You are on page 1of 59

UNIVERSITY OF MINES AND TECHNOLOGY

(UMaT), TARKWA, GHANA

FACULTY OF GEOSCIENCES AND ENVIRONMENTAL STUDIES


(FGES)

COURSE:

Physical and Analytical Chemistry


(Lecture Notes 1)
COURSE CODE:

GL_MN 154
Course Instructor:

Thomas Wi-Afedzi (PhD)

July, 2022
About this Course

Research in Science (Chemistry) and Engineering can be viewed as proceeding along two (2) main
directions: theoretical and experimental.

‘Experimental’ involves the science of observation. A question is posed to nature, corresponding


experiments are conducted and observations are made to know the outcome. Example, if we mix
these two substances what will happen?

‘Theoretical’ on the other hand, is concerned with building general frameworks for placing large
numbers of individual observations into a rational order. This can be achieved by developing
mathematical models, deducing general laws, which can later be used in predicting future
experiments. The tools of theory are usually mathematics and physics applied to chemical
problems.

Without theory, experimental science is just a vast of catalog of irrationalized observations and
without experiments, theory is simply a long list of conjectures (unproven; with no verification or
falsification). With this in mind, the first part of this course, Physical Chemistry, will seek to
review some important principles and concepts bordering mostly macroscopic science and
properties of atoms and molecules whereas the second part, Analytical Chemistry, is a
measurement science.

i
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Table of Contents
About this Course ......................................................................................................................... i
1 PERIODICITY ........................................................................................................................ 4
1.1 Introduction to the periodic table ..................................................................................... 4
1.1.1 Group and Period ...................................................................................................... 4
1.2 Periodicity of physical properties ..................................................................................... 4
1.2.1 Atomic Radius: (Covalent Atomic Radii)................................................................. 5
1.2.2 Ionization Energy ...................................................................................................... 5
1.2.3 Electron Affinity ....................................................................................................... 6
1.2.4 Electronegativity ....................................................................................................... 6
2 PROPERTIES OF GASES, LIQUIDS AND SOLIDS ........................................................... 8
2.1 Gases ................................................................................................................................ 8
2.1.1 Boyle’s Law .............................................................................................................. 8
2.1.2 Charles’ Law ............................................................................................................. 8
2.1.3 Avogadro’s Law........................................................................................................ 9
2.1.4 The Ideal Gas Law .................................................................................................... 9
2.1.5 Gas Stoichiometry (Molar Volume) ....................................................................... 10
2.1.6 Molar Mass (M) ...................................................................................................... 11
2.1.7 Dalton’s Law of Partial Pressures ........................................................................... 11
2.1.8 Mole Fraction (  ) .................................................................................................. 12
2.1.9 Facts from the Kinetic Molecular Theory of Gases ................................................ 13
2.1.10 Effusion and Diffusion ............................................................................................ 14
2.1.11 Graham’s Law of Diffusion .................................................................................... 14
2.1.12 Real Gases ............................................................................................................... 15
2.2 Liquids............................................................................................................................ 16
2.2.1 Introduction ............................................................................................................. 16
2.2.2 Surface Tension (γ, σ or T) ..................................................................................... 16
2.2.3 Capillary Action ...................................................................................................... 16
2.2.4 Viscosity ................................................................................................................. 17
2.2.5 Vapour Pressure and Phase Changes ...................................................................... 17
2.3 Solids .............................................................................................................................. 19
1|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

2.3.1 Introduction ............................................................................................................. 19


2.3.2 Types of Solids ....................................................................................................... 19
2.3.3 Types of Crystalline Solids ..................................................................................... 20
2.3.4 Metal Alloys............................................................................................................ 20
3 THERMOCHEMISTRY ....................................................................................................... 22
3.1 Introduction .................................................................................................................... 22
3.2 Exothermic and Endothermic Reactions ........................................................................ 22
3.3 Standard Enthalpy Change for a Reaction ..................................................................... 23
3.1.1 Standard enthalpy change of formation for a compound, ∆H°f .................................... 25
3.4 Heat Capacity ................................................................................................................. 26
3.5 Second Law .................................................................................................................... 31
3.5.1 Entropy.................................................................................................................... 32
3.5.2 Thermodynamic Equations and Remarks ............................................................... 34
3.6 Free Energy .................................................................................................................... 36
3.7 Third Law ....................................................................................................................... 40
4 CHEMICAL EQUILIBRIUM ............................................................................................... 41
4.1 Introduction .................................................................................................................... 41
4.2 Equilibrium Constant & the Law of Mass Action ......................................................... 41
4.3 Characteristics of the Equilibrium Constant .................................................................. 41
4.3.1 Equilibrium Expressions Involving Pressures ........................................................ 42
4.4 The Concept of Activity ................................................................................................. 42
4.5 Heterogeneous Equilibria ............................................................................................... 42
4.6 Applications of the Equilibrium Constant...................................................................... 44
4.6.1 6.4.1 The Extent of a Reaction................................................................................ 44
4.7 Reaction Quotient ........................................................................................................... 45
4.8 Solving Equilibrium Problems ....................................................................................... 46
4.9 Le Chatelier’s Principle .................................................................................................. 47
4.9.1 Effect of Change in Concentration ......................................................................... 47
4.9.2 Effect of a Change in Pressure ................................................................................ 49
4.9.3 The Effect of a Change in Temperature .................................................................. 50
4.10 Free Energy and Equilibrium ......................................................................................... 51

2|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

4.11 Temperature Dependence of K ...................................................................................... 51


4.12 Ionic Equilibrium ........................................................................................................... 53
4.12.1 Ionic Product, Q ...................................................................................................... 53
4.12.2 Solubility Product (Ksp) .......................................................................................... 53
4.12.3 Ionic Product and Precipitation ............................................................................... 55
4.12.4 Solubility (Product) and pH of Solution ................................................................. 56
4.12.5 Percent Dissociation................................................................................................ 56

3|Page
CHAPTER ONE

1 PERIODICITY

1.1 Introduction to the periodic table

The periodic table was originally constructed to represent the patterns observed in the chemical
properties of the elements. The array of elements and properties observed was bewildering.
Gradually, however, patterns were noticed.

1.1.1 Group and Period


A group is a set of elements in the same vertical column of the periodic table. Elements in the
same group have the same valence electron configuration and often show similar chemical
behaviour. Elements in the same period (horizontal row) of the periodic table (Figure 1.1) have
the same number of electronic shells.

Figure 1.1 The Periodic table of Elements

1.2 Periodicity of physical properties

The physical properties of the elements show a striking periodicity. This is particularly clear when
we examine the variation of atomic sizes and the energies needed to remove electrons from atoms.

4|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

1.2.1 Atomic Radius: (Covalent Atomic Radii)

2r
The radius of an atom (r) is defined as half the distance between the nuclei in a molecule consisting
of identical atoms.

Atomic radius decreases in going from left to right across a period. This decrease can be explained
in terms of the increasing effective nuclear charge (decreasing shielding) in going from left to
right. This means that the valence electrons are drawn closer to the nucleus, decreasing the size of
the atom.

Atomic radius increases down a group, because of the increases in the orbital sizes in successive
principal quantum levels.

1.2.2 Ionization Energy


It is the energy required to remove an electron from the ground state of a gaseous atom or ion,

E (g) E+(g) + e-(g)

Where, the atom or ion is assumed to be in its ground state.

Units: kJ/ mol.

When given in eV per atom, the term ionization potential is used. (1 eV = 1.602 × 10-19 J).

a) Going across a period from left to right, the first ionization energy increases. This is because
electrons are put in the same quantum level (shell) and thus there is no shielding effect. Nuclear
charge and therefore effective nuclear charge increases. We expect electrons to be bound more
firmly in going from left to right across a given period. Thus, ionization energy should increase.

b) First ionization energy values decrease in going down a group. This is because down a group,
nuclear charge increases but electrons are put into new quantum levels. The shielding of outer
electrons results in a decrease in effective nuclear charge. Sizes of atom increase and ease of
removal of electrons thus increases down a group.

5|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

1.2.3 Electron Affinity


It is the energy change associated with the addition of an electron to a gaseous atom.

E (g) + e- → E-(g)

Electron affinity generally increases across the period, from left to right. However, there are
several exceptions to this rule in each period.

Down a group electron affinity generally decreases, since the electron is added at increasing
distances from the nucleus. However, in going down most groups the changes in electron affinity
are relatively small and numerous exceptions occur, e.g. group 7A elements (the halogens).

Table 1.5 Electron affinity values for the halogens

Atom Electron Affinity (kJ/mol)

F -327.8 *

Cl -348.7
Expected trend
Br -324.5

I -292.2

* Smaller energy release attributed to the small size of the 2p orbitals. Because the electrons must
be very close together in these orbitals, there are unusually large electron-electron repulsions.

1.2.4 Electronegativity
An atom is likely to form a cation if it has low ionization energy and a low electron affinity. On
the other hand, it is likely to form an anion if it has a high electron affinity and a high ionization
energy. This suggests the following pattern of behavior.

Ionization energy Electron affinity Behaviour

Low Low Form cation

High High Form anion

The table can be summarized by introducing the concept of electronegativity χ (chi), defined as
6|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

χ = ionization energy + electron affinity

When both ionization energy and electron affinity are low, χ is low; when both are high, χ is high.
This simplifies the table to

Electronegativity Behaviour

Low Form cation

High Form anion

7|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

CHAPTER TWO

2 PROPERTIES OF GASES, LIQUIDS AND SOLIDS

2.1 Gases

The component particles of gases are far apart and are in rapid, random motion, exerting relatively
small forces on each other. Gases have low density, high compressibility, and completely fill a
container.

2.1.1 Boyle’s Law


It states that at constant temperature the volume of a fixed mass of gas is inversely proportional to
the pressure.

1
V∝
P

Or,

k
V=
P

⇒ PV = Constant 2.1

Boyle’s law only approximately describes the relationship between pressure and volume for a gas.
Highly accurate measurements on various gases at a constant temperature have shown that the
product PV is not quite constant but changes with pressure. At pressures much higher than
atmospheric pressure, the changes in PV become very significant. A gas that obeys Boyle’s law is
called an ideal gas. Boyle’s law is only approximately followed for real gases.

It is commonly used to predict the new volume of a gas when the pressure is changed (at constant
temperature), or vice versa.

2.1.2 Charles’ Law


Charles’ law states that, at constant pressure, the volume of a gas is directly proportional to the
temperature.

V∝T

Or,

V = b. T

8|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Absolute Zero

Absolute zero has a special significance, as much evidence suggests that this temperature
cannot be attained. At temperatures below this point, the extrapolated volumes, from a plot of
V = b. T, would be negative. Temperatures of approximately 10-6 K have been produced in
laboratories, but 0K has never been reached.

