ()

] ()

] ()

])
ln(K
C
) = -10.952+2.231
.
I
0.5
R = 0.997
-10.95
-10.94
-10.93
-10.92
-10.91
-10.9
0 0.01 0.02 0.03
l
n
(
K
C
/
(
m
o
l
/
l
i
t
e
r
)
)
I
0.5
, (mol/liter)
0.5
113


The concentrations can be obtained from a stoichiometric table (define A: initial concentration of
acetic acid; N: initial amount of NaCl in the solution, which dissociates completely; X as the amount
of acetic acid, which has dissociated; and assume :
Compound Initial Final
CH
3
COOH(aq) A A-X
CH
3
COO
-
(aq) - X
H
+
(aq) - X
NaCl(aq) N -
Cl
-
(aq) - N
Na
+
(aq) - N
Thus, the ionic strength is given by

The equilibrium constant (

) (

][

]
[



Putting it all together:
(


)


( )
and thus X can be calculated as a function of N (the initial concentration of NaCl added to the
solution). X represents the concentration of H
+
in the solution, and hence the pH = -log(X). The
addition of NaCl to the solution decreases the pH (i.e. more acetic acid is dissociated; see Fig. 7.2). In
this range of ionic strengths, the mean ionic activity coefficient decreases. The activity based
equilibrium constant remains however constant. Thus, the ratio of the product of the dissociation
products relative to the undissociated acetic acid must increase (with decreasing activity coefficient)
in order to keep K
a
constant.

Figure III.7.2: Dependency of the pH of a 10 mmol/liter acetic acid solution on the amount of NaCl
added to the solution due to change in the ionic strength



Example: Influence of ionic strength on pH of aqueous acetic acid solutions - common ion effect
The pH of aqueous acetic acid solutions is dependent on the ionic strength of the solution and the
type of ions present Hence, the pH of the solution can be altered by adding sodium acetate
(Na(CH
3
COO)) to the solution. Calculate the pH of the solution as a function of the ionic strength of
3
3.1
3.2
3.3
3.4
3.5
0 0.5 1 1.5
p
H
Amount of NaCl added, mol/liter
114

the solution for solutions with containing initially 10 mmol acetic acid per liter to which between 0
and 1 mol Na(CH
3
COO) per liter is added.

The concentrations can be obtained from a stoichiometric table (define A: initial concentration of
acetic acid; N: initial amount of Na(CH
3
COO) in the solution, which dissociates completely; X as the
amount of acetic acid, which has dissociated; and assume :
Compound Initial Final
CH
3
COOH(aq) A A-X
CH
3
COO
-
(aq) - X+N
H
+
(aq) - X
NaCH
3
COOaq) N -
Na
+
(aq) - N
The ionic strength of the solution is given by:

(()

] ()

] ()

])
The equilibrium constant (

) (

][

]
[

()



Putting it all together:
(
( )

)


( )

The pH of the solution increases with increasing amount of sodium acetate added to the solution
(see figure 7.3) indicating a decrease in the concentration of H
+
-ions, and thus indicating a decrease
in the extent of dissociation of acetic acid. This is of course due to the presence of acetate ions in the
solution originating from sodium acetate. Thus, at a constant value of K
C
the increase in the acetate
concentration needs to be off-set by a proportional decrease in the H
+
-concentration. However, this
decrease in the extent of dissociation of acetic acid is moderated by the change in the activity
coefficient with increasing ionic strength (vide supra).


Figure III.7.3: Dependency of the pH of a 10 mmol/liter acetic acid solution on the amount of
Na(CH
3
COO) or NaCl added to the solution common ion effect opposed to the effect
of ionic strength



3
3.5
4
4.5
5
5.5
6
6.5
0 0.5 1 1.5
p
H
Amount of NaCl added, mol/liter
Addition of NaCH
3
COO
Addition of NaCl
115

III.7.1 Solubility of salts
The dissolution of salts is often given in terms of a solubility (g of salt dissolved per 100 cm
3
of water)
or in terms of the solubility product K
sp
. The solubility in terms of g of salt per 100 cm
3
of water can
be easily recalculated in terms of the solubility product taking into account the molar mass of the
salt and assuming that the change in the volume upon dissolving the salt is negligibly small. The
solubility product is commonly defined as:



From a thermodynamic viewpoint, the dissolution of a salt can be represented as a chemical
reaction:

()

()

()

The activity based equilibrium constant is given by:

()
)

()
)

()

but the activity of a solid (at not too a high a pressure) equals 1:

()
)

()
)



The molality of a solution is approximately the concentration in mol/liter in a solution (since the
density of the solution is approximately 1 kg/liter):



Hence,



Similarly to the consideration on ionic solutions, the activity coefficient can be modeled using the
Debye-Hckel theory.


