0% found this document useful (0 votes)
23 views

Chapter 1 Diffusion

This chapter introduces concepts of chemical thermodynamics including phase diagrams, potential diagrams, and Gibbs free energy diagrams. It discusses classical thermodynamics which deals with physical properties of systems in equilibrium without assumptions about atomic structure. Statistical thermodynamics provides additional information by relating thermodynamic relationships to molecular behavior. The chapter concludes by introducing commonly used thermodynamic conventions and the laws of thermodynamics, focusing on those essential for understanding energetics in materials science.

Uploaded by

Gitesh Pande
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
23 views

Chapter 1 Diffusion

This chapter introduces concepts of chemical thermodynamics including phase diagrams, potential diagrams, and Gibbs free energy diagrams. It discusses classical thermodynamics which deals with physical properties of systems in equilibrium without assumptions about atomic structure. Statistical thermodynamics provides additional information by relating thermodynamic relationships to molecular behavior. The chapter concludes by introducing commonly used thermodynamic conventions and the laws of thermodynamics, focusing on those essential for understanding energetics in materials science.

Uploaded by

Gitesh Pande
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 38

Chapter 1

Thermodynamics, Phases, and Phase


Diagrams
In this chapter, we will briefly go through the basics of chemical thermodynamics.
It is assumed that the reader is somewhat familiar with the fundamental concepts,
and therefore, they are not discussed in great detail. The emphasis of the chapter is
to build a thermodynamic foundation that can be utilized in the later chapters for
diffusion kinetic analyses. We will put special emphasis on the use of different
types of diagrams to represent thermodynamic data. Therefore, we introduce phase
diagrams, potential diagrams, and Gibbs free energy diagrams in considerable
detail. These ‘‘tools’’ are then used extensively in diffusion kinetic analysis later on
in the book. We will conclude the chapter by introducing some commonly used
thermodynamic conventions.
Classical thermodynamics is a phenomenological theory which deals with the
physical properties of macroscopic systems under equilibrium conditions and the
relations between them. The great importance of classical thermodynamics lies in
its exactness as well as in its generality. It does not make any assumptions concerning
the atomic structure of the system nor the interactions between the atoms.
Even though this can be regarded as being beneficial in many applications, this can
also be regarded as a weakness, especially in the case of solids and their solutions
and compounds. Statistical thermodynamics, on the other hand, strives to obtain
thermodynamic relationships based on the molecular behavior of matter. It provides
additional information that cannot be achieved with classical thermodynamics.
Firstly, statistical thermodynamics shows that the laws of thermodynamics
are a direct consequence of the principles of quantum theory combined with one
very general statistical postulate. Secondly, statistical thermodynamics provides
general relations that cannot be derived from the laws of thermodynamics. Most
importantly, by utilizing statistical thermodynamics, it is possible to obtain a
physical understanding of the properties of solutions and about the reasons for
their behavior. Thus, it is beneficial to utilize both of the approaches described
above to obtain a more fundamental understanding of the behavior of different
material combinations. The subject of thermodynamics is vast, and there are a
large number of excellent books available [1–5]. The following seeks to summarize
those parts of thermodynamics that are considered essential for a basic
understanding of energetics in materials science. Further, the topics in Chap. 1 are
A. Paul et al., Thermodynamics, Diffusion and the Kirkendall Effect in Solids,
DOI: 10.1007/978-3-319-07461-0_1, _ Springer International Publishing Switzerland 2014
1
chosen in such a way to be closely correlated to the use of thermodynamics in the
diffusion calculations from subsequent chapters. The treatment utilized in Chap. 1
partly follows the approach presented in the comprehensive textbook written by
Kivilahti [6].
1.1 Thermodynamics System and Its State
The system is a clearly defined part of a macroscopic space, distinguished from the
rest of the space by a physical boundary. The rest of the space (taking only the part
that can be regarded to interact with the system) is defined as the environment. The
system can be isolated, closed, or open depending on its interactions with the
environment. An isolated system cannot exchange energy or matter, a closed
system can exchange energy, but not matter, and an open system can exchange
both energy and matter with the environment. A system can be homogeneous, thus
thoroughly uniform, or heterogeneous. A homogeneous system is defined as a
phase, which can be either a pure component (element or chemical compound) or a
solution phase. A heterogeneous system, on the other hand, is a phase mixture.
Thermodynamics aims to determine the state of the system under investigation.
From experiment, it is known that when a certain number of macroscopic variables
of the system have been fixed, the values of all other variables are also fixed and
the state of the system becomes fully determined. In thermodynamics, the variables
can be extensive, intensive, and partial. Extensive properties depend on the
size of the system, whereas the intensive properties do not. Partial properties are
the molar properties of a component. Those variables which are chosen to represent
the system are called independent variables. A macrostate of the system is
characterized, for example, by its temperature (T), pressure (p), and composition
(ni) or temperature (T), volume (V), and composition (ni). A macrostate does not
change over time if its observable properties do not change. The system can,
however, go through changes in its state for a number of different reasons. These
changes can be reversible or irreversible. A reversible change is a change that can
be reversed by an infinitesimal modification of a variable, whereas irreversible
processes have a definite direction which cannot be reversed. In a system, the
energy of that system is constantly being redistributed among the particles of that
system. The particles in liquids and gases are constantly redistributing in location
as well as changing in quanta value (the individual amount of energy that each
molecule has). Every specific arrangement of the energy of each molecule in the
whole system at one instant is called a microstate. The nature of every microstate
implicitly contains the important concept of fluctuations in it. It is evident that a
given macrostate can be represented by number of different microstates.
2 1 Thermodynamics, Phases, and Phase Diagrams
1.2 The Laws of Thermodynamics
Thermodynamics is based on a few empirical generalizations, which are stated in
the form of the following laws.
The zeroth law defines temperature such that if two systems are independently
in equilibrium with a third system, they must also be in equilibrium with each
other. Then, they have a common state variable—temperature.
The first law states the principle of conservation of energy such that the
macrostate of a system can be characterized with an extensive variable, called
internal energy E, which is constant in an isolated system. When the system
interacts with the environment and transfers from one macrostate to another, the
infinitesimal change in the internal energy can be stated as
E ¼ dq þ dw ð1:1Þ
where dq and dw are the heat and work transferred into the system during the
change. When the system receives heat from the environment, dq[0, and when
the system gives up heat, dq\0. The same is, of course, true for the work
transferred. If the system does work, dw\0, and if work is done on the system,
dw[0.If the external pressure acting on the systems’ straight interface is p, then
dw ¼ _pdV, if the expansion work is the only form of work. The internal energy
of the system is a state function. This means that dE is an exact differential. During
a change, its value is, therefore, independent of the path between the initial and the
final states. It is to be noted that dq and dw are not exact differentials, but infinitesimal
quantities of heat and work, and thus, they are path functions. Their value,
when integrated, depends on the path between the initial and final states.
The second law gives the criteria for the spontaneous change in nature that
allows the macrostate in equilibrium to be characterized by a variable S, the
entropy, which has the following properties
(i) Entropy, which is defined as

dq
T
__
rev
ð1:2Þ
is a state function. In Eq. 1.2, the subscript rev refers to a reversible process.
Entropy can be expressed as a function of the independent state variables of the
system as S ¼ SðE; Vi; niÞ. The infinitesimal entropy change of a closed system in
an arbitrary reversible process can be thus written as
dS ¼
oS
oE
__
dE þ
oS
oV
__
dV ð1:3Þ
1.2 The Laws of Thermodynamics 3
By utilizing the first law ðdEÞV;ni¼ dq and Eq. 1.2, we obtain
oS
oE
__
V;n
¼
1
T
ð1:4Þ
where T is the absolute temperature.
(ii) The entropy of the system is an extensive property.
(iii) The entropy of the system can change for one of two reasons, either as a
result of the transfer of entropy between the system and the environment or
by the creation of entropy within the system. The entropy change can be
written as
dS ¼ deS þ diS ð1:5Þ
where diS is the entropy created within the system. From the experiment, it is
known that this quantity is always positive. During a totally reversible change, the
entropy change can be zero. When the system is isolated, its entropy can never be
decreased
dS ¼ ðdSÞE;V¼ diS_0 ð1:6Þ
Hence, in real irreversible processes, the entropy of an isolated system always
increases and reaches its maximum at the equilibrium state.
The third law states that the entropy of the system has a property that S ! So
when T ! 0, where So is a constant independent of the structure of the system. At
absolute zero, the entropy of pure, defect-free, crystalline elements has the same
value, So, which has been chosen to be zero.
Thus, the thermodynamics of closed and isolated systems is based on the following
equations
dE ¼ dq þ dw ðfor all changesÞ
dS ¼
dq
T
ðfor reversible changesÞ
dS_0 ðfor changes in isolated systemsÞ
These equations can be combined to give the fundamental equation for a closed
homogeneous system
dE ¼ TdS þ dw or
dE ¼ TdS _ pdV
ð1:7Þ
if only expansion work is considered.
4 1 Thermodynamics, Phases, and Phase Diagrams
Note The concept of entropy is highly ambiguous. Several interpretations
have been given to entropy. The entropy law is a consequence of the fact that
matter is composed of interacting particles that are in motion and which
constantly show a tendency to muddle up and thereby to mix both matter and
energy. Thus, it has been proposed that entropy is the measure of the systems
mixed-upness (Gibbs), or the degree of disorder (Planck). According to
Guggenheim, entropy is the measure of the spread of energy and matter.
Shannon, on the other hand, has defined entropy as the lack of information or
data [6].
The above equations are valid for closed systems with fixed composition. In
order to extend the treatment to open heterogeneous systems, we need to choose a
third variable, one that describes the composition and quantity of the system. This
is the ni being the number of moles of component i.
1.3 Heterogeneous Systems
A heterogeneous system is composed of several homogeneous subsystems,
meaning phases which each have their own energy E/, entropy S/, and composition
ni
/ (i = 1, 2,…k). Consequently, the energy, the entropy, and the number of
moles of substance of the phase mixture are

X
/
E/ ð1:8Þ

X
/
S/ ð1:9Þ

X
/
n/ ¼
X
/
X
i
n/i ð1:10Þ
To exactly determine the state of the phase mixture requires that each phase that
it contains must be described accurately. If we choose the variables (S, V, and ni) to
describe the state of a given phase, all other properties are then necessarily
functions of the chosen variables. This means that especially the internal energy of
a phase can be expressed as E/(S/, V/, ni
/). Its exact differential for an arbitrary
change can be written as
1.2 The Laws of Thermodynamics 5
dE ¼
oE
oS
__
dS þ
oE
oV
__
dV þ
X
i
oE
oni
__
dni ð1:11Þ
When the composition of the phase does not change, Eq. 1.7 is valid and the
first two partial derivatives in Eq. 1.11 are temperature of the phase (T) and its
pressure (p). The last term is defined as the chemical potential of a component i.
The chemical potential is defined formally in the following [1]: ‘‘If to any
homogeneous mass we suppose an infinitesimal quantity of any substance to be
added and its entropy and volume remaining unchanged, the increase of the energy
of the mass divided by the quantity of the substance added is the (chemical)
potential for that substance in the mass considered.’’
Consequently, we obtain an equation for the change in the phase internal energy
dE/ ¼ T/dS/ _ p/dV/ þ
X
i
l/
i
dn/i ð1:12Þ
This equation is the fundamental equation for the independent variables S, V,
and ni. The internal energy E is their characteristic function, the thermodynamic
potential of the phase. In a thermodynamic system, each phase has such a
potential.
Next, a new thermodynamic function is defined with the help of internal energy
and entropy
F ¼ E _ TS ð1:13Þ
By differentiating the function and by substituting Eq. 1.13 into the differential
form of 1.12, we obtain
dF/ ¼ _S/dT/ _ p/dV/ þ
X
i
l/
i
dn/i ð1:14Þ
This equation defines the Helmholtz free energy F, which is a function of the
independent variables T, V, and ni. The properties of this free energy function shall
be discussed in more detail in Sect. 1.5.
Let us further examine the function E/ with independent variables (S/, V/, and
ni
/). Because T, p, and li are intensive variables, they are not dependent on the
amount of phase /. From this, it follows that as the intensive variables remain
constant, Eq. 1.12 can be integrated. This gives the thermodynamic potential of
phase / as
E/ ¼ TS/ _ pV/ þ
X
i
l/
i
n/i ð1:15Þ
Next, we define two new functions, the enthalpy (H) and the Gibbs free energy (G)
6 1 Thermodynamics, Phases, and Phase Diagrams
H _ E þ pV ð1:16Þ
G _ H _ TS ð1:17Þ
By recalling the definition of Helmholtz free energy (1.13), Eq. 1.15 (together
with 1.16 and 1.17) yields a function
G/ ¼
X
i
l/
i
n/i ð1:18Þ
This function (Gibbs free energy) is also a thermodynamic potential of a phase,
and it is an extensive variable. Therefore, the Gibbs free energy of a phase mixture
is given as