2.1.3 Avogadro’s Law


Equal volumes of gases at the same temperature and pressure contain the same number of
“particles”.
Mathematically,
V = a. n 2.3
Where
V = volume of the gas
n = number of moles
a = proportionality constant
Thus, V  n
The equation states that for a gas at constant temperature and pressure, the volume is directly
proportional to the number of moles of gas.
This relationship is obeyed closely by gases at low pressures.

2.1.4 The Ideal Gas Law


The ideal gas law is obtained by combining the three laws considered previously, viz.,
Boyle’s law: V = k / P (@ constant T and n)
Charles’s law: V = b. T (@ constant P and n)
Avogadro’s law: V = a. n (@ constant T and P)
This combination will yield
 T.n 
V = R.   2.4
 P 
Where R is the combined proportionality constant called the universal gas constant. When the
pressure is expressed in atmospheres and the volume in liters, R has the value 0.08206 L atm mol-
1 -1
K .

9|Page
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

NB.
1 L atm mol-1 K-1 = 101.3 J.
R = 0.08206 L atm mol-1 K-1 × 101.3 J L-1 atm-1 = 8.3145 J K-1 mol-1.
Equation 3.4 can be rearranged to the more familiar form of the ideal gas law:
PV = nRT 2.5
The ideal gas law is an equation of state for a gas, where the state of the gas is its condition at a
given time. A particular state of a gas is described by its pressure, volume, temperature, and
number of moles. Knowledge of any three of these properties is enough to completely define the
state of a gas, since the fourth property can then be determined from the equation for the ideal gas
law. It is important to recognize that the ideal gas law is an empirical equation; based on
experimental measurements of the properties of gases. Gases that obey this law are said to behave
ideally. Most gases obey this law closely enough at pressures below 1 atm that only minimal errors
result from assuming ideal behaviour.

Example

A sample of hydrogen gas (H2) has a volume of 8.56 L at a temperature of 0°C and a pressure of
1.5 atm. Calculate the moles of H2 present in this sample.
Solution
From PV = nRT
PV
n=
RT
P = 1.5 atm; V = 8.56 L; T = 0°C = 273. K; R = 0.08206 L atm mol-1 K-1
(1.5atm)(8.56L )
Thus n =
(0.08206L atm mol -1
)
K -1 (273K )

= 0.57 mol

2.1.5 Gas Stoichiometry (Molar Volume)


Suppose we have 1 mole of an ideal gas at 00C (273.2 K) and 1 atm. From the ideal gas law, the
volume of the gas is given by:

nRT (1.0mol )(0.08206Latm−1mol −1 )(273.2 K )


V= = = 22.42 L
P 1.0atm

10 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

The volume of 22.42 L is called the molar volume of an ideal gas. The conditions of 0 °C and 1
atm, called the standard temperature and pressure (STP) are common reference conditions for the
properties of gases.
Many chemical reactions involve gases. By assuming ideal behaviour for these gases, we can carry
out stoichiometric calculations if the pressure, volume, and temperature of the gases are known.

2.1.6 Molar Mass (M)


One very important use of the ideal gas law is in the calculation of the molar (molecular weight)
of a gas from its measured density (ρ).
If the number of moles of is n, then:
massofgas m
n= =
molarmass M

Substituting into the ideal gas equation gives

m
RT
nRT
P= = M
V V

Or,

mRT
P=
V .M

m
But is the gas density, ρ, in units of grams per liter.
V

Thus

RT RT
P= ⇔M= 2.6
M P

Thus, if the density of a gas at a given temperature and pressure is known, its molar mass can be
calculated.

2.1.7 Dalton’s Law of Partial Pressures


For a mixture of gases in a container, the total pressure exerted is the sum of the pressures each
gas would exert if it were alone.
PTOTAL = P1 + P2 + P3 …
The pressures P1 + P2 + P3 … etc., are called partial pressures:

11 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Assuming ideal behaviour for each gas, the partial pressure gas can be calculated from the ideal
gas law:
n1 RT n RT n RT
P1 = ; P2 = 2 ; P3 = 3
V V V
Therefore,
Ptotal = P1 + P2 + P3 + …
n1 RT n RT n RT
= + 2 + 3 +…
V V V
nRT
= (n1 + n2 + n3 + …) ( )
V
nRT
Ptotal = ntotal ( ) 2.7
V
Where ntotal is the sum of the numbers of moles of various gases.
Thus, for a mixture of ideal gases, it is the total number of moles of particles that is important, not
the identity or composition of the individual gas particles.
The fact that the pressure exerted by an ideal gas is not affected by the identity (structure) of the
gas particles reveals two things about ideal gases:
1. The volume of the individual gas particles must not be important; and
2. The forces among the particles must not be important.

2.1.8 Mole Fraction (  )

It is the ratio of the number of moles of a given component in a mixture to the total number of
moles in the mixture.
For a given component in a mixture, the mole fraction (  1) is given by:

n1 n1
1= = 2.8
nTOTAL n1 + n2 + n3 + ...

From the ideal gas equation, we know that the number of moles of a gas is directly proportional to
the pressure of the gas, since
V
n=P( ) 2.9
RT
That is, for each component in the mixture,

12 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

V
n1 = P1 ( )
RT
V
n2 = P2 ( )
RT
V
n3 = P3 ( ), etc.
RT
Therefore, we can represent the mole fraction in terms of pressures:
n1 P1 (V / RT )
1= =
nTOTAL P1 (V / RT ) + P2 (V / RT ) + P3 (V / RT )

P1
=
P1 + P2 + P3 + ...

P1
 1= 2.10
PTOTAL

Thus, the mole fraction of a particular component in a mixture of ideal gases is directly related to
its partial pressure.
Equation 3.10 can be rearranged to give
𝑃1 = 𝜒1 ∙ 𝑃𝑇𝑂𝑇𝐴𝐿 2.11
That is the partial pressure of a particular component of a gaseous mixture is equal to the mole
fraction of that component times the total pressure.

2.1.9 Facts from the Kinetic Molecular Theory of Gases


1. The average kinetic energy of a collection of gas particles is directly proportional to the
Kelvin temperature of the gas.
3
(KE)avg = RT 2.12
2
3  PV 
= .  2.12 b
2 n 

2. Root Mean Square Velocity, Urms


The root mean square velocity of an ideal gas is defined by the expression

3RT 3RT
Urms = = 2.13
N A .m M
13 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

where
m = mass in kilograms of a single gas particle
M = NA. m = mass of a mole of gas particles in kilograms
NA = Avogadro’s number (number of particles in a mole = 6.022 × 1023 particles)
R = 8.3145 J K-1 mol-1
3. The Most Probable Velocity, Ump:
This is the velocity possessed by the greatest number of gas particles

2k BT 2 RT
Ump = = 2.14
m M
where
M = molar mass of the gas particles in kg = 6.022 × 1023 × m

R = gas constant = 6.022 × 1023 × kB

kB = Boltzmann’s constant = 1.38066 × 10-23 J/K

4. Average Velocity, Uavg

8 k BT 8 RT
Uavg = ū = = 2.15
.m .M

NB. These three velocities are not the same. They stand in the ratios
1.00: 1.128: 1.225

2.1.10 Effusion and Diffusion


Diffusion: the term used to describe the mixing of gases.
Effusion: the term used to describe the passage of a gas through a tiny orifice
into an evacuated chamber.

2.1.11 Graham’s Law of Diffusion


He found experimentally that the rate of effusion of a gas is inversely proportional to the square
root of the mass of its particles.

M2
(Rate of effusion for gas 1)/ (Rate of effusion for gas 2) =
M1

14 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

2.1.12 Real Gases


An ideal gas is a hypothetical concept. No gas exactly follows the ideal gas law, although many
gases come very close at low pressures and / or high temperatures.
For an ideal gas PV/nRT equals 1 under all conditions. However, for real gases PV/nRT
approaches 1 only at low pressures (typically ≤ 1 atm.)
The assumptions of the kinetic molecular theory need to be modified in order to fit the behaviour
of real gases.
Johannes van der Waals developed an equation for real gases in 1873. His equation is given by the
expression:
2
n
[Pobs + a   ](V – nb) = nRT 2.16
V 
Corrected pressure Corrected volume
Where:

Pobs = observed pressure

V = volume of the container

nb = volume correction
2
n
a   = pressure correction.
V 

Volume Correction

The ideal gas law assumes a hypothetical gas consisting of volumeless entities that do not interact
with each other. In contrast, a real gas consists of atoms or molecules that have finite volumes.
Thus, the volume available to a given particle in a real gas is less than the volume of the container,
because the gas particles themselves take up some of the space. To account for this discrepancy,
van der Waals represented the actual volume as the volume of the container, V, minus a correction
factor for the volume of the molecules, nb, where n is the number of moles of gas and b is an
empirical constant.

Pressure Correction

Attractions do occur among the particles in a real gas, which make the observed pressure Pobs
smaller than it would be if the gas particles did not interact. The size of the correction factor
n
depends on the concentration of gas molecules,   . For large numbers of particles, the number
V 

15 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

of interacting pairs of particles depends on the square of the number of particles and thus on the
2
n
square of the concentration,  
V 

2.2 Liquids

2.2.1 Introduction
The properties of liquids lie between those of gases and solids but not midway between. Extensive
attractive forces exist among the molecules in the liquid state, similar to but not as strong as those
in the solid state. The liquid and solid states show many similarities and are strikingly different
from the gaseous state. Liquids have low compressibility, lack of rigidity, and high density
compared with gases.

2.2.2 Surface Tension (γ, σ or T)


To increase a liquid’s surface area, molecules must move from the interior of the liquid to the
surface. This requires energy, since some intermolecular forces must be overcome.

The resistance of a liquid to an increase in its surface area is called the surface tension of the
liquid. This property of surface tension is responsible for liquid droplets (e.g. water) assuming a
spherical shape, rise in capillary tube and move through porous materials that they wet. Liquids
with relatively large intermolecular forces tend to have relatively high surface tensions. Surface
tension may be accurately determined by measuring the height to which a liquid will rise in a
capillary tube (capillary action). Mathematically,

1
𝛾 = ℎ𝜌𝑔𝑟
2
Where h = height, ρ = density, g = acceleration due to gravity, r = radius

2.2.3 Capillary Action


It is the spontaneous rise of a liquid in a narrow tube. Two (2) different forces are responsible for
this property:

1. Cohesive forces: the intermolecular forces among molecules of the liquid; and
2. Adhesive forces: the forces between the liquid molecules and their container.

