0% found this document useful (0 votes)
12 views416 pages

Physical Chemistry For MSC

Uploaded by

prachichhillar4
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
12 views416 pages

Physical Chemistry For MSC

Uploaded by

prachichhillar4
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 416

CHAPTER 1

Quantum Mechanics – I
 Postulates of Quantum Mechanics
In modern quantum theory, the postulates of quantum mechanics are simply the step-to-step
procedure to solve a simple quantum mechanical problem. In other words, it is like the manual that must be
followed to retrieve the information about various states of any quantum mechanical system. We will first
learn about the nature and the significance of these postulates, and then we will apply them to some real
problems like the particle in a one-dimensional box or the harmonic oscillator.
 The First Postulate
All time-independent states of any quantum mechanical system can be described mathematically as
long as the function used is single-valued, continuous and finite.
Explanation: The systems around us can be broadly classified into two categories; the first is classical and the
other one as quantum mechanical. The classical systems simply refer to the systems which are governed by
the classical or the Newtonian mechanics. Now because all the macroscopic objects follow Newton’s laws of
motion, they fall in the category of classical systems; for example, a rotating gym dumbbell, the vibrating
spring of steel, or an athlete running in the playground. Every classical system can possess many states which
belong to a continuous domain, and each state can be described mathematically.

However, if the rotating gym dumbbell is replaced by the rotating diatomic molecule, the system
would not remain classical anymore and would start violating classical laws. The states of such microscopic
systems (here it just means the extremely small) belong to a discontinuous domain and can also be described
mathematically. These mathematical descriptions are labeled as ψ1, ψ2, ψ3 ….. ψn and generally called as the
“wave functions”. The term “wave function” is used because as we go from the macroscopic to the microscopic
world i.e. from classical to the quantum mechanical world, things start behaving like waves rather particle. All
of the states are wave-like; and because every wave we see around us is continuous, single-valued and finite;

Copyright © Mandeep Dalal


12 A Textbook of Physical Chemistry – Volume I

only continuous, single-valued and finite expressions can represent those states. For instance, when you drop
a stone in a standstill pond, the waves are generated which travel from the center to the boundary of the pond;
and you don’t see any discontinuity in it.

Hence, if a function is not single-valued, continuous and finite; it will not be able to represent any
wave-like behavior at all. That is why every function that correlates a quantum mechanical state must be single-
valued, continuous and finite; and this function describes the corresponding state completely.
 The Second Postulate
For every physical property like linear momentum or the kinetic energy, a particular operator exists
in quantum mechanics, the nature of which depends upon the classical expression of the same property.
Explanation: In classical mechanics, there are simply straight forward formulas for all physical properties;
like linear momentum can simply be calculated by multiplying the mass with velocity. However, in case of
quantum mechanical systems, the value of a certain physical property for a particular state cannot be calculated
simply by using its classical formula but from an operator. It does sound silly but the classical formulas which
are so well-tested on the scale of time fail in quantum world. For instance, you can use the mv2/2 to calculate
the kinetic energy of a moving particle in classical world by just putting its mass and velocity; but if the mass
of the moving particle is extremely less, you will not get any rational results.
It is also worthy to note it again that though the classical formulas fail to give the value of physical
property, they are still important as they form the basis of the derivations for corresponding quantum
mechanical operators. For instance, the operator for kinetic energy (T) along x-axis can be derived as:

1 (𝑚𝑣)2 𝑝2 (1)
𝐾. 𝐸. (𝑇) = 𝑚𝑣 2 = =
2 2𝑚 2𝑚
Where m and v are mass and the velocity, respectively; and p represents the angular momentum whose squared
operator is:

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 13

−ℎ2 𝜕 2 (2)
𝑝̂𝑥2 =
4𝜋 2 𝜕𝑥 2
Now putting the value of momentum squared from equation (2) into equation (1), we get:

−ℎ2 𝜕 2 (3)
𝑇̂𝑥 =
8𝜋 2 𝑚 𝜕𝑥 2
The expressions of various quantum mechanical operators are given below.

Table 1. Various important physical properties and their corresponding quantum mechanical operators.

Physical property Operator

Name Symbol Symbol Operation

Position x 𝑥̂ Multiplication by x

Position squared x2 𝑥̂ 2 Multiplication by x2

Momentum px 𝑝̂𝑥 ℎ 𝜕
2𝜋𝑖 𝜕𝑥
Momentum squared px 2 𝑝̂𝑥2 −ℎ2 𝜕 2
4𝜋 2 𝜕𝑥 2
Kinetic energy 𝑃2 𝑇̂𝑥 −ℎ2 𝜕 2
𝑇=
2𝑚 8𝜋 2 𝑚 𝜕𝑥 2
Potential energy V(x) 𝑉̂ (𝑥) Multiplication by V(x)

Total energy 𝐸 = 𝑇 + 𝑉(𝑥) ̂


𝐻 −ℎ2 𝜕 2
+ 𝑉(𝑥)
8𝜋 2 𝑚 𝜕𝑥 2

For three dimensional systems, the total operator can be obtained by summing the individual
operators along three different axes. For instance, some important three-dimensional operators are:

−ℎ2 𝜕 2 𝜕2 𝜕2 (4)
𝑇̂ = ( + + )
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

ℎ 𝜕 𝜕 𝜕 (5)
𝑝̂ = ( + + )
2𝜋𝑖 𝜕𝑥 𝜕𝑦 𝜕𝑧

−ℎ2 𝜕 2 𝜕2 𝜕2 (6)
̂=
𝐻 ( + + ) + 𝑉(𝑥, 𝑦, 𝑧)
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

Copyright © Mandeep Dalal


14 A Textbook of Physical Chemistry – Volume I

 The Third Postulate


If ψ is a well-behaved function for the given state of system and  is a suitable operator for a
particular physical property, then the operation on ψ by the operator  gives the function ψ multiplied by the
value of the physical property which can be constant or variable but always real (R). Mathematically, it can
be shown as:

𝐴̂𝜓 = 𝑅𝜓 (7)

Explanation: The third postulate of quantum mechanics actually connects the first and second postulate of
quantum mechanics. The first postulate talks about the possibility of describing a quantum mechanical state
mathematically, while the second postulate says that the values of all physical properties in the quantum world
are obtained by the operator rather than the simple classical formula. Now the third postulate says that if we
operate the operator (from second postulate) over the wave function (from first postulate), we will get the value
of the corresponding physical property.
However, at this point, a new problem arises as we do not know the exact mathematical description
i.e. the wave function of any quantum mechanical state; and the operators need the absolute mathematical
description of the quantum mechanical state to yield any actual result. Now though we know the expressions
of different operators proposed by the second postulate; the first postulate speaks only about the presence of a
single-valued, continuous and finite mathematical function but does not give actual function itself; and without
the knowledge of actual “wave functions”, the operators are pretty much useless. Therefore, one would think
that there must be some route by which the wave functions are obtained first, which would be used as operand
afterward. However, the procedure to find the exact mathematical descriptions of various quantum mechanical
states is somewhat more synergistic. The “magic mystery” is that all the operators need absolute expression of
the wave function that defines the quantum mechanical state except one, the most famous “Hamiltonian
operator”. The special thing about the Hamiltonian operator is that it does not necessarily need the absolute
form but the symbolic form only to yield the value of its physical property i.e. energy. Nevertheless, in the
process of applying the Hamiltonian operator over the symbolic form of the wave function, the absolute
expression is also obtained. Mathematically,

̂ 𝜓 = 𝐸𝜓
𝐻 (8)

After putting the expression of the Hamiltonian operator in equation (8) and then rearranging, we get:

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚(𝐸 − 𝑉)𝜓 (9)


+ + + =0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

The second-order differential equation i.e. equation (9) is the famous Schrodinger wave equation, the
solution of which gives not only the energy but the wave function as well. Now, once the exact expression of
the wave function representing a particular state is known, other operators can be operated over it to find their
values.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 15

 The Fourth Postulate


If the value of the physical property obtained after multiplying the wave function by the
corresponding operator is constant (postulate 3), the value is called as the eigen-value and is directly
reportable; and the wave function will be labeled as the eigen-function of the operator used.
Explanation: The third postulate said that when the wave function of a particular quantum mechanical state
is multiplied by the operator of an observable quantity, we get a real value multiplied by the wave function
itself; however, the value obtained so can be constant or variable. Mathematically,

The constant value of the observable quantity can be reported directly, and the function is called the
eigenfunction of the operator under consideration.
 The Fifth Postulate
If the value of the physical property obtained after multiplying the wave function by the
corresponding operator is variable i.e. non-eigen, the value can be reported only after averaging it over the
whole configurational space.

∮ 𝜓 ∗ 𝑂̂𝜓 𝑑𝜏 (10)
< 𝑎 > 𝑜𝑟 ā =
∮ 𝜓 ∗ 𝜓 𝑑𝜏

Explanation: As we have seen in the fourth postulate that the value obtained by multiplying the Hermitian
operator with any quantum mechanical state can also be variable in nature. For instance, if we multiply a wave
function simply by position operator, we will get

𝑥̂𝜓 = 𝑥𝜓 (11)

or
𝑥𝜓 (12)
𝑥̂ =
𝜓
Now because “x” is a variable number, it must have reported as an average value before any further rational
argument is made.
Therefore, we can say that the fifth postulate is simply an extension of the fourth postulate; i.e. the
fourth postulate is used to obtain the value of a particular physical property if it is an eigenvalue, however, the
fifth postulate is employed to calculate all non-eigenvalues.

Copyright © Mandeep Dalal


16 A Textbook of Physical Chemistry – Volume I

 Derivation of Schrodinger Wave Equation


The Schrodinger wave equation can be derived from the classical wave equation as well as from the
third postulate of quantum mechanics. Now though the two routes may appear completely different, the final
result is just the same indicating the objectivity of the quantum mechanical system.
 The Derivation of Schrodinger Wave Equation from Classical Wave Equation
After the failure of the Bohr atomic model to comply with the Heisenberg’s uncertainty principle and
dual character proposed by Louis de Broglie in 1924, an Austrian physicist Erwin Schrodinger developed his
legendary equation by making the use of wave-particle duality and classical wave equation. In order to
understand the concept involved, consider a wave traveling in a string along the x-axis with velocity v.

Figure 1. The wave motion in a string.

It can be clearly seen that the amplitude of the wave at any time t is the function of displacement x, and the
equation for wave motion can be formulated as given below.

𝜕2𝑦 1 𝜕2𝑦 (13)


=
𝜕𝑥 2 𝑣 2 𝜕𝑡 2
Therefore, we can say that y is a function of x well at t.

𝑦 = 𝑓(𝑥)𝑓′(𝑡) (14)

Where f(x) and f´(t) are the functions of coordinate x and time, respectively. The nature of the function f(x) can
be understood by taking the example of stationary or the standing wave.
A standing wave is created in a string fixed between two points with a wave traveling in one direction,
and when it strikes the other end, it gets reflected with the same velocity but in negative amplitude. This would
create vibrations in that string with or without nodes depending upon the frequency incorporated. We can
create fundamental mode (0 node), first overtone (1 node) or second overtone (2 nodes) just by changing the
vibrational frequency. The nature of these standing or stationary waves can be understood more clearly by the
diagram given below.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 17

Figure 2. Standing waves in a string.

The mathematical description for such a wave motion is

𝑓 ′ (𝑡) = 𝐴 sin 2𝜋𝜈𝑡 (15)

Where A is a constant representing maximum amplitude and ν is the frequency of the vibration. Now putting
the value of f´(t) from equation (15) in equation (14), we get

𝑦 = 𝑓(𝑥) 𝐴 𝑆𝑖𝑛 2𝜋𝜈𝑡 (16)

Differentiating the above equation w.r.t. t, we are left with

𝜕𝑦 (17)
= 𝑓(𝑥) 𝐴2𝜋𝜈 𝐶𝑜𝑠 2𝜋𝜈𝑡
𝜕𝑡
Differentiating again

𝜕2𝑦 (18)
= −𝑓(𝑥) 4𝜋 2 𝜈 2 𝐴 𝑆𝑖𝑛 2𝜋𝜈𝑡
𝜕𝑡 2
𝜕2𝑦 (19)
= −4𝜋 2 𝜈 2 𝑓(𝑥) 𝑓 ′ (𝑡)
𝜕𝑡 2
Now differentiating equation (14) w.r.t. x only, we get

𝜕𝑦 𝜕𝑓(𝑥) (20)
= 𝑓′(𝑡)
𝜕𝑥 𝜕𝑥
Differentiating again

𝜕2𝑦 𝜕 2 𝑓(𝑥) (21)


= 𝑓′(𝑡)
𝜕𝑥 2 𝜕𝑥 2
Now put the value of equation (19) and (21) in equation (13), we get

Copyright © Mandeep Dalal


18 A Textbook of Physical Chemistry – Volume I

𝜕 2 𝑓(𝑥) 1 (22)
𝑓′(𝑡) = ( ) [−4𝜋 2 𝜈 2 𝑓(𝑥) 𝑓 ′ (𝑡)]
𝜕𝑥 2 𝑣2

𝜕 2 𝑓(𝑥) −4𝜋 2 𝜈 2 (23)


= 𝑓(𝑥)
𝜕𝑥 2 𝑣2
The equation (23) is now time-independent; and therefore, shows the amplitude dependence only upon the
coordinate x. Since c = νλ (v = c/λ), the velocity of the wave can also be replaced by the multiplication of
frequency and wavelength i.e. v = νλ.

𝜕 2 𝑓(𝑥) −4𝜋 2 𝜈 2 (24)


= 𝑓(𝑥)
𝜕𝑥 2 𝜈 2 𝜆2
𝜕 2 𝑓(𝑥) −4𝜋 2 (25)
= 𝑓(𝑥)
𝜕𝑥 2 𝜆2
The symbol of the function f(x) is replaced by popular ψ(x) or simply the ψ.

𝜕 2 𝜓 −4𝜋 2 (26)
= 𝜓
𝜕𝑥 2 𝜆2
Also, as we know that λ=h/mv, the equation (26) becomes

𝜕 2 𝜓 −4𝜋 2 𝑚2 𝑣 2 (27)
= 𝜓
𝜕𝑥 2 ℎ2
𝜕 2 𝜓 4𝜋 2 𝑚2 𝑣 2 (28)
+ 𝜓=0
𝜕𝑥 2 ℎ2
Furthermore, as the total energy (E) is simply the sum of the potential (V) and kinetic energy, we can say that

𝑚𝑣 2 (29)
𝐸= +𝑉
2

𝑚𝑣 2 = 2(𝐸 − 𝑉) (30)

After putting the value of mv2 from equation (30) in equation (28), we get

𝜕 2 𝜓 8𝜋 2 𝑚 (31)
− (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 ℎ2
For three-dimension i.e. ψ(x, y, z), the above equation can be extended to following

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (32)
+ + − (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

The above-mentioned second order differential equation i.e. equation (32) is our popular form of the
Schrodinger wave equation.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 19

 The Derivation of Schrodinger Wave Equation from the Postulates of Quantum Mechanics
The Schrodinger wave equation can be derived using the first three postulates of quantum mechanics.
In other words, we can say that the Schrodinger wave equation is nothing but the rearranged form of the
following equation:

̂ 𝜓 = 𝐸𝜓
𝐻 (12)

In order to prove the above claim, consider a single particle having “m” mass that moves with a velocity “v”
in the three-dimensional region. The sum of its kinetic and potential energy can be given as:

𝐸 =𝑇+𝑉 (13)

However, we know that

1 𝑝2 (14)
𝑇 = 𝑚𝑣 2 =
2 2𝑚
Where “p” represents the total linear momentum of the particle under consideration. Furthermore, as we also
know that

𝑝2 = 𝑝𝑥2 + 𝑝𝑦2 + 𝑝𝑧2 (15)

Where px, py and pz are the magnitudes of total linear momentum along x, y and z-axis, respectively. Now
putting the value of p2 from equation (15) into equation (14), we get the following

𝑝𝑥2 + 𝑝𝑦2 + 𝑝𝑧2 (16)


𝑇=
2𝑚
And now put the value of kinetic energy from equation (16) into equation (13). We get

𝑝𝑥2 + 𝑝𝑦2 + 𝑝𝑧2 (17)


𝐸= +𝑉
2𝑚
However, from the second postulate of quantum mechanics, we know that the expressions for linear
momentum operator along three different directions are:

ℎ 𝜕 (18)
𝑝
̂𝑥 =
2𝜋𝑖 𝜕𝑥
ℎ 𝜕 (19)
𝑝
̂𝑦 =
2𝜋𝑖 𝜕𝑦

ℎ 𝜕 (20)
𝑝
̂𝑧 =
2𝜋𝑖 𝜕𝑧
The operator of “V” is simply itself as it is a function of position coordinates only.

Copyright © Mandeep Dalal


20 A Textbook of Physical Chemistry – Volume I

Hence, after putting values of linear momentum operators and potential energy operator in equation (17), the
operator for total energy (Hamiltonian operator) becomes

1 ℎ 𝜕 2 ℎ 𝜕 2 ℎ 𝜕 2 (21)
̂=
𝐻 [( ) +( ) +( ) ]+𝑉
2𝑚 2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑧

ℎ2 𝜕2 𝜕2 𝜕2 (22)
̂=−
𝐻 ( + + )+𝑉
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

ℎ2 (23)
̂=−
𝐻 ∇2 + 𝑉
8𝜋 2 𝑚
Where
𝜕2 𝜕2 𝜕2
∇2 = + +
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2
represents the Laplacian operator.
Now, after putting the value of Hamiltonian operator from equation (22) into equation (12) i.e. given by the
third postulate of quantum mechanics, get

ℎ2 𝜕2 𝜕2 𝜕2 (24)
[− 2 ( 2 + 2 + 2 ) + 𝑉] 𝜓 = 𝐸𝜓
8𝜋 𝑚 𝜕𝑥 𝜕𝑦 𝜕𝑧

ℎ2 𝜕2 𝜕2 𝜕2 (25)
[− ( + + ) + 𝑉] 𝜓 − 𝐸𝜓 = 0
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

ℎ2 𝜕 2 𝜓 ℎ2 𝜕 2 𝜓 ℎ2 𝜕 2 𝜓 (26)
− − − + 𝑉𝜓 − 𝐸𝜓 = 0
8𝜋 2 𝑚 𝜕𝑥 2 8𝜋 2 𝑚 𝜕𝑦 2 8𝜋 2 𝑚 𝜕𝑧 2

Multiplying the equation (26) throughout by


8𝜋 2 𝑚

ℎ2
we get

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 8𝜋 2 𝑚 (27)
+ + − 𝑉𝜓 + 𝐸𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2 ℎ2

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (28)
+ + + (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

Equation (28) is the most popular form of the Schrodinger wave equation for three dimensional systems. In
the case of two and one dimensional systems first three terms can be reduced to two and one, respectively.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 21

 Max-Born Interpretation of Wave Functions


In 1926, a German physicist Max Born formulated a rule which is generally called as the Born law
or Born's rule of quantum mechanics, giving the probability that a measurement on a quantum system will
yield a given result. In other words, it states that the probability density of finding the particle at a given point
is proportional to the square of the magnitude of the particle's wavefunction at that point. The Max-Born
interpretation is one of the key concepts of quantum mechanics to understand wave-particle duality.
Let us suppose that the particle under consideration is an electron whose way of existence is
represented by a mathematical expression ψ which is a function of the electron’s coordinates i.e. x, y, and z.
Max Born actually suggested that because this mathematical expression is single-valued, continuous and finite
i.e. wave-like; one can opt the same route to find it’s intensity as we use in case of light or sound waves. In the
case of light or sound waves, the intensity at any point can simply be obtained by squaring the amplitude i.e.
ψ of the wave at the same point. Therefore, in the case of the electron, the square of the amplitude of electron
wave (ψ2) at a particular point also gives the intensity of the electron wave at the same point. In other words,
the density of electron wave (probability density) at a point for a quantum mechanical state is simply obtained
by the square of the magnitude of the corresponding wavefunction at the same point. Mathematically, we can
show this as:

𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 𝑑𝑒𝑛𝑠𝑖𝑡𝑦 = |𝜓|2 = 𝜓𝜓 ∗ (29)

Where ψ* designates a complex conjugate of the wave function ψ. The reason for using ψ* lies in the fact that
the wave function representing a quantum mechanical state is not always real but be imaginary as well.
However, as the probability density should always be real, ψψ* is more appropriate than simple ψ2. In other
words, if the wave function defining the quantum mechanical state is real, we can use ψ 2 as the probability
density; nevertheless, if the wave function does contain the imaginary part (like ψ = a + ib), ψψ* must be used
to yield real values. This can be explained by taking an imaginary expression ψ and then multiplying it by its
complex conjugate ψ* to yield real value.

𝜓 = 𝑎 + 𝑖𝑏; 𝜓 ∗ = 𝑎 − 𝑖𝑏 (30)

or

𝜓𝜓 ∗ = (𝑎 + 𝑖𝑏) × (𝑎 − 𝑖𝑏) (31)

or

𝜓𝜓 ∗ = 𝑎2 + 𝑏 2 (32)

Moreover, if ψ is real, ψ = ψ*, ψψ* becomes ψ2, the value we have already discussed.
Now though the probability density in space is not a constant parameter (ψ is not constant), in a very
small segment it can be considered constant. Now let us discuss the Max-Born interpretation for one, two and
three dimensional systems.

Copyright © Mandeep Dalal


22 A Textbook of Physical Chemistry – Volume I

 One Dimensional Systems


The probability of finding the particle in any one-dimensional system in the region from x to x+dx
must be obtained by integrating ψ2 from x to x+dx i.e. by finding the area under the curve from x to x+dx.
Now although the ψ2 or ψψ* (because ψ is continuous) varies continuously with x, the decrease or
increase in ψ2 can be neglected and it can be assumed that it remains constant as we move from x to x+dx.

Figure 3. Born interpretation of wave function and probability density in a one-dimensional system.

Therefore, the area of the shaded region and hence the probability of finding the particle can be obtained just
by multiplying the length (or height i.e. ψ2) with the width (dx) of the narrow rectangle. Thus we can say that

𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = |𝜓|2 × (𝑑𝑥) = 𝜓𝜓 ∗ 𝑑𝑥 = 𝜓 2 𝑑𝑥 (33)

Since the chances of finding the particle over whole length (whole configurational space) must be unity, we
get the following

(34)
∮|𝜓|2 × (𝑑𝑥) = ∮ 𝜓𝜓 ∗ 𝑑𝑥 = ∮ 𝜓 2 𝑑𝑥

 Two Dimensional Systems


The probability of finding the particle in an area element dx dy (dA), situated at a distance r distance
from the center, would be ψ (x, y) × ψ* (x, y) × dx × dy; or in short can be written as ψψ*dA. Hence, it must be
obtained by integrating ψ2 from (x, y) to (x+dx, y+dy) i.e. by finding the area under the curve dA. Now although
the ψ2 or ψψ* (because ψ is continuous) varies continuously with coordinates (x y), the decrease or increase in
ψ2 can be neglected and it can be assumed that it remains constant as we move from (x, y) to (x+dx, y+dy).
Therefore, the area of the shaded region and hence the probability of finding the particle can be
obtained just by multiplying the magnitude of the wave function (ψ2) with the area (dA) of the area element.
Thus we can say that

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 23

𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = |𝜓|2 × (𝑑𝐴) = 𝜓𝜓 ∗ 𝑑𝐴 = 𝜓 2 𝑑𝐴 (35)

Since the chances of finding the particle over the whole area (whole configurational space) must be unity, we
get the following

(36)
∮|𝜓|2 × (𝑑𝐴) = ∮ 𝜓𝜓 ∗ 𝑑𝐴 = ∮ 𝜓 2 𝑑𝐴

The pictorial representation of the area element is given below.

Figure 2. Born interpretation of wave function and probability density in two dimensional systems.

 Three Dimensional System


The probability of finding the particle in a volume element dx dy dz (dV), situated at a distance r
distance from the center, would be ψ (x, y, z) × ψ* (x, y, z) × dx × dy× dz; or in short can be written as ψψ*dV.
must be obtained by integrating ψ2 from (x, y, z) to (x+dx, y+dy, z+dz) i.e. by finding the area under the curve
from (x, y, z) to (x+dx, y+dy, z+dz). Now although the ψ2 or ψψ* (because ψ is continuous) varies continuously
with coordinates (x, y, z), the decrease or increase in ψ2 can neglected and it can be assumed that it remains
constant as we move from (x, y, z) to (x+dx, y+dy, z+dz).
Therefore, the probability of finding the particle can be obtained just by multiplying the magnitude
of the wave function (ψ2) with the volume (dV) of the area element. Thus we can say that

𝑃𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = |𝜓|2 × (𝑑𝑉) = 𝜓𝜓 ∗ 𝑑𝑉 = 𝜓 2 𝑑𝑉 (35)

Since the chances of finding the particle over the whole area (whole configurational space) must be unity, we
get the following

(36)
∮|𝜓|2 × (𝑑𝑉) = ∮ 𝜓𝜓 ∗ 𝑑𝑉 = ∮ 𝜓 2 𝑑𝑉

The pictorial representation of the area element is given below.

Copyright © Mandeep Dalal


24 A Textbook of Physical Chemistry – Volume I

Figure 3. Born interpretation of wave function and probability density in three dimensional systems.

 The Heisenberg’s Uncertainty Principle


In quantum mechanical world, the Heisenberg's uncertainty principle (or simply the uncertainty
principle) is one of a variety of mathematical inequalities asserting a fundamental limit to the precision with
which certain pairs of physical properties of a particle, known as complementary variables or canonically
conjugate variables such as position x and momentum p, can be known. The concept was first introduced in
1927, by a German physicist Werner Heisenberg.
The Heisenberg’s uncertainty principle states that the more precisely the position of some particle is
determined, the less precisely its momentum can be known, and vice versa.
The formal inequality relating the standard deviation of position Δx and the standard deviation of
momentum Δpx was derived by Earle Hesse Kennard later that year and by Hermann Weyl in 1928:

ℎ (37)
Δ𝑥. Δ𝑝𝑥 ≥
4𝜋
or

Δx . Δpx ≥ ħ/2 (38)

Where ħ is the reduced Planck’s constant, which is obviously equal to the Planck’s constant divided by 2π.
Besides the equation (37), the is also an energy-time uncertainty relation given by W. Heisenberg which states
that higher the lifetime of a quantum mechanical state, less uncertain would be the energy value.
Mathematically, it can be shown as:

ℎ (39)
Δ𝐸. Δ𝑡 ≥
2𝜋
Where ΔE and Δt represent the uncertainties in the energy and time respectively.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 25

 Position Momentum Uncertainty


Among various kinds of uncertainties, the position-momentum uncertainty is one of the popular kind
that arises as a consequence of wave-particle duality. In order to understand the relation, we first need to study
the effect of wave behavior on the simultaneous measurement of position about x-coordinate and the linear
momentum component along the x-axis for a microscopic particle.
Consider a beam of particles traveling with a momentum “p” along the y-direction, and this beam
finally strikes a narrow slit of width “w”. Now, from the principles of optics, we know that the uncertainty in
the position of the particle along x-axis must be equal to the slit width. In other words, as the width of the slit
is along x-axis, any particle that strikes the detector must have crossed the Δx region i.e. w, the slit width
available. However, we exactly don’t know where it does cross from. It could be along the center of the slit,
or along a line slightly above or below the central trajectory. Therefore, the slit width (w = Δx) would be equal
to a crossing domain that we are uncertain about. However, a diffraction pattern will be observed in the case
of microscopic particles because of their wave-like character. The amplitude of the wave at a particular point
on the detector represents the number of the particles reaching that point. Now because of this diffraction, the
incident beam does not strike only at the central point O but also at the above and below to it. It means that
some particles do reach upward and downward to O, suggesting that the part of their linear momentum is
transferred along x-axis also.

Figure 4. The diffraction of electron waves by single slit systems.

Copyright © Mandeep Dalal


26 A Textbook of Physical Chemistry – Volume I

The x-component of linear momentum of the wave (aka particle) diffracted at an angle α can be obtained by
the rectangular resolution of the linear momentum vector. The particles diffracted upward and downward at
an angle α will yield the x-component as P sinα and −P sinα, respectively. Now because a large number of
particles reach the plate in between +α to −α i.e in between the first minimums, half of the momentum spread
in the central diffraction peak should give the uncertainty in the momentum along x-axis. Mathematically, we
can say that

∆𝑝𝑥 = 𝑃 𝑠𝑖𝑛𝛼 (40)

Multiplying the above equation by the uncertainty in the position i.e. width of the slit used for the measurement
purpose, we get

∆𝑥. ∆𝑝𝑥 = 𝑤. 𝑃 𝑠𝑖𝑛𝛼 (41)

Here, it is very important to recall the fact that the condition which must be satisfied to obtain the first minima
is that the path difference between the waves reaching the minima point should be an integral multiple of λ/2.

Figure 5. The calculation for 1st order diffraction for electron wave in single slit systems.

Hence we have the following equalities from the diagram given above.

𝐴𝑄 = 𝐷𝑄 (42)

𝐶𝑄 = 𝑑𝑖𝑓𝑓𝑒𝑟𝑒𝑛𝑐𝑒 𝑖𝑛 𝑡ℎ𝑒 𝑝𝑎𝑡ℎ 𝑙𝑒𝑛𝑔𝑡ℎ (43)

Now because the distance of the detector is very large as compared to the slit width, AQ and CQ can be
considered parallel to each other i.e. AQ ‖ DQ. Hence, we can say that

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 27

< 𝐴𝐷𝐶 = 90° (44)

< 𝐶𝐴𝐷 = 𝛼 (45)

also
𝑤 (46)
𝐴𝐶 =
2
𝜆 (47)
𝐶𝐷 =
2
From the trigonometric relations, we get

𝐶𝐷 (48)
= 𝑆𝑖𝑛 𝛼
𝐴𝐶
𝐶𝐷 = 𝐴𝐶 𝑆𝑖𝑛 𝛼 (49)

Putting the values of AC and CD from equation (46) and (47) in equation (49), we get

𝜆 𝑤 (50)
= 𝑆𝑖𝑛 𝛼
2 2
𝜆 = 𝑤 𝑆𝑖𝑛 𝛼 (51)

Now, after putting the value of w from equation (51) in equation (41), we get

𝜆 (52)
∆𝑥. ∆𝑝𝑥 = . 𝑃 𝑠𝑖𝑛𝛼
𝑆𝑖𝑛 𝛼
∆𝑥. ∆𝑝𝑥 = 𝜆. 𝑃 (53)

Using the de Broglie relation (λ = h/p) in equation (53), we get

ℎ (54)
∆𝑥. ∆𝑝𝑥 = .𝑃
𝑃
∆𝑥. ∆𝑝𝑥 = ℎ (55)

Now because we didn’t define the uncertainty very precisely, we should not use the “equal” sign. Therefore,
the above equation can be reduced to the following.

∆𝑥. ∆𝑝𝑥 ≈ ℎ (56)

This eventually means that decreasing the uncertainty in the position of the incident particle (decreasing the
slit width) would result in a higher uncertainty in the momentum along x-axis; while the higher slit width does
give more precise momentum but small precision in the calculation of the position of the incident particle.

Copyright © Mandeep Dalal


28 A Textbook of Physical Chemistry – Volume I

 Energy Time Uncertainty


The uncertainty principle doesn’t limit itself to position-momentum only but can also be applied to
some other pairs of conjugate variables. All the variable pairs whose products have the same dimension as the
Plank’s constant h (Js) are said to be a conjugate pair. Besides the position-momentum, another famous
uncertainty is relation energy-time because the product of these two quantities (energy × time) also has the
unit of h (Js).

∆𝐸. ∆𝑡 ≈ ℎ (57)

Where ΔE and Δt are uncertainties in energy and time, respectively. This popular relation can be derived
directly from the concept of wave-particle duality. In the quantum mechanical world, a particle is supposed to
possess a wave packet. Now, let us consider that this wave packet occupies the Δx region along the direction
x-direction and travels with a velocity v. The time it needs to pass a certain point in x-direction has an
uncertainty magnitude of Δt, and can be formulated as:

Δ𝑥 (58)
∆𝑡 =
𝑣
Now because this wave packet occupies the region Δx, the momentum uncertainty along x-axis can be given
by the following relation.

ℎ (59)
∆𝑝𝑥 =
Δ𝑥
or

ℎ (60)
Δ𝑥 =
∆𝑝𝑥

Putting the value of Δx from equation (60) in equation (58), we get

ℎ (61)
∆𝑡 =
𝑣Δ𝑝𝑥

Moreover, we also know that

𝑝𝑥2 (62)
𝐸=
2𝑚
Differentiating the above equation w.r.t px, we get

𝑑𝐸 ∆𝐸 𝑝𝑥 𝑚𝑣 (63)
= = = =𝑣
𝑑𝑝𝑥 ∆𝑝𝑥 𝑚 𝑚

𝑑𝐸 (64)
∆𝐸 = ∆𝑝 = 𝑣. ∆𝑝𝑥
𝑑𝑝𝑥 𝑥

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 29

Multiplying equation (63) and (64), we get

ℎ (65)
∆𝐸. ∆𝑡 = 𝑣∆𝑝𝑥 .
𝑣Δ𝑝𝑥

∆𝐸. ∆𝑡 ≈ ℎ (66)

The physical interpretation of the above relation can be viewed in terms of fluctuating energy level with a total
ΔE uncertainty if the system does not stay in it longer than Δt interval of time i.e. lifetime of the state.

 Quantum Mechanical Operators and Their Commutation Relations


An operator may be simply defined as a mathematical procedure or instruction which is carried out
over a function to yield another function.

(Operator) . (Function) = (Another function) (67)

The function used on the left-hand side of the equation (67) is called as the operand i.e. the function
over which the operation is actually carried out. The operator alone has no significance but when operated over
a certain mathematical description, these operators can provide very detailed insights into those functions.
Some of the simple illustrations of equation (67) are given below.
i) Consider the differential operator d/dx whose operation has to be studied over the function y = x5. The
mathematical treatment is

𝑑𝑦 𝑑 5 (68)
= 𝑥 = 5𝑥 4
𝑑𝑥 𝑑𝑥
The operation of d/dx on y means that the rate of change of function y w.r.t. the variable x. The expression x5
is the operand while the 5x4 is the final result of our differential operator.
ii) Consider the integral operator ʃ (y) dx whose operation has to be studied over the function y = x5. The
mathematical treatment is

𝑥6 (69)
∫ 𝑦(𝑑𝑥) = ∫ 𝑥 5 (𝑑𝑥) =
6

The operation of ʃdx on y means that we can find the function whose derivative is x5. The expression x5 is the
operand while the x6/6 is the final result of our integral operator.
In a similar way, the multiplication of a function by a constant number, or taking the square and cube
roots of any function are also the operators which give some other function after operating them over the
operand. The symbol of the operator typically carries a cap over it (Â) which differentiates it from the function
used in the whole procedure.

Copyright © Mandeep Dalal


30 A Textbook of Physical Chemistry – Volume I

 Algebra of Operators
Just like the normal algebra, the resultants like addition or the multiplication of operators also follow
certain rules; however, these rules are different from the typical algebra. Some of the most important rules of
operator algebra are given below.
1. Addition and subtraction of operators: Let A and B as two different operators; f as the function that has
to be used as the operand. Then, the addition and subtraction of these two operators must be carried out in the
manner discussed below.

(𝐴̂ + 𝐵̂)𝑓 = 𝐴̂𝑓 + 𝐵̂𝑓 (70)

and

(𝐴̂ − 𝐵̂)𝑓 = 𝐴̂𝑓 − 𝐵̂𝑓 (71)

2. Multiplication of operators: If A and B as two different operators; and f as the function that has to be used
as operand. Then, the multiplication of these two operators must be carried out in the manner discussed below.

𝐴̂𝐵̂𝑓 = 𝑓 ′′ (72)

The interpretation of the above equation is that first we need to operate B on f, which would give us another
function f´, which in turn is further used as the operand for operator giving the final result fʺ. In other words,
we can say that when multiplication of two or more operators is used, we should follow from left to right.
Moreover, the square or cube of a particular operator must be considered as double or triple multiplication of
the operator itself; mathematically, it can be shown as given below.

𝐴̂2 𝑓 = 𝐴̂𝐴̂𝑓 (73)

At this point it also very important to discuss one of the most fundamental properties of operator
multiplication, the commutation relation or the commutation rule. Consider two operators, A and B which can
be operated over the function f.

𝑑 (74)
𝐴̂ = ; 𝐵̂ = 𝑥; 𝑓 = 𝑥3
𝑑𝑥
Now

𝑑 𝑑 4 (75)
𝐴̂𝐵̂𝑓 = 𝑥(𝑥 3 ) = 𝑥 = 4𝑥 3
𝑑𝑥 𝑑𝑥
And

𝑑 3 (76)
𝐵̂𝐴̂𝑓 = 𝑥 (𝑥 ) = 𝑥(3𝑥 2 ) = 3𝑥 3
𝑑𝑥
From equation (75) and (76), it the clear that in this case

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 31

𝐴̂𝐵̂𝑓 ≠ 𝐵̂𝐴̂𝑓 (77)

These operators are said to be non-commutating with the commutator given below.

𝐴̂𝐵̂ − 𝐵̂𝐴̂ = 4𝑥 3 − 3𝑥 3 (78)

However, the two operators are said to be commute if their result is the same even after reverting their order
of application. Mathematically, it can be stated as given by equation (79).

𝐴̂𝐵̂𝑓 = 𝐵̂𝐴̂𝑓 (79)

This is quite different from the normal algebra in which the product of two numbers is always the same
irrespective of the order of multiplication (x.y = y.x). Summarizing the commutation rule, it can be concluded
that

̂ 𝐵̂] = 𝐴̂𝐵̂ − 𝐵̂𝐴̂ = 0 → 𝐶𝑜𝑚𝑚𝑢𝑡𝑎𝑡𝑖𝑛𝑔


[𝐴, (80)

and

̂ 𝐵̂] = 𝐴̂𝐵̂ − 𝐵̂𝐴̂ ≠ 0 → 𝑁𝑜𝑛-𝑐𝑜𝑚𝑚𝑢𝑡𝑎𝑡𝑖𝑛𝑔


[𝐴, (81)

3. Linear Operator: An operator  is said to be a linear operator if its application on the sum of two functions
f and g gives the same result as the sum of its individual operations. Mathematically, it can be shown as given
below.

𝐴̂(𝑓 + 𝑔) = 𝐴̂𝑓 + 𝐴̂𝑔 (82)

For example, consider the differential operator A; with f and g as the functions which have to be used as the
operand.

𝑑 (83)
𝐴̂ = ; 𝑓 = 2𝑥 2 ; 𝑔 = 3𝑥 2
𝑑𝑥
or

𝑑 𝑑 (84)
𝐴̂(𝑓 + 𝑔) = (2𝑥 2 + 3𝑥 2 ) = (5𝑥 2 ) = 10𝑥
𝑑𝑥 𝑑𝑥
or

𝑑 𝑑 (85)
𝐴̂𝑓 + 𝐴̂𝑔 = (2𝑥 2 ) + (3𝑥 2 ) = 4𝑥 + 6𝑥 = 10𝑥
𝑑𝑥 𝑑𝑥
Hence, from equation (84) and equation (85), it is clear that the differential operator is clearly linear
in nature. On the other hand, the “square root” operator is not linear as it does not give the same result when
operated individually.

Copyright © Mandeep Dalal


32 A Textbook of Physical Chemistry – Volume I

 Some Important Quantum Mechanical Operators


One of the most basic and very popular operators in quantum mechanics is the Laplacian operator,
typically symbolized as ∇2, and is given by the following expression.

𝜕2 𝜕2 𝜕2 (86)
∇2 = + +
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

The popular form of the Schrodinger equation can be written in terms of Laplacian operator as well.

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (87)
+ + + (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

or

8𝜋 2 𝑚 (88)
∇2 𝜓 + (𝐸 − 𝑉)𝜓 = 0
ℎ2
The Laplacian operator is pronounced as “del squared”. This operator is also a part of the “mighty”
Hamiltonian operator which forms the basis for value evaluation for other operators, as we have already
discussed in the postulates of quantum mechanics. The Hamiltonian operator is typically symbolized as 𝐻
̂ and
is given by the following expression.

ℎ2 𝜕2 𝜕2 𝜕2 (89)
̂=−
𝐻 ( + + )+𝑉
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

or

ℎ2 (90)
̂=−
𝐻 ∇2 + 𝑉
8𝜋 2 𝑚
The popular form of the Schrodinger equation is written in terms of the Hamiltonian operator as well.

̂ 𝜓 = 𝐸𝜓
𝐻 (91)

or

ℎ2 𝜕2 𝜕2 𝜕2 (92)
[− ( + + ) + 𝑉] 𝜓 = 𝐸𝜓
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

or

ℎ2 (93)
(− ∇2 + 𝑉) 𝜓 = 𝐸𝜓
8𝜋 2 𝑚

Furthermore, we know from the third postulate of quantum mechanics that owing to the constant value of E
(eigenvalue) the wave function ψ can be labeled as eigenfunction.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 33

Therefore, the Schrodinger equation is also called as the “eigen value equation”. Simplifying this, we can say
that

(𝐸𝑛𝑒𝑟𝑔𝑦 𝑜𝑝𝑒𝑟𝑎𝑡𝑜𝑟)(𝑊𝑎𝑣𝑒 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛) = (𝐸𝑛𝑒𝑟𝑔𝑦)(𝑊𝑎𝑣𝑒 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛) (94)

The equation (94) is applicable to observables in the quantum mechanical world.


For three dimensional systems, like the Hamiltonian, the operator can be obtained by summing the
individual operators along three different axes. For instance, some important three-dimensional operators are:

−ℎ2 𝜕 2 𝜕2 𝜕2 (95)
𝑇̂ = ( + + )
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

ℎ 𝜕 𝜕 𝜕 (96)
𝑝̂ = ( + + )
2𝜋𝑖 𝜕𝑥 𝜕𝑦 𝜕𝑧

The list of various important quantum mechanical operators in one dimension, along with their mode of
operation is given below.

Table 2. Name and symbols of various important physical properties and their corresponding quantum
mechanical operators.

Physical property Operator

Name Symbol Symbol Operation

Position x 𝑥̂ Multiplication by x

Position squared x2 𝑥̂ 2 Multiplication by x2

Position cubed x3 𝑥̂ 2 Multiplication by x3

Momentum px 𝑝̂𝑥 ℎ 𝜕
2𝜋𝑖 𝜕𝑥
Momentum squared px 2 𝑝̂𝑥2 −ℎ2 𝜕 2
4𝜋 2 𝜕𝑥 2
Kinetic energy 𝑃2 𝑇̂𝑥 −ℎ2 𝜕 2
𝑇=
2𝑚 8𝜋 2 𝑚 𝜕𝑥 2
Potential energy V(x) 𝑉̂ (𝑥) Multiplication by V(x)

Total energy 𝐸 = 𝑇 + 𝑉(𝑥) ̂


𝐻 −ℎ2 𝜕 2
+ 𝑉(𝑥)
8𝜋 2 𝑚 𝜕𝑥 2

Copyright © Mandeep Dalal


34 A Textbook of Physical Chemistry – Volume I

Besides the record of different operators presented in ‘Table 2’, there still many operators which are
extremely important like angular momentum, parity, or the step-up–step-down operators. The discussion of
every operator is beyond the scope of this book; however, a brief discussion of the essential operators in
quantum mechanics is given below.
1. Angular momentum operator: In order to understand the angular momentum operator in the quantum
mechanical world, we first need to understand the classical mechanics of one particle angular momentum. Let
us consider a particle of mass m which moves within a cartesian coordinate system with a position vector “r”.
Hence, we can say that

𝑟 = 𝑖𝑥 + 𝑗𝑦 + 𝑘𝑧 (97)

The coordinates x. y and z are the functions of time, and therefore, we can define the velocity as the time
derivative of the position vector as given below.

𝑑𝑟 𝑑𝑥 𝑑𝑦 𝑑𝑧 (98)
𝑣= =𝑖 +𝑗 +𝑘
𝑑𝑡 𝑑𝑡 𝑑𝑡 𝑑𝑡
or

𝑣 = 𝑣𝑥 + 𝑣𝑦 + 𝑣𝑧 (99)

Now, since we that p = mv, we can say that

𝑝𝑥 = 𝑚𝑣𝑥 ; 𝑝𝑦 = 𝑚𝑣𝑦 ; 𝑝𝑧 = 𝑚𝑣𝑧 (100)

The angular momentum of a particle with mass m and distance r from the origin is given by the following
relation.

Figure 6. The angular momentum vector.

𝐿⃗ = 𝑣 × 𝑚 × 𝑟 (101)

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 35

𝐿⃗ = 𝑝 × 𝑟 (102)

Equation (102) can also be written in the form of a matrix as:

𝑖 𝑗 𝑘 (103)
𝐿=[ 𝑥 𝑦 𝑧]
𝑝𝑥 𝑝𝑦 𝑝𝑧

𝐿𝑥 = 𝑦𝑝𝑧 − 𝑧𝑝𝑦 ; 𝐿𝑦 = 𝑧𝑝𝑥 − 𝑥𝑝𝑧 ; 𝐿𝑧 = 𝑥𝑝𝑦 − 𝑦𝑝𝑥 (104)

Where i, j, k are the unit vectors along x, y, z axis and Lx, Ly, Lz are the component of angular momentum along
x, y, z axis. Moreover, it is also worthy to note that the angular momentum vector is always perpendicular to
the direction of the position vector of the particle i.e. the plane in which the particle is moving.
Now since the mathematical nature of any quantum mechanical operator is dependent upon the
classical expression of the same observable, the angular momentum is not any exception. The quantum
mechanical operator for angular momentum is given below.

ℎ (105)
𝐿̂ = −𝑖 (𝑟 × ∇) = −𝑖ħ(𝑟 × ∇)
2𝜋
The angular momentum can be divided into two categories; one is orbital angular momentum (due to the orbital
motion of the particle) and the other is spin angular momentum (due to spin motion of the particle). Moreover,
being a vector quantity, the operator of angular momentum can also be resolved along different axes.

𝐿̂ = 𝐿̂𝑥 + 𝐿̂𝑦 + 𝐿̂𝑧 (106)

And we know that

ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 𝜕 (107)
𝐿̂𝑥 = 𝑦𝑝𝑧 − 𝑧𝑝𝑦 = 𝑦 ( )−𝑧( )= (𝑦 − 𝑧 )
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑧 𝜕𝑦

or

ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 𝜕 (108)
𝐿̂𝑦 = 𝑧𝑝𝑥 − 𝑥𝑝𝑧 = 𝑧 ( )−𝑥( )= (𝑧 −𝑥 )
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑥 𝜕𝑧
or

ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 𝜕 (109)
𝐿̂𝑧 = 𝑥𝑝𝑦 − 𝑦𝑝𝑥 = 𝑥 ( )−𝑦( )= (𝑥 −𝑦 )
2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑦 𝜕𝑥

ℎ 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (110)
𝐿̂ = [(𝑦 − 𝑧 ) + (𝑧 − 𝑥 ) + (𝑥 − 𝑦 )]
2𝜋𝑖 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑦 𝜕𝑥

It is also worthy to recall that equation (107) to (110) can also be reported in terms of ħ; or by multiplying and
dividing by i, or both.

Copyright © Mandeep Dalal


36 A Textbook of Physical Chemistry – Volume I

2. Ladder operator: These operators are also called as step-up–step-down or rising-lowering operators. The
reason for such terminology lies in the fact that these operators can increase or decrease the eigenvalues.
Moreover, it should also be noted that this increase or decrease is always quantized in nature.

𝐽̂+ = 𝐽̂𝑥 + 𝑖𝐽̂𝑦 (111)

and

𝐽̂+ = 𝐽̂𝑥 − 𝑖𝐽̂𝑦 (112)

The equation (111) and (112) represent the step-up and step-down operators respectively. These operators can
be used to increase or decrease the eigen values.
 Operator Evaluation
The operator evaluation simply means that we need to find the result by applying the operator over a
given function. Some general examples are given below.
i) (d/dx) (x5): In this case d/dx is the operator while the function x5 is the operand.

𝑑 5 (113)
𝑥 = 5𝑥 4
𝑑𝑥
ii) ʃ(x5): In this case, ʃ is the operator while the function x5 is the operand.

𝑥6 (114)
∫ 𝑥5 =
6

iii) (d2/dt2) (ASine 2πνt): In this particular case, (d2/dt2) is the operator while the function (A Sin 2πνt) is the
operand.
Let the function is symbolized by y. Then, we have

𝑦 = 𝐴 𝑆𝑖𝑛 2𝜋𝜈𝑡 (115)

Differentiating with respect to t, we get

𝑑𝑦 (116)
= 𝐴 2𝜋𝜈 𝐶𝑜𝑠 2𝜋𝜈𝑡
𝑑𝑡
Differentiating again

𝑑2 𝑦 (117)
= −𝐴 4𝜋 2 𝜈 2 𝑆𝑖𝑛 2𝜋𝜈𝑡
𝑑𝑡 2
The operator evaluation is frequently used as a part of the commutator calculation and will be
discussed in detail in this chapter.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 37

 Calculation of Resultant Operator


Sometimes the operator is simplified to another form which is easy to apply over a function. This
resultant operator is obtained by the rules of operator algebra. For instance, consider the following cases.
i) Find the resultant expression for the following operator

𝑑 2 (118)
( 𝑥)
𝑑𝑥
In order to find the resultant operator, suppose a function ψ(x) which is used as an operand, then we can say

𝑑 2
𝑑 𝑑 (119)
( 𝑥) 𝜓 = ( 𝑥) ( 𝑥) 𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥
or

𝑑 2
𝑑 𝑑 (120)
( 𝑥) 𝜓 = ( 𝑥) ( 𝑥𝜓)
𝑑𝑥 𝑑𝑥 𝑑𝑥
or

𝑑 2
𝑑 𝑑𝜓 𝑑𝑥 (121)
( 𝑥) 𝜓 = ( 𝑥) (𝑥 +𝜓 )
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑 𝑑𝜓 (122)
( 𝑥) 𝜓 = (𝑥 2 + 𝑥𝜓)
𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑2 𝜓 𝑑𝜓 𝑑𝜓 𝑑𝑥 (123)
( 𝑥) 𝜓 = [𝑥 2 2 + (2𝑥)] + [𝑥 +𝜓 ]
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑2 𝜓 𝑑𝜓 𝑑𝜓 (124)
2
( 𝑥) 𝜓 = 𝑥 + 2𝑥 + 𝑥 +𝜓
𝑑𝑥 𝑑𝑥 2 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑2 𝑑 (125)
( 𝑥) 𝜓 = [𝑥 2 2 + 3𝑥 + 1] 𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥

Removing ψ from both sides, we get

𝑑 2
𝑑2 𝑑 (126)
( 𝑥) = 𝑥 2 2 + 3𝑥 +1
𝑑𝑥 𝑑𝑥 𝑑𝑥
ii) Find the resultant expression for the following operator

𝑑 𝑑 (127)
(𝑥 + )
𝑑𝑥 𝑑𝑥
In order to find the resultant operator, suppose a function ψ(x) which is used as operand, then we can say that

Copyright © Mandeep Dalal


38 A Textbook of Physical Chemistry – Volume I

𝑑 𝑑 𝑑 𝑑𝜓 (128)
[(𝑥 + ) ] 𝜓 = (𝑥 + )
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 𝑑 𝑑𝜓 𝑑2 𝜓
[(𝑥 + ) ]𝜓 = 𝑥 +
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 2
Removing ψ from both sides, we get

𝑑 𝑑 𝑑 𝑑2 (129)
(𝑥 + ) =𝑥 + 2
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥
iii) Find the resultant expression for the following operator

𝑑 2 (130)
( + 𝑥)
𝑑𝑥
In order to find the resultant operator, suppose a function ψ(x) which is used as operand, then we can say that

𝑑 2
𝑑 𝑑 (131)
[( + 𝑥) ] 𝜓 = [( + 𝑥) ( + 𝑥)] 𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑 𝑑𝜓 (132)
[( + 𝑥) ] 𝜓 = ( + 𝑥) ( + 𝑥𝜓)
𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑2 𝜓 𝑑 𝑑𝜓 (133)
[( + 𝑥) ] 𝜓 = 2
+ 𝑥𝜓 + 𝑥 + 𝑥2𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑2 𝜓 𝑑𝜓 𝑑𝑥 𝑑𝜓 (134)
[( + 𝑥) ] 𝜓 = 2
+𝑥 +𝜓 +𝑥 + 𝑥2𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 2
𝑑2 𝜓 𝑑𝜓 (135)
[( + 𝑥) ] 𝜓 = 2
+ 2𝑥 + 𝑥2𝜓 + 𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥

Removing ψ from both sides, we get

𝑑 2
𝑑2 𝑑 (136)
( + 𝑥) = 2 + 2𝑥 + 𝑥2 + 1
𝑑𝑥 𝑑𝑥 𝑑𝑥
iv) Find the resultant expression for the following operator

𝑑 𝑑 (137)
(𝑥 + ) (𝑥 − )
𝑑𝑥 𝑑𝑥
In order to find the resultant operator, suppose a function ψ(x) which is used as operand, then we can say that

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 39

𝑑 𝑑 𝑑 𝑑𝜓 (138)
[(𝑥 + ) (𝑥 − )] 𝜓 = (𝑥 + ) (𝑥𝜓 − )
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 𝑑 𝑑𝜓 𝑑 𝑑2 𝜓 (139)
[(𝑥 + ) (𝑥 − )] 𝜓 = 𝑥𝑥𝜓 − 𝑥 + 𝑥𝜓 − 2
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

𝑑 𝑑 𝑑𝜓 𝑑𝜓 𝑑𝑥 𝑑2 𝜓 (140)
[(𝑥 + ) (𝑥 − )] 𝜓 = 𝑥 2 𝜓 − 𝑥 +𝑥 +𝜓 −
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 2

𝑑 𝑑 𝑑𝑥 𝑑2 𝜓 (141)
[(𝑥 + ) (𝑥 − )] 𝜓 = 𝑥 2 𝜓 + 𝜓 −
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 2

𝑑 𝑑 𝑑𝑥 𝑑2 (142)
[(𝑥 + ) (𝑥 − )] 𝜓 = [𝑥 2 + − 2] 𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥

Removing ψ from both sides, we get

𝑑 𝑑 𝑑2 (143)
(𝑥 + ) (𝑥 − ) = 𝑥 2 + 1 − 2
𝑑𝑥 𝑑𝑥 𝑑𝑥
The resultant operator calculation is frequently used as a part of the commutator calculation and will
be discussed in detail in this chapter.
 Commutation Relations of Various Quantum Mechanical Operators
As we have discussed previously that one of the most fundamental properties of operator
multiplication is the commutation relation or the commutation rule. two operators, A and B, are said to be
commutating or non-commutating depending upon the value of their commutator.

̂ 𝐵̂] = 𝐴̂𝐵̂ − 𝐵̂𝐴̂ = 0 → 𝐶𝑜𝑚𝑚𝑢𝑡𝑎𝑡𝑖𝑛𝑔


[𝐴, (144)

̂ 𝐵̂] = 𝐴̂𝐵̂ − 𝐵̂𝐴̂ ≠ 0 → 𝑁𝑜𝑛-𝑐𝑜𝑚𝑚𝑢𝑡𝑎𝑡𝑖𝑛𝑔


[𝐴, (145)

The physical significance of the commutation relations is that when two operators commute, it means they are
having a simultaneous set of eigenfunctions; and their corresponding physical properties can be calculated
simultaneously and accurately. However, if the commutator is non-zero, the respective physical properties
cannot be obtained simultaneously and accurately. Some important commutation relations are given below.
1. Commutators of some simple operators:
i) Calculate the commutator of the following

𝑑 (146)
[𝑥, ]
𝑑𝑥
Let it be operated over a function ψ. We have

Copyright © Mandeep Dalal


40 A Textbook of Physical Chemistry – Volume I

𝑑 𝑑 𝑑 (147)
[𝑥, ]𝜓 = 𝑥 𝜓− 𝑥𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑 𝑑𝜓 𝑑𝜓 (148)
[𝑥, ]𝜓 = 𝑥 −𝜓−𝑥
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑 (149)
[𝑥, ] 𝜓 = −𝜓
𝑑𝑥
or

𝑑 (150)
[𝑥, ] = −1
𝑑𝑥
ii) Calculate the commutator of the following

𝑑 (151)
[𝑦, ]
𝑑𝑥
Let it be operated over a function ψ. We have

𝑑 𝑑 𝑑 (152)
[𝑦, ]𝜓 = 𝑦 𝜓− 𝑦𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑 𝑑𝜓 𝑑𝜓 𝑑𝑦 (153)
[𝑦, ]𝜓 = 𝑦 −𝑦 −𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑 (154)
[𝑥, ]𝜓 = 0
𝑑𝑥
iii) Calculate the commutator of the following

𝑑 𝑑2 (155)
[ , 2]
𝑑𝑥 𝑑𝑥

Let it be operated over a function ψ. We have

𝑑 𝑑2 𝑑 𝑑2 𝑑2 𝑑 (156)
[ , 2 ]𝜓 = 𝜓 − 𝜓
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥 2 𝑑𝑥 2 𝑑𝑥

or

𝑑 𝑑2 𝑑3 𝜓 𝑑3 𝜓 (157)
[ , 2 ]𝜓 = −
𝑑𝑥 𝑑𝑥 𝑑𝑥 3 𝑑𝑥 3

𝑑 𝑑2 (158)
[ , 2 ]𝜓 = 0
𝑑𝑥 𝑑𝑥

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 41

2. Commutators of position and linear momentum operators:


i) Find the commutator of the following

[𝑥̂, 𝑝̂𝑥 ] (159)

Let it be operated over a function ψ. We have

[𝑥̂, 𝑝̂𝑥 ]𝜓 = 𝑥̂ 𝑝̂𝑥 𝜓 − 𝑝̂𝑥 ̂𝑥 𝜓 (160)

ℎ 𝜕 ℎ 𝜕 (161)
[𝑥̂, 𝑝̂𝑥 ]𝜓 = 𝑥 𝜓− 𝑥𝜓
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥
ℎ 𝜕𝜓 ℎ 𝜕𝜓 ℎ 𝜕𝑥 (162)
[𝑥̂, 𝑝̂𝑥 ]𝜓 = 𝑥 − 𝑥 − 𝜓
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥

[𝑥̂, 𝑝̂𝑥 ]𝜓 = − 𝜓
2𝜋𝑖
ℎ ℎ𝑖 (163)
[𝑥̂, 𝑝̂𝑥 ] = − = = 𝑖ħ
2𝜋𝑖 2𝜋
ii) Find the commutator of the following

[𝑥̂ 𝑛 , 𝑝̂𝑥 ] (164)

Let it be operated over a function ψ. We have

[𝑥̂ 𝑛 , 𝑝̂𝑥 ]𝜓 = 𝑥̂ 𝑛 𝑝̂𝑥 𝜓 − 𝑝̂𝑥 𝑥̂ 𝑛 𝜓 (165)

ℎ 𝜕 ℎ 𝜕 𝑛 (166)
[𝑥̂ 𝑛 , 𝑝̂𝑥 ]𝜓 = 𝑥 𝑛 𝜓− 𝑥 𝜓
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥
ℎ 𝑛 𝜕𝜓 ℎ 𝑛 𝜕𝜓 ℎ (167)
[𝑥̂ 𝑛 , 𝑝̂𝑥 ]𝜓 = 𝑥 − 𝑥 − 𝑛𝑥 𝑛−1 𝜓
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥 2𝜋𝑖
ℎ (168)
[𝑥̂ 𝑛 , 𝑝̂𝑥 ]𝜓 = − 𝑛𝑥 𝑛−1 𝜓
2𝜋𝑖
Removing ψ from both sides, we get

ℎ (169)
[𝑥̂ 𝑛 , 𝑝̂𝑥 ] = − 𝑛𝑥 𝑛−1
2𝜋𝑖
The commutation relations between position and linear momentum can mainly be divided into three
categories as discussed below.
(a) When position and momentum are along the same axis:

[𝑥̂ 𝑛 , 𝑝̂𝑥 ] = 𝑛𝑖ħ𝑥 𝑛−1 (170)

Copyright © Mandeep Dalal


42 A Textbook of Physical Chemistry – Volume I

[ 𝑝̂𝑥 , 𝑥̂ 𝑛 ] = −𝑛𝑖ħ𝑥 𝑛−1 (171)

and

[𝑥̂, 𝑝̂𝑥𝑛 ] = 𝑛𝑖ħ𝑝𝑥𝑛−1 (172)

[𝑝̂𝑥𝑛 , 𝑥̂ ] = −𝑛𝑖ħ𝑝𝑥𝑛−1 (173)

(b) When position and momentum are along different axis:

[𝑥̂, 𝑝̂𝑦 ] = 0 (174)

[𝑥̂, 𝑝̂𝑧 ] = 0 (175)

[𝑦̂, 𝑝̂𝑥 ] = 0 (176)

[𝑦̂, 𝑝̂𝑧 ] = 0 (177)

[𝑧̂ , 𝑝̂𝑥 ] = 0 (178)

[𝑧̂ , 𝑝̂𝑦 ] = 0 (179)

(b) When positions are along the different axis:

[𝑥̂, 𝑦̂] = 0 (180)

[𝑥̂, 𝑧̂ ] = 0 (181)

[𝑦̂, 𝑧̂ ] = 0 (182)

(b) When positions are along the different axis:

[𝑝̂𝑥 , 𝑝̂𝑦 ] = 0 (183)

[𝑝̂𝑥 , 𝑝̂𝑧 ] = 0 (184)

[𝑝̂𝑦 , 𝑝̂𝑧 ] = 0 (185)

3. Commutators of angular momentum operators:


i) The commutator of orbital angular momentum operators along x and y-axis.

[𝐿̂𝑥 , 𝐿̂𝑦 ] = 𝐿̂𝑥 𝐿̂𝑦 − 𝐿̂𝑦 𝐿̂𝑥 (186)

Finding the values of 𝐿̂𝑥 𝐿̂𝑦 , we get

ℎ 𝜕 𝜕 ℎ 𝜕 𝜕 (187)
𝐿̂𝑥 𝐿̂𝑦 = [ (𝑦 − 𝑧 )] [ (𝑧 − 𝑥 )]
2𝜋𝑖 𝜕𝑧 𝜕𝑦 2𝜋𝑖 𝜕𝑥 𝜕𝑧

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 43

ℎ2 𝜕 𝜕 𝜕 𝜕 (188)
= − 2 [(𝑦 − 𝑧 ) (𝑧 − 𝑥 )]
4𝜋 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑧

ℎ2 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (189)
= − 2 (𝑦 𝑧 −𝑧 𝑧 −𝑦 𝑥 +𝑧 𝑥 )
4𝜋 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑧 𝜕𝑦 𝜕𝑧

ℎ2 𝜕 𝜕𝑧 𝜕2 2
𝜕2 𝜕2 𝜕2 (190)
= − 2 (𝑦 + 𝑦𝑧 2 − 𝑧 2 − 𝑦𝑥 2 + 𝑧𝑥 )
4𝜋 𝜕𝑥 𝜕𝑧 𝜕 𝑧𝑥 𝜕 𝑦𝑥 𝜕𝑧 𝜕𝑦𝑧

𝜕 𝜕2 𝜕2 𝜕2 𝜕2 (191)
= −ħ2 (𝑦 + 𝑦𝑧 2 − 𝑧 2 2 − 𝑦𝑥 2 + 𝑧𝑥 )
𝜕𝑥 𝜕 𝑧𝑥 𝜕 𝑦𝑥 𝜕𝑧 𝜕𝑦𝑧

Similarly obtaining the value of 𝐿̂𝑦 𝐿̂𝑥 , we get

ℎ 𝜕 𝜕 ℎ 𝜕 𝜕 (192)
𝐿̂𝑦 𝐿̂𝑥 = [ (𝑧 − 𝑥 )] [ (𝑦 − 𝑧 )]
2𝜋𝑖 𝜕𝑥 𝜕𝑧 2𝜋𝑖 𝜕𝑧 𝜕𝑦

ℎ2 𝜕 𝜕 𝜕 𝜕 (193)
=− 2 [(𝑧 − 𝑥 ) (𝑦 − 𝑧 )]
4𝜋 𝜕𝑥 𝜕𝑧 𝜕𝑧 𝜕𝑦

ℎ2 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (194)
=− 2
(𝑧 𝑦 − 𝑧 𝑧 −𝑥 𝑦 +𝑥 𝑧 )
4𝜋 𝜕𝑥 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑧 𝜕𝑧 𝜕𝑦

ℎ2 𝜕2 2
𝜕2 𝜕2 𝜕2 𝜕 𝜕𝑧 (195)
=− 2 (𝑧𝑦 2
− 𝑧 2
− 𝑥𝑦 2
+ 𝑥𝑧 +𝑥 )
4𝜋 𝜕 𝑥𝑧 𝜕 𝑥𝑦 𝜕𝑧 𝜕𝑧𝑦 𝜕𝑦 𝜕𝑧

2
𝜕2 2
𝜕2 𝜕2 𝜕2 𝜕 (196)
= −ħ (𝑧𝑦 2 − 𝑧 2 − 𝑥𝑦 2 + 𝑥𝑧 +𝑥 )
𝜕 𝑥𝑧 𝜕 𝑥𝑦 𝜕𝑧 𝜕𝑧𝑦 𝜕𝑦

Now putting the values of 𝐿̂𝑥 𝐿̂𝑦 and 𝐿̂𝑦 𝐿̂𝑥 in equation (183), we get the following.

𝜕 𝜕2 𝜕2 𝜕2 𝜕2 (197)
[𝐿̂𝑥 , 𝐿̂𝑦 ] = [−ħ2 (𝑦 + 𝑦𝑧 2 − 𝑧 2 2 − 𝑦𝑥 2 + 𝑧𝑥 )]
𝜕𝑥 𝜕 𝑧𝑥 𝜕 𝑦𝑥 𝜕𝑧 𝜕𝑦𝑧
𝜕2 𝜕2 𝜕2 𝜕2 𝜕
− [−ħ2 (𝑧𝑦 2
− 𝑧 2
2
− 𝑥𝑦 2
+ 𝑥𝑧 + 𝑥 )]
𝜕 𝑥𝑧 𝜕 𝑥𝑦 𝜕𝑧 𝜕𝑧𝑦 𝜕𝑦

𝜕 𝜕 (198)
[𝐿̂𝑥 , 𝐿̂𝑦 ] = −ħ2 (𝑦 −𝑥 )
𝜕𝑥 𝜕𝑦

Taking negative sign common, we get

𝜕 𝜕 (199)
[𝐿̂𝑥 , 𝐿̂𝑦 ] = ħ2 (𝑥 −𝑦 )
𝜕𝑦 𝜕𝑥

Copyright © Mandeep Dalal


44 A Textbook of Physical Chemistry – Volume I

𝜕 𝜕 (200)
[𝐿̂𝑥 , 𝐿̂𝑦 ] = 𝑖ħ [−𝑖ħ (𝑥 − 𝑦 )]
𝜕𝑦 𝜕𝑥

[𝐿̂𝑥 , 𝐿̂𝑦 ] = 𝑖ħ𝐿̂𝑧 (201)

ii) The commutator of orbital angular momentum operators along y and z-axis.

[𝐿̂𝑦 , 𝐿̂𝑧 ] = 𝐿̂𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑦 (202)

Finding the values of 𝐿̂𝑦 𝐿̂𝑧 , we get

ℎ 𝜕 𝜕 ℎ 𝜕 𝜕 (203)
𝐿̂𝑦 𝐿̂𝑧 = [ (𝑧 − 𝑥 )] [ (𝑥 − 𝑦 )]
2𝜋𝑖 𝜕𝑥 𝜕𝑧 2𝜋𝑖 𝜕𝑦 𝜕𝑥

ℎ2 𝜕 𝜕 𝜕 𝜕 (204)
= − 2 [(𝑧 − 𝑥 ) (𝑥 − 𝑦 )]
4𝜋 𝜕𝑥 𝜕𝑧 𝜕𝑦 𝜕𝑥

ℎ2 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (205)
= − 2 (𝑧 𝑥 −𝑥 𝑥 −𝑧 𝑦 +𝑥 𝑦 )
4𝜋 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑥 𝜕𝑧 𝜕𝑥

ℎ2 𝜕 𝜕𝑥 𝜕2 2
𝜕2 𝜕 𝜕2 (206)
= − 2 (𝑧 + 𝑧𝑥 − 𝑥 2 − 𝑧𝑦 2 + 𝑥𝑦 )
4𝜋 𝜕𝑦 𝜕𝑥 𝜕𝑥𝑦 𝜕 𝑧𝑦 𝜕𝑥 𝜕𝑧𝑥

𝜕 𝜕2 𝜕2 𝜕 𝜕2 (207)
= −ħ2 (𝑧 + 𝑧𝑥 − 𝑥 2 2 − 𝑧𝑦 2 + 𝑥𝑦 )
𝜕𝑦 𝜕𝑥𝑦 𝜕 𝑧𝑦 𝜕𝑥 𝜕𝑧𝑥

Similarly obtaining the value of 𝐿̂𝑧 𝐿̂𝑦 , we get

ℎ 𝜕 𝜕 ℎ 𝜕 𝜕 (208)
𝐿̂𝑧 𝐿̂𝑦 = [ (𝑥 − 𝑦 )] [ (𝑧 − 𝑥 )]
2𝜋𝑖 𝜕𝑦 𝜕𝑥 2𝜋𝑖 𝜕𝑥 𝜕𝑧

ℎ2 𝜕 𝜕 𝜕 𝜕 (209)
=− 2 [(𝑥 − 𝑦 ) (𝑧 − 𝑥 )]
4𝜋 𝜕𝑦 𝜕𝑥 𝜕𝑥 𝜕𝑧

ℎ2 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (210)
=− 2
(𝑥 𝑧 −𝑥 𝑥 −𝑦 𝑧 +𝑦 𝑥 )
4𝜋 𝜕𝑦 𝜕𝑥 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑧

ℎ2 𝜕2 2
𝜕2 𝜕2 𝜕2 𝜕 𝜕𝑥 (211)
=− 2 (𝑥𝑧 − 𝑥 − 𝑦𝑧 2
+ 𝑦𝑥 +𝑦 )
4𝜋 𝜕𝑦𝑥 𝜕𝑦𝑧 𝜕𝑥 𝜕𝑥𝑧 𝜕𝑧 𝜕𝑥

2
𝜕2 2
𝜕2 𝜕2 𝜕2 𝜕 (212)
= −ħ (𝑥𝑧 −𝑥 − 𝑦𝑧 2 + 𝑦𝑥 +𝑦 )
𝜕𝑦𝑥 𝜕𝑦𝑧 𝜕𝑥 𝜕𝑥𝑧 𝜕𝑧

Now putting the values of 𝐿̂𝑦 𝐿̂𝑧 and 𝐿̂𝑧 𝐿̂𝑦 in equation (212), we get the following.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 45

𝜕 𝜕2 𝜕2 𝜕 𝜕2 (213)
[𝐿̂𝑦 , 𝐿̂𝑧 ] = [−ħ2 (𝑧 + 𝑧𝑥 2
− 𝑥 2 − 𝑧𝑦 2 + 𝑥𝑦 )]
𝜕𝑦 𝜕𝑥𝑦 𝜕 𝑧𝑦 𝜕𝑥 𝜕𝑧𝑥
𝜕2 𝜕2 𝜕2 𝜕2 𝜕
− [−ħ2 (𝑥𝑧 − 𝑥2 − 𝑦𝑧 2 + 𝑦𝑥 + 𝑦 )]
𝜕𝑦𝑥 𝜕𝑦𝑧 𝜕𝑥 𝜕𝑥𝑧 𝜕𝑧

𝜕 𝜕 (214)
[𝐿̂𝑦 , 𝐿̂𝑧 ] = −ħ2 (𝑧 −𝑦 )
𝜕𝑦 𝜕𝑧

Taking negative sign common, we get

𝜕 𝜕 (215)
[𝐿̂𝑦 , 𝐿̂𝑧 ] = ħ2 (𝑦 −𝑧 )
𝜕𝑧 𝜕𝑦

𝜕 𝜕 (216)
[𝐿̂𝑦 , 𝐿̂𝑧 ] = 𝑖ħ [−𝑖ħ (𝑦 − 𝑧 )]
𝜕𝑧 𝜕𝑦

[𝐿̂𝑦 , 𝐿̂𝑧 ] = 𝑖ħ𝐿̂𝑥 (217)

iii) The commutator of orbital angular momentum operators along z and x-axis.

[𝐿̂𝑧 , 𝐿̂𝑥 ] = 𝐿̂𝑧 𝐿̂𝑥 − 𝐿̂𝑥 𝐿̂𝑧 (218)

Finding the values of 𝐿̂𝑧 𝐿̂𝑥 , we get

ℎ 𝜕 𝜕 ℎ 𝜕 𝜕 (219)
𝐿̂𝑧 𝐿̂𝑥 = [ (𝑥 − 𝑦 )] [ (𝑦 − 𝑧 )]
2𝜋𝑖 𝜕𝑦 𝜕𝑥 2𝜋𝑖 𝜕𝑧 𝜕𝑦

ℎ2 𝜕 𝜕 𝜕 𝜕 (220)
=− 2 [(𝑥 − 𝑦 ) (𝑦 − 𝑧 )]
4𝜋 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑦

ℎ2 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (221)
=− 2
(𝑥 𝑦 − 𝑥 𝑧 −𝑦 𝑦 +𝑦 𝑧 )
4𝜋 𝜕𝑦 𝜕𝑧 𝜕𝑦 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑥 𝜕𝑦

ℎ2 𝜕 𝜕𝑦 𝜕2 𝜕2 2
𝜕2 𝜕2 (222)
=− (𝑥 + 𝑥𝑦 − 𝑥𝑧 − 𝑦 + 𝑦𝑧 )
4𝜋 2 𝜕𝑧 𝜕𝑦 𝜕𝑦𝑧 𝜕𝑦 2 𝜕𝑥𝑧 𝜕𝑥𝑦

𝜕 𝜕2 𝜕2 𝜕2 𝜕2 (223)
= −ħ2 (𝑥 + 𝑥𝑦 − 𝑥𝑧 2 − 𝑦 2 + 𝑦𝑧 )
𝜕𝑧 𝜕𝑦𝑧 𝜕𝑦 𝜕𝑥𝑧 𝜕𝑥𝑦

Similarly obtaining the value of 𝐿̂𝑥 𝐿̂𝑧 , we get

ℎ 𝜕 𝜕 ℎ 𝜕 𝜕 (224)
𝐿̂𝑥 𝐿̂𝑧 = [ (𝑦 − 𝑧 )] [ (𝑥 − 𝑦 )]
2𝜋𝑖 𝜕𝑧 𝜕𝑦 2𝜋𝑖 𝜕𝑦 𝜕𝑥

Copyright © Mandeep Dalal


46 A Textbook of Physical Chemistry – Volume I

ℎ2 𝜕 𝜕 𝜕 𝜕 (225)
= − 2 [(𝑦 − 𝑧 ) (𝑥 − 𝑦 )]
4𝜋 𝜕𝑧 𝜕𝑦 𝜕𝑦 𝜕𝑥

or

ℎ2 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (226)
= − 2 (𝑦 𝑥 −𝑦 𝑦 −𝑧 𝑥 +𝑧 𝑦 )
4𝜋 𝜕𝑧 𝜕𝑦 𝜕𝑧 𝜕𝑥 𝜕𝑦 𝜕𝑦 𝜕𝑦 𝜕𝑥

ℎ2 𝜕2 2
𝜕2 𝜕2 𝜕 𝜕𝑦 𝜕2 (227)
= − 2 (𝑦𝑥 −𝑦 − 𝑧𝑥 2 + 𝑧 + 𝑧𝑦 )
4𝜋 𝜕𝑧𝑦 𝜕𝑧𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑦 𝜕𝑦𝑥

𝜕2 𝜕2 𝜕2 𝜕 𝜕2 (228)
= −ħ2 (𝑦𝑥 − 𝑦2 − 𝑧𝑥 2 + 𝑧 + 𝑧𝑦 )
𝜕𝑧𝑦 𝜕𝑧𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑦𝑥

Now putting the values of 𝐿̂𝑧 𝐿̂𝑥 and 𝐿̂𝑥 𝐿̂𝑧 in equation (218), we get the following.

𝜕 𝜕2 𝜕2 𝜕2 𝜕2 (229)
[𝐿̂𝑧 , 𝐿̂𝑥 ] = [−ħ2 (𝑥 + 𝑥𝑦 − 𝑥𝑧 2 − 𝑦 2 + 𝑦𝑧 )]
𝜕𝑧 𝜕𝑦𝑧 𝜕𝑦 𝜕𝑥𝑧 𝜕𝑥𝑦
2
𝜕2 2
𝜕2 𝜕2 𝜕 𝜕2
− [−ħ (𝑦𝑥 −𝑦 − 𝑧𝑥 2 + 𝑧 + 𝑧𝑦 )]
𝜕𝑧𝑦 𝜕𝑧𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑦𝑥

𝜕 𝜕 (230)
[𝐿̂𝑧 , 𝐿̂𝑥 ] = −ħ2 (𝑥 −𝑧 )
𝜕𝑧 𝜕𝑥
Taking negative sign common, we get

𝜕 𝜕 (231)
[𝐿̂𝑧 , 𝐿̂𝑥 ] = ħ2 (𝑦 −𝑧 )
𝜕𝑧 𝜕𝑦

𝜕 𝜕 (232)
[𝐿̂𝑧 , 𝐿̂𝑥 ] = 𝑖ħ [−𝑖ħ (𝑦 − 𝑧 )]
𝜕𝑧 𝜕𝑦

[𝐿̂𝑧 , 𝐿̂𝑥 ] = 𝑖ħ𝐿̂𝑦 (233)

iv) The commutator of total orbital angular momentum squared operator and orbital angular momentum along
one of the three-axis.

[𝐿̂2 , 𝐿̂𝑧 ] = [𝐿̂2𝑥 + 𝐿̂2𝑦 + 𝐿̂2𝑧 , 𝐿̂𝑧 ] (234)

= [𝐿̂2𝑥 𝐿̂𝑧 + 𝐿̂2𝑦 𝐿̂𝑧 + 𝐿̂2𝑧 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑥 − 𝐿̂𝑧 𝐿̂2𝑦 − 𝐿̂𝑧 𝐿̂2𝑧 ] (235)

= [(𝐿̂2𝑥 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑥 ) + (𝐿̂2𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑦 ) + (𝐿̂2𝑧 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑧 )] (236)

[𝐿̂2 , 𝐿̂𝑧 ] = [𝐿̂2𝑥 , 𝐿̂𝑧 ] + [𝐿̂2𝑦 𝐿̂𝑧 ] + [𝐿̂2𝑧 𝐿̂𝑧 ] (237)

Now finding [𝐿̂2𝑥 , 𝐿̂𝑧 ] first, we get

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 47

[𝐿̂2𝑥 , 𝐿̂𝑧 ] = 𝐿̂2𝑥 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑥 (238)

= 𝐿̂𝑥 𝐿̂𝑥 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑥 𝐿̂𝑥 (239)

= [𝐿̂𝑥 𝐿̂𝑥 𝐿̂𝑧 − 𝐿̂𝑥 𝐿̂𝑧 𝐿̂𝑥 ] − [𝐿̂𝑧 𝐿̂𝑥 𝐿̂𝑥 − 𝐿̂𝑥 𝐿̂𝑧 𝐿̂𝑥 ] (240)

= 𝐿̂𝑥 [𝐿̂𝑥 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑥 ] − [𝐿̂𝑧 𝐿̂𝑥 − 𝐿̂𝑥 𝐿̂𝑧 ]𝐿̂𝑥 (241)

= 𝐿̂𝑥 [𝐿̂𝑥 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑥 ] + [𝐿̂𝑥 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑥 ]𝐿̂𝑥 (242)

= 𝐿̂𝑥 [−𝑖ħ𝐿̂𝑦 ] + [−𝑖ħ𝐿̂𝑦 ]𝐿̂𝑥 (243)

= −𝑖ħ𝐿̂𝑥 𝐿̂𝑦 − 𝑖ħ𝐿̂𝑦 𝐿̂𝑥 = −𝑖ħ[𝐿̂𝑥 𝐿̂𝑦 + 𝐿̂𝑦 𝐿̂𝑥 ] (244)

Similarly,

[𝐿̂2𝑦 , 𝐿̂𝑧 ] = 𝐿̂2𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑦 (245)

= 𝐿̂𝑦 𝐿̂𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑦 𝐿̂𝑦 (246)

= [𝐿̂𝑦 𝐿̂𝑦 𝐿̂𝑧 − 𝐿̂𝑦 𝐿̂𝑧 𝐿̂𝑦 ] − [𝐿̂𝑧 𝐿̂𝑦 𝐿̂𝑦 − 𝐿̂𝑦 𝐿̂𝑧 𝐿̂𝑦 ]

= 𝐿̂𝑦 [𝐿̂𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑦 ] − [𝐿̂𝑧 𝐿̂𝑦 − 𝐿̂𝑦 𝐿̂𝑧 ]𝐿̂𝑦

= 𝐿̂𝑦 [𝐿̂𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑦 ] + [𝐿̂𝑦 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑦 ]𝐿̂𝑦

= 𝐿̂𝑦 [𝑖ħ𝐿̂𝑥 ] + [𝑖ħ𝐿̂𝑥 ]𝐿̂𝑦

= 𝑖ħ𝐿̂𝑦 𝐿̂𝑥 + 𝑖ħ𝐿̂𝑥 𝐿̂𝑦 = 𝑖ħ[𝐿̂𝑦 𝐿̂𝑥 + 𝐿̂𝑥 𝐿̂𝑦 ] (247)

Similarly,

[𝐿̂2𝑧 , 𝐿̂𝑧 ] = 𝐿̂2𝑧 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂2𝑧

= 𝐿̂𝑧 𝐿̂𝑧 𝐿̂𝑧 − 𝐿̂𝑧 𝐿̂𝑧 𝐿̂𝑧

[𝐿̂2𝑧 , 𝐿̂𝑧 ] = 0 (248)

Now putting the value of 𝐿̂2𝑥 𝐿̂𝑧 , 𝐿̂2𝑦 𝐿̂𝑧 and 𝐿̂2𝑧 𝐿̂𝑧 in equation (237), we get

[𝐿̂2 , 𝐿̂𝑧 ] = −𝑖ħ[𝐿̂𝑥 𝐿̂𝑦 + 𝐿̂𝑦 𝐿̂𝑥 ] + 𝑖ħ[𝐿̂𝑦 𝐿̂𝑥 + 𝐿̂𝑥 𝐿̂𝑦 ] + 0

[𝐿̂2 , 𝐿̂𝑧 ] = 0 (249)

Also

Copyright © Mandeep Dalal


48 A Textbook of Physical Chemistry – Volume I

[𝐿̂2 , 𝐿̂𝑦 ] = 0; 𝑎𝑛𝑑 [𝐿̂2 , 𝐿̂𝑥 ] = 0 (250)

Hence, the commutation relations of angular momentum operators along two different directions do not
commute with each other and hence cannot give eigenvalues simultaneously and accurately. One the other
hand, total angular momentum squared and angular momentum along one axis do commute with each other.
The commutation relations between angular momentum operators can be mainly divided into four
categories as discussed below.
(a) Orbital angular momentum commutation:

[𝐿̂𝑥 , 𝐿̂𝑦 ] = 𝑖ħ𝐿̂𝑧 ; [𝐿̂𝑦 , 𝐿̂𝑥 ] = −𝑖ħ𝐿̂𝑧 (251)

[𝐿̂𝑦 , 𝐿̂𝑧 ] = 𝑖ħ𝐿̂𝑥 ; [𝐿̂𝑧 , 𝐿̂𝑦 ] = −𝑖ħ𝐿̂𝑥 (252)

[𝐿̂𝑧 , 𝐿̂𝑥 ] = 𝑖ħ𝐿̂𝑦 ; [𝐿̂𝑥 , 𝐿̂𝑧 ] = −𝑖ħ𝐿̂𝑦 (253)

[𝐿̂2 , 𝐿̂𝑥 ] = 0; [𝐿̂𝑥 , 𝐿̂2 ] = 0 (254)

[𝐿̂2 , 𝐿̂𝑦 ] = 0; [𝐿̂𝑦 , 𝐿̂2 ] = 0 (255)

[𝐿̂2 , 𝐿̂𝑧 ] = 0; [𝐿̂𝑧 , 𝐿̂2 ] = 0 (256)

(b) Spin angular momentum commutation:

[𝑆̂𝑥 , 𝑆̂𝑦 ] = 𝑖ħ𝑆̂𝑧 ; [𝑆̂𝑦 , 𝑆̂𝑥 ] = −𝑖ħ𝑆̂𝑧 (257)

[𝑆̂𝑦 , 𝑆̂𝑧 ] = 𝑖ħ𝑆̂𝑥 ; [𝑆̂𝑧 , 𝑆̂𝑦 ] = −𝑖ħ𝑆̂𝑥 (258)

[𝑆̂𝑧 , 𝑆̂𝑥 ] = 𝑖ħ𝑆̂𝑦 ; [𝑆̂𝑥 , 𝑆̂𝑧 ] = −𝑖ħ𝑆̂𝑦 (259)

[𝑆̂ 2 , 𝑆̂𝑥 ] = 0; [𝑆̂𝑥 , 𝑆̂ 2 ] = 0 (260)

[𝑆̂ 2 , 𝑆𝑦 ] = 0; [𝑆̂𝑦 , 𝑆̂ 2 ] = 0 (261)

[𝑆̂ 2 , 𝑆̂𝑧 ] = 0; [𝑆̂𝑧 , 𝑆̂ 2 ] = 0 (262)

(c) Total angular momentum commutation:

[𝐽̂𝑥 , 𝐽̂𝑦 ] = 𝑖ħ𝐽̂𝑧 ; [𝐽̂𝑦 , 𝐽̂𝑥 ] = −𝑖ħ𝐽̂𝑧 (263)

[𝐽̂𝑦 , 𝐽̂𝑧 ] = 𝑖ħ𝐽̂𝑥 ; [𝐽̂𝑧 , 𝐽̂𝑦 ] = −𝑖ħ𝐽̂𝑥 (264)

[𝐽̂𝑧 , 𝐽̂𝑥 ] = 𝑖ħ𝐽̂𝑦 ; [𝐽̂𝑥 , 𝐽̂𝑧 ] = −𝑖ħ𝐽̂𝑦 (265)

[𝐽̂2 , 𝐽̂𝑥 ] = 0; [𝐽̂𝑥 , 𝐽̂2 ] = 0 (266)

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 49

[𝐽̂2 , 𝐽𝑦 ] = 0; [𝐽̂𝑦 , 𝐽̂2 ] = 0 (267)

[𝐽̂2 , 𝐽̂𝑧 ] = 0; [𝐽̂𝑧 , 𝐽̂2 ] = 0 (268)

(d) Total angular momentum commutation:

[𝐿̂𝑥 , 𝑆̂𝑥 ] = 0; [𝑆̂𝑥 , 𝐿̂𝑥 ] = 0 (263)

[𝐿̂𝑥 , 𝑆̂𝑦 ] = 0; [𝑆̂𝑦 , 𝐿̂𝑥 ] = 0 (264)

[𝐿̂𝑥 , 𝑆̂𝑧 ] = 0; [𝑆̂𝑧 , 𝐿̂𝑥 ] = 0 (265)

[𝐿̂𝑦 , 𝑆̂𝑥 ] = 0; [𝑆̂𝑥 , 𝐿̂𝑦 ] = 0 (266)

[𝐿̂𝑦 , 𝑆̂𝑦 ] = 0; [𝑆̂𝑦 , 𝐿̂𝑦 ] = 0 (267)

[𝐿̂𝑦 , 𝑆̂𝑧 ] = 0; [𝑆̂𝑧 , 𝐿̂𝑦 ] = 0 (268)

[𝐿̂𝑧 , 𝑆̂𝑥 ] = 0; [𝑆̂𝑥 , 𝐿̂𝑧 ] = 0 (269)

[𝐿̂𝑧 , 𝑆̂𝑦 ] = 0; [𝑆̂𝑦 , 𝐿̂𝑧 ] = 0 (270)

[𝐿̂𝑧 , 𝑆̂𝑧 ] = 0; [𝑆̂𝑧 , 𝐿̂𝑧 ] = 0 (271)

4. Commutators of Ladder operators:


i) Find the commutator of the following

[𝐽̂2 , 𝐽̂+ ] (272)

Let

[𝐽̂2 , 𝐽̂+ ] = [𝐽̂2 , 𝐽̂𝑥 + 𝑖𝐽̂𝑦 ] (273)

= 𝐽̂2 (𝐽̂𝑥 + 𝑖𝐽̂𝑦 ) − (𝐽̂𝑥 + 𝑖𝐽̂𝑦 )𝐽̂2 (274)

= 𝐽̂2 𝐽̂𝑥 + 𝑖𝐽̂2 𝐽̂𝑦 − 𝐽̂𝑥 𝐽̂2 − 𝑖𝐽̂𝑦 𝐽̂2 (275)

= [𝐽̂2 𝐽̂𝑥 − 𝐽̂𝑥 𝐽̂2 ] + 𝑖[𝐽̂2 𝐽̂𝑦 − 𝐽̂𝑦 𝐽̂2 ] (276)

= [𝐽̂2 , 𝐽̂𝑥 ] + 𝑖[𝐽̂2 , 𝐽̂𝑦 ] (277)

= 0 + 𝑖(0) = 0 (278)

Hence

[𝐽̂2 , 𝐽̂+ ] = 0 (279)

Copyright © Mandeep Dalal


50 A Textbook of Physical Chemistry – Volume I

Similarly

[𝐽̂2 , 𝐽̂− ] = 0 (280)

ii) Find the commutator of the following

[𝐽̂+ , 𝐽̂𝑧 ] (281)

Let

[𝐽̂+ , 𝐽̂𝑧 ] = [ 𝐽̂𝑥 + 𝑖𝐽̂𝑦 , 𝐽̂𝑧 ] (282)

= (𝐽̂𝑥 + 𝑖𝐽̂𝑦 )𝐽̂𝑧 − 𝐽̂𝑧 (𝐽̂𝑥 + 𝑖𝐽̂𝑦 ) (283)

= 𝐽̂𝑥 𝐽̂𝑧 + 𝑖𝐽̂𝑦 𝐽̂𝑧 − 𝐽̂𝑧 𝐽̂𝑥 − 𝐽̂𝑧 𝑖𝐽̂𝑦 (284)

= 𝐽̂𝑥 𝐽̂𝑧 − 𝐽̂𝑧 𝐽̂𝑥 + 𝑖𝐽̂𝑦 𝐽̂𝑧 − 𝑖𝐽̂𝑧 𝐽̂𝑦 (285)

= [𝐽̂𝑥 𝐽̂𝑧 − 𝐽̂𝑧 𝐽̂𝑥 ] + 𝑖[𝐽̂𝑦 𝐽̂𝑧 − 𝐽̂𝑧 𝐽̂𝑦 ] (286)

= [𝐽̂𝑥 , 𝐽̂𝑧 ] + 𝑖[𝐽̂𝑦 , 𝐽̂𝑧 ] (287)

= −𝑖ħ𝐽̂𝑦 + 𝑖(𝑖ħ𝐽̂𝑥 ) = −𝑖ħ𝐽̂𝑦 − ħ𝐽̂𝑥 (288)

= −ħ(𝐽̂𝑥 + 𝑖𝐽̂𝑦 ) = −ħ𝐽̂+ (289)

[𝐽̂+ , 𝐽̂𝑧 ] = −ħ𝐽̂+ (290)

Similarly

[𝐽̂− , 𝐽̂𝑧 ] = ħ𝐽̂− (291)

iii) Find the commutator of the following

[𝐽̂+ , 𝐽̂− ] (292)

Let

[𝐽̂+ , 𝐽̂− ] = (𝐽̂𝑥 + 𝑖𝐽̂𝑦 )(𝐽̂𝑥 − 𝑖𝐽̂𝑦 ) − (𝐽̂𝑥 − 𝑖𝐽̂𝑦 )(𝐽̂𝑥 + 𝑖𝐽̂𝑦 ) (293)

= 𝐽̂𝑥 𝐽̂𝑥 − 𝑖𝐽̂𝑥 𝐽̂𝑦 + 𝑖𝐽̂𝑦 𝐽̂𝑥 + 𝐽̂𝑦 𝐽̂𝑦 − (𝐽̂𝑥 𝐽̂𝑥 + 𝑖𝐽̂𝑥 𝐽̂𝑦 − 𝑖𝐽̂𝑦 𝐽̂𝑥 + 𝐽̂𝑦 𝐽̂𝑦 ) (294)

= 𝐽̂𝑥 𝐽̂𝑥 − 𝑖𝐽̂𝑥 𝐽̂𝑦 + 𝑖𝐽̂𝑦 𝐽̂𝑥 + 𝐽̂𝑦 𝐽̂𝑦 − 𝐽̂𝑥 𝐽̂𝑥 − 𝑖𝐽̂𝑥 𝐽̂𝑦 + 𝑖𝐽̂𝑦 𝐽̂𝑥 − 𝐽̂𝑦 𝐽̂𝑦 (295)

= −𝑖𝐽̂𝑥 𝐽̂𝑦 + 𝑖𝐽̂𝑦 𝐽̂𝑥 − 𝑖𝐽̂𝑥 𝐽̂𝑦 + 𝑖𝐽̂𝑦 𝐽̂𝑥 (296)

= −𝑖[𝐽̂𝑥 𝐽̂𝑦 − 𝐽̂𝑦 𝐽̂𝑥 ] + 𝑖[𝐽̂𝑦 𝐽̂𝑥 − 𝐽̂𝑥 𝐽̂𝑦 ] (297)

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 51

= −𝑖[𝐽̂𝑥 , 𝐽̂𝑦 ] + 𝑖[𝐽̂𝑦 , 𝐽̂𝑥 ] (298)

= −𝑖[𝑖ħ𝐽̂𝑧 ] + 𝑖[−𝑖ħ𝐽̂𝑧 ] (299)

= ħ𝐽̂𝑧 + ħ𝐽̂𝑧 = 2ħ𝐽̂𝑧 (300)

The commutation relations between angular-momentum and Ladder operators can be mainly divided
into three categories as discussed below.
(a) Ladder operator and total angular momentum commutation:

[𝐽̂2 , 𝐽̂+ ] = 0; [ 𝐽̂+ , 𝐽̂2 ] = 0 (301)

[𝐽̂2 , 𝐽̂− ] = 0; [ 𝐽̂− , 𝐽̂2 ] = 0 (302)

[𝐽̂+ , 𝐽̂𝑧 ] = −ħ𝐽̂+ ; [ 𝐽̂𝑧 , 𝐽̂+ ] = ħ𝐽̂+ (303)

[𝐽̂− , 𝐽̂𝑧 ] = ħ𝐽̂− ; [ 𝐽̂𝑧 , 𝐽̂− ] = −ħ𝐽̂− (304)

[𝐽̂+ , 𝐽̂− ] = 2ħ𝐽̂𝑧 ; [ 𝐽̂− , 𝐽̂+ ] = −2ħ𝐽̂𝑧 (305)

(b) Ladder operator and orbital angular momentum commutation:

[𝐿̂2 , 𝐿̂+ ] = 0; [ 𝐿̂+ , 𝐿̂2 ] = 0 (306)

[𝐿̂2 , 𝐿̂− ] = 0; [ 𝐿̂− , 𝐿̂2 ] = 0 (307)

[𝐿̂+ , 𝐿̂𝑧 ] = −ħ𝐿̂+ ; [ 𝐿̂𝑧 , 𝐿̂+ ] = ħ𝐿̂+ (308)

[𝐿̂− , 𝐿̂𝑧 ] = ħ𝐿̂− ; [ 𝐿̂𝑧 , 𝐿̂− ] = −ħ𝐿̂− (309)

[𝐿̂+ , 𝐿̂− ] = 2ħ𝐿̂𝑧 ; [ 𝐿̂− , 𝐿̂+ ] = −2ħ𝐿̂𝑧 (310)

(b) Ladder operator and spin angular momentum commutation:

[𝑆̂ 2 , 𝑆̂+ ] = 0; [ 𝑆̂+ , 𝑆̂ 2 ] = 0 (311)

[𝑆̂ 2 , 𝑆̂− ] = 0; [ 𝑆̂− , 𝑆̂ 2 ] = 0 (312)

[𝑆̂+ , 𝑆̂𝑧 ] = −ħ𝑆̂+ ; [ 𝑆̂𝑧 , 𝑆̂+ ] = ħ𝑆̂+ (313)

[𝑆̂− , 𝑆̂𝑧 ] = ħ𝑆̂− ; [ 𝑆̂𝑧 , 𝑆̂− ] = −ħ𝑆̂− (314)

[𝑆̂+ , 𝑆̂− ] = 2ħ𝑆̂𝑧 ; [ 𝑆̂− , 𝑆̂+ ] = −2ħ𝑆̂𝑧 (315)

Copyright © Mandeep Dalal


52 A Textbook of Physical Chemistry – Volume I

 Hermitian Operators – Elementary Ideas, Quantum Mechanical Operator


for Linear Momentum, Angular Momentum and Energy as Hermitian
Operator
It is a quite well-known fact that all the physical properties are actually real quantities, and therefore
are bound to have real values. It means that any operator which is used to represent a physical property must
yield real values. In this section, we will discuss the elementary idea of Hermitian operators (named in honor
of a great mathematician Charles Hermite), and will also prove that many important operators in quantum
mechanics like linear momentum, angular momentum and Hamiltonian are Hermitian in nature.
 Elementary Idea of Hermitian Operator
Every physical property must have real eigen or expectation values, which therefore implies that the
corresponding operators should have some special characteristics. One of the most important special
characteristics includes a feature that the Hermitian conjugate of such an operator should be itself. In other
words, if the Hermitian conjugate of an operator is itself, the operator is called as Hermitian; however, if the
Hermitian conjugate of an operator is equal to its negative expression, the operator is called as anti-Hermitian
or skew-Hermitian. Mathematically, we can say that

𝑖𝑓 𝐴† = 𝐴; 𝐴 𝑖𝑠 𝐻𝑒𝑟𝑚𝑖𝑡𝑖𝑎𝑛 (316)

𝑖𝑓 𝐴† = −𝐴; 𝐴 𝑖𝑠 𝑎𝑛𝑡𝑖‒ 𝐻𝑒𝑟𝑚𝑖𝑡𝑖𝑎𝑛 (317)

Where A is an operator whose Hermitian conjugate is represented by A†.


However, the obvious question regarding the aforementioned definition would be “what is a
Hermitian conjugate and how is it obtained”. The answer is “the operator A† will be called as the Hermitian
conjugate (or adjoint) of operator A if the operation of A† on the complex conjugate of function ψ gives the
same result as when the A is operated over ψ”. Mathematically, we can say that
+∞ (318)
⟨𝜓|𝐴|𝜓⟩ = ∫ 𝜓 ∗ (𝑥)𝐴𝜓(𝑥)𝑑𝑥 = ⟨𝜓|𝐴𝜓⟩ = ⟨𝐴† 𝜓|𝜓⟩
−∞

or

⟨𝐴† 𝜑|𝜓⟩ = ⟨𝜑|𝐴𝜓⟩ (319)

1. Hermitian conjugates of different operators: The Hermitian conjugates of different operators can be
studied in three different categories.
i) Hermitian conjugates of quantum mechanical operators:
Let Q be any quantum mechanical operator, then by the definition of Hermitian conjugates operator,
we have the following condition.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 53

⟨𝜑|𝑄𝜓⟩ = ⟨𝑄 † 𝜑|𝜓⟩ (320)

If Q is the momentum operator, then we can proceed as discussed below.

(321)
∫ 𝜓 ∗ 𝑝̂𝑥 𝜓𝑑𝑥 = ∫ 𝜓𝑝̂𝑥 𝜓∗ 𝑑𝑥

ℎ 𝜕 ℎ 𝜕 † ∗ (322)

∫𝜓 ( ) 𝜓𝑑𝑥 = ∫ 𝜓 ( ) 𝜓 𝑑𝑥
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥

ℎ 𝜕 † ∗ ℎ † 𝜕 † ∗ (323)
∫𝜓( ) 𝜓 𝑑𝑥 = ∫ 𝜓 ( ) ( ) 𝜓 𝑑𝑥
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥

ℎ 𝜕 † ∗ ℎ 𝜕 (324)
∫𝜓( ) 𝜓 𝑑𝑥 = ∫ 𝜓 (− ) (− ) 𝜓 ∗ 𝑑𝑥
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥

ℎ 𝜕 † ∗ ℎ 𝜕 (325)
∫𝜓( ) 𝜓 𝑑𝑥 = ∫ 𝜓 ( ) 𝜓 ∗ 𝑑𝑥
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥

Therefore, we can say that the Hermitian conjugate of the linear momentum operator is itself, and hence it is
a Hermitian operator. Now from the most primitive definition of Hermitian operators, that all operators which
correspond to observable quantities, we can say that the Hermitian conjugates of the following operator are
themselves.

Operator Hermitian conjugate

𝑥̂ 𝑥̂

𝑥̂ 2 𝑥̂ 2

𝑝̂𝑥 𝑝̂𝑥

𝑝̂𝑥2 𝑝̂𝑥2

𝑇̂𝑥 𝑇̂𝑥

𝑉̂ (𝑥) 𝑉̂ (𝑥)

̂
𝐻 ̂
𝐻

ii) Hermitian conjugates of a constant operator:


There are some operators which are complex numbers. The Hermitian conjugates of such operators
are actually their complex conjugates. Let we have the operator A

𝐴̂ = 𝑎 + 𝑖𝑏 (326)

Copyright © Mandeep Dalal


54 A Textbook of Physical Chemistry – Volume I

and since the definition of Hermitian operator is

⟨𝜑|𝐴𝜓⟩ = ⟨𝐴† 𝜑|𝜓⟩ (327)

gives the integer as

⟨𝜑|(𝑎 + 𝑖𝑏)𝜓⟩ = ⟨(𝑎 − 𝑖𝑏)𝜑|𝜓⟩ = (𝑎 + 𝑖𝑏)⟨𝜑|𝜓⟩ (328)

Hence, the Hermitian conjugates of constant operators are their complex conjugates. The Hermitian conjugates
of some operators are given below.

Operator Hermitian conjugate

(𝑎 + 𝑖𝑏) (𝑎 + 𝑖𝑏)† = (𝑎 − 𝑖𝑏)

(+𝑖𝑏) (+𝑖𝑏)† = (−𝑖𝑏)

𝑖 𝑖 † 𝑖
(+ ) (+ ) = (− )
4 4 4
iii) Hermitian conjugates of a mathematical operator:
The Hermitian conjugates of mathematical operators can be obtained by obtaining their respective
integrals as discussed below. Let we have a mathematical operator A

𝑑 (326)
𝐴̂ =
𝑑𝑥
We use the following integral to derive the result
+∞ (327)
𝑑 𝑑𝜓(𝑥)
⟨𝜑| 𝜓⟩ = ∫ 𝜑∗ (𝑥) 𝑑𝑥
𝑑𝑥 𝑑𝑥
−∞

Integrating the above equation by part, we get

𝑑𝜑∗ (𝑥)
+∞ (328)
𝑑
⟨𝜑| 𝜓⟩ = [𝜑∗ (𝑥)𝜓(𝑥)] − ∫ 𝜓(𝑥) 𝑑𝑥
𝑑𝑥 𝑑𝑥
−∞

𝑑 (329)
= 0 − ⟨ 𝜑|𝜓⟩
𝑑𝑥

𝑑 (330)
= − ⟨ 𝜑|𝜓⟩
𝑑𝑥

Hence, the Hermitian conjugate of d/dx operator is −d/dx. Similarly, we can prove that the Hermitian conjugate
of d2/dx2 is d2/dx2.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 55

2. Properties of Hermitian conjugates: From the definition and properties of scalar product, adjoints or
Hermitian conjugate show the following properties.
i) Let C a constant and A as an operator.

(𝐶𝐴)† = 𝐶 ∗ 𝐴† (331)

For example

𝑖 𝜕 † 𝑖 † 𝜕 † (332)
( ) =( ) ( )
4 𝜕𝑥 4 𝜕𝑥

𝑖 𝜕 † 𝑖 𝜕 (333)
( ) = (− ) (− )
4 𝜕𝑥 4 𝜕𝑥

𝑖 𝜕 † 𝑖 𝜕 (334)
( ) =
4 𝜕𝑥 4 𝜕𝑥
ii) Let A and B as two operators.

