You are on page 1of 44

Module 1 – Phase Equilibrium 1

Engr. Caesar Pobre Llapitan

TOPICS
I. The Fundamental Fact of Phase Equilibrium
II. The Gibbs Phase Rule
III. Equilibria in Single-Component Systems
IV. Equilibria in Multiple-Component Systems

Intended Learning Outcome


At the end of this module, the students must be able to
1. formulate the fundamental principles of phase equilibrium given the conditions
imposed on the system.
2. evaluate the application of the Gibbs phase rule for various conditions of equilibrium
between phases.
3. analyze the phase equilibrium of a one-component systems.
4. analyze the phase equilibrium of a multi-component systems.

I. THE FUNDAMENTAL FACT OF PHASE EQUILIBRIUM

A. Introduction
The occurrence of a single stable phase or the coexistence of several phases is predictable with
the application of the principles of thermodynamics. Our purpose here is to establish the
thermodynamic relations obeyed by two or more phases that are at equilibrium with each other.

A phase refers to a form of matter that is uniform with respect to chemical composition and
the state of aggregation on both microscopic and macroscopic length scales.

▪ The principal kinds of phases are solids, liquids, and gases, although plasmas (ionized
gases), liquid crystals, and glasses are sometimes considered to be separate types of
phases.
▪ Solid and liquid phases are called condensed phases and a gas phase is often called a
vapor phase.

To describe a system thermodynamically, we need to know what’s in the system; that is, the
components of the system. We define component (c) as a unique chemical substance that has
definite properties.

For example, a system composed of pure UF6 has a single chemical component: uranium
hexafluoride.
o UF6 is composed of two elements, uranium and fluorine, but each element lost its
individual identity when the compound UF6 was formed

The phrase “chemically homogeneous” can be used to describe single-component systems.

On the other hand, a mixture of iron filings and sulfur powder is composed of the two
components iron and sulfur.
o close inspection reveals two distinct materials in the system that have their own
unique properties.

This Fe/S mixture is therefore a two-component system. The phrase “chemically


inhomogeneous” is used to describe multicomponent systems.

A phase transition occurs when a pure component changes from one phase to another. Table
1 lists the different types of phase transitions, which you are already familiar with.
Module 1 – Phase Equilibrium 2
Engr. Caesar Pobre Llapitan

Table 1: Phase transitions


Term Transition
Melting (or fusion) Solid → Liquid
Boiling (or vaporization) Liquid → Gas
Sublimation Solid → Gas
Condensation Gas → Liquid
Condensation (or deposition) Gas → Solid
Solidification (or freezing) Liquid → Solid

Phase transitions occur widely in the world around us, from the boiling of water in a teakettle
to the melting of Arctic glaciers. The water cycle of evaporation, condensation to form clouds,
and rainfall plays a key role in the ecology of the planet.

Laboratory and industrial applications of phase transitions abound, and include such processes
as distillation, precipitation, crystallization, and adsorption of gases on the surfaces of solid
catalysts.

The universe is believed to have undergone phase transitions in its early history as it expanded
and cooled after the Big Bang (M. J. Rees, Before the Beginning, Perseus, 1998, p. 205), and some
physicists have speculated that the Big Bang that gave birth to the universe was a phase
transition produced by random fluctuations in a preexisting quantum vacuum (A. H. Guth, The
Inflationary Universe, Perseus, 1997, pp. 12–14 and chap. 17).

There are also phase transitions between different solid forms of a chemical component, which
is a characteristic called polymorphism.
o elemental carbon exists as graphite or diamond (carbon exhibit solid-phase
allotropy)
o helium exhibits allotropy in the liquid phase – as a normal liquid or as a superfluid
o solid H2O can actually exist as at least six structurally different solids, depending on
the temperature and pressure (water has at least six polymorphs)
o calcium carbonate (mineral) exists either as aragonite or calcite, depending on the
crystalline form of the solid

Note:
 In application to elements, we use the word allotrope. With compounds, this
phenomenon is called polymorphism.

B. Equilibrium between Phases


There are three aspects of equilibrium. Thermal equilibrium implies that all phases have the
same temperature, mechanical equilibrium implies that all phases have the same pressure, and
phase equilibrium implies that the chemical potential of every substance has the same value in
every phase.

Under most conditions of constant volume, amount, pressure, and temperature, a single-
component system has a unique stable phase.

For example, a liter of H2O at atmospheric pressure and 25°C is normally in the liquid phase.
However, under the same conditions of pressure but at 125°C, a liter of H2O would exist as a gas.
These are the phases that are thermodynamically stable under these conditions.

If the state variables of the system are constant, then the system is at equilibrium. Therefore, it
is possible for two or more phases to exist in a system at equilibrium.
Module 1 – Phase Equilibrium 3
Engr. Caesar Pobre Llapitan

The criterion for stability at constant temperature and pressure is that the Gibbs energy,
G(T, P, n), is minimized.

Consider an equilibrium two-phase simple system containing several substances is depicted


schematically in Figure 1.

Figure 1: A Two-Phase Simple System (Retrieved from Mortimer, 2008)

- The system is closed to the surroundings, but each phase is open to the other
and both phases are at the same temperature and pressure.

If the contribution of the surface area between the phases is negligible (a good approximation
in most cases), the Gibbs energy of the system is the sum of the Gibbs energies of the two phases:

G = G( ) + G( )
I II
(1)

where we denote the two phases by the superscripts (I) and (II).

Since the system is closed, if component number i moves out of one phase it must move into
the other phase:

dni( ) = − dni( )
I II
(2)

For an infinitesimal change of state involving changes in T and P and a transfer of matter from
one phase to the other,

dG = dG ( ) + dG II
I

c c (3)
 i( ) dni( ) − S ( II ) dT + V ( II ) dP +  i(
II )
= − S ( ) dT + V ( ) dP + dni( )
I I I I II

i =1 i =1

If T and P are constant, the fundamental criterion of equilibrium implies that dG must vanish
for an infinitesimal change that maintains equilibrium:

c c
dG =  i( I ) dni( I ) +  i( II ) dni( II )
i =1 i =1

Using Eq. (2):


Module 1 – Phase Equilibrium 4
Engr. Caesar Pobre Llapitan

c
dG =   i( I ) − i( II )  dni( I ) = 0 (at equilibrium) (4)
i =1

Assume that we can find a semipermeable membrane that will selectively allow component i to
pass, but not the others. The term of the sum for component i must vanish since all of the other
dn’s vanish. Since dni is not necessarily equal to zero, the other factor in the term must vanish,
and we can write

i( ) = i( )
I II
(at equilibrium) (5-a)

If more than two phases are present at equilibrium, we conclude that the chemical potential of
any substance has the same value in every phase in which it occurs.

i( ) = i( )
 
(system at equilibrium) (5-b)

where the superscripts () and () designate any two phases of a multiphase system.

The properties of a system at equilibrium do not depend on how the system arrived at
equilibrium. Therefore, Eq. (5) is valid for any system at equilibrium, not only for a system that
arrived at equilibrium under conditions of constant T and P.

We call it the fundamental fact of phase equilibrium:

 In a multiphase system at equilibrium the chemical potential of any substance has the
same value in all phases in which it occurs.

Note:
 The mass or size of each phase does not affect the phase-equilibrium position, since the
equilibrium position is determined by equality of chemical potentials, which are
intensive variables.

Because we are considering a closed system with a single component, there are two other
implicit conditions for a system at equilibrium:

Tphase 1 = Tphase 2
pphase 1 = pphase 2

If an equilibrium is established and then the temperature or the pressure is changed, the
equilibrium must shift: that is, the relative amounts of the phases must change until equation 5
is reestablished.

C. Nonequilibrium Phases
What if the chemical potentials of the phases are not equal? Then one (or more) of the phases is
not the stable phase under those conditions. The phase with the lower chemical potential is the
more stable phase. For example:
- at 210°C, solid H2O has a lower chemical potential than liquid H2O, whereas at 110°C,
liquid H2O has a lower chemical potential than solid H2O
- at 0°C at normal pressure, both solid and liquid H2O have the same chemical potential.
They can therefore exist together in the same system, at equilibrium.
Module 1 – Phase Equilibrium 5
Engr. Caesar Pobre Llapitan

Assuming that the nonequilibrium state of the system can be treated as a metastable state, the
criterion for spontaneous processes is given by

dG  0 (T and P constant) (6)

Since dT and dP vanish and since the system as a whole is closed,

c
dG =   i( I ) − i( II )  dni( I )  0 (7)
i =1

Each term separately must be negative since the introduction of semipermeable membranes
would show each term to obey the inequality separately. The two factors in each term of the
sum in Eq. (7) must be of opposite signs:

i( )  i( ) dni( )  0


I II I
implies ` (8)

i( )  i( ) dni( )  0


I II I
implies (9)

 At constant temperature and pressure, any substance tends to move spontaneously from
a phase of higher chemical potential to a phase of lower chemical potential.

Example
Determine whether the chemical potentials of the two substances listed are the same or
different. If they are different, state which one is lower than the other.
a) Liquid mercury, Hg (,), or solid mercury, Hg (s), at its normal melting point of 238.9°C
b) H2O (l) or H2O (g) at 99°C and 1 atm
c) H2O (l) or H2O (g) at 100°C and 1 atm
d) H2O (l) or H2O (g) at 101°C and 1 atm
e) Solid lithium chloride, LiCl, or gaseous LiCl at 2000°C and normal pressure (The boiling
point of LiCl is about 1350°C.)
f) Oxygen, O2, or ozone, O3, at STP

Solution:
a) At the normal melting point, both solid and liquid phases can exist in equilibrium.
Therefore, the two chemical potentials are equal.
b) At 99°C, the liquid phase of water is the stable phase, so H2O,l < H2O,g.
c) 100°C is the normal boiling point of water, so at that temperature, the chemical
potentials are equal.
d) At 101°C, the gas phase is the stable phase for H2O. Therefore, H2O,g < H2O,l. (See what
a difference 2° makes?)
e) Because the stated temperature is above the boiling point of LiCl, the chemical potential
of gas-phase LiCl is lower than solid-phase LiCl.
f) Because diatomic oxygen is the stablest allotrope of oxygen, we expect that O2 < O3.
Note that this example doesn’t involve a phase transition.

D. Transport of Matter in a Nonuniform Phase


Consider a one-phase system that has a uniform temperature and pressure but a nonuniform
composition. Imagine dividing the system into small regions (subsystems), each one of which
is small enough so that the concentration of each substance is nearly uniform inside it. Any
substance will move spontaneously from a subsystem of higher value of its chemical potential
Module 1 – Phase Equilibrium 6
Engr. Caesar Pobre Llapitan

to an adjacent subsystem with a lower value of its chemical potential. The fundamental fact of
phase equilibrium for nonuniform systems can be stated:

 In a system with uniform temperature and pressure, any substance tends to move from a
region of higher chemical potential to a region of lower chemical potential.