Adhesive forces between a polar liquid and a given surface are strongest when the surface is
made of a substance that has polar bonds. For example, glass contains many oxygen atoms with
partial negative charges that are attractive to the positive end of a polar molecule such as water.
This ability of water to “wet” glass makes it creep up the walls of the tube where the water surface
touches the glass. This, however, tends to increase the surface area of the water, which is opposed
by the cohesive forces that try to minimize the surface area. Thus, because water has both strong
16 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

cohesive (intermolecular) forces and strong adhesive forces to glass, it “pulls itself” up a glass
capillary tube to a height where the weight of the column of water just balances the water’s
tendency to be attracted to the glass surface. The meniscus of water in a glass tube has a concave
shape indicating that water’s adhesive forces toward glass are stronger than its cohesive forces. A
non-polar liquid such as mercury shows a convex meniscus in a glass tube. This behaviour is
characteristic of a liquid in which the cohesive forces are stronger than the adhesive forces toward
the glass.

2.2.4 Viscosity
Another property of liquids that is strongly dependent on intermolecular forces is viscosity.
Viscosity is a measure of a liquid’s resistance to flow.

Liquids with large intermolecular forces tend to be highly viscous. For example, glycerol whose
structure is below (fig. 3.1) has an unusually high viscosity, mainly due to its high capacity to form
hydrogen bonds. Molecular complexity also leads to higher viscosity because very large molecules
can become entangled with each other.

Figure 3.1 Molecular structure of glycerol

2.2.5 Vapour Pressure and Phase Changes

2.2.5.1 Vaporization or Evaporation


One very familiar example of a change in state occurs when a liquid evaporates from an open
container. This is clear evidence that the molecules of a liquid can escape from the liquid’s surface
and form a gas. This process is called vaporization or evaporation and is endothermic, since
energy is required to overcome the relatively strong intermolecular forces in the liquid. The
endothermic nature of vaporization has great practical significance:

• Water acts as a coolant. Because of the strong hydrogen bonding among its molecules in
the liquid state, water has an unusually large heat of vaporization (40.7 kJ mol-1). A
significant portion of the sun’s energy that reaches the earth is spent evaporating water
from the oceans, lakes, and rivers rather than warming the earth. The vaporization of

17 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

water is also crucial to the body temperature control system through evaporation of
perspiration.

2.2.5.2 Standard Enthalpy of Vaporization

This is the energy required to vaporize 1 mole of liquid at a pressure of 1 atm. It is usually
symbolized as  0vap .

NB: Vapour is the usual term for the gas phase of a substance that exists as a solid or liquid at 25
°C and 1 atm.

2.2.5.3 Vapour Pressure


It is the pressure of a vapour present at equilibrium. At this pressure, the rate of condensation
equals the rate of evaporation.

The vapour pressure of a liquid can be measured by a simple barometer.

The vapour pressures of liquids vary widely:

• Liquids with high vapour pressures are said to be volatile.


• The vapour pressure of a liquid is principally determined by the size of the
intermolecular forces in the liquid. The greater the intermolecular forces the lower the
vapour pressure.
• Generally, substances with large molar masses have relatively low vapour pressures,
mainly because of the large dispersion forces.
• Vapour pressure increases significantly with temperature.

2.2.5.4 Dependence of Vapour Pressure on Temperature


The plot of vapor pressure as a function of temperature is not linear. However, using the Clausius-
Clapeyron equation (Eqn 3.24) a linear relationship is established between the natural log of vapor
pressure and the reciprocal of absolute temperature.

H vap  1 
ln (Pvap) = -   +C 2.24
R T 

where

∆Hvap = enthalpy of vaporization, R = universal gas constant, T = absolute temperature, C = a


constant characteristic of a given liquid.

C can be interpreted in terms of ∆S, the entropy of vaporization for a given liquid.

18 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Equation 3.24 is important for several reasons. For example, we can determine the heat of
vaporization for a liquid by measuring Pvap at several temperatures and then evaluating the
slope of the graph. On the other hand, if we know the values of ∆Hvap and Pvap at one
temperature, we can use equation 3.24 to calculate Pvap at another temperature, since C does
not depend on temperature.

At two temperatures T1 and T2:

H vap H vap
ln (Pvap) T1 + = C = ln (Pvap) T2 + 2.25
RT1 RT2

From which

 Pvap
T1
 H vap  1 1 
ln T2  =
  −  2.26
P  R  T2 T1 
 vap 

Like liquids, solids have vapour pressures. Iodine, for instance, sublimes at 25 ºC and 1 atm
(i.e. it goes directly from the solid state to the gaseous state without passing through the liquid
state.)

Sublimation also occurs with dry ice (solid carbon dioxide) under these conditions.

2.3 Solids

2.3.1 Introduction
Solids have much greater densities, are compressible only to a very slight extent, and are rigid;
a solid maintains its shape irrespective of its container. These properties indicate that the
components of a solid are close together and exert large attractive forces on each other.

2.3.2 Types of Solids


Solids can be classified in several ways, but the broadest categories are:
• Crystalline solids and
• Amorphous solids.
Crystalline solids are those with a highly regular arrangement of their components. At the
microscopic level this regular arrangement produces beautiful, characteristic shapes of
crystals.

The positions of the components in a crystalline solid are usually represented by a lattice, a
three-dimensional array of points designating the centers of the components (atoms, ions, or
molecules) that shows the repetitious pattern of the components.

19 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

The smallest repeating unit of the lattice is called the unit cell.

Amorphous solids are those with considerable disorder in their structures. An example is
common glass, which is best pictured as a solution whose components are “frozen in place”
before they can achieve an ordered arrangement.

2.3.3 Types of Crystalline Solids


We can have three (3) types of crystalline solids, viz.,

i. Atomic solids, illustrated by elements such as graphite, diamond, and


buckminsterfullerene (all pure carbon), boron, silicon, and all metals.
ii. Molecular solids, represented by sucrose
iii. Ionic solids, represented by NaCl

2.3.4 Metal Alloys


An alloy is defined as a substance that contains a mixture of elements and has metallic properties.

Alloys can be classified into two types, viz., substitutional alloys and interstitial alloys.

2.3.4.1 Substitutional Alloys


In substitutional alloys, some of the host metal atoms are replaced by other metal atoms of similar
size. For example, in brass zinc atoms replace approximately one-third of the atoms in the host
copper metal, as shown in fig. 3.1.

Copper

Zinc

Figure 2.1 Brass, an example of a substitutional alloy.

Sterling silver (93% silver and 7% copper), pewter (85% tin, 7% copper, 6% bismuth, and 2%
antimony) and plumber’s solder (67% lead and 33% tin) are other examples of substitutional
alloys.

2.3.4.2 Interstitial Alloys


An interstitial alloy is formed when some of the interstices (holes) in the closest packed metal
structure are occupied by small atoms as shown in fig 3.2.
20 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Steel is the best-known interstitial alloy and it contains carbon atoms in the holes of an iron crystal.
The presence of the interstitial atoms changes the properties of the host metal.

Carbon

Iron

Figure 2.2 Steel, an example of an interstitial alloy.

21 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

CHAPTER THREE

3 THERMOCHEMISTRY

3.1 Introduction

The specific application of the first law of thermodynamics to study chemical reactions is referred to
as thermochemistry. Thermochemistry is concerned with the measurement or calculation of the heat
absorbed or given out in chemical reactions. Precisely therefore, thermochemistry is the part of
thermodynamics dealing with “enthalpy (i.e., heat content) changes” accompanying chemical
reactions. In order words, thermochemistry is the study of heat energies, H in a reaction and
as such provides important information regarding heat balances and fuel requirements in a
metallurgical process. The unit of heat energy is the joule or kilojoule (J or kJ).
In this context, it will be useful to refer to some of the important terms associated with thermal
effects.

3.2 Exothermic and Endothermic Reactions

Chemical reactions are carried out commercially either to obtain a useful product or as
a way of obtaining energy, e.g. combustion of coal:
𝐶 + 𝑂2 → 𝐶𝑂2 + ℎ𝑒𝑎𝑡
NB: This process involves conversion of chemical energy into heat energy.

Every known chemical reaction involves an energy change; most reactions occur with
an evolution of heat and are called exothermic reactions, a few reactions (which will
normally only take place at high temperatures) absorb heat and are called endothermic
reactions. If heat is lost by the system, i.e. given out to the surroundings (exothermic
reaction) a negative sign (–) is placed in front of the stated quantity of heat. Conversely
if the system gains heat from the surroundings (endothermic reaction) a positive sign is
used. Two examples are given below:
298𝐾
2𝐴𝐿 + 32𝑂2 → 𝐴𝑙2 𝑂3 ; −1700 𝑘𝐽 𝑚𝑜𝑙 −1

Where the heat gained or lost is per mole of Al 203 formed. For any given reaction the

22 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

heat liberated or absorbed is constant under specified conditions.


The top of a mould into which metal is cast may be made of a material (e.g. powdered
aluminium and an oxidising agent) which evolves heat when in contact with a source
of heat such as molten metal. This allows the metal at the top of the mould to remain
molten thus feeding the contraction cavity.
1373 𝐾
𝑍𝑛𝑂 + 𝐶 → 𝑍𝑛 + 𝐶𝑂; +350 𝑘𝐽
This extraction reaction for zinc needs to overcome the difficulties of heat supply and
conservation in order to be commercially operable at 1373 K, i.e. 350 kJ of heat energy
must be added to the system in order that one mole of carbon will reduce one mole of
zinc oxide at 1373 K. (Thermodynamic relationships refer to the absolute temperature
scale, therefore K rather than °C will be used extensively in this chapter).

3.3 Standard Enthalpy Change for a Reaction

The enthalpy, H, of a substance can be defined as its heat (energy) content. Consider
the enthalpy changes for exothermic and endothermic reactions as follows:

Table 3.1 Enthalpy changes for exothermic and endothermic reactions

Exothermic reaction Endothermic reaction


Heat is evolved Heat is absorbed

Energy content (H) is lower at the end of the Energy content (H) is higher at the end of
reaction, i.e. reactants products the reaction, i.e. reactants products

H1 H2 H1 H2

(initial) (final) (initial) (final)

H2 < H1 H2 > H1

Now the change in enthalpy, ∆H, is given by*


∆𝐻 = 𝐻2 − 𝐻1
Eqn 3.1
= 𝐻𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 − 𝐻𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠

23 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Equation above is written this way conventionally (instead of ∆H= H reacts – Hprods).
Hence, for an exothermic reaction, as H2 < H1, then ∆H is negative, i.e. heat is given out
during the reaction (at constant pressure). While for an endothermic reaction, as H 2 > H 1
then ∆H is positive and heat is absorbed during the reaction (at constant pressure). Fig. 5.1
gives typical heat energy profiles for exothermic and endothermic reactions. Before a reaction
can occur energy is required, called activation energy (K), to surmount an initial energy
barrier; without this energy the reaction cannot proceed.
As stated earlier the enthalpy change, ∆H, for a reaction is constant for a given reaction under
specified conditions. ∆H is dependent upon:
a) Temperature
b) Pressure
c) Physical states of reactants and products
d) Amounts of substances reacting
The standard enthalpy change for a reaction, ∆Hᵒ, is defined as the change in enthalpy
referring to the masses (in moles) of the reactants and products shown in the equation for the
reaction at 101 325 Pa * (1 atm) and at a stated temperature, T (given as a subscript) with
the substances in the physical states normal under these conditions.