(𝐴 + 𝐵)† = 𝐴† + 𝐵† (335)

For example

𝜕 𝜕2 𝜕 †

𝜕2
† (336)
( + 2) = ( ) + ( 2)
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥

𝜕 𝜕2 𝜕

𝜕2 (337)
( + 2 ) = (− ) + ( 2 )
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥

𝜕 𝜕2 𝜕

𝜕2 (338)
( + 2 ) = (− + 2)
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝜕𝑥

iii) Let A and B as two operators, then

(𝐴𝐵)† = 𝐴† 𝐵† (339)

For example

𝜕 𝜕2

𝜕 † 𝜕2
† (340)
( ) = ( ) ( 2)
𝜕𝑥 𝜕𝑥 2 𝜕𝑥 𝜕𝑥

𝜕 𝜕2

𝜕 𝜕2 (341)
( ) = (− ) ( )
𝜕𝑥 𝜕𝑥 2 𝜕𝑥 𝜕𝑥 2

Copyright © Mandeep Dalal


56 A Textbook of Physical Chemistry – Volume I

𝜕 𝜕2

𝜕3 (342)
( ) = (− )
𝜕𝑥 𝜕𝑥 2 𝜕𝑥 3

iv) Let A be the operators, then



(𝐴† ) = 𝐴 (343)

For example

𝜕 †

𝜕 (344)
[( ) ] = ( )
𝜕𝑥 𝜕𝑥

It should also be noted that the multiplication to an anti-hermitian operator by i makes it Hermitian, while the
vice-versa is also equally true for adjoints.
v) For any operator A and its adjoint, the product (AA†) is Hermitian. For instance

𝜕 𝜕 𝜕2 (343)
( ) (− ) = − 2
𝜕𝑥 𝜕𝑥 𝜕𝑥
vi) For any operator A and its adjoint, the sum (A+A†) is Hermitian. For instance

(𝑥 + 𝑥 † ) = 2𝑥 (343)

vii) For any operator A and its adjoint, then AA†+A†A is Hermitian. For instance

(𝑖3)(𝑖3)† + (𝑖3)† (𝑖3) = (𝑖3)(−𝑖3) + (−𝑖3)(𝑖3) = 9 + 9 = 18 (343)

3. Characterization of Hermitian operator: We know that the average value of any operator (say Â) in
quantum mechanics is calculated by the equation given below.

(344)
𝐴̅ = ∫ 𝜓 ∗ 𝐴̂𝜓𝑑𝑥

Where ψ is the wave function representing any quantum mechanical state and ψ* is its complexes conjugate.
Now because of the fact that the average value of any physical observable must be a real value, we can say
that the operator used in equation (344) must follow the following condition.

𝐴̅ = 𝐴̅∗ (345)

(346)
∫ 𝜓 ∗ 𝐴̂𝜓𝑑𝑥 = [∫ 𝜓 ∗ 𝐴̂𝜓𝑑𝑥]

∗ (347)
∫ 𝜓 ∗ 𝐴̂𝜓𝑑𝑥 = ∫(𝜓 ∗ )∗ (𝐴̂𝜓) 𝑑𝑥

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 57

∗ (348)
∫ 𝜓 ∗ 𝐴̂𝜓𝑑𝑥 = ∫ 𝜓(𝐴̂𝜓) 𝑑𝑥

Every linear operator that satisfies the equation (348) for all quantum-mechanically acceptable wave functions
is called the Hermitian operator.
Besides the form given by equation (348), one more popular definition of a Hermitian operator is
also given below.

∗ (349)
∫ 𝑓 ∗ 𝐴̂𝑔𝑑𝑥 = ∫ 𝑔(𝐴̂𝑓) 𝑑𝑥

From the equation, we can state that a Hermitian operator must fulfill the condition for the well-behaved
functions f and g. It can be clearly seen that on the left side of the equation (349), Â is operated over the function
g; while on the right side, the  is operated over the function f. However, if we put f = g, the equation (349) is
also reduced to equation (348); indicating that both definitions are correct.
4. Properties of Hermitian operators: The important properties of Hermitian operators are discussed below.
i) The eigenvalues of Hermitian operators are always real:
Let  be a Hermitian operator with a well-behaved wavefunction ψ representing a quantum
mechanical state, then we can say that

Â𝜓 = 𝑎𝜓 (350)

Each side of equation (350) can be expressed as an imaginary and a real part as well; with left-hand real part
equal to the right-hand real part, while left side imaginary part equal to right imaginary one. After taking the
complex conjugate of equation (350), the imaginary parts would reverse sign but still holding the condition of
equivalence.

Â∗ 𝜓 ∗ = 𝑎∗ 𝜓 ∗ (351)

Multiplying the equation (350) by ψ* and integrating over the whole configurational space, we get

(352)
∫ 𝜓 ∗ 𝐴̂𝜓𝑑𝑥 = 𝑎 ∫ 𝜓 ∗ 𝜓𝑑𝑥

Similarly, multiplying the equation (351) by ψ and integrating over the whole configurational space, we get

(353)
∫ 𝜓𝐴̂∗ 𝜓 ∗ 𝑑𝑥 = 𝑎∗ ∫ 𝜓𝜓 ∗ 𝑑𝑥

Now because left-hand sides of equation (352) and (353) are equal to each other (owing to the Hermitian nature
of the operator), the right-hand sides are also equivalent; therefore, we can say that

(354)
𝑎∗ ∫ 𝜓𝜓 ∗ 𝑑𝑥 = 𝑎 ∫ 𝜓 ∗ 𝜓𝑑𝑥

Copyright © Mandeep Dalal


58 A Textbook of Physical Chemistry – Volume I

(355)
0 = (𝑎 − 𝑎∗ ) ∫ 𝜓 ∗ 𝜓𝑑𝑥

Since the wave function is a square-integrable, the integral part of the equation (355) cannot be zero and left
us with the only possibility given blow.

(𝑎 − 𝑎∗ ) = 0 (356)

𝑎 = 𝑎∗ (357)

The physical interpretation of the result given by equation (357) is that a must be real in order to yield zero
from equation (356).
ii) Non-degenerate eigenfunctions of Hermitian operators are always orthogonal to each other:
Let ψm and ψn be two square-integrable eigenfunctions of a Hermitian operator Â; therefore, we say

𝐴̂𝜓𝑚 = 𝑎1 𝜓𝑚 (358)

also

𝐴̂∗ 𝜓𝑛∗ = 𝑎2 𝜓𝑛∗ (359)

Multiplying the equation (358) by ψn* and integrating over the whole configurational space, we get

(360)
∫ 𝜓𝑛∗ 𝐴̂𝜓𝑚 𝑑𝑥 = 𝑎1 ∫ 𝜓𝑛∗ 𝜓𝑚 𝑑𝑥

Similarly, multiplying the equation (359) by ψm and integrating over the whole configurational space, we get

(361)
∫ 𝜓𝑚 𝐴̂∗ 𝜓𝑛∗ 𝑑𝑥 = 𝑎2 ∫ 𝜓𝑚 𝜓𝑛∗ 𝑑𝑥

Now because left-hand sides of equation (360) and (361) are equal to each other (owing to the Hermitian nature
of the operator), the right-hand sides are also equivalent; therefore, we can say that

(362)
𝑎1 ∫ 𝜓𝑛∗ 𝜓𝑚 𝑑𝑥 = 𝑎2 ∫ 𝜓𝑚 𝜓𝑛∗ 𝑑𝑥

(363)
(𝑎1 − 𝑎2 ) ∫ 𝜓𝑚 𝜓𝑛∗ 𝑑𝑥 = 0

Since the wave functions used are non-degenerate i.e. a1 ≠ a2; the only possibility we are left with for the
equation to be true is given below.

(364)
∫ 𝜓𝑚 𝜓𝑛∗ 𝑑𝑥 = 0

Hence, we can say that ψm and ψn are definitely orthogonal to each other.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 59

iii) If two Hermitian operators commute, their product is also a Hermitian operator:
Let ψ1 and ψ2 be two well-behaved functions; while 𝐴̂ and 𝐵̂ as two Hermitian operators. Therefore,
we can say that

(365)
∫ 𝜓1∗ 𝐴̂𝐵̂𝜓2 𝑑𝑥

Since 𝐴̂ is Hermitian, we can say that

(366)
∫ 𝜓1∗ 𝐴̂𝐵̂𝜓2 𝑑𝑥 = ∫ 𝜓1∗ 𝐴̂(𝐵̂𝜓2 )𝑑𝑥

(367)
∫ 𝐴̂∗ 𝜓1∗ 𝐵̂𝜓2 𝑑𝑥 = ∫ 𝜓1∗ 𝐴̂(𝐵̂𝜓2 )𝑑𝑥

Since 𝐵̂ is also Hermitian, therefore

(368)
∫(𝐴̂∗ 𝜓1∗ )𝐵̂𝜓2 𝑑𝑥 = ∫ 𝐵̂∗ 𝐴̂∗ 𝜓1∗ 𝜓2 𝑑𝑥

From equation (366) and (368), we get

(369)
∫ 𝜓1∗ 𝐴̂𝐵̂𝜓2 𝑑𝑥 = ∫ 𝐵̂ ∗ 𝐴̂∗ 𝜓1∗ 𝜓2 𝑑𝑥

If the operator 𝐴̂ and 𝐵̂ commute with each other, we have

𝐴̂𝐵̂ = 𝐵̂𝐴̂ 𝑜𝑟 𝐴̂∗ 𝐵̂∗ = 𝐵̂∗ 𝐴̂∗ (370)

Therefore, equation (369) becomes

(371)
∫ 𝜓1∗ 𝐴̂𝐵̂𝜓2 𝑑𝑥 = ∫ 𝐴̂∗ 𝐵̂∗ 𝜓1∗ 𝜓2 𝑑𝑥

Which is the condition for the product operator to act as Hermitian.


iv) If two Hermitian operators do not commute, their commutator operator is anti-Hermitian in nature:
Let 𝐴̂ and 𝐵̂ as two Hermitian operators; therefore, we can say that their commutation must follow
the following condition.
∗ ∗
[𝐴̂, 𝐵̂] = (𝐴̂𝐵̂) − (𝐵̂𝐴̂)

(372)

or

[𝐴̂, 𝐵̂] = 𝐴̂∗ 𝐵̂∗ − 𝐵̂∗ 𝐴̂∗ = −(𝐵̂∗ 𝐴̂∗ − 𝐴̂∗ 𝐵̂∗ ) (373)

or

Copyright © Mandeep Dalal


60 A Textbook of Physical Chemistry – Volume I


[𝐴̂, 𝐵̂] = −[𝐵̂, 𝐴̂]

(374)


[𝐴̂, 𝐵̂] = −[𝐵̂, 𝐴̂]

(375)

For instance, consider the commutator of position and momentum operator

ℎ (376)
[𝑥̂, 𝑝̂𝑥 ] = 𝑖
2𝜋
The commutator iħ is antihermitian in nature.
 The Linear Momentum Operator as Hermitian
In order to prove the linear momentum operator as the Hermitian, we must find its Hermitian
conjugate first. The general expression of linear momentum operator is

ℎ 𝜕 (377)
𝑝̂𝑥 =
2𝜋𝑖 𝜕𝑥

Let 𝑝̂𝑥† be the Hermitian conjugate which can be calculated as follows:

ℎ † 𝜕 † (378)
𝑝̂𝑥† =( ) ( )
2𝜋𝑖 𝜕𝑥
or

ℎ 𝜕 (379)
𝑝̂𝑥† = (− ) (− )
2𝜋𝑖 𝜕𝑥
or

ℎ 𝜕 (380)
𝑝̂𝑥† = ( ) ( )
2𝜋𝑖 𝜕𝑥
Comparing equation (377) and (380), we can see that the Hermitian conjugate of linear momentum operator is
exactly equal to the linear momentum operator i.e. 𝑝̂𝑥† = 𝑝̂𝑥 ; proving that it is defiantly a Hermitian operator.
 The Angular Momentum Operator as Hermitian
In order to prove the angular momentum operator as Hermitian, we must find its Hermitian conjugate
first. The general expression of the angular momentum operator is

ℎ 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (381)
𝐿̂ = [(𝑦 − 𝑧 ) + (𝑧 − 𝑥 ) + (𝑥 − 𝑦 )]
2𝜋𝑖 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑦 𝜕𝑥

Let 𝐿̂†𝑥 be the Hermitian conjugate which can be calculated as follows:

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 61

† (382)
ℎ 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕
𝐿̂†𝑥 = [ [(𝑦 − 𝑧 ) + (𝑧 − 𝑥 ) + (𝑥 − 𝑦 )]]
2𝜋𝑖 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑦 𝜕𝑥

ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 †
=[ 𝑦 − 𝑧 + 𝑧 − 𝑥 + 𝑥 − 𝑦 ]
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥

or

ℎ † 𝜕 † ℎ † 𝜕 † ℎ † 𝜕 † (383)
𝐿̂†𝑥 = ( ) (𝑦)† ( ) − ( ) (𝑧)† ( ) + ( ) (𝑧)† ( )
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥
ℎ † †
𝜕 † ℎ † †
𝜕 † ℎ † †
𝜕 †
− ( ) (𝑥) ( ) + ( ) (𝑥) ( ) − ( ) (𝑦) ( )
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥

or

ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 (384)
𝐿̂†𝑥 = (− ) (𝑦) (− ) − (− ) (𝑧) (− ) + (− ) (𝑧) (− )
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥
ℎ 𝜕 ℎ 𝜕
− (− ) (𝑥) (− ) + (− ) (𝑥) (− )
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦
ℎ 𝜕
− (− ) (𝑦) (− )
2𝜋𝑖 𝜕𝑥
or

ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 ℎ 𝜕 (385)
𝐿̂† = 𝑦 − 𝑧 + 𝑧 − 𝑥 + 𝑥 − 𝑦
2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑧 2𝜋𝑖 𝜕𝑦 2𝜋𝑖 𝜕𝑥

ℎ 𝜕 𝜕 𝜕 𝜕 𝜕 𝜕 (386)
𝐿̂† = [(𝑦 − 𝑧 ) + (𝑧 − 𝑥 ) + (𝑥 − 𝑦 )]
2𝜋𝑖 𝜕𝑧 𝜕𝑦 𝜕𝑥 𝜕𝑧 𝜕𝑦 𝜕𝑥

Comparing equation (381) and (386), we can see that the Hermitian conjugate of the angular momentum
operator is exactly equal to the angular momentum operator i.e. 𝐿̂† = 𝐿̂; proving that it is defiantly a Hermitian
operator.
 The Hamiltonian or Energy Operator as Hermitian
In order to prove the energy operator as Hermitian, we must find its Hermitian conjugate first. The
general expression of the energy operator is

−ℎ2 𝜕 2 (387)
̂=
𝐻 + 𝑉(𝑥)
8𝜋 2 𝑚 𝜕𝑥 2

Let 𝐻
̂ † be the Hermitian conjugate which can be calculated as follows:

Copyright © Mandeep Dalal


62 A Textbook of Physical Chemistry – Volume I

−ℎ2 𝜕 2
† (388)
̂† = [
𝐻 + 𝑉(𝑥) ]
8𝜋 2 𝑚 𝜕𝑥 2

or

−ℎ2 𝜕 𝜕
† (389)
̂† = [
𝐻 + 𝑉(𝑥) ]
8𝜋 2 𝑚 𝜕𝑥 𝜕𝑥

−ℎ2

𝜕 † 𝜕 † (390)
̂† = (
𝐻 ) ( ) ( ) + (𝑉(𝑥))†
8𝜋 2 𝑚 𝜕𝑥 𝜕𝑥

or

−ℎ2 𝜕 𝜕 (391)
̂† = (
𝐻 ) (− ) (− ) + (𝑉(𝑥))
8𝜋 2 𝑚 𝜕𝑥 𝜕𝑥

−ℎ2 𝜕 2 (392)
̂† =
𝐻 + 𝑉(𝑥)
8𝜋 2 𝑚 𝜕𝑥 2
Comparing equation (387) and (392), we can see that the Hermitian conjugate of energy operator is exactly
equal to the energy operator i.e. 𝐻 ̂ ; proving that it is defiantly a Hermitian operator.
̂† = 𝐻

 The Average Value of the Square of Hermitian Operators


The expectation value of the square of every Hermitian operator is always positive. In other words,
we can say that if A is a Hermitian operator, then

〈𝐴2 〉 > 0 (393)

This can be proved by taking a well-behaved function ψ as discussed below.

∫ 𝜓 ∗ 𝐴2 𝜓𝑑𝜏 (394)
〈𝐴2 〉 =
∫ 𝜓 ∗ 𝜓𝑑𝜏

The right-hand side of equation (394) will be positive only if the numerator as well as denominator, both are
either positive or negative. Since the wave-function is well-behaved (normalized), the value of denominator is

(395)
∫ 𝜓 ∗ 𝜓𝑑𝜏 = 1

Since the denominator is positive, the numerator must also be positive. Now owing to the Hermitian nature of
operator A, we can evaluate the numerator as given below.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 63

(396)
∫ 𝜓 ∗ 𝐴2 𝜓𝑑𝜏 = ∫ 𝜓 ∗ 𝐴𝐴∗ 𝜓 𝑑𝜏

(397)
= ∫(𝜓 ∗ 𝐴∗ )𝐴𝜓 𝑑𝜏

or

(398)
= ∫|𝐴𝜓|2 𝑑𝜏

Hence, the value of numerator given by equation (398) is greater than zero i.e. positive, making the average
value of the square of the Hermitian operator (A) also positive.

 Commuting Operators and Uncertainty Principle (x & p; E & t)


One of the most important properties of operator multiplication is the commutation relation or the
commutation rule. Two operators, A and B, are said to be commutating or non-commutating depending upon
the magnitude of their commutator.

̂ 𝐵̂] = 𝐴̂𝐵̂ − 𝐵̂𝐴̂ = 0 → 𝐶𝑜𝑚𝑚𝑢𝑡𝑎𝑡𝑖𝑛𝑔


[𝐴, (399)

and

̂ 𝐵̂] = 𝐴̂𝐵̂ − 𝐵̂𝐴̂ ≠ 0 → 𝑁𝑜𝑛-𝑐𝑜𝑚𝑚𝑢𝑡𝑎𝑡𝑖𝑛𝑔


[𝐴, (400)

The physical significance of the commutation relations implies in the fact that when two operators commute,
they possess simultaneous set of eigenfunctions; and their respective physical properties can be evaluated
simultaneously and accurately. However, if the commutator is non-zero, the respective physical properties
cannot be obtained simultaneously and accurately; which is actually the popular uncertainty principal. Two of
the most common uncertainty systems; position-momentum and energy-time; can also be proved from
commutation relations.
 Position-Momentum Uncertainty (x & p)
The position-momentum uncertainty can be justified only if the commutation of their operators is
non-zero. Therefore, we need to find the following.

[𝑥̂, 𝑝̂𝑥 ] (401)

Let it be operated over a function ψ. We have

[𝑥̂, 𝑝̂𝑥 ]𝜓 = 𝑥̂ 𝑝̂𝑥 𝜓 − 𝑝̂𝑥 ̂𝑥 𝜓 (402)

or

Copyright © Mandeep Dalal


64 A Textbook of Physical Chemistry – Volume I

ℎ 𝜕 ℎ 𝜕 (403)
[𝑥̂, 𝑝̂𝑥 ]𝜓 = 𝑥 𝜓− 𝑥𝜓
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥
ℎ 𝜕𝜓 ℎ 𝜕𝜓 ℎ 𝜕𝑥 (404)
[𝑥̂, 𝑝̂𝑥 ]𝜓 = 𝑥 − 𝑥 − 𝜓
2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥 2𝜋𝑖 𝜕𝑥
ℎ (405)
[𝑥̂, 𝑝̂𝑥 ]𝜓 = − 𝜓
2𝜋𝑖
ℎ ℎ𝑖 (406)
[𝑥̂, 𝑝̂𝑥 ] = − = = 𝑖ħ
2𝜋𝑖 2𝜋
Equation (406) proves that we cannot determine the position and momentum of a particle along one axis
simultaneously and accurately.
 Energy-Time Uncertainty (E & t)
The energy-time uncertainty can be justified only if the commutation of their operators is non-zero.
Therefore, we need to find the following.

[𝑡̂, 𝐸̂ ] (407)

Let it be operated over a function ψ(t). We have

[𝑡̂, 𝐸̂ ]𝜓 = 𝑡̂ 𝐸̂ 𝜓 − 𝐸̂ 𝑡̂ 𝜓 (408)

or

ℎ 𝜕 ℎ 𝜕 (409)
[𝑡̂, 𝐸̂ ]𝜓 = 𝑡 𝜓− 𝑡𝜓
2𝜋𝑖 𝜕𝑡 2𝜋𝑖 𝜕𝑡
ℎ 𝜕𝜓 ℎ 𝜕𝜓 ℎ 𝜕𝑡 (410)
[𝑡̂, 𝐸̂ ]𝜓 = 𝑡 − 𝑡 − 𝜓
2𝜋𝑖 𝜕𝑡 2𝜋𝑖 𝜕𝑡 2𝜋𝑖 𝜕𝑡
ℎ (411)
[𝑡̂, 𝐸̂ ]𝜓 = − 𝜓
2𝜋𝑖
ℎ (412)
[𝑡̂, 𝐸̂ ] = −
2𝜋𝑖
ℎ𝑖 (413)
[𝑡̂, 𝐸̂ ] =
2𝜋

[𝑡̂, 𝐸̂ ] = 𝑖ħ (414)

The equation (412) proves that higher the lifetime of the state lower will be energy fluctuation i.e. uncertainty
ΔE, and the vice-versa is also true.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 65

 Schrodinger Wave Equation for a Particle in One Dimensional Box


In the first section of this chapter, we discussed the postulates of quantum mechanics i.e. the step-by-
step procedure to solve a quantum mechanical problem. Now it’s the time to implement those rules to the
simplest quantum mechanical problem i.e. particle in a one-dimensional box. Consider a particle trapped in a
one-dimensional box of length “a”, which means that this particle can travel in only one direction only, say
along x-axis. The potential inside the box is V, while outside to the box it is infinite.

Figure 7. The particle in a one-dimensional box.

One other popular depiction of the particle in a one-dimensional box is also given in which the potential is
shown vertically while the displacement is projected along the horizontal line.

Figure 8. The second representation particle in a one-dimensional box.

So far we have considered a quantum mechanical system of a particle trapped in a one-dimensional box. Now
suppose that we need to find various physical properties associated with different states of this system. Had it
been a classical system, we would use simple formulas from classical mechanics to determine the value of
different physical properties. However, being a quantum mechanical system, we cannot use those expressions
because they would give irrational results. Therefore, we need to use the postulates of quantum mechanics to
evaluate various physical properties.
Let ψ be the function that describes all the states of the particle in a one-dimensional box. At this
point we have no information about the exact mathematical expression of ψ; nevertheless, we know that there
is one operator that does not need the absolute expression of wave function but uses the symbolic form only,
the Hamiltonian operator. The operation of Hamiltonian operator over this symbolic form can be rearranged
to give to construct the Schrodinger wave equation; and we all know that the wave function as well the energy,
both are the obtained as this second-order differential equation is solved. Mathematically, we can say that

Copyright © Mandeep Dalal


66 A Textbook of Physical Chemistry – Volume I

̂ 𝜓 = 𝐸𝜓
𝐻 (415)

After putting the value of one-dimensional Hamiltonian in equation (415), we get

−ℎ2 𝜕 2 (416)
[ 2 + 𝑉] 𝜓 = 𝐸𝜓
8𝜋 𝑚 𝜕𝑥 2

or

−ℎ2 𝜕 2 𝜓 (417)
+ 𝑉𝜓 = 𝐸𝜓
8𝜋 2 𝑚 𝜕𝑥 2
−ℎ2 𝜕 2 𝜓 (418)
+ 𝑉𝜓 − 𝐸𝜓 = 0
8𝜋 2 𝑚 𝜕𝑥 2
𝜕 2 𝜓 8𝜋 2 𝑚 8𝜋 2 𝑚 (419)
+ 𝐸𝜓 − 𝑉𝜓 = 0
𝜕𝑥 2 ℎ2 ℎ2
or

𝜕 2 𝜓 8𝜋 2 𝑚 (420)
+ (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 ℎ2
The above-mentioned second order differential equation is the Schrodinger wave equation for a particle
moving along one dimension only. Since the conditions outside and inside the box are different, the equation
(420) must be solved separately for both cases.
1. The solution of Schrodinger wave equation for outside the box: After putting the value of potential
outside the box in equation (420) i.e. V = ∞, we get

𝜕 2 𝜓 8𝜋 2 𝑚 (421)
+ (𝐸 − ∞)𝜓 = 0
𝜕𝑥 2 ℎ2
Since E is negligible in comparison to the ∞, the above equation becomes

𝜕2𝜓 (422)
− ∞𝜓 = 0
𝜕𝑥 2
𝜕2𝜓 (423)
∞𝜓 =
𝜕𝑥 2
1 𝜕2𝜓 (424)
𝜓= =0
∞ 𝜕𝑥 2
The physical significance of the equation (424) is that the particle cannot go outside the box, and is always
reflected back when it strikes the boundaries. In other words, as the function describing the existence of
particles is zero outside the box, the particle cannot exist outside the box.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 67

2. Solution of Schrodinger wave equation for inside the box: After putting the value of potential inside the
box in equation (420) i.e. V = 0, we get

𝜕 2 𝜓 8𝜋 2 𝑚 (425)
+ (𝐸 − 0)𝜓 = 0
𝜕𝑥 2 ℎ2
or

𝜕 2 𝜓 8𝜋 2 𝑚𝐸 (426)
+ 𝜓=0
𝜕𝑥 2 ℎ2
Now consider

8𝜋 2 𝑚𝐸 (427)
𝑘2 =
ℎ2
After using the value from equation (427) in equation (426), we get

𝜕2𝜓 (428)
+ 𝑘 2𝜓 = 0
𝜕𝑥 2
The general solution of the above equation is

𝜓 = 𝐴 𝑆𝑖𝑛 𝑘𝑥 + 𝐵 𝐶𝑜𝑠 𝑘𝑥 (429)

Hence, from just the symbolic form we have obtained some kind of expression for the wave function defining
quantum mechanical states. However, the function given by equation (429) cannot be used to find different
physical properties or the nature of corresponding quantum mechanical states. The reason is that this
expression does have some unknown parameters like A, B and k. Since the function describing any quantum
mechanical state must be single-valued, finite and continuous; the function ψ must also follow these conditions
to become a “wave-function”. Therefore, these boundary conditions are fulfilled only if the magnitude of ψ is
zero at the start and at the end of the box (function outside is zero).
i) The first boundary condition: ψ must vanish when x = 0 i.e.

0 = 𝐴 𝑆𝑖𝑛 𝑘(0) + 𝐵 𝐶𝑜𝑠 𝑘(0) (430)

0 = 0 + 𝐵 𝐶𝑜𝑠 𝑘(0) (431)

𝐵=0 (432)

So, the function ψ is acceptable only if the value of the constant B is zero. After putting the value of B in
equation (429), we get

𝜓 = 𝐴 𝑆𝑖𝑛 𝑘𝑥 + (0) 𝐶𝑜𝑠 𝑘𝑥 (433)

𝜓 = 𝐴 𝑆𝑖𝑛 𝑘𝑥 (434)

Copyright © Mandeep Dalal


68 A Textbook of Physical Chemistry – Volume I

ii) The second boundary condition: ψ must vanish when x = a, i.e.,

0 = 𝐴 𝑆𝑖𝑛 𝑘𝑎 (435)

𝑆𝑖𝑛 𝑘𝑎 = 0 (436)

Moreover, as we know that

𝑆𝑖𝑛 0 = 0 𝑜𝑟 𝑆𝑖𝑛 0𝜋 = 0 (437)

𝑆𝑖𝑛 180 = 0 𝑜𝑟 𝑆𝑖𝑛 1𝜋 = 0 (438)

𝑆𝑖𝑛 360 = 0 𝑜𝑟 𝑆𝑖𝑛 2𝜋 = 0 (439)

𝑆𝑖𝑛 540 = 0 𝑜𝑟 𝑆𝑖𝑛 3𝜋 = 0 (440)

or

𝑆𝑖𝑛 𝑛𝜋 = 0 (441)

Where n = 0, 1, 2, 3, 4, 5 …. ∞. Comparing equation (436) and equation (441), we conclude that

𝑆𝑖𝑛 𝑘𝑎 = 𝑆𝑖𝑛 𝑛𝜋 = 0 (442)

Which eventually means that

𝑘𝑎 = 𝑛𝜋 (443)
𝑛𝜋 (444)
𝑘=
𝑎
After putting the value of k in equation (434), we get
𝑛𝜋𝑥 (445)
𝜓 = 𝐴 𝑆𝑖𝑛
𝑎
The only parameters that is still unknown in equation (445) is A, which can also be obtained by the condition
of normalization i.e. the function must define the state completely. Therefore, we can say that
𝑎 𝑎
𝑛𝜋𝑥 (446)
∫ 𝜓 = 𝐴 ∫ 𝑆𝑖𝑛2 (
2 2
)=1
𝑎
0 0

𝑎 (447)
𝐴2 . =1
2
(448)
2
2
2
𝐴 = 𝑜𝑟 𝐴 = √
𝑎 𝑎

After putting the value of A in equation (445), we get

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 69

(449)
2 𝑛𝜋𝑥
𝜓 = √ 𝑆𝑖𝑛
𝑎 𝑎

Since the function ψ also depends upon the discrete variable n, it is better to write the above equation given as

(450)
2 𝑛𝜋𝑥
𝜓𝑛 = √ 𝑆𝑖𝑛
𝑎 𝑎

The equation (450) represents all the quantum mechanical states of a particle in one-dimensional box. We can
obtain functions for individual states just by putting different values of “n” allowed by the boundary conditions.
For first quantum mechanical state i.e n = 1

(451)
2 𝜋𝑥
𝜓1 = √ 𝑆𝑖𝑛
𝑎 𝑎

For second quantum mechanical state i.e n = 2

(452)
2 2𝜋𝑥
𝜓2 = √ 𝑆𝑖𝑛
𝑎 𝑎

For third quantum mechanical state i.e n = 3

(453)
2 3𝜋𝑥
𝜓3 = √ 𝑆𝑖𝑛
𝑎 𝑎

Similarly, we can write the expression for ψ4, ψ5, ψ6 and so on. It is also worthy to note that even though the n
= 0 is permitted by the boundary condition, we still didn’t use it in equation (450); which is obviously because
it makes the whole function to collapse to zero.
One of the most remarkable results of this procedure that we have not discussed yet is the correlation
of equation (427) and equation (444).

8𝜋 2 𝑚𝐸 𝑛2 𝜋 2 (454)
𝑘2 = = 2
ℎ2 𝑎
𝑛2 ℎ2 (455)
𝐸𝑛 =
8𝑚𝑎2
The energy of different quantum mechanical states can be obtained by putting n = 1, 2, 3.... ∞ in equation
(455). Hence, we have obtained the wave-function as well as the energy for a particle in one-dimensional box.

Copyright © Mandeep Dalal


70 A Textbook of Physical Chemistry – Volume I

 Evaluation of Average Position, Average Momentum and Determination of


Uncertainty in Position and Momentum and Hence Heisenberg’s
Uncertainty Principle
The third postulate of quantum mechanics states that when the wave-function of a particular quantum
mechanical state is multiplied by the operator of an observable quantity, we get a real value multiplied by the
wave function itself. However, the value obtained this way can be constant or variable. Mathematically, the
constant value of the observable quantity can be reported directly, and the function is called an eigenfunction
of the operator under consideration. If the value of the physical property obtained after multiplying the wave
function by the corresponding operator is variable i.e. non-eigen, the value can be reported only after averaging
it over the whole configurational space.

∮ 𝜓 ∗ 𝑂̂𝜓 𝑑𝜏 (456)
<𝑎 >=
∮ 𝜓 ∗ 𝜓 𝑑𝜏

Since the wave function ψ is normalized, the denominator becomes unity; therefore, equation (456) is reduced
to the following

(457)
< 𝑎 > = ∮ 𝜓 ∗ 𝑂̂𝜓 𝑑𝜏

Since the operation by the Hamiltonian over the symbolic form has already given the absolute expressions for
different quantum mechanical states, now we can operate other operators to evaluate their average values. In
this section, we will determine the average values of position, position-squared, momentum and momentum-
squared; which in turn will be used to prove the Heisenberg’s uncertainty finally.
 Evaluation of Average Position
The quantum mechanical operator for the position of a particle in one-dimensional is 𝑥̂; while the
general form of wave function is

(458)
2 𝑛𝜋𝑥
𝜓𝑛 = √ 𝑆𝑖𝑛
𝑎 𝑎

Using this in equation (457), we get

(459)
< 𝑥 > = ∮ 𝜓 ∗ 𝑥 𝜓 𝑑𝜏

or

(460)
< 𝑥 > = ∮ 𝑥 𝜓 2 𝑑𝑥

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 71

2
𝑎
𝑛𝜋𝑥 2 𝑛𝜋𝑥
𝑎
(461)
< 𝑥 > = ∫ 𝑥. 𝑆𝑖𝑛2 ( ) 𝑑𝑥 = ∫ 𝑥 𝑆𝑖𝑛2 ( ) 𝑑𝑥
𝑎 𝑎 𝑎 𝑎
0 0

𝑎
1 − 𝐶𝑜𝑠 (
2𝑛𝜋𝑥
) (462)
2 𝑎 ] 𝑑𝑥
= ∫𝑥[
𝑎 2
0

1
𝑎
2𝑛𝜋𝑥 (463)
= ∫ (𝑥 − 𝑥 𝐶𝑜𝑠 ) 𝑑𝑥
𝑎 𝑎
0

1
𝑎
2𝑛𝜋𝑥
𝑎
(464)
= [∫ 𝑥 𝑑𝑥 − ∫ 𝑥 𝐶𝑜𝑠 ( ) 𝑑𝑥 ]
𝑎 𝑎
0 0

1 𝑎2 𝑎 (465)
= [ − 0] =
𝑎 2 2

 Evaluation of Average Position-Squared


The quantum mechanical operator for the position-squared of a particle in one-dimensional is 𝑥̂ 2 ;
Using this in equation (457), we get

(466)
< 𝑥 2 > = ∮ 𝜓 ∗ 𝑥 2 𝜓 𝑑𝑥

𝑎
2 𝑛𝜋𝑥 2 𝑛𝜋𝑥
𝑎
(467)
< 𝑥 2 > = ∫ 𝑥 2 . 𝑆𝑖𝑛2 ( ) 𝑑𝑥 = ∫ 𝑥 2 𝑆𝑖𝑛2 ( ) 𝑑𝑥
𝑎 𝑎 𝑎 𝑎
0 0

𝑎 2𝑛𝜋𝑥
1 − 𝐶𝑜𝑠 ( 𝑎 ) (468)
2 2
= ∫𝑥 [ ] 𝑑𝑥
𝑎 2
0

2 𝑎3 𝑎3 1 𝑎3 𝑎3 (469)
= [ − 2 2] = [ − 2 2 ]
𝑎 6 4𝑛 𝜋 𝑎 3 2𝑛 𝜋

𝑎2 𝑎2 (470)
= − 2 2
3 2𝑛 𝜋

 Evaluation of Average Momentum


The quantum mechanical operator for the position-squared of a particle in one-dimensional is 𝑝̂𝑥 ;
Using this in equation (457), we get

Copyright © Mandeep Dalal


72 A Textbook of Physical Chemistry – Volume I

ℎ 𝜕 (471)
< 𝑝̂𝑥 > = ∮ 𝜓 ∗ 𝜓 𝑑𝑥
2𝜋𝑖 𝜕𝑥
𝑎 (472)
2 𝑛𝜋𝑥 ℎ 𝜕 2 𝑛𝜋𝑥
< 𝑝̂𝑥 > = ∫ [√ 𝑆𝑖𝑛 ( )] [√ 𝑆𝑖𝑛 ( )] 𝑑𝑥
𝑎 𝑎 2𝜋𝑖 𝜕𝑥 𝑎 𝑎
0

ℎ 2
𝑎
𝑛𝜋𝑥 𝑛𝜋 𝑛𝜋𝑥 (473)
= [ ] ∫ 𝑆𝑖𝑛 ( ) ( ) 𝐶𝑜𝑠 ( ) 𝑑𝑥
2𝜋𝑖 𝑎 𝑎 𝑎 𝑎
0

ℎ 2 𝑛𝜋 𝑛𝜋𝑥
𝑎
𝑛𝜋𝑥 (474)
= [ ] ( ) ∫ 𝑆𝑖𝑛 ( ) 𝐶𝑜𝑠 ( ) 𝑑𝑥
2𝜋𝑖 𝑎 𝑎 𝑎 𝑎
0

< 𝑝̂𝑥 > = 0 (475)

 Evaluation of Average Momentum-Squared


The quantum mechanical operator for the position-squared of particle in one-dimensional is 𝑝̂𝑥2 ;
Using this in equation (457), we get

ℎ2 𝜕 2 (476)
< 𝑝̂𝑥2 > = ∮ 𝜓 ∗ (− ) 𝜓 𝑑𝑥
4𝜋 2 𝜕𝑥 2

𝑎 (477)
2 𝑛𝜋𝑥 ℎ2 𝜕 2 2 𝑛𝜋𝑥
< 𝑝̂𝑥2 > = ∫ [√ 𝑆𝑖𝑛 ( )] (− 2 2 ) [√ 𝑆𝑖𝑛 ( )] 𝑑𝑥
𝑎 𝑎 4𝜋 𝜕𝑥 𝑎 𝑎
0

ℎ2 2
𝑎
𝑛𝜋𝑥 𝑛𝜋 2 𝑛𝜋𝑥 (478)
= − 2 ( ) ∫ 𝑆𝑖𝑛 ( ) [(−) ( ) 𝑆𝑖𝑛 ( )] 𝑑𝑥
4𝜋 𝑎 𝑎 𝑎 𝑎
0

ℎ2 2 𝑛𝜋 2 𝑛𝜋𝑥
𝑎
(479)
= 2 ( ) ( ) ∫ 𝑆𝑖𝑛2 ( ) 𝑑𝑥
4𝜋 𝑎 𝑎 𝑎
0

𝑛 2 ℎ2
𝑎
𝑛𝜋𝑥 (480)
= 3
∫ 𝑆𝑖𝑛2 ( ) 𝑑𝑥
2𝑎 𝑎
0

𝑎
1 − 𝐶𝑜𝑠 (
2𝑛𝜋𝑥
) (481)
𝑛 2 ℎ2 𝑎 ] 𝑑𝑥
= ∫ [
2𝑎3 2
0

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 73

2𝑛𝜋𝑥 𝑎 (482)
2 2
𝑛 ℎ 𝑥 − 𝑆𝑖𝑛 ( )
= [ 𝑎 ]
2𝑎3 2𝑛𝜋
𝑎 0

or

𝑛 2 ℎ2 𝑎 (483)
= ( )
2𝑎3 2
or

𝑛 2 ℎ2 (484)
< 𝑝̂𝑥2 > =
4𝑎2

 The Heisenberg’s Uncertainty


In order to prove the Heisenberg’s uncertainty principle from for the quantum mechanical system of
a particle in one-dimensional box, we first need to find the uncertainties in position and momentum. Once both
uncertainties are known, we can simply multiply both to yield final result.
1. Uncertainty in position: The uncertainty in position is simply the difference between the square root of the
uncertainty in the position-squared. Mathematically, we can say that

∆𝑥 = (< 𝑥 2 > −< 𝑥 >2 )1/2 (485)

After putting the values of average position and position-squared from equation (465) and (470) in equation
(485), we get

𝑎2 𝑎2 𝑎 2
1/2 (486)
∆𝑥 = [( − 2 2 ) − ( ) ]
3 2𝑛 𝜋 2

or

𝑎2 𝑎2
1/2 (487
∆𝑥 = [( − 2 2 )]
12 2𝑛 𝜋

or

1 1 1/2 (488)
∆𝑥 = 𝑎 ( − 2 2 )
12 2𝑛 𝜋
2. Uncertainty in momentum: The uncertainty in momentum is simply the square root of the difference
between the uncertainty in momentum and uncertainty in the momentum-squared. Mathematically, we can say
that

Copyright © Mandeep Dalal


74 A Textbook of Physical Chemistry – Volume I

∆𝑝𝑥 = (< 𝑝𝑥2 > −< 𝑝𝑥 >2 )1/2 (489)

After putting the values of average position and position-squared from equation (475) and (484) in equation
(489), we get

𝑛2 ℎ2
1/2 (490)
∆𝑝𝑥 = [( 2 ) − (0)2 ]
4𝑎

or

𝑛ℎ (491)
∆𝑝𝑥 =
2𝑎
Now multiplying equation (488) and (491), we get

1 1 1/2
𝑛ℎ (492)
∆𝑥. ∆𝑝𝑥 = [𝑎 ( − 2 2) ] ( )
12 2𝑛 𝜋 2𝑎

or

𝑛ℎ 1 1 1/2 (493)
= ( − )
2 12 2𝑛2 𝜋 2
Multiply and divide the above equation by 2nπ

𝑛ℎ 2𝑛𝜋 1 1 1/2 (493)


∆𝑥. ∆𝑝𝑥 = . ( − )
2 2𝑛𝜋 12 2𝑛2 𝜋 2
or

𝑛ℎ 1 4𝑛2 𝜋 2 4𝑛2 𝜋 2
1/2 (494)
= . ( − 2 2)
2 2𝑛𝜋 12 2𝑛 𝜋

or

ℎ 𝑛2 𝜋 2
1/2 (495)
∆𝑥. ∆𝑝𝑥 = ( − 2)
4𝜋 3

Since n2π2/3 is always greater than 2, we can conclude that

ℎ (496)
∆𝑥. ∆𝑝𝑥 >
4𝜋
Which is the famous Heisenberg’s uncertainty principle.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 75

 Pictorial Representation of the Wave Equation of a Particle in One


Dimensional Box and Its Influence on the Kinetic Energy of the Particle in
Each Successive Quantum Level
The solution of the Schrodinger wave equation for a one-dimensional box gives the wave function as
well as the energy of the system. The general form of wave-function representing various quantum mechanical
states is given below.

(497)
2 𝑛𝜋𝑥
𝜓𝑛 = √ 𝑆𝑖𝑛
𝑎 𝑎

The energy of the system is given by equation (498) as:

𝑛2 ℎ2 (498)
𝐸𝑛 =
8𝑚𝑎2
The general depiction of a particle trapped in a one-dimensional box with zero potential inside, along with the
conditions outside, is shown below.

Figure 9. The graphical and pictorial representation of various wave-functions of the particle trapped in a
one-dimensional box.

Copyright © Mandeep Dalal


76 A Textbook of Physical Chemistry – Volume I

The pictorial representation of the wave-functions in different quantum mechanical states and the
corresponding energies are shown below.

Figure 10. The graphical and pictorial representation of various wave-functions of the particle trapped in a
one-dimensional box.

It can be seen clearly from the figure given above that as the number of nodes in wave-function defining a
particular quantum mechanical state increases, the energy of the state also increases.
Furthermore, we can also comment on the symmetry of different wave functions w.r.t the center of
the box. The symmetry of different states can be classified mainly into two categories as given below.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 77

𝑆𝑦𝑚𝑚𝑒𝑡𝑟𝑖𝑐 → 𝐸𝑣𝑒𝑛 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛 → 𝜓𝑜𝑑𝑑

and

𝐴𝑛𝑡𝑖𝑠𝑦𝑚𝑚𝑒𝑡𝑟𝑖𝑐 → 𝑂𝑑𝑑 𝑓𝑢𝑛𝑐𝑡𝑖𝑜𝑛 → 𝜓𝑒𝑣𝑒𝑛

Hence, function like ψ1, ψ3, ψ5 are symmetric while ψ2, ψ4, ψ6 are antisymmetric. Some of the important results
wavefunction and energy analysis for the particle in a one-dimensional box are listed below.
 Quantization of Energy
Owing to the discrete domain of n i.e. 1, 2, 3 …. ∞; the kinetic energy associated with the particle,
that is trapped in a one-dimensional box, can also have discrete or quantized values only. Therefore, the
quantized variable is also popularly called as the “quantum number.

Figure 11. The quantized or discrete energy levels a particle of mass m, confined in a one-dimensional box
of length a.

It is also worthy to note that the energy gap between successive energy levels shows a linear divergence with
the increasing value of the quantum number n. Moreover, the energy of particle also depends inversely upon
the mass and the box length; which eventually means that the energy levels would become continuous if the
mass or length of the box becomes very large, proving the Bohr’s correspondence principle.

Copyright © Mandeep Dalal


78 A Textbook of Physical Chemistry – Volume I

 Non-Quantization of the Energy of the Particle


If the walls of the box are removed, the boundary conditions will no longer be applicable, and the
particle would become free to move. In other words, the constant A, B and k can have any value; and therefore,
states of the particle are not quantized anymore. The general expression for the energy of the particle is

𝑛2 ℎ2 (498)
𝐸𝑛 =
8𝑚𝑎2
Hence, in such a case, a freely moving particle like an electron has restrictions and gives a continuous energy
spectrum.
 Box length and the Wave Function at the Walls
We have already studied that the magnitude of the wave function at the ends of the box must be equal
to zero to maintain its continuity. This is possible only if the length of the box is an integral multiple of half of
the wavelength. This can be proved as

𝑛 2 ℎ2 (499)
𝐸𝑛 =
8𝑚𝑎2
Also

1 𝑚2 𝑣 2 𝑝2 (500)
𝐸 = 𝑚𝑣 2 = =
2 2𝑚 2𝑚
Using the de-Brogli relation (λ = h/p) in equation (500), we get

𝑝2 (ℎ/𝜆)2 ℎ2 (501)
𝐸= = =
2𝑚 2𝑚 2𝑚𝜆2
Now from equation (499) and (501), we conclude

𝑛2 ℎ2 ℎ2 (502)
=
8𝑚𝑎2 2𝑚𝜆2
or

𝑛2 1 (503)
2
= 2
4𝑎 𝜆
or

𝜆 (504)
𝑎 = 𝑛( )
2
This result of equation (504) also proves that the number of nodes in nth quantum mechanical state are n–1.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 79

 The Probability Density


The wave density of simply the probability density in the one-dimensional box is not the same at all
the points. It is more noticeable when the quantum number defining the state is small. However, it becomes
more and more uniform as n increases.

Figure 12. The graphical and pictorial representation of the probability density of a particle with mass m
and confined in a one-dimensional box of length a.

The increasing uniformity of with increasing value of n is in accordance with the Bohr’s
correspondence principle which states that the results of quantum mechanics approach classical values at very
high quantum numbers.

Copyright © Mandeep Dalal


80 A Textbook of Physical Chemistry – Volume I

 Lowest Energy of the Particle


As we have already discussed that the wave function and energy, both are obtained as the solution of
the Schrodinger wave equation for a particle in a one-dimensional box. The general forms of wave-function
and energy for various quantum mechanical states are given below.

(505)
2 𝑛𝜋𝑥 𝑛 2 ℎ2
𝜓𝑛 = √ 𝑆𝑖𝑛 𝑎𝑛𝑑 𝐸𝑛 =
𝑎 𝑎 8𝑚𝑎2

We can write the expressions for ψ1, ψ2, ψ3, ψ4, ψ5, ψ6 and so on; however, it is also worthy to note that even
though the n = 0 is permitted by the boundary condition, we cannot use it because this would make the whole
function to collapse to zero.

Figure 13. All the energy levels a particle in a one-dimensional box of including the “lowest energy of the
particle”.

Hence, the minimum acceptable value of the quantum number n is 1 rather than 0; which makes the minimum
energy of the particle non-zero.

ℎ2 (506)
𝐸1 =
8𝑚𝑎2

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 81

This non-zero value is popularly called as the zero-point energy and is a function of the mass of the particle
and length of the box.

Figure 14. The plot of the wave function (left) and probability for the lowest energy state a particle in one-
trapped in dimensional box.

Hence, in order to create the lowest energy, the particle must occupy the whole box without any node, having
the highest probability at the center.

Copyright © Mandeep Dalal


82 A Textbook of Physical Chemistry – Volume I

 Problems
Q 1. State and explain the third postulate of quantum mechanics.
Q 2. Why should the function representing a quantum mechanical state be continuous, single-valued and finite?
Q 3. Why don’t we report non-eigenvalues directly? What is the need for their expectation values?
Q 4. Derive Schrodinger wave equation from the postulates of quantum mechanics.
Q 5. What is the Max-Born interpretation of “wave function”? Explain in detail by taking the example of one-
dimensional systems.
Q 6. What is position-time uncertainty? How would you prove it for the photons passing through a slit of
length d?
Q 7. What is operator commutation? Evaluate [𝑥̂ 2 , 𝑝
̂].
𝑥

Q 8. Explain the energy-time uncertainty for a particle traveling along x-axis. Also, support your argument
from the results of operator algebra.
Q 9. What are Hermitian operators? Prove that the operators for linear momentum and angular momentum are
Hermitian in nature.
Q 10. Can the average value for the square of the Hermitian operator be negative? If not, explain why?
Q 11. Derive and solve the Schrodinger wave equation for a particle moving in a one-dimensional box.
Q 12. Prove the Heisenberg’s uncertainty principle for the particle trapped in a one-dimensional box of length
a. Also, comment on its validity in other systems.
Q 13. Give the pictorial representation of the first three quantum mechanical states of a particle in a one-
dimensional box. Also, formulate the corresponding symmetry and number of nodes.
Q 14. Derive the relation between the box length and the wavelength of the particle in the 1-dimensional box.
Q 15. What is zero-point energy? How is it created by a particle of mass m which is trapped in a one-
dimensional box of length a.
Q 16. What is the average position? How is it different from the “most probable position”?
Q 17. State and explain the Bohr’s correspondence principle.

Copyright © Mandeep Dalal


CHAPTER 1 Quantum Mechanics – I 83

 Bibliography
[1] B. R. Puri, L. R. Sharma, M.S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2018.
[2] I. N. Levine, Quantum Chemistry, Pearson Prentice Hall, New Jersey, USA, 2009.
[3] D. A. McQuarrie, Quantum Chemistry, University Science Books, California, USA, 2008.
[4] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[5] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[6] M. Reed, B. Simon Functional Analysis, Elsevier, Amsterdam, Netherlands, 2003.
[7] G. E. Bowman, Essential Quantum Mechanics, Oxford University Press, Oxford, UK, 2008.
[8] W. Heisenberg, Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik,
Zeitschrift für Physik, 1927, 172.
[9] E. Schrödinger, Die gegenwärtige Situation in der Quantenmechanik, Naturwissenschaften. 1935, 807.
[10] P. Alberto, C. Fiolhaisdag, V. Gil, Relativistic particle in a box, European Journal of Physics, 1996, 17.
[11] R. K. Prasad, Quantum Chemistry, New Age International Publishers, New Delhi, India, 2010.

Copyright © Mandeep Dalal


CHAPTER 2
Thermodynamics – I
 Brief Resume of First and Second Law of Thermodynamics
There are four laws of thermodynamics that define the fundamental physical quantities like
temperature, energy, and entropy that characterize thermodynamic systems at thermal equilibrium. These laws
describe how these quantities behave under different conditions and rule out the possibility of some phenomena
the perpetual motion. The zeroth law of thermodynamics states that If two systems are each in thermal
equilibrium with a third system, they are in thermal equilibrium with each other; therefore, this law helps to
define the concept of temperature. In this section, we will discuss the elementary ideas and mutual correlation
between the first and second laws of thermodynamics.
 First Law of Thermodynamics
The first law of thermodynamics states that the energy can neither be created nor destroyed, but can
be converted from one form to another.
The first law of thermodynamics is obtained on the experimental basis. In other words, we can say
that the energy of an isolated system is always constant, which means that whenever some energy disappears
from the system, an equal amount of energy in some other form is also produced. In 1847, Helmholtz explained
this situation in his famous words, “it is impossible to construct a perpetual machine”. The term perpetual
machine refers to a device that can work continuously without any energy consumption. Furthermore, we all
know that heat is always produced whenever some mechanical work is done. These correlations were studied
by Joule (1840 – 1880); and he found that mechanical work is directly proportional to the heat produced.
Mathematically, we can say that

𝑊 ∝ 𝑄 𝑜𝑟 𝑊 = 𝐽𝑄 (1)

Where J represents the proportionality constant and is called as “Joule’s mechanical equivalent of heat”. If Q
= 1, W = J; making J as the amount of mechanical work required to produce one calorie of heat. The
experimental value for J was found to be 4.184 joules, which is a very popular relation (1 calorie = 4.184
joule). The first law of thermodynamics can also be deduced from the equivalence of heat and work. Suppose
there is now an equivalence between the work and heat; and let Q heat is converted into work. Now when the
same amount of work is done to produce the heat Q′; considering Q ≠ Q′, we can say that Q is either greater
or less than Q′. This would eventually mean that a certain amount of energy has been destroyed or created in
this process, which is against the first law of thermodynamics.
The mathematical formulation of the first law thermodynamics can be obtained from the increase in
the internal energy of the system. The internal energy of the system can be increased in two ways; one is doing
work on the system, and the second one involves the supply of heat. Suppose that the initial internal energy of

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 85

the system is E1, after supplying heat q and doing work w on the system, the final amount of internal energy
can be formulated as:

𝐸2 = 𝐸1 + 𝑞 + 𝑤 (2)

𝐸2 − 𝐸1 = 𝑞 + 𝑤 (3)

∆𝐸 = 𝑞 + 𝑤 (4)

𝑞 = ∆𝐸 − 𝑤 (5)

If the work is done by the system, putting w = −PΔV in equation (5), we get

𝑞 = ∆𝐸 − (−𝑃Δ𝑉 ) (6)

𝑞 = ∆𝐸 + 𝑃Δ𝑉 (7)

The physical significance of the equation (7) is that heat absorbed by a given system is converted work done
by the system and to raise its internal energy.

Figure 1. The pictorial representation of the first law of thermodynamics.

It is also worthy to note that the general form of the first law of thermodynamics is applicable only
in the case of chemical thermodynamics or physical processes. In 1905, Albert Einstein showed that energy
and mass are just the faces of the coin, and can be transformed within each other. In other words, his findings
showed that the mass can be converted into energy and the energy can back be converted into mass.
Mathematically, the formulation is

𝐸 = 𝑚𝑐 2 (8)

Where m is the mass and c is the velocity of light. Since the velocity of light is a very large quantity (so the
square), even the small disappearance of mass would generate a huge amount of energy. Such observations are
pretty common in case of nuclear reactions and can be neglected here. Therefore, in a broad sense, it is the
“law of energy-mass conservation”.

Copyright © Mandeep Dalal


86 A Textbook of Physical Chemistry – Volume I

 Second Law of Thermodynamics


The second law of thermodynamics states that it is impossible to convert the heat completely into
work without leaving some effect elsewhere.
The second law of thermodynamics is actually a rational solution to the limitations of the first law.
For instance, the first law talks about the exact equivalence between heat and work, but it is quite far from
reality. In 1824, a French scientist Sadi Carnot showed that for every heat engine there is an upper limit to the
efficiency of conversion of heat to work. In order to illustrate Carnot’s conclusion, consider a locomotive
engine that is supplied with a certain amount of heat; however, all of that heat will not be used to move the
train but a part of it will always be consumed in some other processes like overcoming the friction. Let q2 be
the heat absorbed by the heat engine at temperature T2, and w is the amount of the work done by the system;
while q1 is the heat returned to the sink at temperature T1, then the Carnot’s formulation can be given as:

𝑤 𝑞2 − 𝑞1 𝑇2 − 𝑇1 (9)
𝜂= = =
𝑞2 𝑞2 𝑇2

Where η is the efficiency of the heat engine and is always less than one. Ideally, η = 1, which means that such
a heat engine would convert 100% of the heat absorbed into work.
One more limitation of the first law is that it does not tell about the feasibility of the process, like
whether the heat can flow from cold terminal to the hot one or not. It simply talks if the heat gained or heat
lost but not the direction of the process. The second law of thermodynamics states that all the spontaneous
processes are thermodynamically irreversible. The word “spontaneous” simply means a process that occurs by
itself and external drive is required. In other words, we can also say that heat cannot flow from a cold body to
hot, the water cannot uphill without any external drive.

Figure 2. The pictorial representation of the second law of thermodynamics.

The 2nd law of thermodynamics also states that the total entropy of an isolated system can never
decline with time; in other words, combined entropy of a system and surroundings remains constant in ideal

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 87

cases where the system is undergoing a reversible process. In all processes, including spontaneous processes,
that occur, the total entropy of the system and surroundings increases and the process is irreversible in the
thermodynamic frame. The entropy-increase accounts for the irreversibility of all the natural processes, and
the asymmetry between the past and the future. Overall, the 2nd law of thermodynamics can be labeled as an
empirical finding that was accepted as a truism of thermodynamic theory. The microscopic origin of the law
can be explained by statistical mechanics.

 Entropy Changes in Reversible and Irreversible Processes


In order to understand the entropy change in reversible and irreversible processes, we need to
understand the concept of entropy first. For a Carnot heat engine working at T 1 and T2, it has been observed
that the heat absorbed (q2) and heat returned (q1) are related as given below.

𝑞2 − 𝑞1 𝑇2 − 𝑇1 (10)
=
𝑞2 𝑇2

or

𝑞1 𝑇1 (11)
1− =1−
𝑞2 𝑇2

or

𝑞1 𝑇1 (12)
=
𝑞2 𝑇2
𝑞1 𝑞2 (13)
=
𝑇1 𝑇2
𝑞 (14)
= 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
𝑇
Therefore, we can say that for any particular system, the ratio of heat absorbed or lost isothermally and
reversibly to the absolute temperature at which this takes place is a constant parameter. If we consider q1 as
the heat absorbed at T1, equation (13) becomes
𝑞2 𝑞1 (15)
=−
𝑇2 𝑇1
𝑞2 𝑞1 (16)
+ =0
𝑇2 𝑇1

or
𝑞 (17)
∑ =0
𝑇

Copyright © Mandeep Dalal


88 A Textbook of Physical Chemistry – Volume I

Consider a reversible cyclic process, consists of many Carnot cycles. In going from point A to B and
then back A, all the closed paths cancel each other that results in parent path ABA.

Figure 3. A cyclic process made of many Carnot cycles.

For each cycle, we have equation (17), for an infinite number of cycles

𝛿𝑞 (18)
∑ =0
𝑇
Where δq is the extremely small amount of heat absorbed at temperature T during the course of an isothermal
and reversible process. Moreover, the total entropy change of the cyclic process ABA can be fragmented into
two components as:

𝛿𝑞 𝛿𝑞 𝛿𝑞 (19)
∑ = ∑ +∑ =0
𝑇 𝑇 𝑇
𝐴→𝐵 𝐵→𝐴

𝛿𝑞 𝛿𝑞 (20)
∑ =−∑
𝑇 𝑇
𝐴→𝐵 𝐵→𝐴

or
𝑞 𝑞 (21)
( ) = −( )
𝑇 𝐴→𝐵 𝑇 𝐵→𝐴

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 89

The physical significance of the equation (21) is that the value of q/T for the path A→B is the same as the path
B→A; which eventually means that the quantity q/T is actually a state function. This quantity i.e. q/T is called
as entropy, and is generally labeled as S. If SA and SB are the entropies at point A and B, respectively; then we
can say that
𝑞 (22)
𝑆𝐴 − 𝑆𝐵 =
𝑇
𝑞 (23)
𝛥𝑆 =
𝑇
Hence, the entropy change may be defined as the amount of the heat absorbed isothermally and reversibly
divided by the temperature at which the heat is absorbed. Being a state function, the change in entropy always
depends upon the initial and final state and not upon the path followed. Moreover, since the heat is absorbed
reversibly, it is better to use qrev in equation (23) instead of simply q, therefore
𝑞𝑟𝑒𝑣 (24)
𝛥𝑆 =
𝑇
For an extremely minute change, the above equation becomes
𝑞𝑟𝑒𝑣 (25)
𝑑𝑆 =
𝑇
The entropy is an extensive property measured in joule per Kelvin per mole (JK−1mol−1). The most
important significance of entropy is that it can be used to measure the randomness in the system.
 Entropy Changes in Reversible Processes
Suppose that the heat absorbed by the system and heat lost by the surrounding are under completely
reversible conditions. In other words, qrev is the heat absorbed and lost by the surrounding at temperature T,
then we can say that the entropy change in the system will be given by the following relation.
𝑞𝑟𝑒𝑣 (26)
𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 =
𝑇
Similarly, the entropy change in the surrounding will be
𝑞𝑟𝑒𝑣 (27)
𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = −
𝑇
Therefore, the total entropy change will be
𝑞𝑟𝑒𝑣 𝑞𝑟𝑒𝑣 (28)
𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = − =0
𝑇 𝑇
Hence, we can conclude that the entropy change in an isolated system is always zero i.e. the sum of entropy
change in system and entropy change in the surrounding is zero under reversible conditions.

Copyright © Mandeep Dalal


90 A Textbook of Physical Chemistry – Volume I

 Entropy Changes in Irreversible Processes


Every reversible process becomes irreversible even if only one part of it becomes irreversible. To
understand this, let us suppose that qirrev heat lost by the surrounding. Although this heat would be absorbed
by the system, the entropy of the system depends upon the heat absorbed reversibly. Therefore, entropy change
of the system at an absolute temperature T will be
𝑞𝑟𝑒𝑣 (29)
𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 =
𝑇
Similarly, the entropy change of the surrounding will be
𝑞𝑖𝑟𝑟𝑒𝑣 (30)
𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = −
𝑇
The total entropy of the isolated system (system + surrounding) will be
𝑞𝑟𝑒𝑣 𝑞𝑖𝑟𝑟𝑒𝑣 (31)
𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = −
𝑇 𝑇
Furthermore, as we know that the wrev > wirrev, and internal energy is a state function that is independent of
whether the process is reversible or irreversible. Mathematically, it is

∆𝐸 = 𝑞𝑟𝑒𝑣 − 𝑤𝑟𝑒𝑣 = 𝑞𝑖𝑟𝑟𝑒𝑣 − 𝑤𝑖𝑟𝑟𝑒𝑣 (32)

we can conclude that,

𝑞𝑟𝑒𝑣 > 𝑞𝑖𝑟𝑟𝑒𝑣 (33)


𝑞𝑟𝑒𝑣 𝑞𝑖𝑟𝑟𝑒𝑣 (34)
>
𝑇 𝑇
or
𝑞𝑟𝑒𝑣 𝑞𝑖𝑟𝑟𝑒𝑣 (35)
+ >0
𝑇 𝑇
After comparing equation (35) with equation (31), we get

𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 > 0 (36)

The physical interpretation of the above equation lies in the fact that all irreversible processes occur via a net
increase in entropy. In other words, the total entropy change of an isolated system in any irreversible process
is always greater than zero. The results of equation (28) and equation (36) can be combined to give a more
generalized form as:

𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 ≥ 0 (37)

Where the sign of ʻ=ʼ stands from reversible and ʻ>ʼ stands for irreversible phenomena, respectively.

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 91

The entropy change in the irreversible process like the flow of heat from the hot end (T 2) to cold end (T1) can
be calculated using the following relation.

𝑞𝑟𝑒𝑣 𝑞𝑟𝑒𝑣 1 1 (38)


𝛥𝑆𝑡𝑜𝑡𝑎𝑙 = − + = 𝑞𝑟𝑒𝑣 ( − )
𝑇2 𝑇1 𝑇1 𝑇2

Where qrev is the amount of heat transferred from T2 to T1.


Furthermore, the Clausius inequality can also be proved from equation (36) which governs that all
the irreversible processes are accompanied by a net increase in the entropy.

𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 > 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 (39)

This can be illustrated with the help of the following examples.


i) Irreversible isothermal expansion of an ideal gas: Suppose an ideal gas expanding isothermally and
irreversibly against vacuum. The condition isothermal means ΔT = 0, which in turn implies that ΔU = 0. Hence
from the first law of thermodynamics, we have dq = 0 i.e. no heat is transferred from, or to the surrounding
giving dSsurrounding = 0.
However, since the dSsystem = R ln (V2/V1). Since V2 is definitely greater than V1, therefore

𝑑𝑆𝑡𝑜𝑡𝑎𝑙 = 𝑑𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝑑𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = 𝑑𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 0 ≥ 0 (40)

ii) Heat flow from hot to the cold end: Suppose dq is the amount of heat transferred from temperature T2 to T1.
The entropy change at the source will be

𝑑𝑞 (41)
𝑑𝑆𝑠𝑜𝑢𝑟𝑐𝑒 = −
𝑇2

Similarly, the entropy change at the sink will be

𝑑𝑞 (42)
𝑑𝑆𝑠𝑖𝑛𝑘 =
𝑇1

The total entropy change of the system can be calculated as

𝑑𝑞 𝑑𝑞 (43)
𝑑𝑆𝑡𝑜𝑡𝑎𝑙 = 𝑑𝑆𝑠𝑜𝑢𝑟𝑐𝑒 + 𝑑𝑆𝑠𝑖𝑛𝑘 = −
𝑇1 𝑇2

or

1 1 (44)
𝑑𝑆𝑡𝑜𝑡𝑎𝑙 = 𝑑𝑞 ( − )
𝑇1 𝑇2

Now because the source is always at a higher temperature than sink (T 2 > T1), dStotal will be positive for sure,
proving that transfer of heat from a hot body to the cold body occurs via a net entropy increment.

Copyright © Mandeep Dalal


92 A Textbook of Physical Chemistry – Volume I

 Variation of Entropy with Temperature, Pressure and Volume


Consider one mole of ideal gas filled in a chamber fitted with a weightless and frictionless piston. If
the system absorbs a very small amount of heat δqrev isothermally and reversibly at temperature T, the total
entropy change in the system can be given by the following relation.

𝛿𝑞𝑟𝑒𝑣 (45)
𝑑𝑆 =
𝑇
Also, from the first law of thermodynamics

𝛿𝑞 = 𝑑𝐸 − 𝛿𝑤 (46)

The above equation may be written in the following form if the process is carried out under reversible
conditions.

𝛿𝑞𝑟𝑒𝑣 = 𝑑𝐸 − 𝛿𝑤 (47)

After putting the value of work of expansion (−PdV), the equation (47) becomes

𝛿𝑞𝑟𝑒𝑣 = 𝑑𝐸 − (−𝑃𝑑𝑉) = 𝑑𝐸 + 𝑃𝑑𝑉 (48)

Using equation (48) in equation (45), we get

𝑑𝐸 + 𝑃𝑑𝑉 (49)
𝑑𝑆 =
𝑇
or

𝑇𝑑𝑆 = 𝑑𝐸 + 𝑃𝑑𝑉 (50)

Moreover, for one mole of an ideal gas, the value of heat capacity at constant volume (Cv) is

𝑑𝐸 (51)
𝐶𝑣 =
𝑑𝑇
𝑑𝐸 = 𝐶𝑣 𝑑𝑇 (52)

Also

𝑃𝑉 = 𝑅𝑇 (53)

𝑅𝑇 (54)
𝑃=
𝑉
Where R is the gas constant; while P, T and V are the pressure, temperature and volume, respectively. Now
putting the value of equation (52) and (54) in equation (50), we get

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 93

𝑅𝑇 (55)
𝑇𝑑𝑆 = 𝐶𝑣 𝑑𝑇 + 𝑑𝑉
𝑉
𝑑𝑇 𝑑𝑉 (56)
𝑑𝑆 = 𝐶𝑣 +𝑅
𝑇 𝑉
Now consider that the initial state (T1, V1) is converted to the final state (T2, V2), the total entropy will be
calculated by the following relation
𝑆2 𝑇2 𝑉2 (57)
𝑑𝑇 𝑑𝑉
∫ 𝑑𝑆 = ∫ 𝐶𝑣 +∫𝑅
𝑇 𝑉
𝑆1 𝑇1 𝑉1

𝑆2 𝑇2 𝑉2 (58)
𝑑𝑇 𝑑𝑉
∫ 𝑑𝑆 = 𝐶𝑣 ∫ +𝑅 ∫
𝑇 𝑉
𝑆1 𝑇1 𝑉1

or

𝑇2 𝑉2 (59)
𝛥𝑆 = 𝐶𝑣 ln + 𝑅 ln
𝑇1 𝑉1

Since, from the ideal gas equation, the following also true for the initial and final state of the system.

𝑃1 𝑉1 = 𝑅𝑇1 (60)

𝑃2 𝑉2 = 𝑅𝑇2 (61)

Dividing equation (61) by equation (60), we get

𝑃2 𝑉2 𝑅𝑇2 (62)
=
𝑃1 𝑉1 𝑅𝑇1

or

𝑉2 𝑃1 𝑇2 (63)
=
𝑉1 𝑃2 𝑇1

After putting the value of (63) in equation (59), we get

𝑇2 𝑃1 𝑇2 (64)
𝛥𝑆 = 𝐶𝑣 ln + 𝑅 ln
𝑇1 𝑃2 𝑇1

As we know that Cp − Cv = R; therefore, after putting Cv = Cp – R in equation (64), we get

𝑇2 𝑃1 𝑇2 (65)
𝛥𝑆 = (𝐶𝑝 − 𝑅) ln + 𝑅 ln
𝑇1 𝑃2 𝑇1

Copyright © Mandeep Dalal


94 A Textbook of Physical Chemistry – Volume I

𝑇2 𝑇2 𝑃1 𝑇2 (66)
= 𝐶𝑝 ln − 𝑅 ln + 𝑅 ln + 𝑅 ln
𝑇1 𝑇1 𝑃2 𝑇1

or

𝑇2 𝑃1 (67)
𝛥𝑆 = 𝐶𝑝 ln + 𝑅 ln
𝑇1 𝑃2

The equation (59) and equation (67) formulate the variation of entropy with temperature-volume and
temperature-pressure, respectively.
The variation of entropy with temperature, pressure or volume in isothermal, isobaric and isochoric
processes is discussed below.
1. Entropy change in isothermal process: Provided that the temperature of the system is kept constant (T2 =
T1 = T), equation (59) and (67) are reduced to the following.

𝑉2 𝑃1 (68)
𝛥𝑆 = 𝑅 ln = 𝑅 ln
𝑉1 𝑃2

2. Entropy change in isochoric process: Provided that the volume of the system is kept constant (V2 = V1 =
V), equation (59) is reduced to the following.

𝑇2 (69)
𝛥𝑆 = 𝐶𝑣 ln
𝑇1

3. Entropy change in isobaric process: Provided that the pressure of the system is kept constant (P2 = P1 =
P), equation (67) is reduced to the following.

𝑇2 (70)
𝛥𝑆 = 𝐶𝑝 ln
𝑇1

It is also worthy to note that if this reversible expansion takes place adiabatically, then qrev = 0, and therefore,
the change entropy will also be zero.

 Entropy Concept as a Measure of Unavailable Energy and Criteria for the


Spontaneity of Reaction
The entropy change during the course of a process can be used to rationalize the unavailable energy
as well as its spontaneity or feasibility. In this section, we will first discuss the entropy concept as a measure
of unavailable energy and then we will study the spontaneity of a process in terms of entropy change.
 Entropy Concept as a Measure of Unavailable Energy
In order to understand the connection between entropy and unavailable energy, consider a heat source
at temperature T2 placed in an atmosphere or surrounding at temperature T0. The surrounding plays the role of

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 95

heat sink here. Now if Q is the heat transferred from the source to a Carnot heat engine working between T 2
and T0, then from Carnot’s theorem, we have

𝑊𝑚𝑎𝑥,𝑇2 ,𝑇0 𝑇2 − 𝑇0 (71)


=
𝑄 𝑇2

𝑇0 (72)
𝑊𝑚𝑎𝑥,𝑇2 ,𝑇0 = 𝑄 (1 − )
𝑇2

Where 𝑊𝑚𝑎𝑥,𝑇2 ,𝑇0 is the maximum work that can be obtained from a heat engine working between T 2 and T0
temperatures. It is worthy to note that only a part of the heat transferred can be turned into work i.e. available
to be used as work.
Now suppose that the same amount of heat is first transferred directly from the source at T2 to another
source at T1 < T2. In other words, the heat Q is now first transferred from hot source to a cold source without
any Carnot bridging. Now the maximum work that can be obtained from a heat engine working between T 1
and T0 temperatures will be

𝑇0 (73)
𝑊𝑚𝑎𝑥,𝑇1 ,𝑇0 = 𝑄 (1 − )
𝑇1

Where 𝑊𝑚𝑎𝑥,𝑇1 ,𝑇0 is the maximum work that can be obtained after the direct heat transfer, since the heat engine
is now working between T1 and T0 temperatures. Now before the irreversible heat transfer (T2→T1), the
magnitude of energy that could have been transformed into work is:

𝐸𝑢𝑛𝑎𝑣𝑎𝑖𝑙𝑎𝑏𝑙𝑒 = ( 𝑊𝑚𝑎𝑥,𝑇2 ,𝑇0 ) − (𝑊𝑚𝑎𝑥,𝑇1 ,𝑇0 ) (74)

𝑇0 𝑇0 (75)
= [𝑄 (1 − )] − [𝑄 (1 − )]
𝑇2 𝑇1

𝑇0 𝑇0 (76)
= 𝑄 [(1 − ) − (1 − )]
𝑇2 𝑇1

𝑇0 𝑇0 (77)
= 𝑄( − )
𝑇1 𝑇2

𝑄 𝑄 (78)
= 𝑇0 ( − )
𝑇1 𝑇2

Since −Q/T2 and Q/T1 are the entropy loss and gain of sources at T2 and T1, respectively; the equation (78)
reduces to the following.

𝐸𝑢𝑛𝑎𝑣𝑎𝑖𝑙𝑎𝑏𝑙𝑒 = 𝑇0 (𝛥𝑆𝑠𝑜𝑢𝑟𝑐𝑒 𝑎𝑡 𝑇1 + 𝛥𝑆𝑠𝑜𝑢𝑟𝑐𝑒 𝑎𝑡 𝑇2 ) (79)

= 𝑇0 𝛥𝑆𝑖𝑟𝑟𝑒𝑣𝑒𝑟𝑠𝑖𝑏𝑙𝑒 ℎ𝑒𝑎𝑡 𝑡𝑟𝑎𝑛𝑠𝑓𝑒𝑟 (80)

Copyright © Mandeep Dalal


96 A Textbook of Physical Chemistry – Volume I

Where 𝐸𝑢𝑛𝑎𝑣𝑎𝑖𝑙𝑎𝑏𝑙𝑒 is lost work or the energy which is no longer available to do any useful work.

Figure 4. The pictorial representation of the entropy–unavailable-energy.

The physical significance of equation (80) can also be demonstrated using a live example. Let us
consider a Carnot engine working between 800K and 200K is supplied with a heat (Q) of 5000 joules. The
efficiency of such an engine will be

𝑇2 − 𝑇0 800 − 200 (81)


𝜂= = = 0.75
𝑇2 800

The maximum work obtained will be

𝑊 = 𝜂𝑄 = 0.75 × 5000 𝐽 = 3750 (82)

Now if this 5000 joules of heat is first transferred to the reservoir at 500K, and then it is fed to a Carnot engine
working obviously between “500K” and 200K. The efficiency of such an engine will be

𝑇1 − 𝑇0 500 − 200 (83)


𝜂= = = 0.6
𝑇1 500

The maximum work obtained will be

𝑊 = 𝜂𝑄 = 0.6 × 5000 𝐽 = 3000 (84)

Comparing the results of equation (82) and (84), we can say that there is 750 J less work from the same heat
transfer in the second process.

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 97

 Entropy Concept as the Criteria for the Spontaneity of Reaction


The second law of thermodynamics states that all the spontaneous processes are thermodynamically
irreversible, and accompanied by a net increase in entropy. The entropy change for a reversible process is,

𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = 0 (85)

And the entropy change for a spontaneous or irreversible process is

𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 > 0 (86)

The results of equation (85) and equation (86) can be combined to give a more generalized form as:

𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 ≥ 0 (87)

Where the sign of ʻ=ʼ stands from reversible and ʻ>ʼ stands for irreversible or spontaneous reaction,
respectively.
Nevertheless, the criteria from equation (87) is not very much practical as 𝛥𝑆𝑠𝑦𝑠𝑡𝑒𝑚 and
𝛥𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 are not easy to determine. Therefore, some other criteria must be deduced which requires only
the knowledge of the change in some simple thermodynamic parameters. In order to do so, write down the
equation (87) for infinitesimally small changes as given below.

𝑑𝑆𝑠𝑦𝑠𝑡𝑒𝑚 + 𝑑𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 ≥ 0 (88)

If δq is the heat lost by the surrounding reversibly and isothermally at temperature T, then we can say that
𝑞𝑟𝑒𝑣 (89)
𝑑𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = −
𝑇
Also, from the first law of thermodynamics, we have

𝛿𝑞𝑟𝑒𝑣 = 𝑑𝑈 + 𝑃𝑑𝑉 (90)

After putting the value from equation (90) in equation (89), we have

𝑑𝑈 + 𝑃𝑑𝑉 (91)
𝑑𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 = −
𝑇
Now putting the value of 𝑑𝑆𝑠𝑜𝑟𝑟𝑜𝑢𝑛𝑑𝑖𝑛𝑔 from equation (91) into equation (88), we get

𝑑𝑈 + 𝑃𝑑𝑉 (92)
𝑑𝑆𝑠𝑦𝑠𝑡𝑒𝑚 − ≥0
𝑇
𝑇𝑑𝑆𝑠𝑦𝑠𝑡𝑒𝑚 − 𝑑𝑈 − 𝑃𝑑𝑉 ≥ 0 (93)

𝑇𝑑𝑆 ≥ 𝑑𝑈 + 𝑃𝑑𝑉 (94)

Copyright © Mandeep Dalal


98 A Textbook of Physical Chemistry – Volume I

Where 𝑑𝑆𝑠𝑦𝑠𝑡𝑒𝑚 = 𝑑𝑆 i.e. change in the entropy of the system. For a spontaneous or irreversible process 𝑇𝑑𝑆
must be greater the sum of internal energy change and pressure-volume work.
However, the above relation is reduced to the following if the internal energy and volume of the
system are kept constant i.e. dU = 0 and dV = 0.

(𝑇𝑑𝑆)𝑈,𝑉 ≥ 0 (95)

or

(𝑑𝑆)𝑈,𝑉 ≥ 0 (96)

Here it is worthy to recall that the sign ʻ=ʼ is for reversible reactions whereas ʻ>ʼ condition is applicable to
spontaneity or irreversibility of a reaction.
Besides, the spontaneity of equation (94) can also be written in the form of internal energy change,

𝑑𝑈 ≤ 𝑇𝑑𝑆 − 𝑃𝑑𝑉 (97)

the above relation is reduced to the following if the internal energy and volume of the system are kept constant
i.e. dS = 0 and dV = 0.

(𝑑𝑈)𝑆,𝑉 ≤ 0 (98)

Again it should be noted that the sign ʻ=ʼ is for reversible reactions whereas ʻ<ʼ condition is applicable to
spontaneity or irreversibility of a reaction.

 Free Energy, Enthalpy Functions and Their Significance, Criteria for


Spontaneity of a Process
The thermodynamic free energy and enthalpies are extremely useful concepts in the field of chemical
thermodynamics. The decrease or increase in free energy is the maximum amount of work that a
thermodynamic system can perform in a process at a constant temperature. The sign of thermodynamic free
energy simply indicates whether a process is thermodynamically forbidden or feasible. The thermodynamic
free energy is a state function, like internal energy and enthalpy. In this section, we will discuss the significance
of enthalpy and free energy functions, and the corresponding spontaneity of a process.
 Enthalpy or Heat Content
As most of the reactions are carried out in open vessels where the atmospheric pressure remains the
same, it must be very interesting to study the heat change that occurs during the course of a chemical reaction.
The work-done by gas in the piston-fitted chamber against constant pressure (P) with ΔV volume change is

𝑤 = −𝑃𝛥𝑉 (99)

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 99

From the first law of thermodynamics, we know that

𝑞 = 𝛥𝑈 − 𝑤 (100)

Where q is the heat absorbed and ΔU is the change in the internal energy of the system respectively. After
putting the value of work of expansion from equation (99) in equation (100), we get

𝑞𝑝 = 𝛥𝑈 − (−𝑃𝛥𝑉) (101)

𝑞𝑝 = 𝛥𝑈 + 𝑃𝛥𝑉 (102)

The symbol qp is used instead of q because the whole process is carried out at constant pressure. Now suppose
that this heat absorbed increases the internal energy from U1 to U2 and volume from V1 to V2 i.e.

𝛥𝑈 = 𝑈2 − 𝑈1 (103)

𝛥𝑉 = 𝑉2 − 𝑉1 (104)

After putting the values from equation (103) and (104) into equation (102), we get

𝑞𝑝 = (𝑈2 − 𝑈1 ) + 𝑃(𝑉2 − 𝑉1 ) (105)

or

𝑞𝑝 = (𝑈2 + 𝑃𝑉2 ) − (𝑈1 + 𝑃𝑉1 ) (106)

The quantity U + PV is now defined as the enthalpy of the system; and since U, P and V all are state functions,
the quantity U + PV must also be a state function. Mathematically, the enthalpy can be shown as

𝐻 = 𝑈 + 𝑃𝑉 (107)

Therefore, if the enthalpy changes initial state H1 to the final state H2 at constant pressure, we have

𝐻1 = 𝑈1 + 𝑃𝑉1 (108)

𝐻2 = 𝑈2 + 𝑃𝑉2 (109)

After putting the values from equation (108) and (109) in equation (106), we get

𝑞𝑝 = 𝐻2 − 𝐻1 (110)

𝑞𝑝 = 𝛥𝐻 (111)

Hence, we can conclude that the enthalpy change of a system is simply the amount of heat absorbed at constant
pressure. Besides, if we compare equation (102) and equation (111), we have

𝛥𝐻 = 𝛥𝑈 + 𝑃𝛥𝑉 (112)

Copyright © Mandeep Dalal


100 A Textbook of Physical Chemistry – Volume I

Therefore, the enthalpy change of a reaction may also be defined as the sum of its internal energy change and
pressure-volume work done.
The physical significance of enthalpy: There is a certain amount of energy stored in every substance or
material called internal energy. This energy comprises of many forms like kinetic, rotational, vibrational or
inter-particle interactions. Moreover; like pressure, volume and internal energy; the enthalpy function is also
an extensive property. Now recalling the expression of enthalpy again

𝐻 = 𝑈 + 𝑃𝑉 (113)

We can say that the enthalpy is nothing but the internal energy that is available for the conversion into heat;
which is why the enthalpy is also called as “heat content”. Furthermore, like internal energy, the heat content
or enthalpy is also not obtainable absolutely. However, like some other thermodynamic quantities, the
parameter that is needed in the various analysis is enthalpy change which can be derived experimentally.
Enthalpy as the criteria for the spontaneity of reaction: The general expression for the enthalpy of a system
is given below.

𝐻 = 𝑈 + 𝑃𝑉 (114)

After differentiating the equation (114), we get

𝑑𝐻 = 𝑑𝑈 + 𝑉𝑑𝑃 + 𝑃𝑑𝑉 (115)

𝑑𝐻 − 𝑉𝑑𝑃 = 𝑑𝑈 + 𝑃𝑑𝑉 (116)

Also as we know that, for the spontaneity of a process

𝑇𝑑𝑆 ≥ 𝑑𝑈 + 𝑃𝑑𝑉 (117)

Now after putting the value of 𝑑𝑈 + 𝑉𝑑𝑃 from equation (116) in equation (117), we get

𝑇𝑑𝑆 ≥ 𝑑𝐻 − 𝑉𝑑𝑃 (118)

or

𝑑𝐻 ≤ 𝑇𝑑𝑆 + 𝑉𝑑𝑃 (119)

For a spontaneous or irreversible process 𝑑𝐻 must be less than the sum of multiplication of temperature and
entropy change, and pressure-volume work.
However, the above relation is reduced to the following if the entropy and pressure of the system are
kept constant i.e. dS = 0 and dP = 0.

(𝑑𝐻)𝑆,𝑃 ≤ 0 (120)

Here it is worthy to recall that the sign ʻ=ʼ is for reversible reactions whereas ʻ<ʼ condition is applicable to
spontaneity or irreversibility of a reaction.

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 101

 Helmholtz Free Energy or Work Function


The Helmholtz free energy is typically denoted by the symbol ʻAʼ, which is derived from the German
word “Arbeit” meaning work. The Helmholtz free energy can be defined mathematically as

𝐴 = 𝑈 − 𝑇𝑆 (121)

Where T, S and U are temperature, entropy, and internal energy, respectively. Moreover; like U, T and S; free
energy A is also a state function. Since A is independent of its previous state, we can say that

𝐴1 = 𝑈1 − 𝑇𝑆1 (122)

𝐴2 = 𝑈2 − 𝑇𝑆2 (123)

Where the subscript 1 and 2 represent the initial and final state. The change in Helmholtz free energy in going
from initial to final state can obtain by subtracting equation (122) from equation (123) as

𝐴2 − 𝐴1 = (𝑈2 − 𝑇𝑆2 ) − (𝑈1 − 𝑇𝑆1 ) (124)

𝛥𝐴 = (𝑈2 − 𝑈1 ) − 𝑇(𝑆2 − 𝑆1 ) (125)

𝛥𝐴 = 𝛥𝑈 − 𝑇𝛥𝑆 (126)

Where ΔU and ΔS are the change in internal energy and entropy, respectively. Therefore, the Helmholtz free
energy change can be described as the difference of internal energy change and the multiplication of entropy
change multiplied with the temperature at which the reaction is actually carried out.
The physical significance of Helmholtz free energy: From the definition of entropy change, we know that
𝑞𝑟𝑒𝑣 (127)
𝛥𝑆 =
𝑇
Also, from the first law of thermodynamics, for the work of expansion we have

∆𝑈 = 𝑞𝑟𝑒𝑣 − 𝑤𝑚𝑎𝑥 (128)

Putting the values of ΔS and ΔU from equation (127) and (128) in equation (126), we get
𝑞𝑟𝑒𝑣 (129)
𝛥𝐴 = (𝑞𝑟𝑒𝑣 − 𝑤𝑚𝑎𝑥 ) − 𝑇 ( )
𝑇
= 𝑞𝑟𝑒𝑣 − 𝑤𝑚𝑎𝑥 − 𝑞𝑟𝑒𝑣

or

−𝛥𝐴 = 𝑤𝑚𝑎𝑥 (130)

Hence, the decrease in Helmholtz free energy at constant temperature is equal to the maximum work done by
the system; that is why the Helmholtz free energy is also called as work function.

Copyright © Mandeep Dalal


102 A Textbook of Physical Chemistry – Volume I

Helmholtz free energy as the criteria for the spontaneity of reaction: The general expression for the
enthalpy of a system is given below.

𝐴 = 𝑈 − 𝑇𝑆 (131)

After differentiating the equation (131), we get

𝑑𝐴 = 𝑑𝑈 − 𝑇𝑑𝑆 − 𝑆𝑑𝑇 (132)

𝑇𝑑𝑆 = 𝑑𝑈 − 𝑆𝑑𝑇 − 𝑑𝐴 (133)

Also as we know that, for the spontaneity of a process

𝑇𝑑𝑆 ≥ 𝑑𝑈 + 𝑃𝑑𝑉 (134)

Now after putting the value of 𝑇𝑑𝑆 from equation (133) in equation (134), we get

𝑑𝑈 − 𝑆𝑑𝑇 − 𝑑𝐴 ≥ 𝑑𝑈 + 𝑃𝑑𝑉 (135)

or

−𝑆𝑑𝑇 − 𝑑𝐴 ≥ 𝑃𝑑𝑉 (136)

𝑑𝐴 ≤ −𝑃𝑑𝑉 − 𝑆𝑑𝑇 (137)

For a spontaneous or irreversible process, 𝑑𝐴 must be less than the negative sum of multiplication of pressure
and volume change with the multiplication of entropy and temperature change.
However, the above relation is reduced to the following if the volume and temperature of the system
are kept constant i.e. dV = 0 and dT = 0.

(𝑑𝐴)𝑉,𝑇 ≤ 0 (138)

Here it is worthy to recall that the sign ʻ=ʼ is for reversible reactions whereas ʻ<ʼ condition is applicable to
spontaneity or irreversibility of a reaction.
 Gibbs Free Energy or Gibbs Function
The Gibbs free energy is typically denoted by the symbol ʻGʼ, and can be defined mathematically as

𝐺 = 𝐻 − 𝑇𝑆 (139)

Where T, S and H are temperature, entropy, and enthalpy, respectively. Moreover; like H, T and S; the Gibbs
free energy G is also a state function. Since G is independent of its previous state, we can say that

𝐺1 = 𝐻1 − 𝑇𝑆1 (140)

𝐺2 = 𝐻2 − 𝑇𝑆2 (141)

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 103

Where the subscript 1 and 2 represent the initial and final state. The change in Gibbs free energy in going from
initial to final state can obtain by subtracting equation (141) from equation (140) as

𝐺2 − 𝐺1 = (𝐻2 − 𝑇𝑆2 ) − (𝐻1 − 𝑇𝑆1 ) (142)

𝛥𝐺 = (𝐻2 − 𝐻1 ) − 𝑇(𝑆2 − 𝑆1 ) (143)

𝛥𝐺 = 𝛥𝐻 − 𝑇𝛥𝑆 (144)

Where ΔH and ΔS are the change in enthalpy and entropy, respectively. Therefore, the Gibbs free energy
change can be described as the difference of enthalpy change and the multiplication of entropy change
multiplied with the temperature at which the reaction is actually carried out.
The physical significance of Helmholtz free energy: From the definition of entropy change, we know that
𝑞𝑟𝑒𝑣 (145)
𝛥𝑆 =
𝑇
Also, at constant pressure, we have

∆𝐻 = 𝛥𝑈 + 𝑃𝛥𝑉 (146)

Putting the values of ΔS and ΔH from equation (145) and (146) in equation (144), we get
𝑞𝑟𝑒𝑣 (147)
𝛥𝐺 = (𝛥𝑈 + 𝑃𝛥𝑉) − 𝑇 ( )
𝑇
= (𝛥𝑈 − 𝑞𝑟𝑒𝑣 ) + 𝑃𝛥𝑉 (148)

Also, from the first of thermodynamics, for the work of expansion we have

∆𝑈 − 𝑞𝑟𝑒𝑣 = −𝑤𝑚𝑎𝑥 (149)

Now, after putting the value of equation (149) in equation (148), we get

𝛥𝐺 = −𝑤𝑚𝑎𝑥 + 𝑃𝛥𝑉 (150)

−𝛥𝐺 = 𝑤𝑚𝑎𝑥 − 𝑃𝛥𝑉 (151)

Hence, the decrease in Gibbs free energy for the process occurring at constant pressure and constant
temperature is equal to the “maximum net work” that can be obtained from the process. The term “maximum
net work” refers to maximum work other than the work of expansion.
Gibbs free energy as the criteria for the spontaneity of reaction: The general expression for the enthalpy
of a system is given below.

𝐺 = 𝐻 − 𝑇𝑆 (152)

Since 𝐻 = 𝑈 + 𝑃𝑉, equation (152) takes the form

Copyright © Mandeep Dalal


104 A Textbook of Physical Chemistry – Volume I

𝐺 = 𝑈 + 𝑃𝑉 − 𝑇𝑆 (153)

After differentiating the equation (153), we get

𝑑𝐺 = 𝑑𝑈 + 𝑃𝑑𝑉 + 𝑉𝑑𝑃 − 𝑇𝑑𝑆 − 𝑆𝑑𝑇 (154)

𝑇𝑑𝑆 = 𝑑𝑈 + 𝑃𝑑𝑉 + 𝑉𝑑𝑃 − 𝑆𝑑𝑇 − 𝑑𝐺 (155)

Also as we know that, for the spontaneity of a process

𝑇𝑑𝑆 ≥ 𝑑𝑈 + 𝑃𝑑𝑉 (156)

Now after putting the value of 𝑇𝑑𝑆 from equation (155) in equation (156), we get

𝑑𝑈 + 𝑃𝑑𝑉 + 𝑉𝑑𝑃 − 𝑆𝑑𝑇 − 𝑑𝐺 ≥ 𝑑𝑈 + 𝑃𝑑𝑉 (156)

or

𝑉𝑑𝑃 − 𝑆𝑑𝑇 − 𝑑𝐺 ≥ 0 (157)

𝑑𝐺 ≤ 𝑉𝑑𝑃 − 𝑆𝑑𝑇 (158)

For a spontaneous or irreversible process, 𝑑𝐺 must be less than the sum of negative multiplication of volume
and pressure change with the multiplication of entropy and temperature change.
However, the above relation is reduced to the following if the pressure and temperature of the system
are kept constant i.e. dP = 0 and dT = 0.

(𝑑𝐺)𝑃,𝑇 ≤ 0 (159)

Here it is worthy to recall that the sign ʻ=ʼ is for reversible reactions whereas ʻ<ʼ condition is applicable to
spontaneity or irreversibility of a reaction.

 Partial Molar Quantities (Free Energy, Volume, Heat Concept)


So far we have discussed the variation of various thermodynamic properties with respect to
temperature and pressure while the composition of the system was kept constant (closed system). In 1907,
G.N. Lewis started the study of open systems i.e. the variation of various thermodynamic properties with
respect to the composition of one or more components. In other words, he studied the behavior of a particular
thermodynamic property of the system when a component is removed from or added to the system under
consideration. Now since a variation like this is observable only for an extensive property, the general
definition of partial molar properties can be given as given below.
A partial molar property may simply be defined as a thermodynamic quantity which indicates how
an extensive property of a solution or mixture changes with the variation in the molar composition of the
mixture at constant temperature and pressure.

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 105

Basically, it is the partial derivative of the extensive property with respect to the number of moles of
the component under consideration. All extensive properties of a mixture have corresponding partial molar
properties. In this section, we will discuss some very important partial molar quantities like partial molar free
energy (𝐺̅𝑖 ), partial molar volume (𝑉 ̅𝑖 ) and partial molar enthalpy (𝐻
̅̅̅𝑖 ).

 Partial Molar Free Energy or Chemical Potential


In order to derive the expression for partial molar free energy, consider a system that comprises of n
types of constituents with n1, n2, n3, n4 … moles. So, being an extensive property, the partial molar free energy
depends upon not only the temperature and pressure but also on the number of moles of different components.
Mathematically, we can say that

𝐺 = 𝑓(𝑇, 𝑃, 𝑛1 , 𝑛2 , 𝑛3 … ) (160)

Now let us assume a small change in the temperature, pressure and amount of different components, this would
impart a variation in partial molar free energy as given below.

𝜕𝐺 𝜕𝐺 𝜕𝐺 (161)
𝑑𝐺 = ( ) 𝑑𝑇 + ( ) 𝑑𝑃 + ( ) 𝑑𝑛1 + ⋯
𝜕𝑇 𝑃,𝑛1 ,𝑛2 … 𝜕𝑃 𝑇,𝑛1 ,𝑛2 … 𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

The first term on the right-hand side gives the change in the free energy with temperature at constant pressure
and compositions; while the second term gives the change in the free energy with pressure at constant
temperature and compositions. The terms afterward represent the variation in free energy with the amount of
one component while the temperature, pressure and all other compositions are kept constant.
However, if the temperature and pressure of the system are kept constant i.e. dT = 0, dP = 0, the
equation (161) takes the form

𝜕𝐺 𝜕𝐺 𝜕𝐺 (162)
(𝑑𝐺) 𝑇,𝑃 = ( ) 𝑑𝑛1 + ( ) 𝑑𝑛2 + ( ) 𝑑𝑛3 …
𝜕𝑛1 𝑇,𝑃,𝑛 𝜕𝑛 2 𝑇,𝑃,𝑛 𝜕𝑛3 𝑇,𝑃,𝑛
2 ,𝑛3 … 1 ,𝑛3 … 1 ,𝑛2 …

Every term on the right-hand side of the equation (162) is partial molar free energy and is symbolized by a
“bar” over it i.e.

(𝑑𝐺) 𝑇,𝑃 = ̅̅̅


𝐺1 𝑑𝑛1 + ̅̅̅
𝐺2 𝑑𝑛2 + ̅̅̅
𝐺3 𝑑𝑛3 … (163)

For the ith component, we can say that

𝜕𝐺 (164)
𝐺̅𝑖 = ( )
𝜕𝑛𝑖 𝑇,𝑃,𝑛
2 ,𝑛3 …

The equation (164) gives the general expression for “partial molar free energy” or the “chemical potential” of
the ith species.

Copyright © Mandeep Dalal


106 A Textbook of Physical Chemistry – Volume I

 Partial Molar Volume


In order to derive the expression for partial molar volume, consider a system that comprises of n types
of constituents with n1, n2, n3, n4 … moles. So, being an extensive property, volume depends upon not only the
temperature and pressure but also on the number of moles of different components. Mathematically, we can
say that

𝑉 = 𝑓(𝑇, 𝑃, 𝑛1 , 𝑛2 , 𝑛3 … ) (165)

Now let us assume a small change in the temperature, pressure and amount of different components, this would
impart a variation in partial molar volume as given below.

𝜕𝑉 𝜕𝑉 𝜕𝑉 (166)
𝑑𝑉 = ( ) 𝑑𝑇 + ( ) 𝑑𝑃 + ( ) 𝑑𝑛1 + ⋯
𝜕𝑇 𝑃,𝑛1 ,𝑛2 … 𝜕𝑃 𝑇,𝑛1 ,𝑛2 … 𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

The first term on the right-hand side gives the change in the volume with the temperature at constant pressure
and compositions; while the second term gives the change in the volume with pressure at constant temperature
and compositions. The terms afterward represent the variation in volume with the amount of one component
while the temperature, pressure and all other compositions are kept constant.
However, if the temperature and pressure of the system are kept constant i.e. dT = 0, dP = 0, the
equation (166) takes the form

𝜕𝑉 𝜕𝑉 𝜕𝑉 (167)
(𝑑𝑉) 𝑇,𝑃 = ( ) 𝑑𝑛1 + ( ) 𝑑𝑛2 + ( ) 𝑑𝑛3 …
𝜕𝑛1 𝑇,𝑃,𝑛 𝜕𝑛2 𝑇,𝑃,𝑛 𝜕𝑛3 𝑇,𝑃,𝑛
2 ,𝑛3 … 1 ,𝑛3 … 1 ,𝑛2 …

Every term on the right-hand side of the equation (167) is partial molar volume and is symbolized by a “bar”
over it i.e.

𝜕𝑉 (168)
𝑉̅1 = ( )
𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

̅̅̅
𝜕𝑉 (169)
𝑉2 = ( )
𝜕𝑛2 𝑇,𝑃,𝑛
1 ,𝑛3 …

After putting the values from equations like (168 – 169) in equation (167), we get

(𝑑𝑉) 𝑇,𝑃 = 𝑉̅1 𝑑𝑛1 + ̅̅̅


𝑉2 𝑑𝑛2 + ̅̅̅
𝑉3 𝑑𝑛3 … (170)

For the ith component, we can say that

̅𝑖 = (
𝜕𝑉 (171)
𝑉 )
𝜕𝑛𝑖 𝑇,𝑃,𝑛
2 ,𝑛3 …

The equation (171) gives the general expression for the “partial molar volume” of the ith species.

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 107

 Partial Molar Enthalpy or Partial Molar Heat Content


In order to derive the expression for partial molar enthalpy, consider a system that comprises of n
types of constituents with n1, n2, n3, n4 … moles. So, being an extensive property, volume depends upon not
only the temperature and pressure but also on the number of moles of different components, i.e.,

𝐻 = 𝑓(𝑇, 𝑃, 𝑛1 , 𝑛2 , 𝑛3 … ) (172)

Now let us assume a small change in the temperature, pressure and amount of different components, this would
impart a variation in molar enthalpy as given below.

𝜕𝐻 𝜕𝐻 𝜕𝐻 (173)
𝑑𝐻 = ( ) 𝑑𝑇 + ( ) 𝑑𝑃 + ( ) 𝑑𝑛1 + ⋯
𝜕𝑇 𝑃,𝑛1 ,𝑛2 … 𝜕𝑃 𝑇,𝑛1 ,𝑛2 … 𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

The first term on the right-hand side gives the change in the enthalpy with the temperature at constant pressure
and compositions; while the second term gives the change in the enthalpy with pressure at constant temperature
and compositions. The terms afterward represent the variation in enthalpy with the amount of one component
while the temperature, pressure and all other compositions are kept constant.
However, if the temperature and pressure of the system are kept constant i.e. dT = 0, dP = 0, the
equation (173) takes the form

𝜕𝐻 𝜕𝐻 𝜕𝐻 (174)
(𝑑𝐻) 𝑇,𝑃 = ( ) 𝑑𝑛1 + ( ) 𝑑𝑛2 + ( ) 𝑑𝑛3 …
𝜕𝑛1 𝑇,𝑃,𝑛 𝜕𝑛2 𝑇,𝑃,𝑛 𝜕𝑛3 𝑇,𝑃,𝑛
2 ,𝑛3 … 1 ,𝑛3 … 1 ,𝑛2 …

Every term on the right-hand side of the equation (174) is partial molar enthalpy and is symbolized by a “bar”
over it i.e.

̅𝐻̅̅1̅ = (
𝜕𝐻 (175)
)
𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

̅𝐻̅̅2̅ = (
𝜕𝐻 (176)
)
𝜕𝑛2 𝑇,𝑃,𝑛
1 ,𝑛3 …

After putting the values from equations like (175 – 176) in equation (174), we get

(𝑑𝐻) 𝑇,𝑃 = ̅𝐻̅̅1̅ 𝑑𝑛1 + ̅𝐻̅̅2̅ 𝑑𝑛2 + ̅𝐻̅̅3̅ 𝑑𝑛3 … (177)

For the ith component, we can say that

̅̅̅
𝜕𝐻 (178)
𝐻𝑖 = ( )
𝜕𝑛𝑖 𝑇,𝑃,𝑛
2 ,𝑛3 …

The equation (178) gives the general expression for the “partial molar enthalpy” of the ith species.

Copyright © Mandeep Dalal


108 A Textbook of Physical Chemistry – Volume I

 Gibb’s-Duhem Equation
In order to derive the expression for Gibbs’s-Duhem equation, consider a system that comprises of n
types of constituents with n1, n2, n3, n4 … moles. So, being an extensive property, the partial molar free energy
depends upon not only the temperature and pressure but also on the number of moles of different components.
Mathematically, we can say that

𝐺 = 𝑓(𝑇, 𝑃, 𝑛1 , 𝑛2 , 𝑛3 … ) (179)

Now let us assume a small change in the temperature, pressure and amount of different components, this would
impart a variation in partial molar free energy as given below.

𝜕𝐺 𝜕𝐺 𝜕𝐺 (180)
𝑑𝐺 = ( ) 𝑑𝑇 + ( ) 𝑑𝑃 + ( ) 𝑑𝑛1 + ⋯
𝜕𝑇 𝑃,𝑛1 ,𝑛2 … 𝜕𝑃 𝑇,𝑛1 ,𝑛2 … 𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

The first term on the right-hand side gives the change in the free energy with the temperature at constant
pressure and compositions; while the second term gives the change in the free energy with pressure at constant
temperature and compositions. The terms afterward represent the variation in free energy with the amount of
one component while the temperature, pressure and all other compositions are kept constant.
However, if the temperature and pressure of the system are kept constant, i.e., dT = 0, dP = 0, the
equation (180) takes the form

𝜕𝐺 𝜕𝐺 𝜕𝐺 (181)
(𝑑𝐺) 𝑇,𝑃 = ( ) 𝑑𝑛1 + ( ) 𝑑𝑛2 + ( ) 𝑑𝑛3 …
𝜕𝑛1 𝑇,𝑃,𝑛 𝜕𝑛2 𝑇,𝑃,𝑛 𝜕𝑛3 𝑇,𝑃,𝑛
2 ,𝑛3 … 1 ,𝑛3 … 1 ,𝑛2 …

Every term on the right-hand side of the equation (181) is partial molar free energy or simply the “chemical
potential” i.e.

𝜕𝐺 (182)
𝜇1 = ( )
𝜕𝑛1 𝑇,𝑃,𝑛
2 ,𝑛3 …

𝜕𝐺 (183)
𝜇2 = ( )
𝜕𝑛2 𝑇,𝑃,𝑛
1 ,𝑛3 …

𝜕𝐺 (184)
𝜇3 = ( )
𝜕𝑛3 𝑇,𝑃,𝑛
1 ,𝑛2 …

𝜕𝐺 (185)
𝜇4 = ( )
𝜕𝑛4 𝑇,𝑃,𝑛
1 ,𝑛2 …

𝜕𝐺 (186)
𝜇5 = ( )
𝜕𝑛5 𝑇,𝑃,𝑛
1 ,𝑛2 …

After putting the values from equations like (182 – 186) in equation (181), we get

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 109

(𝑑𝐺) 𝑇,𝑃 = 𝜇1 𝑑𝑛1 + 𝜇2 𝑑𝑛2 + 𝜇3 𝑑𝑛3 + 𝜇4 𝑑𝑛4 + 𝜇5 𝑑𝑛5 … (187)

Now, if the system composition is definite, the integration of the above equation gives

𝐺𝑇,𝑃,𝑁 = 𝜇1 𝑛1 + 𝜇2 𝑛2 + 𝜇3 𝑛3 + 𝜇4 𝑑𝑛4 + 𝜇5 𝑑𝑛5 … (188)

The differentiation of equation (188) at constant temperature and constant pressure but changing composition
gives the following relation

(𝑑𝐺) 𝑇,𝑃 = (𝑛1 𝑑𝜇1 + 𝜇1 𝑑𝑛1 ) + (𝑛2 𝑑𝜇2 + 𝜇2 𝑑𝑛2 ) + (𝑛3 𝑑𝜇3 + 𝜇3 𝑑𝑛3 ) (189)
+ (𝑛4 𝑑𝜇4 + 𝜇4 𝑑𝑛4 ) + (𝑛5 𝑑𝜇5 + 𝜇5 𝑑𝑛5 ) …

or

(𝑑𝐺) 𝑇,𝑃 = (𝜇1 𝑑𝑛1 + 𝜇2 𝑑𝑛2 + 𝜇3 𝑑𝑛3 + 𝜇4 𝑑𝑛4 + 𝜇5 𝑑𝑛5 … ) (190)


+ (𝑛1 𝑑𝜇1 + 𝑛2 𝑑𝜇2 + 𝑛3 𝑑𝜇3 + 𝑛4 𝑑𝜇4 + 𝑛5 𝑑𝜇5 … )

After comparing the equation (187) and (190), we can conclude that the content included in the second bracket
must be equal to zero. Mathematically, we can say that

𝑛1 𝑑𝜇1 + 𝑛2 𝑑𝜇2 + 𝑛3 𝑑𝜇3 … = 0 (191)

or

∑ 𝑛𝑖 𝑑𝜇𝑖 = 0 (192)

Which is the popular Gibbs-Duhem equation, and is applicable to the systems under constant temperature-
pressure conditions.
The physical significance of the Gibbs-Duhem equation can be understood by taking the example of
binary solutions i.e. a system of two components only. The Gibbs-Duhem equation for such systems is

𝑛1 𝑑𝜇1 + 𝑛2 𝑑𝜇2 = 0 (193)

or

𝑛1 𝑑𝜇1 = −𝑛2 𝑑𝜇2 (194)

or
𝑛2 (195)
𝑑𝜇1 = − 𝑑𝜇2
𝑛1

Hence, the chemical potential of one constituent is not independent of another component in binary solutions.
In other words, the chemical potentials or partial molar free energies of two components of the binary system
are mutually dependent; and if the one increases the other one decreases.

