Metallurgical Thermodynamics - Ghosh
Metallurgical Thermodynamics - Ghosh
Metallurgical Thermodynamics - Ghosh
AIIlNDRA GHOSH
Formerly Professor
Materials and Metallurgical Engineering
Indian Institute of Technology Kanpur
C 2003 by PHI Leaming Private Limited, Delhi. All rights reserved. No part of this book may be
reproduced in any form , by mimeograph or any other means, without pennission in writing from
the publisher.
ISBN-978-81 -203-2091-8
The export rights of this book are vested solely with the publisher.
Published by Asoke K. Ghosh, PHI Leaming Private Limited, Rimjhim House, 111 , Patparganj
Industrial Estate, Delhi-110092 and Printed by V.K. Batra at Peart Offset Press Private Limited,
New Delhi-110015.
ThxTBooK OF ivlATERIAI.S AND
METAILURGICAL 'fHERMODYNAMia>
To
(Late) Shri Sajoy Panchanan Ghosh
(Late) Shri Gobinda Lal Sarkar
Shrimati Binapani Biswas
Contents
Preface xv
List of Symbols with Units xvii
1. INTRODUCTION 1-8
1.1 Historical Perspective 1
1.2 History of Classical Thermodynamics 2
1.2.1 General 2
1.2.2 Chemical Thermodynamics 2
1.2.3 MetaJlurgical Thermodynamics 3
1.3 Introductory Concepts and Definitions 3
1.3.1 System and Surrounding 4
1.3.2 State of a System 4
1.3.3 Thermodynamic Equilibrium 5
1.3.4 Characterization of Systems-Some elucidations 6
1.4 Values of Some Physical Constants 7
1.5 Summary 8
vii
Contents
Bibliography 271-272
Index 277-280
Preface
In the beginning of the 20th Century, metallurgy was primarily an art. Application of
pure and engineering sciences gave it recognition as a branch of engineering by 1950s, and
contributed to the development of metallurgical sciences, including metallurgical thermo-
dynamics. Atomic and solid state physics also matured by 1930s/1940s. In the decade of
1940s, major advances were also initiated in high technology areas such as nuclear, space
and aviation technologies, and solid state electronics. These required a variety of new
special purpose materials. Metallurgical sciences, solid state physics etc. provided scientific
foundation for the same. Earlier, metallurgical engineers were primarily concerned with
metals and alloys, and ceramists with traditional oxide ceramics, such as refractories. However,
the new materials, both metallic and non-metallic, changed the character of academic
programmes. It was realized that the fundamentals of all these are the same, and hence can
be taught in a generalized way. This led to the growth of Materials Science and Engineering.
In the llTs and engineering colleges in India, Bachelor's degree programme is offered
in metallurgical engineering or in a combined progranpne of metallurgical and materials
engineering. This text aims to cater to the basic one-semester undergraduate course on
thermodynamics. There are examples and problems on both metallic and inorganic non-
metallic materials. Chapter 14 has an additional brief ,,resentation on thermodynamics of
surfaces as well as interfaces and defects in solids of importance to materials science in
general.
Chapters 1-5 cover general thermodynamics and thermochemistry. Chapter 12 presents
some introductory discussions on statistical thermodynamics. The remaining chapters
(i.e. Chapters 6-11 and 13) deal with standard topics of chemical and metallurgical/materials
thermodynamics. As stated earlier, the principal emphasis is on high temperature systems
and processes of interest to metallurgy and materials science.
I wish to express my gratitude to the All India Council for Technical Education
(AICTE) for providing Emeritus Fellowship and miscellaneous financial assistance during
preparation of the manuscript. I am aJso indebted to the Indian Institute of Technology
Kanpur for making infrastructural facilities available, which has made this work possible.
I wish to thank Mr. J.L. Kuril for typing the manuscript, Mr. B.K. Jain for tracing the
diagrams, and Mr. A. Sharma for assistance in many ways. Finally, I sincerely thank my
wife, Radha and other family members without whose cooperation and encouragement the
writing of this book would not have been possible.
AHINDRA GHOSH
List of Symbols with Units
("-" indicates a dimensionless quantity; mo! means gram mole (i.e. mole); J means Joule)
Greek Symbols
a coefficient of volumetric thermal expansion
a. Darken 's a-function for component i in a solution
/3 isothermal compressibility of a material
r ratio of Cp/Cv
Yi activity coefficient of component i in a solution
rP Henry's Law constant for solute i in a binary solution
surface excess of component i
List of Symbols with Units
Superscripts
0 denotes values at standard state for extensive state
properties (e.g. SJ, d, JfJ)
m indicates change of value of a quantity due to mixing
(e.g. G{" =Gi - GP, fl.Gm =G - G°)
XS value of a state property in excess of that in ideal
solution (e.g. c;xs = Gi - Giid)
Subscripts
i, j, for component i, j ,
m for melting
f for formation of a compound from elements
T at temperature T
Tr for phase transformation
v for vapourization
Others
[ ] metallic phase
( ) nonmetallic condensed phase (e.g. oxide, etc.)
Introduction
2. Statistical thermodynamics
Statistical thermodynamics originated from the Kinetic Theory of Gases, which related
pressure to average kinetic energy of molecules in an ideal gas. The application of probability
theory, quantum theory and statistical mechanics allowed it to arrive at macroscopic
thermodynamic relations from atomistic (i.e. microscopic) point of view. It was founded by
Maxwell, Gibbs and Boltzmann in the late 19th Century, and was developed further in the
1920s and 1930s.
3. In-eversible thermodynamics
Irreversible thermodynamics deals with the application of thermodynamics to irreversible
processes, and was first proposed by I. Prigogine in 1942.
Textbook of Materials and Metallurgical Thermodynamics
1.2.1 General
The literal meaning of thennodynamics is that it is a subject dealing with the relation
between heat and motion. The invention of the steam engine in the 18th Century triggered
the issue of the relation between heat and work amongst engineers and physicists. In 1798,
Count Rumford observed that the heat generated during boring of cannons was approximately
proportional to the work done. In 1840. Joules' classic and precise experiments provided
confirmation of the above and led to the concept of Mechanical Equivalent of Heat. It was
later generalized into Law of Conservation of Energy, which states that "energy can neither
be created nor destroyed; it can be only transfonneJ from one fonn into another". These
developments eventually led to the generalized definition of thermodynamics as a subject
dealing with energy and its transformation from one form to another. This constituted che
basis for the First Law of Thermodynamics.
Although mechanical energy can be completely converted into heat, it was observed that
the reverse is not true, i.e. heat cannot be completely converted into work in a straightforward
fashion. Quantitative explanation came first fiom the famous Carnot's cycle, derived by
S. Carnot in 1824. These Jed to the enunciation of the Second Law of Thermodynamics,
employing the concept of reversible and irreversible processes. Later in the mid-19th Century,
Clausius and J.J. Thomson (Lord Kelvin) proposed entropy as the thermodynamic parameter
for quantitative application of the Second Law.
In 1906, W. Nernst proposed •his Heat Theorem on the basis of experimental evidence.
It was later generalized by Max Planck as "the entropy of any homogeneous substance
(which is at comp!P.te internal equilibrium), may be taken as zero at 0 K". This is known
as the Third Law of Thenrwdynamics.
into existence. The 19th Century further witnessed major developments in chemistry,
such as
• Dalton's atomic theory and reaction stoichiometry
• Avogadro's hypothesis and concept of g-mole
• The concept of chemical equilibrium
• The heats of reaction
• Faraday's Laws of .Electrolysis
These provided the foundation for Physical Chemistry and Chemical Thermodynamics.
In the late 19th Century, J.W. Gibbs proposed his famous free energy function, known as
Gibbs free energy. The systematic application of this to thermodynamics by Gibbs and
others firmly established the foundation of chemical thermodynamics. The subject attained
maturity by early 20th Century.
Chemical thermodynamics is also based on the three laws of thermodynamics. However,
it employs many other auxilliary relations also, mostly derived from Gibbs free energy
co~siderations as well as other physico-chemical relations.
large-scale applications. Therefore, this text will be concerned almost exclusively with this.
The following introductory concepts and definitions are the starting points for classical
thermodynamics.
where n is number of moles (i.e. g-moles), R the universal gas constant, T the absolute
temperature, and v the molar volume.
Now we have to distinguish between property and variable. The molar volume is a
property of the substance. Of course, it can be varied. In that sense, it is a variable too.
But pressure and temperature are not properties. They are variables, generally imposed by
the surrounding. Therefore, variables may or may not be properties. It may also be noted
here that state properties are also known as state functions .
The volume of a substance is proportional to its mass (or number of moles) . On the
other hand, P and T are independent of the mass of the system, and are known as intensive
variables. Volume V is an extensive property. It can be made intensive by dividing it by
mass or number of moles. Hence, the molar volume (v) is an intensive property. Density
is an intensive property.
It is obvious that, if an equation contains both extensive property and intensive variable,
then there must be a term denoting mass or number of moles. If the latter are missing, then
there is an implicit assumption that the extensive property is for a fixed mass (say, a closed
system). The general convention in chemical thermodynamics is to go for molar properties,
which are intensive. This way, the relationships and functions become independent of the
quantity of matter, and hence of more general applicability [e.g. Eq. (1.2)].
v = -v =f(P, T) (1.4)
n
It should be emphasized here th.at Eqs. (1.3) and (1.4) are applicable to a substance of
constant composition and structure, i.e. its nature remains the same during change of
pressure and temperature. For every substance, except for ideal gas, the function is different.
When the system is at equilibrium with respect to an intensive variable, the magnitude
of the variable throughout the system is the same (i.e. uniform). The term equilibrium has
originated from the Latin word "acquilibrium", meaning equal weight (e.g. as in a weighing
balance). In general, it refers to a state, which is not changing with time.
The mechanical interaction of the system with the surrounding is represented by pressure,
Textbook of Materials and Metallurgical Thermodynamics
in the absence of a field of force (electric, magnetic etc.). Here, mechanical equilibrium
means pressure equilibrium, i.e. uniformity of pressure throughout the ~ystem. Thus, there
is no tendency for the pressure to change with time anywhere in the system.
Similarly, at themial equilibrium, temperature should be uniform throughout the "System.
Only then there will be no tendency for heat to flow from one part of the system to another,
i.e. no thermal change takes place in the system.
For purely pl )Sica] processes, such as expansion/compression of a substance, thermodynamic
equilibrium conusts of mechanical and thermal equilibria. But for physico-chemical processes
and chemically reactive systems, thermodynamic equilibrium would also require attainment
of physico-chemicallchemical equilibrium in addition to mechanical and thermal equilibria.
This means attainment of uniform chemical potential besides uniformity of pressure and
temperature in the system. Detailed discussions of chemical potential will be taken up in
Chapter 10. ,
T he system is at its most stable state when it is at thermodynamic equil ibrium. It will
stay that way unless disturbed. An open or a closed system interacts with its surrounding.
There, the equilibrium is always with respect to the surrounding. If the surrounding changes,
the state of equilibrium of the system will also change.
Sometimes a system is not at complete thermodynamic equil ibrium, but at partial or
pseudo equilibrium. Thermodynamics is capable of handling such cases. One example is
metastable state, where the system undergoes change so slowly that there will be no perceptible
change in a long period of time (say, several years/centuries/millennia). For instance, if a
mixture of hydrogen and oxygen gas is kept at room temperature, very little detectable
reaction would occur even in years, although it is not at chemical equilibrium.
Cementite (Fe3C) in iron-carbon system is metastable, but can stay for perhaps several
centuries or millennia at room temperature. In both the above cases, raising the temperature
increases the rate enormously, and allows attainment of chemical equilibria in a short period
of time.
Another example is partial chemical equi libria in some complex reactive systems, where
some reactions attain equilibrium whereas some others do not.
We have broadl y classified system-as open, closed and isolated already. It has also been
stated that processes may be purely physical or physico-chemical/chemical. In the former
case, we are dealing with a nonreactive system, whereas, when physico-chemical processes
and/or chemical reactions occur, the system is a reactive one. Some further comments are
now noted to elaborate the above observations.
A system consisting of one phase only (say, water) is a single-phase system, which is
known as a homogeneous system. At 0°C, ice and water can co-exist at equil ibrium. Then
it is a heterogeneous system. However, the above classifications do not provide full description
of a system. Further classifications are required for application of the laws of thermodynamics.
A substance in a closed or isolated system not only has fixed mass, but has fixed overall
composition as well. Consider water in a closed vessel. On cooling below 0°C, it is
Introduction
transformed into ice. Again, on heating above 100°C, it is transformed into a gas. Therefore,
the state of aggregation of H 20 (broadly speak ing, structure of H2 0 ) depends on temperature.
Similarly, above 910°C, the structure of pure iron changes from a-Fe (i.e. BCC) to y-Fe
(i.e. FCC). These changes are known as phase clianges, and are physico-chemical processes.
The system has one-compo11e11t only (H 20 or Fe), and is also referred to as Unary system.
A multicompo11e11r system obviously has more than one component. An aqueous solution
of NaCl, for example, has two components, (viz. NaCl and H 20 ), and is known as a binary
system. Similarly, we have ternary, quartemary etc., all coming under multicomponent
system. If the NaCl solution is cooled below 0°C, pure ice crystals would start forming. As
temperature is lowered further, more and more ice forms. This increases concentration of
NaCl in residual water. Hence, in this two-phase system, composition o; water is changing.
However, composition of the entire system in terms of number of moles of H 20 and NaCl
does not change, if it is a closed system.
In contrast, in an open system, both mass and composition of the systein can change.
For example, if we consider only the aqueous NaCl solution as our system (even in a closed
vessel) in the two-phase ice + aqueous solution situation, then both mass and composition
of the solution are changing with change of temperature. If pure water is heated in an open
beaker, its mass will keep decreasing with vaporization. If it is an aqueous solution , then
besides mass, concentration of NaCl in solution will keep increasing with vaporization.
If water is heated to a very high temperature, the steam dissociates into hydrogen and
oxygen, and we have to be concerned with this chemical reaction in the gaseous state. Then,
if the system is closed, its total mass will remain constant. Not only that, its composition
in terms of total number of moles of hydrogen and oxygen (in elemental form and combined
as H 20) will remain constant. Therefore, the meaning of the term composition has to be
interpreted that way.
To sum up, in classical thermodynamics, a system can be classified as:
1. Open, closed or isolated
2. Unary or multicomponent
3. Reactive or nonreactive
4. Homogeneous or heterogeneous
These are required for application of thermodynamics. Some of these will again be
elucidated in subsequent chapters. However, it is hoped that these introductory comments
would be useful to readers.
1.5 Summary
1. Thermodynamics may be broadly classified as: classical (i.e. macroscopic or
phenomenological), statistical, and irreversible.
2. Most applications are based on classical thermodynamics. Hence, the present text is
primarily concerned with it. Also, generally, the term themwdynamics denotes classical
thermodynamics.
3 . The three laws of thermodynamics constitute the foundations of thermodynamics.
4. Thermodynamics deals with energy and its transformation from one form to another.
5. Metallurgical Thermodynamics, or its later generalization, Thermodynamics of Materials,
belongs to chemical thermodynamics, which deals with reactive systems.
6. A system is classified as: open, closed and isolated. The rest of the Universe is its
surrounding.
7 . State variables are extensive or intensive.
8. Molar properties are intensive properties (or variables).
9. Pressure, temperature and volume are common state variables.
10. A state can be defined by state variables only at thermodynamic equilibrium. It then
constitutes a thermodynamic substance.
11. A system needs characterization before its thermodynamic treatment.
Chapter2
First Law of Thermodynamics
Two important thermodynamic quantities have been proposed on the basis of the First Law
of Thermodynamics, viz. Internal Energy U and Enthalpy H. Both are energy terms and are
state properties.
dU = fxJ - OW (2 .2)
The concept of internal energy was proposed assuming the validity of the Law of Conservation
First Law of Thermodynamics
of Energy. AU is just the difference between q and W. There is no direct proof of existence
of U as a state property. However, there are plenty of indirect experimental evidence of its
existence and is a well-established quantity.
In the early days of thermodynamics in the 19th Century, when the First Law was proposed,
the nature of U as a state property was inferred from experiments on cyclic processes, where
the system was changed but ultimately brought back to the same initial state. It was found
that for such cyclic processes,
provided q and W are expressed in the same energy unit (say, joule). This means that
:E AU= O, i.e. U is a state property.
cycle As discussed in Chapter 1, in conventional classical thermodynamics, we are generally
concerned with P, V, T as state variables. In chemical thermodynamics, the chemical
composition and structure of substance in the system also are variables. However, variation
of thermodynamic quantities with composition and structure will be taken up later. For the
time being, we shall assume the substance to retain its initial composition and structure all
throughout the process (i.e. a nonreactive system).
In Chapter 1, we have further stated that P, V, T for a system of fixed mass are
interdependent, with only two of them acting as independent variables. Hence, for a system
of fixed mass (imagine a closed system for easy comprehension),
U = f(V, D (2.4)
or
U = f(P , D (2.5)
or
U = f(P, V) (2.6)
du = (au)
av r
dV + (au)
ar v
dT (2.7)
or
du = (auJ
aP T
dP + (auJ
ar p
dT (2.8)
or
dU = (auJ
aP v
dP + (auJ
av p
dV (2.9)
Textbook of Materials and Metallurgical Thermodynamics
t State 1
Idealized
reversible
(T _ T ) State 2
_c:~;·~--
r sur
Reversible Irreversible
Time---+
Fig. 2.1 Temperature vs. time curves for isothermal processes (schematic); path is hypothetical
for irreversible process.
Process, on the other hand, occurs reversibly if the pressure is decreased a little to P - dP,
where dP is an infinitesimally small change in pressure. Then sufficient time is allowed for
the gas to attain Tsur• and again the pressure is decreased a little, and time allowed. Here
First Law of Thermodynamics
we are carrying out the process slowly through a series of equilibrium stages. It should be
noted that the temperature is always close to T5ur in the reversible process. The mathematically
limiting situation for curve 2 is the constant temperature line at T = Tsur• which is the
idealized version of an isothermal reversible process. Figure 2.2 illustrates these.
(a) Irreversible
expansion
(b) Reversible
expansion
Fig. 2.2 Irreversible and reversible expansion of a gas at any instant of time.
(State 1)
A Irreversible
I
I Idealized reversible path
I (PV ::;; nRT ::;; constant)
I
I
I
I
I
---r----~-p~;(..~ B
: Irreversible : (State 2)
1
contraction I
V1 V2
Volume---.
Fig. 2.3 Pressure-volume diagram for isothermal expansion-compression cycle of an ideal gas
in a cylinder-piston system (schematic); P is inside pressure; path is hypothetical for
irreversible cycle.
Textbook of Materials and Metallurgical Thermodynamics
which is the same as Eq. (1.1), and the corresponding curve is for isothermal reversible
process at temperature T. The irreversible expansion and compres~ion cannot be depicted on
the P-V diagram since their paths are undefined. Moreover, for irreversible process, the
pressure of gas in the cylinder is not expected to be uniform. However, it is qualitatively
correct to say that, during expansion, the gas pressure inside the cylinder (i.e. in the system)
will be higher than the external pressure. The reverse is true for irreversible compression.
On this basis, some hypothetical curves are shown.
When the gas expands, it performs work against external pressure. Suppose the cylinder
moves by a distance dl. Then,
f dV y;
f
V2 V2
W1 = P dV =nRT - =nRT lo ---1. (2.13)
~ ~ v Vi
(ii) The work done by the gas during compression (B ~ A) = W2, where
v, v;
J~
~
W2 = P dV =nRT In -1.. = - nRT In ---1. = - W1 (2.14)
~ Vi
Thus, net work done in the cycle A ~ B ~ A is
L
cycle
w =WI+ W2 = 0 (2.15)
Since the system comes back to the initial state, as stated in section 2.2.2,
I: i1U = O (2.16)
cycle
This means that there is no net interaction of the system with the surrounding for a
cyclic reversible process. Hence, it does not leave any permanent change in system or
surrounding during reversing. This is an important conclusion.
First Law of Thermodynamics
It is obvious that for irreversible processes, the above is not true and some net heat and
work interaction with the surrounding would occur for cycle A ~ B ~ A. The shaded area
is the net work done during the irreversible cycle A ~ B ~ A. The presence of friction
between the piston and the cylinder will introduce irreversibility as well. Therefore, for
reversibility, friction should be absent.
It should be noted again that the path is not known for an irreversible process. Hence,
it cannot be shown in a diagram. Therefore, the irreversible processes shown in Figs. 2.1
and 2.3 are hypothetical.
(2.19)
and would depend on the P-V relationship. Let us consider the following reversible process
paths:
(i) At constant volume (i.e. isochoric), W = O; hence from Eq . (2.1),
LiU = q (2.20)
(ii) At constant pressure (i.e. isobaric),
On the basis of experimental measurements by Joule and others in the 19th Century, it
was proposed that U is a function of temperature only for an ideal gas. Later, kinetic theory
of gases also predicted the same. Hence for the isothermal expansion-compression of an
ideal gas,
11U = 0, i.e. W =q (2.22)
It should be kept in mind that Eq. (2.22) is applicable only to a fixed mass of gas since
any change in mass will change internal energy even at a constant temperature.
(iv) Adiabatic:
q = 0, i.e. 11U =- W (2.23)
2.5 Enthalpy
Enthalpy H is defined as
H = U + PV (2.24)
Hence, by definition, H is a state property, since U and V are state properties and P is a
state variable. Differentiating Eq. (2.24), we get
dH = dU + P dV + V dP (2 .25)
i.e.
dH = dU + P dV at constant P (2.26)
If bW = 0, then by combining Eq. (2.26) with Eq. (2.18), we obtain
q
I
=
st ale I
dH =H 2 - HI =.&1 (2.28)
Since most processes occur or are carried out under conditions where Eq. (2.28) is
exactly or approximately valid, Eq. (2.28) constitutes the basis for process heat balance,
which is very important for manufacturing industries as well as analysis of processes.
Energy is costly, and should be accounted for and saved as much as possible. The bulk of
the energy supply goes to meet process heat requirements. Without process heat balance,
energy audit is not possible. Further, Eq. (2.28) allows experimental determination of L1H
of a process by measurement of q by a calorimeter. Since H is a state property, LJH
(i.e. H 2 - H 1) can be tabulated in thermodynamic data sources and used by all.
Many texts, especially the older ones, therefore, refer to enthalpy as heat content of a
substance due to the above equality. However, the latter terminology is not scientifically
correct since heat is energy in transition and not a state property.
First Law of The"nodynamics
(au)r =O
dP
(2.29)
(aH) ='(au)
dP r aP T
+ [a(PV)]
ap T
(2.30)
In ideal gas at constant T, PV = constant; so, d(PV) = 0. From all these above,
(aH) _(au) _0
ap T ap T
(2.31)
(E.2.1 )
(;~l = v; b (E.2.2)
(E .2.4)
Textbook of Materials and Metallurgical Thermodyn~mics
C= 8q (2.32)
dT
0q is heat required to raise the temperature of l mole by dT. Since 0q depends on path, C
is a meaningless quantity unless the path is specified. The following have proved to be
useful.
The molar heat capacity at constant volume is given by
cv -(oq) -(au)
- dT v - dT v (2.33)
since at a constant volume, 0q = dU. Similarly, molar heat capacity at constant pressure
(Cp) is given by
C -("dq)
P -dT
-("dH)
CJT
P - p
(2.34)
since at constant P, &, = dH. In Eqs. (2.33) and (2.34), U and H are molar values of
internal energy, and enthalpy, respectively.
From Eq. (2.34),
dH = Cp dT (2.35)
Integrating Eq. (2.35) between temperatures T1 and T2, we get
(2.36)
It should be noted that U and H are molar qµantities here. Earlier, they referred to
internal energy and enthalpy of the entire system. Both conventions will be followed. So, U
and H may be molar values or for the entire system, depending on what is being discussed.
Cp > Cv since Cp includes heat required to do work against pressure also, besides raising
temperature. From Eqs. (2.33) and (2.34),
First Law of Thermodynamics
Cp _ Cv = (aH)
ar
_(au)
ar v
p
= [a<u + PV)] _ (au)
ar ar v
p
_(au)
ar
_(au)
ar
+P
p
(av)
ar v p
(2.37)
(2.38)
(2.40)
(au)
av
=O
r
(2.41)
Cp - Cv= R (2.42)
i.e.
R/Cv
i = ( ~: )
(2.46)
R
-=r-1 (2.47)
Cv
(2.48)
P2
Pi
=(V1
V2
)r, i.e. pyr =m =a constant (2.50)
Figure 2.4 compares P-V relationships for isothermal and adiabatic reversible expansion
of an ideal gas on the P-V diagram, starting from the same initial state A(P1> V 1, T1) to the
same final pressure P2 . The final volume is larger for the isothermal process.
(2.51)
First Law of Thermodynamics
I
I
I Isothermal
I / reversible
I
I Adiabatic
I reversible
I
I
I
---T------
1 (P2, V2, T2) C
I I
I I
V2 V3
Volume--.
Fig. 2.4 P-V diagrams for isothermal and adiabatic reversible expansion of an ideal gas from
the same initial state (schematic).
=m
vt-r -
2
v,1-r)
I
( (2.52)
1-r
Noting from Eq. (2.50) that
(2.53)
we obtain
W = P2 V2 - Pi \'i (2.54)
1- r
From Eq. (2.45),
EXAMPLE 2.2 Consider a three-step cyclic process, as shown in Fig. 2.5. The working
substance is 1 kg-mot of a diatomic ideal gas initially at 300 K and 1 atm. The gas is heated
at constant volume to 500 K, then expanded adiabatically to the initial temperature of
300 K, and finally compressed isothermally to initial pressure of 1 atm. Each step occurs
reversibly.
Textbook of Materials and Metallurgical Thermodynamics
State 1
P 1 = 1 atm State 2
T1 = 300 K LIV= 0 Ti= 500 K
State 3
T3 = 300 K
Fig. 2.5 Stage-wise cyclic process corresponding to the data given in Example 2.2.
(a) Find out the initial as well as the other two states after each step, in terms of
pressure, volume and temperature.
(b) Verify the following data table.
Path dU &I w q
(all values in kilojoules)
500
State 2: T2 = 500 K, V2 = 24.62 m 3, P 2 = 1 x
300
= 1.6 atm
State 3: T3 = 300 K
From Eq. (2.48),
7
For a diatomic gas, y = - = 1.4
5
Fi. st Law of Thermodynamics
Hence,
P dV = nRT In -1..
v. (from Eq. (2.13))
V3
24 62
= 103 x 8.314 x 3oo(tn · ) x 10-3
88.29
= - 3185 kJ
2.8 Summary
1. The statements of the First Law of Thermodynamics are:
L1U = q - W, for a finite process
dU = liq - bW, for an infinitesimal process.
2. U is internal energy of the system, and is a state property. U is a function of
temperature only for a fixed mass of an ideal gas.
3. q is heat absorbed by the system from the surrounding, and W is work done by the
system on the surrounding.
4. q and W are energy in transition, are path dependent and, not state properties/state
variables.
S. bW = P dV + bW', where P dV is work done against pressure, and bW' is any other
form of work (electric, magnetic, surface etc.). For all topics up to Chapter 12,
bW' = 0, and OW= PdV.
6. For a reversible process, the path is known. Hence, W and q can be calculated. For
a finite reversible process going from state 1 to state 2, if bW' = 0, then
V2
W =
f
v,
P dV, where P is the pressure in the system.
PROBLEMS
2.1 Calculate the difference between ,1H and L1U per mole (i.e. g-mole) of the reaction:
2H2 (g) + 02 (g) = 2H20 (g)
at 500°C and 1 atm pressure.
First Law of Thermodynamics
2.2 One mole of an ideal monatomic gas is compressed from an initial volume of 20
litres to 10 litres by a reversible polytropic process, following the law:
PVl. 2 = constant. The initial temperature of the gas was 100°C.
Calculate:
(a) W, q for the process
(b) LJU, &I between the final and initial states of the gas.
2.3 3.5 moles of an ideal diatomic gas is initially at 500 K and 18 atm pressure. It
undergoes reversible isothermal expansion from 18 to 8 atm pressure, and then
further reversible adiabatic expansion from 8 atm to 1.6 atm. Calculate: W, q, LJU,
&I for each stage.
2.4 Calculate W, q, .dU, &I for reversible expansion of 2 moles of oxygen gas by the
adiabatic and isobaric paths. Assume oxygen to behave as ideal gas. At the initial
state, P = 2 atm, T = 273 K. The final volume is double that of the initial volume.
2.S A system comprises one mole of an ideal monatomic gas at 0°C and 1 atrn. The
system is subjected to the following processes, each of which is conducted reversibly:
(a) 10-fold increase in volume at constant temperature
(b) 100-fold adiabatic increase in pressure
(c) return to the initial state along a straight line path in the P-V. diagram.
Calculate the work done by the system in step (c), and the total heat added to the
system as a result of the cyclic process.
Heat Capacity and Enthalpy-
Auxiliary relations
and applications
molar heat capacities of all solid elements may be taken as 3R per°C. Further measurements
subsequently showed this rule to be approximate and very limited in scope. Figure 3.1
presents Cv vs. T experimental data for a few solid elements at low temperature. It shows
that:
25
20
,,.....
~0 15
s
-;;;
u
:; 10
0
0
:::..
t>
5
Temperature (K)
Fig. 3.1 Constant-volume molar heat capacities of Pb, Cu, Si and diamond as function of
temperature.
(3.1)
Combining Eqs. (3.1) and (3.2) and using the relation between n; and Avogadro's
number (N0), we get
(3.3)
1
I: exp (-hvilk 8 T) = I:x; = 1 +x +x 2 + ... = - - if x « (3.4)
i 1 - x
Following Einstein's derivation steps (being skipped), we get
3 3N hv
U =-Nohv
2
+ (e h vlk T0
B 1)-
(3.5)
and
c -(au)
v - oT v -
-3R(BE)z_e_Br;tr_
T (elJelT - 1)2 (3.6)
(since R = N0 k8 ), where
hv
lk = Einstein temperature = (3.7)
ks
At very low temperatures (near 0° K), the Debye equation may be simplified as
Cv = 464.5(T/B0 ) 3 (3 .8)
The Debye equation gives better fit with experimental data, provided the best fitted
values of 8 0 are employed. Even then, it is not satisfactory. Moreover, it is concerned with
elements only, that too in the crystalline state. Hence, due to its limited applicability, it is
a normal practice to experimentally determine heat capacity vs. temperature relationships
for solids and liquids.
Heat Capacity and Enthalpy-Auxiliary relations and applications
Experimental data consist of Cp as a function of temperature. For solids and liquids, the
PV term is very small. Hence, H is taken as equal to U, and Cp as equal to Cv. In other
words, no distinction is made between Cp and Cv so far as applications are concerned. It
has been found that experimental Cp vs. T data for elements and compounds fit best with
an equation of the type:
Cp =a + bT + cr2 (3.9)
where a, b, c are empirically fitted constants and differ from substance to substance.
The last term is the smallest and, therefore, often ignored. In some cases, such as liquid
metals, both bT and cr-2 are usually ignored. Equation (3.9) is employed for correlation
of experimental Cp data of diatomic and polyatomic gases as well. For alloys, where
experimental data are not available, approximate estimates are done assuming heat capacity
to vary linearly with atom fraction.
HT - H298 = f T
298
Cp dT (3.10)
298 K (strictly 298.15) is known as reference temperature. Combining Eq. (3.10) with
Eq. (3.9), we get
HT - H 29s f
= T,,.
298
Cp(S) dT + &lm + JTb Cp(l) dT +&Iv +
T,,.
JT
Tb
Cp(g) dT (3.12)
where Cp(s), Cp(I) and Cp(g) are Cp's for solid, liquid and gaseous A, respectively. If the
Heat Capacity and Enthalpy-Auxiliary relations and applications
substance is not a pure element or compound, then it does not have a single melting or
boiling point, and hence Eq. (3.12) is not applicable.
Sometimes an element or a compound undergoes phase transfonnation at the solid state.
Then these terms are also to be added. An example is iron which exhibits the following
structures at the solid state.
8-iron melts at 1809 K. Therefore, the sensible heat of iron at 2000 K is given as:
1033 f 1186
H'lJXXJ -H298 =
f 298
Cp(a)dT+L1111r(a ~ /3)+
1033
Cp(/3)dT
+ tlH m + f 2000
1809
Cp(l) dT (3. 13)
EXAMPLE 3.1 Calculate L1ll and L1.U when one mole of water at 25 °C at 1 atm pressure
is conv.erted into steam at 130°C and 2 atm.
Some relevant data (in J/mol) are now noted:
Cp of water = 75.3 per K, Latent heat of vaporization of water (L111v) = 40.64 x 103,
Cp of steam between 100 - 130°C = 30.11 + 9.937 x 10- 3 i-per K.
100+273 f 130+273
= f 25+273
Cp(water) dT + L111v +
100+273
Cp(steam) dT
403
=75.3 x 75 + 40.64 x 10 + [30.11 T + 9·937 x 10-3r 2]
3
2 373
= 47 .3 1 x 103 J/mol
Textbook of Materials and Metallurgical Thermodynamics
373
=22.4 x 10- 3 x - =30.6 x 10-3 m 3
273
Vw "' 0 since it is very small.
Hence,
P(V5 - Vw) = 1 X 30.6 X 10-3 m3 atm
= 30.6 x 10- 3 x 1.013 x 105 = 3100 joules
[Ll(PV)]m = [Ll(RT)]~~ =8.314(403 - 373) =250 joules
Therefore,
LlU = 47 .31 x 103 - 3100 - 250 = 43.96 x 103 J/mol
Some comments
For calculation of enthalpy changes, reactions, dissolutions and phase transformations have
been assumed to occur isothermally. Since enthalpy is a state property, L1H depends only
on initial and final states, and not the path. The above assumption provides the most
convenient path for calculation. Figure 2.1 has schematically shown variation of temperature
with time when an isothermal process occurs. For a reversible isothermal process, the
temperature remains constant all through . If the process is not reversible, then temperature
at beginning and end of a process would be the same. In between, the temperature can vary
significantly.
For calculation of enthalpy changes, the isothennal processes are general (i.e. irreversible
or reversible). Only the initial and final temperatures are assumed to be the same. This is
illustrated by Fig. 3.2 with the example of chemical reaction (3.14).
t
u
3
(U A+ BC AB+ C
:U Ti----.....---
c..
E
t
u Initial state Exothermic Final state
E-< (state l) reaction (state 2)
begins
Process ~
Time---•
Fig. 3.2 Variation of temperature with progress of an irreversible exothermic isothermal reaction
(schematic).
Textbook of Materials and Metallurgical Thermodynamics
from its elements. Standard heat of formation at a temperature T, i.e. [L111f(Dl is nothing
but L1111 at T, when elements and their compound are aJI at their respective standard states.