Figure 3.1: Heat energy profile for (a) an exothermic reaction (b) an endothermic reaction

For example, ∆H 298 refers to the enthalpy change at standard pressure (1 atm) and a
temperature of 298 K. The data usually is in the standard state for both reactants and
products, i.e. the forms in which they are stable under a pressure of 1 atm at a stated
temperature (usually 298 K or 25 °C). The thermodynamic standard state for gaseous

24 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

reactants and products is normally taken as 1 atm partial pressure (i.e. ∆H °) while that for
solids or liquids is normally taken as the pure state (i.e. ∆Hᵒ). For this reason all standard
states in this text will be denoted by the symbol “ᵒ” from now on.
The previous example equations can now be rewritten as full thermochemical equations:
a) C(graphite) + O 2 (g) CO 2 (g) ∆Hᵒ 298 = -393 kJ mo1 -1
b) 2 A 1 ( s ) + 32O 2 ( g ) Al2O3(s) ∆Hᵒ 298 = -1700 kJ (mol Al 2 0 3 ) -1

c) ZnO(s) + C(graphite) Zn(g) + CO(g) ∆Hᵒ 1373 = +350 kJ mol -1


Note that reaction (b) could be written as:
4A1(s) + 302(g) 2Al2O3(s) ∆Hᵒ298 = -3400 kJ
i.e. 3400 kJ of heat are released when 4 moles of solid Al react with 3 moles of gaseous O2
to produce 2 moles of solid Al 2 O 3 . In this case the units are kJ not kJ mol -1 since all
amounts differ from 1 mol.

There are various ways in which enthalpy changes for reactions may be stated more specifically.
Read on the following topics:
(i) Standard enthalpy change of combustion, ∆Hᵒc, of a substance
(ii) Enthalpy change of atomisation, ∆HᵒA, for an element
(iii)Standard enthalpy change of transformation, ∆Hᵒtrans
(iv) Standard enthalpy change of solution, ∆Hᵒsol
(v) Standard enthalpy change of dilution, ∆Hᵒdil
(vi) Standard enthalpy change of neutralisation, ∆Hᵒneut

3.1.1 Standard enthalpy change of formation for a compound, ∆H°f


This is the change in enthalpy when one mole of a compound is formed from its elements
under standard conditions. An example is given by reaction (a) above. The standard enthalpy
change of formation for carbon dioxide may also be written:
∆Hᵒ f [CO 2 (g)] = -393 kJ mol -1
Standard (atmospheric) pressure = 1 atm
= 101 325 N m-2
= 101 325 Pa (i.e. 1 N m-2 = 1Pa = 760 mmHg)

25 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

All of the oxides formed are termed exothermic compounds except those with positive enthalpy
change which is an endothermic compound. The enthalpy change of formation for a compound is
one of the factors which determines its stability to heat and chemical attack; in fact, the more
negative the value of ∆H°f, the greater the stability of that compound since this amount of heat
must be added to decompose it.
It will be explained later that other factors (i.e. free energy and entropy) must influence the
feasibility of a reaction otherwise: (a) endothermic reactions would never occur spontaneously:
and (b) reversible chemical reactions would not be possible since
 
A ⎯−⎯
H
⎯→ B + H
A ⎯ ⎯⎯ B Laplace rule

3.4 Heat Capacity

When a substance is heated, its temperature typically rises. For a specified energy, q, transferred by
heating, the size of the resulting temperature change, ΔT, depends on the ‘heat capacity’ of the
substance. The heat capacity, C, of a system is defined as the heat introduced or withdrawn to raise
or lower the temperature of the system by one degree celsius. Thus,
𝑞
𝐶=
𝑑𝑇
It follows that the heat absorbed or released by a system can simply be measured if we measure
a temperature change and then use the appropriate value of the heat capacity of the system:
𝑞 = 𝐶𝑑𝑇
The concept of heat capacity is used only when the introduction of heat to or the removal of heat
from, the system brings about a corresponding change in temperature. The concept is not used in
certain situations such as those involving phase changes. For example, if the system is a mixture of
solid iron and liquid iron under 1 atm pressure and at 1535°C (melting point of iron), then the
addition of heat simply melts some of the iron and no temperature change results. In such a situation
the heat capacity, as the definition goes, would be infinite.

The expression for heat capacity brings out the fact that it is an indefinite quantity even when mass
is specified. This is no longer the case when certain conditions, particularly constant volume or
constant pressure conditions, are specified. The heat capacity then becomes a definite quantity as a
consequence of becoming a definite quantity.

26 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

It is usual to encounter in studies of energy changes associated with reactions of chemical interest,
a great variety of chemical materials and transformations. There are many ways by which
transformations are implemented but it is convenient to consider two (2) conditions that are special
and occur frequently:
i. The volume of the system is kept constant.
ii. The pressure on the system is held constant. The second scenario, for example, is that
occurring for reactions or other processes carried out in containers that are open to
atmosphere.
The heat capacity at constant volume and at constant pressure is given as (Cv) and (Cp) respectively.
For a finite process in which the system changes from state 1 to state 2, the work done at constant
pressure is
ʃdw = Pʃdv
The first law expression may now be written as
dU = q-pʃdv
so that
U2 – U1 = (q2 – q1) – P (V2 – V1)
On rearrangement,
qp = (U2 + P V2) – (U1 + P V1)
Where qp = q2 — q1 is the heat absorbed by the system at constant pressure in going from state 1 to
state 2.
A thermodynamic function called enthalpy or heat content, represented by H, is now defined as
follows;
H=U+PV
Using this function in the earlier expression for heat absorbed at constant pressure,
qp = H 1 – H 2 = ΔH

For an infinitesimal change


=
Cp (dT)p (dT)p
And
dH = Cp dT
It is thus seen that heat capacity at constant volume is the rate of change of internal energy with
temperature, while heat capacity at constant pressure is the rate of change of enthalpy with

27 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

temperature.
Like internal energy, enthalpy and heat capacity are also extensive properties. The heat capacity
values of substances are usually expressed per unit mass or mole. For instance, the specific heat
which is the heat capacity per gram of the substance or the molar heat, which is the heat capacity per
mole of the substance, is generally considered. The heat capacity of a substance increases with
increase in temperature. This variation is usually represented by an empirical relationship such as
Cp =a+bT+cT2
Where a, b, and c are constants (for a given substance) and T is the temperature in degrees Kelvin.
The unit for Cp is calories per degree per mole (Cal/ᵒC/mol) or joules per degree per mole
(J/ᵒC/mol).
Metallurgical processes are mostly carried out at constant pressure and, therefore, enthalpy assumes
the role of a very important thermodynamic parameter in metallurgical thermochemistry. The heat
absorbed in a process at constant pressure is equal to ΔH, the increase in the enthalpy of the system.
It can thus be said that the heat change accompanying a chemical reaction is equal to the difference
between the total heat content of the products and that of the reactants, at constant pressure and
temperature conditions. This quantity is called the heat of reaction, ΔH, and can be expressed as
follows;
ΔH = ΔH (products) - ΔH (reactants)
Where, all the enthalpy values refer to the specified temperature and a pressure of one atmosphere.

Two (2) types of situation may generally arise in respect of this equation. In the first, the enthalpy
of the products exceeds that of the reactants (ΔH is positive), while in the second the converse
happens (ΔH is negative). A reaction that conforms to the former situation is called an exothermic
reaction and a reaction that corresponds to the latter situation is called an endothermic reaction.

An exothermic reaction is accompanied by evolution of heat. An endothermic reaction, in contrast,


occurs with absorption of heat. Enthalpy changes are associated with different types of reactions
in forms such as the heat of formation of a compound, the heat of combustion of a substance, the
heat of transformation, and the heat of solution. In defining these enthalpy changes, one clearly
stipulates the amounts of the reacting substances. The heat of formation of a compound is the
change in enthalpy that results when one mole of the compound is formed from its constituent
elements. Since the absolute value of the heat content of a substance is not known, the convention in
thermochemistry is to take the heat contents of elements in their standard states, i.e., in their
28 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

stable forms under 1 atm pressure-to be zero, at 25 °C (298 K). On this basis, the heat of formation
of one mole of a compound at 25 °C, as defined above, turns out to be the same as the heat content
of one mole of the compound at 25 °C.

There are two (2) important relationships in thermochemistry which are very useful in the
calculation of enthalpies of reactions. These are known as Hess’s law and Kirchoff’s equation.
Hess’s law states that the overall change in enthalpy in a reaction is the same whether the reaction
takes place in one step or through a number of intermediate steps. This law can also be regarded
as a consequence of the fact that enthalpy is a state function so that the enthalpy difference between
the final state (products) and the initial state (reactants) is the same, irrespective of the reaction path
(sequence in which the reaction takes place). As an example, let the following reaction be considered,

C + O2 = CO2
For this reaction the standard heat of reaction at 298 K, ΔHᵒ 298, equals -393510 J. Now let the
following two reactions, which also finally result in the formation of CO2, be considered
C + 0.5O 2 = CO (ΔH° 298 = –110530 J)
CO + 0.5 O2 = CO2 (ΔH°298 = –282980 J)
Adding,
C + O 2 = CO 2 (ΔH° 298 = –110530 J – 282980 J = –393510 J)
The overall enthalpy change is the same as that indicated earlier. A useful consequence of Hess’s
law is that thermochemical equations can be added and subtracted just like algebraic equations. This
facilitates the calculation of enthalpy changes for reactions which cannot be studied
experimentally.