X
/
X
i
l/
i
n/i ð1:19Þ
When the Gibbs free energy function for a phase is known, all other thermodynamic
properties of a given phase can be expressed with the help of this
potential and its derivatives. The properties of the Gibbs free energy function are
discussed in more detail in Sect. 1.5.
1.4 Commonly Used Terms and First Glance at Phase
Diagrams
Thermodynamics is an exact discipline. Therefore, it is of great importance to
define a few more key terms which will be frequently encountered later on in the
text. A component refers to independent species in the system under investigation,
giving the minimum number of substances which must be available in the laboratory
in order to make up any chosen equilibrium mixture of the system in
question. A phase is a region of uniformity in a system under investigation, as
already stated. It is a region of uniform chemical composition and uniform
physical properties. A phase is also distinguished from other dissimilar regions by
an interface.
To illustrate the concepts of compound and a phase, we will consider a simple
example H2O. Ice, water, and water vapor are all different phases of the compound
H2O that exist in different temperature and pressure ranges, as shown in Fig. 1.1.
The diagram shown in Fig. 1.1 is called a unary phase diagram and is shown for
water in the figure. The point marked as C is called the critical point. When temperature
rises above that critical point, the gas phase (water vapor) cannot be
liquefied by increasing the pressure. The curve TC gives the equilibrium vapor
pressure of the liquid as a function of temperature up to the critical point. At point
T (called the triple point), all three phases of water are in equilibrium with each other.
1.3 Heterogeneous Systems 7
Based on the Gibbs phase rule (derived later on) at this point, the number of degrees
of freedom is zero. The equilibrium can therefore be attained only at a specific
temperature and pressure. The curve ST gives the equilibrium vapor pressure of the
solid (ice) as a function of temperature. The curve TM gives the change in the
melting point of ice as a function of pressure. It is to be noted here that the curve TM
for the system H2O is highly unusual as the TM curve here is descending, whereas in
most of the systems, it is ascending. This is a result of the fact that the molar volume
of solid water (ice) is larger than that of liquid water (in Sect. 1.7 is introduced the
Clausius–Clapeyron equation that can be used to calculate this). In most systems,
however, the opposite is true. Another unary system exhibiting this type of behavior
(i.e., larger volume in solid than in liquid) is bismuth (Bi).
Different pure elements, for example, Cu or Ni, also have three different phases:
solid, liquid, and gas. Similarly, two allotropic forms, solid gray tin and white tin,
which have a different crystal structure and properties, are considered as distinct
phases. To show the example of phases with two different components, we consider
the Ag–Cu binary phase diagram, which is shown in Fig. 1.2. All phases are
made of the two components, Ag and Cu. The different phases a, b, and liquid are
stable within a certain temperature and composition (expressed here as weight
percentage) range. Note that the phase diagram shown in Fig. 1.2 is determined at
constant pressure. The a-phase is basically a solid solution Ag(Cu), that is, Ag
(with its face-centered cubic (FCC) structure) with a limited amount of dissolved
Cu, whereas the b-phase is a solid solution Cu(Ag), that is, Cu (with FCC structure)
with a limited amount of dissolved Ag. Different notations (a and b) are used
to differentiate solid solutions from pure elements. The solvus curve separates the
single solid-phase region a from the solid two-phase region a + b. Similarly,
another solvus curve separates the one-phase solid region b from that of the solid
two-phase region a + b. The solidus curve separates the solid one-phase a-region
from the two-phase region where the solid a and the liquid are in equilibrium.
Similarly, another solidus curve separates the solid one-phase region b from the
two-phase region b + liquid. The liquidus curve, on the other hand, separates the
two-phase a + liquid and b + liquid areas from the liquid one-phase area L. In
Fig. 1.1 The pressure–
temperature diagram of H2O
8 1 Thermodynamics, Phases, and Phase Diagrams
Fig. 1.2, there is a horizontal line of specific importance. It represents the so-called
eutectic reaction, where liquid L reacts to form two new solid phases a and b. At
the line, there are three phases L, a, and b which are in equilibrium with each
other. According to the Gibbs phase rule (note that here the pressure is constant),
such an equilibrium in a binary system can exist only at a specific temperature and
only with specific compositions of the three phases participating in the equilibrium.
It is common practice to show the stability of phases in a single-component
system in different temperature and pressure ranges as shown in Fig. 1.1. In a
binary system case, the stability of the phases is shown in a different temperature
and composition range under constant pressure. Unless mentioned, a binary phase
diagram (shown in Fig. 1.2) is commonly determined at atmospheric pressure.
Note that at different pressure, the binary temperature–composition phase diagram
will be different since the equilibrium transition temperature between different
phases changes with pressure. It is also to be noted that typically, especially in the
case of metals, the vapor region is not shown in the binary phase diagram as it
typically exists at relatively high temperatures under atmospheric pressure.
Finally, it is important to realize that one cannot obtain any information about
kinetics or the morphology of the phase mixture from the phase diagram. The
diagram only gives information about the phases that can be in equilibrium under
certain composition–temperature combinations. Although there are three different
species present in a system, there are times when the phase diagram is presented as
a binary phase diagram. For example, as Fig. 1.3 shows, the MgO–Al2O3 phase
diagram is presented as a binary phase diagram, where MgO and Al 2O3 are
considered as the components. The reason for this is clear. Even though there are
three species (Mg, O, and Al) in the system, there are only two components (MgO
and Al2O3). Only the amounts of these components can be changed independently.
This is called a pseudobinary phase diagram.
Fig. 1.2 Binary phase
diagram of Ag–Cu
1.4 Commonly Used Terms and First Glance at Phase Diagrams 9
In a ternary system (Fig. 1.4), where three elements are mixed, the phase
diagrams take the standard form of a prism which combines an equilateral triangular
base (ABC) with three binary system ‘‘walls’’ (A–B, B–C, and C–A). This
three-dimensional form allows the three independent variables to be specified
(two-component concentrations and temperature). In practice, determining different
sections of the diagram from these kinds of graphical models is difficult and,
therefore, horizontal (isothermal) sections through the prism are used (Fig. 1.4b).
The isothermal section is a triangle at a given temperature, where each corner
represents the pure element, each side represents relevant binary systems, and
areas of different phases can be determined inside the triangle. In addition to the
isothermal section, also vertical sections (isopleths) can be taken from a space
diagram of a given ternary system. We will return to these diagrams and their uses
in Sects. 1.12 and 1.13.
Another commonly used term, as already mentioned, is composition. Composition
can be expressed in terms of mole fraction, atomic fraction or atomic percentage,
and weight fraction or weight percentage. It should be pointed out that in a
binary (not pseudobinary) or multicomponent system, the mole fraction is equal to
the atomic fraction. This can be shown very easily for a system of total 1 mol,
where XA and XB are mole fractions of A and B, respectively. This can be written as
XA þ XB ¼ 1 ð1:21Þ
If nA and nB are the total number of atoms of A and B, respectively, we can
write
XA ¼ nA=No and XB ¼ nB=No ð1:22Þ
where No (=6.022 9 1023 atoms/mole) is the Avogadro number.
Fig. 1.3 Pseudobinary phase
diagram of MgO and Al2O
10 1 Thermodynamics, Phases, and Phase Diagrams
This means that the atomic fraction of A ðNAÞ and B ðNBÞ, with the help of
Eq. 1.21, can be expressed as
NA ¼
nA
nA þ nB
¼
XANo
XANo þ XBNo
¼
XA
XA þ XB
¼ XA ð1:23aÞ
NB ¼
nB
nA þ nB
¼
XBNo
XANo þ XBNo
¼
XB
XA þ XB
¼ XB ð1:23bÞ
Although in the previous example, we considered the one-mole system (which
will be useful in the proceeding section), it can be shown that the mole fraction is
always equal to the atom fraction, even if the system has a total more or less than
one mole of atoms. For example, we consider the system of total x mole, where the
mole of A and B are xA and xB, respectively. This can be written as
xA þ xB ¼ x ð1:24Þ
The mole fraction of A, XA, can be expressed as
XA ¼
xA
x
¼
xA
xA þ xB
ð1:25Þ
Consequently, the atomic fraction of A, NA, can be expressed as
NA ¼
nA
nA þ nB
¼
xANo
xANo þ xBNo
¼
xA
xA þ xB
¼ xA ð1:26Þ
A similar expression can be derived for B.
Fig. 1.4 a Ternary system, and b isothermal section at T = x _C
1.4 Commonly Used Terms and First Glance at Phase Diagrams 11
Concentration can be expressed as molal concentration, that is, ci = number of
moles (g-atoms, g-ions, etc.) of the solute i per 1,000 g of solution, or as volume
concentration, that is, the number of moles per cubic meter (m3). It is to be noted
that the latter definition is valid only at constant temperature. When describing the
composition of the liquid solution, for example, it is expedient to use as the two
other independent variables (in addition to composition regardless of how it is
expressed) temperature and pressure, so that differentiation with respect to temperature
implies constant pressure. Thus, we have
oC
oT
__
¼ _aCS
where a is the thermal expansivity and Cs is the concentration of the species of
interest. The relation above shows that if Cs is chosen as a variable, it will not be
an independent variable [2]. Further, when we consider the solid state, it becomes
evident that in order to use volume concentrations, we should have knowledge
about the molar volume as a function of composition of the phase under investigation.
This is why volume concentrations are not always convenient variables
and, for this reason, will not typically be used later on in the text.
1.5 Spontaneous Change
Entropy is the basic fundamental concept when the direction of natural change is
considered as discussed in Sect. 1.2. Unfortunately, the use of entropy as the
criteria for spontaneous change requires that changes in both the system and
the environment are investigated. As the environment is not always easily defined,
the entropy criterion is not convenient to use in many practical cases. However, if
we concentrate on the system, we may lose some generality but gain a lot in the
sense that the environment no longer needs to be considered. Next, we will look in
greater detail how this can be achieved. Consider a system in thermal equilibrium
with its surroundings at a temperature T. When a change in the system occurs, the
second law of thermodynamics states (the Clausius inequality)
dS _
dq
T
_0 ð1:27Þ
Depending on the conditions under which the process occurs, this inequality
can be developed in two ways.
(i) Heat transfer at constant volume
In the absence of non-expansive work, it is possible to write dqV ¼ dE. This is
because as volume is kept constant and only expansion work is considered, the
work done by or to the system must be zero. Thus, we can write
12 1 Thermodynamics, Phases, and Phase Diagrams
dE ¼ dq
and utilizing Eq. 1.27, the following is obtained
dS _
dE
T
_0 ð1:28Þ
It is to be noted that here the criteria of spontaneity is expressed in terms of
state functions only. Equation 1.28 can be rearranged as
TdS_dE ðV constant; no additional workÞ ð1:29Þ
At either constant internal energy (dE = 0) or constant entropy (dS = 0),
Eq. 1.29 can be expressed as
dSE;V _0 or dES;V _0
The first inequality states that entropy increases in a spontaneous change in a
system with constant volume and constant internal energy. The second inequality
states that given the constant entropy and volume of a system, its internal energy
decreases during spontaneous change. This is, in fact, a statement about entropy
since it states that if the entropy of the system remains unchanged in the transformation,
there must be an increase in the entropy of the environment caused by
the outflow of heat from the system.
(ii) Heat transfer at constant pressure
Again, in the absence of non-expansive work, we may write dqp ¼ dH and
obtain
TdS_dH ðp constant; no additional workÞ ð1:30Þ
At constant enthalpy or entropy, the following inequalities are obtained
dSH;p _0 or dHS;p _0
which can be interpreted in a similar fashion as inequalities concerning heat
transfer at constant V.
Unfortunately, transformations where E and V, H and p, S and V, or S and p are
constant are rare. Far more frequently, transformations take place under conditions
where V and T, or even more typically, p and T, are constant.
Equations 1.29 and 1.30 can be written as
dE _ TdS_0 and dH _ TdS_0 ð1:31Þ
The Helmholtz and Gibbs free energy functions were defined as follows
(Sect. 1.3)
1.5 Spontaneous Change 13
F ¼ E _ TS and G ¼ H _ TS ð1:32Þ
At constant temperature, the differentials of the functions F and G are
ðdFÞT;V¼ dE _ TdS ð1:33Þ
ðdGÞT;p¼ dH _ TdS ð1:34Þ
where the entropies of the phases have been replaced by the temperature of the
system. We get two new inequalities for a spontaneous change with frequently
observed variables
ðdFÞT;V _0 ð1:35Þ
ðdGÞT;p _0 ð1:36Þ
(iii) Expansion work is not the only form of work
How shall the above-derived conditions for spontaneity change if the expansion
work is no longer the only form of work? The second law of thermodynamics
states that dE ¼ dq þ dwtot, where dwtot ¼ dw0 _ pdV is the total work and dw0
takes into account all other forms of work except expansion work. By solving dq,
we get
dq¼ dE _ dw0 þ pdV ð1:37Þ
and utilizing the fact that dq _ TdS_0, we obtain
dE _ TdS _ dw0 þ pdV _0 ð1:38Þ
By utilizing the definition of the Helmholtz free energy, we obtain
ðdFÞT _dw0 _ pdV ¼ dwtot ð1:39Þ
Thus, at constant T, change occurs spontaneously when the change in Helmholtz
energy is smaller than the total amount of work. If the volume is constant
dV = 0, then
ðdFÞT;V _dw0 ð1:40Þ
which is equal to Eq. 1.39 when the expansion work is the only form of work.
14 1 Thermodynamics, Phases, and Phase Diagrams
From the definition of enthalpy (H = E + pV) and from dE ¼ dq þ dw0 _ pdV
under constant pressure, it follows that
ðdHÞp¼ dE þ pdV ¼ dq þ dw0 _ pdV þ pdV ð1:41Þ
which gives
dqp ¼ ðdHÞp_dw0 ð1:42Þ
Combining this with Eq. 1.31 results in
dH _ TdS _ dw0 _0 ðconstant pressureÞ ð1:43Þ
and finally,
ðdGÞT;p _dw0 ð1:44Þ
At constants T and p, the change is spontaneous if the change in Gibbs energy is
less than the additional work done. Equations 1.44 and 1.40 can be stated also as
_DGis the maximum amount of work (other than expansion work) that the system
can release during spontaneous change at constant temperature and pressure. The
value _DFis the maximum amount of total work that the system can release during