•See 0 . Kubaschewski, P.J. Spencer. and C.B. Alcock, Materials Thermochemistry, 6th ed., Pergammon
Press, Oxford, 1993; see also earl ier editions by 0 . Kubaschewski, E.LI. Evans, and C.B. Alcock). Some
standard data sources have been given in the Appendix to the book.
Textbook of Materials and Metallurgical Thermodynamics
Subtraction of Eq. (3.19) from Eq. (3.18) yields Eq. (3. 14), i.e.
A (pure) + BC (pure) = AB (pure) + C (pure); L1l-1°(7) (3.20)
According to Hess' Law, therefore,
L1JJO (T) =L1H?, AB (T) - L1H?, BC (T) (3 .21 )
Hence, L1l-I° of any reaction can be calculated, if the experimental values of m? of the
relevant compounds are available.
Enthalpy is a state property. Hence, we may write
or
:E L1H
cycle
=O for finite stages (3.22)
For reaction (3.14), application of Eq. (3.22), which is the basis of Hess' Law, is
illustrated by Fig. 3.3, where
L L1Ho
cycle
=M-1~ BC (T) + mo (T) -
'
m~ AB (T)
'
=0 (3.23)
i.e. L1H 0 (T)=t1H?,AB(T) - L1H~Bc (T) , which is Hess' Law [given in Eq. (3.21)].
A+B+C
(at 1)
A+ BC AB+ C
(at 1) (at 7)
Fig. 3.3 Illustration of thermodynamic basis of Hess' Law; A, B, C, AB, BC are pur& and one
mole each .
Heat Capacity and Enthalpy-Auxiliary relations and applications
Cyclic processes and the concept of cycle have been employed in thermodynamics and
other subjects. In connection with computer programming, the term loop is more popular.
Hence, the latter term is also being advocated now. Johnson and Stracher (1995)* have
widely employed loops for problem solving.
EXAMPLE 3.2 Calculate .1H~98 for the reaction: 2Ca0 + Si02 = Ca 2Si04
Given: t1H~ for the formation of CaO, Si02 and Ca2Si04 from their elements at 298 K
are -634, - 910.9 and - 2305.3 kilojoules per mole (kJ mole- 1), respectively.
Solutio11 Mi~98 = Ml?, 298 (Ca2Si04) - 2 Ml?, 298 (CaO) - Ml?, 298 (Si0 2)
= -2305 .3 - (-2 x 634 - 910.9] = - 126.4 kJ per mole of reaction
t1H for an isothermal process depends on the temperature at which the process occurs. The
relationship of t1Ff of a reaction with temperature and other quantities is known as Kirchhoff's
Law. Again, like Hess' law, it is based on the definition of enthalpy as a state property.
The derivation is being illustrated for reaction (3.14) by application of the loop in Fig. 3.4.
(3.24)
or
(3.25)
Equation (3.25) is the generalized version of Kirchhoff s Law. If the reactants and products
do not have any phase changes between T1 and T2, then Eq. (3.25) can be simplified as
M1~2 = m~1 + J Ti
7j
[ roducl:
p t
c~ - l:
reactant
c~J dT = m~ I
+ J L1C~
r
7j
2
dT (3.26)
If the reactants and products are pot pure and not in their standard states, then also
Kirchhoffs Law can be employed. However, the compositions of the solutions should not
change with temperature. For melting and other phase transformations, the equation gets
simplified. For example, for melting of solid,
*D.L. Johnson and G.B. Stracher, Thermodynamic Loop Applications in Materials Systems, Minerals, Metals
Materials Society, Warrendale (Pennsylvania), 1995.
Textbook of Materials and Metallurgical Thermodynamics
T-----r
Mio
A + BC Tz AB + c
T2
t
...
t>
(Ho
T2 (Ho -
T2 H~.>c
......
:J
J______
E
t>
E-o
r,
__ J_____
A + BC Mio
T1
AB + c
Process path - - +
Fig. 3.4 Illustration for derivation of Kirchhoff's Law.
EXAMPLE 3.3 Calculate the adiabatic flame temperature when C 2H4 gas is ignited at
298 K with stoichiometric amount of oxygen.
Given: (i) Heat of combustion (&/) of C 2H 4 at 298 K is -1411 .3 kJ per mole of C 2H4
(with gaseous reference state of H 20 at 298 K).
(ii) Cp of gases (J/mol/K)
Solution If a fuel is assumed to burn under adiabatic condition (i.e. no heat loss to
surrounding), then the resulting temperature of the products of combustion is known as
adiabatic flame temperature (Ta). The consequent heat balance equation may be written as:
Sensible heats of reactants at 298 K + hear released due to combustion at 298 K = heat
required to raise temperature of products from 298 K to Ta. The sensible heats of reactants
(i.e. C 2H 4 and 0 2) at 298 K are zero by definition of sensible heat. The combustion
reaction is:
(E.3.1)
Ta
1411.3 x 103 + O =
f
298
[2Cp(COi) + 2Cp (H20)] dT (E.3 .2)
Substituting the values and rearranging the terms in Eq. (E.3.2), and on integration, we get
Heat Capacity and Enthalpy-Auxiliary relations and applications
3
10
1411.3 x 103 = 113.9 (T" - 298) + l04.4 x - (T"2 - 2982 ) (E.3.3)
2
Equation (E.3.3) is a quadratic equation. By rearranging the terms, it becomes
= 4290 K
[Note: H 20 is a stable liquid at 298 K. If that is employed as reference state at 298 K,
the value· of Ml for combustion is to be adjusted as: Ml =-
1411.3 - 2Mlv, where Mlv
is the molar heat of vaporization of H 20. On the RHS also 2M:lv is to be subtracted. Thus,
these will automatically get cancelled, and the answer will be the same.]
The final temperature was 1000°C, and the fina1 product mixture consisted of AI20 3(s),
Cr(s), and some unreacted Cr20 3(s). Calculate the weight of Cr20 3 addition.
= lOOO
27
[J
298
932
Cp [Al(s)]dT+.1Hm(Al)+f
913
932
Cp [Al(l)]dT]
(E.3 .5)
Since the atomic mass of Al = 27, and M.P. of Al = 932 K.
Noting
Cp[AI (s)] = 20.67 + 12.38 x 10-3 T J/mol/K, Cp[AI (I) ] = 29.29 J/mol/K,
and
Lllim(AI) = 10.46 x 103 J/mol ,
we have sensible heat of reactants = 10.96 x 105 j oules
Textbook of Materials and Metallurgical Thermodynamics
1000
=- =37.04
27
(c) The product has excess Cr20 3, but no excess Al. Hence,
1
nc,.,., 0
.• 3
(added)= x > -2 nA1
So,
. od 1 37.04
nA 1 0 (m pr uct) = - nA1 = - - = 18.52
2 3 2 2
J298
{ 18.52 Cp[Al 2 0 3 (s)] + 37.04 Cp[Cr(s)] + (x - 18.52)jCp[Cr20 3(s)] }dT
(E.3.6)
2
Cp = a + bT + c1 J/mol/K, with the following values of a, b, c:
a bx 103 C X 10-S
Putting in the values in Eq. (E.3.6), and carrying out integration, we obtain
Sensible heat of products = 1.23 x 105 x + 9.58 x 105
Hence, the heat balance is
10.96 x 105 + 103. 15 x 105 = 1.23 x I0 5x + 9.58 x 105
i.e. 10.96 + 103.15 - 9.58 = l.23x or x = 85 g-moles
or weight of Cr20 3 added = 85 x (2 x 52 + 3 x 16) x 10- kg 3
= 13.0 kg.
3.4 Summary
1. For ideal monatomic gases and ideal diatomic gases at not too high temperature,
molar heat capacities at constant volume (Cv) and at constant pressure (Cp) are
given by the Kinetic Theory of Gases.
2. For all other substances, the values of Cp as a function of temperature are determined
experimentally, and are expressed by the relation
Cp =a + bT + cr-2
where a, b and c are empirical constants.
3. For solids and liquids, Cp and Cv are approximately the same.
4. Enthalpy changes are due to
(a) change of temperature of a substance, and
(b) other reactions and processes, e.g. phase transformations (i ncluding melting and
vapourization), reaction, and mixing; these are all treated as isothermal processes.
Textbook of Materials and Metallurgical Thermodynamics
PROBLEMS
3.1 Assuming no heat is lost to the surroundings, calculate the amount of heat required
to just melt 1 kg of Pb initiaJly at 15°C.
3.2 Consider oxidation of tungsten carbide (WC) as follows:
1
Mn (I) + 2 0 2 (g) = MnO (s)
Note the following transformation steps for Mn from 298 K:
Mn(a) ~ Mn(,8) ~ Mn(n ~ Mn(b) ~ Mn(l),
3.5 Calculate .11.f' and LJv° for the following reaction at 1000 K:
3.6 Which of the following processes releases more heat at 1000 K and by how much?
(a) Oxidation of graphite to CO
(b) Oxidation of diamond to CO
I
3.9 Calculate the heat of reaction at 1200 K per mole for t}\e following reaction:
I
ZnO (s) + C (graphite) = Zn (g) + CO (g)
[Note: From this chapter onwards, for problem solving, data are to be picked
up from the Data Tables in Appendix, as required.]
J
I
Textbook of Materials and Metaffurgicaf Thermodynamics
S. Carnot (1824) conceptualized an ideal heat engine which operates reversibly between
a heat source of constant temperature T2 and heat sink of constant temperature T 1• Such an
ideal reversible cycle is known as Carnot Cycle (CS), which laid the fou11datio11 of the
Second Law of Thermodynamics.
It may be emphasized here that, for reversibility, not only the process should be carried
out very slowly, but there should be no friction between the cylinder and the piston since
frictional heat dissipation is irreversible.
Figure 4.2 shows the Carnot cycle. It consists of four stages: two isothermal and two
adiabatic. The temperature of isothermal stage l ---+ 2 is T2, and that of 3 ---+ 4 is Ti.
T 2 > Ti. All stages are reversible.
t
Adiabatic
reversible
Volume-
Fig. 4.2 P-V diagram for Carnot cycle (schematic).
Stage 1 ~ 2 (isothermal)
L1U = 0 for isothermal expansion/contraction of an ideal gas (section 2.5.1). Hence, from
the First Law, and Eq. (2.13),
Stage 2 ~ 3 (adiabatic)
q2~3 = O; hence,
(4.5)
Heat Capacity and Enthalpy-Auxiliary relations and applications
3,7 Calculate .l!JJ° per mole of the following reaction, occurring at 900 K.
2Al (s) + Cr20 3 (s) = Al20 3 (s) + 2Cr (s)
3.8 Calculate the adiabatic flame temperature for combustion of the following gas
mixture: 25% CO, 12.5% C0 2, and 62.5% N2 (by volume). The initial tem?eratures
of the gas and air are 298 K. The theoretical amount of air was used. Assume that
air contains 21 % by volume 0 2, rest N2•
3.9 Calculate the heat of reaction at 1200 K per mole for t~e following reaction:
ZnO (s) + C (graphite) = Zn (g) + CO (g)
[Note: From this chapter onwards, for problem solving, data are to be picked
up from the Data Tables in Appendix, as required.]
Second Law of Thermodynamics
and Entropy
4.1 Introduction
The First Law of Thermodynamics merely states that if a process occurs, an exact equivalence
exists amongst various forms of energy changes. It provides no information regarding the
feasibility of the process. The Second Law of Thermodynamics provides means to predict
whether a particular process/reaction would take place under certain specified conditions,
and thus is of great importance. Another important aspect of the Second Law, which is
really fundamental to the problem enunciated above, is concerned with conversion of heat
into work.
A spontaneous process occurs without external intervention of any kind. Examples are
flow of heat from higher to lower temperature, diffusion of a species from higher concentration
to lower concentration, mixing, and acid-base reactions. All natural processes, i.e. processes
occurring in nature without external intervention, are spontaneous. In a refrigerator, heat is
pumped out from lower to higher temperature by an artificial device. This device consumes
electrical energy provided from outside.
4.2.1 Introduction
A device utilizing heat to generate mechanical work and operating in a cycle is called a heat
engine. The substance contained in it is called the working substance. Its operating feature
is illustrated in Fig. 4.1.
q)
Heat q2 Heat Heat
source at -
~
engine sink at
•
T2 -
q•
Ti
•
Work(W)
Fig. 4.1 Heat engine (schematic).
(4.2)
(Note: q denotes heat absorbed by the system, i.e. the engine; hence by convention
q) = - q1)
From Eqs. (4.1) and (4.2),
w q2 + q.
1] =-=~-~
q2 q2
(4.3)
.•
Textbook of Materials and Metallurgical Thermodynamics
S. Carnot (1824) conceptuali zed an ideal heat engine which operates reversibly between
a heat source of constant temperature T2 and heat sink of constant temperature T 1• Such an
ideal reversible cycle is known as Carnot Cycle (CS), which laid the foundation of the
Second Law of Thermodynamics.
It may be emphasized here that, for reversibility, not only the process should be carried
out very slowly, but there should be no friction between the cylinder and the piston since
frictional heat dissipation is irreversible.
Isothermal
i
Adiabatic
reversible
Volume-
Fig. 4.2 P-V diagram for Carnot cycle (schematic) .
q 2 _,. 3 = O; hence,
W2...,.3 = - (L1Uh...,.3 = - nCv(T1 - T2) (4.5)
Second Law of Thermodynamics and Entropy
Hence,
(4 .9)
(~r
T,
= ~ (4. 10)
r.
(~J' = Ti
T,
(4.11)
Hence,
( 4.12)
W T2 - T1
17 =-= ~-~ (4.14)
qi T2
Equation (4.14) shows that the efficiency of this reversible cycle depends only on T2
and r,.
Textbook of Materials and Metallurgical Thermodynamics
------
4.2.3 Efficiency of Generalized Carnot Cycle
In a generalized Carnot cycle, the working substance is any thermodynamic substance. It is
not restricted to ideal gas only. Section 1.3.3 has defined a thermodynamic substance which
is at thermodynamic equilibrium and obeys Eq. ( 1.3). Since a reversible process proceeds
through equilibrium states, it essentially has a working thermodynamic substance.
One of the statements of Generalized Carnot Theorem is:
All reversible cycles operating between the same upper and lower temperatures must
have the same efficiency. The proof of this statement can be given in the following way.
Figure 4.3 shows two engines working in cycles between the same heat reservoirs.
Heat reservoir at T2
Heat reservoir at T1
Engine I is working with ideal gas and transferring heat from reservoir at T2 to that at T1•
Engine II is working with another working substance and transferring heat in the reverse
direction (i.e. T1 ~ T2 ). Both engines are reversible.
Assume that
Efficiency of engine I, 711 > efficiency of engine II, 7111 (4.15)
Then ,
W1 qi (I) + qi (I)
711 =--
qi (I)
= qi (I)
(4.16)
or
W1 + W11 > 0 (4.20)
Therefore, combining Eqs. (4.19) and (4.20), we obtain
(4.23)
EXAMPLE 4.1 Two bodies ( 1 and 2) of equal heat capacity and initial temperatures T1
and T 2 form an adiabatically closed system. What would be the final temperature if one lets
the system come to thermal equilibrium (i) freely, and (ii) reversibly? What is the maximum
work that can be obtained from the system?
Solution (i) Since the bodies are in adiabatic enclosure, no heat is exchanged with the
surrounding. Therefore,
(E.4.1)
where m is mass of a body and mCp its heat capacity, and Tl is final temperature after
attainment of thermal equilibrium . .
From Eq. (E.4.1),
_1j+T2
TI -
2
Textbook of Materials and Metallurgical Thermodynamics
(E.4.4)
(E.4.5)
or
T1 = (T1T2)112
Work done by the reversible engine (Wrev) is the maximum work that can be obtained
from the system.
Since the system is adiabatic,
Total heat converted into work = Wrev = mCp(T1 - T1) - mCp(T1 - T2)
= mCp[(T1 + T2) - 2(T1Ti) 112 ]
= mCp(VT1 - VT2) 2
4.2.4 Thermodynamic Temperature Scale
In Carnot cycle for ideal gas, the temperature Twas the absolute temperature as arrived at
from Ideal Gas Laws. Since the efficiency of all reversible engines operating in cycle
between the same temperatures are the same, Kelvin proposed that the absolute temperature
scale be based on this thermodynamic conclusion. He called it the thennodynamic temperature
scale. However, it was made the same as that based on ideal gas by fixing the freezing
temperature of pure water (0°C) as 273.16 and that of boiling water as 373.16. In his
honour, the name of the unit is known as Kelvin (K).
Second Law of Thermodynamics and Entropy
4.3 Entropy
q2 + q1 = T2 - 1j
(4.24)
q2 T2
i.e.
(4.25)
for two isothermal stages. For adiabatic stages, q = 0. Hence for the entire cycle,
L 51..=0 (4.26)
cycle T
Consider any arbitrary cycle. If it is reversible, then it may be regarded as consisting
of a large number of tiny CS, as shown in Fig. 4.4. A ~ B ~A is the arbitrary reversible
cycle. The dashed lines divide it into many small Carnot cycles. The arrows indicate path
directions for the small CS. In an interior cycle, its own path directions and those of its
surrounding neighbours are opposite. Hence they cancel one another. Consider another CS
at periphery. Here, the peripherial arrows are not opposed. Therefore, the net uncancelled
lines of all the Carnot cycles are shown by solid lines.
Illustration of cancellation
effect for interior tiny
Carnot cycles
Volume--•
Fig. 4.4 P-V diagram for an arbitrary reversible cycle, divided into several tiny Carnot cycles.
Textbook of Materials and Metallurgical Thermodynamics
(4.27)
since it is the sum of several Carnot cycles. If CS are made infinitesimally small (i.e.
differential CS), then the solid lines would merge with the cycle ABA as the limiting case,
and Eq. (4.27) may be written as
f
cycle
i =O (4.28)
for the cycle ABA. Since it is a reversible cycle, heat is absorbed reversibly. Hence
Eq. (4.28) may be rewritten as
(4.29)
or
f
cycle
dS =0 (4.30)
where
dS = 8qrev (4.31)
T
EXAMPLE 4.2 Calculate Af for each step of the three step cyclic process of expansion/
compression of the ideal gas of Example 2.2.
Solution Step I: Reversible isocltoric
= 10.617 x 103 J K- 1
Second Law of Thermodynamics and Entropy
Hence,
(M)cycle = .1S1 + Af11 + Afm = 103 (10.617 - 10.617) = 0
1. Entropy changes are to be calculated only through a reversible path. This restriction
is not there for any other state property, including enthalpy.
2. T he absolute val ue of entropy of a substance (S) can be determined. Thermodynamic
data sources provide the values of S~ for pure substances. This is a consequence of
the Third L~w of Thermodynamics, which will be presented in Chapter 12. This is
in contrast with energy where only changes are available in data sources.
The e ntropy changes are to be calculated in the following ways, which are analogous
to those for enthalpy.
dS = Sqrev = Cv dT = C d In T (4.33)
T T v
At constant pressure,
(b) Phase transformations can be made to occur slowly and reversibly. For a pure
substance, reversible phase changes (i.e. melting, boiling etc.) at a constant pressure occurs
at a constant temperature . Equation (4.31) is for an infinitesimal prcx:ess at const T. It can
be easily converted into a finite difference form at constant T. Then dS becomes /.l.S and
&Jrev becomes qrev (i.e. 6.H). Therefore, for melting of a pure substance,
• A uO
L1So = _LJJ1_m (4.35)
,,, T
m
(4.36)
(4.37)
(c) Combining the above, for substance A and for the process
A (solid) A (liquid) A (gas) A (gas)
at 298 K at Tm at Tb at T
we have
s~ - S~9s = J Tm
298
c~T(s) dT + /.l.S~ + J Tb
T.,
c~T(I) dT + 6.S~ + J
T
Tb
c~T(g) dT (4.38)
(d) For a solution, L1Sm etc. depend on the composition of the solution as well. Moreover,
equilibrium (i.e. reversible) transformation temperature keeps changing as transformation
progresses (see Section 10.4). Therefore, simple equations like Eqs. (4.35)-(4.38) are to be
modified.
EXAMPLE 4.3 Calculate entropy of one mole of liquid Pb at 1000 K from that at 298 K.
Solution
600
f 1000
600
3
[32.4 - 3.1 x 10- T] dT
T
Second Law of Thermodynamics and Entropy
600
= 64.9 + 23.6 In + 9.75 x 10- 3(600 - 298)
298
1000
+ 8.017 + 32.4 In - - 3.1 x 10-3(1 000 - 600)
600
= 107 .69 J K- 1 mo1- 1•
By nature, reactions and mixing are irreversible processes. Like enthalpy changes, L1S and
L1Smix are to be considered only for isothermal processes. However, they cannot be calculated
from L1ll and L111mix using equations of the type (4.35)-(4.37) in view of irreversibility. For
example, consider reaction (3.14) occurring isothermally at temperature T. For this,
(4.40)
EXAMPLE 4.4 Calculate the entropy per mole of the following reaction at 1000 K from
that at 298 K.
Solution From Eq. (4.40), AS'° at 1000 K for the above reaction is related to AS'~98
as follows:
I
= 67.36 - 64.85 - 2x 205.0 (from data sources)
= - 100.0 J K - 1
-
- f
298
600 6.C;
--dT
T
4810
-- -+
600
f 1000
600
6.C"
__
T
Pr!T (E.4.8)
where
L1H?000 =L1H~98 +
f600
LJC; dT - &!~(Pb) +
f 1000
LJC;: dT (E.4.9)
298 600
Now, L1H~98 = L1HJ(Pb0) at 298 K = - 219.2 x 103 J/mol PbO (from data sources).
Other data have already been listed above. Carrying out calculations, it is found that
L1H?000 = -217 x 10 3 J
As stated in this section, chemical reactions are irreversible. Therefore, L1S~ ~ &!~IT .
This is confirmed, since here,
The various interpretations of entropy evolved gradually over a period of several decades.
These are enumerated as follows:
1. dS = /jqrevlT for an infinitesimal , isothermal, reversible process.
2. Entropy is Time's Arrow, i.e. a fundamental indicator of time.
3. Entropy has relation with heat not available for work.
4. Entropy is a measure of disorder of a system or a substance, which is under consideration.
The first interpretation has already been dealt with . The second and third interpretations
will be arri ved at following further discussions in this chapter. The fourth interpretation will be
elaborated in relation to the Third Law and statistical interpretati on of entropy in Chapter 12.
Textbook of Materials and Metallurgical Thermodynamics
Conclusions regarding entropy changes for reversible and irreversible processes have been
arrived at in various ways. Two procedures for derivation are presented here.
Derivation procedure I
Consider absorption of an infinitesimal quantity of heat (8q) by a closed system at a
temperature T syst from its surroundings. Assume that both system and surrounding have
large heat capacities. Their temperatures remain essentially constant even after transfer of &,.
Hence, the system absorbs heat isothermally and reversibly. The surrounding also gives out
heat isothermally and reversibly. Thus,
oq (4.41)
(dS)system = +
~yst
8q
(dS)surrounding =- T (4.42)
SUIT
So,
1 1
(dS)syst + surr = 8q (-- - - -) (4.43)
Tsy st Tsurr
The reversible exchange of&, requires that the rate of heat transfer be very slow. This
is possible if
(4.44)
Hence, for reversible heat exchange, from Eqs. (4.43) and (4.44),
If Tsurr is significantly higher than Tsyst• then the heat transfer is irreversible. From
Derivation procedure II
Consider the heat engine in Fig. 4.1 as the system and the two heat reservoirs as its surrounding.
SBCOnd Law of Thermodynamics and Entropy
That is, the engine and reservoirs are insulated from outside except for the heat and work
interactions indicated in the figure. Since entropy is a state property, for one cycle,
(M)engine =0 (4.49)
(L1S)heat reservoin -
_ _ qz + q1 _ _ q2 + q,
T
2
T., - T
I
T. ( )2 I
(4.50)
T2 - 1i
T/irr < T/rev• i.e. T/irr < (4.53)
Tz
i.e .
Tz - 1i
-W = qz + qi < -=--____.. (4.54)
q2 q2 T2
or
(4.55)
i.e.
q2 + !11.. < 0
(4.56)
T2 1i
Combining Eq. (4.56) with Eq. (4.50), we get
EXAMPLE 4.5 Calculate the entropy change of the Universe in isothermal freezing of
1 g-mole of supercooled liquid gold at 1250 K, from the following data for gold.
Solution
(i) LtSsyst
Actual freezing process is isothermal but irreversible, since it is not occurring at the equilibrium
freezing temperature. But ..dSsyst is to be calculated only along a reversible path. It makes
no difference as to which reversible path we choose since entropy is a state property.
The simplest reversible path is
!;.Ssyst = S2
- SI = f,1336 Cp(l) dT - Ml~ +
12S0 T 1336
f 12SO Cp(S) dT
1336 T
1336
29.69 12.36 X 10 3 + f (23.68 + 5.19 10- 3 T) dT
=
f
mo
- - dT- - - - -
T 1336 1336
12SO
T
X
= - 9.327 J K- 1
(ii) ..dSsurr
Both the initial and final states of the system have been assumed to be 1250 K. This
is possible only if the surrounding is at 1250 K, and it has infinite heat capacity (i.e. the
surrounding absorbs heat released during freezing reversibly at 1250 K). Hence,
Lllisyst =H2 - H1 =f
1336
29.69 dT - 12.36 x 103 +
f1250
(23.68 + 5.19 x 10-
3
T) dT
12.50 1336
or
LiS'universc = - 9.327 + 9.967 = + 0.64 J/ K
LiS'universe is positi ve since the process is irre versible, as the derivation in Eq. (4.48) shows.
Note: In Example 4.4 also, the reaction is isothermal at 1000 K, but irreversible. Hence,
the surrounding is at 1000 K and has large heat capacity. Hence the surrounding may be
assumed to absorb heat released by the reaction reversibly at 1000 K.
4.5 Summary
1. A heat engine converts heat into work and operates in cycle.
Efficiency of an engine = Tf = Wlq 2, where W is work done by the engine and q2
is heat absorbed by it from the source, in one cycle.
2. For a reversible Carnot cycle operating between source temperature (T2) and sink
temperature (T1), for any working substance,
T2 - 7;
TJ =~-~
T2
This is the basis for the thermodynamic temperature scale, proposed by Kelvin.
Textbook of Materials and Metallurgical Thermodynamics
dS = 8qttv
T
where 8qrt v is an infinitesimal quantity of heat absorbed by the system reversibly at
temperature T. For a finite process from state 1 to state 2,
S2 - S1 = f 2 8qrev
1 T
Hence, although S is a state property, change of entropy due to change of state has
to be calculated along reversible path only. For an isothennal process at T, AS = qrejT.
4. From the Third Law of Thermodynamics (not discussed yet), entropy of a perfectly
ordered crystalline solid is zero at 0 K. This allows determination and tabulation of
absolu~e values of entropy of substances at various temperatures.
PROBLEMS
4.4 Calculate the standard entropy of formation of TiC from titanium and graphite at
1200 K.
4.5 Calculate LiS for both system and surrounding per mole of the following isothermal
process at 800 K: Al (I) = Al (s). Assume aluminium to be pure. Is the process
irreversible?
4.6 Calculate LJS> at 1000 K for the reaction in Problem 3.5. What is the value of LiS
of Universe?
4.7 One gram of liquid Th02 at 2900°C is mixed with 5 g of Th0 2 at 3400°C adiabatica1ly.
(a) What is the final temperature?
(b) What is the entropy change of the system and the surrounding combined?
(c) Is the process spontaneous?
Assume Cp to be independent of temperature.
4.8 1 kg of a metallic powder initially at 25°C is mixed with 1 kg of the same liquid
metal, initially at 150°C, in a container insulated from the surroundings. Cp of both
liquid and powder are 4.0 J/g/K. Calculate the entropy change of the Universe for
this process.
Auxiliary Functions and Relations,
Criteria for Equilibrium
Since U, S, H are state properties and T is a state variable, A and G are also state properties
by definition.
dA = dU - T dS - S dT (5.7)
Combining Eq. (5.7) with Eq. (5.5), we get
dA = T dS - POV - OW' - T dS - S dT
= - S dT - P dV - OW' (5.8)
Differentiating Eq. (5.2), we obtain
dG = dH - T dS - S dT (5.9)
Combining Eq. (5.9) with Eq. (5.6), we have
dG = (T dS + V dP - OW) - T dS - S dT
= - S dT + V dP - OW' (5.10)
Equations (5.5), (5.6), (5.8) and (5.10) have been derived by combining the First and
Second Laws of Thermodynamics with the assumption of reversible process, and for fixed
mass and composition. If it is further assumed that the only work done is against pressure
(i.e. bW' = 0), then these equations get further simplified into the following equations:
dU = T · dS - P dV (5.11)
dH = T dS + V dP (5.12)
dA = - S dT - P dV (5.13)
dG =- S dT + V dP (5.14)
Equations (5.11)-(5.14) are four basic differential equations in thermodynamics, and
are valid only under the assumptions stated above. Closed and isolated systems have fixed
mass and composition. The equations are applicable to open systems as well, provided we
consider a fixed mass (1 mole or 1 kg, for example), and constant composition.
Textbook of Materials and Metallurgical Thermodynamics
(5.17)
Applying formula (5.17), we obtain from Eqs. (5. 11)-(5. 14) the following four differential
equations:
,(5.18)
(5.21)
Equations (5.18)-(5.21) are known as Maxwell's relations. These are the four basic
equations. By combining these with other equations, it is possible to arrive at many other
thermodynamic relations in the form of differential equations. They have been found to be
very valuable in science and engineering.
Another property of function (5.15) is
(oz)
8x Y
(ox) (oy) __ 1
oy z oz x
(5.22)
(2.39)
(8U)
8V T
=r(8S)
8V
_ T
p (5.23)
(8U)
8V
= r(8P) _p
T 8T v
(5.24)
C - Cv =r(8V)
P 8T
(8T8p) p v
(5.25)
(5.27)
a=..!_
v 8T
(8V) p
(5.28)
vra 2
Cp -Cv =p (5.30)
Equation (5.30) is valid only for a thermodynamic substance since Eq. (5.11), which has been
used for derivation, is applicable to reversible processes (i.e. for a thermodynamic substance).
circumstances, we wish to find the relationship of (:;:)q with physical variables. The
dU = (~l dV + (~l dT
which is the same as Eq. (2.7). Hence, at liq = 0,
(5.31)
Noting that (~)v = Cv• and differentiating Eq. (5.31) with respect to P, Eq. (5.31) can
be rewritten as
(5.32)
(Cp - Cv)l
(!;:). ~ (~),
= [ (:),
(5.33)
Auxiliary Functions and Relations, Criteria for Equilibrium
(5.34)
Again,
dV = (bV)
Sf
dT + (bV)
,5P
p
dP T
(5.36)
On the basis of definitions of a and p [see Eqs. (5.28) and (5.29)], we have
dV = Va dT - Vp dP (5.37)
or
bV)
(5P q
= va(sr)
fipq
- V/J (5.38)
(Sf)
fipq =[
vra vra
vra2] = Cp
(5.39)
Cv+p
u. Since
If we consider the effect of uniaxial tensile stress (o). then P may be substituted by
the coefficient of linear expansion is 1!3a. and the sign of tensile stress is opposite to
pressure, which is compressive in nature, Eq. (5.39) gets modified into
vra (5 .40)
3Cp
Equation (5.40) gives the thermoelastic coefficient (:J in terms of physical variables.
Experimental measurements have shown approximate agreement with this equation. The
Textbook of Materials and Metallurgical Thermodynamics
relationship between Cp and Cv, and the derivation of thermoelastic coefficient are two
examples of the usefulness of Maxwell' s relations. Many other relations have been obtained,
and are listed in some standard texts.*
Since it is easy to maintain temperature and pressure constant, the Gibbs free energy
criterion { Eq. (5.44 )] is ernployed in chemical and metallurgical thermodynamics. However,
in other areas, other criteria are also employed.
*See, for example, R.A. Swalin, Thermodynamics of Solids, Wiley, New York, 1964, p. 32.
Auxiliary Functions and Relations, Criteria for Equilibrium
(5.49)
(5.50)
or
(5.51)
q (rev) q (irr)
-1 - + -2- < 0 (5.52)
Ti Tz
In general, for a cyclic process A -7 B -7 A, let A -7 B be irreversible and B -7 A
be reversible. Then, with similar logic as given above,
(5.53)
(5.54)
Textbook of Materials and Metallurgical Thermodynamics
i.e.
S - S > :E 8q(irr)
D A A --+B T (5.55)
or
dS > 8q(irr) (5.56)
T
i.e. for an irreversible process,
8q < T dS (5.57)
Combining Eq. (5 .57) with Eq. (2.2) and Eq. (2.17) and for OW = 0, for irreversible process,
dU < T dS - P dV (5.58)
i.e.
(dU)sv < 0, i.e. (~U)s, v < 0 for finite process (5.59)
Similarly,
(dH)s,P < 0 or (~s.P < 0 (5.60)
then the process under consideration is not feasible. However, it may be noted that the
process would tend to proceed spontaneously in the backward (i.e. reverse) direction, since
for reverse process, the sign of (dG)r,P will be reverse, i.e. (dG)r,p < 0.
The above considerations have led to the minimum free energy criterion for equilibrium.
It is illustrated schematically in Fig. 5.1. It is evident from Fig. 5.1 that the system would
tend towards state 2 either from the left or from the right. At state 2, (dG)r,P = 0 since
it is at minimum.
Forward Backward
process process
l I
I
I
I
I
I
l
I
------ I
I
(dG)r,P < 0
for 1
(dG)r,P < 0
for
lI
I I I
1 State 1 ~ State 2 1 State 2 ~ State 3 1
Figure 5.1 constitutes the basis for the free energy minimization technique of searching
equilibrium state of a system. This is the general method for complex equilibrium calculations.
Figure 5.1 is two dimensional. For a three-dimensional arrangement, (G)T,P -is a surface
(like a bowl), and the minimum is at the bottom of the bowl.
5.8 Summary
1. Helmholtz free energy A = U - TS
Gibbs free energy, G =H - TS = U + PV - TS
By definition, free energy is a state propert) .
2. For a system of fixed mass and composition, and OW' = 0, the following criteria were
arrived at by combining the First and Second Laws of Thermodynamics.
Textbook of Materials and Metallurgical Thermodynamics
Of these, the criteria based on Gibbs free energy are employed in chemical and
metallurgical thermodynamics. For finite processes, dG etc. are to be replaced by L1G
etc. At equilibrium, (G)r,P is at its minimum.
3. The four differential equations for reversible processes constitute the basis for the
four basic Maxwell's relations.
2
4. Cp - Cv = vra , where V is molar volume, a is the coefficient of volumetric
f3
thermal expansion, and f3 is the isothermal compressibility of a substance.
As stated in Section 1.3, the state of a system decides its energy content. Again, the state
depends on many variables. Free energy is no exception. This issue is again being elaborated
for Gibbs free energy (G) with the intention of eventual understanding of its various
applications in chemical thermodynamics.
