Thermodynamics Merged

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 539

Advanced Chemical Engineering

Thermodynamics
(CHC502)

Suman Dutta
Course Objectives and Learning
Outcomes
Course Objectives: The course aims to impart advanced knowledge on
chemical engineering thermodynamics, particularly phase equilibria,
thermodynamics of solution and chemical equilibria including molecular
thermodynamics.

Learning Outcomes: Students will be able to formulate solutions to


phase and chemical equilibria problems for complex systems based on
classical and molecular thermodynamics and to relate thermodynamic
concepts to applications in separation and chemical reactions
encountered in chemical process industries.
List of suggested book
Text Books:
1. Sandler, S.I. (2017). Chemical, Biochemical and Engineering
Thermodynamics. 5th Ed., Wiley.
2. Smith, J. M., van Ness, H. C., and Abbott, M. M. (2004). Introduction
to Chemical Engineering Thermodynamics. 7th Ed., McGraw–Hill.
Reference Books:
1. Prausnitz, J. M., Lichtenthaler, R. N., and de Azevedo, E. G. (1999).
Molecular Thermodynamics of Fluid–Phase Equilibria. Prentice
Hall.
2. Haile, J. M. (1992), Molecular Dynamics Simulation: Elementary
Methods. Wiley.
Topic to be discussed

Review of basic concept:


Introduction to advanced chemical engineering thermodynamics.
Laws of thermodynamics
Laws of thermodynamics

Zeroth Law of thermodynamics: thermal equilibrium


First Law of thermodynamics: conservation of energy
Second Law of thermodynamics: concept of
irreversibility, concept of entropy
Third Law of thermodynamics: absolute value of
entropy
Zeroth Law of thermodynamics

The Zeroth law of thermodynamics states


that if two bodies are individually in thermal
equilibrium with a separate, third body, the
first two are also in thermal equilibrium with
each other.
First Law of thermodynamics

The first law of thermodynamics states that


the change in internal energy of a system
equals the net heat transfer into the system
minus the net work done by the system. In
equation form, the first law of thermodynamics
is ∆U = Q − W.
Second Law of thermodynamics

The Kelvin-Planck Statement: It is impossible to construct a


device which operates on a cycle and produces no other effect
than the transfer of heat from a single body in order to produce
work.

The Clausius Statement: It is impossible to construct a device


which operates on a cycle and produces no other effect than the
transfer of heat from a cooler body to a hotter body.
Carnot's Theorem
Carnot's Theorem: No heat engine operating between two heat
reservoirs can be more efficient than a reversible heat engine operating
between the same two reservoirs.
Corollary 1 of Carnot's Theorem: All reversible heat engines
operating between the same two heat reservoirs must have the same
efficiency.
Corollary 2 of Carnot's Theorem: The efficiency of a reversible heat
engine is a function only of the respective temperatures of the hot and
cold reservoirs. It can be evaluated by replacing the ratio of heat
transfers QL and QH by the ratio of temperatures TL and TH of the
respective heat reservoirs.
Third Law of thermodynamics

The entropy of a pure, perfect crystalline


substance at 0 K is zero.
Thank you
Advanced Chemical Engineering
Thermodynamics
(CHC502)

Suman Dutta
Course Objectives and Learning
Outcomes
Course Objectives: The course aims to impart advanced knowledge on
chemical engineering thermodynamics, particularly phase equilibria,
thermodynamics of solution and chemical equilibria including molecular
thermodynamics.

Learning Outcomes: Students will be able to formulate solutions to


phase and chemical equilibria problems for complex systems based on
classical and molecular thermodynamics and to relate thermodynamic
concepts to applications in separation and chemical reactions
encountered in chemical process industries.
List of suggested book
Text Books:
1. Sandler, S.I. (2017). Chemical, Biochemical and Engineering
Thermodynamics. 5th Ed., Wiley.
2. Smith, J. M., van Ness, H. C., and Abbott, M. M. (2004). Introduction
to Chemical Engineering Thermodynamics. 7th Ed., McGraw–Hill.
Reference Books:
1. Prausnitz, J. M., Lichtenthaler, R. N., and de Azevedo, E. G. (1999).
Molecular Thermodynamics of Fluid–Phase Equilibria. Prentice
Hall.
2. Haile, J. M. (1992), Molecular Dynamics Simulation: Elementary
Methods. Wiley.
Topic to be discussed

Review of basic concept:


Introduction to advanced chemical engineering thermodynamics.
Laws of thermodynamics
Laws of thermodynamics

Zeroth Law of thermodynamics: thermal equilibrium


First Law of thermodynamics: conservation of energy
Second Law of thermodynamics: concept of
irreversibility, concept of entropy
Third Law of thermodynamics: absolute value of
entropy
Zeroth Law of thermodynamics

The Zeroth law of thermodynamics states


that if two bodies are individually in thermal
equilibrium with a separate, third body, the
first two are also in thermal equilibrium with
each other.
First Law of thermodynamics

The first law of thermodynamics states that


the change in internal energy of a system
equals the net heat transfer into the system
minus the net work done by the system. In
equation form, the first law of thermodynamics
is ∆U = Q − W.
Second Law of thermodynamics

The Kelvin-Planck Statement: It is impossible to construct a


device which operates on a cycle and produces no other effect
than the transfer of heat from a single body in order to produce
work.

The Clausius Statement: It is impossible to construct a device


which operates on a cycle and produces no other effect than the
transfer of heat from a cooler body to a hotter body.
Carnot's Theorem
Carnot's Theorem: No heat engine operating between two heat
reservoirs can be more efficient than a reversible heat engine operating
between the same two reservoirs.
Corollary 1 of Carnot's Theorem: All reversible heat engines
operating between the same two heat reservoirs must have the same
efficiency.
Corollary 2 of Carnot's Theorem: The efficiency of a reversible heat
engine is a function only of the respective temperatures of the hot and
cold reservoirs. It can be evaluated by replacing the ratio of heat
transfers QL and QH by the ratio of temperatures TL and TH of the
respective heat reservoirs.
Third Law of thermodynamics

The entropy of a pure, perfect crystalline


substance at 0 K is zero.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Statistical Thermodynamics
Microscopic level/molecular level: most probable state

Classical Thermodynamics
Macroscopic level: bulk average condition
Statistical thermodynamics

Statistical thermodynamics is based on the fundamental assumption that all


possible configurations of a given system, which satisfy the given boundary
conditions such as temperature, volume and number of particles, are equally
likely to occur.

The overall system will therefore be in the statistically most probable


configuration.

The entropy of a system is defined as the logarithm of the number of


possible configurations multiplied with Boltzmann’s constant. While such
definition does not immediately provide insight into the meaning of entropy, it
does provide a straightforward analysis since the number of configurations
can be calculated for any given system.
Statistical thermodynamics

The distribution of a given amount energy E over N identical systems. Or


perhaps better: to determine the distribution of an assembly of N identical
systems over the possible states in which this assembly can find itself, given
that the energy of the assembly is a constant E.

There is a weak interaction between them, so weak that the energy of


interaction can be disregarded, that one can speak of the ‘private’ energy of
every one of them and that the sum of their ‘private’ energies has to equal E.

Erwin Schrödinger; Statistical Thermodynamics; Dover


Classical thermodynamics

Classical thermodynamics provides the same concepts. However,


those were obtained through experimental observation. The
classical analysis is therefore more tangible compared to the
abstract mathematical treatment of the statistical approach.
PVT correlation and Equation of state
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Improved Cubic Equations of State
Equation of states

The van der Waals equation of state (1873) is the simplest and realistic cubic
equation of state

It has two pure component parameters a and b. The parameter a is a measure


of the attractive forces between the molecules, and b is related to the size of
the molecules. Since the van der Waals equation of state is cubic in volume,
three volumes exist for any given temperature and pressure.
Equation of states

Redlich-Kwong Equation of state

The parameters a and b are usually expressed as


Equation of states
Soave-Redlich-Kwong

where ω is the acentric factor.


Equation of states
Peng and Robinson EoS

Peng and Robinson (1976) redefined a(T) as


Equation of states
Beattie-Bridgeman equation of state

The Beattie-Bridgeman equation of state is given by

where
Equation of states
Benedict, Webb and Rubin equation of state

The Benedict, Webb and Rubin equation of state is an empirical


equation written as

It has eight adjustable parameters


Equation of states

Virial Equations of State

The parameters in the equation (B,C,D = ci) are called "virial coefficients".
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Compressibility factor

The deviation from ideal-gas behaviour at a given temperature and


pressure can be accounted for by the introduction of a correction
factor called the compressibility factor Z.

Compressibility factor is the ratio of the actual molar volume of the


gas to the calculated molar volume at the same temperature and
pressure. Z=Vreal/Videal.
Compressibility factor

The ideal gas law is defined as:

and the ideal gas law corrected for non-ideality is defined is:
Compressibility factor
Compressibility factor

There are three regimes that affect the compressibility factor:


1.the value of Z tends toward 1 as the gas pressure approaches 0, where
all gases tend toward ideal behavior
2.the value of Z is less than 1 at intermediate pressures because the
intermolecular forces of attraction cause the actual volumes to be less
than the ideal values
3.the value of Z is greater than 1 and ultimately tends toward infinity at
high pressures because the intermolecular repulsive forces cause the
actual volumes to be greater than the ideal values
Compressibility factor

The gases seem to obey the principle of corresponding states reasonably


well.
The following observations can be made from the generalized
compressibility chart:
1. At low pressures (PR<<1), gases behave as an ideal gas regardless of
temperature.
2. At high temperature (TR>2), ideal gas behavior can be assumed with
good accuracy regardless of pressure (except when PR>>1).
3. The deviation of a gas from ideal gas behavior is greatest in the vicinity
of the critical point
Compressibility factor
we now have an equation for determining Z by using the van
der Waals parameters a and b:

And using the Redlich-Kwong EoS parameters


Compressibility factor
Using Virial equation of state we have Z as given below

The parameters B, C and D are referred to as the second, third and fourth
virial coefficients, respectively. The coefficients are not constants since
they vary from one gas to another as well as the temperature of the gas
under consideration. They are sometimes written as B(T), C(T) and D(T)
to denote that they are functions of temperature. The numerical value of
the coefficients must be determined experimentally.
Compressibility factor

The second virial coefficient provides the largest part of the correction for
the non-ideal behavior of a gas. For that reason, the virial equation is
sometimes truncated after the term containing the second coefficient.
However, when the third coefficient is available, the equation is usually
truncated after the term containing the third coefficient. The fourth
coefficient is rarely available.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Mixing rule

In real life application, most of the fluids are mixture of various


components, not pure one. Therefore, a suitable mixing rule is essential
for estimation of PVT correlation of fluids.

Van der Waal’s mixing rule

Where kij is binary interaction parameter, which is usually recovered by


using the experimental data.
Mixing rule

The composition of natural gas is given in the following table

Composition Tc (°C) Pc (bar)


(%)
Methane CH4 95 -82.6 46.5
Carbon 5 31.2 73.8
CO2
Dioxide

We need to store 1 kg of natural gas at a temperature 40 °C and pressure 20 atm.


Calculate the volume required using van der Waals EOS and mixing rule.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Problem 2: Predict the pressure of nitrogen gas at T = 175 K and v = 0.00375 m3/kg
on the basis of (a) ideal gag equation of state
(b) the van der Waals equation of stare
(c) Beattie-Bridgeman equation of state

Compare the values obtained to the experimentally determined value of 10,000 kPa.

The constants in the Beattie-Bridgeman equation are as follows

A= 102.29
B= 0.05378
c= 4.2 × 104
Solution hints:

Find out the value Tc and Pc of nitrogen

Calculate the value of a and b for nitrogen

Calculate ideal gas pressure

Calculate pressure using van der Walls EoS

Calculate pressure using Beattie-Bridgeman equation of state


Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Intermolecular forces

Thermodynamic properties of any pure substance are determined by


intermolecular forces that operate between the molecules of that
substance. Similarly, thermodynamic properties of a mixture depend
on intermolecular forces that operate between the molecules of the
mixture.

The case of a mixture, however, is necessarily more complicated


because consideration must be given not only to interaction between
molecules belonging to the same component, but also to interaction
between dissimilar molecules.
Intermolecular forces

We can use our knowledge of intermolecular forces in only an


approximate manner to interpret and generalize phase-equilibrium
data. The theory of intermolecular forces gives us no more than a
qualitive, or perhaps semiquantitative, basis for understanding phase
behavior, but even such a limited basis can be useful for understanding
and correlating experimental results.
Intermolecular forces

When a molecule is in the proximity of another, forces of attraction and


repulsion strongly influence its behavior. If there were no forces of attraction,
gases would not condense to form liquids, and in the absence of repulsive
forces, condensed matter would not show resistance to compression. The
configurational properties of matter can be considered as a compromise
between those forces that pull molecules together and those that push them
apart; by configurational properties we mean those properties that depend on
interactions between molecules rather than on the characteristics of isolated
molecules. For example, the energy of vaporization for a liquid is a
configurational property, but the specific heat of a gas at low pressure is not.
Intermolecular forces

There are many different types of intermolecular forces, these forces may
be classified under the following arbitrary but convenient headings:

Electrostatic forces between charged particles (ions) and between


permanent dipoles, quadrupoles, and higher multipoles.