spontaneous change at constant temperature.
Given that G = G(T, P, n1, n2,…) in an open system, with ni being the number
of moles of component i, the derivative of the Gibbs energy function yields
dG ¼ _SdT þ Vdp þ
X
i
lidni ð1:45Þ
where li is the chemical potential of component i. At a constant value of the
independent variables P, T, and nj(j 6¼ i), the chemical potential equals the partial
molar Gibbs free energy, (qG/qni)P,T,j6¼i. The chemical potential (partial Gibbs
energy) has an important function analogous to temperature and pressure.
A temperature difference determines the tendency of heat to flow from one body
into another, while a pressure difference, on the other hand, determines the tendency
toward a bodily movement. A chemical potential can be regarded as the
cause of a chemical reaction or the tendency of a substance to diffuse from one
phase to another.
As shown before in Eq. 1.17, the Gibbs free energy can be expressed as
G ¼ H _ TS
1.5 Spontaneous Change 15
where H (J/mole) is the enthalpy, T (Kelvin, K) is the absolute temperature, and
S (J/mole K) is the entropy of the system. Further, H, the total heat content or total
energy of the system, was defined in Eq. 1.16 as
H ¼ E þ pV
where E is the internal energy, P is the pressure, and V is the volume of the system.
In general, the contribution of PV in Eq. 1.16 is very small in the solid and
liquid states if the pressure is not exceptionally high. Therefore, while working
with condensed phases (solid and liquid), the PV term can, in most cases, be
neglected. Hence, the change in internal energy of the system can be approximated
to be equal to its enthalpy
H ¼ E ð1:46Þ
The internal energy of the system consists of the potential and kinetic energies
of the atoms within the system. The kinetic energy of solids and liquids is caused
by the vibration of atoms at their position. In liquids and gases, the translational
and rotational movement of the atoms (or molecules), within the system, provides
an additional contribution to the kinetic energy. Every atom vibrates with different
energy at its position with degrees of freedom in x, y, and z directions with very
high frequency that is temperature dependent. The frequency spectrum starts from
0 and goes up to a maximum value of mD, which is called the Debye frequency. By
utilizing the vibration frequencies, it is possible to calculate the heat capacity of a
given solid. Above a certain temperature (hD, the Debye temperature), all atoms
are essentially vibrating with their corresponding maximum Debye frequency. For
metals at room temperature, they are typically above their Debye temperature,
which makes it possible to use single (maximum) frequency values when considering
the diffusion of atoms, for instance. The average total energy (=3NkT
where k is the Boltzmann constant and N is the number of atoms in a crystal) of
atoms is fixed with respect to a particular temperature. Moreover, the vibration of
any atom depends on the vibration of neighboring atoms because of inter-atomic
bonding. This coupling produces an elastic wave with quantized energy. The
quantum of energy in an elastic wave is called a phonon. For example, sound
waves and thermal vibrations in crystals are phonons. The other part of internal
energy in solids, the potential energy, depends on the inter-atomic bonding
between the atoms. In a single-component system, the potential energy depends on
one type of bonding, but, in a binary or multicomponent system, the potential
energy depends on the type, number, and magnitude of the different bonds between
the atoms within the system. This is explored further in Sect. 1.9 for binary systems
cases. The entropy of a crystal is composed of two terms: thermal entropy
and the configurational entropy. The first part is concerned with the distribution of
energy over the available energy states in the crystal (system) and the latter part
with the distribution of atoms or particles within the crystal (system).
16 1 Thermodynamics, Phases, and Phase Diagrams
Note By utilizing the Gibbs free energy, all forms of work (excluding
expansion work) can be taken into account 2(DG)p,T C w0 = Rlini +
cA + zFU + ___, where the first term is the chemical part, the second is the
surface energy contribution, the third is the electrical component, etc. Thus,
Gibbs energy gives the amount of maximum additional (non-expansion)
work that the system can perform. For all spontaneous processes, the change
in Gibbs energy must be negative. It should also be noted that the temperature
and pressure of the system do not have to be constant during the whole
process. It is adequate that they are the same at the initial and final stages. An
example is an exothermic reaction taking place at temperature T, where the
reaction heat is transferred to the environment at the end of the reaction, thus
making Tinitial equal to Tfinal. This is, of course, a consequence of the fact that
the Gibbs energy is a state function and its value is only dependent on the
initial and final states, not the path between them.
The Helmholtz free energy of a closed system, on the other hand, is a
function of temperature and volume. Helmholtz free energy (F) is maximum
free energy, which can be used to do work at constant volume and temperature
and can be expressed as
F ¼ E _ TS
where E is the internal energy. The main difference between Gibbs free
energy (i.e., the change in energy at constant pressure and temperature) and
Helmholtz free energy (i.e., the change in energy at constant volume and
temperature) is ‘‘PV’’. This comes from the fact that there is need for extra
work to accommodate the volume change. Thus, the Helmholtz free energy
is the maximum amount of any kind of work the system can do and is,
therefore, sometimes called the maximum work function. The change in
Helmholtz free energy must also always be negative for a spontaneous
change.
With the help of the Gibbs free energy function derived above, the equilibrium
state of the system can be investigated. There is the relation between the chemical
potential of components and the total Gibbs energy of the system, as expressed in
Eq. 1.19 (Sect. 1.3) through
Gtot ¼
X
/
X
i
l/
i
n/i
ffi _
The Gibbs energy function can be utilized from the component level to the
system level and back again. Hence, Eq. 1.19 provides the very important connection
between component and system level properties.
1.5 Spontaneous Change 17
Three stable equilibrium states to be considered here are (i) complete or global
thermodynamic equilibrium, (ii) local thermodynamic equilibrium, and (iii) partial
thermodynamic equilibrium. When the system is at complete equilibrium, its
Gibbs free energy (G) function has reached its minimum value
dG ¼0 or lai
¼ lbi
¼ _ _ _ ¼ l/
i ; ði ¼ A; B; C; . . .Þ ð1:47Þ
and then, the system is in mechanical, thermal, and chemical equilibrium with its
surroundings. Consequently, there are no gradients inside the individual phases
and no changes in the macroscopic properties of the system are to be expected.
Local equilibrium, on the other hand, is defined in such a way that the equilibrium
exists only at the interfaces between the different phases present in the system. This
means that the thermodynamic functions are continuous across the interface and the
compositions of the phases right at the interface are very close to those indicated by
the equilibrium phase diagram. This also indicates that there are activity gradients in
the adjoining phases. These gradients, together with the diffusivities, determine the
diffusion of components in the various phases of a joint region.
Partial equilibrium means that the system is in equilibrium only with respect to
certain components. It is generally found that some processes taking place in the
system can be rapid, while others are relatively slow. If the rapid ones occur
quickly enough to fulfill the requirements for stable equilibrium (within the limit
of error) and the slow ones are slow enough that they can be ignored, then it is
quite proper to treat the system as being in equilibrium with respect to the rapid
processes alone [7].
It is also possible that the global energy minimum of the system is not accessible
owing to different restrictions. In such cases, we are dealing with metastable
equilibrium, which can be defined as a local minimum of the total Gibbs energy of
the system. In order to obtain global stable equilibrium, some forms of activation
(e.g., thermal energy) must be brought into the system. It is to be noted that
metastable equilibrium can also be complete, local, or partial; the local metastable
equilibrium concept, in any case, will be used frequently in the following sections.
Very often, one or more interfacial compounds, which should be thermodynamically
stable at a particular temperature, are not observed between two materials
and, then, these interfaces are in local metastable equilibrium. Another situation
commonly encountered occurs in solid/liquid reaction couples, where during the
few first seconds, the solid material is in local metastable equilibrium with the
liquid containing the dissolved atoms, before the intermetallic compound(s) is
formed at the interface. In fact, a principle commonly known as Ostwald’s rule
states that, when a system undergoing reaction proceeds from a less stable state,
the most stable state is not formed directly but rather the next more stable state is
formed, and so on, step by step until (if ever) the most stable is formed. It is a fact
that most materials used in everyday life have not been able to reach their absolute
minimum energy state and are, therefore, in metastable equilibrium. It should be
noted that a system at metastable equilibrium has thermodynamic properties,
which are exactly determined, just as a system at stable equilibrium.
18 1 Thermodynamics, Phases, and Phase Diagrams
1.6 Free Energy and Phase Stability
of Single-Component System
Different phases of a single element can be stable at a different temperature range
under a particular pressure (we consider atmospheric pressure). For example,
below the melting point, a solid phase is stable, whereas above the melting point, a
liquid phase is stable. In general, at a particular temperature, the phase with the
lowest Gibbs free energy will be the stable one. If at a particular temperature, the
free energy of two phases is the same, then both phases are stable at that temperature.
This takes place, for example, at the melting point where the solid and
the liquid phases exist together. This also means that the system is in equilibrium
and there is no driving force for change. To explain the stability of phases at
different temperatures, we need to know the change in their free energies as a
function of temperature. Consequently, (following Eq. 1.18) in order to determine
free energy at a particular temperature, it is necessary to determine the enthalpy
and the entropy at that particular temperature. Both properties can be determined
from the knowledge of specific heat at constant pressure, CP. The specific heat or
specific heat capacity CP (J/mole K) is defined as the amount of heat required to
increase the temperature of a system by one Kelvin under constant pressure.
The absorption or release of heat, dq, in a reversible process, at constant
pressure from the system to the surrounding area is equal to the enthalpy change,
dH, of the system. We can write
dq ¼ dH ð1:48Þ
Further, from the definition of Cp, the equation can be written
Cp ¼
dq
dT
ð1:49Þ
From Eqs. 1.48 and 1.49, follows
dH ¼ CpdT ð1:50Þ
By integrating Eq. 1.50, it can be expressed as
ZH
o
dH ¼
ZT
o
CPdT
HT ¼ Ho þ
ZT
o
CPdT
ð1:51Þ
where HT and Ho are enthalpy at temperature T and 0 K, respectively.
1.6 Free Energy and Phase Stability of Single-Component System 19
The enthalpy at room temperature 298 K is often known, and Eq. 1.51 can be
written as
HT ¼ H298 þ
ZT
298
CPdT ð1:52Þ
Further, from the definition of entropy for a reversible process, we know
dS ¼
dq
T
¼
CPdT
T
ð1:53Þ
By integrating Eq. 1.53, we get
ST ¼ So þ
ZT
o
CP
T
dT ¼
ZT
o
Cp
T
dT ð1:54Þ
where So is the entropy at 0 K. However at 0 K, the entropy of a defect-free pure
element is, by definition, zero (according to the third law of thermodynamics).
Moreover, if the entropy at 298 K is known, then Eq. 1.54 can be written as
ST ¼ S298 þ
ZT
298
CP
T
dT ð1:55Þ
Note We have considered above a pure element with a defect-free structure.
However, it is to be emphasized that it is impossible to obtain a defect-free
structure at temperatures above 0 K. There will always be a certain amount
of point defects, such as vacancies and impurities present in the structure
under the equilibrium condition. The free energy of a phase including the
contribution from defects can be expressed as
Gm ¼ G þ DGd
Gm is the free energy of a single-component material with point defects; G is
the free energy of the defect-free material, and DGd is the free energy change
because of the presence of defects. As will be shown later on, vacancies, for
instance, are always present with a certain equilibrium concentration above
0 K. However, since the concentration of defects, in general, is small
compared to the number of atoms, we can in many cases neglect the contribution
from DGd.
20 1 Thermodynamics, Phases, and Phase Diagrams
In general, the CP values for different phases can be experimentally determined
and are available in the literature. The way that CP typically varies with temperature
is shown in Fig. 1.5a. From the knowledge of CP, it is possible to calculate
H and S at a particular temperature T and consequently determine the variation of
free energy G as a function of temperature. If there are phase transformations
within the temperature range of interest, the enthalpies and entropies of the corresponding
transformation must be added, at the appropriate T, and the integration
must continue with the Cp value of the new phase, to obtain the correct H and S at
the required temperature. The enthalpy of the formation of all pure elements under
atmospheric pressure and with their most stable form at room temperature (298 K)
has been defined to be zero at all temperatures. These are called the standard
enthalpies of formation. And from these, the enthalpy change as a function of
temperature can be determined as HT ¼ RT
298
CPdT. The typical change in enthalpy,
entropy, and free energy is shown in Fig. 1.5b. There are a few important points
that should be noted here. It is clear from Eq. 1.50 that the slope of the enthalpy
curve dH/dT is equal to CP. Since the value of CP always increases with temperature,
the slope of the enthalpy curve will also increase continuously with rising
temperature. Further, from standard thermodynamic relation, we know that
dG = Vdp - SdT. Since transformations at constant pressure are under consideration,
we can write dG = - SdT. Hence, the slope of the free energy curve dG/
dT is equal to -S. Since entropy always increases with temperature, the slope of
the free energy, G, should always decrease with rising temperature.
Now, let us consider the stability of the solid and liquid phases of a metal. To
do this, we will first need to determine the change in free energy with temperature
for both solid and liquid phases separately. From Eqs. 1.17, 1.50, and 1.55, we can
write the expressions for free energy for solid and liquid phases as
GS ¼ HS