Copyright © Mandeep Dalal


110 A Textbook of Physical Chemistry – Volume I

It is also worthy to note that if the number of moles of different constituents remains constant (closed
system), i.e., 𝑑𝑛𝑖 = 0; equation (180) reduces to the following.

𝜕𝐺 𝜕𝐺 (196)
𝑑𝐺 = ( ) 𝑑𝑇 + ( ) 𝑑𝑃
𝜕𝑇 𝑃,𝑁 𝜕𝑃 𝑇,𝑁

Also, for a closed system, we know that

𝑑𝐺 = 𝑉𝑑𝑃 − 𝑆𝑑𝑇 (197)

Which means that

𝜕𝐺 𝜕𝐺 (198)
( ) = −𝑆 𝑎𝑛𝑑 ( ) =𝑉
𝜕𝑇 𝑃,𝑁 𝜕𝑃 𝑇,𝑁

Which is the variation of free energy with temperature and pressure in closed systems. These two relations can
further be used to deduce the variations of chemical potentials with temperature and volume.

Copyright © Mandeep Dalal


CHAPTER 2 Thermodynamics – I 111

 Problems
Q 1. Define the first law of thermodynamics with pictorial description.
Q 2. Discuss the limitations of the first law of thermodynamics and the need for the second law.
Q 3. What is entropy? How does it behave in reversible and irreversible reactions?
Q 4. Comment on the statement “all spontaneous processes are thermodynamically irreversible”.
Q 5. Derive the relation showing the variation of entropy with temperature and pressure.
Q 6. What do you mean by “unavailable energy” or the “lost work”?
Q 7. Discuss the spontaneity of a process in terms of entropy change.
Q 8. What is Gibbs free energy? What are its significances?
Q 9. Define the work function and discuss the related spontaneity science.
Q 10. What are partial molar quantities? Discuss with the special reference of chemical potential.
Q 11. Derive Gibbs-Duhem equation.

Copyright © Mandeep Dalal


112 A Textbook of Physical Chemistry – Volume I

 Bibliography
[1] G. Raj, Thermodynamics, Krishna Prakashan, Meerut, India, 2004.
[2] B. R. Puri, L. R. Sharma, M. S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2008.
[3] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[4] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[5] S. Carnot, R. Fox, Reflections on the Motive Power of Fire: a Critical Edition with the Surviving Scientific
Manuscripts, Manchester University Press, New York, USA, 1986.
[6] P. A. Rock, Chemical Thermodynamics, University Science Books, California, USA, 1983.
[7] E. B. Smith, Basic Chemical Thermodynamics, Imperial College Press, London, UK, 2014.

Copyright © Mandeep Dalal


CHAPTER 3
Chemical Dynamics – I
 Effect of Temperature on Reaction Rates
The temperature of the system shows a very marked effect on the overall rate of the reaction. In fact,
it has been observed that the rate of a chemical reaction typically gets doubled with every 10°C rise in the
temperature. However, this ratio may differ considerably and may reach up to 3 for different reactions. Besides,
this ratio also varies as the temperature of the reaction increases gradually. The ratio of rate constant at two
different temperatures is called as “temperature coefficient” of the reaction. Although we can determine the
temperature coefficient between any two temperatures for any chemical reaction, generally it is calculated for
10°C difference.

𝑘 𝑇+10 (1)
𝑇𝑒𝑚𝑝𝑒𝑟𝑎𝑡𝑢𝑟𝑒 𝐶𝑜𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡 = =2−3
𝑘𝑇

Where 𝑘 𝑇 and 𝑘 𝑇+10 are rate constants at temperature T and T+10, respectively. Now, if once the temperature
coefficient is known, you can determine the relative increase or decrease in the overall reaction-rate by using
the following relation.

𝑅2 𝑘2 𝑇2 −𝑇1 (2)
= = (𝑇𝑒𝑚𝑝𝑒𝑟𝑎𝑡𝑢𝑟𝑒 𝐶𝑜𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑡) ∆𝑇𝑡𝑐
𝑅1 𝑘1

Where R2 and R1 are the reaction-rates at temperatures T2 and T1, respectively. The ∆𝑇𝑡𝑐 is the temperature
range for the temperature coefficient.
In order to illustrate the dominance of the effect of temperature change on the reaction rate, consider
a reaction in which the temperature of the system is raised from 310°C to 400°C. Now, if the temperature
coefficient for 10°C temperature-rise is 2, the relative increase in the rate constant or rate will be

𝑅2 𝑘2 400−310 (3)
= = (2) 10
𝑅1 𝑘1

𝑅2 𝑘2 (4)
= = (2)9 = 512
𝑅1 𝑘1

𝑅2 (5)
= 512
𝑅1

Hence, a 90°C rise in temperature increases the rate of reaction 512 times, which is definitely huge. Now the
question arises, why is it so? What did the temperature do that made this happen? In this section, we will
answer these questions.

Copyright © Mandeep Dalal


114 A Textbook of Physical Chemistry – Volume I

 Fundamentals of Temperature-Rate Correlation


Before we discuss the effect of temperature on the reaction rate, we must understand the cause of a
reaction itself first. The primary requirement for a reaction to occur is the collision between the reacting
molecules. In other words, the reactant molecules must collide with each other to form the product. Therefore,
if we assume that every collision results in the formation of the product, the rate of reaction should simply be
equal to the collision frequency of the reacting system.
For a reaction between A and B, the collision frequency (Z) is the number of collisions between A and
B occurring in the container per unit volume per unit time.

(6)
8𝑘𝐵 𝑇
𝑍 = 𝑛𝐴 𝑛𝐵 𝛩𝐴𝐵 √
𝜋𝜇𝐴𝐵

Where nA and nA are the number densities (in the units of m−3) of particles A and B, respectively. The term ΘAB
is the reaction cross-section (in m2) when particle A with radius rA and B with radius rB collide with each other
i.e. ΘAB = π (σAB)2 = π(rA + rB)2 = π(σA/2 + σB/2)2. kB is the Boltzmann's constant (m2 kg s−2 K−1). T represents
the temperature of the system. The term μAB represents the reduced mass of the reactants A and B i.e. μAB =
mAmB/mA+mB. From equation (6), it follows that when we heat the substance, the particles collide more
frequently and hence increase the collision frequency. Now one may think that this collision frequency would
result in a larger rate of reaction, and therefore, the mystery is solved. However, this isn’t sufficient to
rationalize the experimental observations. For instance, if we increase the temperature from 300 K to 310 K,
the relative increase in the collision frequency (nZ), and hence in reaction rate, from equation (6) can be
determined as given below.

(7)
310
𝑛𝑍 = √ = 1.0165
300

This is only 1.65% increase for a 10° rise in temperature. This is pretty far from the reality i.e. reaction-rate
almost gets double (100% increase). So, the actual mechanism is still behind the scene and must be understood.
At this point we must introduce the concept of activation energy, otherwise, the concept cannot be
discussed further. The collision of reacting molecules would result in the chemical reaction only if they possess
a certain amount of minimum energy i.e. threshold energy. Since every molecule does have some energy, the
energy it needs to reach the threshold is less than the actual threshold energy. The energy required by reactant
molecules to cross the barrier is called the activation energy or the enthalpy of activation for the reaction. A
simple equation can be used to deduce their relationships as given below.

Activation energy = Threshold energy – Energy actually possessed by the molecules

The rate of a chemical reaction is inversely proportional to the magnitude of the activation energy i.e. larger
the activation energy, slower will be the reaction and vice-versa.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 115

Figure 1. The reaction coordinate diagram for a typical chemical reaction.

Hence, we can say that only effective collisions would result in the chemical reactions, but how can we find
the number of molecules having energy high enough to react with each other. For this, we need to go into the
basics of energy distribution among a large number of particles i.e. Maxwell’s distribution of energies.

Figure 2. The Maxwell-Boltzmann distribution of energies at temperature T.

After marking the activation energy on the Maxwell-Boltzmann distribution curve, the particle with sufficient
energy to react can easily be found from the area under the corresponding curve i.e. dashed area. The undashed
area at a particular temperature is quite large, and therefore, represents the number particle whose collision
would not result in any reaction chemical change. It can be clearly seen that most of the particles don't have

Copyright © Mandeep Dalal


116 A Textbook of Physical Chemistry – Volume I

enough energy, and hence, are unable to yield the product. The reaction-rate will be very small If there are
very few particles with enough energy at any time.
However, if the temperature is raised, the maxima of the Maxwell-Boltzmann distribution curve shifts
towards higher energy. This makes the number of “efficient particles” to increase and thereby increases the
number of effective collisions too. Consider the Maxwell-Boltzmann energy distributions at temperature T and
T+10.

Figure 3. The Maxwell-Boltzmann distribution of energies at temperature T and T+10°C.

Now although the area under the whole curve remains the same, the dashed area is doubled. Thus,
the primary reason for almost 100% rise in the overall rate of reaction for every 10°C is the 100% increase in
the number of effective collisions.
 The Arrhenius Equation
In 1884, the famous Dutch chemist Jacobus Henricus Van't Hoff realized that his equation (Van't
Hoff equation) could also be used to suggests a formula for the rates of both forward and backward reactions.
In 1889, Svante Arrhenius immediately noticed the importance of this invention and proposed an empirical
equation based on Van't Hoff’s work. This equation is extremely useful in the modeling of the temperature
variation of many chemical reactions. The equation proposed by Arrhenius is

𝑘 = 𝐴 𝑒 −𝐸𝑎/𝑅𝑇 (8)

Where the symbol k, R and T represent rate constant, gas constant and temperature, respectively. A is popularly
known as the pre-exponential factor or Arrhenius constant with the units identical to those of the rate constant
used, and therefore, will vary depending on the order of the reaction. The term Ea represents the activation
energy measured in joule mole−1.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 117

Another popular form of the Arrhenius equation is

𝑘 = 𝐴 𝑒 −𝐸𝑎/𝑘𝐵 𝑇 (9)

The only difference in the equation (8) and equation (9) is the energy units of Ea; the former one uses energy
per mole, which is more common in chemistry, while the latter form uses energy per molecule directly, which
is common in physics. The different units are accounted for in using either the gas constant, R, or the Boltzmann
constant, kB, as the multiplier of temperature T. If the reaction is first order, A will have the units of s−1 and can
be called as collision frequency or frequency factor.
The physical significance of k is that it represents the number of collisions that result in a reaction
per second; A is the number of collisions (leading to a reaction or not) per second occurring with the proper
orientation to react. The exponential factor is the probability that any given collision will result in a reaction.
It can also be seen that either increasing the temperature or decreasing the activation energy (for example
through the use of catalysts) will result in an increase in the rate of reaction. Taking the natural logarithm of
both side of equation (8), we get

ln 𝑘 = ln 𝐴 + ln 𝑒 −𝐸𝑎/𝑅𝑇 (10)

𝐸𝑎 (11)
ln 𝑘 = ln 𝐴 −
𝑅𝑇
Rearrange the above equation, we get

𝐸𝑎 (12)
ln 𝑘 = − + ln 𝐴
𝑅𝑇
The equation (12) has the same form as the equation of straight line i.e. 𝑦 = 𝑚𝑥 + 𝑐; which means that if we
plot “ln k” vs 1/T, the slope and intercept will yield “−Ea/R” and “ln A”, respectively.

Figure 4. The Arrhenius plot of ln k vs 1/T.

Copyright © Mandeep Dalal


118 A Textbook of Physical Chemistry – Volume I

In addition to the equation (12), one of the more popular forms of the Arrhenius equation can be derived by
converting it to the common logarithm as given below.

𝐸𝑎 (13)
2.303 log 𝑘 = − + 2.303 log 𝐴
𝑅𝑇
or

𝐸𝑎 (14)
log 𝑘 = − + log 𝐴
2.303 𝑅𝑇
The equation (14) also has the same form as the equation of straight line i.e. 𝑦 = 𝑚𝑥 + 𝑐; which means that if
we plot “log k” vs 1/T, the slope and intercept will yield “−Ea/2.303R” and “log A”, respectively.

Figure 5. The Arrhenius plot of log k vs 1/T.

To obtain the integrated form of the Arrhenius equation, differentiate the equation (12) as given below.

𝑑 ln 𝑘 𝐸𝑎 (15)
=
𝑑𝑇 𝑅𝑇 2
Now integrating the above equation between temperature T1 and T2, we get

𝑘2 𝐸𝑎 1 1 𝑘2 𝐸𝑎 1 1 (16)
ln = [ − ] 𝑜𝑟 log = [ − ]
𝑘1 𝑅 𝑇1 𝑇2 𝑘1 2.303𝑅 𝑇1 𝑇2

The above equation can be used to find the activation energy if rate constants are known at two different
temperatures.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 119

 Rate Law for Opposing Reactions of Ist Order and IInd Order
A reaction will be called as the opposing or reversible reaction if the reactants react together to form
a product and the products also react to yield the reactants simultaneously under the same conditions.
In a simple context, we can say these reactions proceed not only in the forward direction but also in
the backward direction. These reactions can be classified into the following categories based upon the kinetic
order of the reactions involved.
 First Order Opposed by First Order
In order to understand the kinetic profile of first-order reactions opposed by the first order, consider
a general reaction in which the reactant A forms product B i.e.

𝑘𝑓 (17)
𝐴 ⇌ 𝐵
𝑘𝑏

Now, if kf ⋙ kb, kb can be neglected. However, kf and kb have comparable values, a rate law depending upon
both the constants can be written. To do so, suppose that a is the initial concentration of the reactant A and x
is the decrease in the concentration of A after ʻtʼ time. The concentration of the product after the same time
would also be equal to x. Hence, the rates of forward reaction (Rf) and backward reaction (Rb) can be given as:

𝑅𝑓 = 𝑘𝑓 [𝐴] = 𝑘𝑓 (𝑎 − 𝑥) (17)

𝑅𝑏 = 𝑘𝑏 [𝐵] = 𝑘𝑏 𝑥 (18)

The net reaction rate i.e. rate of formation of the product can be given as

𝑑𝑥 (19)
= 𝑘𝑓 (𝑎 − 𝑥) − 𝑘𝑏 𝑥
𝑑𝑡
However, when the equilibrium is attained, the rate of forward reaction will be equal to the rate of backward
reaction i.e. Rf = Rb. Therefore, the will take the form

𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) = 𝑘𝑏 𝑥𝑒𝑞 (20)

Where xeq is the concentration of product B or the decrease in the concentration of reactant A at equilibrium.
Now putting the value of kb from equation (20) into equation (19), we get

𝑑𝑥 𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) (21)
= 𝑘𝑓 (𝑎 − 𝑥) − 𝑥
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 (𝑎 − 𝑥𝑒𝑞 ) (22)
= 𝑘𝑓 [(𝑎 − 𝑥) − 𝑥]
𝑑𝑡 𝑥𝑒𝑞

or

Copyright © Mandeep Dalal


120 A Textbook of Physical Chemistry – Volume I

𝑑𝑥 (𝑎 − 𝑥)𝑥𝑒𝑞 − (𝑎 − 𝑥𝑒𝑞 )𝑥 (23)


= 𝑘𝑓 [ ]
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 𝑎𝑥𝑒𝑞 − 𝑎𝑥 𝑘𝑓 𝑎(𝑥𝑒𝑞 − 𝑥) (24)


= 𝑘𝑓 =
𝑑𝑡 𝑥𝑒𝑞 𝑥𝑒𝑞

or
𝑥𝑒𝑞 (25)
𝑑𝑥 = 𝑘𝑓 𝑎𝑑𝑡
(𝑥𝑒𝑞 − 𝑥)

Integrating equation (25), we get

−𝑥𝑒𝑞 ln(𝑥𝑒𝑞 − 𝑥) = 𝑘𝑓 𝑎𝑡 + 𝐶 (26)

Where C is the constant of integration. When t = 0, x = 0; putting these values in equation (26), we get

−𝑥𝑒𝑞 ln 𝑥𝑒𝑞 = 𝐶 (27)

Using the value of C form equation (27) in (26)

−𝑥𝑒𝑞 ln(𝑥𝑒𝑞 − 𝑥) = 𝑘𝑓 𝑎𝑡 − 𝑥𝑒𝑞 ln 𝑥𝑒𝑞 (28)

or

𝑥𝑒𝑞 ln 𝑥𝑒𝑞 − 𝑥𝑒𝑞 ln(𝑥𝑒𝑞 − 𝑥) = 𝑘𝑓 𝑎𝑡 (29)

𝑥𝑒𝑞 𝑥𝑒𝑞 (30)


𝑘𝑓 = ln
𝑎𝑡 𝑥𝑒𝑞 − 𝑥

Using equation (30), the rate constant for the forward reaction can easily be determined by measuring simple
quantities like a, t, xeq and x. Now rearranging equation (20) for kb

𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) (31)
𝑘𝑏 =
𝑥𝑒𝑞

𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) 𝑎 𝑥𝑒𝑞 𝑎 (32)


𝑘𝑏 = = 𝑘𝑓 ( − ) = 𝑘𝑓 ( − 1)
𝑥𝑒𝑞 𝑥𝑒𝑞 𝑥𝑒𝑞 𝑥𝑒𝑞

Now putting the value kf from equation (30) in equation (32), we get

1 𝑥𝑒𝑞 𝑥𝑒𝑞 𝑥𝑒𝑞 (33)


𝑘𝑏 = ( ln )− ln
𝑡 𝑥𝑒𝑞 − 𝑥 𝑎𝑡 𝑥𝑒𝑞 − 𝑥

Hence, the value of the rate constant for backward reaction can also be obtained just by measuring t, xeq and x;
or in other words, the kf eventually yields the kb also.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 121

Alternatively, the values of kf and kb can also be obtained a slightly different route. Rearranging
equation (30), we get

𝑎 𝑘𝑓 1 𝑥𝑒𝑞 (34)
= ln
𝑥𝑒𝑞 𝑡 𝑥𝑒𝑞 − 𝑥

Also rearranging equation (20), we have

𝑘𝑓 𝑎 (35)
= 𝑘𝑏 + 𝑘𝑓
𝑥𝑒𝑞

Equating the right-hand sides of equation (34) and (35), we get

1 𝑥𝑒𝑞 (36)
ln = 𝑘𝑏 + 𝑘𝑓
𝑡 𝑥𝑒𝑞 − 𝑥
𝑥𝑒𝑞 (37)
ln = (𝑘𝑏 + 𝑘𝑓 )𝑡
𝑥𝑒𝑞 − 𝑥

or

𝑥𝑒𝑞 (𝑘𝑏 + 𝑘𝑓 ) (38)


log = 𝑡
𝑥𝑒𝑞 − 𝑥 2.303

Equation (37) and (38) are the equations of straight line (y = mx + c) with zero intercept.

𝑥𝑒𝑞 𝑥𝑒𝑞
Figure 6. The plot of ln and log vs time for first order opposed by the first order.
𝑥𝑒𝑞 −𝑥 𝑥𝑒𝑞 −𝑥

Now, finding kf +kb from slope and kf/kb from equilibrium constant, kf and kb can easily be obtained from the
𝑥𝑒𝑞 −𝑥 𝑥𝑒𝑞 −𝑥
elimination method. It should also be noted that if plot ln 𝑥𝑒𝑞
and log 𝑥𝑒𝑞
, the slopes will become negative.

Copyright © Mandeep Dalal


122 A Textbook of Physical Chemistry – Volume I

 First Order Opposed by Second Order


In order to understand the kinetic profile of first-order reactions opposed by second-order, consider
a general reaction in which the reactant A forms product B and C i.e.

𝑘𝑓 (39)
𝐴 ⇌ 𝐵+𝐶
𝑘𝑏

Now, if kf ⋙ kb, kb can be neglected. However, kf and kb have comparable values, a rate law depending upon
both the constants can be written. To do so, suppose that a is the initial concentration of the reactant A and x
is the decrease in the concentration of A after ʻtʼ time. The concentration of both the products after same time
would also be equal to x. Hence, the rates of forward reaction (Rf) and backward reaction (Rb) can be given as:

𝑅𝑓 = 𝑘𝑓 [𝐴] = 𝑘𝑓 (𝑎 − 𝑥) (40)

𝑅𝑏 = 𝑘𝑏 [𝐵][𝐶] = 𝑘𝑏 𝑥 2 (41)

The net reaction rate i.e. rate of formation of the product can be given as

𝑑𝑥 (42)
= 𝑘𝑓 (𝑎 − 𝑥) − 𝑘𝑏 𝑥 2
𝑑𝑡
However, when the equilibrium is attained, the rate of forward reaction will be equal to the rate of backward
reaction i.e. Rf = Rb. Therefore, the equation (42) will take the form
2
𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) = 𝑘𝑏 𝑥𝑒𝑞 (43)

Where xeq is the concentration of product B and C or the decrease in the concentration of reactant A at
equilibrium. Now putting the value of kb from equation (43) into equation (42), we get

𝑑𝑥 𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) 2 (44)
= 𝑘𝑓 (𝑎 − 𝑥) − 2 𝑥
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 (𝑎 − 𝑥𝑒𝑞 ) 2 (45)
= 𝑘𝑓 [(𝑎 − 𝑥) − 2 𝑥 ]
𝑑𝑡 𝑥𝑒𝑞

or

𝑑𝑥 2
(𝑎 − 𝑥)𝑥𝑒𝑞 − (𝑎 − 𝑥𝑒𝑞 )𝑥 2 (46)
= 𝑘𝑓 [ 2 ]
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 2
𝑎𝑥𝑒𝑞 2
− 𝑥𝑥𝑒𝑞 − 𝑎𝑥 2 + 𝑥𝑒𝑞 𝑥 2 (47)
= 𝑘𝑓 [ 2 ]
𝑑𝑡 𝑥𝑒𝑞

or

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 123

2
𝑥𝑒𝑞 (48)
2 2
𝑑𝑥 = 𝑘𝑓 𝑑𝑡
(𝑎𝑥𝑒𝑞 − 𝑥𝑥𝑒𝑞 − 𝑎𝑥 2 + 𝑥𝑒𝑞 𝑥 2 )

Integrating equation (48), and then rearranging

𝑥𝑒𝑞 𝑎𝑥𝑒𝑞 + 𝑥𝑒𝑞 (𝑎 − 𝑥𝑒𝑞 ) (49)


𝑘𝑓 = ln
𝑡(2𝑎 − 𝑥𝑒𝑞 ) 𝑎(𝑥𝑒𝑞 − 𝑥)

Using equation (49), the rate constant for the forward reaction can easily be determined by measuring simple
quantities like a, t, xeq and x. Now, we know that the equilibrium constant for the first order opposed by second-
order will be

[𝐵][𝐶] (50)
𝐾=
[𝐴]

also

𝑘𝑓 (51)
𝐾=
𝑘𝑏

Now putting the value of kf from equation (49) in equation (51) and the rearranging for kb, we get

𝑥𝑒𝑞 𝑎𝑥𝑒𝑞 + 𝑥𝑒𝑞 (𝑎 − 𝑥𝑒𝑞 ) (52)


𝑘𝑏 = ln
𝑡(2𝑎 − 𝑥𝑒𝑞 )𝐾 𝑎(𝑥𝑒𝑞 − 𝑥)

Hence, the value of the rate constant for backward reaction can also be obtained just by measuring t, xeq and x
and the equilibrium constant from equation (50).
 Second Order Opposed by First Order
In order to understand the kinetic profile of second-order reactions opposed by first order, consider a
general reaction in which two reactants A and B form product C i.e.

𝑘𝑓 (53)
𝐴+𝐵 ⇌ 𝐶
𝑘𝑏

Now, if kf ⋙ kb, kb can be neglected. However, kf and kb have comparable values, a rate law depending upon
both the constants can be written. To do so, suppose that a is the initial concentration of both the reactant A
and B; while x is the decrease in the concentrations of both reactants after ʻtʼ time. The concentration of the
product after the same time would also be equal to x. Hence, the rates of forward reaction (Rf) and backward
reaction (Rb) can be given as:

𝑅𝑓 = 𝑘𝑓 [𝐴][𝐵] = 𝑘𝑓 (𝑎 − 𝑥)2 (54)

𝑅𝑏 = 𝑘𝑏 [𝐶] = 𝑘𝑏 𝑥 (55)

Copyright © Mandeep Dalal


124 A Textbook of Physical Chemistry – Volume I

The net reaction rate i.e. rate of formation of the product can be given as

𝑑𝑥 (56)
= 𝑘𝑓 (𝑎 − 𝑥)2 − 𝑘𝑏 𝑥
𝑑𝑡
However, when the equilibrium is attained, the rate of forward reaction will be equal to the rate of backward
reaction i.e. Rf = Rb. Therefore, the equation (56) will take the form
2
𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) = 𝑘𝑏 𝑥𝑒𝑞 (57)

Where xeq is the concentration of the product C or the decrease in the concentration of reactant A or B at
equilibrium. Now putting the value of kb from equation (57) into equation (56), we get

𝑑𝑥 𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 )
2 (58)
= 𝑘𝑓 (𝑎 − 𝑥)2 − 𝑥
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 (𝑎 − 𝑥𝑒𝑞 )
2 (59)
= 𝑘𝑓 [(𝑎 − 𝑥)2 − 𝑥]
𝑑𝑡 𝑥𝑒𝑞

or

𝑑𝑥 (𝑎 − 𝑥)2 𝑥𝑒𝑞 − (𝑎 − 𝑥𝑒𝑞 ) 𝑥


2 (60)
= 𝑘𝑓 [ ]
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 (𝑎2 + 𝑥 2 − 2𝑎𝑥)𝑥𝑒𝑞 − (𝑎2 + 𝑥𝑒𝑞


2
− 2𝑎𝑥𝑒𝑞 )𝑥 (61)
= 𝑘𝑓 [ ]
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 𝑎2 𝑥𝑒𝑞 + 𝑥 2 𝑥𝑒𝑞 − 2𝑎𝑥𝑥𝑒𝑞 − 𝑎2 𝑥 − 𝑥𝑒𝑞


2
𝑥 + 2𝑎𝑥𝑒𝑞 𝑥 (62)
= 𝑘𝑓 [ ]
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 𝑎2 𝑥𝑒𝑞 + 𝑥 2 𝑥𝑒𝑞 − 𝑎2 𝑥 − 𝑥𝑒𝑞


2
𝑥 (63)
= 𝑘𝑓 [ ]
𝑑𝑡 𝑥𝑒𝑞

or
𝑥𝑒𝑞 (64)
2
𝑑𝑥 = 𝑘𝑓 𝑑𝑡
(𝑎2 𝑥𝑒𝑞 + 2
𝑥 𝑥𝑒𝑞 − 𝑎2 𝑥 − 𝑥𝑒𝑞 𝑥)

Integrating equation (64), and then rearranging

𝑥𝑒𝑞 𝑥𝑒𝑞 (𝑎2 − 𝑥𝑒𝑞 𝑥) (65)


𝑘𝑓 = 2 ) ln 𝑎2 (𝑥
𝑡(𝑎2 − 𝑥𝑒𝑞 𝑒𝑞 − 𝑥)

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 125

Using equation (66), the rate constant for the forward reaction can easily be determined by measuring simple
quantities like a, t, xeq and x. Now, we know that the equilibrium constant for a second-order reaction opposed
by first order will be

[𝐶] (66)
𝐾=
[𝐴][𝐵]

𝑘𝑓 (67)
𝐾=
𝑘𝑏

Now putting the value of kf from equation (65) in equation (67) and the rearranging for kb, we get

𝑥𝑒𝑞 𝑥𝑒𝑞 (𝑎2 − 𝑥𝑒𝑞 𝑥) (68)


𝑘𝑏 = 2 )𝐾 ln
𝑡(𝑎2 − 𝑥𝑒𝑞 𝑎2 (𝑥𝑒𝑞 − 𝑥)

Hence, the value of the rate constant for backward reaction can also be obtained just by measuring t, xeq and x
and the equilibrium constant from equation (66).
 Second Order Opposed by Second Order
In order to understand the kinetic profile of second-order reactions opposed by second-order, consider
a general reaction in which two reactants A and B form product C and D i.e.

𝑘𝑓 (69)
𝐴+𝐵 ⇌ 𝐶+𝐷
𝑘𝑏

Now, if kf ⋙ kb, kb can be neglected. However, kf and kb have comparable values, a rate law depending upon
both the constants can be written. To do so, suppose that a is the initial concentration of both the reactant A
and B; while x is the decrease in the concentrations of both reactants after ʻtʼ time. The concentration of the
products after the same time would also be equal to x. Hence, the rates of forward reaction (Rf) and backward
reaction (Rb) can be given as:

𝑅𝑓 = 𝑘𝑓 [𝐴][𝐵] = 𝑘𝑓 (𝑎 − 𝑥)2 (70)

𝑅𝑏 = 𝑘𝑏 [𝐶][𝐷] = 𝑘𝑏 𝑥 2 (71)

The net reaction rate i.e. rate of formation of the product can be given as

𝑑𝑥 (72)
= 𝑘𝑓 (𝑎 − 𝑥)2 − 𝑘𝑏 𝑥 2
𝑑𝑡
However, when the equilibrium is attained, the rate of forward reaction will be equal to the rate of backward
reaction i.e. Rf = Rb. Therefore, the equation (72) will take the form
2 2
𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) = 𝑘𝑏 𝑥𝑒𝑞 (73)

Copyright © Mandeep Dalal


126 A Textbook of Physical Chemistry – Volume I

Where xeq is the concentration of the product C and D or the decrease in the concentration of reacant A or B at
equilibrium. Now putting the value of kb from equation (73) into equation (72), we get

𝑑𝑥 𝑘𝑓 (𝑎 − 𝑥𝑒𝑞 ) 2
2 (74)
= 𝑘𝑓 (𝑎 − 𝑥)2 − 2 𝑥
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 (𝑎 − 𝑥𝑒𝑞 ) 2
2 (75)
= 𝑘𝑓 [(𝑎 − 𝑥)2 − 2 𝑥 ]
𝑑𝑡 𝑥𝑒𝑞

or

𝑑𝑥 (𝑎 − 𝑥)2 𝑥𝑒𝑞
2
− (𝑎 − 𝑥𝑒𝑞 ) 𝑥 2
2 (76)
= 𝑘𝑓 [ 2 ]
𝑑𝑡 𝑥𝑒𝑞

𝑑𝑥 𝑎2 𝑥𝑒𝑞
2
+ 𝑥 2 𝑥𝑒𝑞
2 2
− 2𝑎𝑥𝑥𝑒𝑞 − 𝑎2 𝑥 2 − 𝑥𝑒𝑞
2 2
𝑥 + 2𝑎𝑥𝑒𝑞 𝑥 2 (78)
= 𝑘𝑓 [ 2 ]
𝑑𝑡 𝑥𝑒𝑞

or
2
𝑥𝑒𝑞 (80)
2 2
𝑑𝑥 = 𝑘𝑓 𝑑𝑡
(𝑎2 𝑥𝑒𝑞 − 2𝑎𝑥𝑥𝑒𝑞 − 𝑎2 𝑥 2 + 2𝑎𝑥𝑒𝑞 𝑥2)

Integrating equation (80), and then rearranging

𝑥𝑒𝑞 𝑥(𝑎 − 2𝑥𝑒𝑞 ) + 𝑎𝑥𝑒𝑞 (81)


𝑘𝑓 = ln
2𝑎𝑡(𝑎 − 𝑥𝑒𝑞 ) 𝑎(𝑥𝑒𝑞 − 𝑥)

Using equation (81), the rate constant for the forward reaction can easily be determined by measuring simple
quantities like a, t, xeq and x. Now, we know that the equilibrium constant for a second-order reaction opposed
by second-order will be

[𝐶][𝐷] (82)
𝐾=
[𝐴][𝐵]

𝑘𝑓 (83)
𝐾=
𝑘𝑏

Now putting the value of kf from equation (81) in equation (83) and the rearranging for kb, we get

𝑥𝑒𝑞 𝑥(𝑎 − 2𝑥𝑒𝑞 ) + 𝑎𝑥𝑒𝑞 (84)


𝑘𝑏 = ln
2𝑎𝑡(𝑎 − 𝑥𝑒𝑞 )𝐾 𝑎(𝑥𝑒𝑞 − 𝑥)

Hence, the value of the rate constant for backward reaction can also be obtained just by measuring t, xeq and x
and the equilibrium constant from equation (84).

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 127

 Rate Law for Consecutive & Parallel Reactions of Ist Order Reactions
In addition to the opposing or reversible reactions, two other types of simultaneous reactions are
consecutive and parallel reactions. In this section, we will discuss the kinetic profiles of these two types of
reactions up to the first order only.
 Consecutive Reactions
In many complex reactions, the order of the reaction has not been found equal to the molecularity
noted from the stoichiometry. So, these reactions must take place in multiple steps rather than a single step.
These multiple steps are individually labeled as consecutive reactions.
The consecutive reactions may be defined as the single-step reactions which can be written to
represent an overall reaction.
In order to understand the kinetic profile of consecutive reactions, consider two first-order reactions
in which reactant A converts to B which in turn converts to product C.

𝐴→
𝑘1
𝐵 →
𝑘2
𝐶 (85)

Where k1 and k2 are the rate constants for the first and second steps, respectively. In other words, A is the
reactant, B is simply the intermediate and C is the final product.
However, kf and kb have comparable values, a rate law depending upon both the constants can be written. Now
suppose that the initial concentrations of reactant A is C0; while the concentrations of A, B and C after time t
are CA, CB and CC, respectively. So, we can say that

𝐶0 = 𝐶𝐴 + 𝐶𝐵 + 𝐶𝐶 (86)

Now, the rate can be deduced in terms of CA, CB and CC as given below.
1. Rate law in terms of CA: The rate of disappearance reactant of A in the given reaction can be given by the
following relation.

𝑑[𝐶𝐴 ] (87)
− = 𝑘1 𝐶𝐴
𝑑𝑡
or

𝑑[𝐶𝐴 ] (88)
− = 𝑘1 𝑑𝑡
𝐶𝐴

Integrating both sides, we get

− ln 𝐶𝐴 = 𝑘1 𝑡 + I (89)

Where I is the constant of integration. However, when t = 0, CA = C0, the equation (89) takes the form

Copyright © Mandeep Dalal


128 A Textbook of Physical Chemistry – Volume I

− ln 𝐶0 = I (90)

Using the value of integration constant from equation (90) in equation (89), we get

− ln 𝐶𝐴 = 𝑘1 𝑡 − ln 𝐶0 (91)

or

ln 𝐶0 − ln 𝐶𝐴 = 𝑘1 𝑡 (92)

or

𝐶𝐴 (93)
−ln = 𝑘1 𝑡
𝐶0

or

𝐶𝐴 (94)
ln = −𝑘1 𝑡
𝐶0

or

𝐶𝐴 (95)
= 𝑒 −𝑘1 𝑡
𝐶0

or

𝐶𝐴 = 𝐶0 𝑒 −𝑘1 𝑡 (96)

2. Rate law in terms of CB: The rate of formation of intermediate B can be given by the following relation.

𝑑[𝐶𝐵 ] (97)
= −𝑘2 𝐶𝐵 + 𝑘1 𝐶𝐴
𝑑𝑡
or

𝑑[𝐶𝐵 ] (98)
= 𝑘1 𝐶𝐴 − 𝑘2 𝐶𝐵
𝑑𝑡
After putting the value of CA from equation (96) in equation (98), we get a linear differential equation of first
order i.e.

𝑑[𝐶𝐵 ] (99)
= 𝑘1 𝐶0 𝑒 −𝑘1 𝑡 − 𝑘2 𝐶𝐵
𝑑𝑡
Integrating and then rearranging equation (99), both side, we get

𝑘1 (100)
[𝐶𝐵 ] = 𝐶0 ( ) (𝑒 −𝑘1 𝑡 − 𝑒 −𝑘2 𝑡 )
𝑘2 − 𝑘1

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 129

3. Rate law in terms of CC: The overall rate of formation of the product C in the given reaction can be given
by the following relation.

𝑑[𝐶𝐶 ] (101)
= 𝑘2 𝐶𝐵
𝑑𝑡
After putting the value of CA and CB from equation (96) and equation (100) into equation (86), we get the
following result.

𝑘1 (102)
𝐶𝐶 = 𝐶0 − 𝐶0 𝑒 −𝑘1 𝑡 − 𝐶0 ( ) (𝑒 −𝑘1 𝑡 − 𝑒 −𝑘2 𝑡 )
𝑘2 − 𝑘1

or

𝑘1 (103)
= 𝐶0 [1 − 𝑒 −𝑘1 𝑡 − ( ) (𝑒 −𝑘1 𝑡 − 𝑒 −𝑘2 𝑡 )]
𝑘2 − 𝑘1

or

𝑘1 𝑒 −𝑘1 𝑡 𝑘1 𝑒 −𝑘2 𝑡 (104)


= 𝐶0 [1 − 𝑒 −𝑘1 𝑡 − + ]
𝑘2 − 𝑘1 𝑘2 − 𝑘1

or

𝑘1 𝑒 −𝑘1 𝑡 𝑘1 𝑒 −𝑘2 𝑡 (105)


= 𝐶0 [1 − (𝑒 −𝑘1 𝑡 + − )]
𝑘2 − 𝑘1 𝑘2 − 𝑘1

or

(𝑘2 − 𝑘1 )𝑒 −𝑘1 𝑡 + 𝑘1 𝑒 −𝑘1 𝑡 − 𝑘1 𝑒 −𝑘2 𝑡 (106)


= 𝐶0 [1 − ( )]
𝑘2 − 𝑘1

𝑘2 𝑒 −𝑘1 𝑡 − 𝑘1 𝑒 −𝑘1 𝑡 + 𝑘1 𝑒 −𝑘1 𝑡 − 𝑘1 𝑒 −𝑘2 𝑡 (107)


= 𝐶0 [1 − ( )]
𝑘2 − 𝑘1

𝑘2 𝑒 −𝑘1 𝑡 − 𝑘1 𝑒 −𝑘2 𝑡 (108)


= 𝐶0 [1 − ( )]
𝑘2 − 𝑘1

or

1 (109)
𝐶𝐶 = 𝐶0 [1 − (𝑘 𝑒 −𝑘1 𝑡 − 𝑘1 𝑒 −𝑘2 𝑡 )]
(𝑘2 − 𝑘1 ) 2

The equation (96), (100) and (109) can be used to plot the time-dependent variation of CA, CB and CC,
respectively.

Copyright © Mandeep Dalal


130 A Textbook of Physical Chemistry – Volume I

Figure 7. The plot CA, CB and CC as a function of time in a typical consecutive reaction.

It can be clearly seen that the concentration of A decreases exponentially, while the concentration of
B increases first and then declines. The concentration of C increases continuously and finally becomes equal
to the concentration of A.
Maxima in the concentration of B: In addition to the time-dependent concentration variation of different
species, one more important parameter to measure is the maximum B concentration. Since, for this the value
of [dCB]/dt = 0, the differentiation of equation (100) and then putting equal to zero gives

𝑑[𝐶𝐵 ] 𝑘1 (110)
= 𝐶0 ( ) (−𝑘1 𝑒 −𝑘1 𝑡 + 𝑘2 𝑒 −𝑘2 𝑡 ) = 0
𝑑𝑡 𝑘2 − 𝑘1

−𝑘1 𝑒 −𝑘1 𝑡 + 𝑘2 𝑒 −𝑘2 𝑡 = 0 (111)

or

𝑘1 𝑒 −𝑘1 𝑡 = 𝑘2 𝑒 −𝑘2 𝑡 (112)

or

𝑘1 𝑒 −𝑘2 𝑡 (113)
= −𝑘 𝑡
𝑘2 𝑒 1

𝑘1 (114)
= 𝑒 (𝑘1 −𝑘2 )𝑡
𝑘2

Taking logarithm both side, we get

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 131

𝑘1 (115)
ln = ln 𝑒 (𝑘1 −𝑘2 )𝑡
𝑘2

𝑘1 (116)
ln = (𝑘1 − 𝑘2 )𝑡
𝑘2

1 𝑘1 (117)
𝑡𝑚𝑎𝑥 = ln
(𝑘1 − 𝑘2 ) 𝑘2

Now putting the value of t from equation (117) in equation (100), we get

𝑘1 −𝑘1 ln(𝑘1 /𝑘2 ) −𝑘2 ln(𝑘1 /𝑘2 ) (118)


[𝐶𝐵 ] = 𝐶0 ( ) [exp { } − exp { }]
𝑘2 − 𝑘1 𝑘1 − 𝑘2 𝑘1 − 𝑘2

Simplifying and then rearranging the above equation, we get

𝑘2 (119)
[𝐶𝐵 ]𝑚𝑎𝑥 = 𝐶0 ( ) ek2 /(𝑘1 −𝑘2 )
𝑘1

Rate law in special cases: In addition to the typical consecutive reaction i.e. k1 = k2, two special cases also
arise from the nature of the step reactions discussed below.
i) When k2 ⋙ k1: In these types of reactions, the value of k1 can be neglected. Therefore, the equation (109)
takes the form

𝐶𝐶 = 𝐶0 (1 − 𝑒 −𝑘1 𝑡 ) (120)

Graphically,

Figure 8. The plot CA, CB and CC vs time in a typical consecutive reaction when k2 ⋙ k1.

Copyright © Mandeep Dalal


132 A Textbook of Physical Chemistry – Volume I

It can be clearly seen that the concentration of the intermediate practically remains constant, and therefore, the
steady-state approximation can be applied in this case.
ii) When k1 ⋙ k2: In these types of reactions, the value of k2 can be neglected. Therefore, the equation (109)
takes the form

𝐶𝐶 = 𝐶0 (1 − 𝑒 −𝑘2 𝑡 ) (121)

Graphically,

Figure 9. The plot CA, CB and CC vs time in a typical consecutive reaction when k1 ⋙ k2.

 Parallel Reactions
In many reactions, the reactant reacts to form more than one product simultaneously. If the amount
of one the reaction product is very large in comparison to the others, then we can simply neglect these other
reactions. However, if the amount of the product formed by other reactions are significant, we must refine the
overall rate equation to represent this.
The parallel or side reactions may simply be defined as the reactions in which initial species react to
give multiple products simultaneously.
In order to understand the kinetics of parallel reactions of the first order, suppose that a reactant A
reacts to form product B and C simultaneously. A typical depiction of the parallel or side reaction with two
pathways is given below.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 133

Where k1 and k2 are the rate contents. Now suppose that a is the initial concentration of the reactant A, while x
is the decrease in the concentrations of the reactant after ʻtʼ time. Hence, the rates of first (R1) and second
reaction (R2) can be given as:

𝑑[𝐵] (122)
𝑅1 = = 𝑘1 [𝐴] = 𝑘1 (𝑎 − 𝑥)
𝑑𝑡
or

𝑑[𝐶] (123)
𝑅2 = = 𝑘2 [𝐴] = 𝑘2 (𝑎 − 𝑥)
𝑑𝑡
The overall reaction rate can be obtained by adding equation (122) and equation (123) as

𝑑𝑥 𝑑[𝐵] 𝑑[𝐶] (124)


= + = 𝑘1 (𝑎 − 𝑥) + 𝑘2 (𝑎 − 𝑥)
𝑑𝑡 𝑑𝑡 𝑑𝑡
or

𝑑𝑥 (125)
= (𝑘1 + 𝑘2 )(𝑎 − 𝑥)
𝑑𝑡
or

𝑑𝑥 (126)
= (𝑘1 + 𝑘2 )𝑑𝑡
(𝑎 − 𝑥)

Integrating the equation (126) and then rearranging, we get

1 𝑎 (127)
𝑘1 + 𝑘2 = ln
𝑡 (𝑎 − 𝑥)

Also, dividing equation (122) by (123), we get

𝑅1 𝑘1 (𝑎 − 𝑥) (128)
=
𝑅2 𝑘2 (𝑎 − 𝑥)

Which implies that

𝑅1 𝑘1 (129)
=
𝑅2 𝑘2

Copyright © Mandeep Dalal


134 A Textbook of Physical Chemistry – Volume I

Hence, the value of rate constants involved, i.e., k1 and k2 can easily be obtained from the use of equation (127)
and equation (129).

Figure 10. The variation of the concentrations of reactants and products as a function of time in a typical
parallel reaction.

It should also be noted from the equation (129) that the ratio of the concentration of products remains the same
with time. Furthermore, the percentage of both products can also be obtained from the knowledge of rate
constants using the relations given below.

𝑘1 (130)
𝐹𝑟𝑎𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑄𝑢𝑎𝑛𝑡𝑢𝑚 𝑌𝑖𝑒𝑙𝑑 𝑜𝑓 𝐴 =
𝑘1 + 𝑘2

also

𝑘2 (131)
𝐹𝑟𝑎𝑐𝑡𝑖𝑜𝑛𝑎𝑙 𝑄𝑢𝑎𝑛𝑡𝑢𝑚 𝑌𝑖𝑒𝑙𝑑 𝑜𝑓 𝐵 =
𝑘1 + 𝑘2

The percentage is obtained by multiplying corresponding fractional quantum yields by 100. Similarly, parts
per thousand (ppt) and parts per million (ppm) can be obtained by multiplying equations (130) and (131) by
103 and 106, respectively.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 135

 Collision Theory of Reaction Rates and Its Limitations


In 1916, a German chemist Max Trautz proposed a theory based on the collisions of reacting
molecules to explain reaction kinetics. Two years later, a British chemist William Lewis published similar
results, however, he was completely unaware of Trautz’s work. The remarkable work of these two gentlemen
was extremely beneficial in explaining the rate of many chemical reactions.
The collision theory states that when the right reactant particles strike each other, only a definite
fraction of the collisions induce any significant or noticeable chemical change; these successful changes are
called successful collisions and are possible only if reacting molecules have sufficient energy at the moment
of impact to break the pre-existing bonds and form all new bonds.
The minimum energy required to make a collision successful is called as the activation energy, and
these types of collisions result in the products of the reaction. The rise in reactant concentration or increasing
the temperature, both result in more collisions and hence more successful collisions, and therefore, increase
the reaction rate. Sometimes, a catalyst is involved in the collision between the reactant molecules that
decreases the energy required for the chemical change to take place, and so more collisions would have
sufficient energy for the reaction to happen. In this section, we will discuss the collision theory of bimolecular
and unimolecular reactions in the gaseous phase.
 Collision Theory for Bimolecular Reactions
In order to understand the collision theory for bimolecular reactions, we must understand the cause
of a reaction itself first. The primary requirement for a reaction to occur is the collision between the reacting
molecules. Therefore, if we assume that every collision results in the formation of the product, the rate of
reaction should simply be equal to collision frequency (Z) of the reacting system i.e. the number of collisions
occurring in the container per unit volume per unit time. Mathematically, we can say that

𝑅𝑎𝑡𝑒 = 𝑍 (132)

However, the actual rate would be much less than what is predicted by the equation (132); which is obviously
due to the fact that all the collisions are not effective. Therefore, equation (132) must be modified to represent
this factor. If f is the fraction of the molecules which are activated, the rate expression can be written as given
below.

𝑅𝑎𝑡𝑒 = 𝑍 × 𝑓 (133)

Now, according to the Maxwell-Boltzmann distribution of energies, the fraction of the molecules having
energy greater than a particular energy E is

𝛥𝑁 (134)
𝑓= = 𝑒 −𝐸/𝑅𝑇
𝑁
Where N is the total number of molecules while ΔN represents the number of molecules having energy greater
than E. However, if E = Ea, the fraction of activated molecules can be written as

Copyright © Mandeep Dalal


136 A Textbook of Physical Chemistry – Volume I

𝛥𝑁 (135)
𝑓= = 𝑒 −𝐸𝑎/𝑅𝑇
𝑁
Where R is the gas constant and T is the reaction temperature. After putting the value of f from equation (135)
into equation (133), we get

𝑅𝑎𝑡𝑒 = 𝑍𝑒 −𝐸𝑎/𝑅𝑇 (136)

At this point, two possibilities arise; one, when the colliding molecules are similar and other, is when
the colliding molecules are dissimilar. We will discuss these cases one by one.
1. Rate of reaction when the colliding molecules are dissimilar: Consider a bimolecular reaction between
different molecules A and B yielding product P as

𝐴+𝐵 →𝑃 (137)

The number of collisions between A and B occurring in the container per unit volume per unit time can be
given by the following relation.

(138)
2 √
8𝜋𝑘𝐵 𝑇
𝑍= 𝑛𝐴 𝑛𝐵 𝜎𝐴𝐵
𝜇𝐴𝐵

Where nA and nA are the number densities (in the units of m−3) of particles A and B, respectively. The term σAB
is simply the average collision diameter i.e. σAB = (σA + σB)/2. kB is the Boltzmann's constant (m2 kg s−2 K−1).
T represents the temperature of the system. The term μAB represents the reduced mass of the reactants A and B
i.e. μAB = mAmB/mA+mB.
The equation (138) can also be expressed in terms of molar masses by putting mA = MA/N, mB = MB/N
and k = R/N; where MA and MB are molar masses of the reactants, R is the gas constant and N represents to
Avogadro number. Therefore, equation (138) takes the form

𝑅 𝑀 𝑀 (139)
8𝜋 ( ) 𝑇 ( 𝐴 + 𝐵 )
2 √
𝑍 = 𝜎𝐴𝐵 𝑁 𝑁 𝑁 𝑛 𝑛
𝑀𝐴 𝑀𝐵 𝐴 𝐵
×
𝑁 𝑁

or

(140)
2 √
8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 )
𝑍= 𝜎𝐴𝐵 𝑛𝐴 𝑛𝐵
𝑀𝐴 𝑀𝐵

Also, as we know that the reaction rate can be written in terms of molecules of reactants reacting per cm3 per
second as

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 137

𝑑𝑛𝐴 𝑑𝑛𝐵 (141)


𝑅𝑎𝑡𝑒 = − =− = 𝑍𝑒 −𝐸𝑎/𝑅𝑇
𝑑𝑡 𝑑𝑡
or

(142)
𝑑𝑛𝐴 𝑑𝑛𝐵 2 √
8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 )
𝑅𝑎𝑡𝑒 = − =− = 𝜎𝐴𝐵 × 𝑛𝐴 𝑛𝐵 × 𝑒 −𝐸𝑎/𝑅𝑇
𝑑𝑡 𝑑𝑡 𝑀𝐴 𝑀𝐵

Now, in order to express the rate in terms of molar concentrations, we need to recall some typical relations like

𝑁[𝐴] 𝑁[𝐵] (143)


𝑛𝐴 = and 𝑛𝐵 =
103 103
also

𝑁 𝑁 (144)
𝑑𝑛𝐴 = 3
𝑑[𝐴] and 𝑑𝑛𝐵 = 3 𝑑[𝐵]
10 10
Using the results of equation (143) and (144) in equation (142), we get

(145)
𝑁 𝑑[𝐴] 𝑁 𝑑[𝐵] 2 √
8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 ) 𝑁[𝐴] 𝑁[𝐵] −𝐸 /𝑅𝑇
𝑅𝑎𝑡𝑒 = − 3 =− 3 = 𝜎𝐴𝐵 𝑒 𝑎
10 𝑑𝑡 10 𝑑𝑡 𝑀𝐴 𝑀𝐵 103 103

or

(146)
𝑑[𝐴] 𝑑[𝐵] 𝑁 2 8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 )
𝑅𝑎𝑡𝑒 = − =− = 3 𝜎𝐴𝐵 √ [𝐴][𝐵] 𝑒 −𝐸𝑎/𝑅𝑇
𝑑𝑡 𝑑𝑡 10 𝑀𝐴 𝑀𝐵

Comparing equation (146) with general rate law expressed in molar concentrations i.e. Rate = k[A][B], we get

(147)
𝑁 2 8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 ) −𝐸 /𝑅𝑇
𝑘 = 3 𝜎𝐴𝐵 √ 𝑒 𝑎
10 𝑀𝐴 𝑀𝐵

Comparing equation (147) with Arrhenius rate constant i.e. 𝑘 = 𝐴𝑒 −𝐸𝑎/𝑅𝑇 , we get

(148)
𝑁 2 8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 )
𝐴= 𝜎 √
103 𝐴𝐵 𝑀𝐴 𝑀𝐵

2. Rate of reaction when the colliding molecules are similar: Consider a bimolecular reaction between
similar molecules A and A yielding product P as

𝐴+𝐴 → 𝑃 (149)

Copyright © Mandeep Dalal


138 A Textbook of Physical Chemistry – Volume I

The number of collisions between A and A occurring in the container per unit volume per unit time can be
given by the following relation.

(150)
4𝜋𝑘𝐵 𝑇
𝑍= 𝑛𝐴2 𝜎 2√
𝑚𝐴

Where nA is the number density (in the units of m−3) of particle A. The term σ is simply the average collision
diameter. kB is the Boltzmann's constant (m2 kg s−2 K−1). T represents the temperature of the system. The term
mA represents the mass of the reactants A.
The equation (150) can also be expressed in terms of molar masses by putting mA = MA/N and k =
R/N; where MA is the molar mass of the reactant, R is the gas constant and N represents to Avogadro number.
Therefore, equation (150) takes the form

𝑅 (151)
4𝜋 (𝑁) 𝑇
𝑍 = 𝜎2√ 𝑛𝐴2
𝑀𝐴
𝑁

or

(152)
2√
4𝜋𝑅𝑇 2
𝑍= 𝜎 𝑛𝐴
𝑀𝐴

Also, as we know that the reaction rate can be written in terms of molecules of reactants reacting per cm3 per
second as

1 𝑑𝑛𝐴 (153)
− = 𝑍𝑒 −𝐸𝑎/𝑅𝑇
2 𝑑𝑡
or

(154)
𝑑𝑛𝐴 4𝜋𝑅𝑇
𝑅𝑎𝑡𝑒 = − = 2 (𝜎 2 √ × 𝑛𝐴2 × 𝑒 −𝐸𝑎/𝑅𝑇 )
𝑑𝑡 𝑀𝐴

Now, in order to express the rate in terms of molar concentrations, we need to recall some typical relations like

𝑁[𝐴] (155)
𝑛𝐴 =
103
also

𝑁 (156)
𝑑𝑛𝐴 = 𝑑[𝐴]
103

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 139

Using the results of equation (155) and (156) in equation (154), we get

(157)
𝑁 𝑑[𝐴] 4𝜋𝑅𝑇 𝑁 2 [𝐴]2
𝑅𝑎𝑡𝑒 = − 3 = 2 (𝜎 2 √ × × 𝑒 −𝐸𝑎/𝑅𝑇 )
10 𝑑𝑡 𝑀𝐴 106

or

(158)
𝑑[𝐴] 4𝑁 2 𝜋𝑅𝑇
𝑅𝑎𝑡𝑒 = − = 3𝜎 √ × [𝐴]2 × 𝑒 −𝐸𝑎/𝑅𝑇
𝑑𝑡 10 𝑀𝐴

Comparing equation (158) with general rate law expressed in molar concentrations i.e. Rate = k[A]2, we get

(159)
4𝑁 𝜋𝑅𝑇 −𝐸 /𝑅𝑇
𝑘 = 3 𝜎2√ 𝑒 𝑎
10 𝑀𝐴

Comparing equation (159) with Arrhenius rate constant i.e. 𝑘 = 𝐴𝑒 −𝐸𝑎/𝑅𝑇 , we get

(160)
4𝑁 𝜋𝑅𝑇
𝐴 = 3 𝜎2√
10 𝑀𝐴

 Collision Theory for Unimolecular Reactions


In order to understand the collision theory for unimolecular reactions, we must understand the root
cause of these reactions. In a typical unimolecular reaction, a single molecule converts into the product by
simply rearranging itself. However, the question that arises here is how these molecules get activated. The
mystery was solved by a British physicist, Frederick Alexander Lindemann, who proposed a time-leg between
activation and actual reaction. In other words, when ordinary molecules collide with each other, some of them
get activated, and the rate depends only upon these molecules but not the ordinary ones i.e.

Where k1, k2 and k3 are the rate constants for the different processes; while A and A* are the ordinary and
activated molecule. The overall rate of formation of the product can be given as

𝑑[𝑃] (161)
= 𝑘3 [𝐴∗ ]
𝑑𝑡
Now, since the molar concentration of [A*] is unknown we must apply the steady-state approximation on [A*].

Copyright © Mandeep Dalal


140 A Textbook of Physical Chemistry – Volume I

At steady state

Rate of formation of [A*] = Rate of disappearance of [A*]

𝑘1 [𝐴]2 = 𝑘2 [𝐴∗ ][𝐴] + 𝑘3 [𝐴∗ ] (162)

or

𝑘1 [𝐴]2 = (𝑘2 [𝐴] + 𝑘3 )[𝐴∗ ] (163)

𝑘1 [𝐴]2 (164)
[𝐴∗ ] =
𝑘2 [𝐴] + 𝑘3

Using the value of [A*] from equation (164) in equation (161), we get

𝑑[𝑃] 𝑘3 𝑘1 [𝐴]2 (165)


𝑅𝑎𝑡𝑒 = =
𝑑𝑡 𝑘2 [𝐴] + 𝑘3

Equation (165) gives rise to two possibilities discussed below.


1. If the concentration of reactant A is very high: In this situation, k2[A] ⋙ k3, and k3 can be neglected,
therefore, the equation (165) takes the form

𝑑[𝑃] 𝑘3 𝑘1 [𝐴]2 (166)


𝑅𝑎𝑡𝑒 = =
𝑑𝑡 𝑘2 [𝐴]

or

𝑑[𝑃] 𝑘3 𝑘1 (167)
= [𝐴]
𝑑𝑡 𝑘2

𝑑[𝑃] (168)
= 𝑘𝑜 [𝐴]
𝑑𝑡
Where ko is the overall rate constant. It is clear from the above result that the unimolecular reactions follow
first-order kinetics in such cases.
2. If the concentration of reactant A is very low: In this situation, k3 ⋙ k2[A] and k2[A] can be neglected,
therefore, the equation (165) takes the form

𝑑[𝑃] 𝑘3 𝑘1 [𝐴]2 (169)


𝑅𝑎𝑡𝑒 = =
𝑑𝑡 𝑘3

𝑑[𝑃] (170)
= 𝑘1 [𝐴]2
𝑑𝑡
Where k1 is the overall rate constant. It is clear from the above result that the unimolecular reactions follow
second-order kinetics in such cases.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 141

 Limitations of Collision Theory


The collision theory of reaction rate is extremely successful in rationalizing the kinetics of many
reactions, however, it does suffer from some serious limitations discussed below.
1. This theory finds application only to reactions occurring in the gas phase and solution having simple reactant
molecules.
2. The rate constants obtained by employing collision theory are found to be comparable to what has been
obtained from the Arrhenius equation only for the simple reactions but not for complex reactions.
3. This theory tells nothing about the exact mechanism behind the chemical reaction i.e. making and breaking
of chemical bonds.
4. The collision theory considers only the kinetic energy of reacting molecules and just ignored rotational and
vibrational energy which also plays an important role in reaction rate.
5. This theory did not consider the steric factor at all i.e. the proper orientation of the colliding molecules
needed to result in the chemical change.

 Steric Factor
One of the most glaring limitations of the collision is that the predicted values of rate constants for
many reactions were found to be considerably different from the values obtained experimentally. Moreover, it
was also noticed that more the complexity, the higher was the deviation. This happened because the collision
theory supposed that the particles participating in the chemical reaction are completely spherical, and thus, are
able to react in every direction. However, this is far from the truth since the orientation of the collisions is not
always appropriate to result in the chemical change. For instance, in the hydrogenation of ethylene, the
dihydrogen molecule must approach the bonding zone between the atoms, and not all the possible collisions
would be able to satisfy this requirement. For more clear view, consider the formation of CO2 as shown below.

To solve this problem, the concept of steric factor (ρ) was introduced, which is simply the ratio of
experimental value to the predicted value of the rate constant. In other words, the steric factor may be defined
as the ratio between the frequency factor and the collision frequency i.e.

𝐴𝑜𝑏𝑠𝑒𝑟𝑣𝑒𝑑 (171)
𝜌=
𝑍𝑐𝑎𝑙𝑐𝑢𝑙𝑎𝑡𝑒𝑑

Copyright © Mandeep Dalal


142 A Textbook of Physical Chemistry – Volume I

It is worthy to note that the value of the steric factor most of the cases is less than unity. Typically, it has been
seen that more the complex the reactant molecules are, the lower is the steric factor. However, some reactions
do have steric factors higher than unity; for instance, the harpoon reactions in which atoms involved exchange
electrons generating ions. The deviation from unity may arise due to different reasons such as non-spherical
shape of reacting molecules, or the partial delivery of kinetic energy, the presence of a solvent when applied
to solutions.
In order to derive the expression for modified collision theory that does consider the reactants steric,
recall the rate constant calculated from simple collision theory first i.e.

𝑅𝑎𝑡𝑒 = 𝑍𝑒 −𝐸𝑎/𝑅𝑇 (172)

For experimental rate, multiply the equation (172) by the probability or steric factor i.e.

𝑅𝑎𝑡𝑒 = 𝜌𝑍𝑒 −𝐸𝑎/𝑅𝑇 (173)

Now considering both the possibilities i.e. whether the reacting species are the same or different, we can
simplify the above equation in two ways:
i) For dissimilar molecules:
If the colliding molecules are not the same, the exponential part in equation (173) takes the form

(174)
𝑁 2 8𝜋𝑅𝑇(𝑀𝐴 + 𝑀𝐵 )
𝐴 = 𝜌 3 𝜎𝐴𝐵 √
10 𝑀𝐴 𝑀𝐵

Substituting the values of different constants, we get

(175)
2 √
𝑇(𝑀𝐴 + 𝑀𝐵 )
𝐴 = 2.753 × 1029 × 𝜌 × 𝜎𝐴𝐵
𝑀𝐴 𝑀𝐵

ii) For similar molecules:


If the colliding molecules are the same, the exponential part in equation (173) takes the form

(176)
𝑁 𝜋𝑅𝑇
𝐴 = 𝜌 3 4𝜎 2 √
10 𝑀𝐴

(177)
𝑇
𝐴 = 3.893 × 1029 × 𝜌 × 𝜎 2 √
𝑀𝐴

This modified collision theory can account for probability factors up to 10−4 but not less than that. This
limitation can be overcome by “transition state theory” discussed in the next section.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 143

 Activated Complex Theory


In 1935, an American chemist Henry Eyring; alongside two British chemists, Meredith Gwynne
Evans and Michael Polanyi; proposed a new theory to rationalize the rate of different chemical reactions which
was based upon the formation of an activated intermediate complex. This theory is also known as the
"transition state theory", "theory of absolute reaction rates”, and "absolute-rate theory".
The activated complex theory states that the rates of various elementary chemical reactions can be
explained by assuming a special type of chemical equilibria (quasi-equilibrium) between reactants and
activated complexes.
Before the development of activated complex theory, the Arrhenius rate law was popularly used to
determine energies for the potential barrier. However, the Arrhenius equation was based on empirical
observations rather than mechanistic investigations as if one or more intermediates are involved in the
conversion or not. For that reason, more development was essential to know the two factors present in the
Arrhenius equation, the activation energy (Ea) and the pre-exponential factor (A). The Eyring equation from
transition state theory successfully addresses these two issues and therefore contributed significantly to the
conceptual understanding of reaction kinetics.

Figure 11. The variation of free energy as the reaction proceeds.

To explore the concept mathematically, consider a reaction between reactant A and B forming a product P via
the activated complex X* as:

𝐾∗ 𝑘 (178)
𝐴 + 𝐵 ⇌ 𝑋∗ → 𝑃

The equilibrium constant for reactants to activated complex conversion is


[𝑋 ∗ ] (179)
𝐾 =
[𝐴][𝐵]

Copyright © Mandeep Dalal


144 A Textbook of Physical Chemistry – Volume I

Henry Eyring showed that the rate constant for a chemical reaction with any order or molecularity can be given
by the following relation.

𝑅𝑇 ∗ (180)
𝑘= 𝐾
𝑁ℎ
Where K* is the equilibrium constant for reactants to activated complex conversion at temperature T. Whereas,
R, N and h represent the gas constant, Avogadro number and Planck’s constant, respectively. Now, as we know
from thermodynamics

𝛥𝐺 ∗ = −𝑅𝑇 ln 𝐾 ∗ (181)

𝛥𝐺 ∗ (182)
− = ln 𝐾 ∗
𝑅𝑇
𝛥𝐺 ∗ (183)
𝐾 ∗ = 𝑒 − 𝑅𝑇

Where ΔG* is the free energy of activation for reactants to activated complex conversion step. Using the value
of K* from equation (183) into equation (180), we get

𝑅𝑇 −𝛥𝐺∗ (184)
𝑘= 𝑒 𝑅𝑇
𝑁ℎ
If we put the thermodynamic value of free energy i.e. ΔG* = ΔH* − TΔS* in equation (184), we get

𝑅𝑇 −(𝛥𝐻∗ −𝑇𝛥𝑆 ∗ ) (185)


𝑘= 𝑒 𝑅𝑇
𝑁ℎ
or

𝑅𝑇 𝛥𝑆 ∗ 𝛥𝐻 ∗ (186)
𝑘= × 𝑒 𝑅 × 𝑒 − 𝑅𝑇
𝑁ℎ
Where ΔH* and ΔS* are enthalpy change and the entropy change of the activation step. Equation (186) is
popularly known as the Eyring equation. Now, since the equation (186) contains very fundamental factors of
the reacting species, that is why this theory got its name of “theory of absolute reaction rates”.
 Significance of Entropy of Activation and Enthalpy of Activation
As far as the equation (184) is concerned, it can easily be seen that as the free energy change of the
activation step increases, the rate constant would decrease. However, if we look at the simplified form i.e.
equation (186), we find three factors; one is RT/Nh which is constant if the temperature is kept constant. The
second factor involves ΔS*, and therefore, we can conclude that the reaction rate would show exponential
increase if the entropy of activation increases. The third factor includes ΔH*, and therefore, we can conclude
that the reaction rate would show exponential decrease if the enthalpy of activation increases. It is also worthy
to note that the first two terms collectively make the frequency factor.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 145

 Comparison with Arrhenius Rate Constant


Like the collision theory, the validity of the “activated complex theory” must also be checked against
the results of the Arrhenius rate equation. In order to do so, recall the Arrhenius equation i.e.
𝐸𝑎 (187)
𝑘 = 𝐴 𝑒 −𝑅𝑇

By looking at the analogy with equation (187), we get

𝑅𝑇 𝛥𝑆 ∗ (188)
𝐴= ×𝑒 𝑅
𝑁ℎ
and
𝐸𝑎 𝛥𝐻 ∗ (189)
𝑒 −𝑅𝑇 = 𝑒 − 𝑅𝑇

Now we can use equation (188) and (189) to determine the value of entropy of activation and enthalpy of
activation.
1. Calculation of entropy of activation: In order to determine the entropy change of the activation step, take
the natural logarithm of the equation (188) i.e.

𝑅𝑇 𝛥𝑆 ∗ (190)
ln 𝐴 = ln + ln 𝑒 𝑅
𝑁ℎ
𝑅𝑇 𝛥𝑆 ∗ (191)
ln 𝐴 = ln +
𝑁ℎ 𝑅
or

𝛥𝑆 ∗ 𝑅𝑇 (192)
= ln 𝐴 − ln
𝑅 𝑁ℎ
𝑅𝑇 (193)
𝛥𝑆 ∗ = 𝑅 ln 𝐴 − 𝑅 ln
𝑁ℎ
Thus, higher is the value of the frequency factor, larger will be the entropy of activation.
2. Calculation of enthalpy of activation: In order to determine the entropy change of the activation step, we
must look at the equation (189) which suggests that the enthalpy change of the activation step as exactly equal
to the activation energy of the reaction dictating the rate considerably i.e. ΔH* = Ea. However, it has been
observed that the exact value of activation energy is slightly different than the enthalpy of activation. In order
to prove the aforementioned statement, rewrite the equation (186) as:

𝑅 𝛥𝑆 ∗ −
𝛥𝐻 ∗ (194)
𝑘= 𝑒 𝑅 × 𝑇 × 𝑒 𝑅𝑇
𝑁ℎ
or

Copyright © Mandeep Dalal


146 A Textbook of Physical Chemistry – Volume I

𝛥𝐻 ∗ (195)
𝑘 = 𝐶 × 𝑇 × 𝑒 − 𝑅𝑇

Where C is another constant. Now taking natural logarithm both side, equation (195) takes the form
𝛥𝐻 ∗ (196)
ln 𝑘 = ln 𝐶 + ln 𝑇 + ln 𝑒 − 𝑅𝑇

or

𝛥𝐻 ∗ (197)
ln 𝑘 = ln 𝐶 + ln 𝑇 −
𝑅𝑇
Now, differentiating both side w.r.t temperature, we get

d ln 𝑘 1 𝛥𝐻 ∗ (198)
= +
𝑑𝑡 𝑇 𝑅𝑇 2
We also know from the Arrhenius equation that
𝐸𝑎 (199)
𝑘 = 𝐴 𝑒 −𝑅𝑇

Now taking natural logarithm both side, equation (199) takes the form
𝐸𝑎 (200)
ln 𝑘 = ln 𝐴 + ln 𝑒 −𝑅𝑇

or

𝐸𝑎 (201)
ln 𝑘 = ln 𝐴 −
𝑅𝑇
Now, differentiating both side w.r.t temperature, we get

d ln 𝑘 𝐸𝑎 (202)
=
𝑑𝑡 𝑅𝑇 2
Comparing (198) and (202), we get

1 𝛥𝐻 ∗ 𝐸𝑎 (203)
+ 2
=
𝑇 𝑅𝑇 𝑅𝑇 2
𝑅𝑇 + 𝛥𝐻 ∗ 𝐸𝑎 (204)
2
=
𝑅𝑇 𝑅𝑇 2
𝛥𝐻 ∗ = 𝐸𝑎 − 𝑅𝑇 (205)

If n is the change in the number of moles of gas in going from reactant to activated complex, the result of
equation (205) takes the form 𝛥𝐻 ∗ = 𝐸𝑎 − 𝛥𝑛𝑔 𝑅𝑇.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 147

 Ionic Reactions: Single and Double Sphere Models


It is a quite well-known fact that the rate of ionic reactions is generally small, which is obviously due
to the larger magnitude of activation energies arising from the very strong nature of electrostatic interactions.
The magnitude of the frequency factor in ionic reactions is a function of ionic charges. The frequency factors
have larger values if the charges on the participating ions are opposite, while smaller values are obtained in
the case of like-charged ions. This behavior can be explained in terms of the kinetic theory of gases; which
suggests that oppositely charged ions are more prone to collision due to attraction than the ions colliding with
same charges (repulsive forces). Besides the collision theory, the activated complex theory also provides an
alternate explanation for the ionic reactions. In this section, we will discuss the rationalization of ionic reactions
on the basis of the single-sphere model and the double-sphere model in detail.
 Double Sphere Model
Before we discuss the double sphere model of the ionic reactions, a simplified surrounding must be
assumed. Although it would be an oversimplification of the actual situation, it is highly beneficial as far as
conceptual and quantitative understanding is concerned. To do so, the solvent is considered as continuous
surrounding with a ε as the dielectric constant.
According to this model, two ions, which can same or opposite charges, combine together to form an
activated complex. In the initial state, the ions are considered as discrete; while in the final state, they assumed
to form a dumbbell like coordination with ʻrʼ as the distance of separation between their centers.

Figure 12. The pictorial depiction of the double-sphere model of ionic reactions.

Now, if ZA and ZB are the charge numbers of the participating ions and x as the distance of separation, the force
of electrostatic interaction (FAB) between them can be given from the Coulomb’s law as:

Copyright © Mandeep Dalal


148 A Textbook of Physical Chemistry – Volume I

𝑍𝐴 𝑍𝐵 𝑒 2 (206)
𝐹𝐴𝐵 =
4𝜋𝜀0 𝜀𝑥 2

Where ε0 and ε are permittivities of the vacuum (8.854 × 10−12 C2 N−1 m−2) and the dielectric constant of the
solvent used, respectively. The symbol e represents the elementary charge and has a value equal to 1.6 × 10−19
C. The value of parameter varies from ∞ to r with the mutual approach of two ions. The amount of work done
in moving the two ions closer by an extant dx will be

𝑤𝑜𝑟𝑘 = 𝑓𝑜𝑟𝑐𝑒 × 𝑑𝑖𝑠𝑝𝑙𝑎𝑐𝑒𝑚𝑒𝑛𝑡 (207)

𝑑𝑤 = 𝐹𝐴𝐵 × 𝑑𝑥 (208)

𝑍𝐴 𝑍𝐵 𝑒 2 (209)
𝑑𝑤 = − 𝑑𝑥
4𝜋𝜀0 𝜀𝑥 2

The negative sign is an indicator of decreasing separation i.e. distance is reduced by dx. The total amount of
work done in moving the two ions from x = ∞ to x = r will be
𝑟
𝑍𝐴 𝑍𝐵 𝑒 2 (210)
𝑤=−∫ 𝑑𝑥
4𝜋𝜀0 𝜀𝑥 2

𝑍𝐴 𝑍𝐵 𝑒 2 (211)
𝑤=
4𝜋𝜀0 𝜀 𝑟

The work given the above equation is actually the potential energy of the system which would have a negative
sign for oppositely charged ions and positive sign if the ions have same charges. Furthermore, we can also say
that this work is the free energy change due to electrostatic interactions, therefore, multiplying it by Avogadro
number (N) would give the value of the corresponding molar free energy change (𝛥𝐺𝐸𝐼 ∗
) i.e.


𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (212)
𝛥𝐺𝐸𝐼 =
4𝜋𝜀0 𝜀 𝑟

Correcting the above equation for non-electrostatic contribution 𝛥𝐺𝑁𝐸𝐼



, the total molar free energy change for
the whole process can be given by the following relation.

∗ ∗ ∗
𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (213)
𝛥𝐺 ∗ = 𝛥𝐺𝑁𝐸𝐼 + 𝛥𝐺𝐸𝐼 = 𝛥𝐺𝑁𝐸𝐼 +
4𝜋𝜀0 𝜀 𝑟

Also, from the activated complex theory, we know that

𝑅𝑇 −𝛥𝐺∗ (214)
𝑘= 𝑒 𝑅𝑇
𝑁ℎ
After putting the value of ΔG* from equation (213) into equation (214), we get

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 149

𝑅𝑇 −(𝛥𝐺𝑅𝑇

𝑁𝐸𝐼 + 𝑁𝑍𝐴 𝑍𝐵 𝑒 )
2
(215)
𝑘= 𝑒 𝑅𝑇4𝜋𝜀0 𝜀 𝑟
𝑁ℎ

𝑅𝑇 −𝛥𝐺𝑁𝐸𝐼 −
𝑁𝑍𝐴 𝑍𝐵 𝑒 2
(216)
𝑘= 𝑒 𝑅𝑇 . 𝑒 𝑅𝑇4𝜋𝜀0 𝜀 𝑟
𝑁ℎ
Taking natural logarithm both side of equation (216), we get

𝑅𝑇

𝛥𝐺𝑁𝐸𝐼 −
𝑁𝑍𝐴 𝑍𝐵 𝑒 2
(217)
ln 𝑘 = ln + ln 𝑒 − 𝑅𝑇 + ln 𝑒 𝑅𝑇4𝜋𝜀0 𝜀 𝑟
𝑁ℎ
or

𝑅𝑇 𝛥𝐺𝑁𝐸𝐼 𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (218)
ln 𝑘 = ln − −
𝑁ℎ 𝑅𝑇 𝑅𝑇4𝜋𝜀0 𝜀 𝑟

Which can also be expressed as

𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (219)
ln 𝑘 = ln 𝑘0 −
𝑅𝑇4𝜋𝜀0 𝜀 𝑟

Where k0 represents the magnitude of the rate constant for the ionic reaction carried out in a solvent of infinite
dielectric constant so that the electrostatic interactions become zero.
 Single Sphere Model
Besides the double-sphere model, another theoretical model that is quite rationalizing is a single-
sphere model. Just like the double-sphere model, the solvent is also considered as a continuum with a ε as the
dielectric constant. However, the primary differentiating aspect of this model is that it considers the two ions,
which can same or opposite charges, to form a single-sphere activated complex.

Figure 12. The pictorial depiction of the single-sphere model of ionic reactions.

Copyright © Mandeep Dalal


150 A Textbook of Physical Chemistry – Volume I

In the initial state, the ions are considered as discrete; while in the final state, they assumed to form a single-
sphere activated complex with ʻr*ʼ as the overall radius. The rate law for this case was derived by Born by
considering the energy required to charge an ion in solution. Now suppose that we need to charge a conducting
sphere of radius r from an initial value of zero to the final value Ze. This can be visualized as a process in
which a very small charge is e.dλ (λ = 0 − Z) is carried from infinite to this sphere.
Now, if ZA and ZB are the charge numbers of the participating ions and x as the distance of separation
between the sphere and the “increment” at any time, the force of electrostatic interaction (dF) between them
can be given from the Coulomb’s law as:

𝜆𝑒 2 𝑑𝜆 (220)
𝑑𝐹 =
4𝜋𝜀0 𝜀𝑥 2

Where ε0 and ε are permittivities of the vacuum (8.854 × 10−12 C2 N−1 m−2) and the dielectric constant of the
solvent used, respectively. The symbol e represents the elementary charge and has a value equal to 1.6 × 10−19
C. The amount of work done in moving the “increment” closer by an extant dx will be

𝑑𝑤 = 𝑑𝐹 × 𝑑𝑥 (221)

𝜆𝑒 2 𝑑𝜆 (222)
𝑑𝑤 = 𝑑𝑥
4𝜋𝜀0 𝜀𝑥 2

The total amount of work done can be obtained by carrying out the double integration with respect to x = ∞ –
r and λ = 0 – Z i.e.
𝑍 𝑟 (223)
𝑒2 𝜆
𝑤= ∫ ∫ 2 𝑑𝜆 𝑑𝑥
4𝜋𝜀0 𝜀 𝑥
0 ∞

𝑍2𝑒2 (224)
𝑤=
8𝜋𝜀0 𝜀 𝑟

The work given the above equation is actually the contribution of the electrostatic interactions to the Gibbs
energy of the ion i.e.

𝑍2𝑒 2 (225)
𝐺𝐸𝐼 =
8𝜋𝜀0 𝜀 𝑟

In the light of the above correlation, the electrostatic contribution to the Gibbs free energy of discrete ions and
activated complex can be written as

𝑍𝐴2 𝑒 2 (225)
𝐺𝐸𝐼 (𝐴) =
8𝜋𝜀0 𝜀 𝑟𝐴

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 151

𝑍𝐵2 𝑒 2 (226)
𝐺𝐸𝐼 (𝐵) =
8𝜋𝜀0 𝜀 𝑟𝐵


(𝑍𝐵 + 𝑍𝐵 )2 𝑒 2 (227)
𝐺𝐸𝐼 =
8𝜋𝜀0 𝜀 𝑟 ∗

Hence, the change in electrostatic contribution can be obtained simply by subtracting the sum of individual
contributions from the overall contribution i.e.


(𝑍𝐵 + 𝑍𝐵 )2 𝑒 2 𝑍𝐴2 𝑒 2 𝑍𝐵2 𝑒 2 (228)
𝛥𝐺𝐸𝐼 = − −
8𝜋𝜀0 𝜀 𝑟 ∗ 8𝜋𝜀0 𝜀 𝑟𝐴 8𝜋𝜀0 𝜀 𝑟𝐵

or


𝑒 2 (𝑍𝐵 + 𝑍𝐵 )2 𝑍𝐴2 𝑍𝐵2 (229)
𝛥𝐺𝐸𝐼 = [ − − ]
8𝜋𝜀0 𝜀 𝑟∗ 𝑟𝐴 𝑟𝐵

Correcting the above equation for non-electrostatic contribution 𝛥𝐺𝑁𝐸𝐼



, the total molar free energy change for
the whole process can be given by the following relation.

∗ ∗ ∗
𝑁𝑒 2 (𝑍𝐵 + 𝑍𝐵 )2 𝑍𝐴2 𝑍𝐵2 (230)
𝛥𝐺 ∗ = 𝛥𝐺𝑁𝐸𝐼 + 𝛥𝐺𝐸𝐼 = 𝛥𝐺𝑁𝐸𝐼 + [ − − ]
8𝜋𝜀0 𝜀 𝑟∗ 𝑟𝐴 𝑟𝐵

Also, from the activated complex theory, we know that

𝑅𝑇 −𝛥𝐺∗ (231)
𝑘= 𝑒 𝑅𝑇
𝑁ℎ
After putting the value of ΔG* from equation (230) into equation (231), we get

𝑅𝑇 −(

𝛥𝐺𝑁𝐸𝐼
+
𝑁𝑒 2 2
(𝑍𝐵 +𝑍𝐵 )2 𝑍𝐴 𝑍2
− − 𝐵 ]) (232)
𝑅𝑇 𝑅𝑇8𝜋𝜀0 𝜀 [ 𝑟∗ 𝑟𝐴 𝑟𝐵
𝑘= 𝑒
𝑁ℎ
Taking natural logarithm both side of equation (232) and rearranging, we get

𝑅𝑇 𝛥𝐺𝑁𝐸𝐼 𝑁𝑒 2 (𝑍𝐵 + 𝑍𝐵 )2 𝑍𝐴2 𝑍𝐵2 (233)
ln 𝑘 = ln − − [ − − ]
𝑁ℎ 𝑅𝑇 𝑅𝑇8𝜋𝜀0 𝜀 𝑟∗ 𝑟𝐴 𝑟𝐵

Which can also be expressed as

𝑁𝑒 2 (𝑍𝐵 + 𝑍𝐵 )2 𝑍𝐴2 𝑍𝐵2 (234)


ln 𝑘 = ln 𝑘0 − [ − − ]
𝑅𝑇8𝜋𝜀0 𝜀 𝑟∗ 𝑟𝐴 𝑟𝐵

Where k0 represents the magnitude of the rate constant for the ionic reaction carried out in a solvent of infinite
dielectric constant so that the electrostatic interactions become zero.

Copyright © Mandeep Dalal


152 A Textbook of Physical Chemistry – Volume I

 Influence of Solvent and Ionic Strength


The rate of reaction in the case of ionic reactions is strongly dependent upon the nature of the solvent
used and the ionic strength. The single and double-sphere treatment of these reactions enables us to study their
effect in detail. In this section, we discuss the application and validity of solvent influence and ionic strength
on the reaction rate.
 Influence of the Solvent
In order to study the influence of solvent on the rate of ionic reactions, recall the rate equation derived
using the double sphere model i.e.

𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (235)
ln 𝑘 = ln 𝑘0 −
𝑅𝑇4𝜋𝜀0 𝜀 𝑟

Where ε0 and ε are permittivities of the vacuum (8.854 × 10−12 C2 N−1 m−2) and the dielectric constant of the
solvent used, respectively. The symbol e represents the elementary charge and has a value equal to 1.6 × 10−19
C. k0 represents the magnitude of the rate constant for the ionic reaction carried out in a solvent of infinite
dielectric constant so that the electrostatic interactions become zero. ZA and ZB are the charge numbers of the
participating ions. The symbol N and R represents the Avogadro number and gas constant, respectively.
Rearranging equation (235), we get

𝑁𝑍𝐴 𝑍𝐵 𝑒 2 1 (236)
ln 𝑘 = − + ln 𝑘0
𝑅𝑇4𝜋𝜀0 𝑟 𝜀

Which is clearly the equation of the straight line (y = mx + c) with a negative slope and positive intercept.
Therefore, it is obvious that the logarithm of the rate constant shows a linear variation with the reciprocal of
dielectric constant.

Figure 13. The plot of ln k vs 1/ε for a typical ionic reaction.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 153

It is quite obvious from the plot that equation (236) holds very good over a wide range of dielectric
contents; however, as the large deviations are observed at lower values of ε. Moreover, if ʻmʼ is the
experimental slope then from equation (236), we have

𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (237)
𝑚=−
𝑅𝑇4𝜋𝜀0 𝑟

Every term in the above equation is known apart from r, suggesting its straight forward determination from
the slope of ln k vs 1/ε. The values of r obtained from equation (237) are found to be quite comparable to other
methods, which in turn suggests its practical application.
Besides the calculation of r, the influence of dielectric constant of solvent can also be used to explain
the entropy of activation. In order to do so, recall from the principles of thermodynamics

𝜕𝐺 (238)
( ) = −𝑆
𝜕𝑇 𝑃

Also, the electrostatic contribution to the Gibbs free energy using the double sphere model is


𝑁𝑍𝐴 𝑍𝐵 𝑒 2 (239)
𝛥𝐺𝐸𝐼 =
4𝜋𝜀0 𝜀 𝑟

However, the only quantity which is temperature-dependent in the above equation is ε. Therefore,
differentiating equation (239) with respect to temperature at constant pressure gives


∗ )
𝜕(𝛥𝐺𝐸𝐼 𝑁𝑍𝐴 𝑍𝐵 𝑒 2 𝜕(1/𝜀) (240)
𝛥𝑆𝐸𝐼 = −( ) =− ( )
𝜕𝑇 𝑃
4𝜋𝜀0 𝑟 𝜕𝑇 𝑃

or


𝑁𝑍𝐴 𝑍𝐵 𝑒 2 𝜕𝜀) (241)
𝛥𝑆𝐸𝐼 = ( )
4𝜋𝜀0 𝜀 2 𝑟 𝜕𝑇 𝑃

or


𝑁𝑍𝐴 𝑍𝐵 𝑒 2 𝜕 ln 𝜀) (242)
𝛥𝑆𝐸𝐼 = ( )
4𝜋𝜀0 𝜀 𝑟 𝜕𝑇 𝑃

Therefore, knowing the dielectric constant of the solvent and r, the entropy of activation can be obtained.
Moreover, it is also worthy to note that the entropy of activation is negative and decreases with an increase in
ZAZB.
One more factor that affects the entropy of activation is the phenomena of “electrostriction” or the
solvent binding. This can be explained by considering the combination of two ions as of same and opposite
charges. If the ions forming activated complex are having one-unit positive charge each, the double-sphere
will have a total of two-unit positive charge. This would result in a very strong interaction between the activated

Copyright © Mandeep Dalal


154 A Textbook of Physical Chemistry – Volume I

complex and the surrounding solvent molecules. This would eventually result in a restriction of free movement
and hence decreased entropy. On the other hand, If the ions forming activated complex possess opposite
charges, the double-sphere would have less charge resulting in decreased electrostriction, and therefore,
increased entropy.

Figure 14. The dependence of entropy of activation on the solvent electrostriction in case of (left) same
charges and (right) opposite charges.

 Influence of Ionic Strength


In order to study the influence of ionic strength (I) on the rate of ionic reactions, we need to recall
the quantity itself first i.e.

1
𝑖=𝑛 (243)
𝐼 = ∑ 𝑚𝑖 𝑧𝑖2
2
𝑖=1

Where mi and zi are the molarity and charge number of ith species, respectively. For instance, the value of z
for Ca2+ and Cl– in CaCl2 are +2 and −1, respectively. It has been found that an increase the ionic strength
increases the rate of reaction if charges on the reacting species are of the same sign. On the other hand, the
reaction rate has been found to follow a declining trend with increasing ionic strength if reaction ions are of
opposite sign. The mathematical treatment of the abovementioned statement is discussed below.
To rationalize the effect of ionic strength of the solution on the rate of reaction in case of ionic
reactions, consider a typical case i.e.

𝐴 𝑍𝐴 + 𝐵 𝑍𝐵 → 𝑋𝑍𝐴 +𝑍𝐵 → 𝑃 (244)

A Danish physical chemist, J. N. Brønsted, proposed the rate equation relating reaction-rate (R) and activity
coefficient as
𝑦𝐴 𝑦𝐵 (245)
𝑅 = 𝑘0 [𝐴][𝐵]
𝑦𝑋

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 155

Where yA, yB and yX are the activity coefficients for the reactant A, B and the activated complex X, respectively.
Brønsted collectively labeled the term yAyB/yX as the “kinetic activity factor”, and it was found be quite accurate
with experimental data. Now, rearranging equation (245), we get

𝑅 𝑦𝐴 𝑦𝐵 (246)
= 𝑘0
[𝐴][𝐵] 𝑦𝑋

Since the left-hand side simply equals to a second-order rate constant, the above equation takes the form
𝑦𝐴 𝑦𝐵 (247)
𝑘 = 𝑘0
𝑦𝑋

Taking logarithm both side, we get


𝑦𝐴 𝑦𝐵 (248)
log 𝑘 = log 𝑘0 + log
𝑦𝑋

or

log 𝑘 = log 𝑘0 + log 𝑦𝐴 + log 𝑦𝐵 − log 𝑦𝑋 (249)

Now the correlation of mean ionic activity coefficient with the ionic strength is given by famous Debye-Huckel
theory i.e.

log 𝑦𝑖 = −𝐵𝑧𝑖2 √𝐼 (250)

Where B is the Debye-Huckel constant. Using the concept of equation (250) in equation (249), we get

log 𝑘 = log 𝑘0 − 𝐵𝑧𝐴2 √𝐼 − 𝐵𝑧𝐵2 √𝐼 + 𝐵𝑧𝑋2 √𝐼 (251)

or

log 𝑘 = log 𝑘0 − 𝐵𝑧𝐴2 √𝐼 − 𝐵𝑧𝐵2 √𝐼 + 𝐵(𝑧𝐴 + 𝑧𝐵 )2 √𝐼 (252)

log 𝑘 = log 𝑘0 + 𝐵[(𝑧𝐴 + 𝑧𝐵 )2 − 𝑧𝐴2 − 𝑧𝐵2 ]√𝐼 (253)

log 𝑘 = log 𝑘0 + 𝐵[𝑧𝐴2 + 𝑧𝐵2 + 2𝑧𝐴 𝑧𝐵 − 𝑧𝐴2 − 𝑧𝐵2 ]√𝐼 (254)

or simply

log 𝑘 = log 𝑘0 + 2𝐵𝑧𝐴 𝑧𝐵 √𝐼 (255)

Rearranging the above equation, we get

log 𝑘 = 2𝐵𝑧𝐴 𝑧𝐵 √𝐼 + log 𝑘0 (256)

Which is clearly the equation of straight line (y = mx + c) with a positive slope (2𝐵𝑧𝐴 𝑧𝐵 ) and positive intercept
(log 𝑘0 ). For aqueous solutions at 25°C, 2B = 1.02, equation (256) takes the form

Copyright © Mandeep Dalal


156 A Textbook of Physical Chemistry – Volume I

log 𝑘 − log 𝑘0 = 1.02 𝑧𝐴 𝑧𝐵 √𝐼 (257)

or

𝑘 (258)
log = 1.02 𝑧𝐴 𝑧𝐵 √𝐼
𝑘0

Where k0 is the rate constant at zero ionic strength and can be obtained by the extrapolation of log k0 vs square
root of the ionic strength.

Figure 15. The plot of log (k/k0) vs (I)1/2 for different ionic reactions in aqueous solution at 25°C.

The above equation can also be extended to explain the dependence of reaction-rate on ionic strength
for third-order reactions. To do so, consider

𝐴 𝑍𝐴 + 𝐵 𝑍𝐵 + 𝐶𝑍𝐶 → 𝑋𝑍𝐴 +𝑍𝐵 +𝑍𝐶 → 𝑃 (259)

Following the same route as in second-order reactions, we will get

log 𝑘 = log 𝑘0 + 𝐵[(𝑧𝐴 + 𝑧𝐵 + 𝑧𝐶 )2 − 𝑧𝐴2 − 𝑧𝐵2 − 𝑧𝐶2 ]√𝐼 (260)

log 𝑘 = log 𝑘0 + 2𝐵(𝑧𝐴 𝑧𝐵 + 𝑧𝐵 𝑧𝐶 + 𝑧𝐶 𝑧𝐴 )√𝐼 (261)

Therefore, a negative slope will be observed if (𝑧𝐴 𝑧𝐵 + 𝑧𝐵 𝑧𝐶 + 𝑧𝐶 𝑧𝐴 ) is negative while a positive slope is
expected for a positive value of (𝑧𝐴 𝑧𝐵 + 𝑧𝐵 𝑧𝐶 + 𝑧𝐶 𝑧𝐴 ).

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 157

 The Comparison of Collision and Activated Complex Theory


After studying the collision as well as the transition state theories in detail, it is time to highlight the
key points of similarities and differences between the two. A comparative analysis of both theories is quite
beneficial as far as the practicality is concerned.

Table 1. The side-by-side comparison between the collision theory and transition state theory.

Collision Theory Activated Complex Theory

1. According to the collision theory, the chemical 1. According to the transition state theory, the
reactions occur when the reactant molecules collide primary cause of the reaction is actually the
with a sufficient amount of kinetic energy. formation of an activated complex or the transition
state, which in turn, converts to the final product.

2. It is based upon the kinetic theory of gases. 2. It is derived from the fundamentals of
thermodynamics.

3. This theory considers the activation energy as the 3. This theory assumes the activation energy as the
minimum energy required to make the collision difference between the energy of the reacting
effective. molecules and the energy of the activated complex.

4. This theory tells nothing about the entropy of 4. The transition state theory enables us to measure
activation. the entropy of activation.

5. Collision theory is applicable to simple chemical 5. This theory provided reasonable predictions even
reactions and large deviations with experimental for the complex reactions.
results are observed as the complexity increases.

6. The incorporation of the correction factor in 6. The incorporation of correction factor was
modified collision theory was arbitrary. justified in terms of entropy of activation i.e. ΔS*.

7. This theory tells nothing about the mechanism 7. The formation of the activated complex is very
involved. much correlated with the actual mechanism going
on.

Copyright © Mandeep Dalal


158 A Textbook of Physical Chemistry – Volume I

 Problems
Q 1. Discuss the fundamental concept of the effect of temperature on the reaction-rate with special reference
to Maxwell-Boltzmann distribution of energies.
Q 2. Derive Arrhenius equation for the rate of reactions. How it can be used to determine the activation energy?
Q 3. Deduce the rate expression for the first order opposed by first-order reactions. How it can be used to yield
the value of backward reaction too?
Q 4. Derive and discuss the rate law for second-order opposed by second order.
Q 5. What are the consecutive reactions? Discuss the condition required to apply the steady-state
approximation to a typical consecutive reaction.
Q 6. Define concurrent or parallel reactions. Derive the rate law for the first-order case.
Q 7. Discuss the collision of the theory bimolecular reaction when the colliding particles are dissimilar.
Q 8. What are unimolecular reactions? How they are treated in the collision framework?
Q 9. Discuss the limitations of collision theory and the incorporation of the steric factor?
Q 10. Derive the rate law in the activated complex framework. How it can be used to determine the entropy of
activation?
Q 11. What is the single-sphere model of ionic reactions? How is it different from the double sphere model?
Q 12. Discuss the influence of ionic strength on the rate of ionic reactions of the third order.
Q 13. Give five points of difference between the collision and activated complex theory.

Copyright © Mandeep Dalal


CHAPTER 3 Chemical Dynamics – I 159

 Bibliography
[1] B. R. Puri, L. R. Sharma, M. S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2008.
[2] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[3] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[4] S. A. Arrhenius, Über die Dissociationswärme und den Einfluß der Temperatur auf den Dissociationsgrad
der Elektrolyte. Z. Phys. Chem. 4 (1889) 96–116.
[5] S. A. Arrhenius, Über die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Säuren Z.
Phys. Chem. 4 (1889) 226–248.
[6] K. J. Laidler, Chemical Kinetics, Harper & Row, New York, USA, 1987.
[7] K. J. Laidler, The World of Physical Chemistry, Oxford University Press, Oxford, UK, 1993.
[8] H. Eyring, The Activated Complex in Chemical Reactions, J. Chem. Phys. 3 (1935) 107-115.

Copyright © Mandeep Dalal


CHAPTER 4
Electrochemistry – I: Ion-Ion Interactions
 The Debye-Huckel Theory of Ion-Ion Interactions
When an ion goes into the solution, it interacts not only with the solvent molecules but also with other
ions that are produced as a result of electrolytic dissociation. The interaction of this ion with its solvent-
surrounding is quite important as it helps the ion to get stabilized in that medium. Nevertheless, the other types
of interactions, i.e., the ion-ion interactions, are also of very much significant value to the overall understanding
of the system because these interactions affect a number of properties of these solutions such as partial molar
free energy change or the electrolytic conductance. Now, since the phenomenon of ion-ion interactions is a
function of interionic separations, which in turn depends upon the population density of these ions; the first
step to know this phenomenon must be the rationalization of ionic population density.
In the case of weak electrolytes, the degree of dissociation is very small at high concentrations
yielding a very low population density of charge carriers. This would result in almost zero ion-ion interactions
at high concentrations. Now, although the degree of dissociation increases with dilution, which in turn also
increases the total number of charge carriers, the population density remains almost unchanged since the extra
water has also been added for these extra ions. Thus, we can conclude that there are no ion-ion interactions in
weak electrolytes, neither at higher nor at the lower concentrations. On the other hand, in the case of strong
electrolytes, the degree of dissociation is a hundred percent even at high concentrations yielding a very high
population density of charge carriers. This would result in very strong ion-ion interactions at high
concentrations. Now, when more and more solvent is added, the total number of charge carriers remains the
same but the population density decreases continuously, creating large interionic separations. This would result
in a decrease in the magnitude of ion-ion interaction with increasing dilution.
In 1913, a British physicist Samuel Milner proposed the first quantitative explanation of ion-ion
interactions in a statistical framework. After that, in 1918, an Indian Chemist, Sir Jnan Chandra Ghosh
proposed a model to compute the energy of Coulombic interaction of the ions with an assumption that the ions
in the solution are fixed just like in the crystals. The Milner’s model was quite complicated mathematically,
while Ghosh’s model did not consider the distorting effect of thermal motion on the ionic distribution in the
solution phase of electrolyte under consideration.
In 1923, Peter Debye and Erich Huckel proposed an extremely important idea to quantify the ion-ion
interactions in the solutions of strong electrolytes. In this model, the solute, i.e., the strong electrolyte, is
assumed to dissociate completely. They considered the ions as perfect spheres that cannot be polarized by the
surrounding electric field while the solvation of ions was ignored. The solvent plays no role other than
providing a medium of the constant relative permittivity (dielectric constant). A reference ion was thought to
be suspended in solvent-continuum of dielectric constant ε with zero electrostriction, and this ion is surrounded
by oppositely charged ionic cloud.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 161

 Fundamental Concept
It is quite interesting to recall the fact that like many other coupling effects, the phenomena of ion-
ion interaction can also be quantified in terms of free energy change. In this case, it can be achieved by
assuming an initial state with no ion-ion interaction and a final state in which the ion-ion interactions do exist;
and then calculating the free energy change in going from initial to the final state.
The initial state, i.e., the state with no ion-ion interaction, can be created hypothetically by taking
ions in the vacuum. However, since the final state possesses ions in solution, the initial state should also have
ions in solution and not in the vacuum, otherwise, two states would differ not only by ion-ion interaction but
also in respect of ion-solvent interaction. Conversely, if we imagine some discharged ions in the solution
(discharged ions mean no ion-ion interaction) and then charge them up fully, free energy change of ion-ion
interaction would be the amount of work done in doing so.

Figure 1. The pictorial representation of the fundamental idea for calculating the free energy of ion-ion in
the Debye-Huckel model of ionic interactions.

It should also be noted that the procedure depicted in ‘Figure 1’ gives the work of charging (i.e., the
free energy of ion-ion interaction) for all the ions present in the solution whether they are positive or negative.
However, the scientific analysis of electrolytic solutions typically needs the free energy of ionic interactions
for individual species. This partial free energy change is called as chemical potential change and is typically
denoted as Δμi-I. The value of Δμi-I for ith species of radius 𝑟𝑖 can be obtained by computing the work of
charging this sphere from a neutral state to its full charge of Zie0. The work of charging (w) a spherical
conductor can be given by the following relation (from electrostatics):

1 (1)
𝑤 = [𝐶ℎ𝑎𝑟𝑔𝑒 𝑜𝑛 𝑡ℎ𝑒 𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑜𝑟 × 𝐸𝑙𝑒𝑐𝑡𝑟𝑜𝑠𝑡𝑎𝑡𝑖𝑐 𝑃𝑜𝑡𝑒𝑛𝑡𝑖𝑎𝑙 𝑜𝑓 𝑐𝑜𝑛𝑑𝑢𝑐𝑡𝑜𝑟]
2
Since the Debye-Huckel model also considers the ions as spherical conductors, the work of charging an ion of
radius r will be

Copyright © Mandeep Dalal


162 A Textbook of Physical Chemistry – Volume I

1 (2)
𝑤 = [𝑍𝑖 𝑒0 × 𝜓]
2
Where ψ represents the electrostatic potential of the reference ion due to the influence on it by the Coulombic
forces of the neighboring ions. Multiplying equation (2) by Avogadro number, we can obtain the amount of
work done required to charge one mole of such ions i.e. the partial molar free energy change.

𝑁𝐴 (3)
𝛥𝜇𝑖−𝐼 = 𝑁𝐴 𝑤 = [𝑍 𝑒 × 𝜓]
2 𝑖 0
The only thing that is unknown in the above expression is ψ. Therefore, the partial molar free energy change
due to the interaction of ith species with the rest of the solution requires the determination of electrostatic
potential exerted at the reference ion by its ionic-surrounding. Now, from the law of superposition of potentials,
we know that the potential at any point due to an assembly of charges is just the sum of the individual potentials
exerted at that point by the individual charges. Since the electrostatic potentials depend upon the distances
between the charges under consideration, the time-averaged distribution of all ions around must be known
(distances from the reference ion are the function of the population distribution of other ions).
Although we know that the total electrostatic potential at the reference ion can be obtained by
knowing the distance of each surrounding ion from the reference ion, the number of ions is extremely large
making it almost impossible to follow the conventional route to do so. So, instead of dealing with individual
ions, Debye and Huckel proposed the revolutionary idea of considering all the surrounding ions as a continuous
ionic cloud that surrounds the reference ions. They assigned the discrete charge only to the reference ion which
is situated in the solvent’s continuum of dielectric constant ε, i.e., water in this case.

Figure 2. The pictorial representation of reference ion in its ionic cloud.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 163

Hence, the Debye-Huckel’s idea transformed the complex problem of computing the population
density around the reference to the determination of excess charge density (ρ). At this point, one might think
why the excess charge density is not zero everywhere because the solution as a whole is actually electroneutral.
The reason for this behavior lies in the fact that positive centers attract negative density while a negative center
tries to accumulate a positive cloud around it. Therefore, in a volume element near positive ion, there should
be a more negative charge, i.e., excess negative charge density and vice-versa. This excess negative charge
density near a positive center is counterbalanced by excess positive charge density near a negative center,
keeping the overall electroneutrality maintained. It should also be mentioned here that since the thermal forces
in the solution are always trying to randomize the ionic distribution, the excess charge density diminishes in
moving away from the center.
 Mathematical Development
So far we have discussed the fundamentals aspects of how we can determine the partial molar free
energy change generated by the ionic interactions of ith species. All we need is the electrostatic potential (ψ)
of the reference ion due to the influence on it by the Coulombic forces of the surrounding. Since ψ, at a distance
r from the reference ion, depends upon the excess charge density at r, the first thing we need to develop it
mathematically is the correlation between these two parameters, i.e., ψr and ρr. One such relation for spherically
symmetric charge distribution is the Poisson’s equation in electrostatics that can be given as:

1 𝑑 2 𝑑𝜓𝑟 4𝜋 (4)
2
(𝑟 ) = − 𝜌𝑟
𝑟 𝑑𝑟 𝑑𝑟 𝜀
Where ψr and ρr are electrostatic potential and excess charge density in a very small volume element dV situated
at distance r from the reference ion.

Figure 3. The depiction of electrostatic potential and excess charge density in dV volume element at a
distance r from the reference ion.

Copyright © Mandeep Dalal


164 A Textbook of Physical Chemistry – Volume I

Rearranging the Poisson’s equation for excess charge density, we get

𝜀 1 𝑑 2 𝑑𝜓𝑟 (5)
𝜌𝑟 = − [ (𝑟 )]
4𝜋 𝑟 2 𝑑𝑟 𝑑𝑟
Also, the “linearized Boltzmann distribution” can be used to prove that

𝑛𝑖0 𝑍𝑖2 𝑒02 𝜓𝑟 (6)


𝜌𝑟 = − ∑
𝑘𝑇
𝑖

Where 𝑛𝑖0 is the bulk concentration of the ith species and k is simply the Boltzmann constant. Equating the
Poisson’s expression with the Boltzmann Formula, i.e., from equation (5) and (6), we get the linearized Poisson
Boltzmann equation as:

𝜀 1 𝑑 2 𝑑𝜓𝑟 𝑛𝑖0 𝑍𝑖2 𝑒02 𝜓𝑟 (7)


− [ 2 (𝑟 )] = − ∑
4𝜋 𝑟 𝑑𝑟 𝑑𝑟 𝑘𝑇
𝑖

or

1 𝑑 2 𝑑𝜓𝑟 4𝜋 (8)
2
(𝑟 )=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 ) 𝜓𝑟
𝑟 𝑑𝑟 𝑑𝑟 𝜀𝑘𝑇
𝑖

Now assume a constant κ2 with value

4𝜋 (9)
𝜅2 = ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02
𝜀𝑘𝑇
𝑖

Using the value of equation (9) in equation (8), we get

1 𝑑 2 𝑑𝜓𝑟 (10)
(𝑟 ) = 𝜅 2 𝜓𝑟
𝑟 2 𝑑𝑟 𝑑𝑟
The above differential equation is the popular form of linearized Poisson-Boltzmann expression and its
solution gives

𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 (11)
𝜓𝑟 =
𝜀 𝑟
Now comparing equation (4) and equation (10), we get

4𝜋 (12)
𝜅 2 𝜓𝑟 = − 𝜌
𝜀 𝑟
or

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 165

𝜀𝜅 2 𝜓𝑟 (13)
𝜌𝑟 = −
4𝜋
Now putting the value of electrostatic potential from equation (11) in equation (13), we get the expression for
excess charge density as a function of r.

