Ge CH 11
Ge CH 11
Ge CH 11
All rights reserved. No part of this Study material be reproduced in any form without
permission in writing from Netaji Subhas Open University.
Kishore Sengupta
3 Registrar
Netaji Subhas UG : CHEMISTRY
(HCH)
Open University
5
Unit 1 Chemical Energetic
Contents
1.0. Objectives
1.1. Introduction
1.2. Brief review of thermodynamics and the Laws of Thermodynamics
1.2.1. System and surrounding
1.2.2. State of a system
1.2.3. Properties of a system:
1.2.4. Types of processes
1.2.5. Work, heat and heat capacity
1.2.6. The Zeroth law of thermodynamics
1.2.7. The First law of thermodynamics
1.2.8. Internal energy
1.2.9. Enthalpy
1.3. Important principles and definitions of thermochemistry
1.3.1. Energy changes accompanying chemical reactions
1.3.2. Thermochemical equations
1.3.3. Standard enthalpy of reactions
1.3.4. Hess's law of constant heat summation
1.4. Standard enthalpies of formations
1.5. Enthalpies of solution and dilution
1.6. Bond energy, bond dissociation energy and resonance energy
1.7. Variation of enthalpy of a reaction with temperature - Kirchhoff's equation
1.8. Third Law of thermodynamics and calculation absolute entropies of substances
1.9. Summary
1.10. Questions
8 NSOU GE-CH-11
1.0 Objectives
After reading this unit, we will be able to:
define the commonly used terms in thermodynamics and laws of
thermodynamics;
state important principles of thermochemistry;
explain standard state and standard enthalpies of formations;
explain enthalpies of solution and dilution;
solve numerical problems based on the enthalpy changes;
define bond enthalpy and bond dissociation enthalpy and resonance energy;
variation of enthalpy of a reaction with temperature – Kirchhoff’s equation;
state the third law of thermodynamics and calculate the absolute entropies of
substances.
1.1 Introduction
When a chemical reaction occurs, it is accompanied by an energy change which may
take any of several different forms. For example, the energy change involved in the combustion
of fuels like kerosene, coal, wood, natural gas, etc., takes the form of heat and light.
Electrical energy is obtained from chemical reactions in batteries. The formation of glucose,
C6H12O6 by the process of photosynthesis requires the absorption of light energy from the
sun. Thus, various forms of energy are interrelated and under certain conditions, these may
be transformed from one form into another. The study of these energy transformations
forms the subject matter of thermodynamics. In this unit we will review some of the
fundamental concepts of energy and heat and the relation between them. We will begin the
study of thermodynamics, which treats the energetic aspects of change in general, and will
finally apply this specifically to chemical change.
different condition. If the transfer of heat is done under the conditions of constant volume,
the associated heat capacity is called as the heat capacity at constant volume (Cv) and if
the transfer of heat is done under the conditions of constant pressure, the associated heat
capacity is called as the heat capacity at constant pressure (Cp). The values of Cp and Cv
for a given solid or liquid systems are not much different. However for gaseous systems
the difference in their values is significant and for an ideal gas it is given by the following
expression
C p – Cv = R
where R is the gas constant.
all the atoms, molecules or ions contained in the system. It is a state variable. It is not
possible to measure the absolute values of internal energy. However, we can calculate the
change in internal energy. If the internal energy of the system in the initial state is U1 and
that in the final state is U2, then change in internal energy ΔU is independent of the path
taken from the initial to the final state. We can write this change as:
Δ U = U 2 – U1
The internal energy of the system can be changed in two ways:
(i) either by allowing heat to flow into the system or out of the system; and
(ii) by work done on the system or by the system.
products as Hproducts. The difference between these enthalpies, ΔrH, is the enthalpy of the
reaction
ΔrH = Hproducts – Hreactants
When Hproducts is greater than Hreactants then ΔrH is positive and heat is absorbed in the
reaction, and the reaction will be endothermic. For example,
H2(g) + I2(g) → 2HI (g); ΔrH = 52.5 kJ
When Hproducts is less than Hreactants then ΔrH is negative and heat is evolved in the
reaction, and the reaction will be exothermic. For example,
CH4(g) + 2O2(g) → CO2(g) + 2H2O (l); ΔrH = – 890.4 kJ
Enthalpy of a reaction changes with pressure and temperature.
conducted in a laboratory. A particular reaction can also be imaginary. The only requirement
is that the individual chemical reactions in the sequence must balance and add up to the
equation for a particular reaction. Also Hess’s law enables arithmetic operations of chemical
equations. This law is helpful in calculating enthalpies of reactions which cannot be
experimentally determined; what is required is to select a correct sequence of reactions.
We give below the use of Hess’s law in calculating the enthalpy of conversion of graphite
to diamond which is very difficult to determine.
C (graphite) + O2 (g) → CO2 (g); ΔrH1 = – 393.5 kJ
C (diamond) + O2 (g) → CO2 (g); ΔrH2 = –395.4 kJ
Then subtracting, we get
C (graphite) – C (diamond) = ΔrH3 = ΔrH1 – ΔrH2 = –393.5 kJ – (–395.4 kJ) =
1.9 kJ
Hence, we can write
C (graphite) → C(Diamond) ΔrH3 = 1.9 kJ
change observed on dissolving the substance in an infinite amount of solvent when the
interactions between the ions (or solute molecules) are negligible.
When an ionic compound dissolves in a solvent, the ions leave their ordered positions
on the crystal lattice. These ions are now more free in solution. But solvation of these ions
(hydration in case solvent is water) also occurs at the same time. This is shown
diagrammatically, for an ionic compound, AB (s)
The enthalpy of solution of AB(s), ΔsolH0, in water is, therefore, determined by the
selective values of the lattice enthalpy, ΔlatticeH0 and enthalpy of hydration of ions, ΔhydH0
as
Δsol H 0
= ΔlatticeH 0 + ΔhydH 0
For most of the ionic compounds, ΔsolH0 is positive and the dissociation process is
endothermic. Therefore the solubility of most salts in water increases with rise of temperature.
If the lattice enthalpy is very high, the dissolution of the compound may not take place at
all.
It is known that enthalpy of solution is the enthalpy change associated with the addition
of a specified amount of solute to the specified amount of solvent at a constant temperature
and pressure. This argument can be applied to any solvent with slight modification. Enthalpy
change for dissolving one mole of gaseous hydrogen chloride in 10 mol of water can be
represented by the following equation. For convenience we will use the symbol aq. For
water
HCl (g) + 10 aq. → HCl.10 aq. ΔH = –69.01 kJ / mol
Let us consider the following set of enthalpy changes:
(S-1) HCl (g) + 25 aq. → HCl.25 aq. ΔH = –72.03 kJ / mol
(S-2) HCl (g) + 40 aq. → HCl.40 aq. ΔH = –72.79 kJ / mol
(S-3) HCl (g) + aq. → HCl. aq. ΔH = –74.85 kJ / mol
NSOU GE-CH-11 17
We can observe here that each of these equation represent the dissociation of C-H
bond but the energies required for them are different. Though all the C-H bond in CH4 are
equivalent, the dissociation energies are not same. This is due to the fact that once a C-
H bond dissociated, the remaining species has different electronic distributions.
We can add these four equations to get
CH4 (g) → C (g) + 4H (g); ΔrH0 = 1663.4 kJ/mol
In this reaction the methane molecule has got converted into its constituent atoms
therefore the enthalpy change is called enthalpy of atomization of methane. If we divide the
enthalpy of atomization of methane by 4, we get an average value (1663.4/4 = 415.9 kJ/
mol) for the bond enthalpy for C-H bond.
Now if we want to determine the bond enthalpy in ethane (C2H6), here two types of
bond exist, C-H and C-C. We assume that the bond enthalpy for C-H bond in ethane is
same as that of methane. The atomization of ethane is given by the expression
C2H6 (g) → 2C (g) + 6H (g); ΔrH0 = 2842.4 kJ/mol
The enthalpy of atomization of ethane would be equal to the bond enthalpies of six C-
H bond and one C-C bond. If we subtract the bond enthalpies of six C-H bonds from
the enthalpy of atomization we can get the bond enthalpy for C-C bond. Thus, bond
enthalpy for a given bond refers to the average bond dissociation energy for the same bond
in a number of molecules having the said bond. Using Hess’s law, bond enthalpies can be
calculated. Again bond enthalpy data can be used to calculate ΔrH for a chemical reaction
occurring in gaseous state by making use of difference in energy absorbed in breaking the
bonds in reactants and energy released in formation of bonds in products.
ΔrH = Σ B.E. (reactants) – Σ B.E. (products).
Using bond energy and enthalpy of formation data resonance energy in a molecule can
be calculated. When a compound shows resonance, there is considerable difference between
the heat of formation as calculated from bond energies and that determined experimentally.
Resonance energy = Experimental or actual heat of formation - Calculated heat of
formation.
For example the resonance energy of N2O can be calculated if observed ΔfHo (N2O)
is 82 kJ mol–1, B.E of N ad N is 946 kJ mol–1, B.E of N N is 418 kJ mol–1, B.E of
O = O is 498 kJ mol–1; B.E of N = 0 is 607 kJ mol–1.
NSOU GE-CH-11 19
1
So we can write, N 2 (g) O 2 (g) N 2O
2
1
In another way we can write, N N O O N N O
2
So, calculated ΔfHo (N2O) = [B.E (Na=N)
+ B.E (O = O)
] – [B.E (N=N)
+ B.E (N=O)
]
498
= 946 – [418 + 607] = + 170 kJ/mole
2
d
dT
r H 0
d 0
dT
H (products)
d 0
dT
H (reactants)
d 0
As H C0p we can write
dT
d
dT
r H 0 C0p (products) C0p (reactants)
20 NSOU GE-CH-11
d
dT
r H 0 C0p
Writing the equation in differential form
d r H 0 C0p dT
T2
0 0
T1 d r H T1 CpdT
2 T
Rearranging,
This is the Kirchhoff’s equation that gives the temperature dependence of enthalpy
change for a reaction.
1.9. Summary
In this unit we have started our discussion by revisiting the important terms and laws
NSOU GE-CH-11 21
1.10. Questions
1. What is standard enthalpy of a reaction?
2. Why the formation enthalpy of Graphite is taken as zero?
3. What is an exothermic and endothermic reaction?
4. Describe Hess’s law and explain.
5. Derive the Kirchhoff’s equation relating to the effect of temperature on the
reaction enthalpy of a reaction.
6. Calculate ΔrHo for the reaction C (graphite) + O2 (g) → CO (g) at 298 K.,
using the following thermochemical equations:
C (graphite) + O2 (g) → CO2 (g); ΔrHo = – 393.5 kJ
CO (g) + 1/2O2 (g) → CO2 (g); ΔrHo = – 283.5 kJ
(Ans. -110 kJ)
Unit 2 Chemical Equilibrium
Contents
2.0. Objectives
2.1. Introduction
2.2. Free energy change in a chemical reaction
2.3. Law of mass action
2.4. Thermodynamic derivation of the law of chemical equilibrium
2.5. Distinction between ΔG and ΔG0
2.6. Relationships between Kp, Kc and Kx for reactions involving ideal gases
2.7. Le Chatelier's principal
2.7.1. Effect of change of concentration
2.7.2. Effect of change of pressure
2.7.3. Effect of change of temperature
2.7.4. Effect of catalyst
2.8. Summary
2.9. Questions
2.0. Objectives
After going through this unit, we will be able to know about
reversible and irreversible reactions.
free energy change in a chemical reaction.
law of mass action.
relation between free energy and equilibrium constant
express the equilibrium constant in different ways
Le Chatelier’s principle and its applications.
2.1. Introduction
In general, chemical reactions can be divided into two types - reversible and irreversible.
NSOU GE-CH-11 23
Irreversible reactions are said to go to completion, implying thereby that the reaction
proceeds until at least one of the reactants is completely used up. So in the irreversible
reactions, the same reactants cannot be formed back from the products under normal set
of experimental conditions. On the other hand, all those reactions which appear not to
proceed beyond a certain stage after sometime even when the reactants are still available
are reversible reactions. This stage of the system is the dynamic equilibrium and the rates
of the forward and reverse reactions become equal. It is due to this dynamic equilibrium
stage that there is no change in the concentrations of various species in the reaction
mixture. The present unit discusses the meaning of this state of equilibrium and how the
equilibrium composition of the reactants and products are correlated with change in Gibb’s
free energy.
If neither energy factor nor entropy factor predominates then ΔH is numerically equal
to TΔS. Then the value of ΔG will be zero. Under these circumstances, the reaction will
be in a state of equilibrium, i.e., no net reaction will occur in any direction.
Standard free energy change (Δ ΔGo): The standard free energy change of a reaction
is defined as the free energy change of a reaction when the reactants and products are in
their standard states (25 oC, 1 atm pressure). At this condition the equation 1 may be
written as
ΔGo = ΔHo – TΔSo (2)
where ΔHo is the standard enthalpy change of the reaction and ΔSo is the standard entropy
change of the reaction at the temperature T.
the rate of the forward reaction, rf [A]a [B]b i.e. rf k f [A]a [B]b
NSOU GE-CH-11 25
and the rate of the reverse reaction, rr M m [N]n i.e, rr k r [M]m [N]n
where kf and kr are proportionality constants and square brackets represent the molar
concentration of the entities enclosed. The constant k-f is rate constant of the forward
reaction and kr is the rate constant of the reverse or backward reaction. Now at equilibrium,
the rate of forward reaction is equal to the rate of the reverse reaction, that is,
kf [M]m [N]n
K eq (3)
kr [A]a [B]b
The constant Keq is called the equilibrium constant of the reaction. Equilibrium constant
being the ratio of the concentrations raise to the stoichiometric coefficients. Therefore, the
n
unit of the equilibrium constant = Mole L1 .
aA + bB mM + nN
where the reactants and the products are assumed to be ideal gases.
We know that chemical potential (i.e., Gibbs free energy) of reactants consisting of a
moles of A and b moles of B is given by the expression
Greactants = aμA + bμB (4)
where μA and μB are the chemical potentials of the species A and B, respectively. Similarly,
for the products we have
Gproducts = mμM + nμN (5)
In each case, pressure and temperature are constant. The free energy of the reaction
is equal to the difference between the free energy of the products and that of the reactants.
