You are on page 1of 201

Title: Wigner-Eckart theorem

Page-1

In this lecture, we will see the relation between the rotation operator and angular
momentum

We will learn the calculation of matrix elements of scalar and vector quantities.

We will see that these matrix elements knowledge can be extrapolated to form the general
Wigner-Eckart theorem.

We will also understand the use of Wigner-Eckart theorem to evaluate the transition
selection rules.
Page-2
Angular momentum operators have a close relationship to rotations. Let us consider a
coordinate system a ≡ ( X , Y , Z ) as shown in figure-16.1. This system rotates at an
infinitesimal δα around the Z axis to form the new axis system a′ ≡ ( X ,′ Y ,′Z ) ′.

Y
Y′

δα X

Z, Z′′′ X′
Figure-16.1

The rotation operator R is related to the angular momentum as R = 1 − i δ α J ⋅ u where u is


the unit vector along the rotation axis Z.

Any scalar quantity is invariant under the rotation of coordinate system. Thus scalar
operator commutes with the rotation oprator.
J ′m ′ RA Jm = J m′ AR ′ Jm

Thus

J '− J = 0 ⇒ ∆ J = 0
m '− m = 0 ⇒ ∆ m = 0
Page-3

Y
Y′

A′
δA

δα X

Z, Z′′′ X′
Figure-16.2

Now any vector A in the old coordinate system a ≡ ( X , Y , Z ) as shown in figure-16.2 is


related to the same vector A′ in the new coordinate system a′ ≡ ( X ,′ Y ,′Z ) ′as
(
A′ = A + δ A = A + δα u × A )
In the following, we will establish a relationship between angular momentum operator
and the vector.
We are removing the arrow sign for the sake of simplicity, but remember A is a vector

a' A a' = a A a +δ α u× a A a

a ' R −1 A R a ' = a A a + δ α u × a A a

(
R −1 A R = A + δ α u × A )
Page-4

Now substituting the value of the rotation operator,

(1 + i δ α J ⋅ u ) A (1 − i δ α J ⋅ u ) = A + δ α ( u × A)
⇒ A − i δ α A ( J ⋅ u ) + i δ α ( J ⋅ u ) A = A + δ α ( u × A)
⇒ − i δ α  A, ( J ⋅ u )  = −i 2δ α ( u × A )
⇒  A, ( J ⋅ u )  = i ( u × A )

Now using this relation we can establish the commutation relations.

 i j k
 
 ux uy uz 
 
 Ax Ay Az 

(
 Az , ( J ⋅ u )  = i u x Ay − u y Ax )
 Ax , ( J ⋅ u )  = i ( u y Az − Ay u z )
 Ay , ( J ⋅ u )  = i ( Axu z − u x Az )
 

[ Az , J z ] = 0 [ Ax , J z ] = −iAy  Ay , J z  = i Ax
 
[ Az , J x ] = i Ay [ Ax , J x ] = 0  Ay , J x  = −iAz
 
 Az , J y  = −i Ax  Ax , J y  = i Az  Ay , J y  = 0
     
Page-5

Az And J z commuts.
Az J z − J z Az = 0

When J = J '

Jm ' | Az J z | Jm − Jm ' | J z Az | Jm = 0

∑ Jm ' | Az | J '' m '' J '' m '' | J z | Jm − ∑ Jm ' | J z | J '' m '' J '' m '' | Az | Jm = 0
J '' m '' J '' m ''

Jm ' | Az | Jm ( m − m ') = 0

When m = m ', Jm ' | Az | Jm ≠ 0

Now the commutation with A+

1
[ A+, J z ] = [ Ax + iAy , J z ]
2
1
= {[ Ax , J z ] + i[ Ay , J z ]}
2
1
= [−iAy + i ( iAx )]
2
1
=− [ Ax + iAy ] = − A+
2
Page-6

So,

A+ J z − J z A+ = − A+
A+ J z − J z A+ + A+ = 0
when J = J ''

Jm ' | A+ J z | Jm − Jm ' | J z A+ | Jm + Jm ' | A+ | Jm = 0


m Jm ' | A+ | Jm − m ' Jm ' | Az | Jm + Jm ' | A+ | Jm = 0
Jm ' | A+ | Jm ( m − m '+ 1) = 0
so m ' = m + 1

Now, the relation between A and J’s matrix elements

[ A+ , J + ] = [ Ax + iAy , J x + iJ y ] = 0

when J = J '

Jm ' | A+ J + | Jm − Jm ' | J + A+ | Jm
Jm + 2 A+ Jm + 1 Jm + 1 J + Jm = Jm + 2 J + Jm + 1 Jm + 1 A+ Jm

Jm + z | A+ | Jm + 1 Jm + z | J + | Jm + 1
= =a
Jm + 1| A+ | Jm Jm + 1| J + | Jm

This is true only if Jm + 1| A+ | Jm = a Jm + 1| J + | Jm


Where a is a constant.
Page-7

Now,
[ Az , J + ] = [ Az , J x + iJ y ]
= iAy + i ( −iAx )
= iAx + iAy = A+

Az J + − J + Az − A+ = 0
Jm ' | Az J + | Jm − Jm ' | J + Az | Jm − Jm ' | A+ | Jm = 0
Jm + 1| Az | Jm + 1 Jm + 1| J + | Jm − Jm + 1| J + | Jm Jm | Az | Jm − a Jm + 1| J + | Jm = 0

[ Jm + 1| Az | Jm + 1 − Jm | Az | Jm − a] Jm + 1| J + | Jm = 0
Jm + 1| Az | Jm + 1 − Jm | Az | Jm = a

This relation is true if we take


J m | Az | J m = a m + b
J m + 1| Az | J m + 1 = a ( m + 1) + b

J m + 1| Az | J m + 1 − J m | Az | J m = a m + a + b − a m − b = a

What is the value of b?


+J +J
TraceAz = ∑ J m | Az | J m = ∑ a m + b = ( 2 J + 1) b
m =− J m =− J
1 1
(
Az =  Ax , J y  = Ax J y − J y Ax
i i
)
1
(
Trace Az = Tr Ax J y − J y Ax
i
) [Tr AB = Tr BA]

=0
So, ( 2 J + 1) b = 0 or , b = 0

J m | Az | J m = a m = a J m | J z | J m
J m + 1| A+ | J m = a J m + 1| J + | J m
J m − 1| A− | J m = a J m − 1| J − | J m

Page-8

In general form:
J m ' | A1q | J m = a J m ' | J q | J m
m' = m when q = 0 

m ' = m +1 when q = +1
m ' = m −1 when q = −1

J ' m ' | A 'q | J m = a J ' m ' | U 'q | J m

J ' m ' | A 'q | J m


⇒ =a
J ' m ' | U 'q | J m

From the above matrix elements relations we write the general form of Wigner Eckart
theorem.

Wigner Eckart theorem:


J ' m ' | Tqk | J m = J ' k m ' q | j ' k j m J ' | T k | J
Coupling Reduced
coefficient / matrix
Symmetry element /
part Physical
interaction

then J ' m ' | U qk | J m = J ' k m ' q | J ' k j m J ' | U k | J

Application:
J ' m ' | A1q | J m = J ' | A | J J '1 m ' q | J '1 J m

J ' m ' | Jq | J m = J ' | J | J J '1 m ' q | J '1 J m

 J ' | A | J 
J ' m ' | A1q | J m =   J ' m ' | Jq | J m
 J ' | J | J 
= a J ' m ' | Jq | J m

Page-9
General Statement:

1
J ' m ' | Tqk | J m = J 'k m'q | J 'k J m J | T k | J
2J + 1
J −m 1  J' k J  k
= ( −1)   J || T || J
2 j + 1  −m ' q m 

Here Tqk is a tensor and T k is the reduced tensor.


If we know the three j properties, we can evaluate selection rule for which matrix element
will be zero.

Properties of 3j:
m' = m + q
 j1 + j2 + j = n
 j + j − j≥0

∆ ( j1 j2 j ) :  1 2
 j1 − j2 + j ≥ 0
 − j1 + j2 + j ≥ 0

For scalar
J ' m ' | T00 | J m = 0 unless J '+ 0 + J = n
J'= J J '+ 0 + J ≥ 0 
  J '− J = 0 ⇒ ∆ J = 0 
m' = m J '+ 0 + J ≥ 0   
  m '− m = 0 ⇒ ∆ m = 0 
J '+ J ≥ 0 J '+ 0 + J ≥ 0 

For vector
J ' m ' | Tq1 | J m = 0 unless ∆ J = 0, ±1
J '+ J ≥ 1
m′ = m − q (q = 1, 0, −1)
∆ m = 0, ±1

Page-10

Now we will evaluate the selection rules for the transitions

We know that in the electric dipole approximation the dipole matrix elements
ψ ′ er ψ should be evaluated. If the dipole matrix element is nonzero the transition is
allowed, otherwise, it is forbidden.

Let us take the case of central field orbitals. The wavefunctions are characterized by the
quantum numbers n, l , ml , s, ms
so
s′ms ′ l m′ l er
′ lml sms = s ms′ sm ′ s l ml ′er ′lml
The dipole being the space operator, does not operator on spin functions. So ∆s = 0 .
Now er is a vector. So, we evaluate the vector matrix elements
l ' m ' | Tq1 | l m = 0 unless ∆ l = 0, ±1
l '+ l ≥ 1
∆ m = 0, ±1

Although it shows ∆ l = 0, ±1 , but l ′ r l = 0 if ∆ l = 0 due to parity. The er connects


only opposite parity states. Thus the selection rules are

∆ l = ±1
l '+ l ≥ 1
∆ m = 0, ±1

Now for the transitions between the terms. The wavefunctions are characterized by
L, M L , S , M S

S ′M s ′ L M
′ L ′er LM L SM S = S M′s sm
′ s L M L′ er ′ LM L

Using the same argument


L′M ′| Tq1 | LM L = 0 unless ∆ L = 0, ±1, ∆ S = 0
L '+ L ≥ 1
∆ M L = 0, ±1

Page-11

How to evaluate a:
J ⋅ A = J x Ax + J y Ay + J z Az
1
= [ J + A− + J − A+ ] + J z Az
2
1 1
J m| J ⋅ A| J m = ∑ 2
J m | J + | J m " J m " | A− | J m +
2 ∑ J m | J − | J m " J m " | A+ | J m
m" m"
+ ∑ J m | J z | J m " J m " | Az | J m
m"
1 1
= a
2 m"∑J m | J+ | J m" J m" | J− | J m + a
2 m" ∑
J m | J− | J m " J m " | J+ | J m

+ ∑ J m | J z | J m " J m " | Az | J m
m"
1
= a J m| ( J+ J− + J− J+ ) + J z J z | J m
2
= a J m | J x2 + J y2 + J z2 | J m

= a J m | J 2 | J m = a J ( J + 1)

J m | A⋅ J | J m = ∑ J m | A| J m' J m'| J | J m
m'

= a J ( J + 1)
J m | A⋅ J | J m
⇒a=
J ( J + 1)
J m | A⋅ J | J m
So, J m ' | A | J m = J m'| J | J m
J ( J + 1)
This is a special form of W-E theorem known as Lande’ formula. We will use this
relation later.

Page-12

In this lecture, we understood the calculation of matrix elements of scalar and vector
quantities.
We use these matrix elements to extrapolate the general form of Wigner-Eckart theorem.

This is used to evaluate the transition selection rules.

We will also use the special form of W-E theorem known as Lande’ formula later.
Title :Fine structure : Spin-orbit coupling

Page-0

In this lecture, we will concentrate on the fine structure of the one electron atoms.

The cause of this fine structure is the interaction between the orbital angular momentum
and the spin angular momentum.

We will review the origin for this interaction and the effect of the coupling between
orbital and spin on the spectral lines.

We will discuss here mainly hydrogen atom and also the other alkali atoms.
Page-1

It is observed in the sodium spectrum (shown in the figure below) that D-line (yellow
emission) is split into two lines
D1 = 5895.93 Å or 589.59 nm, D2 = 5889.9 Å or 588.99 nm

The reason of observing this doublet is that the energy levels split into two for the
terms except s-level (l = 0).

Many lines of the other alkali atoms are also doublets.

With high resolution spectrometer shows that even hydrogen atom lines are also
having this doublet nature.

The Coulomb interactions between the nucleus and electron in the outermost orbit for
alkali atoms can not explain this observation

Not only alkali atoms, but also in the other multielectron atoms the transitions between
the terms split into more number of transitions. This is known as Fine structure.

Emission from Sodium lamp


4000
5895.93 Å
3500 5889.96 Å
3000

2500

2000 Series1

1500

1000

500

0
580 585 590 595

Wavelength (nm) →
Page-2
In the previous lectures, we have discussed that the Hamiltonian for describing one
electron atoms (Alkali atoms) is

2
− ∇ 2ψ ( r ,θ , φ ) + V ( r )ψ ( r ,θ , φ ) = Eψ ( r , θ , φ )

where V ( r ) is the Coulomb interaction between the nucleus and electron.
We have also discussed the Hamiltonian needed to explain multielectron atoms

H = H * + H1
2 N N
where H * = −
2me
∑ ∇i2 + ∑U ( ri )
i =1 i =1

N
Z e2 N
e2
where ∑U ( ri ) = −∑
i i =1 ri
+ ∑
i < j rij

N
e2 N
e2
and H1 = ∑ − ∑ Non-Spherical part only
i< j rij i < j rij

However, these descriptions are not enough to explain the observed splitting. We need to
include the spin-orbit coupling into the Hamiltonian.

In the following we will understand the origin of this spin orbit interaction.
Page-3

Orbital magnetic dipole moment:


Let us consider that an electron is moving with velocity V in a circular Bohr orbit of
radius r that produces a current. This current loop will produce a magnetic field with the
magnetic moment,

eω 2 1 µl
µl = iA = − π r = − eω r 2
2π 2
v
eω r
Where i = − , ω is the angular velocity, A is the area.
2π -e

Magnitude of orbital angular momentum l

l = meVr = mω r 2
Where, me is the mass of the electron. Substituting we get, the orbital magnetic moment
e
µl = − l
2m
Now, the Bohr magneton is defined as the magnetic moment of the first Bohr orbital. So,
e
the Bohr magneton µ B = = 9.27x10-24 J/T
2m

The Orbital magnetic moment in terms of Bohr magneton is thus, in the vector form
gl µ B µB
µl = − l and putting g l = 1, µl = − l
z
Substituting the value of the angular momentum ml l
µB
µl = l (l + 1) = µ B l (l + 1)
l = l (l + 1)
Along the Z-direction, the component of lZ = ml , so

µB µB
µl = −
z
lz = − ml = − µ B ml
Page-4
When this dipole moment is placed in an external magnetic field along the Z-direction, it
experiences a torque which can be expressed as τ = µ × B
The potential energy
∆E = − µˆ B ⋅ Bˆ

For a static magnetic moment, this torque tends to line up the magnetic moment with the
magnetic field B, so that it reaches its lowest energy configuration.

Here, the magnetic moment arises from the motion of an electron in orbit around a
nucleus and the magnetic moment is proportional to the angular momentum of the
electron.

The torque exerted then produces a change in angular momentum. This change is
perpendicular to that angular momentum, causing the magnetic moment to precess
around the direction of the magnetic field rather than settle down in the direction of the
magnetic field.

This precession is known as Larmor precession.

∆φ
Z BZ l Sinθ

l
θ

e
µl = − l
2me

Figure-17.1
Page-5

When a torque is exerted perpendicular to the angular momentum l, it produces a change


in angular momentum ∆l which is perpendicular to l. Referring to figure-17.1, the torque
is given by,

∆l l sin θ∆φ
τ= = = lω sin θ ………………………….(17.1)
∆t ∆t

And also

µB
τ = µ × B = µl B sin θ = lB sin θ ……………………(17.2)

µB
lB sin θ = lω sin θ
So equating equations 17.1 and 17.2 we get
µB
=> ω = B

This is known as Larmor frequency.

Spin magnetic moment :

Similar to orbital dipole moment, electron also produces the magnetic moment due to the
spin angular momentum.
The spin dipole moment, in terms of spin Lande g-factor g s . Z
s
g s µB
µs = − s

µ s = − g s µ B ms
z

Where, µ SZ the component in the Z-direction µS


Page-6

As we discussed above, that the fine-structure in atomic spectra cannot be explained by


Coulomb interaction between nucleus and electron.

We have to consider magnetic interaction between orbital magnetic moment and the
intrinsic spin magnetic moment. This is known as Spin-Orbit interaction.
Let us understand how this interaction takes place.

If we consider the reference frame of electron then nucleus moves about electron. This
current j = − ZeV produces magnetic field at the electron.

According to Ampere’s Law, the magnetic field at electron due to nucleus is


µ j × r − Zeµ0 V × r
B= 0 3 =
4π r 4π r3

Since l = r × mV = −mV × r
v
+ Ze r
− Zeµ0 mV × r Zeµ0 l −e
B= =
4π m r3 4π m r 3

If we take the average field, then

Zeµ0 1
B= l ……………..(17.3)
4π m r 3
Page-7

Now, the orientation potential energy of magnetic dipole moment is ∆ESpin −Orbit = − µ s ⋅ B
gs µB
We know that µs = − s
g s µB
∆ESpin −Orbit = s ⋅B

Transforming back to reference frame with nucleus, must include the factor of 2 due to
Thomas precession (Reference : Eisberg & Resnick):

1 gs µB
∆Eso = s ⋅B
2
This is the spin-orbit interaction energy. Substituting the value of B from equation 17.3,
we get

g s µ B Zeµ0 1
∆Eso = s ⋅ l = ASO l ⋅ s …………………………..(17.4)
2 4π m r 3
Page-8

Let us first discuss about the alkali atoms. The Hamiltonian needed for calculating the
energy levels is

2
H =− ∇ 2 + V ( r ) + H Spin−Orbit

= H 0 + H Spin −Orbit

We already know the energies calculated for H0 . We can treat H Spin−Orbit as perturbation
on the energies of H0.

But first let us see the effect of this interaction.


This interaction couples the spin and orbital
angular momentum to form the total angular j =l +s
momentum J. J
The figure 17.2 represents the vector diagram
and accordingly we write
j = l + s (for one electron atoms).

Because of this coupling, the one electron


wavefunctions l ml s ms which are the l
eigenfunctions of H0 are no more the s
eigenfunctions of the total Hamiltonian H.
The eigenfunctions of the H will be J mJ
which are the coupled wavefunctions and can be
derived from the uncoupled wavefunctions
l ml s ms , as described in previous lectures.

Figure-17.2
So the perturbation energy
∆ESO = J mJ ASO l .s J mJ
Now, j = l + s
j 2 = l 2 + s 2 + 2l .s
1
l .s =  j 2 − l 2 − s 2 
2

And thus, the energy correction due to the spin-orbit interaction is


A
∆ESO = J mJ ASO l .s J mJ = SO [ j ( j + 1) − l (l + 1) − s ( s + 1)] ………………….(17.5)
2
Page-9

Let us first look at the Sodium energy levels.

The terms arising from the ground state configuration is 2 S .


So l = 0, s = 1 . According to coupling of angular momenta j = 1
2 2
2 s +1
The notation we will use is LJ
According to this the ground state energy level is 2 S 1 . Using the relation given in
2
equation 17.5, we get

( )2
A
∆ESO 2 S 1 = 1  1 ( 1 + 1) − 0(0 + 1) − 1 ( 1 + 1) 
2  2 2 2 2 
=0

Now for the first excited state, terms is 2 P

So l = 1, s = 1 . According to coupling of angular momenta j = 1 , 3


2 2 2
Now this terms will split into two energy levels 2 P1 and 2 P3
2 2

∆ESO ( P ) = A2  12 ( 12 + 1) −1(1 + 1) − 12 ( 12 + 1)


2
1
2
2

= − A2

And

∆ESO ( P ) = A2  3 2 ( 3 2 + 1) −1(1 + 1) − 12 ( 12 + 1)


2
3
2
2

A2
=
2
Page-10

So the construction of the energy levels is


2
P3
2
2
P A2
2
A2
2
P1
2
E0
5889.96 Å 5895.93 Å
2 (1) (2) 2 S
S 1
2

Now let us calculate the spin orbit constant for 2P level.

−1 A2 108
For (1) the transition energy : ν 1 (cm ) = E0 + = = 16978.04 cm −1
2 5889.96

108
For (2) the transition energy : ν 2 (cm −1 ) = E0 − A2 = = 16960.85 cm −1
5895.93

Solving these two we get A2 = 11.46 cm −1 and E0 = 16972.31 cm −1


Page-11

Similar to sodium atom, Hydrogen atom also shows doublet.

Spectral lines of H found to be composed of closely spaced doublets. Splitting is due


to interactions between electron spin s and the orbital angular momentum l

Hα line is single line according to the Bohr or Schrödinger theory. occurs at 656.47
nm for Hydrogen and 656.29 nm for Deuterium (isotope shift, λ∆~0.2 nm).

Spin-orbit coupling produces fine-structure splitting of ~0.016 nm corresponds to an


internal magnetic field on the electron of about 0.4 Tesla.

Orbital and spin angular momenta couple together via the spin-orbit interaction

Internal magnetic field produces torque which results in precession of l and s about
their sum, the total angular momentum:

This kind of coupling is called L-S coupling or Russell-Saunders coupling

2
P3
2
2
P

2
P1
10.2 eV 2
121.6 nm splitting
4.5×10-5 eV
2
S
2
S1
2
Page-12

Relativistic kinetic energy correction :


According to special relativity, the kinetic energy of an electron of mass m and velocity v
is:
p2 p4
T≈ − 3 2 where p is the momentum
2m 8m c

The first term is the standard non-relativistic expression for kinetic energy. The second
term is the lowest-order relativistic correction to this energy.