Induction forces between a permanent dipole (or quadrupole) and an


induced dipole, that is, a dipole induced in a molecule with polarizable
electrons.
Intermolecular forces

Forces of attraction (dispersion forces) and repulsion between


nonpolar molecules.

Specific (chemical) forces leading to association and solvation, i.e., to


the formation of loose chemical bonds; hydrogen bonds and charge-
transfer complexes are perhaps the best examples.
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces

For simplicity, we assume that ions, atoms, or molecules are in free


space (vacuum). For electrostatic forces, extension to a medium
other than vacuum is made through introduction of the relative
permeability (or dielectric constant) of the medium.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Electrostatic Forces
Electrostatic Forces
Electrostatic Forces
Electrostatic Forces
Electrostatic Forces
Electrostatic Forces

Electrostatic forces between ions are inversely proportional to the square of the
separation and therefore they have a much longer range than most other
intermolecular forces that depend on higher powers of the reciprocal distance.
These electrostatic forces make the dominant contribution to the configurational
energy of salt crystals and are therefore responsible for the very high melting
points of salts. In addition, the long-range nature of ionic forces is, at least in
part, responsible for the difficulty in constructing a theory of electrolyte
solutions.
Electrostatic Forces
Electrostatic Forces
Electrostatic Forces

Table: Permanent dipole moments


Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Intermolecular forces
Intermolecular forces

Figure 1: Orientation of two dipoles.


Intermolecular forces

If the distance between the dipoles is large compared to and , the potential
energy is

Γ =− 2 cos cos − sin sin cos −

The orientation making the potential energy a maximum is that


corresponding to when the dipoles are in the same straight line, the positive
end of one facing the positive end of the other.
Intermolecular forces
In an assembly of polar molecules, the relative orientations of these molecules
depend on the interplay of two factors: The presence of an electric field set up by
the polar molecules tends to line up the dipoles, whereas the kinetic (thermal)
energy of the molecules tends to toss them about in a random manner.

We expect, therefore, that as the temperature rises, the orientations become more
random until in the limit of very high temperature, the average potential energy
due to polarity becomes vanishingly small. This expectation is confirmed by
experimental evidence; whereas at low and moderate temperatures the behavior of
polar gases is markedly different from that of nonpolar gases, this difference tends
to disappear as the temperature increases. It was shown by Keesom (1922) that at
moderate and high temperatures, orientations leading to negative potential energies
are preferred statistically.
Intermolecular forces
Intermolecular forces
In addition to dipole moments, it is possible for molecules to have quadrupole
moments due to the concentration of electric charge at four separate points in the
molecule. The difference between a molecule having a dipole moment and one
having a linear quadrupole moment is shown schematically.
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Table: Quadrupole moments for selected molecules

A.D. Buckingham, 1967, Adv. Chem. Phys., 12


D.E. Stogryn and A.D. Stogryn, 1966, Mol. Phys., 11: 371
Intermolecular forces
Intermolecular forces

Whereas the scientific literature on dipole moment is extensive, considerably


less is known about quadrupole moment and little work has been done on higher
multipoles such as octapoles and hexadecapoles. The effect of quadrupole
moments on thermodynamics properties is already much less than that of dipole
moments and the effect of higher multipoles is usually negligible. This relative
rank of importance follows because intermolecular forces due to multipoles
higher than dipoles are extremely short range; for dipoles, the average potential
energy is proportional to the sixth power of the inverse distance of separation
and for quadrupoles the average potential energy depends on the tenth power of
the reciprocal distance. For higher multipoles, the exponent is larger.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Intermolecular forces
Polarizability and Induced Dipoles
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces

For molecules with a permanent dipole moment, the potential energy due to
induction is usually small when compared to the potential energy due to
permanent dipoles; and similarly, for molecules with a permanent
quadrupole moment, the induction energy is usually less than that due to
quadrupole-quadrupole interactions.
Intermolecular forces
Table: Average polarizabilities

C.G. Gray and K.E. Gubbins, 1984, Theory of Molecular Fluids, Vol. 1,
Oxford: Clarendon Press.
Intermolecular forces

Intermolecular Forces between Nonpolar Molecules

The concept of polarity has been known for a long time but until about 1930
there was no adequate explanation for the forces acting between nonpolar
molecules. It was very puzzling, for example, why such an obviously nonpolar
molecule as argon should nevertheless show serious deviations from the ideal-
gas laws at moderate pressure. In 1930 it was shown by London that so-called
nonpolar molecules are, in fact, nonpolar only when viewed over a period of
time; if an instantaneous photograph of such a molecule were taken, it would
show that, at a given instant, the oscillations of the electrons about the nucleus
had resulted in distribution of the electron arrangement sufficient to cause a
temporary dipole moment.
Intermolecular forces

This dipole moment, rapidly changing its magnitude and direction, averages
zero over a short period of time; however, these quickly varying dipoles
produce an electric field which then induces dipoles in the surrounding
molecules. The result of this induction is an attractive force called the induced
dipole-induced dipole force.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Intermolecular forces

Intermolecular Forces between Nonpolar Molecules

The concept of polarity has been known for a long time but until about 1930
there was no adequate explanation for the forces acting between nonpolar
molecules. It was very puzzling, for example, why such an obviously nonpolar
molecule as argon should nevertheless show serious deviations from the ideal-
gas laws at moderate pressure. In 1930 it was shown by London that so-called
nonpolar molecules are, in fact, nonpolar only when viewed over a period of
time; if an instantaneous photograph of such a molecule were taken, it would
show that, at a given instant, the oscillations of the electrons about the nucleus
had resulted in distribution of the electron arrangement sufficient to cause a
temporary dipole moment.
Intermolecular forces

This dipole moment, rapidly changing its magnitude and direction, averages
zero over a short period of time; however, these quickly varying dipoles
produce an electric field which then induces dipoles in the surrounding
molecules. The result of this induction is an attractive force called the induced
dipole-induced dipole force.
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces

Equations (3) and (4) give the important result that the potential energy
between nonpolar molecules is independent of temperature and varies
inversely as the sixth power of the distance between them. The attractive
force therefore varies as the reciprocal seventh power. This sharp decline
in attractive force as distance increases explains why it is much easier to
melt or vaporize a nonpolar substance than an ionic one where the
dominant attractive force varies as the reciprocal second power of the
distance of separation.
Intermolecular forces
Intermolecular forces

Equation (6) gives some theoretical basis for the frequently applied
geometric mean rule, which is so often used in equations of state for has
mixtures and in theories of liquid solutions.
Intermolecular forces
Intermolecular forces
Table: Relative magnitude of intermolecular forces between two
identical molecules at 0°C
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Intermolecular forces
Intermolecular forces
Intermolecular forces
Mie’s potential energy function for nonpolar molecules
Intermolecular forces
Intermolecular forces
Intermolecular forces

Figure: Three forms of Mie’s potential for simple, nonpolar molecules.


Intermolecular forces
Intermolecular forces

Mie’s potential applies to two nonpolar, spherically symmetric molecules that


are completely isolated. In nondilute system, and especially in condensed
phases, two molecules are not isolated but have many other molecules in their
vicinity. By introducing appropriate simplifying assumptions, it is possible to
construct a simple theory of dense media using a form of Mie’s two-body
potential such as that of Lennard-Jones.
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces
Intermolecular forces

Structural Effects

Intermolecular forces of nonspherical molecules depend not only on the center-


to-center distance but also on the relative orientation of the molecules. The
effect of molecular shape is most significant at low temperatures and when the
intermolecular distances are small, i.e., especially in the condensed state. For
example, there are significant differences among the boiling points of isomeric
alkanes that have the same carbon number; a branched isomer has a lower
boiling point than a straight chain, and the more numerous the branches, the
lower the boiling point.
Intermolecular forces

Figure: Boiling point (in °C) of some alkane isomers


Intermolecular forces
Intermolecular forces

Specific (Chemical) Forces

In addition to physical intermolecular forces briefly described,


there are specific forces of attraction which lead to the formation of
new molecular species; such forces are called chemical forces. A
good example of such a force is that between ammonia and
hydrogen chloride; in this case, a new species, ammonium chloride,
is formed.
Intermolecular forces
Hydrogen Bond

The most common chemical effect encountered in the thermodynamics


of solutions is that due to the hydrogen bond.

Electron donor-electron acceptor complexes

Chemical effect may also result from other kinds of bonding forces leading
to loose complex formation between electron donors and electron
acceptors, sometimes called charge-transfer complexes.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Molecular theory of corresponding states

Classical or macroscopic theory of corresponding states was derived by van


der Waals based on his well-known equation of state. It can be shown,
however, that van der Waals’ derivation is not tied to a particular equation but
can be applied to any equation of state containing two arbitrary constants in
addition to gas constant R.

From the principle of continuity of gaseous and liquid phases, van der
Waals showed that at the critical point

(1)
Molecular theory of corresponding states

These relations led van der Waals to the general result that for variables v
(volume), T (temperature), and P (pressure) there exists a universal function
such that

(2)

is valid for all substances; subscript c refers to critical point. Another way of
stating this result is to say that, if the equation of state for any one fluid is
written in reduced coordinates (i.e., v/vc, T/Tc, P/Pc), that equation is also
valid for any other fluid.
Molecular theory of corresponding states

Classical theory of corresponding states is based on mathematical properties


of the macroscopic equation of state. Molecular or microscopic theory of
corresponding states, however, is based on mathematical properties of the
potential-energy function.

Intermolecular forces of a number of substances are closely approximated by


the inverse-power potential function. The independent variable in this
potential function is the distance between molecules. When this variable is
made dimensionless, the potential function can be rewritten in a general way
such that the dimensionless potential is a universal function F of the
dimensionless distance of separation between molecules:
Molecular theory of corresponding states

(3)

where ℇi is an energy parameter and σi is a distance parameter characteristic of


the interaction between two molecules of species i. For example, if function F
is given by the Lennard-Jones potential, then ℇi is the energy (times minus one)
at the potential-energy minimum, and σi is the distance corresponding to zero
potential energy. However, equation (3) is not restricted to the Lennard-Jones
potential, nor is it restricted to an inverse-power function. Equation (3) merely
states that the reduced potential energy (Гii/ℇi) is some universal function of the
reduced distance (r/σi).
Molecular theory of corresponding states

Once the potential-energy function of a substance is known, it is possible, at


least in principle, to compute the macroscopic configurational properties of
that substance by the techniques of statistical mechanics. Hence a universal
potential-energy function, equation (3), leads to a universal equation of
state and to universal values for all reduced configurational properties.

To obtain macroscopic thermodynamic properties from statistical


mechanics, it is useful to calculate the canonical partition function of a
system depending on tempertaure, valume, and number of molecules. For
fluids containing small molecules, the partition function Q is expressed as a
product of two factors,
Molecular theory of corresponding states

(4)

where the translational contributions to the energy of the system are


separated from all others, due to other degrees of freedom such as
rotation and vibration. It is assumed that contributions from rotation
and vibration depend only on temperature. These contributions are
called internal because (by assumption) they are independent of the
presence of other near-by molecules.
Molecular theory of corresponding states

In the classical approximation, the translational partition function, Qtrans,


splits into a product of two factors, one arising from the kinetic energy and
the other from the potential energy. For a one-component system of N
molecules, Qtrans is given by

(5)
Molecular theory of corresponding states

where m is the molecular mass, k is Boltzmann’s constant, h is Planck’s


constant, and Гt(r1, ….rN) is the potential energy of the entire system of
N molecules whose positions are described by vectors r1, ….., rN. For a
given number of molecules and known molecular mass, the first factor
depends only on the temperature. The second factor, called the
configurational integral, ZN, depends on temperature and volume:

(6)
Molecular theory of corresponding states

Hence the configurational part provides the only contribution that depends
on intermolecular forces. However, ZN is not unity for an ideal gas (Гt =
0). For an ideal gas, ZNid = VN.

The equation of state is obtained from Q.

(7)
Molecular theory of corresponding states

The equation of state depends only on ZN when equation (4) and (5) are valid.
Therefore, the main problem in applying statistical mechanics to real fluids
lies in the evaluation of the configurational partition function.