Gibbs free energy G of a substance is a function of temperature and pressu,re. But the
L1G criterion is useful for application only at constant temperature (T) and pressure (P) for
fixed mass of a system of constant composition (refer Chapter 5). Thus, let us first keep
T and P constant, and consider the dependence of G of a substance on other variables.
Free energy is an extensive property and is proportional to mass (or mole) of the
substance. Therefore, we should have a convention about the quantity of matter we are
talking about. By Universal convention in thermodynamics, tht!'tsymbol G generally refers io
Gibbs free energy per g-mole of a substance. However, G also refers to free energy of the
entire system. These will be specified in subsequent discussions. However, it is understood
that if someone wishes to adopt some basis other than g-mole, it would not make any
difference to conclusions in most cases.
The free energy of a substance of fixed mass also depends on
(a) chemical composition of the substance;
(b) state of ·aggregation of the substance (solid, liquid, gas or plasma);
(c) struc.~ure of the material, if it is a solid, such as crystal structure, microstructure,
substructure, macrostructure;
75
Textbook of Materials and Metallurgical Thermodynamics
95
= 0 +RT In t. +0
2.149
= - 212.5 J mo1- 1 at T = 263 K
Note: L1G 1 and L1G3 are zero because these steps are at equilibrium.
The standard state of a substance has already been defined in Section 3.3 in connection with
enthalpy . The same statement is applicable to free energy, i.e. the standard state is pure
element or compound at I atm pressure and at its stablest state at the temperature under
consideration (i.e. at ambient temperature). It may be stated further that it should be ideal,
if it is gaseous.
It has also been mentioned in Section 3.3 that the thermodynamic quantities at standard
state of a substance are distinguished by superscript zero (JfJ, SJ, G°, ...).
Now, in Eq. (6.6), if P 1 = 1 atm and P2 = P (i.e. a variable pressure), and G 2 = G,
then
L1G = G - G° = RT In P (6.8)
Simil arly, for nonideal gas, Eq. (6.7) may be integrated to yield
G - Go = RT ln(flt°) (6.9)
where J° is fugacity at standard state.
It is to be noted further that any gas tends to behave ideally if P ~ 0. i.e.
if P ~ 0, then f ~ P , flP ~ 1 (6. 10)
Gases tend to deviate from ideal behaviour more if pressure is increased and/or temperature
is decreased. The permanent gases exhibit significant departure from ideality only at extre mely
low (sub-zero) temperatures and/or very high pressures. In metals and materials processing,
these are rare. Most processing are at high temperature and moderate pressure. Hence
departures from ideal behaviour are small.
V= RT - a
p ( 6.11)
where a is a temperature dependent parameter. Hence, from Eqs. (6.7) and (6. 11),
6. 1 Introductory Comments
As stated in Section 1.3, the state of a system decides its energy content. Again, the state
depends on many variables. Free energy is no exception. This issue is again b_e ing elaborated
for Gibbs free energy (G) with the intention of eventual understanding of its various
applications in chemical thermodynamics.
Gibbs free energy G of a substance is a function of temperature and pressu,re. But tht
LiG criterion is useful for application only at constant temperature (7) and pressure (P) for
fixed mass of a system of constant composition (refer Chapter 5). Thus, let us first keep
T and P constant, and consider the dependence of G of a substance on other variables.
Free energy is an extensive property and is proportional to mass (or mole) of the
substance. Therefore, we should have a convention about the quantity of matter we are
talking about. By Universal convention in thennodynamics, thtasymbol G generally refers io
Gibbs free energy per g-mole of a substance. However, G also refers to free energy of the
entire system. These will be specified in subsequent discussions. However, it is understood
that if someone wishes to adopt some basis other than g-mole, it would not make any
difference to conclusions in most cases.
The free energy of a substance of fixed mass also depends on
(a) chemical composition of the substance;
(b) state of ·aggregation of the substance (solid, liquid, gas or plasma);
(c) struc.~ure of the material, if it is a solid, such as crystal structure, microstructure,
substructure, macrostructure;
75
Textbook of Materials and Metallurgical Thermodynamics
It is a lso to be kept in mind that we are concerned with change of free energy (i.e. L1G),
and not its absolute value. In chemical thermodynamics, strengths of other energy fields
[i.e. item (d) above] are assu med to remain constant during a process, and hence, its
contribution to L1G is negligible.
With the above discussion in mind, we shall consider the application of Gibbs free
energy in thermodynamics of materials, which is based on chemical thermodynamics as
stated in Chapter 1. It is to be further noted that quantitative application is possible only
for thermodynamic equilibria in view of the criterion (L1G)r. p = 0. Of course, this criterion
is also applicable to reversible processes. However, it is beyond the scope of any introductory
text on thermodynamics.
For irreversible processes, the criterion (L1G)r,P < 0 is qualitative in nature. It can be
employed for assessment of feasibility of a process only, and the same will be occasionally
illustrated. Hence, for the reasons cited above, any text on chemical/metallurgical/material
thermodynamics is basically concerned with the study of thermodynamic equilibria by
which, we mean (a) phase equilibria in heterogeneous system, or (b) chemical equilibria in
either homogeneous or heterogeneous systems.
In this chapter, we shall take up the si mplest system of constant composition , as generally
applicable to the one-component system (a pure element or pure compound). In subsequent
chapters, more complex situations involving chemical and phase equilibria in multicomponent
systems, where compositions of phases are variable, will be taken up one after another.
Systems involving changes of electrical and surface e nergy, i.e. thermodynamics of electro-
chemical cells and thermodynamics of surfaces, will be dealt with in Chapters 13 and 14.
As illustrated in Chapter 5, in solid state processes where volume remains approximately
constant, internal energy (U) and Helmholtz free energy (A) criteria are more appropriate.
In a way, these topics also belong to thermodynamics of materials but traditionally they are
not covered in a subject like the present one.
G i • Hi · Si and Gi, H-1, Si are free energy, enthalpy and entropy values at states 1 and 2,
respectively. Then,
(6.2)
then what should be L1G? If A is solid or liquid, L1G .. ,.,, . 0, unless the change of pressure
is very large since pressure does not have much effect on their energy (see Section 3.3).
However, if A is a gas, then change of pressure would have a significant effect on G. From
Eq. (5.14), at const T, dG = V dP. If A is an ideal gas, then
The unit of pressure in chemical thennodynamics is one standard atmosphere (760 mm'Hg},
simply designated as atm.
If the gas is not ideal, then, in analogy with Eq. (6.5), a .function known as .fugacity
(/) has been defined as follows:
dG = RT d(Jn f) (6.7)
It may be noted here that the above definition of .fugacity is a general one applicable to
all thennodynamic substances, and not restricted to nonideal gases only.
EXAMPLE 6.1 The equilibrium vapour pressure of ice at - 10°C is 1.95 mmHg, and that
of supercooled water at the same temp~rature is 2.149 mmHg. Calculate the free energy
change accompanying the change of 1 mole of supercool'ed water to ice at -10°C.
Solution Free energy is a state property. Hence we can choose any path for calculation
with the same result. Let us choose the following path:
The standard state of a substance has already been defined in Section 3.3 in connection with
enthalpy. The same statement is applicable to free energy, i.e. the standard state is pure
element or compound at I atm pressure and at its stablest state at the temperature under
consideration (i.e. at ambient temperature). It may be stated farther that it should be ideal,
if it is gaseous.
It has also been mentioned in Section 3.3 that the thermodynamic quantities at standard
state of a substance are distinguished by superscript zero (I.fl, SJ, G°, ... ).
Now, in Eq. (6.6), if P 1 = 1 atm and P 2 = P (i.e. a variable pressure), and G2 = G,
then
L1G = G - G° = RT In P (6.8)
Similarly, for nonideal gas, Eq. (6.7) may be integrated to yield
G - G° = RT lnlfl.f) (6.9)
where .f is fugacity at standard state.
It is to be noted further that any gas tends to behave ideally if P ~ 0. i.e.
if P ~ 0, then f ~ P, PP ~ 1 (6.10)
Gases tend to deviate from ideal behaviour more if pressure is increased and/or temperature
is decreased. The permanent gases exhibit significant departure from ideality only at extremely
low (sub-zero) temperatures and/or very high pressures. In metals and materials processing,
these are rare. Most processing are at high temperature and moderate pressure. Hence
departures from ideal behaviour are small.
V =RT - a (6.11 )
p
where a is a temperature dependent parameter. Hence, from Eqs. (6. 7) and (6.11),
i.e.
!!._ dP (6.13)
RT
/) aP
In ( p = - RT + C (6.14)
(6.15)
RT =P (6.17)
v I
where P; is pressure of ideal gas corresponding to the values of V and T in Eq. (6.16).
Combining Eqs. (6.16) and (6.17), we get
f ::: p
(6.18)
p Pi
From experimental values of a, some values of f were calculated for H 2 , 0 2 , CH4 at
P = 50 atm at a few temperatures. These are given in Table 6.1. It may be noted that the
gases are fairly close to ideal, especially at 200°C. So at lower pressure, they may be
assumed to be ideal without any significant error.
From Eq . (5.14),
(8G)
oT p
= -S (6.19)
G =H - TS =H + r( 0G)
8T P
(6.20)
i.e.
1
T 8T
(8G) P -
G= -
T2
H
T2
(6.21)
or
[8(~T)l = - -~ (6.22)
Equation (6.22) is one form of G-H equation. The alternative form is derived through
differentiation by parts as follows:
So,
.8(GIT)]
[ 8(1/T) r
= -T2 _!_
T
(8G) + G= _
8T P
T2 [_!_
T
(8G) _!}__]
8T y2
P (6.24)
Gibbs Free Energy and One-Component Systems
o(GIT)]
[ 0(1/T) p
= _ yi (- !!_)
T2
=H (6.25)
dG = dG2 - dG 1 = d(L1G)
= CV2 - V1)dP - (S2 - S1) dT
= L1V dP - L1S dT (6.26)
So,
(6.27)
Since Eq. (6.27) is analogous to Eq. (6. 19), the following relations can be derived:
(6.28)
where L1G is change of Gibbs free energy for an isothermal process. Similarly, L1V, L1S, nH
also refer to changes in V, S, H due to occurrence of the isothermal process.
(6.34)
(6.37)
(4.37)
j
Gibbs Free Energy and One-Component Systems
Therefore,
(6.38)
transfomzations involving only condensed phases (i.e. solids and liquids). ( ~~ )eq is the rate
of change of equilibrium phase transformation temperature with change of pressure . ..1V1r>
L1S1r, ..1H1r are difference, respectively, of molar volume, entropy and enthalpy of A between
phases II and I.
Change of pressure has negligible effect on L:1V1r and ..1H1r. Also, change of temperature
is typically very small. Hence, these are taken as constants and equal to their values at the
standard pressure of 1 atm., i.e. at standard states of A in both phases. As discussed in
Section 3.3.2, values at standard state are designated by superscript 'O'. Therefore,
Eq. (6.38) may also be written as
(dT)
dP cq
..1V~· ~
= ..111~ (6.39)
For example, the Clapeyron equation for melting/freezing of a pure element or compound
may be written as
During melting, a substance absorbs heat, i.e. ..111~ is positive. For most solids, the volume
per mole increases upon melting, i.e. ..1\1! is positive. These make the sign of ( ";; )eq
positive as per Eq. (6.40), i.e. Tm increases with increase of pressure. Exceptions to this are
Textbook of Materials and Metallurgical Thennodynamics
Tu
Temperature - - - -
(b)
Fig. 6.1 Variation of: (a) GA (I), GA (II), and (b) LiG, with temperature (schematic).
H20 , bismuth etc., where Li~ is negative and the contraction of volume occurs upon melting.
Tm will decrease with increase of P. Figure 6.2 presents the phase diagram for H20 ; S, L
.•••
l
I •
v .•
•'
----~1H
· i~~---t11~~1~
+0.0024°C +0.0098°C 100°c 374°C
Temperature
Fig. 6.2 Phase diagram for H2 0 ; dashed lines show metastable extensions. (P. Saha, in
Elements of Ceramic Science, D. Ganguly and S. Kumar (Eds.), Indian Institute of
Ceramics, Calcutta, 1982, p. 94.)
Gibbs Free Energy and One-Component Systems
and V denote regions of stability of solid, liquid and vapour, respectively. Line SL is solid-
liquid transition boundary obeying the Clapeyron equation. It may be noted that increasing
P lowers T for SL.
(i) 910°C is the equilibrium transformation temperature for this at 1 atm pressure.
(ii) At 910°C, LiH~~r = 912 J moi- 1, densities of a- and yiron are 7.571 and 7.633
g cm-3, respectively.
(iii) Ignore change of volume with pressure.
Solution
dT)
(dP = LiV~. ~ (6.39)
eq LiH2
since R = 82.06 cm3.atm/(mol. K). Substituting the values in Eq. (6.39), we get
6
dT) = .,_ 0.06 x 1183 x 10- = _7 .88 x 10_3 K/atm
( dP eq 9.0 x 10- 3
Thus, change of T1r (LiT1r) due to change of pressure from 1 atm to 100 atm is
dT
LiTir = - (100 - 1) = - 0.78 K
dP
EXAMPLE 6.3 Carbon has two allotropes, graphite and diamond. At 25°C and I atm
pressure, graphite is the stable form . Calculate the minimum pressure which must be applied
to graphite at 25°C in order to bring about its transformation to diamond.
Given:
(i) Hf98 (graphite) - Hf98 (diamond) = - 1900 J/mol
.
(ii) ~98 (J/mol.K) are 5.735 and 2.428 for graphite and diamond, r~spectively.
(iii) Densities of graphite and diamond at 25°C are 2.22 and 3.515 gm/cm3, respectively.
dG = V dP - S dT
where LlV = Vdiamond - Vgraphite• where V is molar volume. This is how change of pressure
affects relative stabilities of graphite and diamond.
At 1 atm and 25°C (i.e. 298 K),
At the minimum pressure required for transformation (Pmin), LIG = 0. Hence, the integral
form of Eq. (E.6.1) is
0
f d (LIG) = JPmin (LIV) dP (E.6.2)
Jo.028.5 I
Now,
1 1
(
LIV= 12 x 10-6(--
3.515
- - -)
2.22
= -1.99 x lo- 6 m3 .mo1-1
(6.42)
j
Gibbs Free Energy and One-Component Systems
Again, vaporization and sublimation can occur over a wide range of temperature. For
example, water vapour co-exists at equilibrium with water from 0°C to 100°C. The equilibrium
vapour pressure (i.e. saturated vapour pressure) is a function of temperature. As Figure 6.2
shows, it is about 5 mm Hg at 0°C and increases with rise in temperature. Therefore, ,1J/1r
and T1r are variables, and hence would be denoted as L1H and T, respectively. In normal
phase equilibria involving vapours, the total pressure in the system is not very large and is
of the order of a few atmospheres at the most. Therefore, ,1J/ may be taken as equal to L1H°,
i.e. ,1H at standard state of 1 atm.
With the above modifications, from Eqs. (6.39) and (6.42),
dP) - -
L1H- 0
(6.43)
(-
dT cq TVv
where L1H° is either &-t;, i.e. heat of vaporization or ~, i.e. heat of sublimation, of a
pure substance.
Assuming the vapour to behave as an ideal gas, we have
V
v
= RT0 (6.44)
PA
where p~ is vapour pressure of pure A in atm. Unit of gas constant R would depend on
the unit of volume. Suppose Vv is in m3, then the unit of R would be m 3.atm mo1-1 K- 1.
The assumption of vapour as ideal gas is reasonable for metallurgy and materials science
since vapour pressures are typicaJly very low.
From Eqs. (6.43) and (6.44),
1 (dP) (6.45)
p~ dT cq
d(ln p~)cq
---~= (6.48)
d(l/T)
Either Eq. (6.47) or Eq. (6.48) is Clausius-Clapeyron (CC) equation i11 differential
form. The subscript "eq" on LHS can be omitted as Jong as we understand that we are
referring to equilibrium vapour pressure in the CC equation.
The total pressure P in the system should have effect on the value of p~ in the following
ways:
(a) Properties of condensed phase depend on P.
(b) If the system is at a very high pressure (say, abov~ 50 atm or so), then the vapour
would not behave as an ideal gas (see section 6.2.3). In that case, p~ should be
replaced by fugacity of pure A (i.e. J2 ).
This effec t can be neglected for chang._
of P in normal ranges.
Although properties of condensed phase depend on P, they have riegligible effect on
vapour pressure, as the following analysis shows.
Consider the situation where liquid A is kept in a closed chamber at temperature T and
inert gas pressure P. After some time, the pressure of vapour of A in the empty space would
attain the equilibrium value ( p~ ). Now, let us change inert gas pressure P to P + dP in
the chamber. Then after some time, a new vapour-liquid equilibrium will be established
with saturated vapour pressure as P~ + d P~ .
For change of P to P + dP, at new equilibrium, dG of condensed phase, viz.
(6.5 1)
Since Vvap is approximately 103 to 105 times that of Vcond per mole of A, dp~ ldP 1s
0
neghgibly small. i.e. change of total pressure has negligible effect on PA.
Gibbs Free Energy and One-Component Systems
!1H 0
lnp~ = - - - + / (6.52)
RT
where I is integration constant. At normal boi ling point, P~ = 1 atm, i.e. In P~ = 0 and
T = ~. Substitutjon of these in Eq. (6.52) yields
1np~ = - fjz 0
(~ - ~) (6.53)
(6.54)
where
L).~ =~(v) - ~(!) (6.55 )
In general, for both liquid A and vapour of A , Eq. (3 .11 ) is applicable indjviduall y, i.e.
where A, 8, C, D are lumped parameters, and are empirical constants. Equation (3. 11 ) has
been deri ved from
L1~ = (a + bT + cr 2
)vap - (a + bT + cr-2) 1iq
=& + & · T + LJ.c • r-2 (6.57)
On this basis. we may rewrite Eq. (3.11) as
(6.59)
Textbook of Materials and Metallurgical Thermodynamics
where m, n and I are lumped constants obtained from empirical Cp, Lill and vapour pressure
data. Equation (6.61) is the most widely employed equation. Since we are dealing with equil\-
brium vapour pressure, the subscript "eq" is generally omitted. Figure 6.3 shows vapour
pressure vs. temperature curves for some liquid metals. The unit of vapour pressure is atm.
-2
-3
-4
Oil
::i::
E 10-
2
-5 ~
E t(
...::sa) -6 ~
..."'"'
Q)
p..
10--4 -7
-8
-9
- 10
It may be mentioned that, if all terms in Eq. (6.58) are retained, then we obtain the
most general vapour pressure vs. temperature relation as
n1 k
In p~ = - + n In T +IT+ - +I (6.62)
T T2
where l and k are two additional empirical constants. However, this equation is rarely
employed.
Figure 6.2 has presented phase diagram for H 20. There, the point 'O' is the common
intersection of the lines for all three equilibria: (i) solid-vapour (SY), (ii) solid-liquid (SL),
and (iii) liquid-vapour (LV). 0 is known as the triple point. Figure 6.4 presents the phase
diagram for zinc. Both in Figs. 6.2 and 6.4, pressure means vapour pressure for SV and
LV equilibria which involve the vapour. For solid-liquid equilibria, pressure means total
pressure P above 0. At triple point, it is vapour pressure for SL also. The existence of a
unique triple point for unary systems can be predicted on the basis of Gibbs Phase Rule,
which will be discussed in Chapter 10.
l atm
t Solid
Vapour
Temperature ___.
Fig. 6.4 Phase diagram for zinc (schematic}.
Quite often, phase transformations occur very slowly. There are situations where
thermodynamically stable states (i.e. ultimate equilibrium) are not attained even in centuries
or even in long geologic periods. These are called metastable phases. They have stability,
but not the ultimate stability. Some relevant examples are now given:
l. Cementite (Fe 3C) is most common in microstructures of iron-carbon alloys. However,
cementite is metastable, and graphite is the stablest state of carbon.
Textbook of Materials and Metallurgical Thermodynamics
2. Glass is an amorphous supercooled liquid. The crystalline state is the stablest. But
glass can remain as such for ages at room temperature. Only upon heating for a
prolonged period above a critical temperature does it crystaJlize.
3. A mixture of hydrogen and oxygen can be kept at room temperature for a very very
long period without reaction. However, it is metastable. If it is ignited by a match
stick, H2 and 0 2 react with explosive speed to form H20.
Thermodynamics is capable of quantitative treatment of metastable phases just like
stable phases. In Fig. 6.2, for example, the dashed lines are extensions of solid lines, and
represent metastable equilibria.
EXAMPLE 6.4 The vapour pressure of liquid zinc above its melting point (692 K) is
given by the approximate formula:
The heat of fusion of zinc is equal to 6673 J mo1- 1. Derive a formula for the vapour
pressure over solid zinc below 692 K.
Solution At 692 K, solid zinc is at equilibrium with liquid zinc. Hence, at 692 K,
15.568 x 103
-8.532 = - +I'
692
Gibbs Free Energy and One-Component Systems
i.e.
I' = l 3.965
Thus,
3
0 ()
InpZn s =- 15.568 x 10 + 13.965
T
6.5 Summary
1. For an isothermal process (reversible or irreversible) at temperature T,
= =
(a) L1G G 2 - G 1 (H2 - TS2) - (H 1 - TS 1) =
L1H - T • L1S, where 2, I designate
state 2 and state l, respectively .
(b) For change of pressure,
dG = RTd (ln P), for ideal gas
= RTd (ln j), for nonideal gas, and condensed substances
where f is known as fugacity. A gas tends to be more and more ideal, as pressure
is lowered. Then, f ~ P. At room temperature and above, and pressure not too high
(say, lower than 10 atm) , the common permanent gases may be taken as ideal.
2. The two alternative forms of Gibbs-Helmholtz (G-H) equation may be stated as:
where L1G, L1H are changes of G and H, respectively, during isothermal processes.
3. Change of equilibrium phase transformation temperature of a pure substance in
condensed state, due to change of pressure, is given by Clapeyron equation, viz.
where L1V1r and L1H1r are changes in molar volume and enthalpy, respectively due to
transformation.
4. For phase equilibria involving a pure condensed phase and its vapour, the Clausius-
Clapeyron (CC) equation is applicable. The two forms of CC equation are
ln(p~)eq = m + n In T + I
T
where m, 11, I are empirical constants.
6. The temperature at which solid, liquid and vapour of a pure substance co-exist at
equilibr·1m, is invariant and is known as the triple point.
PROBLEMS
6.1 Find the change in free energy associated with isothermal expansion of 1 mole of
an ideal gas from 2 atm to 1 atm pressure at 300 K.
6.2 Vapour pressure of water at 25°C is 23.76 mm Hg. Calculate the free energy change
for transfer of 1 mole of water to vapour at 1 atm pressure at 25°C.
6.3 The variation of the volume of a liquid with pressure is given by: V = V0(1 - ftP),
where pis the compressibility coefficient and V0 is the volume at low (i.e. virtually
zero) pressure. Derive an expression of L1G for a liquid due to change of pressure
from P 1 to P 2. If, for water at 25 °C, V = 1.0029 cm 3 g- 1 at 1 atm, and p = 49 x
10-6 atm- 1, calculate L1G per mole of water due to compression from 1 to 10 atm
at 25°C. Would it make appreciable difference if pis assumed as zero?
6.4 Solid bismuth has a density of 9.80 g cm-3 at 25°C. Its coefficient of linear thermal
expansion is 14.6 x 10-6/°C. Density of liquid bismuth is 10.07 g cm- 3. The atomic
mass of Bi is 209. What would be the change in melting temperature at a pressure
of 1000 atm?
6.5 Calculate the minimum hydrostatic pressure required to stabilize Fe3C at 298 K.
Given:
(a) 3Fe(s) + C(graphite) = Fe 3C(s), LIG° = 19,045 J mo1-1 at 298 K.
(b) The molar volumes of Fe3C, graphite and Fe are 24.24, 5.34 and 7.10 cm3 • mo1- 1,
respectively.
(c) Assume the molar volumes to be independent of pressure.
6.6 From the vapour pressure vs. temperature relationship for liquid silver, calculate
(a) the heat of vaporization of liquid silver at the normal boiling temperature of
2147 K, and (b) heat capacity difference between liquid and gaseous silver.
6.7 Below the triple point (-56.2°C), the vapour pressure of solid C02 is given as
3ll6
In p(atm) =--T
- + 16.01
The molar heat of melting of C02 is 8330 joules. Calculate the vapour pressure of
liquid C02 at 25°C, and explain why solid C02 evaporates rather than melts in an
open~_£onJainer.
Gibbs Free Energy and One-Component Systems
6.8 The chemical reaction inside the retort for ex.traction of zinc at high temperature,
above boiling point of Zn, may be written as
The gaseous products of reaction enter a condenser (for condensing Zn(g) into
liquid Zn) at 950°C and leave the condenser at 450°C and l atm. Assuming that
the Zn vapour coming out of the condenser is at equilibrium with liquid Zn at
450°C, calculate the efficiency of recovery of Zn in the cond~nser.
------
Chapter 7
Activity, Equilibrium Constant and
Standard Free Energy of Reactions
7.1 Introduction
Solution
A solution is a phase containing more than one component, where the components are mixed
(i.e. dispersed) at atomic/molecular level. Aqueous solution with dissolved salts, acids etc.
is an example. A NaCl solution in water has two components: NaCl, H 2 0 . It is a binary
solution. If the sol ution contains H 20, NaCl and KN0 3, then it is a ternary solution, and so
on.
96
Activity, Equilibrium Constant and Standard Free Energy of Reactions
Amongst various composition parameters, the mole fraction has been adopted as the fundamental
and generally l1sed one in chemical thermodynamics for nonaqueous solutions. Components
in a solution (e.g. H20 , NaCl etc. in aqueous solution) would be designated by subscripts
l, 2, ... , i, j , ... . The general symbol is i.
where n; denotes the number of moles of component i in the solution, nr is the total number
of moles in the solution, and X; is the mole fraction of component i in the solution. By
definition,
:E Xi= 1 (7.2)
I
A gas mixture is also a solution, and an ideal gas mixture obeys Dalton ~s Law of partial
pressures, viz.
x. = J!.L (7.4)
' Pr
.
dQ' 80.')
= (--
on1 P.T.ni ....,n;.
(0<2'
dn 1 + - -
0"1,
) n1, ...,n;,
(7 .7)
except n1 except "2
Let
(80.')
on; P. T , n , •••,except n;
1
=Q; (7.8)
Combining Eq. (7.9) with Eqs. (7.1) and (7.5), we obtain II'
II•
I :I: Q•dn.
dQ = -dQ
nr
= I
-'----
nr
I I
= :I:1 Q1 dX1 -
(7.10)
(dG')P,r = ~ G1 dn1,
I
(dG)p, r = ~ G1 dX1
I
(7.11)
where (f = Q per mole for pure i, and Q,m = change of Q upon dissolution of 1 mole of ,.
pure i. From Eq. (7.12), for Gibbs free energy per g-mole of i,
(7 .13)
where G1 = partial molar free energy of i in the solution and G;" =partial molar free energy
of mixing of i in the solution.
Textbook of Materials and Metallurgical Thermodynamics
From the above discussions, it is evident that G, is nothing but G per mole of component
i in a solution. Therefore, in analogy with Eq. (6.9), we may write
(7.14)
where
f; = fugacity of component i in solution
]; = fugacity of pure i(i.e. at standard state of I)
Activity of component i in a solution ( ai) is defined as
0
ai = f;ffi (7.15)
From Eqs. (7.14) and (7. 15),
(7.16)
Therefore, from Eqs. (7.16) and (7.18), for ideal gas mixture,
a;= P; (7.19)
However, it is to be noted that ai is dimensionless and the unit of Pi is standard atmosphere.
Therefo re, a; is only numerically equal to the value of Pi in atm, for ideal gases.
The validity of Eq. (7.18) has also been verified by experiments. A closed container was
partitioned into two sections with a semi-permeable membrane, which allowed passage to
oxygen only. One section was evacuated. Air at I atmosphere was filled into the other
section. Then sufficient time was allowed for attainment of equilibrium by permeation of
oxygen through the membrane. It was then found that P0z in 0 2 + N 2 mixture was equal
to pressure of pure oxygen <P0i) in the section, which was initially evacuated. In other
words, in an ideal gas mixture, it is the partial pressure, which constitutes the thennodynamic
parameter, not the total pressure.
where L, M, ... and Q, R, ... are general symbols for reactants and products and l, m, ...
and q, r, ... denote number of moles. Then,
where LIG = actual change of free energy as a result of the reaction and Lid = value of LIG
if reactants and products are at their respective standard states at the temperature T of
reaction. From Eqs. (7.21) and (7.22),
=RT In ~I .aR···]
[
m =RT In J (7.24)
aL • t2M .•.
(L1G)p,T = 0 (5.44)
and hence,
Lid = - RT In [1) 31 eq = - RT In K (7 .25)
Since the standard state of a substance is at l atm pressure, L1G° of a reaction is function
of temperature only. Therefore, the equilibrium constant K is also a function of temperature
only.
Free energy is a state property. Hence, Hess ' Law (see Section 3.3.4) is applicable. Thus,
where tJd/ is standard Gibbs free energy of formation of a compound from its elements at
temperature T. Obviously, L1dr for an element is zero at any temperature, by definition.
Therefore, if we have values of Liq> s, we can calculate L1G° of a reaction. Hence, we shall
concentrate on variation of L1Gj> with T in the following sections, and also present a
graphical method of representation of the same, known as the Ellingham diagrams.
8(L1GIT)] = __
fill_
[ (6.28)
8T P T2
d(L1Gj>IT) __ L1H?
(7 .29)
dT T2
where L1H? = standard enthalpy change for formation of a compound from its elements at
temperature T. Since pressure is constant, Eq. (7.29) has been simplified and written as exact
differential equation.
If there is no phase change, then on the basis of Kirchhoffs Law, we may write
where
:E
elements
cp0 (7.31)
By Eq. (6.57), we have
= - ; 2
[ L1Hf (298) + Lia(T - 298) ll
•
+ -L1b
2
(T 2 - 2982 ) - L1c - (1T 1)]
- -
298
(7.33)
.I
1[
- 2 &lf (298)
r 2
1]
- L1a • 298 - -L1b • 2982 + L1c • -
298
where k .is a constant. This is because Aflf (298) is also a constant. Integrating Eq. (7.34), IT•
we get
1
2 [,1//f (298) +Lia • T - Lia • 298] (7 .37)
T
'
"
{
l
,,,, Textbook of Materials and Metallurgical Thermodynamics
L'.1Gf = I + mT + gT In T (7 .38)
Most thermochemical data books express temperature dependence of u:ft in the form of
Eq. (7 .38). For many reactions, ~ is approximately independent of temperature (i.e.
Li~ .., o) over a limited temperature range. Then Eq. (7.38) gets further simplified into
Noting the expressions for the variation of entropy with temperature from Section 4.3.2,
if there is no phase change, then for a pure substance,
~
_!!_ dT (7.41)
T
I
On the basis of the above, we may write
293
LiCO
_
T
P dT (7.42)
I
I
£JG?cn =[&?<298) + J: ~ ar] - r[ &1<298) + J: '7 ar] (7.43)
Combining Eq. (7.43) with Eq. (7.32), we can arrive at Eq. (7.36) as well. This is the
alternative procedure for derivation of Eq. (7.36). If Li~ = 0, then both ~ and L1.S'? are
independent of temperature, and Eq. (7 .39) is obtained. Then, combining Eqs. (7 .39) and
(7 .40), we get
I
EXAMPLE 7.1 Calculate LiG for the process: Al(s) = Al(!) at 1000 K.
f
Gived: for aluminium, melting point = 933 K; heat of fusion at melting point = 10.9
x 103 J mo1- 1; Cp = 32.5 and 29.3 J mo1- 1K- 1 for solid and liquid, respectively.
'.:
l
Activity, Equilibrium Constant and Standard Free Energy of Reactions
Solution
LIG1000 = LIH1000 - lOOO(LIS)1000 (E.7. 1)
1000
Llli1000 = Llli933 +
J.933 (LICp) dT
1000 LICp
LlS1000 = LlS933 + f933 - - dT
T
EXAMPLE 7.2 (i) Calculate Lie° at 600 K for one mole of the reaction:
NiO (s) + Co (s) = Coo (s) + Ni (s)
(i i) How much error is involved if LICp for this reaction is assumed to be zero?
Given: (i) LIHJ at 298 K for CoO and NiO are: -239 and -244.6 kJ mo1- 1, respectively.
(ii) ~98 values are 30.0, 52.9, 29.8, 38.l J mo1-1K- 1 for Co, CoO, Ni, NiO, respectively.
(ii i) The values of Cp (J mo1- 1K- 1) are as follows:
Co: 21.4 + 14.3 x 10- 3 T - 0.88 x 105 T - 2
CoO: 48.28 + 8.54 x 10- 3 T + 1.7 x 105 T-2
Ni: 32.6 - 1.97 x 10- 3 T - 5.59 x 105 T-2
NiO: -20.9 + 157.2 x 10-3 T + 16.3 x 105 T- 2
Solution (i) From Eq. (6.3),
(E.7.3)
From Eq. (7.42),
o o J (i()() LICp dT
LIS(,()() =L1S29s + 298 T (E.7 .4)
Textbook of Materials and Metallurgical Thermodynamics
[Note: If there are phase transformations, then the more general form of Kirchhoff equation,
as given in Chapters 3 and 4, are to be employed in Eqs. (E.7.3) and (E.7.4).]
Now,
L1H~s =L1HJ. 29g(CoO) - L1HJ. 298(Ni0), (from Hess' Law)
l
~98 = ~(Co0)+~93(Ni)j - l~(Ni0)+~98 (Co)j
= 52.9 + 29.8 - 38.1 - 30.0 = 14.6 J moi- 1K-1
LiCp = [Cp(CoO) + Cp(Ni)] - [Cp(NiO) + Cp(Co)] (E.7 .5)
From the data provided and from Eq. (E.7.5),
LiCp = 80.38 - 164.9 x 10-3T - 19.31 x H>5r2 J mo1- 1K- 1
Solving Eqs. (E.7.3) and (E.7.4), we get
Therefore,
Lid' = 5600 - 600 x 14.6 = - 3160 joules
Combining these with the relation: Lid' =- RT In K, the variation of K with T is given by
either
d(ln K ) LJHO
(7.47)
dT = RT2
or
d(ln K) LJHo
d ( l/T)
= - -R- (7.48)
The above are alternative forms of the Van ' t Hoff equation.
Activity, Equilibrium Constant and Standard Free Energy of Reactions
(7.49)
where the values of x and y will depend on the specific compound. M and MxOy are general
symbols for metal and metal oxide, respec ti vely. They can be solid, liquid or gaseous,
depending on temperature. Specific examples are:
2Ni + 0 2(g) = 2Ni0 (7 .50)
Free energy is an extensive property. Hence, the value of t!J.G of a reaction would depend
on the number of moles involved. In the Ellingham diagram, I mole of oxygen constitutes
the basis in the oxide formation reaction. Thus, the values of !J(j/ are per mole of 0 2.