Using perturbation theory, it can be show that

Z 2α 4  1 3 
∆Erel = − 3
mc 2  − 
n  2l + 1 8n 

This energy correction does not split the energy levels, it only produces an energy shift
comparable to spin-orbit effect.

So the total energy correction for the fine structure

∆EFS = ∆Eso + ∆Erel

As En = -Z2E0/n2, where E0 = 1/2α2mc2, we can write

Z 2 E0  Z 2α 2  1 3 
2 
EH −atom = − 1+  − 
n  n  j + 1/ 2 4n  
Page-13

So we found out that


Energy correction only depends on j, which is of the order of α2 ~ 10-4 times smaller
that the principle energy splitting.

All levels are shifted down from the Bohr energies.

For every n>1 and l, there are two states corresponding to j = l ± 1/2.

States with same n and j but different l, have the same energies i.e., they are
degenerate.

L=l = 0 L=l = 1 L=l = 2


n=3 2
∆EFS 2
P3 D5
3 2S 1 2 2 2
2
2
P1 D3
2 2

n=2 2
2
∆EFS P3
2 S1 2
2
2
P1
2

n =1
∆EFS
1 2S 1
2
Page-14

Here, we learnt that due to the orbital motion of the electron, the charge of the nucleus
creates a magnetic field at the electron and this interacts with the spin of the electron.

This interaction essentially couples the orbital and the spin angular momenta and the
effect is the splitting of energy levels.

This transition pattern is known as fine structure.

The coupled angular momentum J is the good quantum number when spin-orbit coupling
is introduced for energy level calculation.

Classical explanation of spin-orbit interaction is not enough for the level having l = 0.
The quantum explanation reveals that for describing the l = 0 level, the Fermi contact
term is needed which does not have any classical analogue.

Relativistic correction is needed to predict the accurate energies.


Title : Fine Structure : multi-electron atoms
Page-0

In this lecture we will concentrate on the fine structure of the multielectron atoms.

As discussed in the previous lecture that the fine structure arises due to spin orbit coupling.

Here we will discuss about the spin orbit coupling in case of more than one electron and their
effect on terms transitions.
Page-1

The Hamiltonian for the multielectron atoms

H = H * + H1 + H Spin−Orbit
2 N N J
where H = −*

2me
∑ ∇ + ∑U ( r )
i =1
2
i
i =1
i

Z e2 N N
e2
where ∑ U ( ri ) = −∑ + ∑
i i =1 ri i < j rij

N
e2 N
e2
and H1 = ∑ − ∑r Non-Spherical part only
i< j rij i< j ij

H Spin −Orbit = ASO ∑ li .∑ si L

The Spin-orbit interaction couples the total spin and total S


orbital angular momentum (L) to form the total angular
momentum J.

The figure-18.1 represents the vector diagram and l1 s2


accordingly we write l2

s1
J = L + S (for one electron atoms).

Figure 18.1
Because of this coupling, the one electron wavefunctions
e2
L ml S ms which are the eigenfunctions of H CentralField +∑
i , j rij e2
are no more the eigenfunctions of the total Hamiltonian H.
H = H CF + ∑ + AL.S
i, j rij
l1ml 1 s1ms 1 LmL JmJ
l1ml 2 s2 ms 2 Sms
Page-2

The eigenfunctions of the H will be J mJ which are the coupled wavefunctions and can be
derived from the uncoupled wavefunctions L ml S ms , as described in previous lectures.

According to the coupling of angular momentum the value of J = J = L + S ,......... L − S and the
2 S +1
notation is LJ where L = 0,1, 2,3,.....are S , P, D, F respectively.

So the perturbation energy

∆ESO = J mJ ASO L.S J mJ

Now, J = L + S

J 2 = L2 + S 2 + 2 L.S
1
L.S =  J 2 − L2 − S 2 
2

And thus, the energy correction due to the spin-orbit interaction is


A
∆ESO = J mJ ASO L.S J mJ = SO [ J ( J + 1) − L( L + 1) − S ( S + 1)] …………………(18.1)
2
Page-3

2S + 1 is the known as multiplicity. Indicates the degeneracy of the level due to spin.

If S = 0 => multiplicity is 1: singlet term.

If S = 1/2 => multiplicity is 2: doublet term.

If S = 1 => multiplicity is 3: triplet term and so on.

In the following we will discuss the effect of spin-orbit interaction on terms to terms transitions.

Singlet levels

S =0; J = L+S = L
ASO
∆ESO =  J ( J + 1) − L ( L + 1)  = 0
2 

1 1
P P1

1 1
S S0

Selection Rules : ∆l = ±1, ∆L = 0, ±1, ∆S = 0, ∆J = 0, ±1


Page-4

Doublet levels

Transition from 2 D to 2 P

For 2 D L = 2, S = 1 , J = 5 , 3
2 2 2

Using equation-18.1

A′
 2 2  2 2
( ) 2 2 (
∆ESO  2 D5  = SO  5 5 + 1 − 2 ( 2 + 1) − 1 1 + 1  = ASO

′ )
A′
 2 2 2 ( ) 2 2 (
∆ESO  2 D3  = SO  3 3 + 1 − 2 ( 2 + 1) − 1 1 + 1  = − ASO
2  
3
2
′ )
For 2 P L = 1, S = 1 , J = 3 , 1
2 2 2

 A′ ′

 2 2 2 2( ) 2 2 (
A′ ′
∆ESO  2 P3  = SO  3 3 + 1 − 1(1 + 1) − 1 1 + 1  = SO
 2 )
A′ ′
 2 2  2 2
( ) 2 2 (
∆ESO  2 P1  = SO  1 1 + 1 − 1(1 + 1) − 1 1 + 1  = − ASO

′′ )
So the construction of energy levels

2
D5

ASO
2

3
2
D ′
ASO 2
2 D3
ν0 2
ν1 ν2 ν3
2
P3
′′
ASO 2

2
2
P ′′
ASO
2
P1
2
Page-5

It is important to know the degeneracy lifted by this spin-orbit interaction. The degeneracy for
 1 
2
D level is 10 (2 L + 1)(2 S + 1) = (2.2 + 1)(2. + 1) = 10 
 2 

When this level splits into 2 D5 and 2 D3 , then


2 2

2
D5 the degeneracy is 6 because J = 5 , mJ = 5 , 3 , 1 , − 1 , − 3 , − 5 . All mJ levels will
2 2 2 2 2 2 2 2
be degenerate. The mJ degeneracy will not be lifted until the external magnetic field is applied.

Similarly, 2 D3 the degeneracy is 4 because J = 3 , mJ = 3 , 1 , − 1 , − 3 .


2 2 2 2 2 2

 1 
The degeneracy for 2 P level is 6 (2 L + 1)(2S + 1) = (2.1 + 1)(2. + 1) = 6 
 2 

When this level splits into 2 P3 and 2 P1


2 2

2
P3 the degeneracy is 4 because J = 3 , mJ = 3 , 1 , − 1 , − 3 . All mJ levels will be
2 2 2 2 2 2
degenerate. The mJ degeneracy will not be lifted until the external magnetic field is applied.

Similarly, 2 P1 the degeneracy is 2 because J = 1 , mJ = 1 , − 1 .


2 2 2 2

We will discuss the later the effect of magnetic field on these levels.

Transitions : as shown in the adjacent figure-18.2


ν→
′′
ASO
′ −
For transition 1→ ν 1 = ν 0 + ASO
2 ν0

3
For transition 2→ ν 2 = ν 0 − ′ + ASO′ ′
ASO
2 ν3 ν1 ν2

3 ′′ Figure 18.2
For transition 3→ ν 3 = ν 0 − ′ − ASO
ASO
2 2

′ ′ > ASO′ so transition pattern will be ν 2 > ν 1 > ν 3 .


In general ASO
Page-6

3
D → 3P transition

S =1 J = 3, 2, 1
For 3D term So,
L=2 3
D3 , 3 D2 , 3 D1

S =1 J = 2, 1, 0
For 3P term
L =1 3
p2 , 3 p1, 3 p0

Now, let us calculate the perturbation energies for these levels using equation-18.1. We take the
spin-orbit constants A2 and A1 for 3D and 3P respectively.

∆ESO ( 3 D3 ) = A22 [12 − 6 − 2] = 2 A2 , ∆ESO ( 3 D2 ) = A22  6 − 6 − 2 = − A2


∆ESO ( 3 D1 ) = A22  2 − 6 − 2  = −3A2
∆ESO ( 3 P2 ) = A21 [6 − 2 − 2] = A1 , ∆ESO ( 3 P1 ) = A21  2 − 2 − 2  = − A1
∆ESO ( 3 P1 ) = A21 [0 − 2 − 2] = −2 A1

3
Transition energies are D3

ν1 = ν 0 + 2A2 − A1 3
D2
3
D
ν 2 = ν 0 − A2 − A1
3
D1
ν0
ν1 ν 2 ν 3 ν 4 ν
5 ν
ν 3 = ν 0 − A2 + A1 6
3
P2
ν 4 = ν 0 − 3A2 − A1
3 3
P P1
ν 5 = ν 0 − 3A2 + A1
3
ν 6 = ν 0 − 3 A2 + 2 A1 P0
Page-7

The Observed transition of Calcium is given in this figure below.

ν5

ν6

ν1
ν4
ν2
ν3

Figure-18.3

Observed Observed Observed Calculated


transitions transitions transitions transitions
(nm) (cm-1) (cm-1)

ν1 = ν 0 + 2A2 − A1 445.47 22448.20 22447.46

ν 2 = ν 0 − A2 − A1 445.59 22442.16 22442.16

ν 3 = ν 0 − A2 + A1 443.50 22547.91 22547.91

ν 4 = ν 0 − 3A2 − A1 445.66 22438.63 22438.63

ν 5 = ν 0 − 3A2 + A1 443.57 22544.36 22544.38

ν 6 = ν 0 − 3 A2 + 2 A1 442.54 22596.68 22597.26


Page-8

From transitions ν 2 and ν 3

E0 − A2 − A1 = 22442.16
E0 − A2 + A1 = 22547.91
−−−−−−−−−−−
2 A1 = 22547.9 − 22442.26
= 105.7
A1 = 52.875 cm −1

From transitions ν 2 and ν 4

E0 − 3 A2 − A1 = 22438.63
E0 − A2 − A1 = 22442.16
−−−−−−−−−−−
2 A2 = 22442.16 − 22438.63
3.53
⇒ A2 = = 1.765 cm−1
2

And ν 0 = 22496.8 cm −1

The other transitions calculated from these values are given in the table.
Page-9

Let us take another example, the transitions of Copper [3d94s4p] 4F → [3d94s5s] 4D

J = 7 ,5 ,3 , 1
4 S=3 2 2 2 2
For D term 2 So, 4
L=2 D7 , 4 D5 , 4 D3 , 4 D 1
2 2 2 2

J = 9 ,7 ,5 , 3
4 S=3 2 2 2 2
For F term 2
4
L=3 F9 , 4 F7 , 4 F5 , 4 F3
2 2 2 2

For 4 F
9 A  9 11 3 5 9
∆E ( J = ) = 1  . − 3.4 − .  = A1
2 2 2 2 2 2 2

7 A 63 − 63
∆E ( J = ) = 2 [ ]=0
2 2 4

5 A 35 − 63 A 28 7
∆E ( J = ) = 1 [ ] = − 1 . = − A1
2 2 4 2 4 2

3 A 15 − 63 A 48
∆E ( J = ) = 1 [ ] = − 1 . = −6 A1
2 2 4 2 4

For 4 D

7 A 63 − 39 A 24
∆E ( J = ) = 2 [ ] = 2 . = 3 A2
2 2 4 2 4

5 A 35 − 39 A 4 A
∆E ( J = ) = 2 [ ]= 2 . =− 2
2 2 4 2 4 2

3 A 15 − 39 A 24
∆E ( J = ) = 2 [ ] = − 2 . = −3 A2
2 2 4 2 4

3 A 3 − 39 A 36 9
∆E ( J = ) = 2 [ ] = − 2 . = − A2
2 2 4 2 4 2
Page-10

The constructed energy level diagram is shown in this figure below.


4
F9
2
4
F 4
F7
2
4
F5
2

4
F3
2

ν0

4
D7
2
4
D
4
D5
2
4
D3
2

4
D1
2

Selection rules: derived from Wigner-Eckart theorem discussed earlier

∆l = ±1, ∆S = 0, ∆J = 0, ±1 . The J = 0 → J = 0 is forbidden.

The transition energies are

9
(1) v (4 F9 → 4 D7 ) = ν o − 3 A2 + A1
2 2 2

(2) v (4 F7 → 4 D7 ) = ν o − 3 A2
2 2

A2
(3) v (4 F7 → 4 D5 ) = ν o +
2 2 2

7
(4) v (4 F5 → 4 D7 ) = ν o − A1 − 3 A2
2 2 2
Page-11

7 A
(5) v (4 F5 → 4 D5 ) = ν o − A1 + 2
2 2 2 2

7
(6) v (4 F5 → 4 D3 ) = ν o − A1 + 3 A2
2 2 2

A2
(7) v (4 F3 → 4 D5 ) = ν o − 6 A1 +
2 2 2

(8) v (4 F3 → 4 D3 ) = ν o − 6 A1 + 3 A2
2 2

9
(9) v (4 F3 → 4 D1 ) = ν o − 6 A1 + A2
2 2 2

The experimental set up for recording the copper 4 F → 4 D spectrum.

Cu electrodes

High voltage
Fiber optic coupled
Cu arc power supply
spectrometer

The copper 4 F → 4 D spectrum.


80000
Copper (4F -> 4D)
465.23
Intensity (arb. units)

60000 453.28
460.86
474.03

40000
458.9
470.57

451.07
20000

0
450 460 470 480 490 500
Wavelength (nm)
Page-12

Assigning the observed three transitions as follows.

Transitions Observed wavelength (nm) Observed transition energy


(cm-1)

4 F7 → 4 D5 458.90 21791.2
2 2

4 F9 → 4 D7 465.23 21494.7
2 2

4 F3 → 4 D 1 470.57 21250.8
2 2

So we have three equations below to solve for obtaining three parameters.

9
ν o − 3 A2 + A1 = 21494.7
2
ν o − 3 A2 = 21250.8
A2
νo + = 21791.2
2

From these above equations, we get

A2
νo + = 21791.2
2
ν o − 3 A2 = 21250.8
7
A2 = 540.4
2
540.4 X 2
A2 = = 154.4 cm −1
7
ν o = 21250.8 + 3*154.4 = 21714
9
A1 = 21494.7 − 21250.8 = 243.9
2
243.9* 2
A1 = = 54.2 cm −1
9
Page-13

Transitions Transition Calculated Calculated observed


energy transition transitions (nm) transitions (nm)
-1
energy (cm )
4
F9 → 4 D7 ν o − 3 A2 +
9
A1
21494.7 465.23 465.23
2 2
2
4
F7 → 4 D7 ν o − 3 A2 21250.8 470.57 470.57
2 2

4
F7 → 4 D5 νo +
A2 21791.2 458.90 458.90
2 2
2
4
F5 → 4 D7 7 21061.1 474.8 474.03
2 2 ν o − A1 − 3 A2
2
4
F5 → 4 D5 7
ν o − A1 +
A2 21601.5 462.93 460.86
2 2
2 2
4
F5 → 4 D3 7
ν o − A1 + 3 A2
21987.5 454.80 453.28
2 2
2
4
F3 → 4 D5 A2 21464.8 465.88 -----
2 2 ν o − 6 A1 +
2
4
F3 → 4 D3 ν o − 6 A1 + 3 A2 21850.8 457.65 ------
2 2

4
F3 → 4 D 1 9 22082.4 452.84 451.07
2 2 ν o − 6 A1 + A2
2
Page-14

Summary of Atomic Energy levels

Gross structure of the atomic energy levels:

It covers largest interactions within the atom:

(a) Kinetic energy of electrons in their orbits.

(b) Attractive electrostatic potential between positive nucleus and negative electrons

(c) Repulsive electrostatic interaction between electrons in a multi-electron atom.

These interactions give energies in the 1-10 eV range and upwards. It determine the
spectrum range whether a photon is IR, visible, UV or X-ray.

Fine structure:

Spectral lines often come as multiplets. E.g., Hα line.

This smaller interactions within atom, called spin-orbit interaction.

The origin of this interaction is

Electrons in orbit about nucleus give rise to magnetic moment which interacts with spin of the
electrons. This introduces splitting of the energy levels and produces small shift in energy.

Hund’s rule :

(1) Of the terms given by equivalent electrons, those with greatest multiplicity lie deepest,
and of these the lowest is that with the greatest L

(2) Multiplets formed equivalent electrons are regular when less than half the shell is
occupied, but inverted when more than half shell is occupied.
Page-15

Recap

In this lecture we have seen the effect of spin orbit coupling on the fine structure of the
multielectron atoms.

We have discussed the transitions between the various spin multiplicity energy levels, sometime
called as multiplets.

We have also gone through the experiments in the copper transitions and also learnt how to
assign the transitions.

We also learnt how to evaluate the spin-orbit constant term for a level from the observed
transitions.
Title: j-j coupling

Page-1

It is seen that for heavier atoms, the nuclear charge causes the spin-orbit interactions to be
strong enough the force between the individual l and s .

For large Z atoms, the electron-electron repulsion becomes weaker than the spin-orbit
interaction of the individual electrons.

In this case, the L-S coupling scheme can not explain the observed transitions from these
atoms.

The calculated energy levels from L-S coupling scheme fails to predict energy levels for
these atoms.

In the lecture, we will describe the j-j coupling scheme that describes the observed
transitions from these heavy atoms.
Page-2

In previous lectures, we have described the Hamiltonian as

2 N N N
e2 N
e2
H =−
2me
∑ ∇i2 + ∑U ( ri ) + ∑
i =1 i =1 i< j rij
− ∑
i < j rij
+ ASO L.S
spin-orbit interaction
Central field electron-electron repulsion

Here, the first term is the central field term, second is the nonspherical part of the electron
electron repulsion and the third part is the spin-orbit interaction.
Since for the low Z-atoms, the electron-electron repulsion is greater, it is treated as
perturbation before the spin orbit interaction.
For High Z atoms, the electron-electron repulsion is lower, it should be treated as
perturbation after the spin orbit interaction.
Thus the Hamiltonian becomes,

2 N N
Z e2 N
e2
H =−
2me
∑ ∇i2 − ∑
i =1 i =1 ri
+ ASO li .si + ∑
i < j rij
spin-orbit interaction
electron-electron repulsion

This is known as j-j coupling


Page-3

In this coupling scheme, Hamiltonian is


2 N N
Z e2 N
e2
H =− ∑ i ∑
2me i =1
∇ 2

i =1 ri
+ ASO li .si + ∑
i < j rij
spin-orbit interaction
electron-electron repulsion

Due to the spin orbit interaction, the spin and orbital angular momenta of the individual
electron get coupled as shown in the figure-20.1.
j =l +s
j 2 = l 2 + s 2 + 2l .s
1 2 2 2 J
l .s =  j − l − s 
2
Thus at this level, j , m j are the good quantum
numbers. Hence to calculate the energy for this
perturbation, we have

∆ESO = jm j ASO l .s jm j
ASO
= jm j j 2 − l 2 − s 2 jm j j1
2
A
= SO [ j ( j + 1) − l (l + 1) − s ( s + 1)] j2
2

l1 l2
s1

s2

Figure-19.1
Page-4

When the electron-electron interaction takes place the individual coupled angular
momenta get coupled to get the total angular momentum J.
So for two electron system,

J = j1 + j2
J 2 = j12 + j 22 + 2 j1. j2
1 2
j1. j2 =  J − j12 − j 22 
2
This interaction energy then

∆Eee = JmJ Aee j1. j2 JmJ


Aee
= JmJ J 2 − j12 − j 22 JmJ
2
A
= ee [ J ( J + 1) − j1 ( j1 + 1) − j2 ( j2 + 1) ]
2
Page-5

Now, let us take some examples.

The electronic configuration 1p2p (non equivalent electrons)

l1 = 1, s1 = 1 , j1 = 3 , 1
2 2 2
l2 = 1 s2 = 1 , j2 = 3 , 1
2 2 2

Now, the notation for the coupling states are { j1 , j2 }J

{ 3 2 , 3 2}
3,2,1,0
{ 3 2 , 1 2} { 1 2 , 3 2} { 12 , 12}
2,1 2,1 1,0

The J states are degenerate in this level of consideration


Next when the electron-electron interaction takes place all these level will be splitted.
{ }
3 ,3
2 2 3
, 3 ,3
2 2 2 { } {
, 3 ,3 , 3 ,3
2 2 1 2 2 0 } { }
{ 3 2 , 12} , 2
{ 3 2 , 12}
1

{ 12 , 3 2} , 2
{ 12 , 3 2}
1

{ 12 , 12} ,
1
{ 12 , 12}0

For the equivalent electrons such as p2 configuration.


The rules are
1. When j1 ≠ j2 , the value of J in { j1 , j2 }J is found by the rules for the coupling of
angular momentum
2. When j1 = j2 , the allowed values in { j1 , j2 }J are given by J = 2 j − 1, 2 j − 3,......
So for p2, the possible states in j-j coupling are

{ 3 2 , 3 2} 2,0
, { 3 2 , 12} 2,1
, { 1 2 , 12}
0
Page-6

In the following the comparison of the levels between the L-S coupling scheme and the j-
j coupling scheme is given. It is to be noted that although the total number of final states
are same, but their relative energies are different.