There are four assumptions that lead to the molecular theorem of


corresponding states, clearly stated by Pitzer (1939) and Guggenheim (1945).
They are:
Molecular theory of corresponding states

1. The partition function is factored according to equation (4), where Qint is


independent of the volume per molecule.
2. The classical approximation equation (5) is used for Qtrans.
3. The potential energy Гt is represented as the sum of the interactions
Гij(rij) of all possible pairs of molecules. For a given ij pair, Гij depends
only on the distance rij between them:

(8)

4. The potential energy of a pair of molecules, reduced by a characteristic


energy, is represented as a universal function of the intermolecular
distance, reduced by a characteristic length, i.e., equation (3).
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Molecular theory of corresponding states

(1)
Molecular theory of corresponding states

(2)
Molecular theory of corresponding states

Because configurational Helmholtz energy is given by

(3)

(4)
Molecular theory of corresponding states

Equations (3) and (4) imply that the configurational integral must be of the form

(5)

(6)
Molecular theory of corresponding states

Introducing the reduced variables

(7)

(8)

(9)
Molecular theory of corresponding states

Equation (8) expresses the molecular (or microscopic) theory of


corresponding states. This theory is analogous to the macroscopic theory
of corresponding states; the difference lies in the reducing parameters.
Molecular theory of corresponding states
Molecular theory of corresponding states

The connection between the macroscopic and the microscopic theories


of corresponding states can be established by substituting equation (8)
into the relations given by

(10)

(11)
Molecular theory of corresponding states

(12)

(13)
Molecular theory of corresponding states

For that particular case, we have, approximately,

(14)
Molecular theory of corresponding states

An important advantage of the molecular theory, relative to the classical


theory of corresponding states, is that the former permits calculation of
other macroscopic properties (e.g. transport properties) in addition to
those which may be calculated by classical thermodynamics from an
equation of state. For the purposes of phase-equilibrium
thermodynamics, however, the main advantage of the molecular theory is
that it can be meaningfully extended to mixtures, thereby providing some
aid in typical phase-equilibrium problems.
Molecular theory of corresponding states

(15)

and on the basis of a hard-sphere model for molecules interaction,

(16)

Equations (15) and (16) provide a basis for obtaining some


properties of a variety of mixtures.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
The Maxwell relations

A number of second derivatives of the fundamental relation have clear


physical significance and can be measured experimentally. For example:
Isothermal compressibility

The property of the energy (or entropy) as being a differential function


of its variables gives rise to a number of relations between the second
derivatives, e. g. :
The Maxwell relations

is equivalent to:

Since this relation is not intuitively obvious, it may be regarded as a success


of thermodynamics.

Given the fact that we can write down the fundamental relation employing
various thermodynamic potentials such as F, H, G, … the number of second
derivative is large. However, the Maxwell relations reduce the number of
independent second derivatives.
The Maxwell relations
Our goal is to learn how to obtain the relations among the second derivatives
without memorizing a lot of information.

For a simple one-component system described by the fundamental relation


U(S,V,N), there are nine second derivatives, only six of them are independent
from each other. In general, if a thermodynamic system has n independent
coordinates, there are n(n+1)/2 independent second derivatives.
The Maxwell relations
The Maxwell relations: a single-component system

For a single-component system with conserved mole number, N, there are


only three independent second derivatives:

Other second derivatives, such as

can all be expressed via the above three, which are trivially related to
three measurable physical parameters
The Maxwell relations
The Maxwell relations

A general procedure for reducing any derivative to a combination of α, κT,


and cP is possible; it is tedious but straightforward.
The Maxwell relations
Stability of thermodynamic systems

The entropy maximum principle states that in equilibrium:

Let us consider what restrictions these two conditions imply for the functional
form of the dependence of S on extensive parameters of a thermodynamic system.
The Maxwell relations
Stability of thermodynamic systems

Let us consider two identical systems with the following dependence of entropy
on energy for each of the systems:
The Maxwell relations

Stability of thermodynamic systems

Consider a homogeneous system with the fundamental relation shown below


and divide it on two equal sub-systems

Initially both sub-systems have energy U0 and entropy S0. Let us consider what
happens to the total entropy of the system if some energy ∆U is transferred
from one sub-system to another. In this case, the entropy of the composite
system will be S(U0 + ∆U)+ S(U0 - ∆U) > 2S0. Therefore the entropy of the
system increases and the energy will flow from one sub-system to the other,
creating a temperature difference. Therefore, the initial homogeneous state of
the system is not the equilibrium state.
The Maxwell relations
Phase Separation

If the homogeneous state of the system is not the equilibrium state, the system
will spontaneously become inhomogeneous, or will separate into phases.
Phases are different states of a system that have different macroscopic
parameters (e. g. density).

Phase separation takes place in phase transitions. When a system undergoes a


transition from one macroscopically different state to another it goes through a
region of phase separation (e. g. water-to-vapor phase transition).
The Maxwell relations
Stability of thermodynamic systems

To have a stable system, the following condition has to be met:

For infinitesimal dU, the above condition reduces to:

Similar argument applies to spontaneous variations of any extensive


variable, e. g. V:
The Maxwell relations

Fundamental relations of this shape are frequently obtained from statistical


mechanics or purely thermodynamic models (e. g. van der Waals fluid). These
fundamental relations are called the underlying fundamental relations. These
relations carry information on system stability and possible phase transitions.
The Maxwell relations
Stability: higher dimensions

If we consider simultaneous fluctuations of energy and volume, the stability


condition is that:

for any values of ∆U and ∆V.

Local stability criteria are now a little more complex:

Determinant of Hessian matrix


The Maxwell relations
Stability conditions in the energy representation

From the fact that the internal energy of the system should be at minimum
in equilibrium, we can derive the condition for stability in the energy
representation:

The local stability criteria in the energy representation are:


The Maxwell relations
Stability conditions for other thermodynamic potentials

Equations used in Legendre transformations:

Here P is a generalized intensive parameter conjugate to the extensive


parameter X and U(P) is the Legendre transform of U(X).
The Maxwell relations

Therefore, the sign of the second derivative of the Legendre transform of


the internal energy is opposite to that of the internal energy.
The Maxwell relations
The Maxwell relations
Physical consequences of stability conditions

Stability conditions result in limitations on the possible values assumed


by physically measurable quantities.

The molar heat capacity at constant volume must be positive.


The Maxwell relations
Physical consequences of stability conditions

Any of the just written stability conditions generates a constraint on


some physical quantity, e. g.:

The isothermal compressibility of a stable thermodynamic system is


non-negative.
The Maxwell relations
Physical consequences of stability conditions

From the Maxwell relations we can also derive:


Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
One component phase diagram

Complete P-T diagram for pure-component systems


One component phase diagram
One component phase diagram
Gibbs Phase Rule

The phase rule is given by the following equation

F=C–P+2

Where
F is degree of freedom
C is the number of components
P is the number of phases
One component phase diagram

P-V diagram for two-phase system showing isotherms


One component phase diagram

Specific volumes at constant temperature and states within the vapor dome in a
liquid-vapor system
One component phase diagram

The states a and c denote the conditions at which all the fluid is in the
liquid state and the gaseous state respectively.
The specific volumes corresponding to these states are

vf = specific volume of liquid phase


vg = specific volume of gas phase
Binary system phase diagram
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Theory of solution

For calculating fugacities from volumetric data, we need

(2)

and

(3)
Theory of solution
Theory of solution

The mathematical relation between volume, pressure, temperature, and


composition is called the equation of state and most forms of the equation
of state are pressure explicit. Therefore, equation (3) is frequently more
useful than equation (2). At low or moderate densities, it is often possible to
describe volumetric properties of a gaseous mixture in a volume explicit
form; in that case equation (2) can be used. At high densities, however,
volumetric properties are much better represented in pressure explicit form,
requiring the use of equation (3).
Theory of solution

Equation (2) and (3) are exact and if the information needed to evaluate
the integrals is at hand, then the fugacity coefficient can be evaluated
exactly. The problem of calculating fugacities in the gas phase, therefore,
is equivalent to the problem of estimating volumetric properties.
Techniques for estimating such properties must come not from
thermodynamics, but rather from molecular physics.
Theory of solution

The Lewis fugacity rule is a particularly simple equation and is


therefore widely used for evaluating fugacities of components in gas
mixture. However, it is not reliable because it is based on the severe
simplification introduced by Amagat’s law.
Theory of solution

The assumption on which the rule rests states that at constant


temperature and pressure, the molar volume of the mixture is a
linear function of the mole fraction.

The fugacity of component i in a gas mixture can be related to the


fugacity of pure gas i at the same temperature and pressure by the
exact relation

(8)
Theory of solution
Theory of solution

The Lewis rule assumes that at constant temperature and pressure, the
fugacity coefficient of i is independent of the composition of the mixture
and is independent of the nature of the other components in the mixture.
These are drastic assumptions. On the basis of our knowledge of
intermolecular forces, we recognize that for component i, deviations from
ideal-gas behavior depend not only on temperature and pressure, but also
on the relative amount on the chemical nature of these other components
that interact with component i.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
The virial equation of state

To calculate with confidence fugacities in a gas mixture, it is advantageous to


use an equation of state where the parameters have physical significance, i.e.
where the parameters can be related directly to intermolecular forces. One
equation of state that possesses this desirable ability is the virial equation of
state.
The virial equation of state

(1)
The virial equation of state

The compressibility factor is sometimes written as a power series in the


pressure:

(2)
The virial equation of state

Relations between the coefficients in (1) and (2) are given below

(3)

(4)

(5)
The virial equation of state

(6)
The virial equation of state

(7)

Reduction of P-V-T data to yield second and third coefficients is illustrated


in Figure, taken from the work of Douslin (1962) on methane. The
coordinates of the figure follow from rewriting the virial equation in the
form

(8)
The virial equation of state

Figure: Reduction of P-V-T data for methane to yield second and third virial coefficients
The virial equation of state
The virial equation of state

For higher densities, the virial equation is little practical interest. Experimental as
well as theoretical methods are not sufficiently developed to obtain useful
quantitative results for fourth and higher virial coefficients. The virial equation is,
however, applicable to moderate densities as commonly encountered in many
typical vapor-liquid and vapor-solid equilibria.

The significance of the virial coefficients lies in their direct relation to


intermolecular forces. In an ideal gas, the molecules exert no forces on one another.
in the real world, no ideal gas exists, but when the mean distance between
molecules becomes very large (low density), all gases tend to behave as ideal
gases. This is not surprising because intermolecular forces diminish rapidly with
increasing intermolecular distance and therefore forces between molecules at low
density are weak.
The virial equation of state

However, as density rises, molecules come into closer proximity with one
another and, as a result, interact more frequently. The purpose of the virial
coefficients is to take these interactions into account. The physical
significance of the second virial coefficient is that it takes into account
deviations from ideal behavior that result from interactions between two
molecules. Similarly, the third virial coefficient takes into account deviations
from ideal behavior that result from the interaction of three molecules. The
physical significance of each higher virial coefficient follows in an analogous
manner.
The virial equation of state

(9)
The virial equation of state

and

(10)
The virial equation of state

Similar expressions can be written for the fourth and higher virial coefficients.
While equations (9) and (10) refer to simple, spherically symmetric molecules, we
do not imply that the virial equation is applicable only to such molecules; rather, it
is valid for essentially all stable, uncharged (electrically neutral) molecules, polar
or nonpolar, including those with complex molecular structure. However, in a
complex molecule the intermolecular potential depends not only on the distance
between molecular centers but also on the spatial geometry of the separate
molecules and on their relative orientation. In such cases, it is possible to relate the
virial coefficients to the intermolecular potential, but the mathematical expressions
corresponding to equation (9) and (10) are necessarily more complicated.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
The virial equation of state
Extension to mixtures
The most important advantage of the virial equation of state for application
to phase equilibrium problems lies in its direct extension to mixtures. This
extension requires no arbitrary assumptions. The composition dependencies
of all virial coefficients are given by a generalization of the statistical-
mechanical derivation used to derive the virial equation for pure gases.
The virial equation of state

(1)
The virial equation of state

(2)
The virial equation of state

(3)
The virial equation of state

(4)

(5)
The virial equation of state
The virial equation of state

(6)

where
The virial equation of state
The virial equation of state

If experimental data are available for several compositions, the cross


coefficients can be obtained from

(7)

and

(8)
The virial equation of state

Equation (3) and (5) are rigorous results from statistical mechanics and are
not subject to any assumptions other than those upon which the virial
equation itself is based. The physical significance of equation (3) and (5) is
not difficult to understand because these equations are a logical
consequence of the physical significance of the individual virial
coefficients; each of the individual virial coefficients describes a particular
interaction and the virial coefficient of the mixture is a summation of the
individual virial coefficients, apparently weighted with respect to
composition.
The virial equation of state

Extension of the virial equation to mixture is based on theoretical rather


than empirical grounds and it is this feature of the virial equation that
makes it useful for phase-equilibrium problems. Empirical equations of
state that contain constants having only empirical significance are useful
for pure components but cannot be extended to mixtures without the use
of somewhat arbitrary mixing rules for combining the constants.
Extension of the virial equation to mixtures, however, follows in a simple
and rigorous way from the theoretical nature of that equation.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of gases in liquids

(1)
Solubilities of gases in liquids
The ideal solubility of a gas

The simplest way to reduce equation (1) to a more useful form is to rewrite it
in a manner suggested by Raoult’s law. In doing so, we introduce several
drastic but convenient assumptions. Neglecting all gas-phase nonidealities as
well as the effect of pressure on the condensed phase (Poynting correction),
and also neglecting any nonidealities due to solute-solvent interactions, the
equation of equilibrium can be much simplified by writing

(2)
Solubilities of gases in liquids
Solubilities of gases in liquids

Figure: Convenient but arbitrary extrapolation of liquid saturation


pressure into the hypothetical liquid region
Solubilities of gases in liquids

The ideal solubility, as calculated by equation (2) and the extrapolation scheme
indicated in figure, usually gives correct order-of-magnitude results provided
the partial pressure of the gas is not large and provided the solution temperature
is well below the critical temperature of the solvent and not excessively above
the critical temperature of the gaseous solute. In some cases, where the physical
properties of solute and solvent are similar (e.g., chlorine in carbon
tetrachloride), the ideal solubility is remarkably close to the experimental value.