So
(ΔG)reaction = Greactants – Gproducts
= (mμM + nμN) – (aμA + bμB) (6)
Now at equilibrium, the free energy change ΔG = 0. So the equation 6 becomes
(mμM + nμN) – (aμA + bμB)= 0 (7)
The chemical potential of the ith species in the gaseous state is given by
i i0 RT ln pi (8)
where, pi is the partial pressure of the ith component and the i0 is its standard chemical
NSOU GE-CH-11 27
potential (i.e., when the partial pressure of the ith component is unity). So from equation
7 and 8 we can obtain
M M N
m 0 RT ln p n 0 RT ln p
N
A A B
a 0 RT ln p b 0 RT ln p 0
B
(9)
pm n
or RT ln M pN
paA p B
b
m0M n0N a0A b0B
G 0products G 0reactant
G 0 reaction (10)
pm n
M pN
0
or, e G RT (11)
paA p B
b
Since ΔG0 depends only on temperature and R is the gas constant, hence the right
hand side of equation 11 is a constant at constant temperature. Thus
pm n
M pN constant K p (12)
paA p B
b
i i0 RT ln x i (13)
xm n
M xN Kx (14)
x aA x B
b
28 NSOU GE-CH-11
If the chemical potentials of the various species are expressed in terms of molar
concentrations (ci), then
i i0 RT ln ci (15)
cm n
M cN Kc (16)
caA cB
b
Equation 12, 14 and 16 are the expressions of Kp, Kx, and Kc respectively.
van’t Hoff reaction isotherm:
From equation 11 and 12 we can have
0
K eq eG RT (17)
or G 0 RT ln K eq (18)
The equation 18 is called van’t Hoff reaction isotherm. It may be used to calculate the
change in free energy of a reaction in the standard condition (ÄG0) from the equilibrium
constant and vice-versa.
piV = niRT
n
pi i RT ci RT
V
Where, ci (=ni/V) is the molar concentration of the ith component in the mixture of
total volume V.
For the reaction
aA + bB mM + nN
We can write, pA = cART, pB = cBRT, pM = cMRT, pN = cNRT
So from equation 12 we get
m n
pm n
MpN c M RT c N RT m n
cM cN
Kp (RT)(m n) (a b)
p aA p B
b
cA RT a cBRT b caA cB
b
= KcRTΔn
where Δn = (m + n) – (a + b)
Relation between Kp and Kx:
In an ideal gaseous mixture, each component obeys Dalton’s law of partial pressure,
i.e.,
where, P is the total pressure and pi is the partial pressure of the ith component with
mole fraction xi in the mixture.
For the reaction
aA + bB mM + nN
we have pA = xAP, pB = xBP, pM = xMP, pN = xNP
m n
pm n
MpN x M P cN P xm xn
Kp M N (P)(m n) (a b) K x P n (20)
paA pB
b
x A P a c B P b x aA x B
b
where Δn = (m + n) – (a + b)
30 NSOU GE-CH-11
K p K c RT n K x P n (21)
equilibrium has shifted to the right of the equation. In this final state of equilibrium, then,
more amount of HI is formed as compared to earlier equilibrium state. Just the opposite
would be the fate of the reaction if, instead of H2 some HI is added to the system under
equilibrium. In short, by changing the concentration at equilibrium, the reaction will move
forward or in the reverse direction so that the value of the equilibrium constant remains the
same.
2.8. Summary
In this unit we have discussed reversible and irreversible reaction and the free energy
change in a chemical reaction. Then we have discussed the meaning of chemical equilibrium
and the law of mass action. A general expression for the equilibrium constant was derived
from the basic principles of thermodynamics. Different forms of the general expression
NSOU GE-CH-11 33
were then utilised in understanding the equilibrium. We then learnt Le Chatelier’s principle
and its use in predicting the shift in the position of equilibrium by the changes brought about
in concentration, temperature and pressure in the system.
2.9. Questions
1. Describe the Law of mass action.
2. Derive the relation between Kp and Kc.
3. Write a short note on Le Chatelier’s principal.
4. Explain the effect of pressure and temperature in the following reaction
N2 (g) + O2 (g) 2NO (g); ΔH = – 43.2 kJ
5. At 20 oC the Kp of the reaction 2NO (g) + Cl2 (g) 2NOCl (g) is 1.9 ×
10-3 atm-1. Calculate the value of K-c at that temperature.
(Ans. 45649.4 L mol-1)
6. At equilibrium, the concentrations of N2 = 3.0 × 10–3 M, O2 = 4.2 × 10–3 M and
NO = 2.8 × 10–3 M in a sealed vessel at 800 K. What will be Kc for the
reaction
N2 (g) + O2 (g) 2NO (g) (Ans. 0.622)
Unit 3 Ionic Equilibrium
Contents
3.0. Objectives
3.1. Introduction
3.2. Strong and weak electrolytes
3.3. Degree of ionization
3.4. Ionization of weak acids and bases
3.5. Ionization constant and ionic product of water
3.6. pH scale
3.7. Common ion effect
3.8. Salt hydrolysis - hydrolysis constant, degree of hydrolysis and pH for different
salts
3.9. Buffer solutions
3.10. Solubility and solubility product of sparingly soluble salts
3.11. Summary
3.12. Questions
3.0. Objectives
After going through this unit, we will be able to know about
Strong and weak electrolytes.
Degree of ionization.
Ionic product of water.
pH scale and pH calculation.
Common ion effect and its applications.
Salt hydrolysis.
Buffer solution and buffer action.
NSOU GE-CH-11 35
3.1. Introduction
There are numerous equilibria that involve ions only. In the following sections we will
study the equilibrium involving ions. It is well known that the aqueous solution of sugar
does not conduct electricity. However, when common salt (NaCl) is added to water it
conducts electricity. Also, the conductance of electricity increases with increase in
concentration of common salt. Michael Faraday classified the substances into two categories
based on their ability to conduct electricity. One category of substances conduct electricity
in their aqueous solutions and are called electrolytes while the other do not conduct
electricity are, referred to as non electrolytes. Faraday further classified electrolytes into
strong and weak electrolytes. Strong electrolytes on dissolution in water are ionized almost
completely, while the weak electrolytes are only partially dissociated. For example, an
aqueous solution of sodium chloride is comprised entirely of sodium ions and chloride
ions, while that of acetic acid mainly contains unionized acetic acid molecules and only
some acetate ions and hydronium ions. This is because 100% ionization in case of sodium
chloride as compared to less than 5% ionization of acetic acid which is a weak electrolyte.
It should be noted that in weak electrolytes, equilibrium is established between ions and
unionized molecules. This type of equilibrium involving ions in aqueous solution is called
ionic equilibrium.
Examples:
Acid: HCl, HNO3, H2SO4 etc.
Base: NaOH, KOH, Ba(OH)2 etc.
Salt: NaCl, KCl, CaCl2, KNO3, CuSO4 etc.
Weak Electrolytes: These are the substances which dissociate to a small extent in
aqueous solution and hence conduct electricity also to a small extent.
Examples:
Acids: H2CO3, CH3COOH, HCN etc.
Base: NH4OH, Al(OH)3, Zn(OH)2 etc.
Salt: MgCO3, PbCl2, BaCO3 etc.
Consider a weak electrolyte AB, which dissociates in water and forms A+ and B-. So
an equilibrium will be established in solution.
Let the initial concentration of the weak electrolyte, AB is c mol/L and α is the degree
of dissociation. Then at equilibrium the concentration of A+ and B- will be αc mol/L and
concentration of AB will be (c- αc) mol/L.
AB A+ + B-
Initial concentration: c 0 0
Equilibrium concentration: c(1–α) αc αc
NSOU GE-CH-11 37
So according to the law of mass action the equilibrium constant can be written
as
A B
K where K is the dissociation constant or ionization constant.
AB
c c 2c
So, K (2)
c(1 ) 1
K
K 2c or, (3)
c
So if we know the value of the dissociation constant of the weak electrolyte and
the concentration of the weak electrolyte we can find its degree of dissociation or
ionization.
5. Common ion effect: Due to the common ion effect, the degree of ionization of
weak electrolytes is decreased or suppressed. The suppression of degree of
ionization of weak electrolyte by the addition of strong electrolyte having common
ion is called common ion effect.
H A
Ka (4)
HA
Ka is the dissociation constant of the acid. If the initial concentration of the
acid is c and α is the degree of dissociation, then the equilibrium concentrations would
be,
HA H+ + A–
Initial concentration: c 0 0
Equilibrium concentration: c(1–α) αc αc
Substituting these values in equation 4, we get
2c
Ka (5)
1
for a weak acid á << 1; we can take (1 – α) = 1.
Ka
(6)
c
From this equation we can calculate the degree of dissociation of a weak acid.
NSOU GE-CH-11 39
Now the concentration of H+ ion in the solution can be calculated using equation 6
H c K a c K c (7)
a
c
BOH B+ + OH-
B OH
Kb
(8)
BOH
Kb is the dissociation constant of the weak base. If the initial concentration of the
base is c and α is the degree of dissociation, then the equilibrium concentrations would
be,
BOH B+ + OH-
Initial concentration: c 0 0
Equilibrium concentration: c(1-α) αc αc
Substituting these values in equation 8, we get
2c
Kb (9)
1
for a weak base α << 1; we can take (1 – α) = 1.
Kb
(10)
c
From this equation we can calculate the degree of dissociation of a weak base.
Now the concentration of OH- ion in the solution can be calculated using equation 10
OH c K b c K c (11)
b
c
40 NSOU GE-CH-11
H OH
K
(14
H 2O
K[H2O] = [H+] [OH–] (15)
In pure water the concentration of H2O molecules is approximately 55.4 M and since
the dissociation of H2O is negligibly small in comparison with its concentration, we can
safely assume that the concentration of H2O at equilibrium is a constant quantity. Thus, K
[H2O] in eq. 15 can be replaced by a new constant.
Kw = [H+] [OH–] (16)
where Kw is called the dissociation constant of water or, more commonly, the ionic product
of water. Hence ionic product of water may be defined as the product of the molar
concentration of H+ and OH- ions.
The value of Kw at 298 K (25 0C) is experimentally determined as Kw = 1.0 ×
10-14.
Hence, [H+] [OH–] = 1.0 × 10–14 (17)
Since the amount of H+ and OH- produced by the dissociation of pure water is equal,
concentration of each ion in solution is given by
NSOU GE-CH-11 41
3.6. pH scale
From the above discussion we have seen that the equilibrium of water is given by
equation 14
H OH
K
2
H O
Now suppose that we add a small quantity of an acid to water, thereby increasing
the concentration of H’ ions at equilibrium. The equilibrium will immediately shift back to
oppose the effect of this increase by the combination of the added H+ ions with some
OH- ions to form undissociated water till eq.14 is satisfied. Once the equilibrium is re-
established, the concentration of the hydrogen ion will be more than the concentration of
the hydroxyl ion in solution. Hence, at 298 K, whenever the concentration of hydrogen
ion in water is greater than 1.0 × 10-7 M, we call the solution to be acidic and whenever
it is less than 1.0× 10-7 M, we call it a basic solution. It is quite inconvenient to express
these concentrations by using powers of 10. In 1909 a Danish botanist S.P.L. Sorensen
proposed a logarithmic scale (called pH scale) for expressing the concentrations of H+
ions. He defined pH as the negative logarithm of the molar concentration of hydrogen
ions. That is,
pH = – log10 [H+] (19)
42 NSOU GE-CH-11
14
OH K w 1 10 11012
[H] 1 107
or pOH = 12
We can state this in a different way that, in 10-2 M acid solution, the concentration
of OH- ion (10–12 M in the above example) is less than the concentration of H+ ion
(10–2 M) and, the product of the two is always constant, and is equal to 1.0 × 10–14.This
can be expressed as,
pH + pOH = 14 = – log Kw (21)
Thus, in pure water or a dilute solution of an acid or a base, we can express the
concentration of H+ or OH- by simply stating the pH of the solution. We have also studied
that the contribution due to self-ionisation of water is negligible in cases of solution of
strong acids and bases as well as of moderately concentrated solutions of weak acids
and weak bases. However, dealing with very dilute solutions of weak acids and bases,
we cannot neglect the contributions due to self-ionisation equilibrium of water.
NSOU GE-CH-11 43
H CH COO
3
Ka
(22)
CH3COOH
H K CH3COOH
a
or, CH COO (23)
3
Now, suppose that we add some acetate ions in the form of solid sodium acetate
to the above solution. Sodium acetate is added in the solid form so as not to cause any
change in the volume of the solution. Sodium acetate being a strong electrolyte will
dissociate almost completely to give Na+ ions and CH3COO– ions in solution. The acetate
ions so added will disturb the equilibrium of acetic acid. The equilibrium will, therefore,
shift to left producing more of undissociated acetic acid in order to counteract the effect
of added acetate ions according to Le Chatelier’s principle. The net result is that the
dissociation of the acid has been suppressed by the addition of a common ion (acetate
ion in this case) at equilibrium. Thus any ion which is involved in a chemical equilibrium
and comes from two different sources in solution is known as ‘common ion’ and its effect
finds great use in the study of buffer solutions and the solubility of sparingly soluble salts
which we will discuss in the next few sections.
reaction with water. Depending on the behaviour towards hydrolysis there are four different
types of salts.
(i) Salt of strong acid and strong base (eg. HCl + NaOH) NaCl
(ii) Salt of strong acid and weak base (e.g. HCl + NH4OH) NH4Cl
(iii) Salt of weak acid and strong base (e.g. CH3COOH + NaOH)
CH3COONa
(iv) Salt of weak acid and weak base (e.g. CH3COOH + NH4OH)
CH3COONH4
Let us consider the acid base behaviour of the different type of salts.
(i) Salt of strong acid and strong base: The cations of the strong bases and the
anions of the strong acids do not get hydrolysed. Therefore the salts of this category do
not show any acid-base behaviour and are neutral.