𝜀𝜅 2 𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 (14)
𝜌𝑟 = − ×
4𝜋 𝜀 𝑟
𝑍𝑖 𝑒0 𝜅 2 −𝜅𝑟 (15)
𝜌𝑟 = − 𝑒
4𝜋𝑟
Now because the magnitude of ρr is a consequence of the unequal distribution of anions and cations, the above
expression also defines the ionic-population-distribution around the reference ion.

Figure 4. The variation of electrostatic potential and excess charge density as a function of the distance r
from the reference ion.

After knowing that the excess charge density decreases exponentially with r, we can determine the total excess
charge (dq) at the same distance just by the following relation.

𝑑𝑞 = 𝜌𝑟 4𝜋𝑟 2 𝑑𝑟 (16)

Where 4πr2dr is the volume element of a hollow spherical shell of thickness dr with the inner and outer radius
as r and r+dr, respectively. After putting the value of ρr from equation (15) in equation (16), we have

𝑍𝑖 𝑒0 𝜅 2 −𝜅𝑟 (17)
𝑑𝑞 = (− 𝑒 ) 4𝜋𝑟 2 𝑑𝑟
4𝜋𝑟

or

Copyright © Mandeep Dalal


166 A Textbook of Physical Chemistry – Volume I

𝑑𝑞 = −𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 𝜅 2 𝑟𝑑𝑟 (18)

Thus, the exponential part decreases with increasing r while the non-exponential part shows a continuous
increase as we move away from the reference ion.

Figure 5. The depiction of excess charge density in a volume element of thickness ‘dr’ as a function of the
distance r from the reference ion.

In order to find the distance of maximum “excess charge”, we need to differentiate the equation (18) with
respect to r, and then derivative needs to be put equal to zero i.e.

𝑑𝑞 (19)
= −𝑍𝑖 𝑒0 𝜅 2 (𝑒 −𝜅𝑟 − 𝜅𝑟𝑒 −𝜅𝑟 ) = 0
𝑑𝑟
Since −𝑍𝑖 𝑒0 𝜅 2 is non zero for sure, the above equation holds true only if

𝑒 −𝜅𝑟 − 𝜅𝑟𝑒 −𝜅𝑟 = 0 (20)

Which implies

𝑟 = 𝜅 −1 (21)

Also from equation (9), we have


1/2 (22)
4𝜋
𝜅=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 )
𝜀𝑘𝑇
𝑖

Hence, the radius of maximum excess charge from the center of the reference ion can be obtained by putting
the value of κ from equation (22) in equation (21), we get rmax or the “Debye-Huckel length” as

𝜀𝑘𝑇 1
1/2 (23)
𝑟𝑚𝑎𝑥 =( 0 )
4𝜋 ∑𝑖 𝑛𝑖 𝑍𝑖2 𝑒02

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 167

At this point, we have understood most of the ideas of Debye-Huckel’s theory of ion-ion interaction,
and the question we raised initially is near to its end, i.e., to calculate the molar chemical potential change of
ion-ion interaction. Now, although we have calculated the electrostatic potential, we cannot use the same in
equation (3) because what we have obtained contains the contributions from the ionic cloud as well as from
the reference ion itself. Therefore, before we calculate the partial molar free energy of ion-ion interaction, we
need to separate potential from the ionic cloud and from the central ion first. The total electrostatic potential
at a distance r can be fragmented as given below.

𝜓𝑟 = 𝜓𝑖𝑜𝑛 + 𝜓𝑐𝑙𝑜𝑢𝑑 (24)

𝜓𝑐𝑙𝑜𝑢𝑑 = 𝜓𝑟 − 𝜓𝑖𝑜𝑛 (25)

From the formulation of potential due to a single charge at a distance r, we know that

𝑍𝑖 𝑒0 (26)
𝜓𝑖𝑜𝑛 =
𝜀𝑟
Now using the value of ψr from equation (11) and of ψion from equation (26) into equation (25), we get

𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 𝑍𝑖 𝑒0 (27)
𝜓𝑐𝑙𝑜𝑢𝑑 = −
𝜀 𝑟 𝜀𝑟
𝑍𝑖 𝑒0 −𝜅𝑟 (28)
𝜓𝑐𝑙𝑜𝑢𝑑 = (𝑒 − 1)
𝜀𝑟

Now because the value of κ is proportional to ∑𝑖 𝑛𝑖0 𝑍𝑖2 𝑒02 , at very large dilution κ becomes very small and 𝑒 −𝜅𝑟
and can be expended as 1 − 𝜅𝑟 (ex = 1+ x). Therefore, equation (28) takes the form

𝑍𝑖 𝑒0 (29)
𝜓𝑐𝑙𝑜𝑢𝑑 = (1 − 𝜅𝑟 − 1)
𝜀𝑟
𝑍𝑖 𝑒0 𝜅 (30)
𝜓𝑐𝑙𝑜𝑢𝑑 = −
𝜀
Hence, potential due to ionic cloud is independent of the distance r in this case. Now putting the value of ψcloud
from equation (30) into equation (3), we get

𝑁𝐴 𝑍𝑖 𝑒0 𝜅 (31)
𝛥𝜇𝑖−𝐼 = 𝑁𝐴 𝑤 = [𝑍𝑖 𝑒0 × (− )]
2 𝜀

𝑁𝐴 𝑍𝑖2 𝑒02 𝜅 (32)


𝛥𝜇𝑖−𝐼 = −
2𝜀
Hence, the Debye-Huckel’s model gives the theoretical value of 𝛥𝜇𝑖−𝐼 arising from ion-ion interactions. Now
before we check the validity of the Debye-Huckel model, we will discuss the many aspects of mathematical
treatment Debye-Huckel model in more detail in forthcoming sections.

Copyright © Mandeep Dalal


168 A Textbook of Physical Chemistry – Volume I

 Potential and Excess Charge Density as a Function of Distance from the


Central Ion
In order to determine the potential and excess charge density as a function of radial distance r,
consider a very small volume element dV, situated at distance r from the reference ion, in which ψr and ρr are
electrostatic potential and excess charge density, respectively.

Figure 6. The depiction of electrostatic potential and excess charge density in dV volume element at
distance r from the reference ion.

Since we want to study how electrostatic potential and excess charge density depend upon the distance from
the reference ion, the first thing we need to develop it mathematically is a correlation of these two parameters
i.e. ψr and ρr. One such relation for spherically symmetric charge distribution is the Poisson’s equation in
electrostatics which can be given as:

1 𝑑 2 𝑑𝜓𝑟 4𝜋 (33)
2
(𝑟 ) = − 𝜌𝑟
𝑟 𝑑𝑟 𝑑𝑟 𝜀
Now the excess charge density in the volume element dV can be obtained by multiplying the total number of
ions per unit volume with their corresponding charges i.e.

𝜌𝑟 = 𝑛1 𝑍1 𝑒0 + 𝑛2 𝑍2 𝑒0 + 𝑛3 𝑍3 𝑒0 … . . 𝑛𝑖 𝑍𝑖 𝑒0 (34)

or

𝜌𝑟 = ∑ 𝑛𝑖 𝑍𝑖 𝑒0 (35)

Where 𝑛1 is the number of first kind ions with 𝑍1 𝑒0 charge, 𝑛2 is the number of second kind ions with 𝑍2 𝑒0
charge and so on up to ith type.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 169

Now, from the Boltzmann distribution law of classical statistical mechanics, we know that

𝑛𝑖 = 𝑛𝑖0 𝑒 −𝑈/𝑘𝑇 (36)

Where U represents the total change in the potential energy of the ith particle in going from the bulk
concentration 𝑛𝑖0 to the actual concentration of the ith particle i.e. 𝑛𝑖 . At this stage, three cases arise, the
magnitude of U can be zero, positive or negative. If U = 0 i.e. there are no ion-ion interaction, 𝑛𝑖 = 𝑛𝑖0 , which
implies that concentration near the reference ion will be equal to the bulk concentration. If U = −, i.e., there
are attractive ion-ion interaction, 𝑛𝑖 > 𝑛𝑖0, which implies that concentration near the reference ion will be
higher to the bulk concentration. In the third scenario, if U = +, i.e., there are repulsive ion-ion interaction,
𝑛𝑖 < 𝑛𝑖0, which implies that concentration near the reference ion will be less to the bulk concentration. Now
according to the Debye-Huckel model, only simple Coulombic forces need to be considered for very dilute
solutions. Therefore, excluding all other short-range interactions like dispersion ones, the potential of average
force U simply can be written as given below.

𝑈 = 𝑍𝑖 𝑒0 𝜓𝑟 (37)

After using the value of U from equation (37) in equation (36), we get

𝑛𝑖 = 𝑛𝑖0 𝑒 −𝑍𝑖𝑒0 𝜓𝑟 /𝑘𝑇 (38)

Putting the value of 𝑛𝑖 from equation (38) in equation (35), we have

𝜌𝑟 = ∑ 𝑛𝑖0 𝑍𝑖 𝑒0 𝑒 −𝑍𝑖𝑒0 𝜓𝑟 /𝑘𝑇 (39)

Now because the Debye-Huckel model considers the solutions in which ψr is much less than kT, we can
conclude that 𝑍𝑖 𝑒0 𝜓𝑟 ≪ 𝑘𝑇. Therefore, 𝑒 −𝑍𝑖𝑒0 𝜓𝑟 /𝑘𝑇 can be expended as

𝑍𝑖 𝑒0 𝜓𝑟 (40)
𝜌𝑟 = ∑ 𝑛𝑖0 𝑍𝑖 𝑒0 (1 − )
𝑘𝑇
or

𝑛𝑖0 𝑍𝑖2 𝑒02 𝜓𝑟 (41)


𝜌𝑟 = ∑ 𝑛𝑖0 𝑍𝑖 𝑒0 − ∑
𝑘𝑇
The first term of the above equation gives the net charge on the whole of the solution, and it must be zero since
the overall electrical neutrality is maintained. Therefore, the above equation takes the form

𝑛𝑖0 𝑍𝑖2 𝑒02 𝜓𝑟 (42)


𝜌𝑟 = − ∑
𝑘𝑇
The above equation is the “linearized Boltzmann distribution”.

Copyright © Mandeep Dalal


170 A Textbook of Physical Chemistry – Volume I

Rearranging the Poisson’s equation for excess charge density i.e. equation (33), we get

𝜀 1 𝑑 2 𝑑𝜓𝑟 (43)
𝜌𝑟 = − [ (𝑟 )]
4𝜋 𝑟 2 𝑑𝑟 𝑑𝑟

Where 𝑛𝑖0 is the bulk concentration of the ith species and k is simply the Boltzmann constant. Equating the
Poisson’s expression with the Boltzmann Formula i.e. from equation (43) and (42), we get the linearized
Poisson Boltzmann equation as:

𝜀 1 𝑑 2 𝑑𝜓𝑟 𝑛𝑖0 𝑍𝑖2 𝑒02 𝜓𝑟 (44)


− [ 2 (𝑟 )] = − ∑
4𝜋 𝑟 𝑑𝑟 𝑑𝑟 𝑘𝑇
𝑖

or

1 𝑑 2 𝑑𝜓𝑟 4𝜋 (45)
(𝑟 ) = ( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 ) 𝜓𝑟
𝑟 2 𝑑𝑟 𝑑𝑟 𝜀𝑘𝑇
𝑖

Now assume a constant κ2 with value

4𝜋 (46)
𝜅2 = ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02
𝜀𝑘𝑇
𝑖

Using the value of equation (46) in equation (45), we get

1 𝑑 2 𝑑𝜓𝑟 (47)
(𝑟 ) = 𝜅 2 𝜓𝑟
𝑟 2 𝑑𝑟 𝑑𝑟
To solve the above differential equation, assume that ψr is a function of a new variable, called μ, as
𝜇 (48)
𝜓𝑟 =
𝑟
Differentiating equation (48), we get

𝑑𝜓𝑟 𝜇 1 𝑑𝜇 (49)
=− 2+
𝑑𝑟 𝑟 𝑟 𝑑𝑟
Multiplying both sides by r2

𝑑𝜓𝑟 𝑑𝜇 (50)
𝑟2 = −𝜇 + 𝑟
𝑑𝑟 𝑑𝑟
Now first multiplying both sides by d/dr and then by 1/r2, we get

1 𝑑 2 𝑑𝜓𝑟 1 𝑑 𝑑𝜇 (51)
2
(𝑟 ) = 2 (−𝜇 + 𝑟 )
𝑟 𝑑𝑟 𝑑𝑟 𝑟 𝑑𝑟 𝑑𝑟
or

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 171

1 𝑑 2 𝑑𝜓𝑟 1 𝑑𝜇 𝑑2 𝜇 𝑑𝜇 (52)
(𝑟 ) = 2 (− +𝑟 2+ )
𝑟 2 𝑑𝑟 𝑑𝑟 𝑟 𝑑𝑟 𝑑𝑟 𝑑𝑟

1 𝑑 2 𝑑𝜓𝑟 1 𝑑2 𝜇 (53)
(𝑟 ) =
𝑟 2 𝑑𝑟 𝑑𝑟 𝑟 𝑑𝑟 2
Using the value of equation (48) and (53) into equation (47), we have

1 𝑑2 𝜇 𝜇 (54)
2
= 𝜅2
𝑟 𝑑𝑟 𝑟
or

𝑑2 𝜇 (55)
= 𝜅2𝜇
𝑑𝑟 2
The general solution of such an equation may be written as

𝜇 = 𝐴𝑒 −𝜅𝑟 + 𝐵𝑒 +𝜅𝑟 (56)

Where A and B are two unknown constants. Now using the value of μ from equation (56) into equation (48),
we get

𝑒 −𝜅𝑟 𝑒 +𝜅𝑟 (57)


𝜓𝑟 = 𝐴 +𝐵
𝑟 𝑟
Since the potential at r = ∞ must vanish, this boundary condition is satisfied only if B = 0. Therefore, the
acceptable form of the equation (57) should be like this

𝑒 −𝜅𝑟 (58)
𝜓𝑟 = 𝐴
𝑟
To evaluate the value of constant, imagine a situation in which ions are so apart from each other that there are
no ion-ion interactions. Such a situation can be created by diluting the solution to a very large extent. In this
state, potential around the reference ion will simply be due to the reference ion itself i.e.

𝑍𝑖 𝑒0 (59)
𝜓𝑟 =
𝜀𝑟

Furthermore, at such large dilution, the bulk concentration will almost be zero (𝑛𝑖0 = 0) which in turn would
make κ = 0. Thus, the equation (58) in such a scenario will be

𝐴 (60)
𝜓𝑟 =
𝑟
Equating the results of equation (59) and (60), we have

Copyright © Mandeep Dalal


172 A Textbook of Physical Chemistry – Volume I

𝑍𝑖 𝑒0 𝐴 (61)
=
𝜀𝑟 𝑟
or

𝑍𝑖 𝑒0 𝑟 (62)
𝐴=
𝜀𝑟
After putting the value of A from equation (62) in equation (58), the final result for electrostatic potential is

𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 (63)
𝜓𝑟 =
𝜀 𝑟
The expression is the solution of the linearized Poisson-Boltzmann equation.

Figure 7. The variation of electrostatic potential as a function of the distance r from the reference ion.

Now, in order to evaluate the excess charge density as a function of distance from the reference ion, compare
equation (33) and equation (47) i.e.

4𝜋 (64)
𝜅 2 𝜓𝑟 = − 𝜌
𝜀 𝑟
or

𝜀𝜅 2 𝜓𝑟 (65)
𝜌𝑟 = −
4𝜋
Now putting the value of electrostatic potential from equation (63) in equation (65), we get the expression for
excess charge density as a function of r.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 173

𝜀𝜅 2 𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 (66)
𝜌𝑟 = − ×
4𝜋 𝜀 𝑟
𝑍𝑖 𝑒0 𝜅 2 −𝜅𝑟 (67)
𝜌𝑟 = − 𝑒
4𝜋𝑟
Now because the magnitude of ρr is a consequence of the unequal distribution of anions and cations, the above
also defines the ionic population distribution around the reference ion.

Figure 8. The variation of excess charge density as a function of the distance r from the reference ion.

Hence, the magnitude of excess charge density declines exponentially as the distance from the reference
increases. It is also worthy to note that the sign of excess charge around the reference ion is always opposite
to the reference ion. A negatively charged reference ion has a positively charged surrounding atmosphere and
vice-versa.

 Debye-Huckel Reciprocal Length


The concept of Debye-Huckel length can be understood in a better way only after knowing how much
charge actually surrounds the reference ion. For this, assume that the total space around the reference ion is
divided into the infinite number of hollow spherical shells of thickness dr with the inner and outer radii as r
and r+dr, respectively. The volume of each such shell will be 4πr2dr, and the total excess charge (dq) in the
same can be obtained by multiplying this volume element with the excess charge density, i.e.,

𝑑𝑞 = 𝜌𝑟 4𝜋𝑟 2 𝑑𝑟 (68)

Now, as we know that the excess charge density at distance r from the reference ion is

Copyright © Mandeep Dalal


174 A Textbook of Physical Chemistry – Volume I

𝑍𝑖 𝑒0 𝜅 2 −𝜅𝑟 (69)
𝜌𝑟 = − 𝑒
4𝜋𝑟
Putting the value of equation (69) in equation (68), we get

𝑍𝑖 𝑒0 𝜅 2 −𝜅𝑟 (70)
𝑑𝑞 = (− 𝑒 ) 4𝜋𝑟 2 𝑑𝑟
4𝜋𝑟

or

𝑑𝑞 = −𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 𝜅 2 𝑟𝑑𝑟 (71)

Since the exponential part becomes zero only at r = ∞, the total charge around the reference ion can be obtained
by integrating the equation (71) from r = 0 to r = ∞.
𝑟=∞ 𝑟=∞
(72)
𝑞𝑐𝑙𝑜𝑢𝑑 = ∫ 𝑑𝑞 = ∫ −𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 𝜅 2 𝑟𝑑𝑟
𝑟=0 𝑟=0

or

𝑞𝑐𝑙𝑜𝑢𝑑 = −𝑍𝑖 𝑒0 (73)

This is an important result which proves that a reference ion, with the charge +𝑍𝑖 𝑒0 , is surrounded by an
exactly equal but oppositely charged cloud i.e. −𝑍𝑖 𝑒0 .
Now, since the exponential part decreases with increasing r while the non-exponential part shows a
continuous increase as we move away from the reference ion, there should be a distance of maximum charge.

Figure 9. The depiction of excess charge density in a volume element of thickness ‘dr’ as a function of the
distance r from the reference ion.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 175

In order to find the distance of maximum “excess charge”, we need to differentiate the equation (71) with
respect to r, and then derivative needs to be put equal to zero i.e.

𝑑𝑞 (74)
= −𝑍𝑖 𝑒0 𝜅 2 (𝑒 −𝜅𝑟 − 𝜅𝑟𝑒 −𝜅𝑟 ) = 0
𝑑𝑟
Since −𝑍𝑖 𝑒0 𝜅 2 is non zero for sure, the above equation holds true only if

𝑒 −𝜅𝑟 − 𝜅𝑟𝑒 −𝜅𝑟 = 0 (75)

Which implies

𝑟 = 𝜅 −1 (76)

Also as we know that


1/2 (77)
4𝜋
𝜅=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 )
𝜀𝑘𝑇
𝑖

Hence, the radius of maximum excess charge from the center of the reference ion can be obtained by putting
the value of κ from equation (77) in equation (76), we get rmax or the “Debye-Huckel length” as

𝜀𝑘𝑇 1
1/2 (78)
𝑟𝑚𝑎𝑥 =( 0 )
4𝜋 ∑𝑖 𝑛𝑖 𝑍𝑖2 𝑒02

It is also worthy to note that the cloud will tend to fade out ever more with the decreasing concentration.

Figure 10. The variation of total charge enclosed in the dr thickness of the spherical shell as a function of
the distance r from the reference ion.

Copyright © Mandeep Dalal


176 A Textbook of Physical Chemistry – Volume I

 Ionic Cloud and Its Contribution to the Total Potential


According to the Debye-Huckel theory of ion-ion interaction, the molar chemical potential change
(𝛥𝜇𝑖−𝐼 ) of ion-ion interaction can be calculated from the following relation.

𝑁𝐴 (79)
𝛥𝜇𝑖−𝐼 = [𝑍 𝑒 × 𝜓]
2 𝑖 0
Where NA is the Avogadro number and ψ the electrostatic potential at distance r due to ionic cloud. The symbol
𝑍𝑖 𝑒0 is the charge on the ith species. The value of ψr is

𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 (80)
𝜓𝑟 =
𝜀 𝑟
Where ε is the dielectric constant of the surrounding medium and κ is a constant defined by
1/2 (81)
4𝜋
𝜅=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 )
𝜀𝑘𝑇
𝑖

The value of electrostatic potential given by equation (80) cannot be used in equation (79) because contains
the contribution form the ionic cloud as well as from the reference ion itself. Therefore, before we calculate
the partial molar free energy of ion-ion interaction, we need to separate potential from the ionic cloud and from
the central ion first.

Figure 11. The pictorial representation of the superposition of the potentials at distance r from the
reference ion due to the ionic cloud ion and the potential due to the central ion itself.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 177

The total electrostatic potential at a distance r can be fragmented as given below.

𝜓𝑟 = 𝜓𝑖𝑜𝑛 + 𝜓𝑐𝑙𝑜𝑢𝑑 (82)

or

𝜓𝑐𝑙𝑜𝑢𝑑 = 𝜓𝑟 − 𝜓𝑖𝑜𝑛 (83)

From the formulation of potential due to a single charge at a distance r, we know that

𝑍𝑖 𝑒0 (84)
𝜓𝑖𝑜𝑛 =
𝜀𝑟
Now using the value of ψr from equation (80) and ψion from equation (84) into equation (83), we get

𝑍𝑖 𝑒0 𝑒 −𝜅𝑟 𝑍𝑖 𝑒0 (85)
𝜓𝑐𝑙𝑜𝑢𝑑 = −
𝜀 𝑟 𝜀𝑟
or

𝑍𝑖 𝑒0 −𝜅𝑟 (86)
𝜓𝑐𝑙𝑜𝑢𝑑 = (𝑒 − 1)
𝜀𝑟

Now because the value of κ is proportional to ∑𝑖 𝑛𝑖0 𝑍𝑖2 𝑒02 , at very large dilution κ becomes very small and 𝑒 −𝜅𝑟
and can be expended as 1 − 𝜅𝑟 (ex = 1+ x). Therefore, equation (86) takes the form

𝑍𝑖 𝑒0 (87)
𝜓𝑐𝑙𝑜𝑢𝑑 = (1 − 𝜅𝑟 − 1)
𝜀𝑟
or

𝑍𝑖 𝑒0 𝜅 (88)
𝜓𝑐𝑙𝑜𝑢𝑑 = −
𝜀
Hence, potential due to ionic cloud is independent of the distance r in this case. Hence, potential due to ionic
cloud is independent of the distance r in this case. Now putting the value of ψcloud from equation (88) into
equation (79), we get

𝑁𝐴 𝑍𝑖 𝑒0 𝜅 (89)
𝛥𝜇𝑖−𝐼 = 𝑁𝐴 𝑤 = [𝑍𝑖 𝑒0 × (− )]
2 𝜀
or

𝑁𝐴 𝑍𝑖2 𝑒02 𝜅 (90)


𝛥𝜇𝑖−𝐼 = −
2𝜀
Hence, Debye-Hückel model gives the theoretical value of 𝛥𝜇𝑖−𝐼 arising from ion-ion interactions.

Copyright © Mandeep Dalal


178 A Textbook of Physical Chemistry – Volume I

 Debye-Huckel Limiting Law of Activity Coefficients and Its Limitations


So far in this chapter, the idea of ions in electrolytic solutions and their mutual interactions was taken
for granted. However, the conceptual understanding of these electrolytic solutions thrived via a different route.
Initially, the scientific community thought that the electrolytic solutions are the same as the non-electrolytic
ones. In other words, they did not pay any special attention to the charged species, and therefore, treated them
just like the non-electrolytes.
 Derivation of Debye-Huckel Limiting Law
Classically, the partial molar free energy (μi) of the ith species in a non-electrolyte is given by the
following relation.

𝜇𝑖 = 𝜇𝑖0 + 𝑅𝑇 ln 𝑥𝑖 (91)

Where 𝜇𝑖0 is the partial molar free energy when the concentration (𝑥𝑖 ) is unity (standard state). Mathematically,
we can say that is if 𝑥𝑖 = 1, 𝜇𝑖 = 𝜇𝑖0 .
At this point, it worthy to point out that the Coulombic interactions (long-range) can be neglected in
the case of non-electrolytic solutions, which is obviously due to the uncharged or neutral nature of the solute
particles. However, some short-range electrostatic interactions like the London dispersion forces or the dipole-
dipole may play a significant role if the average inter-particle distance is small enough. This means that if the
dilution is large enough, the ion-ion interactions (long-range Coulombic effects) can simply be ignored in a
non-electrolytic solution. Thus, the equation (91) is valid only for those solutions in which no long-range
electrostatic interactions occur, and such solutions are called as ideal solutions. However, when the equation
(91) was applied to the electrolytic solutions, it was found that

𝜇𝑖 − 𝜇𝑖0 ≠ 𝑅𝑇 ln 𝑥𝑖 (92)

Such solutions which show deviation from equation (91) are called as non-ideal solutions. This means that
unlike non-electrolytes, the ion-ion interactions must be considered before any theoretical treatment for the
chemical potential of ionic solutions is carried out.
Therefore, to rationalize this deviation, the scientific community developed a unique approach even
before the Debye-Huckel’s ionic-cloud theory. They tried to quantify the deviation from “idealistic behavior”
by incorporating an empirical parameter fi in the equation (92) as

𝜇𝑖 − 𝜇𝑖0 = 𝑅𝑇 ln 𝑥𝑖 𝑓𝑖 (93)

Where 𝜇𝑖0 is the partial molar free energy when the concentration (𝑥𝑖 ) and correction factor (𝑓𝑖 ) both are unity.
Mathematically, we can say that if 𝑥𝑖 = 1 and 𝑓𝑖 = 1, 𝜇𝑖 = 𝜇𝑖0 ; which is obviously a hypothetical situation in
which standard state behaves ideally. Moreover, it should also be noted that term xi represents the actual
concentration of the ions of ith type which may or may not be equal to expected concentration because weak
electrolytes do not dissociate completely like the strong one. After that, it was assumed that it is not the actual

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 179

concentration (𝑥𝑖 ) but the effective concentration (𝑥𝑖 𝑓𝑖 ) that dictates the chemical potential change. For
simplicity, the effective concentration if ith species was simply labeled as the activity (𝑎𝑖 ). Mathematically,
we can say that

𝑎𝑖 = 𝑥𝑖 𝑓𝑖 (94)

The correction factor 𝑓𝑖 is also called as the “activity coefficient” and has a value of unity for ideal solutions
i.e. when 𝑓𝑖 = 1, 𝑎𝑖 = 𝑥𝑖 .
Consequently, the partial molar free energy change in going from the ideal state to the real state in
ionic solutions can be written as

𝜇𝑖 − 𝜇𝑖0 = 𝑅𝑇 ln 𝑥𝑖 + 𝑅𝑇 ln 𝑓𝑖 (95)

The above equation is an empirical formulation of the behavior of ionic solutions and cannot give any
theoretical result for 𝑓𝑖 . Therefore, in order to find out the physical significance of the activity coefficient, we
need to assume an ionic solution that can be switched from ideal (no ion-ion interaction) to the real situation
(with ion-ion interaction). For ideal solutions, we have

𝜇𝑖 (ideal) = 𝜇𝑖0 + 𝑅𝑇 ln 𝑥𝑖 (96)

For real solutions,

𝜇𝑖 (real) = 𝜇𝑖0 + 𝑅𝑇 ln 𝑥𝑖 + 𝑅𝑇 ln 𝑓𝑖 (97)

The chemical potential change from ion-ion interaction can be obtained by subtracting equation (96) from
equation (97) i.e.

𝜇𝑖 (real) − 𝜇𝑖 (ideal) = 𝛥𝜇𝑖−𝐼 = 𝜇𝑖0 + 𝑅𝑇 ln 𝑥𝑖 + 𝑅𝑇 ln 𝑓𝑖 − 𝜇𝑖0 − 𝑅𝑇 ln 𝑥𝑖 (98)

𝛥𝜇𝑖−𝐼 = 𝑅𝑇 ln 𝑓𝑖 (99)

Hence, the activity coefficient is a measure of chemical potential change due to the interaction of ith types with
the rest of the ionic species. Now, according to the Debye-Huckel theory of ion-ion interaction, the partial
molar free energy change due to ion-ion interaction is

𝑁𝐴 𝑍𝑖2 𝑒02 𝜅 (100)


𝛥𝜇𝑖−𝐼 =−
2𝜀
Where Zi is the charge number on the ith type of ion while e0 is the electronic charge. NA is the Avogadro
number and ε is the dielectric constant of the surrounding medium, i.e., solvent. The symbol κ with 𝑛𝑖0 as the
bulk concentration is

4𝜋
1/2 (101)
𝜅=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 )
𝜀𝑘𝑇
𝑖

Copyright © Mandeep Dalal


180 A Textbook of Physical Chemistry – Volume I

Now, from equation (99) and equation (100), we have

𝑁𝐴 𝑍𝑖2 𝑒02 𝜅 (102)


𝑅𝑇 ln 𝑓𝑖 = −
2𝜀
The above result implies that the Debye-Huckel’s ionic cloud theory enables us to determine the activity
coefficient in a theoretical framework.
Furthermore, if the actual concentration of the ith type of ion is represented in terms of molarity (𝑐𝑖 )
or the molality (𝑚𝑖 ), the equation (96) takes the form

𝜇𝑖 = 𝜇𝑖0 (𝑐) + 𝑅𝑇 ln 𝑐𝑖 (103)

and

𝜇𝑖 = 𝜇𝑖0 (𝑚) + 𝑅𝑇 ln 𝑚𝑖 (104)

Similarly, for real solutions, equation (97) takes the forms

𝜇𝑖 = 𝜇𝑖0 (𝑐) + 𝑅𝑇 ln 𝑐𝑖 + 𝑅𝑇 ln 𝑓𝑐 (105)

and

𝜇𝑖 = 𝜇𝑖0 (𝑚) + 𝑅𝑇 ln 𝑚𝑖 + 𝑅𝑇 ln 𝑓𝑚 (106)

Where 𝜇𝑖0 (𝑐) and 𝜇𝑖0 (𝑚) are corresponding standard chemical potential at a molarity 𝑐𝑖 and molality 𝑚𝑖 ,
respectively.
At this point, some experimental limitations must be discussed before any comparative analysis of
the activity coefficient is carried out. It is a quite well-known fact that we cannot measure the hydration energy,
i.e., ion-solvent interaction of individual ionic species because the addition of only cations or the anions is not
possible practically. Even if it was possible, it would result in a negatively or positively charged solution
depending upon the nature of the ions added, which eventually, would cause undesired interactions. Owing to
similar arguments, it is also practically impossible to measure the activity coefficient 𝑓𝑖 which is also a function
chemical potential change arising from ion-ion interaction. The only way to avoid the situation is to add the
electroneutral electrolytes to the solvent which would eventually produce the positive and negative ions
simultaneously irrespective of whether it is strong or weak. Therefore, one can determine the activity of a net
electrolyte consisted of minimum two ionic species, and we need something that can connect activity
coefficient of electrolyte with the individual ionic species. In other words, the activity coefficient of a single
ionic species can be determined only via a theoretical route. All this resulted in the idea of “mean ionic activity
coefficient”. To illustrate mathematically, consider a NaCl-type univalent electrolyte MA. The chemical
potential of cations (M+) and anions (A−) can be written as
0
𝜇𝑀 + = 𝜇𝑀 + + 𝑅𝑇 ln 𝑥𝑀 + + 𝑅𝑇 ln 𝑓𝑀 + (107)

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 181

and

𝜇𝐴− = 𝜇𝐴0 − + 𝑅𝑇 ln 𝑥𝐴− + 𝑅𝑇 ln 𝑓𝐴− (108)

Combining the equation (107) with equation (108), we get


0
𝜇𝑀+ + 𝜇𝐴− = (𝜇𝑀 0
+ + 𝜇𝐴− ) + 𝑅𝑇 ln (𝑥𝑀 + 𝑥𝐴− ) + 𝑅𝑇 ln (𝑓𝑀 + 𝑓𝐴− ) (109)

The above equation gives the free energy of the system due to two moles of ions i.e. one mole of M+ and mole
of A−; or due to one mole of electroneutral electrolyte. Since we are interested in the average input to the total
free energy due to one mole of ions only, therefore, we must divide the equation (109) by 2, i.e.,
0
𝜇𝑀+ + 𝜇𝐴− 𝜇𝑀 0
+ + 𝜇𝐴− (110)
= + 𝑅𝑇 ln (𝑥𝑀+ 𝑥𝐴− )1/2 + 𝑅𝑇 ln (𝑓𝑀+ 𝑓𝐴− )1/2
2 2
If we consider
𝜇𝑀+ + 𝜇𝐴− (111)
𝜇± =
2

0
0
𝜇𝑀 0
+ + 𝜇𝐴− (112)
𝜇± =
2
and

𝑥± = (𝑥𝑀+ 𝑥𝐴− )1/2 (113)

𝑓± = (𝑓𝑀+ 𝑓𝐴− )1/2 (114)

Where 𝜇± and 𝜇±
0
are the mean chemical potential and standard mean chemical potential, respectively. The
symbol 𝑥± and 𝑓± represent the mean ionic mole fraction and mean ionic activity coefficient, respectively.
Here, it is also worthy to note that 𝜇± and 𝜇±
0
are simply the arithmetic means whereas 𝑥± and 𝑓± are the
geometric mean quantities. Now using values from equations (111−114) into equation (110), we get
0
𝜇± = 𝜇± + 𝑅𝑇 ln 𝑥± + 𝑅𝑇 ln 𝑓± (115)

Since it is for one mole instead two, we can write

1 0 (116)
𝜇 = 𝜇± = 𝜇± + 𝑅𝑇 ln 𝑥± + 𝑅𝑇 ln 𝑓±
2 𝑀𝐴
Hence, the experimental value of 𝑓± can be obtained just by knowing the free energy of one mole of electrolytic
solution at a particular concentration. Once the value of 𝑓± is known, the product of individual activity
coefficients can be obtained using equation (114). The individual values of activity coefficients obtained from
equation (102) can be put into equation (114) to compare with experimentally observed value so that the
Debye-Huckel model can be tested.

Copyright © Mandeep Dalal


182 A Textbook of Physical Chemistry – Volume I

Furthermore, if one mole of electroneutral electrolyte generates ν+ and ν− moles of cations and anions, then
equation (107) and equation (108) will become
0
𝜈+ 𝜇+ = 𝜈+ 𝜇+ + 𝜈+ 𝑅𝑇 ln 𝑥+ + 𝜈+ 𝑅𝑇 ln 𝑓+ (117)

and
0
𝜈− 𝜇− = 𝜈− 𝜇− + 𝜈− 𝑅𝑇 ln 𝑥− + 𝜈− 𝑅𝑇 ln 𝑓− (118)

To find the free energy change due to per mole of cation and anion, add equation (117) to equation (118) and
then divide by 𝜈 = 𝜈+ + 𝜈− .

𝜈+ 𝜇+ + 𝜈− 𝜇− 0
𝜈+ 𝜇+ 0
+ 𝜈− 𝜇− 𝜈 1/𝜈 𝜈 1/𝜈 (119)
= + 𝑅𝑇 ln (𝑥++ 𝑥−𝜈− ) + 𝑅𝑇 ln (𝑓+ + 𝑓−𝜈− )
𝜈 𝜈
If we consider
𝜈+ 𝜇+ + 𝜈− 𝜇− (120)
𝜇± =
𝜈

0
0
𝜈+ 𝜇+ 0
+ 𝜈− 𝜇− (121)
𝜇± =
𝜈
and

𝜈
𝑥± = (𝑥++ 𝑥−𝜈− )
1/𝜈 (122)

𝜈 1/𝜈
𝑓± = (𝑓+ + 𝑓−𝜈− ) (123)

Now using values from equations (120−123) into equation (119), we get
0
𝜇± = 𝜇± + 𝑅𝑇 ln 𝑥± + 𝑅𝑇 ln 𝑓± (124)

Taking logarithm both side of equation (123), we have

1 (125)
ln 𝑓± = (𝜈+ ln 𝑓+ + 𝜈− ln 𝑓− )
𝜈
Now putting the value of ln 𝑓+ and ln 𝑓− using equation (102), we get

1 𝑁𝐴 𝑒02 𝜅 (126)
ln 𝑓± = − [ (𝜈 𝑍 2 + 𝜈− 𝑍−2 )]
𝜈 2𝜀𝑅𝑇 + +

Now owing to the electroneutrality of the solution, 𝜈+ 𝑍+ must be equal to 𝜈− 𝑍− , i.e.,

𝜈+ 𝑍+2 + 𝜈− 𝑍−2 = 𝜈+ 𝑍+ 𝑍− + 𝜈− 𝑍− 𝑍+ (127)

𝜈+ 𝑍+2 + 𝜈− 𝑍−2 = 𝑍+ 𝑍− (𝜈+ + 𝜈− ) (128)

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 183

𝜈+ 𝑍+2 + 𝜈− 𝑍−2 = 𝑍+ 𝑍− 𝜈 (129)

After putting the value of equation (129) into equation (126), we have

𝑁𝐴 𝑒02 𝜅 (130)
ln 𝑓± = − 𝑍 𝑍
2𝜀𝑅𝑇 + −
Furthermore, using the value of κ using equation (101), the above equation takes the form

𝑁𝐴 𝑒02 𝑍+ 𝑍− 4𝜋
1/2 (131)
ln 𝑓± = − ( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 )
2𝜀𝑅𝑇 𝜀𝑘𝑇
𝑖

Since 𝑛𝑖0 = 𝑐𝑖 𝑁𝐴 /1000, the equation (131) becomes


1/2 (132)
𝑁𝐴 𝑒02 𝑍+ 𝑍− 4𝜋 𝑐𝑖 𝑁𝐴 𝑍𝑖2 𝑒02
ln 𝑓± = − ( ∑ )
2𝜀𝑅𝑇 𝜀𝑘𝑇 1000
𝑖

Multiply and divide the equation (132) by 2, and then put 𝑐𝑖 𝑍𝑖2 /2 = 𝐼, i.e., ionic strength, we get

𝑁𝐴 𝑒02 𝑍+ 𝑍− 8𝜋𝑁𝐴 𝑒02


1/2 (133)
ln 𝑓± = − ( ) √𝐼
2𝜀𝑅𝑇 1000𝜀𝑘𝑇

𝑁𝐴 𝑒02 𝑍+ 𝑍− (134)
ln 𝑓± = − 𝐵√𝐼
2𝜀𝑅𝑇
Where the constant B is defined as

8𝜋𝑁𝐴 𝑒02
1/2 (135)
𝐵=( )
1000𝜀𝑘𝑇

Converting the natural logarithm to the common logarithm, the equation (134) becomes

1 𝑁𝐴 𝑒02 𝑍+ 𝑍− (136)
log 𝑓± = − 𝐵√𝐼
2.303 2𝜀𝑅𝑇
After a new constant A as

1 𝑁𝐴 𝑒02 (137)
𝐴= 𝐵
2.303 2𝜀𝑅𝑇
The equation (136) can be further simplified as given below.

log 𝑓± = −𝐴(𝑍+ 𝑍− )√𝐼 (138)

Which is the Debye-Huckel limiting law of activity coefficients.

Copyright © Mandeep Dalal


184 A Textbook of Physical Chemistry – Volume I

 Limitations of Debye-Huckel Limiting Law


The negative sign in the Debye-Huckel limiting law implies the fact that the activity coefficient is
always less than unity. In order to discuss the limitations of this law, recall the popular relationship, i.e.,

log 𝑓± = −𝐴(𝑍+ 𝑍− )√𝐼 (139)

Since A is constant and the product 𝑍+ 𝑍− is also constant for a particular electrolyte, the logarithm of activity
coefficient must decrease linearly with the square root of ionic strength. In other words, the slope log 𝑓± and
√𝐼 can be determined simply by knowing the valences of the ions involved and by the knowledge of some
physical constants. Furthermore, it should also be noted that the slope is independent of the very nature of the
electrolyte and is a function of valences of cations and anions only. For instance, since the value of A for the
water as solvent is 0.509, the slope of for NaCl as well as for KCl must be equal to 1×0.509 only. For uni-
bivalent and the bi-bivalent electrolyte is should be 2×0.509 and 4×0.509. The logarithmic variation of mean
ionic activity coefficient (𝑓± ) with the square root of the ionic strength (√𝐼) for electrolytes of different
valences follows the equation (139) pretty strictly.

Figure 12. The logarithmic variation of 𝑓± with the square root of the ionic strength (√𝐼) for electrolytes of
different valences in very dilute (left) and in solutions up to large concentration (right).

It is obvious from both the graphs that log 𝑓± becomes zero when the dilution is very large indicating
that the activity coefficient is unity, which is according to the Debye-Huckel limiting law. However, it should
also be remembered that any theory is always a simplification of the real problem, and therefore, some
deviations are expected. The same has been observed when the ionic strength is increased. It can be clearly
seen that the deviation of the experimental result increases with the rise in the square root of the ionic strength.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 185

 Ion-Size Effect on Potential


In the Debye-Huckel theory of ion-ion interaction, the central ion was simply treated as a point charge
instead of its actual size. The scientific community realized that it might be one of the reasons behind the
deviations observed from Debye-Huckel limiting law of activity coefficients. In order to understand the effect
of ion size on the potential, recall the expression for the Debye-Huckel length (rmax) or κ−1, i.e.,

𝜀𝑘𝑇 1
1/2 (140)
𝑟𝑚𝑎𝑥 =( 0 )
4𝜋 ∑𝑖 𝑛𝑖 𝑍𝑖2 𝑒02

Where 𝑛𝑖0 is the bulk concentration of the ith species and k is simply the Boltzmann constant. The symbol ε
represents the dielectric constant of the surrounding medium. The symbol Zi shows the charge number of the
ion whereas e0 represents the electronic charge. It is obvious from the equation (140) that the mean thickness
of the ionic cloud (κ−1) is actually inversely proportional to the concentration. This means that at higher
concentrations, the mean thickness of the ionic cloud will be small and cannot outrank the size of the central
ion anymore. In other words, at higher concentration, the size of the reference ion cannot be neglected at all,
and therefore, the point charge approximation is no longer valid. For instance, At a concentration of 0.1N, the
mean thickness of the ionic cloud is only ten times the radius of the reference ion.
Consider 4πr2dr as the volume element of a hollow spherical shell of thickness dr with the inner and
outer radius as r and r+dr, respectively, with the size of the ion represented by the parameter a.

Figure 13. The depiction of excess charge density as a function of the distance r from the reference ion of
finite size a.

Therefore, the ion-size must be considered for a more realistic picture of the actual situation. To do
so, consider the linearized Poisson-Boltzmann equation which is free from both approximations, i.e., ion-size
matters and the point charge consideration.

Copyright © Mandeep Dalal


186 A Textbook of Physical Chemistry – Volume I

1 𝑑 2 𝑑𝜓𝑟 (141)
(𝑟 ) = 𝜅 2 𝜓𝑟
𝑟 2 𝑑𝑟 𝑑𝑟
Where ψr is the potential at a distance r from the reference ion. The symbol κ is defined as

4𝜋
1/2 (142)
𝜅=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 )
𝜀𝑘𝑇
𝑖

The general solution of the above equation is

𝑒 −𝜅𝑟 𝑒 +𝜅𝑟 (143)


𝜓𝑟 = 𝐴 +𝐵
𝑟 𝑟
Since the potential at r = ∞ must vanish, this boundary condition is satisfied only if B = 0. Therefore, the
acceptable form of the equation (143) should be like this

𝑒 −𝜅𝑟 (144)
𝜓𝑟 = 𝐴
𝑟
To evaluate the value of constant A in this scenario, a slightly different route is followed. As we know that the
total charge in the dr thickness of a spherical shell at distance r is

𝑑𝑞 = 𝜌𝑟 4𝜋𝑟 2 𝑑𝑟 (145)

Now, the excess charge density at distance r from the reference ion will be

𝜀 1 𝑑 2 𝑑𝜓𝑟 𝜀𝜅 2 𝜓𝑟 (146)
𝜌𝑟 = − [ 2 (𝑟 )] = −
4𝜋 𝑟 𝑑𝑟 𝑑𝑟 4𝜋
Using the value of ψr from equation (144) into equation (146), we get

𝜀𝜅 2 𝑒 −𝜅𝑟 (147)
𝜌𝑟 = − 𝐴
4𝜋 𝑟
After putting the value of equation (147) into equation (145), we have

𝜀𝜅 2 𝑒 −𝜅𝑟 (148)
𝑑𝑞 = − 𝐴 4𝜋𝑟 2 𝑑𝑟
4𝜋 𝑟
or

𝑑𝑞 = −𝜀𝜅 2 𝐴𝑒 −𝜅𝑟 𝑟𝑑𝑟 (149)

Since the exponential part becomes zero only at r = ∞, the total charge around the reference ion can be obtained
by integrating the equation (149) from an unknown parameter r = a to r = ∞.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 187

𝑟=∞ 𝑟=∞
(150)
2 −𝜅𝑟
𝑞𝑐𝑙𝑜𝑢𝑑 = ∫ 𝑑𝑞 𝑑𝑟 = ∫ −𝜀𝜅 𝐴𝑒 𝑟𝑑𝑟
𝑟=𝑎 𝑟=𝑎

or
𝑟=∞
(151)
𝑞𝑐𝑙𝑜𝑢𝑑 = ∫ 𝑑𝑞 𝑑𝑟 = −𝐴𝜀𝑒 −𝜅𝑎 (1 + 𝜅𝑎)
𝑟=𝑎

Also as we know that the total charge on the ionic cloud must be equal and opposite to the charge on the
reference ion i.e.

𝑞𝑐𝑙𝑜𝑢𝑑 = −𝑍𝑖 𝑒0 (152)

Equating the results of equation (151) and equation (152), we have

−𝐴𝜀𝑒 −𝜅𝑎 (1 + 𝜅𝑎) = −𝑍𝑖 𝑒0 (153)

or

𝑍𝑖 𝑒0 𝑒 𝜅𝑎 (154)
𝐴=
𝜀 (1 + 𝜅𝑎)

Using the value of A from equation (154) in equation (144), we get

𝑍𝑖 𝑒0 𝑒 𝜅𝑎 𝑒 −𝜅𝑟 (155)
𝜓𝑟 =
𝜀 (1 + 𝜅𝑎) 𝑟

 Ion-Size Parameter and the Theoretical Mean - Activity Coefficient in the


Case of Ionic Clouds with Finite-Sized Ions
As we know that the total electrostatic potential at a distance r can simply be fragmented as given
below.

𝜓𝑟 = 𝜓𝑖𝑜𝑛 + 𝜓𝑐𝑙𝑜𝑢𝑑 (156)

or

𝜓𝑐𝑙𝑜𝑢𝑑 = 𝜓𝑟 − 𝜓𝑖𝑜𝑛 (157)

The electrostatic potential at a distance r from an ion of finite size is

𝑍𝑖 𝑒0 𝑒 𝜅𝑎 𝑒 −𝜅𝑟 (158)
𝜓𝑟 =
𝜀 (1 + 𝜅𝑎) 𝑟

Copyright © Mandeep Dalal


188 A Textbook of Physical Chemistry – Volume I

Also, from the formulation of potential due to a single charge at a distance r, we know that

𝑍𝑖 𝑒0 (159)
𝜓𝑖𝑜𝑛 =
𝜀𝑟
After putting the values of 𝜓𝑟 and 𝜓𝑖𝑜𝑛 from equation (158) and equation (159) into equation (157), we have

𝑍𝑖 𝑒0 𝑒 𝜅𝑎 𝑒 −𝜅𝑟 𝑍𝑖 𝑒0 (160)
𝜓𝑐𝑙𝑜𝑢𝑑 = −
𝜀 (1 + 𝜅𝑎) 𝑟 𝜀𝑟

or

𝑍𝑖 𝑒0 𝑒 𝜅(𝑎−𝑟) (161)
𝜓𝑐𝑙𝑜𝑢𝑑 = [ − 1]
𝜀𝑟 (1 + 𝜅𝑎)

At this point, we must recall the relationship of mean activity coefficient and chemical potential change of the
ion-ion interaction, i.e.,

𝑅𝑇 ln 𝑓𝑖 = 𝛥𝜇𝑖−𝐼 (162)

also

𝑁𝐴 𝑍𝑖 𝑒0 (163)
𝛥𝜇𝑖−𝐼 = 𝜓
2
From equation (162) and (163), we have

𝑁𝐴 𝑍𝑖 𝑒0 (164)
𝑅𝑇 ln 𝑓𝑖 = 𝜓
2
𝑁𝐴 𝑍𝑖 𝑒0 (165)
ln 𝑓𝑖 = 𝜓
2𝑅𝑇
Moreover, as the potential at the surface (r = a) of the ion must be

𝜓 = 𝜓𝑐𝑙𝑜𝑢𝑑 (166)

Which implies that 𝜓𝑐𝑙𝑜𝑢𝑑 in equation (161) at r = a should become

𝑍𝑖 𝑒0 1 (167)
𝜓𝑐𝑙𝑜𝑢𝑑 = [ − 1]
𝜀𝑎 (1 + 𝜅𝑎)

or

𝑍𝑖 𝑒0 1 − 1 − 𝜅𝑎 (168)
𝜓𝑐𝑙𝑜𝑢𝑑 = [ ]
𝜀𝑎 (1 + 𝜅𝑎)

𝑍𝑖 𝑒0 1 (169)
𝜓𝑐𝑙𝑜𝑢𝑑 = − −
𝜀𝜅 (1 + 𝜅𝑎)

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 189

After using the value of equation (169) in equation (165), we get

𝑁𝐴 𝑍𝑖2 𝑒02 1 (170)


ln 𝑓𝑖 = − −1
2𝑅𝑇𝜀𝜅 (1 + 𝜅𝑎)

The mean ionic activity coefficient, in this case, will be

𝐴𝑍+ 𝑍− 1/2 (171)


log 𝑓± = − 𝐼
(1 + 𝜅𝑎)

Where I is the ionic strength and A is Debye-Huckel constant. Now since the thickness of the ionic cloud is
defined as

𝜅 = 𝐵𝐼1/2 (172)

After putting the value of κ from equation (172) into equation (171), we have

𝐴𝑍+ 𝑍− 𝐼1/2 (173)


log 𝑓± = −
1 + 𝑎 𝐵𝐼1/2
Which is the equation for the mean ionic activity coefficient when the central ion has a finite size. Now recall
the equation for mean ionic activity coefficient when the central ion was considered as point charge i.e.

log 𝑓± = −𝐴𝑍+ 𝑍− 𝐼1/2 (174)

Hence, the two equations differ only in respect of denominator 1 + 𝜅𝑎. The rearranged form of the equation
(171) is

𝐴𝑍+ 𝑍− (175)
log 𝑓± = − 𝐼1/2
(1 + 𝑎/𝜅 −1 )

At large dilution, 𝑎 ≪ 𝜅 −1 , and the denominator tends to approach unity. Therefore, at very large dilution, the
equation (175) becomes equal to equation (174), proving the correspondence principle.
The most obvious approach to estimate the ion size parameter is the sum of the crystallographic radii
of the cation and anions present in the electrolytic-solution. This is because the two ions cannot come closer
than the sum of their individual ionic radii. However, since ions are hydrated in aqueous solutions, one might
think the ion size parameter as the sum of the hydrated radii instead. Nevertheless, the hydrated radii would
face some compression during the course of the collision of ion. All this suggests that the magnitude of ‘a’
must be greater than the sum of the individual crystallographic radii and should be less than the sum of their
individual hydrated radii. Therefore, the “mean distance of closest approach” should be more appropriate for
this situation. The best way to estimate the ion-size parameter is to calibrated equation (175) to match the
experimental value of mean ionic activity coefficient. After knowing the ion size parameter at one
concentration, the value of mean ionic activity coefficient can easily be determined at other concentrations.

Copyright © Mandeep Dalal


190 A Textbook of Physical Chemistry – Volume I

 Debye-Huckel-Onsager Treatment for Aqueous Solutions and Its


Limitations
It is a well-known fact that the conductance of weak electrolytic solutions increases with the increase
in dilution. This can be easily explained on the basis of Arrhenius's theory of electrolytic dissociation which
says that the magnitude of dissociated electrolyte, and hence the number of charge carriers, increases with the
increase in dilution. However, the problem arises when the strong or true electrolytes show the same trend but
at a much lower scale. We used the word “problem” because even at the higher concentration, the electrolyte
dissociates completely inferring that there is no possibility of further dissociation with dilution. This means
that there should be no increase in the conductance of strong electrolytes with the addition of water.

Figure 14. The typical variation of molar conductance (Λm) with the square root of the concentration (√𝑐)
for strong and weak electrolytes.

The primary reason behind this weird behavior of strong electrolyte is that the conductance of any
electrolytic solution depends not only upon the number of charge carriers but also upon the speed of these
charge carriers. Therefore, if the dilution does not affect the number of charge carriers in strong electrolytes,
it must be affecting the speed of ions to change its conductance. The main factor that is responsible for
governing the ionic mobility is ion-ion interactions. Now since these ion-ion interactions are dependent upon
the interionic distances, they eventually vary with the population density of charge carriers. Higher population
density means smaller interionic distances and therefore stronger ion-ion interactions. On the other hand, the
lesser population density of ions would result in larger interionic separations and hence weaker ion-ion
interactions.
In the case of weak electrolytes, the degree of dissociation is very small at high concentrations
yielding a very low population density of charge carriers. This would result in almost zero ion-ion interactions

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 191

at high concentrations. Now although the degree of dissociation increases with dilution which in turn also
increases the total number of charge carriers, the population density remains almost unchanged since extra
water has been added for these extra ions. Thus, we can conclude that there are no ion-ion interactions in weak
electrolytes neither at high nor at the low concentration; and hence the rise in conductance with dilution almost
a function dissociation only.
In the case of strong electrolytes, the degree of dissociation is a hundred percent even at high
concentrations yielding a very high population density of charge carriers. This would result in very strong ion-
ion interactions at high concentrations, hindering the speed of various charge carriers. Now when more and
more solvent is added, the total number of charge carriers remains the same but the population density
decreases continuously creating large interionic separations. This would result in a decrease in ion-ion
interaction with increasing dilution, and therefore, the charge carriers would be freer to move in the solution.
Thus, we can conclude that though there is no rise in the number of charge carriers with dilution, the declining
magnitude of ion-ion interaction creates faster ions and larger conductance.
 Factor Affecting the Conductance of Strong Electrolytic Solutions
In 1923, Peter Debye and Erich Huckel proposed an extremely important idea to quantify the
conductance of strong electrolytes in terms of these interionic interactions. In this model, a reference ion is
thought to be suspended in solvent-continuum of dielectric constant ε and is surrounded by oppositely charged
ions. Besides the ion-ion interaction, the viscosity of the solvent also affects the overall speed of the moving
ion. The primary effects which are responsible for controlling the ionic mobility are discussed below.
1. Asymmetry effect or the relaxation effect: In the absence of applied electric field, the ionic atmosphere
of the reference ion remains spherically symmetrical. This means that the electrostatic force of attraction on
the reference ion from all the directions would be the same. However, when the electric field is applied, the
ion starts to move towards the oppositely charged electrode. This, in turn, would destroy the spherical
symmetry of the cloud, and more ions would be left behind creating a net backward pull to the reference ion.
This effect, therefore, would slow down the moving ions and the conductance would be decreased.

Figure 15. The asymmetric or relaxation effect in the conductance of strong electrolytes.

Copyright © Mandeep Dalal


192 A Textbook of Physical Chemistry – Volume I

Alternatively, this can also be visualized in terms of cloud-destruction and cloud-building around the reference
ion. In other words, during the movement of ion, the ionic cloud around the reference ion must rebuild itself
to keep things natural. Since this rebuilding is not instantaneous and takes some time called as relaxation time,
the old cloud exerts a backward pull on the reference ion opposing oppressing its speed. All this results in a
diminished magnitude of the conductance.
2. Electrophoretic effect: After the application of the external electric field, the reference ion and ionic could
move in opposite directions. During the course of this movement, the solvent associated with the surrounding
ions also moves in a direction opposite to the central ion. In other words, we can say that the reference ion has
to move against a solvent stream, which makes it somewhat slower than usual. This phenomenon is called as
the electrophoretic effect.

Figure 16. The asymmetric or relaxation effect in the conductance of strong electrolytes.

Since the electrophoretic effect reduces the speed of the ion, the conductance of the electrolytic solution is also
affected considerably.
3. Viscous Effect: In addition to the asymmetric and electrophoretic effects, another type of resistance also
exists which affects the conductance of electrolytic solutions, the “viscous effect”. This is simply the frictional
resistance created by the viscosity of the solvent used. For an ion of given charge and size, the ionic mobility
as well the conductance decrease with the increase in the magnitude of the viscosity of the solvent used. In
other words, less viscous solvents yield higher conductance and vice-versa.
 Mathematical Development of Debye-Huckel-Onsager theory of Strong Electrolytes
It is a well-known fact that the equivalent conductivity (Λ) of an electrolytic solution is correlated to
the ionic mobilities (u) ions involved as

𝛬 = 𝐹(𝑢+ + 𝑢− ) (176)

Where F is the Faraday constant. Now, recall the ionic mobilities of the cation and anion i.e.

0
𝑍+ 𝑒0 𝑒02 𝜔 0 (177)
𝑢+ = 𝑢+ −𝜅( + 𝑢 )
6𝜋𝜂 6𝜀𝑘𝑇 +

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 193

0
𝑍− 𝑒0 𝑒02 𝜔 0 (178)
𝑢− = 𝑢− −𝜅( + 𝑢 )
6𝜋𝜂 6𝜀𝑘𝑇 −

Where 𝑢+ 0
and 𝑢−
0
are the ionic mobilities of the cation and anions at infinite dilution, respectively. The symbol
ε represents the dielectric constant of the medium whereas η is the coefficient of viscosity. 𝑍+ and 𝑍− are
charge numbers of the cation and anion, respectively. The symbol e0 simply shows the electronic charge. The
symbol κ represents (𝑛𝑖0 is the bulk concentration)

1/2 1 (179)
4𝜋 4𝜋𝑍 2 𝑒02 𝑐 2 𝑁𝐴 1/2
𝜅=( ∑ 𝑛𝑖0 𝑍𝑖2 𝑒02 ) =( ) ( )
𝜀𝑘𝑇 𝜀𝑘𝑇 1000
𝑖

The quantity ω is defined as

𝑍+ 𝑍− 2𝑞 𝑍+ 𝑍− 𝜆+ + 𝜆− (180)
𝜔= 𝑤ℎ𝑒𝑟𝑒 𝑞=
1 + √𝑞 𝑍+ + 𝑍− 𝑍+ 𝜆+ + 𝑍− 𝜆−

After putting the values of 𝑢+ and 𝑢− from equation (177) and (178) in equation (176), we get

0
𝑍+ 𝑒0 𝑒02 𝜔 0 0
𝑍− 𝑒0 𝑒02 𝜔 0 (181)
𝛬 = 𝐹 [𝑢+ −𝜅( + 𝑢+ )] + 𝐹 [𝑢− −𝜅( + 𝑢 )]
6𝜋𝜂 6𝜀𝑘𝑇 6𝜋𝜂 6𝜀𝑘𝑇 −

In the case of symmetrical electrolytes, we can put 𝑍+ = 𝑍− = 𝑍, and therefore 𝑍+ + 𝑍− = 2𝑍. Thus, the
above equation for such cases takes the form

0 0)
𝐹𝑍𝜅𝑒0 𝑒02 𝜔𝑘 0 0
(182)
𝛬 = 𝐹(𝑢+ + 𝑢− −[ + 𝐹(𝑢+ + 𝑢− )]
3𝜋𝜂 6𝜀𝑘𝑇

Since 𝐹(𝑢+
0 0)
+ 𝑢− = 𝛬0 , the equation (182) becomes

0
𝐹𝑍𝜅𝑒0 𝑒02 𝜔𝑘 0 (183)
𝛬 =𝛬 −[ + 𝛬 ]
3𝜋𝜂 6𝜀𝑘𝑇

Now expending above equation further by putting the value of κ from equation (179), we get
1 1 (184)
0
𝐹𝑍𝑒0 8𝜋𝑍 2 𝑒02 𝑁𝐴 2 𝑒02 𝜔 8𝜋𝑍 2 𝑒02 𝑁𝐴 2 0
𝛬 =𝛬 −[ ( ) + ( ) 𝛬 ] √𝑐
3𝜋𝜂 1000𝜀𝑘𝑇 6𝜀𝑘𝑇 1000𝜀𝑘𝑇

Define two constant A and B as


1 1 (185)
𝐹𝑍𝑒0 8𝜋𝑍 2 𝑒02 𝑁𝐴 2 𝑒02 𝜔 8𝜋𝑍 2 𝑒02 𝑁𝐴 2
𝐴= ( ) 𝑎𝑛𝑑 𝐵 = ( )
3𝜋𝜂 1000𝜀𝑘𝑇 6𝜀𝑘𝑇 1000𝜀𝑘𝑇

Therefore, the equation (184) can be simplified as

Copyright © Mandeep Dalal


194 A Textbook of Physical Chemistry – Volume I

𝛬 = 𝛬0 − (𝐴 + 𝐵𝛬0 )√𝑐 (186)

or

𝛬 = 𝛬0 − 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡√𝑐 (187)

The equation (184) and equation (186) are the popular forms famous Debye-Huckel-Onsager equation for
electrolyte solutions. The constants A and B can easily be determined from the knowledge of temperature T,
valence type of the electrolyte z, the viscosity of the medium, the dielectric constant and other universal
constants like Avogadro number. From Equation (187), it is obviously a straight line equation which means
the for symmetrical electrolytic solutions, we can plot the conductance vs square root of the concentration for
which the slope will be negative. The intercept after extrapolation gives the value of conductance of such
solutions at infinite dilution.
 Limitations of Debye-Huckel-Onsager Equation
Since the plot of conductance vs square root of the concentration is linear with negative slope and
positive intercept, it seems quite straightforward to study the strong electrolytes. However, it has been observed
that the equation (187) is followed only up low and moderate concentrations.

Figure 17. The comparison of theoretical and experimental conductance as a function of concentration for
some symmetric electrolytes.

It can be clearly seen that the theory and experiment move apart as the concentration increases. This is simply
because some approximation used to derive are Debye-Huckel-Onsager equation are not valid.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 195

 Debye-Huckel-Onsager Theory for Non-Aqueous Solutions


Before we discuss the Debye-Huckel-Onsager theory for non-aqueous solutions, recall the same for
aqueous solutions i.e.

𝛬 = 𝛬0 − 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡√𝑐 (187)

or

𝛬 = 𝛬0 − (𝐴 + 𝐵𝛬0 )√𝑐 (188)

Where the two constants, A and B, are defined as


1 (189)
𝐹𝑍𝑒0 8𝜋𝑍 2 𝑒02 𝑁𝐴 2
𝐴= ( )
3𝜋𝜂 1000𝜀𝑘𝑇

and
1 (190)
𝑒02 𝜔 8𝜋𝑍 2 𝑒02 𝑁𝐴 2
𝐵= ( )
6𝜀𝑘𝑇 1000𝜀𝑘𝑇

Where 𝐹 is the Faraday constant and NA is the Avogadro number. The symbol ε represents the dielectric
constant of the medium whereas η is the coefficient of viscosity. 𝑍 is charge numbers of the cation and anion.
The symbol e0 simply shows the electronic charge. The quantity ω is defined as

𝑍+ 𝑍− 2𝑞
𝜔=
1 + √𝑞

Where q is defined as

𝑍+ 𝑍− 𝜆 + + 𝜆−
𝑞=
𝑍+ + 𝑍− 𝑍+ 𝜆+ + 𝑍− 𝜆−

it is obvious from the Debye-Huckel-Onsager equation that the plot of conductance vs square root of
the concentration will be a straight line with a negative slope and positive intercept. The intercept after
extrapolation gives the value of conductance of such solutions at infinite dilution.
Now, it has been observed that the Debye-Huckel-Onsager equation can also be applied to non-
aqueous solutions up to the fairly good agreement. For instance, consider the variation of equivalent
conductivity as a function of the square root of the concentration for different alkali sulfocyanates in methanol
as the solvent. The theoretical predictions show that the results of the Debye-Hückel-Onsager equation are in
good agreement with the experiment up to 0.002 mol dm−3.

Copyright © Mandeep Dalal


196 A Textbook of Physical Chemistry – Volume I

Figure 18. The variation of equivalent conductivity of alkali sulfocyanates vs c1/2 in CH3OH.

In going from water to nonaqueous solvent, a significant variation in the quantities like dielectric constant of
the medium, the distance of the closest approach, or viscosity is observed. Now since the Debye-Hückel-
Onsager equation does have these quantities, the slope and intercept of the 𝛬 vs c1/2 may also vary drastically.

 The Solvent Effect on the Mobility at Infinite Dilution


As we know that the asymmetry and electrophoretic effects are not active at infinite dilution because
of their dependence on the size of ionic-cloud, the mobility of ions in such cases can be formulated simply
from the Stokes law, i.e.,

0
𝑍𝑒0 (191)
𝑢𝑐𝑜𝑛 =
6𝜋𝜂𝑟

Where r represents the radius of the solvated ion and η is the coefficient of viscosity. 𝑍 is charge numbers of
the cation and anion. The symbol e0 simply shows the electronic charge. Now, if we imagine the same
electrolyte in different electrolytic solutions, we can say
0
𝑢𝑐𝑜𝑛 𝜂𝑟 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (192)

Assuming further that the r is also independent of solvent type, we have


0
𝑢𝑐𝑜𝑛 𝜂 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (193)

Thus, it is obvious from the equation (193) that ionic mobility is inversely proportional to the coefficient of
viscosity. More viscous solvents would result in a slow drift of ions and vice-versa. It has also been found that
the equation (193) is more valid for solvents other than water.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 197

Figure 19. The variation of equivalent conductivity vs c1/2 in different solvents.

Moreover, the radius of solvated ion may vary drastically in going from one solvent to another (in
some cases it gets even double), the result of equation (193) does not find a very large application domain.
This variation in the solvated radius is primarily due to the difference in the size of solvent molecules. For
instance, the size of the solvated ion is larger in ethanol than in methanol, which in turn, is larger than what is
observed in water. All this results in an opposite trend in the ionic mobilities, and therefore, in the equivalent
conductivities as well. Recalling the simple Walden’s rule i.e.

𝛬𝜂 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (194)

Where Λ is the equivalent conductivity of the electrolytic solution at concentration c. This simply means that
equivalent conductivity and the viscosity of the solvent are inversely proportional to each other. However, the
effect of the solvated radius must be considered for more accurate results. Therefore, we must more acceptable
form of Walden’s rule, i.e.,

𝑢0 𝜂𝑟 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (195)

The symbol r represents the radius of the ionic species considered in the solvent under examination.

Copyright © Mandeep Dalal


198 A Textbook of Physical Chemistry – Volume I

 Equivalent Conductivity (Λ) vs Concentration C1/2 as a Function of the


Solvent
In order to understand the variation of equivalent conductivity with the square root of concentration
for different solvents, recall the generalized Walden’s rule, i.e.,

𝑢0 𝜂𝑟 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (196)

The 𝑢0 is the ionic mobility at infinite dilution and symbol r represents the radius of the ionic species
considered in the solvent under examination. The symbol η represents the coefficient of viscosity of the solvent
used. Also, as we know that the equivalent conductivity at infinite dilution can be obtained from the relation
given below.

𝛬0 = 𝐹𝑢0 (197)

Now, if the ionic mobility obtained using equation (196) is used in equation (197) for different nonaqueous
solutions, it has been found that predicted values of equivalent conductivity are quite large and sometimes
even outnumber the equivalent conductivity in water as the solvent. All this suggests that there should be no
problem in using non-aqueous solutions in electrochemical systems. However, the quantity that is more
important for practical applications is the specific conductivity (σ) at a finite concentration rather than the
equivalent conductivity at infinite dilution. In other words, the conjugative relationship of electrode geometry
and specific conductivity dictates the overall electrolytic resistance of the system under consideration. At this
point, one might ask why the electrolytic resistance is important. The answer lies in the fact that the magnitude
of useful power wasted as heat in the electrolytic solution is I2R (I is the current passed); and therefore, it is
the electrolytic resistance that R must possess a lower value for feasibility. The lower value of R implies a
higher value of specific conductivity, and the relation of specific conductivity with equivalent conductivity at
a certain concentration ‘c’ is

𝜎 = 𝛬𝑍𝑐 (198)

Now since the equivalent conductance Λ also depends on concentration, we cannot use equivalent conductance
at infinite dilution in equation (198) to find out the specific conductivity. First of all, one must determine the
equivalent conductivity at concentration ‘c’, which is possible from the Debye-Huckel-Onsager equation if the
values of constants A and B are known. Hence, to proceed further, we must recall the Debye-Huckel-Onsager
equation for the non-aqueous solution first, i.e.,

𝛬 = 𝛬0 − (𝐴 + 𝐵𝛬0 )√𝑐 (199)

Where the two constants, A and B, are defined as


1 (200)
𝐹𝑍𝑒0 8𝜋𝑍 2 𝑒02 𝑁𝐴 2
𝐴= ( )
3𝜋𝜂 1000𝜀𝑘𝑇

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 199

1 (201)
𝑒02 𝜔 8𝜋𝑍 2 𝑒02 𝑁𝐴 2
𝐵= ( )
6𝜀𝑘𝑇 1000𝜀𝑘𝑇

Where 𝐹 is the Faraday constant and NA is the Avogadro number. The symbol ε represents the dielectric
constant of the medium whereas η is the coefficient of viscosity. 𝑍 is charge numbers of the cation and anion.
The symbol e0 simply shows the electronic charge. ω is a parameter defined earlier in this chapter.

Figure 20. The variation of equivalent conductivity vs c1/2 in any arbitrary solvent.

It can be clearly seen that both the constants in the slope have the dielectric constant ε in the
denominator, and therefore, the equivalent conductivity as well as specific conductivity show decrease with
the increasing ε. In other words, we can say that the lower dielectric constant would result in stronger ion-ion
interaction and vice-versa. Therefore, in the case of non-aqueous solvents with very low dielectric constants,
the relative variation in the magnitude of equivalent conductivity with the square root of the concentration is
very large. Consequently, all this result in a very low specific conductivity at any practical electrolytic
concentration in non-aqueous solution than in water. Furthermore, in addition to the larger electrolytic
resistance, non-aqueous solutions are hard to keep so because they can always absorb some moisture from the
atmosphere.

Copyright © Mandeep Dalal


200 A Textbook of Physical Chemistry – Volume I

 Effect of Ion Association Upon Conductivity (Debye-Huckel-Bjerrum


Equation)
According to the Debye-Huckel theory of ion-ion interaction, the ions produced by the dissociation
of electrolytes are randomly distributed in the solution with oppositely charged ionic-cloud surrounding them.
Nevertheless, the possibility that some negative ions might get very close to the positively charged reference
ion was neglected. In such a situation, the translational thermal energy would not be enough to make the ions
to move independently, and an ionic-pair may be formed. Being a combined entity of two equal and opposite
charges, these ion pairs are completely neutral and distributed randomly in the electrolytic solutions. These
neutral ion-pairs are not influenced by the reference ion under consideration. Therefore, the fraction of ions
participating in these ion pairs must be obtained for a more accurate picture of the electrochemical properties
of the solutions.

Figure 21. The probability of finding an oppositely charged ion i.e. Z−e0 at distance r in dr thickness
around the reference ion.

In order to understand the concept, imagine a spherical shell at a distance r from the reference ion
with dr thickness. The probability of finding the anion (Pr) in the spherical shell is proportional to three main
factors. The first one is the ratio 4πr2dr volume element to the total volume V, the second one is the total
number of anions i.e. N− and the third is the e−U/kT (U represents the potential energy of anion situated at distance
r). Mathematically, we can formulate this as

4𝜋𝑟 2 𝑑𝑟 (202)
𝑃𝑟 = ( ) (𝑁− )(𝑒 −𝑈/𝑘𝑇 )
𝑉

Now because 𝑁− /𝑉 represents the anions’ concentration 𝑛−


0
, the above equation takes the form

𝑃𝑟 = 4𝜋𝑟 2 𝑑𝑟 𝑛−
0 −𝑈/𝑘𝑇
𝑒 (203)

Furthermore, after recalling the value of U, i.e.,

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 201

𝑍+ 𝑍− 𝑒02 (204)
𝑈=−
𝜀𝑟
Equation (203) becomes

𝑃𝑟 = 4𝜋𝑟 2 𝑛−
2
0 𝑍+ 𝑍− 𝑒0 /𝜀𝑟𝑘𝑇
𝑒 𝑑𝑟 (205)

Where ε is the dielectric constant of the medium and e0 is the electronic charge. Now define a new parameter
λ as

𝑍+ 𝑍− 𝑒02 (206)
𝜆=
𝜀𝑘𝑇
Therefore, we can write
0
𝑃𝑟 = (4𝜋 𝑛− )𝑒 𝜆/𝑟 𝑟 2 𝑑𝑟 (207)

Similarly, the probability of finding a positively charged ion at distance r in dr thickness around a negatively
charged reference ion will be
0
𝑃𝑟 = (4𝜋 𝑛+ )𝑒 𝜆/𝑟 𝑟 2 𝑑𝑟 (208)

Therefore, the probability of finding ith type of ion at distance r in dr thickness around a kth type of reference
ion can be formulated as

𝑃𝑟 = (4𝜋 𝑛𝑖0 )𝑒 𝜆/𝑟 𝑟 2 𝑑𝑟 (208)

where

𝑍𝑖 𝑍𝑘 𝑒02 (209)
𝜆=
𝜀𝑘𝑇
The variation of probability of finding one type ion around other types shows a strange behavior with distance.

Table 1. The total number of oppositely charged ions as a function of distance.

r (pm) Number of ions in shell × 1022

200 1.77

250 1.37

300 1.22

350 1.17

400 1.21

Copyright © Mandeep Dalal


202 A Textbook of Physical Chemistry – Volume I

When r varies from very small to moderate value, the magnitude of Pr decreases because of the factor
𝑒 𝜆/𝑟 . However, after a certain value, the probability starts to rise due to the dominance of the directly correlated
factor of r2. Hence, we can say that the probability of finding the oppositely charged ion at distance r from the
reference ion first decreases and then increases as we move away from the central ion.

Figure 22. The variation of the overall probability of finding an ion around ion under consideration as a
function of distance.

Since the probability of finding a negative ion within a certain distance around a positive ion is
obtained by applying the lower and upper bounds, we need to find these limits first. Now, because the ion-pair
can be formed only if the participating ions close enough so that the electrostatic attraction can outshine the
thermal forces; let this distance be represented by q.
In other words, we can say that the formation of ion-pair takes place only if the separation between
the cation and anion is less than q. Since the cation and anion cannot approach each other closer than the ionic-
size parameter (a), i.e., (the distance of closest approach), the ion-pair will be formed for an interionic
separation of greater than a but less than q. The probability of the ion-pair formation is the number of ith type
of ion participating in the ion-association to the total number of the same type of ions. Therefore, the probability
of ion-pair formation (θ) can be formulated as
𝑞 𝑞
(210)
𝜃 = ∫ 𝑃𝑟 𝑑𝑟 = ∫(4𝜋 𝑛𝑖0 )𝑒 𝜆/𝑟 𝑟 2 𝑑𝑟
𝑎 𝑎

Owing to the divergent nature of the integral, the upper limit can be set to the overall probability minima (q)
as shown below.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 203

Figure 23. The depiction of the overall probability of finding an oppositely charged ion around reference
ion from a to q.

To find the minima, differentiate equation (208) with respect to r and then put equal to zero, i.e.,

𝑑𝑃𝑟 𝜆 𝜆 𝜆 (211)
= 4𝜋 𝑛𝑖0 𝑒 𝑟 2𝑟 − 4𝜋 𝑛𝑖0 𝑟 2 𝑒 𝑟 2
𝑑𝑟 𝑟
or

2𝑟𝑚𝑖𝑛 − 𝜆 = 0 (212)

or

𝜆 (213)
𝑞 = 𝑟𝑚𝑖𝑛 =
2
After putting the value of λ from equation (206) into (213), we get

𝑍+ 𝑍− 𝑒02 (214)
𝑞=
2𝜀𝑘𝑇
Bjerrum claimed that only Coulombic interactions (short-range) lead to the formation of ion-pair; and if the
separation of cation and anion larger than q, they should be treated as free ions. In other words, we can say
that the formation of ion-pair is feasible if a < q and infeasible if a > q.

Copyright © Mandeep Dalal


204 A Textbook of Physical Chemistry – Volume I

Figure 24. The pictorial representation of ion-pair in the electrolytic solution.

Using the upper limit as q = λ/2, the probability of ion-pair formation (θ) becomes
𝜆/2 (215)
𝜃 = 4𝜋 𝑛𝑖0 ∫ 𝑒 𝜆/𝑟 2
𝑟 𝑑𝑟
𝑎

At this stage, after define a new variable for simplicity 𝑦 = 𝜆/𝑟 = 2𝑞/𝑟 and a new constant 𝑏 = 𝜆/𝑎, the
above equation takes the form

3 𝑏 (216)
𝑍+ 𝑍− 𝑒02
𝜃 = 4𝜋 𝑛𝑖0 ( ) ∫ 𝑒 𝑦 𝑦 −4 𝑑𝑦
𝜀𝑘𝑇
2

Hence, using the above equation we can easily determine the fraction of the ions reserved in pairing and cannot
move freely. Nevertheless, it would be more beneficial if we could use some other simple number to quantify
the magnitude of ion-pair formation instead of θ.
From the Arrhenius theory of electrolytic dissociation, we know that the A+B− type electrolyte would
give monovalent cations and anions in water i.e.

𝐴𝐵 ⇌ 𝐴+ + 𝐵− (217)

Using the law of mass action, the dissociation constant can be given as

𝑎𝐴+ 𝑎𝐵− (218)


𝐾=
𝑎𝐴𝐵

Similarly, if a cation M+ and anion A− form an ion-pair i.e.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 205

𝑀+ + 𝐴− ⇌ 𝐼𝑃 (219)

The law of mass action can also be used to give association constant as
𝑎𝐼𝑃 (220)
𝐾𝐴 = + −
𝑎𝑀 𝑎𝐴

Where 𝑎𝑀 +
, 𝑎𝐴− and 𝑎𝐼𝑃 are the activities of the cation, anion and ion-pair, respectively. Moreover, it can also
be seen that the association constant is just the reciprocal of dissociation constant. Now, θ is the fraction of
ions forming ion-pair, whereas θc and (1 − θ)c are the concentrations of ion-pair and free ions, respectively.
Therefore, the association constant in terms of activity coefficients can be written as

𝜃𝑐𝑓𝐼𝑃 (221)
𝐾𝐴 =
(1 − 𝜃)𝑐𝑓+ (1 − 𝜃)𝑐𝑓−

or

𝜃 1 𝑓𝐼𝑃 (222)
𝐾𝐴 = 2
(1 − 𝜃) 𝑐 𝑓+ 𝑓−

Where f+, f− and fIP are the activity coefficients for cation, anion and ion-pair, respectively. In terms of mean
ionic activity coefficient and analytical concentration ca, the equation (222) can also be written as

𝜃 1 𝑓𝐼𝑃 (223)
𝐾𝐴 =
(1 − 𝜃)2 𝑐𝑎 𝑓±2

Now because the ion pairs are neutral species, and therefore, do not participate in ion-ion interactions; the
activity coefficients ion-pairs can be taken as unity. Therefore, equation (223) can be written as

𝜃 1 1 (224)
𝐾𝐴 =
(1 − 𝜃) 𝑐𝑎 𝑓±2
2

or

𝜃 = 𝐾𝐴 (1 − 𝜃)2 𝑐𝑎 𝑓±2 (225)

At this stage, the only thing that is needed for further treatment is the relation between θ and the conductivity
of the electrolyte used. To do so, recall the correlation between the concentration of free ions and specific
conductivity (σ) i.e.

𝜎 = 𝑍𝐹(𝑢+ + 𝑢− )𝑐𝑓𝑟𝑒𝑒 𝑖𝑜𝑛𝑠 (226)

Where F is the Faraday constant.


𝑐𝑓𝑟𝑒𝑒 𝑖𝑜𝑛𝑠 (227)
𝜎 = 𝑍𝐹(𝑢+ + 𝑢− ) 𝑐𝑎
𝑐𝑎

Copyright © Mandeep Dalal


206 A Textbook of Physical Chemistry – Volume I

The term 𝑐𝑓𝑟𝑒𝑒 𝑖𝑜𝑛𝑠 /𝑐𝑎 is the fraction of ions that are free (not associated), and therefore
𝑐𝑓𝑟𝑒𝑒 𝑖𝑜𝑛𝑠 (228)
=1−𝜃
𝑐𝑎

Using the result of equation (228) in equation (227), we get

𝜎 = 𝑍𝐹(𝑢+ + 𝑢− )(1 − 𝜃)𝑐𝑎 (229)

Furthermore, after converting the specific conductivity to equivalent conductivity (Λ = σ/Zca), the equation
(229) takes the form

𝛬 = 𝐹(𝑢+ + 𝑢− )(1 − 𝜃) (230)

Imagine a situation if no ion-association occurs (θ = 0), then the equation (230) would reduce to

𝛬𝜃=0 = 𝐹(𝑢+ + 𝑢− ) (231)

Using equation (231) in equation (230), we get

𝛬 = 𝛬𝜃=0 (1 − 𝜃) (232)

or

𝛬 (233)
=1−𝜃
𝛬𝜃=0

or

𝛬 (234)
𝜃 =1−
𝛬𝜃=0

Now putting the values of θ and 1−θ from equation (234) and (233) in equation (225), we have

𝛬 𝛬 2 (235)
1− = 𝐾𝐴 ( ) 𝑐𝑎 𝑓±2
𝛬𝜃=0 𝛬𝜃=0

Rearranging, we get

1 1 𝐾𝐴 𝑓±2 (236)
= + 2 𝛬𝑐𝑎
𝛬 𝛬𝜃=0 𝛬𝜃=0

Rearranging equation (229) for 𝐹(𝑢+ + 𝑢− ) and putting in equation (231), we get
𝜎 (237)
𝛬𝜃=0 = 𝐹(𝑢+ + 𝑢− ) =
𝑍(1 − 𝜃)𝑐𝑎

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 207

The above equation shows equivalent conductivity at zero ion-association at (1 − 𝜃)𝑐𝑎 concentration.
Therefore, we must use Debye-Huckel-Onsager equation to express the more appropriate conductivity at small
concentrations i.e.

𝛬𝜃=0 = 𝛬0 − (𝐴 + 𝐵𝛬0 )√1 − 𝜃 √𝑐𝑎 (238)

Where the two constants, A and B, are defined as


1 1 (239)
𝐹𝑍𝑒0 8𝜋𝑍 2 𝑒02 𝑁𝐴 2 𝑒02 𝜔 8𝜋𝑍 2 𝑒02 𝑁𝐴 2
𝐴= ( ) 𝑎𝑛𝑑 𝐵= ( )
3𝜋𝜂 1000𝜀𝑘𝑇 6𝜀𝑘𝑇 1000𝜀𝑘𝑇

Where 𝐹 is the Faraday constant and NA is the Avogadro number. The symbol ε represents the dielectric
constant of the medium whereas η is the coefficient of viscosity. 𝑍 is charge numbers of the cation and anion.
The symbol e0 simply shows the electronic charge. ω is a parameter defined earlier in this chapter. The equation
(238) can be simplified as

𝛬𝜃=0 = 𝛬0 𝑊 (240)

Where W represents a continued fraction i.e.

−1/2 −1/2 (241)


𝑊 = 1 − 𝑤 {1 − 𝑤[1 − 𝑤(… )−1/2 ] }

Provided that

(𝐴 + 𝐵𝛬0 )√𝑐𝑎 √𝛬 (242)


𝑤=
(𝛬0 )3/2

Using the result of equation (240) into (236), we get

1 1 𝐾𝐴 𝑓±2 (243)
= + 𝛬𝑐
𝛬 𝛬0 𝑊 (𝛬0 𝑊)2 𝑎

or

𝑊 𝑊 𝐾𝐴 𝑓±2 𝛬𝑐𝑎 (244)


= +
𝛬 𝛬0 (𝛬0 )2 𝑊

After looking at the above equation, it is obvious that the formation of ion-pairs has changed the variation of
equivalent conductivity with concentration drastically. When the formation of ion-pairs was ignored, the
variation of equivalent conductivity with concentration was empirically explained by the Kohlrausch’s law.
Nevertheless, in the case of non-aqueous solvents with low dielectric constant, a significant amount
of ion-pair formation occurs, and therefore, 𝑊/𝛬 is plotted vs 𝑓±2 𝛬𝑐𝑎 /𝑊 with 𝑊/𝛬0 as intercept and
𝐾𝐴 /(𝛬0 )2 as slope.

Copyright © Mandeep Dalal


208 A Textbook of Physical Chemistry – Volume I

Figure 25. The variation of 𝑊/𝛬 is plotted vs 𝑓±2 𝛬𝑐𝑎 /𝑊 for some typical electrolytic solutions.

Copyright © Mandeep Dalal


CHAPTER 4 Electrochemistry – I: Ion-Ion Interactions 209

 Problems
Q 1. Discuss the Debye-Huckel theory of ion-ion interaction in detail.
Q 2. What is excess charge density? How does it vary with the distance from the central ion?
Q 3. Define Debye-Huckel reciprocal length.
Q 4. Derive the expression for the contribution of the ionic cloud to the total potential at a particular distance
from the reference ion in strong electrolytes.
Q 5. State and explain the Debye-Huckel limiting law of activity coefficient. Also, discuss its limitations.
Q 6. What is the ion-size parameter? How does it affect total potential around the central ion?
Q 7. Discuss the asymmetry effect in the conductance of strong electrolytes?
Q 8. Derive and discuss the Debye-Huckel-Onsager equation for aqueous solutions.
Q 9. Discuss the effect of the nature of the solvent on the ionic mobility at infinite dilution.
Q 10. How does the equivalent conductivity vary with the square root of the concentration if the dielectric
constant of the solvent is very low?
Q 11. Define ion-association in strong electrolytic solutions. How does this affect overall conductivity?

Copyright © Mandeep Dalal


210 A Textbook of Physical Chemistry – Volume I

 Bibliography
[1] J. Bockris, A. Reddy, Modern Electrochemistry – Volume 1: Ionics, Kluwer Academic Publishers, New
York, USA 2002.
[2] B. R. Puri, L. R. Sharma, M. S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2008.
[3] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[4] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[5] M. R. Wright, An Introduction to Aqueous Electrolyte Solutions, John Wiley & Sons Ltd, Sussex, UK,
2007.
[6] P. Debye, E. Hückel, The Theory of Electrolytes. I. Lowering of Freezing Point and Related Phenomena,
Physikalische Zeitschrift., 24 (1923) 185-206.
[7] V. S. Bagotsky, Fundamentals of Electrochemistry, John Wiley & Sons, New Jersey, USA, 2006.
[8] R.R. Netz, H. Orland, Beyond Poisson-Boltzmann: Fluctuation effects and correlation functions, The
European Physical Journal E, 1 (2000) 203-214.

Copyright © Mandeep Dalal


CHAPTER 5
Quantum Mechanics – II
 Schrodinger Wave Equation for a Particle in a Three Dimensional Box
In the first chapter of this book, we derived and discussed the Schrodinger wave equation for a particle
in the one-dimensional box. In this chapter, we will extend that procedure to the particle in a three-dimensional
box. In order to do so, consider a particle trapped in a 3-dimensional box of length, breadth, and height as a, b
and c, respectively. This means that this particle can travel in any direction i.e. along x-, y- and z-axis. The
potential inside the box is 0, while outside to the box it is infinite.

Figure 1. The particle in a three-dimensional box.

So far we have considered a quantum mechanical system of a particle trapped in a three-dimensional box. Now
suppose that we need to find various physical properties associated with different states of this system. Had it
been a classical system, we would use simple formulas from classical mechanics to determine the value of
different physical properties. However, being a quantum mechanical system, we cannot use those expressions
because they would give irrational results. Therefore, we need to use the postulates of quantum mechanics to
evaluate various physical properties.
Let ψ be the function that describes all the states of the particle in a three-dimensional box. At this
point we have no information about the exact mathematical expression of ψ; nevertheless, we know that there
is one operator that does not need the absolute expression of wave function but uses the symbolic form only,
the Hamiltonian operator. The operation of Hamiltonian operator over this symbolic form can be rearranged
to give to construct the Schrodinger wave equation; and we all know that the wave function as well the energy,
both are obtained as this second-order differential equation is solved. Mathematically, we can say that

Copyright © Mandeep Dalal


212 A Textbook of Physical Chemistry – Volume I

̂ 𝜓 = 𝐸𝜓
𝐻 (1)

After putting the value of three-dimensional Hamiltonian in equation (1), we get

−ℎ2 𝜕 2 𝜕2 𝜕2 (2)
[ 2 ( 2 + 2 + 2 ) + 𝑉] 𝜓 = 𝐸𝜓
8𝜋 𝑚 𝜕𝑥 𝜕𝑦 𝜕𝑧

or

−ℎ2 𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 (3)
( + + ) + 𝑉𝜓 = 𝐸𝜓
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

−ℎ2 𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 (4)
( + + ) + 𝑉𝜓 − 𝐸𝜓 = 0
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

or

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (5)
+ + + (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

The above-mentioned second order differential equation is the Schrodinger wave equation for a particle
moving along three dimensions. Since the conditions outside and inside the box are different, the equation (5)
must be solved separately for both cases.
1. The solution of Schrodinger wave equation for outside the box: After putting the value of potential
outside the box in equation (5) i.e. V = ∞, we get

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (6)
+ + + (𝐸 − ∞)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

Since E is negligible in comparison to the ∞, the above equation becomes

𝜕2𝜓 𝜕2𝜓 𝜕2𝜓 (7)


+ + − ∞𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

𝜕2𝜓 𝜕2𝜓 𝜕2𝜓 (8)


∞𝜓 = + +
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

1 𝜕2𝜓 𝜕2𝜓 𝜕2𝜓 (9)


𝜓= ( + + )=0
∞ 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

The physical significance of the equation (9) is that the particle cannot go outside the box, and is always
reflected back when it strikes the boundaries. In other words, as the function describing the existence of
particles is zero outside the box, the particle cannot exist outside the box.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 213

2. The solution of Schrodinger wave equation for inside the box: After putting the value of potential inside
the box in equation (5) i.e. V = 0, we get

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (10)
+ + + (𝐸 − 0)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚𝐸 (11)
+ + + 𝜓=0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

The above equation has three variables and is difficult to solve directly. Therefore, it is better to separate
variable, we already know the steps to solve a one-variable equation. To do so, consider that the wave function
ψ is the multiplication of three individual functions as

𝜓(𝑥, 𝑦, 𝑧) = 𝜓(𝑥) × 𝜓(𝑦) × 𝜓(𝑧) = 𝑋𝑌𝑍 (12)

Using the above expression in equation (11), we get

𝜕 2 𝑋𝑌𝑍 𝜕 2 𝑋𝑌𝑍 𝜕 2 𝑋𝑌𝑍 8𝜋 2 𝑚𝐸 (13)


+ + + 𝑋𝑌𝑍 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

From the rules of partial derivative, the equation (13) takes the form

𝜕2𝑋 𝜕2𝑌 𝜕 2 𝑍 8𝜋 2 𝑚𝐸 (14)


𝑌𝑍 + 𝑋𝑍 + 𝑋𝑌 + 𝑋𝑌𝑍 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

Now divide the above equation by XYZ on both side i.e.

1 𝜕 2 𝑋 1 𝜕 2 𝑌 1 𝜕 2 𝑍 8𝜋 2 𝑚𝐸 (15)
+ + + =0
𝑋 𝜕𝑥 2 𝑌 𝜕𝑦 2 𝑍 𝜕𝑧 2 ℎ2

Assuming

8𝜋 2 𝑚𝐸 (16)
𝑘2 =
ℎ2
The equation (15) becomes

1 𝜕2𝑋 1 𝜕2𝑌 1 𝜕2𝑍 (17)


+ + + 𝑘2 = 0
𝑋 𝜕𝑥 2 𝑌 𝜕𝑦 2 𝑍 𝜕𝑧 2

Also fragmenting the constant 𝑘 2 along three x-, y- and z-axis i.e. 𝑘 2 = 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2, the equation (17) can

1 𝜕2𝑋 1 𝜕2𝑌 1 𝜕2𝑍 (18)


+ + + 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2 = 0
𝑋 𝜕𝑥 2 𝑌 𝜕𝑦 2 𝑍 𝜕𝑧 2

The above equation can be written as the sum of three equations with only one variable in each i.e.

Copyright © Mandeep Dalal


214 A Textbook of Physical Chemistry – Volume I

𝜕2𝑋 (19)
+ 𝑘𝑥2 𝑋 = 0
𝜕𝑥 2
𝜕2𝑌 (20)
+ 𝑘𝑦2 𝑌 = 0
𝜕𝑦 2

𝜕2𝑍 (21)
+ 𝑘𝑧2 𝑍 = 0
𝜕𝑧 2
The equations (19-21) are simple one-dimensional differential equations whose solutions can be obtained just
like in the one-dimensional box. The solution of equation (19) will give the x-dependent wave function as well
the energy distribution along x-axis i.e.

(22)
2 𝑛𝑥 𝜋𝑥 𝑛𝑥2 ℎ2
𝜓𝑛𝑥 (𝑥) = 𝑋 = √ 𝑆𝑖𝑛 𝑎𝑛𝑑 𝐸𝑛𝑥 =
𝑎 𝑎 8𝑚𝑎2

Similarly, the solution of equation (20) will be

(23)
2 𝑛𝑦 𝜋𝑦 𝑛𝑦2 ℎ2
𝜓𝑛𝑦 (𝑦) = 𝑌 = √ 𝑆𝑖𝑛 𝑎𝑛𝑑 𝐸𝑛𝑦 =
𝑏 𝑏 8𝑚𝑏 2

Just like the above two, the solution of equation (21) will be

(24)
2 𝑛𝑧 𝜋𝑧 𝑛𝑧2 ℎ2
𝜓𝑛𝑧 (𝑧) = 𝑍 = √ 𝑆𝑖𝑛 𝑎𝑛𝑑 𝐸𝑛𝑧 =
𝑐 𝑐 8𝑚𝑐 2

After putting the expressions of individual wave functions from equation (22-24) in equation (12), the total
wave function can be obtained i.e.

(25)
8 𝑛𝑥 𝜋𝑥 𝑛𝑦 𝜋𝑦 𝑛𝑧 𝜋𝑧
𝜓𝑛𝑥 𝑛𝑦 𝑛𝑧 (𝑥, 𝑦, 𝑧) = √ 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛
𝑎𝑏𝑐 𝑎 𝑏 𝑐

Since 𝑘 2 = 𝑘𝑥2 + 𝑘𝑦2 + 𝑘𝑧2, the total energy must be the sum of individual energies i.e.

𝑛𝑥2 𝑛𝑦2 𝑛𝑧2 ℎ2 (26)


𝐸𝑛𝑥 𝑛𝑦 𝑛𝑧 = ( + + )
𝑎2 𝑏 2 𝑐 2 8𝑚

Where 𝑛𝑥 , 𝑛𝑦 , 𝑛𝑦 are the discrete variable whose permitted values from boundary conditions can be 0, 1, 2, 3,
4….∞. Nevertheless, it is worthy to note that even though the n = 0 is permitted by the boundary conditions,
we still don’t use it in equation (25); which is obviously because it makes the whole function zero.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 215

 The Concept of Degeneracy Among Energy Levels for a Particle in Three


Dimensional Box

The solution of Schrodinger wave equation for a particle of mass ‘m’ trapped in three dimensional of
sides a, b and c with zero potential inside and infinite potential outside provide the total wave function ψ as

(27)
8 𝑛𝑥 𝜋𝑥 𝑛𝑦 𝜋𝑦 𝑛𝑧 𝜋𝑧
𝜓𝑛𝑥 𝑛𝑦 𝑛𝑧 (𝑥, 𝑦, 𝑧) = √ 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛
𝑎𝑏𝑐 𝑎 𝑏 𝑐

Where 𝑛𝑥 , 𝑛𝑦 , 𝑛𝑦 are the discrete variable whose permitted values from boundary conditions can be 1, 2, 3,
4….∞. The variable x, y and z represent the position of the particle along the corresponding axis. Besides, the
expression for total energy is

𝑛𝑥2 𝑛𝑦2 𝑛𝑧2 ℎ2 (28)


𝐸𝑛𝑥 𝑛𝑦 𝑛𝑧 = ( 2 + 2 + 2)
𝑎 𝑏 𝑐 8𝑚

For a cubical box, all the sides become equal (a = b = c). Using this condition in equation (27), the total wave
function representing different quantum mechanical states take the following form.

(29)
8 𝑛𝑥 𝜋𝑥 𝑛𝑦 𝜋𝑦 𝑛𝑧 𝜋𝑧
𝜓𝑛𝑥 𝑛𝑦 𝑛𝑧 (𝑥, 𝑦, 𝑧) = √ 3 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛
𝑎 𝑎 𝑎 𝑎

Similarly, the energy expression also changes to

ℎ2 (30)
𝐸𝑛𝑥 𝑛𝑦 𝑛𝑧 = (𝑛𝑥2 + 𝑛𝑦2 + 𝑛𝑧2 )
8𝑚𝑎2
Now, in order to define various quantum mechanical states, we need to put valid set quantum numbers. The
expression for first quantum mechanical and corresponding energy can be obtained by putting 𝑛𝑥 = 𝑛𝑦 =
𝑛𝑧 = 1 in equations (29–30) i.e.

(31)
8 𝜋𝑥 𝜋𝑦 𝜋𝑧 3ℎ2
𝜓111 = √ 3 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛 and 𝐸111 =
𝑎 𝑎 𝑎 𝑎 8𝑚𝑎2

Similarly, the next state with energy can be obtained by putting 𝑛𝑥 = 𝑛𝑦 = 1 and 𝑛𝑧 = 2 in equations (29–
30) i.e.

(32)
8 𝜋𝑥 𝜋𝑦 2𝜋𝑧 6ℎ2
𝜓112 = √ 3 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛 and 𝐸112 =
𝑎 𝑎 𝑎 𝑎 8𝑚𝑎2

If 𝑛𝑥 = 𝑛𝑧 = 1 and 𝑛𝑦 = 2; the wavefunction and energy become

Copyright © Mandeep Dalal


216 A Textbook of Physical Chemistry – Volume I

(33)
8 𝜋𝑥 2𝜋𝑦 𝜋𝑧 6ℎ2
𝜓121 = √ 3 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛 and 𝐸121 =
𝑎 𝑎 𝑎 𝑎 8𝑚𝑎2

If 𝑛𝑦 = 𝑛𝑧 = 1 and 𝑛𝑥 = 2, the state with energy becomes

(34)
8 2𝜋𝑥 𝜋𝑦 𝜋𝑧 6ℎ2
𝜓211 = √ 3 𝑆𝑖𝑛 𝑆𝑖𝑛 𝑆𝑖𝑛 and 𝐸211 =
𝑎 𝑎 𝑎 𝑎 8𝑚𝑎2

It can be clearly seen that three quantum mechanical states 𝜓112 , 𝜓121 and 𝜓211 possess the same amount of
energy (i.e. 6ℎ2 /8𝑚𝑎2 ); and therefore, are said to be degenerate. In other words, there are three different ways
of existence of the particle inside the box so that the particle possesses 6ℎ2 /8𝑚𝑎2 energy as total.

Figure 2. The energy level diagram representing different quantum mechanical states (in the units of
ℎ2 /8𝑚𝑎2 ) for a particle trapped in a cubical box.

Hence, the degeneracy of the ground state is one i.e. there is only one way for the particle to exist in the box
to create zero-point energy (3ℎ2 /8𝑚𝑎2 ). On the other hand, the degeneracy of first excited stated is 3 as 𝜓112 ,
𝜓121 and 𝜓211, all have 6 units of energy. Moreover, after careful examination of energy diagram, it can be
concluded that degeneracy is 1 if 𝑛𝑥 = 𝑛𝑦 = 𝑛𝑧 , 3 if 𝑛𝑥 = 𝑛𝑦 or 𝑛𝑦 = 𝑛𝑧 or 𝑛𝑥 = 𝑛𝑧 , and 6 if 𝑛𝑥 ≠ 𝑛𝑦 ≠ 𝑛𝑧 .

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 217

 Schrodinger Wave Equation for a Linear Harmonic Oscillator & Its


Solution by Polynomial Method
A diatomic molecule is the quantum-mechanical analog of the classical version of the harmonic
oscillator. It represents the vibrational motion and is one of the few quantum-mechanical systems for which
an exact solution is available. In this section, we will discuss the classical and quantum mechanical oscillator
and their comparative study.
 The Classical Treatment of Simple Harmonic Oscillator
In order to understand the vibrational states of a simple diatomic molecule, we must understand the
classical oscillator first. In order to do so, consider a spring of length r in which a displacement ‘x’ is
incorporated by expending or compressing it.

Figure 2. The pictorial representation of the displacement-inducing in a typical spiral.

For a moment, imagine that the spiral is extended by a displacement of ‘x’; then the restoring force (F)
developed in the spiral can be obtained using Hook’s law as

𝐹 = −𝑘𝑥 (35)

Where k is the constant of proportionality. The minus sign is because the restoring force and the displacement
both are vector quantity but in the opposite direction. In other words, if we expend the spiral, the spiral will
try to compress itself and vice-versa. From equation (35), it seems that the restoring force depends only upon
displacement induced only, however, it is found that stronger spirals have larger restoring force than the weaker
ones for the same magnitude of displacement, indicating a lager force constant. Therefore, the physical
significance of the force constant lies in the fact that it can be used to comment on the strength of oscillator.
Since the potential energy in this expended state is simply the amount of work done in the process of
incorporating the displacement ‘x’, we need calculate the same for further analysis. The restoring force is
proportional to the displacement, and therefore, is a variable quantity; suggesting that we need to carry out the
integration force curve vs displacement. Suppose that the total displacement “x” is fragmented in very small

Copyright © Mandeep Dalal


218 A Textbook of Physical Chemistry – Volume I

“dx” segments. The amount of work done in inducing ‘dx’ displacement will be ‘dw’ and can be given by the
following relation

𝑑𝑤 = 𝐹. 𝑑𝑥 (36)

The total work from zero displacement to ‘x’ displacement will be


𝑥
(37)
𝑊 = ∫ 𝐹. 𝑑𝑥
0

𝑥
(38)
= ∫ −𝑘𝑥. 𝑑𝑥
0

𝑥2
𝑥
(39)
= −𝑘 [ ]
2 0

1 (40)
𝑊 = − 𝑘𝑥 2
2
Since there is no electrostatic attraction, the potential energy (V) of the system at displacement will simply be

1 (41)
𝑉 = 𝑘𝑥 2
2
The above equation represents a parabolic behavior and shows that the potential energy varies continuously
with the displacement. Larger the displacement, higher will be the potential energy.

Figure 3. The variation of potential energy as a function of displacement in a classical oscillator.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 219

If ‘m’ is the reduced mass of the diatomic molecule, the equilibrium vibrational frequency ‘ν’ of the oscillator
can be given as

(42)
1 𝑘
𝜈= √
2𝜋 𝑚

Where m is the reduced mass defined by the ratio of the product to the sum of individual masses i.e. 𝑚 =
𝑚1 𝑚2 /(𝑚1 + 𝑚2 ). It is obvious that the energy levels of a simple harmonic oscillator in classical mechanics
are continuous (including zero), and have a limit over the expansion and compression for each value.
Furthermore, the classical oscillator is bound to spend most of its time in the extreme state (fully compressed
and fully expended) and the least time in the equilibrium position.
 The Quantum Mechanical Treatment of Simple Harmonic Oscillator
In order to find out the quantum mechanical behavior of a simple harmonic oscillator, assume that
all the vibrational states can be described by a mathematical expression ψ. Since we don’t know the exact
nature of ψ, we need to follow the postulates of quantum mechanics. Therefore, after applying the Hamiltonian
operator over this symbolic wave function, we have

𝐻𝜓 = 𝐸𝜓 (43)

−ℎ2 𝜕 2 (44)
( + 𝑉) 𝜓 = 𝐸𝜓
8𝜋 2 𝑚 𝜕𝑥 2

or

−ℎ2 𝜕 2 𝜓 (45)
+ 𝑉𝜓 − 𝐸𝜓 = 0
8𝜋 2 𝑚 𝜕𝑥 2
Rearranging, we have

𝜕 2 𝜓 8𝜋 2 𝑚 (46)
+ (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 ℎ2
After putting the value of potential energy form equation (41) in equation (46), we get

𝜕 2 𝜓 8𝜋 2 𝑚 1 (47)
2
+ 2
(𝐸 − 𝑘𝑥 2 ) 𝜓 = 0
𝜕𝑥 ℎ 2
Now put the value of k form equation (42) in equation (47) i.e.