ZT
0
CSP
dT _ T
ZT
0
CSP
T
dT ð1:56aÞ
GL ¼ HL

ZT
0
CLP
dT _ T
ZT
0
CLP
T
dT ð1:56bÞ
The superscripts ‘‘S’’ and ‘‘L’’ are denoted for solid and liquid phases,
respectively. In general, the CP of the liquid phase at a particular temperature is
higher than that of the solid phase. The typical variation of CP for solid and liquid
phases is shown in Fig. 1.6a. The corresponding changes in enthalpy and free
energy as a function of temperature of the phases are shown in Fig. 1.6b.
As already discussed, the pV term for both solid and liquid phases is very small
and the enthalpy can be taken to be practically equal to E. Therefore, the Gibbs
energy function can be written as G = E - TS, making the free energy low for a
1.6 Free Energy and Phase Stability of Single-Component System 21
phase with a low internal energy E and/or high entropy S. It is also apparent that at
low temperature, the E term will dominate, whereas at higher temperature, the
term becomes more and more significant. In general, the solid phases have higher
bonding energies compared to those of liquid phases. So the internal energy, i.e.,
enthalpy, of the solid phase is lower than the liquid phase. Further, the entropies of
liquids are typically larger than those of solids. Thus, at higher temperatures, the
liquid phase becomes stable. From Fig. 1.6b, it can be seen, for instance, that
below the melting point the solid phase is stable, whereas above the melting point
the liquid phase is stable. At the melting point, their Gibbs energies are the same,
as discussed in the beginning of this section.
Now, let us turn to consider solid-state transformation between gray tin to white
tin. Gray tin has a diamond crystal structure which is very brittle. White tin, on the
other hand, which is commercially available with a metallic luster has a BCT
(body-centered tetragonal) structure. The Gibbs energy curves for both structures
are shown in Fig. 1.7.
In Fig. 1.7, the molar Gibbs energy of the BCT-Sn has been set at zero for all
temperatures. Thus, the BCT-Sn is the reference state. The Gibbs energies of
different forms of Sn are then compared against this self-chosen reference value.
As can be seen from Fig. 1.7, BCT-Sn should be stable between 13 and 232 _C.
Below 13 _C, Sn with a diamond structure is the most stable form of tin, and above
232 _C, the liquid Sn is the most stable form of tin. It is to be noted that even
though the transition temperature between the diamond and body-centered
tetragonal structures is 13 _C, in practice the transformation requires undercooling
to about -30 _C. This is because at 13 _C, the two crystal structures are in
equilibrium and their Gibbs energies are the same. Thus, the driving force for the
transformation is zero. As temperature decreases, the driving force for the transformation
increases and the kinetics becomes slower. Thus, the optimum conditions
for the transformation are found at -30 _C. This occurrence of the low
Fig. 1.5 a Typical change of CP with temperature. b Arbitrary values of H, S, and G with
temperature. Note that in many metals, the enthalpy is zero at room temperature, 298 K
22 1 Thermodynamics, Phases, and Phase Diagrams
temperature diamond form of Sn should be avoided as it will lead to a phenomenon
called tin pest. Owing to the much larger molar volume of Sn with the
diamond structure (in comparison with the smaller BCT-Sn), the transition fractures
(or even pulverizes) the tin objects going through the transition. One can also
determine the hypothetical melting point of the diamond Sn from the intersection
point of the metastable part of the diamond Gibbs energy curve with that of the
liquid phase. It turns out to be about 160 _C.
Fig. 1.6 a Arbitrary values of specific heat at constant pressure solid and liquid phase. b The
change of enthalpy (HS-enthalpy of solid phase, HL-enthalpy of liquid phase) and free energy
(GS-free energy of solid phase, GL-free energy of liquid phase) of solid and liquid phases with
temperature. L is the latent heat of fusion. Solid line follows the change of enthalpy and free
energy of the system with temperature
Fig. 1.7 The molar Gibbs energies of different phases of Sn as a function of temperature at
1 atm pressure. The figure on the right shows an enlarged part of the figure on the left
1.6 Free Energy and Phase Stability of Single-Component System 23
The increasing importance of the entropy term (-TS) is the reason why in
many metals, we find the phase with a relatively closely packed structure is stable
at a lower temperature, whereas a relatively loosely packed structure is stable at
higher temperature. The reason for this lies in a more loosely packed structure
where there is a higher degree of vibrational freedom. For instance, a-Ti with an
HCP structure is stable at low temperature, whereas b-Ti with a BCC structure is
stable at high temperature.
1.7 Pressure Effect of Single-Component Phase Diagram
Until now, as mentioned earlier, it has been assumed that all transformations occur
under constant, typically atmospheric, pressure. If we consider the Gibbs free
energy at constant temperature but under different pressure, for example, at higher
pressure, then the freedom for vibration of atoms will be decreased in comparison
with the normal pressure. This will result in an increase in free energy, as shown in
Fig. 1.8. We know from the standard thermodynamic relation dG = Vdp - SdT
that the slope of the free energy versus pressure curve is equal to volume V, at
constant temperature. Since the volume of matter generally decreases with
increasing pressure, the slope of the free energy versus pressure curve will be
positive, but it will decrease continuously with increasing pressure. Consequently,
the equilibrium transition temperature from one phase to another will be different
under different pressures and depending on the conditions, the transition temperature
with increasing pressure might increase or decrease. This can be understood
as being based on the Clausius–Clapeyron relation, which can be derived by
considering the equilibrium transition temperature between the a- and c-phases in
an iron system. By using a standard thermodynamic relation, it is possible to write
for the molar Gibbs energy
Fig. 1.8 The variation of
molar Gibbs free energy g
with increase in pressure
24 1 Thermodynamics, Phases, and Phase Diagrams
dga ¼ va
mdP _ sadT ð1:57aÞ
dgc ¼ vc
mdP _ scdT ð1:57bÞ
Since the equilibrium transition between these two phases is under consideration
at equilibrium temperature, one has ga ¼ gc and further, dga ¼ dgc. By
equating Eqs. 1.57a and 1.57b, the following is obtained
oP
oT
__
eq
¼
sc _ sa
vc
m _ va
m
¼
Ds
Dvm
ð1:58Þ
From Eqs. 1.17, we can write
gc ¼ hc _ Tsc ð1:59aÞ
ga ¼ ha _ Tsa ð1:59bÞ
From Eqs. 1.59a and 1.59b and from the consideration of equilibrium transition,
we can write
Dg ¼ gc _ ga ¼ 0 ¼ ðhc _ haÞ _ Tðsc _ saÞ ¼ Dh _ TDs ð1:60Þ
Further, we can write at the transition temperature (Ttr)
Ds ¼
Dh
T
ð1:61Þ
By introducing Eq. 1.61 in Eq. 1.60, we arrive at the following
dP
dT
__
¼
Dh
TDvm
ð1:62Þ
Equation 1.62 is the Clausius–Clapeyron equation, which can be used to calculate,
for example, the TM curve shown in Fig. 1.1.We know that the a-phase has
a BCC structure and the c-phase has an FCC structure. Since FCC is a more closely
packed structure, we can write for the transition a ? c, Dv ¼ vc
m _ va
m\0. On the
other hand, we have seen previously that the enthalpy of a phase which is stable at
higher temperature is higher (less negative) than that of a phase which is stable at
lower temperature. This becomes Dh ¼ hc _ ha[0. From Eq. 1.63, it follows that
dP
dT
__
eq\0 for the equilibrium transformation from the a-phase to the c-phase. This
is the reason why the equilibrium transition temperature decreases with increasing
pressure. Let us consider the case of the equilibrium transition between the c- to
d-phase. Since the c-phase has an FCC structure, whereas the d-phase has a BCC
structure, it is possible to write for the transition c ? d Dv ¼ vd
m_ vc
m[0.
1.7 Pressure Effect of Single-Component Phase Diagram 25
Further, from our previous explanation, we can write Dh ¼ hd _ hc[0. So
dP
dT
__
eq[0 for this transition and equilibrium transition temperature increases with
increasing pressure.
1.8 Free Energy and Stability of Phases in a Binary System
In the previous sections, we have considered mainly single-component systems
(i.e., pure elements). It is common knowledge that most materials in nature consist
of several phases and that these phases themselves are never pure elements. In fact,
based on the second law of thermodynamics, a pure substance exists only in our
minds and represents a limiting state, which we may asymptotically approach but
never actually obtain. Thus, the thermodynamic description of multicomponent
systems is of great importance from the theoretical as well as from the practical
point of view. In the treatment of multicomponent open systems, the most common
process considered in defining the thermodynamic functions for a solution is called
the mixing process, which Guggenheim defines as [2]:
The mixing process is the change in state experienced by the system when appropriate
amounts of the ‘pure’ components in their reference states are mixed together forming a
homogeneous solution brought to the same temperature and pressure as the initial state.
It is to be noted that although the mixing process is strongly influenced by
interaction forces between atoms and molecules (i.e., Dh), the fundamental cause
behind mixing is the entropy (Ds) change of the system.
For our analysis, we shall consider a system with a total of one mole of atoms,
where XA is the mole fraction of element A and XB is the mole fraction of element
B. This translates into
XA þ XB ¼ 1 ð1:63Þ
We define the free energies of pure elements A as GA and that of B as GB at a
particular temperature. The total molar free energy g0 of a purely mechanical
mixture can be written as
g0 ¼ XAgA þ XBgB ð1:64Þ
Now, if we allow interdiffusion to taken place between the elements A and B,
there will be change in the free energy because of mixing, gmix. Consequently, the
total free energy of the system after mixing can be written as
g ¼ g0 þ Dgmix ð1:65Þ
26 1 Thermodynamics, Phases, and Phase Diagrams
From Eqs. 1.17 and 1.65, we find the expression for the free energy change
because of mixing as
DGmix ¼ g _ g0 ¼ h _ Ts _ h0 þ Ts0 ¼ ðh _ h0Þ _ Tðs _ s0Þ
Dgmix ¼ Dhmix _ TDsmix
ð1:66Þ
where ho and h are the total enthalpy of the system before and after mixing. The
values so and s are the entropies of the system before and after mixing. The value
Dhmix ¼ ðh _ h0Þ is the change in enthalpy, and Dsmix ¼ ðs _ s0Þ is the change in
entropy due to mixing. The enthalpy of mixing can be zero, negative, or positive
depending on the system; the entropy of mixing, on the other hand, is always
positive. We shall first briefly discuss the mixing process in general and then look
a little closer at where the different terms in Eq. 1.66 arise.
1.8.1 Change in Free Energy in an Ideal System
Note that in the case of an ideal solution, Dhmix ¼ 0 and the free energy of the
system can be written as
g ¼ g0 þ Dgmix ¼ g0 _ TDsmix ð1:67Þ
The change in free energy with composition is shown in Fig. 1.9a at one
particular temperature T. The straight dotted line represents the total free energy
ðg0 ¼ XAgA þ XBgBÞ of the elements A and B before any mixing (i.e., a purely
mechanical mixture of A and B). The solid curved line represents the free energy
of the system after mixing ðg ¼ g0 _ TDsmixÞ. Further, the change in free energy
with composition at higher temperature, T1, is shown in Fig. 1.9b. The change in
free energy caused by mixing in an ideal solution Dgmix ¼ _TDsmix is naturally
larger at the higher temperature.
Entropy of mixing (Dsmix)
The entropy of mixing originates from two different contributions, thermal and
configurational. If we consider that there are no volume and enthalpy changes
caused by mixing, then the only contribution to the entropy will configurational.
Configurational entropy comes from the possibility of arranging the atoms A and B
in different ways for a particular macrostate. Following statistical thermodynamics,
the configurational entropy can be expressed as
S ¼ k ln w ð1:68Þ
where w is a thermodynamic probability, a kind of measure of randomness. This
means that the molar entropy of mixing can be written as
1.8 Free Energy and Stability of Phases in a Binary System 27
Dsmix ¼ s _ s0 ¼ k lnw _ k ln 1 ¼ k ln w ð1:69Þ
where, as explained before, s0 is the entropy before mixing and s is the entropy
after mixing. Since in the case of a pure element, there is only one way by which
atoms can be arranged (if vacancies are neglected), we can write w = 1. If we
consider the random solid solution, then the number of different ways by which
atoms A and B can be arranged is