Table 1 compares ideal and observed solubilities of four gases in a number of


solvents at 25°C and 1.013 bar partial gas pressure. The ideal solubility is
significantly different from observed solubilities, but it is of the right order of
magnitude.
Solubilities of gases in liquids

Table 1: Solubilities (mole fraction × 104) of gases in several liquid


solvents at 25°C and 1.013 bar pressure.
Solubilities of gases in liquids

The ideal solubility given by equation (2) suffers from two serious defects.
First, it is independent of the nature of the solvent; equation (2) says that a
given gas, at a fixed temperature and partial pressure, has the same
solubility in all solvents. This conclusion is contrary to observations as
illustrated by the data in Table 1. Second, equation (2), coupled with the
extrapolation scheme shown in Figure 1, predicts that at constant partial
pressure, the solubility of a gas always decreases with temperature.
Solubilities of gases in liquids

This prediction is frequently correct but not always; near room temperature
the solubilities of the light gases helium, hydrogen, and neon increase with
rising temperature in most solvents, and at somewhat higher temperatures
the solubilities of gases like nitrogen, oxygen, argon, and methane also
increase with rising temperature in many common solvents. Because of
these two defects, the ideal-solubility equation has severely limited
applicability, it should be used only whenever no more is desired than a
rough estimate of gas solubility.
Solubilities of gases in liquids
Henry’s law and its thermodynamic significance

It was observed many years ago that the solubility of a gas in a liquid is often
proportional to its partial pressure in the gas phase, provided that the partial
pressure is not large. The equation that describes this observation is commonly
known as Henry’s law:

(3)
Solubilities of gases in liquids
Just how small the partial pressure and solubility have to be for equation (3) to
hold, varies from one system to another, and the reasons for this variation will
become apparent later. In general, however, as a rough rule for many common
systems, the partial pressure should not exceed 5 or 10 bar and the solubility
should not exceed about 3 mol %; however, in those systems where solute and
solvent are chemically highly dissimilar (e.g., systems containing helium or
hydrogen) large deviations from equation (3) are frequently observed at much
lower solubilities.

On the other hand, in some systems (e.g., carbon dioxide/benzene near room
temperature), equation (3) appears to hold to large partial pressures and
solubilities, but such cases are rare; the apparent validity of equation (3) at
large solubilities is usually fortuitous due to cancellation of two (or more)
factors that, taken separately, would cause that equation to fail.
Solubilities of gases in liquids

The assumptions leading to equation (3) can readily be recognized by


comparing it with equation (1). A comparison of the left-hand sides
shows that in Henry’s law the gas phase is assumed to be ideal and thus
the fugacity is replaced by the partial pressure. A comparison of the
right-hand sides shows that the fugacity in the liquid phase is assumed to
be proportional to the mole fraction and that the constant of
proportionality is taken as an empirical factor that depends on the
natures of solute and solvent and on the temperature.
Solubilities of gases in liquids

(4)

Thus

(5)

where 1 stands for solvent and 2 stands for solute.


Solubilities of gases in liquids
Solubilities of gases in liquids

(6)
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of gases in liquids
Solubilities of gases in liquids
Solubilities of gases in liquids

(7)

Table 2 presents experimental Henry’s constants for four gases in ethylene


oxide at three temperatures. These results were calculated from experimental
solubility data and from volumetric data shown in Table 3. The second virial
coefficients are needed to calculate vapor-phase fugacity coefficients and the
liquid-phase partial molar volumes of the solutes at infinite dilution are needed
to correct for the effect of pressure.
Solubilities of gases in liquids

Table 2: Henry’s constants (bar) for four gases in ethylene oxide*


Solubilities of gases in liquids
Table 3: Some volumetric properties of four ethylene oxide(1)/gas(2) systems*
Solubilities of gases in liquids
Effect of pressure on gas solubility

(8)
Solubilities of gases in liquids

The thermodynamic definition of Henry’s constant is

(9)

Substitution of equation (9) into equation (8) gives

(10)
Solubilities of gases in liquids

(11)
Solubilities of gases in liquids

(12)

Equation (12) is the Krichevsky-Kasarrnovsky equation (1935), although its


first clear derivation was given by Dodge and Newton (1937). This equation
is remarkably useful for representing solubilities of sparingly soluble gases
to very high pressures.
Solubilities of gases in liquids
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of gases in liquids

Variation of the activity coefficient of the solute with mole fraction can be
taken into account by a suitable method. In the simplest case we may assume
that the activity coefficient of the solvent is given by a two-suffix Margules
equation:

(13)
Solubilities of gases in liquids

(14)

(15)
Solubilities of gases in liquids

and instead of equation (12) we obtain:

(16)

Equation (16) assume that the partial molar volume of the solute is
independent of pressure and composition over the pressure and composition
ranges under consideration.

Equation (16) is the Krichevsky-Ilinskaya equation (1945). Because of the


additional parameter, it has a wider applicability than does equation (12). It is
especially useful for solutions of light gases (such as helium and hydrogen) in
liquid solvents where the solubility is appreciable.
Solubilities of gases in liquids

For example, Orentlicher (1964) found that equation (16) could be used to
correlate solubility data for hydrogen in a variety of solvents at low
temperatures and pressures to about 100 bar.

Table 4 gives parameters reported by Orentlicher. In the systems studied,


the solubility of hydrogen may be as large as 20 mol % and therefore the
data could not be adequately represented by the simpler Krichevsky-
Kasarnovsky equation.
Solubilities of gases in liquids
Table 4: Thermodynamic parameters for correlating hydrogen solubilities.*
Solubilities of gases in liquids
Solubilities of gases in liquids

(17)

Comparison with equation (16) shows that

(18)

(19)
Solubilities of gases in liquids

and

(20)

(21)
Solubilities of gases in liquids

Second, from elementary calculus,

(22)

(23)
Solubilities of gases in liquids
Solubilities of gases in liquids

(24)
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of gases in liquids
Effect of temperature on gas solubility

The temperature derivative of the solubility, as calculated from the


Gibbs-Helmholtz equation, is directly related to either the partial molar
enthalpy or the partial molar entropy of the gaseous solute in the liquid
phase. Therefore, if something can be said about the enthalpy or entropy
change of solution, insight can be gained on the effect of temperature on
solubility. A general derivation of the thermodynamic relations is given
by Hildebrand and Scott, 1962; and Sherwood, 1962.
Solubilities of gases in liquids

Here we consider only the relatively simple case where the solvent is
essentially nonvolatile and where the solubility is sufficiently small to
make the activity coefficient of the solute independent of the mole
fraction. With these restrictions, it can be shown that

(25)

and

(26)
Solubilities of gases in liquids

First, we consider equation (26); if the partial molar entropy change of the
solute is positive, then the solubility increases with rising temperature;
otherwise, it falls. To understand the significance of the entropy change, it is
convenient to divide it into two parts:

(27)
Solubilities of gases in liquids
The first term on the right-hand side of equation (27) is (essentially) the entropy of
condensation of the pure gas and, in general, we expect this term to be negative
because the entropy (disorder) of a liquid is lower than that of a saturated gas at
the same temperature. The second term is the partial molar entropy of solution of
the condensed solute and, assuming ideal entropy of mixing for the two liquids,
we can write

(28)
Solubilities of gases in liquids
Solubilities of gases in liquids
Solubilities of gases in liquids
Solubilities of gases in liquids

Some qualitative insight into the effect of temperature on gas solubility


can also be obtained from the partial molar enthalpy change (equation
25). It is useful to divide this change into two parts:

(29)
Solubilities of gases in liquids

The first term in equation (29) is (essentially) the enthalpy of condensation


of pure solute and, because the enthalpy of a liquid is generally lower than
that of a gas at the same temperature, we expect this quantity to be negative.
The second quantity is the partial enthalpy of mixing for the liquid solute; in
the absence of solvation between solute and solvent, this quantity tends to be
positive (endothermic) and the theory of regular solutions tells us that the
larger the difference between the cohesive energy density of the solute and
that of the solvent, the larger the enthalpy of mixing.
Solubilities of gases in liquids

If there are specific chemical interactions between solute and solvent (e.g.,
ammonia and water), then both terms in equation (29) are negative
(exothermic) and the solubility decreases rapidly as the temperature rises.
Solubilities of gases in liquids

The effect of temperature on solubility is sensitive to the intermolecular


forces of the solute-solvent system. When the partial pressure of the solute
is small, the solubility typically decreases with temperature, goes through a
minimum, and then rises. To illustrate, Figure 8 shows the solubility of
methane in n-heptane over a wide temperature range. For most common
systems the temperature corresponding to minimum solubility lies well
above room temperature but for light gases, minimum solubility is often
observed at low temperature.
Solubilities of gases in liquids

A thorough study of the solubilities of simple gases in water from 0 to 50°C


has been reported by Benson and Krause (1976). Theoretical and empirical
evidence show that, although other expressions may provide reasonable
approximations, the effect of temperature on Henry’s constant over narrow
temperature ranges is best given by an equation of the form

(30)
Solubilities of gases in liquids

Figure 8: Solubility of methane in n-heptane when the vapor-phase


fugacity of methane is 0.01 bar.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of gases in liquids
Effect of temperature on gas solubility

Equation (30) gives good results for the partial entropy and enthalpy of
solution and, most impressive, for the partial heat capacity of the solute.
Solubilities of gases in liquids
Solubilities of gases in liquids

Over a wider range of temperatures, simple equations such as equation


(30) are unable to describe Henry’s constant. Harvey (1996) developed
a semiempirical correlation of Henry’s constants over large
temperature ranges. To illustrate, Figure 9 shows Henry’s constants for
six nonpolar gases in water as a function of temperature obtained from
Harvey’s correlation. In addition to the maximum (corresponding to
the minimum solubility), a striking feature is the significant decrease in
Henry’s constant as the critical temperature of water (647.1 K) is
approached.
Solubilities of gases in liquids
Solubilities of gases in liquids
Solubilities of gases in liquids

Japas and Levelt Sengers (1989) derived the correct functional form for
this divergence. Near the solvent’s critical point, a function of Henry’s
constant is linear in density:

(31)
Solubilities of gases in liquids

(32)
Solubilities of gases in liquids

(33)

While equations (31) and (32) are only asymptotic results, they describe
experimental data over a wide range of conditions. Figure 10 shows
Henry’s constant data for several solutes in water plotted according to
equation (31). The data display striking linearity (more than one has a
right to expect from a result derived only near the critical point) from
near-critical temperature to approximately 100°C.
Solubilities of gases in liquids

Figure 10: Henry’s constants for several gases in water plotted according
to equation (31).
Solubilities of gases in liquids

Although we do not fully understand the reason for such extended linear
behavior, it can be used to developed correlations, for example, Henry’s
constants (Harvey and Levelt Sengers, 1990; Harvey, 1996) and for
infinite-dilution partition coefficients (Alvarez et al., 1994). Because these
correlations are “anchored” with the correct near-critical functional form,
they can be extrapolated to high temperatures with more confidence than
empirical fitting equations. These results demonstrate how we can
improve correlations by choosing the proper independent variable (here
solvent density) and making reasonable use of theoretical boundary
conditions.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of solids in liquids
Calculation of the pure-solute fugacity ratio

(5)

To simplify notation, let

and let
Solubilities of solids in liquids

These two fugacities depend only on the properties of the solute (component
2); they are independent of the nature of the solvent. The ratio of these two
fugacities can readily be calculated by the thermodynamic cycle indicated in
Figure 2. The molar Gibbs energy change for component 2 in going from a to
d is related to the fugacities of solid and subcooled liquid by

(6)

where, for simplicity, subscript 2 has been omitted. This Gibbs energy change
is also related to the corresponding enthalpy and entropy changes by

(7)
Solubilities of solids in liquids

Figure 2: Thermodynamic cycle for calculating the fugacity of a pure subcooled liquid.
Solubilities of solids in liquids

(8)
Solubilities of solids in liquids

(9)

(10)

Which becomes

(11)
Solubilities of solids in liquids

(12)

(13)
Solubilities of solids in liquids
Solubilities of solids in liquids

Figure 3: Fugacity ratio for solid and subcooled liquid carbon dioxide.
Solubilities of solids in liquids
Ideal solubility

(14)

Equation (14) provides a reasonable method for estimating solubilities of solids in


liquids where the chemical nature of the solute is similar to that of the solvent.
Solubilities of solids in liquids
Equation (14) immediately leads to useful conclusions concerning the
solubilities of solids in liquids. Strictly, these conclusions apply only to ideal
solutions but they are useful guides for other solutions that do not deviate
excessively from ideal behavior:

• For a given solid/solvent system, the solubility increases with rising


temperature. The rate of increase is approximately proportional to the
enthalpy of fusion and, to a first approximation, does not depend on the
melting temperature.