𝜕 2 𝜓 8𝜋 2 𝑚 (48)
+ (𝐸 − 2𝜋 2 𝜈 2 𝑚𝑥 2 )𝜓 = 0
𝜕𝑥 2 ℎ2
or

Copyright © Mandeep Dalal


220 A Textbook of Physical Chemistry – Volume I

𝜕2𝜓 8𝜋 2 𝑚𝐸 16𝜋 4 𝑚2 𝜈 2 𝑥 2 (49)


+( − )𝜓 = 0
𝜕𝑥 2 ℎ2 ℎ2

After defining constants

8𝜋 2 𝑚𝐸 4𝜋 2 𝑚𝜈 (50)
𝛼= 𝑎𝑛𝑑 𝛽=
ℎ2 ℎ
The equation (49) takes the form

𝜕2𝜓 (51)
+ (𝛼 − 𝛽 2 𝑥 2 )𝜓 = 0
𝜕𝑥 2
or

𝜕2𝜓 (52)
+ (𝛼 − 𝛽𝛽𝑥 2 )𝜓 = 0
𝜕𝑥 2

Now define a new variable 𝑦 = √𝛽𝑥, then we have the derivative as

𝑑𝑦 (53)
= √𝛽
𝑑𝑥
Squaring both side of the equation (53), and then rearranging

𝑑2 𝑦 (54)
=𝛽 𝑜𝑟 𝑑𝑥 2 = 𝑑2 𝑦/𝛽
𝑑𝑥 2
Now put the value of 𝑑𝑥 2 and 𝛽𝑥 2 in equation (52), we get

𝜕2𝜓 (55)
𝛽 + (𝛼 − 𝛽𝑦 2 )𝜓 = 0
𝜕𝑦 2

Dividing the above equation by β, we get

𝜕2𝜓 𝛼 (56)
2
+ ( − 𝑦2) 𝜓 = 0
𝜕𝑦 𝛽

The equation (56) can be solved asymptotically i.e. at very large values of y. Thus, when 𝑦 ≫ 𝛼/𝛽, the equation
(56) becomes

𝜕2𝜓 (57)
− 𝑦2𝜓 = 0
𝜕𝑦 2

The two possible solutions of the above equation are

𝜓 = 𝑒 ±𝑦
2 /2
(58)

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 221

2
Nevertheless, only one of them is acceptable because for 𝜓 = 𝑒 +𝑦 /2, the wavefunction becomes infinite as y
tends to approach ∞. Therefore, the only single-valued, continuous and finite solution we left with is

𝜓 = 𝑒 −𝑦
2 /2
(59)

Since the acceptable solution given above is valid only at very large values of y, it is quite reasonable to think
that the exact solution may also contain some pre-exponential part to attain validity at all values of y. Therefore,
after incorporating some y-dependent unknown function ‘F(y)’ in equation (59), we get

𝜓 = 𝐹(𝑦) 𝑒 −𝑦
2 /2
(60)

In order to find the value of F(y), differentiate the equation (60) first i.e.

𝑑𝜓 2 𝑑𝐹 −𝑦2 /2 (61)
= −𝑦𝑒 −𝑦 /2 . 𝐹 + 𝑒
𝑑𝑦 𝑑𝑦

Differentiating again, we get

𝑑2 𝜓 −
𝑦2 2 2 𝑑𝐹 (62)
= [−𝑦. (−𝑦)𝑒 2 . 𝐹(𝑦) + (−1. 𝑒 −𝑦 /2 . 𝐹(𝑦)) + (−𝑦. 𝑒 −𝑦 /2 )]
𝑑𝑦 2 𝑑𝑦
2 /2 𝑑𝐹 2 𝑑2 𝐹
+ [−𝑦. 𝑒 −𝑦 + 𝑒 −𝑦 /2 2 ]
𝑑𝑦 𝑑𝑦

or

𝑑2 𝜓 2 −𝑦2 /2 −𝑦 2 /2 −𝑦 2 /2
𝑑𝐹 −𝑦 2 /2
𝑑2 𝐹 (63)
= 𝑦 𝑒 . 𝐹(𝑦) − 𝑒 . 𝐹(𝑦) − 2𝑦. 𝑒 + 𝑒
𝑑𝑦 2 𝑑𝑦 𝑑𝑦 2

or

𝑑2 𝜓 𝑑2 𝐹 𝑑𝐹 2 (64)
2
= [ 2
− 2𝑦. + (𝑦 2 − 1) 𝐹] 𝑒 −𝑦 /2
𝑑𝑦 𝑑𝑦 𝑑𝑦

Now, after using equation (60) and equation (64) in equation (56), we get

𝑑2 𝐹 𝑑𝐹 2 𝛼 2 (65)
[ 2
− 2𝑦. + (𝑦 2 − 1) 𝐹] 𝑒 −𝑦 /2 + ( − 𝑦 2 ) 𝐹(𝑦) 𝑒 −𝑦 /2 = 0
𝑑𝑦 𝑑𝑦 𝛽

or

𝑑2 𝐹 𝑑𝐹 𝛼 2 (66)
[ 2 − 2𝑦. + ( − 1) 𝐹(𝑦)] 𝑒 −𝑦 /2 = 0
𝑑𝑦 𝑑𝑦 𝛽
2
Now because of the quantity 𝑒 −𝑦 /2 will be zero only at 𝑦 = ±∞, the sum of the terms present in the bracket
must be zero at normal y-values i.e.

Copyright © Mandeep Dalal


222 A Textbook of Physical Chemistry – Volume I

𝑑2 𝐹 𝑑𝐹 𝛼 (67)
− 2𝑦. + ( − 1) 𝐹(𝑦) = 0
𝑑𝑦 2 𝑑𝑦 𝛽

The above differential equation is a “Hermit differential equation” and can be solved to find the expression for
the “unknown” function 𝐹(𝑦). The solution of equation (67) can be obtained by the polynomial method by
expressing the function F(y) as a power series in terms of variable ‘y’.

𝐹 = 𝑎0 + 𝑎1 𝑦 + 𝑎2 𝑦 2 + 𝑎3 𝑦 3 + 𝑎4 𝑦 4 … … … … (68)

Differentiating the above equation, we get

𝑑𝐹 (69)
= 𝑎1 + 2𝑎2 𝑦 + 3𝑎3 𝑦 2 + 4𝑎4 𝑦 3 … … … …
𝑑𝑦

Differentiating again, we get

𝑑2 𝐹 (70)
= 2𝑎2 + 6𝑎3 𝑦 + 12𝑎4 𝑦 2 … … … …
𝑑𝑦 2

Using equation (68-70) in equation (67), we get

[2𝑎2 + 6𝑎3 𝑦 + 12𝑎4 𝑦 2 … ] − 2𝑦[𝑎1 + 2𝑎2 𝑦 + 3𝑎3 𝑦 2 + 4𝑎4 𝑦 3 … ] (71)


𝛼
+ ( − 1) [𝑎0 + 𝑎1 𝑦 + 𝑎2 𝑦 2 + 𝑎3 𝑦 3 + 𝑎4 𝑦 4 … ] = 0
𝛽

or

[2𝑎2 + 6𝑎3 𝑦 + 12𝑎4 𝑦 2 … ] − [2𝑎1 𝑦 + 4𝑎2 𝑦 2 + 6𝑎3 𝑦 3 + 8𝑎4 𝑦 4 … ] (72)


𝛼 𝛼 𝛼
+ [( − 1) 𝑎0 + ( − 1) 𝑎1 𝑦 + ( − 1) 𝑎2 𝑦 2 + ⋯ ] = 0
𝛽 𝛽 𝛽

After further rearranging


𝛼 𝛼 (73)
[2𝑎2 + ( − 1) 𝑎0 ] + [6𝑎3 𝑦 − 2𝑎1 𝑦 + ( − 1) 𝑎1 𝑦]
𝛽 𝛽
𝛼
+ [12𝑎4 𝑦 2 − 4𝑎2 𝑦 2 + ( − 1) 𝑎2 𝑦 2 ] + ⋯ = 0
𝛽

or
𝛼 𝛼 (74)
[2𝑎2 + ( − 1) 𝑎0 ] + [6𝑎3 − 2𝑎1 + ( − 1) 𝑎1 ] 𝑦
𝛽 𝛽
𝛼
+ [12𝑎4 − 4𝑎2 + ( − 1) 𝑎2 ] 𝑦 2 + ⋯ = 0
𝛽

The above equation is valid only when coefficients of the individual power of y are zero i.e.
For y0

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 223

𝛼 (75)
2𝑎2 + ( − 1) 𝑎0 = 0
𝛽

For y1
𝛼 (76)
6𝑎3 + ( − 1 − 2) 𝑎1 = 0
𝛽

For y2
𝛼 (77)
12𝑎4 + ( − 1 − 4) 𝑎2 = 0
𝛽

Similarly, for yk
𝛼 (78)
(𝑘 + 1)(𝑘 + 2)𝑎𝑘+2 + ( − 1 − 2𝑘) 𝑎𝑘 = 0
𝛽

The above equation can be rearranged for the coefficient 𝑎𝑘+2 i.e.

𝛼
( − 1 − 2𝑘) 𝑎𝑘 (79)
𝛽
𝑎𝑘+2 =−
(𝑘 + 1)(𝑘 + 2)

Where k is an integer. The expression given above is popularly known as the recursion formula, and allows
one to determine the coefficient 𝑎𝑘+2 of the term yk+2 in terms of 𝑎𝑘 -coefficient of the yk term. In simple words,
we can calculate 𝑎2 , 𝑎4 , 𝑎6 etc. in terms of 𝑎0 if we set 𝑎1 = 0; likewise, the coefficients 𝑎3 , 𝑎5 , 𝑎7 etc. can
be obtained in terms of 𝑎1 if we set 𝑎0 = 0.
However, the power series will still be made up of the infinite number of terms, making function F(y)
infinite at 𝑦 = ∞. Therefore, we must restrict the number of terms so that the function remains acceptable.
This can be made possible if, at a certain value of 𝑘 = 𝑛, the numerator in equation (79) becomes zero i.e.
𝛼 𝛼 (80)
− 1 − 2𝑘 = − 1 − 2𝑛 = 0
𝛽 𝛽

or
𝛼 (81)
= 2𝑛 + 1
𝛽

Where 𝑛 = 0, 1, 2, 3 …. etc. The series that is obtained so contains a finite number of terms and is called as
“Hermit polynomial” i.e. 𝐻𝑛 (𝑦). All these Hermit polynomials are generating-function defined and are given
below.

2 𝑑𝑛 −𝑦2 (82)
𝐻𝑛 (𝑦) = (−1)𝑛 . 𝑒 𝑦 . .𝑒
𝑑𝑦 𝑛

Copyright © Mandeep Dalal


224 A Textbook of Physical Chemistry – Volume I

For instance, some of the Hermit polynomials for n = 0, 1 and 2 are calculated as given below.
For n = 0, the equation (82) becomes

2 𝑑0 −𝑦2 (83)
𝐻0 (𝑦) = (−1)0 . 𝑒 𝑦 . .𝑒
𝑑𝑦 0
2 2
𝐻0 (𝑦) = 1. 𝑒 𝑦 . 1. 𝑒 −𝑦 = 𝑒 𝑦
2 −𝑦2
= 𝑒0 (84)

𝐻0 (𝑦) = 1 (85)

For n = 1, the equation (82) becomes

2 𝑑 −𝑦2 (86)
𝐻1 (𝑦) = (−1)1 . 𝑒 𝑦 . .𝑒
𝑑𝑦
2
𝐻1 (𝑦) = (−1). 𝑒 𝑦 . (−2𝑦). 𝑒 −𝑦
2
(87)

𝐻1 (𝑦) = (−1). (−2𝑦). 𝑒 𝑦


2 −𝑦 2
= 2𝑦. 𝑒 0 (88)

𝐻1 (𝑦) = 2𝑦 (89)

For n = 2, the equation (82) becomes

2 𝑑2 −𝑦2 (90)
𝐻2 (𝑦) = (−1)2 . 𝑒 𝑦 . .𝑒
𝑑𝑦 2

2 𝑑 2 (91)
𝐻2 (𝑦) = (+1). 𝑒 𝑦 . . −2𝑦. 𝑒 −𝑦
𝑑𝑦
2 2
𝐻2 (𝑦) = 𝑒 𝑦 [(−2𝑦)𝑒 −𝑦 (−2𝑦) + (−2)𝑒 −𝑦 ]
2
(92)

2
𝐻2 (𝑦) = 𝑒 𝑦 [(4𝑦 2 − 2)𝑒 −𝑦 ]
2
(93)

𝐻2 (𝑦) = (4𝑦 2 − 2)𝑒 𝑦


2 −𝑦 2
= (4𝑦 2 − 2)𝑒 0 (94)

𝐻2 (𝑦) = 4𝑦 2 − 2 (95)

The total wavefunction: After knowing the unknown part F(y), the complete eigenfunction for a simple
harmonic oscillator in the quantum mechanical world can be written as

𝜓𝑛 (𝑦) = 𝑁𝑛 𝐻𝑛 (𝑦) 𝑒 −𝑦
2 /2
(96)

Where 𝑁𝑛 is the normalization constant for nth state while the symbol 𝐻𝑛 (𝑦) represents the Hermit polynomial
of nth order in terms of y-variable. Once the wavefunctions are obtained in terms of y, they can easily be
converted into x-dependent function by simply putting 𝑦 = √𝛽𝑥.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 225

Table 1. Eigenfunctions representing various quantum mechanical states of a simple harmonic oscillator.

α/β n Hn(y) ψn(y) ψn(x)

1 0 1 𝑁0 . 𝑒 −𝑦
2 /2
𝑁0 . 𝑒 −𝛽𝑥
2 /2

3 1 2𝑦 𝑁1 . 2𝑦. 𝑒 −𝑦
2 /2
𝑁1 (2√𝛽𝑥)𝑒 −𝛽𝑥
2 /2

5 2 4𝑦 2 − 2 𝑁2 (4𝑦 2 − 2)𝑒 −𝑦
2 /2
𝑁2 (4𝛽𝑥 2 − 2)𝑒 −𝛽𝑥
2 /2

7 3 8𝑦 3 − 12𝑦 𝑁3 (8𝑦 3 − 12𝑦)𝑒 −𝑦


2 /2
𝑁3 (8𝛽 3/2 𝑥 3 − 12√𝛽𝑥)𝑒 −𝛽𝑥
2 /2

The normalization constant can be obtained by recalling the fact that every wave function must describe the
corresponding state completely. This means that square of wave function under consideration over the whole
configurational space must be equal to unity i.e.

+∞ +∞ 2 (97)
𝛽𝑥 2
∫ 𝜓𝑛2 𝑑𝑥 = ∫ [𝑁𝑛 𝐻𝑛 (𝑥) 𝑒 − 2 ] =1
−∞ −∞

𝑁𝑛2 (98)
2𝑛 𝑛! √𝜋 = 1
√𝛽

√𝛽 (99)
𝑁𝑛2 =
2𝑛 𝑛! √𝜋
1/2 (100)
√𝛽
𝑁𝑛 = ( )
2𝑛 𝑛! √𝜋

It can be clearly seen from the above equation that the normalization consents are different for different states.
For instance, some of the normalization constants are given below.
1/2 (101)
√𝛽 𝛽 1/4
𝑁0 = ( ) =( )
20 0! √𝜋 𝜋
1/2 (102)
√𝛽 1 𝛽 1/4
𝑁1 = ( ) = ( )
21 1! √𝜋 √2 𝜋
1/2 (103)
√𝛽 1𝛽 1/4
𝑁2 = ( ) = ( )
22 2! √𝜋 2√2 𝜋

Copyright © Mandeep Dalal


226 A Textbook of Physical Chemistry – Volume I

The eigenvalues of energy: Since we have already proved that the total wavefunction for a simple harmonic
oscillator is acceptable only when the following condition of equation (81) is satisfied i.e.
𝛼 (104)
= 2𝑛 + 1
𝛽

Furthermore, we also know that

8𝜋 2 𝑚𝐸 4𝜋 2 𝑚𝜈 (105)
𝛼= 𝑎𝑛𝑑 𝛽=
ℎ2 ℎ
After using the value of α and β, the equation (104) take the form

8𝜋 2 𝑚𝐸 ℎ (106)
2
× 2 = 2𝑛 + 1
ℎ 4𝜋 𝑚𝜈
or

4𝜋 2 𝑚𝜈 ℎ2 (107)
𝐸 = (2𝑛 + 1) × 2
ℎ 8𝜋 𝑚
or

ℎ𝜈 (108)
𝐸𝑛 = (2𝑛 + 1)
2
or

1 (109)
𝐸𝑛 = (𝑛 + ) ℎ𝜈
2
Where n is a discrete variable (vibrational quantum number) with values 0, 1, 2, 3…∞. The symbol En
represents the vibrational energies of different vibrational states.
 The Classical and Quantum-Mechanical Interpretation of Vibrational States
In order to have a relative interpretation of vibrational states in the classical and quantum-mechanical
framework, recall the classical expression for potential energy curve of simple harmonic oscillator i.e.

1 (110)
𝑉 = 𝑘𝑥 2
2
Where k is the force constant and x is the displacement induced. Also, the general expressions for all the
vibrational states and corresponding energies are

2 1 (111)
𝜓𝑛 (𝑥) = 𝑁𝑛 𝐻𝑛 (𝑥) 𝑒 −𝛽𝑥 𝑎𝑛𝑑 𝐸𝑛 = (𝑛 + ) ℎ𝜈
2
Where 𝑁𝑛 and 𝐻𝑛 (𝑥) are the normalization constant and Hermit polynomial for nth state.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 227

Figure 4. The depiction of various vibrational states of a simple harmonic oscillator in the classical and
quantum mechanical framework.

After looking at the figure given above, the following points can be made about the differences and similarities
in the classical and quantum mechanical oscillators.
i) It can be clearly seen that the energy levels of a classical oscillator are continuous including zero while the
energy levels the quantum-mechanical analog is discontinuous with zero-point energy of hν/2. In other words,
the classical oscillator can have zero vibrational energy but the vibrational motion cannot be ceased completely
in case of the quantum mechanical version of the simple harmonic oscillator.
ii) There is always a limit over the compression as well as over the expansion in the classical oscillator to have
a certain amount of energy. On the other hand, since the function becomes zero only at infinite displacement,
there is no limit over the compression and expansion in the quantum oscillator theoretically.

Copyright © Mandeep Dalal


228 A Textbook of Physical Chemistry – Volume I

iii) The classical oscillator spends more time in the extreme states i.e. fully compressed and fully expended,
and spends the least time with equilibrium bond length. As far as the ground vibrational state of the quantum
mechanical oscillator is concerned, it spends most time with equilibrium (because the function is maximum
for requ or x = 0), and probability to spend time in compressed and expended mode decreases as the magnitude
of compression and expansion increases.
iv) If we plot the square of wavefunction vs displacement incorporated, it can be clearly seen that the most
probable bond lengths shift towards the compressed and expanded states as the quantum number increases
which is in accordance with the Bohr’s correspondence principle.

Figure 5. The variation of probability as a function of bond length or displacement in different vibrational
states of a simple harmonic oscillator.

Furthermore, it also worthy to note that all 𝜓𝑒𝑣𝑒𝑛 states are symmetric while all 𝜓𝑜𝑑𝑑 wavefunctions are
asymmetric in nature with 𝑛 nodes.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 229

 Zero Point Energy of a Particle Possessing Harmonic Motion and Its


Consequence
In order to understand the minimum or the zero-point energy of a simple harmonic oscillator, recall
the general wavefunction representing all the vibrational states of a simple harmonic oscillator i.e.

𝜓𝑛 (𝑦) = 𝑁𝑛 𝐻𝑛 (𝑦) 𝑒 −𝑦
2 /2
(112)

Where y is a displacement-based variable with a value equal to √𝛽x. The constant β depends upon the reduced
mass of the oscillator (m) and equilibrium vibrational frequency (ν) as

4𝜋 2 𝑚𝜈 (113)
𝛽=

The symbol 𝑁𝑛 and 𝐻𝑛 (𝑦) are the normalization constant and Hermit polynomial for nth state i.e.
1/2 (114)
√𝛽 𝑦2
𝑑𝑛 2
𝑁𝑛 = ( ) 𝑎𝑛𝑑 𝐻𝑛 (𝑦) = (−1)𝑛 . 𝑒 . 𝑛 . 𝑒 −𝑦
2𝑛 𝑛! √𝜋 𝑑𝑦

Also, the general expression for the energies is given below.

1 (115)
𝐸𝑛 = (𝑛 + ) ℎ𝜈
2
Now, for the ground vibrational state (𝑛 = 0), 𝑁0 and 𝐻0 (𝑦) can be obtained from equation (114) i.e.

𝛽 1/4 (116)
𝑁0 = ( ) 𝑎𝑛𝑑 𝐻0 (𝑦) = 1
𝜋
After using the values of y, 𝑁0 and 𝐻0 (𝑦) in equation (112), the ground state function becomes

𝛽 1/4 2
(117)
𝜓0 (𝑥) = ( ) . 1. 𝑒 −𝛽𝑥 /2
𝜋
Hence, the ground state wave function does not collapse at 𝑛 = 0, which means that corresponding energy can
also be obtained by putting 𝑛 = 0 in equation (115) i.e.

1 (118)
𝐸0 = (0 + ) ℎ𝜈
2
or

1 (119)
𝐸0 = ℎ𝜈
2
The above equation gives the minimum energy which is always possessed by a simple harmonic oscillator.

Copyright © Mandeep Dalal


230 A Textbook of Physical Chemistry – Volume I

Figure 6. The variation of ground vibrational wavefunction and probability as a function of bond length or
displacement in a simple harmonic oscillator.

It is well-known that the classical oscillator spends more time in the extreme states i.e. fully
compressed and fully expended, and spends the least time with equilibrium bond length. However, as far as
the ground vibrational state of the quantum mechanical oscillator is concerned, it spends the most time with
equilibrium (because the function is maximum for requ or x = 0), and probability to spend time in compressed
and expended mode decreases as the magnitude of compression and expansion increases. Moreover, there is
always a limit over the compression as well as over the expansion in the classical oscillator to have a certain
amount of energy; however, since the function becomes zero only at infinite displacement, there is no limit
over the compression and expansion in the quantum oscillator theoretically.
It is also worthy to note that the energy given by the equation (119) is in joules. However, in many
textbooks or papers, it is also reported in terms of wavenumbers. To do so, we need first put the value of
frequency as 𝜈 = 𝑐/𝜆 and then 1/𝜆 = 𝜈̅ in the equation (119) i.e.

1 𝑐 1 (119)
𝐸0 = ℎ = ℎ𝑐𝜈̅
2 𝜆 2
Where c is the velocity of light. Now, to convert the zero-point energy in wavenumbers, divide equation (119)
by ℎ𝑐 i.e.

1 ℎ𝑐𝜈̅ 𝜈̅ (120)
𝜈̅0 = =
2 ℎ𝑐 2

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 231

 Schrodinger Wave Equation for Three Dimensional Rigid Rotator


In order to study the rotational behavior of a diatomic molecule, consider a system two masses m1
and m2 joined by a rigid rod of length “r”. Now assume that this dumbbell type geometry rotates about an axis
that is perpendicular to r and passes through the center of mass.

Figure 7. The pictorial representation of the diatomic rigid rotator in classical mechanics.

If 𝑣1 and 𝑣2 are the velocities of the mass m1 and m2 revolving about the axis of rotation, the total kinetic
energy (T) of the rotator can be given by the following relation.

1 1 (121)
𝑇 = 𝑚1 𝑣12 + 𝑚2 𝑣22
2 2
Since we know that linear velocity 𝑣 is simply equal to the angular velocity 𝜔 multiplied by the radius of
rotation r i.e. 𝑣 = 𝜔𝑟, the equation (121) takes the form

1 1 (122)
𝑇 = 𝑚1 (𝑟1 𝜔)2 + 𝑚2 (𝑟2 𝜔)2
2 2
1 (123)
𝑇 = (𝑚1 𝑟12 + 𝑚2 𝑟22 )𝜔2
2
1 (123)
𝑇 = 𝐼𝜔2
2

Where I is the moment of inertia with definition 𝐼 = ∑ 𝑚𝑖 𝑟𝑖2 . Furthermore, we know from mass-center that

Copyright © Mandeep Dalal


232 A Textbook of Physical Chemistry – Volume I

𝑚1 𝑟1 = 𝑚2 𝑟2 (124)

Now since 𝑟 = 𝑟1 + 𝑟2 , we rearrange equation (124) to give


𝑚2 𝑚1 (125)
𝑟1 = 𝑟 𝑎𝑛𝑑 𝑟2 = 𝑟
𝑚1 + 𝑚2 𝑚1 + 𝑚2

In the two-mass system 𝐼 = 𝑚1 𝑟12 + 𝑚2 𝑟22 , so have

𝑚2 2 𝑚1 2 (126)
𝐼 = 𝑚1 ( 𝑟) + 𝑚2 ( 𝑟)
𝑚1 + 𝑚2 𝑚1 + 𝑚2
𝑚1 𝑚2 (127)
𝐼=( ) 𝑟2
𝑚1 + 𝑚2

𝐼 = 𝜇𝑟 2 (128)

Where 𝜇 = 𝑚1 𝑚2 /𝑚1 + 𝑚2 is the reduced mass of the rigid diatomic system. Since we that the kinetic energy
and linear moment of a particle of mass m moving with velocity v are

1 (129)
𝑇 = 𝑚𝑣 2 𝑎𝑛𝑑 𝑝 = 𝑚𝑣
2
The counterparts in the angular motion can be written as

1 (130)
𝑇 = 𝐼𝜔2 𝑎𝑛𝑑 𝐿 = 𝐼𝜔
2
Multiplying and dividing the rotational kinetic energy by I, we have

𝐼 2 𝜔2 (𝐼𝜔)2 𝐿2 (131)
𝑇= = =
2𝐼 2𝐼 2𝐼
It is clear from the above equation that the kinetic energy of a classical rotator can have any value because the
value-domain of angular velocity is continuous. Moreover, as no external force is working on the rotator, the
potential can be set to zero. In other words, the Hamiltonian for diatomic rigid rotator can be given as

̂ = 𝑇̂ + 𝑉̂
𝐻 (132)

𝐿̂2 (133)
̂=
𝐻 +0
2𝐼

The expression for the operator 𝐿̂2 in polar coordinates is

ℎ2 1 𝜕 𝜕 1 𝜕 (134)
𝐿2 = − 2 [ (𝑆𝑖𝑛 𝜃 ) + 2 ]
4𝜋 𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛 𝜃 𝜕𝜙

Using equation (134) in equation (133), the Hamiltonian operator takes the form

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 233

ℎ2 1 𝜕 𝜕 1 𝜕 (135)
̂=−
𝐻 [ (𝑆𝑖𝑛 𝜃 ) + ]+0
8𝜋 2 𝐼 𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙

Now, let ψ be the function that describes all the rotational states of the diatomic rigid rotator. The operation of
Hamiltonian operator over ψ can be rearranged to give to construct the Schrodinger wave equation; and we all
know that the wave function as well the energy, both are the obtained as this second-order differential equation
is solved. Mathematically, we can say that

̂ 𝜓 = 𝐸𝜓
𝐻 (136)

After putting the expression of the Hamiltonian operator from equation (135) in equation (136), we get

ℎ2 1 𝜕 𝜕 1 𝜕 (137)
− 2 [ (𝑆𝑖𝑛 𝜃 ) + ] 𝜓 = 𝐸𝜓
8𝜋 𝐼 𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙

or

ℎ2 1 𝜕 𝜕𝜓 1 𝜕𝜓 (138)
− 2 [ (𝑆𝑖𝑛 𝜃 ) + ] = 𝐸𝜓
8𝜋 𝐼 𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙

or

1 𝜕 𝜕𝜓 1 𝜕𝜓 8𝜋 2 𝐼𝐸𝜓 (139)
(𝑆𝑖𝑛 𝜃 ) + = −
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 ℎ2

or

1 𝜕 𝜕𝜓 1 𝜕𝜓 8𝜋 2 𝐼𝐸𝜓 (140)
(𝑆𝑖𝑛 𝜃 ) + + =0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 ℎ2

The above differential equation contains two variable 𝜙 and 𝜃, and therefore, is difficult to solve. Thus, we
need to use the same mathematical technique we used to study particle in a 3-dimensional box i.e. the
separation of variables. To do so, consider the total wavefunction as the product of two independent, one θ-
dependent and other as a ϕ-dependent function only i.e.

𝜓(𝜃, 𝜙) = 𝜓(𝜃) × 𝜓(𝜙) = 𝛩 × 𝛷 (141)

After putting the value of equation (141) in equation (140), we get

1 𝜕 𝜕𝛩𝛷 1 𝜕𝛩𝛷 8𝜋 2 𝐼𝐸𝛩𝛷 (142)


(𝑆𝑖𝑛 𝜃 )+ + =0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 ℎ2

Since the first and second terms contain the partial derivatives w.r.t. θ and ϕ, respectively; function Φ and Θ
must be kept constant correspondingly, i.e.,

Copyright © Mandeep Dalal


234 A Textbook of Physical Chemistry – Volume I

1 𝜕 𝜕𝛩 1 𝜕𝛷 8𝜋 2 𝐼𝐸𝛩𝛷 (143)
𝛷 (𝑆𝑖𝑛 𝜃 ) + 𝛩 + =0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 ℎ2

Dividing the above equation by 𝛩𝛷, the equation (143) takes the form

1 1 𝜕 𝜕𝛩 1 1 𝜕𝛷 8𝜋 2 𝐼𝐸 (144)
(𝑆𝑖𝑛 𝜃 ) + + =0
𝛩 𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝛷 𝑆𝑖𝑛2 𝜃 𝜕𝜙 ℎ2

After multiplying equation (144) by 𝑆𝑖𝑛2 𝜃, we get

𝑆𝑖𝑛 𝜃 𝜕 𝜕𝛩 1 𝜕𝛷 8𝜋 2 𝐼𝐸 (145)
(𝑆𝑖𝑛 𝜃 ) + + 𝑆𝑖𝑛2 𝜃 = 0
𝛩 𝜕𝜃 𝜕𝜃 𝛷 𝜕𝜙 ℎ2

Rearranging

𝑆𝑖𝑛 𝜃 𝜕 𝜕𝛩 8𝜋 2 𝐼𝐸 2
1 𝜕𝛷 (146)
(𝑆𝑖𝑛 𝜃 ) + 𝑆𝑖𝑛 𝜃 = −
𝛩 𝜕𝜃 𝜕𝜃 ℎ2 𝛷 𝜕𝜙

At this point, we can set both sides equal to constant m2 i.e.

𝑆𝑖𝑛 𝜃 𝜕 𝜕𝛩 8𝜋 2 𝐼𝐸 2 2
1 𝜕𝛷 (147)
(𝑆𝑖𝑛 𝜃 ) + 𝑆𝑖𝑛 𝜃 = 𝑚 = −
𝛩 𝜕𝜃 𝜕𝜃 ℎ2 𝛷 𝜕𝜙

The equation (147) can be fragmented into two equations, each containing a single variable i.e.

𝜕𝛷 (148)
+ 𝑚2 𝛷 = 0
𝜕𝜙

And

𝜕 𝜕𝛩 8𝜋 2 𝐼𝐸 (149)
𝑆𝑖𝑛 𝜃 (𝑆𝑖𝑛 𝜃 ) + 𝛩 𝑆𝑖𝑛2 𝜃 − 𝑚2 𝛩 = 0
𝜕𝜃 𝜕𝜃 ℎ2

Now dividing the above equation by 𝑆𝑖𝑛2 𝜃, we get

1 𝜕 𝜕𝛩 8𝜋 2 𝐼𝐸 𝑚2 𝛩 (150)
(𝑆𝑖𝑛 𝜃 ) + 𝛩 − =0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 ℎ2 𝑆𝑖𝑛2 𝜃
or

1 𝜕 𝜕𝛩 8𝜋 2 𝐼𝐸 𝑚2 (151)
(𝑆𝑖𝑛 𝜃 ) + ( 2 − )𝛩 = 0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 ℎ 𝑆𝑖𝑛2 𝜃

or

1 𝜕 𝜕𝛩 𝑚2 (152)
(𝑆𝑖𝑛 𝜃 ) + (𝛽 − )𝛩 = 0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 235

Where the constant β is defined as

8𝜋 2 𝐼𝐸 (153)
𝛽=
ℎ2
The solution of Φ equation: Recall the differential equation obtained after separation of variables having ϕ
dependence i.e.

𝜕𝛷 (154)
+ 𝑚2 𝛷 = 0
𝜕𝜙

The general solution of such an equation is

𝛷(𝜙) = 𝑁𝑒 𝑖𝑚𝜙 (155)

Where N represents the normalization constant. The wavefunction given above will be acceptable only if m
has integer value i.e. 0, ±1, ±2, etc. This can be understood in terms of single-valued, continuous and finite
nature of quantum states.
i) The boundary condition for function Φ: If we replace the angle “ϕ” with “ϕ + 2π”, the position of the point
under consideration should remain the same i.e.

𝛷(𝜙 + 2𝜋) = 𝛷(𝜙) (156)

Therefore

𝑁𝑒 𝑖𝑚(𝜙+2𝜋) = 𝑁𝑒 𝑖𝑚𝜙 (157)

or

𝑒 𝑖𝑚(𝜙+2𝜋) = 𝑒 𝑖𝑚𝜙 (158)

𝑒 𝑖𝑚𝜙 . 𝑒 𝑖𝑚2𝜋 = 𝑒 𝑖𝑚𝜙 (159)

𝑒 𝑖𝑚2𝜋 = 𝑒 𝑖𝑚𝜙 𝑒 −𝑖𝑚𝜙 (160)

𝑒 𝑖𝑚2𝜋 = 𝑒 𝑖𝑚𝜙−𝑖𝑚𝜙 = 𝑒 0 (161)

𝑒 𝑖𝑚2𝜋 = 1 (162)

Since we know from the Euler’s expansion 𝑒 𝑖𝑥 = 𝐶𝑜𝑠 𝑥 + 𝑖 𝑆𝑖𝑛 𝑥, the equation (162) takes the form

𝑒 𝑖𝑚2𝜋 = 𝐶𝑜𝑠 2𝜋𝑚 + 𝑖 𝑆𝑖𝑛 2𝜋𝑚 (163)

After putting the value of equation (163) in equation (162), we get

𝐶𝑜𝑠 2𝜋𝑚 + 𝑖 𝑆𝑖𝑛 2𝜋𝑚 = 1 (164)

The relation holds true only when we use 𝑚 = 0, ±1, ±2, ±3, ±4, etc.

Copyright © Mandeep Dalal


236 A Textbook of Physical Chemistry – Volume I

ii) The normalization constant for function Φ: In order to determine the normalization constant for the Φ
function, we must put the squared-integral over whole configuration space as unity i.e.
2𝜋 (165)

∫ 𝛷 𝛷 𝑑𝜙 = 1
0

or
2𝜋 (166)
2 𝑖𝑚𝜙 −𝑖𝑚𝜙
𝑁 ∫ 𝑒 .𝑒 𝑑𝜙 = 1
0

2𝜋 2𝜋 (167)
2 𝑖𝑚𝜙−𝑖𝑚𝜙
𝑁 ∫ 𝑒 𝑑𝜙 = 𝑁 ∫ 𝑒 0 𝑑𝜙 = 1
2

0 0

𝑁 2 [𝜙]2𝜋 2
0 = 𝑁 [2𝜋] = 1 (168)

(169)
1
𝑁=√
2𝜋

After using the value of normalization constant in equation (155), we get

(170)
1
𝛷(𝜙) = √ 𝑒 ±𝑖𝑚𝜙
2𝜋

Solution of Θ equation: Recall the differential equation obtained after separation of variables having θ
dependence i.e.

1 𝜕 𝜕𝛩 𝑚2 (171)
(𝑆𝑖𝑛 𝜃 ) + (𝛽 − )𝛩 = 0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃

After defining a new variable 𝑥 = 𝐶𝑜𝑠 𝜃, we have

𝑆𝑖𝑛2 𝜃 + 𝐶𝑜𝑠 2 𝜃 = 1 (172)

𝑆𝑖𝑛2 𝜃 = 1 − 𝐶𝑜𝑠 2 𝜃 (173)

𝑆𝑖𝑛 𝜃 = √1 − 𝐶𝑜𝑠 2 𝜃 (174)

𝑆𝑖𝑛 𝜃 = √1 − 𝑥 2 (175)

Also, since we assumed 𝑥 = 𝐶𝑜𝑠 𝜃, the first derivative w.r.t. θ will be

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 237

𝜕𝑥 (176)
= −𝑆𝑖𝑛 𝜃
𝜕𝜃
The derivative of Θ function w.r.t. θ can be rewritten as

𝜕𝛩 𝜕𝛩 𝜕𝑥 (177)
= .
𝜕𝜃 𝜕𝑥 𝜕𝜃
After putting the values of 𝜕𝑥/𝜕𝜃 from equation (176) in equation (177), we get

𝜕𝛩 𝜕𝛩 (178)
= −𝑆𝑖𝑛 𝜃
𝜕𝜃 𝜕𝑥
After removing Θ from both sides

𝜕 𝜕 (179)
= −𝑆𝑖𝑛 𝜃
𝜕𝜃 𝜕𝑥
Multiplying both sides of equation (178) by 𝑆𝑖𝑛 𝜃, we have

𝜕𝛩 𝜕𝛩 (180)
𝑆𝑖𝑛 𝜃 = −𝑆𝑖𝑛2 𝜃
𝜕𝜃 𝜕𝑥
𝜕𝛩 𝜕𝛩 (181)
𝑆𝑖𝑛 𝜃 = −(1 − 𝑥 2 )
𝜕𝜃 𝜕𝑥
Now, after putting the values of equation (179) and (181) in equation (171), we get

1 𝜕 𝜕𝛩 𝑚2 (182)
(−𝑆𝑖𝑛 𝜃 ) [−(1 − 𝑥 2 ) ] + (𝛽 − )𝛩 = 0
𝑆𝑖𝑛 𝜃 𝜕𝑥 𝜕𝑥 1 − 𝑥2

or

𝜕 𝜕𝛩 𝑚2 (183)
[(1 − 𝑥 2 ) ] + (𝛽 − )𝛩 = 0
𝜕𝑥 𝜕𝑥 1 − 𝑥2

The equation given above is a Legendre’s polynomial and has physical significance only in the range of 𝑥 =
+1 to − 1. Therefore, consider that one more form of Θ function so that this condition is satisfied i.e.
𝑚
𝛩(𝜃) = (1 − 𝑥 2 ) 2 . 𝑋(𝑥) (184)

Where X is a function depending upon variable x. The differentiation of the above equation w.r.t. x yields

𝜕𝛩 𝑚 𝑚 𝑑𝑋 (185)
= −𝑚𝑥(1 − 𝑥 2 ) 2 −1 . 𝑋 + (1 − 𝑥 2 ) 2 .
𝜕𝑥 𝑑𝑥
After multiplying the above equation by 1 − 𝑥 2 and 𝜕/𝜕𝑥, we get

Copyright © Mandeep Dalal


238 A Textbook of Physical Chemistry – Volume I

𝜕 𝜕𝛩 𝜕 𝑚 𝑚 𝑑𝑋 (186)
[(1 − 𝑥 2 ) ] = [−𝑚𝑥(1 − 𝑥 2 ) 2 . 𝑋 + (1 − 𝑥 2 ) 2 +1 . ]
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝑑𝑥
𝑚 𝑚
= [−𝑚(1 − 𝑥 2 )𝑚/2 + 𝑚2 𝑥 2 (1 − 𝑥 2 ) 2 −1 ] 𝑋 − [2𝑥(𝑚 + 1)(1 − 𝑥 2 ) 2 ] 𝑋 ′ (187)
𝑚
+ [(1 − 𝑥 2 ) 2 +1 ] 𝑋 ′′

Where 𝜕/𝜕𝑥 and 𝜕 2 /𝜕𝑥 2 are represented by the symbol 𝑋 ′ and 𝑋 ” , respectively. Now, after using the value of
equation (184) and equation (187) in equation (183), we get

−1
𝑚 𝑚
(188)
[−𝑚(1 − 𝑥 2 )𝑚/2 + 𝑚2 𝑥 2 (1 − 𝑥 2 ) 2 ] 𝑋 − [2𝑥(𝑚 + 1)(1 − 𝑥 2 ) 2 ] 𝑋 ′
𝑚 𝑚2 𝑚
+ [(1 − 𝑥 2 ) 2 +1 ] 𝑋 ′′ + (𝛽 − 2 ) (1 − 𝑥 2 ) 2 . 𝑋 = 0
1−𝑥

Dividing above expression by (1 − 𝑥 2 )𝑚/2 , we have

(1 − 𝑥 2 )𝑋 ′′ − 2(𝑚 + 1)𝑥𝑋 ′ + [𝛽 − 𝑚(𝑚 + 1)]𝑋 = 0 (189)

or

(1 − 𝑥 2 )𝑋 ′′ − 2𝛼𝑥𝑋 ′ + 𝜆𝑋 = 0 (190)

Where 𝛼 = 𝑚 + 1 and 𝜆 = 𝛽 − 𝑚(𝑚 + 1). Now assume that the function X can be expressed as a power
series expansion as given below.

𝑋 = 𝑎0 + 𝑎1 𝑥 + 𝑎2 𝑥 2 + 𝑎3 𝑥 3 … … … … (191)

𝑋 ′ = 𝑎1 + 2𝑎2 𝑥 + 3𝑎3 𝑥 2 … … … … (192)

𝑋 ′′ = 2𝑎2 + 6𝑎3 𝑥 + 12𝑎4 𝑥 2 … … … … (193)

Putting values of equation (191-193) in equation (190), we get

(1 − 𝑥 2 )(2𝑎2 + 6𝑎3 𝑥 + 12𝑎4 𝑥 2 + 20𝑎5 𝑥 3 ) − 2𝛼𝑥(𝑎1 + 2𝑎2 𝑥 + 3𝑎3 𝑥 2 + 4𝑎4 𝑥 3 ) (194)


+ 𝜆(𝑎0 + 𝑎1 𝑥 + 𝑎2 𝑥 2 + 𝑎3 𝑥 3 ) = 0

or

(2𝑎2 + 𝜆𝑎0 ) + [6𝑎3 + (𝜆 − 2𝛼)𝑎1 ]𝑥 + [12𝑎4 + (𝜆 − 2𝛼 − 2)𝑎2 ]𝑥 2 … … … = 0 (195)

The above equation is satisfied only if each term on the left-hand side is individually equal to zero i.e.
coefficients of each power of x are vanish. The general expression for the coefficients must follow the condition
given below.

(𝑛 + 1)(𝑛 + 2)𝑎𝑛+2 + [𝜆 − 2𝑛𝛼 − 𝑛(𝑛 − 1)]𝑎𝑛 = 0 (196)

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 239

Where 𝑛 = 0, 1, 2, 3 etc. Summarizing the result, we can write

2𝑛𝛼 + 𝑛(𝑛 − 1) − 𝜆 (197)


𝑎𝑛+2 = 𝑎𝑛
(𝑛 + 1)(𝑛 + 2)

After putting values of α and λ in equation (197), we get

𝑎𝑛+2 (𝑛 + 𝑚)(𝑛 + 𝑚 + 1) − 𝛽 (198)


=
𝑎𝑛 (𝑛 + 1)(𝑛 + 2)

Which is the Recursion formula for the coefficients of the power of x. Now, in order to obtain a valid
wavefunction, the power series must contain a finite number of terms which is possible only if numerator
becomes zero i.e.

(𝑛 + 𝑚)(𝑛 + 𝑚 + 1) − 𝛽 = 0 (199)

𝛽 = (𝑛 + 𝑚)(𝑛 + 𝑚 + 1) (200)

Since we know that m as well n both are the whole numbers, their sum must also be a whole number. Therefore,
the sum of n and m can be replaced by another whole number symbolized by l i.e.

𝛽 = 𝑙(𝑙 + 1) (201)

Where 𝑙 = 0, 1, 2, 3 etc. After putting the value of β from equation (201) in equation (183), we get

𝜕 𝜕𝛩 𝑚2 (202)
[(1 − 𝑥 2 ) ] + [𝑙(𝑙 + 1) − ]𝛩 = 0
𝜕𝑥 𝜕𝑥 1 − 𝑥2

The general solution of equation (202) is

𝛩 = 𝑁𝑃𝑙𝑚 (𝑥) = 𝑁𝑃𝑙𝑚 (𝐶𝑜𝑠 𝜃) (203)

Where N is the normalization constant and 𝑃𝑙𝑚 (𝑥) is the associated “Legendre function” which is defined as
given below.

𝑑𝑚 𝑃𝑙 (𝑥) (204)
𝑃𝑙𝑚 (𝑥) = (1 − 𝑥 2 )𝑚/2
𝑑𝑥 𝑚
Where 𝑃𝑙 (𝑥) is the Legendre polynomial given by

1 𝑑𝑙 (𝑥 2 − 1)𝑙 (205)
𝑃𝑙 (𝑥) =
2𝑙 𝑙! 𝑑𝑥 𝑙
In order to proceed further, we must discuss the concept of orthogonality and the normalization of the
“Legendre’s function”.
i) Orthogonality of associated Legendre’s function: The orthogonality of the associated Legendre’s polynomial
follows the conditions given below.

Copyright © Mandeep Dalal


240 A Textbook of Physical Chemistry – Volume I

+1 (206)
∫ 𝑃𝑘𝑚 (𝑥) 𝑃𝑙𝑚 (𝑥) =0 𝑖𝑓 𝑘 ≠ 𝑙
−1

+1 (207)
2 (𝑙 + 𝑚)!
∫ 𝑃𝑘𝑚 (𝑥) 𝑃𝑙𝑚 (𝑥) = 𝑖𝑓 𝑘 = 𝑙
(2𝑙 + 1) (𝑙 − 𝑚)!
−1

ii) Normalization of associated Legendre’s function: The normalization of the associated Legendre’s
polynomial follows the conditions given below.
+1 (208)

∫ 𝛩𝑚,𝑙 𝛩𝑚,𝑙 (𝑑𝜃) = 1
−1

+1 (209)
2
𝑁 ∫ 𝑃𝑘𝑚 (𝑥) 𝑃𝑙𝑚 (𝑥) 𝑑𝑥 =1
−1

After solving the integral, we get

2 (𝑙 + 𝑚)! (210)
𝑁2. =1
(2𝑙 + 1) (𝑙 − 𝑚)!

(211)
(2𝑙 + 1)(𝑙 − 𝑚)!
𝑁=√
2(𝑙 + 𝑚)!

Using the value of normalization constant in equation (203), we get

(212)
(2𝑙 + 1)(𝑙 − 𝑚)! 𝑚
𝛩(𝜃) = √ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃)
2(𝑙 + 𝑚)!

The complete eigenfunction of rigid rotator: The total eigenfunction for the rigid rotator now can be
obtained by simply multiplying the solution of ϕ-dependent and θ-dependent differential equations i.e.
equation (170) and equation (203).

(213)
(2𝑙 + 1)(𝑙 − 𝑚)! 𝑚 1
𝜓𝑙,𝑚 (𝜃, 𝜙) = 𝛩𝑙,𝑚 (𝜃)𝛷𝑚 (𝜙) = √ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃). √ 𝑒 ±𝑖𝑚𝜙
2(𝑙 + 𝑚)! 2𝜋

(214)
1 (2𝑙 + 1)(𝑙 − 𝑚)! 𝑚
𝜓𝑙,𝑚 (𝜃, 𝜙) = √ √ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃). 𝑒 ±𝑖𝑚𝜙
2𝜋 2(𝑙 + 𝑚)!

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 241

 Energy of Rigid Rotator


The energy of a rigid rotator can be understood only after considering its classical and quantum
mechanical aspects. In the previous section of this chapter, we discussed the classical and quantum mechanical
nature of the rigid rotator. consider a system two masses m1 and m2 joined by a rigid rod of length “r”. Now
assume that this dumbbell type geometry rotates about an axis that is perpendicular to r and passes through the
center of mass.
 The energy of Classical Rigid Rotator
If 𝑣1 and 𝑣2 are the velocities of the mass m1 and m2 revolving about the axis of rotation, the total
kinetic energy (T) of the rotator can be given by the following relation.

1 1 (215)
𝑇 = 𝑚1 𝑣12 + 𝑚2 𝑣22
2 2
Since we know that linear velocity 𝑣 is simply equal to the angular velocity 𝜔 multiplied by the radius of
rotation r i.e. 𝑣 = 𝜔𝑟, the equation (215) takes the form

1 1 (216)
𝑇 = 𝑚1 (𝑟1 𝜔)2 + 𝑚2 (𝑟2 𝜔)2
2 2
1 (217)
𝑇 = (𝑚1 𝑟12 + 𝑚2 𝑟22 )𝜔2
2
1 (218)
𝑇 = 𝐼𝜔2
2

Where I is the moment of inertia equal with definition 𝐼 = ∑ 𝑚𝑖 𝑟𝑖2 . Furthermore, the value of I can also be
written as
𝑚1 𝑚2 (219)
𝐼=( ) 𝑟2
𝑚1 + 𝑚2

𝐼 = 𝜇𝑟 2 (220)

Where 𝜇 = 𝑚1 𝑚2 /𝑚1 + 𝑚2 is the reduced mass of the rigid diatomic system. After multiplying and dividing
the rotational kinetic energy by I i.e. equation (218), we have

𝐼 2 𝜔2 (𝐼𝜔)2 𝐿2 (221)
𝑇= = =
2𝐼 2𝐼 2𝐼
Where L is the angular momentum of the rotator. It is clear from the above equation that the kinetic energy of
a classical rotator can have any value because the value-domain of angular velocity is continuous. Moreover,
as now the external force is working on the rotator, the potential can be set to zero. Therefore, we can conclude
that the total energy of a classical diatomic rigid rotator is given by equation (221).

Copyright © Mandeep Dalal


242 A Textbook of Physical Chemistry – Volume I

 The energy of Quantum Mechanical Rigid Rotator


In order to understand the energy of a quantum mechanical rigid rotator, recall the Schrodinger wave
equation for the same first i.e.

1 𝜕 𝜕𝜓 1 𝜕𝜓 8𝜋 2 𝐼𝐸𝜓 (222)
(𝑆𝑖𝑛 𝜃 ) + + =0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 ℎ2

Where ψ is the mathematical expression defining various quantum mechanical states depending upon two
variables θ and ϕ. During the course of the solution of the above equation, a constant β is defined for simplicity
as given below.

8𝜋 2 𝐼𝐸 (223)
𝛽=
ℎ2
However, the boundary conditions that keep the function single-valued, continuous and finite; also proved that
the constant β must satisfy the following condition also.

𝛽 = 𝑙(𝑙 + 1) (224)

Where 𝑙 = 0, 1, 2, 3, 4 etc. After equating the value of β from equation (223) and equation (224), we get

8𝜋 2 𝐼𝐸 (225)
= 𝑙(𝑙 + 1)
ℎ2
ℎ2 (226)
𝐸𝑙 = 2 𝑙(𝑙 + 1)
8𝜋 𝐼
Hence, unlike the classical counterpart, the energy levels of quantum mechanical rigid rotators are
discontinuous.

Figure 8. The energy level diagram of the diatomic rigid rotator in units of ℎ2 /8𝜋 2 𝐼.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 243

 Space Quantization
The solution of the Schrodinger wave equation for the diatomic rigid rotator provided the
mathematical descriptions of all the rotational states along with their corresponding energies. The general form
of total eigenfunction for the rigid rotator is given below.

(227)
1 (2𝑙 + 1)(𝑙 − 𝑚)! 𝑚
𝜓𝑙,𝑚 (𝜃, 𝜙) = √ √ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃). 𝑒 ±𝑖𝑚𝜙
2𝜋 2(𝑙 + 𝑚)!

Where ψ is the mathematical expression defining various quantum mechanical states depending upon two
variables θ and ϕ. Furthermore, the most important property of a rigid rotator after energy is the angular
momentum which can be obtained using the last postulate of quantum mechanics i.e.

(228)
< 𝐿 >= ∮ 𝜓𝑙,𝑚 (𝜃, 𝜙) 𝐿̂ 𝜓𝑙,𝑚 (𝜃, 𝜙)

ℎ (229)
𝐿𝑙 = √𝑙(𝑙 + 1)
2𝜋
Alternatively, we know that the energies of various rotational states of rigid rotators are given by the following
relation.

ℎ2 (230)
𝐸= 𝑙(𝑙 + 1)
8𝜋 2 𝐼
Where 𝑙 = 0, 1, 2, 3, 4 etc. Also, we know that the angular momentum and energy are related classically as

1 (𝐼𝜔)2 𝐿2 (231)
𝐸= 𝐼𝜔2 = =
2 2𝐼 2𝐼
or

𝐿 = √2𝐸𝐼 (232)

After using the value of energy from equation (230) into equation (232), we get

(233)
ℎ2
𝐿𝑙 = √2𝐼. 𝑙(𝑙 + 1)
8𝜋 2 𝐼

or

(234)
ℎ2
𝐿𝑙 = √ 2 𝑙(𝑙 + 1)
4𝜋

or

Copyright © Mandeep Dalal


244 A Textbook of Physical Chemistry – Volume I

ℎ (235)
𝐿𝑙 = √𝑙(𝑙 + 1)
2𝜋
Which is exactly the same as given by equation (229). Since 𝑙 = 0, 1, 2, 3, 4 etc., the quantum mechanically
allowed values of angular momentum (in the units of ℎ/2𝜋) are given below.

𝐿0 = √0(0 + 1) unit = 0 unit (236)

𝐿1 = √1(1 + 1) unit = √2 unit (237)

𝐿2 = √2(2 + 1) unit = √6 unit (238)

𝐿3 = √3(3 + 1) unit = √12 unit (239)

However, there is boundary condition in quantum mechanics that says that only integral effects are allowed
reference direction if the angular momentum is generated by integral quantum number and half-integral effects
are allowed in reference direction if the momentum is generated by half-integral quantum number.
This can be understood by taking the example of a diatomic molecule rotating in the first excited
rotational state i.e. 𝑙 = 1. The angular momentum of such a molecule will be √2 or 1.414 units. However,
since this angular momentum is obtained using an integral quantum number (𝑙 = 1), only integral effects (i.e.
+1, 0, −1) are allowed in reference direction. Now if z-axis is the reference direction, the effect of any vector
𝐴 in the reference direction is calculated by multiplying its magnitude with the cosine of the angle it makes
with reference direction.

Figure 9. The angular momentum of the diatomic rigid rotator (left) and its rectangular resolution.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 245

In tringle OPC, the side OC represents the effect of the angular momentum vector 𝐴 along z-axis, can be
calculated as given below.

𝑂𝐶 (240)
= 𝐶𝑜𝑠 𝜃
𝑂𝑃
𝑂𝐶 = 𝑂𝑃. 𝐶𝑜𝑠 𝜃 (241)

𝐴𝑧 = 𝐴 𝐶𝑜𝑠 𝜃 (242)

Hence, a diatomic molecule in its first rotational state cannot rotate in xy-plane since it will generate √2 or
1.414 units of angular momentum along the z-axis (from right-hand thumb rule). In other words, the √2 units
of angular momentum cannot orient itself along z-axis because this makes 𝜃 = 0° and since 𝐶𝑜𝑠 0 = 1, 𝐴𝑧 =
𝐴 i.e. angular momentum effect along the z-axis is also 1.414 unit which is not allowed quantum mechanically.
The effects of angular momentum allowed in the z-direction are +1, 0, −1; for which angles required are
determined as follows.

1 (243)
+1 = √2 𝐶𝑜𝑠 𝜃 ⇒ 𝜃 = 𝐶𝑜𝑠 −1 = 45°
√2
0 (244)
0 = √2 𝐶𝑜𝑠 𝜃 ⇒ 𝜃 = 𝐶𝑜𝑠 −1 = 90°
√2
−1 (245)
−1 = √2 𝐶𝑜𝑠 𝜃 ⇒ 𝜃 = 𝐶𝑜𝑠 −1 = 135°
√2
Hence, we can say that in order to be allowed, the 1.414 units of angular momentum must orient itself only at
45°, 90° and 135° in space from reference direction (z-axis in this case).

Figure 10. The space quantization of angular momentum of rigid rotator in l = 1 rotational state.

Copyright © Mandeep Dalal


246 A Textbook of Physical Chemistry – Volume I

Since the orientation of angular momentum can orient itself in any direction from the z-axis as far as the
effective angular momentum +1 unit along z-direction; therefore, we should use a cone around the same at
45°. The same is true for 0 and −1 effects with 90° and 135°, respectively.

Figure 11. The space quantization of angular momentum of the rigid rotator in l = 0, 1, 2 and 3 states.

It is also worthy to mention that the concept of space quantization is equally applicable to the angular
momentums of all other systems also like orbital or spin angular momentum of electrons or nuclei.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 247

 Schrodinger Wave Equation for Hydrogen Atom: Separation of Variable in


Polar Spherical Coordinates and Its Solution
In the first section of this chapter, we derived and discussed the Schrodinger wave equation for a
particle in a three-dimensional box. In this section, we will apply the procedure to an electron that exits around
the nucleus. In order to do so, consider an electron at a distance r from the center of the nucleus, and this
electron can travel in any direction i.e. along x-, y- and z-axis. The potential energy of such an electron-nucleus
system will be −𝑍𝑒 2 /𝑟; where 𝑍𝑒 and 𝑒 are charges on nucleus and electron respectively.

Figure 12. An electron around nucleus at r distance.

So far we have considered a quantum mechanical system of an electron around the nucleus. Now suppose that
we need to find various physical properties associated with different states of this system. Had it been a
classical system, we would use simple formulas from classical mechanics to determine the value of different
physical properties. However, being a quantum mechanical system, we cannot use those expressions because
they would give irrational results. Therefore, we need to use the postulates of quantum mechanics to evaluate
various physical properties.
Let ψ be the function that describes all the states of the electron around the nucleus. At this point we
have no information about the exact mathematical expression of ψ; nevertheless, we know that there is one
operator that does not need the absolute expression of wave function but uses the symbolic form only, the
Hamiltonian operator. The operation of Hamiltonian operator over this symbolic form can be rearranged to
give to construct the Schrodinger wave equation; and we all know that the wave function as well the energy,
both are the obtained as this second-order differential equation is solved. Mathematically, we can say that

̂ 𝜓 = 𝐸𝜓
𝐻 (246)

After putting the value of three-dimensional Hamiltonian in equation (1), we get

−ℎ2 𝜕 2 𝜕2 𝜕2 (247)
[ 2 ( 2 + 2 + 2 ) + 𝑉] 𝜓 = 𝐸𝜓
8𝜋 𝑚 𝜕𝑥 𝜕𝑦 𝜕𝑧

or

Copyright © Mandeep Dalal


248 A Textbook of Physical Chemistry – Volume I

−ℎ2 𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 (248)
( + + ) + 𝑉𝜓 = 𝐸𝜓
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

−ℎ2 𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 (249)
( + + ) + 𝑉𝜓 − 𝐸𝜓 = 0
8𝜋 2 𝑚 𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 (250)
+ + + (𝐸 − 𝑉)𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2

After putting the value of potential energy of the electron-nucleus system in equation (250), we get

𝜕 2 𝜓 𝜕 2 𝜓 𝜕 2 𝜓 8𝜋 2 𝑚 𝑍𝑒 2 (251)
+ + + (𝐸 + )𝜓 = 0
𝜕𝑥 2 𝜕𝑦 2 𝜕𝑧 2 ℎ2 𝑟

The above-mentioned second order differential equation is the Schrodinger wave equation for an electron
around the nucleus. However, since it is neither completely in cartesian nor completely in polar coordinates
(contains x, y, z as well as r variable), the solution is very much difficult. Therefore, recall the transformation
of cartesian coordinates to polar coordinates in three dimensions as given below.

Figure 13. Correlation between cartesian and polar coordinates in three dimensions.

In tringle AOP, the side OA is simply the z-coordinate and can be obtained as

𝑂𝐴 (252)
= 𝐶𝑜𝑠 𝜃 ⇒ 𝑂𝐴 = 𝑂𝑃 𝐶𝑜𝑠 𝜃 ⇒ 𝑧 = 𝑟 𝐶𝑜𝑠 𝜃
𝑂𝑃
Similarly, in AOP

𝐴𝑃 (253)
= 𝑆𝑖𝑛 𝜃 ⇒ 𝐴𝑃 = 𝑂𝑃 𝑆𝑖𝑛 𝜃 ⇒ 𝑂𝑄 = 𝑟 𝑆𝑖𝑛 𝜃
𝑂𝑃

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 249

In tringle BOQ, the side OB is simply the x-coordinate and can be obtained as

𝑂𝐵 (254)
= 𝐶𝑜𝑠 𝜙 ⇒ 𝑂𝐵 = 𝑂𝑄 𝐶𝑜𝑠 𝜙 ⇒ 𝑥 = 𝑟 𝑆𝑖𝑛 𝜃 𝐶𝑜𝑠 𝜙
𝑂𝑄

Since the side BQ equal to OC, BQ also represents the y-coordinate and can be obtained as

𝐵𝑄 (255)
= 𝑆𝑖𝑛 𝜙 ⇒ 𝐵𝑄 = 𝑂𝑄 𝑆𝑖𝑛 𝜙 ⇒ 𝑦 = 𝑟 𝑆𝑖𝑛 𝜃 𝑆𝑖𝑛 𝜙
𝑂𝑄

Now using equation (252-254), the equation (251) can be transformed to polar coordinates as given below.

1 𝜕 2 𝜕𝜓 1 𝜕 𝜕𝜓 1 𝜕 2 𝜓 8𝜋 2 𝜇 𝑍𝑒 2 (256)
(𝑟 ) + (𝑆𝑖𝑛𝜃 ) + + (𝐸 + )𝜓 = 0
𝑟 2 𝜕𝑟 𝜕𝑟 𝑟 2 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝑟 2 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2 ℎ2 𝑟

or

1 𝜕 2 𝜕𝜓 1 𝜕 𝜕𝜓 1 𝜕2𝜓 8𝜋 2 𝜇 𝑍𝑒 2 (257)
[ (𝑟 ) + (𝑆𝑖𝑛𝜃 ) + ] + (𝐸 + )𝜓 = 0
𝑟 2 𝜕𝑟 𝜕𝑟 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2 ℎ2 𝑟

Which is the Schrodinger wave equation for hydrogen and hydrogen-like species in polar coordinates.
 Separation of Variables
The wave function representing quantum mechanical states, in this case, is actually a function of
three variable r, θ and ϕ. Now, we know that it is easier to solve three differential equations with one variable
in each rather a single differential equation with three variables. Therefore, in order to separate variables,
consider that the wave function ψ is the multiplication of three individual functions as

𝜓(𝑟, 𝜃, 𝜙) = 𝜓(𝑟) × 𝜓(𝜃) × 𝜓(𝜙) = 𝑅. 𝛩. 𝛷 (258)

After putting the value of equation (258) in equation (257) and then multiplying throughout by r2, we get

𝜕 2 𝜕𝑅 𝛷𝑅 𝜕 𝜕𝛩 𝑅𝛩 𝜕 2 𝛷 8𝜋 2 𝜇𝑟 2 𝑍𝑒 2 (259)
𝛩𝛷 (𝑟 )+ (𝑆𝑖𝑛𝜃 ) + + (𝐸 + ) 𝛩𝛷𝑅 = 0
𝜕𝑟 𝜕𝑟 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2 ℎ2 𝑟

Furthermore, divide equation (259) throughout 𝛩𝛷𝑅 i.e.

1 𝜕 2 𝜕𝑅 1 1 𝜕 𝜕𝛩 1 1 𝜕 2 𝛷 8𝜋 2 𝜇𝑟 2 𝑍𝑒 2 (260)
(𝑟 )+ (𝑆𝑖𝑛𝜃 ) + + (𝐸 + )=0
𝑅 𝜕𝑟 𝜕𝑟 𝛩 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝛷 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2 ℎ2 𝑟

or

1 𝜕 2 𝜕𝑅 8𝜋 2 𝜇𝑟 2 𝑍𝑒 2 1 1 𝜕 𝜕𝛩 1 1 𝜕2𝛷 (261)
(𝑟 )+ (𝐸 + ) = − (𝑆𝑖𝑛𝜃 ) −
𝑅 𝜕𝑟 𝜕𝑟 ℎ2 𝑟 𝛩 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝛷 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2

The above equation holds true if we put both sides equal to a constant β i.e.

Copyright © Mandeep Dalal


250 A Textbook of Physical Chemistry – Volume I

1 𝜕 2 𝜕𝑅 8𝜋 2 𝜇𝑟 2 𝑍𝑒 2 (262)
(𝑟 )+ (𝐸 + )=𝛽
𝑅 𝜕𝑟 𝜕𝑟 ℎ2 𝑟

and

1 1 𝜕 𝜕𝛩 1 1 𝜕2𝛷 (263)
(𝑆𝑖𝑛𝜃 ) + = −𝛽
𝛩 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝛷 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2

The equation (262) contains only r variable, and therefore, is called as the “radial equation”. However, the
equation (263) still contains two variable, and thus, needs further separation. To do so, first multiply equation
(263) throughout by 𝑆𝑖𝑛2 𝜃 i.e.

𝑆𝑖𝑛𝜃 𝜕 𝜕𝛩 1 𝜕2𝛷 (264)


(𝑆𝑖𝑛𝜃 ) + = −𝛽 𝑆𝑖𝑛2 𝜃
𝛩 𝜕𝜃 𝜕𝜃 𝛷 𝜕𝜙 2

or

𝑆𝑖𝑛𝜃 𝜕 𝜕𝛩 1 𝜕2𝛷 (265)


(𝑆𝑖𝑛𝜃 ) + 𝛽 𝑆𝑖𝑛2 𝜃 = −
𝛩 𝜕𝜃 𝜕𝜃 𝛷 𝜕𝜙 2

The above equation also holds true if we put both sides equal to a constant m2 i.e.

𝑆𝑖𝑛𝜃 𝜕 𝜕𝛩 (266)
(𝑆𝑖𝑛𝜃 ) + 𝛽 𝑆𝑖𝑛2 𝜃 = 𝑚2
𝛩 𝜕𝜃 𝜕𝜃
and

1 𝜕2𝛷 (267)
= −𝑚2
𝛷 𝜕𝜙 2

The equation (266) contains only θ variable, and therefore, is called as “theta equation”. Likewise, the equation
(267) contains only ϕ variable, and therefore, is called as “phi equation”.
 Solutions of R(r), Θ(θ) and Φ(ϕ) Equations
The single variable equations obtained after separation of variables can be solved separately to yield
r, θ and ϕ-dependent functions which then are multiplied give total wave function.
1. The solution of Φ(ϕ) equation: Recall and rearrange the differential equation obtained after separation of
variables having ϕ dependence i.e.

1 𝜕2𝛷 𝜕𝛷 (268)
= −𝑚2 ⇒ + 𝑚2 𝛷 = 0
𝛷 𝜕𝜙 2 𝜕𝜙

The general solution of such an equation is

𝛷(𝜙) = 𝑁𝑒 𝑖𝑚𝜙 (269)

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 251

Where N represents the normalization constant. The wavefunction given above will be acceptable only if m
has integer value i.e. 0, ±1, ±2, etc. This can be understood in terms of single-valued, continuous and finite
nature of quantum states.
i) The boundary condition for function Φ: If we replace the angle “ϕ” with “ϕ + 2π”, the position of point under
consideration should remain the same i.e.

𝛷(𝜙 + 2𝜋) = 𝛷(𝜙) (270)

Therefore

𝑁𝑒 𝑖𝑚(𝜙+2𝜋) = 𝑁𝑒 𝑖𝑚𝜙 (271)

𝑒 𝑖𝑚(𝜙+2𝜋) = 𝑒 𝑖𝑚𝜙 (272)

𝑒 𝑖𝑚𝜙 . 𝑒 𝑖𝑚2𝜋 = 𝑒 𝑖𝑚𝜙 (273)

𝑒 𝑖𝑚2𝜋 = 𝑒 𝑖𝑚𝜙 𝑒 −𝑖𝑚𝜙 (274)

𝑒 𝑖𝑚2𝜋 = 𝑒 𝑖𝑚𝜙−𝑖𝑚𝜙 = 𝑒 0 (275)

𝑒 𝑖𝑚2𝜋 = 1 (276)

Since we know from the Euler’s expansion 𝑒 𝑖𝑥 = 𝐶𝑜𝑠 𝑥 + 𝑖 𝑆𝑖𝑛 𝑥, the equation (276) takes the form

𝑒 𝑖𝑚2𝜋 = 𝐶𝑜𝑠 2𝜋𝑚 + 𝑖 𝑆𝑖𝑛 2𝜋𝑚 (277)

After putting the value of equation (277) in equation (276), we get

𝐶𝑜𝑠 2𝜋𝑚 + 𝑖 𝑆𝑖𝑛 2𝜋𝑚 = 1 (278)

The relation holds true only when we use 𝑚 = 0, ±1, ±2, ±3, ±4, etc.
ii) The normalization constant for function Φ: In order to determine the normalization constant for the Φ
function, we must put the squared-integral over whole configuration space as unity i.e.
2𝜋 (279)
∫ 𝛷∗ 𝛷 𝑑𝜙 = 1
0

2𝜋 (280)
𝑁 ∫ 𝑒 𝑖𝑚𝜙 . 𝑒 −𝑖𝑚𝜙 𝑑𝜙 = 1
2

2𝜋 2𝜋 (281)
2 𝑖𝑚𝜙−𝑖𝑚𝜙 2 0
𝑁 ∫ 𝑒 𝑑𝜙 = 𝑁 ∫ 𝑒 𝑑𝜙 = 1
0 0

Copyright © Mandeep Dalal


252 A Textbook of Physical Chemistry – Volume I

𝑁 2 [𝜙]2𝜋 2
0 = 𝑁 [2𝜋] = 1 (282)

or

(283)
1
𝑁=√
2𝜋

After using the value of normalization constant in equation (269), we get

(284)
1
𝛷𝑚 (𝜙) = √ 𝑒 𝑖𝑚𝜙
2𝜋

Which is the complete solution of ϕ-equation.

Table 1. Complex and real forms of some normalized Φ-functions.

|𝑚| Complex form Real form

0
1 1
𝛷0 (𝜙) = √ 𝛷0 (𝜙) = √
2𝜋 2𝜋

1
1 1
𝛷+1 (𝜙) = √ 𝑒 𝑖𝜙 𝛷+1 (𝜙) = √ 𝐶𝑜𝑠 𝜙
2𝜋 𝜋

1 1
𝛷−1 (𝜙) = √ 𝑒 −𝑖𝜙 𝛷−1 (𝜙) = √ 𝑆𝑖𝑛 𝜙
2𝜋 𝜋

2
1 1
𝛷+2 (𝜙) = √ 𝑒 𝑖2𝜙 𝛷+2 (𝜙) = √ 𝐶𝑜𝑠 (2𝜙)
2𝜋 𝜋

1 −𝑖2𝜙 1
𝛷−2 (𝜙) = √ 𝑒 𝛷−2 (𝜙) = √ 𝑆𝑖𝑛 (2𝜙)
2𝜋 𝜋

3
1 1
𝛷+3 (𝜙) = √ 𝑒 𝑖3𝜙 𝛷+3 (𝜙) = √ 𝐶𝑜𝑠 (3𝜙)
2𝜋 𝜋

1 −𝑖3𝜙 1
𝛷−3 (𝜙) = √ 𝑒 𝛷−3 (𝜙) = √ 𝑆𝑖𝑛 (3𝜙)
2𝜋 𝜋

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 253

2. The solution of Θ(θ) equation: Recall and rearrange the differential equation obtained after separation of
variables having θ-dependence i.e.

𝑆𝑖𝑛𝜃 𝜕 𝜕𝛩 (285)
(𝑆𝑖𝑛𝜃 ) + 𝛽 𝑆𝑖𝑛2 𝜃 = 𝑚2
𝛩 𝜕𝜃 𝜕𝜃
or

𝜕 𝜕𝛩 (286)
𝑆𝑖𝑛 𝜃 (𝑆𝑖𝑛 𝜃 ) + 𝛩𝛽𝑆𝑖𝑛2 𝜃 − 𝑚2 𝛩 = 0
𝜕𝜃 𝜕𝜃

Now dividing the above equation by 𝑆𝑖𝑛2 𝜃, we get

1 𝜕 𝜕𝛩 𝑚2 𝛩 (287)
(𝑆𝑖𝑛 𝜃 ) + 𝛩𝛽 − =0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃
or

1 𝜕 𝜕𝛩 𝑚2 (288)
(𝑆𝑖𝑛 𝜃 ) + (𝛽 − )𝛩 = 0
𝑆𝑖𝑛 𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃

After defining a new variable 𝑥 = 𝐶𝑜𝑠 𝜃, we have

𝑆𝑖𝑛2 𝜃 + 𝐶𝑜𝑠 2 𝜃 = 1 (289)

𝑆𝑖𝑛2 𝜃 = 1 − 𝐶𝑜𝑠 2 𝜃 (290)

𝑆𝑖𝑛 𝜃 = √1 − 𝐶𝑜𝑠 2 𝜃 (291)

𝑆𝑖𝑛 𝜃 = √1 − 𝑥 2 (292)

Also, since we assumed 𝑥 = 𝐶𝑜𝑠 𝜃, the first derivative w.r.t. θ will be

𝜕𝑥 (293)
= −𝑆𝑖𝑛 𝜃
𝜕𝜃
The derivative of Θ function w.r.t. θ can be rewritten as

𝜕𝛩 𝜕𝛩 𝜕𝑥 (294)
= .
𝜕𝜃 𝜕𝑥 𝜕𝜃
After putting the values of 𝜕𝑥/𝜕𝜃 from equation (293) in equation (294), we get

𝜕𝛩 𝜕𝛩 (295)
= −𝑆𝑖𝑛 𝜃
𝜕𝜃 𝜕𝑥
After removing Θ from both sides

Copyright © Mandeep Dalal


254 A Textbook of Physical Chemistry – Volume I

𝜕 𝜕 (296)
= −𝑆𝑖𝑛 𝜃
𝜕𝜃 𝜕𝑥
Multiplying both sides of equation (295) by 𝑆𝑖𝑛 𝜃, we have

𝜕𝛩 𝜕𝛩 (297)
𝑆𝑖𝑛 𝜃 = −𝑆𝑖𝑛2 𝜃
𝜕𝜃 𝜕𝑥
𝜕𝛩 𝜕𝛩 (298)
𝑆𝑖𝑛 𝜃 = −(1 − 𝑥 2 )
𝜕𝜃 𝜕𝑥
Now, after putting the values of equation (296) and (298) in equation (288), we get

1 𝜕 𝜕𝛩 𝑚2 (299)
(−𝑆𝑖𝑛 𝜃 ) [−(1 − 𝑥 2 ) ] + (𝛽 − )𝛩 = 0
𝑆𝑖𝑛 𝜃 𝜕𝑥 𝜕𝑥 1 − 𝑥2

𝜕 𝜕𝛩 𝑚2 (300)
[(1 − 𝑥 2 ) ] + (𝛽 − )𝛩 = 0
𝜕𝑥 𝜕𝑥 1 − 𝑥2

The equation given above is a Legendre’s polynomial and has physical significance only in the range of 𝑥 =
+1 to − 1. Therefore, consider that one more form of Θ function so that this condition is satisfied i.e.
𝑚
𝛩(𝜃) = (1 − 𝑥 2 ) 2 . 𝑋(𝑥) (301)

Where X is a function depending upon variable x. The differentiation of the above equation w.r.t. x yields

𝜕𝛩 𝑚 𝑚 𝑑𝑋 (302)
= −𝑚𝑥(1 − 𝑥 2 ) 2 −1 . 𝑋 + (1 − 𝑥 2 ) 2 .
𝜕𝑥 𝑑𝑥
After multiplying the above equation by 1 − 𝑥 2 and 𝜕/𝜕𝑥, we get

𝜕 𝜕𝛩 𝜕 𝑚 𝑚
+1 𝑑𝑋 (303)
[(1 − 𝑥 2 ) ] = [−𝑚𝑥(1 − 𝑥 2 ) 2 . 𝑋 + (1 − 𝑥 2 ) 2 . ]
𝜕𝑥 𝜕𝑥 𝜕𝑥 𝑑𝑥
𝑚
= [−𝑚(1 − 𝑥 2 )𝑚/2 + 𝑚2 𝑥 2 (1 − 𝑥 2 ) 2 −1 ] 𝑋 − [2𝑥(𝑚 + 1)(1 − 𝑥 2 ) 2 ] 𝑋 ′
𝑚
(304)
𝑚
+ [(1 − 𝑥 2 ) 2 +1 ] 𝑋 ′′

Where 𝜕/𝜕𝑥 and 𝜕 2 /𝜕𝑥 2 are represented by the symbol 𝑋 ′ and 𝑋 ” , respectively. Now, after using the value of
equation (301) and equation (304) in equation (300), we get
𝑚
−1
𝑚
(305)
[−𝑚(1 − 𝑥 2 )𝑚/2 + 𝑚2 𝑥 2 (1 − 𝑥 2 ) 2 ] 𝑋 − [2𝑥(𝑚 + 1)(1 − 𝑥 2 ) 2 ] 𝑋 ′
𝑚 𝑚2 𝑚
+ [(1 − 𝑥 2 ) 2 +1 ] 𝑋 ′′ + (𝛽 − ) (1 − 𝑥 2) 2
.𝑋 = 0
1 − 𝑥2

Dividing the above expression by (1 − 𝑥 2 )𝑚/2 , we have

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 255

(1 − 𝑥 2 )𝑋 ′′ − 2(𝑚 + 1)𝑥𝑋 ′ + [𝛽 − 𝑚(𝑚 + 1)]𝑋 = 0 (306)

or

(1 − 𝑥 2 )𝑋 ′′ − 2𝛼𝑥𝑋 ′ + 𝜆𝑋 = 0 (307)

Where 𝛼 = 𝑚 + 1 and 𝜆 = 𝛽 − 𝑚(𝑚 + 1). Now assume that the function X can be expressed as a power
series expansion as given below.

𝑋 = 𝑎0 + 𝑎1 𝑥 + 𝑎2 𝑥 2 + 𝑎3 𝑥 3 … … … … (308)

𝑋 ′ = 𝑎1 + 2𝑎2 𝑥 + 3𝑎3 𝑥 2 … … … … (309)

𝑋 ′′ = 2𝑎2 + 6𝑎3 𝑥 + 12𝑎4 𝑥 2 … … … … (310)

Putting values of equation (308-310) in equation (307), we get

(1 − 𝑥 2 )(2𝑎2 + 6𝑎3 𝑥 + 12𝑎4 𝑥 2 + 20𝑎5 𝑥 3 ) − 2𝛼𝑥(𝑎1 + 2𝑎2 𝑥 + 3𝑎3 𝑥 2 + 4𝑎4 𝑥 3 ) (311)


+ 𝜆(𝑎0 + 𝑎1 𝑥 + 𝑎2 𝑥 2 + 𝑎3 𝑥 3 ) = 0

or

(2𝑎2 + 𝜆𝑎0 ) + [6𝑎3 + (𝜆 − 2𝛼)𝑎1 ]𝑥 + [12𝑎4 + (𝜆 − 2𝛼 − 2)𝑎2 ]𝑥 2 … … … = 0 (312)

The above equation is satisfied only if each term on the left-hand side is individually equal to zero i.e.
coefficients of each power of x are vanish. The general expression for the coefficients must follow the condition
given below.+

(𝑛 + 1)(𝑛 + 2)𝑎𝑛+2 + [𝜆 − 2𝑛𝛼 − 𝑛(𝑛 − 1)]𝑎𝑛 = 0 (313)

Where 𝑛 = 0, 1, 2, 3 etc. Summarizing the result, we can write

2𝑛𝛼 + 𝑛(𝑛 − 1) − 𝜆 (314)


𝑎𝑛+2 = 𝑎𝑛
(𝑛 + 1)(𝑛 + 2)

After putting values of α and λ in equation (314), we get

𝑎𝑛+2 (𝑛 + 𝑚)(𝑛 + 𝑚 + 1) − 𝛽 (315)


=
𝑎𝑛 (𝑛 + 1)(𝑛 + 2)

Which is the Recursion formula for the coefficients of the power of x. Now, in order to obtain a valid
wavefunction, the power series must contain a finite number of terms which is possible only if numerator
becomes zero i.e.

(𝑛 + 𝑚)(𝑛 + 𝑚 + 1) − 𝛽 = 0 (316)

𝛽 = (𝑛 + 𝑚)(𝑛 + 𝑚 + 1) (317)

Copyright © Mandeep Dalal


256 A Textbook of Physical Chemistry – Volume I

Since we know that m as well n both are the whole numbers, their sum must also be a whole number. Therefore,
the sum of n and m can be replaced by another whole number symbolized by l i.e.

𝛽 = 𝑙(𝑙 + 1) (318)

Where 𝑙 = 0, 1, 2, 3 etc. After putting the value of β from equation (318) in equation (300), we get

𝜕 2
𝜕𝛩 𝑚2 (319)
[(1 − 𝑥 ) ] + [𝑙(𝑙 + 1) − ]𝛩 = 0
𝜕𝑥 𝜕𝑥 1 − 𝑥2

The general solution of equation (319) is

𝛩 = 𝑁𝑃𝑙𝑚 (𝑥) = 𝑁𝑃𝑙𝑚 (𝐶𝑜𝑠 𝜃) (320)

Where N is the normalization constant and 𝑃𝑙𝑚 (𝑥) is the associated “Legendre function” which is defined as
given below.

𝑑𝑚 𝑃𝑙 (𝑥) (321)
𝑃𝑙𝑚 (𝑥) = (1 − 𝑥 2 )𝑚/2
𝑑𝑥 𝑚
Where 𝑃𝑙 (𝑥) is the Legendre polynomial given by

1 𝑑𝑙 (𝑥 2 − 1)𝑙 (322)
𝑃𝑙 (𝑥) =
2𝑙 𝑙! 𝑑𝑥 𝑙
In order to proceed further, we must discuss the concept of orthogonality and the normalization of the
“Legendre’s function”.
i) Orthogonality of associated Legendre’s function: The orthogonality of the associated Legendre’s polynomial
follows the conditions given below.
+1 (323)
∫ 𝑃𝑘𝑚 (𝑥) 𝑃𝑙𝑚 (𝑥) =0 𝑖𝑓 𝑘 ≠ 𝑙
−1

+1
(𝑙 + 𝑚)! (324)
2
∫ 𝑃𝑘𝑚 (𝑥) 𝑃𝑙𝑚 (𝑥) = 𝑖𝑓 𝑘 = 𝑙
(2𝑙 + 1) (𝑙 − 𝑚)!
−1

ii) Normalization of associated Legendre’s function: The normalization of the associated Legendre’s
polynomial follows the conditions given below.
+1 (325)

∫ 𝛩𝑚,𝑙 𝛩𝑚,𝑙 (𝑑𝜃) = 1
−1

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 257

+1 (326)
2
𝑁 ∫ 𝑃𝑘𝑚 (𝑥) 𝑃𝑙𝑚 (𝑥) 𝑑𝑥 =1
−1

2 (𝑙 + 𝑚)! (327)
𝑁2. =1
(2𝑙 + 1) (𝑙 − 𝑚)!

(328)
(2𝑙 + 1)(𝑙 − 𝑚)!
𝑁=√
2(𝑙 + 𝑚)!

Using the value of normalization constant in equation (320), we get

(329)
(2𝑙 + 1)(𝑙 − 𝑚)! 𝑚
𝛩𝑙,𝑚 (𝜃) = √ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃)
2(𝑙 + 𝑚)!

Which is the complete solution of Θ-equation.

Table 2. Some normalized Θ-functions and corresponding spherical harmonics.

Θ-functions Spherical harmonics

1
𝛩0,0 = 1 1
√2 𝛶0,0 = .√
√2 2𝜋

3 3 1
𝛩1,0 = √ 𝐶𝑜𝑠 𝜃 𝛶1,0 = √ 𝐶𝑜𝑠 𝜃.
2 2 √2𝜋

3 3 1
𝛩1,±1 = √ 𝑆𝑖𝑛 𝜃 𝛶1,±1 = √ 𝑆𝑖𝑛 𝜃. √ 𝑒 ±𝑖𝜙
4 4 2𝜋

5 5 1
𝛩2,0 = √ (3𝐶𝑜𝑠 2 𝜃 − 1) 𝛶2,0 = √ (3𝐶𝑜𝑠 2 𝜃 − 1).
8 8 √2𝜋

15 15 1
𝛩2,±1 = √ 𝑆𝑖𝑛 𝜃 𝐶𝑜𝑠 𝜃 𝛶2,±1 = √ 𝑆𝑖𝑛 𝜃 𝐶𝑜𝑠 𝜃. √ 𝑒 ±𝑖𝜙
4 4 2𝜋

15 15 1
𝛩2,±2 = √ 𝑆𝑖𝑛2 𝜃 𝛶2,±2 = √ 𝑆𝑖𝑛2 𝜃. √ 𝑒 ±𝑖2𝜙
16 16 2𝜋

Copyright © Mandeep Dalal


258 A Textbook of Physical Chemistry – Volume I

3. The solution of R(r) equation: Recall and rearrange the differential equation obtained after separation of
variables having r-dependence i.e.

1 𝜕 2 𝜕𝑅 8𝜋 2 𝜇𝑟 2 𝑍𝑒 2 (330)
(𝑟 )+ (𝐸 + )=𝛽
𝑅 𝜕𝑟 𝜕𝑟 ℎ2 𝑟

After putting ħ = h/2π and rearranging, we get

1 𝜕 2 𝜕𝑅 2𝜇𝑟 2 (331)
(𝑟 ) + 2 (𝐸 − 𝑉) = 𝛽
𝑅 𝜕𝑟 𝜕𝑟 ħ
After multiplying by R on both sides and then dividing by r2 throughout, we get

1 𝜕 2 𝜕𝑅 2𝜇 𝛽𝑅 (332)
2
(𝑟 ) + 2 (𝐸 − 𝑉)𝑅 = 2
𝑟 𝜕𝑟 𝜕𝑟 ħ 𝑟
Now, as we know from the solution of Θ-equation that 𝛽 = 𝑙(𝑙 + 1), the above equation takes the form

1 𝜕 2 𝜕𝑅 2𝜇 𝑙(𝑙 + 1)𝑅 (333)


2
(𝑟 ) + 2 (𝐸 − 𝑉)𝑅 =
𝑟 𝜕𝑟 𝜕𝑟 ħ 𝑟2
1 𝜕 2 𝜕𝑅 2𝜇 𝑙(𝑙 + 1)𝑅 (334)
2
(𝑟 ) + 2 (𝐸 − 𝑉)𝑅 − =0
𝑟 𝜕𝑟 𝜕𝑟 ħ 𝑟2
or

1 2 𝜕2𝑅 𝜕𝑅 2𝜇 𝑙(𝑙 + 1) (335)


2
[𝑟 2
+ 2𝑟 ] + [ 2 (𝐸 − 𝑉) − ]𝑅 = 0
𝑟 𝜕𝑟 𝜕𝑟 ħ 𝑟2

𝜕 2 𝑅 2 𝜕𝑅 2𝜇 𝑙(𝑙 + 1) (336)
[ 2+ ] + [ 2 (𝐸 − 𝑉) − ]𝑅 = 0
𝜕𝑟 𝑟 𝜕𝑟 ħ 𝑟2

Putting the value of potential energy for atomic hydrogen or hydrogen-like species again in the above equation,
we get

𝜕 2 𝑅 2 𝜕𝑅 2𝜇𝐸 2𝜇𝑍𝑒 2 𝑙(𝑙 + 1) (337)


+ + [ + 2 − ]𝑅 = 0
𝜕𝑟 2 𝑟 𝜕𝑟 ħ2 ħ 𝑟 𝑟2

As we know from the classical mechanics that elliptical orbits represent bound states have energies less than
zero whereas hyperbolic orbits represent unbound states have energies greater than zero. Now assume that
electron around the nucleus is bound somehow i.e.

2𝜇𝐸 𝜇𝑍𝑒 2 (338)


− = 𝛼2 𝑎𝑛𝑑 =𝜆
ħ2 ħ2 𝛼
Using equation (338) in equation (337), we get

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 259

𝜕 2 𝑅 2 𝜕𝑅 2
2𝛼𝜆 𝑙(𝑙 + 1) (339)
+ + [−𝛼 + − ]𝑅 = 0
𝜕𝑟 2 𝑟 𝜕𝑟 𝑟 𝑟2

At this stage, we need to define a new variable 𝜌 = 2𝛼𝑟, so that

𝜕𝜌 (340)
= 2𝛼
𝜕𝑟
Which follows

𝜕𝑅 𝜕𝑅 𝜕𝜌 𝜕𝑅 (341)
= . = 2𝛼
𝜕𝑟 𝜕𝜌 𝜕𝑟 𝜕𝜌

Also

𝜕2𝑅 𝜕 𝜕𝑅 𝜕 𝜕𝑅 𝜕 𝜕𝜌 𝜕𝑅 𝜕𝜌 𝜕 𝜕𝑅 (341)
= [ ] = [2𝛼 ] = [2𝛼 ] = [2𝛼 ]
𝜕𝑟 2 𝜕𝑟 𝜕𝑟 𝜕𝑟 𝜕𝜌 𝜕𝑟 𝜕𝜌 𝜕𝜌 𝜕𝑟 𝜕𝜌 𝜕𝜌

𝜕2𝑅 𝜕 𝜕𝑅 2
𝜕2𝑅 (342)
= 2𝛼 [2𝛼 ] = 4𝛼
𝜕𝑟 2 𝜕𝜌 𝜕𝜌 𝜕𝜌2

After using the values of 𝜕𝑅/𝜕𝑟 and 𝜕 2 𝑅/𝜕𝑟 2 from equation (341) and equation (342) in equation (339), we
get the following.

𝜕2𝑅 2
2
𝜕𝑅 2
2𝛼𝜆 𝑙(𝑙 + 1) (343)
4𝛼 + 2𝛼 + [−𝛼 + − ]𝑅 = 0
𝜕𝜌2 𝑟 𝜕𝜌 𝑟 𝑟2

Now divide the above equation by 4𝛼 2 i.e.

𝜕 2 𝑅 1 𝜕𝑅 1 𝜆 𝑙(𝑙 + 1) (344)
2
+ + [− + − ]𝑅 = 0
𝜕𝜌 𝛼𝑟 𝜕𝜌 4 2𝛼𝑟 4𝛼 2 𝑟 2

Using 𝜌 = 2𝛼𝑟, we get

𝜕 2 𝑅 2 𝜕𝑅 1 𝜆 𝑙(𝑙 + 1) (345)
2
+ + [− + − ]𝑅 = 0
𝜕𝜌 𝜌 𝜕𝜌 4 𝜌 𝜌2

When 𝜌 → ∞, the above equation takes the form

𝜕2𝑅 1 (346)
− 𝑅=0
𝜕𝜌2 4

The general solutions of the differential equation given above are

𝑅(𝜌) = 𝑒 +𝜌/2 𝑎𝑛𝑑 𝑅(𝜌) = 𝑒 −𝜌/2 (347)

The function 𝑅(𝜌) = 𝑒 +𝜌/2 becomes ∞ when ρ = ∞, and hence, is not acceptable. Therefore, we are left with

Copyright © Mandeep Dalal


260 A Textbook of Physical Chemistry – Volume I

𝑅(𝜌) = 𝑒 −𝜌/2 (348)

Since the acceptable solution given above is valid only at very large values of ρ, it is quite reasonable to think
that the exact solution may also contain some pre-exponential part to attain validity at all values of ρ. Therefore,
after incorporating some ρ-dependent unknown function ‘F(ρ)’ in equation (348), we get

𝑅(𝜌) = 𝐹(𝜌) 𝑒 −𝜌/2 (349)

Differentiating above equation with w.r.t ρ at first and second order and then putting the values of 𝑅(𝜌),
𝜕𝑅/𝜕𝜌 and 𝜕 2 𝑅/𝜕𝜌2 in equation (345), we get

𝜕2𝐹 2 𝜕𝐹 1 𝜆 𝑙(𝑙 + 1) (350)


2
+ ( − 1) + [− + − ]𝐹 = 0
𝜕𝜌 𝜌 𝜕𝜌 𝜌 𝜌 𝜌2

For simplification, put 𝜕 2 𝑅/𝜕𝜌2 = 𝐹 ′′ and 𝜕𝑅/𝜕𝜌 = 𝐹 ′ i.e.

2 1 𝜆 𝑙(𝑙 + 1) (351)
𝐹 ′′ + ( − 1) 𝐹 ′ + [− + − ]𝐹 = 0
𝜌 𝜌 𝜌 𝜌2

Hence, the problem has been reduced to the determination of the solution of F which can be assumed as

𝐹(𝜌) = 𝜌 𝑠 𝐺(𝜌) (352)

Where 𝐺(𝜌) represents a power series expansion of ρ i.e.

𝐺(𝜌) = 𝑎0 + 𝑎1 𝜌 + 𝑎2 𝜌2 + 𝑎3 𝜌3 … (353)

Or we can say that


𝑘=∞ (354)
𝑘
𝐺(𝜌) = ∑ 𝑎𝑘 𝜌
𝑘=0

It is also worthy to mention that 𝑎0 ≠ 0. Now differentiating equation (352) w.r.t. ρ, we get

𝐹 ′ (𝜌) = 𝑠𝜌 𝑠−1 𝐺 + 𝜌 𝑠 𝐺 ′ (355)

The double derivative of the same will be

𝐹 ′′ (𝜌) = 𝑠(𝑠 − 1)𝜌 𝑠−2 𝐺 + 2𝑠𝜌 𝑠−1 𝐺 ′ + 𝜌 𝑠 𝐺 ′′ (356)

After putting the values of 𝐹(𝜌), 𝐹 ′ (𝜌) and 𝐹 ′′ (𝜌) from equation (352, 355, 356) into equation (351), we get

2 (357)
𝑠(𝑠 − 1)𝜌 𝑠−2 𝐺 + 2𝑠𝜌 𝑠−1 𝐺 ′ + 𝜌 𝑠 𝐺 ′′ + ( − 1) [𝑠𝜌 𝑠−1 𝐺 + 𝜌 𝑠 𝐺 ′ ]
𝜌
1 𝜆 𝑙(𝑙 + 1) 𝑠
+ [− + − ]𝜌 𝐺 = 0
𝜌 𝜌 𝜌2

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 261

Multiplying throughout by 4𝜌2 , we get

4𝜌2 𝑠(𝑠 − 1)𝜌 𝑠−2 𝐺 + 4𝜌2 . 2𝑠𝜌 𝑠−1 𝐺 ′ + 4𝜌2 . 𝜌 𝑠 𝐺 ′′ + (8𝜌 − 4𝜌2 )[𝑠𝜌 𝑠−1 𝐺 + 𝜌 𝑠 𝐺 ′ ] (358)
+ [−4𝜌 + 4𝜌𝜆 − 4𝑙(𝑙 + 1)]𝜌 𝑠 𝐺 = 0

or

4𝑠(𝑠 − 1)𝜌 𝑠 𝐺 + 8𝑠𝜌 𝑠+1 𝐺 ′ + 4𝜌 𝑠+2 𝐺 ′′ + 8𝑠𝜌 𝑠 𝐺 − 4𝑠𝜌 𝑠+1 𝐺 + 8𝜌 𝑠+1 𝐺 ′ (359)
− 4𝜌 𝑠+2 𝐺 ′ − 4𝜌 𝑠+1 𝐺 + 4𝜆𝜌 𝑠+1 𝐺 − 4𝑙(𝑙 + 1)𝜌 𝑠 𝐺 = 0

or

4𝑠(𝑠 − 1)𝜌 𝑠 𝐺 + 8𝑠𝜌 𝑠 𝐺 − 4𝑠𝜌 𝑠+1 𝐺 − 4𝜌 𝑠+1 𝐺 + 4𝜆𝜌 𝑠+1 𝐺 − 4𝑙(𝑙 + 1)𝜌 𝑠 𝐺 (360)
+ 8𝑠𝜌 𝑠+1 𝐺 ′ + 8𝜌 𝑠+1 𝐺 ′ − 4𝜌 𝑠+2 𝐺 ′ + 4𝜌 𝑠+2 𝐺 ′′ = 0

or

[4𝑠(𝑠 − 1)𝜌 𝑠 + 8𝑠𝜌 𝑠 − 4𝑠𝜌 𝑠+1 − 4𝜌 𝑠+1 + 4𝜆𝜌 𝑠+1 − 4𝑙(𝑙 + 1)𝜌 𝑠 ]𝐺 (361)
+ [8𝑠𝜌 𝑠+1 + 8𝜌 𝑠+1 − 4𝜌 𝑠+2 ]𝐺 ′ + 4𝜌 𝑠+2 𝐺 ′′ = 0

Dividing throughout by 𝜌 𝑠 , we get

[4𝑠(𝑠 − 1) + 8𝑠 − 4𝑠𝜌 − 4𝜌 + 4𝜆𝜌 − 4𝑙(𝑙 + 1)]𝐺 + [8𝑠𝜌 + 8𝜌 − 4𝜌2 ]𝐺 ′ (362)


2 ′′
+ 4𝜌 𝐺 = 0

If 𝜌 = 0, the function 𝐺(𝜌) = 𝑎0 and the above equation takes the form

[4𝑠(𝑠 − 1) + 8𝑠 − 4𝑙(𝑙 + 1)]𝑎0 = 0 (363)

Since 𝑎0 ≠ 0, the quantity that must be equal to zero to satisfy the above result is

4𝑠(𝑠 − 1) + 8𝑠 − 4𝑙(𝑙 + 1) = 0 (364)

𝑠(𝑠 − 1) + 2𝑠 − 𝑙(𝑙 + 1) = 0 (365)

𝑠(𝑠 + 1) − 𝑙(𝑙 + 1) = 0 (366)

𝑠(𝑠 + 1) = 𝑙(𝑙 + 1)

Which implies that

𝑠=𝑙 𝑜𝑟 𝑠 = −(𝑙 + 1) (367)

Now, if we put 𝑠 = −(𝑙 + 1) the first term in the function F(ρ) becomes 𝑎0 /0𝑙+1 at 𝜌 = 0 which infinite, and
hence is not an acceptable solution. Thus, the only we are left with is 𝑠 = 𝑙; after using the same in equation
(362), we get

Copyright © Mandeep Dalal


262 A Textbook of Physical Chemistry – Volume I

[4𝑙(𝑙 − 1) + 8𝑙 − 4𝑙𝜌 − 4𝜌 + 4𝜆𝜌 − 4𝑙(𝑙 + 1)]𝐺 + [8𝑙𝜌 + 8𝜌 − 4𝜌2 ]𝐺 ′ + 4𝜌2 𝐺 ′′ (368)


=0

[−4𝑙𝜌 − 4𝜌 + 4𝜆𝜌]𝐺 + [8𝑙𝜌 + 8𝜌 − 4𝜌2 ]𝐺 ′ + 4𝜌2 𝐺 ′′ = 0 (369)

Dividing the above equation by 4𝜌, we get

[−𝑙 − 1 + 𝜆]𝐺 + [2𝑙 + 2 − 𝜌]𝐺 ′ + 𝜌𝐺 ′′ = 0 (370)

Now differentiating equation (353) at first and second order, we get


𝑘=∞ (371)
′ (𝜌) 1−1 2−1 3−1
𝐺 = 𝑎1 1𝜌 + 𝑎2 2𝜌 + 𝑎3 3𝜌 … = ∑ 𝑎𝑘 . 𝑘. 𝜌𝑘−1
𝑘=0

Similarly
𝑘=∞ (372)
′′ (𝜌) 2−2 3−2
𝐺 = 𝑎2 . 2. (2 − 1)𝜌 + 𝑎3 . 3. (3 − 1)𝜌 … = ∑ 𝑎𝑘 . 𝑘. (𝑘 − 1). 𝜌𝑘−2
𝑘=0

After using the values equation (354, 371, 372) into equation (370), we get
𝑘=∞ 𝑘=∞ 𝑘=∞ (373)
𝑘 𝑘−1
[−𝑙 − 1 + 𝜆] ∑ 𝑎𝑘 𝜌 + [2𝑙 + 2 − 𝜌] ∑ 𝑎𝑘 . 𝑘. 𝜌 + 𝜌 ∑ 𝑎𝑘 . 𝑘. (𝑘 − 1). 𝜌𝑘−2
𝑘=0 𝑘=0 𝑘=0
=0

The above equation holds true only if the coefficients of individual powers of ρ become zero. So, simplifying
equation (373) for two summation terms (ak and ak+1), we have

[−𝑙 − 1 + 𝜆][𝑎𝑘 𝜌𝑘 + 𝑎𝑘+1 𝜌𝑘+1 ] + [2𝑙 + 2 − 𝜌][𝑎𝑘 . 𝑘. 𝜌𝑘−1 + 𝑎𝑘+1 . (𝑘 + 1). 𝜌𝑘 ] (374)
𝑘−2 𝑘−1 ]
+ 𝜌[𝑎𝑘 . 𝑘. (𝑘 − 1). 𝜌 + 𝑎𝑘+1 . (𝑘 + 1). 𝑘. 𝜌 =0

−𝑙𝑎𝑘 𝜌𝑘 − 𝑎𝑘 𝜌𝑘 + 𝜆𝑎𝑘 𝜌𝑘 − 𝑙𝑎𝑘+1 𝜌𝑘+1 − 𝑎𝑘+1 𝜌𝑘+1 + 𝜆𝑎𝑘+1 𝜌𝑘+1 + 2𝑙𝑎𝑘 . 𝑘. 𝜌𝑘−1 (375)
+ 2𝑎𝑘 . 𝑘. 𝜌𝑘−1 − 𝜌𝑎𝑘 . 𝑘. 𝜌𝑘−1 + 2𝑙𝑎𝑘+1 . (𝑘 + 1). 𝜌𝑘 + 2𝑎𝑘+1 . (𝑘
+ 1). 𝜌𝑘 − 𝜌𝑎𝑘+1 . (𝑘 + 1). 𝜌𝑘 + 𝜌. 𝑎𝑘 . 𝑘. (𝑘 − 1). 𝜌𝑘−2
+ 𝜌. 𝑎𝑘+1 . (𝑘 + 1). 𝑘. 𝜌𝑘−1 = 0

Now putting a coefficient of 𝜌𝑘 equal to zero, we get

−𝑙𝑎𝑘 𝜌𝑘 − 𝑎𝑘 𝜌𝑘 + 𝜆𝑎𝑘 𝜌𝑘 − 𝑎𝑘 . 𝑘. 𝜌𝑘 + 2𝑙𝑎𝑘+1 . (𝑘 + 1). 𝜌𝑘 + 2𝑎𝑘+1 . (𝑘 + 1). 𝜌𝑘 (376)


𝑘
+ 𝑎𝑘+1 . (𝑘 + 1). 𝑘. 𝜌 = 0

−𝑙𝑎𝑘 − 𝑎𝑘 + 𝜆𝑎𝑘 − 𝑎𝑘 𝑘 + 2𝑙𝑎𝑘+1 (𝑘 + 1) + 2𝑎𝑘+1 (𝑘 + 1) + 𝑎𝑘+1 (𝑘 + 1)𝑘 = 0 (377)

or

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 263

[−𝑙 − 1 + 𝜆 − 𝑘]𝑎𝑘 + [2𝑙(𝑘 + 1) + 2(𝑘 + 1) + (𝑘 + 1)𝑘]𝑎𝑘+1 = 0 (378)

[2𝑙(𝑘 + 1) + 2(𝑘 + 1) + (𝑘 + 1)𝑘]𝑎𝑘+1 = −[−𝑙 − 1 + 𝜆 − 𝑘]𝑎𝑘 (379)

or

𝑙+1−𝜆+𝑘 (380)
𝑎𝑘+1 = 𝑎
2𝑙(𝑘 + 1) + 2(𝑘 + 1) + (𝑘 + 1)𝑘 𝑘

or

𝑙+1−𝜆+𝑘 (381)
𝑎𝑘+1 = 𝑎
(𝑘 + 1)(2𝑙 + 𝑘 + 2) 𝑘

The equation (384) is the recursion formula where k is an integer. This expression allows one to determine the
coefficient 𝑎𝑘+1 in terms of 𝑎𝑘 which is arbitrary.
Now, since the series 𝐺(𝜌) consists of the infinite number of terms, the function 𝐹(𝜌) becomes
infinite at a very large value of k i.e. infinite. Consequently, the function 𝑅(𝜌) will also become infinite if the
number of terms is not limited to a finite value. Therefore, we must break off the series to a finite number of
terms which is possible only if the numerator becomes zero i.e.