(ii) Salt of strong acid and weak base: the salts of this type dissociate in aqueous
solutions to give a cation of a weak base and the anion belonging to strong acid. For
example,
NH4Cl dissociates as:
NH4Cl (aq) → NH4+ (aq) Cl– (aq)
The anion from the strong acid does not get hydrolysed but the cation get hydrolysed
as per the following equation:
NH4+ (aq) + H2O (l) NH4OH + H+ (aq)
which produces a weak base and excess of hydrogen ions in solution, thus rendering the
solution acidic. Hence, we can say that the salt of a weak base and a strong acid will
produce an acidic solution.
Let c mol/litre is concentration of the NH4Cl and h is its degree of hydrolysis,
then
NH4+ + H2O NH4OH + H+
Initial concentration: c 0 0
Equilibrium concentration: c(1- h) hc hc
NSOU GE-CH-11 45
NH 4OH H
K
NH H O
4 2
NH 4OH H
or, K H 2O
NH
4
NH 4OH H
or, Kh (24)
NH
4
where Kh is the hydrolysis constant of salt NH4Cl or NH4+ ion. Also we have for weak
base NH4OH which will get dissociate in water as
NH4OH NH4+ + OH–
NH OH
4
Kb
(25)
NH 4OH
From equation 24 and 25 we can write
Kh × Kb = [H+] [OH–] = Kw
Kw
So, Kh (26)
Kb
NH 4OH H ch ch h 2c
Kh h 2c
NH (1 h)c 1 h
4
46 NSOU GE-CH-11
Kh Kw
Hence, h (27)
c K b .c
From this equation we can calculate the degree of hydrolysis of the salt.
Kw K w .c
Now, [H ] hc c
K b .c Kb
1 1 1
or, log H log K w log c log K b
2 2 2
1 1 1
or, pH log1014 log c pK b
2 2 2
1 1
or pH 7 pK b log c (28)
2 2
This is the equation for the pH of a solution of salt of strong acid and weak base.
(iii) Salt of weak acid and strong base: the salts of this type dissociate in aqueous
solutions to give a anion of a weak acid and the cation belonging to strong base. For
example, CH3COONa dissociates as:
CH3COONa (aq) → Na+ (aq) + CH3COO– (aq)
in this case the cation does not get hydrolysed but the anion get hydrolysed as per the
following equation.
CH3COO– (aq) + H2O (l) CH3COOH (aq) + OH– (aq)
So in the solution there will be excess OH- ion. Hence the salts of a weak acid and
a strong base, when dissolved in water, will produce a basic solution.
By considering the similar method as above we can show that if c mol/litre is the
concentration of the CH3COONa and h is its degree of hydrolysis then
We can get
Kw
Kh (29)
Ka
NSOU GE-CH-11 47
Kw
h (30)
K a.c
1 1
pH 7 pK a log c (31)
2 2
(iv) Salt of weak acid and weak base: the salts of this type dissociate in aqueous
solutions to give a anion of a weak acid and the cation belonging to a weak base. For
example, ammonium acetate, CH3COONH4 dissociates as:
CH3COONH4 (aq) → NH4+ (aq) + CH3COO– (aq)
In this case both the cation as well as the anion would undergo hydrolysis according
to the following equations
NH4+ (aq) + H2O (l) NH4OH + H+ (aq)
CH3COO– (aq) + H2O (l) CH3COOH (aq) + OH– (aq)
Hydrolysis of one ion gives hydrogen ions, whereas that of the other ion gives hydroxyl
ions in solution. Therefore, the pH of the solution will depend on the extent of the hydrolysis
of the two ions. If NH4+ hydrolyses to a greater extent than CH3COO– ion, the solution
will be acidic and, if the reverse is true, then the solution will be basic. If the extent of
hydrolysis is exactly equal, then the solution should be neutral as if no hydrolysis is taking
place.
By considering the similar method we can establish that if c mol/litre is the concentration
of the CH3COONH4 and h is its degree of hydrolysis then
We can get
Kw
Kh (32)
Ka Kb
Kw
h (33)
Ka Kb
1 1
pH 7 pK a pK b (34)
2 2
48 NSOU GE-CH-11
From the equation 34, we can see that the pH of the solution is independent of the
concentration of the salt.
If pKa = pKb then the pH of the solution is 7.
When pKa >> pKb then the pH > 7, so the solution will be basic.
When pKa << pKb then the pH < 7, so the solution will be acidic.
Buffer solutions are the ones that resist a change in its pH on adding a small
amount of an acid or a base.
In laboratory reactions, in industrial processes and in the bodies of plants and animals,
it is often necessary to keep the pH nearly constant despite the addition of acids and
bases. The oxygen carrying capacity of haemoglobin in our blood and activity of the
enzymes in our cells depends very strongly on the pH of our body fluids. pH of the
blood is very close to 7.4 and pH of saliva is close to 6.8. Fortunately, animals and
plants are protected against sharp changes in pH by the presence of buffers.
There are two kinds of commonly used buffer-solutions
i) A weak acid and a soluble ionic salt of the weak acid forms a buffer with pH
less than 7 are called acidic buffers e.g. acetic acid and sodium acetate.
ii) A weak base and a soluble ionic salt of the weak base forms a buffer with pH
above 7 are called basic buffers. e.g. ammonium hydroxide and ammonium
chloride.
Let us consider a buffer solution containing acetic acid, CH3COOH and sodium
acetate CH3COONa to understand about how buffer resist the change of pH in the
solution. In acetic acid - sodium acetate buffer CH3COOH acts as acid reserve while
CH3COONa (or CH3COO- ions) works as the base reserve. In the solution mixture the
NSOU GE-CH-11 49
added components dissociate as follows. The weak acid dissociates partially while the
salt undergoes complete dissociation.
CH3COOH (aq) + H2O (l) CH3COO– (aq) + H3O+ (aq)
CH3COONa (aq) → Na+ (aq) + CH3COO– (aq)
If we add a strong acid such as HCl to this solution, it produces H3O+. These added
H3O+ (acid) react with an equivalent amount of the base reserve [CH3COO–] to generate
undissociated acetic acid. The reaction being
CH3COO– (aq) + H3O+ (aq) → CH3COOH (aq) + H2O (l)
The net effect of this reaction is that there is a slight increase in the concentration
of the acid reserve and an equivalent decrease in the concentration of the base reserve.
The effective reaction being
CH3COOH (aq) + HCl (aq) CH3COOH (aq) + NaCl (aq)
Similarly, when small amount of a strong base like NaOH is added, it generates OH-
ions. These additional OH– neutralize some of the H3O+ ions present in the solution.
Since one of the products of the acid dissociation equilibrium (eq) is used up, there is
some more ionisation of CH3COOH to re-establish the equilibrium. The net result is the
neutralization of OH- by CH3COOH. In other words we can say that the added OH-
ions (base) react with the acid reserve to produce CH3COO– ions.
OH– (aq) + CH3COOH (aq) → CH3COO– (aq) + H2O (l)
The effective reaction is the reaction of the added base with acid reserve.
NaOH (aq) + CH3COOH (aq) → CH3COONa (aq) + H2O (l)
The net effect of this reaction is that there is a slight increase in the concentration
of the base reserve and an equivalent decrease in the concentration of the acid reserve.
So we can observe that the added acid or the base only cause minor changes in
the concentrations of the weak acid and the salt. The concentration of the hydronium ions
and thereby the pH does not change significantly.
Now the pH of a buffer solution can be calculated from the Henderson’s equation.
Let us derive the expression for an acidic buffer system that we have discussed above.
In acetic acid – sodium acetate buffer the central equilibrium is
CH3COOH (aq) + H2O (l) CH3COO– (aq) + H3O+ (aq)
50 NSOU GE-CH-11
H O CH COO
3
Ka
3
3
CH COOH
Rearranging, we get
H O K CH3COOH
3 a
CH COO
3
The concentration of undissociated acetic acid can be taken as total acid concentration
[Acid] and that of sodium acetate as the total salt concentration [Salt]. Thus the above
equation may be re written as
H O K [Acid]
3 a
[Salt]
[Acid]
log H3O log K a log
[Salt]
[Acid]
or pH pK a log
[Salt]
[Salt]
or, pH pK a log (35)
[Acid]
[Salt]
pOH pK b log (36)
[Base]
NSOU GE-CH-11 51
x y
A y Bx
K
(37)
A x By
We can see that, as compared to the solubility of AgCl in water (1.37 × 10–5 M),
its solubility in presence of 0.01 M AgNO3 is almost 103 times less.
3.11. Summary
In summary we have started the discussion explaining electrolytes and non electrolytes.
Then we focused our discussion on strong and weak electrolytes followed by degree of
dissociation for weak electrolytes and the factors affecting the degree of dissociation.
Ionization of weak acid and weak bases are discussed. After that, ionic product of water
was explained and the concept was extended for establishing the concept of pH scale.
We have found the effect of common ion in dissociation of weak acid and bases. We
have learned about the salt hydrolysis and calculated the hydrolysis constant, degree of
hydrolysis and pH for different types of salts. We have seen that the presence of common
ions in a solution of a weak acid or a weak base suppress its dissociation. Such solutions
act as buffer solutions which resist a change in their pH on addition of small amount of
an acid or a base. The pH of buffer solutions depend on their composition and can be
found by using a simple equation called Henderson’s equation. After that we have discussed
that the product of the concentration of the ions in the solubility equilibrium is a constant
called solubility product (Ksp) and is proportional to the solubility of the sparingly soluble
54 NSOU GE-CH-11
salt. We have found that the presence common ion decreases the solubility of a sparingly
soluble salt. This is called common ion effect and has widespread applications in qualitative
analysis.
3.12. Questions
1. Write a short note on pH scale.
2. What is degree of ionization?
3. What are sparingly soluble salts?
4. What will happen if we pass HCl gas in a saturated solution of NaCl?
5. What is buffer? Give example of one acidic and one basic buffer.
6. Why the pH of a buffer remains constant upon addition of small amount of acid
or base?
7. Why the solution of FeCl3 is acidic?
8. If the solubility of Zinc phosphate in water is S mol/L, then find the expression
for the solubility product of the salt. (Ans. 108 S5)
9. Calculate the pH of 0.1 M NaOH solution. (Ans. 13)
10. If in 1 L of buffer solution 0.2 mol of NH4Cl and 0.1 mol of NH4OH is
dissolved then what will be the pH of the solution? Given that Kb of NH4OH is
1.8 x 10–5. (Ans. 8.96)
Unit 4 Kinetic Theory of Gases
Contents
4.0. Objectives
4.1. Introduction
4.2. Postulates of kinetic theory of gases
4.3. Kinetic gas equation
4.4. Maxwell Boltzmann distribution laws of molecular velocities and molecular
energies
4.5. Average, root mean square and most probable velocities
4.6. Collision Parameters
4.7. Viscosity of gases
4.8. Deviation of real gases from ideal behavior
4.9. Critical phenomena
4.10. Summary
4.11. Questions
4.0. Objectives
After studying this unit, we will be able to:
state the postulates of kinetic theory of gases
1
derive the kinetic gas equation PV = mnc2
3
explain the distribution of molecular speeds and energies
calculate the most probable speed, the average speed and the root mean square
speed
derive an expression to calculate the collision number between gas molecules
calculate the mean free path of molecules
explain the origin and factors responsible for viscosity of gases
56 NSOU GE-CH-11
differentiate between ideal gas and real gas and reasons for deviation from ideal
behaviour
explain the critical phenomenon and critical constants.
4.1. Introduction
The kinetic theory describes a gas as a large number of sub microscopic particles
(atoms or molecules), all of which are in constant, and random motion. The rapidly moving
particles constantly collide with each other and with the walls of the container. Kinetic
theory explains macroscopic properties of gases, such as pressure, temperature, viscosity,
thermal conductivity, and volume, by considering their molecular composition and motion.
The theory depicts that gas pressure is due to the impacts, on the walls of a container,
of molecules or atoms moving at different velocities.
vi. The pressure of the gas is the hits recorded by the molecules on the walls of the
container in which the gas is contained.
vii. The average kinetic energy of gas molecules is directly proportional to absolute
temperature.
viii. This means that the average kinetic energy of molecules is the same at a given
temperature.
Figure 4.1. Velocity of a gas molecule along X-axis placed in a cubic container.
58 NSOU GE-CH-11
Now after collision with the A side of the wall it will reflect and will collide with the
opposite wall at a distant of .
Hence, the number of collision per second for a gas molecule moving along X-axis
u
with a velocity u
l
So for u component of velocity of a molecule change in momentum per second
u 2mu 2
2mu
l l
For the same molecule the change of momentum per second for the v and w
2mv 2 2mw 2
component of the velocity will be and respectively.
l l
Total change of momentum for a molecule per second
2mc2
2m 2
l
2 2
u v w
l
2mnc2
For n number of molecules rate of change of momentum
l
From the Newton’s second law we know that rate of change of momentum is equal
to the force. As the pressure (P) is equal to the force acting per unit area and the total
surface area of the container = 6l2.
1
or, PV mnc 2 (1)
3
This is known as kinetic gas equation.
NSOU GE-CH-11 59
3 mc2
dNc m
2 2kT 2
4 e c dc (2)
N 2kT
3 mc 2
dNc M
2 2RT 2
4 e c dc (3)
N 2RT
1 dN c
The Maxwell distribution is customarily plotted with the function in the y-
N dc
axis and c as the x-axis. The fraction of the molecules in the speed range c to c + dc
dN c
is ; dividing this by dc gives the fraction of the molecules in this speed range per
dc
unit width of the interval. The Maxwell’s distribution of molecular velocities is plotted in
figure 4.2.
We can see from the plot that the fraction of molecules having velocities greater than
zero increases with an increase in velocity, reaches a maximum and then falls off towards
zero at higher velocities. The important features of the curves are as follows:
i) The fraction of molecules with too low or too high velocities is very small.
ii) There is a certain velocity for which the fraction of molecules is maximum. This
is called the most probable velocity.
Effect of temperature on distribution of molecular velocities for a particular gas:
Effect of temperature on distribution of molecular velocities for a N2 gas is plotted
in figure 4.3. The important features are:
i) The entire distribution curve shifts to the right and becomes broader with increase
in temperature
ii) With increase of temperature fraction of the molecules having high velocities
increases considerably.
iii) The most probable velocity increases with increase in temperature but the number
of molecules present at that velocity decreases.