Spin-orbit j-j coupling


coupling
1
S 1
S0 3 3
  3 3
 2 2 0  
 2 2 0,2
1 1
D D2
3 3
2  
p  2 2 2 p2
3 1
3 1  
3
P2    2 2 1,2
 2 2 2
3
P 3 1
3
P1  
 2 2 1
3
P0
1 1  1 1 
   
 2 2 0  2 2 0

Configuration Coulomb spin-orbit Coulomb spin-orbit Configuration


repulsion repulsion
Page-7

In this lecture we understood that for heavier atoms the L-S coupling scheme is not valid.

In this case we have to consider j-j coupling scheme.

The relative energies of the final states for these two schemes are different, but the total
number of splitted levels is same.

Since the equivalent electrons are in the same values of n and l, the electrostatic
interaction is expected to be larger than spin-orbit interaction. In this case L-S coupling is
favoured.

It is for nonequivalent, j-j coupling is important

There are many situations where neither L-S nor j-j coupling are valid approximation. In
this case both have to be treated simultaneously as perturbation. This coupling scheme is
known as intermediate coupling.
Title: Effect of Static Magnetic field on the spectral lines.

Page-0
In this lecture we will discuss the effect of static magnetic field on the spectral lines.

The effect is known as Zeeman effect and the pattern seen after applying the magnetic
field is known as Zeeman pattern.

We will also discuss the Normal and Anomalous Zeeman effect.

We will see also the change of the Zeeman pattern when the magnetic field is increased.
Page-1

When a source of light emitting line spectra is placed under a static magnetic field, it is
observed that the spectral lines split into several components.

Zeeman in 1896 first observed the phenomenon.


Applying the magnetic field his observations were made in two directions with respect to
the magnetic field directions.
1. Observations perpendicular to the magnetic field:
(a) Spectral lines split into three components
(b) Central line has the same frequency as the original line before applying the magnetic
field.
(c) Central line is linearly polarized and parallel to the magnetic field.
(d) The two other components are equally separated on both sides of the central
frequency.
(e) Both of them are also linearly polarized and polarization is perpendicular to the
magnetic field.

2. Observations along the magnetic field direction:


(a) Central line is absent
(b) The other two components are circularly polarized

This effect is known as Normal Zeeman effect. It was also later observed that in some
cases there were more than three lines. To differentiate, it was named as Anomalous
Zeemen effect. In the following, we will discuss this in details.
Page-2
The interaction of atom and magnetic field can be described as the interactions of
magnetic field with (a) orbital angular momentum and (b) spin angular momentum of
electron.

The Orbital magnetic moment in terms of Bohr magneton in the vector form is
gl µ B µB
µl = − l and putting g l = 1, µl = − l
z
Substituting the value of the angular momentum ml l
µB
µl = l (l + 1) = µ B l (l + 1)
l = l (l + 1)

Z
The spin dipole moment, in terms of spin Lande g-factor g s . s

g s µB
µs = − s

µ s = − g s µ B ms
z
µS

So the total magnetic moment of the atom due to the electron is


µelectron = µorbital + µ spin
= µl + µ s
µB g S µB
=− l− s

= − µ B [l + 2s ]/
Since in this expression will be cancelled with the eigenvalue in terms of , we will
drop this and will consider
µelectron = − µ B [l + 2 s ]

And the interaction energy with the applied magnetic field Bz is


Emag = − µelectron .Bz
Page-3

We know that the Hamiltonian for the atom in the L-S coupling scheme is written as

H= H CF + H ee + H so
l1ml1 s1ms1 Lml Sms JmJ
l2 ml2 s2 ms2

The total quantum number J is the good quantum number and arises from the uncoupled
quantum numbers L and S.
We also know that this L and S are the orbital and spin quantum number for the single
electron case [L = l, and S = s].
But for the multielectron case, these are the total orbital and spin quantum number
[ L = ∑ li , S = ∑ si ]

When the magnetic field is applied, the interaction between the magnetic field and the
magnetic moment is treated as perturbation. Depending on the magnitude we have to
write the Hamiltonian.
At the first stage, we consider that the perturbation due to the magnetic field is lesser than
the interaction due to spin orbit (weak field case). In this case, we can write the
Hamiltonian as

H= H CF + H ee + H so + H mag
l1ml1 s1ms1 Lml Sms JmJ
l2 ml2 s2ms2
Z-axis
The meaning of this in the vector diagram is given in
the figure-20.1. Bz

S
In LS-coupling, the spin-orbit interaction couples the
spin and orbital angular momenta to give a total J
L
angular momentum J according to
J = L+S

In an applied magnetic field, J precesses about Bz Figure-20.1

L and S precess more rapidly about J due to spin-orbit interaction. Spin-orbit effect here
is stronger than the magnetic field..
Page -4

Now we will calculate the interaction energies due to the magnetic field perturbation.
The interaction energy is

Emag = − µelectron .Bz


So,
(
H mag = µ B L + 2 S ⋅ Bz )
(
= µ B L + 2 S ⋅ Bz )
= µ B  L + 2 S  ⋅ Bz = µ B [ Lz + 2S z ] Bz

Since the good quantum number is J we have to calculate,

J m 'J | H mag | J mJ = µ B Bz J mJ | ( Lz + 2S z ) | J mJ

Let us first calculate J m 'J | Lz + 2 S z | J mJ .


Using special case of Wigner-Eckart theorem, ( Lande formula)
J m | A⋅ J | J m
J m'| A| J m = J m'| J | J m
J ( J + 1)

J m 'J | Lz + 2 S z | J mJ =
( )
J mJ | J ⋅ L + 2 S | J mJ
J m ' J | J z | J mJ
J ( J + 1)

( )
J mJ | J ⋅ L + 2 S | J mJ = J mJ | J 2 + J ⋅ S | J mJ

J 2 + S 2 − L2
2
= J mJ | J + | J mJ
2
J ( J + 1) + S ( S + 1) − L ( L + 1)
= J ( J + 1) +
2
3 J ( J + 1) + S ( S + 1) − L ( L + 1)
=
2
Page-5

3 J ( J + 1) + S ( S + 1) − L ( L + 1)
J m 'J | Lz + 2 S z | J mJ = J m ' J | J z | J mJ
2 J ( J + 1)
 3 S ( S + 1) − L ( L + 1) 
= +  J m ' J | J z | J mJ
 2 2 J ( J + 1) 
= g J mJ
3 S ( S + 1) − L ( L + 1)
Where g J = + is known as Lande g-factor of electron
2 2 J ( J + 1)
So,
J m 'J | H mag | J mJ = µ B Bz J mJ | ( Lz + 2S z ) | J mJ
Emag = g J µ B Bz mJ

So this energy correction due to the magnetic field tells us that for a particular J, the
degeneracy for mJ will be removed by applying magnetic field.
The meaning is

mJ = 2, Emag = 2 g J µ B Bz

J =2 mJ = 1, Emag = g J µ B Bz
mJ = 0, Emag = 0
mJ = −1, Emag = − g J µ B Bz
mJ = −2, Emag = −2 g J µ B Bz
Page-6

The following block diagram (figure-20.2) represents the experimental arrangement for
the Zeeman effect.

Sample is placed in the magnetic field.

Magnetic field axis is taken as Z-axis in the laboratory frame.

The light emitting from the sample is passed through the analyzer.

A high resolution monochromator with light detector is placed to record the spectrum.

The transitions are observed either perpendicular to the magnetic field direction (x-axis or
Y axis) or the direction of the magnetic field (Z-axis) through the magnet by making a
hole.

In the perpendicular direction, two observations are made (a) placing the analyzer parallel
to the magnetic field, (b) perpendicular to the magnetic field.

Z-axis
Magnet ↑ Bz Analyzer
Parallel to
Z-axis
X-axis High
resolution
spectrometer
Atomic Light
source such
as (sodium Y-axis
lamp)

Analyzer
Parallel to
X-axis
Figure-20.2
Page-7

Let us take example of transition from a singlet (S = 0) state to a singlet (S = 0) state for
this experiment.
3 S ( S + 1) − L ( L + 1)
For singlet states since S = 0 so J = L and thus g J = + =1
2 2 J ( J + 1)
The following figure-20.3 illustrates the transition 1D2 → 1P1
For 1D2 we have S = 0, L = 2 and J = 2, mJ = 2,1,0,-1,-2.
So this level will split into 5 sublevels with energy separation Emag = g J µ B Bz mJ .
mJ = 2, Emag = 2 µ B Bz
mJ = 1, Emag = µ B Bz mJ = 2
mJ = 0, Emag = 0 1
D2 ( J = 2) mJ = 1
mJ = −1, Emag = − µ B Bz mJ = 0
mJ = −2, Emag = −2µ B Bz mJ = -1
mJ = -2
Similarly, 1P1 will split into ∆mJ = 0
∆mJ = +1
3 sublevels ∆mJ = −1

mJ = 1, Emag = µ B Bz mJ = 1
1
P1 ( J = 1)
mJ = 0, Emag = 0 mJ = 0
mJ = −1, Emag = − µ B Bz mJ = -1
σ π σ
Note here that the separation
between the sublevels is same ν0 ν 0 − µ B Bz ν 0 ν 0 + µ B Bz
µ B Bz and depends on the ν
magnitude of the magnetic Before applying After applying
field magnetic field magnetic field

Figure-20.3
Page-8

Now we will discuss the transition ν 0 after applying the magnetic field. As discussed in
previous lectures that the rule for allowed electric dipole transition is to calculate the
nonzero quantity of the dipole matrix element
i.e J ′mJ′ er JmJ ≠ 0
Since the dipole is a vector quantity, we can use the rule obtained from Wigner-Eckart
theorem.
For vector
J ' m ' | Tq1 | J m = 0 unless ∆ J = 0, ±1
J '+ J ≥ 1
m′ = m − q (q = +1, 0, −1)
∆ m = 0, ±1
Referring to the figure-20.2 and figure-20.3,
Using the relation ∆mJ = 0 , all the transitions have transitions energy = ν 0
Using the relation ∆mJ = −1 , all the transitions have transitions energy = ν 0 − µ B Bz
Using the relation ∆mJ = +1 , all the transitions have transitions energy = ν 0 + µ B Bz
So we will observe only three transitions with transition energies
ν 0 − µ B Bz , ν 0 and ν 0 + µ B Bz
This is known as NORMAL ZEEMAN EFFECT.

Now we will focus the observations in two directions.


(a) When these transitions are observed or the detector is placed perpendicular to the
magnetic field (Y-direction in the figure-20.2 for experiment)
The theorem explains that the J ′mJ′ er JmJ ≠ 0 when q = 0, or q = Z
This means that the transitions ∆m = 0 will have the polarization in the Z-direction.
Again J ′mJ′ er JmJ ≠ 0 when q = +1, or X+iY that is circular polarization and
J ′mJ′ er JmJ ≠ 0 when q = -1, or X-iY that is also circular polarization.
So when the observation is made with analyzer parallel to the Z-axis, only the central
component i.e for ∆m = 0 only ν 0 is observed. This is termed as π component
When the observation is made with analyzer parallel to the X-axis, only the X component
of the other two transitions i.e for ∆m = −1 , ν 0 − µ B Bz and ∆m = +1 ν 0 + µ B Bz are
observed. This is termed as σ component.

(b) When these transitions are observed or the detector is placed along the magnetic
field (Z-direction in the figure-20.2) then since the light propagation direction is
Z, ∆m = 0 will not be observed. But the other two componentsν 0 − µ B Bz and
ν 0 + µ B Bz will be observed and the polarization will be left and right circular
respectively.
Page-9

Now let us consider the case of doublet to doublet transition

1 1
2
P1 s = , L = 1, J =
2
2 2 2 mJ mJ.gJ
P1
1 3 2 +1/2 +1/3
. − 1.2 -1/2 -1/3
3 2 2 2
gJ = + =
2 1 3 3
2. .
2 2

1 1
2
S1 s = , L = 0, J = 2
2
2 2 S1 +1/2 +1
2
1 3
. -1/2 -1
3
gJ = + 2 2 = 2
2 2. 1 . 3
2 2 ν0 σ π π σ
ν
For
1 1 Figure-20.4
2
P1 mJ = Emag = µ B BZ
2
2 3
1 1
mJ = − Emag = − µ B BZ
2 3

And for
1
2
S1 mJ = Emag = µ B BZ
2
2
1
mJ = − Emag = − µ B BZ
2
Page-10

Now the transitions are (using 2 PJ ,mJ notation)


(1)
2
P1 1 → 2S 1 1 ∆mJ = 0, π − polarisation
, ,
2 2 22

1
ν 1 = ν 0 + µ B Bz − µ B Bz
3
2
= ν 0 − µ B Bz
3
(2)
2
P1 1 → 2S 1 1 ∆mJ = −1, σ − polarisation
,− ,
2 2 22

1
ν 2 = ν 0 − µ B Bz − µ B Bz
3
4
= ν 0 − µ B Bz
3
(3)
2
P1 1 →2 S 1 1 ∆mJ = 0, π − polarisation
,− ,−
2 2 2 2

1
ν 3 = ν 0 − µ B Bz + µ B Bz
3
2
= ν 0 + µ B Bz
3

(4)
2
P1 1 →2S 1 1 ∆mJ = +1, σ − polarisation
, ,−
2 2 2 2

1
ν 3 = ν 0 + µ B Bz + µ B Bz
3
4
= ν 0 + µ B Bz
3

This example is the case of sodium D line. This 2 P1 → 2 S 1 transition will split into four
2 2
spectral lines as shown in the figure-20.4. The separation between them will depend on
the magnetic field strength.

When the observation is made perpendicular to the magnetic field, ν 2 and ν 3 will be
observed by placing the analyzer parallel to the Z-axis. These two lines will be missing if
the observation is made along the direction of the magnetic field.
Similarly, ν 1 and ν 4 will be observed when the analyzer is placed parallel to the X-axis.
Page-11

Let us take the case of the other D-line i.e. 2 P3 → 2 S 1 .


2 2
For
mJ mJ .gJ
1 3 +3/2 +6/3
2
P3 s = , L = 1, J =
2
2 2 2
P3 +1/2 +2/3
2
1 3
. − 1.2 -1/2 -2/3
3 4
gJ = + 2 2 = -3/2 -6/3
2 3 5 3
2. .
2 2

2
S1 +1/2 +1
2
-1/2 -1
σ σ π π σ σ

ν0
ν
3 6
For 2
P3 mJ = Emag = µ B BZ
2
2 3
1 2
mJ = Emag = µ B BZ
2 3
1 2
mJ = − Emag = − µ B BZ
2 3
3 6
mJ = − Emag = − µ B BZ
2 3

1
For 2
S1 mJ = Emag = µ B BZ
2
2
1
mJ = − Emag = − µ B BZ
2
Page-12

Now the transitions are (using 2 PJ ,mJ notation)

(1)
2
P3 1 → 2S 1 1 ∆mJ = −1, σ − polarisation
,− ,
2 2 2 2

2 5
ν 1 = ν 0 − µ B Bz − µ B Bz = ν 0 − µ B Bz
3 3
(2)
2
P3 3 → 2S 1 1 ∆mJ = −1, σ − polarisation
,− ,−
2 2 2 2

ν 2 = ν 0 − 2µ B Bz + µ B Bz = ν 0 − µ B Bz
(3)
2
P3 1 → 2 S 1 1 ∆mJ = 0, π − polarisation
, ,
2 2 2 2

2 1
ν 3 = ν 0 + µ B Bz − µ B Bz = ν 0 − µ B Bz
3 3
(4)
2
P3 1 → 2S 1 1 ∆mJ = 0, π − polarisation
,− ,−
2 2 2 2

2 1
ν 4 = ν 0 − µ B Bz + µ B Bz = ν 0 + µ B Bz
3 3

(5)
2
P3 3 → 2 S 1 1 ∆mJ = 1, σ − polarisation
, ,
2 2 2 2

ν 5 = ν 0 + 2 µ B Bz − µ B Bz = ν 0 + µ B Bz

(6)
2
P3 1 → 2 S 1 1 ∆mJ = 1, σ − polarisation
, ,−
2 2 2 2

2 5
ν 6 = ν 0 + µ B Bz + µ B Bz = ν 0 + µ B Bz
3 3
Page-13

Triplet to triplet transition ( 3S1 → 3 P1

3
S1 ; S = 1, L = 0, J =1
3 1.2 − 0
gJ = +
2 2.1.2
=2

3
P1 ; S = 1, L = 1, J = 1
3
gJ = +0
2
3
=
2

3
m j = 1 mJ g J = 2
S1
m j = 0 mJ g J = 0
m j = −1 mJ g J = −2

3
P1 m j = 1 mJ g J = 3
2
m j = 0 mJ g J = 0
σσ π π σσ m j = −1 mJ g J = − 3
2

ν0 ν
Page-14

Now the transitions are

(1)
3
3
S1, −1 
→ P1,0 ∆mJ = −1; σ − polarization
ν 1 = ν 0 − 2 µ B Bz + 0
=ν 0 − 2µ B Bz

(2)
3
3
S1,0 
→ P1,1 ∆mJ = −1; σ − polarization
3
ν 2 = ν 0 + 0 − µ B Bz
2
3
=ν 0 − µ B Bz
2

(3)
3
3
S1, −1 
→ P1, −1 ∆mJ = 0; π − polarization
3
ν 3 = ν 0 − 2µ B Bz + µ B Bz
2
1
=ν 0 − µ B Bz
2
(4)
3
3
S1,1 
→ P1,1 ∆mJ = 0; π − polarization
3
ν 4 = ν 0 + 2µ B Bz − µ B Bz
2
1
=ν 0 − µ B Bz
2
(5)
3
3
S1,0 
→ P1, −1 ∆mJ = 1; σ − polarization
3
ν 5 = ν 0 + 0 + µ B Bz
2
3
=ν 0 + µ B Bz
2
(6)
3
3
S1,1 
→ P1,0 ∆mJ = 1; σ − polarization
ν 6 = ν 0 + 2µ B Bz + 0
=ν 0 + 2 µ B Bz
Page-15

Strong field case: Paschen-Back effect

At this stage, we consider that the perturbation due to the magnetic field is stronger than
the interaction due to spin orbit (strong field case). In this case, we can write the
Hamiltonian as

H= H CF + H ee + H mag + H so
l1ml1 s1ms1 Lml Sms
l2 ml2 s2ms2
Z-axis
= H CF + H ee + µ B ( L + 2 S ).Bz + ASO L.S
Bz
In LS-coupling, the spin-orbit interaction is weaker
L
than the magnetic field interaction and can not couple
the spin and orbital angular momenta.
mL
S
If the magnetic field is applied in the Z-axis , L and S mS
precess more rapidly about Bz .

Emag = Lml Sms µ B ( L + 2 S ).Bz Sms Lml


= µ B Bz Lml Sms ( Lz + 2S z ). Sms Lml
= µ B Bz ( ml + 2ms )
And,
ESO = Lml Sms ASO ( L.S ) Sms Lml
= ASO Lml Sms Lx S x + Ly S y + Lz S z Sms Lml
= ASO ml ms

The components Lx S x and Ly S y are averaged out because these are precessing around the
Z-axis and so only Lz S z will survive.
Page-16
Now let us look at the transition of Sodium line under strong magnetic field. The
transition selection rule used is ∆ml = 0, ±1 , since the electric dipole operator can not
change the spin of the electron ∆ms = 0 .

The magnetic field splits the transition to three transitions, ν 0 − µ B Bz , ν 0 and ν 0 + µ B Bz .

The spin orbit interaction term only shifts the energy level a small amount.

In the weak magnetic field case, there was altogether 10 transitions (4 for 2 P1 and 6 for
2
2
P3 ). As the magnetic field increases these 10 lines merge into three lines. So the
2
anomalous Zeeman pattern observed in weak field case slowly converted to normal
Zeeman pattern in the strong field case.

Emag = µ B Bz ( ml + 2ms ) ESO = ASO ml ms

ms = 1
2
Term ml = 1 ms = 1
2
2
P ml = 0 ms = 1 , − 1
2 2
ms = − 1
2
ml = −1
ms = − 1
2

ms = 1
2
2
S
ms = − 1
2

ν 0 − µ B Bz ν 0 ν 0 + µ B Bz

ν
Page-17
Recap

In this lecture we have discussed the magnetic field effect on the spectral lines.

We discussed that in the weak magnetic field the total angular momentum J precesses
around the applied magnetic field direction.

In this condition, the singlet to singlet transitions show Normal Zeeman effect. Any other
transitions show anomalous Zeeman effect.

Transitions corresponding to ∆m = 0 will have the polarizations in the magnetic field


direction.

Transitions corresponding to ∆m = ±1 will have the circular polarizations.

In strong magnetic field, the coupling between spin and orbit breaks down and L and S
precess around the magnetic field direction.
Hyperfine Structure of Spectral Lines:

Page-1

In this lecture, we will go through the hyperfine structure of atoms.

Various origins of the hyperfine structure are discussed

The coupling of nuclear and electronic total angular momentum is explained.


Page-2

When individual multiplet ( J → J transitions) components are examined with spectral apparatus
of the highest possible resolution, it is found that in many atomic spectra each of these
components is still further split into a number of components lying extremely close together.

This splitting is called hyperfine structure.

−1
The magnitude of the splitting is ~ 2 cm .

Hyperfine structure is caused by properties of the atomic nucleus.

Isotopic effect:

Heavier isotopes present 1 in 5000 in ordinary hydrogen.

Therefore, different isotopes of same element have slightly different spectral lines.

Consider 1H (hydrogen) and 2H (deuterium):

1
RH = R∞ = 1.09677 ×107 meter −1
1+ m
MH
1
RD = R∞ = 1.097074 ×107 meter −1
1+ m
MD

The wavelength difference is therefore:

∆λ = λH − λD = λH  1 − D 
λ
 λH 

= λH 1 − H 
R
 RD  Figure shows the Balmer
line of H & D
This is known as isotope shift.

observed calculated
(cm −1 ) (cm−1 )
Hα 1.79 1.787
Hβ 1.33 1.323
Hγ 1.19 1.182
Hδ 1.12 1.117
Page-3

A quantitative explanation of the isotope effect is not simple, exception H atom. For the heavier
elements the effect is traced back to the change of nucleus radius with mass.