𝑙+1−𝜆+𝑘 =0 (382)

Define a new quantum number “n” at this stage as

𝜆 =𝑙+1+𝑘 =𝑛 (383)

Since l and k are integers, n can be 1, 2, 3. 4 …. and so on. Moreover, as 𝑛 ≥ 𝑙 + 1, the largest value that l can
have is 𝑛 − 1. Hence, the value of l has a domain ranging from 0 to 𝑛 − 1. Putting 𝜆 = 𝑛 in equation (370)

[−𝑙 − 1 + 𝑛]𝐺 + [2𝑙 + 2 − 𝜌]𝐺 ′ + 𝜌𝐺 ′′ = 0 (384)

Defining 2𝑙 + 1 = 𝑝 and 𝑛 + 𝑙 = 𝑞, we get

[𝑞 − 𝑝]𝐺 + [𝑝 + 1 − 𝜌]𝐺 ′ + 𝜌𝐺 ′′ = 0 (385)

The solution of the equation given above is the “associated Laguerre polynomial” multiplied by a constant
factor i.e.
𝑝
𝐺(𝑝) = 𝐶𝐿𝑞 (𝜌) = 𝐶𝐿2𝑙+1
𝑛+𝑙 (𝜌) (385)

The constant C can be set as normalization constant and “associated Laguerre polynomial” is
𝑘=𝑛−𝑙−1
(−1)𝑘+1 [(𝑛 + 𝑙)!]2 𝜌𝑘 (385)
𝐿2𝑙+1
𝑛+𝑙 (𝜌) = ∑
(𝑛 − 𝑙 − 1 − 𝑘)! (2𝑙 + 1 + 𝑘)! 𝑘!
𝑘=0

Copyright © Mandeep Dalal


264 A Textbook of Physical Chemistry – Volume I

After using the value of 𝐹(𝜌) from equation (352) in equation (349), we get radial wavefunction as

𝑅(𝜌) = 𝜌 𝑠 𝐺(𝜌) 𝑒 −𝜌/2 (386)

Since 𝑠 = 𝑙 and also using 𝐺(𝑝) from equation (385), the above equation takes the form

𝑅𝑛,𝑙 (𝜌) = 𝐶 𝑒 −𝜌/2 𝜌𝑙 𝐿2𝑙+1


𝑛+𝑙 (𝜌) (387)

Now, after using the value of 𝐿2𝑙+1


𝑛+𝑙 (𝜌) from equation (385) in equation (387), we get

𝑘=𝑛−𝑙−1
(−1)𝑘+1 [(𝑛 + 𝑙)!]2 𝜌𝑘 (388)
−𝜌/2 𝑙
𝑅𝑛,𝑙 (𝜌) = 𝐶 𝑒 𝜌 ∑
(𝑛 − 𝑙 − 1 − 𝑘)! (2𝑙 + 1 + 𝑘)! 𝑘!
𝑘=0

i) The normalization constant for function R(r): In order to determine the normalization constant for the R
function, we must put the squared-integral over whole configuration space as unity i.e.

(389)
2 (𝑟). 2
∫ 𝑅𝑛,𝑙 𝑟 . 𝑑𝑟 = 1
0

The factor r2 is introduced to convert the length dr into a volume around the center of the nucleus. At this
point, recall the value of ρ again but in terms of equation (338, 383) i.e.

2𝜇𝑍𝑒 2 𝑟 2𝜇𝑍𝑒 2 𝑟 2𝑍𝑟 𝜇𝑒 2 (390)


𝜌 = 2𝛼𝑟 = = = .
ħ2 𝜆 ħ2 𝑛 𝑛 ħ2
Since 𝑎0 = ħ2 /𝜇𝑒 2 i.e. the “Bohr radius”, the equation (390) takes the form

2𝑍𝑟 1 (391)
𝜌= .
𝑛 𝑎0

So that
𝑛𝑎0 (392)
𝑟= 𝜌
2𝑍
Also
𝑛𝑎0 (393)
𝑑𝑟 = 𝑑𝜌
2𝑍
After using the values of 𝑅𝑛,𝑙 (𝜌), r and dr from equation (388, 392, 393) in equation (389), we get

𝑛𝑎0 2 𝑛𝑎0 (394)
2
𝐶 ∫ 𝑒 𝜌 . 𝜌2𝑙 . [𝐿2𝑙+1
2
𝑛+𝑙 (𝜌)] . [ 𝜌] . [ ] 𝑑𝜌 = 1
2𝑍 2𝑍
0

or

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 265

2
𝑛𝑎0 3 2𝑛{(𝑛 + 𝑙)!}3 (395)
𝐶 ( ) [ ]=1
2𝑍 (𝑛 − 𝑙 − 1)!

or

(395)
2𝑍 3 (𝑛 − 𝑙 − 1)!
𝐶 = √( ) [ ]
𝑛𝑎0 2𝑛{(𝑛 + 𝑙)!}3

After using the value of normalization constant from above equation into equation (388), we get

𝑘=𝑛−𝑙−1 (396)
2𝑍 3 (𝑛 − 𝑙 − 1)! (−1)𝑘+1 [(𝑛 + 𝑙)!]2 𝜌𝑘
𝑅𝑛,𝑙 (𝜌) = √( ) [ ] 𝑒 −𝜌/2 𝜌𝑙 ∑
𝑛𝑎0 2𝑛{(𝑛 + 𝑙)!}3 (𝑛 − 𝑙 − 1 − 𝑘)! (2𝑙 + 1 + 𝑘)! 𝑘!
𝑘=0

2𝑍𝑟 𝑘 (397)
𝑙 𝑘=𝑛−𝑙−1 (−1)𝑘+1 [(𝑛 + 𝑙)!]2 (
𝑛𝑎0 )
3
2𝑍 (𝑛 − 𝑙 − 1)! 𝑍𝑟 2𝑍𝑟
= √( ) [ 3
] . exp (− ).( ) . ∑
𝑛𝑎0 2𝑛{(𝑛 + 𝑙)!} 𝑛𝑎0 𝑛𝑎0 (𝑛 − 𝑙 − 1 − 𝑘)! (2𝑙 + 1 + 𝑘)! 𝑘!
𝑘=0

Which is the complete solution of R-equation.

Table 3. Some of the initial radial wave functions in terms of distance from the center of the nucleus for
the hydrogen atom and other hydrogen-like species.

n l Radial wave function (𝑅𝑛,𝑙 )

1 0 𝑍 3/2 −𝑍𝑟/𝑎
𝑅1,0 = 2( ) 𝑒 0
𝑎0

2 0 1 𝑍 3/2 𝑍𝑟
𝑅2,0 = ) (2 − ) 𝑒 −𝑍𝑟/2𝑎0
(
2√2 𝑎0 𝑎0

2 1 𝑍 3/2 𝑍𝑟 −𝑍𝑟/2𝑎
1
𝑅2,1 = (
) ( )𝑒 0
2√6 𝑎0 𝑎0

3 0 2 𝑍 3/2 𝑍𝑟 𝑍𝑟 2
𝑅3,0 = ( ) (27 − 18 − 2 ( ) ) 𝑒 −𝑍𝑟/3𝑎0
81√3 𝑎0 𝑎0 𝑎0

3 1 4 𝑍 3/2 𝑍𝑟 𝑍𝑟 2
𝑅3,1 = ( ) (6 ( ) − ( ) ) 𝑒 −𝑍𝑟/3𝑎0
81√6 𝑎0 𝑎0 𝑎0

3 2 1 𝑍 3/2 𝑍𝑟 3/2 𝑍𝑟 2
𝑅3,2 = ) ( ) − ( ) 𝑒 −𝑍𝑟/3𝑎0
(
81√30 𝑎0 𝑎0 𝑎0

Copyright © Mandeep Dalal


266 A Textbook of Physical Chemistry – Volume I

The total wavefunction: After solving the ϕ-, θ- and r-dependent equations, we have Φm(ϕ), Θl,m(θ) and Rn,l(r)
functions. Now, recall the total wave function that depends upon all the three variable i.e.

𝜓𝑛,𝑙,𝑚 (𝑟, 𝜃, 𝜙) = 𝜓𝑛,𝑙 (𝑟) × 𝜓𝑙,𝑚 (𝜃) × 𝜓𝑚 (𝜙) (398)

After putting the values of Φm(ϕ), Θl,m(θ) and Rn,l(r) from equation (397) in equation (398), we get

𝜓𝑛,𝑙,𝑚 (𝑟, 𝜃, 𝜙) = 𝑅𝑛,𝑙 . 𝛩𝑙,𝑚 . 𝛷𝑚 (399)

2𝑍𝑟 𝑘 (400)
𝑙 𝑘=𝑛−𝑙−1 (−1)𝑘+1 [(𝑛 + 𝑙)!]2 (
𝑛𝑎0 )
3
2𝑍 (𝑛 − 𝑙 − 1)! 𝑍𝑟 2𝑍𝑟
= √( ) [ 3
] . exp (− ).( ) . ∑
𝑛𝑎0 2𝑛{(𝑛 + 𝑙)!} 𝑛𝑎0 𝑛𝑎0 (𝑛 − 𝑙 − 1 − 𝑘)! (2𝑙 + 1 + 𝑘)! 𝑘!
𝑘=0

(2𝑙 + 1)(𝑙 − 𝑚)! 𝑚 1


×√ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃) × √ 𝑒 𝑖𝑚𝜙
2(𝑙 + 𝑚)! 2𝜋

Which is the complete expression for all the quantum mechanical states of a single electron around the nucleus.

Table 4. Some of the initial total wave functions for the hydrogen atom and other hydrogen-like species.

n l m Total wave function (𝜓𝑛,𝑙,𝑚 )

1 0 0 1 𝑍 3/2 −𝑍𝑟/𝑎
𝜓1,0,0 = ( ) 𝑒 0
√𝜋 𝑎0
2 0 0 1 𝑍 3/2 𝑍𝑟
𝜓2,0,0 = ( ) (2 − ) 𝑒 −𝑍𝑟/2𝑎0
4√𝜋 𝑎0 𝑎0

2 1 0 1 𝑍 5/2
𝜓2,1,0 = ( ) 𝑒 −𝑍𝑟/2𝑎0 𝑟 𝐶𝑜𝑠𝜃
4√𝜋 𝑎0

2 1 ±1 1 𝑍 5/2 −𝑍𝑟/2𝑎
𝜓2,1,±1 = ( ) 𝑒 0 𝑆𝑖𝑛𝜃 𝑒 −𝑖𝜙
4√𝜋 𝑎0

3 0 0 2𝑍 3/2 𝑍𝑟 𝑍𝑟 2 1
𝜓3,0,0 = ( ) (27 − 18 − 2 ( ) ) 𝑒 −𝑍𝑟/3𝑎0 .
81√3 𝑎0 𝑎0 𝑎0 √4𝜋

3 1 0
4𝑍 3/2 𝑍𝑟 𝑍𝑟 2 3 1
𝜓3,1,0 = ( ) (6 ( ) − ( ) ) 𝑒 −𝑍𝑟/3𝑎0 . √ 𝐶𝑜𝑠 𝜃.
81√6 𝑎0 𝑎0 𝑎0 2 √2𝜋

3 1 ±1
4 𝑍 3/2 𝑍𝑟 𝑍𝑟 2 3 1
𝜓3,1,±1 = ( ) (6 ( ) − ( ) ) 𝑒 −𝑍𝑟/3𝑎0 . √ 𝑆𝑖𝑛 𝜃. √ 𝑒 ±𝑖𝜙
81√6 𝑎0 𝑎0 𝑎0 2 2𝜋

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 267

The eigenvalues of energy: Since the series 𝐺(𝜌) consists of infinite number of terms, the function 𝐹(𝜌)
becomes infinite at a very large value of k i.e. infinite. Consequently, the function 𝑅(𝜌) will also become
infinite if the number of terms are not limited to a finite value. Therefore, we must break off the series to a
finite number of terms which is possible only if the numerator in equation (381) becomes zero i.e.

𝑙+1−𝜆+𝑘 =0 (401)

or

𝜆 =𝑙+1+𝑘 =𝑛 (402)

Where n is the principal quantum number and can have values 1, 2, 3. 4 …. because l and k are integers always.
Now recall the value of λ from equation (338) and then squaring both sides, we get

𝜇2 𝑍 2 𝑒 4 (403)
𝜆2 =
ħ4 𝛼 2
Also putting the value of 𝛼 2 from equation (338) in equation (403), we get

𝜇2 𝑍 2 𝑒 4 𝜇 2 𝑍 2 𝑒 4 ħ2 𝜇𝑍 2 𝑒 4 (404)
𝜆2 = = − . = −
ħ4 𝛼 2 ħ4 2𝜇𝐸 2 𝐸 ħ2

𝜇𝑍 2 𝑒 4 𝜇𝑍 2 𝑒 4 (405)
𝐸𝑛 = − == −
2 𝜆2 ħ2 2 𝑛 2 ħ2
Which is the same as given by the pre-wave-mechanical quantum theory.

Figure 14. The energy level for various quantum mechanical states of the hydrogen atom.

It is also worthy to note that the total number of wave functions that can be written for a given value of n are
n2, and therefore, we can say that the degeneracy of any energy level is also n2.

Copyright © Mandeep Dalal


268 A Textbook of Physical Chemistry – Volume I

 Principal, Azimuthal and Magnetic Quantum Numbers and the Magnitude


of Their Values
The Schrodinger wave equation for hydrogen and hydrogen-like species in the polar coordinates can
be written as:

1 𝜕 2 𝜕𝜓 1 𝜕 𝜕𝜓 1 𝜕2𝜓 8𝜋 2 𝜇 𝑍𝑒 2 (406)
[ (𝑟 ) + (𝑆𝑖𝑛𝜃 ) + ] + (𝐸 + )𝜓 = 0
𝑟 2 𝜕𝑟 𝜕𝑟 𝑆𝑖𝑛𝜃 𝜕𝜃 𝜕𝜃 𝑆𝑖𝑛2 𝜃 𝜕𝜙 2 ℎ2 𝑟

After separating the variables present in the equation given above, the solution of the differential equation was
found to be

𝜓𝑛,𝑙,𝑚 (𝑟, 𝜃, 𝜙) = 𝑅𝑛,𝑙 . 𝛩𝑙,𝑚 . 𝛷𝑚 (407)

2𝑍𝑟 𝑘 (408)
2𝑍 3
(𝑛 − 𝑙 − 1)! 𝑍𝑟 2𝑍𝑟 𝑙 𝑘=𝑛−𝑙−1 (−1)𝑘+1 [(𝑛 + 𝑙)!]2 ( )
𝑛𝑎0
= √( ) [ 3
] . exp (− ).( ) . ∑
𝑛𝑎0 2𝑛{(𝑛 + 𝑙)!} 𝑛𝑎0 𝑛𝑎0 (𝑛 − 𝑙 − 1 − 𝑘)! (2𝑙 + 1 + 𝑘)! 𝑘!
𝑘=0

(2𝑙 + 1)(𝑙 − 𝑚)! 𝑚 1


×√ . 𝑃𝑙 (𝐶𝑜𝑠 𝜃) × √ 𝑒 𝑖𝑚𝜙
2(𝑙 + 𝑚)! 2𝜋

It is obvious that the solution of equation (406) contains three discrete (n, l, m) and three continuous (r, θ, ϕ)
variables. In order to be a well-behaved function, there are some conditions over the values of discrete variables
that must be followed i.e. boundary conditions. Therefore, we can conclude that principal (n), azimuthal (l)
and magnetic (m) quantum numbers are obtained as a solution of the Schrodinger wave equation for hydrogen
atom; and these quantum numbers are used to define various quantum mechanical states. In this section, we
will discuss the properties and significance of all these three quantum numbers one by one.
 Principal Quantum Number
The principal quantum number is denoted by the symbol n; and can have value 1, 2, 3, 4, 5…..∞. The
label “principal” is allotted because valid values of l and m can be defined only after defining an acceptable
value of n. Some of the most important significances of the principal quantum number are given below.
1. The energy of an electron in hydrogen-like systems: The principal quantum number gives the energy of
the electron in all hydrogen and hydrogen-like species by the following relation.

𝜇𝑍 2 𝑒 4 (409)
𝐸𝑛 = −
2 𝑛 2 ħ2
Where μ is the reduced mass of the system while e represents the electronic charge. The symbol Z represents
the nuclear charge of the one-electron system. Now since main shells are nothing but the classification of
different quantum mechanical states of electron on the basis of energy only, we can also say that n tells about
the main shells in the modern wave mechanical model of the atom.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 269

2. Degeneracy in hydrogen-like systems: Since the total number of wave functions that can be written for a
given value of n are n2, we can say that the degeneracy of any energy level is also n2. In other words, we can
say that because the energy depends only upon the value of n, all wave functions with the same value of n must
possess the same energy. For instance, if we 𝑛 = 2, a total of four wave-functions can be written i.e.
𝜓2,0,0 , 𝜓2,1,0 , 𝜓2,1,+1 and 𝜓2,1,−1 . Owing to the same value of n, all of these states are bound to have the same
energy, and thus, are degenerate.
3. The maximum number of electrons per unit cell: Since two electrons can have the same set of principal,
orbital and magnetic quantum numbers via opposite spins, the maximum number of electrons per unit cell will
be 2n2 i.e. the double of the degeneracy. For instance, if we 𝑛 = 2, the maximum number of electrons that can
be filled in the second main shell is 2 × 22 = 8. Similarly, if we 𝑛 = 3, the maximum number of electrons
that can be filled in the third main shell is 2 × 32 = 18.
4. Spectra of elemental hydrogen: In order to understand this concept, recall the energy expression for the
hydrogen atom i.e.

𝑚 𝑒4 4𝜋 2 𝑚 𝑒 4 (410)
𝐸𝑛 = − = −
2 𝑛2 ħ2 2 𝑛2 ℎ2
In the SI system, the above equation needs to be corrected for permittivity factor (4πεo) i.e.

4𝜋 2 𝑚 𝑒 4 4𝜋 2 𝑚 𝑒 4 (411)
𝐸𝑛 = − 𝐽𝑜𝑢𝑙𝑒𝑠 = − 𝐽𝑜𝑢𝑙𝑒𝑠
2 𝑛2 ℎ2 (4𝜋𝜀𝑜 )2 32 𝜋 2 𝑛2 ℎ2 𝜀𝑜2

or

𝑚 𝑒4 (412)
𝐸𝑛 = − 𝐽𝑜𝑢𝑙𝑒𝑠
8 𝑛2 ℎ2 𝜀𝑜2

Converting Joules into cm−1 (dividing by hc) the above equation takes the form

𝑚 𝑒4 𝑅 (413)
𝜈̅𝑛 = − 2 3 2 𝑐𝑚−1 = − 2 𝑐𝑚−1
8 𝑛 ℎ 𝑐 𝜀𝑜 𝑛

Now the selection rules for electronic transitions are

𝛥𝑛 = 𝑎𝑛𝑦𝑡ℎ𝑖𝑛𝑔 𝑎𝑛𝑑 𝛥𝑙 = ±1 (414)

This means that electron can move from s to p-orbital only; and all s–s, p–p, d–d and f–f transitions are Laporte
forbidden. Now assume that electron shows a transition from an initial quantum mechanical state (n1) to the
final quantum mechanical state (n2). The energy of absorption can be formulated as:

𝑅 𝑅 −1
(415)
𝛥𝜈̅ = 𝜈̅𝑛2 − 𝜈̅𝑛1 = (− 2 ) − (− 2 ) 𝑐𝑚
𝑛2 𝑛1

or

Copyright © Mandeep Dalal


270 A Textbook of Physical Chemistry – Volume I

1 1 −1
(416)
𝛥𝜈̅ = 𝑅 ( 2 − 2 ) 𝑐𝑚
𝑛1 𝑛2

The above equation can also be used to determine the wavenumber of the emission spectral line of the hydrogen
atom. All the possibilities and corresponding series are given below.

Figure 15. Energy level diagram of the hydrogen atom in the units of Rydberg constant.

At this stage, we must consider all the possibilities that may arise from the transitioning of the electron from
different quantum mechanical states. These transitions are grouped in various series labeled as Lyman, Balmer,
Paschen, Brackett and Pfund series.

Table 5. Different spectral series in the hydrogen atom.

Series name Lower state (n1) Higher state (n2) Region

Lyman 1 2, 3, 4, 5, …. ∞ UV

Balmer 2 3, 4, 5, 6, …. ∞ Visible

Paschen 3 4, 5, 6, 7, …. ∞ Near IR

Brackett 4 5, 6, 7, 8, …. ∞ Mid IR

Pfund 5 6, 7, 8, 9, …. ∞ Far IR

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 271

 Azimuthal Quantum Number


The azimuthal quantum number is denoted by the symbol l; and can have value n–1, n–2, n–3…..0.
The label azimuthal quantum number is also called as “angular momentum quantum number” because the
values of l also govern the orbital angular momentum of the electron in a particular quantum mechanical state.
Some of the most important significances of the principal quantum number are given below.
1. Orbital angular momentum of the electron: The azimuthal quantum number gives the angular momentum
of the electron in all hydrogen and hydrogen-like species by the following relation.

(417)
< 𝐿 >= ∮ 𝜓𝑛,𝑙,𝑚 (𝑟, 𝜃, 𝜙) 𝐿̂ 𝜓𝑛,𝑙,𝑚 (𝑟, 𝜃, 𝜙)

ℎ (418)
= √𝑙(𝑙 + 1)
2𝜋
After looking at the equation (418), it is obvious that it’s only the ‘l’ quantum number that controls the
magnitude of the orbital angular momentum quantum number. Furthermore, owing to the quantized nature of
‘l’ quantum number, the angular momentum of an electron in an atom is also quantized. For instance, if we
use l = 0, 1, 2, 3 in equation (418), we will get 0, √2, √6 and √12 units of angular momentum, respectively.
2. Subshells in the main shell: The azimuthal quantum number can also be used to classify different quantum
mechanical states on the basis of orbital angular momentum. In other words, subshells are nothing but the
classification degenerate quantum mechanical states on the basis of angular momentum.

Figure 16. Energy level diagram of the hydrogen atom with further classification.

Copyright © Mandeep Dalal


272 A Textbook of Physical Chemistry – Volume I

3. Shape and number of angular nodes in atomic orbital: The angular momentum quantum number, l, also
controls the number of angular nodes that pass through the nucleus. An angular node (planar or conical) is
observed when the angular part of the wave function passes through zero and changes sign.

Figure 17. Different orbitals and the corresponding angular nodes.

4. The energy of different subshells in multi-electron atoms: In hydrogen and H-like atoms (i.e. one-
electron systems), the energy levels depend only upon the principal quantum number. However, these energy
levels also split according to the magnitude of l as well. Quantum states of higher l are placed above than the
states with lower l. For instance, the energy of 2s orbital is lower than 2p, 3d exists at higher position than 3p.

Figure 18. The energy pattern of different subshells in multi-electron atoms.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 273

 Magnetic Quantum Number


The magnetic quantum number is denoted by the symbol m, and can have values +l to –l in unit steps.
In other words, the quantum number m is nothing but the allowed effects of orbital angular momentum in the
z-direction. The label “magnetic quantum number” arises because m affects the energy of the electron in an
externally applied magnetic field. In the absence of such a field, all spherical harmonics corresponding to the
different arbitrary values of m will be equivalent. Some of the most important significances of the principal
quantum number are given below.
1. The orientation of orbital angular momentum: The azimuthal quantum number gives the angular
momentum of the electron in all hydrogen and hydrogen-like species by the following relation:

ℎ (419)
𝐿𝑙 = √𝑙(𝑙 + 1)
2𝜋
Since 𝑙 = 0, 1, 2, 3, 4 … … (𝑛 − 1) etc., the quantum mechanically allowed values of orbital angular momentum
(in the units of ℎ/2𝜋) are given below.

𝐿0 = √0(0 + 1) unit = 0 unit (420)

𝐿1 = √1(1 + 1) unit = √2 unit (421)

𝐿2 = √2(2 + 1) unit = √6 unit (422)

𝐿3 = √3(3 + 1) unit = √12 unit (423)

However, there is boundary condition in quantum mechanics that says that only integral effects are allowed
reference direction if the angular momentum is generated by integral quantum number and half-integral effects
are allowed in reference direction if the momentum is generated by half-integral quantum number.
Since, 𝐿𝑧 = 𝐿 𝐶𝑜𝑠 𝜃, √2 units of orbital angular momentum cannot orient itself along z-axis because
this makes 𝜃 = 0°, and since 𝐶𝑜𝑠 0 = 1, 𝐿⃗𝑧 = 𝐿 i.e. orbital angular momentum effect along the z-axis is also
1.414 unit which is not allowed quantum mechanically. The effects of angular momentum allowed in the z-
direction are +1, 0, −1; for which angles required are determined as follows.

1 (424)
+1 = √2 𝐶𝑜𝑠 𝜃 ⇒ 𝜃 = 𝐶𝑜𝑠 −1 = 45°
√2
0 (425)
0 = √2 𝐶𝑜𝑠 𝜃 ⇒ 𝜃 = 𝐶𝑜𝑠 −1 = 90°
√2
−1 (426)
−1 = √2 𝐶𝑜𝑠 𝜃 ⇒ 𝜃 = 𝐶𝑜𝑠 −1 = 135°
√2

Copyright © Mandeep Dalal


274 A Textbook of Physical Chemistry – Volume I

Hence, we can say that in order to be allowed, the 1.414 units of orbital angular momentum must orient itself
only at 45°, 90° and 135° in space from reference direction (z-axis in this case). Since the orientation of angular
momentum can orient itself in any direction from the z-axis as far as the effective orbital angular momentum
+1 unit along z-direction; therefore, we should use a cone around the same at 45°. The same is true for 0 and
−1 effects with 90° and 135°, respectively.

Figure 19. The space quantization of orbital angular momentum of an electron for l = 0, 1, 2 and 3 states.

It is also worthy to mention that all the orientations of orbital angular momentum are degenerate in the absence
of any externally applied magnetic field.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 275

2. The energy of different orientations: The probability of different orientations of orbital angular
momentum, and also the corresponding magnetic fields, are same. However, after applying the magnetic field
along the z-direction, the situation will not be the same.
The orientation with a maximum negative value of orbital angular momentum along the z-axis will
have magnetic dipole aligned along the applied magnetic field, and thus, will be most stable. Conversely, the
orientations with the maximum positive value of orbital angular momentum along the z-axis will have
magnetic dipole aligned opposite to the applied magnetic field, and thus, will be least stable. Likewise, the
orientations with angular momentum component in between will also have intermediary energies in this case.

Figure 20. Different orientations of the orbital angular momentum of an electron in a d-subshell and
corresponding energies in the applied magnetic field.

To understand this, consider the typical case of d-subshell. The orbital angular momentum for the
corresponding electron can be obtained using equation (419) i.e.

ℎ (427)
𝐿 = √2(2 + 1)
2𝜋
ℎ (428)
𝐿 = √6
2𝜋
The allowed effects of orbital angular momentum in the z-direction are +2, +1, 0, −1, −2 in the units of h/2π.
Therefore, each spectral line arising from the transition to the d-subshell will split in a quintet in the presence
of the externally applied magnetic field.

Copyright © Mandeep Dalal


276 A Textbook of Physical Chemistry – Volume I

 Probability Distribution Function


The probability distribution function is the behavior of ψ2 at various points around the nucleus as a
function of distance r from the nucleus. The plots of such functions are also called as the probability
distribution curves. Nevertheless, since it is only the radial part (𝑅𝑛,𝑙 ) that varies with the distance from the
nucleus, the graphs of ψ2 must behave in the same manner. To understand this more precisely, consider the
plot of the first two quantum mechanical states of an electron in a hydrogen atom.
 Probability Distribution of 𝝍𝟏,𝟎,𝟎 State (1s Orbital)
In order to understand the probability distribution function of the electron in the ground state of the
hydrogen atom, recall the mathematical expression for the same i.e.

1 1 3/2 −𝑟/𝑎 (429)


𝜓1,0,0 = ( ) 𝑒 0
√𝜋 𝑎0
Squaring both sides, we get

2
1 −2𝑟/𝑎 (430)
𝜓1,0,0 = 𝑒 0
𝜋𝑎03

It is obvious from the equation (429, 430) that the pre-exponential part is simply a constant and variation
depends only upon the exponential part.

Figure 21. The variation of electron density vs distance from the center of the nucleus in 1s orbital.

Hence, it can be concluded that the density of the electron wave is highest at the center of the nucleus and
decreases as the distance from the center of the nucleus increases, and becomes zero only at infinite.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 277

 Probability Distribution of 𝝍𝟐,𝟎,𝟎 State (2s Orbital)


In order to understand the probability distribution function of an electron in the 2s state of the
hydrogen atom, recall the mathematical expression for the same i.e.

1 1 3/2 𝑟 (431)
𝜓2,0,0 = ( ) (2 − ) 𝑒 −𝑟/2𝑎0
4√𝜋 𝑎0 𝑎0

Squaring both sides, we get

2
1 𝑟 2 −𝑟/𝑎 (432)
𝜓2,0,0 = (2 − ) 𝑒 0
16𝜋𝑎03 𝑎0

It is obvious from the equation (431, 432) that the pre-exponential part is simply a constant and variation
depends only upon the exponential part.

Figure 22. The variation of electron density vs distance from the center of the nucleus in 2s orbital.

Hence, it can be concluded that the density of electron wave is non zero at the center of the nucleus and
decreases as the distance from the center of the nucleus increases, and becomes zero at 𝑟 = 2𝑎0 . Now since
the wave function changes sign after 2𝑎0 , the density of electron wave after that increases first and then
decreases exponentially and finally becomes zero at infinite distance. Now it’s quite confusing because we
have been told that the electron cannot reside within the nucleus and the probability of finding the electron
inside the nucleus is almost zero, meaning that there must be something else that also governs the probability.

Copyright © Mandeep Dalal


278 A Textbook of Physical Chemistry – Volume I

 Radial Distribution Function


The radial distribution function is the behavior of 𝑅𝑛,𝑙
2
. 4𝜋𝑟 2 𝑑𝑟 as a function of distance r from the
center of the nucleus. These plots solve the problem posed by the simple “probability distribution curves”
which suggested that the probability of finding the electron must be highest at the center of the nucleus in the
ground electronic state. In the radial distribution plots, we assume that the probability of finding the particle at
a distance r from the nucleus depends not only upon the density of electron wave but also varies with the
magnitude of the volume of the spherical shell of dr thickness at the same distance. This is quite rational
because the r can be in any direction around the nucleus.
Consider that the space around the nucleus is divided into an infinite number of concentric shells of
thickness dr. Now though the electron density will show a decrease with increasing r, the volume of the
concentric shells will increase. More volume at distance r means more the chances of finding the electron at
same. The two effects will try to counter each other, and therefore, the resultant probability at distance r must
be the multiplication of the two effects i.e.
2
𝑅𝑎𝑑𝑖𝑎𝑙 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = 𝜓𝑛,𝑙,𝑚 × 𝑑𝑉𝑠ℎ𝑒𝑙𝑙 (433)

Nevertheless, since it is only the radial part (𝑅𝑛,𝑙 ) that varies with the distance from the nucleus, the above
expression for simplicity can be reduced to
2
𝑅𝑎𝑑𝑖𝑎𝑙 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = 𝑅𝑛,𝑙 × 𝑑𝑉𝑠ℎ𝑒𝑙𝑙 (434)

Now as we have already derived the mathematical expression of radial wavefunction hydrogen atom already
in this previously, the only thing we need is the mathematical expression of the volume element also.

Figure 23. The depiction of a concentric shell of thickness dr around the nucleus of a hydrogen atom at a
distance r.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 279

The volume of the shaded portion (spherical shell of thickness dr) can be obtained subtracting the volume of
the inner sphere from the outer sphere i.e.

4 4 (435)
𝑑𝑉 = 𝜋(𝑟 + 𝑑𝑟)3 − 𝜋𝑟 3
3 3
4 4 (436)
= 𝜋(𝑟 3 + 𝑑𝑟 3 + 3𝑟 2 𝑑𝑟 + 3𝑟 𝑑𝑟 2 ) − 𝜋𝑟 3
3 3
4 4 4 (437)
= 𝜋𝑟 3 + 𝜋𝑑𝑟 3 + 4𝜋𝑟 2 𝑑𝑟 + 4𝜋𝑟 𝑑𝑟 2 − 𝜋𝑟 3
3 3 3
4 (438)
𝑑𝑉 = 𝜋𝑑𝑟 3 + 4𝜋𝑟 2 𝑑𝑟 + 4𝜋𝑟 𝑑𝑟 2
3
Since dr is very small, the terms involving square and cube of dr can be neglected for simplicity. All this leaves
us with only one term i.e. 𝑑𝑉 = 4𝜋𝑟 2 𝑑𝑟. After using the value of dV in equation (434), we get
2
𝑅𝑎𝑑𝑖𝑎𝑙 𝑝𝑟𝑜𝑏𝑎𝑏𝑖𝑙𝑖𝑡𝑦 = 𝑅𝑛,𝑙 × 4𝜋𝑟 2 𝑑𝑟 (439)

To understand this more precisely, consider the plot for the ground quantum mechanical state of an electron in
a hydrogen atom i.e. 1s orbital.
 Radial Probability Distribution Curve for Ground State of Hydrogen Atom
The valid values of n, l and m that can be put in the general form of the hydrogenic wavefunction to
obtain ground state are 1, 0 and 0, respectively. Therefore, we can start by writing the mathematical expression
for the same i.e.

1 3/2 −𝑟/𝑎 (440)


𝑅1,0 = 2 ( ) 𝑒 0
𝑎0

The probability distribution function can be obtained by squaring equation (440) i.e.

2
4 −2𝑟/𝑎 (441)
𝑅1,0 = 𝑒 0
𝑎03

After multiplying the “probability distribution function” with “volume element”, the expression for the “radial
distribution function” can be formulated. Mathematically, we can say that

4 −2𝑟/𝑎
0 × 4𝜋𝑟 2 𝑑𝑟
(442)
𝑃(𝑟) = 𝑒
𝑎03

It is obvious from the equation (442) that probability will become zero if we put r = 0 (4𝜋𝑟 2 𝑑𝑟 = 0). Now, if
we increase the r, the radial probability will first increase due to increasing volume element, attaining maxima;
and then it will start declining due to the dominance of 𝑅1,0 2
part. In other words, the density of electron-wave
decreases exponentially but the volume of the concentric shell increases continuously.

Copyright © Mandeep Dalal


280 A Textbook of Physical Chemistry – Volume I

Figure 24. The variation of radial probability as a function of r (1s orbital).

In order to find the radius of maximum probability, we need to put 𝑑𝑃/𝑑𝑟 equal to zero. It has been found that
the radius of maximum probability will come out to be 0.53×10−10 m, which is exactly equal to the radius of
the first Bohr orbit (a0).
 Radial Probability Distribution Curves for Other Hydrogenic Wavefunctions
The other valid sets of n, l can be put in the general form of radial part of the wavefunction, to obtain
2
𝑅𝑛,𝑙 , and hence the corresponding “radial distribution functions”.

Figure 25. The variation of radial probability as a function of distance from the center of the nucleus.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 281

 Shape of Atomic Orbitals (s, p & d)


The wave mechanical model of atom says that there is a non-zero probability of finding the electron
almost everywhere in space excepting the angular and radial nodes. This means that primitive diagrams that
depict the orbital shapes are intended to describe the region encompassing 90−95% probability density. In a
typical drawing of orbital, we first plot the radial wave function and the angular part is superimposed. The
shapes of some typical orbitals are discussed below.
 Shape of s-Orbitals
In order to draw the shape of s-orbital, we first need to recall the radial part of the same and then we
will have to superimpose the angular part. For instance, the radial part of 1s orbital is

1 3/2 −𝑟/𝑎 (443)


𝑅1,0 = 2 ( ) 𝑒 0
𝑎0

It is obvious from the equation (443) that the radial part of the wave function has the largest magnitude when
r = 0, and it decreases as we move away from the nucleus. The function will become zero only at infinite
distance and will never change its sign. All this leads to a spherical-shaped cloud without any radial node. The
angular part of every s-orbital is

1 (444)
𝛶0,0 = = 0.28
√4𝜋
Hence, after multiplying radial wave function by a constant value of the angular part, the magnitude of function
at all the points in space will reduce to 28% of the initial value.

Figure 26. The shape of some lower energy s-orbitals.

Similarly, the shapes of some other s-orbitals are also given below to explain the concept more precisely. It is
worthy to mention that the plots are easy to draw if we treat radial and angular parts consequently.

Copyright © Mandeep Dalal


282 A Textbook of Physical Chemistry – Volume I

 Shape of p-Orbitals
In order to draw the shape of p-orbital, we first need to recall the radial part of the same and then we
will have to superimpose the angular part. For instance, the radial part of 2p orbital is

1 1 3/2 𝑟 (445)
𝑅2,1 = ( ) ( ) 𝑒 −𝑟/2𝑎0
2√6 𝑎0 𝑎0

It is obvious from the equation (445) that the radial part of the wave function has the zero magnitudes when r
= 0; and it increases as we move away from the nucleus, reaches a maximum, and decreases afterward. The
function becomes zero only at infinite distance and will never change its sign. All this leads to a spherical-
shaped cloud without any radial node. The angular part of p-orbitals are

(446)
3 1
𝛶1,0 = √ 𝐶𝑜𝑠 𝜃.
2 √2𝜋

(447)
3 1
𝛶1,±1 = √ 𝑆𝑖𝑛 𝜃. √ 𝑒 ±𝑖𝜙
2 2𝜋

Hence, the full plot for pz-orbital is obtained after multiplying radial wave function by angular part given by
equation (446); and the sign of function above the xy-plane will remain positive whereas a negative sign will
be obtained below xy-plane. Similarly, we can obtain the three-dimensional plots for px and py by multiplying
equation (445) by equation (447).

Figure 27. The shape of 2p-orbitals.

Similarly, the shapes of some other p-orbitals are also given below to explain the concept more precisely. It is
worthy to mention that the plots are easy to draw if we treat radial and angular parts consequently.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 283

Figure 28. The shape of some 3p, 4p and 5p orbitals.

Copyright © Mandeep Dalal


284 A Textbook of Physical Chemistry – Volume I

 Shape of d-Orbitals
In order to draw the shape of d-orbital, we first need to recall the radial part of the same and then we
will have to superimpose the angular part. For instance, the radial part of 3d orbital is

1 𝑍 3/2 𝑍𝑟 3/2 𝑍𝑟 2 (448)


𝑅3,2 = ( ) ( ) − ( ) 𝑒 −𝑍𝑟/3𝑎0
81√30 𝑎0 𝑎0 𝑎0

It is obvious from the equation (448) that the radial part of the wave function has the zero magnitudes when r
= 0; and it increases as we move away from the nucleus, reaches a maximum, and decreases afterward. The
function becomes zero only at infinite distance and will never change its sign. All this leads to a spherical-
shaped cloud without any radial node. The angular part of d-orbitals are

(449)
5 1
𝛶2,0 = √ (3𝐶𝑜𝑠 2 𝜃 − 1).
8 √2𝜋

(450)
15 1
𝛶2,±1 = √ 𝑆𝑖𝑛 𝜃 𝐶𝑜𝑠 𝜃. √ 𝑒 ±𝑖𝜙
4 2𝜋

(451)
15 1
𝛶2,±2 = √ 𝑆𝑖𝑛2 𝜃. √ 𝑒 ±𝑖2𝜙
16 2𝜋

Hence, the full plot for dz2-orbital is obtained after multiplying radial wave function by angular part given by
equation (449); and the sign of function in the xy-plane will become negative whereas a positive sign will be
obtained in two opposite lobes along z-axis. It is also worthy to note that the function becomes completely zero
in two conical surfaces (above and below) before changing its sign. Similarly, we can obtain the three-
dimensional plots for px and py by multiplying equation (448) by equation (450, 451).

Figure 29. Continued on the next page…

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 285

Figure 29. The shape of 3d-orbitals.

Similarly, the shapes of some other d-orbitals are also given below to explain the concept more precisely. It is
worthy to mention that the plots are easy to draw if we treat radial and angular parts consequently.

Figure 30. The shape of 4d orbitals.

Copyright © Mandeep Dalal


286 A Textbook of Physical Chemistry – Volume I

It is also worthy to mention that the radii of maximum probability for 4d orbitals are larger than that
of 3d orbitals. The same is true for s and p orbitals i.e. radius of maximum probability of s and p orbitals follow
the order 3s > 2s > 1s and 4p > 3p > 2p, respectively.

Copyright © Mandeep Dalal


CHAPTER 5 Quantum Mechanics – II 287

 Problems
Q 1. Derive and discuss the Schrodinger wave equation for a particle of mass m trapped inside a cubical box
with side a. Provided that the potential inside the box is zero while outside the box is infinite.
Q 2. Discuss the concept of the degeneracy of quantum mechanical states in a 3D box.
Q 3. What is the zero-point energy of a simple harmonic oscillator? How does it vary with force constant?
Q 4. Derive and discuss the Schrodinger wave equation for a diatomic rigid rotator. Also, draw the energy
level diagram for the same.
Q 5. Define space quantization with special reference to diatomic rigid rotator.
Q 6. What is the difference between the radial and angular wave function for hydrogen atom? Write down both
parts for 3dz2 orbital.
Q 7. What are quantum numbers in the modern wave mechanical model of the atom? Also, discuss the main
significance of the principal quantum number.
Q 8. Write down a short note on “probability distribution functions”.
Q 9. What are “radial distribution functions”? Also, explain how you would determine the radius of maximum
probability for 1s orbital of the hydrogen atom.
Q 10. What are the formulas to find the number of angular and radial nodes?
Q 11. Draw and discuss the shape of 4dxy and 3dz2 in detail

Copyright © Mandeep Dalal


288 A Textbook of Physical Chemistry – Volume I

 Bibliography
[1] B. R. Puri, L. R. Sharma, M.S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2018.
[2] M. Reed, B. Simon Functional Analysis, Elsevier, Amsterdam, Netherlands, 2003.
[3] G. E. Bowman, Essential Quantum Mechanics, Oxford University Press, Oxford, UK, 2008.
[4] W. Heisenberg, Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik,
Zeitschrift für Physik, 1927, 172.
[5] E. Schrödinger, Die gegenwärtige Situation in der Quantenmechanik, Naturwissenschaften. 1935, 807.
[6] P. Alberto, C. Fiolhaisdag, V. Gil, Relativistic particle in a box, European Journal of Physics, 1996, 17.
[7] R. K. Prasad, Quantum Chemistry, New Age International Publishers, New Delhi, India, 2010.
[8] I. N. Levine, Quantum Chemistry, Pearson Prentice Hall, New Jersey, USA, 2009.
[9] D. A. McQuarrie, Quantum Chemistry, University Science Books, California, USA, 2008.
[10] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[11] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.

Copyright © Mandeep Dalal


CHAPTER 6
Thermodynamics – II
 Clausius-Clapeyron Equation
The Clausius-Clapeyron equation was initially proposed by a German physics Rudolf Clausius in
1834 and then further developed by French physicist Benoît Clapeyron in 1850. This equation is extremely
useful in characterizing a discontinuous phase transition between two phases of a single constituent.
 Derivation of Clausius-Clapeyron Equation
In order to derive the Clausius-Clapeyron equation, consider a system at equilibrium i.e. the free
energy change for the ongoing process is zero (𝛥𝐺 = 0). However, we know from the principles of
thermodynamics that the variation of free energy with temperature and pressure can be formulated by the
following differential equation.

𝑑𝐺 = 𝑉𝑑𝑃 − 𝑆𝑑𝑇 (1)

For typical single-constituent equilibria, we have

𝑃ℎ𝑎𝑠𝑒 𝐼 ⇌ 𝑃ℎ𝑎𝑠𝑒 𝐼𝐼 (2)

Where phase I can be solid, liquid, or gas; whereas phase II can be liquid or vapor depending upon the nature
of the transition whether it is melting, vaporization or sublimation, respectively.
Now consider a system at pressure P and temperature T; and the free energies of phase I and phase
II of the same system are G1 and G2, respectively. Moreover, if we change the temperature from T to T+dT and
pressure from P to P+dP; the free energy changes in phase I and phase II will be dG1 and dG2, respectively.
Mathematically, we can formulate the situation as given below.

𝑑𝐺1 = 𝑉1 𝑑𝑃 − 𝑆1 𝑑𝑇 (3)

𝑑𝐺2 = 𝑉2 𝑑𝑃 − 𝑆2 𝑑𝑇 (4)

Where V1 and S1 represent the molar volume and entropy of phase I, while V2 and S2 represent the molar volume
and entropy of phase II. However, if the phase I and phase II are in equilibrium with each other, we have

𝛥𝐺 = 0 (5)

𝐺2 − 𝐺1 = 0 (6)

Now recalling the fact that G1 and G2 are simply the free energies of the phase I and phase II at temperature T
and pressure P; we can say the same must also be true at temperature T+dT and pressure P+dP i.e.

(𝐺2 + 𝑑𝐺2 ) − (𝐺1 + 𝑑𝐺1 ) = 0 (7)

Copyright © Mandeep Dalal


290 A Textbook of Physical Chemistry – Volume I

(𝐺2 − 𝐺1 ) + (𝑑𝐺2 − 𝑑𝐺1 ) = 0 (8)

Since 𝐺2 − 𝐺1 = 0 from equation (6), the above equation takes the form

𝑑𝐺2 − 𝑑𝐺1 = 0 (9)

𝑑𝐺2 = 𝑑𝐺1 (10)

After putting the values of 𝑑𝐺1 and 𝑑𝐺2 from equation (3, 4) in equation (10), we get

𝑉2 𝑑𝑃 − 𝑆2 𝑑𝑇 = 𝑉1 𝑑𝑃 − 𝑆1 𝑑𝑇 (11)

or

𝑉2 𝑑𝑃 − 𝑉1 𝑑𝑃 = 𝑆2 𝑑𝑇 − 𝑆1 𝑑𝑇 (12)

(𝑉2 − 𝑉1 )𝑑𝑃 = (𝑆2 − 𝑆1 )𝑑𝑇 (13)

𝛥𝑉. 𝑑𝑃 = 𝛥𝑆. 𝑑𝑇 (14)

𝑑𝑃 𝛥𝑆 (15)
=
𝑑𝑇 𝛥𝑉
Now if the 𝛥𝐻 is the latent heat of phase transformation occurring at temperature, the entropy change is
or

𝛥𝐻 (16)
𝛥𝑆 =
𝑇
After using the value of 𝛥𝑆 from equation (16) into equation (15), we get

𝑑𝑃 𝛥𝐻 (17)
=
𝑑𝑇 𝑇𝛥𝑉
The equation was first developed by Calpeyron in 1834 and therefore is named after him Calpeyron equation.
Now if phase I is solid while phase II is vapor i.e. solid ⇌ melt equilibria, the equation (17) takes the form

𝑑𝑃 𝛥𝑓𝑢𝑠 𝐻 (18)
=
𝑑𝑇 𝑇𝑓 𝛥𝑉

Where 𝛥𝑓𝑢𝑠 𝐻 is the latent heat of fusion and 𝑇𝑓 is the melting point. The term 𝛥𝑉 = 𝑉𝑙 − 𝑉𝑠 i.e. the difference
in the volume of melt and solid phase. Therefore, equation (18) can also be written as

𝑑𝑃 𝛥𝑓𝑢𝑠 𝐻 (19)
=
𝑑𝑇 𝑇𝑓 (𝑉𝑙 − 𝑉𝑠 )

For vaporisation equilibrium i.e. liquid ⇌ vapour,

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 291

𝑑𝑃 𝛥𝑣𝑎𝑝 𝐻 (20)
=
𝑑𝑇 𝑇𝑏 (𝑉𝑣 − 𝑉𝑙 )

Where 𝛥𝑣𝑎𝑝 𝐻 is the latent heat of vaporization and 𝑇𝑏 is the boiling point. The term 𝛥𝑉 = 𝑉𝑣 − 𝑉𝑙 i.e. the
difference in the volume of vapor and liquid phases. Now since the volume of the liquid is very small in
comparison to the gaseous phase, the term Vl in equation (20) can simply be neglected for simplicity i.e.

𝑑𝑃 𝛥𝑣𝑎𝑝 𝐻 (21)
=
𝑑𝑇 𝑇𝑉𝑣

If the vapor act as an ideal gas i.e. 𝑃𝑉𝑣 = 𝑅𝑇 or 𝑉𝑣 = 𝑅𝑇/𝑃, the above equation can be written as

𝑑𝑃 𝛥𝑣𝑎𝑝 𝐻 (22)
= .𝑃
𝑑𝑇 𝑅𝑇 2
or

1 𝑑𝑃 𝛥𝑣𝑎𝑝 𝐻 (23)
. =
𝑃 𝑑𝑇 𝑅𝑇 2
or

𝑑 𝑙𝑛 𝑃 𝛥𝑣𝑎𝑝 𝐻 (24)
=
𝑑𝑇 𝑅𝑇 2
The above equation is popularly known as the Clausius-Clapeyron equation.
One more popular form of Clausius-Clapeyron Equation is the integrated one which can easily be
derived by using the rearranged form of equation (24) as given below.

𝛥𝑣𝑎𝑝 𝐻 (25)
𝑑 𝑙𝑛 𝑃 = 𝑑𝑇
𝑅𝑇 2
If the temperature changes from T1 to T2 and pressure is varied from P1 to P2, then we can say that
𝑃2 𝑇2 (26)
𝛥𝑣𝑎𝑝 𝐻
∫ 𝑑 𝑙𝑛 𝑃 = ∫ 𝑑𝑇
𝑅𝑇 2
𝑃1 𝑇1

or
𝑇2 (27)
𝑃2 𝛥𝑣𝑎𝑝 𝐻 1
𝑙𝑛 = ∫ 2 𝑑𝑇
𝑃1 𝑅 𝑇
𝑇1

𝑃2 𝛥𝑣𝑎𝑝 𝐻 1 1 (28)
𝑙𝑛 = [ − ]
𝑃1 𝑅 𝑇1 𝑇2

Copyright © Mandeep Dalal


292 A Textbook of Physical Chemistry – Volume I

After converting to the common logarithm, we get

𝑃2 𝛥𝑣𝑎𝑝 𝐻 1 1 (29)
2.303 𝑙𝑜𝑔 = [ − ]
𝑃1 𝑅 𝑇1 𝑇2

or

𝑃2 𝛥𝑣𝑎𝑝 𝐻 1 1 (30)
𝑙𝑜𝑔 = [ − ]
𝑃1 2.303 𝑅 𝑇1 𝑇2

Which is the integrated form of Clausius-Clapeyron equation.


 Applications of Clausius-Clapeyron Equation
There are many applications of Clausius-Clapeyron equation ranging from chemistry to meteorology
and climatology as well. Some of the most important applications in chemistry are given below.
i) For transitions between a gas and a condensed phase (liquid or solid), the Clausius-Clapeyron equation can
be given as

𝐿 1 (31)
𝑙𝑛 𝑃 = − ( ) + 𝐶
𝑅 𝑇
Where C represents a constant. The symbol L represents the specific latent heat for a liquid-gas transition
whereas specific latent heat of sublimation for a solid-gas transition. If the latent heat is known, then knowledge
of one point on the coexistence curve determines the rest of the curve. Conversely, as the plot 𝑙𝑛𝑃 vs 1/T is
linear with the zero intercept, the simple linear fitting is used to estimate the latent heat of phase change. The
latent heat of vaporization of any liquid can also be determined if the vapor pressures of the same are known
at two different temperatures.
ii) The boiling point of a liquid at any temperature can be determined if the boiling point at a particular
temperature and latent heat of vaporization of the same are known.
iii) If the latent heat of vaporization and the vapor pressure of a liquid at a particular temperature are known,
the vapor pressure of the same liquid at any other temperature can be known.
iv) One more application of this equation is to check if a phase transition will occur via compression or
expansion in solid ⇌ melt equilibria. To understand this, recall the

𝑑𝑃 𝛥𝑓𝑢𝑠 𝐻 (32)
=
𝑑𝑇 𝑇𝑓 𝛥𝑉

The sign for 𝑑𝑃/𝑑𝑇 can be positive or negative for expansion and compression, respectively.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 293

 Law of Mass Action and Its Thermodynamic Derivation


According to the law of mass action, the rate of a chemical reaction is directly proportional to the
product of the activities or simply the active masses of the reactants each term raised to its stoichiometric
coefficients.
To understand the law of mass action in mathematical language, consider a reaction in which two
reactants A and B react to form the product C and D i.e.

𝑎𝐴 + 𝑏𝐵 → 𝑐𝐶 + 𝑑𝐷 (33)

Then the law of mass action says the rate of the above conversion should be

𝑅𝑎𝑡𝑒 ∝ [𝐴]𝑎 [𝐵]𝑏 (34)

𝑅𝑎𝑡𝑒 = 𝑘[𝐴]𝑎 [𝐵]𝑏 (35)

Where k is the constant of proportionality and is typically labeled as rate constant of the reaction.
However, the actual rate of the reaction may or may not be equal to what is suggested by the “law of
mass action” because the actual rate law may have powers raised to the active masses different from their
stoichiometric coefficients. Mathematically, the actual rate law for the reaction given by equation (33) is

𝑅𝑎𝑡𝑒 ∝ [𝐴]𝛼 [𝐵]𝛽 (36)

𝑅𝑎𝑡𝑒 = 𝑘[𝐴]𝛼 [𝐵]𝛽 (37)

Now comparing equation (35) and equation (37); the law of mass action and actual rate law will give same
results when 𝑎 = 𝛼 and 𝑏 = 𝛽; whereas different results will be observed when 𝑎 ≠ 𝛼 and 𝑏 ≠ 𝛽.
 Modern Definition of the Law of Mass Action
The law of mass action can be used to study the composition of a mixture in a reversible reaction
under equilibrium conditions. To do so, consider a typical reversible reaction i.e.

𝑎𝐴 + 𝑏𝐵 ⇌ 𝑐𝐶 + 𝑑𝐷 (38)

Now, from the law of mass action, we know that the rate of forward reaction (Rf) and rate backward reaction
(Rb) will be

𝑅𝑓 = 𝑘𝑓 [𝐴]𝑎 [𝐵]𝑏 (39)

𝑅𝑏 = 𝑘𝑏 [𝐶]𝑐 [𝐷]𝑑 (40)

Where 𝑘𝑓 and 𝑘𝑏 are the rate constants for the forward and backward reactions, respectively. After equilibrium
is reached, we have

𝑅𝑓 = 𝑅𝑏 (41)

Copyright © Mandeep Dalal


294 A Textbook of Physical Chemistry – Volume I

𝑘𝑓 [𝐴]𝑎 [𝐵]𝑏 = 𝑘𝑏 [𝐶]𝑐 [𝐷]𝑑 (42)

or

𝑘𝑓 [𝐶]𝑐 [𝐷]𝑑 (43)


=
𝑘𝑏 [𝐴]𝑎 [𝐵]𝑏

Since the 𝑘𝑓 and 𝑘𝑏 are also constant at equilibrium, the ratio of the two is also a constant and is typically
labeled as K or the equilibrium constant. Therefore, equation (43) is modified as

𝑘𝑓 [𝐶]𝑐 [𝐷]𝑑 (44)


𝐾= =
𝑘𝑏 [𝐴]𝑎 [𝐵]𝑏

All this leads to the modern definition of “law of mass action” that the ratio of the multiplication of molar
concentrations of products raised to the power of their stoichiometric coefficients to the multiplication of the
molar concentrations of the reactants raised to the power of their stoichiometric coefficients is constant at
constant temperature and is called as “equilibrium constant”. It is also worthy to mention that equation (44) is
also known as the “law of chemical equilibrium”.
 Thermodynamic Derivation of the Law of Mass Action
In order to derive the law of mass action thermodynamically, recall the general form of a typical
reversible reaction under equilibrium conditions in which reactants and products are ideal gases i.e.

𝑎𝐴 + 𝑏𝐵 ⇌ 𝑐𝐶 + 𝑑𝐷 (45)

Now, as we know that the total free energy of the reactant (𝐺𝑅 ) can be formulated as

𝐺𝑅 = 𝑎𝜇𝐴 + 𝑏𝜇𝐵 (46)

Where 𝜇𝐴 and 𝜇𝐵 are the chemical potentials of reactant A and B, respectively. Similarly, the total free energy
of the products (𝐺𝑃 ) can also be formulated i.e.

𝐺𝑃 = 𝑐𝜇𝐶 + 𝑑𝜇𝐷 (47)

It is also important to mention that the temperature and pressure are kept constant. Moreover, the free energy
change of the whole reaction can be obtained by subtracting equation (46) from equation (47) i.e.

𝛥𝐺𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 = 𝐺𝑃 − 𝐺𝑅 (48)

𝛥𝐺𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛 = (𝑐𝜇𝐶 + 𝑑𝜇𝐷 ) − (𝑎𝜇𝐴 + 𝑏𝜇𝐵 ) (49)

Recalling the fact that the free energy change at equilibrium is zero, equation (49) is reduced to

(𝑐𝜇𝐶 + 𝑑𝜇𝐷 ) − (𝑎𝜇𝐴 + 𝑏𝜇𝐵 ) = 0 (51)

Now recall the expression of the chemical potential of the ith species in gas phase i.e.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 295

𝜇𝑖 = 𝜇𝑖0 + 𝑅𝑇 ln 𝑝𝑖 (52)

Where 𝑝𝑖 and 𝜇𝑖0 are the partial pressure and standard chemical potential of ith species, respectively. Now
using equation (52) in equation (51), we get

[𝑐(𝜇𝐶0 + 𝑅𝑇 ln 𝑝𝐶 ) + 𝑑(𝜇𝐷0 + 𝑅𝑇 ln 𝑝𝐷 )] − [𝑎(𝜇𝐴0 + 𝑅𝑇 ln 𝑝𝐴 ) + 𝑏(𝜇𝐵0 + 𝑅𝑇 ln 𝑝𝐵 )] = 0 (53)

or

𝑐𝜇𝐶0 + 𝑐𝑅𝑇 ln 𝑝𝐶 + 𝑑𝜇𝐷0 + 𝑑𝑅𝑇 ln 𝑝𝐷 − 𝑎𝜇𝐴0 − 𝑎𝑅𝑇 ln 𝑝𝐴 − 𝑏𝜇𝐵0 − 𝑏𝑅𝑇 ln 𝑝𝐵 = 0 (54)

𝑐𝜇𝐶0 + 𝑅𝑇 ln 𝑝𝐶𝑐 + 𝑑𝜇𝐷0 + 𝑅𝑇 ln 𝑝𝐷𝑑 − 𝑎𝜇𝐴0 − 𝑅𝑇 ln 𝑝𝐴𝑎 − 𝑏𝜇𝐵0 − 𝑅𝑇 ln 𝑝𝐵𝑏 = 0 (55)

𝑅𝑇 ln 𝑝𝐶𝑐 + 𝑅𝑇 ln 𝑝𝐷𝑑 − 𝑅𝑇 ln 𝑝𝐴𝑎 − 𝑅𝑇 ln 𝑝𝐵𝑏 = −𝑐𝜇𝐶0 − 𝑑𝜇𝐷0 + 𝑎𝜇𝐴0 + 𝑏𝜇𝐵0

or

𝑅𝑇 ln (𝑝𝐶𝑐 𝑝𝐷𝑑 ) − 𝑅𝑇 ln (𝑝𝐴𝑎 𝑝𝐵𝑏 ) = −[𝑐𝜇𝐶0 + 𝑑𝜇𝐷0 − 𝑎𝜇𝐴0 − 𝑏𝜇𝐵0 ] (56)

(𝑝𝐶𝑐 𝑝𝐷𝑑 ) (57)


𝑅𝑇 ln = −[𝐺𝑃𝑜 − 𝐺𝑅𝑜 ]
(𝑝𝐴𝑎 𝑝𝐵𝑏 )

(𝑝𝐶𝑐 𝑝𝐷𝑑 ) 𝑜
(58)
𝑅𝑇 ln = −𝛥𝐺𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛
(𝑝𝐴𝑎 𝑝𝐵𝑏 )

Where 𝛥𝐺𝑟𝑒𝑎𝑐𝑡𝑖𝑜𝑛
𝑜
is the standard free energy change of the reaction can be simply abbreviated as 𝛥𝐺 𝑜 only.
Therefore, the equation (58) can be rearranged as given below.

(𝑝𝐶𝑐 𝑝𝐷𝑑 ) 𝛥𝐺 𝑜 (59)


ln =−
(𝑝𝐴𝑎 𝑝𝐵𝑏 ) 𝑅𝑇

𝑝𝐶𝑐 𝑝𝐷𝑑 −
𝛥𝐺 𝑜 (61)
= 𝑒 𝑅𝑇
𝑝𝐴𝑎 𝑝𝐵𝑏

Now because 𝛥𝐺 𝑜 is a function of temperature only and R is a constant quantity, the right-hand side can be
put equal to another constant, say ‘Kp’.
𝛥𝐺 𝑜 (62)
𝑒 − 𝑅𝑇 = 𝐾𝑝

From equation (61) and equation (62), we have

𝑝𝐶𝑐 𝑝𝐷𝑑 (63)


𝐾𝑝 =
𝑝𝐴𝑎 𝑝𝐵𝑏

Which is again the modern statement of “law of mass action” but in terms of partial pressures.

Copyright © Mandeep Dalal


296 A Textbook of Physical Chemistry – Volume I

Other forms of equation (63) can also be written depending upon the reactants and products involved.
If the chemical potentials of the reactants and products are in mole fractions (xi) i.e.

𝜇𝑖 = 𝜇𝑖0 + 𝑅𝑇 ln 𝑥𝑖 (64)

Then equation (63) takes the form

𝑥𝐶𝑐 𝑥𝐷𝑑 (65)


𝐾𝑥 =
𝑥𝐴𝑎 𝑥𝐵𝑏

Similarly, If the chemical potentials of the reactants and products are in molar concentrations (ci) i.e.

𝜇𝑖 = 𝜇𝑖0 + 𝑅𝑇 ln 𝑐𝑖 (66)

Then equation (63) takes the form

[𝐶]𝑐 [𝐷]𝑑 (67)


𝐾𝑐 =
[𝐴]𝑎 [𝐵]𝑏

Which is the popular form of “law of mass action”.

 Third Law of Thermodynamics (Nernst Heat Theorem, Determination of


Absolute Entropy, Unattainability of Absolute Zero) And Its Limitation
The third law of thermodynamics states that the entropy of a system approaches a constant value as
its temperature approaches absolute zero.
In order to understand the abovementioned statement fully, we need to discuss some very important
concepts and consequences of third law thermodynamics first such as “Nernst heat theorem” or “absolute
entropies” etc.
 Nernst Heat Theorem
In 1906, Walther Nernst, a German chemist, studied the variation of enthalpy change and free energy
change as a function of temperature. His results gained very much popularity as “Nernst heat theorem” in
chemistry and built the foundation of many other discoveries. For detailed picture, recall the Gibbs-Helmholtz
equation i.e.

𝜕(𝛥𝐺) (68)
𝛥𝐺 = 𝛥𝐻 + 𝑇 ( )
𝜕𝑇 𝑃

It is obvious from the above equation that the free energy change will become equal to the enthalpy change
when the temperature is reduced to absolute zero i.e. 𝛥𝐺 = 𝛥𝐻 at 𝑇 = 0. Besides, Nernst also noted that the
magnitude of 𝜕(𝛥𝐺)/𝜕𝑇 declines gradually and approaches the zero with the decrease of temperature.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 297

In other words, W. Nernst observed that as the temperature is decreased continuously, the Gibbs free
energy change was decreasing while the enthalpy change was increasing gradually with the same magnitude.
Therefore, the change in slope in both curves must become zero near absolute zero i.e.

lim 𝜕(𝛥𝐺) lim 𝜕(𝛥𝐻) (69)


= =0
𝑇 → 0 𝜕𝑇 𝑇 → 0 𝜕𝑇
Which is the mathematical form of Nernst heat theorem. The pictorial representation of these observations is
also given below for a more clear perspective.

Figure 1. The graphical representation of Nernst heat theorem.

It is also worthy to mention that though we have shown 𝛥𝐺 greater than 𝛥𝐻 when the temperature is set greater
than zero, the reverse may also be possible because 𝜕(𝛥𝐺)/𝜕𝑇 can have positive as well as negative values.
Furthermore, we also know that variation of free energy change with the temperature at constant
pressure is equal to the negative of entropy change i.e.

𝜕(𝛥𝐺) (70)
( ) = −𝛥𝑆
𝜕𝑇 𝑃

Also, from the definition of change in the heat capacity, we have

𝜕(𝛥𝐻) (71)
( ) = 𝛥𝐶𝑃
𝜕𝑇 𝑃

Copyright © Mandeep Dalal


298 A Textbook of Physical Chemistry – Volume I

After putting the values of equation (70, 71) in equation (69), we get

lim (𝛥𝑆)
=0 𝑎𝑛𝑑
lim
𝛥𝐶 = 0 (72)
𝑇→0 𝑇→0 𝑃
Now because, no gas or liquid at or in the vicinity of absolute zero, we can apply the Nernst heat theorem to
solids only.
The right-hand side result given in equation (72) is extremely important as far as the third law of
thermodynamics is concerned. Actually, we can say that the “third law of thermodynamics” follows from the
results obtained by the Nernst heat theorem. In order to understand it more clearly, recall one of two results of
Nernst heat theorem i.e.

lim
𝛥𝐶 = 0 (73)
𝑇→0 𝑃
This means that as we approach absolute zero, the heat capacity of reactants becomes equal to the heat capacity
of the product. In other words, we can say that the heat capacities of all substances are equal at absolute zero.
Mathematically, it can be formulated as

lim
𝐶𝑃 (𝑝𝑟𝑜𝑑𝑢𝑐𝑡) =
lim
𝐶 (𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡) (74)
𝑇→0 𝑇→0 𝑃
Since the results of the quantum mechanics say that the heat capacity of tends to zero as the temperature
approaches absolute zero, the above equation takes the form

lim
𝐶 =0 (75)
𝑇→0 𝑃
Similarly recalling the second result of Nernst heat theorem i.e.

lim (𝛥𝑆)
=0 (76)
𝑇→0
Which means that as we approach absolute zero, the entropy of reactants becomes equal to the entropy of the
product. In other words, we can say that the entropy of all substances are equal at absolute zero.
Mathematically, it can be formulated as

lim
𝑆𝑝𝑟𝑜𝑑𝑢𝑐𝑡 =
lim
𝑆 (77)
𝑇→0 𝑇 → 0 𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡
Using the same argument as in case of heat capacity, the above equation takes the form

lim
𝑆=0 (78)
𝑇→0
Therefore, the entropy of all crystalline substances can be taken as zero at absolute zero, which is the general
statement of the third law of thermodynamics.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 299

 Determination of Absolute Entropy


One of the most interesting and important applications of the third law of thermodynamics (or the
Nernst heat theorem) is that it can be used to determine the absolute entropies of different substances at any
temperature. The procedure employs the fact that we can calculate the entropy change easily and if the entropy
of the initial state is zero then this difference will simply be equal to the absolute entropy i.e.

𝛥𝑆 = 𝑆𝑇 − 𝑆0 (79)

Where 𝑆𝑇 and 𝑆0 are the entropies at temperatures T and 0K, respectively. Since we know from the third law
of thermodynamics that 𝑆0 = 0, the equation (79) takes the form

𝛥𝑆 = 𝑆𝑇 − 0 = 𝑆𝑇 (80)

Before we proceed further, we must remember that a substance may change phase when it supplied
with heat. Therefore, we need to discuss the absolute entropies in solid, liquid and gases separately.
1. Absolute entropies in case of solids: To calculate the absolute entropy of a solid at any temperature, we
just need to find the total entropy change in shifting the absolute zero state to that temperature. The very small
entropy change is given by

𝛿𝑞 (81)
𝑑𝑆 =
𝑇
However, the general expression for heat capacity is

𝛿𝑞 (82)
𝐶𝑃 =
𝑑𝑇
𝛿𝑞 = 𝐶𝑃 𝑑𝑇 (83)

After putting the value of 𝛿𝑞 from equation (83) into equation (81), we have

𝐶𝑃 𝑑𝑇 (84)
𝑑𝑆 =
𝑇
In order to find the total entropy change in the same phase (solid in this case) when the temperature is raised
from 0 to 𝑇, we need to integrate the above equation over the range of interest i.e.
𝑇 𝑇 (85)
𝐶𝑃 𝑑𝑇
∫ 𝑑𝑆 = ∫
𝑇
0 0

𝑇 (86)
𝑆𝑇 − 𝑆0 = ∫ 𝐶𝑃 𝑑 ln 𝑇
0

Since 𝑆0 = 0, the above equation becomes

Copyright © Mandeep Dalal


300 A Textbook of Physical Chemistry – Volume I

𝑇 (87)
𝑆𝑇 = ∫ 𝐶𝑃 𝑑 ln 𝑇
0

Thus, the entropy of any solid at temperature T can be obtained by heat capacity at many temperature points
between 0K to T. The total integral of equation (87) can be obtained by measuring the area under the plot of
𝐶𝑃 vs ln T.

Figure 2. The plot of heat capacity vs T and ln T.

2. Absolute entropies in case of liquids: To calculate the absolute entropy of a liquid at any temperature, we
just need to find the total entropy change in shifting the absolute zero state to that temperature. The whole
process can be divided into three steps; one in heating up the solid phase form 0K to it melting or the fusion
point, then phase change from solid to liquid at fusion temperature, and the last step in which liquid is heated
up to temperature T.
i) Entropy change from 0K to Tf :
𝑇𝑓 (88)
𝛥𝑆1 = ∫ 𝐶𝑃 (𝑠) 𝑑 ln 𝑇
0

ii) Entropy change of fusion at Tf :

𝛥𝐻𝑓 (89)
𝛥𝑆2 =
𝑇𝑓

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 301

iii) Entropy change from Tf to T :


𝑇 (90)
𝛥𝑆3 = ∫ 𝐶𝑃 (𝑙) 𝑑 ln 𝑇
𝑇𝑓

The total entropy change for the process can be obtained by simply adding equations (88–90) i.e.

𝑆𝑇 − 𝑆0 = 𝛥𝑆1 + 𝛥𝑆2 + 𝛥𝑆3 (91)


𝑇𝑓 𝑇 (92)
𝛥𝐻𝑓
𝑆𝑇 = ∫ 𝐶𝑃 (𝑠)𝑑 ln 𝑇 + + ∫ 𝐶𝑃 (𝑙) 𝑑 ln 𝑇
𝑇𝑓
0 𝑇𝑓

Where 𝐶𝑃 (𝑠) and 𝐶𝑃 (𝑙) are the heat capacities for solid and liquid phases, respectively. The symbol 𝛥𝐻𝑓
represents the latent heat of fusion.
3. Absolute entropy in case of gases: To calculate the absolute entropy of a gas at any temperature, we just
need to find the total entropy change in shifting the absolute zero state to that temperature. The whole process
can be divided into five steps; one in heating up the solid phase form 0K to it melting or the fusion point, then
phase change from solid to liquid at fusion temperature, followed by the raising the temperature from 𝑇𝑓 to 𝑇𝑏 .
The fourth step includes the vaporization of liquid phase to gaseous phase at boiling point followed by the last
step in which the temperature must be raised from 𝑇𝑏 to the required temperature 𝑇.
i) Entropy change from 0K to Tf :
𝑇𝑓 (93)
𝛥𝑆1 = ∫ 𝐶𝑃 (𝑠) 𝑑 ln 𝑇
0

ii) Entropy change of fusion at Tf :

𝛥𝐻𝑓 (94)
𝛥𝑆2 =
𝑇𝑓

iii) Entropy change from Tf to Tb :


𝑇𝑏 (95)
𝛥𝑆3 = ∫ 𝐶𝑃 (𝑙) 𝑑 ln 𝑇
𝑇𝑓

iv) Entropy change of vaporization at Tb :

𝛥𝐻𝑣𝑎𝑝 (96)
𝛥𝑆4 =
𝑇𝑏

Copyright © Mandeep Dalal


302 A Textbook of Physical Chemistry – Volume I

v) Entropy change from Tb to T :


𝑇 (97)
𝛥𝑆5 = ∫ 𝐶𝑃 (𝑔) 𝑑 ln 𝑇
𝑇𝑏

The total entropy change for the process can be obtained by simply adding equations (88–90) i.e.

𝑆𝑇 − 𝑆0 = 𝛥𝑆1 + 𝛥𝑆2 + 𝛥𝑆3 + 𝛥𝑆4 + 𝛥𝑆5 (98)


𝑇𝑓 𝑇𝑏 𝑇 (99)
𝛥𝐻𝑓 𝛥𝐻𝑣𝑎𝑝
𝑆𝑇 = ∫ 𝐶𝑃 (𝑠)𝑑 ln 𝑇 + + ∫ 𝐶𝑃 (𝑙) 𝑑 ln 𝑇 + + ∫ 𝐶𝑃 (𝑔) 𝑑 ln 𝑇
𝑇𝑓 𝑇𝑏
0 𝑇𝑓 𝑇𝑏

Where 𝐶𝑃 (𝑠), 𝐶𝑃 (𝑙) and 𝐶𝑃 (𝑔) are the heat capacities for solid, liquid and gas phases, respectively. The symbol
𝛥𝐻𝑣𝑎𝑝 represents the latent heat of vaporization.
 The unattainability of Absolute Zero
One more statement of the third law of thermodynamics is that the lowering of the temperature of a
material body to the absolute zero is impossible in the finite number of steps. To understand this claim, recall
the concept of Carnot refrigerator first. A Carnot refrigerator is basically just the reverse of the Carnot heat
engine i.e. Carnot heat engine working in reverse cycle. Since a Carnot heat engine provides work through
reversible isothermal-adiabatic compressions and expansions, a net amount of work must be done in Carnot
refrigerator making it electricity consumer.
Let 𝑞1 be the amount of heat absorbed by a Carnot refrigerator form a body at lower temperature T 1
and 𝑞2 as the amount of heat rejected by the same refrigerator to a body at higher temperature T2. The
coefficient of performance (β) of the reversible Carnot refrigerator can be given by the following relation.

1 (100)
𝛽=
𝑞2 /𝑞1 − 1

Since for a Carnot cycle, we know that


𝑞1 𝑞2 (101)
=
𝑇1 𝑇2

or

𝑞2 𝑇2 (102)
=
𝑞1 𝑇1

After using the value of 𝑞2 /𝑞1 from equation (102) in equation (100), we get

1 (103)
𝛽=
𝑇2 /𝑇1 − 1

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 303

Furthermore, we also know that the coefficient of performance of an ideal Carnot refrigerator is simply the
ratio of the cooling effect to the work done i.e.
𝑞1 (104)
𝛽=
𝑤
Which means that how much heat is removed from the body at lower temperature per unit of work done. Now
from equation (103) and equation (104), we have

𝑞1 1 (105)
=
𝑤 𝑇2 /𝑇1 − 1

𝑞1 𝑇1 (106)
=
𝑤 𝑇2 − 𝑇1

Taking reciprocal of the above result, we get

𝑤 𝑇2 − 𝑇1 (107)
=
𝑞1 𝑇1

Thus, it is obvious from the above equation that as the lower temperature T1 approaches zero, more and more
work will be needed to remove the same amount of heat (w is inversely proportional to T1), which is the
unattainability of the absolute zero.
 Limitation of Third Law of Thermodynamics
Although the third law of thermodynamics is very useful in determining various thermodynamic
properties of various substances, there are some limitations of the same. The reason being is the fact that
entropies obtained using the third law of thermodynamics are thermal entropies and are somewhat smaller than
the entropy values obtained using statistical mechanics. This deviation is typically in range of 3–5 JK−1 mol−1.
From this, we may conclude that there is some entropy present even at the 0K temperature. This
entropy is called as “residual entropy” and can be obtained using the following relation.

𝑆 = 𝑘 ln 𝑊 (108)

Where k is Boltzmann constant and the W is the thermodynamic probability which represents the number of
equally probable orientations of the molecule under consideration. For instance, consider a sample of N number
of carbon monoxide molecules. Since each molecule can have two orientations that are equally probable (CO
CO OC CO OC OC), the thermodynamic probability will be 𝑊 = 2𝑁 . The residual entropy is

𝑆 = 𝑘 ln 𝑊 = 𝑘 ln 2𝑁 = 𝑘𝑁 ln 2 (109)

For one mole of sample 𝑘𝑁 = 𝑅, the above equation takes the form

𝑆 = 𝑅 ln 2 = 2.303𝑅 log 2 = 5.85 JK −1 mol−1 (110)

Copyright © Mandeep Dalal


304 A Textbook of Physical Chemistry – Volume I

 Phase Diagram for Two Completely Miscible Components Systems


Before we discuss the phase diagram for two completely miscible components systems, it is better to
recall the Gibbs phase rule which states that

𝐹 = 𝐶−𝑃+2 (110)

Where F is the number of degrees of freedom, C represents the number of components and P simply gives the
total number of phases. For a two components system, the equation (110) reduces to

𝐹 = 2−𝑃+2= 4−𝑃 (111)

At this point, there are three major possibilities depending upon the number of phases involved which will be
discussed one by one.
 Types of Two-Component Systems Based on Number of Phases
i) When we use 𝑃 = 1 in equation (111), we get

𝐹 =4−1=3 (112)

This means that for a one-phase and two-component system, there are three degrees of freedom. This single-
phase can be gaseous, solid solution, or two completely miscible liquids. Moreover, since the number of
degrees of freedom is three, there are three conditions that can be varied without disturbing the total number
of phases at equilibrium state.

Figure 3. The graphical representation of a one-phase, two-component system.

Hence, the three degrees of freedom can be depicted along the three axes of a cube. However, it is more
common to vary only two variables while keeping the third one as constant. For instance, in case of pressure-
temperature diagrams, composition is kept constant; while in case of pressure-composition curves, the
temperature is kept fixed at a particular value. Furthermore, in case of temperature-composition curves, the
pressure of the system is always kept constant. Now although all the possibilities are there, it is more
convenient to fix the pressure of the system at atmospheric pressure and vary the temperature and composition.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 305

Therefore, we can say that instead of using three variables, it is more popular to keep one constant
and vary the other two. This will also simplify the graphical representation of a simple two-component system.

Figure 4. The two-dimensional variants of the one-phase-two-component system.

ii) When we use 𝑃 = 2 in equation (111), we get

𝐹 =4−2=2 (113)

This means that for a two-phase and two-component system, there are two degrees of freedom. These phases
can be liquid-vapour, liquid-liquid, or solid-liquid. Moreover, since the number of degrees of freedom is two,
there are two variables that can be varied without disturbing total number of phases at equilibrium state.
iii) When we use 𝑃 = 3 in equation (111), we get

𝐹 =4−3=1 (114)

Which means that for a three-phase and two-component system, there is only one degree of freedom. These
phases can be liquid-liquid-vapour, solid-liquid-liquid, or solid-solid-liquid. Moreover, since the number of
degrees of freedom is one, there is only one variable that can be varied without disturbing the total number of
phases at equilibrium state.
Now although a complete phase diagram for two-component systems must be able to show the
different phases up to number three, the situation is slightly simple for “condensed systems”. More specifically,
if all the phases in two-component systems are either solid or liquids only (solid-liquid equilibria), a minor
pressure disturbance will have little to no effect on the system. Therefore, we can conclude that the number of
degree of freedom in such “condensed systems” is actually reduced by one i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (115)

Which is the reduced or condensed phase rule where 𝐹 ′ is the total number of degrees of freedom excluding
pressure. Now because the total number of phases that can exist simultaneously are reduced to 2 (system
become univariant at P =2), The complete phase diagram can be drawn on two dimensional paper with vertical
and horizontal sides representing temperature and pressure, respectively.

Copyright © Mandeep Dalal


306 A Textbook of Physical Chemistry – Volume I

 Types of Completely Miscible Two-Component Systems in Solid-Liquid Equilibria


We know that in a solid-liquid equilibrium, the solid can be one of the constituents or solid solution
which are completely miscible in the liquid phase.
Consider a liquid mixture of two components A and B at temperature T. Now if this liquid mixture is
allowed to cool down below the freezing point of the mixture, the solid will start to separate out. At this point,
assume that we have a number of such mixtures with the same components A and B but with different
compositions (i.e. with different ratios of A and B). The cooling of all the mixtures is carried out in open vessels
so that the pressure remains constant (atmospheric pressure). After that, the only thing we need to do is to plot
these freezing point vs the composition. These completely miscible two-component systems can primarily be
classified into three categories.
1. Eutectic systems: In these types of systems, the components do not react with each other but only the simple
mixing takes place in the solution or in the molten state. The common examples of such systems are lead-silver
system, bismuth-cadmium system, potassium iodide-water system.

Figure 5. General phase diagram of eutectic systems.

2. Systems forming solid compounds AxBy with congruent and incongruent melting points: In these types
of systems, the components do react with each other and the formation of a compound takes place. The
common example of such systems are the Mn-Zn system (forming MgZn2) and Na2SO4-H2O system (forming
Na2SO4.10H2O). These types of systems can further be classified.
i) Systems forming solid compounds with congruent melting point: In these types of systems, the two
components react to give a compound which is quite stable until its melting point is reached. When melted,
the composition remains the same as that of the solid, and such compounds are said to have a congruent melting
point. Some of the common examples of such systems are Mg-Zn and FeCl3-H2O systems.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 307

Figure 6. General phase diagram for systems forming solid compounds with congruent melting points.

ii) Systems forming solid compounds with incongruent melting point: In these types of systems, the two
components react to give a compound which is not stable up to its melting point. When heated, the
decomposition starts before the melting point is reached; and a new solid phase and a solution or melt with a
different composition from the original solid are formed. Such compounds are said to undergo peritectic or
transition reaction are labeled to have a congruent melting point. Some of the common examples of such
systems are Na2SO4-H2O and NaCl-H2O systems. A typical transition can be represented as

𝐶1 ⇌ 𝐶2 + melt or solution (116)

Where C1 is the compound formed by the reaction between participating components whereas C2 represents
the compound formed as a result of decomposition of C1 below its fusion temperature.

Figure 7. General phase diagram for systems forming solid compounds with incongruent melting points.

Copyright © Mandeep Dalal


308 A Textbook of Physical Chemistry – Volume I

3. Systems forming solid solutions: In these types of systems, the components are completely miscible with
each other in solid phase and completely homogeneous solid solutions are produced. The X-ray diffraction
studies are typically employed to check that single crystalline phase is obtained rather than a mixture of two
solid phases. The common example of such systems are Co-Ni system, Au-Ag system and AgCl-NaCl system.

Figure 8. The general phase diagram of systems forming solid solutions.

 Experimental Methods for The Determination of Phase Diagram of Two-Component Systems


The most common approaches for the determination of phase diagram of two-component systems are
“cooling curve” and “thaw melt” methods. These methods are quite popular due to their easiness and
practicability to many systems.
1. Cooling curve method: In this approach, a liquid mixture of two components A and B at temperature T is
allowed to cool down below the freezing point of the mixture, the solid will start to separate out. A number of
such mixtures with the same components A and B but with different compositions (i.e. with different ratios of
A and B) are then allowed to cool down in open vessels so that the pressure remains constant (atmospheric
pressure). The temperature of the system is regularly recorded at different times to obtain temperature vs time
plots. After that, from the breaks and arrests in those plots, important information such as freezing point or
solidification time are obtained for different compositions.
Now in order to understand the concept more clearly, a variation of temperature with time for pure
and a liquid mixture must be discussed. In the case of a liquid mixture, above point b, the system is completely
liquid and it will cool down very rapidly in going from point a to point b. Now because at point b, the
crystallization of the compound A will start releasing a large amount of heat, and therefore, the rate of cooling
will be slightly slower until point c is reached. Now since the system is getting solidified at eutectic temperature
i.e. it is moving from c (solidification of B starts) to d the temperature of the system will remain constant
during this conversion. After point d, the system will cool down as a solid mixture with two degrees of freedom.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 309

Figure 9. The typical cooling curves for pure a pure liquid (left) and a liquid mixture (right).

It is also worthy to mention that in some cases a phenomenon called “supercooling” is observed in which the
break at point ‘b’ is absent and cooling of liquid continues along a-b with sudden rise after some time giving
an abnormal break at b' instead of b. The correct break is found by extrapolating from point b' backward.

Figure 10. The typical cooling curves for a liquid mixture with supercooling.

Now although the supercooling phenomenon does not create a major problem in most of the cases, it can be a
real mess in some systems. Therefore, another method that can overcome these limitations must be discussed.

Copyright © Mandeep Dalal


310 A Textbook of Physical Chemistry – Volume I

2. Thaw-melt method: This method is free from the limitation of the “supercooling phenomenon” posed by
the cooling carve method. In this approach, a solid mixture of two components A and B at temperature T is
heated continuously. This is exactly opposite of what we follow in cooling curve method and we will note
down melting points for different compositions. A number of such solid mixtures with same components A
and B but with different compositions (i.e. with different ratios of A and B) are then heated up in open vessels
so that the pressure remains constant (atmospheric pressure). The temperature of system is regularly recorded
at different times to obtain temperature vs time plots. After that, from the breaks and arrests in those plots,
important information such as melting point or liquefaction time are obtained for different compositions.
Now in order to understand the concept more clearly, a variation of temperature with time for pure
and a liquid mixture must be discussed. In the case of a solid mixture, below point e, the system is completely
solid and it will heat up slowly in going from point e to point d. Now since the system is getting liquefied at
eutectic temperature i.e. it is moving from d to c; the temperature of the system will remain constant during
this conversion. Now because at point c, the melting of A compound will continue requiring a large amount of
heat, and therefore, the rate of heating will be slightly slower until point b is reached. After point b, the system
will heat up as liquid mixture with two degrees of freedom very rapidly.

Figure 11. The typical Thaw-melting heating curves (temperature vs time) for a pure liquid (left) and a
liquid mixture (right).

It is also worthy to mention that the temperature at which the melting of the solid mixture starts is typically
called as “thaw point” and the temperate at which liquefaction ends is called as “melting point”.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 311

 Eutectic Systems (Calculation of Eutectic Point)


In order to understand the eutectic systems in a comprehensive manner, we need to recall the general
phase diagram for eutectic systems i.e.

Figure 12. The general phase diagram of eutectic systems.

It is obvious from the phase diagram given above that the diagram has two curves OQ and QS enclosing the
area ODQ and SQE, respectively. The significance of different parts of the above-depicted diagram is discussed
below.
1. The OQ and SQ curves: The point O and point S represent the freezing point the pure compound A and
compound B, respectively. When component B is added to component A, its freezing point of decreases
regularly along OQ. Likewise, when component A is added to component B, its freezing point decreases
regularly along SQ. In conclusion, we can say that the curve OQ and SQ represent the temperature-conditions
at which solid A and solid B are in equilibrium with the liquid mixture. Since there are only two phases
involved, and both are condensed in nature (solid and liquid), we need to use condensed or reduced phase rule
here i.e. after putting C = 2 and P =2 in equation (114), we get

𝐹′ = 2 − 2 + 1 = 1 (117)

Which means that the system is univariate. In other words, only one condition is needed to be defined to define
the whole system; for instance, if we define a particular composition of A and B, the freezing point of the
mixture is completely fixed. The curve OQ and SQ also called as “liquidus” because they separate the field of
all liquid from that of liquid + solid crystals.

Copyright © Mandeep Dalal


312 A Textbook of Physical Chemistry – Volume I

2. Point Q or the eutectic point: At point ‘Q’, all the three phases (solid A + solid B + liquid) are in
equilibrium with each other; and therefore, zero degrees of freedom exists at this point i.e.

𝐹′ = 2 − 3 + 1 = 0 (118)

This means that we can neither change temperature nor composition without disturbing the mutual equilibrium
of solid A, solid B and liquid. Moreover, below this temperature, the mixture freezes as a whole. This
temperature is called as “eutectic point” and the composition corresponding to this point is called as eutectic
composition.
3. Area ODQ and SQE: In addition to the enclosed area ODQ and SQE, there are also areas above OQS line
and below DQE line as well.
i) Area above OQS: It is obvious from the phase diagram that the area lying above the line OQS only has a
liquid phase, which means

𝐹′ = 2 − 1 + 1 = 2 (119)

Two degrees of freedom means that we can change temperature well as composition without disturbing
completely melt state as long as the temperature-composition coordinates lie above the line OQS. In other
words, we need to define both, temperature as well as composition, to define the system completely in this
area.
ii) Area below DQE: It is also obvious from the phase diagram that the area lying below the line DQE has two
solid phase, which means

𝐹′ = 2 − 2 + 1 = 1 (120)

The only one degree of freedom means that we can change temperature only without disturbing the state of the
system. In other words, we only need to define the temperature to define the system completely in this area.
The curve DQE is also called as “solidus” because they separate the field of all solids from that of liquid +
solid crystals.
iii) Area enclosed by ODQ: It is also obvious from the phase diagram that the area enclosed by ODQ line has
solid A + liquid phases, which means

𝐹′ = 2 − 2 + 1 = 1 (121)

The only one degree of freedom means that we can change temperature only without disturbing the state of the
system. In other words, we only need to define the temperature to define the system completely in this area.
iv) Area enclosed by SQE: It is also obvious from the phase diagram that the area enclosed by SQE line has
solid B + liquid phases, which means

𝐹′ = 2 − 2 + 1 = 1 (122)

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 313

The only one degree of freedom means that we can change temperature only without disturbing the state of the
system. In other words, we only need to define the temperature to define the system completely in this area.
It is also worthy to note that we can find out the equilibrium phase-composition at any temperature
by simply plotting a horizontal line that cuts the line OQ or SQ. For instance, let such a line 1x’ that cuts OQ
at x', or 2y' that cuts OQ at y'. The importance of this line will be disused later in this section.
 Cooling of a Liquid Mixture and the Calculation of Eutectic Point
In order to understand the cooling profile of a liquid mixture and eutectic point, we must state the
following rule first.
During equilibrium melting or crystallization in a closed system, the initial composition of the system
will be identical to the final composition of the system.
Now consider the composition of the liquid mixture in which compound A is 80% while compound
B is 20% (represented by the vertical line X). Therefore, it follows from the above-mentioned rule that after
crystallization we must get the same composition i.e. 80% (solid A) + 20% (solid B).

Figure 13. The cooling of a liquid mixture.

Above temperature T1, the composition X lies in the field of “all liquid”, and therefore, will exist as liquid
completely. However, when the temperature is reduced to T1 (corresponding to point 1) the crystal formation
of A will be initiated. The further lowering of the temperature will make more crystals of A to form. All this
will result in a continuous change of composition of the liquid mixture which is becoming more and more
enriched in B. Therefore, as the temperature is lowered from T1 to T2, the composition of the liquid will change

Copyright © Mandeep Dalal


314 A Textbook of Physical Chemistry – Volume I

from point x to point y (corresponding to point 2). Likewise, the further lowering of temperature from T 2 to T3
and then to eutectic temperature (TE) will change the composition from y to z (corresponding to point 3) and
then to Ce (corresponding to point Q), respectively.
As near point Q, the liquid will be rich in compound B relative to the initial composition, and the
crystallization of compound B will also start at point Q; it is better to imagine the composition in solid mixture
afterward. By recalling the fact that all the three phases coexist at point Q (solid A + solid B + liquid mixture),
even an infinitesimal decrease in temperature will disturb the “eutectic point” and liquid B will also start
converting into a solid phase. This will make the system move from point Q to Q', getting exactly the same
composition x i.e. 80% solid A + 20% solid B (represented by the vertical line X). A further lowering of
temperature will make the system to move downward from point Q' along the vertical line X again.
1. Determination of phase composition:
Now, it is clear from the phase diagram that above point 1 along line X, a 80% A and 20 % B will
exist in the liquid state; whereas, below point Q', a 80% A and 20 % B will exist in solid-state. However,
between point 1 and Q' (i.e. at temperatures between T1 and Te) two phases (liquid mixture of A and B +
crystals of A) will be present in the system. If the crystallization process is at any point, the relative phase
proportion can be obtained by simple lever rule. For this, left-hand side and right-hand side of line X are noted.
For instance, at T2 temperature, the amount of crystals of A and liquid could be found by just measuring the
distances a and b i.e.

𝑏 (123)
% 𝑜𝑓 𝑐𝑟𝑦𝑠𝑡𝑎𝑙𝑠 𝑜𝑓 𝐴 = × 100
𝑎+𝑏
𝑎 (124)
% 𝑜𝑓 𝑙𝑖𝑞𝑢𝑖𝑑 𝑚𝑖𝑥𝑡𝑢𝑟𝑒 = × 100
𝑎+𝑏
Similarly, at T3 temperature, the amount of crystals of A and liquid could be found by just measuring the
distances c and d i.e.

𝑑 (125)
% 𝑜𝑓 𝑐𝑟𝑦𝑠𝑡𝑎𝑙𝑠 𝑜𝑓 𝐴 = × 100
𝑐+𝑑
𝑐 (126)
% 𝑜𝑓 𝑙𝑖𝑞𝑢𝑖𝑑 𝑚𝑖𝑥𝑡𝑢𝑟𝑒 = × 100
𝑐+𝑑
It is also worthy to note that the percentage composition of liquid at T3 65% compound A + 35% compound
B. Now if c = 2 and d = 3 then putting in equation (125, 126), we get 60% of crystal A and 40% of the liquid
mixture. This 60% solid is pure crystalline A while 40% liquid mixture is actually composed of 65% compound
A (26) + 35% compound B (14).
At this point, three possibilities of calculating the eutectic point arise when the cooling curve under
three different conditions. Now we need to study cooling curves at composition X, Y, pure A, and eutectic (let
eutectic be 60% compound A and 40% compound B).

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 315

2. Calculation of eutectic point:


If the composition of the liquid mixture in which compound A is 80% while compound B is 20%
(represented by the vertical line X). Therefore, it follows from the above-mentioned rule that after
crystallization we must get the same composition i.e. 80% (solid A) + 20% (solid B). Now, if the system
allowed to cool down and then the temperatures are plotted against time, we will get the cooling curves for the
same systems. Since above point 1 the system is completely liquid, it will cool down very rapidly in going
from point X to point 1. However, point 1, the crystalline of compound A will also start releasing a large
amount of heat, and therefore, the rate of cooling will be slow until point Q is reached. Now since the system
is getting solidified at eutectic temperature i.e. it is moving from Q to Q'; the temperature of the system will
remain constant during this conversion.

Figure 14. The cooling of a liquid mixture (compound A and compound B) when composition exists left
to the eutectic composition.

After point Q' both compounds will exist as solids, and therefore, the system will cool down below point Q'
along the vertical line X. The same type of curve will be obtained if we start from a composition left to the
“eutectic composition”. The only difference will be that the compound B will separate out first up to point Q,
and then the solidification of compound A will start.
3. Some additional points:
Besides the features we already discussed, there are also some additional remarks which are based
on the experimental results and can also be justified theoretically are given below.
i) Although the temperature corresponding to point ‘1’ depends upon the initial composition, the temperature
corresponding to Q to Q' always remains the same for compositions.
ii) If the composition of the mixture is more close to the eutectic composition, the portion 1–Q will be small
whereas the portion Q–Q'. The reason for this type of behavior lies in the diagram itself. Assume a composition
Y which is closer to the eutectic composition than composition X.

Copyright © Mandeep Dalal


316 A Textbook of Physical Chemistry – Volume I

Now, if the system allowed to cool down and then the temperatures are plotted against time, we will
get the cooling curves for the same systems. Since above point 2, the system is completely liquid, it will cool
down very rapidly in going from point Y to point 2. After point 2, the crystallization of compound A will also
start releasing a large amount of heat, and therefore, the rate of cooling will be slow until point Q is reached.
Now, since the composition X has less A and more B than composition Y, the solidification of A (up to point
Q) will take less time but solidification of B will take more time. This will make portion 2–Q smaller than 1–
Q but will make portion Q–Q' larger.

Figure 15. The cooling of a liquid mixture (compound A and Compound B) when compositions exist at
eutectic point or left to the eutectic composition.

Now since the system is getting solidified at eutectic temperature i.e. it is moving from Q to Q'; the temperature
of the system will remain constant during this conversion.
iii) When the liquid mixture composition is the same as that of “eutectic composition” (shown by line Z), then
the cooling will take place in all-liquid phase up to point Q first, which means that there will be no break from
point Z to Q.
However, after reaching point Q, the solidification of both compounds (A as well as B) will start to
take place. Now since point Q is at eutectic temperature, there will be no change in temperature until this
process completes. This will create a long halt (Q–Q') in the cooling curve since the whole liquid needs more
time to solidify than some fraction left in solid-liquid equilibria.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 317

 Some Typical Eutectic Systems


Some of the typical examples of two-components systems forming eutectic mixtures are given below
for a more comprehensive analysis.
1. Bismuth-cadmium system (Bi-Cd):
The Bi-Cd is a typical case of solid-liquid equilibria in a two-component system that form a eutectic
mixture. The phases involved in this case are solid Bi, solid Cd, liquid mixture of two (Cd + Bi) and gas-phase
also. Now since a minor pressure disturbance will have little to no effect on the system and all the phases in
two-component systems are either solid or liquids only (solid-liquid equilibria), Which is the reduced or
condensed phase rule i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (127)

Now because the total number of phases that can exist simultaneously are reduced to 3 (system becomes non-
variant at P = 3).
The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. Consider a liquid mixture of Bi and Cd at
temperature T. Now if this liquid mixture is allowed to cool down below the freezing point of the mixture, the
solid will start to separate out. Prepare a number of such mixtures but with different compositions (i.e. with
different ratios of Bi and Cd). The cooling of all the mixtures is carried out in open vessels so that the pressure
remains constant (atmospheric pressure). After that, the plot freezing points vs the composition is

Figure 16. The phase diagram of Bi-Cd system.

The discussion on different parts (curve OQ, Curve SQ and the eutectic point) of the phase diagram of Bi-Cd
system is given below.

Copyright © Mandeep Dalal


318 A Textbook of Physical Chemistry – Volume I

i) Curve OQ and SQ: The point O and point S represent the freezing point the pure Bi (271°C) and pure Cd
(321°C), respectively. When cadmium is added to bismuth, its freezing point of decreases regularly along OQ.
Likewise, bismuth is added to cadmium, its freezing point of decreases regularly along OQ. In conclusion, we
can say that the curve OQ and SQ represent the temperature-conditions at which solid Bi and solid Cd are in
equilibrium with the liquid mixture. Since there are only two phases involved, i.e. after putting C = 2 and P =2
in equation (127), we get

𝐹′ = 2 − 2 + 1 = 1 (128)

Which means that the system is univariate. In other words, only one condition is needed to be defined to define
the whole system; for instance, if we define a particular composition of Bi and Cd, the freezing point of the
mixture is completely fixed. Since the points on OQ and SQ represent the initial freezing temperature whereas
the points on solidus DE represent final freezing temperature; OQ and SQ can also be considered as the
solubility curves of Bi in molten Cd and Cd in molten Bi, respectively.
ii) Eutectic point (Q): At point ‘Q’, all the three phases (solid Bi + solid Cd + liquid) are in equilibrium with
each other; and therefore, zero degrees of freedom exist at this point i.e.

𝐹′ = 2 − 3 + 1 = 0 (129)

This means that we can neither change temperature nor composition without disturbing the mutual equilibrium
of solid Bi, solid Cd and liquid. Moreover, below this temperature (144°C), the mixture freezes as a whole.
This temperature is called as “eutectic point” and the composition corresponding to this point is called as
eutectic composition (40% Cd + 60% Bi).
2. Lead-Silver system (Pb-Ag):
The Pb-Ag is a typical case of solid-liquid equilibria in a two-component system that form eutectic
mixture. The phases involved in this case are solid Ag, solid Pb, liquid mixture of two (Ag + Pb) and gas-
phase also. Now since a minor pressure disturbance will have little to no effect on the system and all the phases
in two-component systems are either solid or liquids only (solid-liquid equilibria), Which is the reduced or
condensed phase rule i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (130)

Now because the total number of phases that can exist simultaneously are reduced to 3 (system becomes non-
variant at P = 3).
The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. Consider a liquid mixture of Ag and Pb at
temperature T. Now if this liquid mixture is allowed to cool down below the freezing point of the mixture, the
solid will start to separate out. Prepare a number of such mixtures but with different compositions (i.e. with
different ratios of Ag and Pb). The cooling of all the mixtures is carried out in open vessels so that the pressure
remains constant (atmospheric pressure). After that, the plot freezing points vs the composition is

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 319

Figure 17. The phase diagram of Pb-Ag system.

The discussion on different parts (curve OQ, Curve SQ and the eutectic point) of the phase diagram of Pb-Ag
system is given below.
i) Curve OQ and SQ: The point O and point S represent the freezing point the pure Ag (961°C) and pure Pb
(327°C), respectively. When lead is added to silver, its freezing point of decreases regularly along OQ.
Likewise, silver is added to lead, its freezing point of decreases regularly along OQ. In conclusion, we can say
that the curve OQ and SQ represent the temperature-conditions at which solid Ag and solid Pb are in
equilibrium with the liquid mixture. Since there are only two phases involved, i.e. after putting C = 2 and P =2
in equation (130), we get

𝐹′ = 2 − 2 + 1 = 1 (131)

Which means that the system is univariate. In other words, only one condition is needed to be defined to define
the whole system; for instance, if we define a particular composition of Ag and Pb, the freezing point of the
mixture is completely fixed. Since the points on OQ and SQ represent the initial freezing temperature whereas
the points on solidus DE represent final freezing temperature; OQ and SQ can also be considered as the
solubility curves of Ag in molten Pb and Pb in molten Ag, respectively.
ii) Eutectic point (Q): At point ‘Q’, all the three phases (solid Ag + solid Pb + liquid) are in equilibrium with
each other; and therefore, zero degree of freedom exists at this point i.e.

𝐹′ = 2 − 3 + 1 = 0 (132)

This means that we can neither change temperature nor composition without disturbing the mutual equilibrium
of solid Ag, solid Pb, and liquid. Moreover, below this temperature (303°C), the mixture freezes as a whole.
This temperature is called as “eutectic point” and the composition corresponding to this point is called as
eutectic composition (2.6% Ag + 97.4% Bi).

Copyright © Mandeep Dalal


320 A Textbook of Physical Chemistry – Volume I

3. Potassium iodide-water system (KI-H2O):


The KI-H2O is a typical case of solid-liquid equilibria in a two-component system that form eutectic
mixture. The phases involved in this case are solid KI, ice, liquid mixture of two (KI + H 2O) and gas-phase
also. Now since a minor pressure disturbance will have little to no effect on the system and all the phases in
two-component systems are either solid or liquids only (solid-liquid equilibria), Which is the reduced or
condensed phase rule i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (133)

Now because the total number of phases that can exist simultaneously are reduced to 3 (system becomes non-
variant at P = 3).
The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. Consider a liquid mixture of KI and H2O at
temperature T. Now if this liquid mixture is allowed to cool down below the freezing point of the mixture, the
solid will start to separate out. Prepare a number of such mixtures but with different compositions (i.e. with
different ratios of KI and H2O). The cooling of all the mixtures is carried out in open vessels so that the pressure
remains constant (atmospheric pressure). After that, the plot freezing points vs the composition is

Figure 18. The phase diagram of KI-H2O system.

The discussion on different parts (curve OQ, Curve SQ and the eutectic point) of phase diagram of KI-H2O
system is given below.
i) Curve OQ and SQ: The point O represents the freezing point pure water (0°C). However, the curve SQ is
incomplete because it is impossible to reach the melting point of KI in the presence of water. When KI is added
to water, its freezing point of decreases regularly along OQ. Likewise, when water is added to KI, its freezing
point of decreases regularly along SQ. In conclusion, we can say that the curve OQ and SQ represent the

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 321

temperature-conditions at which solid KI and solid H2O are in equilibrium with the liquid mixture. Since there
are only two phases involved, i.e. after putting C = 2 and P =2 in equation (133), we get

𝐹′ = 2 − 2 + 1 = 1 (134)

Which means that the system is univariate. Since the points on OQ and SQ represent the initial freezing
temperature whereas the points on solidus DE represent final freezing temperature; OQ and SQ can also be
considered as the solubility curves of KI in molten H2O and ice in molten KI, respectively.
ii) Eutectic point (Q): At point ‘Q’, all the three phases (solid KI + solid H2O + liquid) are in equilibrium with
each other; and therefore, zero degree of freedom exists at this point i.e.

𝐹′ = 2 − 3 + 1 = 0 (135)

This means that we can neither change temperature nor composition without disturbing the mutual equilibrium
of solid KI, ice and liquid. Moreover, below this temperature (−22°C), the mixture freezes as a whole. This
temperature is called as “eutectic point” and the composition corresponding to this point is called as eutectic
composition (52% KI + 48% H2O).

 Systems Forming Solid Compounds AxBy with Congruent and Incongruent


Melting Points
In some two-component systems, the participants react together to form solid compounds AxBy. On
the basis of the melting point of the compounds formed, these systems can be further divided.
 Systems Forming Solid Compounds AxBy with Congruent Melting Points
In these systems, the solid compound melts sharply at temperature T with the same composition as
in the initial solid. These compound are said to possess a congruent melting point with the phase diagram as

Figure 19. The general phase diagram of systems forming compounds with congruent melting points.