Distribution of molecular energies:
The Maxwell’s speed distribution, eq. 2, can be converted to an energy distribution.
1 1
1
The kinetic energy of a molecule is mc 2 . Then, c 2 2 . Thus equation 2
2
2 m
can be converted to energy distribution equation as
3 1
dN 1 2 2 kT
2N e d (4)
N kT
where dN , is the number of molecules having kinetic energies between ε and ε +dε.
Energy distribution curve is distinctly different shape compared to that of the speed
distribution (figure 4.4). In particular, the energy distribution curve rises sharply at the
beginning and thus it rises much more quickly than the velocity distribution. After passing
the maximum, the energy distribution falls off more slowly than does the velocity distribution.
As usual, the distribution is broadened at higher temperatures, a greater proportion of the
62 NSOU GE-CH-11
molecules having higher energies. As before, the areas under the curves for different
temperatures must be the same.
n c n c n 3c3 .... n i ci
cav 1 1 2 2
n1 n 2 n 3 .... n
From Maxwell’s equation of distribution of velocity, it can be shown that the average
velocity for a gas with molecular mass M is
8RT
Cav (5)
M
From Maxwell’s equation of distribution of velocity, it can be shown that the root
mean square velocity for a gas with molar mass M at temperature T is
3RT
Crms (6)
M
NSOU GE-CH-11 63
3RT
Cmp (7)
M
satisfactorily account for various gaseous properties such as viscosity, diffusion, thermal
conductivity, mean free path, number of collision the molecules undergo. It can be imagined
that when two molecules collide, the effective area of the target is πσ2. The quantity πσ2
is called the collision cross-section of the molecule because it is the cross-sectional area
of an imaginary sphere surrounding the molecule into which the centre of another molecule
cannot penetrate.
Collision number:
Collision number is the number of collisions suffered by a single molecule per unit
time per unit volume of the gas. It can be imagined that a gas molecule having diameter
σ will collide with the molecules present in the area of πσ2. Now if the molecule is
moving with an average velocity of c and n´ number of molecules are present in unit
volume then in one second total number of collisions suffered by the molecule will be
2 c n . Now as all the gas molecules are moving this velocity should be considered
as relative velocity
2 c . So, the number of molecules with which a single molecule
will collide per unit time is given by Z1 2 2 c n . This is the expression of collision
number.
Now as n´ number of molecules are present per unit volume of the gas, total number
of collisions per unit time per unit volume is = 2 2 c n 2 . Since each collision
involves two molecules, the number of collisions of like molecules occurring per unit time
per unit volume of the gas is given by
Z11
1
2
2 2 c n2 1
2
2 c n2
1
2
2 n2
8RT
M
8RT
as c
M
8RT
Z11 22 n 2
M
This number Z11 represent the number of molecular collisions occurring per unit time
per unit volume of the gas, which is known as collision frequency.
NSOU GE-CH-11 65
Now we already know that if a molecule has the average velocity c then the
molecule will face Z1 2 2 c n number of collisions while travelling c distance.
So, Mean free path,
where, σ = collision diameter and n´ = number of molecules present per unit volume.
1
mn c l where, c is average velocity, n’ is the number of gas molecules
3
present in unit volume and m is the mass of the gas molecules.
Effect of temperature and pressure on coefficient of viscosity:
Effect of pressure: The viscosity of gas is independent of pressure. With increase
of pressure though more number of molecules changes between the layers, but they travel
shorter distance due to higher number of collisions. So the change of pressure does not
affect the viscosity for gases.
Effect of temperature: Viscosity in gases arises from molecules traversing layers of
flow and transferring momentum between layers. The momentum transfer is caused by
66 NSOU GE-CH-11
free motion of gas molecules between collisions. So, increase of thermal agitation of the
molecules results in a larger viscosity. Hence, gaseous viscosity increases with increase
of temperature.
1 1
We know, l =
2
and mn c l
2 n 3
1 1 1 m c 1 m 8kT 2 mkT
So, mn c l 2 (9)
3 2
2 n 3 2 2 3 2 2 m 3 3
Figure 4.6. Compressibility factor Z, plotted against pressure for H2, N2 and CO2 at constant temperature.
From the plot of plot of Z versus p (figure 4.7) for nitrogen at different temperatures
varying between -70°C and 50°C it is seen that as the temperature is raised the dip in
the curve becomes smaller and smaller. At 50°C, the curve seems to remain almost
horizontal (Z=1) at the low pressure region. So it can be said that at this region PV
remain constant and thus it obeys the Boyle’s law within this range of pressure at 50°C.
This temperature is called Boyle temperature. The Boyle temperature is different for
different gases.
Causes of deviation and Van der Waals equation of state for real gases:
In order to explain deviations from ideal behaviour, it is necessary to modify the
kinetic theory of gases. Van der Waals realized that two of the assumptions of the kinetic
molecular theory were questionable. The kinetic theory assumes that gas particles occupy
a negligible fraction of the total volume of the gas. It also assumes that the force of
attraction between gas molecules is zero.
The first assumption works at pressures close to 1 atm. But something happens to
the validity of this assumption as the gas is compressed. Imagine for the moment that the
atoms or molecules in a gas were all clustered in one corner of a cylinder. At normal
pressures, the volume occupied by these particles is a negligibly small fraction of the total
volume of the gas. But at high pressures, this is no longer true. As a result, real gases
are not as compressible at high pressures as an ideal gas. The volume of a real gas is
therefore larger than expected from the ideal gas equation at high pressures.
Van der Waals proposed that we correct for the fact that the volume of a real gas
is too large at high pressures by subtracting a term from the volume of the real gas before
we substitute it into the ideal gas equation. He therefore introduced a constant (b) into
the ideal gas equation that was equal to the volume actually occupied by a mole of gas
NSOU GE-CH-11 69
particles. Because the volume of the gas particles depends on the number of moles of
gas in the container, the term that is subtracted from the real volume of the gas is equal
to the product of number of moles of gas and b.
P(V - nb) = nRT
When the pressure is relatively small, and the volume is reasonably large, the nb term
is too small to make any difference in the calculation. But at high pressures, when the
volume of the gas is small, the nb term corrects for the fact that the volume of a real gas
is larger than expected from the ideal gas equation.
The assumption that there is no force of attraction between gas particles cannot be
true. If it was, gases would never condense to form liquids. In reality, there is a small
force of attraction between gas molecules that tends to hold the molecules together. This
force of attraction has two consequences: (i) gases condense to form liquids at low
temperatures and (ii) the pressure of a real gas is sometimes smaller than expected for
an ideal gas.
To correct for the fact that the pressure of a real gas is smaller than expected from
the ideal gas equation, van der Waals added a term to the pressure in this equation. This
term contained a second constant (a) and has the form: n2a/V2. The complete van der
Waals equation is therefore written as follows.
n 2a
P 2 (V nb) nRT (10)
V
This is the van der Waals equation for n moles of gas which contains a pair of
constants ‘a’ and ‘b’. The’ quantities ‘a’ and ‘b’ are called the van’der Waals constants
or parameters. It may be pointed that ‘b’ is a measure of the molecular size and ‘a’ is
related to the intermolecular interaction. The SI unit of a is N m4 mol-2 and the unit of
b is m3 mol-1.
on the gas or ii) decrease the temperature of the gas. Both these processes will tend to
bring the molecules closer and liquefaction may take place. But it is observed practically
that temperature is the dominant factor for liquefaction and pressure assumes a secondary
role. For every real gas a temperature is observed above which it cannot be liquefied
even on the application of very high pressures. This characteristic temperature of a real
gas above which it cannot be liquefied is called “Critical Temperature” represented by
TC.
The importance of critical temperature of a gas was first discovered by Andrew in
his experiments on pressure-volume relationships (isotherms) of carbon dioxide gas at a
series of temperatures. The isotherm of carbon dioxide determined by him at different
temperatures is shown in figure 4.9.
Critical constants:
Critical temperature, TC is the maximum temperature at which a gas can be liquefied
that is a temperature above which liquid cannot exist.
Critical Pressure, PC is the minimum pressure required for liquefaction to take place
at the critical condition.
Critical volume, VC is the volume occupied by one mole of the gas at critical
temperature and critical pressure.
72 NSOU GE-CH-11
It can be shown that the values of these parameters are given by the following
expressions:
8a
TC
27Rb
a
PC
27b2
VC = 3b
where R is the universal gas constant.
These critical constants are related by the following relation
RTc 8
(11)
Pc Vc 3
4.10. Summary
In this unit, we have discussed some characteristic microscopic features of gases. It
has been shown how a simple kinetic molecular model of the gas can be used to derive
an equation to calculate the pressure exerted by a gas. This equation can be used further
to derive the ideal gas equation. This model is useful in showing how the constant collisions
between molecules are responsible for a distribution of the speed of molecules. Further,
this model helps us in deriving expressions for various kinds of speeds. We have also
evolved a method of calculating the total collision frequency and the mean free path
assuming hard sphere model for the molecules. We have discussed the viscosity of gases
and effect of temperature on the viscosity. The deviation of real gas from ideal behaviour
was explained and van der Waals equation was derived. Finally the critical phenomenon
and critical constants were explained for real gases.
4.11. Questions
1. Write down the postulates of kinetic theory of gases.
2. Derive the expression of pressure for an ideal gas using kinetic molecular theory
of gases.
3. Write down the expression for Maxwell distribution of molecular velocities.
NSOU GE-CH-11 73
4. What is mean free path of a gas? What is the relation between mean free path
and collision diameter?
5. How does the viscosity of gas vary with temperature?
6. What is compressibility factor of a gas?
7. What are the reasons for deviation of real gases from ideal behaviour?
8. Write down the definitions of critical constants.
9. Calculate the average velocity of CO2 molecule present in 1 gm of CO2 gas at
27 0C. (Ans. 3.8 x 104 cm/s)
10. If TC = 304.2 K and PC = 72.8 atm for a gas, then calculate the value of va
der Walls constants. (Ans. a = 3.63 atm L2 mol-2, b = 0.0428 L mol-1)
Unit 5 Liquids
Contents
5.0. Objectives
5.1. Introduction
5.2. Surface tension
5.3. Determination of surface tension using stalagmometer
5.4. Viscosity of a liquid
5.5. Determination of coefficient of viscosity using Ostwald viscometer
5.6. Summary
5.7. Questions
5.0. Objectives
After studying this unit, we will be able to
the surface tension and origin of surface tension of a liquid
effect of temperature in the surface tension
determine the surface tension of a liquid
explain the viscosity of a liquid and effect of temperature on viscosity
determine the viscosity by using Ostwald viscometer.
5.1. Introduction
Intermolecular forces are stronger in liquid state than in gaseous state. Molecules in
liquids are so close that there is very little empty space between them and under normal
conditions liquids are denser than gases. Molecules of liquids are held together by attractive
intermolecular forces. Liquids have definite volume because molecules do not separate
from each other. However, molecules of liquids can move past one another freely, therefore,
liquids can flow, can be poured and can assume the shape of the container in which these
are stored. In the following sections we will look into some of the physical properties
of the liquids such as surface tension and viscosity.
NSOU GE-CH-11 75
Figure 5.1. Forces acting on a molecule on liquid surface and on a molecule inside the liquid.
76 NSOU GE-CH-11
Surface tension is defined as the force acting per unit length perpendicular to the line
drawn on the surface of liquid. It is represented by Greek letter γ (Gamma). It has
dimensions of kg s–2 and in SI unit it is expressed as Nm–1.
Surface Energy: The effect of surface tension is to reduce the area of the surface
to a minimum. If we wish to increase the area of the surface of a liquid, we have to work
against the force of surface tension. The energy required to increase the surface area of
the liquid by one unit is defined as surface energy of the liquid. Its SI unit is Jm–2.
Effect of Temperature on Surface Tension: The magnitude of surface tension of
a liquid depends on the attractive forces between the molecules. When the attractive
forces are large, the surface tension is large. Increase in temperature increases the kinetic
energy of the molecules and effectiveness of intermolecular attraction decreases, so surface
tension decreases as the temperature is raised.
Some Effects of Surface Tension: Some important effects can be explained by surface
tension.
1. Due to presence of surface tension in liquids every liquid try to reduce the area
of the surface to a minimum. Hence drops of a liquid or the drops of rain are spherical
in shape as sphere has minimum surface area for a given volume.
2. As the surface of a liquid tries to contract due to tension, so it behaves like a
membrane. Also the distances between the surfaces of molecules are less. So an iron
needle can float in the water.
3. Capillary action of a liquid is a well known phenomenon which can be explained
in terms of surface tension. If the liquid in the glass tube is water, the water is drawn
slightly up the walls of the tube by adhesive forces between water and glass. The interface
between the water and the air above it, called a meniscus, is concave, or curved in.
With liquid mercury, the meniscus is convex, or curved out. Cohesive forces in mercury,
consisting of metallic bonds between Hg atoms, are strong; mercury does not wet glass.
The effect of meniscus formation is greatly magnified in tubes of small diameter, called
capillary tubes. In the capillary action, the water level inside the capillary tube is
noticeably higher than outside. The soaking action of a sponge depends on the rise of
water into capillaries of a fibrous material, such as cellulose. The penetration of water
into soils also depends in part on capillary action. Conversely, mercury- with its strong
cohesive forces and weaker adhesive forces- does not show a capillary rise. Rather,
mercury in a glass capillary tube will have a lower level than the mercury outside the
capillary.
NSOU GE-CH-11 77
4. When a drop of liquid spreads into a film across a surface, we say that the liquid
wets the surface. Whether a drop of liquid wets a surface or retains its spherical shape
and stands on the surface depends on the strengths of two types of intermolecular forces.
The forces exerted between molecules holding them together in the drop are cohesive
forces, and the forces between liquid molecules and the surface are adhesive forces.
If cohesive forces are strong compared with adhesive forces, a drop maintains its shape.