In general, the heat capacity of any substance changes with change in temperature. Thus, in a
reaction, the heat capacities of reactants and products change with variation in temperature and this
change results in change of ΔH values for reactions as the temperature varies. For example, in the
generalized typical reaction shown below,

Reactants (R) at T1 ΔHT1 Products (P) at T1


CPR (T2 -T1) CPP (T2 - T1)

Reactants (R) at T2 ΔHT2 Products (P) at T2

29 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Starting from reactants at temperature, T1, products at temperature, T2, are obtained. This is
accomplished in two ways.
In the first way, the reaction is conducted at temperature, T1 (heat of reaction = ΔHT1) and product
is also obtained at temperature, T1. The products at T1 are then heated to temperature, T2 (heat
involved = C p (T2 - T1)) to obtain products at temperature, T2. In this way, the heat required to

obtain products at T2 starting from reactants at T1 is;


ΔHT1 + Cp (T2 - T1)
In the second way, reactants at temperature, T1 are heated to the temperature, T2 (heat involved =
C p (T2 — T1 ) to obtain reactants at temperature, T2. Reaction is then conducted to convert reactants

at T2 to products at T2 (heat of reaction = ΔHT2). In this way, the heat required to obtain products
at T2 starting from reactants at T1 is C p (T2-T1) + ΔHT2. Pursuing either of the two ways
described, reactants from a given initial state ultimately reach the same final state.
Then, according to Hess’s law, the heat required in both the ways should be the same. In other words,

ΔHT1 + CP (T2 - T1) = ΔHT2 + CP (T2 - T1)


Or
ΔH T2 - ΔH T2 = (C P - C P ) (T 2 - T 1 )
The Cp values used here, of course, refer to the mean values of Cp in the temperature interval
T1 — T2. The above expression may be rewritten as
𝑑(ΔH)
= Δ𝐶𝑝
𝑑𝑇

Where ΔCp refers to the difference in the specific heat between the products and reactants (i.e.,
ΔC p = C P - C P ). This equation is the well-known Kirchoff’s equation that is more often used in
the integrated form, which is;
T2 T2
T1
d (H ) = H T2 − H T1 =  C p dT
T1

Or
T2
H T2 = H T1 +  C p dT
T1

To calculate the heat of reaction at any temperature T2, when the heat of reaction at a given
temperature T1 is known it is necessary to have information on the dependency of heat capacity
on temperature for all the reactants and products. This information, as has been described earlier,

30 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

is available in the form of equations such as:


Cp = a + b T + cT2 +...
Where a, b, c are constants for a given substance and T is the temperature.
Usually, the temperature T1 is taken as 298 K. Including also the standard conditions, the equation
becomes;
ΔHT = ΔH298 + ʃΔCp dT ± ΔH
The last term in the above equation, AH, refers to the enthalpies of transformation that the
reactants and/or products may undergo in the temperature interval 298 to T. Enthalpies of
transformations are added (the sign is “+”) if products transform and subtracted (the sign is “-”)
if reactants transform. Molar heat capacities of reactants and products do vary with change in
temperature. If the variation is not appreciable, the Cp may be assumed to be constant, with respect
to temperature and the equation becomes
ΔH T = ΔH 2 9 8 + ΔC p (T 2 - T 1 )
And
ΔC p = ΔCP - ΔC R
For example, if the general reaction a A + b B = c C + d D is considered, then;
ΔC p = (c C p C + d C p D) - (a CpA - b CpB)
If the value of ΔHT is to be calculated very accurately, the variation of Cp with temperature must be
factored and this is done using the following expression for ΔCp:
ΔCp = a + b T + c T -2
+ ...
And if there is phase transformation in the temperature range considered, Cp values for each of
the phases should be factored and accordingly different expressions for ΔCp will have to be
calculated in different temperature ranges. Thus,
Tt T2
H T2 = H T1 +  C p dT + H t +  C p dT
T1 T1

Where, for example, the product undergoes an endothermic phase transformation at Tt and its
enthalpy of transformation is ΔHt.

3.5 Second Law

The first law of thermodynamics enunciates that heat and work are mutually convertible, but the
law does not concern itself with the extent of this convertibility. In other words, if one were to be
guided by the first law alone, then every process involving the conversion of energy should be feasible,
31 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

provided that energy is neither created nor destroyed. Thus, according to this law, in every cyclic
process work may be completely converted into heat or heat may be completely converted into
work; heat may flow from a hot to a cold or from a cold to a hot body; a gas may expand from high
pressure to low pressure or contract from low pressure to high pressure. It is known from our
experience that not all of this is true. In natural processes, work is not completely converted into
heat; heat flows from a hot to a cold body, but the reverse flow is not possible; a high-pressure gas
expands to low pressure, but the opposite process does not occur. It is true that these spontaneous
processes can be reversed, but not on their own. Some external agency is required to make the
reverse process occur. A piece of ice when kept in the atmosphere absorbs heat and melts but it will
not solidify and give back heat without any work being performed on it.

It has been seen thus far that the first law, when applied to thermodynamic processes, identifies the
existence of a property called the internal energy. It may in other words be stated that analysis of
the first law leads to the definition of a derived property known as internal energy. Similarly, the
second law, when applied to such processes, leads to the definition of a new property, known as the
entropy. Here again it may in other words be said that analysis of the second law leads to the
definition of another derived property, the entropy. If the first law is said to be the law of internal
energy, then the second law may be called the law of entropy. The three Es, namely energy,
equilibrium and entropy, are centrally important in the study of thermodynamics. It is sometimes stated
that classical thermodynamics is dominated by the second law.

3.5.1 Entropy
Entropy can be described by considering a closed system undergoing a reversible process. The
entropy change, dS, of the system is defined by the relationship:
𝑑𝑞𝑟𝑒𝑣
= 𝑑𝑆𝑟𝑒𝑣
𝑇

Where dqrev is the heat absorbed by the system isothermally at constant pressure and T is the
temperature of the system at which the heat absorption takes place. Elaborating on the above
expression, when a certain amount of heat, q, is applied to a system at temperature, T, its disorder
or entropy increase. Entropy is often visualized in terms of the disorder of the system; the greater
the disorder the larger is the entropy.

The magnitude of the entropy change is related to the heat q and the temperature T as given below:
32 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

ΔS α q
And
1
∆𝑆𝛼 𝑇

The first expression clarifies that entropy of the system increases when it takes up heat. Absorption of
heat results in rise of temperature. Increase in entropy per degree rise in temperature is not the same
at all temperatures: it is more at low temperatures and relatively less at high temperatures. This
is shown by the inverse relationship between the entropy change and temperature. The combined
expression for the variation of entropy change with quantity of heat and temperature therefore
becomes;
𝑞𝑟𝑒𝑣
∆𝑆𝛼
𝑇
The proportionality constant is 1 for reversible thermodynamic processes and
𝑞𝑟𝑒𝑣
∆𝑆 =
𝑇
In terms of the heat and the entropy changes in the system, the second law may be expressed as
follows:
dqrev
dS =
T
The equality applies when heat is transferred reversibly and the inequality refers to irreversible or
natural or spontaneous transfer of heat. Thus,
dqrev
dS =
T
dqirrev
dS 
T
With the entropy term introduced, it is possible to make a combined statement of the first and the
second laws. To do this, one may consider an infinitesimal change of state of a closed system. For this,
the first law gives
dU = dq – PdV
And if this change occurs reversibly, the second law gives
dq
dS =
T
Or

33 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

dq = TdS
Combining the above equations
dU = TdS – PdV
There are some restrictions on the applicability of this equation and these include:
i. That the system is closed.
ii. That the work associated with volume change is the only type of work performed by the
system.
At this stage, it would be worthwhile to consider some processes and the entropy changes associated
with them.

In a system undergoing a reversible adiabatic process, there is no change in its entropy. This is so
because by definition, no heat is absorbed in such a process. A reversible adiabatic process,
therefore, proceeds at constant entropy and may be described as isentropic. The entropy, however,
is not constant in an irreversible adiabatic process.

The simplest process involving a change in entropy is a reversible process occurring at a constant
temperature, T. For such a process, the change in entropy, ΔS, can be expressed as
q rev
S=
T
The change in entropy of a system in such an isothermal reversible process thus equals the heat
absorbed by the system divided by the absolute temperature of the system. A common example of a
reversible isothermal process is a phase change under constant pressure during which the
temperature also remains constant. To implement the change reversibly, the system is brought into
contact with a heat reservoir at a temperature infinitesimally higher than the equilibrium
temperature at the given pressure.

3.5.2 Thermodynamic Equations and Remarks


Table 3.2: Thermodynamic Equations and Remarks
Thermodynamic Equation Remarks

dq = d U + d W This holds good for any process undergone by a closed system

dq = dU + P dV This holds good for a closed system when only PdV work is

34 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

present.

dq = T dS This is true only for a reversible process.

T dS = d U + P dV This holds good for any process reversible or irreversible,


undergone by a closed system, since it is a relationship among
properties which are not dependent on path

T dS = dH – V dP This equation relates only the properties of closed system

There is no path function term in the equation, and therefore it holds good for any process.
The phase change then proceeds very slowly, and at all stages of the process the system is essentially
at the equilibrium temperature T. The heat (q) taken up by the system equals the heat of
transformation, ∆Htr, and the change in entropy, ΔS, can then be written as
H tr
S=
T

In most processes, a reversible absorption of heat is accompanied by a change in temperature, and a


calculation of the corresponding entropy change requires an evaluation of the integral of dq/T. The
term q is related to the heat capacity of the system which is usually expressed as a function of
temperature.

The processes that occur at a finite rate, with finite differences of temperature and pressure between
parts of a system or between a system and its surroundings, are irreversible processes. It has been
shown that the entropy of an isolated system increases in every natural (i.e., irreversible) process.
It may be noted that this statement is restricted to isolated systems and that entropy in this case
refers to the total entropy of the system. When natural processes occur in an isolated system, the
entropy of some portions of the system may decrease and that of other portions may increase. The
total increment, however, is always greater than the total decrement. The entropy of a non-isolated
system may either increase or decrease, depending on whether heat is added to it or removed from
it and whether irreversible processes occur within it. Considered all in all, it is necessary to define
clearly the system under consideration when increases and decreases in entropy are discussed.

35 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

3.6 Free Energy

In thermodynamics, entropy enjoys the status as an infallible criterion of spontaneity. The concept of
entropy could be used to determine whether or not a given process would take place spontaneously.
It has been found that in a natural or spontaneous process there would be an increase in the entropy
of the system. This is the most general criterion of spontaneity that thermodynamics offers; however,
to use this concept one must consider the entropy change in a process under the condition of constant
volume and internal energy. Though infallible, entropy is thus not a very convenient criterion. There
have, therefore, been attempts to find more suitable thermodynamic functions that would be of
greater practical convenience and utility. This has led to the introduction of two other functions,
which incorporate the entropy but are more convenient to use in many cases. These two functions
are known as Helmholtz free energy (also known as work function) and Gibbs free energy.
They lend themselves to the treatment of equilibrium state toward which process moves. In studies
of processes (say of chemical reactions) which are usually performed in the laboratory at a constant
temperature in a thermostat and in either open reactors when P is constant or in closed reactors when
V is constant, the concepts of the two terms as defined below are helpful.
Helmholtz free energy, represented by the symbol A, is defined as:
A=U–TS
In this expression U is the internal energy, T is the absolute temperature and S is the entropy.
Gibbs free energy, represented by the symbol G (the symbol F is also sometimes used in place of
symbol G), is defined as:
G=U–TS+PV
Where, P and V refer, as usual, to the pressure and the volume of the system. This definition may be
expressed in two alternative forms. First, by using the relationship;
A=U—TS
G = A + P V,
And second, by using the relationship;
H = U + PV,
G=H–TS
Before proceeding to describe the utility of Gibbs free energy in discussing equilibria, it is
worthwhile presenting expressions relating the changes in G to changes in the other thermodynamic
parameters of a system.