A nonuniformity of the chemical potential is the driving force for diffusion.

II. THE GIBBS PHASE RULE

A. Introduction and Definitions


The various conditions of equilibrium between phases, such as the number of phases, the
number of components and the degrees of freedom (or the variance), can be correlated with
one another with the help of a general rule, known as the phase rule.

The Gibbs phase rule gives the number of independent intensive variables in a simple system
that can have several phases and several components. It is an important generalization dealing
with the behavior of heterogeneous systems.

This relationship governing all heterogeneous equilibria was first discovered as early as 1874 by
an American physicist Willard Gibbs. It may be stated mathematically as follows:

F =C −P +2 (10)

where F is the number of degrees of freedom, C is the number of components and P is the number
of phases of the system.

In general, it may be said that with the application of phase rule it is possible to predict
qualitatively by means of a diagram the effect of changing pressure, temperature and
concentration on a heterogeneous system in equilibrium.

Before proceeding with the deduction of the rule, it is necessary to define and explain the terms
involved, viz., phase, component and degree of freedom.

Phase
 A phase is defined as any homogeneous and physically distinct part of a system which is
separated from other parts of the system by definite bounding surfaces.

The term homogeneous means that the system has identical physical properties and chemical
composition throughout the whole of the system.
In general, we have:
▪ For gaseous system
Only one phase is possible since gases are completely miscible with one another in all
proportions.

▪ For liquid system


The number of phases is equal to the number of layers present in the system. For
completely miscible liquids, the number of phases is equal to one.

▪ For solid system


In general, every solid constitutes a single phase except when a solid solution is formed.
Thus, the number of phases in the solid system is equal to the number of solids present.
Module 1 – Phase Equilibrium 7
Engr. Caesar Pobre Llapitan

In solid solution, the number of phases is equal to one. Each polymorphic form and
allotropic modification constitutes a separate phase.

Examples:
1. Pure substances. A pure substance (solid, liquid, or gas) made of one chemical species
only, is considered as one phase. Thus oxygen (O2), benzene (C6H6), and ice (H2O) are
all 1-phase systems. It must be remembered that a phase may or may not be continuous.
Thus, whether ice is present in one block or many pieces, it is considered one phase.

2. Mixtures of gases. All gases mix freely to form homogeneous mixtures. Therefore, any
mixture of gases, say O2 and N2, is a 1-phase system.

3. Miscible liquids. Two completely miscible liquids yield a uniform solution. Thus, a
solution of ethanol and water is a 1-phase system.

4. Non-miscible liquids. A mixture of two non-miscible liquids on standing forms two


separate layers. Hence a mixture of chloroform (CHCl3) and water constitutes a 2-phase
system.

5. Aqueous solutions. An aqueous solution of a solid substance such as sodium chloride


(or sugar) is uniform throughout. Therefore, it is a 1-phase system.

However, a saturated solution of sodium chloride in contact with excess solid sodium
chloride is a 2-phase system.

6. Mixtures of solids.
(i) By definition, a phase must have throughout the same physical and chemical
properties. Ordinary sulphur as it occurs in nature is a mixture of monoclinic and
rhombic sulphur. These allotropes of sulphur consist of the same chemical species but
differ in physical properties. Thus, mixture of two allotropes is a 2-phase system.

(ii) A mixture of two or more chemical substances contains as many phases. Each
of these substances having different physical and chemical properties makes a
separate phase. Thus, a mixture of calcium carbonate (CaCO3) and calcium oxide
(CaO) constitutes two phases.

Let us consider the equilibrium system: The Decomposition of Calcium carbonate.


When calcium carbonate is heated in a closed vessel, we have

CaCO 3 CaO + CO2


solid gas
solid

There are two solid phases and one gas phase. Hence it is a 3-phase system

Number of Components
 It is the smallest number of independent chemical constituents by means of which the
composition of each and every phase can be expressed.

To understand the above definition and to use it for finding the number of components of a
system, remember that:
a) The chemical formula representing the composition of a phase is written on LHS.
Module 1 – Phase Equilibrium 8
Engr. Caesar Pobre Llapitan

b) The rest of the chemical constituents existing independently in the system as


represented by chemical formulas are placed on RHS.
c) The quantities of constituents on RHS can be made minus (–) or zero (0) to get at the
composition of the phase on LHS.

The independent chemical constituent is the one whose concentration can be varied
independent of other constituents of the system. Some typical examples are given below.

▪ One-component system
1) Water and sulphur systems are 1-component systems. Water system has three
phases: ice, water, and water vapor. The composition of all the three phases is expressed
in terms of one chemical individual H2O. Thus, water system has one component only.

Sulphur system has four phases: rhombic sulphur, monoclinic sulphur, liquid sulphur
and sulphur vapor. The composition of all these phases can be expressed by one
chemical individual sulphur (S). Hence it is a 1-component system.

As is clear from above, when all the phases of a system can be expressed in terms of one
chemical individual, it is designated as a one-component or 1-component system.

2) Dissociation of Ammonium chloride. Ammonium chloride when heated in a closed


vessel exists in equilibrium with the products of dissociation, ammonia (NH3) and
hydrogen chloride as (HCl).

NH4 Cl NH3 + HCl


gas
solid gas

The system consists of two phases, namely, solid NH4Cl and the gaseous mixture
containing NH3 and HCl. The constituents of the mixture are present in the same
proportion in which they are combined in solid NH4Cl. The composition of both the
phase can, therefore, be expressed in terms of the same chemical individual NH4Cl.

Phase Components
Solid = NH4Cl
Gaseous = x NH3 + x HCl or x NH4Cl

Thus, dissociation of ammonium chloride is a one-component system.

▪ Two-component system
3) Mixture of gases. A mixture of gases, say O2 and N2, constitutes one phase only. Its
composition can be expressed by two chemical substances O2 and N2.

Phase Components
Gaseous mixture = xO2 + yN2

Hence a mixture of O2 and N2 has two components. In general, the number of


components of a gaseous mixture is given by the number of individual gases present.

4) Sodium chloride solution. A solution of sodium chloride in water is a 1-phase system.


Its composition (xNaCl.yH2O) can be expressed in terms of two chemical individuals,
sodium chloride and water.
Module 1 – Phase Equilibrium 9
Engr. Caesar Pobre Llapitan

Phase Components
Aqueous solution
= NaCl + y H2O
of sodium chloride

Therefore, an aqueous solution of sodium chloride or any other solute is a two-


component or 2-component system.

A saturated solution of sodium chloride, in contact with excess solid sodium


chloride has two phase, namely aqueous solution and solid sodium chloride. The
composition of both phases can be expressed in terms of two chemical individuals NaCl
and H2O.

Phase Components
Aqueous solution
= NaCl + y H2O
of sodium chloride
Solid sodium chloride = NaCl + 0 H2O

Hence a saturated solution of sodium chloride or any other solute in contact with solid
solute, is 2-component system.

5) Decomposition of Calcium carbonate. When calcium carbonate is heated in a closed


vessel, the following equilibrium system results.

CaCO 3 CaO + CO2


solid gas
solid

It has three phases: calcium carbonate, calcium oxide, and carbon dioxide. The
composition of all the phase can be expressed in terms of any two of the three chemical
substances in equilibrium.

Let us select calcium oxide (CaO) and carbon dioxide (CO2) as the components. Then
we can write,

Phase Components
CaCO3 = CaO + CO2
CaO = CaCO3 – CO2
CO2 = CaCO3 – CaO

Again, selecting calcium carbonate (CaCO3) and calcium oxide (CaO) as the
components, we have

Phase Components
CaCO3 = CaCO3 + 0 CaO
CaO = 0 CaCO3 + CaO
CO2 = CaCO3 – CaO

Thus, decomposition of calcium carbonate is a 2-components system.

Furthermore, as is clear from the above examples, by the components of a system is


meant the number of chemical individuals and not any particular chemical
substances by name.
Module 1 – Phase Equilibrium 10
Engr. Caesar Pobre Llapitan

Degree of Freedom
 The degree of freedom or variance of the system is the minimum number of independent
variables such as temperature, pressure and concentration, that must be ascertained so
that a given system in equilibrium is completely defined.

A system with F = 0 is known as nonvariant or having no degree of freedom.


A system with F = 1 is known as univariant or having one degree of freedom.
A system with F = 2 is known as bivariant or having two degrees of freedom.

A system is defined completely when it retains the same state of equilibrium (or can be
reproduced exactly) with the specified variables. Let us consider some examples.

1) For a pure gas, F = 2. For a given sample of any pure gas PV = RT. If the values of
pressure (P) and temperature (T) be specified, volume (V) can have only one definite
value, or that the volume (the third variable) is fixed automatically. Any other sample of
the gas under the same pressure and temperature as specified above, will be identical
with the first one. Hence a system containing a pure gas has two degrees of freedom
(F = 2).

2) For a mixture of gases, F = 3. A system containing a mixture of two or more gases is


completely defined when its composition, temperature and pressure are specified. If
pressure and temperature only are specified, the third variable i.e., composition could
be varied. Since it is necessary to specify three variables to define the system completely,
a mixture of gases has three degrees of freedom (F = 3).

3) For water water vapor, F = 1. The system water in equilibrium with water vapor,
has two variables temperature and pressure. At a definite temperature the vapor pressure
of water can have only one fixed value. Thus, if one variable (temperature or pressure)
is specified, the other is fixed automatically. Hence the system water has one degree
of freedom (F = 1).

4) For saturated NaCl solution, F = 1. The saturated solution of sodium chloride in


equilibrium with solid sodium chloride and water vapor.

NaCl NaCl - solution water vapor


solid

Thus, the system is completely defined if we specify temperature only. The other two
variables i.e,. the composition of NaCl-solution (solubility) and vapor pressure have a
definite value at a fixed temperature. Hence the system has one degree of freedom.

5) For ice-water-vapor system, F = 0. In the system ice water vapor, the three
phases coexist at the freezing point of water. Since the freezing temperature of water
has a fixed value, the vapor pressure of water has also a definite value. The system has
two variables (temperature and pressure) and both these are already fixed.

Thus, the system is completely defined automatically, there being no need to specify any
variable. Hence it has no degree of freedom (F = 0).