If adhesive forces are strong enough, the energy requirement for spreading the drop into
a film is met through the work done by the collapsing drop. Water wets many surfaces,
such as glass and certain fabrics. This characteristic is essential to its use as a cleaning
agent. If glass is coated with a film of oil or grease, water no longer wets the surface
and water droplets stand on the glass. Adding a detergent to water has two effects: The
detergent solution dissolves grease to expose a clean surface, and the detergent lowers
the surface tension of water. Lowering the surface tension means lowering the energy
required to spread drops into a film. Substances that reduce the surface tension of water
and allow it to spread more easily are known as wetting agents. They are used in
applications ranging from dish washing to industrial processes.
78 NSOU GE-CH-11
Figure 5.3. Drawing of a stalagmometer and the formation of liquid drop from a capillary tube in a
stalagmometer.
NSOU GE-CH-11 79
where w, is the weight of the drop and 2πr is the outer circumference of the capillary
tube.
From the above expression, it is clear that the surface tension of a liquid can be
determined if the weight of a single drop ‘w’ and the outer radius of the dropping tube,
‘r’ are known.
If, we have two liquids, such that
w1 2 r 1 and w 2 2 r 2
w1 1
then, we can say that (2)
w2 2
If γ1 for one of the liquids is known, γ2 for other liquid can be determined without
needing a measurement of r, the outer radius of the dropping end of the capillary, provided
the weights of the individual drops of two liquids are known. This method of determination
is also known as Drop Weight Method. Alternatively, the surface tension can also be
determined using the Drop Number Method as given below.
Instead of finding the weights of individual drops, it is easier to count the number
of drops formed by equal volumes of two liquids. With two different liquids, the weights
of equal volumes are proportional to their densities. If nl and n2 are the number of drops
formed by the same volume V of the two liquids, then; vl, the volume of a single drop
of first liquid (i.e., liquid one) = V/nl.
Thus, weight of the single drop of the first liquid = w1 = V/n1.d1.g, where dl is the
density of the first liquid.
And the weight of the single drop of the second liquid = w2 = V/n2.d2.g, where d2
is the density of the second liquid.
Substituting the above values of wl and w2 in equation 2, we get
V d1
d1g
1 n n n d
1 1 2 1 (3)
2 V d 2 n1d 2
d 2g
n2 n2
where, γ1 and γ2 are the surface tensions of two individual liquids, and d1 and d2 are
their densities, respectively. Thus, for the determination of surface tension of any liquid,
80 NSOU GE-CH-11
the number of drops produced from equal volume of two liquids and their densities must
be known, in addition to the surface tension of the reference liquid (e.g. water).
When a liquid flows over a fixed surface, the layer of molecules in the immediate
contact of surface is stationary. The velocity of upper layers increases as the distance of
layers from the fixed layer increases. This type of flow in which there is a regular gradation
of velocity in passing from one layer to the next is called laminar flow. If we choose any
layer in the flowing liquid (Fig.5.4), the layer above it accelerates its flow and the layer
below this retards its flow. If the velocity of the layer at a distance dz is changed by a
du
value: ‘du’ then velocity gradient is given by the amount . A force is required to
dz
maintain the flow of layers. This force is proportional to the area of contact of layers and
velocity gradient i.e.
F A (A is the area of contact)
du du
F (where, is velocity gradient; the change in velocity with distance)
dz dz
NSOU GE-CH-11 81
du
F A
dz
du
F A (4)
dz
‘η’ is proportionality constant and is called coefficient of viscosity. Viscosity coefficient
is the force when velocity gradient is unity and the area of contact is unit area. Thus ‘η’
is measure of viscosity. SI unit of viscosity coefficient is 1 Newton second per square
metre (N s m–2) = pascal second (Pa s = 1kg m–1s–1). In CGS system the unit of
coefficient of viscosity is poise (named after great scientist Jean Louise Poiseuille).
1 poise = 1 g cm–1s–1 = 10–1 kg m–1s–1
Greater the viscosity, the more slowly the liquid flows. The viscosity is very much
influenced by the shape, size and the chemical nature of the liquid molecules. The greater
the size of the molecules and the higher the molar mass, the higher will be the viscosity
because the greater will be the intermolecular interactions. The hydrogen bonds also
enhance the coefficient of viscosity to a large extent. It is, indeed, the presence of a
network of hydrogen bonds which accounts for the very high viscosity of glycerol.
Incidentally, the larger the number of hydroxyl groups in a molecule, the more complex
will be the network of hydrogen bonds and the greater will be the resistance of a liquid
to flow. In long chain hydrocarbons or polymeric compounds the viscosity increases with
the increase in the length of the molecular chain. Due to this reason, heavy hydrocarbon
oil and grease (which are used as lubricants) have fairly high viscosity values.
Effect of Temperature on viscosity of Liquids:
In liquids, as the temperature raises, the kinetic energy of the molecules increases
and the intermolecular forces of attraction become weak, resulting in the subsequent
decrease in the viscosity. The value of the coefficient of viscosity appreciably drops as
the temperature of liquid increases such that for each degree rise in temperature there is
about two percent decrease in the viscosity. The viscosity and temperature are related
to each other by the following expression:
A
log B
T
where A and B are constants for a given liquid and T is the absolute temperature.
82 NSOU GE-CH-11
pr 4 t
(5)
8Vl
where = coefficient of viscosity of the liquid
V = volume of the liquid flowing out of the tube
t = time in which the volume V flows
r = radius of the tube
1 = length of the tube
p = driving pressure necessary to maintain uniform rate of flow of volume V, of the
liquid.
This involves the use of Ostwald Viscometer (Figure 5.5) in which a fixed volume
of a liquid is allowed to fall under its own weight or the force of gravity, and the time
required for a given volume of the liquid to flow is noted. Obviously the driving pressure
p is replaced by h.d.g, where h is the height of the liquid and d is its density and g is
the acceleration due to gravity. Therefore,
p = h.d.g
Substituting h.d.g. for p in Poiseuille’s Equation 5, we get,
r 4 .h.d.g. t
(6)
8Vl
If equal volumes of the two liquids (1 and 2) are allowed to fall through the same
capillary tube under identical conditions of temperature and pressure then, from Eq. 5.6
by comparison, we have
1 d1.t1
(7)
2 d 2 .t 2
where η1, d1and t1 are, respectively, the coefficient of viscosity, density and time of flow
for the liquid 1under examination and η2, d2 and t2 the corresponding values for the
reference liquid (liquid 2). Thus, by knowing η2, d2 , t2 and dl and tl the coefficient of
viscosity of first liquid, could be determined.
5.6. Summary
In this unit we have discussed about the characteristics of liquids. Surface tension
and viscosity of liquids were explained and the dependence of these characteristics on
intermolecular forces was highlighted. Also the theory behind experimental determination
of surface tension and viscosity were discussed.
5.7. Questions
1. What is surface tension? How the surface tension of a liquid varies with
Temperature?
2. What is surface energy? What is its unit?
3. Why the drop of a liquid is spherical?
84 NSOU GE-CH-11
4. Write down the theory for determination of surface tension using stalagmometer
by drop count method.
5. Why the viscosity of water is higher than alcohol?
6. How viscosity coefficient of a liquid does vary with temperature?
7. Write down the theory for determination of viscosity of a liquid using Ostwald
viscometer.
8. If 10 ml of water and ether forms 29 and 86 drops respectively in a viscometer
at 293 K, then calculate the surface tension of ether. Given that surface tension
of water is 7.2 x 10-2 N/m and density of ether is 0.7 gm/cc at 293 K.
(Ans. 1.6995 × 10-2 N/m)
9. In an Ostwald viscometer the flow time for water and toluene of equal volume
is 102.2 second and 68.9 second respectively. Find out the viscosity coefficient
of toluene if the viscosity coefficient of water is 0.01009 dyne s cm-2.
(Ans. 5.9 × 10-3 dyne s cm-2)
Unit 6 Solids
CONTENTS
6.0. Objectives
6.1. Introduction
6.2. Forms of solids
6.3. Symmetry elements
6.4. Unit cell
6.5. Crystal systems
6.6. Bravais lattice types
6.7. Law of Crystallography
6.7.1. Law of constancy of interfacial angles
6.7.2. Law of symmetry
6.7.3. Law of rational indices
6.8. Miller indices
6.9. X-Ray diffraction by crystals and Bragg's law
6.10. Structure of NaCl, KCl and CsCl
6.11. Defects in crystals
6.12. Summary
6.13. Questions
6.0. Objectives
After studying this Unit, we will be able to
describe general characteristics of solid state
distinguish between amorphous and crystalline solids
identify the symmetry elements in a solid
define lattice, basis, unit cell, primitive and nonprimitve cells
describe the seven crystal systems and the fourteen Bravais lattices
86 NSOU GE-CH-11
6.1. Introduction
We are mostly surrounded by solids and we use them more often than liquids and
gases. For different applications we need solids with widely different properties. These
properties depend upon the nature of constituent particles and the binding forces operating
between them. In gaseous state we have studied that if thermal energy is much greater
than the forces of attraction then we have matter in gaseous state. Molecules in gaseous
state move with very large speeds and because of very small attraction forces, the gas
molecules move practically independent of one another.
In the liquid state the forces of attraction are greater than the thermal energy. We
have also studied that molecules in liquid state too have kinetic energy, they cannot move
very far away because of the larger forces of attraction amongst them. Because of this
property, liquids have definite volume, but they do not have definite shape. Liquids also
resemble gases in their ability to flow. Gaseous and liquid states are, therefore, both
classified as fluids.
The solids are distinguished from a liquid or gas in terms of their rigidity which makes
them occupy definite volume and have a well defined shape. In solid state, the constituent
particles are in close contact and have strong forces of attraction between them.
A true solid possesses the following characteristics
(a) A sharp melting point
(b) A characteristic heat of fusion
(c) General incompressibility
(d) A definite three-dimensional arrangement
Hence solids are characterised by high density and low compressibility compared to
those of the gas phase. In solids, atoms, ions and molecules are held together by relatively
NSOU GE-CH-11 87
strong chemical forces- ionic bond, covalent bond, or by intermolecular Van der Waal’s
forces. They do not translate although they vibrate to some extent on their fixed positions.
This explains why solids are rigid and have definite shape.
Figure 6.1. Two dimensional structure of (a) quartz and (b) quartz glass.
(crystalline) and quartz glass (amorphous) are shown in Fig. 6.1 (a) and (b) respectively.
While the two structures are almost identical, yet in the case of amorphous quartz glass
there is no long range order.
Figure 6.2. Anisotropy in crystals is due to different arrangement of particles along different directions.
Crystalline solids on the other hand are anisotropic, because their physical properties
are different in different directions. For example the velocity of light through a crystal
varies with the direction in which it is measured. Thus, a ray of light enter such a crystal
may split up into two components each following different velocity. This phenomenon is
known as double refraction. This can be shown in fig 6.2 in which simple two-dimensional
arrangement of only two different kinds of atoms is depicted if the properties are measured
along the direction indicated by the slanting line CD, they will be different from those
measured in the direction indicated by the vertical line AB. The reason is that while in
the first case, each row is made up of alternate types of atoms, in the second case; each
row is made up of one type of atoms only. In amorphous solids, atoms or molecules are
arranged at random and in a disorderly manner and, therefore all directions are identical
and all properties are alike in all directions.
NSOU GE-CH-11 89
Plane of symmetry
When an imaginary plane can divide by a crystal into two parts such that one is the
exact mirror image of the other, the crystal is said to have a plane of symmetry.
Axis of symmetry
An axis of symmetry is a line about which the crystal is rotated such that it presents
the similar appearance more than once during complete rotation i.e. rotation through an
angle of 3600. Depending upon its nature, a crystal may have 2-fold, 3-fold, 4-fold or
6-fold axes of rotation.
Centre of Symmetry
It is a found at the centre of the crystal so that any line drawn through it will meet
the surface of the crystal at equal distance on either side. It may be pointed out that a
crystal may have number of planes or axis of symmetry but it has only one centre of
symmetry.
90 NSOU GE-CH-11
Unit cell is the smallest portion of a crystal lattice which, when repeated in different
directions, generates the entire lattice. A unit cell is characterised by:
(i) its dimensions along the three edges, a, b and c. These edges may or may not
be mutually perpendicular.
(ii) angles between the edges, α (between b and c) β (between a and c) and ã
(between a and b). Thus, a unit cell is characterised by six parameters, a, b, c,
α, β and γ.
Unit cells can be of following types;
(a) Simple or primitive unit cell (P): The simplest unit cell which has the lattice
points at the corners is called a simple or primitive unit cell. It is denoted by P.
NSOU GE-CH-11 91
(b) Non primitive or multiple unit cell: When unit cell contains more than one
lattice points, it is called non primitive or multiple unit cell. It is further divided into the
following three categories:
(i) Face centred unit cell (F): When a unit cell, besides the points present at the
corners of the unit cell, there is one point at the centre of each face, it is called
face centred arrangement or face centred unit cell. It is denoted by F.
(ii) Body centred unit cell (I): When in a unit cell, besides the points at the
corners of the cell, there is one point at the centre with in its body, it is called
body-centred arrangement or body-centred with cell. It is denoted by I.
(iii) Side centre or end face unit cell: When in a unit cell, besides the points at
the corners of the cell, the points are located at the centre of any two parallel
faces of the unit cell, it is called side-centred or end face unit cell. It is denoted
by c.
group or 32 crystal systems. Some of the systems, however, have been grouped together
so that we have only seven different categories, known as the seven basic crystal systems.
These are cubic, orthorhombic, tetragonal, monoclinic, triclinic, hexagonal and rhombohedral
or trigonal (figure 6.5). Crystal systems differ in length of the unit cell edges and the
angles between the unit cell edges.
Let OX, OY and OZ represent the three crystallographic axes and let ABC be a
unit plane (figure 6.8). The unit intercepts will then be a, b and c. According to the above
law, the intercepts of any face such as KLM, on the same three axes will be simple
whole number multiples of a, b and c, respectively. As can be seen from the figure, the
simple multiples in this case are 2, 2 and 3.
intercepts of that face on the various axes. The procedure for determination of Miller
indices for a plane is as follows:
(i) Write the intercepts as multiples of a, b, c say la, mb, nc
(ii) Take the reciprocals of l, m and n
(iii) Clear fraction to get whole numbers h,k,l.
(iv) Miller indices to the plane are (h,k,l).