36 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Keeping in view the expression for G, one obtains, in the differential form,
dG = dU – T dS – S dT + V dP + P dV
It has been seen earlier that by combining the first and the second laws, the following relationship can
be arrived at:
dU = T dS – P dV
Using this relationship, one could write
dG = V dP – S dT
At constant pressure (dP = 0), this can be expressed as
dGP = – (S dT)P
Similarly, at constant temperature (dT = 0),
dGT = (V dP)T
These relationships interrelate the parameters pressure, volume and temperature with the Gibbs
free energy of a system. It may be pointed out that the results embodied in these equations are
applicable to closed systems only.
The free energy change in a chemical reaction is expressed in the same way as the enthalpy of the
reaction is expressed. For example, the free energy change in the reaction
aA+bB=cC+dD is
G r = c G C + d G D – a G A – b GB
Or
Gr = ΔG (products) - ΔG (reactants)
In the case of a system undergoing a simple process that takes it from state A to state B,
Gr = GB – GA
Note that the slope of a plot of Gr against T, i.e., of free energy change against temperature, gives the
negative of the entropy change for the process.
For a finite process occurring at constant temperature T and pressure P, the free energy change can
be expressed as
Gr = Hr – T Sr
The above equation is especially valuable in highlighting the importance of change in free energy
as an explicit criterion for the spontaneity of a chemical reaction. The occurrence of a chemical
reaction is indicated only when there is an overall decrease in the energy of the system. This was
originally identified with only the enthalpy of the system, and the exothermicity of a reaction was

37 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

taken to indicate that it would occur spontaneously. Inadequacy of enthalpy change as the sole arbiter
of the actual energy change in the system was brought by the existence of spontaneous endothermic
reactions. Similarly, reactions in which the entropy change is negative but spontaneous were also
identified. It thus became imperative that the real indicator of the overall decrease in the energy of the
system should combine both the enthalpy and entropy changes. The free energy change incorporates
the changes in enthalpy and entropy in such a way as to express the combined effect to indicate the
direction of the chemical reaction. Spontaneous reactions may thus be endothermic or may occur
with negative entropy change but cannot occur with a positive free energy change. In other words,
in a spontaneous reaction the combination of endothermicity attended with a negative entropy
change is prohibited. The above equation is known as the Gibbs–Helmholtz equation. It enables
the evaluation of the ΔH of a reaction from a knowledge of the free energy change (ΔG) and
of its temperature coefficient [(SΔG)/(ST)] P .
An alternative form of the Gibbs–Helmholtz equation can be obtained as follows. At constant
pressure,
G r dT = H r dT + T d(G r )
Dividing both sides of this equation by T2 and rearranging the terms, one obtains
Td (G ) GdT − HdT
=
T2 T2
Since the left-hand side of this equation is of the form d(x/y) = (y dx — x dy) /y2, one has
d (G / T ) H
= 2
dT T
Expressed in this form, the Gibbs–Helmholtz equation is widely used in experimental
thermodynamics to determine AH, the enthalpy of a reaction, from the experimentally determined
variation of Gr with temperature. The Gibbs–Helmholtz equation may also be written as
d (G / T )
=H
d (r )
The Gibbs–Helmholtz equation is applicable to closed systems of fixed composition undergoing
isobaric (constant pressure) processes.
An important use of the free energy function is to obtain a simple criterion for the occurrence of
spontaneous processes and for thermodynamic equilibrium. According to the second law of
thermodynamics,

38 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

𝑑𝑞
𝑑𝑆 ≥
𝑇
Where the “greater than” sign refers to an irreversible process, and the “equal to” sign to a reversible
process. Using the statement of the first law of thermodynamics, the term q may be replaced and one
may write the following:
𝑑𝑈+𝑤
𝑑𝑆 ≥ 𝑇

If in any process dU and w both are zero, then one obtains:


dS > 0
Now for w to be zero, the process has to be a constant volume process and in that case:
dS U ,V > 0
Where, the subscripts U and V denote constancy of internal energy and volume. In other words, when
the internal energy and the volume are maintained constant, the entropy of a system increases in a
spontaneous (irreversible) process, but remains unaltered when the system undergoes a reversible
process, i.e., when it is in a state of thermodynamic equilibrium. Since the entropy can either increase
or remain constant in a process, one can conclude that the entropy of a system at equilibrium is a
maximum at constant internal energy and volume.
Substituting;
dw = P dV,
𝑑𝑈+𝑃𝑑𝑉
𝑑𝑆 = 𝑇

Or
T dS = dU + P dV
Or
dU – T dS + P dV = 0
At constant temperature and pressure, dG = d U - T dS + P dV, and so for any process,
dGTP ≤ 0
Where, the “less than” sign refers to a spontaneous process and the “equal to” sign to an equilibrium
process.
Since most reactions are performed under constant pressure and temperature conditions (i.e., isobaric
and isothermal conditions), this relationship is very useful to stipulate the conditions for the
occurrence of a spontaneous process or an equilibrium (i.e., reversible) process. Since dGTP is either
less than or equal to zero, depending on whether the system undergoes a spontaneous change or is

39 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

in an equilibrium state, it follows that for a system in equilibrium, at a given temperature and
pressure, the free energy must be a minimum. All the spontaneous processes occurring at constant
temperature and pressure are accompanied by a decrease in the free energy. This result is of
fundamental importance for it gives a simple and convenient criterion for assessing whether a given
process is possible or not. For a finite process, implemented under constant pressure and temperature
conditions, this criterion may be expressed as:
ΔG T P = 0
The above form of the thermodynamic representation of the condition for the occurrence of a
spontaneous process or of equilibrium is frequently used.

3.7 Third Law

The third law of thermodynamics states that, at absolute zero, the entropy of a perfectly
crystalline solid is zero.
Mathematically,
S0k = 0
The second law of thermodynamics has introduced entropy, and this function has been shown to
be important when directions of spontaneous changes are investigated. The second law, moreover,
shows how differences in entropy of two states of a system can be determined. The third law gives
a method for assigning a value to the entropy of a system.
The third law of thermodynamics also, like the first and the second laws, is an expression of our
experience with nature. It has evolved from Nernst’s observations that at very low temperatures the
enthalpy and the free energy of a reaction approach each other asymptotically, that the asymptote is
parallel to the temperature axis and just above absolute zero they virtually cease to vary with
temperature.

40 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

CHAPTER FOUR

4 CHEMICAL EQUILIBRIUM

4.1 Introduction

Chemical equilibrium is the state where the concentrations of all reactants and products remain
constant with time. Any chemical reaction carried out in a closed vessel will reach equilibrium.
For some reactions the equilibrium position so favours the products that the reaction appears to
have gone to completion. In such situations we say the equilibrium position for the reaction lies
far to the right, in the direction of the products. For example, when gaseous hydrogen and oxygen
are mixed in stoichiometric quantities and react to form water vapour, the reaction proceeds
essentially to completion. The amounts of the reactants that remain when the system reaches
equilibrium are so tiny as to be negligible. In contrast, some reactions occur only to a slight extent.
For example, when solid CaO is placed in a closed vessel at 250C, the decomposition to solid Ca
and gaseous O2 is virtually undetectable. In cases like this, the equilibrium position is said to lie
far to the left, in the direction of the reactants.

4.2 Equilibrium Constant & the Law of Mass Action

The law of mass action states that for a general reversible reaction
pP + qQ ⇌ rR + sS 4.1
The equilibrium constant K can be expressed as

K=
Rr S s 4.2
Pp Qq
K is constant for a fixed temperature.

4.3 Characteristics of the Equilibrium Constant

1. The equilibrium expression for a reaction written in reverse is the reciprocal of that for the
original reaction.
2. When a balanced equation for a reaction is multiplied by a factor n, the equilibrium
expression for the new reaction is the original expression raised to the nth factor. Thus Knew
= (Korig.)n

41 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

3. The apparent units for K are determined by the powers of the various concentration terms.
The (apparent) units for K therefore depend on the reaction being considered

4.3.1 Equilibrium Expressions Involving Pressures


For gases,
PV = nRT, or
n
P =   RT = CRT 4.3
V 
Where
n
C =   , represents the molar concentration of the gas.
V 
Generally, for the hypothetical reaction, equation 6.1,
KP = KC. (RT)Δn 4.4
Where Δn = (r + s) – (p + q), is the change in the number of moles.

4.4 The Concept of Activity

The equilibrium constant expression does not simply involve the observed equilibrium pressure or
concentration for a substance but involves the ratio of the equilibrium pressure (or concentration)
for that substance. The ratio is defined as the activity of the substance.
In terms of pressure:
Pi
Activity (of ith component) = ai = 4.5
Pref

Where
Pi = partial pressure of the ith gaseous component
Pref = 1 atm (exactly)
and where ideal behaviour is assumed.

4.5 Heterogeneous Equilibria

Many equilibria involve more than one phase and are called heterogeneous equilibria. For
example, the thermal decomposition of calcium carbonate in the commercial preparation of lime
occurs by a reaction involving both solid and gas phases:

42 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

CaCO3(s) ⇌ CaO (s) + CO2 (g) 4.6


Application of the law of mass action leads to the equilibrium expression
[CO2 ][ CaO ]
K/ = 4.7
[CaCO3 ]

However, experimental results show that the position of a heterogeneous equilibrium does not
depend on the amounts of pure solids or liquids present. This result makes sense when the meaning
of an activity for a pure solid or liquid is understood.
Thus, for the decomposition of CaCO3 considered above, we do not insert [CaCO3] or [CaO] into
the equilibrium expression but rather into the activity of each:
Pure solid
c1 V1 = c2 V2

[CaCO3 ] [CaO]
aCaCO3 = =1 and aCaO = =1
[CaCO3 ] [CaO]

Pure solid (reference state)


Thus, the equilibrium expressions for the decomposition of solid CaCO3 are
[CO2 ](1)
K= = [CO2] and
1
PCO2 (1)
KP = = PCO2
1
In summary, we can make the following general statement: the activity of a pure solid or liquid is
always 1.

If pure solids or pure liquids are involved in a chemical reaction, their concentrations are not
included in the equilibrium expression for the reaction. This simplification occurs only with pure
solids or liquids, not with solutions or gases, because in these last two cases the activity cannot be
assumed to be 1.