ðnA þ nBÞ!
nA!nB!
ð1:70Þ
where nA and nB are the total number of atoms of A and B, respectively.
According to Stirling’s approximation,
ln N! ¼ N ln N _ N ð1:71Þ
Following Stirling’s approximation, Eq. 1.69 can be written as
DSmix ¼ k ln w
¼ ½ðnA þ nBÞ lnðnA þ nBÞ _ ðnA þ nBÞ_ _ ½nA ln nA _ nA_ _ ½nB ln nB _ nB_
ð1:72Þ
Fig. 1.9 Free energy versus
composition diagram in an
ideal case at a low
temperature, T, and b high
temperature, T1
28 1 Thermodynamics, Phases, and Phase Diagrams
From the definition of the mole fraction, we can write
nA ¼ XAN0; nB ¼ XBN0; XA þ XB ¼ 1 ð1:73Þ
Substituting Eq. 1.73 in Eq. 1.72, we get
Dsmix ¼ _kN0½XA ln XA þ XB ln XB_
¼ _R½XA ln XA þ XB ln XB_
ð1:74Þ
It is, therefore, clearly apparent from Eq. 1.74 that the entropy of mixing is
always positive and will vary, as shown in Fig. 1.10. It can also be seen that the
entropy of mixing reaches its maximum at XB = 0.5. To find the slope at different
compositions, we can differentiate Eq. 1.74 (note that XA + XB = 1).
dDsmix
dXB
¼ _R _lnð1 _ XBÞ _ ð1 _ XBÞ
1
ð1 _ XBÞ
þ ln XB þ XB
1
XB
_
¼ _R ln
XB
ð1 _ XBÞ
ð1:75Þ
So the slope at XB = 0.5 is equal to zero, whereas the slope at XB = 0 or 1 is
infinity. Thus, it is evident from the above discussion why a pure substance is just a
limit which we can approach but never achieve—as stated at the beginning of this
section. As the slope goes to infinity at XB = 0, it states that in order to remove the
last B impurity from A, an infinite amount of energy must be used.
Fig. 1.10 The change in
entropy of mixing with the
change in composition
1.8 Free Energy and Stability of Phases in a Binary System 29
1.8.2 Change in Free Energy in a System with Exothermic
Transformation
As discussed above, the enthalpy of mixing in an exothermic transformation is
negative. The free energy of mixing for such a case at different compositions at a
particular temperature T is shown in Fig. 1.11a. The free energy curves at higher
temperature T1 are shown in Fig. 1.11b. At higher temperature, the TDsmix term
will be higher, making Dgmix higher after mixing. In consequence, the total free
energy of the system will change far more drastically with composition compared
to at lower temperature.
1.8.3 Change in Free Energy in a System with Endothermic
Transformation
In endothermic transformation, the enthalpy of mixing Dhmix[0. So, if the
temperature under consideration is reasonably low, the negative contribution to the
Gibbs energy of mixing from TDsmix may be smaller than the positive contribution
from the enthalpy of mixing Dhmix within a certain composition range. In that case,
Fig. 1.11 Free energy versus
composition diagram for a
system with exothermic
transformation at a low
temperature, T, and b high
temperature, T1
30 1 Thermodynamics, Phases, and Phase Diagrams
the free energy of mixing will be positive at a certain composition range and the
total free energy change can vary, for example, as shown in Fig. 1.12 resulting in a
formation of a miscibility gap. However, at higher temperature, T1 at all compositions
Dhmix will be smaller than TDsmix and the free energy of mixing is always
negative, as can be seen in Fig. 1.12b. The next section considers the origin of the
enthalpy of mixing.
Note We have considered a very simplified model to establish the relation of
free energy after mixing in a binary system. We have not considered the
elastic strain that could play an important role. In some systems, where the
size of the atoms is very similar, this factor can be disregarded. However, in
some systems because of a large difference in atomic size, the elastic strain
might play a significant role. Calculations, however, become extremely
complicated if we are to consider the effect of elastic strain and so it is left
outside the scope of this book. It must, nevertheless, be remembered that the
elastic energy (as other forms of work) can be incorporated into the Gibbs
free energy of the system.
Fig. 1.12 Free energy versus
composition diagram in a
system with endothermic
transformation at a low
temperature, T, and b high
temperature, T1
1.8 Free Energy and Stability of Phases in a Binary System 31
1.9 Thermodynamics of Solutions and Phase Diagrams
In the following section, the thermodynamic background necessary for understanding
the phase diagrams introduced briefly in Sect. 1.4 is discussed. After that
the binary and ternary phase diagrams are discussed in greater detail.
1.9.1 The Chemical Potential and Activity in a Binary Solid
Solution
For any heterogeneous system at equilibrium, the chemical potential of a component
i has the same value in all phases of the system, where the component has
accessibility. A general problem for dealing with solutions thermodynamically can
be regarded as one of properly determining the chemical potentials of the components.
Usually, the treatment utilizes the activity function introduced by Lewis
and Randall [8]. The value of the treatment lies in its close relation to composition;
with appropriate choice of reference state, the activity approaches the mole fraction
as the mole fraction approaches unity. Most commonly in the thermodynamics
of solutions, it is not the activity which is used, but rather the activity coefficient
which is defined as the ratio of the activity ai to the mole fraction Xi
ci ¼
ai
Xi
ð1:76Þ
In terms of the chemical potential, the activity can be expressed
lj
i _ loi
¼ RT ln aj
i ¼ RT ln X j
i þ RT ln c j
i ð1:77Þ
where loi
is the chemical potential of pure i in the reference or standard state, lj
i
the chemical potential of i in phase j, aj
i the activity of component i in phase j,
R the gas constant, T the temperature, and (i = A, B,…; j = a, b,…). In the
limiting case of ideal solutions, where the enthalpy (Dh = 0) and volume change
(Dm = 0) of mixing are zero and the only contribution to Gibbs free energy of
mixing arises from the configurational entropy term
Dsm ¼
X
i¼A
Xi ln Xi ð1:78Þ
the activity coefficient in Eq. 1.77 is unity and the activity of the component equals
its mole fraction (i.e., Raoultian behavior, see discussion below). If the equality is
valid for all compositions, the solution is called perfect. Thus, the activity coefficient
represents deviation of the real solutions from this limiting behavior. The
use of activity coefficient instead of activity in Eq. 1.77 clearly indicates the
32 1 Thermodynamics, Phases, and Phase Diagrams
excess energy term RT ln lj
i to be responsible for the non-ideal behavior. This
issue is addressed in more detail in the next section.
As only relative values of thermodynamic functions can be determined, an
agreed reference state has to be established for each element or species in order to
make thermodynamic treatment quantitative (see Fig. 1.7 and related discussion).
In principle, the choice of the reference state is arbitrary as long as the chosen state
is used consequently throughout the analysis. The chosen state is then defined to be
zero and all other possible states of the element are compared against the reference
state to obtain their relative stabilities. It should be noted that there are some
uncertainties related to the usage of reference states in the literature.
1.9.2 Free Energy of Solutions
In Sect. 1.9, we briefly discussed the so-called mixing process as well as the binary
solution phases. In the beginning of this section, we also introduced the concept of
activity, which describes the deviation of the behavior of a solution from ideal
behavior. A statistical approach can be used to provide more insight into the
properties of the phases. The simplest model is one in which the total energy of the
solution is given by a summation of interactions between the nearest neighbor
atoms. If we have a binary system with two types of atoms (A and B), there will be
three interaction energy terms. These are the energy of the A–A pairs, that of the
B–B pairs, and that of the A–B pairs. Here, we assume that the total energy of the
solution arises from the interactions between the nearest neighbors. The binding
energy may be defined by considering that the change in energy as the distance
between a pair of atoms is decreased from infinity to an equilibrium separation.
The change in energy during this process is the binding energy, which for a pair of
A atoms is given as -2eAA, for B atoms as -2eBB, and so forth. Thus, the bond
energies are negative quantities.
The simplest model for real solution phases based on the above-defined nearest
neighbor interaction approach is the so-called regular solution model. It is based
on the following assumptions: (i) Mixing among accessible lattice spaces is
completely random pi ¼ Ni
N ¼ Xi, (ii) atoms interact only with their nearest
neighbors, (iii) the bond energy between dissimilar atoms eij is independent of
composition and temperature, and (iv) there is no change in volume upon mixing.
Let us examine a solution which is formed by two metals (A and B) with an
identical crystal structure. Let us further assume that the system as defined above is
in equilibrium with its surroundings. We will presume that metal A has NA atoms
and metal B has NB atoms and that eAA and eBB are the bond energies of the AA
and BB atom pairs according to the assumptions (ii) and (iii). Then, the configuration
energies of pure metals, with the coordination number z, are E0A
¼
zN0A
eAA=2 and E0B
¼ zN0B
eBB=2, when the atoms are at rest at their equilibrium
lattice points. The number two in the nominator in the above equations is
1.9 Thermodynamics of Solutions and Phase Diagrams 33
introduced in order to prevent calculating the A–A and B–B interactions twice.
The formation of the solution phase starts by removing one atom from each metal
and transferring them to an infinite distance from the metal (and each other). The
work associated with this process is -z(eAA + eBB). By returning an atom A to
metal B and an atom B to metal A, the pure metals are transformed into solutions
with infinite dilution. The total number of A–B bonds with bond energy eAB in
these solutions is 2z. The energy associated with this mixing process (per interchange)
can be described as
dE ¼ 2z eAB _
1
2
ðeAA þ eBBÞ
_
¼ _2zIAB ð1:79Þ
where IAB (per bond) is the interchange energy.
As the energy change associated with the formation of a mixture with a number
of NAB bonds is DE ¼ NABIAB, the internal energy of the solution phase is

z
2
ffi _
NAðeAAÞ þ
z
2
ffi _
NBðeBBÞ þ NABIAB ð1:80Þ
The next step is to identify what the most probable number of AB bonds (NAB)
is with the nominal composition of XoB
. This problem can be resolved by utilizing
the first assumption of the regular solution model, i.e., that the mixing among the
lattice sites is completely random. This means that the probability that an atom A
is in position 1 equals pA(1) = (NA/N) and that atom B is in position 2 equals
pB(2) = (NB/N). However, since pA(1)pB(2) = (NANB/N2) and because
pA(1)pB(2) = pA(2)pB(1), we get
pð1;2Þ
AB ¼
2NANB
N2 ð1:81Þ
In the solution phase, we have total of 0.5zN adjacent lattice site pairs and
therefore,
_pAB ¼ pð1;2Þ
AB ¼
2NANB
N2
zN
2
__
¼z
NANB
NA þ NB
__
ð1:82Þ
Thus, the internal energy of the solution phase can be written as