Alternatively,
 The degree of freedom of the system may be defined as the number of factors, such as
temperature, pressure and concentration, which can be varied independently without
altering the number of phases.
Module 1 – Phase Equilibrium 11
Engr. Caesar Pobre Llapitan

- For a single-phase system, both temperature and pressure can be varied


independently of each other and thus the system is bivariant.
- For two phases in equilibrium, only one variable can be varied as the other one
will automatically have a fixed value and thus the system is univariant.
- When two phases are in equilibrium, this special condition can be attained at a
definite temperature and pressure. The system is, therefore, defined completely
and no further statement of external conditions is necessary.

B. Derivation of Phase Rule


Phase rule is a general rule which is applicable to all types of reactive and nonreactive systems.
In a nonreactive system, we simply have the distribution of various components in different
phases without any complications such as the chemical reaction between the components.

Nonreactive System
We derive the rule, first, for a nonreactive system and then show how the same rule can be used
for a reactive system.

Consider a heterogeneous system in equilibrium of C components in which P phases are present.


We have to determine the degrees of freedom of this system i.e., the number of variables which
must be arbitrarily fixed in order to define the system completely.

Since the state of the system will depend upon the temperature and the pressure, these two
variables are always there.

The system at equilibrium can be completely described if we know the values of the variables
listed in Table 2.

Table 2: Number of Variables to be Known to Define a Given System Completely


Variables Number
i. Temperature of the system 1
ii. Pressure of the system 1
iii. Concentration of each and every component in all the P phases.
For each phase, we will have to specify the values of C
PC
concentration terms and thus for P phases, we will have to specify
in all PC values
Total number of variables that need to be specified PC + 2

Values of these variables can be obtained by solving the equations which are applicable when
the system is at equilibrium. There are two types of equations which are available (Table 3).

Table 3: Types and Number of Equations which are Available


Equations Number
i. Condition of sum of the amount fractions in any phase being
equal to one
x 1 ( 1 ) + x2( 1 ) + ... + xc( 1 ) = 1
x 1 ( 2 ) + x2( 2 ) + ... + xc( 2 ) = 1
P

x 1 ( P ) + x2( P ) + ... + xc( P ) = 1


There will be as many equations as the number of phases.
Thus, their total number will be
Module 1 – Phase Equilibrium 12
Engr. Caesar Pobre Llapitan

ii. Thermodynamic condition of phase equilibria


According to this, the species will be distributed in such a
manner that value of free energy of the system at equilibrium
is minimum. The condition for this is that the chemical
potential of any component will have the same value in all the
P phases, i.e.
1 ( 1 ) = 1 ( 2 ) = ... = 1 ( P ) C(P – 1)
2( 1 ) = 2( 2 ) = ... = 2( P )

c( 1 ) = c( 2 ) = ... = c( P )


For each component, we will have (P – 1) equations and thus
for C components, the number of equations will be C(P – 1)
Total number of equations that are available P + C(P – 1)

Mathematically, we know that the number of variables that can be obtained from a set of
equations is equal to the number of equations.
- If there are as many equations as there are variables, then the temperature, pressure and
composition of the whole system in equilibrium can be determined (nonvariant system).
- If the number of variables exceeds the number of equations by one, then the equilibrium
of the system cannot be determined until one of the variables is arbitrarily chosen
(mono-variant or univariant) and has one degree of freedom.

Thus, we see that the excess of variables over equations, which is called the variance F of the
system, is given as

 Total number of variables   Total number of equations 


Variance =   −  
 that need to be specified   that are available 

Using the results from Tables 2 and 3:

F = ( PC + 2 ) − P + C ( P − 1 ) or

F = C −P +2 (10)

F + P = C +2 (11)

Equation (10 or 11) is the phase rule which connects the number of phases and components with
the variance of the system.

Example:
Find F for a system consisting of solid sucrose in equilibrium with an aqueous solution of
sucrose.

Solution:
The system has two chemical species (water and sucrose), so C = 2.
The system has two phases (the saturated solution and the solid sucrose), so P = 2.

Hence: F=C-P+2=2-2+2=2
Module 1 – Phase Equilibrium 13
Engr. Caesar Pobre Llapitan

Two degrees of freedom make sense, since once T and P are specified, the equilibrium mole
fraction (or concentration) of sucrose in the saturated solution is fixed.

Question 1
Find f for a system consisting of a liquid solution of methanol and ethanol in equilibrium
with a vapor mixture of methanol and ethanol. Give a reasonable choice for the independent
intensive variables. (Answer: 2; T and the liquid-phase ethanol mole fraction.)

Reactive System
Let us consider a heterogeneous system composed of arbitrary amounts of C constituents. For
the sake of simplicity, we assume that four of the constituents are chemically active, capable of
undergoing the reaction.

 1 A1 +  2 A2  3 A3 +  4 A4

We write down the number of variables that are needed to describe the system completely and
the number of equations that are available at equilibrium.

Table 4: Number of Variables to be Stated to Define a Given System Completely


Variables Number
i. Temperature of the system 1
ii. Pressure of the system 1
iii. Concentration of constituents in different phases. CP
Total number of variables that need to be specified CP + 2

Table 5: Types and Number of Equations which are Available


Equations Number
i. Condition of amount fraction:
x 1 ( 1 ) + x2( 1 ) + ... + xc( 1 ) = 1
x 1 ( 2 ) + x2( 2 ) + ... + xc( 2 ) = 1 P

x 1 ( P ) + x2( P ) + ... + xc( P ) = 1


ii. Condition of thermodynamic equilibrium between phases:
1 ( 1 ) = 1 ( 2 ) = ... = 1 ( P )
2( 1 ) = 2( 2 ) = ... = 2( P ) C(P – 1)

c( 1 ) = c( 2 ) = ... = c( P )


iii. Besides above two conditions, there is another condition
which has to be satisfied for the reactive system, that is, the
thermodynamic condition of reaction equilibrium. At
equilibrium, the value of reaction potential, rG is zero. Thus 1
0 = r G =  j  j 
j
Thus, we get one more equation
Total number of equations that are available P + C(P – 1) + 1

From Tables 4 and 5, we have


Module 1 – Phase Equilibrium 14
Engr. Caesar Pobre Llapitan

 Total number of variables   Total number of equations 


Variance =   −  
 that need to be specified   that are available 
F = ( PC + 2 ) − P + C ( P − 1 ) + 1
F = (C − 1 ) − P + 2

Similarly, if two independent reactions are possible, then we get

F = (C − 2 ) − P + 2

Generalizing, we have

F = (C − r ) − P + 2 (12)

where r is the number of independent reactions that are taking place in the system. By
independent chemical reactions, we mean that no reaction can be written as a combination of
the others.

Additional Restricting Conditions


We have considered three kinds of equations amongst the variables T, P, and the x’s:
a) equations of phase equilibria;
b) equations of reaction equilibria; and
c) equations of the type 
xi = 1
i

Sometimes a chemical reaction takes place in such a manner that additional equations
expressing further restriction upon the x’s are to be satisfied.

i. Mole fractions arising from stoichiometric


o Suppose we have a gas-phase system containing only NH3; we then add a catalyst to
establish the equilibrium 2NH3  N2 + 3H2; further, we refrain from introducing any
N2 or H2 from outside.

Since all the N2 and H2 comes from the dissociation of NH3, we must have
nH2 = 3nN2 and xH2 = 3xN2

This stoichiometry condition is an additional relation between the intensive


variables besides the equilibrium relation 2NH3 = N2 + 3H2
.
ii. Electroneutrality conditions
o Suppose that we have a system in which a salt is dissolved in a solvent. Suppose that
the salt dissociates according to the scheme given below:
AB → A + + B −

All the ions remain in the solution and no precipitate is formed. Consequently, we
have the additional restricting equation
x A+ = x B −

If multi-dissociation takes place as shown below.


Module 1 – Phase Equilibrium 15
Engr. Caesar Pobre Llapitan

A → B+ + C -
C- D+ + E2-
E2- F + + G 3-

then three independent restricting equations among the xs are available. These are
x B + = xG 3 − + x E 2 − + xC −
x D+ = xG 3 − + x E 2 −
x F + = xG 3 −

Adding these equations, we get the dependent equation


x B + + xG 3 − + x F + = 3 xG 3 − + 2 x E 2 − + xC −

which expresses the condition of electrical neutrality.

If, in general, a reactive system has r independent reactions and z independent restrictive
conditions, then the number of equations that are available at equilibrium is:

Total number  Equations due to the   Equations due to 


= + 
of equations  conditions of amount fractions   the phase equilibria 
 Equations due to the   Equations due to 
+ + 
 chemical reactions   restricting conditions 
= P + C (P −1) + r + z

Thus, the variance is

F = ( CP + 2 ) − P + C ( P − 1 ) + r + z
Or
F = (C − r − z ) − P + 2 (13-a)
Or
F = C' − P + 2 (13-b)

where C' = C – r – z and is known as the number of components of the system.

▪ It is thus equal to the total number of constituents present in the system less the number
of independent chemical reactions and the number of independent restricting
equations. With this definition, we have put the phase rule in the form that is applicable
for a nonreactive system.

Some Typical Examples to Compute the Number of Components


Having defined the number of components as the minimum number of independent chemical
constituents required to express the composition of each and every phase of a system. This
definition is completely in agreement with the definitions given below for the reactive and the
nonreactive systems.

For nonreactive system:


 Number of components = Number of constituents
Module 1 – Phase Equilibrium 16
Engr. Caesar Pobre Llapitan

For reactive system:


 Number of components = (Number of constituents) – (Number of chemical reactions)
– (Number of restricting conditions)

Illustration 1: Sodium Dihydrogen Phosphate-Water System


 It is a two-component so long as no precipitate is formed by virtue of a reaction between
salt and water.

Neglecting all dissociations


o There are two constituents, no chemical reaction, and no restricting equation. Hence

C' = C – r – Z = 2 - 0 - 0 = 2

Complete dissociation of the salt


o The total number of constituents present in the solution are three: Na+, SO42– and H2O.
Since there is one restricting condition:
xNa+ = 2 xSO2-
4

we have C' = C – r – Z = 3 - 0 - 1 = 2

Partial dissociation of the salt


o The dissociation may be represented as
Na 2SO4 2Na+ + SO42−

Now the system possesses four constituents, namely, Na2SO4, Na+, SO42– and H2O.
There is one reaction and one restricting condition:
xNa+ = 2 xSO2-
4

Thus, we have C' = C – r – Z = 4 - 1 - 1 = 2

Dissociation of water also


o We have the following two equilibrium reactions:
Na 2SO4 2Na+ + SO42−
H2O H+ + OH−

The different constituents are Na2SO4, Na+, SO42–, H2O, H+ and OH-, a total of six.
There are two reactions and two restricting conditions:
xNa+ = 2 xSO2- and xH+ = xOH-
4

Thus, C' = C – r – Z = 6 - 2 - 2 = 2

Illustration 2: Aluminum Chloride-Water System


 It can be shown that this system is a two-component system irrespective of whatever
chemical changes may be considered in the solution except that no precipitate is formed.