Sm150 − Sm152 is double that of Sm152 − Sm154 .

Usual increase is not from Sm150 → Sm152 .

In many cases the isotope effect is not sufficient to explain the hyperfine structure.

The number of hyperfine structure components is often considerably greater than the number of
isotopes.

In particular, elements which have only one isotope in appreciable amount also show hyperfine
structure splitting.

Likewise, the number of components of different lines is frequently quite different for one and the
same element.

These hyperfine structures can be quantitatively explained, when it is assumed that the “atomic
nucleus possess an intrinsic angular momentum with which is associated a magnetic moment”.

This angular momentum can have different magnitudes for different nuclei and also of course, for
different isotope of the same element.

This is known as Nuclear spin

.
Page-4

Magnetic moment & Angular momentum of the nucleus.

Nucleus consists of Proton & Neutron.

Proton:

1
(i) Possesses angular momentum I P described by the spin quantum number ; this angular
2
momentum obeys the general rules of quantization.

1
Component along oz-axis ( I P ) z = ±
2

 1  1 
Magnitude I P =  2  2 + 1 
  

(ii) Possesses a magnetic moment µ P parallel and in the same sense as its angular
momentum I P .

e
Magnetic momentum of proton µ P = ( µ P ) z = 2.79 = 2.79 µ BN
2 MK

e m 1
µ BN = nuclear magneton = ≈ µB = µB
2 MK M 1836

µB
So, µ N =
B
1836

Neutron: also possesses magnetic moment

µ Neu = −1.913 µ BN

But µ Neu + µ P ≠ total magnetic moment of the nucleus.


Page-5

The structure of the nucleus is complex.

Inter nuclear forces are non-central forces involving angles between the magnetic moments and
the radius vector joining the nucleus.

Further more, within the nucleus the nucleus possess an orbital angular momentum which can be
zero for certain nuclei.

The nuclear magnetic moment µ N is related to nuclear angular momentum I,

µN = g I µB I
1
= g IN µ BN I So, g I = g IN
1836

g I or g IN is called nuclear Lande’ factor.

The general adopted sign for g I :

The Lande’ factor is positive when the nuclear magnetic moment and angular momentum
are in the same direction, and is considered negative when in the opposite direction.

I I
µN

µN

g I < 0, µ I < 0 g I > 0, µ I > 0


Page-6

Nuclear spin and Magnetic moment:

(1) All isotopes having an even mass no. A and an even atomic number z , have zero
nuclear spin and zero nuclear magnetic moment.

Example: 42 He ; 16 20
8 O ; 10 Ne ; .....

(2) All isotopes having an even mass number A and an odd atomic number z , have an
integral nuclear spin.

2 6 10
1 D, I =1 ; 3 Li, I =1 ; 5 B, I = 3 ; .....

(3) All isotopes having an odd mass number A , have a half integral nuclear spin.

1
1H, I =1 ; 3
2 He, I =1 ; 39
19 K , I =3
2 2 2
1s 1s 2 s [ _ ] 4s
Page-7

Magnetic field due the orbital motion of electron:

A point charge is q = −e is moving in a classical orbit with a velocity V . At a given instant, the
field it creates at the nucleus, is

µ0 i d × r µ0 −r µ q
B= 3
= qV × 3 = 0 3 m r × V
4π r 4π r 4π m r
µ q l
= 0
4π m r 3

r is directed from the nucleus towards the charge q .

Hence the magnetic field due to the orbital motion of the electron is

µ0 q 1 µ 1
Bl = 3
l = − 0 2µB 3 l
4π m r 4π r

The interaction energy between the nuclear magnetic moment and the orbital motion of electron
µ0 1
El = − µ N ⋅ Bl = 2 g I µ B2 l ⋅ I 3
4π r

Using special case of Wigner-Eckart theorem, ( Lande formula)

J m | A⋅ J | J m
J m'| A| J m = J m'| J | J m
J ( J + 1)

jm|l ⋅ j | jm
j m'|l | j m = j m'| j | j m
j ( j + 1)

l.j
So we can substitute, l = j and we get
j ( j + 1)

µ0 1 µ l.j 1
El =

2 g I µ B2 l ⋅ I 3 = 0 2 g I µ B2
r 4π j ( j + 1) r 3
( I . j ) ……………………..(21.1)
Page-8

The interaction energy between two magnetic dipole moments µ N and µ S separated by r is
given by

µ0  µ N .µS 3 ( µ N ⋅ r )( µ S ⋅ r ) 
ESpin =  +  …………………………………(21.2)
4π  r 3 r5 

Substituting the value of µ N and µ S in this equation, we get

µ0 g I g S µ B2  I . s 3 ( I ⋅ r ) ( s ⋅ r ) 
ESpin = − 3 +  ……………………………..(21.3)
4π  r r5 
 

I .s
Let us take the first term
r3

jm| s ⋅ j | jm
Using the relation j m ' | s | j m = j m'| j | j m
j ( j + 1)

s. j
So we can substitute, s = j and we get
j ( j + 1)

I .s j.s 1 l.s + s 2 1
= (I . j ) = (I . j )
r3 j ( j + 1) r 3 j ( j + 1) r 3

Now let us take the second term


(
3 I ⋅r )(s ⋅ r )
r5

r. j
we can substitute, r = j and we get
j ( j + 1)

(
3 I ⋅r ) ( s ⋅ r ) = 3( I ⋅ j ) ( s ⋅ r ) r. j
5 5
r r j ( j + 1)
Page-9

j .r = (l + s ).r

Now, = l .r + s .r
= 0 + s .r
= s .r

So,
(
3 I ⋅r ) ( s ⋅ r ) = 3( I ⋅ j ) ( s ⋅ r ) 2

r5 j ( j + 1) r 5

Substituting the values in equation-21.3 of


I .s
and
( )
3 I ⋅ r (s ⋅ r )
, we get the interaction
r3 r5
energy for spin

µ0 g I g S µ B2  I . s 3 ( I ⋅ r ) ( s ⋅ r ) 
ESpin = − 3 + 
4π  r r5 
 
µ0 g I g S µ B2  3 ( I ⋅ j ) ( s ⋅ r ) l.s + s 2 1 
2

= − ( I . j )  …………………..(21.4)
4π  j ( j + 1) r 5 j ( j + 1) r 3 
 
µ0 g I g S µ B2 ( I ⋅ j )  3 ( s ⋅ r ) l.s + s 2 
2

=  − 
4π j ( j + 1)  r 5 r3 
 
Page-10

Now we will calculate the total interaction energy due to electron orbital (equation 21.1) and spin
(equation 21.4)

EHF = El + ESpin

µ0 2 g I µ B2 ( I ⋅ j )  3 ( s ⋅ r ) l.s + s 2 
2
µ l.j 1
= 0 2 g I µ B2
4π j ( j + 1) r 3
( ) 4π
I . j + 
j ( j + 1)  r 5

r3 

 

Here we have substitutes g S = 2 for the electron.

So

µ 2 g I µ B2 ( I ⋅ j )  l . j 3 ( s ⋅ r ) l.s + s 2 
2

EHF = 0  + − 
4π j ( j + 1)  r 3 r5 r3 
 
µ 2 g I µ B2 ( I ⋅ j )  l .s + l 2 3 ( s ⋅ r ) l.s + s 2 
2

= 0  + − 
4π j ( j + 1)  r 3 r5 r3 
 
µ 2 g I µ B2 ( I ⋅ j )  l 2 3 ( s ⋅ r ) s 2 
2

= 0  + − 3
4π j ( j + 1)  r 3 r5 r 
 

1 11  3
Substituting s .r = r and s 2 = s ( s + 1) =  + 1 = , we get
2 22  4

µ 2 g I µ B2 ( I ⋅ j )  l (l + 1) 3 ( r ) 3 
2

EHF = 0  3 + − 
4π j ( j + 1)  r 4 r 5 4r 3 
 
µ0 2 g I µ B2 ( I ⋅ j )  l (l + 1) 
= …………………….(21.5)
4π j ( j + 1)  r 3 
µ0 2 g I µ B2 l (l + 1) 1
=
4π j ( j + 1) r 3
(I ⋅ j )
Page-11

So the Hamiltonian including the hyperfine interaction for one electron system is

H = H 0 + H Spin −Orbit + H Hyperfine

Here the hyperfine interaction is coupling the total angular momentum of the electron j and the
nuclear angular momentum I. So we need the new angular momentum F which will be the good
quantum number for the total Hamiltonian.

So we define,

F = j + I and the eigenfunction is F mF which will be the coupled state arising from the

uncoupled states of j m j I mI

′ I⋅j
The interaction energy EHF = AHF

′ → constant, characteristic of the level j and l .


AHF

Note that the value of AHF = 0 for l = 0, i.e. for the s-states.

However, experimentally splitting is observed for the 2 S 1 state of hydrogen. This can not be
2

explained by this classical explanation.

However, starting from the Dirac equation, if one evaluates the Hamiltonian for the hyperfine
interaction including the vector potential (Reference: Atoms and Molecules by M. Weissbluth), it
becomes

µ0 2 g I µ B2  ( + 1) 1 8π 
Hh =  I⋅j+ ( )
δ (r ) I ⋅ s 
 j ( j + 1) r
4π 3 3 

= AHF
( + 1) I ⋅ j + A I ⋅ s = A′ I ⋅ j + A I ⋅ s ……………………(21.6)
F HF F
j ( j + 1)

The first term is the dipole-dipole interaction with corresponding to classical expression as we
derived earlier. The last term is known as Fermi Contact Interaction term, it has no classical
analog and contributes only for s-states. Since ψ ( 0 ) at r = 0 for non-s states is zero, Fermi
contact term goes to zero for non-s states.

2 1
δ ( r ) = ψ ( 0) =
π a03
Where a0 is the Bohr radius.

Page-12
For = 0; i.e. s states first term zero; second term is important. For penetrating orbit
contact term is important.

F = I +s
F 2 = I 2 + s2 + 2 I ⋅ s
F 2 − I 2 − s2
⇒ I ⋅s =
2

AF
EHF = F mF AF I .s F mF =  F ( F + 1) − I ( I + 1) − s ( s + 1) 
2 

For hydrogen, 2 S 1 state, I = 1 , s = 1 , F = 0,1


2 2 2

 AF 1  1  1  1 
EHF ( F = 1) = 1(1 + 1) − 2  2 + 1 − 2  2 + 1 
 2    
So
A  3 A
= F 2 −  = F
2  2 4

AF 1  1  1  1 
EHF ( F = 0) = 0 ( 0 + 1) − 2  2 + 1 − 2  2 + 1 
2    
A  3 3
= F 0 −  = − AF
2  2 4

The hyperfine splitting = ∆EHF = EHF ( F = 1) − EHF ( F = 0) = AF

The calculated values of AF = 0.047 cm-1.

F =1
n =1 AF
2
4
S1 0.047 cm −1
2
3 AF
4
F =0
Page-13

For non s states, We have

EHF = AHF
( + 1) I ⋅ j
j ( j + 1)
F 2 = I 2 + j2 + 2 I ⋅ j
F 2 − I 2 − j2
⇒I⋅j=
2
A ( + 1)  F F + 1 − I I + 1 − j j + 1 
EHF = HF ( ) ( ) ( )
2 j ( j + 1) 

For hydrogen, 2 P3 state, l = 1, I = 1 , j = 3 , F = 2,1


2 2 2

AHF 1(1 + 1)  1  1  3  3 
EHF ( F = 2) =  2 ( 2 + 1) −  + 1 −  + 1 
So,
2 3 3 +1 
2 2 ( ) 2  2  2  2 

A 8  18  2
= F 6 −  = AHF
2 15  4 5

AHF 1(1 + 1)  1  1  3  3 
EHF ( F = 1) = 1(1 + 1) −  + 1 −  + 1 
And
2 3 3 +1 
2 2 ( ) 2  2  2  2 

A 8  18  2
= F  2 −  = − AHF
2 15  4 3

16
The hyperfine splitting = ∆EHF = EHF ( F = 2) − EHF ( F = 1) = AHF
15

F =2
2
2
AHF
P3 5
2
2
AHF
3
F =1
Page-14

For hydrogen, 2 P1 state, l = 1, I = 1 , j = 1 , F = 0,1


2 2 2

AHF 1(1 + 1)  1  1  1  1 
EHF ( F = 1) = 1(1 + 1) −  + 1 −  + 1 
So,
(
2 1 1 +1 
2 2 ) 2  2  2  2 

A 8 3 2
= F  2 −  = AHF
2 3 2 3

And

AHF 1(1 + 1)  1  1  1  1 
EHF ( F = 0) =  0 ( 0 + 1) −  + 1 −  + 1 
2 1 1 +1 
2 2 ( ) 2  2  2  2 

A 8  3
= F 0 −  = −2 AHF
2 3  2

8
The hyperfine splitting = ∆EHF = EHF ( F = 1) − EHF ( F = 0) = AHF
3

F =1
2
2
AHF
P1 3
2
2 AHF

F =0
Page-15

For multielectron atom

The interaction energy is EHF = A ' I ⋅ J

A ' → hyperfine constant, characteristic of the level J and L

F2 = I2 + J2 + 2I ⋅ J
F2 − I2 − J2
⇒ I ⋅J =
2
A'
EHF ( F ) =  F ( F + 1) − I ( I + 1) − J ( J + 1) 
2

EHF ( F + 1) − EHF ( F ) = hyperfine splitting or hyperfine structure.

Hyperfine splitting is very small – measurements can be made to a high degree of precision.

The general conference of weights and measures (1964) defined “atomic second” from
the transition between the hyperfine energy levels F = 4, mF = 0 and F = 3, mF = 0 of the
6
S 1 , Ground state of Cs atom 133
55 Cs .
2

These two sublevels correspond to parallel and anti parallel orientations of the spins S =
1/2 of the valence electron and I = 7/2 of the nucleus of the Cs atom
Page-16

Various Corrections:

Since the hyperfine splitting is very small, a lot of small corrections are needed.

(1) Polarization of the inner shells: For atoms with many electrons, the resultant of the
electron spins in the completed inner subshells cannot be regarded as zero, statistically
each spin has a slight tendency to align parallel to the spins of the valence electrons. In
evaluating the field B0 , it is necessary to take account of this magnetization of the inner
shells 30%.
(2) Relativistic effect: As a result of high electrostatic charge of the nucleus with high atomic
number z , the velocity of the electrons is high in the neighborhood of the nucleus and
corrections are necessary. These corrections can modify the result with far heavy atoms
for levels with small J , by a factor of the order of two.
(3) Volume effects: With increasing t , the approximation of a point nucleus cannot be
preserved.

Electric Quadrupole Effects

The distribution of the charge q N within the nucleus is not spherically symmetric.

In classical theory, if the origin is taken as the center of gravity of the electric charges.
Within the nucleus, the corresponding electric dipole moment is zero, there then remains the
problem as to relative positions of the center of gravity of the electric charges and of the center of
gravity of the masses within the nucleus.

In quantum theory, symmetry rules result in zero dipole moment for the nucleus. The first
term in the multipole moment expansion corresponds to the interaction of electric quadrupole
moment with electric field gradient create by the electrons in the region of the nucleus.

Let’s assume that the nucleus has a cylindrical charge distribution around its own Oz
axis, I is also Oz axis. The electron cloud has cylindrical symmetry around Oz axis (direction of
J ). The electric field gradient,

∂E z ∂ 2V
φzz = − =
∂ z ∂ z2

Q → quadrupole moment of the nucleus


Page-17

The additional energy ∆EQ resulting from quadrupole coupling will be,

e Q φ zz  3 2 1
∆EQ =  cos θ − 
4 2 2

Where θ is the angle between Oz ( I ) and Oz ( J ) .

Define a quadrupole coupling constant = D = eQ φzz

D3 2 1
∆EQ =  cos θ − 
4 2 2

From vector model,

F ( F + 1) − I ( I + 1) − J ( J + 1)
cos θ =
2  I ( I + 1) J ( J + 1) 

Or, using quantum mechanics,

3
D2 (
C C + 1) − 2 I ( I + 1) J ( J + 1)
∆EQ =
4 I ( 2 I − 1) J ( 2 J − 1)

Where, C = F ( F + 1) − I ( I + 1) − J ( J + 1)

F=5
2 D
4
3 A′
J =1 2

I=3 A′
2 D
F=3
2
5 A′
2 5D
4
F=1
2
Page-18

Recap

In this lecture we came to know the origin of hyperfine structure such as isotope effect,
hyperfine interaction etc.

The hyperfine structure is very small and can only be observed with a very high
resolution.

We have understood the interaction of interaction of nuclear magnetic moment and the
total electronic angular moment.

We now know that the ground state hyperfine splitting of hydrogen can not be described
by classical concept.

The Fermi contact term is important to describe this splitting and quite accurately predict
the experimental observation.

We have also understood the various corrections due to quadrupole effect, volume effect
and relativistic effect.
Zeeman effect in Hyperfine structures

Page-0

In the last lecture, we understood the hyperfine structure that originates due to the
interaction between the total angular momentum of electron and the nuclear angular
momentum

In this lecture we will discuss the effect of static magnetic field on this hyperfine
structure.

We will discuss this concept on the basis of the magnetic field strength.
Page-1
By now, we know that the nuclear magnetic moment is

µN = g I µB I
1
= g IN µ BN I So, g I = g IN
1836

As discussed before, the total magnetic moment due to the electron is


µelectron = µorbital + µ spin
= µl + µ s
= −µB [ L + g s S ]
So the total magnetic moment of the atom including the nuclear part is
µtotal = µelectron + µnucleus
= −µB L − g S µB S + g I µB I

And the interaction energy with the applied magnetic field Bz is


Emag = − µtotal .Bz

We know that the Hamiltonian for the atom in the L-S coupling scheme including the
interaction with nuclear magnetic moment is written as

H= H CF + H ee + H so + H hf
l1ml1 s1ms1 Lml Sms JmJ FmF
I mI
l2 ml2 s2 ms2

The total electronic quantum number J and I are not the good quantum number.

Here the hyperfine interaction is coupling the total angular momentum of the electron J
and the nuclear angular momentum I. So the new angular momentum F which will be the
good quantum number for the total Hamiltonian.
So we have
F = J + I and the eigenfunction is F mF which will be the coupled state arising from
the uncoupled states of J mJ I mI
The hyperfine interaction energy
′ I ⋅J
EHF = AHF
Page-2
When we apply the magnetic field, this will interact with the total magnetic moment of
the atom.

There will be three situations of the interaction those depends on the magnitude of the
magnetic field strength compared to the hyperfine interaction.
1. The weak field case
2. The intermediate field and
3. The strong field case.

Since the coupling strength between the nuclear spin and the total angular momentum of
the electron is very small, the magnitude of the applied magnetic field should be very
small.

On the other hand, the field considered to be strong compared to hyperfine interaction
will be small compared to spin-orbit interaction.

Here we will discuss only two these cases.

Case-I

Weak Field case

For this situation we can write the Hamiltonian as

H= H CF + H ee + H so + H hf + H mag ……………………….(22.1)
l1ml1 s1ms1 Lml Sms JmJ FmF
I mI
l2 ml2 s2 ms2

Z-axis
When this magnetic field is applied the total angular
momentum F starts precessing around the magnetic Bz
field.
I
And the interaction energy with the applied magnetic
field Bz is F
J
Emag = FmF − µtotal .Bz FmF
= FmF ( µ B L + g S µ B S − g I µ B I ). Bz FmF
= µ B Bz FmF ( Lz + 2 S z − g I I z ) FmF
Here we have substituted spin Lande g-factor g s = 2 .
Page-3

So we have to calculate

F m 'F | H mag | F mF = µ B Bz F mF | ( Lz + 2 S z ) − g I I z | F mF …….(22.2)

Again using special case of Wigner-Eckart theorem, ( Lande formula)


J m | A⋅ J | J m
J m'| A| J m = J m'| J | J m
J ( J + 1)
So,

F m 'F | Lz + 2 S z | F mF =
( )
F mF | J ⋅ L + 2 S | F mF
F m 'F | J z | F mF ….(22.3)
J ( J + 1)

Now,
( ) ( )
J ⋅ L + 2S = J ⋅ L + S + S = J ⋅ J + S ( )
= J2 + J ⋅S

We have to find J ⋅ S
So,
J = L+S
J 2 + S 2 − 2 J .S = L2
J 2 + S 2 − L2
J .S =
2
Substituting, we get
( )
F mF | J ⋅ L + 2 S | F mF = F mF | J 2 + J ⋅ S | F mF

J 2 + S 2 − L2
= F mF | J 2 + | F mF
2
J ( J + 1) + S ( S + 1) − L ( L + 1)
= J ( J + 1) +
2
3 J ( J + 1) + S ( S + 1) − L ( L + 1)
=
2
Page-4
From equation 22.3
3J ( J + 1) + S ( S + 1) − L ( L + 1)
F m 'F | Lz + 2 S z | F mF = F m 'F | J z | F mF
2 J ( J + 1)
3 S ( S + 1) − L ( L + 1)
= + F m 'F | J z | F mF
2 2 J ( J + 1)
= g J F m 'F | J z | F mF

3 S ( S + 1) − L ( L + 1)
Where g J = + is known as Lande g-factor of electron
2 2 J ( J + 1)

F mF | J ⋅ F | F mF
F m 'F | J z | F mF = F m 'F | Fz | F mF
F ( F + 1)
F ( F + 1) + J ( J + 1) − I ( I + 1)
= mF
F ( F + 1)
= a mF

F mF | F ⋅ I | F mF
F m 'F | I z | F mF = mF
F ( F + 1)
F ( F + 1) + I ( I + 1) − J ( J + 1)
= mF
F ( F + 1)
= b mF

F m 'F | H mag | F mF = ( a g J mF − g I b mF ) µ B Bz
…………(22.4)
= ( a g J − b g I ) mF µ B Bz = g hf mF µ B Bz

Where g hf = ag J − bg I
Thus applying the magnetic field the hyperfine energy levels will split as shown in the
figure mF
2
F =2 1
0
-1
2 -2
P3
2

1
0
F =1 -1
Page-5
Case –II :Strong field case:

For this situation we can write the Hamiltonian as

H= H CF + H ee + H so + H mag + A′I .J ……………..(22.5)


l1ml1 s1ms1 Lml Sms JmJ
I mI
l2 ml2 s2ms2

Z-axis

Bz

mJ
I
mI

In LS-coupling, the spin-orbit interaction is stronger than the magnetic field interaction
and hyperfine interaction is weaker so hyperfine interaction can not couple J and I.