z
2
ffi _
NAðeAAÞ þ
z
2
ffi _
NBðeBBÞ þ z
NANB
NA þ NB
__
IAB ð1:83Þ
By assuming that the chemical potentials of pure metals A and B can be
approximated as l0
A ffi 12
ZðeAAÞ and l0
B ffi 12
ZðeBBÞ and utilizing the definition of
the chemical potential, the Gibbs energy of the regular solution phase becomes
34 1 Thermodynamics, Phases, and Phase Diagrams
G ¼ NAlo
A þ NBlo
B þ kTX
i
Ni ln
Ni
NA þ NB
__
þz
NANB
NA þ NB
__
IAB ð1:84Þ
from which the molar Gibbs energy is obtained as
g ¼ Xo
AlA þ Xo
BlB þ RTX
i
Xi ln Xi þ LABXAXB ð1:85Þ
where LAB (=zNIAB) is the molar interaction energy, i.e., the interaction
parameter.
The relationship between activity and the interaction parameter can be written
as
ai ¼ ciXi ¼ Xi exp
Lijð1 _ XiÞ2
RT
"#
ð1:86Þ
Consequently, the sign of the interaction parameter determines whether the
formation of mixture is favored or hindered. When |eAA + eBB|\|2eAB| and the
interaction parameter is negative (remember that the bonding energies are negative),
the solution will have a larger than random probability of bonds between
unlike atoms, and thus, mixing or compound formation is favored, as in Fig. 1.13.
The converse is true when the interaction parameter is positive (|eAA + eBB|
[ |2eAB|) since atoms then prefer to be neighbors to their own kind and form
clusters. From Eq. 1.86, it is also seen how the activity coefficient depends on both
the sign and magnitude of the interaction parameter. Activity is eventually
determined by the interactions between different types of atoms in the solution
phase. It is also helpful to notice that the excess term in Eq. 1.86 LABXAXB can be
identified with the enthalpy of mixing in Eq. 1.66.
Justification of Eq. 1.86
Deviations from ideal behavior are commonly expressed in the form of excess
functions. The excess Gibbs energy of mixing can be expressed as
Dgxs
mix ¼ Dhxs
mix _ TDsxs
mix. This is the extra energy of mixing resulting from
the formation of a real instead of an ideal solution. In a regular solution model,
the entropy of mixing is defined to be the same as that of an ideal solution.
Fig. 1.13 Schematic
presentation of the effect of
the sign of the enthalpy of
mixing and that of the
interaction parameter on the
formation of a solution phase
1.9 Thermodynamics of Solutions and Phase Diagrams 35
Thus, the excess entropy of mixing Dsxs
mix is zero and Dgxs
mix ¼ Dhxs
mix in the
case of regular solutions. Therefore, the expression LABXAXB in Eq. 1.86 can
be equated to the excess enthalpy of mixing. It is to be noted that the
expression LABXAXB is also the simplest possible expression for the excess
energy of mixing, as it is required that the excess energy of mixing goes to
zero when XA = 0 or XA = 1. There is a standard relation between the partial
molar properties of a component and the total properties of a phase that can be
expressed for the excess enthalpy of mixing and the partial excess enthalpy of
the mixing of component B as
Dhxs
B;mix ¼ Dhxs
mix þ XA
dDhxs
mix
dXB
We have the above-defined Dhxs
mix ¼ LABXAXB ¼ LABð1 _ XBÞXB ¼
LABðXB _ X2B
Þ
Thus, dDhxs
mix
dXB
¼ LABð1 _ 2XBÞ
and we obtain
Dhxs
B;mix ¼ LABXAXB þ XALABð1 _ 2XBÞ
¼ XALABðXB þ 1 _ 2XBÞ ¼ XALABð1 _ XBÞ ¼ LABX2
A
The excess enthalpy of the mixing of component B can also be equated (in
the case of regular solution model) to RT ln cB as discussed above.
Thus, we can write as follows:
RT ln cB ¼ LABX2
A
cB ¼ exp
LABð1 _ XBÞ2
RT
"#
As aB ¼ XBcB, we finally obtain
aB ¼ XBcB ¼ XB exp
LABð1 _ XBÞ2
RT
"#
36 1 Thermodynamics, Phases, and Phase Diagrams
In Fig. 1.14, the effect of the sign and magnitude of the interaction parameter
on the formation of a solution phase is shown [9]. When there is no preferred
interaction between the atoms in the system, the interaction parameter LAB = 0
(eAA + eBB = 2eAB), the integral heat of mixing is zero, and the free energy of
mixing is given by curve I. As the interaction parameter is made more positive, it
can be seen how the enthalpy of mixing becomes more positive and the free energy
of mixing becomes less negative. When a certain magnitude of positive interaction
is reached, it can be seen that the system is about to enter the state where the
solution phase becomes unstable. When LAB is increased to even more positive
values, one can see how the Gibbs energy curve changes its sign of curvature at the
middle region and the so-called miscibility gap is formed. This is associated with
the formation of two separate phase regions—one rich in A and another rich in B.
Fig. 1.14 The effect of interaction parameter on the stability of a solution phase [9]
1.9 Thermodynamics of Solutions and Phase Diagrams 37
It was shown that one may find a situation where the interaction parameter is
zero and there is no net interaction between A and B (i.e., eAA + eBB = 2eAB).
This type of behavior is associated with the above-defined ideal systems and is
described by Raoult’s law. The Raoultian solution was shown above to be the one
where the activity coefficient (Eq. 1.76) is unity and the activity of the component
equals its mole fraction. Such behavior is shown in Fig. 1.15. If Raoult’s law is
obeyed by the solution phase through the whole composition range, the solution is
called perfect. This type of solution does not exist in reality, but it does provide a
convenient reference state to which the behavior of real solutions can be compared.
In Fig. 1.15, another limiting law (Henry’s law) is also shown. This limiting
law can be understood by utilizing the regular solution model and Eq. 1.86. When
one approaches the limit where Xi ? 0, i.e., the solution becomes dilute, it can be
seen from Eq. 1.87 that the activity coefficient becomes concentration independent
as
1ci ¼ exp
Lij
RT
_
ð1:87Þ
This defines the Henry’s law line seen in Fig. 1.15. The limiting laws shown in
Fig. 1.15 provide the reference states to which real solutions can be compared. As
we approach pure substance (Xi ? 1), the solution behavior necessarily approaches
Raoultian behavior no matter how ‘‘non-ideally’’ it otherwise behaves. This is
true also for Henry’s law, as all solutions approach it as the solution becomes
dilute enough. It is also to be noted that if the solute follows Henry’s law, then the
solvent necessarily follows Raoult’s law. Furthermore, whereas perfect solutions
do not exist, ideal solution behavior is commonly encountered in practice within
restricted composition limits.
Fig. 1.15 Raoult’s and Henry’s laws in a binary solution
38 1 Thermodynamics, Phases, and Phase Diagrams
Historically, activity measurements have been carried out mainly by measuring
the changes in the partial pressure of a given substance upon alloying with respect
to the values of the pure component. As this type of approach also gives an easily
accessible alternative route to derive the above-defined limiting laws, we shall
briefly consider Raoult’s and Henry’s laws from this point of view. Consider a
pure liquid A in a closed vessel (initially evacuated) at temperature T. It will
spontaneously evaporate until the pressure in the vessel is equal to the saturated
vapor pressure of liquid A (po
A) at temperature T. At this point, the rate of evaporation
reðAÞ and the rate of condensation rcðAÞ are equal. In order for an atom to
escape the surface of the liquid and enter the gas phase, it must overcome the
attractive forces exerted on it by its neighbors (i.e., overcome the activation energy
E barrier). The magnitude of E determines the intrinsic evaporation rate. The
condensation rate is proportional to the number of A atoms in the vapor phase,
which strike (and stick) the liquid surface in unit time. For a fixed temperature, the
condensation rate is proportional to the pressure of the vapor rcðAÞ ¼ kpo
A which is
equal to reðAÞ at equilibrium. A similar situation holds for a liquid B. If we now add
a small amount of liquid B to liquid A, what happens? If the mole fraction of A in
the resulting binary mixture is XA and assuming that the atomic diameters of A and
B are comparable and there is no surface excess, the fraction of the surface area
occupied by A atoms is XA. It is a natural assumption that atom A can evaporate
only from a site where it is present and, therefore, reðAÞ is decreased by a factor of
XA and the equilibrium pressure exerted by A is decreased from po
A to pA
reðAÞXA ¼ kpA ð1:88Þ
and by utilizing the above-defined equality between evaporation rate and equilibrium
pressure, we obtain
pA ¼ XApo
A ð1:89Þ
which is Raoult’s law. A similar equation holds for component B. The law states
that the vapor pressure exerted by a component i in a solution is equal to the
product of the mole fraction of i in the solution and the vapor pressure of i at the
temperature of the solution.
While deriving Raoult’s law, it was assumed that there is no change in the
intrinsic evaporation rates. This requires that the magnitudes of the A–A, B–B, and
A–B interactions are balanced so that the depth of the potential energy well of an
atom at the surface site is independent of the types of atoms surrounding it (see
discussion below). If we take that the A–B interaction is much stronger than that
between identical atoms and consider a solution of A in B which is sufficiently
dilute in such a way that every A atom is surrounded only by B atoms, in this case,
the activation energy for an A atom to evaporate from the surface is higher than
without B and thus, the intrinsic evaporation rate will be smaller ðr0e
ðAÞ\reðAÞÞ and
equilibrium occurs when
1.9 Thermodynamics of Solutions and Phase Diagrams 39
r0e
ðAÞXA ¼kpA ð1:90Þ
which results in
pA ¼
r0e
ðAÞ
reðAÞ
XApo
A ð1:91Þ
and as ðr0e
ðAÞ\reðAÞÞ, pA is a smaller quantity than that in Eqs. 1.88, 1.91 can be
written as
pA ¼ k0
AXA ð1:92Þ
If the XA of the solution is increased, it becomes more probable that not all of
the A atoms at the surface are surrounded only by the B atoms. This will have an
effect on the activation energy (depth of the potential energy well), and thus, after
a certain critical value of XA, the intrinsic evaporation rate becomes composition
dependent and Eq. 1.93 no longer holds. Equation 1.93 is, of course, Henry’s law
(a similar equation holds for B atoms also). Note also that Raoultian and Henrian
activity coefficients have different reference states. More information about the use
of these standard states as well as changing between them can be found, for
example, from Refs. [10, 11].
1.10 Lever Rule and the Common Tangent Construction
When two materials, especially metals, are mixed together, they either form a
homogeneous solution or separate into a mixture of phases, as already discussed.
Let us consider an alloy X in Fig. 1.16 in the binary system A–B to separate into a
mixture of two phases a and b (under a particular temperature and pressure).
We shall assume that there are N atoms of alloy X and that the fraction of atoms
in the a-phase is (1-x) and in the b-phase is x. The number of B atoms in alloy X is
nX
B, the number of B atoms in the a-phase is na
B, and the number of B atoms in the
b-phase is nb
B. We can change these to atomic fractions by dividing by the total
number of atoms N to get
XB ¼
nx
B
N
XaB
¼
na
B
Nð1 _ xÞ
Xb

nb
B
Nx
ð1:93Þ
40 1 Thermodynamics, Phases, and Phase Diagrams
Since nX
B ¼ na
B þ nb
B, then
NXB ¼ NXaB
ð1 _ xÞ þ NXb
Bx ð1:94Þ
where

XB _ XaB
Xb
B _ XaB
¼
m
mþn
ð1:95Þ
and
1_x¼
Xb
B _ XB
Xb
B _ XaB
¼
n
mþn
ð1:96Þ
and
x
1_x
¼
m
n
ð1:97Þ
Equation 1.97 is called the lever rule, which enables us to calculate the relative
amounts of phases in a phase mixture in terms of the alloy composition and the
phases into which it separates. The free energy of a phase mixture can also be
determined by using the lever rule. If alloy X separates into phases a and b, the free
energy of an alloy will be unchanged by the separation. The free energy of alloy
X is, therefore, equal to the sum of the free energies of the a- and b-phases. Since
alloy X consists of an amount of the a-phase equal to Xb
B_XB
Xb
B_XaB