Neglecting all dissociations


o There are two constituents, no chemical reaction and no restricting equation. Hence

C' = C – r – Z = 2 - 0 - 0 = 2
Single dissociation of the salt
Module 1 – Phase Equilibrium 17
Engr. Caesar Pobre Llapitan

o We may represent the dissociation as


NaH2PO4 Na + + H2PO4−

There are four constituents (NaH2PO4, H2O, Na+ and H2PO4–), one chemical reaction
and one restricting equation:
xNa+ = xH PO-
2 4

Hence: C' = C – r – Z = 4 - 1 - 1 = 2

Multiple dissociation of the salt


o The multistep dissociation may be represented as
NaH2PO4 Na + + H2PO4−
H2PO4− H+ + HPO24−
HPO24− H+ + PO43 −

There are seven constituents, NaH2PO4, Na+, H2PO4–, H+, HPO42–, PO43– and H2O.
Number of chemical reactions = 3
Number of independent restricting equations = 2:
xNa+ = x 3- + x 2- + x -
(i)
PO4 HPO4 H2PO4

xH+ = 2 x 3-
+x (ii)
PO4 HPO2-
4

Note: One may consider the charge balance expression


xNa+ + xH+ = 3 x 3- + 2 x 2- + x -
(iii)
PO4 HPO4 H2PO4

Equation (ii) may be obtained by subtracting Eq. (i) from the above expression.

Thus, we have C' = C – r – Z = 7 - 3 - 2 = 2

Dissociation of water also


o Besides the multiple dissociation, we also have
H2O H+ + OH−

There are total eight constituents: NaH2PO4, Na+, H2PO4–, H+, HPO42–, PO43–, H2O, and
OH-.

Number of chemical reactions = 4


Number of restricting equations = 2:
xNa+ = x 3- + x 2- + x
PO4 HPO4 H2PO-4

xH+ = 2 x 3-
+x
PO4 HPO2-
4

Thus, we have C' = C – r – Z = 8 - 4 - 2 = 2

Illustration 3: Aluminum Chloride-Water System

In this case, AlCl3 combines with water according to the equation:


Module 1 – Phase Equilibrium 18
Engr. Caesar Pobre Llapitan

AlCl 3 + 3H2O Al(OH)3 + 3HCl

o Some of the Al(OH)3 is precipitated out. The dissociation of the various species will be
as follows:
AlCl 3 Al 3+ + 3Cl − Al(OH)3 Al 3+ + 3OH−
H2O H+ + OH− HCl H+ + Cl −

Total number of constituents = 8:


AlCl3, H2O, Al(OH)3, HCl, Al3+, Cl–, H+ and OH–

Number of chemical reactions = 4

Since some of the Al(OH)3 has precipitated out, there is only one restricting equation,
namely, that expressing the electrical neutrality of the solution
3x Al3+ + xH+ = xOH- + xCl-

Thus, the number of components C' = C - r - Z = 8 – 1 – 4 = 3

o If we assume complete dissociation of AlCl3 and that of HCl, we will have

C = 6, r = 2 and Z = 1

Thus, C' = C – r – Z = 6 – 1 – 2 = 3

Sample Problems
1. Show that an aqueous system containing K+, Na+ and Cl– is a three-component system
whereas K+, Na+, Cl– and Br– is a four-component system. What are the number of
components if the salts are present in equal amounts?

Solution:
(a) K+, Na+, Cl– and H2O system:
Number of reactions, r = 3:
H2O H+ + OH −
KCl K + + Cl −
NaCl Na + + Cl −

Number of constituents, C = 8:
H2O, KCl, NaCl, H+, OH–, K+, Cl– and Na+

Number of restricting equations, Z = 2:


Electrical neutrality: xNa+ + xK + + xH+ = xCl- + xOH-
Water dissociation: xH+ = xOH-

Hence, Number of components, C' = C – r – Z = 8 – 3 – 2 = 3

o If the salts are present in equal amounts, then one more restricting equation exists, i.e.
xNa + = xK +

Hence, Number of components, C' = C – r – Z = 8 – 3 – 3 = 2


Module 1 – Phase Equilibrium 19
Engr. Caesar Pobre Llapitan

o Alternatively, leaving KCl and NaCl on the basis that they are strong electrolytes, we
have
Number of constituents, C = 6:
H2O, H+, OH–, K+, Cl– and Na+

Number of reactions, r = 1:
H2O H+ + OH−

Number of restricting equations, Z = 2:


xH+ = xOH- and xNa+ + xK + = xCl-

Hence, Number of components, C' = C – r – Z = 6 – 1 – 2 = 3

o If the salts are present in equal amounts, then one more restricting equation exists, i.e.
xNa + = xK +

Hence, Number of components, C' = C – r – Z = 6 – 1– 3 = 2

(b) K+, Na+, Cl–, Br– and H2O system:


Number of reactions, r = 5;
H2 O H+ + OH − NaCl Na + + Cl −
KCl K + + Cl − KBr K + + Br −
NaBr Na + + Br −

Number of constituents, C = 11:


H2O, H+, OH–, KCl, K+, Cl–, NaBr, Na+, Br–, NaCl and KBr

Number of restricting equations, Z = 2:


Electrical neutrality: xNa+ + xK + + xH+ = xCl- + xBr- + xOH-
Water dissociation: xH+ = xOH-

Hence, Number of components, C' = C – r – Z = 11 – 5 – 2 = 4

o If the salts are present in equal amounts, then we have three independent restricting
equations:
xNa+ = xK + ; xH+ = xOH- ; and xCl- = xBr-

Hence, Number of components, C' = C – r – Z = 11 – 5 – 3 = 3

o Alternatively, leaving KCl, NaBr, NaCl and KBr as they are strong electrolytes, we have:
Number of constituents, C = 7:
H2O, H+, OH–, Na+, K+, Cl– and Br–

Number of reactions, r = 1:
H2O H+ + OH−

Number of restricting equations, Z = 2:


xH+ = xOH- and xNa+ + xK + = xCl- + xBr-
Module 1 – Phase Equilibrium 20
Engr. Caesar Pobre Llapitan

Hence, Number of components, C' = C – r – Z = 7 – 1 – 2 = 4

o If the salts are present in equal amounts, then we have three independent restricting
equations:
xNa+ = xK + ; xH+ = xOH- ; and xCl- = xBr-

Hence, Number of components, C' = C – r – Z = 7 – 1 – 3 = 3

2. Show that NH4Cl(s) – NH3(g) – HCl(g) system in which pNH3 = pHCl is a one-component
system whereas when pNH3 ≠ pHCl is a two-component system.

Solution:
(a) NH4Cl(s) – NH3(g) – HCl(g) when pNH3 = pHCl. The condition of pNH3 = pHCl would arise
only when the gases are obtained by the sublimation of NH4Cl(s). Thus, we have
Number of reactions, r = 1:
NH4 Cl (s) NH3 (g) + HCl (g)

Number of constituents, C = 3:
NH4Cl(s), NH3(g) and HCl(g)

Number of restricting equations, Z = 1:


pNH3 = pHCl

Hence, Number of components, C' = C – r – Z = 3 – 1 – 1 = 1

(b) NH4Cl(s) – NH3(g) – HCl(g) system and pNH3 ≠ pHCl. Here, we have
Number of reactions, r = 1
Number of constituents, C = 3
Number of restricting equations, Z = 0

Hence, Number of components, C' = C – r – Z = 3 – 1 – 0 = 2

3. Determine the number of components in a system containing NH4Cl (s), NH4+ (aq), Cl– (aq),
H2O (l), H3O+ (aq), H2O (g), NH3 (g), OH– (aq), and NH4OH (aq).

Solution:
We have
Number of constituents, C = 9

Number of equilibrium reactions, r = 5:


NH4 Cl (s) + aq NH4+ (aq) + Cl − (aq)
NH4+ (aq) + 2H2O NH4 OH (aq) + H 3O + (aq)
NH3 (g) + H2O (l ) NH4 OH (aq)
2H2O (l ) H3 O+ (aq) + OH − (aq)
H 2 O (l ) H2O (g)

Number of restricting conditions, Z = 1:


xNH+ + xH O+ = xCl- + xOH-
4 3

Hence, Number of components, C' = C – r – Z = 9 – 5 – 1 = 3


Module 1 – Phase Equilibrium 21
Engr. Caesar Pobre Llapitan

4. Consider a homogeneous mixture of four ideal gases capable of undergoing the reaction
 1 A1 +  2 A2  3 A3 +  4 A4

Determine the components if we start with (a) arbitrary amounts of A1 and A2 only, (b)
arbitrary amounts of all the four gases, and (c) ν1 moles of A1 and ν2 moles of A2 only.

Solution:
(a) Number of constituents, C = 4:
A1, A2, A3 and A4

Number of reactions, r = 1
Number of restrictions, Z = 1:
 A3  :  A4  : : v3 : v4

Hence, Number of components, C' = C – r – Z = 4 – 1 – 1 = 2

(b) Number of constituents, C = 4


Number of reactions, r = 1
Number of restrictions, Z = 0

Number of components, C' = C – r – Z = 4 – 1 – 0 = 3

(c) Number of constituents, C = 4


Number of reactions, r = 1
Number of restrictions, Z = 2:
 A1  :  A2  : : v1 : v2 and  A3  :  A4  : : v3 : v4

Hence, Number of components, C' = C – r – Z = 4 – 1 – 2 = 1

Classroom Activity 1
Compute the number of components in the following systems:
1. A solution containing Na+, Cl–, Ag+, NO3–, AgCl(s) and H2O.
2. A solution containing H+, OH–, Na+, Cl–, Ag+, NO3–, AgCl(s) and H2O.
3. A solution containing H2O, Na+, Cl–, K+, NO3–, NH4+, NH3, H+ and OH–.
4. An aqueous solution containing H3PO4, H2PO4, HPO4–, PO43–, Na+, and H+ at 1 atm
pressure.
5. NH4Cl(s), NH4+(aq), Cl–(aq), H2O(l), H3O+(aq), H2O(g), NH3(g), OH–(aq), and
NH4OH(aq).
Answers: (1) 4, (2) 4, (3) 6, (4) 3, (5) 3)

Exercises 1
Problems from Section 5.2

Reference: Physical Chemistry, 3rd Edition by Robert G. Mortimer


Module 1 – Phase Equilibrium 22
Engr. Caesar Pobre Llapitan

III. EQUILIBRIA IN SINGLE-COMPONENT SYSTEMS

A. One-Component Phase Diagram


We use projections of the PVT surface for a one-component system onto the PV-plane and the
PT– plane to describe equilibria with a single-component system. Figure 2 describe a typical PT-
diagram.