If the magnetic field is applied in the Z-axis , J and I precess more rapidly about Bz .

This is known as Back-Goudsmit Effect which is strong field case.

Emag = µ B Bz J mJ | I mI | ( Lz + 2 S z ) | I mI | J mJ
= µ B Bz  J m 'J | Lz + 2 S z | J mJ − I m 'I | g I I z | I mI 
= µ B Bz g J mJ − µ B Bz g I mI

And the correction term = J mJ | I mI | A′I .J | I mI | J mJ = A ′mI mJ

So Emag+ Ehf = ( mJ g J − mI g I ) µ B Bz + A ' mI mJ


………………….(22.6)
Page-6

Here we take an example of the ground state of hydrogen 2 S 1 .


2

For this, L = 0, S = 1 , J = 1 , I = 1 . The energy level structure after applying the


2 2 2
magnetic field is shown in the following diagram. It is to be noted that, each Zeeman
level splits into (2I+1) level.

µ B Bz g J mJ − µ B Bz g I mI A ' mI mJ

mI = − 1 mJ = 1 , mI = 1
mJ = 1 2 2 2
2

mI = 1 mJ = 1 , mI = − 1
2 2 2

2
S1
2
E.P.R

mJ = − 1 , mI = − 1
2 2
mI = − 1
mJ = − 1 2
2 N.M.R

mI = 1 mJ = − 1 , mI = 1
2 2 2

The transitions shown in blue lines corresponding to electron spin resonance (E.P.R) and
the transitions in reds correspond to nuclear magnetic resonance.

Since these transitions are within the same electronic configuration, these are not electric
dipole transitions. They are magnetic dipole transitions.

In the next two lectures, we will go through the details of E.P.R and N.M.R.
Page-7

In this lecture we have gone through the interaction of the magnetic field and the total
magnetic moment of the atom

In the weak field case each hyperfine level splits into (2F+1) levels

In the strong magnetic field the hyperfine interaction does not couple I and J. Instead, I
and J precess around the magnetic field.

This phenomenon is used to develop the electron spin resonance and nuclear spin
resonance spectroscopy those are widely used for various fields of research.
Title: Electron Spin resonance spectroscopy
Page-0

In this lecture, we will learn about the applications of the magnetic field effect on atoms
which we have learnt in previous lectures.

Here, the electron Spin resonance spectroscopy (EPR) or sometimes known as electron
spin resonance (ESR) will be discussed.

Since the magnetic field interacts with the spin, the spin of the system has to be nonzero.
Because of this reason, the system having free electrons or unpaired electrons is required
for this spectroscopic study.

We will start with the basic understanding of this spectroscopy and then we will discuss a
few applications of it.
Page-1

Electron Paramagnetic Resonance ( EPR ), sometimes referred to as Electron Spin


Resonance ( ESR ), is a widely accepted spectroscopic technique in various research
fields.

The resonance occurs between the spin of the unpaired electron and the electromagnetic
field under a static magnetic field.

This technique is used to study paramagnetic centers on various oxide surfaces, which are
frequently encountered in heterogeneous catalysis.

The observed paramagnetic centers include surface defects, inorganic or organic radicals,
metal cations or supported metal complexes and clusters.

Each of these paramagnetic species will produce a characteristic EPR signature.

Diamagnetic oxide materials can also be studied using suitable paramagnetic probes,
including nitroxides and transition metal ions.
Page-2
Let us start with ground state 2 S 1 of hydrogen atom having single electron. Here, L =0,
2
S = ½, and J = ½. Under the magnetic field this level splits into two energy levels,
essentially due to the spin projection ms = ½ and ms = - ½ because mL = 0.
The splitting energy due to the magnetic field is Emag  g J B Bz as shown in figure 23.1.
The value of gJ = 2.
The transition between this level is governed by the magnetic dipole selection rule
l  0, S  1. This transition changes the spin projection by 1.

B Bz g J mJ Absorption
mJ  1
2

2
B Bz
S1 Bz 
2 h  2B Bz
First derivative

B Bz

mJ   1 Bz 
2
Figure-23.1

Since,
h  gS B Bz  2B Bz ……………………………………(23.1)
Here, g J  gs because L = 0

h  g S  B Bz  2 B
 2 B 2  9.27 1021 erg / gauss
   2.8 MHz / gauss
Bz h 6.626 1027 erg  sec

It can be seen that the frequency required for the transition to occur is about 2.8 MHz per
Gauss of applied field.
This means that for the magnetic field usually employed in the laboratory, the radiation
required belongs to the microwave region.
When the magnetic field used is about 3400 Gauss, the corresponding applied frequency
required is in the microwave ~9 to 10 GHz).
This corresponds to a wavelength of about 3.4 cm and is known as the X - band
frequency.
Page-3

The following figure-23.2 represents the block diagram of the experimental set up for
E.S.R

Modulation
Sample Klystron

Magnet
Detector
Amplifier

Detector

Magnet Sweep
power supply

Modulator Data acquisition


system
Figure-23.2

The experimental set up consists of

1. Microwave radiation source is Klystrons or gun oscillator

2. A sample cell is a cavity where the microwave is transmitted through waveguide

3. A d.c magnetic field and a sweeping circuit for changing the magnetic field.

4. A detection system basically a rectifier crystal for measuring the absorption

5. A data acquisition and a phase sensitive signal processing unit

6. A computer to control the magnet as well as storing the data


Page-4
The ratio of the electron population in the mJ  1 state n1 to the mJ   1 state n 1
2 2 2 2

n1  E
g  B
 J B z
at a given temperature is given by 2
 e kT  e kT
n 1
2

To improve the sensitivity of the measurement, either Bz increases or temperature


decreases

For common EPR experiments one of the frequency is selected for microwave radiation

Band  0 (MHz)  (cm) Bz (Gauss)


X 9,500 3 3,400
K 36,000 0.8 13,000

mJ   1
2

h 0  g J B Bz
Energy
Bz 
For resonance condition, (figure-23.3)
instead of varying 0 , the magnetic
mJ   1
field is varied. 2

At resonance condition, the absorption Absorption


takes place. So the graph obtained is
the absorption versus the magnetic
field.

To measure the value of the magnetic


field correctly, the first derivative of Bz 
the absorption curve is recorded, as
shown in the figure. First derivative

Bz 

Figure-23.3
Page-5
Now we will consider the interaction of electron magnetic moment and the nuclear spin.
As discussed in the previous lecture, the interaction energy (Back-Goudsmit Effect) is
 mJ g J  mI g I   B Bz  A ' mI mJ …………………..(23.2)

when the magnetic field strength is higher than the hyperfine interaction.
Using this relation, we see that the electron Zeeman levels split further into four levels
due to the nuclear spin.

B Bz g J mJ B Bz g I mI A ' mI mJ

mI   1 mJ  1 , mI  1
mJ  1 2 Ea 2 2
2

mI  1 Eb mJ  1 , mI   1
2 2 2

2
S1
2
E.P.R

Ec mJ   1 , mI   1
2 2
mI   1
mJ   1 2
2

mI  1 Ed mJ   1 , mI  1
2 2 2
g J B Bz

Figure-23.4

A A
2 2

The E.P.R Selection rule : mJ  1, mI  0 .


The reason for this is that these transitions are magnetic dipole transitions of electron
where the mJ changes by 1. But this cannot change the nuclear spin simultaneously. So
the mI will be unchanged during the transition.
Applying this E.S.R. selection rule two transitions are observed.

Page-6
We calculate now the energies of these levels and see the energies of these two
transitions. Substituting the value of mJ   1 and mI   1 in equation 23.2, we get
2 2

1 A 1 1
Ea  g J  B Bz  g I  B Bz  mJ  , mI 
2 4 2 2

1 A 1 1
Eb  g J  B Bz  g I  B Bz  mJ  , mI  
2 4 2 2

1 A 1 1
Ec   g J  B Bz  g I B Bz  mJ   , mI  
2 4 2 2

1 A 1 1
Ed   g J  B Bz  g I  B Bz  mJ   , mI 
2 4 2 2

And the two transition energies are

A
 1  Ea  Ed  g J B Bz 
2

A
 2  Eb  Ec  g J  B Bz 
2

So these two transitions are separated by the hyperfine constant A between them. This is
a direct measure of the Fermi contact term (AF) for the hydrogen atom discussed in
previous lectures.
The intensity of both the transitions will be equal as shown in figure 23.4.

Note:
Since in the transition energy expression g I B Bz does not appear, from now on we will
drop this term for the discussion as well as we will not include it in the diagram.
Page-7

Let us consider that the electron is interacting with two similar protons with I = ½ and ½ .
According to coupling of angular momenta, we have the value of coupled I = 1, 0 and the
value of mI = 1, 0, -1
As shown in figure 23.5, each electron Zeeman level splits into three levels. Using E.S.R
selection rule mJ  1, m1I  0, m 2I  0 , there will be three transitions. The
separation between the two transitions is A . The intensity ratio is 1:2:1, because the
middle line is consisting of two transitions as shown in figure 23.5

AmI2 mJ
Am1I mJ
mJ  1 , M I  1, m1I  1 , mI2  1
B Bz g J mJ 2 2 2

mI  1 mJ  1 , M I  0, m1I  1 , m 2I   1
2 2 2 2
mJ  1 mJ  1 , M I  0, m I   1 , m I  1
1 2
2 2 2 2
mI   1
2 mJ  1 , M I  1, m1I   1 , mI2   1
2 2 2
2
S1
2

mI   1 mJ   1 , M I  1, m1I   1 , m2I   1


2 2 2 2
mJ   1 , M I  0, m1I  1 , m 2I   1
2 2 2
mJ   , M I  0, m I   , m I 
1 1 1 2 1
mJ   1 2 2 2
2 mI  1
2 mJ   1 , M I  1, m1I  1 , mI2  1
2 2 2
1 1 2
1 1

Figure-23.5
Page-8

Let us take an example of Benzene radical.


Electron delocalized to all six C  H bonds

Here all protons will be interacting with the electron same way, so the hyperfine
interaction will be same.

So let us predict the E.S.R. spectrum of benzene radical. There will be (6+1) = 7 lines.

The intensities of the lines of the spectrum follow the binomial expansion (1+x)n.

For six equivalent protons, the intensity ratios 1:6:15:20:15:6:1

This can be determined by Pascal’s triangle given below in figure-23.6.

1 1 I  1 1
2

1 2 1 I  2 1
2

1 3 3 1 I  3 1
2

1 I  4 2
1
1 4 6 4

1 5 10 10 5 1 I  5 1
2

1 6 15 20 15 6 1 I  6 1
2

Figure-23.6
Page -9

For nonequivalent protons, we write the interaction energy as


Emag  mJ g J B Bz  A1m1I mJ  A2 m2 mJ I
The energy level diagram is shown in figure-23.7

A2 mI2 mJ
A1m1I mJ
mJ  1 , m1I  1 , mI2  1
B Bz g J mJ 2 2 2

mI  1 mJ  1 , m1I  1 , m I2   1
2 2 2 2
mJ  1 mJ  1 , m1I   1 , m I2  1
2 2 2 2
mI   1
2 mJ  1 , m1I   1 , m2I   1
2 2 2
2
S1
2

mI   1 mJ   1 , m1I   1 , m2I   1
2 2 2 2
mJ   , m I 
1 1 1 ,m  
2 1
2 2 I 2
mJ   1 mJ   , m I   , m I 
1 1 1 2 1
2 2 2 2
mI  1
2 mJ   1 , m1I  1 , m2I  1
2 2 2
A1

A2 A2
Figure-23.7

Here we consider that the interaction with proton-1 is grater than proton-2, i.e. A1  A2 .
So the first, the transition splits into two because of the first proton-1, the separation is
A1 . And then each of these transitions further split into two, a total of four transition. The
separation between the transitions is A2
Page-10

Let us predict the E.S.R spectrum of the radical fragment CH  CH 2

It should contain six lines - a large doublet arising from the interaction with the CH
fragment (producing two lines) and a smaller triplet due to the electron interacting more
weakly with the two remote protons of the CH2 fragment (producing three lines).
The resulting diagram is shown in Figure.

In the case where the radical is CH  CH 2  , the resulting pattern is also predicted to
contain six lines, now triplet will have large compare to doublets as shown in Figure 23.8.


CH  CH 2 CH  CH 2 

Figure-23.8
Page-11

Here we will focus some applications of E.S.R

(a) Structural determination of radicals.


Methyl radical  CH 3 has three equivalent 
C  H bonds. This provides four
lines. The splitting between the lines is 25 Gauss.

The structure can be planar or tetrahedral.

For most planar aromatic radicals, the value of A is 25 Gauss.

The theoretical value of the splitting in planar form is ~ 41 Gauss

The theoretical value of the splitting in tetrahedral form is ~ 300 Gauss.

Therefore, the structure of the methyl radical is planar.

(b) Benzoquinone anion radical:


Four identical C  H bonds provide five lines. As discussed before,

1 proton – splits into 2 lines 1:1


2 protons split into 3 lines 1:2:1
3 protons split into 4 lines 1:3:3:1
4 protons split into 5 lines 1:4:6:4:1

.
O
H H

H H
O-
Page-12

(c) E.S.R for solid sample

In crystal, the crystal field can interact strongly with L and splits into 2L 1 sublevels.
Each of these sublevels further split into 2S  1 sublevels when the external magnetic
field is applied.
The magnetic dipole selection rule

mJ  1
In another case, the spin degeneracy is removed in the crystal when weak spin orbit
coupling is present. In this case the spin degeneracy is removed without external
magnetic field.
This is illustrated in two different cases
For Mn++, spin orbit coupling is zero. On applying magnetic field the ground state (6S5/2)
5
splits into (2   1  6) levels. Since the separation between the sublevels is same, all
2
the resonance occur at the same frequency.
5
2
3
2 h 0
Splitting 
1
2
B
1
2
3
2
5
2
Page-13
In case of Cr+++, spin degeneracy is removed by crystal field, and the sublevels are not
equally placed. So the resonances occur at different frequencies.

3
2
1
2
Splitting  h 0

B

1
2
3
2

Further, if the paramagnetic ion has nuclear spin, each sublevel will further split and the
separation between the splitted sublevels provides information of the hyperfine constant.
This study provide information about
(i) Type of ion.
(ii) Symmetry of the crystal.
Page-14

Recap

In this lecture we have learnt, the Electron Paramagnetic Resonance ( EPR ), sometimes
referred to as Electron Spin Resonance ( ESR ) spectroscopic technique which is a widely
accepted tool to determine the structure, conjugations and spin in various research fields.

This technique is used to study paramagnetic centers on various oxide surfaces, which are
frequently encountered in heterogeneous catalysis such as surface defects, inorganic or
organic radicals, metal cations, metal complexes and clusters.

Diamagnetic oxide materials can also be studied using suitable paramagnetic probes,
including nitroxides and transition metal ions.

In biology, spin labeling and conformational studies are carried out in biomolecules with
suitable probe such as nitroxide radicals etc.
Title : Nuclear magnetic resonance
spectroscopy (N.M.R)
Page-0

In this lecture we will describe the static magnetic field interaction with the nuclear spin.

The spectroscopy developed on this concept is known as Nuclear magnetic resonance


spectroscopy (N.M.R).

We will go through the theory of this interaction, its consequence and the applications.

The fine structure seen in high resolution spectrum is very useful to determine the
structure of molecule.

This spectroscopy has wide application in various fields of science.


Page-1

In 1944, Rabi (Columbia) won the Nobel prize "for his resonance method for recording
the magnetic properties of atomic nuclei".

Nuclear magnetic resonance spectroscopy: It is commonly referred to as NMR, and is a


technique which exploits the magnetic properties of certain nuclei to study physical,
chemical, and biological properties of matter. The most interesting isotope 12C has
nuclear spin 0 and does NOT exhibit magnetic resonance. The same goes for 16O and 32S
(the naturally occurring isotopes). Fortunately, the most abundant isotope 1H has I = ½
and so it does exhibit magnetic resonance. And also 13C, 19F, 31P and 15N exhibit nuclear
magnetic resonance. Thus the wide applications of N.M.R. are based on 1H and 13C.
Page-2

When hydrogen (1H) or a proton having I = ½ is introduced into a magnetic field, the
energy level splits into two corresponding to the two projections of its spin mI   1
2
and mI   1 . This way the degeneracy of mI is removed due to the magnetic field.
2

As described in previous lecture, the interaction energy of proton with the magnetic field
is described as

Emag   I mI g I  B I .B0 I mI
  g I  B B0 I mI I z I mI …………………………..(24.1)
  g I  B B0 mI   g N  BN B0 mI

Where g N is the nuclear Lande g-factor,  BN is the nuclear Bohr magneton and B0 is the
applied magnetic field in the Z-direction of the laboratory frame.

It is clear from equation-24.1 that the energy difference corresponding to mI   1 and


2
mI   1 depends on the magnetic field strength. Referring to the following figure-24.1,
2
it is seen that when an electromagnetic field is incident, the magnetic dipole transition
takes place between mI   1 and mI   1 .
2 2

g N  5.5857 mI   1
2
And

mI   1 0
BN  5.05 1027 joule / Tesla 2
B0
0 g N  BN 5.585  5.05 1027
 
B0 h 6.626 1034 mI   1
2
 4.257 107 Absorption
 42.57 MHz / Tesla

For
B0  1.141T ,  0  60 MHz Figure-24.1 B0

And for B0  7.04T ,  0  300 MHz , so for the transition the radio frequency (r.f) is
needed.
Page-3

The block diagram of the experiment setup is shown in figure-24.2


It consist of
1. A source of radio frequency radiation.
2. A receiver coil.
3. A d.c magnetic field.
4. A sweep generator for varying the magnetic field.
5. A data acquisition system.

Sample

Superconducting
r.f input magnet
oscillator

r.f receiver
Superconducting
magnet power
Data supply
acquisition

Figure-24.2
Page-4
For experiment,
 Sample is placed between the poles of a powerful magnet, made of superconductor.
 The radio frequency (r.f) field is generated in the coil connected to the r.f oscillator.
 The detector coil is placed at right angles to both the direction of the magnetic field and the
transmitter coil.
 The magnet is provided with the sweep coils which are used to vary the magnetic field.
In general, the r.f. frequency is kept fixed. The magnetic field is varied until the resonance
condition is reached. The nuclear magnetic moment transition induces an emf in the detector
coil, which is amplified and then recorded as resonance absorption.
 Homogeneity or the magnetic field is achieved by spinning the sample tube with an
appropriate frequency. In this way, all the nuclei experience an average magnetic field.
 The sensitivity is achieved by cooling the sample because,

n 1
2  e g N BN B / kT
n 1
2

Where the electron population in the mI   1 state n1 and that in mI   1 state n 1


2 2 2 2
Page-5

 According to the theory, we should expect that all protons in 1H have the resonance the
same magnetic field.
 However, in actual case, this is not observed.
For example, when experiment is done with ethanol (CH3-CH2-OH), three peaks were
observed with relative intensity 1:2:3 corresponding to one, two and three protons as shown
in figure-24.3.

c
Absorption 3
HC

b
Hb
HC C 2
a
C 1
HC
O
Hb Ha
B0

Figure-24.3

This implies that the actual magnetic fields seen by the protons are different from the applied
magnetic field Bo.
So, if we write the effective field
B  Bo (1   )

Where  = Screening constant (105 for protons)


The value of  depends on the chemical environment around the nucleus.
Page-6
Let us consider, two types of environment of the nucleus and because of that two
resonance lines for nucleus of type A and type B. So,

BB  BA  Bo (1   B )  Bo (1   A )
 Bo ( A   B )
 Bo AB
In terms of resonant frequency
 B  A   o (1   B )  o (1   A )
  o ( A   B )
  o AB
This  AB   A   B is known as chemical shift. It is expressed as
 B  A
 AB   106 ppm ……………………………….(24.2)
o

What influences the chemical shift?


(a) Shielding effects:
When an atom is placed in a magnetic field, its electrons circulate about the
direction of the applied magnetic field. This circulation causes a small magnetic
field at the nucleus which opposes the externally applied magnetic field as shown
in figure 24.4.
This ultimately decreases the effective magnetic field felt by the proton,
shifting the signal to the higher magnetic field. This is called local diamagnetic
shielding.
Induced field
External
magnetic
field

Circulating proton
electrons

Figure-24.4
Page-7

(b) De-shielding effect:


(a) When H atom is bonded with an electronegative atom, this electronegative
atom attracts the electron towards it. Thus the density of the electron cloud
decreases and as a result de-shields the nucleus. In this cause the resonance
occur at a lower magnetic field. The magnitude of the de-shielding rapidly
decreases as the distance between the proton and electronegative atom
increases.
(b) In a paramagnetic substance the electron distribution is anisotropic about the
atom. In this case, the electron circulation about the nucles generates a
magnetic field. This result in de-shielding the nucleus.
(c) In aromatic protons, such as in benzene molecule, the circulating current due
to the delocalized  electrons produces a magnetic dipole. The induced field at the
proton is parallel to the applied field and the result is therefore de-shielding.
The following figure explains the resonance condition for shielding and de-
shielding condition.