__
and similarly
XB_XaB
Xb
B_XaB

__
of the b-phase, the molar free energy of alloy X will be
g¼a
Xb
B _ XB
Xb
B _ XaB
!
þb
XB _ XaB
Xb
B _ XaB
!
ð1:98Þ
Fig. 1.16 The lever rule
1.10 Lever Rule and the Common Tangent Construction 41
where a and b represent the free energies of the a- and b-phases at the given
temperature and pressure, as seen from Fig. 1.17. We can further rearrange
Eq. 1.98 in order to obtain the free energy of alloy X
g¼a
ðXb
B _ XaB
Þ _ ðXB _ XaB
Þ
Xb
B _ XaB
"#
þb
XB _ XaB
Xb
B _ XaB
!
¼ a þ ðb _ aÞ
XB _ XaB
Xb
B _ XaB
!
¼aþc
ð1:99Þ
Hence, alloy X which separates into two phases of composition XaB
and Xb
B with
the free energies a and b has a free energy given by the point x on the straight line
connecting a and b (the common tangent). This is depicted in Fig. 1.17.
As an example of the use of common tangent construction, the calculation of
two-phase equilibrium is presented in Fig. 1.18. The condition for chemical
equilibrium is that the chemical potentials of the components are equal in the
phases that are in equilibrium. In the beginning, the a-phase with composition a1 is
contacted with the b-phase with a composition b1. As seen from Fig. 1.18 (at the
moment in question), the chemical potential (partial molar Gibbs energy) of
component A in the a-phase is la1
A and in the b-phase lb1
A , whereas that of component
B in the a-phase is la1
A and in the b-phase it is lb1
A , which are hardly equal.
Thus, there is a driving force D1lBðla1
B _ lb1
B Þ which drives the diffusion of the B
atoms to the a-phase (from composition b1 to a1) and D1lAðlb1
A _ la1
A Þ driving the
A atoms in the opposite direction. As the diffusion proceeds, the driving force for
diffusion gradually decreases (D2lB and D2lA) and vanishes when the chemical
potentials of the components (A and B) become equal in both phases. This takes
place when the two Gibbs energy curves for the a- and b-phases have a common
tangent and the equilibrium has been established.
Fig. 1.17 The common
tangent construction
42 1 Thermodynamics, Phases, and Phase Diagrams
1.11 The Gibbs Phase Rule
The Gibbs phase rule explains the number of phases that will be present in a
system in equilibrium and is expressed as
F þ P ¼ C þ N ð1:100Þ
where F is the number of degrees of freedom (always C 0), P is the number of
phases (liquid phase, a-phase, b-phase), C is the number of components, and
N corresponds to the non-compositional variable. In our case, there are two noncompositional
variables present, temperature and pressure. This means that
Eq. 1.100 can be written as
F þ P ¼ C þ 2 ð1:101Þ
Now for the purpose of explaining the stability of the phases in equilibrium, a
single-component system that is presented in Fig. 1.19a is considered. First, let us
examine the single-phase, solid, liquid or gas phase, region which can be written in
this region C = 1, P = 1. Also from Eq. 1.101, we can write F = 2. This means
that in this region, there are two degrees of freedom, temperature and pressure.
Accordingly, temperature and pressure can be varied independently within the
region. Consequently, to determine the state of the phase in a single-phase region,
Fig. 1.18 Use of the common tangent construction to determine the phase equilibria
1.11 The Gibbs Phase Rule 43
Fig. 1.19 a single-component phase diagram b binary phase diagram
both temperature and pressure must be fixed. Next, we consider the phase
boundary, along which the two phases are in equilibrium. For this can be written
C = 1, P = 2. Further, from Eq. 1.100, we can write F = 1, meaning that there is
only one degree of freedom and only temperature or pressure can be varied along
that tie-line. The other variable is automatically fixed for a particular temperature
or pressure. Let us further consider the invariant point A, where all three phases
can exist together. At this point, C = 1, P = 3. Therefore, following Eq. 1.100,
F = 0, meaning that there are no degrees of freedom at that point and all three
phases can exist only at one particular pressure and temperature.
Let us further consider the binary phase diagram shown in Fig. 1.19b. Since
binary phase diagrams are measured at constant (in general, atmospheric) pressure,
there is only one non-compositional variable present, which is temperature.
Equation 1.100 can thus be written as
F þ P ¼ C þ 1 ð1:102Þ
Now, if we consider the single-phase region, then we can write C = 2, P = 1.
Following Eq. 1.102, we find F = 2 in the single-phase a-region. This means that
to determine the state of an alloy inside a single-phase region, both temperature and
composition must be fixed. Further, if we consider a two-phase region such as
(L + a), then we can write P = 2, C = 2. So following Eq. 1.102, we find the
number of degrees of freedom F = 1. To determine the state of an alloy inside this
region, we need to fix only one variable, either T, XA, or XB. Since, if we fix any one
of these variables, other variables will be fixed automatically. Take, for instance, T2
where the composition of liquid and a-phases have fixed values. Next, if we consider
the eutectic point E, where the three phases, a, b, and liquid exist together, we
can write P = 2, C = 2. Following Eq. 1.103, we find F = 0, meaning that there
are no degrees of freedom and T, XA, and XB are all fixed at this point.
The phase rule is a convenient tool to check that experimentally determined
phase diagrams are correct. With its help, it is possible to point out anomalies in
phase diagrams and to offer corrections.
44 1 Thermodynamics, Phases, and Phase Diagrams
Let us briefly look at one example. In Fig. 1.20, there is a hypothetical binary
A–B phase diagram. It contains four errors. Let us next look what they are and
produce two versions of the corrected diagram.
Error 1 Two-phase region in a binary diagram, thus F = 1 (pressure is fixed).
If the temperature is fixed, the compositions are unambiguously determined. In the
diagram, this is not so. If we choose temperature conveniently, the tie-line enters
the single-phase region and returns back to the two-phase region. Correction
remove the bend from the liquidus.
Error 2 Pure element, thus one component F = 1 + 1 - P = 2 - P. At the
melting point, there are two phases in equilibrium F = 2 - 2 = 0. Hence, phase
transformation for a pure element takes place at one particular temperature.
Correction: Liquidus and solidus curves must meet at the same point.
Error 3 Eutectic line represents three-phase equilibrium, thus F = 0. Temperature
must be constant. Correction Eutectic line must be horizontal.
Error 4 There are four phases in equilibrium at the eutectic isotherm, and thus,
F = -1. The number of degrees of freedom must not be negative and, therefore,
four-phase equilibrium in a binary system, with constant pressure, is not possible.
Correction 1 c-phase must be removed to obtain the necessary degree of freedom
to make F = 0. Correction 2 If there is first a two-phase a + b region below the
eutectic isotherm and after that a peritectoid reaction takes place, we can preserve
the c-phase.
The two versions of the corrected phase diagrams are shown in Fig. 1.21.
1.12 Correlation of Free Energy and Phase Diagram
in Binary Systems
We shall first start with the simplest possible binary system, where elements A and
B are completely miscible in both solid and liquid state. This requires that the
elements A and B have (i) the same crystal structure, (ii) their size difference is
less than 15 %, and (iii) their electronegativities have similar values. These are
Fig. 1.20 A hypothetical
erroneous phase diagram
1.11 The Gibbs Phase Rule 45
known as the Hume-Rothery rules. An example of such a system is Cu–Ni above
355 _C (i.e., above the solid state, miscibility gap caused at least partly by the
ferromagnetism of Ni). Thus, the A–B system shown in Fig. 1.22 exhibits ideal
behavior and atoms will not have any preference to select neighboring atoms. To
give a mental picture of what complete miscibility in solid state means, we can
trace the following line of thought. If one begins with pure A, with its own crystal
structure, and starts to replace A atoms with B atoms, then, in the case of complete
solubility, one can eventually replace all A atoms with B atoms without any
change in crystal structure or formation of new phases and reach pure element B,
with its own crystal structure (which has to be, by definition, the same as A’s). To
illustrate this behavior, we have to consider the change in free energy with
composition for two different phases, solid (gs) and liquid (gL), to find the stability
of the phases at different temperatures and compositions. Here, we assume that the
melting point of element A (TA
m ) is higher than the melting point of element B
(TB
m). First, if we consider a relatively high temperature, as shown in Fig. 1.22a, we
know from our previous discussion that the liquid phase will be stable because of a
high contribution of entropy. Now, if we start to decrease the temperature to a
certain extent, two factors shall be mainly noticed, which shall change in the free
energy diagram. We have seen before that the free energy of a liquid phase
decreases faster than that of a solid phase. So with the decrease in temperature, the
difference in free energy of both liquid and solid phases will decrease. Further,
because of the decrease in temperature, the contribution of Dgmix _TDsmix will
decrease, which means that the curvature of both curves will recede. If we
decrease the temperature up to the melting point of element A (TA
m), then we shall
find, as shown in Fig. 1.22b, that gL and gS will intersect at XB = 0. If the temperature
is decreased further, then the curves are found to intersect somewhere in
the middle, as shown in Fig. 1.22c. The diagram can be separated into three
different composition range of 0 _ XSB
_ _, XSB
_ XLB
_ _, and XLB
_ _ 1_. Here, especially
the composition range of XSB
_ XLB
_ _draws attention for further discussion.
Fig. 1.21 The two correct versions of the diagram shown in Fig. 1.20
46 1 Thermodynamics, Phases, and Phase Diagrams
For the sake of discussion, if we consider the nominal composition X
B in this
composition range, then it is apparent that the liquid phase (with this composition)
cannot be stable at this temperature, since it does not correspond to the minimum
free energy of the system. At the first instance, the solid phase with this composition
seems to be the stable one. However, the system always tries to minimize its
free energy if possible. In this case, there is a possibility to further minimize the
free energy, if both solid and liquid phases exist together, the Gibbs energy value
then sits on the common tangent as defined in Sect. 1.10. It is clear from Fig. 1.22c
that the system will have minimum free energy when the solid phase with the
composition of XSB
exists with the liquid phase having the composition of XLB
.
Fig. 1.22 Free energy versus composition diagram at different temperatures a–e for an
isomorphous system and f the corresponding phase diagram
1.12 Correlation of Free Energy and Phase Diagram in Binary Systems 47
Since X
B is the average composition of the alloy, the mole fraction of the solid
phase, following the lever rule (defined in Sect. 1.8), will be ðXLB
_X
B Þ=ðXLB
_ XSB
Þ
and the mole fraction of the liquid phase will be ðX
B _ XSB
Þ=ðXLB
_ XSB
Þ. It should
be noted that with the change in average composition within the range of
ðXLB
_ XSB
Þ, the composition of the solid and liquid phases will not change, but, only
the relative amount of the phases will change. It can also be understood from
Fig. 1.22c that in this range, the liquid and solid phases with the composition XSB
and XLB
will exist together, since the chemical potential or activity of elements A
and B in both the phases must be the same (lS
A ¼ lL
A and lS
B ¼ lL
B). In other words,
once solid and liquid phases reach their stable composition, there is no further
driving force for change. It is also clear that in the composition range of 0 _ XSB
_ _,
the solid phase will be stable since it has minimum free energy in that composition
range, whereas in the composition range of XLB
_ _ 1_, only the liquid phase will be
stable since it has minimum free energy in that composition range. If we decrease
the temperature to TB
m, the free energy curves of the phases, as depicted in
Fig. 1.22d, will intersect at XB = 1. If we decrease the free energy of the system
even further, the free energy of the solid phase, as shown in Fig. 1.22e, will be
lower than the free energy of the liquid phase in all compositions and the solid
phase will be stable. The corresponding phase diagram can be seen from Fig. 1.22f.
Next, let us consider a solution with a positive heat of mixing (endothermic
behavior), which means that there is a miscibility gap in the system. Further, it is
assumed that the miscibility gap occurs only in the solid state but not in the liquid
phase. Thus, at low temperatures, the free energy of the mixing of the solid phase will
be positive because of the positive heat of mixing. At higher temperature, however,
the free energy of mixing becomes negative, because of the growing importance of
the entropy term ð_TDsmixÞ.At a reasonably high temperature, the free energy of the
solid and liquid phases might vary with composition, as shown in Fig. 1.23a.
With a further decrease in temperature to TA
m, the free energy curves of solid
(gS) and liquid (gL) phases will intersect, as is presented in Fig. 1.23b. Any further
decrease in temperature down to T2, because of the difference in curvature, gL
intersects gS at two points, so there are five-phase regions stable at a different
composition range, as given in Fig. 1.23c. With a further decrease in temperature
to T3, we shall find that gS is lower than gL at all compositions so that only the solid
phase is stable at this temperature, as shown in Fig. 1.23d. It also should be noted
that because of the endothermic nature of transformation, with the decrease in
temperature, the curvature of gS is decreased very rapidly and with the further
decrease in temperature to T4, the free energy of mixing becomes positive and the
curvature of the free energy curve will become positive in a certain composition
range in the middle.
This is the reason why in this composition range, the solid cannot be present as a
single stable phase but will spontaneously dissociate into two different phases with
compositions a1 and a2, as can be seen from Fig. 1.23e. The area between compositions
a1 and a2 in Fig. 1.23e is called the spinodal region. It is to be noted that
48 1 Thermodynamics, Phases, and Phase Diagrams
decomposition of the phase inside the spinodal does not require nucleation. Only after
reaching the inflection points of the Gibbs energy curve, must nucleation precede the
formation of a new phase. Figure 1.23f presents the corresponding phase diagram.
There are systems in which the enthalpy of mixing is positive and of such a high
magnitude that the free energy curve will have positive curvature within a certain
composition range up to a reasonably high temperature. In such cases, the free
energy curves of the phases will change as a function of temperature in a way
shown in Fig. 1.24a–e. The corresponding phase diagram is presented in Fig. 1.24f.
Here, the main difference will be that at temperature TE and at a particular
Fig. 1.23 Free energy versus composition diagram of a system which goes through endothermic
transformation because of mixing and corresponding phase diagram
1.12 Correlation of Free Energy and Phase Diagram in Binary Systems 49
composition XE, the three phases, a, b, and L, can all exist together. Point E, plotted
in Fig. 1.24f, corresponds to eutectic transformation. At the eutectic isotherm, the
number of degrees of freedom is zero and thus, the equilibrium can occur only at a
specific temperature and with fixed compositions of a, b, and L.
In all the above cases, we have considered systems where the crystal structure
of elements A and B were similar and thus, there has been only one free energy
curve for the solid phase. However, in a system where the elements have different
crystal structure, we need to consider different free energy curves for different
elements, as shown in Fig. 1.25. Let us designate, A(B), i.e., element A with a
particular crystal structure, alloyed with some B as the a-phase. Similarly, B(A) is
Fig. 1.24 Free energy versus composition diagram of a system which goes through endothermic
transformation with very high enthalpy of mixing and corresponding phase diagram
50 1 Thermodynamics, Phases, and Phase Diagrams
designated as the b-phase. Note in Fig. 1.25a that element A with the crystal