• For a given solvent and a fixed temperature, if two solids have similar
enthalpies of fusion, the solid with lower melting temperature has the
higher solubility; similarly, if the two solids have nearly the same
melting temperature, the one with the lower enthalpy of fusion has the
higher solubility.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Solubilities of solids in liquids
Calculation of the pure-solute fugacity ratio

(5)

To simplify notation, let

and let
Solubilities of solids in liquids

These two fugacities depend only on the properties of the solute (component
2); they are independent of the nature of the solvent. The ratio of these two
fugacities can readily be calculated by the thermodynamic cycle indicated in
Figure 2. The molar Gibbs energy change for component 2 in going from a to
d is related to the fugacities of solid and subcooled liquid by

(6)

where, for simplicity, subscript 2 has been omitted. This Gibbs energy change
is also related to the corresponding enthalpy and entropy changes by

(7)
Solubilities of solids in liquids

Figure 2: Thermodynamic cycle for calculating the fugacity of a pure subcooled liquid.
Solubilities of solids in liquids

(8)
Solubilities of solids in liquids

(9)

(10)

Which becomes

(11)
Solubilities of solids in liquids

(12)

(13)
Solubilities of solids in liquids
Solubilities of solids in liquids

Figure 3: Fugacity ratio for solid and subcooled liquid carbon dioxide.
Solubilities of solids in liquids
Ideal solubility

(14)

Equation (14) provides a reasonable method for estimating solubilities of solids in


liquids where the chemical nature of the solute is similar to that of the solvent.
Solubilities of solids in liquids
Equation (14) immediately leads to useful conclusions concerning the
solubilities of solids in liquids. Strictly, these conclusions apply only to ideal
solutions but they are useful guides for other solutions that do not deviate
excessively from ideal behavior:

• For a given solid/solvent system, the solubility increases with rising


temperature. The rate of increase is approximately proportional to the
enthalpy of fusion and, to a first approximation, does not depend on the
melting temperature.

• For a given solvent and a fixed temperature, if two solids have similar
enthalpies of fusion, the solid with lower melting temperature has the
higher solubility; similarly, if the two solids have nearly the same
melting temperature, the one with the lower enthalpy of fusion has the
higher solubility.
Solubilities of solids in liquids

(15)
Solubilities of solids in liquids
Solubilities of solids in liquids

Figure 4: Solubility of aromatic solids in benzene.


Solubilities of solids in liquids

Equation (14) gives the ideal solubility of solid 2 in solvent 1. By interchanging


subscripts, we may use the same equation to calculate the ideal solubility of
solid 1 in solvent 2; by repeating such calculations at different temperatures, we
can then obtain the freezing diagram of the binary system as a function of
composition, as shown in Figure 5 taken from Prigogine and Defay (1954).

In these calculations we assume ideal behavior in the liquid phase and


complete immiscibility in the solid phase. The left side of the diagram
represents equilibrium between the liquid mixture and solid o-
chloronitrobenzene, while the right side represents equilibrium between the
liquid mixture and solid p- chloronitrobenzene. At the point of intersection,
called the eqtectic point, three phases are in equilibrium.
Solubilities of solids in liquids

Figure 5: Freezing point for the system o-chloronitrobenzene (1)/ p-chloronitrobenzene (2)
Solubilities of solids in liquids

Nonideal solutions
Solubilities of solids in liquids
Solubilities of solids in liquids

Figure 6: Solubility of cholesterol (2) in polar solvents. Solid line is the best fit through
the experimental points; the dashed line is the ideal-solubility curve equation (15).
Solubilities of solids in liquids
As Figure 6 shows, the experimental solubility curves (Bar et el., 1984) deviate
significantly from the ideal-solubility curve as given by equation (15). In
contrast to the previous example concerned with the solubility of aromatic solids
in benzene, for cholesterol/polar-solvent systems, all molecular interactions
(solute-solute, solute-solvent, and solvent-solvent) are relative complex. For
these systems, activity coefficients for cholesterol are far in excess of unity.
Solubilities of solids in liquids

Figure 7: Solubility of aromatic solids in carbon tetrachloride.


Solubilities of solids in liquids

A comparison of Figure 4 and Figure 7 shows that at the same temperature,


the solubilities in carbon tetrachloride are lower than those in benzene; in
other words, the activity coefficients of the solutes in carbon tetrachloride are
larger than those in benzene. If the activity coefficients of the solutes in
carbon tetrachloride are represented by the simple empirical relation

(16)
Solubilities of solids in liquids
As for solutions of liquid components, there is no general method for
predicting activity coefficients of solid solutes in liquid solvents. For
nonpolar solutes and solvents, however, a reasonable estimate can frequently
be made with the Scatchard-Hildebrand relation

(17)

is the volume fraction of the solvent.


Solubilities of solids in liquids
The molar liquid volume and solubility parameter of the solvent can be
determined from the thermodynamic properties of the solvent, but it is necessary
to use a thermodynamic cycle (as illustrated in Figure 2) to calculate these
functions for the subcooled liquid solute.

(18)

where subscript t refers to the triple-point temperature.

(19)
Solubilities of solids in liquids

(20)
Solubilities of solids in liquids
Solubilities of solids in liquids

The square of the solubility parameter is defined as the ratio of the energy
of complete vaporization to the liquid volume; therefore, if the vapor
pressure of the subcooled liquid is large, it is necessary to add a vapor-
phase correction to the energy of vaporization given by equation (20). Such
a correction, however, is rarely required and for most cases of interest the
solubility parameter of the subcooled liquid is given with sufficient
accuracy by

(21)
Solubilities of solids in liquids
The regular-solution theory of Scatchard-Hildebrand can be significantly
improved when the geometric-mean assumption is not used; in that event,
equation (17) becomes
Solubilities of solids in liquids

Figure 8: Solubility parameters of subcooled liquids.


Solubilities of solids in liquids

Figure 9: Molar volumes of subcooled liquids.


Solubilities of solids in liquids

A more fundamental extension of the Scatchard-Hildebrand equation was


introduced by Myers (1965) in his study of the solubility of solid carbon
dioxide in liquefied light hydrocarbons. Because carbon dioxide has a
large quadrupole moment, separate consideration was given to the
contributions of dispersion forces and quadrupole forces to the cohesive
energy density of subcooled liquid carbon dioxide. The energy of
vaporization was divided into two parts:

(23)
Solubilities of solids in liquids

As a result, two cohesive-energy densities can now be computed, corresponding


to the two types of intermolecular forces:

(24)

(25)

where superscript L has been omitted. These cohesive-energy densities for


carbon dioxide are shown as a function of temperature in Figure 10.
Solubilities of solids in liquids

Figure 10: Cohesive-energy density due to dispersion forces and due


to quadrupole forces for subcooled liquid carbon dioxide.
Solubilities of solids in liquids

The activity coefficient of component 2, the solute, dissolved in a nonpolar


solvent, is now written

(26)

Splitting the cohesive-energy density into a dispersion part end a quadrupole


part has an important effect on the calculated solubility of solid carbon dioxide.
Although the contribution from quadrupole forces is small, appreciable error is
introduced by not separately considering this contribution.
Solubilities of solids in liquids

To calculate the cohesive-energy density due to quadrupole forces, Meyer


derived the relation

(27)
Solubilities of solids in liquids
If the solvent, component 1, also has a significant quadrupole moment, then an
additional term must be added to the bracketed quantity in equation (26) to
account for quadrupole forces between the dissimilar components; further, the
geometric-mean term must be modified to include only the dispersion
cohesive-energy density of component 1. The bracketed term in equation (26)
then becomes

Using the theory of intermolecular forces, Myers showed that

(28)
Solubilities of solids in liquids

(29)
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
High-Pressure phase equilibrium
Numerous chemical processes operate at high pressure and, primarily for economic
reasons, many separation operations (distillation, adsorption) are conducted at high
pressures; further, phase equilibria at high pressures are of interest in geological
exploration, such as in drilling for petroleum and natural gas.

In the early days of phase-equilibrium thermodynamics, such an attempt could not


be made because of computational complexity; realistic thermodynamic
calculations for high-pressure equilibrium are difficult without computers.

The adjective high-pressure is relative; in some areas of technology (e.g., outer


space research) 1 mm of mercury is a high pressure whereas in others (e.g., solid-
state research) a pressure of a few hundred bars is considered almost a vacuum.
Figure 1 (Schneider, 1976) shows a rough comparison between pressures observed
in nature and pressures of some common industrial processes.
High-Pressure phase equilibrium

Figure 1: Pressure scale for natural (left) and chemical (right) processes (Schneider, 1976).
High-Pressure phase equilibrium

For the description of these and similar processes, it is necessary to


understand the thermodynamic properties of fluids at high pressures. We
designate here as “high-pressure” any pressure sufficiently large to have an
appreciable effect on the thermodynamic properties of all the phases under
consideration. In vapor-liquid equilibria, a high pressure may be anywhere
between (about) 20 to 1000 bar, depending on the system and on the
temperature; only in rare cases does the pressure exceed 1000 bar, because in
most cases of common interest the vapor-liquid critical condensation pressure
of the system is below 1000 bar. In liquid-liquid equilibria or in gas-gas
equilibria the pressure may be considerably larger, although experimental
studies are rare for fluid mixtures at pressures beyond 1000 bar.
High-Pressure phase equilibrium

Fluid mixtures at high pressures

For vapor-liquid behavior of a typical simple system, consider mixtures of


ethane and n-heptane; the critical temperature of ethane is 32.3°C and that
of n-heptane is 267.0°C. Figure 2 shows the relation between pressure and
composition at 149°C.
High-Pressure phase equilibrium

Figure 2: Pressure-composition diagram for the system ethane/n-heptane at


149°C. The critical point of the mixture is at C (Mehta and Thodos, 1965).
High-Pressure phase equilibrium
The left-hand line gives the saturation pressure (bubble pressure) as a function
of liquid composition and the right-hand line gives the saturation pressure
(dew pressure) as a function of vapor composition. The two lines meet at the
critical point where the two phases become identical. At 149°C the critical
composition is 76 mol% ethane and the critical pressure is 88 bar.

At this temperature and composition, therefore, only one phase can exist at
pressure higher than 88 bar; further, regardless of pressure, it is not possible to
have at 149°C a coexisting liquid phase containing more than 76 mol% ethane.

To characterize vapor-liquid equilibria for a binary system, measurements like


those shown in Figure 2 must be repeated for other temperatures; for each
temperature, there is a critical composition and critical pressure.
High-Pressure phase equilibrium

Figure 3: Critical temperatures and pressures for the system ethane/n-heptane (Mehta and Thodos, 1965).
High-Pressure phase equilibrium
Figure 3 gives experimentally observed critical temperatures and pressures
as a function of mole fraction for the ethane/n-heptane system. While the
critical temperature of this system is a monotonic function of composition,
the critical pressure goes through a maximum; many, but by no means all,
binary systems behave this way.
High-Pressure phase equilibrium

Because Figures 2 to 4 are for simple systems, they do not indicate the
variety of phase behavior that is possible in binary systems. That variety
increase substantially when we consider also partial immiscibility of
liquids and possible occurrence of solid phases.
High-Pressure phase equilibrium

Figure 4: K factors for the methane/propane system (Sage and Lacey, 1938).
High-Pressure phase equilibrium
Phase behavior at high pressure

To understand phase behavior at high pressure, we need to know how to


calculate and interpret phase diagrams.

Calculation of phase diagrams is based on solving the equation of phase


equilibrium; for vapor-liquid equilibrium,
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
High-Pressure phase equilibrium
Numerous chemical processes operate at high pressure and, primarily for economic
reasons, many separation operations (distillation, adsorption) are conducted at high
pressures; further, phase equilibria at high pressures are of interest in geological
exploration, such as in drilling for petroleum and natural gas.

In the early days of phase-equilibrium thermodynamics, such an attempt could not


be made because of computational complexity; realistic thermodynamic
calculations for high-pressure equilibrium are difficult without computers.

The adjective high-pressure is relative; in some areas of technology (e.g., outer


space research) 1 mm of mercury is a high pressure whereas in others (e.g., solid-
state research) a pressure of a few hundred bars is considered almost a vacuum.
Figure 1 (Schneider, 1976) shows a rough comparison between pressures observed
in nature and pressures of some common industrial processes.
High-Pressure phase equilibrium

Figure 1: Pressure scale for natural (left) and chemical (right) processes (Schneider, 1976).
High-Pressure phase equilibrium

For the description of these and similar processes, it is necessary to


understand the thermodynamic properties of fluids at high pressures. We
designate here as “high-pressure” any pressure sufficiently large to have an
appreciable effect on the thermodynamic properties of all the phases under
consideration. In vapor-liquid equilibria, a high pressure may be anywhere
between (about) 20 to 1000 bar, depending on the system and on the
temperature; only in rare cases does the pressure exceed 1000 bar, because in
most cases of common interest the vapor-liquid critical condensation pressure
of the system is below 1000 bar. In liquid-liquid equilibria or in gas-gas
equilibria the pressure may be considerably larger, although experimental
studies are rare for fluid mixtures at pressures beyond 1000 bar.
High-Pressure phase equilibrium

Fluid mixtures at high pressures

For vapor-liquid behavior of a typical simple system, consider mixtures of


ethane and n-heptane; the critical temperature of ethane is 32.3°C and that
of n-heptane is 267.0°C. Figure 2 shows the relation between pressure and
composition at 149°C.
High-Pressure phase equilibrium

Figure 2: Pressure-composition diagram for the system ethane/n-heptane at


149°C. The critical point of the mixture is at C (Mehta and Thodos, 1965).
High-Pressure phase equilibrium
The left-hand line gives the saturation pressure (bubble pressure) as a function
of liquid composition and the right-hand line gives the saturation pressure
(dew pressure) as a function of vapor composition. The two lines meet at the
critical point where the two phases become identical. At 149°C the critical
composition is 76 mol% ethane and the critical pressure is 88 bar.