Deshielded mI   1
2

Shielded

0 0 0
mI   1
2
B0

mI   1
2

Figure-24.5 Dehielded Shielded B0


Page-8

From the above de-shielding effect discussion we note that, in NMR spectrum
(1) Number of peaks = no. of H environments
(2) Position of the peak depends on the chemical environment.

Area under the peak= relative number of H in the chemical environment

Referring to figure 24.3 for the case of ethanol,

Since, H a  close to oxygen atom  more de-shielding  resonance at lowest


magnetic field. Also only one H contribution to the intensity
H b  Two bonds away  middle peak, two H atoms contribution to the
intensity, and
H c  Far away  less de-shielding  highest magnetic field, three H in the
methyl group, thus intensity is thrice that of intensity of Ha.

In practice, chemical shift is measured with reference to tetramethylsilane.


It is chemically inert so,

 sample  TMS
 sample,TMS  106 ppm …………………………….(24.3)
o

This will come as ‘-‘ve number, so,  scale is introduced,


  10.00   sample,TMS

Tetramethyl silane (TMS) [Si-(CH3)4] is used as reference because it is soluble in most


organic solvents, is inert, volatile, and has 12 equivalent 1H and 4 equivalent 13C. TMS
signal is set to 0.

Si
Page-9

Spin-Spin Coupling :
When N.M.R spectrum is recorded under high resolution, it is observed that ;
1) The peak corresponding to Hb splits into four peaks with intensity pattern
1:3:3:1
2) The peak corresponding to Hc splits into three peaks with intensity pattern
1:2:1

The splitting of the lines arises as a result of the nuclear spin-spin interaction with the
neighboring protons. Such coupling between pairs of nuclear spins is an important feature
of nuclear magnetic resonance (NMR) spectroscopy as it can provide detailed
information about the structure and conformation of molecules.
In the following we will discuss the theory of this spin-spin coupling.
For a proton mI   1 . Let us denote mI   1 as  and mI   1 as , then possible
2 2 2
combinations for the CH2 group and CH3 group are given in table-24.1 and table-24.2
respectively.
Table-24.1
CH2 group  mI degeneracy
 +1 1
,  0 2
 -1 1

Table-24.2
CH3 group  mI degeneracy
 +3/2 1
, ,  +1/2 3
, ,  -1/2 3
 -3/2 1
Page-10

The interaction of the two types of nuclei A and B with the applied magnetic field is
given by
 Zeeman   g N  BN I zA B0 1   A   g N  BN I zB B0 1   B  ……………….(24.4)

Now we will introduce the spin-spin interaction term in the Hamiltonian


 sc  J  A   B
Where J is the spin-spin coupling constant.

Substituting g N  BN B0   0 , we get
   0 1   A  I zA  0 1   B  I zB  J  A   B ……………………….(24.5)
Now,
 A  B  I xA I xB  I yA I yB  I zA I zB
The energies can be obtained by solving the Schrodringer equation
  E
Where,    ci i
i
The possible combinations are,
1    A   B  2    A   B 
3    A   B  4    A   B 

Let us take that the  is normalized, then


E    

We construct the determinant to solve E


11  E 12 13 14
 21  22  E  23  24
0
 31  32  33  E  34
 41  42  43  44  E
…………………………(24.6)
Page-11

I   Ix  i I y
Now
I   Ix  i I y
I   0 I  
We know that,
I    I   0
And,
  
I A I B   I xA I xB  I yA I yB  i I yA I xB  I xA I yB 
I A I B    I xA I xB  I yA I yB   i  I yA I xB  I xA I yB 

I I  I I  I I  I I 
A B A B 1 A B A B
x x y y
2
Substituting in equation-24.5


1 
   0 1   A  I zA  0 1   B  I zB  J  I zA I zB  I A I B   I A I B  
2 
 
The various terms can easily be deduced as
  J
11   0 1   A   0 1   B  
2 2 4
  J
 22  0 1   A   0 1   B  
2 2 4
0 0 J
 33   1   A   1   B  
2 2 4
0 0 J
 44  1   A   1   B  
2 2 4
J
 23   32 
2
All other terms will be zero. So the determinant in equation-24.6 will be

0 0 J
 1   A   1   B   E 0 0 0
2 2 4
0 0 J J
0 1   A   1   B   E 0
2 2 4 2 0
J 0 0 J
0  1   A   1   B   E 0
2 2 2 4
0 0 J
0 0 0 1   A   1   B   E
2 2 4
Page-12

From the determinant we get,


    J
E1   0  1  A  B  
 2 2  4
    J
E4   0 1  A  B  
 2 2  4
J
E1   M 
4
J    
E4  M  where M   0 1  A  B 
4  2 2 
To get the other two roots we have to solve the following determinant

0 J
1   A 
2 2
0 J
 1   B   E
2 4 0
J 0
 1   A 
2 2
0 J
 1   B   E
2 4

We obtain,

J  02  2 J 2
E  
4 4 4
Where,    A   B .

So we have now,

J 1 2 2
E2    0   J 2
4 2
J 1  J2 
    0  1  2 2     
4 2  2  
 0 
J 1 2 2
E3     0   J 2
4 2
J 1  J2 
    0  1  2 2     
4 2  2  
 0 
Page-13
The energy level diagram is shown in figure-24.6

Bare nuclei 0th order 1st order


1
  0 ( A   B )
 2 J 4

4

1
 0 ( A   B ) J
2 
4
 ,  1
3
  0 ( A   B ) J
2  2
4

J

4
1
 1
 0 ( A   B )
J   0
  0, J  0
2
J 0
Figure-24.6
The
transitions are
J 1  J2 
E2  E1    M   0  1  2 2     
2 2  2  
 0 
J 1  J2 
E3  E1    M   0  1  2 2     
2 2  2  
 0 
J 1  J2 
E4  E2   M   0  1  2 2     
2 2  2  
 0 
J 1  J2 
E4  E3   M   0  1  2 2     
2 2  2  
 0 

Figure -24.7 shows the splitting of Bare nuclei


the transitions when J  0 . 12 13
It should be noted that the interaction 24
34
between the two nonequivalent
protons (having different chemical 0th order
shift) gives rise to the splitting of the
transitions. The separation between 12 34 13 24
these transitions depends on the J J
1st order
interaction term J.
Figure-24.7
Page-14
The splitting structure follows the Pascal’s triangle discussed in the previous lecture. In
figure24.8 describes this method.

1 1 I  1 1
2

1 2 1 I  2 1
2

1 3 3 1 I  3 1
2

1 I  4 2
1
1 4 6 4

1 5 10 10 5 1 I  5 1
2

1 6 15 20 15 6 1 I  6 1
2

Figure-24.8

So if we again look at the fine structure of the ethanol given in figure-24.9, the line
marked as b splits into four having intensity ratio 1:3:3:1 because of the CH3 group and
the line marked as c splits into three with intensity 1:2:1 due to the CH2 group.
c
Absorption 3

HC b
2
a
Hb
HC C 1

HC C
O B0
Hb Ha
Jbc Jbc
Figure-24.9
Page-15

Let us take another example.

For a group given in figure-24.10,


Here, Ha, Hb and Hc are nonequivalent protons having different chemical shift. The low
resolution spectrum is shown in this figure. Jab is the coupling constant between Ha and
Hb and Jbc is the coupling constant between Hb and Hc.

Absorption c
Jab Jbc 2
Ha Hb Hc a b
1 1
C C C
Jbc
Hc
B0

Figure-24.10

There are two situations


(a) Jab > Jbc or (b) Jab < Jbc . The fine structure is shown in figure-24.11

c c
a b b
a
Jbc Jbc
Jab Jab Jbc
Jab
Jbc Jab
Jbc

B0 B0
Case-(a) Case-(a)

Figure-24.11
Page-16

There are few points need to be remembered from the above discussions.
1. When two nuclei have the same chemical shift, only one line is observed
even there is spin-spin interaction between them. This is why for the
equivalent protons, no splitting is observed.
2. The major application of NMR spectroscopy is to determine the structure
of the molecule. This requires the determination of the two quantities,
chemical shift and the spin-spin coupling constant.
3. The chemical shift provides the information about the nature of the
chemical group
4. The spin-spin coupling constant and the pattern help us to determine the
arrangement of the atoms
5. The relative intensity of the peaks tells us about the ratio of the equivalent
atoms in different chemical environment
6. There are more to learn about the complex structure and the kinetics and
those are beyond the scope of this course.
Page-17

Recap

In this lecture, we have learnt that due to the shielding and de-shielding of the protons,
there is chemical shift.
This Chemical shift depends on :
• Electronegativity of nearby atoms
• Hybridization of adjacent atoms
• diamagnetic effects
• paramagnetic effects
• solvent effect

The fine structure of these transitions depends on the coupling of the intrinsic angular
momentum of different protons.

Such coupling between pairs of nuclear spins is an important feature of nuclear magnetic
resonance (NMR) spectroscopy as it can provide detailed information about the structure
and conformation of molecules.
Title: Fundamentals of the Quantum Theory of molecule
formation
Page-0

In the last module, we have discussed about the atomic structure and atomic physics to
understand the spectrum of atoms.

However, many atoms can combine to form particular molecules, e.g. Chlorine (Cl) and
Sodium (Na) atoms form NaCl molecules.

Bonding between ions, as in the negative charged chlorine ion and the positively charged
sodium ion, could be understood in the light of coulomb interaction (attraction) between
oppositely charged bodies.

But atoms of the same type can also form bonds, as for example in the case of H 2 .

It remained, however, inexplicable that two similar atoms, which are electrically neutral,
could form a bound state.

In this lecture, we will understand the formation of molecule from atom in the quantum
mechanical framework.
Page-1

It only became possible with the aid of quantum mechanics to attain a fundamental
understanding for the formation of molecule.

Even in the case of ionic bonding, basic new insights have been obtained through
quantum theory.

First, for example, it must be understood why the ions form in the first place, and why the
electron which is transferred from sodium to chlorine thus finds an energetically more
favorable state.

In the following, we will develop some important basic ideas for the quantum theory of
chemical bonding. However, Physics & Chemistry are still far away from a complete
solution to these problems.

To understand chemical bonding, the interactions of several particles must be taken into
account: given n atomic nuclei and m electrons, one would have to find the complete
wavefunction and the corresponding energies of the total system.

It is useful to keep in mind that the nuclear masses are much greater than those of the
electrons. Thus the electronic motions are much faster than nucleus.

Then we may ignore the motion of the nuclei and treat them as fixed.

In atomic physics, we were able to obtain much information from spectroscopic


observations and could direct our attention to both the ground states and the excited
states.

In the study of chemical bonding, the determination of the wavefunction of the ground
state of the particular molecule plays a more important role.
Page-2

The Hydrogen – Molecule Ion H +2

Certainly the simplest case of chemical bonding occurs in the hydrogen molecule
+
ion H . This species is observed as a bound state in gas discharges in a hydrogen
2

atmosphere, in such a gas discharge, the hydrogen molecule loses one electron.

The bonding energy, equivalent to the dissociation energy, has been determined to be
2.65 eV.

We are dealing with two nuclei (a and b in figure-26.1) and one electron. If the
nuclei are far removed from one another, we can imagine that the electron is localized on
one nucleus or other.

The wavefunctions are those of the hydrogen atomic ground state.

So the Hamiltonian for nucleus a ,


-e
 2
e2 
    a  ra   Ea a  ra  ….. (26.1)
2 0
 rb
 2m ra  ra
Ha

And correspondingly for nucleus b , a Rab b


Figure-26.1
H b b  rb   Eb0 b  rb  ….. (26.2)

So, Ea0  Eb0  E 0

If we let the nuclei approach one another, the electron, which was, for example, at first
attached to nucleus a , will experience the attractive Coulomb force of nucleus b .

Conversely, an electron which was at first bound to nucleus b will experience the
attractive Coulomb force of nucleus a . We must therefore set up a Schrodinger Equation
which contains the Coulomb potential for both.

Further in order to calculate the total energy of the system, we must take into account the
Coulomb repulsion between nuclei. The additional energy  e  is not directly related
2

 Rab 
to the energy of the electron, it will only produce a constant shift of all the energy eigen
values. We will introduce it at the end.
Page-3

Referring to the figure 26.1, let us define,

Rab  Ra  Rb ra  r  Ra rb  r  Rb

Thus, Schrödinger equation, then

 2
e2 e2 
   2
    E ..... (26.3)
 2 m ra rb 

Here E and  are energy eigen value and the wavefunction for the whole system
respectively and are yet to be calculated.

We will use the perturbation method. In principle, the electron could be in the
neighborhood of nucleus a or of nucleus b , with the same energy. These two states a
and b are thus degenerate. Now however, the other nucleus, from which the electron is
by chance more distant, acts, as a perturbation to the electronic state. We thus expect that
the degeneracy will be lifted by this perturbation.

In the presence of the degeneracy, we take the total wavefunction is the linear
combination of the hydrogen wavefunctions defined in equation 26.1 and 26.2 and
defined as,

  c1 a  c2 b

Where c1 and c2 are the coefficients. This is generally known as linear combination of the
atomic orbital (LCAO) and  is known as the molecular orbital.

So, putting this, in equation 26.3, and rearranging we get,

 e2   e2 
 a
H   
 1 a  b
c H   c2 b  E  c1 a  c2 b  ….. (26.5)
 rb   ra 

2
e2 2
e2
Where H a   2  and H b   2 
2m ra 2m rb
Page-4

Substituting the values from equations 26.1 and 26.2 in 26.5, we get,

 0 e2   0 e2 
 E   c1 a   E   c2 b  E c1 a  E c2 b
 rb   ra 
….. (26.6)
 e2   e2 
  E 0  E   c1 a   E 0  E   c2 b  0
 E rb   E ra 

While a and b are functions of position, the coefficients are independent of positions.

We assume that the functions a and b are real, as in the case of hydrogen atom
ground state wavefunctions, and the functions a and b are not orthogonal since the
electron is associated with two different nuclei a and b. So,

 a b dV  S and  a a dV   b b dV  1 ….. (26.7)

Now, Equation 26.6 is multiplied by a

 e2   e2 
 a  E 


 1 a
rb 
c dV  c2 a 


E   b dV  0
ra 
e2 e2
 E c1  c1  a  ra  a  ra  dV  c2  E   a b dV  c2  a  ra  b  rb  dV  0
rb ra
  E c1  c1 C    E c2 S  D c2   0
  E  C  c1   E  S  D  c2  0

………….(26.8)

Where,

e2
  a  ra  a  ra  dV  C
rb
e2
  a  ra  b  rb  dV  D
ra
Page-5

Equation 26.6 is multiplied by b and following the same procedure we get,

 E  S  D  c1   E  C  c2  0 …….. (26.9)

So, to get the solution, following determinant from equations 26.8 and 26.9 should
vanish.

The determinant:

 E  C E  S  D   c1 
    0
 E  S  D E  C  c2 

So,

 E  C    E  S  D  0
2 2

 E 2  C 2  2 E C  E 2 S 2  D 2  2 E DS  0
 E 2 1  S 2   2  C  DS  E   C 2  D 2   0

2  C  DS   4  C  DS   4 1  S 2  C 2  D 2 
2

 E 
2 1  S 2 

  C  DS   C 2  D 2 S 2  2 C DS  C 2  D 2  C 2 S 2  D 2 S 2

1  S 
2

  C  DS    D  CS 
2


1  S  2

  C  DS    D  CS 

1  S  2

Two solutions:

C  DS  D  CS CD
i  
1 S 2
1 S
C  DS  D  CS CD
 ii  
1 S 2
1  S …………………………………….(26.10)

CD
So, E  
1 S
Page-6

If we take, solution (i) in equation 26.10 and then,

 C  D
Substituting  E   1  S  in Eqn.(26.8)

 E  C  c1   E  S  D  c2  0
 CD   C  D  S 
   C  c1     D  c2  0
 1 S   1 S 
  C  D  C  CS   CS  DS  D  DS 
  c1    c2  0
 1 S   1 S 
 CS  D   D  CS 
  c1    c2  0
 1 S   1 S 
 CS  D   CS  D 
  c1    c2  0
 1 S   1 S 

So, c1  c2  c and thus,    c a  b 

E 0  E  E
CD
 E  E0   E ……………………………………………(26.11)
1 S
Now if we take, solution (ii) in equation 26.10 and

 C  D
Substituting  E   1  S  in Equation.(26.9)

 E  S  D  c1   E  C  c2  0
 C  D  S   CD 
   D  c1     C  c2  0
 1 S   1 S 
 CS  DS  D  DS    C  D  C  CS 
  c1    c2  0
 1 S   1 S 
 D  CS   D  CS 
  c1    c2  0
 1 S   1 S 

So, c1  c2  c and thus    c a  b 

CD
E  E0   E
1 S
Page-7

Let us understand the meaning of this wavefunctions   and   . Figure-26.2 shows the
plotting of wavefunctions a and b as well as a  b  .

   c a  b 

a b

a b
X
Density distribution of electron

Figure -26.2

The following observations can be made.

(a) Symmetric wavefunction   is formed by a  b . Because of the overlap, the


occupation probability for   between two nuclei is increased.
(b) Thus the density distribution of electron in the   state shown in the lower part of
figure 26.2 increases as the distance between a and b decreases.
Page-8

Antisymmetric wavefunction   is formed from a  b . The occupation probability is


clearly zero in the plane of symmetry.

   c a  b 

a
b

a b X

Density distribution of electron

Figure-26.3

As can be seen from figures-26.2 and 26.3 that the distance between a and b is an
important factor to have finite overlap between the a and b .

Now we will concentrate on the energy values E and E . To do that we have to


understand the meaning of the quantities S, C and D. We will try to evaluate these
quantities through the diagrams.
Page-9

The quantity  a b dV  S

From figure 26.4 it is clear that this quantity S is considerable when Rab is sufficiently
close and S increases when Rab decreases. The quantity S is known as overlap integral.

(i)

a b

a b X

1.2
(ii)
(iii)

1.0

a b
0.8

0.6

0.4

0.2

0.0

X 0 1 2 3 4
a b radial distance

Figure-26.4

Figure-26.4(i) represents the 1s atomic orbital of hydrogen for two nuclei a and b.

Figure-26.4(ii) shows that when the nuclei are close enough the overlap increases.

Figure-26.4(iii) plots the value of overlap integral S with respect to the internuclear
distance Rab. When both the nuclei are coming closer the overlap increases and the value
of S increases. When they fall on each other, the overlap integral becomes 1. However,
this situation does not arise due to the nuclear-nuclear repulsion term. We will visualize
this picture later.
Page-10

e2
C    a  ra  a  ra  dV
rb
e2
   b  rb  b  rb  dV
ra

This term is known as the direct coulomb term. Figure-26.5(i) shows the orbital a2 and
the coulomb attraction potential. As the two nuclei approaches to each other, the overlap
between them increases. Figure-26.5(ii) plots the term C with respect to the internuclear
distance Rab. Here also we see that, this term is appreciable when the both the nuclei are
close enough.

(i)

a2

a b
X

e2

1.0

0.8
(ii)
4 0 rb
0.6

0.4

0.2

0.0
0 1 2 3 4

radial distance

Figure-26.5
Page-11

e2
D    a  ra  b  rb  dV
ra
e2
   a  ra  b  rb  dV
rb

(i)
a b
1.0

0.8
(ii)
0.6
a b
X 0.4

0.2

2
e
 0.0

4 0 rb
0 1 2 3 4

radial distance

Figure-26.6

The term D is known as cross integral term. The meaning of this is that there should be
e2
overlap between a , b and coulomb attraction term (  ) as shown in figure
4 0 rb
26.6(i). The value of this integral is plotted in figure 26.6(ii). It is clear from this figures
that when there is no overlap between a and b , the value of D goes to zero.
Page-12

In figure 26.7(i), the quantity S , C, D are plotted together to have a comparison. It can be
seen that all these quantities depend on the Rab .

1.2 curve 1
1.0
(i) curve 2
1.0
S (ii) CD
curve 3 
1 S
0.5
0.8

0.6
C 0.0

0.4 D
CD

0.5
0.2

0.0 1 S
0 1 2 3 4 1.0
0 1 2 3 4
radial distance
radial distance

Figure-26.7

CD CD
In figure 26.7(ii), the quantities  and  are plotted with respect to the
1 S 1 S
internuclear distance Rab. As can be seen, from this plot that these quantities are
degenerate when the two nuclei are far away from each other. If two nuclei approach one
another the quantities split up in a way depending on whether symmetric or
antisymmetric / bonding or antibonding. In Symmetric case energy reduces whereas in
Antisymmetric case energy increases.
Page-13

Now, if we add nuclear energy, the energy diagram is shown in figure-26.8 symmetric
case there is a minimum, so it forms a bound state. Antibonding is not a stable state.

1.0 Anti-bonding
Equilibrium
internuclear
distance
0.5 Bonding

E0 0.0

0.5
0 1 2 3 4

radial distance

Figure-26.8

Physical meaning: in the bound state the probability density of electron is quite large in
the middle, profit Coulomb attraction from both the nuclei, energy goes down. Whereas
in the antibonding case, only electron experiences attractive force of one nucleus at a
time. The internuclear distance where the energy is having a minimum is known as
equilibrium internuclear distance. Thus, this molecule H2+ - ion is stable with the bonding
energy lower than the individual hydrogen atom.
Page-14

Recap

From this lecture we have developed the basic idea of the quantum theory of chemical
bonding.