Copyright © Mandeep Dalal


322 A Textbook of Physical Chemistry – Volume I

Consider two components A and B which also form a chemical compound AB by reacting with each
other. Therefore, in the complete solid-state, there will be three phases named solid A, solid B and solid AB.
Accordingly, there should also be three different freezing point curves i.e. OP, PQR and RS. In addition to this
liquidus, there will be two solidus CE and DE as well. The significance of different parts of the above-depicted
diagram is discussed below.
1. Curve OP and RS: The point O and point S represent the freezing point the pure compound A and
compound B, respectively. When compound AB is added to component A, its freezing point decreases
regularly along OP. Likewise, when component AB is added to component B, its freezing point decreases
regularly along SR. In conclusion, we can say that the curve OP and SR represent the temperature-conditions
at which solid A and solid B are in equilibrium with the liquid mixture. Since there are only two phases
involved, and both are condensed in nature (solid and liquid), we need to use condensed or reduced phase rule
here i.e. after putting C = 2 and P =2 in equation (114), we get

𝐹′ = 2 − 2 + 1 = 1 (136)

Which means that the system is univariate. In other words, only one condition is needed to be defined to define
the whole system; for instance, if we define a particular composition of A and AB (or B and AB), the freezing
point of the mixture is completely fixed.
2. Curve PQR: When compound A is increased in compound AB, its freezing point of decreases regularly
along QP. Likewise, when component B is added to compound AB, its freezing point decreases regularly along
QR. In conclusion, we can say that the curve PQR represents the temperature-conditions at which compound
AB is in equilibrium with the liquid mixture. Since there are only two phases involved, and both are condensed
in nature (solid and liquid), we need to use condensed or reduced phase rule here i.e. after putting C = 2 and P
=2 in equation (114), we get

𝐹′ = 2 − 2 + 1 = 1 (137)

However, it is also worthy to mention that at the point Q also represents the congruent melting point of
compound AB because liquid and solid phases have the same compositions. Consequently, we can also
conclude that the system becomes one-component system at this point because the solid as well liquid phases
contain only the compound AB alone. Furthermore, the congruent melting point of compound AB may lie
above or below the congruent melting points of component A and component B.
3. Eutectic points P and R: In order to explain that the liquid phase can have two different compositions in
equilibrium with the solid phase i.e. x and x', the curve PQR is divided into two parts by the vertical line QQ'.
This makes the concept very simple as the left and right parts can be treated as simple eutectic systems
separately. The right half of the diagram is a eutectic system with components B and AB (solidus CD); whereas
left hand side is a eutectic system with components A and AB (solidus is EF). The eutectic temperature and
eutectic composition for the left-hand side portion are given by point P. Similarly, the eutectic temperature
and eutectic composition for the right-hand side portion are given by point R.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 323

 Some Typical Examples of Systems Forming Compounds with Congruent Melting Points
Some of the typical examples of two-components systems forming compounds with congruent
melting points are given below for a more comprehensive analysis.
1. Magnesium-zinc system (Mg-Zn):
The Mg-Zn is a typical case of solid-liquid equilibria in a two-component system that form
compounds with congruent melting points. The phases involved in this case are solid Mg, solid Zn, solid
MgZn2, liquid mixture of three (Mg + Zn + MgZn2) and vapor phase too. Now since a minor pressure
disturbance will have little to no effect on the system and all the phases in two-component systems are either
solid or liquids only (solid-liquid equilibria), which is the reduced or condensed phase rule i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (138)

The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. Consider a liquid mixture of Mg and Zn at
temperature T. Now if this liquid mixture is allowed to cool down below the freezing point of the mixture, the
solid will start to separate out. Prepare a number of such mixtures but with different compositions (i.e. with
different ratios of Mg and Zn). The cooling of all the mixtures is carried out in open vessels so that the pressure
remains constant (atmospheric pressure). After that, the plot freezing points vs the composition is

Figure 20. The phase diagram of Mg-Zn system.

Copyright © Mandeep Dalal


324 A Textbook of Physical Chemistry – Volume I

The discussion on different parts (curve OQ, Curve SQ and the eutectic point) of the phase diagram
of Bi-Cd system is given below.
i) Curve OP and point P: The point O represents the freezing point of the pure Zn (420°C). When Mg is added
to zinc, its freezing point of decreases regularly along OP and Zn separates out simultaneously. Since there are
only two phases involved, i.e. after putting C = 2 and P =2 in equation (138), we get

𝐹′ = 2 − 2 + 1 = 1 (139)

Which means that the system is univariate. However, after the point P is reached, the compound MgZn2 is
formed and also starts separating out as solid. We can say that there are three phases that coexist at point P i.e.
solid Zn, solid MgZn2 and melt which makes the system invariant (for P = 3, F' = 0).
ii) Curve PQ and point Q: The point Q represents the freezing point of the compound MgZn2 (590°C). When
Mg is added to zinc after point P, it combines with zinc to form MgZn2 which keeps on separating and its
freezing point increases regularly until point Q is reached (33% magnesium). Since there are only two phases
involved, i.e. after putting C = 2 and P =2 in equation (138), we get

𝐹′ = 2 − 2 + 1 = 1 (140)

Which means that the system is univariate. Now since the liquid and solid phases have the same composition
at point Q, the corresponding temperature can be called a congruent melting point of compound MgZn2.
Furthermore, as the number of components becomes one at point Q i.e. solid MgZn 2 and melt MgZn2, the
system becomes invariant at Q (for P = 3, F' = 0).
iii) Curve QR and point R: When Mg is further added to zinc after point Q, it goes into melt MgZn2 separating
out as solid, and therefore, the freezing point of MgZn2 decreases regularly until point R is reached. Since there
are only two phases involved along curve QR, i.e. after putting C = 2 and P =2 in equation (138), we get the
following

𝐹′ = 2 − 2 + 1 = 1 (141)

Which means that the system is univariant. Moreover, we can also conclude here that there are three phases
that coexist at point R i.e. solid Mg solid MgZn2 and melt which makes the system invariant (for P = 3, F' =
0).
iv) Curve SR: When Mg is further added to zinc after point R, it starts separating out as solid, and therefore,
the freezing point of Mg increases regularly until point S is reached. Since there are only two phases involved
along curve SR, i.e. after putting C = 2 and P =2 in equation (138), we get

𝐹′ = 2 − 2 + 1 = 1 (142)

Which means that the system remains univariant along curve SR. in reverse we can also say that the freezing
point of Mg decreases regularly along curve SR until point R is reached i.e. solid Mg solid MgZn2 and melt
which makes the system invariant (for P = 3, F' = 0).

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 325

2. Ferric chloride-water system (FeCl3-H2O):


The FeCl3-H2O is another typical case of solid-liquid equilibria in a two-component system that forms
stable compounds with congruent melting point. The phases involved in this case are solid FeCl3, ice, solid
Fe2Cl6.12H2O, Fe2Cl6.7H2O, Fe2Cl6.5H2O, Fe2Cl6.4H2O, liquid mixture and vapor phase too. Now since a
minor pressure disturbance will have little to no effect on the system and all the phases in two-component
systems are either solid or liquids only (solid-liquid equilibria), the reduced phase rule can be used i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (143)

The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. The cooling of all the mixtures is carried out in
open vessels so that the pressure remains constant (atmospheric pressure). After that, the plot of freezing points
vs the composition is obtained.

Figure 21. The phase diagram of FeCl3-H2O system.

i) Curve OP and point P: The point O represents the freezing point of the pure water (0°C). When FeCl3 is
added to water, the freezing point of water decreases regularly along OP and ice separates out simultaneously.
Since there are only two phases involved along curve OP (C = 2 and P =2), the equation (143) gives

𝐹′ = 2 − 2 + 1 = 1 (144)

Which means that the system is univariant along curve OP. However, after reaching the point P, the liquid
phase becomes saturated with compound Fe2Cl6.12H2O which also starts separating out as solid afterward. We
can say that there are three phases that coexist at point P i.e. solid ice, solid Fe2Cl6.12H2O and solution which

Copyright © Mandeep Dalal


326 A Textbook of Physical Chemistry – Volume I

makes the system invariant (for P = 3, F' = 0). In other words, the point P is the “eutectic point for water and
Fe2Cl6.12H2O.
ii) Curve PQ and point Q: The point Q represents the congruent melting point of the compound Fe2Cl6.12H2O
(37°C). When FeCl3 is added to the water after point P, it combines with water to form Fe2Cl6.12H2O which
keeps on separating and its freezing point increases regularly until point Q is reached. Since there are only two
phases involved along curve PQ, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (145)

Which means that the system is univariant. Furthermore, as the number of components becomes one at point
Q i.e. solid Fe2Cl6.12H2O and solution, the system becomes invariant at Q (for P = 2, F' = 0).
iii) Curve QR and point R: When FeCl3 is further added to the water after point Q, the freezing point of
Fe2Cl6.12H2O decreases regularly until point R is reached. Since there are only two phases involved along
curve QR, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (146)

Which means that the system is univariant. After point R, the compound Fe2Cl6.7H2O is formed and also starts
separating out as solid. However, we can say that there are three phases that coexist at point R i.e. solid
Fe2Cl6.12H2O, solid Fe2Cl6.7H2O and solution which makes the system invariant (for P = 3, F' = 0).
iv) Curve RS and point S: The point S represents the congruent melting point of the compound Fe2Cl6.7H2O
(37°C). When FeCl3 is added to the water after point R, it combines with water to form Fe2Cl6.7H2O which
keeps on separating and its freezing point increases regularly until point S is reached. Since there are only two
phases involved, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (147)

Which means that the system is univariant. Furthermore, as the number of components becomes one at point
S i.e. solid Fe2Cl6.7H2O and solution, the system becomes invariant at S (for P = 2, F' = 0).
v) Curve ST and point T: When FeCl3 is further added to the water after point S, the freezing point of
Fe2Cl6.7H2O decreases regularly until point T is reached. Since there are only two phases involved along curve
ST, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (148)

Which means that the system is univariant. After point T, the compound Fe2Cl6.5H2O is formed and also starts
separating out as solid. However, we can say that there are three phases that coexist at point T i.e. solid
Fe2Cl6.7H2O, solid Fe2Cl6.5H2O and solution which makes the system invariant (for P = 3, F' = 0).
vi) Curve TU and point U: The point U represents the congruent melting point of the compound Fe2Cl6.5H2O
(37°C). When FeCl3 is added to the water after point T, it combines with water to form Fe 2Cl6.5H2O which

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 327

keeps on separating and its freezing point increases regularly until point U is reached. Since there are only two
phases involved, i.e. after putting C = 2 and P =2 in equation (138), we get

𝐹′ = 2 − 2 + 1 = 1 (149)

Which means that the system is univariant. Furthermore, as the number of components becomes one at point
U i.e. solid Fe2Cl6.5H2O and solution, the system becomes invariant at U (for P = 2, F' = 0).
vii) Curve UV and point V: When FeCl3 is further added to the water after point U, the freezing point of
Fe2Cl6.5H2O decreases regularly until point V is reached. Since there are only two phases involved along curve
UV, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (150)

Which means that the system is univariant. After point V, the compound Fe2Cl6.4H2O is formed and also starts
separating out as solid. However, we can say that there are three phases that coexist at point V i.e. solid
Fe2Cl6.5H2O, solid Fe2Cl6.4H2O and solution which makes the system invariant (for P = 3, F' = 0).
viii) Curve VW and point W: The point W represents the congruent melting point of the compound Fe2Cl6.4H2O
(37°C). When FeCl3 is added to the water after point V, it combines with water to form Fe 2Cl6.4H2O which
keeps on separating and its freezing point increases regularly until point W is reached. Since there are only
two phases involved, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (151)

Which means that the system is univariant. Furthermore, as the number of components becomes one at point
W i.e. solid Fe2Cl6.4H2O and solution, the system becomes invariant at Q (for P = 2, F' = 0).
ix) Curve WX and point X: When FeCl3 is further added to the water after point W, the freezing point of
Fe2Cl6.4H2O decreases regularly until point X is reached. Since there are only two phases involved along curve
WX, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (152)

Which means that the system is univariant. After point X, anhydrous Fe2Cl6 is formed and also starts separating
out as solid. However, we can say that there are three phases that coexist at point R i.e. solid Fe2Cl6.4H2O,
anhydrous Fe2Cl6 and solution which makes the system invariant (for P = 3, F' = 0).
x) Curve XY: The point Y represents the freezing point of the compound Fe2Cl6. When FeCl3 is added to the
water after point X, it combines with water to yield anhydrous Fe2Cl6 which keeps on separating and its
freezing point increases regularly along XY is reached. Since there are only two phases involved along curve
XY, i.e. after putting C = 2 and P =2 in equation (143), we get

𝐹′ = 2 − 2 + 1 = 1 (153)

Which means that the system is univariant.

Copyright © Mandeep Dalal


328 A Textbook of Physical Chemistry – Volume I

 Systems Forming Solid Compounds AxBy with Incongruent Melting Points


In these types of systems, the two components react to give a compound which is not stable up to its
melting point. When heated, the decomposition starts before the melting point is reached; and a new solid
phase and a solution or melt with a different composition from the original solid are formed. Such compounds
are said to undergo peritectic or transition reaction are labeled to have congruent melting point. A typical
transition can be represented as

𝐶1 ⇌ 𝐶2 + melt or solution (154)

Where C1 is the compound formed by the reaction between participating components whereas C2 represents
the compound formed as a result of decomposition of C1 below its fusion temperature.

Figure 22. The general phase diagram of systems forming compounds with incongruent melting points.

Consider two components A and B which also form a chemical compound AB2 by reacting with each
other. The cooling of all the mixture-compositions is carried out in open vessels so that the pressure remains
constant (atmospheric pressure). After that, the plot of freezing points vs the composition is obtained. The
significance of different parts of the above-depicted diagram is discussed below.
1. Point O, S, Q and R: The point O and point S represent the freezing point the pure compound A and
compound B, respectively. The point Q, however, is quite strange because it represents the transition
temperature (incongruent melting point of AB2) where compound AB2 decomposes into compound B. If does
not decompose at this temperature, its congruent melting point would be R. In other words, we can say that the
point R represents the hypothetical congruent melting point of compound AB2.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 329

2. Curve OP, SQ and QP: When compound B is added to component A, its freezing point decreases regularly
along OP. In other words, OP is the fusion curve of compound A along which solid A is in equilibrium with
melt or solution. The line SQ is the fusion curve of compound B along which solid B is in equilibrium with
melt or solution. Similarly, QP is the fusion curve of compound AB2 along which solid AB2 is in equilibrium
with melt or solution.
When a liquid mixture with composition X is allowed to cool down, it will do so by keeping its
composition the same until point 1 is reached where the compound A will just start to separate out as solid.
Further cooling will lead to a change in composition along the line 1P. When point P is attained, the formation
of compound AB will be started; and since three phases coexist at this point (solid A, solid AB and liquid), the
reduced phase rule gives

𝐹′ = 𝐶 − 𝑃 + 1 (155)

𝐹′ = 2 − 3 + 1 = 0 (156)

Which means that there will be no degree of freedom at point P (P is non-variant). On the other hand, if liquid
mixture with composition Y is allowed to cool down, it will do so by keeping its composition the same until
point 2 is reached where the compound B will just start to separate out as solid. Further cooling will lead to a
change in composition along the line 2Q. When point Q is attained, the following meritectic reaction will take
place

𝑆𝑜𝑙𝑖𝑑 𝐵 + 𝑆𝑜𝑙𝑢𝑡𝑖𝑜𝑛 ⇌ 𝑆𝑜𝑙𝑖𝑑 𝐴𝐵2 (157)

Therefore, the formation of compound AB2 will be started; and since three phases coexist at this point (solid
B, solid AB and liquid), the point Q also become invariant. Furthermore, it is also worthy to mention that the
transformation of compound B to compound AB2 occurs at a constant temperature, and therefore, the point Q
is also called a peritectic point.
 Some Typical Examples of Systems Forming Compounds with Incongruent Melting Points
Some of the typical examples of two-components systems forming compounds with incongruent
melting points are given below for a more comprehensive analysis.
1. Sodium chloride-water system (NaCl-H2O):
The NaCl-H2O is a typical case of solid-liquid equilibria in a two-component system that forms
compounds with an incongruent melting point. The phases involved in this case are solid NaCl, solid
NaCl.2H2O, ice, liquid mixture, and vapour phase too. Now since a minor pressure disturbance will have little
to no effect on the system and all the phases in two-component systems are either solid or liquids only (solid-
liquid equilibria), which is the reduced or condensed phase rule i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (158)

The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. Consider a liquid mixture of water and NaCl at

Copyright © Mandeep Dalal


330 A Textbook of Physical Chemistry – Volume I

temperature T. Now if this liquid mixture is allowed to cool down below the freezing point of the mixture, the
solid will start to separate out. Prepare a number of such mixtures but with different compositions (i.e. with
different ratios of H2O and NaCl). The cooling of all the mixtures is carried out in open vessels so that the
pressure remains constant (atmospheric pressure). After that, the plot freezing points vs the composition is

Figure 23. The phase diagram of NaCl-H2O system.

i) Point O, S and Q: The point O and point S represent the freezing point the pure water and pure NaCl,
respectively. The point Q, however, is quite strange because it represents the transition temperature
(incongruent melting point of NaCl.2H2O) where compound NaCl.2H2O decomposes into pure NaCl.
ii) Curve OP, SQ and QP: When NaCl is added to water, its freezing point decreases regularly along OP. In
other words, OP is the fusion curve of water along which the ice is in equilibrium with the solution. The line
SQ is the fusion curve of pure NaCl along which solid NaCl is in equilibrium with the brine solution. Similarly,
QP is the fusion curve of compound NaCl.2H2O along which solid NaCl.2H2O is in equilibrium with solution.
When NaCl is added to pure water, ice will just start to separate out as solid along path OP. This will
lead to a change in composition along the line OP. When point P is attained, the formation of compound
NaCl.2H2O will start; and since three phases coexist at P (ice, NaCl.2H2O and liquid), the reduced phase rule

𝐹′ = 2 − 3 + 1 = 0 (159)

Which means that there will be no degree of freedom at point P (P is non-variant). On the other hand, the curve
SQ is the fusion curve of NaCl. The further cooling will lead to a change in composition along the line SQ.
When point Q is attained, the following meritectic reaction will take place

𝑆𝑜𝑙𝑖𝑑 𝑁𝑎𝐶𝑙 + 𝑆𝑜𝑙𝑢𝑡𝑖𝑜𝑛 ⇌ 𝑆𝑜𝑙𝑖𝑑 𝑁𝑎𝐶𝑙. 2𝐻2 𝑂 (160)

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 331

Therefore, the formation of compound NaCl.2H2O will be started; and since three phases coexist at this point
(solid NaCl, solid NaCl.2H2O and liquid), the point Q also become invariant. Furthermore, it is also worthy to
mention that the transformation of compound NaCl to compound NaCl.2H2O occurs at a constant temperature,
and therefore, the point Q is also called as peritectic point.
2. Sodium sulphate-water system (Na2SO4-H2O):
The Na2SO4-H2O is another typical case of solid-liquid equilibria in a two-component system that
forms compounds with an incongruent melting point. The phases involved in this case are solid Na2SO4, solid
Na2SO4.10H2O, ice, liquid mixture, and vapor phase too. Now since a minor pressure disturbance will have
little to no effect on the system and all the phases in two-component systems are either solid or liquids only
(solid-liquid equilibria), which is the reduced or condensed phase rule i.e.

𝐹′ = 𝐶 − 𝑃 + 1 (161)

The complete phase diagram can be drawn on two dimensional paper with vertical and horizontal
sides representing temperature and composition, respectively. Consider a liquid mixture of water and Na2SO4
at temperature T. Now if this liquid mixture is allowed to cool down below the freezing point of the mixture,
the solid will start to separate out. Prepare a number of such mixtures but with different compositions (i.e. with
different ratios of H2O and Na2SO4). The cooling of all the mixtures is carried out in open vessels so that the
pressure remains constant (atmospheric pressure). After that, the plot freezing points vs the composition is

Figure 24. The phase diagram of Na2SO4-H2O system.

i) Point O and Q: The point O represents the freezing point the pure water. The point Q, however, is quite
strange because it represents the transition temperature (incongruent melting point of Na 2SO4.10H2O) where
compound Na2SO4.10H2O decomposes into pure Na2SO4.

Copyright © Mandeep Dalal


332 A Textbook of Physical Chemistry – Volume I

ii) Curve OP, SQ and QP: When Na2SO4 is added to water, its freezing point decreases regularly along OP. In
other words, OP is the fusion curve of compound water along which ice is in equilibrium with the solution.
The line SQ is the fusion curve of pure Na2SO4 along which Na2SO4 is in equilibrium with the solution.
Similarly, QP is the fusion curve of compound Na2SO4.10H2O along which solid Na2SO4.10H2O is in
equilibrium with the solution.
When Na2SO4 is added to water, ice will start to separate out along OP. This will lead to a change in
composition along the line OP. When point P is attained, the formation of compound Na2SO4.10H2O will be
started; and since three phases coexist at P (ice, Na2SO4-H2O and solution), the reduced phase rule gives

𝐹′ = 2 − 3 + 1 = 0 (162)

Which means that there will be no degree of freedom at point P (P is non-variant). On the other hand, the curve
SQ is fusion curve of Na2SO4. The further cooling will lead to a change in composition along the line SQ.
When point Q is attained, the following meritectic reaction will take place

𝑆𝑜𝑙𝑖𝑑 𝑁𝑎2 𝑆𝑂4 + 𝑆𝑜𝑙𝑢𝑡𝑖𝑜𝑛 ⇌ 𝑆𝑜𝑙𝑖𝑑 𝑁𝑎2 𝑆𝑂4 . 10𝐻2 𝑂 (163)

Therefore, the formation of compound Na2SO4.10H2O will be started; and since three phases coexist at this
point (solid Na2SO4, solid Na2SO4.10H2O and liquid), the point Q also become invariant. Furthermore, it is
also worthy to mention that the transformation of compound Na2SO4 to compound Na2SO4.10H2O occurs at a
constant temperature, and therefore, the point Q is also called a peritectic point.

 Phase Diagram and Thermodynamic Treatment of Solid Solutions


In these types of systems, the components are completely miscible with each other in solid phase and
completely homogeneous solid solutions are produced. The X-ray diffraction studies are typically employed
to check that single crystalline phase is obtained rather than a mixture of two solid phases. In order to draw
and understand the phase diagrams of solid solutions, we need to discuss the same in a comprehensive
thermodynamic framework first.
 General Thermodynamic Treatment of Solid Solutions
The general thermodynamic treatment of solid solutions includes the shift in the solvent’s freezing
point during the crystallization and the nature of the cooling curve as well. The necessary discussion on both
concepts is given below.
1. The shifting of solvent’s freezing point during crystallization: The thermodynamic expression for the
shift of solvent’s freezing point when the solid begins to solidify during cooling can be obtained by assuming
the solid solution as an ideal solution. Therefore, the chemical potential for ith constituent (μi) can be given by
the following relation.

𝜇𝑖 = 𝜇𝑖∗ + 𝑅𝑇 ln 𝑥𝑖 (164)

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 333

Where 𝑥𝑖 represents the fractional composition of the constituent whereas 𝜇𝑖∗ is the chemical potential of the
pure solid. After setting the primary condition for the phase equilibria between solid and liquid, we have

𝜇1 (𝑠) = 𝜇1 (𝑙) (165)

Considering both as ideal, we can write above equation using equation (164) as

𝜇1∗ (𝑠) + 𝑅𝑇 ln 𝑥1 (𝑠) = 𝜇1∗ (𝑙) + 𝑅𝑇 ln 𝑥1 (𝑙) (166)

After rearranging, we get

𝑅𝑇 ln 𝑥1 (𝑠) − 𝑅𝑇 ln 𝑥1 (𝑙) = 𝜇1∗ (𝑙) − 𝜇1∗ (𝑠) (167)

𝑅𝑇[ln 𝑥1 (𝑠) − ln 𝑥1 (𝑙)] = 𝜇1∗ (𝑙) − 𝜇1∗ (𝑠) (168)

𝑥1 (𝑠) 𝜇1∗ (𝑙) − 𝜇1∗ (𝑠) (169)


𝑅 ln =
𝑥1 (𝑙) 𝑇

Now since 𝜇1∗ (𝑙) − 𝜇1∗ (𝑠) represents the molar free energy of the fusion for the pure solvent at p pressure and
T temperature, the equation can also be written as

𝑥1 (𝑠) 𝛥𝑓𝑢𝑠 𝜇1∗ (170)


𝑅 ln =
𝑥1 (𝑙) 𝑇

Now putting the value of molar free energy of fusion (Δ𝑓𝑢𝑠 𝜇1∗ = Δ𝑓𝑢𝑠 𝐻1,𝑚
∗ ∗
− 𝑇Δ𝑓𝑢𝑠 𝑆1,𝑚 ) in the above
equation, we get

𝑥1 (𝑠) Δ𝑓𝑢𝑠 𝐻1,𝑚 ∗
− 𝑇Δ𝑓𝑢𝑠 𝑆1,𝑚 (171)
𝑅 ln =
𝑥1 (𝑙) 𝑇

𝑥1 (𝑠) Δ𝑓𝑢𝑠 𝐻1,𝑚 ∗ (172)
𝑅 ln = − Δ𝑓𝑢𝑠 𝑆1,𝑚
𝑥1 (𝑙) 𝑇

Recalling the entropy of fusion at for the pure solvent at a temperature 𝑇1∗ (melting point) i.e.



Δ𝑓𝑢𝑠 𝐻1,𝑚 (173)
Δ𝑓𝑢𝑠 𝑆1,𝑚 =
𝑇1∗

Using the above result in equation (172), we get



𝑥1 (𝑠) Δ𝑓𝑢𝑠 𝐻1,𝑚 ∗
Δ𝑓𝑢𝑠 𝐻1,𝑚 (174)
𝑅 ln = −
𝑥1 (𝑙) 𝑇 𝑇1∗

𝑥1 (𝑠) ∗
1 1 (175)
𝑅 ln = Δ𝑓𝑢𝑠 𝐻1,𝑚 [ − ∗]
𝑥1 (𝑙) 𝑇 𝑇1

or

Copyright © Mandeep Dalal


334 A Textbook of Physical Chemistry – Volume I

𝑥1 (𝑠) ∗
𝑇1∗ − 𝑇 (176)
𝑅 ln = Δ𝑓𝑢𝑠 𝐻1,𝑚 [ ]
𝑥1 (𝑙) 𝑇𝑇1∗

After putting 𝑇 − 𝑇1∗ = ∆𝑇𝑓 , the above equation takes the form

𝑥1 (𝑠) ∗
−∆𝑇𝑓 (177)
𝑅 ln = Δ𝑓𝑢𝑠 𝐻1,𝑚 [ ]
𝑥1 (𝑙) 𝑇𝑇1∗

If the solution is very dilute, the melting point 𝑇1∗ will be quite close to temperature T; therefore, 𝑇𝑇1∗ can be
replaced by 𝑇1∗ 2 . Therefore, the equation (177) takes the form


−∆𝑇𝑓 (178)
𝑅[ln 𝑥1 (𝑠) − ln 𝑥1 (𝑙)] = Δ𝑓𝑢𝑠 𝐻1,𝑚 [ ∗2 ]
𝑇1

Furthermore, for very dilute solutions, we have

ln 𝑥1 (𝑠) = ln [1 − 𝑥2 (𝑠)] ≈ −𝑥2 (𝑠) (179)

ln 𝑥1 (𝑙) = ln [1 − 𝑥2 (𝑙)] ≈ −𝑥2 (𝑙) (180)

Using equation (179, 180) in equation (178), we get


−∆𝑇𝑓 (181)
𝑅[−𝑥2 (𝑠) + 𝑥2 (𝑙)] = Δ𝑓𝑢𝑠 𝐻1,𝑚 [ ∗2 ]
𝑇1

or

𝑅 𝑇1∗ 2 (182)
−∆𝑇𝑓 = ∗ [𝑥2 (𝑙) − 𝑥2 (𝑠)]
Δ𝑓𝑢𝑠 𝐻1,𝑚

𝑅 𝑇1∗ 2 𝑥2 (𝑠) (183)


−∆𝑇𝑓 = ∗ 𝑥2 (𝑙) [1 − ]
Δ𝑓𝑢𝑠 𝐻1,𝑚 𝑥2 (𝑙)

Recalling the expression for molality i.e.

𝑛2 (𝑙) 𝑛2 (𝑙) 𝑛2 (𝑙) 𝑛2 (𝑙) (184)


𝑥2 (𝑙) = ≈ = =[ ] 𝑀 = 𝑚𝑀1
𝑛1 (𝑙) + 𝑛2 (𝑙) 𝑛1 (𝑙) 𝑚1 (𝑙)/𝑀1 𝑚1 (𝑙) 1

Using the above result in equation (183), we have

𝑅 𝑇1∗ 2 𝑀1 𝑥2 (𝑠) (185)


−∆𝑇𝑓 = ∗ 𝑚 [1 − ]
Δ𝑓𝑢𝑠 𝐻1,𝑚 𝑥2 (𝑙)

or

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 335

∆𝑇𝑓 = −𝐾𝑓 𝑚[1 − 𝐾] (186)

Where K is the distribution coefficient given by 𝑥2 (𝑠)/𝑥2 (𝑙). In other words, the distribution coefficient the
ratio of the fractional amount of the solute in the solid phase to the fractional amount of the solute in the
solution phase.
Therefore, we can conclude that the value of ∆𝑇𝑓 can be negative or positive depending upon the
magnitude of K.
i) When ∆𝑇𝑓 is negative: The magnitude of ∆𝑇𝑓 will be negative if K < 1; which means that 𝑇 < 𝑇1∗ . The
physical significance is that depression in the freezing point of the solvent will be observed.
ii) When ∆𝑇𝑓 is positive: The magnitude of ∆𝑇𝑓 will be positive if K > 1; which means that 𝑇 > 𝑇1∗ . The
physical significance is that a less elevation in the freezing point of the solvent will be observed.
The two above-mentioned conditions can also be can be summarized in one statement that if the addition of a
component induces a decline in the freezing point of the solvent, it fractional amount must be greater in liquid
phase than in the solids phase, and vice-versa.
2. The nature of cooling curve of the solid solution: Now because the two components of a solid solution
are completely miscible with each other, the maximum number of phases which can coexist is only two (solid
solution + liquid). Therefore, recalling reduced phase

𝐹′ = 𝐶 − 𝑃 + 1 (187)

After putting C = 2 and P = 2, we get

𝐹′ = 2 − 2 + 1 = 1 (188)

Which means that there is only one degree of freedoms i.e. univariant system. In other words, only one variable
is needed to be defined to define the system completely which can either be temperature or the composition.
Therefore, we can conclude that the solid-liquid equilibria can exist at different conditions of temperatures.
Also, the composition of the two solutions will be fixed at a given temperature.
Furthermore, there will be two breaks in the cooling curve; one at the start of the freezing of the solid
solution, and another at the end of the freezing of solid solution. However, if the composition of the solid
solution is exactly same to the solid solution, it will become a one-component system, and therefore, after
putting C = 1 and P = 2 in equation (187), we get

𝐹′ = 1 − 2 + 1 = 0 (189)

Which means that there is no degree of freedoms i.e. non-variant system. In other words, we will not be able
to change temperature or composition without disturbing the state of equilibrium. Furthermore, the melting or
freezing process at this point will occur at constant temperature, which in turn, would result in arrests in
corresponding cooling curve.

Copyright © Mandeep Dalal


336 A Textbook of Physical Chemistry – Volume I

 General Discussion on the Phase Diagrams of Solid Solutions


Depending upon the value of 𝐾𝐴 𝑖𝑛 𝐵 and 𝐾𝐵 𝑖𝑛 𝐴 , the phase diagrams of very dilute solid solutions of
component A in B or B in A can primary be classified into three categories. A general discussion on these
three types of solid solutions is given below.
1. Ascending solid solutions:
In these types of solutions, 𝐾𝐵 𝑖𝑛 𝐴 > 1 (B is solute and A is solvent i.e. 𝑥𝐵 (𝑠)/𝑥𝐵 (𝑙)) and 𝐾𝐴 𝑖𝑛 𝐵 <
1 (A is solute and B is solvent i.e. 𝑥𝐴 (𝑠)/𝑥𝐴 (𝑙)). Furthermore, the solid solutions with 𝐾𝐵 𝑖𝑛 𝐴 < 1 and
𝐾𝐴 𝑖𝑛 𝐵 > 1 also fall into this category. The freezing points of such solutions exist in-between the freezing
points of pure solvents. Now, in order to draw the phase diagram of such solid solutions, we need to understand
nature of the solidus and liquidus in the thermodynamic framework first. To do so, rearrange equation (175).

𝑥1 (𝑙) ∗
1 1 (190)
𝑅 ln = −Δ𝑓𝑢𝑠 𝐻1,𝑚 [ − ∗]
𝑥1 (𝑠) 𝑇 𝑇1

or

𝑥1 (𝑙) (191)
= 𝑒 −𝑎
𝑥1 (𝑠)

Where the parameter a is defined as



Δ𝑓𝑢𝑠 𝐻1,𝑚 1 1 (192)
𝑎= [ − ∗]
𝑅 𝑇 𝑇1

Likewise, the second exponent can be written as

𝑥2 (𝑙) (193)
= 𝑒 −𝑏
𝑥2 (𝑠)

with

Δ𝑓𝑢𝑠 𝐻1,𝑚 1 1 (194)
𝑏= [ − ∗]
𝑅 𝑇 𝑇2

For solid and liquid phases, we have

𝑥1 (𝑙) = 1 − 𝑥2 (𝑙) (195)

and

𝑥1 (𝑠) = 1 − 𝑥2 (𝑠) (196)

Now putting the value of equation (196) in equation (191), we get

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 337

𝑥1 (𝑙) (197)
= 𝑒 −𝑎
1 − 𝑥2 (𝑠)

Now putting the value of 𝑥2 (𝑠) from equation (193), we get

𝑥1 (𝑙) (198)
= 𝑒 −𝑎
1 − 𝑥2 (𝑙)/𝑒 −𝑏

After putting the value of 𝑥2 (𝑙) from equation (195), we get

𝑥1 (𝑙) (199)
= 𝑒 −𝑎
1 − [1 − 𝑥1 (𝑙)]/𝑒 −𝑏

The solution of the above equation for 𝑥1 (𝑙) gives

𝑒 −𝑎 (𝑒 −𝑏 − 1) (200)
𝑥1 (𝑙) =
𝑒 −𝑏 − 𝑒 −𝑎
Using the above result in equation (191), we get

𝑒 −𝑏 − 1 (201)
𝑥1 (𝑠) = −𝑏
𝑒 − 𝑒 −𝑎
Similarly, we can solve for 𝑥2 (𝑙) and 𝑥2 (𝑙) as

𝑒 −𝑏 (𝑒 −𝑎 − 1) (202)
𝑥2 (𝑙) = 1 − 𝑥1 (𝑙) =
𝑒 −𝑎 − 𝑒 −𝑏
and

𝑒 −𝑎 − 1 (203)
𝑥2 (𝑠) = 1 − 𝑥1 (𝑠) =
𝑒 −𝑎 − 𝑒 −𝑏
It is obvious from the equation (200-203) that none of them is linear; and the expressions for 𝑥1 (𝑙) and 𝑥1 (𝑠)
are different from each other excepting at 𝑇1∗ and 𝑇2∗ only. At any temperature T that lies in the range of 𝑇1∗ and
𝑇2∗, a is negative (𝑒 −𝑎 > 1) and b is positive (𝑒 −𝑏 < 1). Therefore, at this point, we can conclude from equation
(200) that

𝑒 −𝑎 (𝑒 −𝑏 − 1) 𝑁𝑒𝑔𝑎𝑡𝑖𝑣𝑒 (204)
𝑥1 (𝑙) = = = 𝑃𝑜𝑠𝑖𝑡𝑖𝑣𝑒 𝑣𝑎𝑙𝑢𝑒
𝑒 −𝑏 − 𝑒 −𝑎 𝑃𝑜𝑠𝑖𝑡𝑖𝑣𝑒
Similarly, the equations (201-203) can also be proved to have positive values. The physical significance of
these results can be summarized by the statement that the physically meaningful compositions are obtained if,
and only if, the temperature of the system (T) lies in-between the 𝑇1∗ and 𝑇2∗.
i) If 𝑇1∗ ≤ 𝑇 ≤ 𝑇2∗ : The equations (191, 193) result in the following conclusions.

Copyright © Mandeep Dalal


338 A Textbook of Physical Chemistry – Volume I

𝑥1 (𝑙) (205)
>1
𝑥1 (𝑠)

𝑥2 (𝑙) (206)
<1
𝑥2 (𝑠)

ii) If 𝑇2∗ ≤ 𝑇 ≤ 𝑇1∗: The equations (191, 193) result in the following conclusions.

𝑥1 (𝑙) (207)
<1
𝑥1 (𝑠)

𝑥2 (𝑙) (208)
>1
𝑥2 (𝑠)

Hence, we can conclude that if the liquid is richer that solid, a depression in the freezing point will be observed;
whereas if solid is richer than liquid, an elevation in the freezing point will be observed. The general phase
diagram is for such systems is given below.

Figure 25. The phase diagram of ascending solid solutions.

The overall phase diagram can be easily understood by analyzing the cooling curve given in the left-hand side
of the diagram. Consider a liquid composition represented by the vertical line abcde which is allowed to cool
down. The system will maintain its liquid state until point b is attained where the crystals of the solid solution
will start to form. The point corresponding to the crystal formation can be obtained by the tie line b-b''. Now
because the solid solution is richer in B while the liquid phase is less rich in B, the compositional point of the

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 339

liquid phase will move towards left. Continuing the process of crystallization, the temperature of the system
will decrease and the composition of the liquid phase will move along bc'd'O. When the overall system is at
point c, we have a solid solution with composition c'' in equilibrium with the liquid solution with composition
c'. These two points are obtained by the tie line through point c easily; while the relative amounts in solid and
liquid phase can be calculated via lever rule. Furthermore, it is also very obvious from the phase diagram that
as the system goes from point b to d, the left part of the tie lines is increasing whereas right hand side decreases
continuously. This means that in going from point b to point d, the amount of solid solution increases
continuously. When the temperature is lowered to point d, almost all of the liquid is transformed into solid
which has the same composition as of the starting liquid solution.
The point O and S represent the freezing points of pure A and pure B, respectively. The curve ObS
is the freezing point curve of the liquid solution whereas the curve OdS represents the fusion point curve of
the solid solution. In the area above the curve ObS, the system is completely liquid; while it is completely solid
below the curve OdS. In the area between the curve ObS and OdS, the liquid solution is in equilibrium with
the solid solution.
2. Minimum-type solid solutions:
In these types of solutions, 𝐾𝐵 𝑖𝑛 𝐴 < 1 (B is solute and A is solvent i.e. 𝑥𝐵 (𝑠)/𝑥𝐵 (𝑙)) and 𝐾𝐴 𝑖𝑛 𝐵 <
1 (A is solute and B is solvent i.e. 𝑥𝐴 (𝑠)/𝑥𝐴 (𝑙)). The freezing points of such solutions exist below the freezing
points of pure solvents. In other words, the freezing points are depressed in these cases.

Figure 26. The phase diagram of minimum-type solid solutions.

Copyright © Mandeep Dalal


340 A Textbook of Physical Chemistry – Volume I

The point O and S represent the freezing points of pure A and pure B, respectively. The point Q is
the minimum freezing point of the liquid solution where solid and liquid solutions have the same composition.
The system becomes non-variant at point Q and freezing takes place at a constant temperature with an arrest
in cooling curve.
The curve ObQb'S is the freezing point curve of the liquid solution whereas the curve OcQc'S
represents the fusion point curve of the solid solution. In the area above the curve ObQb'S, the system is
completely liquid; while it is completely solid below the curve OcQc'S. In the area ObQcO, the liquid solution
(composition on ObQ) is in equilibrium with the solid solution (composition on OcQ). In the area Sb'Qc'S, the
liquid solution (composition on Sb'Q) is in equilibrium with the solid solution (composition on Sc'Q).
3. Maximum-type solid solutions:
In these types of solutions, 𝐾𝐵 𝑖𝑛 𝐴 > 1 (B is solute and A is solvent i.e. 𝑥𝐵 (𝑠)/𝑥𝐵 (𝑙)) and 𝐾𝐴 𝑖𝑛 𝐵 >
1 (A is solute and B is solvent i.e. 𝑥𝐴 (𝑠)/𝑥𝐴 (𝑙)). The freezing points of such solutions exist above the freezing
points of pure solvents. In other words, the freezing points are elevated in these cases.

Figure 27. The phase diagram of maximum-type solid solutions.

The point O and S represent the freezing points of pure A and pure B, respectively. The point Q is
the maximum freezing point of the liquid solution where solid and liquid solutions have the same composition.
The system becomes non-variant at point Q and freezing takes place at a constant temperature with an arrest
in cooling curve.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 341

The curve ObQb'S is the freezing point curve of the liquid solution whereas the curve OcQc'S
represents the fusion point curve of the solid solution. In the area above the curve ObQb'S, the system is
completely liquid; while it is completely solid below the curve OcQc'S. In the area ObQcO, the liquid solution
(composition on ObQ) is in equilibrium with the solid solution (composition on OcQ). In the area Sb'Qc'S, the
liquid solution (composition on Sb'Q) is in equilibrium with the solid solution (composition on Sc'Q).

Copyright © Mandeep Dalal


342 A Textbook of Physical Chemistry – Volume I

 Problems
Q 1. Derive and discuss the Clausius-Clapeyron equation.
Q 2. What is the law of mass actions? How is it different from actual rate law?
Q 3. Derive the law of mass action in the thermodynamic framework.
Q 4. State and discuss the Nernst heat theorem.
Q 5. What is the third law of thermodynamics? How it can be used to obtain the absolute entropies?
Q 6. What do you mean by reduced phase rule? Discuss the same for two-component systems.
Q 7. Draw and discuss the phase diagram of eutectic systems with reference to the cooling curve method.
Q 8. Define congruent and incongruent melting points.
Q 9. Draw and discuss the phase diagram of lead-silver system.
Q 10. Draw and discuss the phase diagram of NaCl-H2O system.
Q 11. What are solid solutions? Discuss the thermodynamic treatment for the same with special reference to
ascending-type solid solutions.

Copyright © Mandeep Dalal


CHAPTER 6 Thermodynamics – II 343

 Bibliography
[1] B. R. Puri, L. R. Sharma, M. S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2008.
[2] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[3] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[4] K. L. Kapoor, A Textbook of Physical Chemistry Volume 3, Macmillan Publishers, New Delhi, India, 2012.
[5] G. Raj, Thermodynamics, Krishna Prakashan, Meerut, India, 2004.
[6] S. Carnot, R. Fox, Reflections on the Motive Power of Fire: a Critical Edition with the Surviving Scientific
Manuscripts, Manchester University Press, New York, USA, 1986.
[7] P. A. Rock, Chemical Thermodynamics, University Science Books, California, USA, 1983.
[8] E. B. Smith, Basic Chemical Thermodynamics, Imperial College Press, London, UK, 2014.

Copyright © Mandeep Dalal


CHAPTER 7
Chemical Dynamics – II
 Chain Reactions: Hydrogen-Bromine Reaction, Pyrolysis of Acetaldehyde,
Decomposition of Ethane
A German chemist, Max Bodenstein, proposed the idea of chemical chain reactions in 1913. He
suggested that if two molecules react with each other, some unstable molecules may also be formed along-
products that can further react with the reactant molecules with a much higher probability than the initial
reactants. Another German chemist, Walther Nernst, explained the quantum yield phenomena in 1918 by
suggesting that the photochemical reaction between hydrogen and chlorine is actually a chain reaction. He
proposed that only one photon of light is actually accountable for the formation of 106 molecules of the final
product. W. Nernst proposed that the incident photon breaks a chlorine molecule into two individual Cl atoms,
each of which initiates a long series of stepwise reactions giving a large amount of hydrochloric acid.
A Danish scientist, Christian Christiansen; alongside a Dutch chemist, Hendrik Anthony Kramers;
observed the polymer-synthesis in 1923 and conclude that a chain reaction doesn’t need a photon always but
can also be started by the violent collision of two molecules. They also concluded that if two or more unstable
molecules are produced during this reaction, the reaction chain could branch itself to grow enormously
resulting in an explosion as well. These ideas were the very initial explanations for the mechanism responsible
for the chemical explosions. A more sophisticated theory was proposed by Soviet physicist Nikolay Semyonov
in 1934 to explain the quantitative aspects. N. Semyonov received the Nobel Prize in 1956 for his work (along
with Sir Cyril Norman Hinshelwood for his independent developments).
Steps Involved in a Typical Chain Reaction: The primary steps involved in a typical chain reaction are
discussed below.
i) Initiation: This step includes the formation of chain carriers or simply the active particles usually free
radicals in a photochemically or thermally induced chemical change.
ii) Propagation: This step may include many elementary reactions in a cycle in which chain carriers react to
forms another chain carrier that continues the chain by entering the next elementary reaction. In other words,
we can label these chain carriers as a catalyst for the overall propagation.
iii) Termination: This step includes the elementary chemical change in which the chain carriers lose their
activity by combining with each other.
It is also worthy to note that the average number of times the propagation cycle is repeated is equal
to the ratio of the overall reaction rate to the rate of initiation, and is called as “chain length”. Furthermore,
some chain reactions follow very complex rate laws with mixed or fractional order kinetics. In this section, we
will discuss nature and kinetics some of the most popular chain reactions such as ethane’s decomposition.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 345

 General Kinetics of Chain Reactions


On the basis of the chain carriers produced in each propagation step, the chain reactions can primarily
be classified into two categories; non-branched or the stationary reactions and branched or the non-stationary
reactions. A typical chain reaction can be written as given below.

𝑘1 (1)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: 𝐴→ 𝑅∗

𝑘2 (2)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: 𝑅∗ + 𝐴 → 𝑃 + 𝑛𝑅 ∗

𝑘3 (3)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: 𝑅∗ → 𝑑𝑒𝑠𝑡𝑟𝑢𝑐𝑡𝑖𝑜𝑛

Where P, R* and A represent the product, radical (or chain carrier) and reactant molecules, respectively. The
symbol represents a number that equals unity for stationary chain reactions and greater than one for non-
stationary or the branched-chain reactions. Furthermore, the destruction of the radical in the termination step
can occur either via its collision with another radical (in gas phase) or by striking the walls of the container.
The steady-state approximation can be employed to determine the concentration of intermediate or
the radical involved in the propagation step. The general procedure for which is to put the overall rate of
formation equals to zero. In other words, the rate of formation of R * must be equal to the rate of decomposition
of the same i.e.

𝑅𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑅 ∗ = 𝑅𝑎𝑡𝑒 𝑜𝑓 𝑑𝑖𝑠𝑎𝑝𝑝𝑒𝑎𝑟𝑎𝑛𝑐𝑒 𝑜𝑓 𝑅 ∗ (4)

𝑘1 [𝐴] = 𝐾3 [𝑅 ∗ ] − 𝑘2 (𝑛 − 1)[𝑅 ∗ ][𝐴] (5)

𝑑[𝑅 ∗ ] (6)
𝑘1 [𝐴] − 𝑘3 [𝑅∗ ] + 𝑘2 (𝑛 − 1)[𝑅 ∗ ][𝐴] = 0 =
𝑑𝑡
or

−𝑘3 [𝑅∗ ] + 𝑘2 (𝑛 − 1)[𝑅 ∗ ][𝐴] = −𝑘1 [𝐴] (7)

𝑘1 [𝐴] (8)
[𝑅 ∗ ] =
𝑘2 (1 − 𝑛)[𝐴] + 𝑘3

Now because the destruction of the radical in the termination step can occur either via its collision with another
radical (in the gas phase) or via striking the walls of the container, the rate constant k3 can be replaced by the
sum of the rate constants of two i.e. 𝑘3 = 𝑘𝑤 + 𝑘𝑔 . After using the value of 𝑘3 in equation (7), we have

𝑘1 [𝐴] (9)
[𝑅 ∗ ] =
𝑘2 (1 − 𝑛)[𝐴] + 𝑘𝑤 + 𝑘𝑔

Now we are ready to apply the concept on different types of chain reactions.

Copyright © Mandeep Dalal


346 A Textbook of Physical Chemistry – Volume I

 Hydrogen-Bromine Reaction
The hydrogen-bromine or the H2-Br2 reaction is a typical case of stationary type chain reactions (n =
1) for which the overall reaction can be written as given below.

H2 + Br2 ⟶ 2HBr (10)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (11)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: Br2 → 2Br

𝑘2 (12)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: Br + H2 → HBr + H

𝑘3 (13)
H + Br2 → HBr + Br

𝑘4 (14)
𝐼𝑛ℎ𝑖𝑏𝑖𝑡𝑖𝑜𝑛: H + HBr → H2 + Br

𝑘5 (15)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: Br + Br → Br2

The net rate of formation of HBr must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[HBr] (16)
= 𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟]
dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the H
and Br first i.e.

d[H] (17)
= 0 = 𝑘2 [𝐵𝑟][𝐻2 ] − 𝑘3 [𝐻][𝐵𝑟2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟]
dt
Similarly,

d[Br] (18)
= 0 = 2𝑘1 [𝐵𝑟2 ] − 𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] − 2𝑘5 [𝐵𝑟]2
dt
Taking negative both side of equation (17), we have

−𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] = 0 (19)

Using the above result in equation (18), we get

2𝑘1 [𝐵𝑟2 ] + 0 − 2𝑘5 [𝐵𝑟]2 = 0 (20)

2𝑘5 [𝐵𝑟]2 = 2𝑘1 [𝐵𝑟2 ] (21)

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 347

𝑘1 1/2 (22)
[𝐵𝑟] = ( ) [𝐵𝑟2 ]1/2
𝑘5

Similarly, rearranging equation (17) again

𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] = 𝑘2 [𝐵𝑟][𝐻2 ] (23)

𝑘2 [𝐵𝑟][𝐻2 ] (24)
[𝐻] =
𝑘3 2 ] + 𝑘4 [𝐻𝐵𝑟]
[𝐵𝑟

Now using the value of [Br] from equation (22), the above equation takes the form

𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (25)


[𝐻] =
𝑘3 [Br2 ] + 𝑘4 [HBr]

Now rearranging equation (17) again in different mode i.e.

𝑘2 [𝐵𝑟][𝐻2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟] = 𝑘3 [𝐻][𝐵𝑟2 ] (26)

Using the above result in equation (16), we have

𝑑[HBr] (27)
= 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘3 [𝐻][𝐵𝑟2 ]
dt
𝑑[HBr] (28)
= 2𝑘3 [𝐻][𝐵𝑟2 ]
dt
After putting the value of [𝐻] from equation (24), the equation (28) takes the form

𝑑[HBr] 2𝑘3 𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]3/2 [𝐻2 ] (29)


=
dt 𝑘3 [Br2 ] + 𝑘4 [HBr]

Taking 𝑘3 [Br2 ] as common in the denominator and then canceling out the same form numerator, we get

𝑑[HBr] 2𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (30)


=
dt 1 + (𝑘4 /𝑘3 )[HBr]/[𝐵𝑟2 ]

Now consider two new constants as

𝑘 ′ = 2𝑘2 (𝑘1 /𝑘5 )1/2 𝑎𝑛𝑑 𝑘 ′′ = 𝑘4 /𝑘3 (31)

Using the above results in equation (30), we get

𝑑[HBr] 𝑘 ′ [𝐵𝑟2 ]1/2 [𝐻2 ] (32)


=
dt 1 + 𝑘 ′′ [HBr]/[𝐵𝑟2 ]

The initial reaction rate expression can be obtained by neglecting [𝐻𝐵𝑟] i.e. 1 + 𝑘 ′′ [HBr]/[𝐵𝑟2 ] ≈ 1 as

Copyright © Mandeep Dalal


348 A Textbook of Physical Chemistry – Volume I

𝑑[HBr] (33)
[ ] = 𝑘 ′ [𝐵𝑟2 ]1/2 [𝐻2 ]0
dt 0

Hence, the order of the hydrogen-bromine reaction in the initial stage will be 1.5 only i.e. first-order w.r.t.
hydrogen and half w.r.t. bromine.
 Pyrolysis of Acetaldehyde
The pyrolysis of acetaldehyde is another typical case of stationary type chain reactions (n = 1) for
which the overall reaction can be written as given below.

CH3 CHO ⟶ CH4 + CO (34)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (35)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: CH3 CHO → ˙CH3 + ˙CHO

𝑘2 (36)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: ˙CH3 + CH3 CHO → CH4 + ˙CH2 CHO

𝑘3 (37)
˙CH2 CHO → ˙CH3 + CO

𝑘4 (38)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: ˙CH3 + ˙CH3 → CH3 CH3

The net rate of formation of CH4 must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[CH4 ] (39)
= 𝑘2 [˙CH3 ][CH3 CHO]
dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the
[˙CH3 ] and [˙CH2 CHO] first i.e.

d[˙CH3 ] (40)
= 0 = 𝑘1 [CH3 CHO] − 𝑘2 [˙CH3 ][CH3 CHO] + 𝑘3 [˙CH2 CHO] − 2𝑘4 [˙CH3 ]2
dt
Similarly,

d[˙CH2 CHO] (41)


= 0 = 𝑘2 [˙CH3 ][CH3 CHO] − 𝑘3 [˙CH2 CHO]
dt
Taking negative both side of equation (41), we have

−𝑘2 [˙CH3 ][CH3 CHO] + 𝑘3 [˙CH2 CHO] = 0 (42)

Using the above result in equation (40), we get

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 349

𝑘1 [CH3 CHO] + 0 − 2𝑘4 [˙CH3 ]2 = 0 (43)

𝑘1 1/2 (44)
[˙CH3 ] = ( ) [CH3 CHO]1/2
2𝑘4

After putting the value of [˙CH3 ] from equation (44) in equation (39), we have

𝑑[CH4 ] 𝑘1 1/2 (45)


= 𝑘2 ( ) [CH3 CHO]1/2 [CH3 CHO]
dt 2𝑘4

𝑑[CH4 ] 𝑘1 1/2 (46)


= 𝑘2 ( ) [CH3 CHO]3/2
dt 2𝑘4

Now consider a new constant as

𝑘1 1/2 (47)
𝑘 = 𝑘2 ( )
2𝑘4

Using in equation (46), we get

𝑑[CH4 ] (48)
= 𝑘[CH3 CHO]3/2
dt
The kinetic chain length for the same can be obtained by dividing the rate of formation of the product by rate
of initiation step i.e.

𝑅𝑝 (49)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑅𝑖

Where 𝑅𝑝 and 𝑅𝑖 are the rate of propagation and rate of initiation respectively. Now since the rate of
propagation is simply equal to the overall rate law i.e. equation (48), the rate of initiation can be given as

𝑅𝑖 = 𝑘1 [CH3 CHO] (50)

After using the values of 𝑅𝑝 and 𝑅𝑖 from equation (48, 50) in equation (49), we get the expression for kinetic
chain length as

𝑘[CH3 CHO]3/2 (51)


𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑘1 [CH3 CHO]

or

𝑘 (52)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ = [CH3 CHO]1/2
𝑘1

Copyright © Mandeep Dalal


350 A Textbook of Physical Chemistry – Volume I

 Decomposition of Ethane
The decomposition of ethane is another typical case of stationary type chain reactions (n = 1) for
which the overall reaction can be written as given below.

CH3 CH3 ⟶ CH2 = CH2 + H2 (53)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (54)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: CH3 CH3 → ˙CH3 + ˙CH3

𝑘2 (55)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: ˙CH3 + CH3 CH3 → CH4 + ˙CH2 CH3

𝑘3 (56)
˙CH2 CH3 → CH2 = CH2 + ˙H

𝑘4 (57)
CH3 CH3 + ˙H → ˙CH2 CH3 + H2

𝑘5 (58)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: ˙CH2 CH3 + ˙H → CH3 CH3

The net rate of decomposition of ethane must be equal to the rate of formation ethylene i.e.

𝑑[C2 H6 ] 𝑑[C2 H4 ] (59)


− = = 𝑘3 [˙CH2 CH3 ]
dt dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the
[˙H], [˙CH3 ] and [˙CH2 CH3 ] first i.e.

d[˙CH3 ] (60)
= 0 = 2𝑘1 [CH3 CH3 ] − 𝑘2 [˙CH3 ][CH3 CH3 ]
dt
Similarly,

d[˙CH2 CH3 ] (61)


= 0 = 𝑘2 [˙CH3 ][CH3 CH3 ] − 𝑘3 [˙CH2 CH3 ] + 𝑘4 [˙H][CH3 CH3 ] − 𝑘5 [˙H][˙CH2 CH3 ]
dt
Similarly,

d[˙H] (62)
= 0 = 𝑘3 [˙CH2 CH3 ] − 𝑘4 [˙H][CH3 CH3 ] − 𝑘5 [˙H][˙CH2 CH3 ]
dt
Rearranging equation (60), we get

2𝑘1 − 𝑘2 [˙CH3 ] = 0 (63)

2𝑘1 (64)
[˙CH3 ] =
𝑘2

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 351

Rearranging equation (62), we get

𝑘4 [˙H][CH3 CH3 ] + 𝑘5 [˙H][˙CH2 CH3 ] = 𝑘3 [˙CH2 CH3 ] (65)

𝑘3 [˙CH2 CH3 ] (66)


[˙H] =
𝑘4 [CH3 CH3 ] + 𝑘5 [˙CH2 CH3 ]

After putting the value of [˙CH3 ] and [˙H] from equation (64, 66) in equation (61), we have

2𝑘1 𝑘3 [˙CH2 CH3 ] (67)


𝑘2 ( ) [CH3 CH3 ] − 𝑘3 [˙CH2 CH3 ] + 𝑘4 ( ) [CH3 CH3 ]
𝑘2 𝑘4 [CH3 CH3 ] + 𝑘5 [˙CH2 CH3 ]
𝑘3 [˙CH2 CH3 ]
− 𝑘5 ( ) [˙CH2 CH3 ] = 0
𝑘4 [CH3 CH3 ] + 𝑘5 [˙CH2 CH3 ]

or

𝑘3 𝑘5 [˙CH2 CH3 ]2 − 𝑘1 𝑘5 [˙CH2 CH3 ][CH3 CH3 ] − 𝑘1 𝑘4 [CH3 CH3 ]2 = 0 (68)

The equation (68) is quadric in nature and can be solved to give

𝑘1 𝑘5 + (𝑘12 𝑘52 + 4𝑘1 𝑘5 𝑘3 𝑘4 )


1/2 (69)
[˙CH2 CH3 ] = [CH3 CH3 ]
2𝑘5 𝑘3

Now putting the result in equation (59), we get

𝑑[C2 H6 ] 𝑑[C2 H4 ] 𝑘1 𝑘5 + (𝑘12 𝑘52 + 4𝑘1 𝑘5 𝑘3 𝑘4 )


1/2 (70)
− = = 𝑘3 [CH3 CH3 ]
dt dt 2𝑘5 𝑘3

𝑑[C2 H6 ] 𝑘1 𝑘5 + (𝑘12 𝑘52 + 4𝑘1 𝑘5 𝑘3 𝑘4 )


1/2 (71)
− = [CH3 CH3 ]
dt 2𝑘5

At this stage, defining a new constant as

𝑘1 𝑘5 + (𝑘12 𝑘52 + 4𝑘1 𝑘5 𝑘3 𝑘4 )


1/2 (72)
𝑘=
2𝑘5

Now owing to the very small rate of initiation step, all the terms 𝑘1 𝑘5 and 𝑘12 𝑘52 can be neglected i.e.

(4𝑘1 𝑘5 𝑘3 𝑘4 )1/2 𝑘1 𝑘3 𝑘4 1/2 (73)


𝑘= =( )
2𝑘5 𝑘5

the equation (71) takes the form

𝑑[C2 H6 ] (74)
− = 𝑘[CH3 CH3 ]
dt

Copyright © Mandeep Dalal


352 A Textbook of Physical Chemistry – Volume I

The kinetic chain length for the same can be obtained by dividing the rate of formation of the product by rate
of initiation step i.e.

𝑅𝑝 (75)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑅𝑖

Where 𝑅𝑝 and 𝑅𝑖 are the rate of propagation and rate of initiation respectively. Now since the rate of
propagation is simply equal to the overall rate law i.e. equation (74), the rate of initiation can be given as

𝑅𝑖 = 𝑘1 [CH3 CH3 ] (76)

After using the values of 𝑅𝑝 and 𝑅𝑖 from equation (74, 76) in equation (75), we get the expression for kinetic
chain length as

𝑘[CH3 CH3 ] (77)


𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑘1 [CH3 CH3 ]

or

𝑘 1 𝑘1 𝑘3 𝑘4 1/2 𝑘3 𝑘4 1/2 (78)


𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ = = ( ) =( )
𝑘1 𝑘1 𝑘5 𝑘1 𝑘5

 Photochemical Reactions (Hydrogen-Bromine & Hydrogen-Chlorine


Reactions)
There are many chemical reactions which occur also when the reactants are exposed to light. These
reactions are called as photochemical reactions. Now before we discuss the nature and types of photochemical
reactions, we need to discuss two laws first; the first is Grotthuss-Draper law while the second one is Stark-
Einstein law.
The first law of photochemistry was given by Theodor Grotthuss and John W. Draper, and therefore,
got its unique name.
The first law of photochemistry states that when the light is allowed to strike the reactions mixture,
it can be partially transmitted, reflected and absorbed; and it is the absorbed portion of the incident light
which is responsible to carry out any chemical change.
The second law of photochemistry was given by Johannes Stark and Albert Einstein, and therefore,
is popularly known as Stark-Einstein law.
The second law of photochemistry states that one photon of light must be absorbed for one molecule
to get activated in a photochemical reaction by a chemical system.
The quantum yield of a reaction is simply the ratio of number of molecules reacting in a given time to the
number of photons absorbed in the same time i.e.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 353

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡 𝑟𝑒𝑎𝑐𝑡𝑖𝑛𝑔 𝑝𝑒𝑟 𝑠𝑒𝑐𝑜𝑛𝑑 (79)


𝑄𝑢𝑎𝑛𝑡𝑢𝑚 𝑦𝑖𝑒𝑙𝑑(𝜙) =
𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑒𝑖𝑛𝑠𝑡𝑒𝑖𝑛𝑠 𝑎𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑝𝑒𝑟 𝑠𝑒𝑐𝑜𝑛𝑑

Where one einstein represents one mole of photons (6.022×1023). The quantum yield is of two types in
photochemical reactions; one for the primary process and the second one for the secondary process. During
the course of the primary process, light is absorbed by the reactant and the quantum yield of this part is always
unity. Therefore, we can say that the number of radiation-absorbing molecules consumed per unit time in the
primary process is equal to intensity of absorbed radiation i.e. Iab (number of photons striking the sample per
unit time). For instance, consider the photo-decomposition

𝑘 (80)
Br2 +ℎ𝜈 → 2Br

Then, the quantum yield will be

𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑚𝑜𝑙𝑒𝑠 𝑜𝑓 𝐵𝑟2 𝑑𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑒𝑑 𝑝𝑒𝑟 𝑢𝑛𝑖𝑡 𝑡𝑖𝑚𝑒 (81)


𝜙=
𝑁𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑒𝑖𝑛𝑠𝑡𝑒𝑖𝑛𝑠 𝑎𝑏𝑠𝑜𝑟𝑏𝑒𝑑 𝑝𝑒𝑟 𝑢𝑛𝑖𝑡 𝑡𝑖𝑚𝑒

or

−𝑑[𝐵𝑟2 ]/𝑑𝑡 (82)


𝜙=
𝐼𝑎𝑏

Since 𝜙 = 1 for primary process, we have

−𝑑[𝐵𝑟2 ]/𝑑𝑡 (83)


1=
𝐼𝑎𝑏

or

𝑑[𝐵𝑟2 ] (84
− = 𝐼𝑎𝑏
𝑑𝑡
𝑑[𝐵𝑟2 ] (85)
− = 𝑘[𝐵𝑟2 ] = 𝐼𝑎𝑏
𝑑𝑡
If rate of formation of Br is asked, then

𝑑[𝐵𝑟2 ] 1 𝑑[𝐵𝑟] (86)


− =+ = 𝑘[𝐵𝑟2 ] = 𝐼𝑎𝑏
𝑑𝑡 2 𝑑𝑡
𝑑[𝐵𝑟] (87)
= 2𝑘[𝐵𝑟2 ] = 2𝐼𝑎𝑏
𝑑𝑡
Two of the most common examples of photochemical reactions are hydrogen-bromine and hydrogen-
chlorine reactions whose kinetics and nature will be discussed in this section.

Copyright © Mandeep Dalal


354 A Textbook of Physical Chemistry – Volume I

 Kinetics of Photochemical Reaction Between Hydrogen and Bromine


The hydrogen-bromine or the H2-Br2 reaction is a typical case of photochemical reactions for which
the overall reaction can be written as given below.

ℎ𝜈 (88)
H2 + Br2 → 2HBr

Since it is a chain reaction, the elementary steps for the same can be proposed as

𝑘1 (89)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: Br2 + ℎ𝜈 → 2Br

𝑘2 (91)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: Br + H2 → HBr + H

𝑘3 (92)
H + Br2 → HBr + Br

𝑘4 (93)
𝐼𝑛ℎ𝑖𝑏𝑖𝑡𝑖𝑜𝑛: H + HBr → H2 + Br

𝑘5 (94)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: Br + Br → Br2

The net rate of formation of HBr must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[HBr] (95)
= 𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟]
dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the H
and Br first i.e.

d[H] (96)
= 𝑘2 [𝐵𝑟][𝐻2 ] − 𝑘3 [𝐻][𝐵𝑟2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟] = 0
dt
Similarly,

d[Br] (97)
= 2𝑘1 [𝐵𝑟2 ] − 𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] − 2𝑘5 [𝐵𝑟]2 = 0
dt
Taking negative both side of equation (96), we have

−𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] = 0 (98)

Using the above result in equation (97), we get

2𝑘1 [𝐵𝑟2 ] + 0 − 2𝑘5 [𝐵𝑟]2 = 0 (99)

2𝑘5 [𝐵𝑟]2 = 2𝑘1 [𝐵𝑟2 ] (100)

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 355

𝑘1 1/2 (101)
[𝐵𝑟] = ( ) [𝐵𝑟2 ]1/2
𝑘5

Similarly, rearranging equation (96) again

𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] = 𝑘2 [𝐵𝑟][𝐻2 ] (102)

𝑘2 [𝐵𝑟][𝐻2 ] (103)
[𝐻] =
𝑘3 2 ] + 𝑘4 [𝐻𝐵𝑟]
[𝐵𝑟

Now using the value of [Br] from equation (101), the above equation takes the form

𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (104)


[𝐻] =
𝑘3 [Br2 ] + 𝑘4 [HBr]

Now rearranging equation (96) again in different mode i.e.

𝑘2 [𝐵𝑟][𝐻2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟] = 𝑘3 [𝐻][𝐵𝑟2 ] (105)

Using the above result in equation (95), we have

𝑑[HBr] (106)
= 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘3 [𝐻][𝐵𝑟2 ]
dt
𝑑[HBr]/dt = 2𝑘3 [𝐻][𝐵𝑟2 ] (107)

After putting the value of [𝐻] from equation (104), the equation (107) takes the form

𝑑[HBr] 2𝑘3 𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]3/2 [𝐻2 ] (108)


=
dt 𝑘3 [Br2 ] + 𝑘4 [HBr]

Taking 𝑘3 [Br2 ] as common in the denominator and then canceling out the same form numerator, we get

𝑑[HBr] 2𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (109)


=
dt 1 + (𝑘4 /𝑘3 )[HBr]/[𝐵𝑟2 ]

Now since 2𝑘1 [𝐵𝑟2 ] = 2𝐼𝑎𝑏 , then


1/2
𝑘1 [𝐵𝑟2 ]1/2 = (𝐼𝑎𝑏 )1/2 (110)

Using the above result in equation (109), we get

𝑑[HBr] 2𝑘2 (𝐼𝑎𝑏 /𝑘5 )1/2 [𝐻2 ] (111)


=
dt 1 + (𝑘4 /𝑘3 )[HBr]/[𝐵𝑟2 ]

Hence, the rate of hydrogen-bromine reaction is directly proportional to the square root of the intensity of
absorbed radiation.

Copyright © Mandeep Dalal


356 A Textbook of Physical Chemistry – Volume I

 Kinetics of Photochemical Reaction Between Hydrogen and Chlorine


The hydrogen-bromine or the H2-Cl2 reaction is a typical case of photochemical reactions for which
the overall reaction can be written as given below.

ℎ𝜈 (112)
H2 + Cl2 → 2HCl

Since it is a chain reaction, the elementary steps for the same can be proposed as

𝑘1 (113)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: Cl2 + ℎ𝜈 → 2Cl

𝑘2 (114)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: Cl + H2 → HCl + H

𝑘3 (115)
H + Cl2 → HCl + Cl

𝑘4 (116)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: 2Cl(at the walls) → Cl2

The net rate of formation of HCl must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[HCl] (117)
= 𝑘2 [𝐶𝑙][𝐻2 ] + 𝑘3 [𝐻][𝐶𝑙2 ]
dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the H
and Cl first i.e.

d[H] (118)
= 𝑘2 [𝐶𝑙][𝐻2 ] − 𝑘3 [𝐻][𝐶𝑙2 ] = 0
dt
Similarly,

d[Cl] (119)
= 2𝑘1 [𝐶𝑙2 ] − 𝑘2 [Cl][𝐻2 ] + 𝑘3 [𝐻][𝐶𝑙2 ] − 2𝑘4 [𝐶𝑙] = 0
dt
Taking negative both side of equation (118), we have

−𝑘2 [𝐶𝑙][𝐻2 ] + 𝑘3 [𝐻][𝐶𝑙2 ] = 0 (120)

Using the above result in equation (119), we get

2𝑘1 [𝐶𝑙2 ] + 0 − 2𝑘4 [𝐶𝑙] = 0 (121)

or

2𝑘1 [𝐶𝑙2 ] = 2𝑘4 [𝐶𝑙] (122)

or

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 357

𝑘1 (123)
[𝐶𝑙] = [𝐶𝑙2 ]
𝑘4

Similarly, rearranging equation (118) again

𝑘2 [𝐶𝑙][𝐻2 ] − 𝑘3 [𝐻][𝐶𝑙2 ] = 0 (124)

or

𝑘2 [𝐶𝑙][𝐻2 ] (125)
[𝐻] =
𝑘3 [𝐶𝑙2 ]

Now using the value of [Cl] from equation (123), the above equation takes the form

𝑘2 [𝐻2 ] 𝑘1 𝑘1 𝑘2 [𝐻2 ] (126)


[𝐻] = [𝐶𝑙2 ] =
[𝐶𝑙
𝑘3 2 𝑘4] 𝑘3 𝑘4

Using values of [𝐶𝑙] and [𝐻] from equation (123, 126) in equation (117), we have

𝑑[HCl] 𝑘1 𝑘1 𝑘2 [𝐻2 ] (127)


= 𝑘2 [𝐶𝑙2 ][𝐻2 ] + 𝑘3 [𝐶𝑙2 ]
dt 𝑘4 𝑘3 𝑘4

or

𝑑[HCl] 𝑘1 𝑘2 [𝐶𝑙2 ][𝐻2 ] 𝑘1 𝑘2 [𝐻2 ][𝐶𝑙2 ] (128)


= +
dt 𝑘4 𝑘4

Now since 2𝑘1 [𝐶𝑙2 ] = 2𝐼𝑎𝑏 , then

𝑘1 [𝐶𝑙2 ] = 𝐼𝑎𝑏 (129)

Using the above result in equation (129), we get

𝑑[HCl] 𝑘2 𝐼𝑎𝑏 [𝐻2 ] 𝑘2 𝐼𝑎𝑏 [𝐻2 ] (130)


= +
dt 𝑘4 𝑘4

or

𝑑[HCl] 2𝑘2 𝐼𝑎𝑏 [𝐻2 ] (131)


=
dt 𝑘4

Hence, we can conclude that the rate of hydrogen-bromine reaction is directly proportional to the intensity of
absorbed radiation. It is also worthy to note that the quantum yield of the hydrogen-chlorine reaction is much
higher than that of hydrogen-bromine reaction which may simply be attributed to the exothermic nature of the
second step in H2-Cl2 reaction which makes it spontaneous in nature.

Copyright © Mandeep Dalal


358 A Textbook of Physical Chemistry – Volume I

 General Treatment of Chain Reactions (Ortho-Para Hydrogen Conversion


and Hydrogen-Bromine Reactions)
In this section, we will discuss the general treatment of some common chain reactions like ortho-para
hydrogen conversion and hydrogen-bromine reactions. To do so, we may also recall some concepts discussed
earlier in this chapter.
 Ortho-Para Hydrogen Conversion
The ortho-hydrogen can convert into para-hydrogen and can easily be measured in the temperature
range of 700−800°C. The conversion is completely homogeneous in nature and the order of the conversion is
1.5 as total. The widely accepted mechanism is given below.

𝑝-H2 ⇌ 2H (132)

𝑘2 (133)
𝐻 + 𝑝-H2 → H + 𝑜-H2

The equilibrium constant for the above-mentioned molecular-atomic equilibria given by equation (132) can be
written as

[𝐻]2 (134)
𝐾=
[𝑝-H2 ]

[𝐻]2 = 𝐾[𝑝-H2 ] (135)


1 (136)
[𝐻] = 𝐾 1/2 [𝑝-H2 ]2

The net rate of conversion of p-H2 must be equal to the sum of the rate of formation and the rate of
disappearance of the same i.e.

𝑑[𝑝-H2 ] (137)
= 𝑘2 [𝑝-H2 ][𝐻]
𝑑𝑡
After using the value of [𝐻] from equation (136) in equation (137), we get

𝑑[𝑝-H2 ] 1 (137)
= 𝑘2 [𝑝-H2 ]𝐾 1/2 [𝑝-H2 ]2
𝑑𝑡
or

𝑑[𝑝-H2 ] (137)
= 𝑘2 𝐾 1/2 [𝑝-H2 ]3/2
𝑑𝑡
The overall activation energy of the conversion is 𝐸 = 𝐸2 + 𝐷/2 where E2 is the activation energy of the
second step while D is the dissociation energy of dihydrogen molecule.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 359

 Hydrogen-Bromine Reactions
The hydrogen-bromine or the H2-Br2 reaction is a typical case of stationary type chain reactions (n =
1) for which the overall reaction can be written as given below.

H2 + Br2 ⟶ 2HBr (138)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (139)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: Br2 → 2Br

𝑘2 (140)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: Br + H2 → HBr + H

𝑘3 (141)
H + Br2 → HBr + Br

𝑘4 (142)
𝐼𝑛ℎ𝑖𝑏𝑖𝑡𝑖𝑜𝑛: H + HBr → H2 + Br

𝑘5 (143)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: Br + Br → Br2

The net rate of formation of HBr must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[HBr] (144)
= 𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟]
dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the H
and Br first i.e.

d[H] (145)
= 0 = 𝑘2 [𝐵𝑟][𝐻2 ] − 𝑘3 [𝐻][𝐵𝑟2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟]
dt
Similarly,

d[Br] (146)
= 0 = 2𝑘1 [𝐵𝑟2 ] − 𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] − 2𝑘5 [𝐵𝑟]2
dt
Taking negative both side of equation (145), we have

−𝑘2 [𝐵𝑟][𝐻2 ] + 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] = 0 (147)

Using the above result in equation (146), we get

2𝑘1 [𝐵𝑟2 ] + 0 − 2𝑘5 [𝐵𝑟]2 = 0 (148)

2𝑘5 [𝐵𝑟]2 = 2𝑘1 [𝐵𝑟2 ] (149)

Copyright © Mandeep Dalal


360 A Textbook of Physical Chemistry – Volume I

𝑘1 1/2 (150)
[𝐵𝑟] = ( ) [𝐵𝑟2 ]1/2
𝑘5

Similarly, rearranging equation (145) again

𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘4 [𝐻][𝐻𝐵𝑟] = 𝑘2 [𝐵𝑟][𝐻2 ] (151)

𝑘2 [𝐵𝑟][𝐻2 ] (152)
[𝐻] =
𝑘3 2 ] + 𝑘4 [𝐻𝐵𝑟]
[𝐵𝑟

Now using the value of [Br] from equation (150), the above equation takes the form

𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (153)


[𝐻] =
𝑘3 [Br2 ] + 𝑘4 [HBr]

Now rearranging equation (145) again in different mode i.e.

𝑘2 [𝐵𝑟][𝐻2 ] − 𝑘4 [𝐻][𝐻𝐵𝑟] = 𝑘3 [𝐻][𝐵𝑟2 ] (154)

Using the above result in equation (144), we have

𝑑[HBr] (155)
= 𝑘3 [𝐻][𝐵𝑟2 ] + 𝑘3 [𝐻][𝐵𝑟2 ]
dt
𝑑[HBr] (156)
= 2𝑘3 [𝐻][𝐵𝑟2 ]
dt
After putting the value of [𝐻] from equation (152), the equation (156) takes the form

𝑑[HBr] 2𝑘3 𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]3/2 [𝐻2 ] (157)


=
dt 𝑘3 [Br2 ] + 𝑘4 [HBr]

Taking 𝑘3 [Br2 ] as common in the denominator and then canceling out the same form numerator, we get

𝑑[HBr] 2𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (158)


=
dt 1 + (𝑘4 /𝑘3 )[HBr]/[𝐵𝑟2 ]

Now consider two new constants as

𝑘 ′ = 2𝑘2 (𝑘1 /𝑘5 )1/2 𝑎𝑛𝑑 𝑘 ′′ = 𝑘4 /𝑘3 (159)

Using in equation (158), we get

𝑑[HBr] 𝑘 ′ [𝐵𝑟2 ]1/2 [𝐻2 ] (160)


=
dt 1 + 𝑘 ′′ [HBr]/[𝐵𝑟2 ]

The initial reaction rate expression can be obtained by neglecting [𝐻𝐵𝑟] i.e. 1 + 𝑘 ′′ [HBr]/[𝐵𝑟2 ] ≈ 1 as

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 361

𝑑[HBr] (161)
[ ] = 𝑘 ′ [𝐵𝑟2 ]1/2 [𝐻2 ]0
dt 0

Hence, the order of the hydrogen-bromine reaction in the initial stage will be 1.5 only i.e. first-order w.r.t.
hydrogen and half w.r.t. bromine.
However, if the initiation occurs by the photon of the incident light, the expression for the overall
rate can slightly be modified. Recall the initiation step again via photochemical decomposition i.e.

𝑘 (162)
Br2 +ℎ𝜈 → 2Br

Then, the quantum yield for this primary change will be

−𝑑[𝐵𝑟2 ]/𝑑𝑡 (163)


𝜙=
𝐼𝑎𝑏

Where 𝐼𝑎𝑏 is the intensity of absorbed radiation. Since 𝜙 = 1 for primary process, we have

−𝑑[𝐵𝑟2 ]/𝑑𝑡 (164)


1=
𝐼𝑎𝑏

𝑑[𝐵𝑟2 ] (165)
− = 𝐼𝑎𝑏
𝑑𝑡
𝑑[𝐵𝑟2 ] (166)
− = 𝑘[𝐵𝑟2 ] = 𝐼𝑎𝑏
𝑑𝑡
For the rate of formation of Br, we have

𝑑[𝐵𝑟2 ] 1 𝑑[𝐵𝑟] (167)


− =+ = 𝑘[𝐵𝑟2 ] = 𝐼𝑎𝑏
𝑑𝑡 2 𝑑𝑡
𝑑[𝐵𝑟] (168)
= 2𝑘[𝐵𝑟2 ] = 2𝐼𝑎𝑏
𝑑𝑡
Now since 2𝑘1 [𝐵𝑟2 ] = 2𝐼𝑎𝑏 , then
1/2
𝑘1 [𝐵𝑟2 ]1/2 = (𝐼𝑎𝑏 )1/2 (169)

Using the above result in equation (158), we get

𝑑[HBr] 2𝑘2 (𝐼𝑎𝑏 /𝑘5 )1/2 [𝐻2 ] (170)


=
dt 1 + (𝑘4 /𝑘3 )[HBr]/[𝐵𝑟2 ]

Hence, the rate of hydrogen-bromine reaction is directly proportional to the square root of the intensity of
absorbed radiation.

Copyright © Mandeep Dalal


362 A Textbook of Physical Chemistry – Volume I

 Apparent Activation Energy of Chain Reactions


In order to understand the activation energy profile of chain reactions, consider a typical chain reaction
such as the synthesis of hydrogen bromide from its elements.

H2 + Br2 ⟶ 2HBr (171)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (172)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: Br2 + M → 2Br

𝑘2 (173)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: Br + H2 → HBr + H

𝑘3 (174)
H + Br2 → HBr + Br

𝑘4 (175)
𝐼𝑛ℎ𝑖𝑏𝑖𝑡𝑖𝑜𝑛: H + HBr → H2 + Br

𝑘5 (176)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: 2Br + M → M ∗ + Br2

It is obvious from the stepwise mechanism that the reaction is actually initiated by the formation of a free
radical when one Br2 molecule collides with some other molecule M (most another Br2). All this just requires
a sufficiently high temperature of the gas. Another mode of formation of Br free radicals is the photochemical
activation in which the sample is exposed to light. In second and third reactions, the consumption of a free
radical generates another, and therefore, propagates the chain. The same happens in the fourth step, but the
molecule of the product is destroyed, and thus inhibited the whole process. Now if these first four steps occur,
then the chain would continue for infinite. However, during the course of the 5th step, two Br combine together,
and thus terminating the chain. The molecule M* in the last step represents a thermally-excited which dissipates
its energy to other molecules quickly.
The overall rate laws for such chain reactions are quite complex and usually have non-integral orders.
The net rate of formation of HBr can be given as

𝑑[HBr] 2𝑘2 (𝑘1 /𝑘5 )1/2 [𝐵𝑟2 ]1/2 [𝐻2 ] (177)


=
dt 1 + (𝑘4 /𝑘3 )[HBr]/[𝐵𝑟2 ]

Now consider two new constants as

𝑘 ′ = 2𝑘2 (𝑘1 /𝑘5 )1/2 𝑎𝑛𝑑 𝑘 ′′ = 𝑘4 /𝑘3 (178)

Using in equation (177), we get

𝑑[HBr] 𝑘 ′ [𝐵𝑟2 ]1/2 [𝐻2 ] (179)


=
dt 1 + 𝑘 ′′ [HBr]/[𝐵𝑟2 ]

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 363

The initial reaction rate expression can be obtained by neglecting [𝐻𝐵𝑟] i.e. 1 + 𝑘 ′′ [HBr]/[𝐵𝑟2 ] ≈ 1 as

𝑑[HBr] (180)
[ ] = 𝑘 ′ [𝐵𝑟2 ]1/2 [𝐻2 ]0
dt 0

Hence, the order of the hydrogen-bromine reaction in the initial stage will be 1.5 only i.e. first-order w.r.t.
hydrogen and half w.r.t. bromine. Expanding equation (180, we get

𝑑[HBr] 𝑘1 1/2 (181)


[ ] = 2𝑘2 ( ) [𝐵𝑟2 ]1/2 [𝐻2 ]0
dt 0 𝑘5

Hence the overall rate constant can be written as

𝑘1 1/2 (182)
𝑘𝑜𝑣𝑒𝑟𝑎𝑙𝑙 = 2𝑘2 ( )
𝑘5

Now assuming E1, E2 and E5 are the activation energies for first (initiation), second (propagation) and fifth
(termination) steps; we can write general expressions for individual rate constants as

𝑘1 = 𝐴𝑒 −𝐸1 /𝑅𝑇 (183)

𝑘2 = 𝐴𝑒 −𝐸2 /𝑅𝑇 (184)

𝑘5 = 𝐴𝑒 −𝐸5 /𝑅𝑇 (185)

After putting values of 𝑘1 , 𝑘2 and 𝑘5 from equation (183-185) in equation (182), we get

𝐴𝑒 −𝐸1 /𝑅𝑇
1/2
1/2
(186)
−𝐸2 /𝑅𝑇
𝑘𝑜𝑣𝑒𝑟𝑎𝑙𝑙 = 2 𝐴𝑒 ( −𝐸 /𝑅𝑇 ) = 2 𝐴𝑒 −𝐸2 /𝑅𝑇 (𝑒 −𝐸1 +𝐸5 /𝑅𝑇 )
𝐴𝑒 5
−𝐸2 −𝐸1 +𝐸5 −2𝐸2 −𝐸1 +𝐸5 1 (2𝐸2 +𝐸1 −𝐸5 ) (187)
𝑘𝑜𝑣𝑒𝑟𝑎𝑙𝑙 = 2 𝐴𝑒 𝑅𝑇 𝑒 2𝑅𝑇 = 2 𝐴𝑒 2𝑅𝑇 = 2 𝐴𝑒 −2 𝑅𝑇

Comparing the above result with Arrhenius rate constant 𝑘 = 𝐴𝑒 −𝐸𝑎/𝑅𝑇 , we get

1 1 (188)
𝐸𝑎 = (2𝐸2 + 𝐸1 − 𝐸5 ) = 𝐸2 − (𝐸1 − 𝐸5 )
2 2
Recalling that 𝐸1 , 𝐸2 and 𝐸5 are the activation energies for initiation, propagation and termination steps; the
equation (188) can be written in more general form i.e.

1 (189)
𝐸𝑎 = 𝐸𝑝 − (𝐸𝑖 − 𝐸𝑡 )
𝑛
Where n is the order of termination step w.r.t. chain carrier.

Copyright © Mandeep Dalal


364 A Textbook of Physical Chemistry – Volume I

 Chain Length
The chain length for any chemical chain reaction may simply be defined as the average number of
times that the closed cycle of chain propagation steps is repeated.
In other words, the chain length of any chain reaction is equal to the ration of the rate of the overall
reaction to the rate of the initiation step in which the active particles are generated. Mathematically,

𝑅𝑝 (190)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑅𝑖

Where 𝑅𝑝 and 𝑅𝑖 are the rate of propagation and rate of initiation respectively. The concept can be well
understood by taking the example of thermal decomposition of acetaldehyde and ethane.
 Chain length in Case of Thermal Decomposition of Acetaldehyde
The thermal decomposition of acetaldehyde is a typical case of stationary type chain reactions for
which the overall reaction can be written as

CH3 CHO ⟶ CH4 + CO (191)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (192)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: CH3 CHO → ˙CH3 + ˙CHO

𝑘2 (193)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: ˙CH3 + CH3 CHO → CH4 + ˙CH2 CHO

𝑘3 (194)
˙CH2 CHO → ˙CH3 + CO

𝑘4 (195)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: ˙CH3 + ˙CH3 → CH3 CH3

The net rate of formation of CH4 must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[CH4 ] (196)
= 𝑘[CH3 CHO]3/2
dt
Using equation (192), the rate of initiation can be given as

𝑅𝑖 = 𝑘1 [CH3 CHO] (197)

Using the values of 𝑅𝑝 and 𝑅𝑖 from equation (196, 197) in equation (190), we get the kinetic chain length as

𝑘[CH3 CHO]3/2 𝑘 (198)


𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ = = [CH3 CHO]1/2
𝑘1 [CH3 CHO] 𝑘1

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 365

 Chain Length in Case of Dehydrogenation of Ethane


The dehydrogenation of ethane is another typical case of stationary type chain reactions for which
the overall reaction can be written as given below.

CH3 CH3 ⟶ CH2 = CH2 + H2 (199)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (200)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: CH3 CH3 → ˙CH3 + ˙CH3

𝑘2 (201)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: ˙CH3 + CH3 CH3 → CH4 + ˙CH2 CH3

𝑘3 (202)
˙CH2 CH3 → CH2 = CH2 + ˙H

𝑘4 (203)
CH3 CH3 + ˙H → ˙CH2 CH3 + H2

𝑘5 (204)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: ˙CH2 CH3 + ˙H → CH3 CH3

The net rate of decomposition of ethane will be

𝑑[C2 H6 ] 𝑘1 𝑘3 𝑘4 1/2 (205)


− =( ) [CH3 CH3 ]
dt 𝑘5

Since the rate of propagation is simply equal to the overall rate law i.e. equation (205), the rate of initiation is

𝑅𝑖 = 𝑘1 [CH3 CH3 ] (206)

After using the values of 𝑅𝑝 and 𝑅𝑖 from equation (205, 206) in equation (190), we get the expression for
kinetic chain length as
1 (207)
𝑘 𝑘 𝑘 2
( 1 3 4 ) [CH3 CH3 ]
𝑘5
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑘1 [CH3 CH3 ]

or

1 𝑘1 𝑘3 𝑘4 1/2 [CH3 CH3 ] (208)


𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ = ( )
𝑘1 𝑘5 [CH3 CH3 ]

𝑘3 𝑘4 1/2 (209)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ = ( )
𝑘1 𝑘5

Copyright © Mandeep Dalal


366 A Textbook of Physical Chemistry – Volume I

 Rice-Herzfeld Mechanism of Organic Molecules Decomposition


(Acetaldehyde)
In 1934, Frank O. Rice and Karl F. Herzfeld performed an extensive study on the chain reactions and
developed special mechanics to account the observed rate laws. In the honor of researchers, this mechanism is
popularly known as Rice-Herzfeld mechanism. The conceptual foundation of this mechanism can easily be
understood by taking the example of an organic compound, acetaldehyde.
The pyrolysis of acetaldehyde is a typical case of stationary type chain reactions (n = 1) for which
the overall reaction can be written as given below.

CH3 CHO ⟶ CH4 + CO (210)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (211)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: CH3 CHO → ˙CH3 + ˙CHO

𝑘2 (212)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: ˙CH3 + CH3 CHO → CH4 + ˙CH2 CHO

𝑘3 (213)
˙CH2 CHO → ˙CH3 + CO

𝑘4 (214)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: ˙CH3 + ˙CH3 → CH3 CH3

The net rate of formation of CH4 must be equal to the sum of the rate of formation and the rate of disappearance
of the same i.e.

𝑑[CH4 ] (215)
= 𝑘2 [˙CH3 ][CH3 CHO]
dt
Now, in order to obtain the overall rate expression, we need to apply the steady-state approximation on the
[˙CH3 ] and [˙CH2 CHO] first i.e.

d[˙CH3 ] (216)
= 0 = 𝑘1 [CH3 CHO] − 𝑘2 [˙CH3 ][CH3 CHO] + 𝑘3 [˙CH2 CHO] − 2𝑘4 [˙CH3 ]2
dt
Similarly,

d[˙CH2 CHO] (217)


= 0 = 𝑘2 [˙CH3 ][CH3 CHO] − 𝑘3 [˙CH2 CHO]
dt
Taking negative both side of equation (217), we have

−𝑘2 [˙CH3 ][CH3 CHO] + 𝑘3 [˙CH2 CHO] = 0 (218)

Using the above result in equation (216), we get

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 367

𝑘1 [CH3 CHO] + 0 − 2𝑘4 [˙CH3 ]2 = 0 (219)

𝑘1 1/2 (220)
[˙CH3 ] = ( ) [CH3 CHO]1/2
2𝑘4

After putting the value of [˙CH3 ] from equation (220) in equation (215), we have

𝑑[CH4 ] 𝑘1 1/2 (221)


= 𝑘2 ( ) [CH3 CHO]1/2 [CH3 CHO]
dt 2𝑘4

𝑑[CH4 ] 𝑘1 1/2 (222)


= 𝑘2 ( ) [CH3 CHO]3/2
dt 2𝑘4

Now consider a new constant as

𝑘1 1/2 (223)
𝑘 = 𝑘2 ( )
2𝑘4

Using in equation (222), we get

𝑑[CH4 ] (224)
= 𝑘[CH3 CHO]3/2
dt
The kinetic chain length for the same can be obtained by dividing the rate of formation of the product by rate
of initiation step i.e.

𝑅𝑝 (225)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑅𝑖

Where 𝑅𝑝 and 𝑅𝑖 are the rate of propagation and rate of initiation respectively. Now since the rate of
propagation is simply equal to the overall rate law i.e. equation (224), the rate of initiation can be given as

𝑅𝑖 = 𝑘1 [CH3 CHO] (226)

After using the values of 𝑅𝑝 and 𝑅𝑖 from equation (224, 226) in equation (225), we get the expression for
kinetic chain length as

𝑘[CH3 CHO]3/2 (227)


𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ =
𝑘1 [CH3 CHO]

or

𝑘 (228)
𝐾𝑖𝑛𝑒𝑡𝑖𝑐 𝑐ℎ𝑎𝑖𝑛 𝑙𝑒𝑛𝑔𝑡ℎ = [CH3 CHO]1/2
𝑘1

Copyright © Mandeep Dalal


368 A Textbook of Physical Chemistry – Volume I

 Branching Chain Reactions and Explosions (H2-O2 Reaction)


On the basis of the chain carriers produced in each propagation step, the chain reactions can primarily
be classified into two categories; non-branched or the stationary reactions and branched or the non-stationary
reactions. A typical chain reaction can be written as given below.

𝑘1 (229)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: 𝐴→ 𝑅∗

𝑘2 (230)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: 𝑅∗ + 𝐴 → 𝑃 + 𝑛𝑅 ∗

𝑘3 (231)
𝑇𝑒𝑟𝑚𝑖𝑛𝑎𝑡𝑖𝑜𝑛: 𝑅∗ → 𝑑𝑒𝑠𝑡𝑟𝑢𝑐𝑡𝑖𝑜𝑛

Where P, R* and A represent the product, radical (or chain carrier) and reactant molecules, respectively. The
symbol represents a number that equals unity for stationary chain reactions and greater than one for non-
stationary or the branched-chain reactions. Furthermore, the destruction of the radical in the termination step
can occur either via its collision with another radical (in gas phase) or by striking the walls of the container.
The steady-state approximation can be employed to determine the concentration of intermediate or
the radical involved in the propagation step. The general procedure for which is to put the overall rate of
formation equals to zero. In other words, the rate of formation of R * must be equal to the rate of decomposition
of the same i.e.

𝑅𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 𝑅 ∗ = 𝑅𝑎𝑡𝑒 𝑜𝑓 𝑑𝑖𝑠𝑎𝑝𝑝𝑒𝑎𝑟𝑎𝑛𝑐𝑒 𝑜𝑓 𝑅 ∗ (232)

𝑘1 [𝐴] = 𝑘3 [𝑅 ∗ ] − 𝑘2 (𝑛 − 1)[𝑅 ∗ ][𝐴] (233)

𝑑[𝑅 ∗ ] (234)
𝑘1 [𝐴] − 𝑘3 [𝑅∗ ] + 𝑘2 (𝑛 − 1)[𝑅 ∗ ][𝐴] = 0 =
𝑑𝑡
or

−𝑘3 [𝑅∗ ] + 𝑘2 (𝑛 − 1)[𝑅 ∗ ][𝐴] = −𝑘1 [𝐴] (235)

𝑘1 [𝐴] (236)
[𝑅 ∗ ] =
𝑘2 (1 − 𝑛)[𝐴] + 𝑘3

Now because the destruction of the radical in the termination step can occur either via its collision with another
radical (in gas phase) or via striking the walls of the container, the rate constant k3 can be replaced by the sum
of the rate constants of two i.e. 𝑘3 = 𝑘𝑤 + 𝑘𝑔 . After using the value of 𝑘3 in equation (236), we have

𝑘1 [𝐴] (237)
[𝑅 ∗ ] =
𝑘2 (1 − 𝑛)[𝐴] + 𝑘𝑤 + 𝑘𝑔

or

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 369

𝑘1 [𝐴] (238)
[𝑅 ∗ ] =
−𝑘2 (𝑛 − 1)[𝐴] + 𝑘𝑤 + 𝑘𝑔

For non-stationary or branched chain reactions, more and more radicals are generated in each successive step
(𝑛 > 1). In other words, for every radical consumed in a chain propagation step, more than one chain carriers
or radicals are generated. Therefore, the possibility of explosion arises when

𝑘2 (𝑛 − 1)[𝐴] = 𝑘𝑤 + 𝑘𝑔 (239)

The situation can be explained in terms of equation (238) because the abovementioned condition will make
the denominator zero, and therefore, making radical concentration to approach infinite. Now because the rate
is usually proportional to radical’s concentration, a very high rate may lead to an explosion.
 Conditions for Different Types of Explosion limits
However, it is also worthy to mention that the occurrence of explosion depends upon the experimental
temperature and pressure. Typically, three explosion limits are observed as the pressure of the reacting system
is raised. To understand the different explosion limits, the behavior of the denominator in equation (238) must
be analysed with pressure.
1. The first explosion limit: When the pressure is very low, the movement of chain carriers towards the wall
of the container is very fast resulting in a very large rate of radicals’ destruction at walls i.e. high 𝑘𝑤 .
Conversely, the probability of radicals colliding with each other at very low pressure resulting in a very low
value of 𝑘𝑔 . Therefore, we can conclude that the denominator in equation (238) has a sufficiently large positive
value at low pressure giving smooth progression of the reaction without any explosion. However, with the rise
in pressure, 𝑘𝑤 declines very rapidly than the increase in 𝑘𝑔 . When a certain pressure value is achieved, the
explosion condition is satisfied i.e.

𝑘2 (𝑛 − 1)[𝐴] = 𝑘𝑤 + 𝑘𝑔 (240)

Which is the first explosion limit.


2. The second explosion limit: The first explosion limit exists over a wide range of pressure. However, if the
pressure is raised continuously, the movement of chain carriers towards the wall of the container is more and
more hindered resulting in a very small rate of radicals’ destruction at walls i.e. low 𝑘𝑤 . Conversely, the
probability of radicals colliding with each other further increases with pressure resulting in the very large value
of 𝑘𝑔 . Eventually, we can conclude that the denominator in equation (238) again becomes sufficiently positive
giving a steady progression of the reaction. Which is the second explosion limit.
3. The third explosion limit: After the second explosion limit, the steady reaction-rate continues over a range
of pressure. However, if the pressure is raised continuously, the heat produced in various propagating steps
would not be able to leave the system at a rate equal to the rate at which it is produced. Therefore, this thermal
effect will keep supporting the rate, eventually leading to a thermally-induced explosion. Which is the third
explosion limit.

Copyright © Mandeep Dalal


370 A Textbook of Physical Chemistry – Volume I

Figure 1. The variation of reaction with pressure in branching chain reactions (left) and variation of
relative pressure with time on a logarithmic scale (right).

 Reaction Profile of H2-O2 Explosion


The reaction between H2 and O2 is a typical example in which all three explosion limits are observed.
The net reaction is

2H2 + O2 ⟶ 2H2 O (241)

Furthermore, the elementary steps for the same can be proposed as

𝑘1 (242)
𝐼𝑛𝑖𝑡𝑖𝑎𝑡𝑖𝑜𝑛: H2 + O2 → HO˙2 + H˙

𝑘2 (243)
𝑃𝑟𝑜𝑝𝑎𝑔𝑎𝑡𝑖𝑜𝑛: H2 + HO˙2 → HO˙ + H2 𝑂

𝑘3 (244)
H2 + HO˙ → H˙ + H2 𝑂

𝑘4 (245)
H˙ + O2 → 𝐻𝑂˙ + O˙

𝑘5 (246)
O˙ + H2 → 𝐻𝑂˙ + H˙

The step 3rd and 4th produce more radicals than they consume, producing an explosion.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 371

 Kinetics of (One Intermediate) Enzymatic Reaction: Michaelis-Menten


Treatment
In 1901, a French Chemist, Victor Henri, proposed that enzyme reactions are actually initiated by a
bond formed between the substrate and the enzyme. The concept was further developed by German researches
Leonor Michaelis and Canadian scientist Maud Menten, who examined the kinetics of an enzymatic reaction
mechanism of invertase. The final model of Michaelis and Menten's treatment was published in 1913 and
gained immediate popularity among the scientific community.

𝑘1 𝑘2 (247)
𝐸+𝑆 ⇌ 𝐸𝑆 ⟶ 𝑃
𝑘−1

Where E is the enzyme that binds to a substrate S, and forms a complex ES. This enzyme-substrate complex
then turns into product P. The overall rate of product formation should be

𝑑[𝑃] (248)
+ = 𝑟 = 𝑘2 [𝐸𝑆]
𝑑𝑡
Since [𝐸𝑆] is intermediate, the steady-state approximation can be applied i.e.

𝑅𝑎𝑡𝑒 𝑜𝑓 𝑓𝑜𝑟𝑚𝑎𝑡𝑖𝑜𝑛 𝑜𝑓 [𝐸𝑆] = 𝑅𝑎𝑡𝑒 𝑜𝑓 𝑑𝑒𝑐𝑜𝑚𝑝𝑜𝑠𝑖𝑡𝑖𝑜𝑛 𝑜𝑓 [𝐸𝑆] (249)

𝑘1 [𝐸][𝑆] = 𝑘−1 [𝐸𝑆] + 𝑘2 [𝐸𝑆] (250)

𝑘1 [𝐸][𝑆] = (𝑘−1 + 𝑘2 )[𝐸𝑆] (251)

𝑘1 [𝐸][𝑆] (252)
[𝐸𝑆] =
𝑘−1 + 𝑘2

Now, defining a new parameter named as Michaelis- Menten constant (𝐾𝑚 ) with the following expression,

𝑘−1 + 𝑘2 (253)
𝐾𝑚 =
𝑘1

equation (252) takes the form

[𝐸][𝑆] (254)
[𝐸𝑆] =
𝐾𝑚

Now, if we put the [𝐸𝑆] concentration given above into equation (248), we will get the rate of enzyme-
catalyzed reaction in terms of substrate concentration and enzyme left unused i.e. [𝐸]. This is again a problem
that can be solved if we obtain the rate expression in terms of total enzyme concentration i.e. [𝐸0 ]; which
would make the calculation of catalytic efficiency much easier. In order to do so, recall the total enzyme
concentration mathematically i.e.

Copyright © Mandeep Dalal


372 A Textbook of Physical Chemistry – Volume I

[𝐸0 ] = [𝐸] + [𝐸𝑆] (255)

Where [𝐸] is the enzyme concentration left unused whereas [𝐸𝑆] represents the enzyme concentration that is
bound with the substrate. Rearranging equation (253), we get

[𝐸] = [𝐸0 ] − [𝐸𝑆] (256)

After using the value of [𝐸] from equation (256) in equation (254), we have

{[𝐸0 ] − [𝐸𝑆]}[𝑆] (257)


[𝐸𝑆] =
𝐾𝑚

[𝐸0 ][𝑆] − [𝐸𝑆][𝑆] (258)


[𝐸𝑆] =
𝐾𝑚

Rearranging for [𝐸𝑆] again

𝐾𝑚 [𝐸𝑆] = [𝐸0 ][𝑆] − [𝐸𝑆][𝑆] (259)

𝐾𝑚 [𝐸𝑆] + [𝐸𝑆][𝑆] = [𝐸0 ][𝑆] (260)

[𝐸0 ][𝑆] (261)


[𝐸𝑆] =
𝐾𝑚 + [𝑆]

After putting the value of [𝐸𝑆] from above equation into equation (248), we get

𝑘2 [𝐸0 ][𝑆] (262)


𝑟=
𝐾𝑚 + [𝑆]

If the rate of the enzyme-catalyzed reaction is recorded in the very initial stage, the substrate concentration can
also be replaced by the initial substrate concentration i.e. [𝑆0 ]. Therefore, the initial enzyme-catalyzed rate (𝑟0 )
should be

𝑘2 [𝐸0 ][𝑆0 ] (263)


𝑟0 =
𝐾𝑚 + [𝑆0 ]

Which is the well-known Michaelis-Menten equation.


Although the equation (263) is the complete form of the Michaelis-Menten equation that has
measurable quantities like[𝐸0 ] and [𝑆0 ], the more popular form of the same includes the maximum initial rate.
At very large initial substrate concentration, the 𝐾𝑚 can simply be neglected in comparison to [𝑆0 ] in the
denominator, which makes the equation (263) to take the form

𝑘2 [𝐸0 ][𝑆0 ] (264)


𝑟𝑚𝑎𝑥 = = 𝑘2 [𝐸0 ]
[𝑆0 ]

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 373

After using the 𝑘2 [𝐸0 ] = 𝑟𝑚𝑎𝑥 in equation (263), we have

𝑟𝑚𝑎𝑥 [𝑆0 ] (265)


𝑟0 =
𝐾𝑚 + [𝑆0 ]

Which is the most popular form of the Michaelis-Menten equation. The plot of the initial reaction rate for the
enzyme-catalyzed reaction is given below.

Figure 2. The variation of initial reaction rate as a function of substrate concentration for an enzyme-
catalyzed reaction.

 Order of Enzyme Catalysed Reactions


The reaction kinetics of the enzyme-catalyzed reactions can be well understood by looking at the
general form of Michaelis-Menten equation.
i) At high [S0]: When initial substrate concentration is very high, Km can be neglected, and the equation (263)
takes the form

𝑘2 [𝐸0 ][𝑆0 ] (266)


𝑟0 = = 𝑘2 [𝐸0 ]
[𝑆0 ]

Hence, the reaction will be zero-order w.r.t. substrate concentration and first order w.r.t the total enzyme
concertation.

Copyright © Mandeep Dalal


374 A Textbook of Physical Chemistry – Volume I

ii) At low [S0]: When initial substrate concentration is low, [S0] can be neglected, and the equation (263) takes
the form

𝑘2 (267)
𝑟0 = [𝐸 ][𝑆 ]
𝐾𝑚 0 0

Hence, the reaction will be the first-order w.r.t. substrate concentration and first order w.r.t the total enzyme
concertation as well.
 Calculation of Catalytic Efficiency
The catalytic efficiency of enzyme-catalyzed reactions is simply the maximum overall reaction rate
per unit of enzyme concentration. When initial substrate concentration is very high, Km can be neglected, and
the equation (263) takes the form

𝑘2 [𝐸0 ][𝑆0 ] (268)


𝑟𝑚𝑎𝑥 = = 𝑘2 [𝐸0 ]
[𝑆0 ]
𝑟𝑚𝑎𝑥 (269)
𝑘2 =
[𝐸0 ]

Hence, k2 is simply equal to the catalytic efficiency of enzyme-catalyzed reactions.