Figure 2: The Coexistence Curves for a Typical Pure Substance


(Retrieved from Mortimer, 2008)

▪ For a one-component one-phase system the number of independent intensive variables


at equilibrium is

F=1–1+2=2 (one component, one phase)

 The temperature and pressure can both be independent, and any point in the area can
represent a possible intensive state of the system. Therefore, in order to define the
condition of the phase both pressure and temperature must be stated.

▪ For one component and two phases, the number of independent intensive variables at
equilibrium is

F=1–2+2=1 (one component, two phases)

The pressure must be a function of the temperature:

P = P(T) (C = 1, P = 2)
Module 1 – Phase Equilibrium 23
Engr. Caesar Pobre Llapitan

This function is represented by a coexistence curve in the phase diagram. Figure 2 shows a
solid–vapor curve, a solid–liquid curve, and a liquid–vapor curve. The equilibrium pressure
when a liquid phase or a solid phase is equilibrated with a vapor (gas) phase is called the
vapor pressure of that phase. Figure 3 shows the equilibrium vapor pressure of liquid water
as a function of temperature.

 Along any of three lines on the phase diagram when one variable (pressure or
temperature) is specified, the other is fixed automatically.

Figure 3: The Equilibrium Vapor Pressure of Water as a Function of Temperature.


(Retrieved from Mortimer, 2008)

▪ If one component and three phases are present, F = 0 and there is no choice about the
temperature, the pressure, or any other intensive variable. This three-phase state is
represented by a triple point at which three coexistence curves intersect.
 At the triple point both pressure and temperature on the diagram are fixed and,
therefore, the system is nonvariant. This implies that if we try to change temperature or
pressure, the equilibrium will be disturbed. For example, if we lower the pressure on the
system, all the liquid will vaporize, leaving only two phases.

B. The Clapeyron Equation


We now discuss a theoretical basis for the coexistence curves that separate different single-
phase regions in the P–T phase diagram. Along the coexistence curves, two phases are in
equilibrium.

Consider a one-component system with two phases ( and ). At equilibrium, the pressure,
temperature, and chemical potential must be the same in the two phases. For the chemical
potentials,

 =   (14)

▪ When the temperature is changed at constant pressure, or the pressure is changed at


constant temperature, one of the phases will disappear.
▪ If the temperature and pressure are both changed in such a way as to keep the two
chemical potentials equal to each other, the two phases will continue to coexist.
▪ The plot of pressure versus temperature along which the two phases coexist is referred
to as the coexistence curve (see Fig. 4).
Module 1 – Phase Equilibrium 24
Engr. Caesar Pobre Llapitan

Figure 4: Coexistence curve for a one-component system. (Retrieved from Silbey, 2005)

The necessary relation for dP/dT was derived by Clapeyron. For a change of pressure and
temperature along the coexistence curve,

d  = d   (15)

since the chemical potential is equal to the molar Gibbs energy for a one-component system.
Thus, using the equation d  = dG = VmdP − SmdT , we obtain

Vm, dP − Sm, dT = Vm,  dP − Sm,  dT

dP Sm,  − Sm, Sm


= = (16-a)
dT Vm,  − Vm, Vm

This equation is referred to as the Clapeyron equation, and it may be applied to vaporization,
sublimation, fusion, or the transition between two solid phases of a pure substance.

The Clapeyron equation also works for liquid-gas and solid-gas phase transitions, but as we will
see shortly, some approximations can be made that allow us to use other equations with
minimal error.

For a reversible phase change at constant pressure ΔG = 0, so that

Hm
Sm =
T

The Clapeyron equation can be written

dP Hm
= (another version of the Clapeyron equation) (16-b)
dT T Vm

This expression relates changes in phase-change conditions, but in terms of the molar quantities
transHm and transVm.

Note that dP/dT is a total derivative, not a partial derivative; however, there is a constraint: the
two phases remain in equilibrium so that G = 0.
Module 1 – Phase Equilibrium 25
Engr. Caesar Pobre Llapitan

• The Clapeyron equation allows us to calculate the slope of the coexistence curves in a P–T
phase diagram if Sm and Vm for the transition are known. This information is necessary
when constructing a phase diagram.

Illustration
Interpret the curves in the phase diagram of water, see Figure 5, that appear to be vertical
and horizontal line segments.

Figure 5 The Phase Diagram of Water (From B. Kamb, in E. Whalley, S. Jones, and
L. Gold, eds., Physics and Chemistry of Ice, University of Toronto Press, Toronto, 1973.)
(Retrieved from Mortimer, 2008)

 A horizontal line segment corresponds to zero value of dP/dT, implying that ΔHm = 0 and
ΔSm = 0 for the phase transition. For example, between ice VI and ice VII it appears that
ΔSm = 0 and ΔVm ≠ 0.
 A vertical line segment corresponds to an undefined (infinite) value for dP/dT, implying
that ΔVm = 0 and ΔSm ≠ 0. For example, ice VII and ice VIII appear to have the same
molar volume.

The molar enthalpies of sublimation, fusion, and vaporization at the triple point are related by

 sub H =  fus H +  vap H (17)

This is expected since enthalpy is a state function and the same amount of heat is required to
vaporize a solid directly as to go through an intermediate melting stage to reach the final vapor
state.

Water is unusual in that it expands on freezing, so that V and thus dP/dT for melting is negative.

• The Clapeyron equation can be applied to substances under extreme conditions of


temperature and pressure, because it can estimate the conditions of phase transitions – and
therefore the stable phase of a compound – at other than standard conditions. Such
Module 1 – Phase Equilibrium 26
Engr. Caesar Pobre Llapitan

conditions might exist, say, at the center of a gas giant planet like Saturn or Jupiter. Or,
extreme conditions might be applied in various industrial or synthetic processes.

❖ Further Reading:
Sections 8.1 – 8.5: Engel, Thomas, Philip Reid and Warren Hehre (2013). Physical
Chemistry, 3rd Edition

Sample Problems:
1. Estimate the pressure necessary to melt water at -10.0°C if the molar volume of liquid
water is 18.01 mL and the molar volume of ice is 19.64 mL. Sm for the process is +22.04
J/K and you can assume that these values remain relatively constant with temperature.
You will need the conversion factor: 1 L-bar = 100 J.

Solution:
The change in molar volume for the reaction
H2O (s) H2O (l)

Vm = 18.01 mL - 19.64 mL = -1.63 mL = -1.63  10-3 L


T = -10.0 0C = -10.0 K
Sm = +22.04 J/K

Therefore,
dP P Sm 22.04 J/K
= = =
dT −10.0 K Vm -1.63  10-3 L

P =
( 22.04 J/K )( -10.0 K ) 
1 L  bar
= 1.35  10 3 bar
-1.63  10-3 L 100 J

o Because 1 bar equals 0.987 atm, it takes about 1330 atm to lower the melting point of
water to 210°C. This is an estimate, because Vm and Sm would be slightly different at
210° than at 0°C (the normal melting point of ice) or at 25°C (the common
thermodynamic reference temperature). However, it is a very good estimate, because
both DV and DS do not vary much over such a small temperature range.

2. Estimate the pressure necessary to make diamond from graphite at a temperature of


2298 K, that is, with T = (2298 - 298) K = 2000 K. Use the following information:

C (s, graphite) C (s, diamond)


Sm (J/K) 5.69 2.43
Vm (L) 4.14 × 10-3 3.41 × 10-3

Solution:
Using the Clapeyron equation, we find that
P (2.43 - 5.69) J/K 1 L  bar
= 
2000 K (3.41 - 4.41)  10 L
-3 100 J
P = 65,200 bar

3. What pressure is necessary to change the boiling point of water from its 1.000-atm value
of 100°C (373 K) to 97°C (370 K)? The enthalpy of vaporization of water is 40.7 kJ/mol.
The density of liquid water at 100°C is 0.958 g/mL and the density of steam is 0.5983 g/L.
You will have to use the relationship 101.32 J = 1 Latm.
Module 1 – Phase Equilibrium 27
Engr. Caesar Pobre Llapitan

Solution:
First, we calculate the change in volume.
For 1.00 mole of water: Vliq = 18.01/0.958 = 18.8 mL
For 1.00 mole of steam: Vvap = 18.01/0.5983 = 30.10 L

Vm = 30.10 L - 18.8 mL = 30.08 L per mole of water

dP Hm
Using Eq. (16-b): =
dT T Vm

Separating the variables, we obtain


Hm dT
dP =
Vm T

Solving,
Pf Hm Tf dT
Pi
dP =
Vm T
i T
Hm T f
P = ln
Vm Ti
40,700 J 370 K 1 L  atm
= ln 
30.08 L 373 K 101.32 J
P = -0.108 atm

o This is the change in pressure from the original pressure of 1.000 atm; the actual pressure
at which the boiling point is 97 °C is therefore (1.000 - 0.108) atm = 0.892 atm. This would
be the pressure about 1000 meters above sea level, or about 3300 feet. Because many
people live at that altitude or higher around the world, substantial populations
experience water with a boiling point of 97 °C.

4. What is the change in the boiling point of water at a 100 0C per Pa change in atmospheric
pressure?

Solution:
The molar enthalpy of vaporization is 40.69 kJ mol-1, the molar volume of liquid water is
0.019 × 10-3 m3/mol, and the molar volume of steam is 30.199 × 10-3 m3/mol, all at 100 0C and
1.01325 bar:
dP  vap Hm  vap Hm
= =
(
dT T  vapVm T Vm, g − Vm, l )
(40 690 J mol-1 )
=
(373.15 K)(30.180  10-3 m 3 mol-1 )
dP
= 3613 Pa K -1
dT

Thus,
dT 1
= -1
= 2.768  10-4 K Pa-1
dP 3613 Pa K
Module 1 – Phase Equilibrium 28
Engr. Caesar Pobre Llapitan

5. Calculate the change in pressure required to change the freezing point of water 1 0C. At
0 0C the heat of fusion of ice is 333.5 J/g, the density of water is 0.9998 g/cm3, and the
density of ice is 0.9168 g/cm3.

Solution:
The reciprocals of the densities, 1.0002 and 1.0908, are the volumes in cubic centimeters of
1 g.

The volume change on freezing (Vm, l – Vm, s) is therefore -9.06 × 10-3 m3/g.