(ii) The tJ.d/ vs. T curve for a reaction consists of some interconnected straight lines.
As long as there is no phase change either in the element or its oxide, !J(j/ is assumed to
vary linearly with temperature in accordance with Eq. (7.44). The slope of a line changes
if a phase change occurs in the element or the oxide. M, B denote, respectively melting and
boiling points of the element. For the oxides, these are denoted by [M], [B}.
Let us take the example of the formation of ZnO, as shown in Fig. 7.1. Liquid zinc boils
.6 at 907°C i.e. 1180 K (at point B), where the slope of the line changes upwards. At point
B, liquid Zn and gaseo us Zn at 1 atm pressure co-exist at equilibrium, and from the
equilibrium criteria, i.e. Eqs. (5.44) and (6.31), for
,.
Zn(I) = Zn(g, l atm), /JGJ = dzn(g) - dzn (l) =0 (7.53)
(7 6])
Textbook of Materials and Metallurgical Thermrv'
0 200 400 ~
or-----oH:'~u.;., r>
_.,,.
-l
-~~
-lt 0 ~
ABSOLUTE ZEf\
(7.49)
where the values of x and y will depend on the specific compound. Mand MxOy are general
symbols for metal and metal oxide, respectively. They can be solid, liquid or gaseous,
depending on temperature. Specific examples are:
2Ni + 0 2(g) = 2Ni0 (7 .50)
Si + 0 2 (g) = Si0 2 (7 .51)
4 2
3Al + 02(g) = 3Al203 (7 .52)
Free energy is an extensive property. Hence, the value of L1G of a reaction would depend
on the number of moles involved. In the Ellingham diagram, 1 mole of oxygen constitutes
the basis in the oxide formation reaction. Thus, the values of L1Cf/ are per mole of 0 2.
(ii) The L1d/ vs. T curve for a reaction consists of some interconnected straight lines.
As long as there is no phase change either in the element or its oxide, L1cft is assumed to
vary linearly with temperature in accordance with Eq. (7.44). The slope of a line changes
if a phase change occurs in the element or the oxide. M, B denote, respectively melting and
boiling points of the element. For the oxides, these are denoted by [M], [B].
Let us take the example of the formation of ZnO, as shown in Fig. 7.1. Liquid zinc boils
at 907°C i.e. 1180 K (at point B), where the slope of the line changes upwards. At point
B, liquid Zn and gaseous Zn at 1 atm pressure co-exist at equilibrium, and from the
equi\i.brium criteria, i.e. Eqs. (5.44) and (6.31), for
-·
I
-·
Fig. 7.1 Ellingham diagram giving standard free energies for formation of oxides from their
respective elements, as function of temperature [F.D. Richardson and J.H.E. Jeffes,
J. Iron Steel Inst. , 160, 261 (1948)].
Activity, Equilibrium Constant and Standard Free Energy of Reactions
is the same for both Zn(!) and Zn(g). Thus, there is 1w discontinuity in the curve at B. With
similar arguments, there will be no discontinuity at points M, [M] , [B].
(iii) The changes in slopes of lines at M, B etc. can be explained as follows. From
Eq. (7 .44), the slope of the line of .LiGf
vs. T is - ~ and the intercept of the line at
absolute zero (i.e. 0 K) is AH?. For ZnO formation below 1180 K, let A.s1 =Li.sf
and
LJH? = &I~ . Then,
LLS'? = 2 ~nO(s) - 2 ~n(I) - sg2 (g)
(7.55)
(7.56)
Combining Eq. (7.55) with Eq. (7.57), and Eq. (7.56) with Eq. (7.58), we get
(7 .60)
where ~ and &I; are entropy and enthalpy of vaporization of liquid zinc. Since ~ is
a positive quantity, ~ - &f <0, i.e. ~< &f.
As, the slope of .LiGf
vs. T curve is
-A.s1 and - ~ > - &f, the change of slope is positive at point B. Similarly, at point M,
the slopes of lines change in the upward direction, whereas for points [M] and [B], they are
opposite.
In Fig. 7 .1, all lines slope upwards except those for the formation of CO and C02• The
explanations for this are as follows: For formation of a metal oxide such as
2M (s, I) + 0 2 (g) = 2 MO (s, I) (7.61)
Now, entropy is a measure of disorder of a substance (see Section 4.4.2). More the
disorder, the higher is its entropy. Since the gas is a highly disordered. state in comparison
to solid or liquid, entropy of oxygen gas is much larger than the entropy of metal and oxide
in Eq. (7.62). Therefore, L15'° "" -S~<&>' and hence the slope(= -LIS°} is positive. For,
C (s) + 0 2(g) = C02(g) (7 63)
Textbook of Materials and Metallurgical Thermodynamics
one mole of 0 2 disappears, but 1 mole of gaseous C0 2 forms. Hence, L1s° is very small and
the line is horizontal in the Ellingham diagram. For,
2C(s) + Oi(g) = 2CO(g) (7.64)
The situation is reverse of Eq. (7 .62). One extra mole of gas is produced. Hence, L1s° for
the reac tion of CO is positive and the slope is downwards.
The above features of CO and C0 2 have very important consequence in extraction and
refining of metals as well as in chemical industries.
L1G65 = L1G25 = G2;0z - G2; - c&i = - 678 kJ per mole 02 (approximate) (7.66)
As points of clarification, L1G65, L1G64 denote L1G for reactions (7.64) and (7.65), respectively.
They are equal to L1~5 and L1~ respectively, since all reactants and products are at their
respective standard states at the temperature under consideration. That is why solids and
liquids have to be pure and gases at 1 atm.
Subtracting reaction (7 .65) from (7 .64), we obtain
2C(s, pure) + Si02 (s, pure) = 2CO (g, 1 atm) + Si (pure, s) (7.67)
for which L1~7 = L1~ -L1~5 = + 230 kJ per .mole of reaction (7.67).
As discussed in Section 5.6, a positive value of L1G (in this case, L1G°) means that
reaction (7 .67) is not feasible at constant' temperature and pressure. Therefore, pure carbon
cannot reduce pure Si0 2 at 1000°C and Pco = 1 atm. In other words, Si02 is more stable
as compared to CO at their respective standard states at 1000°C. This is what is meant by
relative stabilities of compounds.
In the Ellingham diagram, the line for Si02 is below that of CO at 1000°C. To
generalize, the line for a stabler oxide would be below that of a less stable oxide. Therefore,
Al 2 0 3 is stabler than Si02 .
At 2000°C, Fig. 7.1 shows that CO is stabler than Si02. Hence, carbon can reduce Si02•
Activity, Equilibrium Constant and Standard Free Energy of Reactions
This interesting feature of the oxide system allows carbon to reduce any oxide into metal
if we go to a sufficiently high temperature. This is the thermodynamic basis for extraction
of metals from oxides, with carbon as reducing agent.
It is to be specially noted that the actual criterion for feasibility of a process/reaction is
(L1G)r. p < 0 [see Eq. (5.62)]. L1G° is equal to L1G only when all reactants and products are
at their respective standard states, as illustrated above. More general criterion (i.e. L1G) for
aJl states, both non-standard and standard, will be taken up in subsequent chapters.
Since all reactants and products are pure and under 1 atm pressure, L1G for reaction (E.7.6)
is the same as L1G°. Hence, if the reaction is at equilibrium at a temperature T, then
(E.7.7)
The forward reaction would occur if L1(j/. < 0 , in which case, BN would have more stability
than Si 3N4 • On the other hand, if L1(j/. > O, then Si 3N4 will be stabler than BN.
From Hess' Law, at temperature T,
B(s) + .!_ N 2(g) = BN(s), L1G° = - 108, 800 + 40.6T J mo1-1 at 1200-2300 K.
2
3Si(s) + 2N2(g) = Si 3N.(s), L1G° = -753, 200 + 336.4T J mo1-1 at 298-1680 K.
Substituting the above values into Eq. (E. 7 .8), the equilibrium temperature (Teq) is
obtained from the following equation:
0 = 4(-108, 800 + 40.6Teq) - (-753, 800 + 336.4Teq)
yielding Teq =
1831 K. However, Si melts at 1680 K, and this temperature is above that.
Therefore, this is an invalid solution.
(ii) Situation 2 Assume Si to be a liquid.
3Si(l) + 2N2(g) = Si~.(s), for which L1d = -892, 950 + 419.3T J mo1- 1 at 1680-1800 K
Combining this with L1G,.> of BN, we get
0 = 4(-108, 800 + 40.6Tcq) -(- 892, 950 + 419.3Teq) = +457, 750- 256.9Teq
Textbook of Materials and Metallurgical Thermodynamics
yielding Tcq = 1782 K. This is a valid solution. If T > Tcq• then .1dj. < 0. Hence, BN would
be relatively stable above 1782 K. Below 1782 K, Si 3N4 would be relatively stable.
7.6 Summary
1. The free energy criterion is useful for application at constant pressure and temperature
(see Chapter 5). Of these, the constant pressure restriction is not significant for
condensed phases. But the process must be isothermal. Hence, from now on, constant
temperature will be assumed in all derivations in this and subsequent chapters, unless
otherwise stated.
2. In a solution, various components have been designated as I, 2, ... , i, j, ... , the
general symbol being i. Mole fraction is employed as the general composition parameter
in thermodynamics of solution of non-aqueous systems.
n. n.
Mole fraction of component i -- X; -- -1:n1- -
-
-'
"r
1
I
- = (8G'
G; on. ) ' P,T,n 1,n2 , ..., except llj
= partial molar free energy of component i
4. (Jm
I
= RT In (iL_)
/;0 =RT Ina. I
we have
a3 x a~ x ...
L1G = fjd + RT In J, where J = acti vity quotie nt = a~ x aZ x
At equilibrium at constant P and fj,d = - /ff In CJJeq = - RT In K, where K is the
equilibrium constant.
6. According to Hess' Law,
L1G° I: .16'? -
= products I: L1G'?
reactants
where .16'? denotes free energy of formation of a compound from its elements.
7. .1<J? is a function of temperature and the most common form of the relation is:
.1<J? = I + mT + gT In T
where / , m, g are empirical constants.
8. Ellingham diagrams are specially constructed nomograms, giving variation of Lid
with temperature. Their properties and usefulness have been discussed by tiling the
example of formation of inorganic oxides from elements.
PROBLEMS
7.1 Calculate the equilibrium constant for the following reaction at 298 K:
7.2 Calculate the molar value of L1G for Si(s) ~ Si(l) at 1500°C. The equilibrium melting
point of solid silicon is 1410°C. Assume L1Cp 0. =
7.3 Calculate: (i) the entropy of transformation of Ti( a ) to Ti(/3) at the normal transformation
temperature of 882°C, and (ii) L1G of transformation at 930°C.
7.4 For 1 mole of the reaction: Ni(s) + 1/2 0 2(g) = NiO(s), calculate &fJ, tj,(}J, L1S° and
tj,d at 500 K.
7 .5 The standard free energy of formation of one mole of MgO is given as
.1<J? = - 603.9 x 103 - 5.355T In T + 393. l T joules
Calc~late the enthalpy and entropy of formation of MgO at 300 K.
7.6 Using the Ellingham diagram for oxides, show which metal will not get oxidized in
superheated steam at 1000°C- Ni or Cr? The conclusions should be based on quantitative
values of Lid.
Textbook of Materials and Metallurgical Thermodynamics
7.7 Predict the relative stability of Cr(s) and Cr23 C6 (s) at 1500 K.
12 2 3
Given: for Cr2 0 3 (s) + - C(s) = - Cr23 C6 (s) + - 0 2 (g)
23 23 2
.tKfJ = 1084.5 x 103 - 263. l 7T J moi- 1
7.8 For dissociation of silver oxide, i.e. Ag20(s) = 2Ag(s) + 21 0 2(g), the equilibrium
constant is 0.8 1 at 450 K. Calculate &I°, ..1S° and .tKf' for the reaction at 300 K for
one mole of the dissociation reaction .
7.9 The free energy of formation ( ..1dr') of NiO is given as - 150.25 kJ/mol, . and
- 133.43 kJ/mol at 1000 K and 1200 K, respectively. From these data, calculate
(a) equilibrium constant at each temperature
(b) ..1H~ and~ of NiO.
Equilibria Involving Ideal Gases
and Pure Condensed Phases
8.1 Introduction
In Section 7 .3, we have deri ved the basic equations for equilibrium calculations as well as
prediction of feasibility for chemical reactions. For reaction (7 .20),
q r
QQ_•aR ••.
l = (8 .1)
aL·~ ...
Also, K = [1Ja1 eq (8.2)
115
Textbook of Materials and Metallurgical Thermodynamics
X. = P; (7 .4)
I f'..
r
where
Xi = mole fraction of component i in the mixture
= volume fraction of component i in the mixture,
Pr = total pressure of all components, i.e.
EXAMPLE 8.1 Hr H20 and CO-C02 gas mixtures are frequently employed in high temperature
expenments in laboratory to obtain a gas phase with known partial pressure of oxygen <P0i.).
These gaseous compounds along with others are also common in metallurgical and materials
processing as well as combustion of fuels at high temperature.
we have
£i = - 239.534 x 103 + 8.139T In T - 9.247T, J mo1- 1 (E.8.2)
LiG? = - RT In K1 = - RT In [ °" 20
= - RT lo [ Pu o 2
Pu x p,'12
2
l
Oi eq
(E.8.3)
From Eqs. (E.8.2) and (E.8.3), and noting that R = 8.316 J mo1-1K-1, we have
As a sample calculation, at 2000 K, if we rr.quire POz = 10- 10 atm, then from Eq. (E.8.4),
1
CO (g) + 20 2 (g) = C02 (g) (E.8.5)
= - RT In K5 = - RT In Pcai
112
[ Pco x Pai
l eq
(E.8.6)
Therefore,
EXAMPLE 8.2 A gas mixture consisting of 1 mole S0 2 and 0.6 mole 0 2 is introduced
into a furnace at 1000 K. The total pressure (Pr) = 1 atm. Calculate the composition after
the mixture attains equilibrium in the furnace.
Solution For the reaction
S02 (g) + 112 02 (g) = S0 3 (g) (E.8.8)
we have
Thus, at 1000 K,
Ps<>J
Ks = 1.87 = (E.8.10)
Ps0i X P~
Let y mole of S0 3 be present in the equilibrium gas mixture. Then, the gas would consist of
(1-y) mole S0 2, (0.6 - 0.5y) mole 0 2, y mole S03
Hence,
11r =1 - y + 0.6 - 0.5y + y = 1.6 - 0.5y mole (E.8.11 )
Combining Eq. (E.8 .11) with Eqs. (7.1), (7.3) and (7.4), we get
L.6 ! o.5y]
---~----..;;;...._ ___ = 1.87
112
(E.8.14)
- y ] [0.6 - 0.5y]
0.5y 1.6 - 0.5y
Solving Eq. (E.8.14), y = 0.49 mole, and hence the equilibrium gas mixture will consist
of
37.5 vol% S02. 26.5 vol% 0 2 and 36.0 vol % 503 (Ans.)
[Note: In a gas mixture, composition is expressed in volume per cent, which is 100 x
vol. fraction (i.e. 100 x mole fraction as per Gas, Laws).]
EXAMPLE 8.3 A gas mixture consisting of 20% CO, 20% C02, 10% H 2, and 50% N 2
is fed into a furnace at 900°C. Find the equilibrium composition of the gas inside the
furnace.
Solution As already mentioned, percentage composition in a gas mixture is by convention
volume per cent, which is the same as mole percent.
At high temperature, chemical reaction would generate H20(g) as seen from the following
relation:
C02 (g) + H2 (g) = CO (g) + H20 (g) (E.8.15)
For solution, more equations are required. These are obtained from materials balance,
e.g.
Equilibria Involving Ideal Gases and Pure Condensed Phases
(E.8.16)
Let us consider 1 mole of the gas mixture at room temperature. Therefore, the number
of moles at room temperature (i.e. initial) are
(E.8.17)
From the Law of Conservation of Elements, the initial number of moles of each element is
the same as that after equilibrium. Hence, ·
nco =0.4 - ncai , ~20 =0.2 - ncai , nH2 =ncai - 0.1 (E.8.22)
Again,
L1G° (for E.8.15) = L1G0 (E.8.1) - L1G° (E.8.5) (E.8.24)
from Hess' Law.
From the values in Eqs. (E.8.2) and (E.8.6) at 900°C,
Pco x PH20
[ Peoz x PH2
l
eq
= l.ll (E.8 .25)
Tl;
P; = X;Pr = -·Pr (E.8.26)
nr
Combining Eqs. (E.8.22), (E.8.25) and (E.8.26), we obtain
Equations (8.6) and (8.7) are the same as the alternative forms of Clausius-Clapeyron
equation, as already derived in section 6.4.2.
(8.10)
where l Pai jeq is Pai in equilibrium with pure M and pure MO.
Figure 8.1 presents this relation graphically. On the equilibrium Pai line, both pure Cu
and pure Cu20 co-exist with 0 2• Below the line only Cu is stable solid, and above the line,
Cu 20 is stable solid. This can be predicted from common sense. However, for quantitative
prediction, Eq. (7 .26) is to be employed. For reaction (8.11 ),
l= - l ]
[
Pai actual
Textbook of Materials and Metallurgical Thermodynamics
-20
&'-40
-60
-80
4 8 12 16 20
l/T x 104 , K- 1
Fig. 8.1 Relative stabilities of Cu and Cu20 as function of Po2 and temperature.
J (pOi. )cq
= ---=-~ (8.14)
If (pOi. )actual < (po2 )cq , JIK > 1, and oxidation of Cu is not possible. So, Cu is stable solid
below the equilibrium line. The reverse is true above the line.
EXAMPLE 8.4 Determine the lowest temperature at which copper oxide (Cu 20) can
dissociate in a vacuum of 10-5 mm Hg.
Solution For dissociation of Cu 20, LiG for reaction (8.11) should be positive.
Noting that: L1G = L1G° + RT ln J
and combining this with Eq. (8.12), the thermodynamic criterion for dissociation is
-3.39 x 105 - 14.25T ln T + 247T + 8.314T In (760 x 105) > 0
Equilibria Involving Ideal Gases and Pure Condensed Phases
i.e.
-3 .39 x 105 - 14.25T In T + 397.87T > 0 (E.8.28)
In Eq. (E.8.28), the coefficient of T is (397 .87 - 14.25 In 7). Since In T is of the order
of 7 to 8 maximum, the coefficient of Tis positive (actually, it has to be positive for a valid
solution). Hence, the higher the value of T, the more positive will be the LHS. Hence, there
will be a lowest temperature (TmiJ above which dissociation of Cu 20 into Cu and 0 2 occurs.
At T = Tmin•
LHS = -3.39 x 105 - 14.25Tmin In Tmin + 397.87 Tmin = 0 (E.8.29)
This equation is implicit and can be solved graphically, or numerically through an iterative
procedure, or simply by trial and error. The solution of Eq. (E.8.29) yields a value of
Tmin = 1140 K . Therefore, Cu 20 will dissociate at T > 1140 K.
EXAMPLE 8.5 The thin layer of silicon can be deposited over a metallic substrate through
vapour deposition technique at high temperature, e .g.
Assume the temperature is 1200°C and pressure 1 atm. Calculate partial pressures of SiC14 ,
H2 and HCI gas at equilibrium with pure solid silicon. Assume that the molar ratio of
SiC1 4 : H2 in inlet gas is 1:2, as per stoichiometry of the reactants . .tK f' of the reaction is
- 4183 J mo1- 1 at this temperature (Johnson and Stracher, Chapter 3).
Solution
(E.8.30)
Assume the initial number of moles introduced as 1 mole SiC1 4 gas and 2 moles of H2 gas.
Let, upon attainment of equilibrium, n moles of SiC14 the reacted. Then the situation would
be as follows: ·
1- /1 2 - 211 4n
Psic14 =-3-+n-, PH2 = 3 +fl ' PHc1 =- -
3+n
(E.8.31)
Textbook of Materials and Metallurgical Thermodynamics
4
4n )
( 3+n
_ _,,____.:___ = 1.4
2 (E.8.32)
2 - 2n] 1- n
[ 3+n 3+n
i.e.
65.4n4 - 8.4n2 + l 1.2n - 4.2 = 0 (E.8.33)
P SiCl4 = 0.188 atm, PHz = 0.375 atm, PHa = 0.437 atm, Pr= 1 atm (check)
log (pn...)
L1d,
oc - - oc -
PM AQ
oc - (8. 16)
~L. eq T OP OQ
Since OQ is fixed, AQ oc log ( Po 2 )cq, and with a calibrated scale, AQ gives the actual
value of p 0 2 in equilibrium with Mn + MnO.
Next, consider the reaction
2Mn (s) + 2C02 (g) = 2Mn0 (s) + 2CO (g) (8.17)
10-2
-200
10~
N
104
B
io-8
0 .1
u c 10-10
-b0
e
..,),(
....
800
1<>6
1
10-12
10-14
~
1<>8
1 lo-•6
1000
' 10-18
A
1200 10-20
10-22
400 800 1200 1600 2000
Temperature, °C 10-24
Fig. 8.2 Procedure for finding out Po2 and PcolPc0i in equilibrium with a metal and its oxide,
from the Ellingham diagram.
Since OR' is fixed, with proper calibration, BR becomes equal to log [p00/Pcaileq·
Textbook of Materials and Metallurgical Thermodynamics
1
Ni(s) + z ·0 2(g) = NiO(s) (8.23)
NiS(s) + z
3
0 2(g) = NiO(s) + S02(g) (8.24)
1
NiO(s) + S02(g) + 20 2(g) = NiSOis) (8.25)
Since solids are pure, their activities are 1. Hence, the equilibrium relations for the
above reactions can be written as
Ni- NiO: log Po 2 =- 2 log K13 (8.27)
3
NiS-NiO: log Ps0z = log K 14 + 2log Po2 (8.28)
Here, Ps0z , p 0 2 are values in equilibrium with respective solids as shown above. Figure 8.3
presents the phase stability diagram for Ni-S-0 system at 1000 K. The plotting is based
on Eqs. (8.27)- (8.30). The straight lines correspond to two-phase equilibria. The areas
correspond to zones of stability of single solids. Such diagrams were termed as predominance
area diagram by T.R. Ingraham (1967) who first constructed some of these, including the
Ni-S-0 system.
+8
T=lOOOK
+4
-8
-12'--~~__........_~~--"-~~~....i......~~~.L-~~__J
- 20 - 16 - 12 -8 -4 0
log Po2
Fig. 8.3 Phase stability diagram for nickel-sulphur-oxygen system at 1000 K [T.R. Ingraham, in
Application of Fundamental Thermodynamics to Metallurgical Processes, Gordon &
Breach, New York, 1967, p. 187].
Textbook of Materials and Metal/urgi.::al Thermodynamics
8.4 Summary
1. For an ideal gas mixture, X; = p;fPr , where p; is the partial pressure of component
i, and Pr is the total pressure as given by Pr = :E P;. Again, as derived in Chapter 7,
a; = p;, in an ideal gas mixture. '
2. On the basis of the above, and equilibrium relations derived in Chapter 7, some
numeric· I examples of equilibrium calculations involving reactions of ideal gases
have been presented. It is to be noted that, in reality too, gaseous equilibria are
attained rapidly at high temperatures.
=
3. From the definition of activity in Chapter 7, a; l if component i is at its standard
state, which is pure i at its stablest state at the process temperature, for solids and
liquids (i.e. condensed phases). On this basis, some examples of equilibrium relations
have been derived and discussed. These are:
(i) The derivation of Clausius-Clapeyron equation
(ii) Pure M + pure metal oxide + 0 2 equilibria, for which
PROBLEMS
8.1 Predict whether a sheet of metallic iron, being annealed in a furnace at 760°C, will
get oxidized by the reaction:
Fe (s) + C02 (g) = FeO (s) + CO (g)
if the atmosphere contains 10% CO, 2% C0 2 and 88% N 2.
8.2 A gas mixture of 50% H2S and 50% 0 2 is introduced into a furnace at 1000 K.
Calculate the gas composition after the equilibrium for the following reaction is attained:
1 1
H2S (g) + 202 (g) = H20 (g) + 2S2 (g)
The total pressure is 3 atm.
8.3 (a) A gas mixture containing 50% C02 and 50% H 2 is introduced into a furnace at
1600°C. What is the value of partial pressure of oxygen in the gas after attainment
of equilibrium?
(b) Using the Ellingham diagram, predict whether this gas will oxidize a piece of
titanium kept in the furnace. Note down the procedure you employed.
Equilibria Involving Ideal Gases and Pure Condensed Phases
8.4 How much heat is evolved when 1 mole of S02 and 0.5 mole of 0 2, each at 1 atm
pressure, react to form the equilibrium SOrS02-0 2 mixture at 1000 K and 1 atm
pressure?
8.5 A partial pressure of Cl 2 of 0. i atrn is to be obtained at 500 K and total pressure
of l atm by establishing equilibrium of the reaction:
8.6 Calculate the pressure of inert gas which must be applied to liquid lead at 1000°C
in order to triple the equilibrium vapour pressure of Pb. The density of liquid Pb
at 1000°C is 9.79 g-cm-3.
8.7 Solid Mo0 2 is to be reduced by dry hydrogen gas at 1400°<: to metallic molybdenum.
The total pressure 1s 1 atm.
(a) What percentage of H 2 will be utilized for reduction assuming attainment of
equilibrium of gas with Mo and Mo02 mixture?
(b) Would there be any change if PT 10 atm? =
8.8 Predict whether a gas mixture consisting of 10% H2S and 90% H 20 can convert
solid Cn20 into solid Cu:!S at 1200 K.
8.9 Predict whether a gas mixture containing 50% H 2, 7% HCl, and 43% Ar at 900 K
and 1 atm total pressure is at equilibrium with a mixture of liquid Sn and liquid SnCJ 2.
8.10 One kilogram of solid calcium carbonate is kept in an evacuated chamber of one
cubic meter volume at room temperature, and the system is heated. Calculate:
(a) The highest temperature up to which solid CaC0 3 would be present
(b) The pressure inside the chamber at 1000 K
(c) The pressure inside the chamber at 1500 K.
8.11 Two of the oxides of iron are FeO and Fe30 4 . Solid iron exists in equilibrium with
one of these oxides at low temperatures and with the other one at high temperatures.
Predict which oxide would be at equilibrium with Fe at room temperature and the
maximum temperature for this equilibrium co-existence.
8.12 Liquid Mn is kept in a furnace at 1900 K. The furnace atmosphere consists of
2.5% H 2S, 50% H2 , and the remaining N2 . Predict whether MnS will form.
8.13 Solid ZnO and solid ZnS are equilibrated at 2000 K with a H 2S-H20-H 2 gas mixture
having P11~0 =0.5 and P11 2 = 0.0421 atm. Calculate the equilibrium partial pressures
of 0 2, H2S, S 2 and Zn gas in the atmosphere.
[Note: Gas corr,po~icion is given in volume per cent as per standard convention.]
Thermodynamics of Solutions
9.1 Introduction
For thermodynamic analysis, we can broadly classify solutions into aqueous and nonaqueous
solutions. Aqueous solutions are extensively dealt with in inorganic and physical chemistry
texts. Also, as stated in Chapter l, metallurgical and materials processing of interest to us
are mostly carried out at high temperatures. Here, the solutions are all nonaqueous and
inorganic, and hence, these are of special interest to us. Therefore, this chapter covers
inorganic nonaqueous solutions, especially for high temperature systems. These may be
further classified inco:
1. Metallic solutions
2. Non-metallic solutions-oxide solutions, sulphide solutions, halide solutions, etc.
Again , all these may be further sub-classified into solid solutions and liquid solutions.
Single-phase alloys (Ag-Cu, Fe-Ni etc.) are metallic solutions. A molten slag is an oxide
solution, and may contain compounds like Si02, CaO, and Al 20 3 •
No material (either element or compound) is absolutely pure. It will have some impurities,
may be in trace quantities. Even at such low concentrations they may affect some properties
significantly. Some examples are now given below:
l . Liquid steel dissolves hydrogen to the extent of few parts per million (ppm). Even
this low concentration of H tends to cause fine cracks on surfaces of solid steel
during hot forging, thus damaging its properties.
2. Liquid copper dissolves some oxygen during processing. Even a concentration less
than O. l % decreases ductility and electrical conductivity of Cu. Hence, for high
conductivity copper, oxygen should be removed to a very low level.
3. In semiconductors, even traces of impurities of the order of parts per billion may
affect performance.
130
Thermodynamics of Solutions
K -_PA -po
- A (9.1)
aA
i.e. '
aA_ -PA
-
(9.2)
p~
where aA is the activity of component A in the liquid solution . PA and P1 are equilibrium
vapour pressures over the liquid solution and pure A respectively at T.
Equation (9.2) can be understocxl by imagining two closed containers, kept at temperature T.
One contains a liquid solution having species A as a component. The other has pure A. If
the containers are kept at T for a long time, vapour-liquid equilibrium will be attained in
each of them. In one, vapour pressure of A will be PA• and in the other, it will be p~. If,
for a solution,
(9.3)
where XA is mole fraction of A in the solution, then the solution obeys Raoult's Law, and
is called an ideal solution. Solutions which do not obey Raoult 's Law are known as
nonideal solutions. To handle thermodynamics of such solutions, a parameter known as
activity coeff icient (y) is employed.
Y; =-X;Oi , 1.e.
. Oi =r, x; (9.4)
where
Y; =
Activity coefficient of component i in a solution, and X; is mole fraction of i,
defined in Eq. (7. 1).
I f,
r; = 1, solution is ideal.
Y; > 1, solution exhibits positive departure from Raoult's Law.
Y; < l , solution exhibits negative departure from Raoul t's Law.
In a solution , a; < I , in contrast to pure i , where a; = l. This difference is due to the
following two effects:
l. Dilution of component i in a solution due to the presence of other components
(dilution effect).
2. Interaction of component i with other components in a solution (interaction effect).
Textbook of Materials and Metallurgical Thermodynamics
Partial molar properties have already been defined and explained in section 7 .2.2. Partial
molar quantities of mixing also have been defined there. It may be noted here further that
all thermodynamic relations derived amongst extensive properties in Chapters 2-5, are applicable
here also. For example, in analogy with Eqs. (5.12) and (5.14), we may write
(9.8)
Similarly,
- - -
G; =H; - TS; (9.9)
- (OG'1
G. = - ·
I Jn.
1 ,P , ni , 112, ..., exceptn,
, - (JS'1
s. = -Jn.
I
1 , P,n-.n2 , ..., except n;
(9.11)
Thermodynamics of Solutions
- (8H'l
H.=
I
-8n.
' , P,n1, "2· .... except n;
' -=(8V'
v;
I
- l 8n.
1 • P,llJ ·"2· .... except n;
(9.12)
(9.13)
where ni. 11 2, ... , n; denote number of moles of components 1, 2, ... , i, ... etc. in the solution.
dQ ' =(OQ'
-8 l
n1 . "3 ..... "l except "1
dfli + - (OQ'
8ni
) ni. II). ····"l except nz
dni + ... + - (OQ'
8nk
J "1 • "3· ... except "t
dnk
(9.14)
nr = n 1 + n2 + .. . + n; + .. . + nk = l:n;j
(9.15)
Imagine preparation of this solution. One method of preparation is to keep adding small
quantities of components (dni. dn 2 , • .• ) in stages at constant T and P such that the overall
composition of the solution remains the same in all stages. Since the composition is always
constant, the values of Q1, Q2 , .. ., Qk would also remain constant. This is because these are
functions of T, P and composition of solution only.
Therefore, after the additions are complete, we get
(9.16)
Since Q is a state property, it does not matter as to how the solution is actually prepared.
It is the final state that matters. Hence, Eq. (9.16) will be always valid. Differentiating
Eq. (9.16), we get
(9.17)
r. x.dn. =o
. I lC.j (9.20)
I
Equations (9.1 9 and (9.20) are alternative forms of the Gibbs-Duhem equation (G-H
equation), which provide the principal foundation to thermodynamics of solutions.
For Gibbs free energy, the G-H equation may be written as
r. ni dG; =0
l
or ~ X; dG;
I
(9.21)
For pure components, the values of Q, viz. cf., dJ, ..., di. are constants at constant T and
P. So, d(f, ddj , ... , ddi_ are zero. Hence,
Subtracting Eq. (9.22) from Eq. (9.19) or (9.20), and from definition of Q;m in Eq. (7.12),
we get
- 0
!.1i;d(Q; - Q; )
j
=0, i.e.
-
r.j n; dQt•
=0, r. x .dQ!"
I
. I I
=o (9.23)
(9.24)
where the superscript 'O' denotes properties at the respective standard states of components,
as already noted in Section 7.2.2.
Subtracting Q0 from Q, we obtain
,1Qm =Q - Qo = ~X;(Q; l
- Qo)= ~X;Qt
l
(9.26)
LiQm is known as the integral molar value for mixing of the property Q. This represents
Thermodynamics of Solutions
the change in the value of Q per mole of solution when pure components are mjxed to form
the solution. Again, L1Q"' is a function of composition of the solution besides temperature
and pressure.
On the basis of Eq. (9.26), we may write
where L1Gm = integral molar .Gibbs free energy of mixing of the solution.
EXAMPLE 9.1 If the integral molar enthalpy of mixing (L1/f") of a binary solution is
LX 1X2, where Lis a constant independent of composition, find the expressions for Hf1 and Hi
in terms of mole fractions .
Solution
(E.9.1)
(E.9.2)
1
Hf1 =L [ 'Ii + n1ni(- l) x 1]
n1 +ni (n, + ni)2
Since a; = y;X;,
(9.29)
and,
iJ:n
~ = R(ln X; + In y;) (9.30)
The application of Gibbs- Helmholtz equation [Eq. (6.28)] to partial molar quantities
leads to
jj-rn
=- -·-
T2 (9.31)
In Eq. (9.31), the composition as also pressure is fixed since partial molar properties
depend on composition as well. Since X; is a composition parameter, and is independent of
temperature, we have
8(ln X; )] =O (9.32)
[
OT P.com
[8GilOP T,comp
=vm 1 (9.36)
Thermodynamics of Solutions
It is being emphasized again that Gr', V;m are functions of composition besides T and P.
Thus, constant composition restriction is required in Eq. (9.36), besides temperature.
For ideal solution, from Eq. (9.34), it follows that GI'1 is not a function of pressure.