 Calculation of Michaelis-Menten Constant
The Michaelis-Menten constant of enzyme-catalyzed reactions can simply be calculated once the
maximum initial reaction rate is known. When initial substrate concentration is very high, Km can be neglected,
and the equation (263) takes the form

𝑘2 [𝐸0 ][𝑆0 ] (270)


𝑟𝑚𝑎𝑥 = = 𝑘2 [𝐸0 ]
[𝑆0 ]

After using the 𝑘2 [𝐸0 ] = 𝑟𝑚𝑎𝑥 in equation (263), we have

𝑟𝑚𝑎𝑥 [𝑆0 ] (271)


𝑟0 =
𝐾𝑚 + [𝑆0 ]

When the initial reaction rate is half of the maximum initial rate i.e. 𝑟0 = 𝑟𝑚𝑎𝑥 /2, we have

𝑟𝑚𝑎𝑥 𝑟𝑚𝑎𝑥 [𝑆0 ] (272)


=
2 𝐾𝑚 + [𝑆0 ]

𝐾𝑚 + [𝑆0 ] = 2[𝑆0 ] (273)

𝐾𝑚 = [𝑆0 ] (274)

Hence, the concentration at which the initial rate in half of the maximum initial rate, will be equal to 𝐾𝑚 .

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 375

 Evaluation of Michaelis's Constant for Enzyme-Substrate Binding by


Lineweaver-Burk Plot and Eadie-Hofstee Methods
The conventional approach to determine the Michaelis's constant (Km) involve the plot of initial
reaction rate vs initial substrate concentration. Then the Michaelis-Menten constant of enzyme-catalyzed
reactions can simply be calculated once the maximum initial reaction rate is known. The typical Michaelis-
Menten is used to fit the data is given below.

𝑟𝑚𝑎𝑥 [𝑆0 ] (275)


𝑟0 =
𝐾𝑚 + [𝑆0 ]

Where 𝑟𝑚𝑎𝑥 = 𝑘2 [𝐸0 ] is the maximum initial rate at [𝐸0 ] enzyme concentration whereas [𝑆0 ] is the initial
substrate concentration. Finally, the concentration at which the initial rate in half of the maximum initial rate,
will be equal to 𝐾𝑚 .
This conventional method, however, suffers from serious limitations like the need for lots of data
points until the initial rate becomes constant. Therefore, two other methods are quite popular to find Michaelis's
constant without getting the actual plateau in the initial reaction rate. In this section, we will discuss two such
methods for the evaluation of Michaelis's constant for enzyme-substrate binding.
 The Lineweaver-Burk Plot to Evaluate Michaelis's Constant
In order to understand the Lineweaver-Burk method for the determination of the maximum initial
rate (𝑟𝑚𝑎𝑥 ) and Michaelis's constant, we need to rearrange the general Michaelis-Menten equation first i.e. the
reciprocal of equation (275) as given below.

1 𝐾𝑚 + [𝑆0 ] (276)
=
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ]

or

1 𝐾𝑚 [𝑆0 ] (277)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥 [𝑆0 ]

or

1 𝐾𝑚 1 1 (278)
= +
[𝑆
𝑟0 𝑟𝑚𝑎𝑥 0 ] 𝑟𝑚𝑎𝑥

Which is the equation of the straight line (𝑦 = 𝑚𝑥 + 𝑐). Therefore, if we plot the reciprocal of initial reaction
rate vs the reciprocal of initial substrate concentration; the slope and intercepts will give 𝐾𝑚 /𝑟𝑚𝑎𝑥 and 1/𝑟𝑚𝑎𝑥 ,
respectively.

Copyright © Mandeep Dalal


376 A Textbook of Physical Chemistry – Volume I

Figure 3. The Lineweaver-burk plot for enzyme-catalyzed reactions to evaluate the maximum initial rate
(𝑟𝑚𝑎𝑥 ) and Michaelis's constant.

It is also worthy to mention that if the intercept is further extrapolated, it will lead to the intercept on the x-axis
that equals to −1/𝐾𝑚 .
 The Eadie-Hofstee Plot to Evaluate Michaelis's Constant
In order to understand the Eadie-Hofstee method for the determination of the maximum initial rate
(𝑟𝑚𝑎𝑥 ) and Michaelis's constant, we need to rearrange the general Michaelis-Menten equation first i.e. the
reciprocal of equation (275) as given below.

1 𝐾𝑚 + [𝑆0 ] (279)
=
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ]

or

1 𝐾𝑚 [𝑆0 ] (280)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥 [𝑆0 ]

or

1 𝐾𝑚 1 1 (281)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥

Multiplying both sides by 𝑟0 , we get

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 377

𝑟0 𝐾𝑚 𝑟0 𝑟0 (282)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥

Rearranging for 𝑟0 /[𝑆0 ], we get

𝑟0 𝑟0 𝐾𝑚 𝑟0 (283)
− =
𝑟0 𝑟𝑚𝑎𝑥 𝑟𝑚𝑎𝑥 [𝑆0 ]
𝑟0 𝑟0 𝑟𝑚𝑎𝑥 𝑟0 𝑟𝑚𝑎𝑥 (284)
= −
[𝑆0 ] 𝑟0 𝐾𝑚 𝑟𝑚𝑎𝑥 𝐾𝑚
𝑟0 𝑟𝑚𝑎𝑥 𝑟0 (285)
= −
[𝑆0 ] 𝐾𝑚 𝐾𝑚

𝑟0 1 𝑟𝑚𝑎𝑥 (286)
=− 𝑟0 +
[𝑆0 ] 𝐾𝑚 𝐾𝑚

Which is the equation of the straight line (𝑦 = 𝑚𝑥 + 𝑐). Therefore, if we plot the ratio of initial reaction rate
to initial substrate concentration vs the initial reaction rate; the slope and intercepts will give −1/𝐾𝑚 and
𝑟𝑚𝑎𝑥 /𝐾𝑚 , respectively.

Figure 4. The Eadie-Hofstee plot for enzyme-catalyzed reactions to evaluate the maximum initial rate
(𝑟𝑚𝑎𝑥 ) and Michaelis's constant.

It is also worthy to mention that if the intercept is further extrapolated, it will lead to the intercept on the x-axis
that equals to 𝑟𝑚𝑎𝑥 .

Copyright © Mandeep Dalal


378 A Textbook of Physical Chemistry – Volume I

 Competitive and Non-Competitive Inhibition


An enzyme inhibitor is a compound that binds to an enzyme and decreases its overall activity, and
the phenomenon is typically known as “enzyme inhibition”. Two of the most common enzyme inhibition
processes will be discussed in this section.
 Competitive Inhibition
In the case of competitive enzyme inhibition, the binding of an inhibitor prevents the binding of the
substrate and the enzyme. This type of behavior is actually achieved by blocking the binding site of the target
molecule (the active site) by some means. The competitive enzyme inhibition can be classified into two types
as discussed below.
1. Fully competitive inhibition: The fully competitive inhibition occurs when an enzyme (E) binds with the
substrate (S) and inhibitor (I) separately, and it is only the enzyme-substrate complex (ES) that will convert
into the product. This whole process can be described mathematically as

𝑘1 𝑘2 (287)
𝐸+𝑆 ⇌ 𝐸𝑆 ⟶ 𝐸 + 𝑃
𝑘−1

𝑘3 (288)
𝐸+𝐼 ⇌ 𝐸𝐼
𝑘−3

After applying the steady-state approximation on ES, we have

𝑑[𝐸𝑆] (289)
= 𝑘1 [𝐸][𝑆] − 𝑘−1 [𝐸𝑆] − 𝑘2 [𝐸𝑆] = 0
𝑑𝑡
𝑘1 [𝐸][𝑆] = 𝑘−1 [𝐸𝑆] + 𝑘2 [𝐸𝑆] (290)

𝑘1 [𝐸][𝑆] (291)
[𝐸𝑆] =
𝑘−1 + 𝑘2

When [𝑆0 ] ≫ [𝐸0 ], we can assume [𝑆0 ] ≈ [𝑆], and also

[𝐸0 ] = [𝐸] + [𝐸𝑆] + [𝐸𝐼] (292)

[𝐸] = [𝐸0 ] − [𝐸𝑆] − [𝐸𝐼] (293)

Now recalling the equilibrium constant for inhibition equilibria i.e.

𝑘3 [𝐸𝐼] (294)
𝐾3 = =
𝑘−3 [𝐸][𝐼]

1 [𝐸][𝐼] (295)
𝐾I = =
𝐾3 [𝐸𝐼]

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 379

Hence, we can say

[𝐸][𝐼] (296)
[𝐸𝐼] =
𝐾I

After using the value of [𝐸𝐼] from equation (296) into equation (293), we have

[𝐸][𝐼] (297)
[𝐸] = [𝐸0 ] − [𝐸𝑆] −
𝐾I

𝐾I [𝐸0 ] − 𝐾I [𝐸𝑆] − [𝐸][𝐼] (298)


[𝐸] =
𝐾I

𝐾I [𝐸] = 𝐾I [𝐸0 ] − 𝐾I [𝐸𝑆] − [𝐸][𝐼] (299)

𝐾I [𝐸] + [𝐸][𝐼] = 𝐾I [𝐸0 ] − 𝐾I [𝐸𝑆] (300)

𝐾I [𝐸0 ] − 𝐾I [𝐸𝑆] (301)


[𝐸] =
𝐾I + [𝐼]

Using the above-derived result in equation (291), we have

𝑘1 [𝑆] 𝐾I [𝐸0 ] − 𝐾I [𝐸𝑆] [𝑆] 𝐾I [𝐸0 ] − 𝐾I [𝐸𝑆] (302)


[𝐸𝑆] = =
𝑘−1 + 𝑘2 𝐾I + [𝐼] 𝐾𝑚 𝐾I + [𝐼]

𝐾I [𝐸0 ][𝑆] − 𝐾I [𝐸𝑆][𝑆] (303)


[𝐸𝑆] =
𝐾𝑚 𝐾I + 𝐾𝑚 [𝐼]

Now rearranging further for [𝐸𝑆], we get

𝐾𝑚 𝐾I [𝐸𝑆] + 𝐾𝑚 [𝐼][𝐸𝑆] = 𝐾I [𝐸0 ][𝑆] − 𝐾I [𝐸𝑆][𝑆] (304)

𝐾𝑚 𝐾I [𝐸𝑆] + 𝐾𝑚 [𝐼][𝐸𝑆] + 𝐾I [𝐸𝑆][𝑆] = 𝐾I [𝐸0 ][𝑆] (305)

𝐾I [𝐸0 ][𝑆] (306)


[𝐸𝑆] =
𝐾𝑚 𝐾I + 𝐾𝑚 [𝐼] + 𝐾I [𝑆]

Since the rate of formation of product is

𝑟0 = 𝑘2 [𝐸𝑆] (307)

Using the value of [𝐸𝑆] from equation (306) in equation (307), we get

𝑘2 𝐾I [𝐸0 ][𝑆] (308)


𝑟0 =
𝐾𝑚 𝐾I + 𝐾𝑚 [𝐼] + 𝐾I [𝑆]

Since 𝑘2 [𝐸0 ] is 𝑟𝑚𝑎𝑥 , the equation (308) takes the form

Copyright © Mandeep Dalal


380 A Textbook of Physical Chemistry – Volume I

𝑟𝑚𝑎𝑥 𝐾I [𝑆] (309)


𝑟0 =
𝐾𝑚 𝐾I + 𝐾𝑚 [𝐼] + 𝐾I [𝑆]

Taking the reciprocal to get Lineweaver-Burk plot

1 𝐾𝑚 𝐾I + 𝐾𝑚 [𝐼] + 𝐾I [𝑆] (310)


=
𝑟0 𝑟𝑚𝑎𝑥 𝐾I [𝑆]

1 𝐾𝑚 𝐾I 𝐾𝑚 [𝐼] 𝐾I [𝑆] (311)


= + +
𝑟0 𝑟𝑚𝑎𝑥 𝐾I [𝑆] 𝑟𝑚𝑎𝑥 𝐾I [𝑆] 𝑟𝑚𝑎𝑥 𝐾I [𝑆]

1 𝐾𝑚 𝐾𝑚 [𝐼] 1 1 (312)
=( + ) +
𝑟0 𝑟𝑚𝑎𝑥 𝑟𝑚𝑎𝑥 𝐾I [𝑆] 𝑟𝑚𝑎𝑥

Rearranging further and using initial substrate concentration, we get

1 𝐾𝑚 [𝐼] 1 1 (313)
= (1 + ) +
𝑟0 𝑟𝑚𝑎𝑥 𝐾I [𝑆0 ] 𝑟𝑚𝑎𝑥

After comparing with equation without enzyme inhibition i.e.

1 𝐾𝑚 1 1 (314)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥

it is clear that the intercepts in both the equations are the same whereas the slope has been increased from
𝐾𝑚 /𝑟𝑚𝑎𝑥 to (𝐾𝑚 /𝑟𝑚𝑎𝑥 )(1 + [𝐼]/𝐾𝐼 ). This implies that enzyme inhibition changed the 𝐾𝑚 to 𝐾𝑚 (1 + [𝐼]/𝐾𝐼 ).

Figure 5. Lineweaver-Burk plot for enzyme-catalyzed reaction with fully competitive enzyme inhibition.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 381

2. Partially competitive inhibition: The partially competitive inhibition occurs when the enzyme (E) binds
with the substrate (S) and inhibitor (I) simultaneously, and complex ES and EI also combine with I and S to
give EIS. Also, besides the enzyme-substrate complex (ES), the complex EIS will also convert into the product
with the same rate of reaction. This whole process can be described mathematically as

𝐾1′ 𝑘2 (315)
𝐸 + 𝑆 ⇌ 𝐸𝑆 ⟶ 𝐸 + 𝑃

𝐾2′ (316)
𝐸 + 𝐼 ⇌ 𝐸𝐼

𝐾3′ (317)
𝐸𝐼 + 𝑆 ⇌ 𝐸𝐼𝑆

𝐾4′ (318)
𝐸𝑆 + 𝐼 ⇌ 𝐸𝐼𝑆

𝑘2 (319)
𝐸𝐼𝑆 ⟶ 𝐸𝐼 + 𝑃

According to classical Michaelis-Menten equation 𝐾𝑚 = (𝑘2 + 𝑘−1 )/𝑘1 ; however, if 𝑘2 ≪ 𝑘−1 , we have
𝐾𝑚 = 𝑘−1 /𝑘1 or 𝐾𝑚 = 1/𝐾1′ . Now, set the following results

1 [𝐸][𝑆] (320)
𝐾𝑚 = ′ =
𝐾1 [𝐸𝑆]

1 [𝐸][𝐼] (321)
𝐾2 = ′ =
𝐾2 [𝐸𝐼]

1 [𝐸𝐼][𝑆] (322)
𝐾3 = ′ =
𝐾3 [𝐸𝐼𝑆]

1 [𝐸𝑆][𝐼] (323)
𝐾4 = ′ =
𝐾4 [𝐸𝐼𝑆]

Following enzyme conservation, we have

[𝐸0 ] = [𝐸] + [𝐸𝑆] + [𝐸𝐼] + [𝐸𝐼𝑆] (324)

Using values of [𝐸𝑆], [𝐸𝐼] and [𝐸𝐼𝑆] from equation (320-322) into equation (324), we get

[𝐸][𝑆] [𝐸][𝐼] [𝐸][𝐼][𝑆] (325)


[𝐸0 ] = [𝐸] + + +
𝐾𝑚 𝐾2 𝐾2 𝐾3

𝐾𝑚 [𝐸0 ] (326)
[𝐸] =
𝐾𝑚 (1 + [𝐼]/𝐾2 ) + [𝑆](1 + 𝐾𝑚 [𝐼]/𝐾2 𝐾3 )

The overall reaction rate of product formations should be

Copyright © Mandeep Dalal


382 A Textbook of Physical Chemistry – Volume I

𝑟0 = 𝑘2 [𝐸𝑆] + 𝑘2 [𝐸𝐼𝑆] (327)

After using values of [𝐸𝑆] and [𝐸𝐼𝑆] from equation (320, 322) into equation (327), we get

[𝐸][𝑆] [𝐸][𝐼][𝑆] (328)


𝑟0 = 𝑘2 + 𝑘2
𝐾𝑚 𝐾2 𝐾3

Substituting the value of [𝐸] from equation (326) in (328), and rearranging at initial substrate concentration

𝑘2 [𝐸0 ][𝑆0 ] (329)


𝑟0 =
{𝐾𝑚 (1 + [𝐼]/𝐾2 )/(1 + 𝐾𝑚 [𝐼]/𝐾2 𝐾3 )} + [𝑆0 ]

Since 𝑘2 [𝐸0 ] is 𝑟𝑚𝑎𝑥 , equation (329) takes the following form after the reciprocal to get Lineweaver-Burk plot

1 {𝐾𝑚 (1 + [𝐼]/𝐾2 )/(1 + 𝐾𝑚 [𝐼]/𝐾2 𝐾3 )} 1 1 (330)


= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥

After comparing with equation without enzyme inhibition i.e.

1 𝐾𝑚 1 1 (331)
= +
[𝑆
𝑟0 𝑟𝑚𝑎𝑥 0 ] 𝑟𝑚𝑎𝑥

it is clear that intercepts in both the equations are the same whereas the slope has been varied from 𝐾𝑚 /𝑟𝑚𝑎𝑥
to{𝐾𝑚 (1 + [𝐼]/𝐾2 )/(1 + 𝐾𝑚 [𝐼]/𝐾2 𝐾3 )}/𝑟𝑚𝑎𝑥 . This implies that enzyme inhibition changed the 𝐾𝑚 to
{𝐾𝑚 (1 + [𝐼]/𝐾2 )/(1 + 𝐾𝑚 [𝐼]/𝐾2 𝐾3 )}.

Figure 6. The Lineweaver-Burk plot for partially competitive enzyme inhibition.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 383

 Non-Competitive Inhibition
Non-competitive inhibition may simply be defined as the enzyme inhibition where the inhibitor
decreases the activity of the enzyme catalysis and binds equally well to the enzyme whether or not it has
already bound the substrate. The non-competitive enzyme inhibition can be classified into two types as
discussed below.
1. Fully non-competitive inhibition: The fully non-competitive inhibition occurs when an enzyme (E) binds
with inhibitor (I), and complex ES and EI also combine with I and S to give EIS. However, it is only the
enzyme-substrate complex (ES) that converts into the product. This whole process can be described
mathematically as

𝐾1′ 𝑘2 (332)
𝐸 + 𝑆 ⇌ 𝐸𝑆 ⟶ 𝐸 + 𝑃

𝐾2′ (333)
𝐸 + 𝐼 ⇌ 𝐸𝐼

𝐾1′ (334)
𝐸𝐼 + 𝑆 ⇌ 𝐸𝐼𝑆

𝐾2′ (335)
𝐸𝑆 + 𝐼 ⇌ 𝐸𝐼𝑆

According to classical Michaelis-Menten equation 𝐾𝑚 = (𝑘2 + 𝑘−1 )/𝑘1 ; however, if 𝑘2 ≪ 𝑘−1 , we have
𝐾𝑚 = 𝑘−1 /𝑘1 or 𝐾𝑚 = 1/𝐾1′ . Now, set the following results

1 [𝐸][𝑆] [𝐸][𝑆] (336)


𝐾𝑚 = ′ = =
𝐾1 [𝐸𝑆] [𝐸𝐼𝑆]

1 [𝐸][𝐼] [𝐸𝑆][𝐼] (337)


𝐾I = ′ = =
𝐾2 [𝐸𝐼] [𝐸𝐼𝑆]

Following enzyme conservation, we have

[𝐸0 ] = [𝐸] + [𝐸𝑆] + [𝐸𝐼] + [𝐸𝐼𝑆] (338)

[𝐸] = [𝐸0 ] − [𝐸𝑆] − [𝐸𝐼] − [𝐸𝐼𝑆] (339)

Using values of [𝐸𝑆], [𝐸𝐼] and [𝐸𝐼𝑆] from equation (336-337) into equation (338), we get

[𝐸][𝑆] [𝐸][𝐼] [𝐸][𝐼][𝑆] (340)


[𝐸0 ] = [𝐸] + + +
𝐾𝑚 𝐾I 𝐾𝑚 𝐾I

[𝐸0 ] (341)
[𝐸] =
(1 + [𝑆]/𝐾𝑚 )(1 + [𝐼]/𝐾𝐼 )

The overall reaction rate of product formations should be

Copyright © Mandeep Dalal


384 A Textbook of Physical Chemistry – Volume I

𝑟0 = 𝑘2 [𝐸𝑆] (342)

After using values of [𝐸𝑆] from equation (336) into equation (342), we get

[𝐸][𝑆] (343)
𝑟0 = 𝑘2
𝐾𝑚

Substituting the value of [𝐸] from equation (341) in (343), and rearranging at initial substrate concentration

𝑘2 [𝐸0 ][𝑆0 ] (344)


𝑟0 =
(𝐾𝑚 + [𝑆0 ])(1 + [𝐼]/𝐾𝐼 )

Since 𝑘2 [𝐸0 ] is 𝑟𝑚𝑎𝑥 , the equation (344) takes the form

𝑟𝑚𝑎𝑥 [𝑆0 ] (345)


𝑟0 =
(𝐾𝑚 + [𝑆0 ])(1 + [𝐼]/𝐾𝐼 )

Taking the reciprocal to get Lineweaver-Burk plot

1 𝐾𝑚 [𝐼] 1 1 [𝐼] (346)


= (1 + ) + (1 + )
𝑟0 𝑟𝑚𝑎𝑥 𝐾I [𝑆0 ] 𝑟𝑚𝑎𝑥 𝐾I

After comparing with equation without enzyme inhibition i.e.

1 𝐾𝑚 1 1 (347)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥

it can be clearly seen that intercept is increased from 1/𝑟𝑚𝑎𝑥 to (1/𝑟𝑚𝑎𝑥 )(1 + [𝐼]/𝐾𝐼 ) whereas the Km has
remained the same.

Figure 7. The Lineweaver-Burk plot for fully non-competitive enzyme inhibition.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 385

2. Partially non-competitive inhibition: The partially non-competitive inhibition occurs when an enzyme (E)
binds with inhibitor (I), and complex ES and EI also combine with I and S to give EIS. However, unlike fully
non-competitive inhibition, besides the enzyme-substrate complex (ES), the complex will also convert into the
product. This whole process can be described mathematically as

𝐾1′ 𝑘2 (348)
𝐸 + 𝑆 ⇌ 𝐸𝑆 ⟶ 𝐸 + 𝑃

and

𝐾2′ (349)
𝐸 + 𝐼 ⇌ 𝐸𝐼

and

𝐾1′ (350)
𝐸𝐼 + 𝑆 ⇌ 𝐸𝐼𝑆

and

𝐾2′ (351)
𝐸𝑆 + 𝐼 ⇌ 𝐸𝐼𝑆

and

𝑘′ (352)
𝐸𝐼𝑆 ⟶ 𝐸𝐼 + 𝑃

According to classical Michaelis-Menten equation 𝐾𝑚 = (𝑘2 + 𝑘−1 )/𝑘1 ; however, if 𝑘2 ≪ 𝑘−1 , we have
𝐾𝑚 = 𝑘−1 /𝑘1 or 𝐾𝑚 = 1/𝐾1′ . Now, set the following results

1 [𝐸][𝑆] (353)
𝐾𝑚 = =
𝐾1′ [𝐸𝑆]

and

1 [𝐸][𝐼] (354)
𝐾𝐼 = ′ =
𝐾2 [𝐸𝐼]

and

1 [𝐸𝐼][𝑆] (355)
𝐾𝑚 = ′ =
𝐾1 [𝐸𝐼𝑆]

and

1 [𝐸𝑆][𝐼] (356)
𝐾𝐼 = ′ =
𝐾2 [𝐸𝐼𝑆]

Copyright © Mandeep Dalal


386 A Textbook of Physical Chemistry – Volume I

Following enzyme conservation, we have

[𝐸0 ] = [𝐸] + [𝐸𝑆] + [𝐸𝐼] + [𝐸𝐼𝑆] (357)

Using values of [𝐸𝑆], [𝐸𝐼] and [𝐸𝐼𝑆] from equation (353-356) into equation (357), we get the following
expression.

[𝐸][𝑆] [𝐸][𝐼] [𝐸][𝐼][𝑆] (358)


[𝐸0 ] = [𝐸] + + +
𝐾𝑚 𝐾𝐼 𝐾𝑚 𝐾𝐼

or

[𝐸0 ] (359)
[𝐸] =
(1 + ([𝑆]/𝐾𝑚 )(1 + [𝐼]/𝐾𝐼 )

The overall reaction rate of product formations should be

𝑟0 = 𝑘2 [𝐸𝑆] + 𝑘 ′ [𝐸𝐼𝑆] (360)

After using the values of [𝐸𝑆] and [𝐸𝐼𝑆] from equation (353-356) into equation (360), we get the following
result.

[𝐸][𝑆] [𝐸][𝐼][𝑆] (361)


𝑟0 = 𝑘2 + 𝑘′
𝐾𝑚 𝐾𝑚 𝐾𝐼

Now, substituting the value of [𝐸] from equation (359) in (361), and rearranging at the initial substrate
concentration

(𝑘2 [𝐸0 ][𝑆0 ] + 𝑘 ′ [𝐸0 ][𝑆0 ][𝐼]/𝐾𝐼 )/(1 + [𝐼]/𝐾𝐼 ) (362)


𝑟0 =
𝐾𝑚 + [𝑆0 ]

Using 𝑘2 [𝐸0 ] is 𝑟𝑚𝑎𝑥 and then taking the reciprocal to get the Lineweaver-Burk plot, we get the following
relation.

1 𝐾𝑚 (1 + [𝐼]/𝐾I ) 1 (1 + [𝐼]/𝐾I ) (363)


= +
𝑟0 (𝑟𝑚𝑎𝑥 + 𝑘 [𝐸0 ][𝐼]/𝐾𝐼 ) [𝑆0 ] (𝑟𝑚𝑎𝑥 + 𝑘 ′ [𝐸0 ][𝐼]/𝐾𝐼 )

After comparing the result given above with the general equation without enzyme inhibition, i.e., we can
conclude some important points.

1 𝐾𝑚 1 1 (364)
= +
𝑟0 𝑟𝑚𝑎𝑥 [𝑆0 ] 𝑟𝑚𝑎𝑥

it can be clearly seen that intercept is increased from 1/𝑟𝑚𝑎𝑥 to (1 + [𝐼]/𝐾𝐼 )/(𝑟𝑚𝑎𝑥 + 𝑘 ′ [𝐸0 ][𝐼]/𝐾𝐼 ) whereas
the Km has remained the same.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 387

Figure 8. The Lineweaver-Burk plot for partially non-competitive enzyme inhibition.

Copyright © Mandeep Dalal


388 A Textbook of Physical Chemistry – Volume I

 Problems
Q 1. What are chemical chain reactions? Discuss their general kinetics.
Q 2. Derive and discuss the rate law for the decomposition of ethane.
Q 3. Define photochemical reactions. How they are different from thermochemical reactions?
Q 4. What is the photochemical quantum yield of a reaction? Derive and discuss the same for photochemical
combination H2-Br2 case.
Q 5. Derive and discuss the Rice-Herzfeld mechanism of decomposition of Acetaldehyde.
Q 6. Define the chain length.
Q 7. Calculate the expression for the chain length for the dehydrogenation of ethane.
Q 8. What are the branching chain reactions? How they lead to the situation of the explosion?
Q 9. Derive and discuss the conventional Michaelis-Menten equation for enzyme-catalyzed reactions.
Q 10. How would you treat Michaelis-Menten equations to get Lineweaver-Burk plot?
Q 11. Discuss the Eadie-Hofstee method for the evaluation of Michaelis-Menten constant in enzyme-catalyzed
reactions.
Q 12. What is enzyme inhibition? Discuss the fully competitive enzyme inhibition in detail.

Copyright © Mandeep Dalal


CHAPTER 7 Chemical Dynamics – II 389

 Bibliography
[1] B. R. Puri, L. R. Sharma, M. S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2008.
[2] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[3] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[4] S. A. Arrhenius, Über die Dissociationswärme und den Einfluß der Temperatur auf den Dissociationsgrad
der Elektrolyte. Z. Phys. Chem. 4 (1889) 96–116.
[5] S. A. Arrhenius, Über die Reaktionsgeschwindigkeit bei der Inversion von Rohrzucker durch Säuren Z.
Phys. Chem. 4 (1889) 226–248.
[6] K. J. Laidler, Chemical Kinetics, Harper & Row, New York, USA, 1987.
[7] K. J. Laidler, The World of Physical Chemistry, Oxford University Press, Oxford, UK, 1993.
[8] H. Eyring, The Activated Complex in Chemical Reactions, J. Chem. Phys. 3 (1935) 107-115.
[9] K. L. Kapoor, A Textbook of Physical Chemistry Volume 5, Macmillan Publishers, New Delhi, India, 2011.

Copyright © Mandeep Dalal


CHAPTER 8
Electrochemistry – II: Ion Transport in Solutions
 Ionic Movement Under the Influence of an Electric Field
In order to imagine the conduction process in electrolytic solutions at the atomic level, we can follow
two approaches which are somewhat different in their initial assumptions.
The first approach includes the visualization of ionic movements governed by the diffusion
phenomenon first and then studying the perturbation of this ionic movement by an externally applied electric
field. Since it is a well-known fact that the diffusion of ions is simply the movement of ions from a high-
numbered region to a low-numbered region. In other words, we can say the ionic diffusion is the result of a
concentration-gradient in which a particular type of ions travel from a high concentration region towards a low
concentration region until a homogeneity in the concentration is reached. Now, although the net movement of
ions stops after the loss of concentration gradient, the individual ionic movement still happens but with zero
mean displacements. In other words, we can say that in a homogeneous ionic solution, the ions can move
randomly in any direction resulting in a zero net diffusion.

Figure 1. The movement of positive ions from higher concentration to lower concentration.

Now since the ions are charged particles, the movements of these ions are strongly affected when an electric
field is applied. From the laws electrostatic interactions, we can conclude that the cations will prefer to move
towards the negative electrode whereas the anions will prefer to move towards the positive electrode. More
specifically, the application of an electric field makes the ions to adopt a single direction in space, which is a
direction along or opposite to the direction of the applied field. Therefore, the ions drift under the applied field
and stop their random walk.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 391

The second approach to study the phenomenon of electrolytic conduction at the atomic level includes
the framing of the drift of only one ion under the externally applied field. The electric field would make the
ion to accelerate as per Newton’s second law. Now if the ion is in the vacuum, it would show an acceleration
until it strikes with the respective electrode. However, it will not happen since a large number of other ions
also present in the same electrolytic solution along with the solvent as well. Consequently, the ion is almost
bound to collide with other ions or solvent particles in its journey. The ion will stop for some time and then
will start to accelerate again. This stop-start phenomenon will impart a discontinuity in the speed and direction
of this moving ion. It means that ionic movement is not very much smooth but actually a resistance is offered
by the surrounding medium. Therefore, we can say that the application of an external electric field will make
the ion move towards the oppositely charged electrode but in a stops-starts and zigzag fashion.

Figure 2. The general depiction movement of ions under the influence of the external electric field.

Since an ion starts moving towards the positively charged electrode only after the application of the
electric field; the initial velocity before that can simply be neglected because it arises from random collisions,
which can be in any random direction. However, after applying the electric field, the ion feels a force that
makes it move in the same direction, i.e., the direction of the electrostatic force. In other words, the electric
field will create an additional velocity component on the ion under consideration that drifts the ion to the
oppositely charged electrode. Now, let 𝐹 be the force vector that imparts a drift velocity 𝑣𝑑 ; and then using
Newton’s second law of motion states that this force divided by the particle’s mass is simply equal to the
acceleration. From the general expressions for acceleration, we have

𝑑𝑣 (1)
𝑎=
𝑑𝑡
And

Copyright © Mandeep Dalal


392 A Textbook of Physical Chemistry – Volume I

𝐹 (2)
𝑎=
𝑚
From equation (1) and equation (2), we have

𝐹 𝑑𝑣 (3)
=
𝑚 𝑑𝑡
Now although the time between two collisions may vary significantly, we can use a mean time τ for simplicity
which can be formulated as (if N collisions take place in ‘t’ time) given below.
𝑡 (4)
𝜏=
𝑁
Now because the drift velocity is imparted to the ion by the external force, its value must be equal to the
product of meantime and the acceleration due to force, i.e.,

𝑑𝑣 (5)
𝑣𝑑 = 𝜏
𝑑𝑡
Using the value of 𝑑𝑣/𝑑𝑡 from equation (3) in equation (5), we get

𝐹 (6)
𝑣𝑑 = 𝜏
𝑚
It is obvious from the above relation that the meantime is related to the drift velocity showing that the jumps
between collisions affect ionic movement. Besides, it is also clear that the drift velocity is directly proportional
to the driving force of the applied electric field. The ionic flux can be formulated in terms of drift velocity as
given below.

𝐹𝑙𝑢𝑥 = 𝐼𝑜𝑛𝑖𝑐 𝑐𝑒𝑛𝑐𝑒𝑛𝑡𝑟𝑎𝑡𝑖𝑜𝑛 × 𝐷𝑟𝑖𝑓𝑡 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦 (7)

Hence, since the drift velocity is directly proportional to the electric force simulating conduction, the flux must
also be proportional to the magnitude of the electric field, i.e.,

𝐹𝑙𝑢𝑥 ∝ 𝐸𝑙𝑒𝑐𝑡𝑟𝑖𝑐 𝑓𝑖𝑒𝑙𝑑 (8)

The nature of equation (6) also unveils the situation where the flux or the drift velocity no longer holds the
direct proportionality with the applied electric field. For equation (6), it is very important to assume that during
the collision, the velocity component imparted to the ion by applied electric field vanishes completely and the
ion starts as a full-fresher each time. If this is not satisfied, these leftover velocity components would add up
after every collision and the real velocity, in that case, would be much greater than the calculation given by
equation (6). In other words, equation (6) will no longer be valid. Therefore, we can conclude that for a
reasonable guess for drift velocity, the magnitude of the applied electric field must be very small.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 393

 Mobility of Ions
It has already been discussed in the previous section that the ions in a homogeneous electrolytic
solution move randomly with zero net displacements, and the situation changes when the external electric field
is applied. The applied field imparts a directive velocity component to the ion under consideration and makes
it move towards the oppositely charged electrode. This ion collides with other ions and drifts towards the
oppositely charged electrode with a stop-start and zig-zag fashion. The drift velocity (𝑣𝑑 ) of such ion is given
by the following relation.
𝜏 (9)
𝑣𝑑 = 𝐹
𝑚

Where 𝐹 is the force exerted upon the ion by applied field and m is the mass of the ion. The symbol 𝜏 represents
the mean lifetime between to collisions during the ionic drift.
It is obvious from the equation (9) that drift velocity is proportional to forces exerted by the electric
field and 𝜏/𝑚 is the constant of proportionality. The physical significance of the proportionality constant lies
in the fact that it becomes equal to the drift velocity when the force is unity, and therefore, represents the
“mobility nature” of the ion considered. In other words, we can say that the proportionality constant in equation
(9) represents the absolute mobility (𝑢̅𝑎𝑏𝑠 ) of the ion i.e.
𝜏 𝑣𝑑 (10)
𝑢̅𝑎𝑏𝑠 = =
𝑚 𝐹
The units of absolute mobility are cm s−1 dyne−1. Now, since the functional electric force is equal to the electric
force per unit charge; or the electric field (X) multiplied with the charge on the ion (𝑧𝑖 𝑒0 ) i.e.

𝐹 = 𝑧𝑖 𝑒0 𝑋 (11)

After using the value of 𝐹 from equation (11) in equation (10), we get
𝑣𝑑 (12)
𝑢̅𝑎𝑏𝑠 =
𝑧𝑖 𝑒0 𝑋

𝑣𝑑 = 𝑢̅𝑎𝑏𝑠 𝑧𝑖 𝑒0 𝑋 (13)

When, 𝑋 = 1 𝑣𝑜𝑙𝑡, the above equation takes the form

(𝑣𝑑 )1volt cm−1 = 𝑢̅𝑎𝑏𝑠 𝑧𝑖 𝑒0 = 𝑢̅𝑐𝑜𝑛𝑣 (14)

Where 𝑢̅𝑐𝑜𝑛𝑣 represents the conventional mobility of the ion with units cm2 Volt−1 s−1. Now although the
expressions of both types of mobilities are quite similar, it is worthy to note that the “absolute mobility” has a
broader domain of application because any force that governs the drift velocity can be used. On the other hand,
the conventional mobility is pretty much limited to the electric force only.

Copyright © Mandeep Dalal


394 A Textbook of Physical Chemistry – Volume I

 Ionic Drift Velocity and Its Relation with Current Density


In this section, we will try to explain how the ionic movement is quantitatively related to current
density flowing through an electrolytic solution under the influence of an applied electric field. To do so, we
need to understand the ionic drift velocity and then its relationship with current density.
 Ionic Drift Velocity
When an ion in the electrolytic solution is placed under the externally applied electric field, the
electric field will make the ion to accelerate as per Newton’s second law. Now if the ion is in the vacuum, it
would show an acceleration until it strikes with the respective electrode. However, it will not happen since a
large number of other ions also present in the same electrolytic solution along with the solvent as well.
Consequently, the ion is almost bound to collide with other ions or solvent particles in its journey. The ion will
stop for some time and then will start to accelerate again. This stop-start phenomenon will impart a
discontinuity in the speed and direction of this moving ion. It means that ionic movement is not very much
smooth but actually a resistance is offered by the surrounding medium. Therefore, we can say that the
application of the external electric field will make the ion move towards the oppositely charged electrode but
in a stops-starts and zigzag fashion. An ion starts moving towards the positively-charged electrode only after
the application of the electric field. The initial velocity before that can simply be neglected because it arises
from random collisions which can be in any random direction. However, after applying the electric field, the
ion feels a force that makes the ion move in the same direction, i.e., the direction of the electrostatic force.
Now let F be the force vector that imparts drift velocity 𝑣𝑑 , then form Newton’s second law of motion states
that this force divided by the particle’s mass is simply equal to the acceleration.

𝐹 𝑑𝑣 (15)
=
𝑚 𝑑𝑡
Now although the time between two collisions may vary significantly, we can use a mean time τ for simplicity
which can be formulated as (if N collisions take place in ‘t’ time) given below.
𝑡 (16)
𝜏=
𝑁
Now because the drift velocity is imparted to the ion by the external force, its value must be equal to product
meantime and the acceleration due to this force i.e.

𝑑𝑣 (17)
𝑣𝑑 = 𝜏
𝑑𝑡
Using the value of 𝑑𝑣/𝑑𝑡 from equation (15) in equation (17), we get

𝐹 (18)
𝑣𝑑 = 𝜏
𝑚
It is obvious that the drift velocity is directly proportional to the driving force of the applied electric field.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 395

 The Relationship between Ionic Drift Velocity and Current Density


It is obvious from the equation (18) that drift velocity is proportional to forces exerted by the electric
field and 𝜏/𝑚 is the constant of proportionality. The physical significance of the proportionality constant lies
in the fact that it becomes equal to the drift velocity when the force is unity, and therefore, represents the
“mobility nature” of the ion considered. In other words, we can say that the proportionality constant in equation
(18) represents the absolute mobility (𝑢̅𝑎𝑏𝑠 ) of the ion, i.e.,
𝜏 𝑣𝑑 (19)
𝑢̅𝑎𝑏𝑠 = =
𝑚 𝐹
The units of absolute mobility are cm s−1 dyne−1. Now, since the functional electric force is equal to the electric
force per unit charge; or the electric field (X) multiplied with the charge on the ion (𝑧𝑖 𝑒0 ), i.e.,

𝐹 = 𝑧𝑖 𝑒0 𝑋 (20)

After using the value of 𝐹 from equation (20) in equation (19), we get
𝑣𝑑 (21)
𝑢̅𝑎𝑏𝑠 =
𝑧𝑖 𝑒0 𝑋

𝑣𝑑 = 𝑢̅𝑎𝑏𝑠 𝑧𝑖 𝑒0 𝑋 (22)

When, 𝑋 = 1 𝑣𝑜𝑙𝑡, the above equation takes the form

(𝑣𝑑 )1volt cm−1 = 𝑢̅𝑎𝑏𝑠 𝑧𝑖 𝑒0 = 𝑢̅𝑐𝑜𝑛𝑣 (23)

Where 𝑢̅𝑐𝑜𝑛𝑣 represents the conventional mobility of the ion with units cm2 Volt−1 s−1. Now although the
expressions of both types of mobilities are quite similar, it is worthy to note that the “absolute mobility” has a
broader domain of application because any force that governs the drift velocity can be used. On the other hand,
the conventional mobility is pretty much limited to the electric force only.
Now consider a plane with unit area perpendicular to the direction of ionic movement under the
influence of the externally applied field. Now although the cations and anions move in opposite directions,
both types of ions will pass through this transit. If 𝑣+ is the drift velocity of cation, then all the cations present
within 𝑣+ cm of this transit plane will pass through it. Let 𝑗+ be the cationic flux that represents the total
number of mole of ions passing through this area in every second. Therefore, the cationic flux must be equal
to the multiplication of the corresponding volume (1 cm2 × 𝑣+ cm) and the concentration of cations (𝑐+ ) in
moles cm−3. Mathematically, we can say that

𝑗+ = 𝑐+ 𝑣+ (24)

The current density (𝐽+ ) or the charge flowing through this transit plane per second due to this cationic flux
can simply be obtained by multiplying the cationic flux by the charge carried by 1 mole of ions (𝑧+ 𝐹), i.e.,

Copyright © Mandeep Dalal


396 A Textbook of Physical Chemistry – Volume I

𝐽+ = 𝑐+ 𝑣+ 𝑧+ 𝐹 (25)

Similarly, If 𝑣− is the drift velocity of anion and 𝑐− the concentration of anions in moles cm−3, the anionic
current density may simply be written as

𝐽− = 𝑐− 𝑣− 𝑧− 𝐹 (26)

Where 𝑧− is the charge number of the anion and F is Faraday constant. In general, we can write for the ith
species as

𝐽𝑖 = 𝑐𝑖 𝑣𝑖 𝑧𝑖 𝐹 (27)

Where 𝑧𝑖 and 𝑐𝑖 are the charge number and concentration in moles cm−3 of the ith type of ion.

Figure 3. The general depiction ionic movement through transit plane under applied field.

Hence, the total current density from all the ionic species must be the summation of their individual current
densities, i.e.,

𝐽 = ∑ 𝐽𝑖 = ∑ 𝑐𝑖 𝑣𝑖 𝑧𝑖 𝐹 (28)
𝑖 𝑖

For univalent electrolytes like NaCl, we have 𝑧+ = 𝑧− = 𝑧 and 𝑐+ = 𝑐− = 𝑐; therefore, we can write

𝐽 = 𝑧𝑐𝐹(𝑣+ + 𝑣− ) (29)

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 397

Now we can start to relate the ionic drift velocity with more generally measurable quantities in the phenomenon
of electrolytic conductance like molar conductivity, equivalent conductivity or simply the conductance. To do
so, use the fundamental expression for the ionic drift velocity form equation (23) in equation (28), i.e.,

𝐽 = ∑ 𝑧𝑖 𝐹𝑐𝑖 (𝑢𝑐𝑜𝑛𝑣 )𝑖 𝑋 (30)


𝑖

Now since the 𝐽/𝑋 is equal to specific conductivity (σ), the above equation takes the form

𝐽 (31)
𝜎= = ∑ 𝑧𝑖 𝐹𝑐𝑖 (𝑢𝑐𝑜𝑛𝑣 )𝑖
𝑋
𝑖

Therefore, for univalent electrolytes, we have

𝐽 (32)
𝜎= = 𝑧𝐹𝑐[(𝑢𝑐𝑜𝑛𝑣 )+ + (𝑢𝑐𝑜𝑛𝑣 )− ]
𝑋
It is clear from the above expression that the specific conductivity is directly proportional to the concentration
of the electrolytic solution, which can be explained in terms of the fact that the number of ions per unit volume
changes with dilution.
Now, in order to connect the ionic drift velocities with molar conductivity (𝛬𝑚 ) for monovalent
electrolytes, recalling the expression for molar conductivity here, i.e.,
𝜎 (33)
𝛬𝑚 =
𝑐
After putting the value of specific conductivity from equation (32), the above equation takes the form

𝑧𝐹𝑐[(𝑢𝑐𝑜𝑛𝑣 )+ + (𝑢𝑐𝑜𝑛𝑣 )− ] (34)


𝛬𝑚 =
𝑐
𝛬𝑚 = 𝑧𝐹[(𝑢𝑐𝑜𝑛𝑣 )+ + (𝑢𝑐𝑜𝑛𝑣 )− ] (35)

For equivalent conductivity (𝛬𝑒𝑞 ) of monovalent electrolytes, recall its correlation with molar conductivity i.e.

𝛬𝑚 (36)
𝛬𝑒𝑞 =
𝑧
Substituting the value of molar conductivity from equation (35) in equation (36), we get

𝑧𝐹[(𝑢𝑐𝑜𝑛𝑣 )+ + (𝑢𝑐𝑜𝑛𝑣 )− ] (37)


𝛬𝑒𝑞 =
𝑧
𝛬𝑒𝑞 = 𝐹[(𝑢𝑐𝑜𝑛𝑣 )+ + (𝑢𝑐𝑜𝑛𝑣 )− ] (38)

Hence, 𝛬𝑒𝑞 will be independent of concentration only if the ionic mobility doesn’t vary with concentration.

Copyright © Mandeep Dalal


398 A Textbook of Physical Chemistry – Volume I

 Einstein Relation Between the Absolute Mobility and Diffusion Coefficient


Since it is a well-known fact that the diffusion of ions is simply the zig-zag walking of ions from a
high-numbered region to a low-numbered region. In other words, we can say the ionic diffusion is the result
of a concentration gradient in which a particular type of ions travel from a high concentration region towards
a low concentration region until a homogeneity in concentration is reached. On the other hand, the conduction
or the ionic migration is a result of the drift velocity component imparted to the ions by the electric force.
However, it is important here to recall the fact that this velocity component does not stop the zig-zag walk of
diffusion but actually gets superimposed on it. Albert Einstein understood this and formulated a relation
between ionic mobility (𝑢̅𝑎𝑏𝑠 ) and diffusion coefficient (D).
Now, since the conduction, as well as the diffusion, are irreversible processes, they cannot be treated
by equilibrium statistical mechanics or by the equilibrium thermodynamics. However, the situation can be
considered as a pseudo-equilibrium if the conduction and diffusion take place in the opposite direction but
with same rates. To do so, consider an electrolytic solution of salt MX in which some of the cations are
radioactive in nature. Now assume that M+ ions are present in higher concentrations in one region and in lower
concentration in some other region. In other words, the tracer ions are present with a concentration gradient.
According to Fick’s law of diffusion, the overall diffusion flux (𝐽𝐷 ) must be

𝑑𝑐 (39)
𝑗𝐷 = −𝐷
𝑑𝑥
After applying the electric field, the tracer ions will feel the field and will start to move towards the opposite
electrode. The drift velocity can be given as

𝑣𝑑 = 𝑢̅𝑎𝑏𝑠 𝐹 (40)

The current density produced by this drift velocity is

𝐽 = 𝑧+ 𝑐𝐹𝑣𝑑 (41)

The conduction flux can be obtained by dividing the current density by charge carried by one mole of ions i.e.

𝑧+ 𝑐𝐹𝑣𝑑 (42)
𝑗𝑐 =
𝑧+ 𝐹

or

𝑗𝑐 = 𝑐𝑣𝑑 (43)

After using the expression of drift velocity from equation (40) in the above expression, we get

𝑗𝑐 = 𝑐 𝑢̅𝑎𝑏𝑠 𝐹 (44)

The strength of the applied electric field is varied in such a way that the conduction flux and diffusion flux are
equal and opposite. Mathematically, it should be like

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 399

𝑗𝑐 = −𝑗𝐷 (45)

𝑗𝑐 + 𝑗𝐷 = 0 (46)

After using values of 𝑗𝐷 and 𝑗𝑐 from equations (39, 44) in the above expression, we get

𝑑𝑐 (47)
𝑐 𝑢̅𝑎𝑏𝑠 𝐹 − 𝐷 =0
𝑑𝑥
or

𝑑𝑐 𝑐𝑢̅𝑎𝑏𝑠 𝐹 (48)
=
𝑑𝑥 𝐷
Since there is no net flow of ions, and therefore, this pseudo-equilibrium can be studied by Boltzmann law.

Figure 4. The pseudo-equilibrium when diffusion flux and conduction flux are equal and opposite.

Owing to the x-dependent variation, recall the ionic concentration at distance x, i.e.,

𝑐 = 𝑐0 𝑒 −𝑈/𝑘𝑇 (49)

Where 𝑐0 is the ionic concentration in the zero potential region while 𝑈 is the potential energy of the ion under
consideration in the externally applied electric field. Differentiating the above equation w.r.t. x, we have

𝑑𝑐 1 𝑑𝑈 (50)
= −𝑐0 𝑒 −𝑈/𝑘𝑇
𝑑𝑥 𝑘𝑇 𝑑𝑥

Replacing 𝑐0 𝑒 −𝑈/𝑘𝑇 by c i.e. using equation (49), we get

Copyright © Mandeep Dalal


400 A Textbook of Physical Chemistry – Volume I

𝑑𝑐 𝑐 𝑑𝑈 (51)
=−
𝑑𝑥 𝑘𝑇 𝑑𝑥
Since the force is 𝐹 = −𝑑𝑈/𝑑𝑥, the above equation takes the form

𝑑𝑐 𝑐 (52)
= 𝐹
𝑑𝑥 𝑘𝑇
From equation (48) and equation (52), we get

𝑐𝑢̅𝑎𝑏𝑠 𝐹 𝑐 (53)
= 𝐹
𝐷 𝑘𝑇
𝑢̅𝑎𝑏𝑠 1 (54)
=
𝐷 𝑘𝑇
or

𝐷 = 𝑢̅𝑎𝑏𝑠 𝑘𝑇 (54)

Which is the famous Einstein relation between the absolute mobility and diffusion coefficient.
Furthermore, from the phenomenological treatment of the diffusion coefficient, it is also a quite well-
known correlation that

𝐷 = 𝐵𝑅𝑇 (55)

Where B represents the undetermined phenomenological coefficient and R is the gas constant. Now, comparing
equation (54) and equation (55), we have

𝑢̅𝑎𝑏𝑠 𝑘𝑇 = 𝐵𝑅𝑇 (56)

or

𝑢̅𝑎𝑏𝑠 𝑘𝑇 𝑢̅𝑎𝑏𝑠 𝑘 (57)


𝐵= =
𝑅𝑇 𝑅
Since 𝑁 = 𝑅/𝑘, the above equation can also be written as

𝑢̅𝑎𝑏𝑠 (58)
𝐵=
𝑁
It is obvious from the above equation that the phenomenological coefficient B can simply be defined as the
ratio of absolute mobility to the Avogadro number. Furthermore, The Einstein relation also connects the
phenomena of diffusion with force arising from viscous drag and force of electric field on the ion during its
drifting movement. Therefore, the formulation also forms the basis of Stokes-Einstein (viscosity and diffusion)
and Nernst–Einstein relation (equivalent conductivity and diffusion).

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 401

 The Stokes-Einstein Relation


Albert Einstein realized that an ion moving in an electrolytic solution is somewhat analogous to a
macroscopic sphere moving in a liquid medium. A macroscopic sphere can travel very fast if it is outside of
water or any other liquid; however, its velocity will definitely be affected if it is put in some liquid. The reduced
velocity of the sphere in liquid can be attributed to an opposing force exerted upon it by the diameter of the
sphere (d), viscosity of the medium (η), density of the medium (ρ) and the speed of the ion itself (𝑣). All these
factors are correlated mathematically to give “Reynolds number (Re)” as
𝜌 (59)
𝑅𝑒 = 𝑣𝑑
𝜂

If the hydrodynamics has a profile that makes the 𝑅𝑒 ≪ 1, Stokes proved that the dragging force (F) on the
sphere can be formulated by the following relation.

𝐹 = 6𝜋𝑟𝜂𝑣 (60)

Which is the famous Stokes’ law.


Now although the equation (60) is very useful in case of a macroscopic sphere moving in water or in
any other liquid, the applicability of the same to microscopic ions requires some testing first. One condition
that must be satisfied for the practicality of Stokes law to ion in solution is the very small value of Reynolds
number i.e. 𝑅𝑒 ≪ 1. After using the ionic diameter and typical drift velocity of ion, it is found that Reynolds
number is very small in comparison to unity, and therefore, suggests that Stokes law can be applied to ionic
movement. However, equation (60) would show deviations from experimental results if the ions tagged are
not completely spherical, and therefore, should be modified for such particles. It has been shown that for
cylindrical particles, the factor 6π must be replaced by 4π to get reasonable results. It is also worthy to note
that the Stokes law fails to explain the viscous drag on extremely small ions, justifying the need for some other
advanced models. Furthermore, besides the viscous drag, the presence of other ions also creates collisions and
stop-start zig-zag movement which makes it very difficult to apply Stokes’ law.
Albert Einstein developed a modified approach to correlate the viscosity with the diffusion coefficient
by suggesting that a driving force (−𝑑𝜇/𝑑𝑥) operates on the particles during diffusion which can be formulated
as given below.

𝑑𝜇 (61)
− = 6𝜋𝑟𝜂𝑣𝑑
𝑑𝑥
Where 𝑣𝑑 is the steady-state velocity of the ion under consideration. The right-hand side of the above equation
means that the driving force we assumed must be opposed by an equal and opposite resistive force given by
Stokes’ law.
Besides, when a charged particle moves in a polar solvent, solvent dipoles surround it from oppositely
charged ends; and this surrounding environment is destroyed and built up again and again due to movement,

Copyright © Mandeep Dalal


402 A Textbook of Physical Chemistry – Volume I

and takes time for it. In this relaxation process, a relaxation force is in operation which can be considered as
an additional frictional force on the tagged ion. Therefore, the dragging force expression is modified to
𝑠 (62)
𝐹 = 6𝜋𝜂𝑣𝑟 − 6𝜋𝜂𝑣
𝜀

Where 𝑠 = (4/9)(𝜏/6𝜋𝜂)𝑒02 /𝑟 3 whereas ε represents the dielectric constant of the solvent. The correction
factor sometimes can be very large but will be neglected for the simplicity of the derivation Stokes-Einstein
law. Now, from the definition of absolute mobility, i.e.,
𝑣𝑑 (63)
𝑢̅𝑎𝑏𝑠 =
𝐹
The drift velocity can be divided either by the diffusional driving force or by the equal and opposite viscous
force given by Stokes law. Therefore, using equation (61) in equation (63), we get

𝑣𝑑 𝑣𝑑 1 (64)
𝑢̅𝑎𝑏𝑠 = = =
−𝑑𝜇/𝑑𝑥 6𝜋𝑟𝜂𝑣𝑑 6𝜋𝑟𝜂

At this stage, recall the famous Einstein’s relation between the absolute mobility and diffusion coefficient, i.e.,

𝐷 = 𝑢̅𝑎𝑏𝑠 𝑘𝑇 (65)

Substituting the value of 𝑢̅𝑎𝑏𝑠 from equation (64) in the above expression, we have

𝑘𝑇 (66)
𝐷=
6𝜋𝑟𝜂

Which is the famous Stokes-Einstein’s relation between the viscosity and diffusion coefficient.
The Stokes-Einstein relation inspired the pioneering work of Perrin who studied the random walk of
a colloidal particle using an ultramicroscope and found that the mean square distance (< 𝑥 2 >) covered in ‘t’
time is correlated with the diffusion coefficient as given below.

< 𝑥2 > (67)


𝐷=
2𝑡
From the knowledge of the weight of colloidal particles and corresponding density, the magnitude of radius r
can be obtained, which in turn can be employed (along with medium’s viscosity) to find the value of Boltzmann
constant from the rearranged form of equation (66)

6𝜋𝑟𝜂𝐷 (68)
𝑘=
𝑇
Since 𝑘 = 𝑅/𝑁𝐴 , the value of the Avogadro number can be obtained from 𝑁𝐴 = 𝑅/𝑘. Furthermore, the Stokes-
Einstein law can also be used to find the value of conventional ionic mobility (𝑢̅𝑐𝑜𝑛𝑣 ). To do so, recall the
expression for conventional mobility i.e.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 403

𝑢̅𝑐𝑜𝑛𝑣 = 𝑢̅𝑎𝑏𝑠 𝑧𝑖 𝑒0 (69)

Now using the value of 𝑢̅𝑎𝑏𝑠 from equation (64), we have


𝑧𝑖 𝑒0 (70)
𝑢̅𝑐𝑜𝑛𝑣 =
6𝜋𝑟𝜂

The mobility given is typically labeled as Stokes mobility. The physical significance of the above equation lies
in the fact that it shows the correlation of conventional mobility with the charge on the ion, radius of the ion
and viscosity of the solvent used. It should also be noted that the equation does not explain the concentration
dependence of ion-ion interaction, and therefore, is an oversimplified approach.

 The Nernst-Einstein Equation


Just like the Stokes-Einstein’s relation found the connection between the viscosity and the diffusion
coefficient; another important equation, popularly called as Nernst-Einstein relation, correlated the diffusion
coefficient with equivalent conductivity. In order to derive the Nernst-Einstein’s equation, recall the expression
for equivalent conductivity (𝛬𝑒𝑞 ) in terms of conventional mobilities for 𝑧: 𝑧 valent electrolyte, i.e.,

𝛬𝑒𝑞 = 𝐹[(𝑢𝑐𝑜𝑛𝑣 )+ + (𝑢𝑐𝑜𝑛𝑣 )− ] (71)

Where F is the Faraday constant; whereas (𝑢𝑐𝑜𝑛𝑣 )+ and (𝑢𝑐𝑜𝑛𝑣 )− are the conventional mobilities of cation
and anion, respectively. Now, since for 𝑧: 𝑧 electrolyte 𝑧+ = 𝑧− = 𝑧, the conventional mobilities are

(𝑢𝑐𝑜𝑛𝑣 )+ = 𝑧+ 𝑒0 (𝑢̅𝑎𝑏𝑠 )+ = 𝑧𝑒0 (𝑢̅𝑎𝑏𝑠 )+ (72)

(𝑢𝑐𝑜𝑛𝑣 )− = 𝑧− 𝑒0 (𝑢̅𝑎𝑏𝑠 )+ = 𝑧𝑒0 (𝑢̅𝑎𝑏𝑠 )− (73)

Using equation (72, 73) in equation (71), we get

𝛬𝑒𝑞 = 𝐹[𝑧𝑒0 (𝑢̅𝑎𝑏𝑠 )+ + 𝑧𝑒0 (𝑢̅𝑎𝑏𝑠 )− ] (74)

𝛬𝑒𝑞 = 𝑧𝑒0 𝐹[(𝑢̅𝑎𝑏𝑠 )+ + (𝑢̅𝑎𝑏𝑠 )− ] (75)

From Einstein’s relation, we know that

𝐷+ (76)
(𝑢̅𝑎𝑏𝑠 )+ =
𝑘𝑇
also

𝐷− (77)
(𝑢̅𝑎𝑏𝑠 )− =
𝑘𝑇
Using equation (76, 77) in equation (75), we get

Copyright © Mandeep Dalal


404 A Textbook of Physical Chemistry – Volume I

𝐷+ 𝐷− (78)
𝛬𝑒𝑞 = 𝑧𝑒0 𝐹 [ + ]
𝑘𝑇 𝑘𝑇
𝑧𝑒0 𝐹 (79)
𝛬𝑒𝑞 = (𝐷+ + 𝐷− )
𝑘𝑇
Which is the popular Nernst-Einstein relation that allows us to find the value of equivalent conductivity just
by knowing the diffusion coefficient of cation and anion only.
Another popular form of the Nernst-Einstein equation can be obtained by multiplying and dividing
the right-hand side of equation (79) by Avogadro number as given below.

𝑧𝑒0 𝐹𝑁𝐴 (80)


𝛬𝑒𝑞 = (𝐷+ + 𝐷− )
𝑘𝑇𝑁𝐴

Since 𝑒0 𝑁𝐴 = 𝐹 and 𝑘𝑁𝐴 = 𝑅, the above equation takes the form

𝑧𝐹 2 (81)
𝛬𝑒𝑞 = (𝐷 + 𝐷− )
𝑅𝑇 +
It is also worthy to note that although nature is same, the equation (81) is more popular in electrochemical
literature than equation (79).

 Walden’s Rule
The Walden’s rule states that the product of the equivalent conductivity and the viscosity of the
solvent for a specific electrolyte at a given temperature is constant.
The Stokes-Einstein relation found the connection between the viscosity (𝜂) and the diffusion
coefficient (D); whereas the Nernst-Einstein relation correlates the equivalent conductivity (𝛬) and diffusion
coefficient. Therefore, a remarkable possibility is to eliminate the diffusion coefficient to correlate the viscosity
of the solvent with the equivalent conductivity. In order to do so, recall the Stokes-Einstein relation first i.e.

𝑘𝑇 (82)
𝐷=
6𝜋𝑟𝜂

Where k is the Boltzmann constant and η is the coefficient of viscosity. The symbol r represents the radius of
the ion and T is the temperature of the electrolytic solution. Now, recall the Nernst-Einstein relation i.e.

𝛬𝑒𝑞 𝑅𝑇 (83)
𝐷=
𝑧𝐹 2
Where z is the charge number and F is the Faraday constant. From equation (82) and equation (83), we get

𝑘𝑇 𝛬𝑒𝑞 𝑅𝑇 (84)
=
6𝜋𝑟𝜂 𝑧𝐹 2

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 405

𝑘𝑇𝑧𝐹 2 (85)
𝛬𝑒𝑞 =
6𝜋𝑟𝜂𝑅𝑇

Since 𝑒0 𝑁𝐴 = 𝐹 and 𝑘𝑁𝐴 = 𝑅, the above equation can be written in the following form

𝑘𝑇𝑧𝐹𝑒0 𝑁𝐴 𝑧𝐹𝑒0 (86)


𝛬𝑒𝑞 = =
6𝜋𝑟𝜂𝑇𝑘𝑁𝐴 6𝜋𝑟𝜂

𝑧𝐹𝑒0 (87)
𝛬𝑒𝑞 𝜂 =
6𝜋𝑟
Putting 𝑧𝐹𝑒0 /6𝜋 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡, the above equation can also be written as
𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 (88)
𝛬𝑒𝑞 𝜂 =
𝑟
Now, if the radius of the ion the solvated ion is same in solvents of different viscosities, the equation (88) is
reduced to

𝑧𝐹𝑒0 (89)
𝛬𝑒𝑞 𝜂 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 =
6𝜋𝑟
Which is the empirical Walden’s rule. The experimental data for potassium iodide in various solvents is given
for more clear picture is given below.

Table 1. The experimental data for KI in different solvents.

Solvent used Equivalent conductivity Viscocity of the solvent 𝛬𝑒𝑞 𝜂


(𝛬𝑒𝑞 ) (𝜂)

Acetophenone 39.8 0.01620 0.64476

Ethanol 50.9 0.01096 0.55786

Pyridine 71.4 0.00958 0.68401

Methanol 114.5 0.00545 0.62403

Propanone 185.4 0.00316 0.58586

Acetonitrile 198.3 0.00345 0.68414

It is obvious for the data listed in ‘Table 1’ that the product of equivalent conductivity and viscosity
of the solvent is almost constant with slight deviation, which can be attributed to different solvated radii in
different solvents.

Copyright © Mandeep Dalal


406 A Textbook of Physical Chemistry – Volume I

 The Rate-Process Approach to Ionic Migration


In order to understand the rate-process approach to ionic migration, recall the fundamental relation
between the ionic drift velocity (𝑣𝑑 ) and current density (𝐽) i.e.

𝐽 = 𝑧𝑐𝐹𝑣𝑑 (90)

Where z is the charge number and c is the concentration of the ions. The symbol F represents the Faraday
constant. The drift velocity (𝑣𝑑 ) is related to the macroscopic force
𝜏 (91)
𝑣𝑑 = 𝐹
𝑚
Where m is the mass of the ion. The symbol 𝜏 represents the mean lifetime between to collisions during the
ionic drift. The symbol 𝐹 is the viscose or electric force that can be formulated as

𝐹 = 6𝜋𝑟𝜂𝑣 𝑜𝑟 𝐹 = 𝑧𝑒0 𝑋 (92)

Where 𝜂 is the coefficient of viscosity and r is the radius of the cation. In the right-hand relation, z is the charge
number of the ion under consideration, 𝑒0 is the electronic charge and X is the applied electric field.
Furthermore, the drift velocity can also be assumed as the resultant velocity of the velocity of ions in
the direction of the force field (𝑣 ) and the velocity of ions in the opposite direction (𝑣⃖). Mathematically, we
can say as given below.

𝑣𝑑 = 𝑣 − 𝑣⃖ (93)

Now, since the velocity is also the ratio of average jump distance (l) to the average time between two successive
jumps (τ), we can also write as

𝑙 (94)
𝑣=
𝜏
Furthermore, the jump frequency i.e. number of jumps per unit of time (𝑘 = 1/𝜏) is simply the reciprocal of
the mean time between successive jumps. Therefore, the velocities can also be written as

𝑣 = 𝑙 𝑘⃗ (95)

or

𝑣⃖ = 𝑙 𝑘⃖⃗ (96)

Now in case of diffusion, the jumps of ions can be considered as a rate phenomenon for which the participating
ion must possess a minimum amount of free energy to activation to do so. It was found that

𝑘𝑇 −𝛥𝐺 ∗ /𝑅𝑇 (97)


𝑘⃗ = 𝑒

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 407

For the diffusion only (from right to left), the above equation can be relabelled with subscript D i.e.

𝑘𝑇 −𝛥𝐺 ∗ /𝑅𝑇 (98)


𝑘⃖⃗𝐷 = 𝑒 𝐷

However, when the electric field is applied, the diffusion of ions from right to left will be opposed. Therefore,
the work-done (w) in moving the ion from the equilibrium state to the barrier-maximum will be equal to the
product of the ionic charge (𝑧+ 𝑒0 ), and the potential difference between the activated state and equilibrium
position. Now assume that this potential difference is simply a fraction β of the overall active potential (Xl).

𝑤 = 𝑧+ 𝑒0 𝛽𝑋𝑙 (99)

Where X is the electric field and l is the distance between the two equilibrium states. For one mole of ions, the
equation (99) takes the form

𝑊 = 𝑁𝐴 𝑧+ 𝑒0 𝛽𝑋𝑙 (100)

or

𝑊 = 𝑧+ 𝐹𝛽𝑋𝑙 (101)

The overall depiction of the rate-process approach to the ionic migration phenomenon under the influence of
an external electric field is shown below.

Figure 5. The general depiction of the rate-process approach to ionic migration under the influence of an
external electric field.

Copyright © Mandeep Dalal


408 A Textbook of Physical Chemistry – Volume I

The process of activation can be described by the diagram given below.

Figure 6. The general depiction ionic movement from the right to left facing the activation barrier under
the influence of an external electric field.

The work done obtained in the abovementioned route will induce a free-energy change (𝛥𝐺𝑒∗) that
will contribute to the free energy of activation, i.e.,

𝛥𝐺𝑡𝑜𝑡𝑎𝑙 = 𝛥𝐺𝐷∗ + 𝛥𝐺𝑒∗ (102)

or

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 409


𝛥𝐺𝑡𝑜𝑡𝑎𝑙 = 𝛥𝐺𝐷∗ + 𝑧+ 𝐹𝛽𝑋𝑙 (103)

After using 𝛥𝐺𝑡𝑜𝑡𝑎𝑙



from equation (103), the equation (97) for “right to left” jumps in the presence of applied
electric field take the form

𝑘𝑇 −(𝛥𝐺 ∗ +𝑧 𝐹𝛽𝑋𝑙)/𝑅𝑇 (104)


𝑘⃖⃗ = 𝑒 𝐷 +

or

𝑘𝑇 −𝛥𝐺 ∗ /𝑅𝑇 −𝑧 𝐹𝛽𝑋𝑙/𝑅𝑇 (105)


𝑘⃖⃗ = 𝑒 𝐷 𝑒 +

Recalling expression for 𝑘⃖⃗𝐷 from equation (98), the equation (105) can also be written as

𝑘⃖⃗ = 𝑘𝐷 𝑒 −𝑧+𝐹𝛽𝑋𝑙/𝑅𝑇 (106)

Similarly, the jump frequency for the “left to right” movement may be found. However, since the cations are
moving along the field (left to right) in case, their movement is favored. The fraction of the barrier these ions
need to climb will be (1 − 𝛽). Finally, it is very important to note that the electrical work of activation should
be negative in “left to right” because field supports the ion. Therefore, the jump frequency for left to right
movement should look like

𝑘⃗ = 𝑘𝐷 𝑒 𝑧+𝐹(1−𝛽)𝑋𝑙/𝑅𝑇 (107)

Now if 𝛽 = 1/2, then 1 − 𝛽 will also be equal to 1/2. This transforms equation (106, 107) as

𝑘⃖⃗ = 𝑘𝐷 𝑒 −𝑝𝑋 (108)

and

𝑘⃗ = 𝑘𝐷 𝑒 𝑝𝑋 (109)

Where 𝑝 = 𝑧+ 𝐹𝑙/2𝑅𝑇. From equation (108), it is very obvious that 𝑘⃖⃗ < 𝑘𝐷 whereas equation (109) implies
that 𝑘⃗ > 𝑘𝐷 . Hence 𝑘⃗ > 𝑘⃖⃗.
Hence, we can conclude that the jumping frequency becomes anisotropic in the presence of externally
applied filed. The jumping frequency in the applied field's direction is higher than what is against.
Nevertheless, in the absence of field, the value of jump frequency is equal in all directions which is a
characteristic feature of the random walk. When the field is applied, this isotropy is abolished and the walk is
not completely random anymore. The ions start moving along the field creating a net current density. In other
words, we can say that ionic drift due to the field is the multiplication of perturbation induced by the field and
the random walk in the field’s absence.

Copyright © Mandeep Dalal


410 A Textbook of Physical Chemistry – Volume I

 The Rate-Process Equation for Equivalent Conductivity


The rate-process equation for equivalent conductivity can be derived by recalling the fundamental
relation between the ionic drift velocity (𝑣𝑑 ) and current density (𝐽) for cation first i.e.

𝐽 = 𝑧+ 𝑐𝐹𝑣𝑑 (110)

Where z is the charge number and c is the concentration of the ions. The symbol F represents the Faraday
constant. Also, the drift velocity can be assumed as the resultant velocity of the velocity of ions in the direction
of the force field (𝑣 ) and the velocity of ions in the opposite direction (𝑣⃖). Mathematically, we can say as given
below.

𝑣𝑑 = 𝑣 − 𝑣⃖ (111)

Now, since the velocity is also the ratio of average jump distance (l) to the average time between two successive
jumps (τ), we can also write as

𝑙 (112)
𝑣=
𝜏
Furthermore, the jump frequency i.e. number of jumps per unit of time (𝑘 = 1/𝜏) is simply the reciprocal of
the mean time between successive jumps. Therefore, the velocities can also be written as

𝑣 = 𝑙 𝑘⃗ (113)

𝑣⃖ = 𝑙 𝑘⃖⃗ (114)

From the rate-process approach to ionic migration, the values of 𝑘⃗ and 𝑘⃖⃗ were found to be

𝑘⃖⃗ = 𝑘𝐷 𝑒 −𝑝𝑋 (115)

and

𝑘⃗ = 𝑘𝐷 𝑒 𝑝𝑋 (116)

Where 𝑘𝐷 is the jumping frequency for diffusion and X is simply the electric field. The expression for symbol
𝑝 = 𝑧+ 𝐹𝑙/2𝑅𝑇. Using values of 𝑘⃗ and 𝑘⃖⃗ in equation (113, 114), we get

𝑣 = 𝑙 𝑘𝐷 𝑒 𝑝𝑋 (117)

𝑣⃖ = 𝑙 𝑘𝐷 𝑒 −𝑝𝑋 (118)

Now substituting the value of 𝑣 and 𝑣⃖ in equation (111), we have

𝑣𝑑 = 𝑙 𝑘𝐷 𝑒 𝑝𝑋 − 𝑙 𝑘𝐷 𝑒 −𝑝𝑋 (119)

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 411

𝑣𝑑 = 𝑙 𝑘𝐷 (𝑒 𝑝𝑋 − 𝑒 −𝑝𝑋 ) (120)

Since 𝑒 𝑝𝑋 − 𝑒 −𝑝𝑋 = 2 𝑆𝑖𝑛ℎ 𝑝𝑋, the equation (120) can also be written as

𝑣𝑑 = 2𝑙 𝑘𝐷 2 𝑆𝑖𝑛ℎ 𝑝𝑋 (121)

Now using the value of drift velocity from equation (121) in equation (110), we get the current density as

𝐽 = 𝑧+ 𝑐𝐹 (2𝑙 𝑘𝐷 2 𝑆𝑖𝑛ℎ 𝑝𝑋) (122)

Therefore, it is obvious from the above expression that the current density varies hyperbolically with the
applied electric field.

Figure 7. The variation of current density with the applied electric field.

Now since the 𝐽/𝑋 is equal to specific conductivity (σ), the above equation takes the form

𝐽 (2𝑙 𝑘𝐷 2 𝑆𝑖𝑛ℎ 𝑝𝑋) (123)


𝜎= = 𝑧+ 𝑐𝐹
𝑋 𝑋
Since molar conductivity is 𝛬𝑚 = 𝜎/𝑐 and equivalent conductivity is 𝛬𝑒𝑞 = 𝛬𝑚 /𝑧+ ; the equation (123) gives

𝛬𝑚 𝜎 (2𝑙 𝑘𝐷 2 𝑆𝑖𝑛ℎ 𝑝𝑋) (124)


𝛬𝑒𝑞 = = = 𝑧+ 𝑐𝐹
𝑧+ 𝑧+ 𝑐 𝑧+ 𝑐𝑋

(2𝑙 𝑘𝐷 2 𝑆𝑖𝑛ℎ 𝑝𝑋) (125)


𝛬𝑒𝑞 = 𝐹
𝑋
Which is the equation for equivalent conductivity.

Copyright © Mandeep Dalal


412 A Textbook of Physical Chemistry – Volume I

 Total Driving Force for Ionic Transport: Nernst-Planck Flux Equation


The rate perspective of the conduction process has already been discussed in the previous sections of
this chapter. During the whole of the discussion, we assumed that the composition of the electrolyte was
uniform throughout. However, the case will become somewhat different if we assume a concentration gradient
w.r.t. tracer ions (cations in this case). Let the concentration of tracer cations is (𝑐+ )𝑥 at a distance x on the left
of the barrier-maximum whereas the concentration (𝑐+ )𝑥+𝑙 on the right of the barrier maximum. Now, if we
assume that (𝑐+ )𝑥+𝑙 > (𝑐+ )𝑥 , we can say that

𝑑(𝑐+ )𝑥 (126)
(𝑐+ )𝑥+𝑙 = (𝑐+ )𝑥 + ×𝑙
𝑑𝑥
Owing to the decreasing concentration gradient of tracer ions from right to left, we cannot use the simple
expression for current density.

Figure 8. The general depiction of the ionic movement from right to left due to diffusion facing activation
barrier under the influence of the external electric field.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 413

Therefore, the fundamental relation between the ionic drift velocity (𝑣𝑑 ) and current density (𝐽) for cation must
be recalled first i.e.

𝐽 = 𝑧+ 𝑐𝐹𝑣𝑑 (127)

Where z is the charge number and c is the concentration of the ions. The symbol F represents the Faraday
constant. Also, the drift velocity can be assumed as the resultant velocity of the velocity of ions in the direction
of the force field (𝑣 ) and the velocity of ions in the opposite direction (𝑣⃖). Mathematically, we can say as given
below.

𝑣𝑑 = 𝑣 − 𝑣⃖ (128)

Using the above result in equation (127), we get

𝐽 = 𝑧+ 𝑐𝐹(𝑣 − 𝑣⃖) (129)

𝐽 = 𝑧+ 𝑐𝐹𝑣 − 𝑧+ 𝑐𝐹𝑣⃖

Since concentration for left and on right are (𝑐+ )𝑥 and (𝑐+ )𝑥+𝑙 ; the above equation can also be written as

𝐽 = 𝑧+ (𝑐+ )𝑥 𝐹𝑣 − 𝑧+ (𝑐+ )𝑥+𝑙 𝐹𝑣⃖ (130)

Using the value of (𝑐+ )𝑥+𝑙 from equation (126) in equation (130), we get

𝑑(𝑐+ )𝑥 (131)
𝐽 = 𝑧+ (𝑐+ )𝑥 𝐹𝑣 − 𝑧+ ((𝑐+ )𝑥 + × 𝑙) 𝐹𝑣⃖
𝑑𝑥

For simplicity, the label (𝑐+ )𝑥 with 𝑐+ i.e.

𝑑𝑐+ (132)
𝐽 = 𝑧+ 𝑐+ 𝐹𝑣 − 𝑧+ (𝑐+ + 𝑙) 𝐹𝑣⃖
𝑑𝑥
Recalling the values of 𝑣 and 𝑣⃖ i.e.

𝑣 = 𝑙 𝑘𝐷 𝑒 𝑝𝑋 (133)

𝑣⃖ = 𝑙 𝑘𝐷 𝑒 −𝑝𝑋 (134)

Where 𝑘𝐷 is the jumping frequency for diffusion and X is simply the electric field. The expression for symbol
𝑝 = 𝑧+ 𝐹𝑙/2𝑅𝑇. When the field strength is very low, 𝑝𝑋 ≪ 1; and therefore, equations (133, 134) can be
expended as given below.

𝑣 = 𝑙 𝑘𝐷 (1 + 𝑝𝑋) (135)

𝑣⃖ = 𝑙 𝑘𝐷 (1 − 𝑝𝑋) (136)

Now, after rearranging equation (132) and then using equations (135, 136), we have

Copyright © Mandeep Dalal


414 A Textbook of Physical Chemistry – Volume I

𝑑𝑐+ (137)
𝐽 = 𝑧+ 𝑐+ 𝐹𝑣 − 𝑐+ 𝑧+ 𝐹𝑣⃖ − 𝑙𝑧 𝐹𝑣⃖
𝑑𝑥 +
𝑑𝑐+ (138)
𝐽 = 𝑧+ 𝑐+ 𝐹𝑙 𝑘𝐷 (1 + 𝑝𝑋) − 𝑐+ 𝑧+ 𝐹𝑙 𝑘𝐷 (1 − 𝑝𝑋) − 𝑙𝑧 𝐹𝑙 𝑘𝐷 (1 − 𝑝𝑋)
𝑑𝑥 +
𝑑𝑐+ (139)
𝐽 = 2𝑧+ 𝑐+ 𝐹𝑙 𝑘𝐷 𝑝𝑋 − 𝑧+ 𝐹𝑙 2 𝑘𝐷 (1 − 𝑝𝑋)
𝑑𝑥
Neglecting pX in comparison to one for low-field approximation, we have

𝑑𝑐+ (140)
𝐽 = 2𝑧+ 𝑐+ 𝐹𝑙 𝑘𝐷 𝑝𝑋 − 𝑧+ 𝐹𝑙 2 𝑘𝐷
𝑑𝑥
Since 𝑝 = 𝑧+ 𝐹𝑙/2𝑅𝑇, the equation (140) can be transformed to

𝑧+ 𝐹𝑙 𝑑𝑐+ (141)
𝐽 = 2𝑧+ 𝑐+ 𝐹𝑙 𝑘𝐷 𝑋 − 𝑧+ 𝐹𝑙 2 𝑘𝐷
2𝑅𝑇 𝑑𝑥
𝑙 2 𝑘𝐷 𝑑𝑐+ (142)
𝐽 = 𝑧+2 𝑐+ 𝐹 2 𝑋 − 𝑧+ 𝐹𝑙 2 𝑘𝐷
𝑅𝑇 𝑑𝑥
Since 𝑙 2 𝑘𝐷 = 𝐷+ , the equation (142) takes the form

𝐷+ 𝑑𝑐+ (143)
𝐽 = 𝑧+2 𝑐+ 𝐹 2 𝑋 − 𝑧+ 𝐹𝐷+
𝑅𝑇 𝑑𝑥
Now recalling the correlation between current density (𝐽+ ) and flux (𝑗+ ) of positive ions i.e.

𝐽+ (144)
𝑗+ =
𝑧+ 𝐹

Now because the current density given by equation (143) is only from cations, the corresponding flux can be
obtained putting value of 𝐽+ from equation (143) in equation (144), we get

𝑧+2 𝑐+ 𝐹 2 𝐷+ 𝑋 𝑧+ 𝐹𝐷+ 𝑑𝑐+ (145)


𝑗+ = −
𝑧+ 𝐹𝑅𝑇 𝑧+ 𝐹 𝑑𝑥

𝑐+ 𝐷+ 𝑑𝑐+ (146)
𝑗+ = 𝑧+ 𝐹𝑋 − 𝐷+
𝑅𝑇 𝑑𝑥
After multiplying and dividing the second term by 𝑐+ 𝑅𝑇, we have

𝑐+ 𝐷+ 𝐷+ 𝑐+ 𝑅𝑇 𝑑𝑐+ (147)
𝑗+ = 𝑧+ 𝐹𝑋 −
𝑅𝑇 𝑅𝑇 𝑐+ 𝑑𝑥

𝑐+ 𝐷+ 𝐷+ 𝑐+ 𝑑(𝑅𝑇 ln𝑐+ ) (148)


𝑗+ = 𝑧+ 𝐹𝑋 −
𝑅𝑇 𝑅𝑇 𝑑𝑥

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 415

𝑐+ 𝐷+ 0
𝐷+ 𝑐+ 𝑑(𝜇+ + 𝑅𝑇 ln 𝑐+ ) (149)
𝑗+ = 𝑧+ 𝐹𝑋 −
𝑅𝑇 𝑅𝑇 𝑑𝑥
𝑐+ 𝐷+ 0
𝐷+ 𝑐+ 𝑑(𝜇+ + 𝑅𝑇 ln 𝑐+ ) (150)
𝑗+ = 𝑧+ 𝐹𝑋 −
𝑅𝑇 𝑅𝑇 𝑑𝑥
Since 𝜇+
0
+ 𝑅𝑇 ln 𝑐+ = 𝜇+ , the above equation becomes

𝑐+ 𝐷+ 𝐷+ 𝑐+ 𝑑𝜇+ (151)
𝑗+ = 𝑧+ 𝐹𝑋 −
𝑅𝑇 𝑅𝑇 𝑑𝑥
It is a well-known fact in electrochemical theory that the electric field is simply equal to negative of the gradient
of electrostatic potential i.e. 𝑋 = −𝑑𝜓/𝑑𝑥. Therefore, the equation (151) takes the form

𝑐+ 𝐷+ 𝑑𝜓 𝐷+ 𝑐+ 𝑑𝜇+ (152)
𝑗+ = 𝑧+ 𝐹 (− ) −
𝑅𝑇 𝑑𝑥 𝑅𝑇 𝑑𝑥
or

𝑐+ 𝐷+ 𝑑𝜓 𝑑𝜇+ (153)
𝑗+ = − (𝑧+ 𝐹 + )
𝑅𝑇 𝑑𝑥 𝑑𝑥
Since the −𝑑𝜇+ /𝑑𝑥 and −𝑧+ 𝐹 𝑑𝜓/𝑑𝑥 are the driving forces for pure diffusion and pure conduction
phenomena, respectively; the total driving force for ionic transport must be equal to the negative gradient of
chemical potential and electrostatic potential. The sum of the two potentials is called as electrostatic-chemical
potential (𝜇̅ + ), and is defined by

𝜇̅ + = 𝑧+ 𝐹𝜓 + 𝜇+ (154)

Taking negative both sides and then differentiating w.r.t. x, we have

𝑑𝜇̅ + 𝑑𝜓 𝑑𝜇+ (155)


− = − (𝑧+ 𝐹 + )
𝑑𝑥 𝑑𝑥 𝑑𝑥
Utilizing the above result in equation (153), we get

𝑐+ 𝐷+ 𝑑𝜇̅+ (156)
𝑗+ = −
𝑅𝑇 𝑑𝑥
Since the Einstein relation is 𝐷+ = (𝑢̅𝑎𝑏𝑠 )+ 𝑘𝑇, the above equation becomes

𝑐+ (𝑢̅𝑎𝑏𝑠 )+ 𝑘𝑇 𝑑𝜇̅ + (157)


𝑗+ = −
𝑅𝑇 𝑑𝑥
Moreover, the relationship between conventional (𝑢̅𝑐𝑜𝑛𝑣 )+ and absolute mobilities (𝑢̅𝑎𝑏𝑠 )+ is

(𝑢̅𝑐𝑜𝑛𝑣 )+ = (𝑢̅𝑎𝑏𝑠 )+ 𝑧+ 𝑒0 (158)

Copyright © Mandeep Dalal


416 A Textbook of Physical Chemistry – Volume I

Therefore, the use of equation (158) in equation (157) gives

𝑐+ (𝑢̅𝑐𝑜𝑛𝑣 )+ 𝑘 𝑑𝜇̅ + (159)


𝑗+ = −
𝑧+ 𝑒0 𝑅 𝑑𝑥

Since 𝑅 = 𝑁𝐴 𝑘, the above equation becomes

𝑐+ (𝑢̅𝑐𝑜𝑛𝑣 )+ 𝑑𝜇̅+ (160)


𝑗+ = −
𝑧+ 𝑒0 𝑁𝐴 𝑑𝑥

Putting 𝑒0 𝑁𝐴 = 𝐹, we have

(𝑢̅𝑐𝑜𝑛𝑣 )+ 𝑑𝜇̅ + (161)


𝑗+ = − 𝑐+
𝑧+ 𝐹 𝑑𝑥

Which is the famous Nernst-Planck flux equation that relates the total driving force for the ionic transport with
the overall flux.

 Ionic Drift and Diffusion Potential


In order to understand the link between ionic drift and diffusion potential, consider a solution of
monovalent electrolyte with concentration c. Now assume that this solution is brought in contact with pure
water and the boundary of contact is assumed to be 𝑥 = 0. Owing to the concentration gradient, both cation as
well anions will start moving into pure water immediately.

Figure 9. The general depiction electrolytic solution and concentration in contact with water in start.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 417

Now owing to different absolute ionic mobilities of cations and anions (say 𝑢̅+ > 𝑢̅− ), the Einstein
relation can be written for cations and anions as given below.

𝐷+ = 𝑢̅+ 𝑘𝑇 (162)

and

𝐷− = 𝑢̅− 𝑘𝑇 (163)

Where 𝐷+ and 𝐷− are the diffusion coefficients for cations and anions, respectively. The symbol k is simply
the Boltzmann constant and T is the temperature. The symbol 𝑢̅+ and 𝑢̅− represent the absolute ionic mobilities
for cation and anion, respectively. Since we have assumed that 𝑢̅+ > 𝑢̅−, the following must be true

𝐷+ > 𝐷− (164)

This implies that the cations will move faster in comparison to anions, will lead the anions in their diffusion
race. Now consider two unit-volume-elements at distance −𝑥1 and −𝑥2 in the water phase with −𝑥2 on more
left than −𝑥1 . Since the cations are moving faster than anions, the concentration ratio of the two (𝑐+ /𝑐−) will
be higher in volume element at −𝑥2 than in the volume element at −𝑥1 distance. In other words, the ratio
𝑐+ /𝑐− will increase as we move from the boundary to the waterside of the system.

Figure 10. The development of diffusion potential due to different ionic mobilities.

Consequently, a situation will arise in which the positive and negative charges are separated with a
negative layer on the left and positive layer on the right. All this will lead to the development of a potential
difference that will oppose the faster movement of cations and will reinforce the slower movement of anions.
This potential is generally called as the “diffusion potential” and tries to level the ionic mobilities of cations
and anions; and hence, tries to maintain the electroneutrality in different parts of the solution. It should also be
noted that diffusion potential is also called as the “liquid junction potential” in the case of concentration cells
and “membrane potential” if the two solutions are separated by an uncharged membrane.

Copyright © Mandeep Dalal


418 A Textbook of Physical Chemistry – Volume I

 The Onsager Phenomenological Equations


Since the ionic mobilities of cations are not independent but affect each other as we have studied in
the previous section, the expression for the flux must also be modified. In order to understand the concept,
recall the general form of Nernst-Planck equation for the ionic flux of ith species (𝑗𝑖 ) i.e.

𝑢̅𝑖 𝑑𝜇̅ 𝑖 (165)


𝑗𝑖 = − 𝑐
𝑧𝑖 𝐹 𝑖 𝑑𝑥

Where 𝑢̅𝑖 is the conventional ionic mobility of ith species and 𝑐𝑖 represents the corresponding concentration.
The symbol 𝑧𝑖 and 𝐹 are the charge number and Faraday constant, respectively. The symbol 𝑑𝜇̅ 𝑖 /𝑑𝑥 is the
total driving force for ionic transport. Correcting for the coupling between ionic mobilities arising from
diffusion potential, we have

𝑢̅𝑖 𝑑𝜇̅ 𝑖 (166)


𝑗𝑖 = − 𝑐 + 𝑐𝑜𝑢𝑝𝑙𝑖𝑛𝑔 𝑐𝑜𝑟𝑟𝑒𝑐𝑡𝑖𝑜𝑛
𝑧𝑖 𝐹 𝑖 𝑑𝑥

These mutually interactive flows can be treated by the methods of near-equilibrium thermodynamics in a
phenomenological or macroscopic framework. This procedure is simple and can be covered in the postulates
given below.
Statement 1: When the system is in near-equilibrium and ionic mobilities are independent of each other, the
fluxes can be treated as proportional to the driving forces (Nernst-Planck flux equation). Mathematically, we
can say about the ionic flux of 1st species (𝑗1 ) that

𝑗1 = 𝐿11 𝐹1 (167)

Where 𝐿11 = 𝑢̅1 𝑐1 /𝑧1 𝐹 is phenomenological constant and 𝐹1 = 𝑑𝜇̅1 /𝑑𝑥 is the corresponding driving force.
Statement 2: If the coupling between ionic mobilities is considered, the fluxes should be treated as
proportional to the driving forces (Nernst-Planck flux equation) plus the contribution from the coupling.
Mathematically, we can say about the ionic flux of 1st species (𝑗1 ) that

𝑗1 = 𝐿11 𝐹1 + 𝑐𝑜𝑢𝑝𝑙𝑖𝑛𝑔 𝑐𝑜𝑟𝑟𝑒𝑐𝑡𝑖𝑜𝑛 (168)

Now if there are many types of species then above equation takes the form

𝑗1 = 𝐿11 𝐹1 + 𝐹𝑙𝑢𝑥 𝑜𝑓 1 𝑑𝑢𝑒 𝑡𝑜 𝑑𝑟𝑖𝑣𝑖𝑛𝑔 𝑓𝑜𝑟𝑐𝑒 𝑜𝑓 2𝑛𝑑 𝑠𝑝𝑒𝑐𝑖𝑒𝑠 (169)


+ 𝐹𝑙𝑢𝑥 𝑜𝑓 1 𝑑𝑢𝑒 𝑡𝑜 𝑑𝑟𝑖𝑣𝑖𝑛𝑔 𝑓𝑜𝑟𝑐𝑒 𝑜𝑓 3𝑟𝑑 𝑠𝑝𝑒𝑐𝑖𝑒𝑠 + ⋯

Statement 3: The proportionality of fluxes is also valid for the contributions of the forces from other ionic
species. Hence, equation (169) can also be written as

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 419

𝑗1 = 𝐿11 𝐹1 + [𝐿12 𝐹2 + 𝐿13 𝐹3 + 𝐿14 𝐹4 … . . 𝐿1𝑛 𝐹𝑛 ] (170)

Where 𝐿12 , 𝐿13 𝐿14 ….. 𝐿1𝑛 are the phenomenological constants for the interactions on flux from the flux of
other ionic species. The symbol 𝐹1 , 𝐹2 , 𝐹3 … . . 𝐹𝑛 are the corresponding driving forces.
Statement 4: If a monovalent electrolyte like NaCl is dissolved into water, the fluxes of cation (𝑗+ ), anion (𝑗− )
and water (𝑗0 ) can be written as

𝑗+ = 𝐿++ 𝐹+ + 𝐿+− 𝐹− + 𝐿+0 𝐹0 (171)

𝑗− = 𝐿−− 𝐹− + 𝐿−+ 𝐹+ + 𝐿−0 𝐹0 (172)

𝑗0 = 𝐿00 𝐹0 + 𝐿0+ 𝐹+ + 𝐿0− 𝐹− (173)

The equations (171-173) are typically known as the Onsager phenomenological equations.
Statement 5: According to Onsager’s reciprocity relation, all symmetrical coefficients are equal.
Mathematically, we can say that

𝐿𝑖𝑗 = 𝐿𝑗𝑖 (174)

Which is an experimentally proved result.

 The Basic Equation for the Diffusion


The basic equation for the diffusion potential can be obtained by using the Onsager phenomenological
equations very easily. To do so, imagine an electrolytic solution with M 𝑧+ type cations, A𝑧− type anions and
water as the solvent with 𝑗+ , 𝑗− and 𝑗0 fluxes, respectively. Now if the solvent is assumed to be non-moving,
its flux can simply be put equal to zero i.e. 𝑗0 = 0. Such a solution will only have two types of ionic fluxes
which can be formulated as given below.

𝑗+ = 𝐿++ 𝐹+ + 𝐿+− 𝐹− (175)

𝑗− = 𝐿−− 𝐹− + 𝐿−+ 𝐹+ (176)

Where 𝐿++ and 𝐿−− are coefficients for mutually independent flow, whereas 𝐿+− and 𝐿−+ are for the mutually
coupled flow.
Now, under steady-state approximation, the magnitude of positive charge flowing through the unit
volume must be equal to the amount of the negative charge flowing through the same element but in the
opposite direction. Mathematically, we can say that

𝑧+ 𝐹𝑗+ = −(𝑧− 𝐹𝑗− ) (177)

Where F is Faraday constant. The minus sign is for the mutually opposite directions.

Copyright © Mandeep Dalal


420 A Textbook of Physical Chemistry – Volume I

Figure 11. The flow of charge under steady-state approximation.

Rearranging equation (177) and then using values of 𝑗+ and 𝑗− form equation (175-176), we get

𝑧+ 𝐹(𝐿++ 𝐹+ + 𝐿+− 𝐹− ) + 𝑧− 𝐹(𝐿−− 𝐹− + 𝐿−+ 𝐹+ ) = 0 (178)

Putting 𝑧+ 𝐹 = 𝑞+ and 𝑧− 𝐹 = 𝑞− for simplicity, we have

𝑞+ (𝐿++ 𝐹+ + 𝐿+− 𝐹− ) + 𝑞− (𝐿−− 𝐹− + 𝐿−+ 𝐹+ ) = 0 (179)

𝑞+ 𝐿++ 𝐹+ + 𝑞+ 𝐿+− 𝐹− + 𝑞− 𝐿−− 𝐹− + 𝑞− 𝐿−+ 𝐹+ = 0 (180)

𝐹+ (𝑞+ 𝐿++ + 𝑞− 𝐿−+ ) + 𝐹− (𝑞+ 𝐿+− + 𝑞− 𝐿−− ) = 0 (181)

Using short symbols 𝑝+ = 𝑞+ 𝐿++ + 𝑞− 𝐿−+ and 𝑝− = 𝑞+ 𝐿+− + 𝑞− 𝐿−−, we have

𝑝+ 𝐹+ + 𝑝− 𝐹− = 0 (182)

Now recalling the expressions for driving forces i.e.

𝑑𝜓 𝑑𝜇+ (183)
𝐹+ = 𝑞+ +
𝑑𝑥 𝑑𝑥
𝑑𝜓 𝑑𝜇− (184)
𝐹− = 𝑞− +
𝑑𝑥 𝑑𝑥
Where 𝑑𝜇+ /𝑑𝑥 and 𝑞+ 𝑑𝜓/𝑑𝑥 are the driving forces for pure diffusion and pure conduction phenomena for
the cations; whereas, 𝑑𝜇− /𝑑𝑥 and 𝑞− 𝑑𝜓/𝑑𝑥 are the driving forces for pure diffusion and pure conduction
phenomena for the anions. Using equation (183, 184) in equation (182), we get

𝑑𝜓 𝑑𝜇+ 𝑑𝜓 𝑑𝜇− (185)


𝑝+ 𝑞+ + 𝑝+ + 𝑝− 𝑞− + 𝑝− =0
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥
or

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 421

𝑑𝜓 𝑑𝜓 𝑑𝜇+ 𝑑𝜇− (186)


−𝑝+ 𝑞+ − 𝑝− 𝑞− = 𝑝+ + 𝑝−
𝑑𝑥 𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑𝜓 𝑑𝜇+ 𝑑𝜇− (187)
− (𝑝+ 𝑞+ + 𝑝− 𝑞− ) = 𝑝+ + 𝑝−
𝑑𝑥 𝑑𝑥 𝑑𝑥
𝑑𝜓 𝑝+ 𝑑𝜇+ 𝑝− 𝑑𝜇− (188)
− = +
𝑑𝑥 𝑝+ 𝑞+ + 𝑝− 𝑞− 𝑑𝑥 𝑝+ 𝑞+ + 𝑝− 𝑞− 𝑑𝑥

Since we know that


𝑝+ 𝑡+ (189)
=
𝑝+ 𝑞+ + 𝑝− 𝑞− 𝑧+ 𝐹

or
𝑝− 𝑡− (190)
=
𝑝+ 𝑞+ + 𝑝− 𝑞− 𝑧− 𝐹

Where 𝑡+ and 𝑡− are transport numbers of cation and anion, respectively. Using equation (189, 190) in equation
(188), we get

𝑑𝜓 𝑡+ 𝑑𝜇+ 𝑡− 𝑑𝜇− (191)


− = +
𝑑𝑥 𝑧+ 𝐹 𝑑𝑥 𝑧− 𝐹 𝑑𝑥

Or in general, we can conclude as

𝑑𝜓 𝑡𝑖 𝑑𝜇𝑖 (192)
− =∑
𝑑𝑥 𝑧𝑖 𝐹 𝑑𝑥

The minus sign of the electric field is because it is in the opposite direction to the chemical potential gradients
of ionic diffusion. Furthermore, equation (192) can also be written as

𝑑𝜓 1 𝑡𝑖 𝑑𝜇𝑖 (193)
− = ∑
𝑑𝑥 𝐹 𝑧𝑖 𝑑𝑥

or

1 𝑡𝑖 (193)
−𝑑𝜓 = ∑ 𝑑𝜇𝑖
𝐹 𝑧𝑖

In terms of activity (ai), the above equation can be written as

𝑅𝑇 𝑡𝑖 (194)
−𝑑𝜓 = ∑ 𝑑 ln 𝑎𝑖
𝐹 𝑧𝑖

Which is the basic equation of diffusion potential.

Copyright © Mandeep Dalal


422 A Textbook of Physical Chemistry – Volume I

 Planck-Henderson Equation for the Diffusion Potential


The basic equation for diffusion potential is applicable only if the potential difference (𝑑𝜓) is
considered over a very small distance (dx). However, the problem of obtaining an overall potential difference
(𝛥𝜓 = 𝜓 0 − 𝜓 𝑙 ) that develops from 𝑥 = 0 to 𝑥 = 𝑙 was still there.

Figure 12. The overall potential difference across the complete interphase domain between electrolytes
with concentration 𝑐0 and 𝑐𝑙 .

This was overcome by Planck-Henderson equation which can be obtained by recalling the basic equation for
diffusion first i.e.

1 𝑡𝑖 (195)
−𝑑𝜓 = ∑ 𝑑𝜇𝑖
𝐹 𝑧𝑖

Where 𝑡𝑖 and 𝑧𝑖 are the charge number of ith species whereas F represents the Faraday constant. Integrating
equation (195), we get
𝑥=𝑙 (196)
0
1 𝑙
𝑡𝑖 𝑑𝜇𝑖
−𝛥𝜓 = 𝜓 − 𝜓 = ∑ ∫ 𝑑𝑥
𝐹 𝑧𝑖 𝑑𝑥
𝑖 𝑥=0

or
𝑥=𝑙 (197)
𝑅𝑇 𝑡𝑖 𝑑 ln 𝑎𝑖
−𝛥𝜓 = ∑ ∫ 𝑑𝑥
𝐹 𝑧𝑖 𝑑𝑥
𝑖 𝑥=0

or

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 423

𝑥=𝑙 (198)
𝑅𝑇 𝑡𝑖 1 𝑑 (𝑓𝑖 𝑐𝑖 )
−𝛥𝜓 = ∑ ∫ 𝑑𝑥
𝐹 𝑧𝑖 𝑓𝑖 𝑐𝑖 𝑑𝑥
𝑖 𝑥=0

At this stage, the things we need to evaluate the equation (198) are the concentration of all species in
the interphase region, the variation of activity coefficient and transport number with concentration. For
simplicity, the activity coefficients can be taken as unity and transport numbers as constant. In addition to these
assumptions, the variation of concentration of ith species with distance is considered as linear i.e.

𝑐𝑖 (𝑥) = 𝑘𝑖 𝑥 + 𝑐𝑖 (0) (199)

For constant 𝑘𝑖 , differentiate above equation i.e.

𝑑𝑐𝑖 𝑐𝑖 (𝑙) − 𝑐𝑖 (0) (200)


= 𝑘𝑖 =
𝑑𝑥 𝑙
Now using equation (199, 200) in equation (198), we get
𝑥=𝑙 (201)
𝑅𝑇 𝑡𝑖 𝑘1
−𝛥𝜓 = ∑ ∫ 𝑑𝑥
𝐹 𝑧𝑖 𝑐𝑖 (0) + 𝑘1 𝑥
𝑖 𝑥=0

or
𝑥=𝑙 (202)
𝑅𝑇 𝑡𝑖 𝑑[𝑘1 𝑥 + 𝑐𝑖 (0)]
−𝛥𝜓 = ∑ ∫
𝐹 𝑧𝑖 𝑘1 𝑥 + 𝑐𝑖 (0)
𝑖 𝑥=0

or

𝑅𝑇 𝑡𝑖 𝑥=𝑙 (203)
−𝛥𝜓 = ∑ {ln [𝑘1 𝑥 + 𝑐𝑖 (0)]}
𝐹 𝑧𝑖
𝑖 𝑥=0

or

𝑅𝑇 𝑡𝑖 𝑐𝑖 (𝑙) (204)
−𝛥𝜓 = ∑ ln
𝐹 𝑧𝑖 𝑐𝑖 (0)
𝑖

Which is the general form of the Planck-Henderson equation for diffusion potential. Using 𝑐+ = 𝑐+ = 𝑐 and
𝑧+ = 𝑧− = 𝑧 for 𝑧: 𝑧 electrolyte, we have

𝑅𝑇 𝑐𝑖 (𝑙) (205)
−𝛥𝜓 = (𝑡+ − 𝑡− ) ln
𝑧𝐹 𝑐𝑖 (0)

Furthermore, putting 𝑡+ + 𝑡− = 1, the equation (205) takes the form

Copyright © Mandeep Dalal


424 A Textbook of Physical Chemistry – Volume I

𝑅𝑇 𝑐𝑖 (𝑙) (206)
−𝛥𝜓 = (2𝑡+ − 1) ln
𝑧𝐹 𝑐𝑖 (0)

Which is the another form of Planck-Henderson equation for simple systems.

Copyright © Mandeep Dalal


CHAPTER 8 Electrochemistry – II: Ion Transport in Solutions 425

 Problems
Q 1. Discuss the ionic movement under the influence of an electric field.
Q 2. What are absolute and conventional ionic mobilities? How they are related?
Q 3. Derive and discuss the relationship between ionic drift velocity and current density.
Q 4. Define the diffusion coefficient. How it is related to the absolute mobility.
Q 5. Derive Stokes-Einstein relation.
Q 6. What is the Nernst-Einstein relation? What is its significance?
Q 7. State and explain Walden’s rule.
Q 8. Explain the rate process approach to ionic migration in detail.
Q 9. Derive the relationship between ionic drift and diffusion potential.
Q 10. What are Onsager phenomenological equations? How they can be used to derive the basic equation of
diffusion potential?
Q 11. Write down the Planck-Henderson equation for monovalent electrolytes.

Copyright © Mandeep Dalal


426 A Textbook of Physical Chemistry – Volume I

 Bibliography
[1] E. Steiner, The Chemistry Maths Book, Oxford University Press, Oxford, UK, 2008.
[2] M. R. Wright, An Introduction to Aqueous Electrolyte Solutions, John Wiley & Sons Ltd, Sussex, UK,
2007.
[3] P. Debye, E. Hückel, The Theory of Electrolytes. I. Lowering of Freezing Point and Related Phenomena,
Physikalische Zeitschrift., 24 (1923) 185-206.
[4] V. S. Bagotsky, Fundamentals of Electrochemistry, John Wiley & Sons, New Jersey, USA, 2006.
[5] J. Bockris, A. Reddy, Modern Electrochemistry – Volume 1: Ionics, Kluwer Academic Publishers, New
York, USA 2002.
[6] B. R. Puri, L. R. Sharma, M. S. Pathania, Principles of Physical Chemistry, Vishal Publications, Jalandhar,
India, 2008.
[7] P. Atkins, J. Paula, Physical Chemistry, Oxford University Press, Oxford, UK, 2010.
[8] A. Fick, On liquid diffusion, Annalen der Physik und Chemie., 94 (1855) 59.

Copyright © Mandeep Dalal

You might also like