structure of element B (ga
B) will have a much higher free energy than its stable free
energy (ga
A); thus, it is a metastable crystal structure of A. Similarly, ga
B is much
higher than ga
B. This system, under our consideration, also has an eutectic transformation
as in the previous example.
Let us next consider the case, where the formation of the c-phase in the system
is associated with strong exothermic transformation, as can be noted from
Fig. 1.26a. This implies that there is considerable difference in the electronegativities
of the elements A and B. Note also that with only a slight change in
Fig. 1.25 Free energy versus composition diagram of a system where elements have different
crystal structures and corresponding phase diagram
1.12 Correlation of Free Energy and Phase Diagram in Binary Systems 51
composition, the free energy of this phase increases very rapidly. Thus, the
composition limits of the stability region of this phase are strictly limited. The
corresponding phase diagram at a particular temperature T is given in Fig. 1.26b.
As will be discussed in Sect. 2.6, different atoms in the ordered phase try to
occupy particular lattice positions in the crystal to maximize the number of A–B
bonds depending on the average composition of the phase. These ordered phases
are known as intermetallic compound or intermediate phases, and in general, these
phases have a different crystal structure than the crystal structure of the pure
element. In the example shown in Fig. 1.26, it is to be observed that the c-phase
has a very narrow homogeneity range. However, in some cases, as presented in
Fig. 1.27, the ordered phase can, in fact, have a wide homogeneity range, where
the change in free energy (because of the small change in composition) is not very
striking, unlike as in the previous example. These phases can deviate from their
stoichiometric composition because of the presence of defects, as discussed in
Sect. 1.11.
Fig. 1.26 Free energy versus composition diagram and the corresponding phase diagram where
the phase (ordered c-phase) goes through strong exothermic transformation
Fig. 1.27 Free energy versus composition diagram and the corresponding phase diagram where
the ordered c-phase has a wide homogeneity range
52 1 Thermodynamics, Phases, and Phase Diagrams
1.13 Ternary Phase Diagrams
According to the Gibbs phase rule, the number of degrees of freedom in a
homogeneous ternary phase under constant pressure is three. Thus, we need to
specify three independent variables (two-component mole fractions and temperature)
in order to fix an equilibrium in a ternary solution phase. This leads to a
three dimensional (T, XA, XB) presentation. As already discussed, it is of common
practice to utilize an equilateral triangular (the Gibbs triangle) base (ABC) with
three binary system ‘‘walls’’ (A–B, B–C, C–A) and temperature as the vertical
axis. Next, we will briefly discuss ternary space diagrams as well as the isothermal
and vertical sections taken from those diagrams.
Figure 1.28 shows the simplest possible ternary system, where there is complete
solid and liquid solubility in the system (ABC). This ternary space model is
very simple and easy to interpret, but as the systems become more complex, the
space model becomes harder and harder to use. Therefore, it is common practice to
utilize different sections and projections from the space model to yield more easily
accessible information. As an example, the liquidus and solidus projections from
the ABC system are shown in Fig. 1.29.
These types of projections are typically made with constant temperature
intervals and can therefore be interpreted similarly as the contour lines in a map.
Accordingly, the closer the spacing of the projection lines, the steeper is the
projected surface. Isothermal sections are the most commonly used types of presentation
of ternary equilibria. Figure 1.30 shows the isothermal section at temperature
T3 from the ABC system given in Fig. 1.28.
The plane intersects the liquidus surface at T3 along the curve l1l2 and the
solidus surface along the curve s1s2. On the left-hand side of the curve l1l2, there is
Fig. 1.28 Ternary space
diagram with complete solid
and liquid solubilities in all
three binary systems
1.13 Ternary Phase Diagrams 53
a single-phase liquid region and on the right-hand side of the curve s1s2, there is a
single-phase solid region. Between these two curves, there is a two-phase liquid
and a solid region. The compositions of the phases in two-phase equilibrium are
obtained at the end points of the tie-line and the amounts by the lever rule, as in
binary phase diagrams. The directions of the tie-lines lying within the figure vary
fan-like, so that there is a gradual transition from the direction of one bounding tieline
to that of the other. No two tie-lines at the same temperature may ever cross.
This is a direct result of the Gibbs phase rule. Beyond these considerations,
nothing can be said about the direction of tie-lines, except that they must run from
liquidus to solidus. Other than those tie-lines on the edges of the diagram, none of
them point toward a corner of the diagram unless by mere coincidence or due to a
complete lack of solubility with the element at the given corner. Therefore, it is
necessary to determine the position and direction of the tie-lines experimentally. It
should be noted that the activity of a given component has the same value at each
end of a tie-line.
Vertical sections (isopleths) from ternary space diagrams can also be taken.
Figure 1.31 presents some ways in which this can be achieved. Afterward, these
isopleths are shown in Fig. 1.32.
Even though the isopleths appear quite like binary phase diagrams, they must
not be confused with them. In general, tie-lines cannot be used with isopleths and
they only show the temperature composition regions of the different phases.
Fig. 1.29 Solidus and
liquidus projections from the
space diagram in Fig. 1.28
Fig. 1.30 Isothermal section
at temperature T from the
diagram seen in Fig. 1.28
54 1 Thermodynamics, Phases, and Phase Diagrams
When there are three phases in equilibrium in a ternary system under constant
pressure, there is still one degree of freedom left. Thus, three-phase equilibrium in a
ternary system exists within a certain temperature range and not at a single temperature
as in binary systems under constant pressure. Three-phase equilibrium in a
ternary phase diagram is represented by a tie-triangle and as the temperature
changes these tie-triangles form a ‘‘stack’’ of tie-triangles (Fig. 1.33). The composition
of phases participating in the three-phase equilibrium can be found from
the corners of the tie-triangle and the amount by applying the lever rule three times.
In this hypothetical ternary system in AC, there is complete solid and liquid
solubility, whereas AB and BC are eutectic systems. Point M is the eutectic point
of system AB, which is at a higher temperature than N, which is the eutectic point
of the system BC. Thus, in both binary systems (AB and BC), an eutectic reaction
l,a + b takes place. In Fig. 1.34, the surfaces AMNC ja MNB are the liquidus
Fig. 1.31 Different ways one
can take an isopleth
Fig. 1.32 Two isopleths
taken along the lines ab and
Ac
Fig. 1.33 Stack of tietraingles
1.13 Ternary Phase Diagrams 55
surfaces and thus determine the solubilities of a and b to liquid. These surfaces
meet at the eutectic valley MN. DG and EF are curves joining the points representing
the respective compositions of the a and b phases formed in the eutectic
reactions in the binary systems. DME and GNF are horizontal lines that represent
these eutectic reactions. The surfaces ADGC and BEF are the solidus surfaces. The
three curves MN, DG, and EF do not lie in the same plane. The curve MN lies
above the surface DEFG, in such a way that there are three curved surfaces
DMNG, MEFN, and DEFG, which enclose a three-phase space where a, b, and the
liquid are in equilibrium. Each of these surfaces is made up of tie-lines representing
l + a, l + b, and a + b equilibria. The surfaces DMNG, MEFN, and
DEFG separate the three-phase space from the liquid + a, liquid + b , and a + b
regions, respectively. Where the three-phase region terminates in the binary systems
AB and BC, it shrinks to the binary eutectic lines. In a ternary system, the
eutectic reaction l ? a + b occurs over a range of temperature. If we have an
alloy with nominal composition of X, as in Fig. 1.34, the solidification takes place
as follows. During the solidification, the first solid phase to form is primary b,
when the liquidus surface is first met. The composition of the liquid then changes
along a path on the liquidus surface and that of the solid b along a path on the
solidus surface as the temperature decreases. Then, at a certain temperature T1,
before solidification is completed, the liquid composition reaches a point on the
curve MN and solid composition a point on the curve EF. The situation at T1 is
given in Fig. 1.35a where a tie-triangle is drawn. At this temperature, the nominal
composition is seen to lie on the lb tie-line. When the temperature is decreased to
T2, the three-phase equilibrium is established, as the nominal composition now lies
inside the tie-triangle Fig. 1.35b. The compositions of the liquid, a, and b are
given by the points l2, a2, and b2, and their respective amounts can be obtained by
applying the lever rule three times
Fig. 1.34 ABC ternary
system, where binary system
AC has complete solid
solubility and binary systems
AB and BC are eutectic ones
56 1 Thermodynamics, Phases, and Phase Diagrams
%liquidðl2Þ ¼
Xl0
2
l2l0
2
_ 100; %b2 ¼
Xb0
2
b2b0
2
_ 100 and %a2 ¼
Xa0
2
a2a0
2
_ 100
Solidification ends at T3 when the a3b3 tie-line is encountered.
As the ternary systems become more complicated, the analysis shown above
becomes increasingly difficult. Thus, it is of common practice to utilize different
sections taken from the space model to provide information in a more accessible
form. Next, we shall consider isothermal sections a little more, as they, in general,
provide the most useful information on the ternary system. When working with the
isothermal section and naming the phase regions, it is helpful to remember that the
sides of the three-phase triangles must always face the two-phase regions, and at
the corners of the tie-triangle, single-phase regions exist. These rules are based on
the more general Palatnik-Landau theorem. Figure 1.36 contains an example of the
AuPbSn system at 200 _C with all the phase regions clearly marked in the figure. It
is to be noted that in the case of systems with stoichiometric compounds, the twophase
regions between three-phase regions may be reduced to just a single tie-line.
As an example of utilization of isothermal sections, we shall consider the following
case. An interesting behavior has been observed in the Cu/SnBi eutectic
system during soldering at temperatures above 200 _C. In particular, when the
solder volume is small, reactions can result in drastic changes to the microstructure
when soldering times are increased. Since bismuth does not react with Cu, only tin is
consumed during the reactions. This will eventually lead to a shift in the liquid solder
composition toward the Bi-rich corner. When the isothermal section of the SnBiCu
equilibrium phase diagram is investigated, it is to be noted that when the enrichment
of liquid with bismuth increases and the composition of the solder is around
60 at-% bismuth, the local equilibrium condition changes, as shown in Fig. 1.37.
Fig. 1.35 The process of solidification of an alloy with nominal composition X in the system
shown in Fig. 1.34 is exemplified in this figure as a function of temperature. a Temperature at
which liquid composition reaches a point on the curve MN and solid composition a point on the
curve EF, b temperature where three-phase equilibrium is established, and c temperature when
solidification ends
1.13 Ternary Phase Diagrams 57
Cu6Sn5 cannot exist in local equilibrium with solder enriched with Bi at this
temperature (*200 _C, shown as contact line 2). Cu3Sn can, however, exist in local
equilibrium even with pure Bi. Therefore, the Cu6Sn5 should transform into the
Cu3Sn layer. This has indeed been experimentally verified to take place [12].
The thing of special interest was that the Cu3Sn layer maintained the original Cu6Sn5
morphology that it replaced [12].
From the thermodynamic data, one can also calculate the so-called phase
fraction diagram (NP)-diagrams, which can be utilized to investigate, for example,
Fig. 1.36 Isothermal section from the Au–Pb–Sn ternary system at 200 _C
Fig. 1.37 Isothermal section
from the Bi–Cu–Sn phase
diagram at 200 _C
58 1 Thermodynamics, Phases, and Phase Diagrams
solidification. They show, as a function of temperature, the changes in the fractions
of phase with a given nominal composition. An example is shown in Fig. 1.38.
As an example of the use of phase fraction diagrams, let us consider the next
case, where identical SnAgCu solder alloy is used to solder components on two
types of printed wiring boards (PWB’s)—one with an Ni(P)/Au and one with a
Cu(OSP) surface finish. Under the reflow conditions typically used in lead-free
soldering, the solidification structure is generally cellular, where the small Cu 6Sn5
and Ag3Sn phases are dispersed between the large primary Sn grains [13]. If
protective Au surface finishes are used, some small needle-like AuSn4 can also be
found inside the solder matrix at the high-angle boundaries. An example of the
microstructure formed in the interconnections soldered with the Sn0.5Ag0.5Cu
alloy on electrochemical Ni(P) with a thin flash Au on top (denoted Ni(P)|Au in the
following) is shown in Fig. 1.39.
Both the Cu6Sn5 and the Ag3Sn particles are uniformly distributed around the
relatively large Sn grains. Figure 1.40 shows a micrograph taken from a sample
Fig. 1.38 NP-diagram of the
phase formation during
solidification of a SnAgCu
solder, when Cu/(OSP)
metallization is used. Note
that owing to the high
fraction of Sn, the diagram
has been enlarged and
therefore relative amount of
phases goes only up to 15 %
Fig. 1.39 Solder
microstructure after reflow
when Ni(P) metallization is
used [13]
1.13 Ternary Phase Diagrams 59
soldered with the same solder alloy but this time on the boards with organic
solderability preservative (OSP) on the Cu pads (noted Cu|OSP). The resulting
microstructure seems to be different even though the same solder alloy was used
relative to Ni(P)|Au, interconnections formed on the Cu|OSP contain more and
larger Cu6Sn5 intermetallic particles dispersed inside the solder.
What is the reason behind these differences observed experimentally? Let us
take a closer look at what takes place during soldering. The thin layer of Au on top
of Ni(P) dissolves instantly and completely into the molten solder, and the Ni starts
dissolving next into the melt. The OSP coating partially evaporates and the rest
dissolves into the solder flux during soldering. In the case of the Cu|OSP boards, it
is the Cu pad that starts dissolving into the solder alloy. The dissolution rate of Cu in
Sn0.5Ag0.5Cu (wt%) is about 0.07 lm/s. Based on this, the amount of Cu dissolution
at the entire area of the soldering pad during the typical 40–45 s time above
217 _C is enough to lift the Cu concentration in the soldered interconnections close
to 1 wt%, even when taking the amount of Cu bonded into the intermetallic layers
on both sides of the interconnections into account. The dissolution rate of Ni is
about 50 times smaller than that of Cu and thus, the dissolution of Ni to the solder is
insignificant. All Ni that is dissolved at the interface is bonded to the (Cu, Ni) 6Sn5
layer. Taking into account the amount of Cu bonded to the intermetallic layers on
both sides of the interconnections, the nominal composition of the interconnections
soldered on the Ni(P)|Au-coated pads will result in about Sn0.5Ag0.3Cu, whereas
the final composition on the interconnection on Cu was about Sn0.5Ag1.0Cu.
An important consequence of higher Cu content is that solidification process is
different in interconnections soldered on Ni from those soldered on Cu.
Figures 1.38 and 1.41 present the phase fraction diagrams, where the amount of
different phases in the relative number of moles can be presented as a function of
temperature. The interconnections soldered on Ni(P)|Au PWB have the
Sn0.5Ag0.3Cu composition, whereas the interconnections soldered on Cu|OSP
have the Sn0.5Ag1.0Cu.
As can be seen from Fig. 1.41, the solidification of the liquid interconnections
soldered on Ni(P)|Au boards starts with the formation of the primary Sn phase when

You might also like