So, from Eq. (9.36),
G!'l
-'-
T =R In X. I
-:t:. f(T) (9.38)
(9.40)
and, hence,
LlSm = - R ~ X; In X; (9.41)
'
Figure 9.1 presents activity vs mole fraction curves schematically for binary A-B solution
at constant temperature. It shows Raoult's Law (RL) lines for components A and B. Positive
and negative departures from Raoult"s Law are illustrated for a 8 . It may be noted that at
small values of X 8 , the variation of a 8 with X8 is linear. This is the basis for Henry's Law
(HL) for binary solutions. Henry 's Law may be rigorously stated as follows: If X8 ~ 0, then
~ ~ a constant ('fa). In other words, in Henry 's Law region for B,
Textbook of Materials and Metallurgical Thermodynamics
\
Raoult' s Law lines
o _____________
for B _
A ___.. B B
Mole fraction of B(X8 ) XB XB = 1
(a) Positive deviation (b) Negative deviation
Fig. 9.1 (a) Positive, and (b) negative deviations from the Raoult's Law for component B in the
binary A-B solution; Henry's Law lines for B are also ,shown (schematic).
(9.42)
where 'fa
is a constant. B is the solute, and A is known as solvent.
Similarly, if XA --+ 0,
(9.43)
One way of classifying solution is division into dilute (low concentration of solute) and
concentrated solution. On this basis, it may also be stated that the solute in binary dilute
solution obeys Henry 's lAw. There is no fixed demarcation between dilute and concentrated
solutions. From a fund.amental point of view, Henry ' s Law is expected to be obeyed if solute
atoms are so far apart from one another ~hat we can ignore solute-solute interactions, i.e.
B-B or A-A interactions, as the case may be.
It may also be observed in Fig. 9.1 that the curves for a 8 merge with Raoult's Law line
if X8 --+ l. Similarly curves for aA merge with Raoult's Law if XA --+ l. In other words,
in the region where B obeys Henry's Law, A tends to obey Raoult's Law, and vice versa.
The thermodynamic basis for the above observation is now derived.
From the Gibbs- Duhem equation, viz. Eq. (9.23),
(9.44)
i.e.
· R1lXAd(ln aA) + X 8 d(ln a 8 )] =0 (9.45)
by combining Eq. (9.44) with Eq . (7.16).
.
'"
Thennodynamics of Solutions
(9.47)
Again, XA + Xe = 1, and so
dXA + dXa = 0 (9.48)
Combining Eqs. (9.47) and (9.48), we get
(9.49)
EXAMPLE 9.2 The following partial pressures of Zn have been determined in Cu-Zn
alloys at 1060°C:
Xz.u 1.0 0.45 0.3 0.2 0.15 0.1 0.05
Pz.u (mm Hg) 3040 970 456 180 90 45 22
(i) Does the system obey Raoult's Law?
(ii) In what ranges of composition, does Zn obey Henry's Law?
(iii) What is the free energy change when 1 gm. atom of pure Zn at 1060°C dissolves
in a large quantity of the alloy at X'ln = 0.3? ·
Solution
(ii) In Henry's Law region, Yzn =constant. It is valid for XZn less than 0.1 to 0 .15 (XZn
< 0.12 as obtained by the graph plotting of Yz0 ).
(iii) When l g-atom of Zn dissolves in a large quantity of the alloy at 1060°C,
the free energy change = Gu, -
~ = GEt
(see Section 7.2.2 and Section 9.3).
9.5.2 Variation of .16"1, .1JITI and .15"1 with Composition for Ideal
Binary Solutions
For an ideal solution, Af-f" = 0 from Eq . (9.39). Further, if the solution is a binary A-B,
then from Eqs. (9.35) and (9.4 1),
Figure 9.2 shows variations of Af-f", L1S"1 and L1G"1 with composition at constant temperature
for an ideal binary solution. The curves of L1Gm, L1Sm, are symmetric with respect to A and
B. L1Sm is positive and is at maximum at XA = X8 = 0.5 . L1Gm is negative and is at minimum
at XA = X8 = 0.5.
A XA = X8 = 0.5 B
Xa-
Fig. 9.2 Variations of e.Hm, aS"' , aG"' with composition in an ideal binary solution at constant
temperature (schematic).
Thermodynamics of Solutions
(9.55)
Multiplying both sides of Eq. (9.55) by X8 , then adding LiGm, and finally combining with
Eq. (9.53), we obtain
d(Li<?) - -
..1(1lll + Xs = Xs(Gr;: - 0:) + L1Gm
dXA
(9.56)
Similarly,
(9.57)
Figure 9.3 presents a schematic curve of LiGm vs. composition for a binary solution at
constant temperature. It also demonstrates the graphical procedure to find out {J; and Gg1
at a certain composition from this curve.
=
By definition, N?1 0 at XA = I and X8 = 1 (i.e. for pure A and B). The zero of LtG"1 is
at the top of the diagram since LiGm for a solution is negative. The point q corresponds to &11
at X8 = X8 , i.e. the composition for which Gr; and GB1 are to .be determined, sqt is tangent
to the curve at q. The graphical ·procedure gives values of Gr; and Ga
as intercepts of the
tangent at XA = l and X8 = l , respectively. The basis for the same is explained as follows:
os = or + rs =pq + rs
- Ar'm X d(L1Gm) - ;:;m [from Eq. (9.56)]
-LA.I + B -VA
dXA
vt = vy - ty = pq - (- y t) = pq + yt
A B
Xa-+
Fig. 9.3 Graphical procedure for determination of (}:_, ~ from LIG"' vs. composition curve in
a binary solution.
.,..o = - RT In K = - RT In
LK.Jss 58
[ 1x P~o 2
]
(9.59)
[Gsi la11oy Pco, cq
since Osio = 1 .
'
Rearrangement of Eq. (9.59) yields
In [Gsi 1anoy = ~8
RT
+ 2 In ( Pco )
Pco
(9.60)
I CC)
"
Th6rmodynamics of Solutions ,,
9.6.1 Method 1
Equation (9.44) may be rewritten as the following equation using standard relation between
Gr and~:
(9.61)
A common logarithm has been chosen in the above equation for convenience of graph
plotting.
Integrating Eq. (9.61) for calculating aA from experimental values of a 8 , we get
(9.62)
The reason for this lower limit is because at XA = 1, aA = 1 and log aA = 0. Hence the
integration constant becomes zero.
Figure 9.4 is a graph (schematic) of X8/XA as function of log a 8. The figure is self-
explanatory. At the base line, XA = l, X8 = 0. Hence a 8 = 0, i.e. log a 8 ~ oo. This poses
difficulty for proper evaluation of the shaded area. The curve also tends to ~ asymptotic
to the y-axis towards the left-hand side, since XA ~ 0, X8 /XA ~ oo. For these reasons, this
:
method is not used.
Textbook of Materials and Metallurgical Thermodynamics
log aA at XA = XA is
given by the shaded area
-log as__...
Jog as at XA = XA
Fig. 9.4 Variation of log a8 with Xef XA in a binary solution, and calculation of log aA by Gibbs-
Duhem integration (schematic}.
9.6.2 Method 2
(9.63)
XA d log (;:) + Xsd( log i:) = XAd(log rAJ + Xsd(log )h) = 0 (9.65)
(9.66)
Xs/XA plot against log y8 is similar to Fig. 9.4 having the following features:
• At base line, XA = 1, Xs = 0. In view of Henry 's Law, log JB = log ~, and hence
has a finite value. Therefore, there is no problem at the lower limit.
• However, for XA ~ 0, X8 /XA ~ oo, and we shall have asymptotic behaviour as in
Fig. 9.4. Thus the shaded area will be somewhat erroneous for small values of XA-
Thermodynamics of Solutions
a. = In r·
I (9.67)
I (1 - X;)2
a - In r· In YA
ae - In Ye (9.68)
- -xi-
I -
A - (1 - X;)2 - -X-~-,
A
Xe
d(ln YA) = - - d(ln re)
XA
Xe 2
= - -XA (2aeXA dXA + XA daa)
(9.72)
(9.73)
where x and y are two variables. On the basis of Eq. (9.73), we have
(9.74)
Textbook of Materials and Metallurgical Thermodynamics
Substituting Eq. (9.74) into Eq. (9.72), and for integration limits XA = 1 to XA = XA,
we get
(9.75)
500°C. The finite values of £XccJ at XPb = 0 and XPb = 1 are demonstrated. The integral in
Eq. (9.75) is the shaded area, and can be reliably found out in the entire composition range.
Thus, this method is generally employed for G-D integration.
1.2
1.0
,.......
"'
~ 0.8
-l
........
0.6
0.4
00
~
0.2
0
0 0.2 0.4 0.6 0.8 1.0
XPb
Fig. 9.5 Determination of lt>b from J'cd for the Cd-Pb system at S00°C by use of a-Function; the
shaded area represents contribution of the final term in Eq. (9.75).
Procedures are also available for G-D integration in ternary solutions, and even quartemaries.
However, they are beyond the scope of this text.
Thermodynamics of Solutions
EXAMPLE 9.3 In liquid Fe-Ni solution at 1873 K, the activity of nickel as a function of
composition is known (noted in the solution that follows).
(i) Calculate the activity of iron at mole fraction of Fe of 0.6 by the Gibbs-Duhem
integration.
(ii) Calculate the following at XFe = 0.6: ~. C't ..dam, Gxs.
SolutWn (i) Given the values of aNi at various XNi as follows:
XNi = 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
a Ni = 0.067 0.137 0.208 0.287 0.376 0.492 0.620 0.776 0.89
We have the following calculations:
lNi = 0.67 0.685 0.69 0.72 0.75 0.82 0.986 0.96 0.99
(l-XNi)2= 0.81 0.64 0.49 0.36 0.25 0.16 0.09 0.04 0.01
llt.r; = -0.49 -0.59 -0.75 -0.92 -1.14 -1.24 -1.35 -1.07 -1.01
By method 3 of the Gibbs-Duhem integration (Eq. 9.75),
The graphical evaluation of the integral (as in Fig. 9.5) yields the value of integral as -0.08.
Hence,
In Jf;e = - (-
0.92 x 0.6 x 0.4 - 0.08) = 0.301, i.e.
(iii) GN\ =RT lo ~i = 8.314 x 1873 In (0.237) = -19438 J mo! at XFe = 0.6
In this connection, the regular solution model, first proposed by Hildebrand, is of great
practical utility for various uses. Again, it is an approximation, but is much more consistent
with behaviours of a large number of solutions.
Hildebrand defined a regular solution as the one in which
s.m=s.
l I
01
• id ' but 11~
I
*o (9.76)
i.e.
(9.77)
where the superscipt 'id' refers to ideal solution. In other words, a regular solution has the
same entropy as an ideal solution of the same composition. But enthalpy is different from
ideal solution.
(9.79)
Since the values of Q and Q; are the same (i.e. Q0 , (f) for standard states for both ideal
and nonideal solutions, we may also write
Qxs = (Q _ Q°) _ (Qid _ QO) = LlQm _ LlQm.id (9.80)
(9.85 )
Thermodynamics of Solutions
Again,
axs = LiGm - LiGm,id = RT"LX; In a; - RT"LX; In X; = RT"LX; In 11 (9.86)
i j j
Hildebrand showed that a;, as defined in Eq. (9.67), is not a function of composition in a
binary regular solution. If, in Eq. (9.75) a8 is a constant, then
In YA = - XaXAaB - aa(XA - 1)
= - XaXAaa + aaXa
= aaXa(l - XA) = aa x~ (9.87)
H
-m
B -- G-XS
s -- RT ln Ya -- RTa XA2 nx 2
-- :.~. A (9.91)
As discussed earlier, '1H of reactions and processes mostly do not vary much with
temperature. Similar has been the observations on &m, ii;" of solutions, which have been
found to be approx.imately constant in a limited temperature range (of course, for a fixed
composition). Hence, from the above equations, fl may be assumed to be independent of
temperature. Therefore,
a= -
n cc
1
(9.92)
RT T
Equation (9.92) is often employed for estimating the effect of temperature on activity etc.
where experimental data are not available.
From Eq. (9.89), if Lilr is positive, then a is positive. Then, on the basis of Eqs. (9.90'
Textbook of Materials and Metallurgical Thermodynamics
and (9.91), In YA and In 1h are positive, i.e. YA > 1, 1h > I, i.e. we have positive departures
from Raoult's Law. Similarly, negative &F1 means negative a and negative departures of
aA, a 8 from Raoult's Law. Again, an ideal solution may be looked upon as a regular solution
with a= 0 .
Figure 9.6 presents variations of LJJfD, L1Gm and LI.Sm for some hypothetical regular
binary solutions with a = + 1 and a = - 1. It also compares these with the ideal solution.
Of course, all these curves are at the same temperature.
L1.S"1 = LIS'11-id
- ve
A B
XA =Xe= 0.5
Xe--+
Fig. 9.6 LJHm, LJSm, LJGm as functions of composition at constant temperature for regular binary
solutions with different values of a (schematic).
EXAMPLE 9.4 Liquid lead-tin solution exhibits regular solution behaviour. At 746 K, it
is given that
In 1f>ti = - 0.737(1 - XPb) 2
(i) What is the corresponding expression for ~n at 746 K?
(ii) If 1 mole of solid Pb at 298 K is added to a large quantity of this liquid alloy of
composition XPb = 0.5, kept in a thermostat at 746 K, calculate:
(a) The quantity of heat required to be transferred from the thermostat to bring the
final temperature of alloy back to 746 K
(b) Entropy changes of system and surrounding
(c) Activity of Pb at 1000 K.
Solution (i) Since it is a regular solution, lXsn = aPb = - 0.737, and hence,
In ~n = - 0.737(1 - Xs 0 ) 2
Thermodynamics of Solutions
(E.9.3)
Xsn
=- (-0.737)XsaX'Pb -
f
1
(-0.737)dXsn
(E.9.5)
= f 600
298
[23.55 + 9.75 x 10-3 T] dT + 4212
746
+
f 600
[32.43 - 3.1 x 10-3 T] dT =17 ,673 J
Therefore,
Llli = 17,673 - 1142 = 16,531 J
(b) Similarly, L1S of the alloy as a result of addition is given by
s2
l'b(ll. 746
- {'() -
"'Pb(s>. 298 -
f 600
298
CP, l'b(s) dT + L1Hm (Pb) +
T Tm (Pb)
f146
600
CP,l'b(I) dT
T (E.9.7)
Therefore,
(.1S)alloy = (LlS)syst = 34.78 + 5.76 = 40.54 J K- 1
ln(yl'b)1000 746
=
In <rl'bh46 1000
Since at XPb = 0.5, Ji'b = 0.83, we have Ji'b at 1000 K = 0.88 and al'b = 0.~.
Let us assume that the solution contains NA atoms of A and N8 atoms of B. Each A-B bond
has one atom of A, whereas each A-A bond has two atoms of A. If Z is the coordinati on
number, then the number of A type atoms in A-B bonds is PAa!Z. On the other hand, the
number of A atoms in A-A bonds is 2PAA/Z. Hence,
(9.95)
Thus, PAA and Pas can be expressed in terms of PAB• NA> Na and Z. Substituting these into
Eq. (9.93), we obtain
NA atoms in pure A contain only A-A bonds, whose numbers are 1/2ZNA since each bond
has two A atoms. Similarly, N8 atoms in pure B have 1/2ZN8 B-B bonds. Hence,
If the total number of atoms (i.e. Avogadro's number) in a mole of solution is N0 , then
x =NA - Na
Xs- (9.101)
A N ' No
0
The probability of an A atom to occupy one atomic site is XA. The probability of a B atom
to occupy a specific site is X8 . For random mixing, the joint probability of an A atom and
a B atom to occupy neighbouring sites forming A-B bond would, therefore, be 2XAX8 . The
factor 2 occurs because A and B atoms can interchange sites mutuaJly also. Since the total
number of bonds is 1/2ZN0 , we have
1
PAB = 2XAXa 2ZN0 = XAX8 ZN0 (9.102)
1
Lilf" = XAXaZNo [HAB - 2<HAA + Haa)J (9.103)
(9.104)
(9.105)
Expanding the exponential and ignoring higher order terms, we get
(9.106)
9.9 Summary
1. This chapter deals with thermodynamics of condensed solutions, i.e. solid solutions
or liquid solutions. Temperature is assumed constant, unless otherwise stated.
2. A solution, which obeys Raoult's Law, i.e. a; = X;, is known as an ideal solution.
3. Activity coefficient is defined as: y1 = !!!.__ • '}'; = 1 for ideal solutions, but °* 1 for
nonideal solutions. x,
4. All properties of solutions are functions of composition at a constant temperature.
5. All standard thermodynamic relations derived earlier for pure substances are applicable
to properties of solution. For example, in analogy with equation : dG = - S dT +
V dP, we may write
~ n1 dQ=0,
I
L.X1 dQ
I
=0, r.i n1 dQt =0, L.X,dQr
I
=0
where Q is an extensive property of state (e.g. H, G, V, S).
(iv) Integration of the Gibbs-Duhem equation allows evaluation of aA(or YA) from
experimental values of a 8 (or rs) as a function of composition at constant
temperature. Darken's method is generally employed. It uses a-function, where
11. In a regular solution, entropy of mixing is equal to that of an ideal solution, i.e.
-s-m.id
S-m
; - ,. '
AC"IJI __ AC"IJl.id
LJt.) LlJ
However,
ii;n "# 0, mm "# 0
12. Excess property of a solution is defined as
PROBLEMS
9.1 Show that in a dilute binary solution A-B at a given temperature, ii;: and Yem are
independent of composition. Here, the subscript B refers to the solute component.
9.2 One mole of solid Cr at 1873 K is added to a large quantity of liquid Fe-Cr solution
at 1873 K with Xer = 0.2. If Fe and Cr form ideal solution, calculate the heat and
entropy changes in the system resulting from the addition. Ignore the difference of
Cp between solid and liquid chromium.
9.3 In a liquid Cu-Sn alloy at 1400 K and at X50 = 0.4, the values of activity of Sn
and Cu are 0 .333 and 0.362, respectively. Calculate PZn• Peu; Ysn• reu; G;!. ~.
L1Gm, and Gxs.
Thermodynamics of Solutions
9.4 Evaluate the thermodynamic feasibility of removal of aluminium from liquid silver
under vacuum at 1000°C. The maximum residual Al in the solution should be
1 mole percent. Henry's Law constant for Al( /1i) = 0.11.
9.5 The following data are given for liquid Cu-Sn alloys at 800°C:
Note: aSn values were experimental; aeu values calculated by Gibbs-Duhem integration with hypothetical
liquid Cu at 800°C as standard state.
(i) Do Sn and Cu obey Raoult's Law?
(ii) Find out the composition ranges for the validity of Henry's Law for Sn and Cu.
(iii) Calculate the following at Xsn = 0.6: PSn• Pcu; iJSn, L1Gm, Gxs. Gc:i.
(iv) By the Gibbs-Duhem integration, verify the value of acu at Xsn = 0.6.
(v) Is it a regular solution?
9.6 The following data are provided for liquid Cu-Ag alloys at 1300 K:
tXz,, =- - -,
19,250 w here R 1
. s .in J mo 1-t K-'
RT
Textbook of Materials and Metallurgical The""odynamics
10.1.1 Definition
In Section 1.3, we discussed about thermodynamic equilibrium. It was mentioned there that
for physico-chemical processes and chemically reactive systems, thermodynamic equilibrium
requires attainment of physico-chemical equilibrium (also referred to as chemical equilibrium)
besides mechanical and thermal equilibria. This means that chemical potential should be
uniform in the entire system in addition to uniformity of pressure and temperature.
J.W. Gibbs first proposed the term 'chemical potential' in analogy with other potentials
(e.g. thermal potential and gravitational potential). The meaning of the term will become
clear as we go along.
In Chapter 7, a distinction was made between a molar quantity (Q) and the same for
system as a whole (Q') vide Eq. (7.5), for convenience of derivation. By adopting this
convention, let us consider a system with variable composition. Then Eqs. (5.11)-(5.14) may
be modified as follows:
dU' = T dS' - P dV + -
8n1(oU') 5• v•
• ·"2' ···
dn1 + - (8U')
0"2 S'. v·.ni . ...
except "2
(oU')
d"2 + ... + -
On; S', V'.n1• ...
except n;
dn, + ...
except n 1
159
Textbook of Materials and Metallurgical Thermodynamics
Similarly,
/.4 = -
(8U'
8n;
J S', v'.ni . nz ... except n;
(1 0.5)
Therefore,
dU'=TdS'-PdV+ :tµ;dn; (10.6)
i
Again,
dH' = dU' + P dV + Y dP (2.25)
Combining Eqs. (10.6) and (2.25), we obtain
/.4 = -
[8H'
8n,
)
S',P."t, "2• ... except n;
(10.8)
Similarly, by combining Eqs. (10.3) and (10.4) with Eq. (10.6), the following relations can
be obtained:
/.4 (8A'l
= -8n.
1
,
,V ,n1, "2 ···· exceptn;
(10.9)
11; =(-8G'l
on, ( 10.10)
,P. n1, "2, ... except n;
In accordance with the above equations, therefore, there are four different definitions of
chemical potemial. In chemical thermodynamics, we employ Gibbs free energy criterion for
equilibrium. Hence, in this text as well as in other texts dealing with metallurgical thermodynamics,
the definition of chemical potential, as given in Eq. (10.10), will be employed. This way,
µ; becomes the same as the partial molar free energy of component i (i.e. q) in a solution,
as defined and discussed in section 7.2.2.
(10.11)
dG ,,l _
-µI
I dn'I + ... +J.liI d11;I + ... (10.1 2)
II d
. n,II + ... +J.liII d n;11 + ...
dG ,,ll -µ,
_
(10. 13)
(10.14)
Since,
(10.15)
Therefore,
where dG' is the total free energy change in the system as a result of exchange of dn 1
between phases I and II. For equilibrium at constant T and P,
dG' = 0, and hence from Eq. (10.16),
µf =Jlill (10.17)
Equation (10.17) can be generalizeo for all components with similar arguments. Hence,
at constant T and P, if phases I and II are at equilibrium, then
µ,1_11
- µj ' J..121-_J..12
u 1_n
' ... , µ; - µ; ' ... (10.18)
Similarly, Eq. (10.17) can be generalized for any number of phases at equilibrium as
Textbook of Materials and Metallurgical Thermodynamics
µ II -_ µIII -_ µIll _ _ µP
I - ... - I
(10.19)
[£1oi]
K - -- (10.20)
20 - [~]2
Here, [ ) denotes components in metallic solution. K20 has a nonzero value. If it is zero, then
..1~ = - RT In K20 would be infinity, which is not possible. Thus, [ lloil can not be zero.
To take another example, suppose a piece of pure Si0 2 is brought to equilibrium with
a molten Cu-Ag alloy. How can we think of µAg in Si0 2 and µSi in the alloy? The
explanation can be given with the help of the following reactions:
Si02(s) + 4[Ag) = [Si) + 2(Ag 20) (10.21)
(Ag20) = 2(Ag) + (0) (10.22)
Si02(s) + 4[Cu] = [Si] + 2(Cu20) (10.23)
In the above equations, ()denotes oxide phase. Some Si would dissolve in the alloy, through
reactions (10.21) and (10.23). Its concentration would be very small. Similarly, through
reactions (10.2 1) and (10.22), a very very small concentration of Ag in oxide phase can be
predicted from thermodynamic equilibria considerations as in the case of 0 2 in Eq. (10.20).
(10.24)
For Eq. (10.16), the proqess under consideration is transfer of component 1 from phase II to
phase I. [Note: dn 1 has been taken as positive. Hence dni is positive, i.e. phase I receives
dn 1 and phase II loses it.]
Equation (10.24) can be generalized to state that a component would tend to be transferred
from a higher to a lower chemical potential. This makes µ; analogous to thermal potential
etc. For example, heat flows from higher thermal potential (i.e. higher temperature) to lower
thermal potential.
Also, species i will be transferred from phase I to phase II, if aJ > llin .
Textbook of Materials and Metallurgical Thermodynamics
EXAMPLE 10.2 Alloy Phase Diagrams. Here, the system is non-reactive, the environment
of the alloy being either vacuum or an inert gas (Ar, He etc.). Again, pressure has hardly any
influence. So, Eq. (10.31) gets simplified into
°C Tin (atomic%) op
10 20 30 40 50 60 70 80 90
350 I I -
' ' I I I
300
~ - 600
'\· ~ -
L
I\ ~ - 500
250 ............. 232°
100
I a+P
-
- 200
50
I -
- 100
j
0 -
Pb 10 20 30 40 50 60 70 80 90 Sn
Tin (wt.%)
Fig. 10.1 Lead-tin binary phase diagram.
Textbook of Materials and Metallurgical Thermodynamics
of the liquid solution on liquidus and that of the solid solution on the solidus are fixed since
these correspond to equilibrium co-existence of two-phases.
At the eutectic point, there are thr~e-phases co-existing at equilibrium. Therefore,
F = 0 here, and is an invariant.
Reading components
If components are reacting amongst themselves, then the reaction equilibria constitute additional
constraints. However, all possible reactions are not independent. Only the number of independent
reactions are to be taken into account. In such a situation,
C=N-R (10.32)
where N is the number of species (elements and compounds) and R is the number of
independent reactions.
EXAMPLE 10.4 Consider a system of pure metal (M), oxygen and pure metal oxide.
There is only one possible ·reaction, viz.
l
M(s) + -02(g)
2
= MO(s) (E.10.2)
N =3 (M, 0 2, MO), R = 1, and P = 3. The gas phase is involved. So both T and P are
variables. Hence, F = (3 - 1) - 3 + 2 = l. If T is fixed ,. then P0i. at equilibrium is fixed.
M and MO are pure anyway. Therefore, the state of the system gets fixed completely by
fixing T. Figure 8.1 on the Cu-Cu 20-02 system illust~tes this.
Suppose, in the above system, M is present as a liquid solution M-X, where X is a noble
metal. Also, MO is present as a separate oxide solution, MO-BO, where BO is more stable
than MO. Then, in this case too, the only significant reaction is oxidation of M to MO. X
would not form oxide significantly, and BO would not be reduced to B significantly. Here,
N = 5 (M, X, 0 2 , MO, BO), R = 1, P = 3 (liquid metallic solution, gas, liquid oxide
solution). Hence, F = 3. This means that even after fixing temperature and pressure, compos.itions
of phases cannot be fixed.
Now, in addition to fixing T, let us fix compositions of metallic solution and oxide
solution independently. Then, all three degrees of freedom are utilized, and P0i of the gas
phase also gets fixed at equilibrium. This may also be explained from the use of equilibrium
constants as follows. For Eq. (E.10.2),
Chemical Potential and Equilibria amongst Phases of V~riable Compositions
Si02 (wt. %)
100 90 80 70 60 50 40 30 20 10 0
MELTING POINT OF SILICA (CRISTOBALITE), l 724uc , 3135-r
"''" .Tll'I r. POINT OF ALUMINA !CORUNDUM), 2050"C, 3722°F
3800
3600 ·-
Cl
3
0
"] ~
...--- 2050
- 2000
~
+
LIQ um VCORUNDUM
....
-;
3400 ,___ <
/I LIO+u111 - 1900
0
"«)
~
i2
a:l
I -
/ p ~' 17°F
-
'O
...
c<S
t:
... 3200 - ~
0 __... ~I I I E-5.i' 1800 .~
~
u
Ix/ I ..J c0
3Al2~ 2Si°'2
c<S ..J
I.I.. 3135
- 1700 u
y
::::>
Vl
0 / • (MUU.ITE)+ ,!QUID 6 Vl
0
0
@i
3000 ·v 2903°F I
Q"'
CORUNl>UM ...
0
t)I)
0 ' en +
JA~L· 2Si02
- 1600 0
5 .5%A~ca
0
2800
- 94.5% i ,, I I N
L TE) 0
CRISTOBAUrE + 3Al2 0 3 • 2Si0:2 (MUU.ITE)
•2578°Fl I I I I~ <(
- 1500
I
2600
i "'
,YMITE i 3Al, O,, . r 0:2 (MrITE) 71.8 %Al 10,
/8.2% S102
- 1400
2400
0 10 20 30 40 50 60 70 80 90 100
Al20 3 (wt. %)
Fig. 10.2 Silica-alumina binary phase diagram.
(E.10.3)
K is fixed by temperature, aMo and aM are fixed by temperature and composition. Thus, from
Eq. (E.10.3), (PO]. )eq is automatically fixed.
EXAMPLE 10.~ (i) Calculate the chemical potential of carbon of a gas mixture of CO and
C0 2 at 1000 K and 2 atm. total pressure. Assume PcdPc0i. = 103 .
(ii) Suppose this gas mixture is brought in contact with zirconium metal, would ZrC form?
I
C(s) + 202(g) = CO(g); ..1~ = -199.4 x 103 J mo1- 1 (E.10.6)
The values of ..1~, ..1cfi have been obtained from the data table in the Appendix. Hence,
=- RT In (
Pc0i. x
2
ac) (E.10.8)
Pco cq
2 1
Noting that PcdPc0i. = 103, Pco = 2 atm, Pc0i. =
1000
= 500 atm., and substituting these
values in Eq. (E.10.8), we get
ac = activity of carbon at equilibrium of reaction (E.10.5) = 7.78 x 10-4
Ile = RT In ac = -59.52 kJ mo1- 1
[Note : Actually, RT In ac =
fie - µg.
However, by convention, is zero [see µg
Eq. (10.26)). Also, as elucidated in section 10.1.3, although carbon is not present in the
gas, there will be a finite value of Ile in the gas.)
(ii) For, Zr(s) + C(s) = ZrC(s); ..1cft = - 41.5 kJ mo1- 1 at 1000 K (E.10.9)
Combining the above equations, and noting that the activities of ZrC and Zr are 1 since they
are pure, we get
1
RT ln(ac)eq = (µc)at eq with zr+ZrC = - 41.5 kJ mol-
Since
{µc)c0-co 2 gas < {µc)zr+ZrC• ZrC will not form.
As elucidated in Section 5.7, any system tends to lower its Gibbs free energy at constant T, P.
The equilibrium state corresponds to minimum (G)r.P· Therefore, the nature of a phase
diagram would depend on variation G with composition at various temperatures. As discussed
several times earlier, the constant pressure restriction is not important.
Section 9.5 has dealt with thermodynamics of binary solutions. Figure 9.2 has schematically
shown variation of L1Gm, Mf11, L1Sm with mole fraction for an ideal binary solution at
constant temperature. Figure 9.6 has schematically shown variation of L1Gm for regular
binary solution for both positive and negative values of a.
For a regular binary solution A-B, where A and B are general symbols of the two
components. From Eqs. (9.86) and (9.89), we have
(10.33)
i.e.
( 10.34)
From Eq. (10.34), L1Gm/RT as function of mole fraction can be calculated for different
assumed values of a. This shows that, if a > 2.0, the curve exhibits the maxima-minima
behaviour. This is shown in Fig. 10.3(a). The curve is symmetric with respect to X8 =
0.5 .
Textbook of Materials and Metallurgical Thermodynamics
XA = X8 = 0.5
+ve
.-----------*----------.. .
a b
stable I
-ve (a)
ly
0 --.--
I
I Xa
I
I
I
(b)
(c)
0
A B
Xa - -~
Fig. 10.3 Variation of: (a) free energy, (b) chemical potentials of A and B, and (c) activities of
A and B, at constant temperature, for binary system A-B exhibiting maxima-minima
type behaviour in .i:IG"' (schematic).
Chemical Potential and Equilibria amongst Phases of Variable Compositions II.ii
Positive a means positive departure from Raoult' s Law (see section 9.7.2). An arbitrary
solution, which exhibits such positive departure but is not a regular solution, would not be
symmetric with respect to X8 = 0.5. Figure 10.3(a) also shows such a behaviour schematically.
Let us now construct the common tangent amnb to the curve mpn. Let us consider an
arbitrary composition, X8 . The point p is L1Gm for the solution. Point q is L1Gm for a mixture
of two phases, whose compositions are fixed at m and n, respectively. Since q is lower than
p , the two-phase mixture a + f3 would be stabler than the solution in this region.
The intercepts of the tangent at pure A and pure B provide the values of G:,:: and GB''
respectively (see Section 9.5.3). Therefore,
=
G;: for a-phase at composition corresponding to m G;:: for ,8-phase at composition
corresponding to 11, i.e.
(G: )a at = (G:: )pat "
Ill (10.35)
Similarly,
'
(10.36)
The free energy vs. c.omposition diagram of Fig. 10.3 schematically illustrates the thermodynamic
basis regarding stabilities of the two-terminal solutions a and f3 and formation of a+ f3 two-
'
phase mixtures. There are phase diagrams which exhibit what is known as 'miscibility gap'.
i This is schematically illustrated in Fig. 10.4(a). Here, above the critical temperature Tw
there is only one phase. But below Tw the single phase is transformed into two-phase
mixture, which is the miscibility gap.
Assuming the solutions to be regular, L1Gm is given by Eq. (10.33). Differentiating .1cm
w.r.t. X8 , we obtain (noting XA = 1 - X 8 , and dXA = - dX 8 )
Textbook of Materials and Metallurgical Thermodynamics
Liquid
solution
(a)
~(s)
+ve
i
!
-ve
(b)
jAt T0 j1
I
I
(c)
I (At T0 I
I
I
i (d)
A B
Xa-----+
Fig. 10.5 (a) Phase diagram, (b) free en1rgy-composition diagram at T = T0 , and (c), (d) activity-
composition diagrams of B and A, respectively at T = T0 , for system with complete
miscibility in liquid and solid phase (schematic).
Chemical Potential and Equilibria amongst Phases of Variable Compositions
Positive a means positive departure from Raoult's Law (see section 9.7.2). An arbitrary
solution, which exhibits such positive departure but is not a regular solution, would not be
=
symmetric with respect to Xa 0.5. Figure 10.3(a) also shows suc h a behaviour schematically.
Let us now construct the common tangent amnb to the curve mpn. Let us consider an
arbitrary composition, X 6 • The point p is L1Gm for the solution . Point q is L1Gm for a mixture
of two phases, whose compositions are fixed at m and n, respectively. Since q is lower than
p, the two-phase mixture a + f3 would be stabler than the solution in this region.
The intercepts of the tangent at pure A and pure B provide the values of G~ and G:;1,
respectively (see Section 9.5.3). Therefore,
a~ for a-phase at composition corresponding to m = a;: for /J.-phase at composition
corresponding to n, i.e.
( G~ )a at = ( G~ )13 at
/11 11 (10.35)
Similarly,
(10.36)
This is in conformity with the principle of equality of chemical potentials for phases at
equilibrium (Section 10.1.2). Hence, it is confirmed that the two-phase mixture would have
fixed compositions of a and f3. It also follows that only a or f3 single phase solutions would
be stable at the two-terminal regions since the L1Gm curves are lower than straight lines
j oining any two points on the curve in the regions 'am' and ' nb' .
From Eqs. (10.38) and (10.39), µA and µ 8 should be constant in two-phase mixture.
Similarly, on the basis of Eq. ( 10.28), aA and a 8 also should be constant in the two-phase
field a+ f3. These are shown in Figures ·10.3(b) and 10.3(c). It may be noted that µA =
0
at XA = 1, µ 8 = 0 at X8 = 1, as .per convention already discussed in section 10.1.3.