To understand chemical bonding, the interactions of several particles should be taken into
account: given n atomic nuclei and m electrons, one would have to find the complete
wavefunction and the corresponding energies of the total system.

It is useful to keep in mind that the nuclear masses are much greater than those of the
electrons. Thus the electronic motions are much faster than nucleus. This is known as
Born-Oppenheimer approximation.

In the study of chemical bonding, the determination of the wavefunction of the ground
state of the particular molecule plays a more important role.

The degenerate atomic energy levels split due to the interaction between the atomic
orbitals. The no. of molecular orbital is same the no. of atomic orbital.

The molecule will be stable only if a minimum is achieved during the calculation of
ground state energy of the molecule.

The minimum arises because the coulomb attractions bring the nuclei closer on the other
hand the nuclear-nuclear repulsion restrict them to come too close. An optimum distance,
the two energies form a minimum for the molecule and thus equilibrium is achieved. The
inter-nuclear distance at equilibrium is the bond length of the molecule.
Title: Understanding of Molecular Orbital

Page-1

In this lecture we will understand how the molecular orbitals are formed from the
interaction of atomic orbitals.

We will see how the electrons occupy these molecular orbitals.

To start, we take up examples of simple diatomic homonuclear molecules and then we


will discuss some heteronuclear molecules.

As the atomic orbitals are important in case of atoms, the molecular orbitals are needed to
understand the structure of molecule.
Page-2
Molecular orbitals are formed from the combinations of atomic orbitals.
Since orbitals are wavefunctions, they can combine either constructively to form bonding
molecular orbitals or destructively to form antibonding molecular orbitals

Why do we need Molecular Orbital theory?


Valence bond theory generally fails to explain the bonding in simple molecules. On the
other hand, molecular orbital theory is better approach for the molecules those are having
extended  systems.
With the help of molecular orbital one can understand the electronic transitions in
molecules.
Page-3
We have seen in the previous lecture that H2+-ion is stable because of the bonding
between the two hydrogen atoms which form a bound state with one electron.
Similar to this, if we consider two hydrogen atoms to form hydrogen molecule, we expect
that the wavefunctions of two hydrogen atoms  (a) and  (b) will interact constructively
or destructively.
If they interact constructively then the molecular wavefunction
1
  1s (a)  1s (b) 
2
This will form bonding orbital. And if they interact destructively, it will form antibonding
orbital
1
  1s (a)  1s (b) 
2
We have also realized that the degenerate levels of the two hydrogen atoms will be
separated as the two nuclei come closer to each other.
As shown in figure 27.1, the ground state of two hydrogen atomic orbitals split into two
energy levels – bonding   and anitibonding   .


1s (a) 1s (b)



Figure-27.1
Page-4
Figure-27.2 explains the energies of bonding and anti-bonding molecular orbitals for first
row diatomic molecules.
Two electrons in H2 occupy bonding molecular orbital, with anti-parallel spins. If
irradiated by UV light, molecule may absorb energy and promote one electron into its
anti-bonding orbital.

The filling of lower molecular orbital indicates that the molecule is stable compared to
two individual atoms.

“+” and “-“ sign indicates the sign of the wavefunction.

1s is symmetric with respect to the center of inversion (middle point of the bond), this

sometimes called as symmetric or gerade and designate as  1s ,g

1s
is not symmetric with respect to the center of inversion (middle point of the bond),

this sometimes called as asymmetric or ungerade and designate as 


1 s ,u

+ -

 1s

1s (a) 1s (b)


1s

+
a b

Figure-27.2
Page-5

Molecular orbital form when atomic orbitals with similar energies and proper symmetry
spatially overlap each other.
For homonuclear molecule such as H2, N2, …e.t.c
s-orbitals combine with s-orbitals and p-orbitals combine with p-orbitals. In figure27.2
we have seen the interaction of s orbital with another s orbital.
For p- orbitals :
There are three px, py and pz orbitals. If we consider X-axis coincides with internuclear
axis then for px orbitals, there will be end-on-end overlap along the bond axis as shown in
figure 27.3. It forms  bonding and nonbonding orbitals.

For  bonding orbital, it is symmetric about the center of inversion and thus it is  2 p ,g

For  antibonding orbital, it is asymmetric about the center of inversion and thus it is

2 p ,u

+ - + -
 2 p

2 p (a)
x
2 p (b)
x


Atomic 2p Atomic
Orbital Orbital
- + -
a b
Molecular
Orbital
Figure-27.3
Page-6
The other px, and py orbitals overlap side-by-side. This is known as -bonding. This
overlap is less than the overlap along the bond axis.
Note that, the overlap is more  it will be more stable and thus energy will be lower.
Thus, the -bonding orbital will be higher in energy than that of the -orbital.

-orbitals are asymmetric with respect to the bond axis as shown in figure-27.4. No
electron density surrounding the bond axis. It has node along the internuclear axis.

 2 p is asymmetric about the center of inversion, and thus  2 p ,u


 2 p is symmetric about the center of inversion, and thus  2 p , g

 
 
 2 p
2 p (a)
y
2 p (b)
y

2p
2 py


Figure-27.4
Page-7
Figure-27.5 shows the total energy level diagram for p-orbitals interaction.
Note that -orbital is the lowest in energy due to maximum overlap.

 2 p

 2 p  2 p
2p 2p
Atomic orbital 2p 2p Atomic orbital

2p
Molecular orbital

Figure-27.5

Note:
Electron can occupy molecular orbital that are lower in energy.
Molecular orbitals may be empty or contain one or two electrons.
For two electrons occupied orbitals, electrons must be spin paired due to Pauli exclusion
principle.
When occupying molecular orbitals, electrons occupy separate orbitals with parallel spin
before pairing.
The highest occupied molecular orbital is known as HOMO.
The lowest unoccupied molecular orbital is known as LUMO.
Page-8.
Let us take an example of oxygen molecule.
Oxygen atom is having electron configuration 1s22s22p4. So the oxygen molecule is
having total of 16 electrons.

1s ,1s  1s   g
2 2 2p

 1s   u

2s 2 , 2s 2   2 s   g  2 p  2 p
  2s   u
2 p4 2 p4
2 p2 , 2 p2   2 p   g 2p 2p
 2 p  u
  2 p   g 2p
  2 p   u
 2s

The figure-27.6 shows that 2s 2 2s 2


for oxygen molecule, there  2s
are two unpaired electrons.
1s
2
1s 1s 2
(a) 1s (b)
Atomic orbital Atomic orbital
Molecular orbital

Figure-27.6

Bond order indicates the strength of a bond and is defined as


Bond order = (no. of electrons in bonding orbitals )/2
– (no. of electrons in the antibonding orbitals )/2
For oxygen, considering only the valence electrons (2p orbital)
Bond order = (8-4)/2 = 2. So the result is a double bond.
Bond order =1 (single bond); Bond order = 2 (double bond) e.t.c
Page-9
For heteronuclear atoms, due to the small energy gap between s and p orbitals, s and p
orbitals overlap is possible. Figure-27.7 shows the overlap between s and p orbital.
This interaction could be constructive or destructive.

s-orbital p-orbital
Figure-27.7

Let us take an example of HF molecule. Figure 27.8 shows the energy level diagram of
the HF molecule. Hydrogen 1s orbital interact with 2s and 2px orbitals of fluorine. 2py
and 2pz orbitals do not have the proper symmetry to interact with 1s orbital. So these are
nonbonding orbital.

3

H1s

F2p

1

2 F2s

1
Figure 27.8
Page-10.
Let us take another example of CO as shown in figure 27.9.
Generally to simplify the things, interactions of the orbitals containg valence electrons
are considered to form the molecular orbitals.
The molecular orbital 2 is the highest filled orbital and is HOMO
The molecular orbital 3 is the lowest unfilled orbital and is LUMO.

4

LUMO
3
C2p

O2p
HOMO 2

C2s 2

1
O2s

1
Figure 27.9
Page-11
Recap

In this lecture we learnt the formation of molecular orbital from the atomic orbitals.
We also came to know about their symmetry such as gerade (g) and ungerade (u)
depending on electronic distribution about the center of inversion.
This arises only for the symmetric molecules that have inversion symmetry.
We have also understood how to distribute electrons into the molecular orbitals.
We have also learnt that the overlap is more, orbital will be more stable and thus energy
will be lower.
Thus, the -bonding orbital will be higher in energy than that of the -orbital.
We have also learnt that for heteronuclear molecule due to the small energy gap between
s and p orbitals, s and p orbital overlap is possible.
Title: Diatomic Molecule : Vibrational and Rotational spectra

Page-0

In this lecture we will understand the molecular vibrational and rotational spectra of
diatomic molecule

We will start with the Hamiltonian for the diatomic molecule that depends on the nuclear
and electronic coordinate.

Then we will use the Born-Oppenheimer approximation, to separate the nuclear and
electronic wavefunctions

We will derive the eigen energy values to understand the rotational and vibrational
spectra of the ground electronic state of diatomic molecules.

At the end we will discuss the rotational and vibrational spectra of some diatomic
molecules.
Page-1

For a diatomic molecule A - B with n electrons as shown in figure-28.1, the


Schrodringer equation can be written as

 2 2 n

 2   2
 V    E
 2
i
2me i 1  ………………………………….(28.1)

Where,  is the total electronic and nuclear wavefunction & E is the total energy.

1st term: K.E. of the relative motion of the nuclei with reduced mass   mA mB
mA  mB

-e
nd
2 term: K.E. of all electrons.
ra rb

And the potential, R


A B
n 2 n 2 n 2 2
e Z e Z e Z Z e
V   A  B  A B Figure-28.1
i  j 1 ri j i 1 ri A i 1 ri B R
 Vee  Ven  Vnn

The molecular wavefunction,  depends on both electron & nuclear coordinates,


    x1 , y1 , z1 , x2 , y2 , z2 ,..... xn , yn , zn , X , Y , Z     r, R 

w.r.t. origin say nucleus A


Since the wavefunction depends on both electron and nucleus coordinates, it is difficult to
solve this problem even for a simple molecule. We would like to separate electronic and
nuclear motion.

Because of the different masses of the electrons and nuclei, we can consider the nuclei to
be stationary to solve the electronic problem (Born-Oppenheimer approximation).
Page-2

The essence of the Born-Oppenheimer approximation is to decompose nuclear and


electronic motions based on the large disparity of the masses of nuclei and mass of
electron.
If we need to solve the Schrodringer equation given in equation-28.1 using Born-
Oppenheimer approximation, we have to follow five steps.

Step-1
Let us assume that the nuclei are clamped in fixed positions. This approximation is
almost close to the reality because the electronic motion is so fast that it will see the
nuclear motions almost stationary. This will eliminate the nuclear kinetic energy term in
the Hamiltonian in equation 28.1

Step-2
Under the assumption in step-1, we can write equation 28.1 as

 2 n 2 
  i  V  r , R    r , R   Ee  R    r , R 
 2m i 1  ……………………………..(28.2)
Here nuclear K.E. is zero, and R appears only as a parameter and   r , R  is the electronic

wavefunction.
We can follow the procedure to fix R  R1

 2 
Solve   i2  V  r , R1    r , R1   Ee  R1    r , R1 
 2m 

Then, fix R  R2

 2 
Solve   i2  V  r , R2    r , R2   Ee  R2    r , R2 
 2m 
and do it for the entire range of R .
Page-3

Step-3
Having obtained the electronic wavefunction   r , R  , and E  R  we can write the total

molecular wavefunction,

  r, R     r, R    R  ………………………………..(28.3)

where,   R   Nuclear wavefunction

Step-4
Now we can substitute equation 28.3 in equation 28.1 and we get,
 2 2 2

     V  r , R   r , R    R 
2

 2
i
2me i 
 2
  2

   2    r , R    R     i2  V  r , R     r , R    R 
 2   2me i 
 E   r, R    R 

Step-5
Since,  i2 operates only on   r , R  . We can write,

 2

   2    r , R    R   Ee  R    R   E   r , R    R 
 2 
Now,  2 is the difference w.r.t. R , so can operate on both   r , R  and   R  .

As a concequence, the electronic wavefunction is relatively insensitive to changes in the


nuclear positions and momenta, and is therefore capable of adjusting itself quasi-
statically to the nuclear motion. This is known as the adiabatic approximation.
Assumption: Electronic wavefunction varies slowly with the inter-nuclear distance,
Page-4
So we get,
2   r, R    R    r, R  2   R 

 2 2 
So,     Ee  R     R   E   R  ……………………………(28.4)
 2 

This is the Schrödinger equation of nuclear motion of diatomic molecule. It is


independent of electronic motion whose effect appears only through Ee  R  , which is

electronic energy as a function of R and act as a potential energy for the motion of the
nuclei.
This is the central approximation of the Born-Oppenheimer approach. Its justification
stems from the fact that nuclear velocities are small compared to electronic velocities.
Now we are ready to solve the equation-28.4. This is simple as we have solved it for
hydrogen atom.

Let us introduce spherical polar coordinate  R,  ,   of one nucleus with respect to other

as origin in equation 28.4, and we get

 2
  2   1     1 2  
 
 R  R  sin   
R sin    2
 Ee  R     R, ,    E   R, ,  
 2 R       sin   
2 2


  R, ,      R  S  ,  
Now, ……………………………(28.5)

Where, S  ,    S JM  ,   are specified by the molecular total angular momentum

quantum number J and Z - component M , analogous to H - atom quantum no. & m.


Page-5
The angular part,
 1  2
 1 2  J  J  1 2
  sin    JM
S  ,    S JM  ,  
2 R 2  sin    sin 2   2  2 R 2
 1   1 2 
2  JM 
 2 sin   2 S  ,    J  J  1 2 S JM  ,  
 sin    sin   
 M 2 S JM  ,    J  J  1 2
S JM  ,  
……(28.6)
This is nothing but the rotational motion as shown in figure-28.2
Putting it in the equation,
 2
  2  1   1 2  
 2 
R  sin   2 
 E  R     R  S JM  ,  
 2 R  R R sin    sin   
2

 2
  2   
  2 
R  J  J  1   E  R     R  S JM  ,  
 2 R  R R   , Rotation
 E , J   R  S JM  ,  

where,   Vibrational quantum number R


Vibration
J  Rotational quantum number
Figure-28.2
  2   J  J  1
2 2

 R    R   E  R    R   E , R   R 
2 R R  R 
2
2 R 2
……….(28.7)
This depends only on inter-nuclear distance R and is called the radial equation for the
nuclear motion.
  R   Vibrational Wavefunction

S JM  ,    Rotational Wavefunction
Page-6
Rotational Spectra of Diatomic Molecule

Simple model is rigid rotor i.e. R is fixed as R e at equilibrium. In classical mechanics,

the magnitude of angular momentum J of such a molecule rotating about center of mass

with angular velocity  .


J   Re  I e 2

From 28.6, rotational energy:


2 R  Re
EJ  J  J  1 putting,
2 R 2  Re2  I e
2
EJ  J  J  1 This is in energy unit joule. We will convert it to cm-1.
2Ie
2
EJ h
F J   1
cm  J  J  1  2
hc 2hcI e 8 c Re2 ……………..(28.8)
 Be J  J  1

Energy Levels as shown in figure 28.3


J 0 ; F  0  0 

J 1 ; F  0   2 Be 
 S JM  PJM  cos   eiM 
J 2 ; F  0   6 Be 
J 3 ; F  0   12 Be 

F(J) = 12Be
J=3

F(J) = 6Be
J=2
F(J) = 2Be
J=1
J=0 F(J) = 0
Rotational energy levels

Figure-28.3
Page-6
Now we will derive the selection rule for the transitions.
We know that we have to calculate the electric dipole transition moment integrals to
derive the selection rule. We take el as dipole moment of the molecule. So the transition
moment integral R is
R   S JM e S J M  d

The components of the dipole moment el

 
e x
 0 sin  cos 

 
e y
 0 sin  sin  d  sin  d d

 
e z
 0 cos 
 2
RZ  0  PJM cos  PJM  sin  d  eiM  e iM  d
0 0


M  M
Spherical Harmonics
 J  M  M  J  M 1 M 
cos  PJM  cos      PJ 1   PJ 1 
 2 J  1   2 J  1 
J  M J  M 1 M M 
RZ  20 
 2J 1
 PJM1 PJM  sin  d 
2J 1  PJ 1 PJ  sin  d 

P PJM2 sin  d  0 if J1  J 2
M
J1

J 1  J  J  1 ; M  0 

J 1  J   1
……….(28.9)
Page-7
Using the selection rule derived in equation 28.9, the transitions are shown in the
following figure-28.4.

F(J) = 12Be
J=3

F(J) = 6Be
J=2
F(J) = 2Be
J=1
J=0 F(J) = 0
Rotational energy levels

Figure-28.4

The transition energies are


J  1 Intensity
F  J  1  Be  J  1 J  2 
F  J   Be J  J  1
  J   F  J  1  F  J 
 Be  J  1  J  2  J  2Be 4Be 6Be 8Be  cm1
 2 Be  J  1 2Be 2Be 2Be 2Be

Figure-28.5
Where J = 0, 1, 2, …..

The expected rotational spectrum is shown in figure-28.5

So we see that under non-rigid approximation, the transition are equidistant with value
2Be and depends on the value of Be.
Page-8
Example:
Rotational Spectrum of HCl molecule shown in figure-28.6

2 Be  21.18 cm1
2
1 h2
Be  10.59 cm  
2 I e hc 8 2  Re2 hc
h

8  cRe2
2

Intensity

21.18 42.36 63.54 84.72  cm1


H  1.0079 amu
Cl  35.453 amu 2Be 2Be 2Be 2Be

1 amu  1.66 1024 gm Figure-28.6


  1.63 1024 gm

6.6256 1027 erg. sec. cm.


Re2 
8  3.142 1.63 1024 gm 10.59  3 1010 cm
sec
16
 1.622 10 cm
o
Re  1.27 108 cm  1.27 A
So the equilibrium bond length of the HCl molecule is 1.27Å.
Thus from the rotational spectrum of a molecule, we can derive the bond length and
structure.
Note: For molecules without permanent dipole moments such as H2 , the electric dipole
pure rotation transition probability is zero. In that system the moment of inertia &
internucler distance can be found out from analysis of rotational structure of electronic
absorption bands or rotational Raman Spectra. We will discuss these two spectra in later.
Page-9

Vibrational Spectra of Diatomic Molecule


Radial equation 28.7,
d  2 d 
2
 R   EROT   R   Ee  R    R   E , J   R 
2 R dR  2
dR 

J  J  1 2
J  J  1 2
Where  EROT
2 R 2 2 Re2

The total energy, E , J  EROT  Evib  Enucl

d  2 d 
2
 R   EROT   R   Ee  R    R    EROT  Evib    R 
2 R dR  dR 
2

d  2 d 
2
 R   Ee  R    R   Evib   R 
2 R dR  dR 
2

1
Putting   R     R
R
2
d2
   R   E  R    R   Evib   R 
2 dR 2
Ee  R  is the electronic energy as shown in the following figure-28.7 which behaves as

the potential energy of the nuclear motion.

Energy

Rab

E(R)
Bonding

Figure-28.7
Page-10

Expand this function in power series about R  Re ,

 dE  1  d 2E 
Ee  R   E  Re      R  Re    2   R  Re   .....
2

 dR  Re 2  dR  R
e

Choose zero of the energy at E  Re   0 minimum because this point is minimum

 dE 
   0.
 dR  Re

1  d 2E  1  d 3E 
So, Ee  R    2   R  Re    3   R  Re   .....
2 3

2  dR  R 3!  dR  R
e e

Energy
Putting R  Re  

1
E  K e  2  ..... Harmonic
2 potential
 d 2E 
Where K e   2 
 dR  Re
Re
Rab

E(R)
Bonding
If neglect the higher power
Figure-28.8
2
d2 1
      Ke  2     Evib    
2 d  2
2
Harmonic oscillator equation with potential as shown in figure-28.8, so the energy value
 1
Evib  E     e   vibrational quantum number  0,1, 2,3,.....
 2
1
1  Ke  2
 e  Classical frequency   
2   
Page-11
Wave Function

 
1
  x2
  N e 2
H x
4 2  ose

h

where, H  
 x  Hermite polynomial  1
The wavefunctions are shown in the figure-28.9 with
 0
dashed curves. The solid curves are the  2
Re
Dipole moment
Internuclear distance
  0  1x
where, 0  Permanent

1  Induced due to vibration 7


e
Selection rule :  3 2
5
Rx  0  *   dx  1  x *   dx e
 2 2
   3
e
     ;0       1  1 2
      1
1
1  0 e
E  0  e 2
2
3
E 1  e
2
 E 1  E  0   e

So only one transition is predicted under


harmonic oscillator approximation Figure- e
28.9). However, many transitions are seen in
Figure-28.9
the infra red vibrational spectra of diatomic
molecule. So harmonic approximation is note the good approximation.
Note: Dipole moment of molecule with equal nucleus is always zero. So for N2 ,O2 ,.....
no infrared spectrum.
Page-12

Anharmonic Oscillator
Introducing anharmonicity

Ee  R   f  r  re   g  r  re   .....
2 3
Condition: g f


Morse Potential as shown in the figure-28.10 : E  R   De 1  e   r re  
2

2 3
 1  1  1
Energy levels: E  e     e xe     e ye     .....
 2  2  2

Convert in cm1
Energy
2 3
 1  1  1
G    e     e xe     e ye   
 2  2  2

where e , e xe , e ye in terms of cm1 . Morse


potential
  Vibrational quantum number
Re
Here e  e xe  e ye Ra
E(R)
Selection rule, Bonding
   1,  2,  3
G(0) Figure-28.10
1 1 1
Zero point energy: G  0   e  e xe  e ye
2 4 8
If the energy referred to the lowest level,
G0    G    G  0 
  1  1
2
 1
2
  1 1 1 
 e     e xe     e ye     .....   e  e xe  e ye  .....
  2  2  2   2 4 8 
 3   3 
  e  e xe  e ye    e xe  e ye  .....  2  e ye  ..... 3
 4   2 
 0  0 x0  0 y0
2 3

3
0  e  e xe  e ye
4
3
where, 0 x0  e xe  e ye  .....
2
0 y0  e ye  .....
Page-13
Transitions between the vibrational levels
 abs  G0    G  0 
 G0  

Thus the observed absorption gives directly the positions of the energy levels.
Example: Observed Vibration Frequency of HCl

 abs (cm -1 ) G  2G Cal. Harmonic Approx.