Example: Calculation of solubility product from solubility data influence of ionic strength
The solubility of silver chloride (AgCl) in aqueous solutions of KNO3 has been determined at 25
o
C (S.
Popoff, E.W. Neumann, J. Phys. Chem. 34 (1930), 1853; E.W. Neumann, J. Amer. Chem. Soc. 54
(1932), 2195).
[KNO
3
],
mmol/liter
[AgCl]
saturation
,
mol/liter
0.0 1.273
.
10
-5

0.509 1.311
.
10
-5

9.931 1.427
.
10
-5

16.431 1.469
.
10
-5

40.144 1.552
.
10
-5

(a) Calculate the solubility product of silver chloride in each of these solutions
(b) Calculate the theoretically predicted solubility product of AgCl from thermodynamic data
Data:
A
f
H (298.15 K), kJ/mol A
f
G (298.15 K), kJ/mol
AgCl(s) -127.1 -109.8
Ag
+
(aq) 105.90 77.11
Cl
-
(aq) -167.46 -131.17

(c) Compare the solubility product for AgCl in these solutions with the theoretically predicted
solubility products
116


The solubility product of AgCl is defined as:

] [

]
Assuming complete dissociation of AgCl [Ag
+
] = [Cl
-
] = [AgCl]

[]



[KNO
3
],
mmol/liter
[AgCl]
saturation
,
mol/liter
K
sp
,
(mol/liter)
2

0.0 1.273
.
10
-5

1.62E-10
0.509 1.311
.
10
-5

1.72E-10
9.931 1.427
.
10
-5

2.04E-10
16.431 1.469
.
10
-5

2.16E-10
40.144 1.552
.
10
-5

2.41E-10

The theoretical estimate of the solubility product is given by:

( )


The activity based equilibrium constant is given in terms of the Gibbs free energy of reaction
(dissolution)


with

()

()

()


The activity coefficient can be modeled using the limited Debye-Hckel model seeing the low
concentrations and low ionic strength:

( )


with the ionic strength:

(()

] ()

] ()

] ()

])
[] [

[][

]
)

( )


[KNO
3
],
mmol/liter
[AgCl]
saturation
,
mol/liter
K
sp
,
(mol/liter)
2

I,
mol/liter
K
sp,theoretical
,
(mol/liter)
2

0 1.27E-05
1.62E-10 1.27E-05 1.73E-10
0.509 1.31E-05
1.72E-10 5.22E-04 1.81E-10
9.931 1.43E-05
2.04E-10 9.95E-03 2.17E-10
16.431 1.47E-05
2.16E-10 1.64E-02 2.32E-10
40.144 1.55E-05
2.41E-10 4.02E-02 2.75E-10

The predicted solubility product is between 7 and 14% too large, and thus the estimated solubility is
between 3 and 7% too larger. Hence, the theoretical approach can yield a good insight in the
solubility of a compound, in the absence of experimental data.





117


Example: Common ion effect solubility of AgCl in the presuence of TlCl
The solubility of silver chloride (AgCl) in water at 25
o
C is 1.273
.
10
-5
mol/liter, and that of thallium
chloride (TlCl) is 0.144 mol/litre. Estimate the simultaneous solubility of AgCl and TlCl in water at
25
o
C.

Step 1: Calculate Ka for the dissolution of the pure salts
The activity based equilibrium constant is defined as


For a pure AgCl-solution

[]


For a pure TlCl-solution

[]



Using the extended Debye-Hckel theory (with z
+
=1 and z
-
=-1)
(



For a pure AgCl-solution []


For a pure TlCl-solution []



Hence, the activity coefficient according to the extended Debye-Hckel theory
for a pure AgCl-solution


for a pure TlCl-solution



For the simultaneous solubility of TlCl and AgCl, we need to set-up a stoichiometric table. Define A as
the amount of AgCl dissolved and T as the amount of TlCl dissolved:
Compound Final concentration
Ag
+
(aq) A
Cl
-
(aq) A+T
Tl
+
(aq) T
The variables A and T can be determined from the two equilibrium constants



which have to be solved simultaneously. Writing out in terms of the variables


[

] [

] (

( ) (

] [

] (

( ) (



The ionic strength of the solution is now . A first estimate is that the ionic strength is
dominated by the concentration of thallium chloride and that the solubility of thallium chloride is
unaffected by the presence of silver chloride.


This results in a concentration of Ag
+
: [

](


118

The assumption that the concentration of Ag+ is negligibly small seems justified, and the
concentration of these ions will not contribute to the solubility of TlCl. Hence, the solubility of TlCl is
not affected by the presence of AgCl, but the solubility of AgCl is strongly depressed.

Das könnte Ihnen auch gefallen