The free energy vs. c'omposition diagram of Fig. 10.3 schematically illustrates the thermodynamic
basis regarding stabilities of the two-terminal solutions a and f3 and formation of a+ f3 two-
phase mixtures. There are phase diagrams which exhibit what is known as 'miscibility gap'.
This is schematically illustrated in Fig. 10.4(a). Here, above the critical temperature Tw
there is only one phase. But below Tero the single phase is transformed into two-phase
mixture, which is the miscibility gap.
Assuming the solutions to be regular, L1Gm is given by Eq. (10.33). Differentiating L1Gm
w.r.t. X8 , we obtain (noting XA = l - X8 , and dXA = - dX8 )
Textbook of Materials and Metallurgical Thermodynamics
(a)
I
I
I
i I•
I
----------------
Single phase
.
I
Phase I Phase II
Phase I
+
Phase 11
A B
A B
(b)
+ve
i
£D 0
-ve
at T = T 1
0
XA = X8 = 0.5
Xs
Fig. 10.4 (a) Phase diagram, ilnd (b) corresponding free energy-composition diagrams tor a
binary system exhibiting miscibility gap, assuming regular solution behaviour (schematic).
Chemical Potential and Equilibria amongst Phases of Variable Compositions
(10.40)
(10.41)
(10.42)
83(.1(Jm)
From Eq. (10.42), we get s:
uX83
=0 at XA = X8 = 0.5.
.
Smce at T = Tcr• b oth 83(.1Gm) and 82(.1Gm) s h ou Id be zero, th e maxima
. of t he nusc1
. 'b"l '
1 1ty
8X83 8X82
gap would occur at X8 = 0.5, provided the solution is a regular one.
Putting XA = X8 = 0.5 in Eq. (10.41), we get
52(L.\Gm)
8
X2 = 0 at 2 a = 4, i.e. a =2 (10.43)
B
Since RTa = .Q,
.Q
Tcr = (10.44)
2R
for a binary regular solution.
From Eq. (9.89), we see that, since L.\lf11 is approximately independent of temperature
at a fixed composition, .Q may be taken as a constant, and not as a function of either
temperature or composition for a particular binary regular solution. It has already been
emphasized once, and concluded that
n
a = - oc
1
RT T
which is the same as equation (9.92).
Thus, at T > Tw a < 2, and the solution is single phase. On the other hand, at T < Tcr•
a> 2, and the miscibility gap ap'pears. Figure 10.4(b) schematically shows the L.\Gm vs. X8
curves at different temperatures.
f To (a)
T If
1
Solid
solution
B
~(S)
dscn
+ve
f
AGm ds(s)
!
- ve
(b)
jAt T0 ,l 1
I
I
(c)
I [At T0 I
I
i
I
(d)
A B
Xa---+
Fig. 10.5 (a) Phase diagram, (b) free en1rgy-composition diagram at T = T0 , and (c), (d) activity-
composition diagrams of B and A, respectively at T = T0 , for system with complete
miscibility in liquid and solid phase (schematic).
Chemical Potential and Eqwlibria amongst Phases of Variable Compositions
we often encounter situations where one of them is liquid and the other solid at the same
temperature. An example is Fig. 10.5(a), where at temperature T0 , pure A is liquid, but pure
B is solid. In such a situation, for construction of free energy-composition diagrams, conversion
of standard state either from liquid to solid or vice-versa would be required. The thermodynamic
basis for doing the same is derived now.
Suppose free energy of pure solid i is d/
(s) and that of pure liquid i is d/
(I) at
temperature T. Then,
RT In [a ; (s)] = G; - d/ (s) (10.45)
where a; (s) and a; (I) refer to activity of i in the solution with respect to solid and liquid
standard states, respectively.
Combining Eqs. (10.45) and (10.46), we get
RT In [a;(s)]
. a;(l)
=d;(I) - d,(s) =L1G2. (10.47)
where .1G~ is the free energy change of pure i upon melting at temperature T. Now,
(10.48)
(10.50)
i.e.
If it is assumed that Cp(l) = CP(s), then Eq. (10.51) gets simplified into
L1~ :::: ~,(at Tm)• (Tm - T) = Li~ iat T,n) • (T,,, - T) (10.52)
m
i.e. LlG~ changes linearly with change of temperature.
Textbook of Materials and Metallurgical Thermodynamics
EXAMPLE 10.7 Gold and silicon are mutually insoluble in the solid state and form a
binary eutectic system with eutectic temperature of 636 Kand composition of Xsi = 0.186
=
and XAu 0.814. Calculate the free energy of the eutectic melt relative to (i) unmixed (i.e.
pure) solid Au and solid Si, and (ii) unmixed liquid Au and liquid Si. Assume Cp is the same
for solid and liquid for both Au and Si.
Solution (i) Free energy of the eutectic melt, relative to pure Au and pure Si solids
at 636 K, is .dGm, where
l j l
LiGm = XAu GAu (eut) - ~u (s) + Xsi Gsi (eut) - OJi (s) j
(E.10.11)
Since the eutectic melt co-exists at equilibrium with both solid gold and solid silicon at
636 K,
(E.10.12)
Therefore, .dGm = 0.
(ii) For free energy of the eutectic melt relative to pure liquid Au and pure liquid Si
at 636 K,
(E.10.13)
Since Cp of solid and liquid are assumed as the same, from Eq. (10.52),
Figure 10.S(a) shows the phase diagram schematically. An example is iron-nickel system.
Tm(A) and Tm(B) are melting points of pure A and pure B, respectively. At T = T0 , pure
A is liquid, but pure B is solid.
Figure 10.S(b) presents the free energy diagrams at T0 . It is to be noted that, in reality,
we do not have either liquid or solid solution in the entire composition range. Hence, both
curves are partly hypothetical and partly real at T = T0 . Since the stablest states of A and
B are liquid and solid, respectively, ~(I) and da(s) have been arbitrarily taken as zero.
~ (s) and daO) are values at metastable states. Hence, they correspond to hypothetical
standard states of A and B, and the values can be calculated by employing Eqs. (10.47) and
(10.52); ef is the common tangent.
.. '
Figures 10.5(c) and 10.S(d) show activity vs. composition relations for Band A, respectively
1
at T0 . Curves corresponding to liquid solution have liquid A and B have standard states,
whereas curves corresponding to those of solid solutions have solid standard states.
Figures 10.6(a) and 10.7(a) are the schematic phase diagrams. The terminal solid solutions
are a and /J. Figures 10.6(b) and 10.6(c) present free energy vs. composition and activity
vs. composition diagrams at T = Ti. Figures 10.7(b) and 10.7(c) are those at T = T2. Curves
in both Figs. 10.6(b) and 10.7(b) for the liquid solution have liquid standard states (real or
hypothetical, as the case may be). Curves for a and /3 solid solutions, respectively, have solid
A and B as standard states. The activity vs. composition curves have employed the real (i.e.
stablest) standard states for each component.
A B
Liquid
(a)
T
+ve at T 1 ]
t b
d
LIG'11
a c
l
- ve
(b)
a
(c)
0 ........
~...._~~~~ ........
A B
A B
Liquid
t
T
(a)
b
a t T2
+ve
c
t
LlG'11 a d
(b)
i
- ve I
I
I
aJ--a+p--1p
QA
as
(c)
a
at T2
OA B
Xe ._..
Fig. 10.7 (a) Phase diagram , (b) free energy-composition diagram at T = T2 , and (c) activity-
composition diagram at T = T2 , for a binary eutectic system with terminal solid
solubilities (schematic) .
Textbook of Materials and Metallurgical Thermodynamics
a Primary Inter-
solid mediate
0 solution phase 0
(/J)
t (a)
t
(b)
a+P
/J+-r
0 0 .5 1.0
Xs
Fig. 10.8 (a) LiG'"-composition diagram at temperature T1 (schematic), for (b) binary phase
diagram with an intermediate phase.
in a and p phases, respectively. Since at these compositions a and P phases are co-existing
at equilibrium, according to Eq. (10. 19), we have
Let us assume that B obeys Henry' s Law in a-phase since its concentration is small.
Therefore, based on the discussion in section 9.5.l , B will obey Raoult's Law in ,B'.phase.
Hence,
14 [at x: (sat)] = ~ +RT In a:[at x: (sat)]
=~ + RT In [fa · x: (sat)]
= µ& + RT In fa + RT In [X: (sat)] (10.54)
fa= xecsat)
x:csat) (10.57)
Suppose now that there is no solid solubility in ,8-phase, as shown in Fig. 10.9. Then
x: (sat)= 1. Hence,
l
x:csat) =fa (10.58)
L
rTJA)
L+ B
T
Metastable extensions
a+B
A Xa X:(sat) B
Xe ---+
Fig. 10.9 Hypothetical binary phase diagram illustrating derivation of solubility equation for B
in a-phase; dashed lines show metastable extensions.
(10.60)
Textbook of Materials and Metallurgical Thermodynamics
At fixed composition X8 , on the basis of Gibbs-Helmholtz equation i.e. Eqs. (6.28), (6.29)
and their applications to partial molar properties [Eqs. (9.31), (9.33))
where A is a constant.
Equation (10.63) thus provides a quantitative relation between solubility of B in a-phase
with temperature. Figure 10.9 shows the metastable extensions of curves by dashed lines.
It should be noted that the solubility vs. temperature relation would be governed by
Eq. (10.62) even in the metastable state.
In [XFc (sat)] = In A - ~
nm (E.10.17)
RT
Putting in the values of XFe(sat) at 1473 Kand 673 Kin Eq. (E.10.17), we obtain
ii~
-7.738 =In A - (E.10.18)
Rx 1473
n~ (E.10.19)
- 12 ..212 = In A -
Rx 673
Solving the above equations, In A = -3.97, n~ = 46143 J mo1- 1• Putting the values of
the above equations, XFe (sat) at 453 K = 9 x 10-8 •
[Note: A may be considered as equal to exp (Sf'e/R) which gives Sr:"! = -40.65 J mo1- 1 K- 11
Chemical Potential and Equilibria amongst Phases of Variable Compositions
l
..1.ve
r'
I
I
I
1- Solubility of
cementite
Fe
Xe~
c
Fig. 10.10 Schematic diagram of Fe·C system explaining why solubility of carbon in austenite
is higher at equilibrium with cementite as compared to that for graphite.
10.6 Summary
1. There are several definitions of chemical potential (µ;) of a component i in a
solution. However, the commonly employed definition in chemical and metallurgical
thermodynamics is in terms of Gibbs free energy, as follows:
(g(;'
Ji;= -
on.
)
r P,T , 11] , n2 • ...• except n;
This way, /1; becomes the same as the partial molar free energy of component i (G;).
Textbook of Materials and Metallurgical Thermodynamics
µ, - ~ =µ, : : : G[" = RT In a;
PROBLEMS
10.1 Calculate the chemical potential of carbon in a gas mixture containing 30% CO,
20% C02 , 50% N2 at 1300 K. Will this gas saturate a piece of steel with carbon?
The total pressure is 5 atm.
Chemical Potentiar and Equilibria amongst Phases of Variable Compositions
10.2 Pure solids of Si, Si02 and Si 3N4 are equilibrated with a 0 2 + N2 gas mixture at
1000 K. How many degrees of freedom does this equilibrium have? What is the
value of chemical potential of N 2 at equilibrium? While maintaining the gas composition
constant, if the temperature is raised to 1200 K, what happens?
10.3 How many degrees of freedom does the system consisting of FeO-MnO solution +
Fe-Mn solution + OrN2 gas mixture have?
10.4 Calculate the chemical potential of S2 in a gas mixture consisting of 2% H2S, 50%
H 2 , and the rest N 2 at 1200 K. From chemical potential considerations, will this
mixture convert metallic copper into Cu2S?
10.S Calculate the chemical potential of oxygen gas (i.e. Oi) in a system consisting of ( .
pure liquid iron and a liquid slag containing FeO at equilibrium at 1600°C; the
activity of FeO in the slag is given as 0.45.
10.6 Derive an expression for the free energy of fusion of iron (LiGm) as function of
temperature. Assume that Cp of liquid Fe is larger than that of solid Fe by
1.3 J mo1-1 K- 1• Heat of fusion of Fe is 15,360 J mo1-1 at its melting point of 1535°C.
10.7 The solubility of carbon in liquid aluminium is 6 ppm at 960°C and 12.5 ppm at
1000°c.
(i) Calculate the solubility at melting point of Al, i.e. at 660°C. Assume dilute
solution behaviour.
(ii) Predict whether solid aluminium carbide (Al 4C 3) will form at carbon saturation
of liquid Al at 660°C. It is given that, for the reaction:
The above reactions are all independent since none of them can be arrived at by combining
the others. Section 10.2 has discussed phase rule and its application to reaction systems as:
Degree of freedom F =C - P+2
C=N-R
which are the same as Eqs. (10.31) and (10.32). In this example,
N = 7 (Si, Si02, SiC, C, CO, C02, SiO)
R = No. of independent reactions = 4
Hence, if P and Tare fixed, then the maximum number of phases co-existing at equilibrium
will be 3 (i.e. for F = 0).
The relations for equilibrium constants for the above reactions are
K4 -_ PcoPs;o (11.5)
Pc0i
Figure 11.1 presents calculated log Pc0i vs. log Ps;o curves at equilibria with reactions
=
(11.1)-(11.4), at 1700 K and P 1 atm on the basis of Eqs. (11.5) and (11.6) (Ref.: A.
Ghosh and G.R St. Pierre, Trans. Met. Soc. AIME, 245, 1969, p. 2106), and from the
standard free energies of formation of compounds.
Derivation of the ternary phase diagram from Fig. 11.1 is based on the following
principles:
1. Phase rule allows equilibrium co-existence of a maximum of three phases at fixed
P and T.
2. The gas compositions in equilibrium with single condensed phases are given by the
curves in Fig. 11. l.
3. At constant Ps;o. Si0 2 is unstable with respect to a gas composition which contains
less C02 than that corresponding to the SiO:reurve. On the other hand, at a constant
PSiO• C, Si and SiC are unstable if the gas contains more C02 than those corresponding
to their equilibrium curves. These statements follow directly from Eqs. (11.1)-{1 l.4).
Textbook of Materials and Metallurgical Thermodynamics
-2
-3
B
Carbon
-5
-6
- 7 L-.--...L...'-----'-----"'----'-----'---'
-5 -4 -3 -2 -1 0
log PSiO
Fig. 11.1 Log Pco-i vs. log PsiO in the gas phase in equilibrium with various condensed phases at
1700 K, and at Pr= 1 atm. [A. Ghosh and G.R. St. Pierre, Trans. AIME, 245, 2106, (1969.)J
Using the above principles, all regions (A, B, C etc.) and all points of intersection in
Fig. 11.1 were examined to find out the stable co-existing phases in the system. Thus we
arrive at Table 11.1.
Si + SiC + Si02
Si SiO 0
Flg.11 .2 Si-C-0 phase diagram at 1700 K and 1 atm [A. Ghosh and G.R. St. Pierre, Trans.
AIME, 245, 2106 (1969)] .
Atom % oxygen
50 52 54 56 58 60
Liquid iron Liquid o xid
1600 + oxygen
R Magnetite
+ oxygen
y z
1400
c:
0
><
0
1200 +
u
0
y-lron B
";:J
<IS
0
.....
+ e
0
::I wustite :I:
.....
~
.....
4) 1000
c..
E Wustite Magnetite
4)
L + +
E-< magnetite hematite
800
a-Iron
+
wustite
600
Q a-Iron + magnetite
FeO Fe304 Fe203
400
0 0.422 24 26 28 30
Wt.% oxygen
Flg.11.3 Iron-oxygen phase diagram [L.S. Darken and R.W. Gurry, J.Am. Chem. Soc., 68, 798
(1946)].
If the solution is saturated with pure solid CuS04 , then as per accepted convention,
(11.8)
(11.11)
(11.12)
Figure 11.4 presents activity vs. composition data in the Fe-0 system at 1100°C. It may
be noted that tre minimum value of a 0 is l. This is due to the procedure adopted.
0.8 40
0.6 30
~
~
i::s
0
i::s
tf
i::s
0.4 20
0 .2 10
(11.16)
(11.17)
From the Ellingham diagram (Fig. 8.2), at 1200 K, P°oz = 5 x 10-20 atm. Let [aMn1
= 0.3 in a Mn-Fe solid solution and (aMno) =·0.45 in a Mn0-Si02 solid solution. Then, Pai
at equilibrium with these will be 11.25 x 10- 20 atm.
A major part of thermodynamic measurements on metals, alloys and other materials
consists of experimental determination of activities of components in various solid and liquid
solutions at high temperature. With extensive measurements by physical chemists, metallurgists
and materials scientists over a period of 60 to 70 years, activity versus composition data are
available in most solutions of interest.
Textbook of Materials and Metallurgical Thermodynamics
£li.eo x PH2 )
K = ~ x PH20 eq
( (E.11.1)
Since FeO is pure, aFe0 = 1. Hence the above equation may be rewritten as
From Eq. (E.11.2), aFc in the alloy= 0.713. Since XFc = 0.721, it is an ideal solution (within
error limit).
EXAMPLE 11.2 The PH/PH2o ratio in an HrH20 gas mixture in equilibrium with pure
liquid Pb and lead silicate melt, at XPbO = 0.7, is 5.66 x 10-4 at 900°C. The corresponding
value at 1100°C is 1.2 x 10-3. Calculate the partial molar heat of mixing of liquid PbO in
this lead silicate melt.
Solution 900°C = 1173 K, 1100°c = 1373 K
The reaction is:
Pb(l) + H20(g) = (Pb0); 0 melt + H2(g) (E.11.3)
for which
Hence,
uJJ, = 82,754 J , K3 = 2.06 x 10- 4 at 1173 K,
LYJ1 = 87,614 J, K3 = 4.64 x 10- 4 at 1373 K,
Noting that a Pb = 1,
(E.11.5)
Putting in the values given in Eq. (E.11.5), we get aPbO = 0.364 and 0.387, respectively at
1173 K and 1373 K.
Now,
8(~) (E.11.7)
8(~)
P, comp
(E.11.8)
EXAMPLE 11.3 Calculate the partial pressure of nitrogen gas over an eutectic liquid
solution of Al-Si alloy at the eutectic temperature, at equilibrium with solid Si 3N 4 (Johnson
and Stracher, 1995, p. 134).
(E.11.9)
Textbook of Materials and Metallurgical Thermodynamics
for which
(E.11.10)
..1Gf1 = -753,190 + 336.43T J mo1-1 for 3Si(s) + 2N2(g) = Si3N 4(s) (E.11.11)
..1~ = 50,630 - 29.91 T J mo1- 1 for Si(s) = Si(l) (E.11.12)
From Hess' Law,
From literature sources, at 850 K and Xsi = 0.122 in liquid Al-Si solution, with reference
to liquid Si as standard state,
allows removal of Cu present as impurity in molten lead. The solid sulphides are present as
pure compounds, and Pb is insoluble in solid Cu. The solubility of Cu in liquid Pb is given as
8060.5
In [Xculsai =- - T
- + 5.207 (E .11.14)
('1cu2s> x [aPb]}
~3 =- RT In K13 =- RT In { 12 (E.11.15)
["cuJ X (~ ) eq
Since Cu 2S and PbS are pure, Ocu2s = aPbS = 1. Concentration of Cu in liquid Pb would
be very small (can be verified later). Hence Pb is almost pure, and the activity of Pb may
be assumed as 1. Therefore,
In metallurgy and materials science. we have to ofren deal with multicomponent solutions-
an important issue. In concentrated multicomponent solutions, the mathematical procedure
for dealing with such interactions is lengthy. and 1s beyond the scope of an introductory
thermodynamics course. For example, during extraction and refining of liquid metals, we
mostly deal with dilute multicomponent solutions. An important example 1s liquid steel,
where Fe is the solvent, and several solutes (C, Si, Mn, S, P etc.) are present in very low
concentrations. Fnr such dilute multicomponent solutions, Carl Wagner ( 1952) proposed a simple
but elegant ana {tical method of handling solute-solute interactions, which is discussed now.
Let A designate the solvent, and l, 2, ... i, j, ... are the symbols for the solutes. Wagner
expressed In y, (i = general symbol of solute) as function of XI> X2, .•. by the Taylor series
expansion around pure solvent A. The equation is as follows:
2 2
+ -l X12 (8 [lny,]) + ... + -1 X1 (8 [1ny,Jl + ...] (1 J.18)
2
2 :!
[2 8X1 xA-+ I 2 8X) XA4 I
In Y; =In r? + X1 (8[ln Y, ])
8X1 x1-+ 0
+ ... + X, (8[1n Y, ])
8X, x,-+ 0
+ X1 [ 8lln Y,
8X1
]l + ...
X 1 -+0
(l l.20)
The influence of one solute on y of another solute was termed as interaction coefficient
by Wagner. This allows us to rewrite of Eq. ( 11.20) as
Reaction Equilibria Involving Condensed Phases with Variable Compositions
(1 1.21)
where
(11.23)
- (/j(;'
G. = -
I t5i
) n, "I .... excep1 n;
' - (/j(;']
G1= -
on . (11.24)
J "! .... CXcepl nJ
Again,
cm =(o(G'Ol- G~)J
I '
G'!1
J
=(o(G' - G~)J
on . (11.25)
n; "I .. ., except n; 1 "I .... excep1 nj
where G~ is G' when components 1, 2, ... , i, j, ... are present as pure in mixture of the
same overall composition.
From Eq. (11.25),
(oGf")
& .
J Xr -+ 0
=RT [o(ln a;)]
& .
J Xr-+O
= RT {<>(In X;) + o(ln r;)}
~ ~.
J Xr-+O
(1 1.27)
Textbook of Materials and Metallurgical Thermodynamics
oGmJ
_ J_ RT .
=-E'·
( on. 1
X;-+0
n,. J
(11.28)
(11.29)
As already stated, activity versus composition data are to be basically obtained from experimental
measurements. For a binary solution, it is the simplest. The activity of one component is
measured, and from G-D integration, the activity of the other can be calculated. This
approach works for ternary also. For multicomponent solutions, however, much larger
experimental programme is required since there are several composition variables. This is
where the concept of Interaction coefficient for dilute solutions comes in handy.
Let us consider the multicomponent A, 1, 2, ... i, j, .. . again. For knowing y,, what we
have to do now is to make measurements in binary A-i, and all ternaries, i.e. A-i-1 ,
A-i-2, .. . A-i-j , .... Experimental data in A-i allows determination of Y/. If experimental
data in A-i and A-i-j, are combined, we can find out cf since in the ternary A-i-j ,
(11.30)
(11.31)
11.S.1 Introduction
The conventional standard state is also known as Raoultian standard state, where the
standard state (SS) is pure solid or pure liquid, whichever is the stablest at the temperature
under consideration. Moreover, Raoult's Law serves as the reference.
Thermodynamics permits use of any other alternative standard state.· Such a state may
be real or hypothetical. However, for such a change of SS, the thermodynamic data employed
for calculation should conform to the new SS. For example, ..16'° for reactions in general
data sources are all for the conventional SS. They are to be changed if a new SS is employed.
In extraction and refining of liquid metals as well as for a variety of other uses, we deal
with dilute solutions. Also, industries prefer to find out answers directly in weight percent.
In order to take care of these, Chipman proposed the " I wt.% Standard State". Its features
are:
• Composition scale is in wt.%, not in mole fraction.
• Henry' s Law, rather than Raoult's Law, is the basis.
• The answer for equilibrium calculation would be directly in wt.%.
Consider the multicomponent solution with 1, 2, ... , i, j , ... as components. Then
conversion of wt.% into mole fraction can be done with the help of the following equation:
_ WJM;
X1 - (11.32)
l:(W;IM;)
where
W; = wt.% of i,
M; = molecular mass of i.
The reverse conversion is as follows:
(11.33)
8 (11.34)
h = (::. ""L
s -+ 0
Ira = Wa at W8 ~0 (11.35)
Raoult's
t as vs X8
curve /
/
/
t
as. ha
/
as /
/
J,, . .
1fa
/
as = as ------7--- ,., ...
/ 'P .... h8 = 1 ---<
/
,,,,,
I .,..
""1S
as = a8 - - ; 7- -.:. ... ::..-17 .... .... .... I
.... I
0
§/
A Ws = 1
~
t
:
B
l A
I
I
Wa = 1
We= Wa X8 = 1 Wa-
X8 = 0
W8 = 0 We= 100
(a) (b)
Real 1 wt.% SS at point S Hypothetical l wt.% SS at point S
Fig. 11.5 Illustration for 1 wt. % SS: (a) real 1 wt.% SS, and (b) hypothetical 1 wt.% SS.
(11.36)
( 11.37)
Since at p, a 13 = raX8 , the combination of Eqs. (11.36) and (11.37) yields
(l l.38)
(l l.39)
For binary A-B, and W8 ~ 0. WA ~ 100, Eq. (l 1.32) gets simplified into
(11.40)
Ll~ = - RT In Kh = - RT In [ (aee>i)
[he] P0z
leq
(11.43)
where Ll~, K1r are LlG° and K for l wt.% SS for component B. The subscript 'h' denotes
Hennan since HL is reference here. The corresponding equation in the Raoultian SS are:
(~Jj
-<1)
Textbook of Materials and Metallurgical Therm
Equation ( 11.46) r -
thermochemical ·
only addit=
are -
f;' is
Eq. (1 .
The commm
The relationship of
Combining Eq. l
i.e.
(l l.37)
(l l.38)
(11.39)
(11.40)
-118 = -100M
-=-8
(l 1.41)
aa ygMA
Consider the following reaction at temperature T for a dilute binary metallic solution A-B,
where B is the solute.
(l l.42)
Lle'2 =- RT In K 11 =- RT In [ (aaai)
[ha) Pai
l
eq
(1 1.43)
where Lle'2, K 11 are LlG° and K for 1 wt.% SS for component B. The subscript 'h ' denotes
Henrian since HL 1s reference here. The corresponding equation in the Raoultian SS are:
RT In (Kh)
K
=- 0
RT In ( a) =RT In
ha
(!ta )
l an (ll.45)
Textbook of Materials and Metallurgical Thermodynamics
f is Henry's Law constant for I wt. % SS. Since f = 1 by definition, log t' = 0. Hence,
Eq. (11 .47) gets simplified into
(11.48)
(11.49)
The relationship of e/ with ej may be derived on the basis of Section 11.4, as follows:
Combining Eq. (11.48) with Eq. (11.38), the former may be rewritten as
'Yi] .
In ft =In
(~ =2.303 7W;ef (11.50)
i.e.
In 'Y; = In i/ + 2.303 L. Wie/
)
(ll.51)
(l l.52)
Reaction Equilibria Involving Condensed Phases with Variable Compositions
(11.53)
. MA .
e'.1 = t 1. (11.54)
230.3 M;
230.3 e( Mj M1 j
e'. = MA el = MA ----=-e ; (11.55)
} 230.3 M; I 230.3 M, MA M;
EXAMPLE 11.5 A Hr H2S gas mixture with H 2/lI2S ratio of 10 was equilibrated with
pure liquid iron at 1850 K. On analysis of quenched sample, the iron was found to contain
0.8 wt.% sulphur. Calculate the free energy change for the following reaction at 1850 K:
z
l
S2 (g, 1 atm) = [S]1 wt.% ss in liq. Fe (E.11.18)
(E. l l.19)
Solution
H2(g) + [S]1 wt%SS = H2S(g); Liejl0 (E. 11.20)
(E.ll.2 1)
Textbook of Materials and Metallurgical Thermodynamics
l.O
0.8
0.6
0.4
0.2 Co, Ni
~ Sn w
e.o
0
0
Mo
-0.2
- 0.4
-0.6
-0.8
-1.0
0 2 4 6 8 10 12 14
Alloying element, wt.%
Fig.11.6 Activity coefficients of nitrogen in Fe-N-j ternary liquid solutions at 1600°C, with iron
as the solvent (A.G. Ward, Physical Chemistry of Iron and Steelmaking, Edward Arnold,
London, 1962).
Now.
A~ 20
= - RT In {
PH2
2
PH s
x [/is]
}
eq
=- 1
RT ln ( -- )
10 [/ts] eq
(E.11 .22)
Again,
hs = fsWs = (10--0.ois x
08
· ) • 0.8 = 0.76
Putting the above value of hs in Eq. (E.11.22), we obtain
A~ = 31,190 J mo1- 1, ,1~ 1 = 690 - 31,190 = - 30,500 J mo1- 1
EXAMPLE 11.6 From experimental measurements, it has been determined that, for Si (pure, I)
= (Si]1 wt% SS in liq. Fe•
(i) (E.11.24)
since 1 wt.% SS is, by definition, on Henry's Law line. Again, ~om Eq. (11.40), at Wsi = 1,
Xs·
'
= Ms;MFe =-56- =0.02.
x 100 2800
Solving Eq. (E.11.24), we get J1; = 1.25 x io-3. At Xsi· = 0 .01, /§i = 0 .0014; from
Eq. (11.50),
[Note: Here, Henry's Law is not obeyed at Xsi = 0.01 since/si::;; 1. Hence, obviously HL
would not be obeyed at 1 wt.% (i.e. Xsi = 0.02). This means that the 1 wt.% SS is not real
bur hypothetical.]
(ii) H-m
Si
-
- (E.11.25)
fill! =
1 119.24 x 10 + 24.0T
T
3
1
= - 119 24 x 1<>3 J mo1- 1
& ~~) .
EXAMPLE 11.7 Liquid iron containing 2 wt. % silicon, 2 wt.% chromium, and 0.5 wt.%
vanadium was kept in a graphite crucible at 1600?C for a long time till equilibrium was
established between the liquid and the crucible. Calculate the wt.% of carbon in iron after
the experiment is over.
Given:
(i) Solubility of graphite in pure liquid iron = 5.4 wt.% at 1600°C
(ii) For Fe-C binary, log /~ = 4350 [l + 4 x 10-4(T - 1770)) (l - XFJ 2 (E.11.26)
T
(iii) e~i =0.1, ~r = - 0.024, ~ = -0.033
Textbook of Materials and Metallurgical Thermodynamics
X _ Wc/12 5.4/12
c - We WFe = 5.4 100 - 5.4 = 0. 2 l
- +- - + - -- -
12 56 12 56
Noting that
4350
1 - XFe =Xe. log/cc= [1 + 4 x 10-4 (1873 - 1770)] (0.21)2 = 0.1067
1873
[Note: If in Fe-C binary and in the multicomponent solution are different since the value
of XFe will be different. Since the interaction effects of Si, Cr, V have been separated, from
a thermodynamic point of view, approximately speaking, (1 - XFe) can be substituted by Xe
in Eq. (E.11.26)]. Hence, from Eq.(E.11.26),
We/12
Xe = --------'=------- (E.11.31)
Wsi Wcr
- + - +We
- +Wv WFe
- +--
28 52 12 51 56
Since Wsi = 2, Wer = 2 and Wv = 0.5, we have
Combining the above equations, and further graphical solution yields We = 4.2 wt.% after
crucible-melt equilibrium is attained.
Reaction Equilibria Involving Condensed Phases with Variable Compositions
(11.56)
for which
K
h
= ([/;.]l112
PN2
(11.57)
(11.58)
11.7 Summary
1. This chapter contains discussions of some additional topics to supplement the earlier
chapters so that the readers have complete information for solution of aJJ types of
equilibrium calculations.
2. As an extension of Chapter 8, a further analysis of phase stabilities of reactive
stoichiometric compounds has been presented with the example of construction of
Si-C-0 ternary diagram. Wustite (FexO) is the best metallurgical example of a
nonstoichiometric compound. Thermodynamic analysis of this has been discussed in
the chapter.
3. Reaction equilibria calculations involving solutions of condensed phases have been
illustrated by solved examples.
4. In a multicomponent solution, components interact amongst themselves. For a dilute
solution, in connection with interactions amongst solutes, Wagner derived analytical
equation to quantify such interactions, and the equation is
Jn y,.I = Jn r? + I: X1e!
I . I
J
Textbook of Materials and Metallurgical Thennodynamics
el = (<5[1nY; ])
I DX
1 X1 --+ 0
Moreover, ~ =cf.
5. Weight% is commonly used as the composition parameter for practical applications.
In order to obtain answers of calculations directly in wt.%, the 'one weight percent
standard state' was proposed. It is based on Henry's Law (HL) as reference. On HL
line, in the binary A-B, where B is the solute,
Activity of B in l wt.% SS (h 8 ) = W8 , where W 8 is wt.% B.
At a composition where HL is not obeyed,
where
PROBLEMS
11.1 A methane-hydrogen gas mixture at l atm. totaJ pressure is allowed to come to
equilibrium with a steel containing 0.60 wt.% carbon at 925°C: The gas was found
to contain 0.64% CH4 and the rest H 2 by volume. Calculate
(i) the activity of carbon in steel;
(ii) the expected analysis of a CO-C02 gas mixture which would be at equilibrium
with carbon in steel.
Reaction Equilibria Involving Condensed Phases with Variable Compositions
11.2 An alloy with 20 atom% Ni and 80 atom%. Au is allowed to react with a gas
mixture of H 20-H'.! at 1000 K to form pure solid NiO. At equilibrium the gas
contains 0.175% H 2 , 50% N'.!, and the rest H 20. Find the activity of Ni in the alloy.
11.3 A Cr-M solid solution contains 15 atom% Cr, and behaves ideally. Upon reaction
with a H '.!0-H'.! gas mixture, only pure solid Cr'.!03 forms. Calculate the Ptt/Ptt 2o
ratio in the gas mixture at equilibrium with the solid solution and Cr20 3 at 1000°C.
11.4 Consider the reaction between a liquid sulphide solution of Cu 2S-FeS (known as
Matte) and a molten slag of FeO-SiOrCu 2 0 at l250°C. Calculate the activity of
Cu 20 in the slag upon attainment of equilibrium. The matte contains 40 wt.% Cu 2S
and the rest FeS, and behaves as an ideal solution. The activity of FeO in the slag
may be assumed as 0.5 relative to pure liquid FeO.
11.5 Zinc, present as impurity in molten lead, can be selectively removed as ZnCl2 by
treatment with chlorine gas. Find the residual concentration of Zn in liquid Pb upon
attainment of the following reaction equilibrium at 663 K:
[Zn] + (PbCI 2) = (ZnCl 2) + Pb(!)
Given: (i) ZnCli-PbCl 2 forms an ideal liquid solution.
(ii) ..1G° for formation ~f PbCl 2 and ZnCl 2 are -257.5 kJ mo1- 1
and -320.7 kJ mo1- 1, respectively at 663 K.
(iii) XZnc12 in salt phase = 0 .983 .
0 2798
(iv) In [YZn Jin Pb= - -- 0.962
T
11.6 It is also possible to lower Zn content in liquid Pb by vacuum distillation. Calculate
the maximum pressure that can be tolerated in the vacuum system at 900 K for the
following conditions:
( i) Zn as only solute in liquid Pb, and Zn concentration is to be lowered to 0.01 atom%.