At this temperature and composition, therefore, only one phase can exist at
pressure higher than 88 bar; further, regardless of pressure, it is not possible to
have at 149°C a coexisting liquid phase containing more than 76 mol% ethane.

To characterize vapor-liquid equilibria for a binary system, measurements like


those shown in Figure 2 must be repeated for other temperatures; for each
temperature, there is a critical composition and critical pressure.
High-Pressure phase equilibrium

Figure 3: Critical temperatures and pressures for the system ethane/n-heptane (Mehta and Thodos, 1965).
High-Pressure phase equilibrium
Figure 3 gives experimentally observed critical temperatures and pressures
as a function of mole fraction for the ethane/n-heptane system. While the
critical temperature of this system is a monotonic function of composition,
the critical pressure goes through a maximum; many, but by no means all,
binary systems behave this way.
High-Pressure phase equilibrium

Because Figures 2 to 4 are for simple systems, they do not indicate the
variety of phase behavior that is possible in binary systems. That variety
increase substantially when we consider also partial immiscibility of
liquids and possible occurrence of solid phases.
High-Pressure phase equilibrium

Figure 4: K factors for the methane/propane system (Sage and Lacey, 1938).
High-Pressure phase equilibrium
Phase behavior at high pressure

To understand phase behavior at high pressure, we need to know how to


calculate and interpret phase diagrams.

Calculation of phase diagrams is based on solving the equation of phase


equilibrium; for vapor-liquid equilibrium,
High-Pressure phase equilibrium
Interpretation of phase diagrams

To interpret phase diagrams we apply the Gibbs phase rule. For nonreacting
systems, that rule is expressed by the simple relation,

(1)
High-Pressure phase equilibrium
Pressure and temperature are the most convenient independent variables for the
measurement and study of phase equilibrium in fluid systems. In fluid systems,
changes in pressure and temperature produce large charges in the phase behavior
and a three-dimensional diagram in pressure, temperature and a third variable is
required for a complete description of a two-component system (see Table 1).

Table 1: Geometric constraints imposed by the phase rule on phase equilibria.


High-Pressure phase equilibrium
High-Pressure phase equilibrium
Classification of phase diagrams for binary mixtures
High-Pressure phase equilibrium

Van Konynenburg and Scott (1980) showed that almost all known types of
binary fluid-phase equilibria (vapor-liquid, liquid-liquid, and gas-gas
equilibria) can be qualitatively predicted using the van der Waals equation of
state and quadratic mixing rules. Their calculations suggested the classification
of binary fluid-phase behavior into six types, based on the shape of the mixture
critical lines and on the absence or presence of three-phase lines.

However, due to limitations of the equation of state used, they were able to
generate only five of these six types with the van der Waals equation of state. A
brief description of each of these types of phase behavior is presented below.

Figure 5 shows a few schematic pressure-temperature diagrams for binary


systems.
High-Pressure phase equilibrium

Figure 5: Six types of phase behavior in binary fluid systems. C = critical point; L
= liquid; V = vapor; UCEP = upper critical end point; LCEP = lower critical end
point. Dashed curves are critical lines and hatching marks heterogeneous regions.
High-Pressure phase equilibrium
Type I-phase behavior

Figure 6 represents type I-mixture


High-Pressure phase equilibrium
In the simple case shown in Figure 6, the line ending at C1 is the vapor-liquid
coexistence curve for pure 1 while the line ending at C2 is the vapor-liquid coexistence
curve for pure 2; C1 and C2 are the critical points. The dashed line joining these points
is the critical locus; each point on that line is the critical point for a mixture of fixed
composition. The continuous vapor-liquid critical line and the absence of liquid-liquid
immiscibility is often observed for mixtures where the two components are chemically
similar and/or their critical properties are comparable. Typical examples are
methane/ethane, carbon dioxide/n-butane, and benzene/toluene.

The critical locus may (but need not) show either a minimum or a maximum. The
latter is an indication of large positive deviations from Raoult’s law and hence of
relatively weak, unlike intermolecular interactions; such behavior is found for binary
mixtures of, e.g., a polar with a non-polar fluid, such as methanol/n-hexane.
High-Pressure phase equilibrium

Type II-phase behavior

Type II is similar to type I except that, at low temperatures, liquid mixtures of


components 1 and 2 are not miscible in all proportions. There is thus one
additional critical line. The line labeled LLV gives the locus of a three-phase
line where one vapor phase is in equilibrium with two liquid phases. This locus
ends at the upper critical end point (UCEP) where the two liquid phases merges
into one liquid phase; the presence of the UCEP depends on temperature as
shown in Figure 5 by the dashed line curving upward, in this case, with a
negative slope. The negative slope indicates that the upper critical solution
temperature decreases with rising pressure. The phase diagrams of type II-
mixtures may be complicated by the presence of an azeotropic line.
High-Pressure phase equilibrium

However, the essential feature of this type (and also of type VI-mixtures) is
the continuous vapor-liquid critical line that is distinct from the liquid-
liquid critical line. Examples of type II are carbon dioxide/n-octane and
ammonia/toluene. As shown, the critical line starting at C2 has an initial
negative slope but that is not always so; the slope may initially be positive
and become negative at higher pressures. Further, 3-phase line LLV may in
some cases reside above the vapor pressure-curve of component 1, as
observed for mixtures of hydrocarbons with the corresponding fully- or
near fully-fluorinated fluorocarbons, e.g. methane/trifluoromethane.
High-Pressure phase equilibrium
Type III-phase behavior

For mixtures with large immiscibility such as water/n-alkane mixtures, the locus
of the liquid -liquid critical lines moves to higher temperatures and may then
interface with the vapor-liquid critical curve. This means that the vapor-liquid
critical locus is not necessarily a continuous line connecting C1 and C2, as
illustrated in type III. In this type, the critical locus has two branches. One
branch goes from the vapor-liquid critical point of the more volatile component,
C1 to UCEP, where the gaseous phase and the liquid phase (richer in the more
volatile component) have the same composition. The other branch starts at C2
and then rises with pressure, perhaps with a positive slope (e.g., helium/water) or
with a negative slope (e.g., methane/toluene) or with a slope that changes sign
(e.g., nitrogen/ammonia and ethane/methanol).
High-Pressure phase equilibrium

The critical line starting from C2 with a positive slope indicates the
existence of what is called gas-gas equilibria: two phases at equilibrium at a
temperature larger than the critical temperature of either pure component.
Gas-gas immiscibility is conveniently classified as either first kind (almost
exclusively confined to mixtures containing helium as one component, e.g.,
water/helium and helium/xenon) where the critical curve extends directly
from the critical point of the less volatile component with a positive slope;
or second kind, where the critical line passes first through a temperature
minimum and then goes steeply to increasing temperatures and pressures
(e.g., methane/ammonia and water/propane).
High-Pressure phase equilibrium

Type IV-phase behavior

Type IV is similar to type V. In both, the vapor-liquid critical line starting at


C2 ends at a LCEP, a terminal point of the LLV three-phase line. However, in
type IV the LLV locus has two parts; this means that there is limited liquid-
phase miscibility at low temperatures, ending at UCEP. As the temperature
and pressure rise, there is a second region of limited miscibility from LCEP to
UCEP. Such phase behavior is shown by ethane/1-propanol and carbon
dioxide/nitrobenzene.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
High-Pressure phase equilibrium
Numerous chemical processes operate at high pressure and, primarily for economic
reasons, many separation operations (distillation, adsorption) are conducted at high
pressures; further, phase equilibria at high pressures are of interest in geological
exploration, such as in drilling for petroleum and natural gas.

In the early days of phase-equilibrium thermodynamics, such an attempt could not


be made because of computational complexity; realistic thermodynamic
calculations for high-pressure equilibrium are difficult without computers.

The adjective high-pressure is relative; in some areas of technology (e.g., outer


space research) 1 mm of mercury is a high pressure whereas in others (e.g., solid-
state research) a pressure of a few hundred bars is considered almost a vacuum.
Figure 1 (Schneider, 1976) shows a rough comparison between pressures observed
in nature and pressures of some common industrial processes.
High-Pressure phase equilibrium

Figure 1: Pressure scale for natural (left) and chemical (right) processes (Schneider, 1976).
High-Pressure phase equilibrium

For the description of these and similar processes, it is necessary to


understand the thermodynamic properties of fluids at high pressures. We
designate here as “high-pressure” any pressure sufficiently large to have an
appreciable effect on the thermodynamic properties of all the phases under
consideration. In vapor-liquid equilibria, a high pressure may be anywhere
between (about) 20 to 1000 bar, depending on the system and on the
temperature; only in rare cases does the pressure exceed 1000 bar, because in
most cases of common interest the vapor-liquid critical condensation pressure
of the system is below 1000 bar. In liquid-liquid equilibria or in gas-gas
equilibria the pressure may be considerably larger, although experimental
studies are rare for fluid mixtures at pressures beyond 1000 bar.
High-Pressure phase equilibrium

Fluid mixtures at high pressures

For vapor-liquid behavior of a typical simple system, consider mixtures of


ethane and n-heptane; the critical temperature of ethane is 32.3°C and that
of n-heptane is 267.0°C. Figure 2 shows the relation between pressure and
composition at 149°C.
High-Pressure phase equilibrium

Figure 2: Pressure-composition diagram for the system ethane/n-heptane at


149°C. The critical point of the mixture is at C (Mehta and Thodos, 1965).
High-Pressure phase equilibrium
The left-hand line gives the saturation pressure (bubble pressure) as a function
of liquid composition and the right-hand line gives the saturation pressure
(dew pressure) as a function of vapor composition. The two lines meet at the
critical point where the two phases become identical. At 149°C the critical
composition is 76 mol% ethane and the critical pressure is 88 bar.

At this temperature and composition, therefore, only one phase can exist at
pressure higher than 88 bar; further, regardless of pressure, it is not possible to
have at 149°C a coexisting liquid phase containing more than 76 mol% ethane.

To characterize vapor-liquid equilibria for a binary system, measurements like


those shown in Figure 2 must be repeated for other temperatures; for each
temperature, there is a critical composition and critical pressure.
High-Pressure phase equilibrium

Figure 3: Critical temperatures and pressures for the system ethane/n-heptane (Mehta and Thodos, 1965).
High-Pressure phase equilibrium
Figure 3 gives experimentally observed critical temperatures and pressures
as a function of mole fraction for the ethane/n-heptane system. While the
critical temperature of this system is a monotonic function of composition,
the critical pressure goes through a maximum; many, but by no means all,
binary systems behave this way.
High-Pressure phase equilibrium

Because Figures 2 to 4 are for simple systems, they do not indicate the
variety of phase behavior that is possible in binary systems. That variety
increase substantially when we consider also partial immiscibility of
liquids and possible occurrence of solid phases.
High-Pressure phase equilibrium

Figure 4: K factors for the methane/propane system (Sage and Lacey, 1938).
High-Pressure phase equilibrium
Phase behavior at high pressure

To understand phase behavior at high pressure, we need to know how to


calculate and interpret phase diagrams.

Calculation of phase diagrams is based on solving the equation of phase


equilibrium; for vapor-liquid equilibrium,
High-Pressure phase equilibrium
Interpretation of phase diagrams

To interpret phase diagrams we apply the Gibbs phase rule. For nonreacting
systems, that rule is expressed by the simple relation,

(1)
High-Pressure phase equilibrium
Pressure and temperature are the most convenient independent variables for the
measurement and study of phase equilibrium in fluid systems. In fluid systems,
changes in pressure and temperature produce large charges in the phase behavior
and a three-dimensional diagram in pressure, temperature and a third variable is
required for a complete description of a two-component system (see Table 1).

Table 1: Geometric constraints imposed by the phase rule on phase equilibria.


High-Pressure phase equilibrium
High-Pressure phase equilibrium
Classification of phase diagrams for binary mixtures
High-Pressure phase equilibrium

Van Konynenburg and Scott (1980) showed that almost all known types of
binary fluid-phase equilibria (vapor-liquid, liquid-liquid, and gas-gas
equilibria) can be qualitatively predicted using the van der Waals equation of
state and quadratic mixing rules. Their calculations suggested the classification
of binary fluid-phase behavior into six types, based on the shape of the mixture
critical lines and on the absence or presence of three-phase lines.

However, due to limitations of the equation of state used, they were able to
generate only five of these six types with the van der Waals equation of state. A
brief description of each of these types of phase behavior is presented below.

Figure 5 shows a few schematic pressure-temperature diagrams for binary


systems.
High-Pressure phase equilibrium

Figure 5: Six types of phase behavior in binary fluid systems. C = critical point; L
= liquid; V = vapor; UCEP = upper critical end point; LCEP = lower critical end
point. Dashed curves are critical lines and hatching marks heterogeneous regions.
High-Pressure phase equilibrium
Type I-phase behavior

Figure 6 represents type I-mixture


High-Pressure phase equilibrium
In the simple case shown in Figure 6, the line ending at C1 is the vapor-liquid
coexistence curve for pure 1 while the line ending at C2 is the vapor-liquid coexistence
curve for pure 2; C1 and C2 are the critical points. The dashed line joining these points
is the critical locus; each point on that line is the critical point for a mixture of fixed
composition. The continuous vapor-liquid critical line and the absence of liquid-liquid
immiscibility is often observed for mixtures where the two components are chemically
similar and/or their critical properties are comparable. Typical examples are
methane/ethane, carbon dioxide/n-butane, and benzene/toluene.