For example, in the decomposition of liquid water to gaseous hydrogen and oxygen,
2 H2O(l) ⇌ 2 H2 (g) + O2 (g) 4.8
K = [H2]2 [O2] 4.9
and
43 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

KP = ( PH22 )(PO2 ) 4.10

However, if the reaction were carried out under conditions in which the water is a gas rather than
a liquid,
2 H2O(g) ⇌ 2 H2 (g) + O2 (g) 4.11
Then
[ H 2 ]2 [O2 ]
K= 4.12
[ H 2O ]2
and
( PH22 )( PO2 )
KP = 4.13
PH22 O

4.6 Applications of the Equilibrium Constant

Knowledge of the equilibrium constant for a reaction allows us to predict:


1. The tendency of the reaction to occur (but not the speed of the reaction)
2. Whether a given set of concentrations represents an equilibrium condition.
3. The equilibrium position that will be achieved from a given set of initial concentrations.

4.6.1 6.4.1 The Extent of a Reaction


The inherent tendency for a reaction to occur is indicated by the magnitude of the equilibrium
constant.
• If the equilibrium constant is far greater than 1 (K ≫ 1), then at equilibrium the reaction
system will consist of mostly products  the equilibrium lies to the right; i.e. reactions
with very large K go essentially to completion.

• If the equilibrium constant is far less than 1 (K≪ 1), then at equilibrium the system will
consist mostly of reactants,  the equilibrium position is far to the left. The given reaction
does not occur to any significant extent.
NB!!
The size of K and the time required to reach equilibrium are not directly related. The time required
to achieve equilibrium depends on the reaction rate; the size of K is determined by factors such as
the difference in energy between products and reactants.
44 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

4.7 Reaction Quotient

When reactants and products of a given chemical reaction are mixed, it is useful to know whether
the mixture is at equilibrium, and if it is not, in which direction the system will shift to reach
equilibrium
i. If the concentration of one of the reactants or products is zero, the system will shift in
the direction that produces the missing component.
ii. However, if all the initial concentrations are non-zero, it is more difficult to determine
the direction of the move toward equilibrium.

In such cases, we use the reaction quotient Q to determine the shift.


Q is obtained by applying the law of mass action, but using initial concentrations instead of
equilibrium concentrations.
For example, consider the synthesis of ammonia by the Haber process,
N2 (g) + 3H2 (g) ⇌ 2NH3 (g) 4.14
NH 3 02
Q= 4.15
N 2 0 H 2 30
Where, the subscripts of zero indicate initial concentrations. To determine in which direction a
system will shift to reach equilibrium, we compare the values of Q and K.
Three possible cases may arise from this comparison.
1. Q = K
The system is at equilibrium; no shift will occur
2. Q > K
In this case the ratio of the initial concentrations of products to initial concentrations of
reactants is too large. For the system to reach equilibrium, a net change of products to reactants
must occur. The system shifts to the left, consuming products and forming reactants, until
equilibrium is achieved.
3. Q < K
Ratio of initial concentrations of products to initial concentrations of reactants is too small.
The system must shift to the right, consuming reactants and forming products, to attain
equilibrium.

45 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

4.8 Solving Equilibrium Problems

The following steps are taken when solving equilibrium problems.


1. Write the balanced equation for the reaction.
2. Write the equilibrium expression using the law of mass action
3. List the initial concentrations
4. Calculate Q and determine the direction of the shift to equilibrium
5. Define the change needed to reach equilibrium, and define the equilibrium concentrations
by applying the change to the initial concentrations
6. Substitute the equilibrium concentrations into the equilibrium expression, and solve for the
unknown
7. Check your calculated equilibrium concentrations by making sure they give the correct
value.

Example 1
Consider the following reversible reaction.
2 HI(g) ⇌ H2(g) + I2(g)
(a) The reaction system is in equilibrium in a container of volume V = 1.0 dm3 at a given
temperature T.
Analysis gives the following equilibrium concentrations:
[HI] = 0.16 M, [H2] = 0.020 M and [I2] = 0.020 M
Determine the value of K at this temperature.
(b) At another temperature, K = 2.0*10-3 for this reaction. Into a container with volume V = 1.0
dm3, 1.0 mol HI is injected, heated to the given temperature and equilibrium allowed to establish.
Calculate the equilibrium concentrations of each of the species.
Solution
(a) The expression for the equilibrium constant K for this reaction is:
[ H 2].[ I 2 ]
K=
[ HI ]2
(0.020M )(0.020M )
K= = 0.016
(0.16M ) 2

46 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

(b) Here it is better to tabulate your results in terms of initial amount of material and the
equilibrium amount by representing the amount of H2 produced at equilibrium by x (mol).

Table 4.1 Initial and Equilibrium amounts of species.


nHI (mol) nH 2 ( mol ) nI 2 ( mol )

Initial 1.0 0 0

Equilibrium 1.0 – 2x X X

Since the volume of the container is 1.0 dm3, the equilibrium concentrations are:
[HI] = (1.0 – 2x) M, [H2] = x M and [I2] = x M
To determine x we substitute these values into the equilibrium constant expression:
x2
2.0*10-3 =
(1.0 − 2 x) 2
The R.H.S. of this equation is a perfect square, so we can find the square root of both sides
and ignore the negative values.
x
= 0.045
1 − 2x
From which,
x = 0.041
The equilibrium concentrations are thus
[H2] = [I2] = 0.041 M
[HI] = (1.0 – 2x) = 0.092 M

4.9 Le Chatelier’s Principle

If a change in conditions (a “stress”) is imposed on a system at equilibrium, the equilibrium


position will shift in a direction that tends to reduce that change in conditions.

4.9.1 Effect of Change in Concentration


Example 1: The Haber Process

47 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Consider once again the equilibrium reaction that occurs in the synthesis of ammonia in the Haber
process:
N2 (g) + 3 H2 (g) ⇌ 2 NH3 (g) 4.17
If n mol of N2 is suddenly injected into the system, then in accordance with Le Chatelier’s principle
the system will shift in a direction that consumes nitrogen. This reduces the effect of the addition.
Hence the reaction shifts to the right. If, on the other hand, ammonia had been added instead of
nitrogen, the system would have shifted to the left to consume ammonia.
Conclusion
If a gaseous reactant or product is added to a system at equilibrium, the system will shift away
from the added component. If a gaseous reactant or product is removed, the system will shift
toward the removed component.

Example 2: Extraction of Arsenic from its Ores:


Arsenic can be extracted from its ores by roasting to form solid As4O6, which is then reduced by
carbon:
As4O6 (s) + 6 C (s) ⇌ As4 (g) + 6 CO (g) 4.18
Predict the direction of the shift of the equilibrium position for this reaction in response to each of
the following changes in conditions:
(a) Addition of CO
(b) Addition of C or As4O6
(c) Removal of As4
Solution
(a) Le Chatelier’s principle predicts that the shift will be away from the substance whose
concentration is increased. The equilibrium position will shift to the left / backward / reverse
direction when CO is added.
(b) Since the amount of a pure solid has no effect on the equilibrium position, changing the amount
of C or As4O6 will have no effect.
(c) If gaseous As4 is removed, the equilibrium position will shift to the right to form more products.
In industrial processes the desired product is often continuously removed from the reaction system
to increase the yield.

48 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

4.9.2 Effect of a Change in Pressure


Basically, there are three (3) ways to change the pressure of a reaction system involving gaseous
components at a given temperature:
1. Addition or removal of a gaseous reactant or product.
2. Addition of an inert gas (one not involved in the reaction)
3. Changing the volume of the container.

NB: One above has already been considered.

2. Addition of an Inert Gas


Addition of an inert gas increases the total pressure but has no effect on the concentrations or
partial pressures of the reactants or products (assuming ideal gas behaviour). The system remains
at the original equilibrium position.

3. Changing the Volume of the Container


When the volume of the container is changed, the concentrations (and thus the partial pressures)
of both reactants and products are changed.
In general, when the volume of the container holding a gaseous system is reduced, the system
responds by reducing its own volume. This is done by decreasing the total number of gaseous
molecules in the system.
Illustration
From the ideal gas law,
PV = nRT 4.19
 RT 
⇒V=  .n 4.20
 P 
Or at constant T and P:
V∝n
That is at constant temperature and pressure; the volume of a gas is directly proportional to the
number of moles of gas present.
Example:
Consider a mixture of gaseous sulphur dioxide, oxygen and Sulphur trioxide at equilibrium.

49 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

2 SO2 (g) + O2 (g) ⇌ 2 SO3 (g) 4.21


If we suddenly reduce the volume, what will happen to the equilibrium position?
Solution
The reaction system can reduce its volume by reducing the number of molecules present.
Consequently, the reaction will shift to the right, since in this direction three molecules (two of
sulphur dioxide and one of oxygen) react to produce two molecules (of sulphur trioxide), thus
reducing the total number of gaseous molecules present. The new equilibrium position will be
further to the right than to the original one, i.e. the equilibrium position will shift toward the side
of the reaction involving the smaller number of gaseous molecules in the balanced equation.

4.9.3 The Effect of a Change in Temperature


The changes discussed so far do not alter the equilibrium constant, although they may alter the
equilibrium position (assuming ideal behaviour).
Because the value of K changes with temperature, the effect of temperature on equilibrium is
different. Le Chatelier’s principle can be used to predict the direction of change.
Let us illustrate this with the Haber process once again. The synthesis of ammonia from nitrogen
and hydrogen releases energy (i.e. it is exothermic). We can represent this situation by treating
energy as a product:
N2 (g) + 3 H2 (g) ⇌ 2 NH3 (g) + Energy 4.22
If energy in the form of heat is added to this system at equilibrium, Le Chatelier’s principle predicts
that the shift will be in the direction that consumes energy; in this case to the left. Note that this
shift decreases the concentration of ammonia and increases the concentrations of nitrogen and
hydrogen, thus decreasing the value of K.
Observed values of K for the synthesis of ammonia at various temperatures is illustrated in Table
4.2

Table 4.2: Observed value of K for the synthesis of Ammonia at various Temperatures.
Temperature (K) K (L2 / mol2)
500 90
600 3
700 0.3

50 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

800 0.04

On the other hand, for a reaction that consumes energy (an endothermic reaction), such as the
decomposition of calcium carbonate, an increase in temperature will cause the equilibrium to shift
to the right and the value of K to increase.
CaCO3 (s) + Energy ⇌ CaO(s) + CO2 (g) 4.23
Summary
To study the effect of a temperature change on a system at equilibrium:
1. Energy is treated as:
(a) A reactant in an endothermic process or as
(b) A product in an exothermic process
2. We then predict the direction of the shift as if an actual reactant or product is added or
removed

4.10 Free Energy and Equilibrium

Consider the hypothetical reaction


a A (g) ⇌ b B (g) 4.24
The Free Energy Change ΔG, for this reaction is defined by the expression
ΔG = ΔG0 + RT lnQ 4.25
Where
ΔG0 = Standard Free Energy Change
R = Gas Constant (8.3145 JK-1mol-1)
T = Absolute Temperature
[ B]b ( PB )b
Q = Reaction Quotient = =
[ A]a ( PA ) a
At equilibrium, ΔG = 0 and Q equals K, the equilibrium constant for the reaction. Thus

ΔG = 0 = ΔG0 + RT lnK
ΔG0 = -RT lnK 4.26

4.11 Temperature Dependence of K

The dependence of the equilibrium constant on temperature can be obtained from the relation:
51 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

ΔG0 = -RT lnK = ΔH0 - T ΔS0 4.27


This equation can be rearranged to give:
− H 0 S 0
lnK = + 4.28
RT R
− H 0  1  S 0
lnK =  + 4.28 b
R T  R
This result assumes that both ΔH0 and ΔS0 are independent of temperature over the temperature
range considered. This assumption is good only over a relatively small temperature range.