Example: calculate the Miller indices of crystal planes which cut through the crystal
axes at (i) 2a, 3b, c and (ii) 6a, 3b, 3c
Solutions:
(i) a b c (ii) a b c
2 3 1 intercepts 6 3 3 intercepts
½ 1/3 1 reciprocals 1/6 1/3 1/3 reciprocals
3 2 6 clear fraction 1 2 2 clear fraction
Hence Miller indices are (326) Hence Miller indices are (122)
In a crystal, several planes can be imagined through the lattice points and they are
designated by Miller indices. If the Miller indices of a plane is (h, k, l), then all the planes
parallel to this plane will have the same Miller indices. The interplanar distance between
these parallel planes (dhkl) is given by the following equation:
2 2 2
1 h k l
d hkl 2 a b c
Where h, k, l are the Miller indices of the planes and a, b, c are the dimensions
of the cell.
For a cubic system, a = b = c so the distance between two planes (dhkl) in a cubic
system
a
d hkl
h k2 l2
2
96 NSOU GE-CH-11
The horizontal lines represent parallel planes in the crystal structure separated from
one another by a distance d. Suppose, a beam of x-ray (MA) incident at an angle è falls
on the crystal and reflected from the crystal (AO). Some of them will be reflected from
uppermost plane at the same angle, while the other will be absorbed and get reflected
from successive planes, as shown in figure 6.9. Similarly, ray NB is reflected from the
NSOU GE-CH-11 97
second plane to yield BP. If the rays AO and BP are to reinforce one another for
constructive interference, they must have the same phase; this condition is met if the extra
distance traversed by NBP ray (i.e. QBR) is equal to an integral number of wavelengths
of the x-ray. The extra distance is (QB+ BR), so that (QB + BR) = nλ, where n is an
integer. But from the geometry of the situation, QB = BR = d sin θ. Consequently, in
terms of the interplanar spacing d, the condition for constructive interference becomes
2d sin θ = nλ, n = 1, 2, 3, . . . ,
which is the fundamental law of x-ray crystallography, the Bragg condition, or Bragg’s
law. For a given wavelength of x-rays, the reflected beam will emerge only at those
angles for which the Bragg condition is satisfied.
The structure of KCl crystal is similar (isomorphous) to that of NaCl. But investigating
the KCl crystal, the maximum reflection of x-rays, corresponding to first order reflections
98 NSOU GE-CH-11
are observed to occur at the glancing angles of 5.380, 7.610 and 9.380 for (100), (110)
and (111) faces, respectively which corresponds to a simple cubic lattice structure. The
explanation for this apparent anomaly is very simple and can be explained on the basis
that the x-rays scattering power for an atom or ion is governed by the number of electrons
or atomic number.
The atomic numbers of potassium (K=19) and chlorine (Cl=17) are not very different
and the x-rays are unable to detect any difference between the two kinds of atoms. If
we imagine all the atoms to be identical, it is evident that face-centred arrangement has
become a simple cubic arrangement. This is the reason for the KCl spectrum corresponding
apparently to the simple cubic lattice. With KCl, the structure is face-centred, but the
face-centred characteristics are marked by the fact that the two types of atoms composing
the substance have nearly the same atomic numbers and atomic weights (K=39, Cl=35.5).
But in the case of sodium chloride the atomic numbers differ considerably (Na=11, Cl=17),
and so their scattering powers are different and hence the true structure as two
interpenetrating face-centred lattices become apparent.
Cesium chloride, CsCl, has a body centred cubic (BCC) structure. In its crystal
lattice, each Cs+ ion is surrounded by 8 Cl- ions and its coordination number is 8. The
value of distance between Cs+ ion and Cl- ion as determined by Bragg’s spectrometer is
3.510Å (figure 6.11).
Figure 6.12. Stoichiometric defects: (i) vacancy defect and (ii) interstitial defect.
Vacancy and interstitial defects as explained above can be shown by non-ionic solids.
Ionic solids must always maintain electrical neutrality. Rather than simple vacancy or
interstitial defects, they show these defects as Schottky and Frenkel defects.
100 NSOU GE-CH-11
(iv) Frenkel Defect: This defect is shown by ionic solids. The smaller ion (usually
cation) is dislocated from its normal site to an interstitial site (Fig. 6.13b) of the
crystal. So it creates a vacancy defect at its original site and an interstitial defect
at its new location. Frenkel defect is also called dislocation defect. It does not
change the density of the solid. Frenkel defect is shown by ionic substance in
which there is a large difference in the size of ions, for example, ZnS, AgCl,
AgBr and AgI due to small size of Zn2+ and Ag+ ions.
(b) Impurity Defects: If molten NaCl containing a little amount of SrCl2 is
crystallised, some of the sites of Na+ ions are occupied by Sr2+ (Fig 6.14). Each Sr2+
replaces two Na+ ions. It occupies the site of one ion and the other site remains vacant.
The cationic vacancies thus produced are equal in number to that of Sr2+ ions. Another
similar example is the solid solution of CdCl2 and AgCl.
NSOU GE-CH-11 101
Figure 6.14. Impurity defect in NaCl crystal by presence of Sr2+ ion as impurity.
(c) Non-Stoichiometric Defects: The defects discussed so far do not disturb the
stoichiometry of the crystalline substance. However, a large number of nonstoichiometric
inorganic solids are known which contain the constituent elements in non-stoichiometric
ratio due to defects in their crystal structures. These defects are of two types: (i) metal
excess defect and (ii) metal deficiency defect.
(i) Metal Excess Defect: This type of defects in crystals can be formed by
absence of anion in the lattice site or by the presence of extra cation in the
interstitial position.
Metal excess defect due to anionic vacancies: Alkali halides like NaCl and KCl
show this type of defect. When crystals of NaCl are heated in an atmosphere of sodium
vapour, the sodium atoms are deposited on the surface of the crystal. The Cl– ions diffuse
to the surface of the crystal and combine with Na atoms to give NaCl. This happens by
loss of electron by sodium atoms to form Na+ ions. The released electrons diffuse into
the crystal and occupy anionic sites (Fig 6.15). As a result the crystal now has an excess
of sodium. The anionic sites occupied by unpaired electrons are called F-centres. They
impart yellow colour to the crystals of NaCl. The colour results by excitation of these
electrons when they absorb energy from the visible light falling on the crystals. Similarly,
excess of lithium makes LiCl crystals pink and excess of potassium makes KCl crystals
violet (or lilac).
102 NSOU GE-CH-11
Metal excess defect due to the presence of extra cations at interstitial sites:
Zinc oxide is white in colour at room temperature. On heating it loses oxygen and turns
yellow.
1
Zn 2 O 2 2e
heating
ZnO
2
Now there is excess of zinc in the crystal and its formula becomes Zn1+xO. The
excess Zn2+ ions move to interstitial sites and the electrons to neighbouring interstitial sites
(Fig 6.16).
(ii) Metal Deficiency Defect: There are many solids which are difficult to prepare
in the stoichiometric composition and contain less amount of the metal as
compared to the stoichiometric proportion. In this case some of the metal ions
posses’ higher valences to maintain the electronutrality of the crystal. A typical
example of this type is FeO which is mostly found with a composition of Fe0.95O.
NSOU GE-CH-11 103
Figure 6.16. Metal excess defect due to the presence of extra cations at interstitial sites.
It may actually range from Fe0.93O to Fe0.96O. In crystals of FeO some Fe2+
cations are missing and the loss of positive charge is made up by the presence
of required number of Fe3+ ions.
6.12. Summary
In this unit, we have briefly described different forms of solid substances and found
that they can be distinguished as crystalline and amorphous. We have studied the terms-
lattice, basis and unit cell of crystalline solids. Symmetry elements present in the different
crystal lattice were discussed. Seven crystal systems and fourteen Bravais lattices were
explained. Different laws of crystallography were discussed and indexing of crystal planes
was explained. Diffraction method and its utility in crystal structure determination were
emphasised on the basis of Bragg’s law. Finally different types of crystal defects were
highlighted.
6.13. Questions
1. Define the term amorphous. Give a few examples of amorphous solids.
2. What is lattice?
3. Give definition of unit cell.
4. What are the symmetry elements present in a crystal?
5. What is the Miller indices of a crystal plane which makes intercepts 3a/2, 2b,
c.? [Ans. (4, 3, 6)]
6. What are the separations of the planes with Miller indices (2, 2, 1) for a cubic
lattice of side length 3Å? [Ans. 1 Å]
7. If an X-ray (wavelength = 1.539 Å) gets reflected at an angle of 22.50 from a
set of crystal planes, then find out the distance between the crystal planes.
[Ans. 2.01 Å]
8. Write down the differences in structure between the crystals of NaCl and KCl.
9. What are Frenkel and Schottky defects? Explain with examples.
NSOU GE-CH-11 105
Unit 7 Solutions
Contents
7.0. Objectives
7.1. Introduction
7.2. Units for the expression of concentration of solution
7.3. Ideal Solution
7.4. Raoult's law
7.5. Non-ideal solution and deviation from Roult's law
7.6. Vapour pressure-composition curves of ideal and non-ideal solutions
7.7. Temperature-composition curves of ideal and non-ideal solutions
7.8. Partial miscibility of liquids
7.9. Critical solution temperature
7.10. Principle of steam distillation
7.11. Nernst distribution law and its applications
7.12. Solvent extraction
7.13. Summary
7.14. Questions
7.0. Objectives
After studying this unit, we will be able to
describe the formation of different types of solutions;
express concentration of solution in different units;
state and explain Raoult’s law;
distinguish between ideal and non-ideal solutions;
explain deviations of real solutions from Raoult’s law;
106 NSOU GE-CH-11
7.1. Introduction
Solutions are homogeneous mixtures of two or more than two components. By
homogenous mixture we mean that its composition and properties are uniform throughout
the mixture. Generally, the component that is present in the largest quantity is known as
solvent.
Solvent determines the physical state in which solution exists. One or more components
present in the solution other than solvent are called solutes. Depending upon the nature
of solvent and solute, solutions can be of different types. On the basis of physical state
of solute and solvent it can be categorize in the following ways:
Type of Solution Solute Solvent Common Examples
Gaseous Solutions Gas Gas Mixture of oxygen and nitrogen gas (Air)
Liquid Gas Chloroform mixed with nitrogen gas
Solid Gas Camphor in nitrogen gas
Liquid Solutions Gas Liquid Oxygen dissolved in water
Liquid Liquid Ethanol dissolved in water
Solid Liquid Glucose dissolved in water
Solid Solutions Gas Solid Solution of hydrogen in palladium
Liquid Solid Amalgam of mercury with sodium
Solid Solid Copper dissolved in gold
For example, 1.00 g.equiv L–1 (or 1.00 N) solution of H2SO4 means that 1 gram
equivalent (49 g) of H2SO4 is dissolved in 1 litre of solution.
Molarity: Molarity (M) is defined as number of moles of solute dissolved in one
litre (or one cubic decimetre) of solution.
Moles of solute
Molarity
Volume of solution in litre
For example, 0.25 mol L–1 (or 0.25 M) solution of NaOH means that 0.25 mol of
NaOH has been dissolved in one litre (or one cubic decimetre).
Molality: Molality (m) is defined as the number of moles of the solute per kilogram
(kg) of the solvent and is expressed as:
Moles of solute
Molality
Volume of solvent in kg
For example, 1.00 mol kg–1 (or 1.00 m) solution of KCl means that 1 mol (74.5
g) of KCl is dissolved in 1 kg of water.
Mole fraction: Commonly used symbol for mole fraction is x and subscript used
on the right hand side of x denotes the component. It is defined as:
For example, in a binary mixture, if the number of moles of A and B are nA and nB
respectively, the mole fraction of A will be
nA
xA
nA nB
ni n
xi i
n1 n 2 n 3 ... ... ... n i n i
It can be shown that in a given solution sum of all the mole fractions is unity, i.e.
x1 + x2 + x3 + .................. + xi = 1
108 NSOU GE-CH-11
Mole fraction unit is very useful in relating some physical properties of solutions, say
vapour pressure with the concentration of the solution and quite useful in describing the
calculations involving gas mixtures.
Parts per million (ppm): When a solute is present in trace quantities, it is convenient
to express concentration in parts per million (ppm) and it is defined as:
As in the case of percentage, concentration in parts per million can also be expressed
as mass to mass, volume to volume and mass to volume. A litre of sea water (which
weighs 1030 g) contains about 6 × 10–3 g of dissolved oxygen (O2). Such a small
concentration is also expressed as 5.8 g per 106 g of sea water or 5.8 ppm. The
concentration of pollutants in water or atmosphere is often expressed in terms of μg
mL–1 or ppm.
pressures of the two components A and B respectively. These partial pressures are related
to the mole fractions xA and xB of the two components A and B respectively.
The French chemist, Francois Marte Raoult (1886) gave the quantitative relationship
between them. The relationship is known as the Raoult’s law which states that for a
solution of volatile liquids, the partial vapour pressure of each component of the solution
is directly proportional to its mole fraction present in solution.
Thus, for component A
pA xA
so, pA = pA0.xA
where pA0 is the vapour pressure of pure component A at the same temperature.
Similarly, for component B
pB = pB0.xB
where pB0 represents the vapour pressure of the pure component B.
According to Dalton’s law of partial pressures, the total pressure (ptotal) over the
solution phase in the container will be the sum of the partial pressures of the components
of the solution and is given as:
ptotal = pA + pB
Figure 7.1. The plot of vapour pressure and mole fraction of an ideal solution at constant temperature.
110 NSOU GE-CH-11
A plot of pA or pB versus the mole fractions xA and xB for a solution gives a linear
plot as shown in Fig. 7.1. These lines (I and II) pass through the points for which xA
and xB are equal to unity. Similarly the plot (line III) of ptotal versus xB is also linear (Fig.
7.1). The minimum value of ptotal is pA0 and the maximum value is pB0, assuming that
component B is less volatile than component A, i.e., pA0 > pB0.
Figure 7.2. The vapour pressures of two component system as a function of composition for a
solution that shows positive deviation from Raoult’s law.
Figure 7.3. The vapour pressures of two component system as a function of composition for a
solution that shows negative deviation from Raoult’s law.