For small changes Hfus, T, and (Vm, l – Vm, s) are virtually constant, so that
P  fus Hm 333.5 J/g
= =
(
T T Vm, l − Vm, s )
(273.15 K)(-9.06  10-8 m 3 /g)

= -1.348  107 Pa K -1

The change in the freezing point of water per bar pressure is


T 105 Pa bar-1
= = -0.0075 K bar-1
P -1.348  107 Pa K -1

o This shows that an increase in pressure of 1 bar lowers the freezing point 0.0075 K. The
negative sign indicates that an increase in pressure causes a decrease in temperature.
The change in pressure required to change the freezing point of water 1 0C is
P 1
= - = -133 bar/K
T 0.0075 K bar-1

6. The melting point of mercury is 234.5 K at 1.013 25 bar pressure and it increases 5.033 ×
10–3 K per bar increase in pressure. The densities of solid and liquid mercury are 14.19
and 13.70 g cm–3, respectively, (a) Determine the molar enthalpy of fusion, (b) Calculate
the pressure required to raise the melting point to 273 K.

Solution:
P  fus Hm
a) Since = , we have
T Tm  fusVm

Tm  fusVm
 fus Hm = and
T /P
 1 1 
 fusVm = Vm, l − Vm, s = M  − 
 l  s 

From the given data, we get


 1 1 
 fusVm = ( 200.59 g/mol )  − 
 13.70 g/cm 3 14.19 g/cm 3
 
= 0.51 cm 3 /mol = 0.51  10-3 dm 3 /mol

Hence,
Module 1 – Phase Equilibrium 29
Engr. Caesar Pobre Llapitan

Tm  fusVm (234.5 K)(0.51  10-3 dm 3 /mol)


 fus Hm = =
T /P (5.033  10-3 K/bar)
 100 kPa 
= 23.30 dm 3 bar mol-1   = 23.30  10 J/mol
2
 1 bar 
= 2.33 kJ/mol

b) Substituting the given data in the expression


 fus Hm T
P =
 fusVm Tm

We get
(23.30 dm 3 bar mol-1 )(38.5 K
P = = 7 501 bar
(0.51  10-3 dm 3 mol-1 )(234.5 K)

7. The vapor pressure of a liquid which obeys Trouton’s rule rises by 52 Torr between
temperatures 1 K below and 1 K above the normal boiling point. Determine the value of
the normal boiling point and the molar enthalpy of vaporization of the liquid.

Solution:
The given data are
dP = 52 Torr dT = 2 K
 vap Hm
= 10.5 or  vap Hm = 10.5 RTb (Trouton’s rule)
RTb

The Clapeyron equation becomes


dP
=
Hm ( 10.5 ) R = ( 10.5 ) R = 10.5 P
dT Tb  vapVm Vm, vap RTb / P Tb

This gives
10.5 P (10.5)(760 Torr)
Tb = = = 306.9 K
dP / dT (52 Torr)/(2 K)

Also,
 vap Hm = 10.5 RTb = (10.5)(8.314 J K -1 mol-1 )(306.9 K)
= 26 791 J mol-1

C. The Clausius–Clapeyron Equation


The Clausius–Clapeyron equation is obtained by integrating the Clapeyron equation in the case
that one of the two phases is a vapor (gas) and the other is a condensed phase (liquid or solid).

For vaporization and sublimation Clausius showed how the Clapeyron equation may be
simplified by assuming (a) that the vapor obeys the ideal gas law and (b) by neglecting the molar
volume of the liquid or solid in comparison with the molar volume of the gas.

For a liquid–vapor transition with our approximations:


Module 1 – Phase Equilibrium 30
Engr. Caesar Pobre Llapitan

Vm = Vm( ) − Vm( )  Vm( ) 


gas liq gas RT
P

The same approximation holds for a solid-vapor transition.

Thus, from Eq. (16), we obtain the derivative form of the Clausius–Clapeyron equation. For a
liquid-vapor transition

dP  vap Hm P  vap Hm dP  vap Hm dT


= = or = (18)
dT TVg RT 2 P R T2

where ΔvapHm is the molar enthalpy change of vaporization. For sublimation (a solid-vapor
transition), ΔvapHm is replaced by ΔsubHm, the molar enthalpy changes of sublimation. We can
omit the subscript and apply the equation to either case.

Equation (18) is known as the Clausius-Clapeyron equation. It relates the vapor pressure of a
condensed phase (solid or liquid) with its molar enthalpy of phase transformation (sublimation
or vaporization) and the temperature of transformation.

Rewriting Eq. (18), we obtain

dP P  vap H
= d ln 0 = dT (19)
P P RT 2

where P0 is the standard pressure used. Integrating on the assumption that H vap is
independent of temperature and pressure yields

P  vap H
 d ln P 0 = T
−2
dT (20)
R

P  vap H
ln =− +C (21)
P0 RT

where C is the integration constant. This suggests that a plot of ln(P/P0) versus 1/T should be
linear, and this is borne out by data on both vaporization and sublimation.

Figure 6: Vapor pressure of water (FP = freezing point, BP = boiling point)


(Retrieved from Silbey, 2005)

Vapor pressure data for pure substances can be found in ChE Handbook and other references.
Module 1 – Phase Equilibrium 31
Engr. Caesar Pobre Llapitan

For convenience, Eq. (18) is integrated between limits P2 and T2 and P1 and T1 as follows:

P2 / P 0 trs Hm T2 −2
P
P /P
1
0
P
d ln
R 0 T1
=T dT
P  H  1  1 
ln 2 = trs m  − −  −  
P1 R  T2  T1  

P2 trs H (T2 − T1 )
ln = (22)
P1 RT1T2

 The Clausius-Clapeyron equation is very useful in considering gas-phase equilibria:


- It helps predict equilibrium pressures at differing temperatures.
- It can predict what temperature is necessary to generate a particular pressure.
- Pressure/temperature data can be used to determine the change in enthalpy for
the phase transition.

Note:
o Over narrow ranges of temperatures, the enthalpy of vaporization can be taken to be a
linear function of temperature.
o In calculating vapor pressures over a wider range of temperature, we have to recognize
that the enthalpy of vaporization approaches zero as the temperature approaches the
critical temperature.

If P1 = 1 bar, then T1 represents the normal sublimation (or boiling) point of the solid (or liquid)
phase. Representing this by T*, we get

P  H 1 1 
ln = − trs m  − *  (23)
1 bar R T T 

where the subscript 2 has been dropped.

If we plot ln (p/bar) (or log p/bar) versus 1/T, we get a linear curve with slope equal to
− trs Hm R (or − trs Hm 2.303 R ) as shown in Figure 7.

Figure 7: Plot of log (p/p0) versus 1/T for the condensed phase-vapor equilibria
(Retrieved from Kapoor, 2015)

- The intercept at 1/T = 0 yields the value of trs Hm RT * (or trs Hm 2.303 RT * ).
From the determined values of slope and intercept, we can determine both trs Hm
and T* of the substance under study.
Module 1 – Phase Equilibrium 32
Engr. Caesar Pobre Llapitan

To represent the vapor pressure as a function of temperature over a wide range of temperature,
it is necessary to take the temperature dependence of  vap H into account. This can be
represented by

 vap H = A + BT + CT 2 (24)

Figure 8: The enthalpy of vaporization approaches zero as the temperature


approaches the critical temperature. (Retrieved from Silbey, 2005)

We write Eq. (19) by inserting Eq. (23) to obtain

dPvap  vap H 1 A B 
= 2
dT =  2 + + C  dT (25)
Pvap RT RT T 

Integrating,

1 A 
ln Pvap =  − + B ln T + CT + D  (26)
R T 

where D is the integration constant. Thus, vapor pressures determined experimentally over a
range of temperatures can be represented by values of A, B, C, and D determined by curve fitting
to minimize the sum of the squares of the deviations from the experimental values.

A number of enthalpies and entropies of fusion and vaporization are given in the references.

Trouton’s rule
If the enthalpy change of a particular substance is not known, and one wishes to estimate the
vapor pressure of a liquid at one temperature from knowledge of the vapor pressure at another
temperature, Trouton’s rule can be used as an approximation.

 vap Hm
=  vap Sm  88 J K -1 mol-1 (for liquids) (27)
Tb

 The average of  vap Sm for a large number of liquids that are not appreciably hydrogen-
bonded bears out the value given in Eq. (27)

The effect of hydrogen bonding is seen in the case of water. The abnormally low value
of acetic acid may be explained by its appreciable association in the vapor state. Allowing
for its apparent molecular weight of about 100 in the vapor state, acetic acid will have a
value of  vap Sm of approximately 100 J K-1mol-1, in line with other associated liquids.
Module 1 – Phase Equilibrium 33
Engr. Caesar Pobre Llapitan

Hildebrand rule
An alternative rule, known as the Hildebrand rule (1915, 1918) and named after the American
physical chemist Joel Henry Hildebrand (1881 – 1983), states that the entropies of vaporization of
unassociated liquids are equal, not at their boiling points but at temperatures at which the vapors
occupy equal volumes.

 All liquids with fairly symmetrical molecules possess equal amounts of configurational
entropy. This means that, on the molecular level, maximum molecular disorder exists in
a liquid that adheres to the rule.

Craft’s rule
Normally, boiling points are recorded at 101.325 kPa (1 atm), whereas they seldom are obtained
experimentally under this exact pressure.

A useful expression for correcting the boiling point to the standard pressure was derived in 1887,
by James Mason Crafts (1839 – 1917) by combing Eq. (18) with Trouton’s rule. Assuming that
dP/dT is approximately P/T, we obtain

P  vap Hm P
=  (28)
T Tb RTb

where Tb is the normal boiling point.

 vap Hm
For ordinary liquid, we may substitute  vap Sm = 88 JK-1mol-1 for :
Tb

P 88 (J K -1 mol-1 )  101.325 (kPa) 1072


= = kPa
T 8.3145 (J K -1 mol-1 ) Tb Tb
or
T  9.3  10-4 Tb P/kPa

For associated liquids, a numerical coefficient of 7.5 × 10-4 gives better results than 9.3 × 10-4,
which is used for normal liquids.

Sample Problems:
1. The vapor pressure of toluene is 0.078 8 bar at 313.75 K and 0.398 bar at 353.15 K.
Calculate the molar enthalpy of vaporization.

Solution:
The given data are
T1 = 313.75 K T2 = 353.15 K
P1 = 0.0788 bar P2 = 0.398 bar

Substituting these values in the expression


P  vap H (T2 − T1 )  vap Hm  1 1 
ln 2 = =−  − 
P1 RT1T2 R  T2 T1 

we get
0.398 bar  vap Hm  1 1 
ln =−  − 
0.0788 bar 8.314 J K -1 mol-1  353.15 K 313.75 K 
Module 1 – Phase Equilibrium 34
Engr. Caesar Pobre Llapitan

Thus,
8.314 J K -1 mol-1 (353.15 K)(313.75 K) 0.398 bar
 vap Hm = = ln
(353.15 K − 313.75 K) 0.0788 bar
= 37 866 J mol-1

2. A certain liquid has a boiling point of 338.15 K at 1.013 25 bar pressure. Using Trouton’s
rule (a) estimate the vapor pressure at 325.15 K, (b) estimate the boiling point at a
pressure of 0.267 bar, and (c) obtain an equation for the vapor pressure in bar of this
liquid as a function of temperature.