Example
The overall reaction for the corrosion (rusting) of iron by oxygen is
4 Fe (s) + 3 O2 (g) ⇌ 2 Fe2O3 (s)

Table 4.3 Using the following data, calculate the equilibrium constant for this reaction at 25 0C.
Substance H 0f (kJ mol-1) S0 (J K-1 mol-1)

Fe2O3 (s) -826 90


Fe (s) 0 27
O2 (g) 0 205

Solution
To calculate K for this reaction, we will use the equation
ΔG0 = -RT ln K.
ΔG0 can be calculated from

ΔG0 = ΔH0 - T ΔS0,


Where
ΔH0 = 2 H 0f ( Fe 2 O3 ( s )) - 3 H 0f ( O 2 ( g )) - 4 H 0f ( Fe ( s ))

= 2 mol (-826 kJ /mol) - 0 - 0


= -1652 kJ = -1.652 × 106 J
ΔS0 0
= 2 S Fe 2 O3
- 3 S O0 3 - 4 SFe
0

52 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

= 2 mol (90 J K-1 mol-1) – 3 mol (205 J K-1 mol-1) – 4 mol (27 J K-1 mol-1)
= -543 J K-1
and
T = 273 + 25 = 298 K
Then
ΔG0 = ΔH0 - T ΔS0 = (-1.652 × 106 J) – (298 K) (-543 J K-1)
= -1.490 × 106 J
and
ΔG0 = -RT ln K = - (8.3145 J K-1 mol-1) (298 K) ln K
Thus
1.490  10−6
ln K = = 601 and K = e601
2.48  103

In terms of base 10,


K = 10261
This is a very large equilibrium constant. The rusting of iron is clearly very favourable from a
thermodynamic point of view.

4.12 Ionic Equilibrium

4.12.1 Ionic Product, Q


For a sparingly soluble salt AxBy in solution, the following equilibrium may be established.
AxBy (s) ⇌ x Ay+ (aq) + y Bx- (aq) 6.29
The ionic product Q is defined as:
Q = [Ay+]x [Bx-]y+ 6.30

4.12.2 Solubility Product (Ksp)


When a sparingly soluble salt AxBy is stirred in water, equilibrium is obtained between solid
undissolved salt and ions in the aqueous solution:
AxBy (s) ⇌ x Ay+ (aq) + y Bx- (aq) 6.31
From this reaction, we define the solubility product, KSP, of the salt AxBy as:
KSP (AxBy) = [Ay+]x [Bx-]y 4.32

53 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

Example 1
The solubility product of PbSO4 is 1.6 × 10-8 M2. Calculate the solubility of PbSO4 in g/dm3.
Solution
In a saturated solution of PbSO4, the following equilibrium is established:
PbSO4 (s) ⇌ Pb2+ (aq) + SO42- (aq)
KSP = [Pb2+].[SO42-]
This shows that one mol PbSO4 dissolves to produce one mol Pb2+ one mol SO42- ions. Let the
solubility of PbSO4 = x mol /dm3. Then in a saturated solution of PbSO4,
[Pb2+] = [SO42-] = x M
Substituting in the expression for the solubility product:
X2 = 1.6 × 10-8
⇒ x = 1.3 × 10-4 M
The solubility of PbSO4 is therefore 1.3 × 10-4 mol /dm3.
Since the molar mass of PbSO4 is 303.3 g/mol, the solubility becomes
1.3 × 10-4 mol /dm3 × 303.3 g/mol
= 3.9 × 10-2 g/ dm3.

Example 2
The solubility of CaF2 is 1.5 × 10-2 g/ dm3. Calculate the solubility product of CaF2. Molar mass
of CaF2 is 78.1 g/mol.
Solution
In a saturated solution of CaF2 the following equilibrium exists:
CaF2 ⇌ Ca2+(aq) + 2 F-(aq)
This shows that 1 mol CaF2 dissolves to produce 1 mol Ca2+ ions and 2 mol F- ions.
1.5  10−2
1.5 × 10-2 g/ dm3 CaF2  mol/ dm3 = 1.9 × 10-4 mol/ dm3.
78.1
In a saturated solution:
[Ca2+] = 1.9 × 10-4 M and [F-] = 2 ×1.9 × 10-4 M.
Thus
KSP(CaF2) = [Ca2+].[F-] = (1.9 × 10-4 M).(3.8 × 10-4 M)2.
= 2.7 × 10-11 M3.

54 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

4.12.3 Ionic Product and Precipitation


We use the concept of ionic product to decide whether we will obtain a precipitate (of a sparingly
soluble salt) when two solutions of quite soluble salts are mixed.
Three possibilities may arise under such conditions, viz.,
Q = KSP: ⇒ Saturated solution, we have an equilibrium situation.
Q > KSP: ⇒ Supersaturated solution, we obtain a precipitation
Q < KSP: ⇒ Unsaturated solution, we obtain no precipitation.

Example
Decide whether a precipitate of BaSO4 will be formed when 50 cm3 of 0.0010 M BaCl2 solution is
mixed with 450 cm3 of 1.0 × 10-5 M Na2SO4 solution.
KSP(BaSO4) = 1.0 × 10-10 M2
Solution
The ionic concentrations before mixing the solutions are

[Ba2+]0 = 1.0 × 10-3 M [SO42-]0 = 1.0 × 10-5 M


We can calculate the ionic concentrations that would be obtained if no precipitate were formed,
by using the dilution formula
For Ba2+:
c1 = 1.0 × 10-3 M
V1 = 50 cm3 = 0.050 dm3
V2 = 50 cm3 + 450 cm3 = 500 cm3 = 0.50 dm3
1.0  10−3  0.050
c2 = [Ba2+] = M = 1.0 × 10-4 M
0.50
For SO42-:
c1 = 1.0 × 10-5 M V1 = 450 cm3 = 0.450 dm3 V2 = 0.5 dm3

2- 1.0  10−5  0.450


c2 = [SO4 ] = M = 9.0 × 10-6 M
0.50
The ionic product can then be calculated as:
Q = [Ba2+].[SO42-] = (1.0 × 10-4 M).( 9.0 × 10-6 M)
= 9.0 × 10-10 M2

55 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

We realize that Q > KSP and therefore a precipitate of BaSO4 is obtained.

4.12.4 Solubility (Product) and pH of Solution


The pH of a solution can affect a salt’s solubility quite significantly. For example, magnesium
hydroxide dissolves according to the equilibrium
Mg (OH)2 (s) ⇌ Mg2+(aq) + 2OH-(aq) 4.33
Addition of OH- ions (an increase in pH) will, by the common ion effect, force the equilibrium to
the left, decreasing the solubility of Mg (OH) 2. On the other hand, an addition of H+ ions (a
decrease in pH) increases the solubility, because OH- ions are removed from solution by reacting
with the added H+ ions. In response to the lower concentration of OH-, the equilibrium position
shifts to the right. This explains how a suspension of solid Mg (OH) 2, known as milk of magnesia,
dissolves in the stomach to combat excess acidity.
This idea also applies to salts with other types of anions. For example, the solubility of Ag3PO4 is
greater in acid than in pure water, because the PO43- ion is a strong base that reacts with H+ to form
the HPO42- ion. The reaction:
H+ + PO43- → HPO42- 4.34
occurs in acidic solution, thus lowering the concentration of PO43- and shifting the equilibrium
Ag3PO4 (s) ⇌ 3 Ag+ (aq) + PO43-(aq) 4.35
to the right. This in turn increases the solubility of Ag3 PO43.
Silver chloride (AgCl) however, has the same solubility in acid as in pure water. This is because
Cl- ion is a very weak base, and the addition of H+ to a solution containing Cl- does not affect the
concentration of Cl- and thus has no effect on the solubility of a chloride salt.
The general rule is that if the anion X- is an effective base, i.e. if HX- is a weak acid, the salt MX
will show increased solubility in acidic solution. Examples of common anions that are effective
bases are OH-, S2-, CO32-, C2O42- and CrO42-. Salts containing these anions are much more soluble
in an acidic solution than in pure water.

4.12.5 Percent Dissociation


It is often useful to specify the amount of a substance that has dissociated in achieving equilibrium.
The percent dissociation is defined as follows:

56 | P a g e
GL_MN 154 Physical & Analytical Chemistry Lect. Notes 1

𝑚𝑜𝑙
𝑎𝑚𝑜𝑢𝑛𝑡 𝑑𝑖𝑠𝑠𝑜𝑐𝑎𝑖𝑡𝑖𝑜𝑛 ( 𝐿 )
𝐷𝑖𝑠𝑠𝑜𝑐𝑖𝑎𝑡𝑖𝑜𝑛 % = × 100
𝑚𝑜𝑙
𝑖𝑛𝑖𝑡𝑖𝑎𝑙 𝑐𝑜𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛 ( 𝐿 )

Example
At 46 oC, Kp for the reaction
N2O4 (g) ⇌ 2 NO2 (g)
is 0.66 atm. Compute the percent dissociation of N2O4 at 46 oC and a total pressure of 380 torr.
What are the partial pressures of N2O4 and NO2 at equilibrium?
Solution
(P ) = 0.66
NO2
2

(P )
Kp = (1)
N 2 O4

PN 2 O4 + PNO2 = 380 torr = 0.500 atm (2)

Let PNO2 = p; then PN 2 O 4 = 0.500 – p

Putting these values into equation (1) we obtain


p2
= 0.66
0.500 − p

⇒ p2 + 0.66 p – 0.33 = 0
PNO2 = 0.332 atm; PN 2 O 4 = 0.168 atm

Since each mol of N2O4, which dissociates produces 2 mol of NO2, the percent dissociation is given
by
0.5PNO2 0.5(0.332)atm
= × 100% = 50%
Po
N 2 O4 [0.5(0.332) + 0.168]atm

REFERENCES (Further Reading)

Burdge, J. and Overby, J., (2017), Atoms First, McGraw-Hill, pp. 1200

57 | P a g e

You might also like