The critical locus may (but need not) show either a minimum or a maximum. The
latter is an indication of large positive deviations from Raoult’s law and hence of
relatively weak, unlike intermolecular interactions; such behavior is found for binary
mixtures of, e.g., a polar with a non-polar fluid, such as methanol/n-hexane.
High-Pressure phase equilibrium

Type II-phase behavior

Type II is similar to type I except that, at low temperatures, liquid mixtures of


components 1 and 2 are not miscible in all proportions. There is thus one
additional critical line. The line labeled LLV gives the locus of a three-phase
line where one vapor phase is in equilibrium with two liquid phases. This locus
ends at the upper critical end point (UCEP) where the two liquid phases merges
into one liquid phase; the presence of the UCEP depends on temperature as
shown in Figure 5 by the dashed line curving upward, in this case, with a
negative slope. The negative slope indicates that the upper critical solution
temperature decreases with rising pressure. The phase diagrams of type II-
mixtures may be complicated by the presence of an azeotropic line.
High-Pressure phase equilibrium

However, the essential feature of this type (and also of type VI-mixtures) is
the continuous vapor-liquid critical line that is distinct from the liquid-
liquid critical line. Examples of type II are carbon dioxide/n-octane and
ammonia/toluene. As shown, the critical line starting at C2 has an initial
negative slope but that is not always so; the slope may initially be positive
and become negative at higher pressures. Further, 3-phase line LLV may in
some cases reside above the vapor pressure-curve of component 1, as
observed for mixtures of hydrocarbons with the corresponding fully- or
near fully-fluorinated fluorocarbons, e.g. methane/trifluoromethane.
High-Pressure phase equilibrium
Type III-phase behavior

For mixtures with large immiscibility such as water/n-alkane mixtures, the locus
of the liquid -liquid critical lines moves to higher temperatures and may then
interface with the vapor-liquid critical curve. This means that the vapor-liquid
critical locus is not necessarily a continuous line connecting C1 and C2, as
illustrated in type III. In this type, the critical locus has two branches. One
branch goes from the vapor-liquid critical point of the more volatile component,
C1 to UCEP, where the gaseous phase and the liquid phase (richer in the more
volatile component) have the same composition. The other branch starts at C2
and then rises with pressure, perhaps with a positive slope (e.g., helium/water) or
with a negative slope (e.g., methane/toluene) or with a slope that changes sign
(e.g., nitrogen/ammonia and ethane/methanol).
High-Pressure phase equilibrium

The critical line starting from C2 with a positive slope indicates the
existence of what is called gas-gas equilibria: two phases at equilibrium at a
temperature larger than the critical temperature of either pure component.
Gas-gas immiscibility is conveniently classified as either first kind (almost
exclusively confined to mixtures containing helium as one component, e.g.,
water/helium and helium/xenon) where the critical curve extends directly
from the critical point of the less volatile component with a positive slope;
or second kind, where the critical line passes first through a temperature
minimum and then goes steeply to increasing temperatures and pressures
(e.g., methane/ammonia and water/propane).
High-Pressure phase equilibrium

Type IV-phase behavior

Type IV is similar to type V. In both, the vapor-liquid critical line starting at


C2 ends at a LCEP, a terminal point of the LLV three-phase line. However, in
type IV the LLV locus has two parts; this means that there is limited liquid-
phase miscibility at low temperatures, ending at UCEP. As the temperature
and pressure rise, there is a second region of limited miscibility from LCEP to
UCEP. Such phase behavior is shown by ethane/1-propanol and carbon
dioxide/nitrobenzene.
High-Pressure phase equilibrium

Type V-phase behavior

In type V, the first branch of the critical line goes from C1 to UCEP as in type
III, but the second branch goes from C2 to the lower critical end point (LCEP).
Contrarily to type IV, iv type V mixtures, the liquids are completely miscible
below LCEP. Figure 7 gives a schematic representation of what is visually
observed at the LCEP and at the UCEP.

Two liquid phases exist only for temperatures between those of the critical
endpoints. At the temperature of the LCEP, the meniscus between the two
liquids disappears, whereas at that of the UCEP, it is the gas-liquid interface
that disappears leaving in both cases one liquid phase in equilibrium with a
gas phase.
High-Pressure phase equilibrium

Figure 7: Three-phase behavior and schematic representation of critical end points for the
binary ethane/n-octadecane (Specovious et al., 1985). The dashed line represents a
disappearing meniscus, i.e., a critical points: gas-liquid at the UCEP (39.30°C) and liquid-
liquid at the LCEP (39.14°C).
High-Pressure phase equilibrium

The immiscibility region for ethane/n-octadecane extends over a very narrow


temperature range (0.157°C) and all phases are rich in ethane. This temperature
range increases to 2.927°C for the binary ethane/n-eicosane, (C20H42) while
complete miscibility is observed for ethane/n-heptadecane.
High-Pressure phase equilibrium
Type VI-phase behavior

The two critical curves meet at an upper critical pressure; at higher pressures
two liquids are miscible. Examples of this complex behavior are found in
mixtures where one (or both) component is self-associated through hydrogen
bonding. An example of type VI is water/2-butanol.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Liquid-liquid equilibria

To illustrate the effect of pressure on liquid-liquid equilibria, Figure 13 shows


experimental results for the system carbon tetrafluoride/propane. In this case,
rising pressure increases the size of the two-phase region; in other words,
rising pressure is unfavorable for miscibility. Thermodynamics can help us to
interpret results such as those shown in Figure 13. In particular, we want to
examine how pressure may be used to induce miscibility or immiscibility in a
binary system.
Liquid-liquid equilibria

Figure 13: Pressure (a) and temperature (b) dependence of liquid-liquid equilibria
in the system carbon tetrafluoride/propane (Jeschke and Schneider, 1982).
Liquid-liquid equilibria

(1)

and

(2)
Liquid-liquid equilibria
The effect of pressure on miscibility in binary liquid mixtures is closely
connected with the volume change on mixing, as indicated by the exact relation

(3)
Liquid-liquid equilibria
Liquid-liquid equilibria
Liquid-liquid equilibria
Liquid-liquid equilibria
To illustrate, we consider a simple, symmetric binary mixtures at some fixed
temperature and 1 bar pressure. For this liquid mixture, we assume that

(4)

(5)

(6)
Liquid-liquid equilibria

(7)

(8)

(9)
Liquid-liquid equilibria

(10)

(11)

For the simple mixture described by equation 4 and 5, we can calculate activity
coefficients and substitute into equations 10 and 11 and we then obtain

(12)

(13)
Liquid-liquid equilibria
Liquid-liquid equilibria

(14)

(15)
Liquid-liquid equilibria

(16)

(17)
Liquid-liquid equilibria

(18)
Liquid-liquid equilibria

Table 2: Predicted and measured effect of pressure on critical solution temperatures in


binary systems (Myers et al., 1966; Clerke and Sengers, 1981).
Liquid-liquid equilibria

Equation 15, a purely thermodynamic equation, is of no use by itself. However,


when combined with physical insight, it produces the much more useful
(approximate) equation 18.

The situation under discussion is similar to the familiar salting-out effect in


liquids where a salt, added to an aqueous solution, serves to precipitate one or
more organic solutes. Here, however, instead of a salt, we add a gas; to dissolve
an appreciable quantity of gas, the system must be at elevated pressure.
Liquid-liquid equilibria

Figure 17: Effect of gaseous component (2) on mutual solubility of liquids (1) and (3).
Liquid-liquid equilibria
Suppose now that we dissolve a small amount of component 2 in the binary
mixture; what happens to the critical solution temperature? This equation was
considered by Prigogine (1943), who assumed that for any binary pair that can
be formed from the three components 1, 2, and 3, the excess Gibbs energy
(symmetric convention) is

(19)

(20)
Liquid-liquid equilibria

(21)

and

(22)
Liquid-liquid equilibria

As shown by Balder (1966) thermodynamic considerations may be used to


establish the phase diagram of a ternary system consisting of two miscible liquid
components and a supercritical gas at high pressures. Such diagrams have been
obtained experimentally (Elgin and Weinstock, 1959; Francis, 1963; Chappelear
and Elgin, 1961) for a variety of ternary systems and have led to suggestions for
separations using a high-pressure gas as the phase-splitting agent.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Exergy analysis
Exergy: work potential of energy

When the geothermal well is discovered, the first thing the explorers do is
estimate the amount of energy contained in the source. This information
alone, however, is of little value in deciding whether to build a power plant
on that site. What we really need to know is the work potential of the
source—that is, the amount of energy we can extract as useful work. The
rest of the energy is eventually discarded as waste energy and is not worthy
of our consideration. Thus, it would be very desirable to have a property to
enable us to determine the useful work potential of a given amount of energy
at some specified state. This property is exergy, which is also called the
availability or available energy.
Exergy analysis

The work potential of the energy contained in a system at a specified state is


simply the maximum useful work that can be obtained from the system. You
will recall that the work done during a process depends on the initial state, the
final state, and the process path. That is,

In an exergy analysis, the initial state is specified, and thus it is not a


variable. The work output is maximized when the process between two
specified states is executed in a reversible manner. Therefore, all the
irreversibilities are disregarded in determining the work potential. Finally, the
system must be in the dead state at the end of the process to maximize the
work output.
Exergy analysis
A system is said to be in the dead state when it is in thermodynamic equilibrium with
the environment it is in (Fig. 1). At the dead state, a system is at the temperature and
pressure of its environment (in thermal and mechanical equilibrium); it has no kinetic
or potential energy relative to the environment (zero velocity and zero elevation
above a reference level); and it does not react with the environment (chemically
inert). Also, there are no unbalanced magnetic, electrical, and surface tension effects
between the system and its surroundings, if these are relevant to the situation at hand.
A system has zero exergy at the dead state.

Figure 1: A system that is in


equilibrium with its environment is
said to be at the dead state.
Exergy analysis
Therefore, we conclude that a system delivers the maximum possible work as it
undergoes a reversible process from the specified initial state to the state of its
environment, that is, the dead state. This represents the useful work potential of the
system at the specified state and is called exergy. It is important to realize that
exergy does not represent the amount of work that a work-producing device will
actually deliver upon installation. Rather, it represents the upper limit on the amount
of work a device can deliver without violating any thermodynamic laws. There will
always be a difference, large or small, between exergy and the actual work delivered
by a device. This difference represents the room engineers have for improvement.

Note that the exergy of a system at a specified state depends on the conditions of
the environment (the dead state) as well as the properties of the system. Therefore,
exergy is a property of the system–environment combination and not of the system
alone. Altering the environment is another way of increasing exergy, but it is
definitely not an easy alternative.
Exergy analysis
Exergy associated with kinetic and potential energy

Kinetic energy is a form of mechanical energy, and thus it can be converted to


work entirely. Therefore, the work potential or exergy of the kinetic energy of a
system is equal to the kinetic energy itself regardless of the temperature and
pressure of the environment. That is,

(1)

where V is the velocity of the system relative to the environment.

Potential energy is also a form of mechanical energy, and thus it can be converted
to work entirely. Therefore, the exergy of the potential energy of a system is equal
to the potential energy itself regardless of the temperature and pressure of the
environment. That is,

(2)
Exergy analysis

where g is the gravitational acceleration and z is the elevation of the system


relative to a reference level in the environment.
Therefore, the exergies of kinetic and potential energies are equal to
themselves, and they are entirely available for work. However, the internal
energy u and enthalpy h of a system are not entirely available for work, as
shown later.
Exergy analysis
Exergy of a closed system

In general, internal energy consists of sensible, latent, chemical, and nuclear


energies. However, in the absence of any chemical or nuclear reactions, the
chemical and nuclear energies can be disregarded, and the internal energy can be
considered to consist of only sensible and latent energies that can be transferred to
or from a system as heat whenever there is a temperature difference across the
system boundary. The second law of thermodynamics states that heat cannot be
converted to work entirely, and thus the work potential of internal energy must be
less than the internal energy itself.