Solution:
(a) The given data are
T1 = 338.15 K P1 = 1.013 25 bar
T2 = 325.15 K P2 = ?
vapHm = 10.5 RT1 (Trouton’s rule)

Substituting the given data in the expression


P  vap Hm  1 1 
ln 2 = −  − 
P1 R  T2 T1 

we get
P2  1 1  T 
ln = − ( 10.5 T1 )  −  = − ( 10.5 )  1 − 1 
1.01325 bar  T2 T1   T2 
 338.15 K 
= − ( 10.5 )  − 1  = − 0.42
 325.15 K 
P2 = e −0.42 (1.01325 bar) = 0.666 bar

(b) In this case, we have


0.267 bar  338.15 K 
ln = − (10.5)  −1
1.01325 bar  T2 
 338.15 K 
−1.334 = − (10.5)  − 1
 T2 
338.15 K 1.334 11.834
= +1=
T2 10.5 10.5
 10.5 
T2 =   (338.15 K) = 300.00 K
 11.834 

(c) Substituting the values of T1 and P1 in the expression


P T 
ln 2 = − ( 10.5 )  1 − 1 
P1  T2 

we get
 P2   338.15 K 
ln   = − ( 10.5 )  − 1
 1.01325 bar   T2 
Module 1 – Phase Equilibrium 35
Engr. Caesar Pobre Llapitan

 P  3 550.58 K 3 550.58 K
ln  2  = ln ( 1.01325 ) − + 10.5 = 0.0132 − + 10.5
 bar  T2 T2
3 550.58 K
= 10.51 −
T2

 P  1 541.72 K
Or log  2  = 4.564 −
 bar  T2

3. If the enthalpy of vaporization of water is assumed to be constant at 2 255 kJ g–1, calculate


the temperature at which water will boil under a pressure of 77.0 cmHg, the boiling
point of water being 373.15 K at 76.0 cmHg. The specific volume of water vapor at 373.15
K and 76 cmHg is 1 644 cm3 g–1 and that of liquid water is 1 cm3 g–1.

Solution:
The given data are
T1 = 373.15 K P1 = 76.0 cmHg
T2 = ? P2 = 77.0 cmHg
vapHm = (2 255 J g ) (18 g mol-1) = 40 590 J mol-1
-1

Vm, v = (1 664  18) cm3 mol-1 = 29 952 cm 3 mol-1


= 29.952 dm3 mol-1
Vm, l = 1  18 cm 3 mol-1 = 0.018 dm 3 mol-1

P2  vap Hm  1 1 
Using the expression ln =−  − 
P1 R  T2 T1 

we get
 77.0 cmHg  (40 590 J mol-1 )  1 1 
2.303 log  =− -1 -1  − 
 76.0 cmHg  (8.314 J K mol )  T2 373.15 K 

Solving for T2 we obtain 373.5 K.

4. Iodine boils at 456.15 K and the vapor pressure at 389.65 K is 100 Torr. The enthalpy of
fusion is 15.65 kJ mol–1 and the vapor pressure of the solid is 1 Torr at 311.85 K. Calculate
the triple point temperature and pressure.

Solution:
The given data are
o Liquid-Vapor Equilibrium
T1 = 456.15 K P1 = 760 Torr
T2 = 389.65 K P2 = 100 Torr

o Solid-Liquid Equilibrium
fusHm = 15.65 kJ mol-1

o Solid-Vapor Equilibrium
T = 311.85 K P = 1 Torr

▪ Calculation of vapHm using the Clausius-Clapeyron equation:


Module 1 – Phase Equilibrium 36
Engr. Caesar Pobre Llapitan

P2  vap Hm  1 1 
log =−  − 
P1 2.303 R  T2 T1 
100  vap Hm  1 1 
log =− -1 -1  − 
760 2.303(8.314 J K mol )  389.65 K 456.15 K 
 vap Hm = 45 076 J mol-1

▪ Value of subHm:
We use Eq. (17); thus,

 sub Hm =  fus Hm +  vap Hm = 15.65  10 3 J mol-1 + 45 076 J mol -1


= 60 726 J mol-1

▪ Calculation of Triple Point Temperature and Triple Point Pressure


Let p and T be the pressure and temperature at the triple point, respectively. At this
point, the vapor pressure of liquid will be the same as that of solid. Hence, applying
Clapeyron equation at the triple point to liquid-vapor and solid-vapor equilibria, we get

o Liquid-Vapor Equilibrium
 p  45 076 J mol-1 1 1 
log   = − -1 -1  − 
 100 Torr  2.303 (8.314 J K mol )  T 389.65 K 
(1)
1 1 
= -(2 354.2 K)  − 
 T 389.65 K 

o Solid-Vapor Equilibrium
 p  60 762 J mol-1 1 1 
log  =− -1 -1  − 
 1 Torr  2.303 (8.314 J K mol )  T 311.85 K 
(2)
1 1 
= -(3 171.5 K)  − 
 T 389.65 K 

Solving Eqs. (1) and (2) we get

T = 384.07 K and p = 81.7 Torr

5. The vapor pressures of solid and liquid white phosphorus are given by the expressions

 p  2 875 K  p  2 740 K
log  2  = − + 5.36 and log  1  = − + 4.95
 atm  T  atm  T

respectively. Calculate (a) the temperature and pressure of the triple point of
phosphorus, and (b) the molar enthalpy and molar entropy of fusion of phosphorus at
the triple point.

Solution:
(a) At triple point, the vapor pressures of solid and liquid will be identical. Equating the
expressions of vapor pressure, we have
 p   p 
log  2  = log  1 
 atm   atm 
Module 1 – Phase Equilibrium 37
Engr. Caesar Pobre Llapitan

2 875 K 2 740 K
or − + 5.36 = − + 4.95
T T

2 875 K 2 740 K
or − = 5.36 − 4.95
T T

Solving T, we obtain that T = 329.3 K

The vapor pressure at the triple point can be obtained by substituting this temperature
in either of the vapor pressure expressions. Thus, we have
 p  2 875 K
log  2  = − + 5.36 = -8.732 + 5.26 = -3.372
 atm  329.3 K
p2 = 4.246  10-4 atm

(b) The expression of variation of vapor pressure with temperature is given by Eq. (21)

 P  Hm  P  Hm
ln  0 = − + C or log  = − + C'
 P = 1 atm  RT  1 atm  2.303 RT

Comparing this expression with the given expressions of vapor pressure, we get
 p  2 875 K  p  2 740 K
log  2  = − + 5.36 and log  1  = − + 4.95
 atm  T  atm  T
 sub Hm  vap Hm
= 2 875 K and = 2 740 K
2.303 R 2.303 R

Thus,
sub Hm = (2.3032)(8.314 J K-1 mol-1 )(2 875 K) = 55 048 J mol-1
 vap Hm = (2.3032)(8.314 J K-1 mol-1 )(2 740 K) = 52 463 J mol-1

Hence:
 fus Hm =  sub Hm −  vap Hm = (55 058 - 52 463) J mol-1
= 2 585 J mol-1

Since G = 0 for the fusion at equilibrium temperature, therefore


 fus Hm 2 585 J mol-1
 fus Sm = = = 7.85 J K -1 mol-1
T 329.3 K

6. From the fG0 of Br (g) at 25 C, calculate the vapor pressure of Br (l). The pure liquid at
1 bar and 25 0C is taken as the standard state.

Solution
PBr2
Br2 (l) Br2 (g) K=
P0

PBr2
G0 = − RT ln
P0
Module 1 – Phase Equilibrium 38
Engr. Caesar Pobre Llapitan

PBr2  G 0   3 110 J mol-1 


= exp  −  = exp  − 
P0  RT 
-1 -1
 (8.314 J K mol )(298.15 K) 
PBr2 = 0.285 bar

7. Calculate G0 for the vaporization of water at 0 0C using data in Table C.2 (Silbey, 2005)
and assuming that H0 for the vaporization is independent of temperature. Use G0 to
calculate the vapor pressure of water at 0 0C.

Solution:
H2O (l) H2O (g)

G 0 (298.15 K) = -228.572 + 237.129 = 8.577 kJ mol-1


H 0 (298.15 K) = -241.818 + 285.830 = 44.012 kJ mol-1

 G  H
From the equation d   = − 2 dT :
 T  T

Integration results into


G2 G1  1 1 
− = − H  − 
T2 T1  T1 T2 

Hence,
G1T2  T 
- G2 = + H  1 − 2 
T1  T1 

 273.15 K  -1  273.15 K 
G (00 C) = 8.577 kJ mol-1   + 44.012 kJ mol  1 − 
 298.15 K   298.15 K 
= (7.858 + 3.690) kJ mol-1 = 11.548 kJ mol-1

PBr2
- G0 = − RT ln
P0
PH2O
G 0 (00 C) = − RT ln
P0
 G 0 (00 C)   11.548 
PH2O = P 0 exp  −  = P exp  −
0

 RT  (8.314)(273.15) 
 
= 6.190  10-3 bar

8. The normal boiling point of toluene is 110.62 °C. Estimate its vapor pressure at 80.00 °C
assuming that toluene obeys Trouton’s rule.

Solution:
Trouton’s rule is given by Eq. (27):
 vap Hm
=  vap Sm  88 J K -1 mol-1
Tb
Module 1 – Phase Equilibrium 39
Engr. Caesar Pobre Llapitan

Rearranging, we can solve for  vap Hm at TB= 110.62 °C:

(
 vap Hm = 88 J K -1 mol-1 Tb )
= ( 88 J K -1
mol-1 ) ( 383.77 K )
= 33 771.76 J mol-1

Using this value for  vap Hm , we can obtain Pvap the Clausius-Clapeyron equation.
P2  vap H  1 1 
ln =  − 
P1 R  T1 T2 
  vap H  1 1 
P2 = P1 exp   − 
 R  T1 T2  

where P1= 1 atm under standard conditions. Thus,


 33 771.76 J mol-1  1 1 
P2 = (1 atm) exp  -1 -1  − 
 8.314 J K mol  383.77 K 353.15 K  
= 0.399 4 atm

9. 2-Propanone (acetone) boils at 329.35 K at 1 atm of pressure. Estimate its boiling point
at 98.5 kPa using Crafts’ rule.

Solution:
Crafts’ rule is given by Eq. (28)
P  vap Hm P
= 
T Tb RTb

First, we can solve for ΔT then we will be able to obtain T.