(ii) Liquid Pb contains 2 wt.% cadmium and 2 wt.% copper, besides Zn. The Zn
content is to be lowered to 0.3 wt.%.
Given: (i) ~~ = -0.04, ei;: = -0.07, ~~ =O in liquid Pb
(ii) Y~n in liquid Pb as in Problem 11.5.
11.7 An Fe-Mn solid solution containing XMn = 0.001 is in equilibrium with an FeO-MnO
solid solution and an oxygen containing gaseous atmosphere at 1000 K. Calculate:
(i) Composition of the oxide solid solution, and (ii) P0z in the gaseous atmosphere.
Both the metallic and oxide solution are ideal.
11.8 A mixture of solid CaO, MgO, 3Ca0.AI20 3 and liquid aluminium exerts an equilibrium
vapour pressure of magnesium of 0 .035 atm. at 1300 K. Write the equation for the
appropriate reaction equilibrium. Calculate the standard free energy of formation of
3Ca0.Al:!03 from CaO and Al 2 0 3, and the activity of Af 20 3 in CaO-saturated
3Ca0.Al 2 0 3 at 1300 K.
Textbook of Materials and Metallurgical Thermodynamics
11.9 A liquid iron solution contains 0.5 wt.% silicon and 2 wt.% carbon. Temperature is
1900 K. Calculate (i) hsi> (ii) µ Si· Given: Y~i in binary Fe-Si solution = l.37 x 10- 3;
4 =0.18. ~i =0. ll
11.10 (i) In the binary Fe-Cu liquid solution at 1550°C with Fe as the solvent. it is gi ven
that:
As discussed in Section 4.4, although Clausius proposed entropy in 1850, it took several
more decades to fully establish the fundamental significance of entropy qualitatively and
quantitatively from the atomistic viewpoint. Early work leading to the Third Law of
Thermodynamics was done by T.W. Richards (1902). W. Nemst (1906) generalized these
findings and formulated the statement of the Third Law, which he called as Heat Theorem.
Subsequently, the statement was refined by Max Planck. The Third Law provided some
atomistic interpretation of entropy qualitatively. It specifically dealt with the behaviour of
entropies and specific heats of substances at absolute zero of temperature (i.e. T = 0).
The subject of Statistical Thermodynamics started developing from the late
19th Century and matured in the decade of 1930s. It is a major subject in its own right, and
provides quantitative relations of internal energy, entropy etc. from atomistic considerations.
However, in common applications in the metals and materials science and engineering, it is
not widely employed. Hence, in this chapter, only introductory concepts and a few quantitative
relations would be briefly presented, primary attention being devoted to interpretation of
internal energy and entropy.
(12.1)
Now,
which is the same as Eq. (6.27). Hence, from Eq. ( 12. 1) and (6.27),
$ ~ 0 as T ~ 0 (12.2)
Again,
(6.3)
• for isothermal processes. Differentiating Eq. (6.3) with respect to T at constant P, we get
(12.3)
[o<;:o)t ~ o (12.4)
(12.5)
where
~=L~-L~ (3.26)
product reactant
Hence, from Eqs. (12.4) and (12.5), for reactions involving pure solids and liquids,
~ ~ 0, asT ~ 0 (12.6)
In accordance with Eq. (12.2), Nernst's Heat Theorem states that "for a\\ reactions
involving pure substances in condensed state, .dS is zero at the absolute zero of temperature".
A pure substance is either a pure element or a pure compound. Consider the formation of
a compound AB from elements A and B, i.e.
A(s) + B (s) = AB(s) (12.7)
Third Law of Thermodynamics, Statistical Thermodynamics, and Entropy
for which
$:~B-~-~ (12.8)
(i) ~B =~ +~
or
(ii) ~B• ~. ~ are all individually zero.
From a probability point of view, the first alternative is very unlikely as a general feature.
It may be true, by chance, in a few cases. Hence, alternative (ii) is accepted as of general
=
validity, and it may be concluded that entropies of pure solids at T 0 are zero. Since all
substances are solids at T =
0, we are omitting liquids. Therefore, the absolute value of
entropy of a pure substance at any temperature can be determined by taking T 0, SJ 0= =
as the lower limit. This is in contrast to energy, whose absolute value cannot be found out
within the scope of conventional thermodynamics. This is the basis on which we find as to
how thermochemical data tables provide values of entropy for pure substances. Figure 12.1
presents variation of SJ with temperature for a few substances.
Figure 12.1 shows some increase in entropy during melting. In this connection, it is
worthwhile to note down the following rules which were proposed long back from experimental
observations.
200
~
0 150
e
;::::;
50
Graphite
Diamond
0-=..~--'-~~-'-~~L-.~-L~~-'-~--1.__J
300 500 700 900 1100 1300 1500
Temperature, K
Fig. 12.1 Variation of molar entropies of some elements and compounds with temperature.
Textbook of Materials and Metallurgical Thermodynamics
(i) The Third Law method. Thi s consists of integration of experimental heat capacities
as function of temperature from 0 K to T, assuming SJ to be zero at T = 0. From
Eq. (4.38),
Third Law of Thermodynamics, Statistical Thermodynamics, and Entropy
(12.10)
(ii) The spectroscopic method. In this method, we evaluate entropy of an ideal gas
from spectroscopic data. The procedure for calculation is based on statistical mechanics.
(iii) The Second law method. This consists of evaluation of LIG° of reactions over a
range of temperature. &-' is evaluated from variation of LIG° with temperature on
the basis of Eq. (6.27).
Comparison of predictions of these methods provided experimental verification of the
Third Law. Comparison of results of all three methods is not common. One example is
entropy of chlorine gas at room temperature (i.e. 298 K).
(12.l l)
So,
(12.12)
The experimental values are:
~ = 1.093 J mo1- 1K- 1
II=--
Monoclinic Ttr Rhombohedral
,.
sulphur sulphur
Ttr = 368.5 K
f
e::s
......
..."'
G)
E
0.
G)
f-e
&f
G j\ ~l
OK -
L'.1.S?v
Fig. 12.2 An example of experimental verification of the Third Law of Thermodynamics.
Hence,
In traditional mechanics, one form or the other of Newton's laws is employed to analyze the
dynamics of system, and to follow motions of individual particles or elements (e.g. fluid
elements) in a system as function of time. However, there are systems which consist of a
large number of particles in motion and are too complex to analyze in terms of individual
particles. These systems usually appear disordered because of the large number of particles
involved, and the innumerable ways of sharing energy of the particles amongst themselves.
Thermodynamics deals with such systems more easily.
The basic laws of thermodynamics deal with relationships amongst macroscopic variables,
such as pressure, temperature, volume and internal energy. Such relations are simplest for
ideal gases. These laws and relations do not, however, consider the behaviour of atoms and
molecules which constitute a system.
The alternative approach to explain these macroscopic thermodynamic variables is the
Third Law of Thermodynamics, Statistical Thermodynamics, and Entropy
so-called Kinetic Theory. It developed gradually from the 17th Century and matured jn .the
19th Century. It was successfully applied to ideal gases on the postulate that the atoms/
molecules of a gas are always in motion due to their thermal energy. The thermodynamic
variables are related to this and are macroscopic averages. Newtonian mechanics is the basis
of dealing with the velocities and kinetic energies of individual molecules.
The equations derived by the kinetic theory of ideal gases are the following:
1-
Pressure =P = - 2
pu (12.13)
3
where p is density of gas and u 2 is mean square (i.e. rms) of linear velocities of the
molecules. Again,
(12.14)
where M is molecular mass. That is, the average translational kinetic energy per mole of an
ideal gas is proportional to its temperature. Dividing Eq. (12.14) by Avogadro' s number
(N0), we get
I - 3
- mu2
2
=-k
2 B
T (12. 15)
3
U= -RT (12.16)
2
Classical statistical mechanics (section 12.2.2) arrived at the Principle of Equipartition
of Energy amongst all independent degrees of freedom. In a monatomic gas, there are three
degrees of freedom of translational motion along the three perpendicular coordinates. Hence,
each is associated with energy of 112 RT per mole. In a diatomic gas, there are five degrees
of freedom (viz. three translational and two rotational). Hence,
I 5
U =5 x - RT= - RT
2 2
(12.17)
Classical formulations of statistical mechanics were done in the late 19th century by Maxwell,
Gibbs and Boltzmann. The kinetic theory · predicted average translational kinetic energy of
a gas. However, it could not predict distribution of speed and energy amongst molecules. We
take recourse to statistical mechanics for this purpose. Besides mechanics, it is founded on
the mathematical science of probability and statistics, an emrging field at that period.
Textbook of Materials and Metallurgical Thermodynamics
Maxwell first solved the problem of the distribution of speeds in a gas containing a large
number of molecules in a box. The Maxwellian speed distribution for a sample of gas at
temperature T containing N molecules, each of mass m is
2
n(u)=4trN(-m-)3f2 u2 exp (- mu ) (12.18)
2trk8 T 2k8 T
where ri(u) is the probability distribution function of speed, n(u). du is the number of
molecules in the gas sample having speed between u and u + du. Therefore,
0
2.0
E
......
,..,"'
-
0
x
,,......
1.0
::t
-.._;
s:::
Following the derivation of speed distribution, energy distribution function was subsequently
derived as
n e - -2N
( )- ~
I
(k5T)312
e12 exp ( -ksT
e )
- (12.20)
where n(e) ·de is number of molecules with energy between e and e + de.
Third Law of Thermodynamics, Statistical Thermodynamics, and Entropy
Internal energy and energy distribution in an ideal gas were successfully predicted by
statistical mechanics. Here the molecules were free and moved like distinct particles. However,
in solids and liquids, the atoms/molecules are close and interacting. This gives rise to
complex interaction and energy exchanges between them. In this, the electrons also participate.
It required development of the subject of quantum mechanics to tackle such situations.
In the late 19th Century, there were two empirical laws of radiation, viz. Wein's Law
and Rayleigh-Jean 's Law. They could not be reconciled on the basis of the wave nature of
radiation. In 1901, Planck reconciled them by postulating that radiation is emitted discretely
as energy packets. It was a revolutionary concept. In 1905, Einstein caJled these packets as
quanta. Thus Planck's theory came to be known as Quantum Theory.
The postulate of the quantum theory can be stated as follows: "If a particle is confined
to move within a fixed volume, then its energy is quantized, i.e. discretized". Application
of quantum theory later explained the structure of an atom by assuming that electrons can
only be at some discrete energy levels.
On the basis of quantum theory and the techniques of statistical mechanics, the subjects
of quantum mechanics and statistical thermodynamics were developed by several 20th Century
physicists. An important postulate is that all particles are exactly alike (i.e. indistinguishable
from one another).
into zero net rate. There are always fluctuations in microscale. But these fluctuations are not
perceptible for a large assembly of molecules. Macroscopic quantities (pressure, temperature,
energy etc.) are only averages. Such averaging is to be done only statistically. For such
statistical analysis, the basic postulates of statistical thermodynamks are as follows:
1. Each of the accessible and distinguishable quantum states (i.e. each microstate) of a
system of fi,xed e11ergy is equally probable.
2. The equi ibrium state corresponds to the most probable macrostate. It is a state
which contains the largest number of microstates.
3. As has already been stated, all particles are exactly alike.
Equation (12.20) provides the distribution function for energy of molecules in ideal gas.
It is known as Maxwell-Boltvnann (MB) statistics.
On the basis of the above postulates (i.e. utilizing quantum theory), the following two
energy distribution functions at the most probable macrostate were derived later by physicists
whose names they bear
1. Bose-Einstein quantum statistics
2. Fermi-Dirac quantum statistics.
Each of these has played a very important role, not only in statistical thermodynamics
but also in quantum mechanics quantum physics.
All the above three statistics are applicable to systems, where mass, composition, internal
energy and volume are fixed. They differ in some other assumptions, the discussion of which
is beyond the scope of this text.
(12.22)
From mathematics, the number of .ways N particles can be arranged amongst the various
energy levels are given as:
(12.23)
Since a system consists of a large number of particles (i.e. atoms etc.), say of the order
of Avogadro's number, Stirling 's approximation is valid and, accordingly,
In W = (N In N - N) - .t (11; In n ; - 11;)
;
(12.24)
or
d(ln W) = -.t(ln
I
n; dn1 + !i. dn;
n
- dn;) = - ~In n; dn;
1
(12.25)
1
.tn.d£.. =0 (12.29)
.
I ' '
Combining Eqs. (12.28) and (12.29), we get
dN = I:.dn.
j I
=0 (12.31)
Nexp( - ~)
n =- - --'--'- ( 12.33)
' p
'**'
' Textbook of Materials and Metallurgical Thermodynamics
1
~ where
'
(12.34)
'-'
The above ec ·1ations give a distribution function of n; as proportional to exp (-E/ k8 7),
and is analogous to Maxwell-Boltzmann distribution function [Eq. (12.20)].
It can be derived further (derivation skipped) that, at the most probable macrostate,
In W=N lnP+-
u (12.36)
k8T
or
;
(12.37)
dU =liq (12.38)
where q is heat flow from surrounding into the system. Hence,
. 8q
".
d(ln W) =- (12.39)
k8 T
oq =dS (12.40)
T
Combining Eqs. (12.39) and (12.40) and integrating, we get
S = k8 In W + constant (12.41)
S = k8 In W (12.42)
This equation was proposed by Planck in 1906, and is known as Boltvnann equation.
It relates entropy with the number of arrangements of particles amongst different energy
levels available in a system of fixed mass, volume, number of particles, as al so of fixed ·
'
internal energy and composition.
Third Law of Thermodynamics, Statistical Thermodynamics, and Entropy
Consider mixing of atoms A and B to form a binary solution. The process may be represented
as
A + B (unmixed) ~ A + B (mixed) (12.43)
(state 1) (state 2)
Integral molar entropy of mixing is
(12.45)
In unmixed state, configurations of both A and B are unique since they exist separately.
Hence, Wconf,t = 1 and, therefore,
(12.46)
Let us take 1 mole of solution. The total number of atoms is Avogadro's number (N0 ).
Assume the total number of sites = N0 , and also assume the number of A and B atoms as
"A and n8, respectively. From the theory of combination, nA atoms can be arranged in No
sites in NOc ways assuming random mixing of A and B atoms. Once A atoms are arranged,
DA
B atoms occupy the remaining sites (i.e. one arrangement only). Therefore,
(12.47)
(12.48)
(12.49)
Textbook of Materials and Metallurgical Thermodynamics
(12.50)
Since nAIN0 and 11 8 1N0 are mole fractions of A and B (i.e. XA and X8 ), and k8 N0 = R,
Eq. (12.50) may be rewritten as
(12.51)
Equation (12.51) is identical with Eq. (9.52) for ideal binary solution. Hence, it may
be concluded that an ideal solution is characterized by
(i) random mixing of atoms/molecules, and
(ii) no change in thermal entropy when the solution forms from pure components.
12.4 Summary
1. Max Planck stated the Third Law of Thermodynamics as
The entropy of any homogeneous substance, which is in complete internal equilibrium,
may be taken as zero at 0 K.
This is valid only for a perfectly ordered, pure, crystalline solid.
2. The Third Law was originally proposed on the basis of experimental measurements
at temperatures close to 0 K. It was subsequently verified for several substances
systematically from experimental data.
3. For ideal gases, the following equations were derived on the basis of the kinetic
theory:
Chapter 13
Thermodynamics of
Electrochemical Cells
13.1 Introduction
An electrochemical reaction involves coupling of chemical reaction with flow of electric
current. It is a very important field in metaJlurgy and materials science, as the following
examples will illustrate:
• Many metals are extracted and/or refined by electrolytic process (Zn, Al, Mg etc.).
• Electroplating and anodizing are widely employed for surface protection of metals
and alloys from corrosion.
• Electrochemical reactions occur in corrosion, hydrometallurgy, and slag-metal reactions.
• Batteries are electrochemical cells.
• The electrochemical method is an important tool for thermodynamic measurements,
especially at high temperatures.
Electrochemical cells are broadly classified into:
(i) Galvanic cells (i.e. batteries), where stored chemical energy is converted into electrical
energy
(ii) Electrolytic cells, i.e. cells for electrolysis, where electrical energy is used to do
chemical work.
Thennodynamic studies/predictions/measurements can be properly done only for galvanic
cells since these can be made to operate reversibly. It may be recalled that the term 'galvanic'
originated from the famous Italian scientist Galvani, who was a pioneer in this field.
228
Thermodynamics of Electrochemical Cells
Cu electrode
Porous
diaphragm
Insulated
container
0 0 0 0
000000
Figure 13. 1 shows that both CuS04 and ZnS04 solutions are saturated in a Daniel cell.
The overall reaction may also be written in ionic form as
Zn(s) + Cu 2+(aq) = Cu(s) + Zn 2+(aq) (13 .2)
You are fami liar with aqueous solutions and their chemical and physico-chemical behaviour.
Electrochemical cells with aqueous electrolytes are operated around room temperature. In
contrast, all other electrolytes listed in Table 13. l are for cells operating primarily at higher
temperatures. It is difficult to specify the exact temperature ranges for them. Some approximate
idea only may be given. Molten salts typically consist of solutions of chlorides or fluorides
and are suited to a temperature range 400-900°C. Molten slags consist of solutions of metal
silicates or borates, and are ideal for electrolytes in the temperature range of 1000-1500°C.
Solid electrolytes (SE) are developments in the last 50 years. Electrochemical cells with
these are also operated primarily at high temperatures, the ranges varying with the nature of
the electrolyte. A large number of such electrolytes have been developed. The best known
and the most widely investigated one is stabilized zirconia. Zr02 is an excellent high temperature
ceramic material. But it undergoes some phase transformations during heating and cooling.
This causes thermal stresses and consequent failure in service. Addition of oxides such as
CaO, MgO, and Y20 3 stabilizes the cubic fluorite (i.e. CaF2) structure from room temperature
upto its melting temperature (approx. 2400°C) and thus prevents failure.
Let us take the example of Zr02-Ca0 solid solution. In pure Zr02 , Zr4+ ions occupy
cationic lattice sites and 02- ions anionic sites. On addition of CaO, divalent Ca 2+ occupies
some cationic sites in place of tetravalent zr4+. For maintainance of local electrical neutrality,
replacement of a zr4+ ion by one Ca2+ ion leads to removal of one 0 2- from anionic site.
This creates an oxygen ion vacancy and allows movement of 0 2- ion by vacancy mechanism.
Thus, 0 2- is the only current carrying species since Zr4+ and Ca2+ are almost immobile.
However, since the material is solid, a reasonably large mobility of 0 2- can be attained only
above 700-800°C. Thus the cell is to be operated only at high temperature.
For thermodynamic measurements, ZrOz-CaO solid electrolyte has been widely used.
Other important ones are molten salts· based on KCI and Li Cl mixture as base electrolyte
(as H 20 is base for aqueous solutions), and SEs such as Zr02-Y 20 3, "ThOr Y 20 3, CaF2 doped
with YF3.
it to an external DC source such that the externally imposed voltage (Vext) opposes that of the
Danjel cell (Vceii). ·If Vext < Vcell> then the cell behaves as a galvanic cell with reaction (13.1)
occurring as shown. However. if Vex•> Vcell> then the direction of reaction (13. 1) is reversed,
i.e. the reaction becomes
Cu(s) + ZnS04(aq) = Zn(s) + CuS04(aq) (13.5)
In this situation, external electrical energy is being consumed to force reaction (13.5) to take
place, and the cell behaves as an electrolytic cell.
In Section 2.3, we have discussed the concept of a reversible proce~s. We may recall that
(a) it is very slow and (b) it occurs near equilibrium, and the process can be reversed by
shifting it from equilibrium in the other direction.
In an electrochemjcal cell, chemjcal reaction occurs only if the current is allowed to flow
through the cell. Ideally, no current flows and, hence, we have cell equilibrium, if the circuit
is open. The voltage at thi s condition is known as Electromotive force (i .e. emf) of a
galvanic cell. However, this emf can not be measured without a voltmeter. Use of a voltmeter
of very high resistance makes the current flow negligible and, hence the cell behaves as a
reversible galvanic cell. Another method is to impose voltage from an external source such
that Vex• is very close to the cell emf, and current flow is negligible.
where E = cell emf and /fq = infinitesimal quantity of electrical charge transferred across
the cell due to chemjcal reaction in the cel l.
Again, from Faraday's Laws of electrolysis,
/fq = ZFdn (13.7)
where
Z = valency involved in the chemical reaction
Textbook of Materials and Metallurgical Thermodynamics
RT
E=f!J - - Inf (13.ll )
ZF
We need not sketch an electrochemical cell in order to describe it. There is a standard
convention for its short representation. Accordingly, for example, the Daniel cell may be
represented as
Zn(s) I ZnS04 (aq) II C uS04Caq,. I Cu(s) (13.12)
(sat) (sat)
Single vertical li nes separate electrode from electrolyte. The double vertical line means
porous diaphragm.
An important question is: What should be the sign of the emf of the above cell-positive
or negative? For this, some convention is required. There is the European convention and
also the US convention. We shall adopt the US convention. Accordingly, if the reaction
proceeds spontaneously from left to right, the sign of E is positive. For the Daniel cell,
reaction proceeds spontaneously in the forward direction of Eq. (13.1) if the circuit is closed ,
i.e. Zn dissolves as ZnS04 and Cu is deposited at cathode from CuS04 . Therefore, in the
representation of Daniel cell (13. 12), reaction proceeds spontaneously from left to right, and
hence the emf is positive.
Thermodynamics of Electrochemical Cells
q} and q}' are electrical potentials of electrodes I and II respectively. (ZFl/J) dn; represents
the electrical work to be performed for transfer of dn; moles of component i. The quantity
{µ; + ZF¢) is the sum of chemical and electrical potential energy per mole of i, and is known
as the electrochemical potential of component i.
Proceeding further as in Chapter 10, for exchange of component i between the electrodes,
we have
( 13.16)
The two electrodes are at electrochemical equilibrium with one another (not chemical
equilibrium) when there is no current flow (i.e. cell voltage =emf= E). Under this situation,
dG' = 0. Therefore, from Eq. (13.16), at constant T and P,
(13.17)
In other words, the electrochemical potentials are equal at the two electrodes for reversible
exchange of component i between the electrodes. Again, under this condition,
q} 1 - ¢1 = cell emf = E (13.18)
Combining Eqs. (13.17) and (13.18), we get
µ.l 1 - µ: = -ZF(¢ 11
- ¢1) = -ZFE (13.19)
Equation (13.9) relating E with L1G is of general applicability to any emf cell.
Equation (13.19) is applicable only to concentration cells. Examples of both will be presented
in the following section.
Textbook of Materials and Metallurgical Thermodynamics
Let us recall the following equations from Sections 6.2 and 6.3,
Calculate the activity and the activity coefficient of PbO in the Pb0-Si0 2 electrolyte.
Solution In the above cell, solid platinum serves as the oxygen electrode since its
surface comes to equilibrium with the gas phase quickly without forming oxide. The cell
reaction is reaction (E.13.6), i.e. formation of PbO from Pb and 0 2.
For the emf cell (E.13 .5), therefore,
l
I
Textbook of Materials and Metallurgical Thermodynamics
- 12
10
11
p ~
v ~
ii ~
II
II
-
6
v r..
~
~
";
~ ~ I;
~
~ ~ II
v ~ II
.
~
i,. ~ II
~
i,. ~ II
~ c
II ~ II
~ ii ii
~
II ~ ii ii II
~ ii ~ II
~
II ~
"'
II
~
~
~ II
~ II
[,
ii
..,._ I;
i,.
I
;
''
'
II
II
ii
II
I
• I I
. ;
- 2
7 ; II
~ "'
' ;
-... - -- ----
~
----
-
ii
ii
~
~
I; - - -
I I : ::_ ------ ---
~
--- - -----
- -- ---- ---- ~ ~
3
-- - -- = === =-11
~ ~ - --- ~ ~ ~
8 ~
-·- -- ....,._
-- - - ---- -
- - ------
ii
II - :;;. _ ~ ~ ii
-
~
- - ~ - -- -----·~ ~
4
9 ,_ - ~ ~
ii
~
~ II
~· J J I I I I
- I I I I I
-
I I
-
-·11
I
i,
' ' ' ' ..... .... ..... ' ......... ' "' ' ' .... ii
'
;
;
ii
ii l. Solid electrolyte crucible
;
ii
;
ii
5
ii
~
[, 2. Recrystallized alumina crucible ii
~
~
ii [,
; ii
3. Outer alumina crucible ii
ii [,
ii
[,
4. Reaction tube-inconel [,
[,
~
~ ii
~
5. S upporting al umina tube ii
i,.
ii 6. Thermocouple protection tube ii
ii ii
7. Cermets
ii
;
;
ii 8. Tin-melt + Sn02(S) "I
; 8. Ag-Sn alloy melt+ Sn02(S) ii "
;
; 10 . Kanthal lead wires v"
v LI
,I LI
v 11. Double bore alumina sheath ii
i,.
I'
II
LI'
12. Chrome) alumel thermocouple ii
ii
ii
..
LI' ii ~
Fig . E.1 3.1 EMF cell for thermodynamic measurements in molten tin-silver alloys.
Ti n is much more reacti ve than sil ver. He nce, the chemical reaction in both electrodes is
(E. 13. 19)
S ufficie nt time was allowed fu r attainment of this equilibri um at each electrode, and
Thermodynamics of Electrochemical Cells
acu and azn are activities of Cu and Zn at respective electrodes. Since they are pure, acu =
aZn =1. In a Daniel cell, the electrolytes are saturated solutions. It means that the ZnS04
solution co-exists at equilibrium with solid ZnS04 , and the CuS0 4 solution with solid
CuS04 . In other words, in Daniel cell, the activities of ZnS04 and CuS04 in the electrolytes
are also 1. However, in unsaturated electrolytes, the activities in electrolytes will be less
than 1. Hence, from Eqs. (E.13.2) and (E.13.3), at 298 K,
(Gzn 4 )
E = 1.104 - 0.0128 In so (E.13.4)
(GcuS04)
Calculate the activity and the activity coefficient of PbO in the Pb0-Si0 2 electrolyte.
Solution In the above cell, solid platinum serves as the oxygen electrode since its
surface comes to equilibrium with the gas phase quickly without forming oxide. The cell
reaction is reaction (E.13.6), i.e. formation of PbO from Pb and 0 2 •
For the emf cell (E.13.5), therefore,
E
s
=£1
s
- RT In J
ZF
=FJ -
s
RT In [ (apbO)
ZF [aPb
]'nP0i,
l (E.13.7)
( -88,800) -
r-O -
c,5 - -
- + 0 ·460 VO
It
2 x 96500
EXAMPLE 13.3 Chlorides, fluorides etc. are low melting solids. Hence, for thermodynamic
measurements, emf cells, with these molten salts as electrolytes, can be fabricated with glass.
An example is the following cell:
(E.13.11)
RT
E= -2F ln[llcdlauoy (E.13.12)
For this, L1G = partial molar free energy of mixing of Cd in the alloy = GQi
J
= Cllcd Luoy =[Ocd Jalloy
(Ocd] pure
which is the same as Eq. (E.1 3.12). Therefore, this example demonstrates how either
Eq. (13.19) or Eq. (13.11 ) can be employed for solving problems of concentration cells.
In the solid oxide electrolyte, 0 2- is the current carrying ion. The emf of such a cell is due
to difference in chemical potentials of oxygen in the two electrodes. Hence, it is a concentration
cell, with concentration difference of oxygen of the electrodes being the cause of emf.
A general representation of these cells with ZrOrCaO electrolyte is
11!.. - 11!! RT p1
E = rvi rvi =- In ~ (E. 13..17)
4F 4F P8i
Such cells have found widespread use in industry for measurement of oxygen dissolved .•
in liquid steel, copper, furnace atmosphere control, combustion control and so on. These emf
cells have also been and are still being used widely for precise thermodynamic measurements
at high temperature. In fact, such cells act as oxygen meters (i.e. oxygen sensors) with
capabilities to measure oxygen concentration or partial pressure in a wide range of temperature
and P0z . These will be illustrated by the following examples.
EXAMPLE 13.4 Roy Chowdhury and Ghosh (Metallurgical Transactions, 2, 217 1-74,
1971) made thermodynamic measurements in liquid tin-silver alloys usi ng solid electrolyte
cell as follows:
Cermet I Sn(l) + Sn02(s) I Zr02-CaO I Sn-Ag(l) + Sn02 (s) I Cermet
(electrode I, P0i =Pb2 ) electrolyte pg2 )
(electrode II, P0i = (E. 13. 18)
The emf cell is shown in Fig. E.13.1. The apparatus and technique had been employed
by Ghosh and co-workers earlier. The electrolyte was in the form of a crucible. Pure Sn melt
with Sn02 was in outer alumina crucible. The alloy melt with Sn02 was inside the zirconia
crucible. The cermets served as nonreacting leads, and the entire assembly was at constant
temperature inside a furnace.
Textbook of Materials and Metallurgical Thermodynamics
- 12
10
11
" ~
"
II
~
~
~
"' ~
6
,, ~
~
"'
I\
I\
' ~
'
'' ~ p '' ;
~
I
v v ' ~
~
,, ' ~ ~
' ,."
~
' ; ~
' I ~ ~
,, ''
I .I ~ ~
I '' ,. I I I I 1
-
~
~
' . - 2
' ,,
~
7 ~ ' ~
"
~
1::=
- - - - -. ------ --- -- - -- -
-
-
- - -... - - -- "-- :_,,.:_~_':.
- ____
~
v
3
- -- --- --- -
~
- - ---
~
- - ----
II ~ ~
~
8
--- - -·- -- - --
-~ ~
-- _--- ~ ~
v
I\
~
I - --
- ----- ----- -· ~ 4
-,,,,.,,,,,,,,,,,, , ,,,,,,,. ,,,,,,.,,- ,, ,,.,,._,
~
9 ~ ~
;
~
I\ ~ ~-·
k-" _,,, _,,_,_,,,, ~
.I
I\
' ... , ........ , , , .... ,, ' '' .............. """.,. ..... "" .... ' ..
~
1. Solid e lectrolyte crucible I ~
5
~ v
,,
; 2. Recrystallized alumina crucible
;
~
~
3. Outer alumina crucible ~
..v
~
.I
~
.I
;
~
v 5. Supporting alumina tube
.. ~
.I
.I 6. Thermocouple protection tube .. II..v
.I ,.
.I
v v
~
II
.I 7. Cermets v ..
; v
II 8. Tin-melt + Sn02(S) II
..v
II
"v ..
II
8. Ag-S n alloy melt + Sn02(S) k-"
II ..
II 10. Kanthal lead wires "
II
II
II v 11. Double bore alumina sheath ,,
LI
1,.1 ~
12. ,,
II
LI
.. Chrome! a lumel thermocouple
.. .I
LI
Fig. E.13.1 EMF cell for thermodynamic measurements in molten tin-silver alloys.
Tin is much more reactive than silver. Hence, the chemical reaction in both electrodes is
(E.1 3.1 9)
Sufficient time was allowed fc,r attainment of this equili brium at each e lectrode, and
I
E qui·1·b .
1 num constant
K = { (llsoOz)
] } (E.13.20)
Cllsn P0z eq
Oso0z = 1 since it is pure solid, a 50 =1 for pure liquid tin. Hence, from Eq. (E.13.20),
K=-1-= __ 1_ (E.13.21)
P&i [llsn] P~
i.e. ll
P,I
i
P0z
= Casnla11oy (E.13.22)
RT
E =4 F In Casolanoy (E.13.23)
(Note that the sign of E is negative. The sign of measured E depends on how the voltmeter
is connected. It has to be taken as negative for calculation purposes.)
Figure E.13.2 presents activity vs. composition relationship in liquid Sn-Ag solution
at 1100 K. The activity of tin was experimentally measured. The activity of silver was
0 .8
0.6
0.4
0.2
calculated from experimental data by the Gibbs-Duhem integration using Darken 's method
(see Section 9.6.3). Pure 'silver is a solid at 1100 K. For GD integration, hypothetical liquid
Ag standard state was employed. The procedure and equations for this conversion have been
presented in section 10.3.3.
EXAMPLE 13.5 During steel making, measurement of dissolved oxygen content of liquid
steel is often required. It is done by oxygen sensor, which is an emf cell with SE electrolyte.
The sensor is immersed into liquid steel. Since it gets damaged, it cannot be reused. The
sensor also carries a thermocouple. The readings of emf and temperature are to be completed
within a minute.
Several commercial designs are available. A common design is:
Molybdenum and iron serve as leads. Since they are dissimilar, emf reading requires correction
of the themwcouple effect. The solid electrolyte is in the form of a small tube open at one
end.
From Eq. (E.13.16),
Since Cr and Cr20 3 are pure, we obtain the following relation on the basis of Eq.(8 .10):
1
202(g) = [O]dissolved in liquid steel (E.13.27)
where K 27 is equilibrium constant for reaction (E.13 .27) and ho is the activity of dissolved
oxygen in steel in 1 wt.% standard state (see Section 11.5). Hence,
=2RT In [ho]
K 11
(E.13.29)
1
<
At 1873 K, !ff/ of Cr20 3 = -422.7 x 103 J/mol 0 2, K27 = 2615. If E = - 0.153 V, then
\
from Eq. (E.13.30), ho = 0.0005. If the behaviour of oxygen in steel is assumed to be
Henrian, then
weight % 0 = W0 = ho = 0.0005 wt.% (i.e. 5 ppm) [Ans.] (E.13.31)
' "
Textbook of Materials and Metallurgical Thermodynamics
are at their respective standard states, as in cell (13.22), where a~11 , az112+, a"+ , PH2 are all
equal to 1. Here, the question arises- what do we mean by a i+ = 1 in an aqueous solution ?
Zn
az112+ simply c an never by measured. Hence, by convention, if the concentration of the salt
(ZnCl2) is 1 molal (i .e . 1 gm-mole in 1000 g water), then Eis taken as E0. This is indicated
in Table 13.2. If the ion exhibits Henrian behaviour, i.e. activity is proportional to molality,
then the calculation of E at concentrations other than l molal is simple. However, it is not
true in most cases and, hence, specific information of behaviour is required. Further discussion
of this topic is beyond the scope of this text.
Table 13.2 Standard single electrode potentials in aqueous solution at 298 K, 1 atm
(standard state is 1 molal)
The overall reaction in a Daniel cell is given by Eq. (13.2). It is the sum of Eqs. (13.3)
and (13.4). Hence, for the Daniel cell,
If the electrode and electrolyte are not at their standard states then, on the basis of
Eq. (13.11), the reduction potential for Zn 2+1zn is given as
E=Efl - -
RT
2F
In [-az,.-
aZn2+
l (13.26)
In Eq. (13.26), Zn is the reduced state and Zn2+ is the oxidized state. We may generalize
it as
Oxidized state + 2e- = reduced state (13.27)
On the basis of the above equations, the generalized equation is
E =t - RT In (aR
ZF a0
l (13.28)
Zno~-(aq)
a
Zn02_
= io-yr1'..~
Zn 2+(aq)
y=6 2 0
I Ozn2+ = 10-x J
Zn(OH}i
x=O 4 6
Eo = -0.763 v
/ x=O
-1
Zn
0 2 4 6 8 10 12 14 16
pH
Fig. 13.2 Simplified potential-pH diagram for Zn/H20 system at 298 K.