112 NSOU GE-CH-11
chloroform and benzene, acetic acid and pyridine shows negative deviation. A mixture of
chloroform and acetone shows negative deviation from Raoult’s law, because chloroform
molecule forms hydrogen bond with acetone molecule which decreases the escaping
tendency of molecules for each component. Thus vapour pressure decreases resulting in
negative deviation from Raoult’s law. Figure 7.3 represents the curve showing negative
deviation.
Figure 7.4. Liquid and vapour composition curves for an ideal solution.
NSOU GE-CH-II 113
phase curve (curve II) lies below the liquid phase curve (curve I). The line ab is called
a tie line and it gives us the composition of the solution in the liquid and vapour phases
in equilibrium at a particular total vapour pressure.
The curves I and I1 of Fig. 7.4 are obtained in the case of solutions obeying Raoult’s
law. In the case of solutions showing positive deviation from Raoult’s law, the liquid and
vapour composition curves are of the type shown in Fig. 7.5. Note that there is a maximum
point, M, where both the liquid and vapour phases have the same composition.
Figure 7.5. Liquid and vapour composition curves for a liquid mixture showing positive deviation.
In the case of a solution showing negative deviation from Raoult’s law, the liquid and
vapour composition curves are of the type shown in Fig. 7.6. Note that the curves meet
at the minimum point M where both the liquid and vapour phases have the same
composition.
Figure 7.6. Liquid and vapour composition curves for a liquid mixture showing negative deviation.
pressure, it will start boiling when the total vapour pressure becomes equal to the
atmospheric pressure. If p represents the atmospheric pressure, then the condition for
boiling is
p = pA + pB
where pA and pB are the partial pressures of the two components A and B. Since different
compositions of a solution have different vapour pressures, the various solutions will not
reach a total vapour pressure equal to the atmospheric pressure at the same temperature.
Hence, the solutions of different compositions will boil at different temperatures. In general,
solutions of low vapour pressure will boil at temperatures higher than those of solutions
for which the vapour pressures are high. It is because solutions of high vapour pressure
can have the total pressure equal to the atmospheric pressure at relatively lower
temperatures as compared to solutions for which vapour pressures are low. Hence it is
possible to draw temperature-composition diagrams which will correspond to the three
general types of vapour pressure composition diagrams.
For an ideal solution of two miscible liquid A and B the boiling point-composition
curve is shown in figure 7.7. Let the vapour pressure of pure A be higher than that of
pure B. Consequently at constant pressure, the boiling point of A (TA) will be lower than
NSOU GE-CH-11 115
that of B (TB). Now as compared to the liquid mixture, the vapour is richer in the more
volatile component. So as in the present case A is more volatile than B, hence, the
vapour composition at any temperature must lie closer to A than the corresponding liquid
composition. In other words, in the composition against temperature plot, the vapour
composition curve must lie above the liquid composition curve. From this plot it can be
understood that by boiling the solution from a mixture of two miscible liquids pure liquids
can be separated. This process is known as fractional distillation.
boiling unchanged). They resemble pure compounds in their boiling behaviour. However,
changes in pressure produce changes in the composition as well as the boiling point of
the azeotropes. The azeotropes are not chemical compounds but are rather mixtures
resulting from the interplay of intermolecular forces in solution.
Figure 7.8. (I) Boiling point-composition diagram of a liquid mixture showing positive deviation; (II)
Boiling point-composition diagram of a liquid mixture showing negative deviation.
There are some liquid pairs (e.g., triethylamine-water) for which mutual solubilities
decrease with rise in temperature. As the temperature is decreased, the mutual solubilities
118 NSOU GE-CH-11
increase and below the consolute temperature, the two liquids become miscible in all
proportions. Such systems possess lower consolute temperature. The variation of mutual
solubility of triethylamine and water with temperature is shown in Fig. 7.10.
There are a few liquids pairs, e.g., nicotine and water which show both the upper
and lower consolute temperatures. These liquid pairs are completely miscible above a
certain temperature (upper consolute temperature) and also below a certain temperature
(lower consolute temperature). The variation of mutual solubilities of nicotine and water
with temperature is shown in Fig. 7.11.
The presence of an impurity, dissolved in one or both of the phases, changes the
CST values as well as the liquid composition at CST. Substance soluble in only one of
the liquids raises the upper CST and lowers the lower CST.
For example, chlorobenzene has a boiling point of 405 K. A mixture of water and
chlorobenzene distils at a constant temperature of 363.3 K, when the external pressure
in 9.8 × 104 Pa, by passing steam through it. Let us explain the procedure for purifying
an organic liquid using steam distillation. The apparatus used for steam distillation is as
shown in Figure 7.12.
The impure organic compound is taken in a round-bottomed flask (A) and a small
quantity of water is added. The flask must be kept in a slanting position to prevent the
impure liquid from splashing up into the condenser. The flask A is then heated gently.
Now, steam from container B is bubbled through the contents of the flask A. Vapours
of the organic compound mix with steam and escape into the water condenser C. The
condensate thus obtained in the flask F is a mixture of water and the organic compound.
This mixture can then be separated by means of a separating funnel.
Concentration of X in A
a constant
Concentration of X in B
Nernst (1891) studied the distribution of several solutes between different appropriate
pairs of solvents and gave a generalization which governs the distribution of a solute
between two non-miscible solvents. This is called Nernst’s Distribution law or Nernst’s
Partition law or simply Distribution law or Partition law.
It states that,
“if a solute X distributes itself between two immiscible solvents A and B at constant
temperature and X is in the same molecular condition in both solvents.”
Concentration of X in A
KD
Concentration of X in B
NSOU GE-CH-11 121
C1
KD
C2
C1 S1
KD
C 2 S2
where S1 and S2 are the solubilities of the solute in the two solvents.
From this expression we get the amount of the organic compound that remained
KV
unextracted after first extraction (w 1 ) w
v kV
Similarly, after the second extraction, the mass of the organic compound that remains
2
KV
unextracted is, w 2 w
v kV
In general, the mass of the organic compound that remains unextracted after n
n
KV
extractions is given by, w n w
v kV
7.13. Summary
In this unit we have learnt about the solutions and different type of solutions. Then
we have learnt that completely miscible liquid pairs may be ideal or non-ideal. Ideal
solutions obey Raoult’s law. Non-ideal solutions either show positive or negative deviation
from
Raoult’s law. We have learnt to draw the vapour pressure-composition and
temperature-composition curves for ideal and non-ideal solutions. We have discussed
about the partially miscibility of liquids and critical solution temperature. Then we have
seen that a pair of immiscible liquids boil at a temperature lower than the boiling points
of any of the liquids. This fact is made use of in steam distillation. We have learned about
the principle of steam distillation. Finally we have learned about the Nernst distribution
law for a system when a solute is added to a pair of immiscible liquids and its applications
in solvent extraction.
7.14. Questions
1. Give an example of a solid solution in which the solute is a gas.
2. Define the following terms: (i) Mole fraction (ii) Molality (iii) Molarity.
3. Write down the differences between ideal and non-ideal solution.
4. What is meant by positive and negative deviations from Raoult’s law?
124 NSOU GE-CH-11
5. Draw the phase diagram of phenol-water system and explain the plot.
6. Write down the principle of steam distillation.
7. Describe Nernst distribution law and state its limitations.
8. If the solubility of I2 in water is 0.345 g/L at 30 0C then what is the solubility
of I2 in CCl4 at 30 0C? Given that the distribution coefficient for I2 in CCl4 and
water is 86. (Ans. 29.67 g/L)
Unit 8 Chemical Kinetics
Contents
8.0. Objectives
8.1. Introduction
8.2. The concept of reaction rates
8.3. Factors affecting reaction rates
8.4. Order and molecularity of a reaction
8.5. Derivation of integrated rate equations
8.5.1. Zero order reaction
8.5.2. First order reaction
8.5.3. Second order reaction
8.6. Half-life of a reaction
8.7. General methods for determination of order of a reaction
8.8. Concept of activation energy and its calculation from Arrhenius equation
8.9. Theories of Reaction Rates
8.9.1. Collision theory
8.9.2. Activated complex theory of bimolecular reactions
8.10. Summary
8.11. Questions
8.0. Objectives
After studying this unit, we should be able to:
define rate law, rate constant of reaction,
discuss the dependence of rate of reactions on pressure, temperature and
catalyst
126 NSOU GE-CH-11
8.1. Introduction
The branch of physical chemistry which deals with the speed or rate at which a
reaction occurs is called chemical kinetics. Chemical Kinetics is the study of rate of a
reaction under different conditions like different concentrations, pressures, temperatures,
catalyst, pH, dielectric constant of the medium, free radical scavengers, neutral salts etc
and suggesting a suitable mechanism for the reaction.
Chemical kinetics constitutes an important topic in physical chemistry. It concerns
itself with measurement of rates of reactions proceedings under given condition of
temperature, pressure and concentration.
The study of chemical kinetics has been highly useful in determining the factors which
influence rate of reaction as well as in understanding mechanism of a number of chemical
reactions. The experimental data have led to the development of the modern theories of
chemical reactivity of molecules.
In this Unit, we shall be dealing with rate of reaction and the factors affecting these.
Some elementary ideas about the collision theory and activated complex theory of reaction
rates are also given. However, in order to understand all these, let us first learn about
the reaction rate.
dCA
r (1)
dt
where –dcA is very small decrease in concentration of A in a very small time interval dt.
Now the concentration of product B increases with time. Hence rate of reaction can also
be expressed in terms of increase in concentration of the product B as well.
Thus
dCB
r (2)
dt
where dcB is very small increase in the concentration of product B in a very small time
interval of time dt.
Now it is from (1) and (2)
dCA dCB
r (3)
dt dt
So for a reaction
A + B → M + N
the rate can be expressed
dCA dC dCM dC N
r B (4)
dt dt dt dt
1 dC A 1 dC B 1 dCC 1 dCD
r kCaA CB
b
(5)
a dt b dt c dt d dt
dCA
r kC A (6)
dt
where k is the rate constant or the velocity constant of the reaction at the given temperature.
This form of rate equation is known as the differential rate equation. If concentration of
A is unity, i.e., CA = 1, then, evidently, r = k. For a general reaction of the type:
aA + bB + cC → Products
The rate of the reaction is given by the rate law expression
r kCaA CB
b c
CC (7)
of the reactant, the number of collisions will increase and the rate of reaction will increase
and on decreasing the concentration the rate will decrease.
(b) Effect of nature of reactants: Reactions between polar or ionic molecules
occur almost instantaneously. Those reactions in which the bonds are arranged or electrons
are transferred takes a comparatively longer time than the reaction between ionic molecules.
We can cite the examples of neutralisation reactions or double displacement reactions
which are very fast while the oxidation reduction reactions are slower.
(c) Effect of catalyst: A catalyst can increase or decrease the rate of a chemical
reaction. For example the combination of hydrogen and oxygen to form water is slow
at ordinary temperature, while it proceeds rapidly in presence of platinum.
(d) Effect of surface area of reactant: Surface area of reactants is of importance
only for heterogeneous reactions. With the decrease in the particle size, surface area of
the reactant for the same mass increases. The smaller particle thus reacts more rapidly
than the larger particles. For example, burning of coal dust in air takes place more rapidly
than large lump of coal.
(e) Effect of temperature: With the exception of few reactions, it has been found
that generally an increase of temperature increases the rate of reaction. The ratio of rate
constants of a reaction at two temperatures differing by 100C is known as temperature
coefficient of the reaction. The temperatures usually selected for this purpose are 250C
and 350C. Thus
Rate at 350 C k 35
Temperature coefficient 0
Rate at 25 C k 25
The value of temperature coefficient for most of the reactants is close to 2 and in
some cases it approaches to 3.
a, b and c indicate how sensitive the rate is to the change in concentration of A and B.
Sum of these exponents, i.e., a + b + c gives the overall order of a reaction
130 NSOU GE-CH-11
whereas a, b and c represent the order with respect to the reactants A, B and C
respectively.
Hence, the sum of powers of the concentration of the reactants in the rate law
expression is called the order of that chemical reaction.
Order of a reaction can be 0, 1, 2, 3 and even a fraction. A zero order reaction
means that the rate of reaction is independent of the concentration of reactants.
A balanced chemical equation never gives us a true picture of how a reaction
takes place since rarely a reaction gets completed in one step. The reactions taking
place in one step are called elementary reactions. When a sequence of elementary
reactions (called mechanism) gives us the products, the reactions are called complex
reactions.
The number of reacting species (atoms, ions or molecules) taking part in an elementary
reaction, which must collide simultaneously in order to bring about a chemical reaction
is called molecularity of a reaction. The reaction can be unimolecular when one reacting
species is involved as for example, decomposition of ammonium nitrite.
NH4NO2 → N2 + 2H2O
Bimolecular reactions involve simultaneous collision between two species, for example,
dissociation of hydrogen iodide.
2HI → H2 + I2
Trimolecular or termolecular reactions involve simultaneous collision between three
reacting species, for example,
2NO + O2 → 2NO2
The probability that more than three molecules can collide and react simultaneously
is very small. Hence, reactions with the molecularity three are very rare and slow to
proceed.
It is, therefore, evident that complex reactions involving more than three molecules
in the stoichiometric equation must take place in more than one step.
The molecularity of any process can only be small position integers, while order of
reaction can have zero as well as fractional values.
NSOU GE-CH-11 131
dCA
Rate r kC0A (8)
dt
As any quantity raised to power zero is unity
dC A
kC0A k 1 k
dt
dcA = – k dt
Let the initial concentration at initial time of the reaction, t= 0 be c0. Subsequently
at any other time t, the concentration will be ‘c’. On integration we obtain
c t
c 0
dc A kdt
0
c – c0 = – kt
c = – kt + c0 (9)
This is the integrated rate equation for zero order reaction. If we plot concentration
(c) against t, we get a straight line (Fig. 8.1) with slope = –k and intercept equal
to c0.
132 NSOU GE-CH-11
c0 c
k (10)
t
Photochemical reaction between H2 and Cl2 over water (saturated with HCl) surface
is an example of zero order reaction.