 RT   Tb 
T =  b    P
 P    vap Hm 
 (329.35 K)(8.314 J K -1 mol -1   1 
T =     (98 500 - 101 325) Pa
 98.5  10 Pa3 -1 -1
   88 J K mol 
T = -0.892 469 K

T is given by:
T = (329.35 – 0.892 469) K = 328 K

Classroom Activity 2
1. The density of rhombic sulfur at 200C is 2.07 g cm−3, and the density of monoclinic sulfur
at 200C is 1.957 g cm−3. Assume that you can use these values at 250C. The Gibbs energy
of formation of monoclinic sulfur is 0.096 kJ mol−1, and the Gibbs energy of formation of
rhombic sulfur is equal to zero by definition.
a. Estimate ΔG for producing 1.000 mol of monoclinic sulfur from rhombic sulfur at
101.0 bar and 298.15 K. Assume that ΔVm is independent of pressure.
b. Is there a pressure at which the two forms could coexist at equilibrium at 298.15K?
Module 1 – Phase Equilibrium 40
Engr. Caesar Pobre Llapitan

2. a. The vapor pressure of solid water (ice I) is equal to 2.149 torr at −100C and to 4.579 torr
at 00C. Find the average enthalpy change of sublimation of ice for this range of
temperature.
b. The vapor pressure of water is equal to 23.756 torr at 250C. Calculate the average
enthalpy change of vaporization for the range of temperature from 00C to 250C.
c. Find the enthalpy change of fusion of water.
3. The triple point of ammonia is at 196.2K and 49.42 torr. The molar enthalpy change of
vaporization is equal to 24.65 kJ mol−1 at this temperature.
a. Find the normal boiling temperature of ammonia.
b. The actual boiling temperature is −330C. Find the average value of the molar enthalpy
change of vaporization for the range between the triple point and the normal boiling
temperature.

Exercise 2
- To be chosen from Mortimer (2008) and Silbey (2005)

D. Classification of Transitions in Single-Component Systems


A given phase transition of a substance may be classified into first-order or second-order
depending upon the following characteristics.

First-Order Phase Transition


A first-order phase transition of a substance is accompanied with changes in the values of
enthalpy and volume of the substance, i.e. trsH and trsV are nonzero. Examples include familiar
phase transitions such as solid to liquid, solid to vapors, and liquid to vapors.

For a substance B, we have:

  B    B 
  = Vm, B and   = − Sm, B
 p T  T  P

If the substance B undergoes phase transition  → , we will have

 B   B (  )   B ( ) 


trsVm, B =   =   −  0 (29)
 p T  p T  p T
and
  B 
trs Hm, B = T trs Sm, B = − T   (30)
 T  p

     
These implies that  B  and  B  have different values on either side of the transition
 p T  T  p
or in other words, these derivatives are discontinuous.

Thus, the main characteristics of first-order transitions are:


1. The chemical potential is a continuous function of temperature.
2. The derivatives of chemical potential are discontinuous.

Figure 9 depicts these characteristics.


Module 1 – Phase Equilibrium 41
Engr. Caesar Pobre Llapitan

Figure 9: Typical variations of (a) Vm versus T, (b) Hm versus T, (c) – μ versus T,


(d) Sm versus T, and (e) Cpm versus T of first-order transition (Retrieved from Kapoor, 2015)

The fact that ( B T ) p is discontinuous at the transition point leads to another important
characteristic that the first-order transition is accompanied by an infinite heat capacity at the
transition point. This may be understood from the following analysis.

 H B 
The heat capacity of the substance B is given by the relation C p, B =   .
 T  p

Now since the transition takes place at constant temperature, the heat supplied to the system
is used in driving the transition (that is why H is discontinuous) rather than in raising the
temperature of the system. Thus, at the transition point, we have
( H B ) p = positive
( T )T = 0

Hence,
 H B  positive
C p, B =   = =
 T  p 0

Second-Order Phase Transition


In a second-order phase transition of a substance B, the chemical potential and its derivatives
(  B p )T and ( B T ) p are continuous functions of temperature whereas the second
derivatives are discontinuous.

The fact that the first derivatives are continuous implies that changes in the values of enthalpy
and volume during transition are zero. In other words, the entropy, the enthalpy and the volume
of the system do not change when the transition occurs.

Although the entropy of the substance changes in a continuous manner with temperature, yet
the rate of change in disorder (i.e. dSB) alters at the phase transition point. Hence, the first
derivative ( Sm,B/T)p is a discontinuous function of temperature. Since
Module 1 – Phase Equilibrium 42
Engr. Caesar Pobre Llapitan

  2 B   Sm, B  C pm, B
−  =   =
 T 2
  p  T p T

it follows that the second derivative of μB and Cp,m,B are discontinuous functions of temperature.

Similarly, from the relations

  2 B   Vm,B 
−
 p2  = −   = T Vm,B (κT is isothermal compressibility)
 p  p T
 2  B  Vm,B 
=  =  Vm,B (α is cubic expansion coefficient)
T p  T  p

and the fact that the rates of variation of Vm with T and p alter at the phase transition, indicate
that the second derivatives (2μB/p2) and (2μB/Tp) are discontinuous.

However, C p ,m does not become infinite at the critical point as in the case of first-order
transitions. The sudden appearance of superconductivity in certain metals when they are cooled
to low temperatures is an example of such a second-order transition.

The various characteristics of second-order transition discussed above are illustrated in Fig. 9.
The heat capacity versus temperature curve often has the shape of Greek , and because of this
reason the second-order transition is frequently referred to as a -transition.

Figure 10: Typical variations of (a) Vm versus T, (b) Hm versus T, (c) – μ versus T, (d) Sm versus T, and
(e) Cpm versus T of the second-order transition (Retrieved from Kapoor, 2015)

E. Variation of Vapor Pressure with External Pressure


The vapor pressure that we have discussed thus far is measured with no other substances
present. We are often interested in the vapor pressure of a liquid that is open to the atmosphere.
The other gases in the atmosphere exert an additional pressure on the liquid that modifies its
vapor pressure.
Module 1 – Phase Equilibrium 43
Engr. Caesar Pobre Llapitan

If an external pressure is applied in addition to the saturated vapor pressure, the vapor pressure
becomes a function of pressure as well as of the temperature. The increased pressure may be
applied by adding an inert gas that is insoluble in the liquid or through the action of a piston
that is permeable to the gas.

Consider the piston and cylinder assembly containing water at 25°C shown in Figure 10.

Figure 11: A piston and cylinder assembly at 298 K is shown with the contents being pure water
(a) Pexternal = 1.00 bar; (b) Pexternal = 0.0316 bar; and (c) Pexternal = 1.00 bar (Retrieved from Engel, 2013)

(a) The weightless piston is loaded with a mass sufficient to generate a pressure of 1 bar,
the system is in the single-phase liquid region of the phase diagram
(b) The mass is reduced so that the pressure is exactly equal to the vapor pressure of
water. The system now lies in the two-phase liquid–gas region of the phase diagram
described by the liquid-gas coexistence curve. The temperature of the system
remains constant
(c) Keeping the temperature constant, enough argon gas is introduced into the cylinder
such that the sum of the argon and H2O partial pressures is 1 bar.

Question: What is the vapor pressure of water in this case, and does it differ from that for the
system shown in Figure 10b?

For a closed system, with T constant and the total pressure Pt equal to the pressure exerted on
the liquid, the Gibbs equation may be written as

dP Vm (l )
Vm (v) dP = Vm (l ) dPt or = (31)
dPt Vm (v)

where P is the pressure of the vapor under the total pressure Pt. If we assume that the vapor
phase behaves ideally, we may substitute Vm(v) = RT/P, obtaining

d ln P Vm (v)
= (32)
dPt RT

Since the molar volume of a liquid does not change significantly with pressure, we may assume
Vm(l) to be constant through the range of pressure to obtain, by integrating,

P Vm (l )
ln = ( Pt − Pv ) (33)
Pv RT

where Pv is the saturated vapor pressure of the pure liquid without an external pressure. Note
that in this case the vapor pressure is now dependent on the total pressure as well as on the
temperature. This is because the two phases exist under different pressures.
Module 1 – Phase Equilibrium 44
Engr. Caesar Pobre Llapitan

Example:
1. The vapor pressure of water at 27.5 °C, a calibration temperature for glassware used in
warmer climates, is 27.536 Torr under its own vapor pressure. Calculate the vapor
pressure of water under an air pressure of 1.00 atm. Assume that air is inert. The density
of water at 27.5 °C is 996.374 g dm–3.

Solution:
To calculate Pvap we may use Eq. (33) which states that:
P V (l )
ln = m ( Pt − Pv )
Pv RT

Rearranging this expression to isolate for the vapor pressure gives,


 V (l ) 
Pvap = Pv exp  m ( Pt − Pv ) 
 RT 

M
where Vm = = 0.018 080 841 1 dm 3 mol-1

Then,
 0.018 080 841 1 dm 3 mol-1 
Pvap = (27.536 Torr)exp 
 (0.082 06 dm 3 K -1 mol-1  760 Torr)(300.65 K) ( 760-27.536 ) Torr 
 
Pvap = 27.555 Torr

This is a small correction, but may be necessary for accurate work.

SPECIAL TOPIC: Effect of Surface Tension on The Vapor Pressure


- Section 8.8: Engel, Thomas, Philip Reid and Warren Hehre (2013). Physical
Chemistry, 3rd Edition. Pearson Education, Inc.
- Section 6.9: Silbey, Robert J., Robert A. Alberty and Moungi G. Bawendi (2005).
Physical Chemistry, 4th Edition. John Wiley & Sons, Inc

IV. EQUILIBRIA IN MULTIPLE-COMPONENT SYSTEMS

- Section 5.6: Mortimer, Robert G. (2008). Physical Chemistry, 3rd Edition. Elsevier
Inc.

References:
1. Ball, David W. (2015). Physical Chemistry, 2nd Edition. Cengage Learning
2. Engel, Thomas, Philip Reid and Warren Hehre (2013). Physical Chemistry, 3rd Edition.
Pearson Education, Inc.
3. K. L. Kapoor (2015). A Textbook of Physical Chemistry, Volume III (Applications of
Thermodynamics), 4th Edition. McGraw Hill Education (India) Private Limited
4. Laidler, Keith J and John S. Meiser (1999). Physical Chemistry, 3rd Edition. Houghton
Mifflin Company
5. Mortimer, Robert G. (2008). Physical Chemistry, 3rd Edition. Elsevier Inc.
6. Silbey, Robert J., Robert A. Alberty and Moungi G. Bawendi (2005). Physical Chemistry,
4th Edition. John Wiley & Sons, Inc

You might also like