Consider a stationary closed system at a specified state that undergoes a reversible


process to the state of the environment (that is, the final temperature and pressure
of the system should be T0 and P0, respectively). The useful work delivered during
this process is the exergy of the system at its initial state (Fig. 2).
Exergy analysis

Figure 2: The exergy of a specified mass at a specified state is the useful work that can
be produced as the mass undergoes a reversible process to the state of the environment.
Exergy analysis

(3)
Exergy analysis

(4)
Exergy analysis

(5)
Exergy analysis

(6)

Integrating from the given state (no subscript) to the dead state (0 subscript)
we obtain

(7)

where Wtotal useful is the total useful work delivered as the system undergoes a
reversible process from the given state to the dead state, which is exergy
by definition.
Exergy analysis
A closed system, in general, may possess kinetic and potential energies, and
the total energy of a closed system is equal to the sum of its internal, kinetic,
and potential energies. Noting that kinetic and potential energies themselves
are forms of exergy, the exergy of a closed system of mass m is

(8)

(9)
Exergy analysis

The exergy change of a closed system during a process is simply the difference
between the final and initial exergies of the system,

(10)

or, on a unit mass basis,

(11)

For stationary closed systems, the kinetic and potential energy terms drop out.
Exergy analysis
When the properties of a system are not uniform, the exergy of the system can be
determined by integration from

(12)

Note that exergy is a property, and the value of a property does not change unless
the state changes. Therefore, the exergy change of a system is zero if the state of the
system or the environment does not change during the process. For example, the
exergy change of steady flow devices such as nozzles, compressors, turbines,
pumps, and heat exchangers in a given environment is zero during steady operation.
Exergy analysis

Figure 3: The exergy of a cold medium is also a


positive quantity since work can be produced by
transferring heat to it.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Evaluation of equilibrium constants
Evaluation of equilibrium constants

R. C. Reid, J. M. Prausnitz, and B. E. Poling, The Properties of Gases and Liquids. 4th ed., chap. 6,
McGraw-Hill, New York. 1987.
Evaluation of equilibrium constants

Example:
Calculate the equilibrium constant for the vapor phase hydration of ethylene at
418.15 to 593.15 K (145 and at 320°C) from data given in App. C.
Evaluation of equilibrium constants
Evaluation of equilibrium constants

For T = 418.15 K, values of the integrals in equation


Evaluation of equilibrium constants
Evaluation of equilibrium constants
Evaluation of equilibrium constants

With these values, the following results are readily obtained:

T/K (t/°C) K
298.15(25) 1 29.366 1 1 29.366
418.15(14 1.4025 29.366 4.985×10-3 0.9860 1.443×10-1
5)
593.15(32 1.9894 29.366 1.023×10-4 0.9794 2.942×10-3
0)
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Relation of equilibrium constants to composition

(1)

becomes:

(2)
Relation of equilibrium constants to composition

The fugacity is related to the fugacity coefficient by Eq. (3):

(3)
Relation of equilibrium constants to composition

Substitution of this equation into Eq. (2) provides an equilibrium expression


displaying the pressure and the composition:

(4)
Relation of equilibrium constants to composition

(5)
Relation of equilibrium constants to composition

(6)
Relation of equilibrium constants to composition

Although Eq. (6) holds only for an ideal-gas reaction, we can base some
conclusions on it that are true in general:
Relation of equilibrium constants to composition
Relation of equilibrium constants to composition
Reactions in Heterogeneous Systems

When liquid and gas phases are both present in an equilibrium mixture of
reacting species, Eq. (1) a criterion of vapor/liquid equilibrium, must be
satisfied along with the equation of chemical-reaction equilibrium. Consider,
for example, the reaction of gas A with liquid water B to form an aqueous
solution C.

(1)
Reactions in Heterogeneous Systems

(2)

(3)
Reactions in Heterogeneous Systems

becomes:

(4)

(5) (6)
Reactions in Heterogeneous Systems

However, all methods theoretically lead to the same equilibrium composition,


provided Henry's law as applied to species C in solution is valid. In practice, a
particular choice of standard states may simplify calculations or yield more
accurate results, because it makes better use of the available data. The nature of
the calculations required for heterogeneous reactions is illustrated in the
following example.
Reactions in Heterogeneous Systems

Liquid-Phase Reactions
For a reaction occurring in the liquid phase, we return to:

(7)

According to the definition of the activity coefficient,

(8)

(9)
Reactions in Heterogeneous Systems

(10)

The difference between these two equations is:

(11)

We can also write

(12)
Reactions in Heterogeneous Systems

From equation (11) and (12)

(13)

(14)

From equation (7), (9) and (14)

(15)
Reactions in Heterogeneous Systems

Except for high pressures, the exponential term is close to unity and may be
omitted. Then,

(16)

and the only problem is determination of the activity coefficients. An equation


such as the Wilson equation or the UNIFAC method can in principle be applied,
and the compositions can be found from Eq. (16) by a complex iterative
computer program.
Reactions in Heterogeneous Systems

(17)

This simple relation is known as the law of mass action. Since liquids often
form nonideal solutions, Eq. (17) can be expected in many instances to yield
poor results.
Reactions in Heterogeneous Systems

The method is based on use of a fictitious or hypothetical standard state for the
solute, taken as the state that would exist if the solute obeyed Henry's law up to
a molality m of unity. In this application, Henry's law is expressed as

(18)

However, one can calculate the properties the solute would have if it obeyed
Henry's law to a concentration of 1 molal, and this hypothetical state often
serves as a convenient standard state for solutes.
Reactions in Heterogeneous Systems

Figure 1: Standard state for dilute aqueous solutions


Reactions in Heterogeneous Systems

The standard-state fugacity is

(19)

Hence, for any species at a concentration low enough for Henry's law to hold,

(20)
and

(21)
Reactions in Heterogeneous Systems
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Exergy analysis
Exergy: work potential of energy

When the geothermal well is discovered, the first thing the explorers do is
estimate the amount of energy contained in the source. This information
alone, however, is of little value in deciding whether to build a power plant
on that site. What we really need to know is the work potential of the
source—that is, the amount of energy we can extract as useful work. The
rest of the energy is eventually discarded as waste energy and is not worthy
of our consideration. Thus, it would be very desirable to have a property to
enable us to determine the useful work potential of a given amount of energy
at some specified state. This property is exergy, which is also called the
availability or available energy.
Exergy analysis

The work potential of the energy contained in a system at a specified state is


simply the maximum useful work that can be obtained from the system. You
will recall that the work done during a process depends on the initial state, the
final state, and the process path. That is,

In an exergy analysis, the initial state is specified, and thus it is not a


variable. The work output is maximized when the process between two
specified states is executed in a reversible manner. Therefore, all the
irreversibilities are disregarded in determining the work potential. Finally, the
system must be in the dead state at the end of the process to maximize the
work output.
Exergy analysis
A system is said to be in the dead state when it is in thermodynamic equilibrium with
the environment it is in (Fig. 1). At the dead state, a system is at the temperature and
pressure of its environment (in thermal and mechanical equilibrium); it has no kinetic
or potential energy relative to the environment (zero velocity and zero elevation
above a reference level); and it does not react with the environment (chemically
inert). Also, there are no unbalanced magnetic, electrical, and surface tension effects
between the system and its surroundings, if these are relevant to the situation at hand.
A system has zero exergy at the dead state.

Figure 1: A system that is in


equilibrium with its environment is
said to be at the dead state.
Exergy analysis
Therefore, we conclude that a system delivers the maximum possible work as it
undergoes a reversible process from the specified initial state to the state of its
environment, that is, the dead state. This represents the useful work potential of the
system at the specified state and is called exergy. It is important to realize that
exergy does not represent the amount of work that a work-producing device will
actually deliver upon installation. Rather, it represents the upper limit on the amount
of work a device can deliver without violating any thermodynamic laws. There will
always be a difference, large or small, between exergy and the actual work delivered
by a device. This difference represents the room engineers have for improvement.

Note that the exergy of a system at a specified state depends on the conditions of
the environment (the dead state) as well as the properties of the system. Therefore,
exergy is a property of the system–environment combination and not of the system
alone. Altering the environment is another way of increasing exergy, but it is
definitely not an easy alternative.
Exergy analysis
Exergy associated with kinetic and potential energy

Kinetic energy is a form of mechanical energy, and thus it can be converted to


work entirely. Therefore, the work potential or exergy of the kinetic energy of a
system is equal to the kinetic energy itself regardless of the temperature and
pressure of the environment. That is,

(1)

where V is the velocity of the system relative to the environment.

Potential energy is also a form of mechanical energy, and thus it can be converted
to work entirely. Therefore, the exergy of the potential energy of a system is equal
to the potential energy itself regardless of the temperature and pressure of the
environment. That is,

(2)
Exergy analysis

where g is the gravitational acceleration and z is the elevation of the system


relative to a reference level in the environment.
Therefore, the exergies of kinetic and potential energies are equal to
themselves, and they are entirely available for work. However, the internal
energy u and enthalpy h of a system are not entirely available for work, as
shown later.
Exergy analysis
Exergy of a closed system

In general, internal energy consists of sensible, latent, chemical, and nuclear


energies. However, in the absence of any chemical or nuclear reactions, the
chemical and nuclear energies can be disregarded, and the internal energy can be
considered to consist of only sensible and latent energies that can be transferred to
or from a system as heat whenever there is a temperature difference across the
system boundary. The second law of thermodynamics states that heat cannot be
converted to work entirely, and thus the work potential of internal energy must be
less than the internal energy itself.

Consider a stationary closed system at a specified state that undergoes a reversible


process to the state of the environment (that is, the final temperature and pressure
of the system should be T0 and P0, respectively). The useful work delivered during
this process is the exergy of the system at its initial state (Fig. 2).
Exergy analysis

Figure 2: The exergy of a specified mass at a specified state is the useful work that can
be produced as the mass undergoes a reversible process to the state of the environment.
Exergy analysis

(3)
Exergy analysis

(4)
Exergy analysis

(5)
Exergy analysis

(6)

Integrating from the given state (no subscript) to the dead state (0 subscript)
we obtain

(7)

where Wtotal useful is the total useful work delivered as the system undergoes a
reversible process from the given state to the dead state, which is exergy
by definition.
Exergy analysis
A closed system, in general, may possess kinetic and potential energies, and
the total energy of a closed system is equal to the sum of its internal, kinetic,
and potential energies. Noting that kinetic and potential energies themselves
are forms of exergy, the exergy of a closed system of mass m is

(8)

(9)
Exergy analysis

The exergy change of a closed system during a process is simply the difference
between the final and initial exergies of the system,

(10)

or, on a unit mass basis,

(11)

For stationary closed systems, the kinetic and potential energy terms drop out.
Exergy analysis
When the properties of a system are not uniform, the exergy of the system can be
determined by integration from

(12)

Note that exergy is a property, and the value of a property does not change unless
the state changes. Therefore, the exergy change of a system is zero if the state of the
system or the environment does not change during the process. For example, the
exergy change of steady flow devices such as nozzles, compressors, turbines,
pumps, and heat exchangers in a given environment is zero during steady operation.
Exergy analysis

Figure 3: The exergy of a cold medium is also a


positive quantity since work can be produced by
transferring heat to it.
Exergy Destruction
Irreversibilities such as friction, mixing, chemical reactions, heat transfer
through a finite temperature difference, unrestrained expansion, non-
quasiequilibrium compression or expansion always generate entropy, and
anything that generates entropy always destroys exergy. The exergy destroyed
is proportional to the entropy generated, is expressed as

Note that exergy destroyed is a positive quantity for any actual process and
becomes zero for a reversible process. Exergy destroyed represents the lost
work potential and is also called the irreversibility or lost work.
EXAMPLE: Exergy Destruction during Heat Conduction

Consider steady heat transfer through a 5-m × 6-m brick wall of a house of
thickness 30 cm. On a day when the temperature of the outdoors is 0°C, the
house is maintained at 27°C. The temperatures of the inner and outer
surfaces of the brick wall are measured to be 20°C and 5°C, respectively, and
the rate of heat transfer through the wall is 1035 W. Determine the rate of
exergy destruction in the wall, and the rate of total exergy destruction
associated with this heat transfer process.
Thank you
Advanced Chemical
Engineering Thermodynamics
(CHC502)

Suman Dutta
Exergy analysis

EXERGY BALANCE: CLOSED SYSTEMS

The nature of exergy is opposite to that of entropy in that exergy can be


destroyed, but it cannot be created. Therefore, the exergy change of a system
during a process is less than the exergy transfer by an amount equal to the
exergy destroyed during the process within the system boundaries. Then the
decrease of exergy principle can be expressed as

Or
Exergy analysis

FIGURE: Mechanisms of exergy transfer.


Exergy analysis

This relation is referred to as the exergy balance and can be stated as the
exergy change of a system during a process is equal to the difference
between the net exergy transfer through the system boundary and the exergy
destroyed within the system boundaries as a result of irreversibilities.

We mentioned earlier that exergy can be transferred to or from a system by


heat, work, and mass transfer. Then the exergy balance for any system
undergoing any process can be expressed more explicitly as
Exergy analysis

where all the quantities are expressed per unit mass of the system. Note that for a
reversible process, the exergy destruction term Xdestroyed drops out from all of the
relations above. Also, it is usually more convenient to find the entropy
generation Sgen first, and then to evaluate the exergy destroyed directly. That is,
Exergy analysis

A closed system does not involve any mass flow and thus any exergy transfer
associated with mass flow. Taking the positive direction of heat transfer to be to
the system and the positive direction of work transfer to be from the system, the
exergy balance for a closed system can be expressed more explicitly as shown in
the figure.
Exergy analysis

FIGURE: Exergy balance for a closed system when the direction of heat
transfer is taken to be to the system and the direction of work from the system.
Exergy analysis

Note that the relations above for a closed system are developed by taking the heat
transfer to a system and work done by the system to be positive quantities.
Therefore, heat transfer from the system and work done on the system should be
taken to be negative quantities when using those relations.
Exergy analysis
Thank you

You might also like