Thermodynamics of Electrochemical Cells
From Table 13.2, £o = - 0.763. For construction of Fig. 13.2, three values of ai~ and
az002- were assumed as 1, 104 and I 0-6 and 1, 10-2 and l 0-6 res pee ti vel y.
13.6 Summary
1. Electrochemical cells may be classified into galvanic cells (i.e. batteries) and electrolytic
cells. Only galvanic cells can be operated reversibly.
2. Besides aqueous solutions at room temperature, molten salts, molten slags (i.e.
oxides) and solid electrolytes are employed as electrolytes in metallurgy and materials
applications at high temperatures.
3. For a reversible galvanic cell, the cell emf(E) is related to thermodynamic quantities
by the two alternate forms of the equation:
(i) L1G = - ZFE
where L1G is free energy of cell reaction, Z is valency involved, and F is Faraday's
constant.
(ii) II I - - ZFE
Jlj-Jlj-
where i is the species exchanged between electrodes I and II, and µ; is the chemical
potential of i. 'i'he first form is used for reaction cells and the second form is
preferred for concentration cells.
RT
4. E =f.!J - -- In J
ZF
where E° is standard emf of the cell, and J is the activity quotient.
5. Some examples of emf cells have been discussed in this chapter. It may be specially
mentioned that, for a solid oxide electrolyte cell, equation in 3(ii) above may be
rewritten as
I
E = RT In Po2
11
4F p 02
where Pbz and p~ are partial pressures of oxygen at electrodes I and II, respectively.
The US Sign convention for emf has been adopted in this text.
6. The single electrode potential of an electrode (£), and electrochemical series in
aqueous solution have been discussed. £ can be taken as oxidation potential or
reduction potential, which are equal and opposite for a specific system. The reduction
potential has been c hosen for application in this text.
7. The fundamentals of Pourbaix diagrams or Potential-pH diagrams, have been illustrated
through the example of Zn-aqueous solution system.
Textbook of Materials and Metallurgical Thermodynamics
PROBLEMS
13.1 The emf of the cell
Pt, H2 (g) I H2S04 I Pb02(s) + PbS04 (s), Pt
aqueous
solution
was measured as 1.63195 and 1.62950 volts, respectively at 45°C and 35°C. Calculate
LIG, LIH and LIS of the cell reaction at 25°C.
13.2 The emf of the cell: Ag(s) I AgCl(s) I Cl 2(1 atm), Pt, was found to be:
£(volts) = 0.977 + 5.7 x 10-4 (350 - t) - 4.8 x 10-7 (350 - t) 2
in the temperature range t = 100°C to 450°C. Calculate the value of LICp for the cell
reaction .
13.3 For commercial production of magnesium, MgC12 is decomposed by electrolysis
according to the following reaction at high temperature:
Calculate the minimum theoretical voltage (to ov'!rcome cell emt) to be applied for
decomposition at 750°C. Assume MgCl 2, Mg as pure.
13.4 The emf of the cell: Al(s) I AlClrNaCl melt I Al-Zn alloy(s)
was measured as 7.43 millivolt at 380°C, and the temperature coefficient of the emf
as 2.9 x 10-5 volt/°C. Calculate aAi. c;:I and ii;:I in the alloy.
13.S Consider the high temperature emf cell in Eq. (E.13.18).
At Xsn = 0.5 in the Sn-Ag alloy, a measured cell emf was fitted with the equation:
£(volt) = 14.69 x 10- 3 - 30.4 x 10-6 T, where Tis temperature in K. Calculate (at
Xsn = 0.5):
(i)
(ii)
14.1 Introduction
By surface, we mean a free surface which separates a solid or liquid from surrounding gas
or vacuum. Ideally, the surrounding should be a vacuum. However, a gas phase has negligible
concentration of atoms/molecules pe r unit volume as compared to that in solid or liquid.
Hence, approximately any interface with gas is also treated as a free surface.
Surfaces play an important role in a variety of phenomena encountered in metallurgy and
materials science. A surface is only a few atomic layers thick. On the surface, there are some
free bonds projecting normal to surface. This causes a difference in structure of the surface
layer from that of the bulk of the solid/liquid.
A surface possesses some extra energy due to the work done to create it. Specific surface
energy is this excess energy per unit surface area (say, in joules per m2 of :mrface). From
an atomistic viewpoint, work is to be done to create free bonds on the surface (i.e. equivalent
to bond breaking). Bond energy is of the nature of enthalpy. Strictly speaking, it is internal
energy, which is approximately the same as enthalpy for condensed phases. Since H = G +TS,
excess surface energy is due to excess surface free energy and excess surface entropy. But
specific surface energy is usually taken as free energy (either Gibbs or Helmholtz free
energy, dependi ng on application) per unit surface area, in excess of that of the bulk.
An interface, by convention, is 'the surface separating two condensed phases. Here, we
do not have the concept of free bonds. Solid surfaces are rough. Hence, the interface between
two solids has only some contact points. Moreover, such contacts depend on surface roughness,
and hence, are unpredictable and irreproducible. Therefore, thermodynamics is not capable
of predicting its properties. Polished interfaces joined by bonding such as adhesives, soldering,
brazing etc., are more amenable to thermodynamic treatment. In contrast, solid/liquid and
liquid/liquid interfaces, especially the latter, can be more defi nitely reproducible. Excess
248
Thermodynamics of Surfaces, Interfaces and Defects
energy at the interface is influenced by the extent of composition mismatch and/or structural
mismatch (such as water and ice). At electrode/electrolyte, slag/metal, metal/aqueous solution
interfaces, we have the electric double layer also, which modifies interfacial energy. Hence,
the phenomena here are more complex than those at free ~urfaces. Grain boundaries in solids
are also interfaces. There, the mismatch in grain orientations gives rise to excess interfacial
energy (i.e. grain boundary energy).
Grain boundaries, stacking faults , point defects (i.e. vacancies and interstitials) and line
defects (i.e. dislocations) play very important roles in phase transformations, diffusion etc.
and in control of mechanical and other properties of solids, as well as their processing. Fine
particles have very high surface energy per unit volume, which are responsible for their
special behaviour pattern.
In all the above arl!as, it is thermodynamics again, which constitutes a principal scientific
foundation. However, other courses in the area of metallurgy and materials are offered ,
especially on structure, properties and processing of solid materials, where thermodynamic
aspects are also dealt with in detail. Hence, for the sake of basic information and understanding,
in this introductory text, onl y a brief mention will be made about thermodynamic aspects.
dG' =µ 1 dn1 + J1?. d"'i. + ... + µ; dn; + ... = "f.i µ, dn; (10. 11)
where 1, 2, ... are components, in the absence of any other work. In the presence of any
other work (OW'),
dG' = "f.µ;
i
dn; - 8W' (14.1)
dG' = "f.µ;
i
dn, + adA (14.3)
8W' is nothing but [- dG;]. which is the change in surface free energy. Hence,
dG; = adA (14.4)
Textbook of Materials and Metallurgical Thermodynamics
G'
G~ = aA, i.e. a= ; = GsA (14.5)
where G~ is surface free energy and GsA is specific surface energy per unit area. O" is a
constant for a surface, provided the surface is isotropic, i.e. it has no directionality. Liquid
surfaces are isotropic, whereas solid surfaces may or may not be isotropic.
You are familiar with surface tension, which is force per unit length of a surface. This
originated from Young's model, which considered the mechanical behaviour of a surface.
The actual interface is replaced by an imaginary infinitesimally thin elastic membrane, called
surface of tension. Dimensions of specific surface energy and surface tension are the same
(Mr2) . For isotropic surfaces, numerically also they are equal. Table 14.1 presents some
values of <1.
These are according to the names of the scientists who proposed them.
G
I I
G + S =S (14.6)
I
where G is the symbol for a monatomic gas molecule, S indicates a free surface
G
I
site, S is a surface site with adsorbed gas atom.
9 (14.7)
-
1 --8 -K
- a p,G
where
1- 0 = fraction of free surface
0 = fraction of surface site with adsorbed atom.
9, (1-0) are measures of surface concentration, i.e. mole fractions of surface sites. Since
the atoms in the adsorbed layer behave ideally, mole fractions are the same as respective
activities. These considerations constitute the basis of Eq . (14.7), which is the Langmuir
adsorption isothenn.
Textbook of Materials and Metallurgical Thermodynamics
1
2G2 (g) = G(g), Ks= li2
PG
(14.8)
Poi
From Eqs. (14.7) and (14.8),
It may be noted that Langmuir isotherm can also be applied to adsorption of a component
from the bulk of the solution. This requires invoking some additional equilibrium relations.
G' = I.G;n;
I
(14.10)
Since we are considering a surface, the surface energy term is to be included in Eq. (14.10).
Hence,
Combining Eqs. (14.13) and (14.3), we obtain the modified Gibbs-Duhem equation for
surfaces as
i.e.
n; da
-=-- (14.16)
A dJL;
By definition,
(14.18)
r..1
== (n;) = da
(14.19)
A surface dJli
Equation (14.19) is the mathematical statement of Gibbs adsorption isothenn. One of the
best examples of its application is for adsorption of oxygen atoms on surface of liquid iron
at high temperature, if we djssolve some oxygen in the liquid.
Figure 14.1 shows variation of awith wt.% oxygen dissolved in liquid iron, as measured
experimentally. Since d<Jldµ; is negative in this case, I'; is positive from Eq. (14.19). Thus,
oxygen is preferentially adsorbed on surface of liquid iron, i.e. it is surface active.
[wt.%, 0)
0.001 0.01 0.1
1600
...
I
0
E
......
,.., 1400
0
0
x
b
1200
1000.__~~--4-~~--'"~~~-'-~~-'-~~---'~~~
-8 -7 -6 -5 -4 -3 -2
In [wt.%, 0]
Fig. 14.1 Variation of surface tension of liquid iron with dissolved oxygen content at 1550°C
[F.A. Holden and W.D. Kingery, J. Phys. Chem., 59, 557, 1955).
Textbook of Materials and Metallurgical Thermodynamics
Figure 14.2 shows fraction of surface covered by adsorbed oxygen as function of wt.%
oxygen in the iron-oxygen solution. It was calculated from the data given in Fig. 14. l
assuming monolayer adsorption. It may be noted that the surface is almost saturated with
oxygen· even at as low as 0 .03% oxygen. In other words, oxygen is strongly surface active
in liquid iron . Sulphur also shows a similar behaviour. It may be mentioned that these
phenomena significantly influence steel making processes.
u
II)
1.0
0.8
Ideal
mono layer
I
/
/
--
~
....
::J I
"'
...... I
'
0
0.6
II)
on
eo;S
I
....
II) t
> I
0
u 0.4
e;;
c:
0
·p
u
.... 0.2
~""
OL-~~_._~~~~~~'--~~-'-~~~
It had been known for centuries that the level of water is higher inside a small diameter glass
tube. Not only that, the surface of water in the tube is concave upwards. This was attributed
to water 'wetting' glass. This is the capillarity phenomenon which you are familiar with. All
these, viz. the height of water inside the tube and the radius of curvature of water surface,
can be quantitatively calculated by force balance involving surface tension values. The
behaviour of mercury is opposite to that of water since it is ' non-wetting' to glass.
In a liquid-gas system, gas pres~ure and c urvature of interface are interrelated due to
surface energy effects. The general tendency of a gas bubble inside a liquid or a liquid
droplet in a gas is to assume spherical shape since spheres have lowest specific surface area
(i.e. surface area/volume ratio ), and hence the lowest surface energy per unit volume of the
liquid (i.e. per mole of liquid).
Thermodynamics of Surfaces, Interlaces and Defects
(14.20)
where L1Gp/V is the excess free energy of bubble per unit volume due to excess pressure (L1P)
inside the bubble. The excess pressure is due to squeezing action of surface tension force on
the bubble. Hence, the 2nd term in Eq. (14.20) containi ng L1GP is negative. At constant
temperature, dG = VdP. Hence,
(14.21)
The system would have the minimum L1Gxs at bubble-liquid equilibrium. Therefore, at
equilibrium,
The value of L1P can be quite large for tiny bubbles. For example, if the bubble radius
is 0.1 mm (10-4m), then L1P = 3.2 x 104 Pa (0.3 atrn) in liquid steel with O' = 1.6 J m-2 ,
at bubble-steel equilibrium.
2a
PA =pA0 + - ( 14.25)
r
0.031 - - - - - - - - - - - - - -
0.030
E 0.029 '
....
"';:,
N
~
0.028
0.027
pfln - - - - - - - - _-:_:-:_:-::_::-=---...J
0.026
0.01 0.1 10
r, micron
Fig. 14.3 Equilibrium vapour pressure of zinc at 900 K as function of radius of liquid Zn droplet.
Figure 14.4 schematically presents the features of Eq. (14.27). As may be noted from
the figure, T1r gets lowered to ~ due to surface energy of phase I. This phenomenon is
encountered during melting of fine particles of solid, or crystals with a highly curved
surface, such as at tips of dendrites, where melting occurs at a lower temperature. It is
obvious that the situation will be reverse and T1r will increase if phase II is of fine size.
Thermodynamics of Surfaces, Interfaces and Defects
(a)
I
t I
I c•
-~(/) ~ G
- - A(l)..,_G,
-- s
+ve (b)
t
L1G 0
........ ......
- - -- °i'~
.......
I 1 ......
I I ....... .......
I I <Jc.........
T~r Trr
Temperature --+
Fig. 14.4 Illustration of effect of surface energy on equilibrium phase transformation temperature
at constant pressure (schematic).
Liquid surfaces are isotropic. Also, atoms at liquid surfaces are mobile, and hence, equilibrium
is attained easily. But these are not generally applicable to solid surfaces. Therefore, the
meaning of surface tension is not clearcut for solids. Also, specific surface energy and
surface tension would not have the same numerical value. Consider a single crystal. It does
not possess a unique specific surface energy since the latter has different values for different
crystal j)lanes. However, for a polycrystalline solid with random orientations at the surface,
Textbook of Materials and Metallurgical Thermodynamics
the statistically averaged values would be isotropic, as in the case of a liquid surface. The
surface energy values of some solids are also included in Table 14.1.
The internal bou.n daries in crystalline solids may be divided into the following:
l. Internal boundaries between crystals of the same phase, resulting from orientation
differences (hereafter referred to as grain boundaries)
2. Internal boundaries separating phases of different structure or compositions or both.
The specific interfacial energy of a grain boundary is known as its grain boundary energy
Ge. Ge would depend on the misorientation, which is also known as the tilt angle (9) of
neighbouring grains. Ge is maximum at 9= 45° since it corresponds to maximum misorientation.
Figure 14.5 presents experimental values of Ge. as a function of 9 in copper at 1338 K.
Besides Ge. the figure also presents values of Ge/GsA• where GsA is surface energy per unit
area, as defined in Eq. (14.5). It may be noted that Ge/GsA is less than 0.4. It has been found
to be less than 0.5 for several metals. GsA corresponds to increase of free energy due to the
presence of free bonds at surface. But Ge is due to misorientation of bonds only. Thus, Ge
is expected to be lower than GsA·
0.40
600
0.32
500
' N
I
< 0.24
Cl)
'
-...
al
M
c.:> 300 0
0.16 \
\ x
200 C)
0.08
\ c.:>
100
0 0
0 10 20 30 40 50 60 70 80 90
9, degrees
Fig . 14.5 Experimental value of grain boundary energy as function of misorientation for a
simple t ilt boundary in copper at 1338 K [N.A. Gjostein and F.N. Rhines, Acta Met.,
7, 319, 1959].
'
'
=
Now, N + nv corresponds to (N+nv)IN0 y moles, where N0 is Avogadro's number. Hence, ,,
the number of moles of vacancies = yXv. Let, 'i•
Ll.S'c = Configurational entropy change due to mixing of 1tv vacancies in lattice sites.
Then, from Chapter 12, J
(14.28)
.
:
For a perfect crystal, the number of arrangements, W = 1. W' is the number of arrangements
of nv vacancies in N + nv lattice sites. From Eq. (12.23) and Stirling's approximation
[Eq.(12.24)],
(14.30)
If L1G is the free energy change due to formation of nv number of vacancies, then !
(14.31)
Figure 14.6 presents the various terms of Eq. (14.32) schematically as function of
vacancy concentration nv· R and No are universal constants. T, N, .&Iv, L1Sv are also assumed
constant in Fig. 14.6.
It may be noted that L1G goes through a minimum. The equilibrium vacancy concentration,
I
ny{eq), corresponds to this minimum. At this concentration, d(L1G)/dnv = 0.
Differentiation of Eq. (14.32) leads to
(14.33)
Textbook of Materials and Metallurgical Thermodynamics
+ve
.1 !2_ (LJH
No v
VJ
...
E
...
~
VJ
::I
0 LJG = ft.n)
·c:
>"'
_R_T (N In N + nv In _nv-'---)
-ve No N + nv N + n.
0
No. of vacancies (nv) --+
Fig. 14.6 Various terms of Eq. (14.32) as function of number of vacancies at constant temperature
(schematic).
Upon rearrangement,
x.(eq) = n.(eq)
N +n_(eq)
=exp (&v) R
exp (- m.)
RT
=exp (- LJG.
RT J
I (14.34)
where '1Gv = free energy of formation of 1 mole of vacancy and, Xv(eq) = equilibrium
vacancy fraction in lattice at temperature r.
Equilibrium concentration of interstitials can also be arrived at following the procedure
adopted above, and, in analogy with Eq. (14.34), we may write
Additional comments
Equations (14.34) and (14.35) are applicable to vacancies and interstitials originating from
thermal motions of atoms. These are known as thennal or intriTlsic vacancies and interstitials.
There is another type, which is fixed by composition and is independent of temperature, i.e.
the extrinsic type. In metals, the best known example is carbon dissolved in iron as extrinsic
interstitial. The carbon atoms are located at interstitial sites in iron lattice due to their small
atomic radii .
..1Sv, ..1S1 values are very small due to very low concentrations of vacancies and interstitials.
-
Henc.;_e, Eqs.
. .. (14.34)
. and (14.35) may be approximated as
Thermodynamics of Surfaces, Interfaces and Defects
(14.36)
& v and &; were calculated theoretically for materials like copper. It was found that
for copper, &/&v is approximately 7 . At 1000 K, from Eqs. (14.36) and (14.37),
Xv(eq) 39
- - ::10
X; (eq)
From this result we may make a generalized statement that, for a close-packed crystal such
as copper, the concentration of interstitials is negligible as compared to that of vacancies.
=
It was also calculated that, for Cu at 1000 K, Xv (eq) 3 x 10-7 This is indeed small.
Hence, vacancies will obey Henry's Law, and all properties related to the same will be
proportional to vacancy concentration. One such property is change of electrical resistivity
(L1p) of a metal due to vacancies. From Eq. (14.36), we obtain
(14.38)
Figure 14.7 presents L1p for gold as function of l/T. From the slope, &v was determined
as 92.5 kilojoules per mole of vacancy. For some other metals also, the values are near about
this. These values also agree reasonably with theoretically calculated ones.
Amongst ionic compounds, the oxides have been most widely studied in view of the fact that
their importance is next only to metals and alloys in metallurgy and materials science.
Compounds may be classified into stoichiometric compounds and non-stoichiometric
compounds.
(i) Stoichiometric compounds. Examples are Si02 , Al 2 0 3 , Zr02 , CaO, ZnO, where
metal/oxygen ratios are simple as in their chemical symbols,
(ii) Nonstoichiometric compounds. FexO is an example, where 0.85 < x < 0.97, and
Fe/O ratio is thus not a simple one. Another example is VOx-
In stoichiometric compounds, point defects, as mentioned in section 14.3.2, originate
either from composition effect or due to thermal disturbances. In ionic crystals, there are two
sub-lattices, viz. anionic and cationic. An anion cannot occupy a cationic sub-lattice, and vice
versa, ·since these require very high energies. Moreover, neighbouring positive charges and
negative charges have to be equal for local charge neutrality.
Textbook of Materials and Metallurgical Thermodynamics
T, °C
950 750 650 550 450
10-S
E
uI
E
.s:::
0
~ 10-9
10-10
Stoichiometric compounds
In Section 13.1, we have discussed ZrO:z-CaO solid electrolyte. It was mentioned there that,
when Ca2+ replaces Zr4+ in cationic sub-lattice, two + charges are less in the cationic site
occupied by Ca 2+. For local charge neutrality, an adjacent 0 2- ion site remains vacant, thus
creating 0 2- vacancy. This is an example of a stoichiometric compound where the point
defects are generated by composition effect.
Point defects arising from thermal effects will always occur in pairs-one positive and
another negative. These may be vacancy-interstitial pair, vacancy-vacancy pair, or pairs
where one is an electronic defect (excess electron or electron hole). For example, NaCl and
KCl exhibit Schottky defects, i.e. equal number of vacancies in M sub-lattice and X sub-
lattice, where M and X denote metal and nonmetal respectively.
As an example, let us consider a crystal MX, where M+ and x- are singly charged. Then
the reaction generating Schottky defects may be written as
The activity of MX is 1 for pure MX. Since o« 1, the activity of M (J-0,• X{l-5) also may
Thermodynamics of Surfaces, Interfaces and Defects
be set equal to l. Therefore, the equilibrium relation for reaction (14.39) may be expressed
as
(14.40)
(14.41)
we have (14.42)
(14.44)
In addition to the above, further relations are obtained from conditions of electroneutrality,
stoichiometry and relation between [e+] and [e- ], as follows:
l vM J + l e- J = l Vx J +le+J (14.45)
(14.47)
J,
In principle, the six unknown concentrations, viz. [VM], [ VM [Vx]. [ Vx].
[e-] and [e+]
can be solved from the four equilibrium relations and two charge and stoichiometric balances.
Non-stoichiometric compounds
Let us take the example of Wustite (FexO) again. In section l l .2.2, we discussed general
issues related to the activities of Fe, 0, "FeO" in the Wustite field and how these are related
to partial pressure of oxygen in the surrounding atmosphere, with which it was assumed to
be at equilibrium. To sum up, the value of x would be governed both by temperature and
p0z . Therefore, defect concentrations in Wustite also will depend on the above variables.
Experiments have established that the dominant point defect is cation vacancy. Moreover, it
is a p-type semiconductor, i.e. it has electron holes. Some other examples of this type of
defect are NiO, Cu2 0, FeS. Figure 14.8 presents log X vs. log P0i in Wustite at 1223 and
1273 K. From the slope, it is found that log X oc p~ approximately, where X is electrical
J
conductivity. Recognizing that Xis proportional to concentration of electron holes, le+ oc p~.
Textbook of Materials and Metallurg1cal Thermodynamics
-0.6
-0.7
~
bl)
0 ........ 8
-0.9
(14.48)
where 0 0 is oxygen in lattice site, V/c- is vacancy in Fe2+ site, and hence is equivalent to
two nega~ive charges. Since 0 0 is very large compared to defect concentration, it hardly
changes and can be assumed as constant. Hence, we may write
( 14.49)
(14.5.1)
14.4 Summary
1. A surface possesses some extra energy due to the work done to create it. The
relationship is
-SW= dG; = adA
Thermodynamics of Surfaces, Interfaces and Defects
where G~ is excess surface free energy, O' is specific surface energy per unit surface
area, and A is surface area.
2. a is lowered by the phenomenon of adsorption at the surface. The adsorption
equilibrium relations at constant temperature are known as adsorption isotherms. We
are primarily concerned with chemisorption. For this, the following two are standard
isotherms:
(i) The Langmuir adsorption isotherm deaJs with monolayer adsorption of ideal
gases on a surface.
(ii) The Gibbs adsorption isotherm deals with adsorption of a component of a
solution onto its surface, and is given as
r.,-- - dO'
dfl;
X,(eq) =exp (- : )
where ,1J/v and .dHi are enthalpies of formation per mole of vacancies and interstitials,
respectively. We have also shown that Xv (eq) » X; (eq).
6. In ionic compounds, all the above defects are present. However, additional features
arise due to electrical charges associated with vacancies and interstitials, as well as
existence of electronic defects. In nonstoichiometric oxides, again, defect concentrations
depend on oxygen potential. In this chapter we have presented some thermodynamic
derivations.
Selected Thermodynamic Data
Cp = a + bT + cr-2, joules/mol/K
Substance a bx 103 c x 10-5 Range, K
Table A-2 Standard Molar Heats of Formation and Molar Entropies at 298 K
Al(s) 0 28.34
Al203(s) -1,674,000 51.1
C(diamond) 1,900 2.44
C(graphite) 0 5.7
CO(g) - 111,700 198.0
C02(g) -394,100 213.9
Cr(s) 0 23.78
Cr203(s) -1,120,300 81.2
FeO(s) -259,600 58.81
Fe304Cs) -1,091,000 151.53
MnO(s) -384,700 59.12
Ni(s) 0 29.8
NiO(s) -244,600 38.0
02(g) 0 205.l
Ti(a) 0 30.56
TiC(s) -183,700 24.3
W(s) 0 33.48
W02(s) -589,800 66.95
W03(s) -844,300 83.3
ZnO(s) -348,300 43.53
Appendix
In p(atm) = -AIT + B In T + CT + D
Metal A B c x 103 D Range, K
Ag(l) 33,200 -0.85 20.31 Tm-Tb
Al(l) 37,884 -1.023 21.83 Tm-Tb
Cu(l) 40,350 -1.21 23.7 Tm-Tb
Sn(l) 35,697 12.32 Tm-Tb
Zn(I) 15,250 - 1.255 21.79 693- 1180
'
•
Bibliography
A. TEXTS
Darken, L.S., and Gurry, R.W., Physical Chemistry of Metals, McGraw-Hill, New York
(1953).
DeHoff, R.T., Thennodynamics in Materials Science, McGraw-Hill, New York (1993).
Denbigh, K., The Principles of Chemical Equilibrium, Cambridge University Press, Cambridge
(1963).
Gaskell, D.R., Introduction to the Thennodynamics of Materials, 3ro ed., Taylor & Francis,
Washington DC (1995).
Johnson, D.L., and Stracher, G.B., Thennodynamic Loop Applications in Materials Systems,
The Minerals, Metals and Materials Society, Warrendale, Pennsylvania (1995).
Kubaschewski, 0., Spencer, P.J. and Alcock, C.B., Materials Thennochemistry, 6th ed.,
Pergamon Press, Oxford (1993).
Swalin, R.A., Themwdynamics of Solids, John Wiley & Sons, New York (1964).
Upadhyaya, G.S., and Dube, R.K., Problems in Metallurgical Thennodynamics and Kinetics,
Pergamon Press, New York (1977).
B. DATA SOURCES
Elliott, J.F. and Gleiser, M., Themwchemistry for Steelmaking, Vol.I, Addison Wesley,
Reading, Mass, (1960).
Elliott, J.F., Gleiser, M., and Ramakrishna, V., Thennochemistry for Steelmaking, Vol. II,
Addison Wesley, Reading, Mass, (1963).
Hultgren, R., Orr, R.L., Anderson, P.O., and Kelley, K.K., Selected Values of Thennodynamic
Properties of Metals and Alloys, John Wiley & Sons, New York (1963).
JANAF Thermochemical Data, The Dow Chemical Co., Midland, Michigan (1962-63).
271
Textbook of Materials and Metallurgical Thermodynamics
Kelley, K.K., Co11tributions to the Data 011 Theoretical Metallurgy XIII, Bulletin 584, U.S.
Govt. Printing Office, Washington DC (1960).
Knacke, 0., Kubaschewski , 0., and Hesselman, K., Thennochemical Properties of Inorganic
Substances, Vols. I and II, 2nd ed., Verlag Stahleisen mbH, Dusseldorf, Germany (1991).
Kubschewski, 0., Spencer, P.J., and Alcock, C.B., (already noted in TEXTS Section)
Schick, Harold L. Thennodynamics of Certain Refractory Compounds, Vol. 2, Academic
Press, New York (1966).
The Japan Society for the Promotion of Science, Steelmaking Data Source Book, The 19th
Committee on Steelmaking, The Gordon and Breach Science Publishers, New York (1984).
Answers to Problems
[Note: Unit of energy is either J (i.e. Joules) or kJ (i.e. kilojoules). Wherever in the
problem, the amount of substance is not mentioned, the values are per mole (i.e. per gm-
mole) of either a substance or per mole of reaction, as the case may be]
CHAPTER 2
1. Lili - L1U = -6424 J
2. (i) w = - 2331 J, q = - 1639 J
(ii) L1U = 692 J, Lili = 1153 J
3. (i) W = q = 11825 J, L1U = Lili = 0 for 151 stage
(ii) W = 13317 J, q = 0, L1U = - 13317 J, Lili= -18800 J
4. (a) W = 2748 J, q = 0, L1U = - 2748 J, Lili = - 3848 J
(b) W = 4539 J, q = Lili = 15888 J, L1U = 11349 J
S. W(step iii) = 4.51 kJ, q(for cycle) = - 8.5 kJ
CHAPTER 3
1. 59.9 kJ
2. -42.2 kJ
3. - 186.7 kJ
4. -404.9 kJ
S. Llli0 = - 288 .2 kJ, L1U0 = - 284.0 kJ
6. Process (ii) releases 906 J more heat.
7. - 562.5 kJ
8. 1589 K
9. 349.4 kJ
CHAPTER 4
1. L1S(syst) = - 53.26 J/K, L1S(surr) = 625 J/K
2. (a) 18.1 J/K; (b) O; (c) -28.72 J/K
273
Textbook of Materials and Metallurgical Thermodynamics
3. -171.61/K
4. - 13.64 J/K
5. L1S(syst) =- 11.64 J/K, L1S(surr)= 12.84 J/K
6. .d.S'°(reaction) = &(syst) = -89.39 J/K, L1S(universe) = L1S(syst + surr) = 198.8 J/K
7. (a) 3589 K; (b) 0.0075 Cp; (c) process spontaneous
8. 118.0 J/K.
CHAPTER 6
1. - 1729 J
2. 8586 J
3. 16.46 J, No
4. -2.42 K
5. 78450 atm
6. (i) 260.6 kJ; (ii) 7.07 J/K
7. 908 atm; as the triple point pressure is 5.14 atrn, the 1-atm isobar does not pass
through the liquid field .
8. 99.95%.
CHAPTER 7
1. 8.5 x 108
2. - 2707 J
3. (i) 3.0 J/K; (ii) - 131.1 J
4. &J° = - 242.7 kJ, L1U0 = - 240.6 kJ, L1S0 = -89.73 J/K, L1G0 = -197.8 kJ
5. L1H° =- 602.4 kJ, L1S° =- 357 .2 J/K
6. Ni shall not be oxidized; L1Gj (NiO) = 108.8 kJ,
L1Gj (Cr20 3) =- 171 .5 kJ per mole of 0 2
7. Cr23C 6 stabler than Cr
8. L1H° = 30.42 kJ; & 0 = 65.9 J/K; L1G0 = 7.35 kJ
9. (i) 7.1 x 107 , 6.4 x HP; (ii) Lllij = - 234.4 kJ, L1Sj = - 84.1 J/K
CHAPTER 8
1. No oxidation
2. H 2S "' 0, 0 2 = 25, H;?O = 50, S 2 = 25 volume %
3. (a) 4.56 x 10-8 atm; (b) Titanium will get oxidized.
4. 43,800 J heat evolved
5. 0 .371
6. 5250 atm
7. (i) 51 %; (ii) No change
8. Conversion of Cu 20 into Cu2S possible
9. Not at equilibrium
10. (i) 1166 K; (ii) 0 .053 atm; (iii) 1.23 atm
(Note: Since 1500 K > 1166 K, no CaC03 will be present and Pc0z. at 1500 K will
be governed by Gas Law.)
Answers to Problems
CHAPTER 9
V,m = Kf[b'(ln
B b'P
rB)]
T, caqi
CHAPTER 10
1. }lc(gas) = - 46.24 kJ/mole C ; cannet saturate steel with carbon.
2. F = l ; 11N2 = - 205.3 kJ; stable phases at 1200 K: Si(s) + gas
3. 3
4. Psi = - 126.3 kJ; Cu will be converted to CuiS.
s. - 320.5 kJ
6. L1G';. = 13f 10 - 1.3T Jn T + 2.554T J
7. (i) 0.003 p1Jm (i.e. 3 ppb); (ii) Al4C3 will form.
8. 1eu (/J) =29.5
CHAPTER 11
1. (i) 0.396; (ii) Proi = 0.049 atm, Pco = 0.951 atm
2. 0.846
3. 1002
4. 1.35 x 10-5
S. [XZn] in Pb = 2.4 x 10-5
6. (i) 2.13 x 10-5 atm; (ii) 1.14 x 10- 3 atm
7. (i) XFe0 = 0.001, XMoO = 0.999; (ii) 2.6 x 10-27 atm
8. -50.0 kJ per mole 3Ca0.Al20 3; 9.79 x 10- 3
9. (i) 1.3; (ii) -163 kJ
10. (i) -36.8 kJ/mole Cu; (ii) -69.6 kJ/mole Mn
11. -0.341
12. 0.0125
13. 0.09 wt.%
14. -567.8 kJ
15. -36.39 kJ
CHAPTER 13
1. LlG = -314.l kJ, &I = -30.l kJ, LlS = 44.l J/K
2. -0.022T J/mole/K
3. 2.47 volts
4. aA1 = 0.673, GAj = -2152 J, H'fi =3315 J
S. (i) asn = 0.43, Gfn = -8414 J, Gl-5 = -1499 I
0
(ii) Ss~ = 11.74 J/K, .SS";; = 5.98 J/K, Hg:, = B:; = 5673 I
6. 0.231 volt
7. (i) - 127.0 kJ; (ii) -177.6 kJ
8. (i) 0.813; (ii) 0.70
Index
A Chemical thermodynamics, 2
Clapeyron equation, 82
Activity, 96 Classical thermodynamics, 1-3
coefficient, 131 Clausius-Clapeyron equation, 87, 120
composition and, 131 , 137 Closed system equations, 65
free energy and, 100-101 Composition parameters, 98
fugacity and, 100 Compressibility, 67
Henrian, 201-202 Configurational entropy, 225
ionic, 191 Constant pressure processes, 15-18
quotient, 101 Constant volume processes, 15-18
Adiabatic processes, 16 Cyclic processes, 11
expansion in, 19
reversible, 20 D
Adsorption isotherms, 250-254
Alloy phase diagrams, 169 Dalton's Law, 98
Alpha function , 145 Daniel cell, 229
Avogadro number, 7, 28, 219-226, 259 Debye equation, 28
Defect thermodynamics, 256-264
B Degrees of freedom, 165
fi:I Index
Joule, J.P., 16 0