Figure 8.1. Variation in the concentration vs time plot for a zero order reaction.
dc A dcP
Rate r kCA
dt dt
Bringing concentration terms in one side and the time on the other side, we can write
the above equation as
dc A
k1dt
cA
Now if the initial concentration at initial time t=0 be c-0, and at any other time t, the
concentration is c then, by integration we get
NSOU GE-CH-11 133
c dcA t
c 0
cA
k1dt
0
c
ln k1t
c0
c c0 e k1t (11)
1 c
k1 ln c0 (12)
t
2.303 c
or, k1 log 0 (13)
t c
This is the expression for the first order rate constant k1.
Sometimes equation 12 is expressed in another format. If initial concentration of the
reactant is a and x moles of it react in time t; then the concentration of the reactant left
behind at time t will be a-x. In such a case the equation 12 can be written as
1
k1 ln a a x (14)
t
Figure 8.2. The plot of concentration versus time (a) and ln(concentration) versus time (b) for a first-
order reaction.
134 NSOU GE-CH-11
From equation 14 it can be seen that the concentration of reactant in a first order
reaction decreases exponentially with time. A plot of ln(concentration) versus time will
give a straight line with slope = -k1. This is shown in figure 8.2.
dc A dcP
Rate r k 2C 2A
dt dt
Where k2 is the second order rate constant. Now if a is the initial concentration of
A, x is the concentration of the product formed after time t and (a-x) is the concentration
of A remaining at time t, then,
dx
Rate r k 2 (a x) 2 (15)
dt
By separating the variables and integrating, we get
x dx t
0 a x 2 0 k 2dt
1 1
k2t
(a x) a
1 1 1 1 x
k2
t (a x) a t a(a x)
(16)
This is the integrated expression for the rate constant of a second order reaction in
which two molecules of the same reactant are involved in the reaction. The most common
NSOU GE-CH-11 135
example of the above type of the second order reaction is the gaseous decomposition
of hydrogen iodide.
2HI (g) → H2 (g) + I2 (g)
The rate expression for this reaction is
d[HI]
Rate r k 2 [HI]2
dt
Case II: When the reactants are different
Consider a second order reaction
A + B → P
where the initial concentration of A is a mol dm–3 and that of B is b mol dm–3. After time
t, x mol dm–3 of A and x mol dm–3 of B react to form x mol dm–3 of the product. Thus
the reactant concentration at time t are (a–x) and (b–x), respectively. The differential rate
expression for the second order reaction is,
dc A dc dc
Rate r B P k 2CA CB
dt dt dt
This can be written as
dx
r k 2 (a x) (b x) (17)
dt
where k2 is the second order rate constant. Separating the variables, we get
dx
k 2dt (18)
(a x)(b x)
1 1 1 1
(19)
(a x)(b x) (a b) b x a x
x dx 1 x 1 x 1 t
0 (a x)(b x) (a b) 0 b x 0 a x k 2 0 dt (20)
136 NSOU GE-CH-11
1 ax 1 a
ln ln k 2 t (21)
(a b) b x (a b) b
1 ax a 1 b(b x)
k2 ln ln ln
b (a b)t a(b x)
(22)
(a b)t b x
This is the integrated expression for the rate constant of a second order reaction.
Here we have assumed that a > b. If we had assumed b > a then the expression becomes
as follows
1 a(b x)
k2 ln
(a b)t b(a x)
c0 c
k
t
c0
At t t1 2 , c
2
c0
c0
Then, t1 2 c0 (23)
k 2k
2
So for a zero-order reaction, the half-life period depends on the initial concentration
of the reactant and the rate constant.
NSOU GE-CH-11 137
2.303 c
k1 log 0
t c
c0
At t t1 2 , c
2
2.303 c
t1 log 0
k1 c0
2
2
2.303
t1 log 2
k1
2
0.693
t1 (24)
k1
2
It can be seen from equation 24, that for a first order reaction, half-life period is
constant, i.e., it is independent of initial concentration of the reacting species. The half-
life of a first order reaction can be readily calculated from the rate constant and vice
versa.
Half life for a second order reaction:
For the second order reaction, we can write from equation 16
1 x
k2
t a(a x)
At x = a/2, t = t1/2
1 a 2 1 a 2 1
k2
t1 2 a(a a 2) t1 2 a a t1 2
a
2
138 NSOU GE-CH-11
1
t1 (25)
k 2a
2
So from equation 25, we can find that t1/2 of a second order reaction is inversely
proportional to the initial concentration of the reactant and thus it does not remain constant
as the reaction proceeds.
2n 1 1
t1 (26)
2
k n (n 1) a 0n 1
Where a0 is the initial concentration of the reactant A and kn is the nth-order rate
constant. From the equation 26 we can see that
1
t1 2 (27)
a 0n 1
Rate in Experiment I1 = r2 = ka m n
2 b1
r1
From the ratio , we can calculate order m, since a1 and a2 are known
r2
m
Rate in exp eriment I r1 ka1m b1n a1
Rate in exp eriment II r2 ka m n
2 b1 a2
r1 a
log m log 1 (28)
r2 a2
Similarly, the rate for one more experiment in which the initial concentration of A is a2
and the initial concentration of B is b2.
m n
So, rate in Experiment III = r3 = ka 2 b 2
n
Rate in exp eriment II r2 ka m2 b1
n
b
m n 1
Rate in exp eriment III r3 ka 2 b 2 b2
r2 b
log n log 1 (29)
r3 b2
Since r2, r3, b1 and b2 are known, n can be calculated. The overall reaction order =
m + n.
140 NSOU GE-CH-11
t1 2 1 a 2 n1
t1 2 2 a1
ln
t1 2 1 (n 1) ln a 2
or,
t1 2 2 a1
ln
t1 2 1
n 1
t1 2 2
or, (30)
a
ln 2
a1
NSOU GE-CH-11 141
Figure 8.3. Diagram showing plot of potential energy vs. reaction coordinate.
All the molecules in the reacting species do not have the same kinetic energy. For
the formation of product the reactant molecules must possess sufficient energy. Increasing
the temperature of the substance increases the fraction of molecules, which collide with
energies greater than Ea. Thus, with increase of the reaction temperature the rate of the
reaction increases. In the Arrhenius equation (31) the factor e-Ea /RT corresponds to the
fraction of molecules that have kinetic energy greater than Ea.
Taking natural logarithm of both sides of equation 31, we get
Ea
ln k ln A (32)
RT
From equation 32, it is evident that a plot of lnk versus 1/T (figure 8.4) gives a straight
line with slope = – Ea/R and intercept = lnA.
By differentiating equation 32 with respect to temperature, we can write
dlnk E
a2 (33)
dT RT
Integrating equation 33 between temperature T1 and T2 when the corresponding rate
constants are k1 and k2, respectively and assuming that Ea is constant over this temperature
range, we obtain
k1 Ea T2 T1
ln (34)
k 2 R T1T2
NSOU GE-CH-11 143
This is the integrated Arrhenius equation. From this equation 34 by knowing the rate
constants at two different temperatures, the energy of activation Ea can be readily determined.
For a reaction involving two different gases A and B, the rate of bimolecular collisions
between unlike molecules is given by
1/2
2 8kT
ZAB 2n A n Bav (36)
where nA and nB are numbers of A and B molecules, respectively, σav is the average
collision diameter defined as (σA+ σB)/2 and μ is the reduced mass defined as μ =
(mAmB)/(mA + mB). The collision number ZAB is given, in terms of molar masses MA and
MB of the two gases, by the expression
12
2 M A M B 8kT
ZAB n A n Bav (37)
MAMB
Let us calculate ZAB for the reaction between H2 and I2 at 700 K and 1 atm pressure,
the quantities of the two gases being 1 mole each. Accordingly, n H2 n I2 1019
molecules cm–3, H2 = 2.2 Å, I2 = 4.6 Å so that σav = 3.4 Å. Hence, according to
Eq. 37,
12
2 254 8 3.14 8.314 107 700
ZAB 10 3.4 10
19 2 8
2 254
The detailed analysis of the dynamics of bimolecular collisions leads to the result that
the number of collisions S-1cm-3 between molecules A and B, when the relative kinetic
energy E along the line of centres is greater than the threshold energy is given by
Assuming that ZAB gives the rate of relative collisions between A and B, we can
write
dn A dt ZAB
12
M A M B 8kT
So, dn A dt 2
n A n Bav e E RT
molecules cm 1s1 (39)
MAMB
Now if the concentrations of the reactants are expressed in mol dm-3, then
103 n A 103 n B
[A] and [A] (40)
nA NA
103 dn A 106
k2 nAnB (41)
N A dt N A 2
NA dn A
Hence, k 2 (42)
103 n A n B dt
1/2
N 2 M M B 8RT
k 2 A 3av A eE RT (43)
10 MAMB
146 NSOU GE-CH-11
Comparing equation 43 with the Arrhenius equation (Eq 31) k = A e –Ea /RT, we find
that the Arrhenius pre-exponential factor is given by
1/2
N 2 M M B 8RT
A A 3av A (44)
10 MAMB
The activation energy Ea, in the Arrhenius equation is thus identified with the relative
kinetic energy E along the line of centres of the two colliding molecules which is required
to cause a reaction between them.
Now let us calculate A and k2 for the H2-I2 reaction at 700 K, considered in the
beginning of this section. Ea has been found to be 167.4 kJ mol-1. Substituting the various
values in equation 44, A comes out to be = 6.0 x 1011 dm3 mol-1 s-1. Hence, from
Arrhenius equation (Eq 31),
k2 = 6.0 × 1011dm3 mol–1 s–1. exp(-167400 J mol–1/ 8.314 J K–1 mol–1 . 700K)
= 0.22 dm3 mol–1 s-1
which is comparable with the experimental value of 0.064 dm3 mol–1 s–1, considering the
uncertainty in the values of the activation energy and average collision diameter.
The collision theory is applicable to simple gaseous reactions. For reactions between
complicated molecules, the observed rate is found to be much smaller than the theoretically
predicted rate. The discrepancy arises due to the facts that the colliding reactant molecules
are treated as hard spheres without any internal energy. Again the spherical model ignores
the dependence of the effectiveness of the collision on the relative orientation of the
colliding molecules. Also the activation energy was treated as though it were related entirely
on translational motion. For this reasons the collision theory is applicable to reactions
between very simple gaseous molecules.
The collision theory can be generalized by introducing the so-called steric factor, p,
into the equation for the bimolecular rate constant in order to take account of the
orientational requirement. Accordingly the equation becomes
k2 = p A e –Ea /RT
(45)
reaction process. It is known as the absolute reaction rate theory or the transition state
theory or more commonly as activated complex theory. According to the activated complex
theory, the bimolecular reaction between two molecules A2 and B2 progresses through
the formation of the so-called activated complex which then decomposes to yield the
product AB. It must be remembered that the activated complex is not merely an
intermediate in the process of breaking or forming of chemical bonds. It is unstable because
it is situated at the maximum of the potential energy barrier separating the products from
the reactants (Figure 8.5). The difference between the energy of the activated complex
and the energy of the reactants is the activation energy, Ea.
Figure 8.5. Energy versus reaction coordinate in activated complex theory in the case of an exothermic
and endothermic reaction.
K# k
(AB) #
A B 2
Products (46)
where (AB)# is the activated complex and K# is the equilibrium constant between the
reactants and activated complex. From classical mechanics, the energy of vibration of the
activated complex (AB)# is given by RT/NA (or kBT where kB is the Boltzmann constant)
whereas from quantum mechanics, it is given by hv so that hv = RT/NA or v = RT/NAh
= kBT/h. The vibrational frequency v is the rate at which the activated complex
molecules move across the energy barrier. Thus, the rate constant k2 can be identified
with v.
148 NSOU GE-CH-11
dA k T
k 2 (AB) # B (AB) # (47)
dt h
where the factor κ, called the transmission coefficient, is a measure of the probability that
a molecule, once it passes over the barrier, will keep on going ahead and not return.
The value of κ is taken to be unity; it is thus omitted from the rate expression. The
concentration of the activated complex , can be obtained by writing the equilibrium
expression
(AB)#
K #
[A] [B]
dA k BT #
K [A] [B] (49)
dt h
Thus the rate constant k2 may be expressed as
k BT #
k2 K (50)
h
The equilibrium constant K# can be expressed in terms of (ΔGo)#, called the standard
Gibbs free energy of activation.
Since for the activated complex, we can write
G H
# # # #
0
RT ln K # and G 0 0
T S0 (51)
Here, (ΔHo)# is the standard enthalpy of activation and (ΔSo)# is the standard entropy
of activation
k BT ( S0 )# 0 #
k2 e R
.e ( H ) RT
h
This is the well known Eyring equation. The application of the activated complex
theory is to reactions in solution is quite complicated because of the participation of the
solvent molecules in the activated complex. Fortunately, the Eyring equation is applicable
for reactions in the solutions phase also.
Taking logs of both sides of equation 53 and differentiating with respect to T we
get
2
d ln k 2 dT ( H 0 ) # RT 1 T ( H 0 ) # RT RT 2 (54)
Also, from the Arrhenius equation we get,
d ln k 2 dT E a RT 2 (55)
E a (H 0 ) # RT
k BT ( S0 )# 0 #
pAe Ea RT
e R
.e ( H ) RT (57)
h
k BT ( S0 ) # R
pA e (58)
h
For a first order gaseous reaction, p = 1 and A = 1010 s–1.
k BT
At room temperature the value of 1013 s 1 .
h
150 NSOU GE-CH-11
Hence,
0 # 1010 s1
e( S ) R
13 1
103
10 s
8.10. Summary
In this unit, we started with the definitions of the terms such as, rate of reaction, rate
law, order and molecularity of reaction. We derived the integrated forms of rate expressions
for first order, second order and zeroth order reactions. We have discussed about the
half life of the different order reactions. We explained the methods of determination of
order of reaction. Finally the concepts of activation energy, Arrhenius equation, collision
theory and the activated complex theory were discussed.
8.11. Questions
1. Rate of reaction depends upon which factors?
2. What are the differences between order and molecularity of a reaction?
3. What do you mean by a zero order reaction? Give example.
4. Show that half life of a first order reaction does not depend on the initial
concentration of the reactants.
5. Describe any method for determination of order of reaction.
6. How the rate of a reaction depends upon temperature?
7. What is activation energy of a reaction?
NSOU GE-CH-11 151
Notes