--------------------------------------------------------------------------------------------------------------
0 0
1 2885.9 2885.9 103.7 2885.9 2885.9
2 5668.0 2782.1 103.2 5668.2 5771.8
3 8346.9 2678.9 102.7 8347.5 8657.7
4 10923.1 2576.2 102.8 10923.6 11543.6
5 13396.5 2473.4 13396.5 14429.5

G0    0  0 x0 2  .....


G0 1  0  0 x0  2885.9
G0  2   20  40 x0  5668.0

20  40 x0  5668.0


20  20 x0  5771.8
( ) (  ) ( )

 20 x0  103.8  0 x0  51.6

G    G0   1  G0  
 0   1  0 x0   1  0  0 x0 2
2

 0  0  0 x0 2  20 x0  0 x0  0  0 x0 2


 0  0 x0  20 x0
 G    G   1  G  
2

 0  0 x0  20 x0   1  0  0 x0  20 x0


 20 x0
which is directly the measured anharmonicity.
Page-14
Recap

In this lecture we have learnt the origin of vibrational and rotational spectra.

We started with the Schrodinger equation for the diatomic molecule.

Then we used the Born-Oppenheimer approximation, to separate the nuclear and


electronic wavefunctions.

We derived the eigen energy values of the rotational and vibrational motions of the
ground electronic state of diatomic molecules.

At the end we have discussed the rotational and vibrational spectra of some diatomic
molecules.
Title: The rotational-vibrational spectra
Page-0
In this lecture, we will find out the rotational-vibrational spectrum of a diatomic
molecule.
The analysis of the rotational spectrum will be carried out using non rigid rotor
approximation.
The rotational structure of the vibrational band will be understood.
The temperature effect of the rotational-vibrational spectrum will also be discussed.
Page-1
In the last lecture we have considered the vibration and rotation separately. We now
consider the actual case in which a diatomic molecule is simultaneously executing both
vibrational and rotational motion.
Classically, the calculated periods of vibration and rotation are of the order of 10-14 and
10-12 sec, respectively. Therefore, a molecule can vibrate one hundred times during one
rotation.
As a first approximation, we separate the two types of motion and calculate the rotational
energy using a mean value of B for the rotational constant that is averaged over the
change in the bond length that occur during much more rapid vibrational motions of the
nuclei.
 1 
2
B   in cm . ……………………………(29.1)
-1

2 hc  R 2 

 1 
Where  2  is the mean value for the  the vibrational state.
 R 

So rotational energy F ( J )  B J  J  1 in cm-1.

Now the radial equation becomes


d  2 d 
2 …………….(29.2)
  R   E   R   E  R    R   E   R 
2 R 2 dR  dR 
ROT

We already
2
d  2 d 
 R   E  R    R   Evib   R  ……………………..(29.3)
2 R dR 
2
dR 
So E  Evib  EROT

And the term value T , J  G0    F  J  in cm-1.


Page-2
The energies for the rotational levels are derived by assuming that there is a rigid fixed
bond between the nuclei. This is known as RIGID ROTOR approximation.
This is not true for an actual molecule undergoing anharmonic oscillations and the
following two modifications must be introduced.
Modification-1:
Average internuclear distance increases slightly with  for an anharmonic oscillator.
 1 
Therefore the mean value of  2  correspondingly decreases.
 R 


So in this case B  Be   e   1
2  where Be is the rotational constant at the

equilibrium internuclear distance Re and e is the small constant and depends on the
shape of the vibrational potential curve.
Modification-2:
In any given vibrational state  the expression for the rotational terms F(J) must include a
correction for the centrifugal distortion of the elastic bond between the nuclei.
Therefore a better model representing as shown in figure-29.1 the rotation is given by
“NON RIGID ROTOR”.

Figure-29.1

Non rigid rotor model can be visualize as two spheres are connected by a spring.
Page-3
Classically, increased rotational energy corresponds to an increased rate of rotation of the
molecule. The centrifugal force which is proportional to the square of the angular
momentum Fe   Re 2 expands the bond against its restoring force, and the work
consumed in this process must be subtracted from the energy calculated in the rigid rotor
approximation.

F ( J )  B J  J  1  D J 2  J  1  ...
2
……………….(29.4)

The constant D is usually smaller that B by a factor of 10-4 or more.


The infra red rotational spectrum
  F ( J  1)  F ( J )
 2 B ( J  1)  4 D ( J  1)3

Rigid rotor
levels

Non rigid rotor levels

Figure-29.2
Page-4
Vibrating Rotor term value
T , J  G0    F  J 
 0  0 x0 2  0 y0 3  B J  J  1
The energy level diagram of these two vibrational levels are shown in figure-29.3
Now let us look at the transitions of rotational lines between the two vibrational levels
 ' and  " . This is known as rotational structure of vibrational transition or rotation-
vibration spectrum.
T ", J "  G0  "  B " J "  J " 1
T ', J '  G0  '  B ' J '  J ' 1

J'5
J'4
J'3
J'2
J ' 1
' J'0

J"5
J" 4
J"3
J" 2
J " 1
" J"0

Vibration-rotation levels

Figure-29.3
Page-5
  1,  2,  3,.......
The transition selection rule is
J  1
  T ' J '  T " J "  G0  '  G0  "  B ' J '  J ' 1  B " J "  J " 1
  ' "  B ' J '  J ' 1  B " J "  J " 1

Let,
J '  J " 1
  J  1
J '  J " 1
When J '  J " 1 , this is known as R branch
For J '  J " 1,
 R   ' "  B '  J " 1 J " 2   B " J "  J " 1

  
  ' "  B ' J "2  3J " 2  B " J "2  J " 
  ' "  2 B '  J "  3B '  B "   J "2  B '  B " 
where J "  0,1, 2,3,.....
……………………(29.5)
When J '  J " 1 , this is known as P branch
For J '  J " 1,
 P   ' "  B '  J " 1 J " B " J "  J " 1

  
  ' "  B ' J "2  J "  B " J "2  J "  ………………………………(29.6)

  ' "  J "  B '  B "   J "2  B '  B " 


where J "  0,1, 2,3,.....

In some cases J '  J " is allowed and it is known as Q-branch.

Take rotationally resolved IR spectrum. Fitting it we get

 1 
B ' & B " ; B '  Be   e  '  
 2  2
 Be 
2
 1  2 R 2 hc
B '  B "  Be   e  " 
2 R 2 hc  2  
Page-6

If B '  B "  B then

 R   0  2 B  2 BJ
 P   0  2 BJ
Vibration-rotation levels
Where  0 is  '   " transition and is
J'5
known as vibrational origin transition.
J'4
The rotational structure of the J'3
J'2
J ' 1
vibrational transition is shown in
' J'0
figure-29.4
If B '  B "
Then we can write a single formula for J"5
the R and P branch transitions. J" 4
J"3
J" 2
J " 1
 m   0  ( B '  B " )m  ( B '  B " )m2
" 
J"0

…………..(29.7)
0
Where m = 1, 2, 3,… give the R P-branch R-branch

branch and Figure-29.4


m = -1, -2, -3, ….give P branch

m = 0 is the missing line

Generally, B ' & B " do not differ much, so the rotational vibrational spectrum consists
of equally spaced lines.
Page-7
Intensities of rotation-vibration spectra:

The intensity of spectral lines depends not only on the transition probability and
frequency of the light but also on the number of the molecules in the initial state.
For the prediction of the intensity, knowledge of the numbers of molecules in the initial
state is necessary, in addition to the transition probabilities.
Since almost all infrared spectra are observed under conditions of thermal equilibrium we
need to consider the distribution of the molecules over the different quantum states in
thermal equilibrium.
Vibration:
Thermal distribution of quantum states.
According to the Maxwell Boltzman Distribution law, the number of molecules d N E
E
between E & E  dE is proportional to e kT dE

G0  hc
The quantities e kT give the relative number of molecules in the different
vibrational levels referred to the number of molecules in the lowest vibrational level. If
the total number of molecule is N then state sum (Boltzman factors over all states)
partition function is given by,
G0 1hc G0  2 hc
Q  1  e kT e kT

Therefore, no. of molecules in state ,

N G0  hc kT
N  e
Q
Page-8
Since
N
N  Q  N  KQ K
Q
if G0 1 is very large,
G0 1hc
No of molecule
then e kT  0

and for Q  1,
G0  h c
N  N e kT

The figure-29.5 gives the relative 0 1 2 3 4


distribution of the no. of molecules in a Energy (E)
Figure-29.5
given vibrational state.

Rotation:
The thermal distribution of the rotational level is not only given by the Boltzman factor
E
e kT dE
, we have to consider the degeneracy of the rotational levels. Each rotational
level J is consisting of (2J+1) of m degenerate levels.

And thus the no. of molecules of a J level is given by


 F  J hc
N J   2 J  1 e kT

Substituting the value of F(J), we get

 B J  J 1hc
N J  K  2 J  1 e kT ………………………..(29.8)
Page-9
To get the rotational quantum number for which the no. of molecule will be maximum,
we have to differentiate equation 29.8 with respect to J, and we get
d NJ
0
dJ J max
 B J  J 1hc
d NJ h c  B J  J 1hc kT
 K 2e kT  K  2 J  1  2 B J  B  e 0
dJ J max
kT

Solving this equation,


 B J  J 1hc
Bh c  B J max  J max 1hc kT
 K  2 J  1
2
02K e kT e
kT
Bh c 2kT
  2 J max  1   2 J max  1 
2 2
2
kT Bh c
2kT 2kT
 2 J max  1   2 J max  1
Bh c Bh c
kT 1 T 1
 J max    0.5896 
2 Bh c 2 B 2
The rotational state sum
2 Bhc 6 Bhc NJ
Qr  1  3e kT  5e kT  .....

 B J  J 1hc
N
NJ   2 J  1 e kT
Qr

For large T & small B,


  Bhc J  J 1
Qr    2J  1 e kT dJ
0 2 4 6 J
0

 B J  J 1h c Figure-29.6
h cB kT  kT
NJ  N  2 J  1 e
kT h cB
 B " J " J "1h c
Cabs 
I abs   2 J " 1 e kT
Qr
where, Cabs  Constant
Page-10
Intensity of the P and R branch

Let us consider B '  B "  B . Then,

 R   ' "  2 B '  J "  3B '  B "   J "2  B '  B " 
  R   ' "  2 B  2 B J "
And,
R P
 P   ' "  J "  B '  B "   J "2  B '  B " 
 P   ' "  2 B J " T = 1000 K

Putting the Jmax,


R J max
  ' "  2 B  2 B J max
P   ' "  2 B J max T = 3000 K
J max

T = 10000 K

Figure 29.7

1 
 R  P  2 B  4 B J max  4 B   J max 
J max
2 
1 kT 1 8BkT
 4B      2.3583 BT
2 2 Bh c 2  hc

So the separation of the P branch and R branch maximum will depend on the value of T
and B. For a given molecule, if T increases then the separation of P and R branch maxima
will increase. The figure shows the intensity pattern of the rotation-vibration spectrum for
different temperatures.
Page-11
Recap

In this lecture we have learnt the non rigid approximation of the rotational energy levels.
Due to the different vibrational frequency the rotational levels get modified.
The intensity of the rotational structure of the vibrational bands depend on the
temperature of the system
A close look at the P and R brach maximum in the rotational structure of the vibrational
bands provide information about the temperature of the system
Title : Raman Scattering :
Page-1
In previous lectures we have learnt that the Infrared spectroscopy is used to record the
absorption spectrum of vibrational and rotational states of diatomic molecules.
The origin of this absorption is the change of the dipole moment due to the vibrational
and rotational motion.
However, homonuclear diatomic molecules neither posses the permanent dipole moment
nor induce dipole moment due to the vibration. Thus these levels are inactive or
forbidden in case of Infra red absorption.
Raman spectroscopy is a tool to overcome this problem. Here we will learn the origin of
the Raman spectrum and its applications.
Page-2
Raman Effect
This inelastic scattering by molecules was first discovered by Sir C. V. Raman in 1928.
He was awarded Nobel Prize for this fundamental work in 1930.
It was predicted by Adolf Smekal in 1923 but named after one of its discoverers in 1928,
the Indian scientist Sir Chandrasekhara Venkata Raman

Raman scattering is the inelastic scattering of a photon – change in photon energy


By nature weak effect (approximately 1 in 107 photons)

Sir Chandrasekhara Venkata Raman

When strong light of 19436 cm-1 is scattered by a material consisting of molecules

containing carbonyl group, its spectrum exhibits two rather weak lines at 17786 cm-1

and 21086 cm-1 apart from the line at 19436 cm-1.


Page-3
Nature of Raman Effect:
When a parallel beam of monochromatic light goes through a gas or liquid or
transparent solid, a part of light is scattered in all directions.
The intensity of scattered light is inversely proportional to the fourth power of
wavelength.
It is found that the scattered light contains exactly the same wavelength as the incident
light. This scattering is called as Rayleigh scattering.
Apart from this wavelength, the scattered light also contains some weak additional lines.
As said earlier that this phenomenon was first discovered by Raman and his collaborators
and is known as Raman Effect.
A comparison of the wave numbers of these additional lines shows that these Raman
lines are independent of the wavelength of the incident light but depends on the nature of
scattering substance.
For different scattering substance, the displacements of the Raman lines from the incident
light are different. Thus, the displacements are characteristics of the scattering substance.
Page-4
Classical theory of Raman Effect:
As discussed in module 1 that when an atom or molecule is brought into an electric field
 
E , an electric dipole moment  is induced in the system. The magnitude of this induced
dipole moment is proportional to the electric field,

   E ………………(30.1)

where  is known as the polarizability.


Except for the case of spherical symmetry, the magnitude of the induced dipole moment
depends on the direction of the electric field.
For example, in case of a diatomic molecule, the induce dipole moment will be higher in
magnitude when the electric field direction is along the inter-nuclear axis than that of the
perpendicular to the inter-nuclear axis.
 
In general, the direction of  does not coincide with the direction of E .So for a
non isotropic molecule, we can write,
 x   xx E x   xy E y   xz E z

 y   yx E x   yy E y   yz E z

 z   zx E x   zy E y   zz E z

E 2  Ex  E y  Ez
2 2 2
Where

 ij is known as polarizability tensor and this tensor is symmetric i.e.  ij   ji .


Using principle axis transformation, we can write,

 x   xx Ex

 y   yy Ey

 z   zz Ez
x, y and z are principle axes
Page-5
If we draw lines in any direction from the origin of the principle axes coordinate system,

of length proportional to 1 , then the locus of the points of lines will form a surface

known as polarizability ellipsoid whose principle axis are x, y and z.
The equation of the ellipsoid as shown in figure-30.1 is given by,
 xx x2   yy y2   zz z 2  1
z

y

x
Figure-30.1

For diatomic molecule


 xx   zz

If a light of frequency  0 is incident then E  E0 sin 2 0t , then the dipole moment will
be from equation 30.1,

   E0 sin 2 0t ………………………………..(30.2)

So the dipole will also oscillate with frequency 0 . According to classical electromagnetic
theory, when a dipole oscillates, it radiates with the same frequency of oscillation. This is
nothing but the Rayleigh scattering as discussed before.
Now, if the molecule vibrates i.e., the internuclear distance changes. And also this
polarizabitity depends on the orientation of the molecule. So the rotation of the molecule
also changes the polarizability.
Page-6
Now we can write the polarizibility for vibration
  0  1 sin 2 t
Where  0 the polariziblity in the equilibrium position is,  is the vibrational frequency

and 1 is the proportionality constant (1  0 ) .


Similarly, for rotation we can write,
  0r  1r sin 2 2 r t
Here the frequency with which the polarizability changes during rotation is twice the
rotational frequency (figure-30.2). The reason for this is that, the polarizabality is same
for that of the opposite field.

Now, Figure 30.2


v   E

 ( 0  1 sin 2  t ) E0 sin 2 0t

  0 E0 sin 2 0t  1 E0 sin 2  t sin 2 0t

1
  0 E0 sin 2 0t  E0 [cos 2 ( 0   )t  cos 2 ( 0   )t ] ……………(30.3)
2
Similarly for rotation
1r
r   0 r E0 sin 2 0t  E0 [cos 2 ( 0  2 r )t  cos 2 ( 0  2 r )t ] ……….(30.4)
2
Thus we see from equation-30.3 that the induced dipole moment not only oscillates with
frequency  0 but also with other displaced frequency ( 0  ) and ( 0  ) in case of
vibration.
For rotation, the displaced frequencies are ( 0  2 r ) and ( 0  2 r ) .
Page-7
So the scattered light contains displaced lines on both the sides of undisplaced line and
the separations in both cases are same.
Frequency on lower side is known as Stokes lines and on higher side is known as anti
stocks lines as shown in figure-30.3.

Stokes
Anti-Stokes

 0  v 0  0  v
Figure-30.3

The experimental Raman spectrum shows that the intensities of the Stokes lines are
higher than that of the anti-Stokes lines.
Classical explanation does not provide information about intensity for both the lines and
thus inadequate for explaining the experimental spectrum.
Page-8
Quantum theory of Raman Effect:

Let us consider for a system, E0 is the ground electronic state and   =0, 1,…… are the

vibrational states of the ground electronic state. If the light frequency  0 incident on this
system, three cases may arises.

Stokes 0
Anti-Stokes

2 f

1
" i
  0 

Stokes Anti-Stokes

Figure-30.4

Case-I:
Molecules absorb the light of frequency h 0 and go to the vibrational state as shown by
dashed line in the figure. The vibratinal state is created by the light and molecular
interaction and exists as long as the light exists. It is not the eigen state of the system but
is a linear combination of all the eigen states of the molecule. Thus by definition, the
lifetime of the virtual state is very very small. Now the molecules will back to the ground
vibrational state (   0) and will emit the same frequency 0 . This is the same Rayleigh
scattering.
Page-9
Case II
Molecules are transferred to the vibrational state by light  0 and these excited molecules

may come back to the higher vibrational state (   1) . In this case the emitted frequency

is 0  . According to the energy conservation, the energy will be lost from the incident

photon energy h 0 to excite the vibrational frequency of the molecule and thus, the
emitted photon energy
h stokes  h 0  h  h(  )

Case III
In thermal equilibrium, the excited vibrational levels are also populated and molecules
there can also absorb light and go to the virtual state. While coming back to the same
vibrational state, this will emit frequency 0 . But if molecules come back to ground

vibrational state (  0) then the emitted frequency will be ( 0  ) . Again, according to
the energy conservation, the photon energy adds up the vibrational energy and emitted

energy will be
h antistokes  h 0  h
Page-10

A typical experimental set up for Raman spectroscopy is shown in figure 30.5

Sample
Laser light

Collection
Laser cutoff lens
filter

Collimating
lens

Monochromator
with detector

Stokes
Anti-Stokes

Figure-30.5

A typical Raman spectrum of CCl4 is shown in figure 30.5


Page-11

Now we will discuss why the Stokes lines are higher in intensity than the anti-Stokes
lines.
As mentioned earlier, the population to these vibrational states depends on the Maxwell
Boltzmann distribution

N G0  hc kT
N  e No of molecule
Q
The figure-30.6 gives the relative
distribution of the no. of molecules in
a given vibrational state.

0 1 2 3 4
Energy (E)
Figure-30.6

The origin of anti-Stokes lines is from the higher vibrational levels and the population is
lower in those states than the ground vibrational level.
As discussed earlier lecture that intensity of transitions lines not only depend on the
transition probability, but also depend on the population of the initial state.
This is the reason for the lower intensity of anti-Stokes lines than the Stokes lines.
For vibrational Raman transitions, the selection rule is   1 . But the parity selection
rule is that it connects the same parity states. This can be understood by considering the
Raman process as two-photon process. Referring to figure-30.4, the two-photon transition
can written as
f er m m er i

Where i , m and f are the initial, middle and final states respectively. Since middle

state m is not an eigen state of the system (virtual state) we have to define as sum of all

eigen states. Thus the transition probability is defined as


R   f er m m er i
m
Page-12
Since the electric dipole operator er connects only opposite parity states, it is clear from
the above equation that for a nonzero transition probability i and f will be having the

same parity whereas m has to be opposite parity states.

So the parity selection rules for which the Raman transitions are allowed

Symmetric  Symmetric
Antisymmetric  Antisymmetric

On the other hand, for Infra red spectroscopy as discussed in the previous lecture that the
transition probability

R  f er i

So i and f will be having the opposite parity states. So IR spectroscopy is the

complementary of the Raman spectroscopy.

For symmetric molecules such as H2, N2 all the vibrational states are symmetric. Thus it
is forbidden in IR transition but allowed only in Raman transition. Hence for this kind of
symmetric molecule, Raman spectroscopy is the only tool to get the information of the
vibrational levels of the ground electronic state.
Page-13
For diatomic molecules with identical nuclei such as H2, N2, for an exchange of the
nuclei, the rotational wavefunctions will either remain unchanged or changes its sign. In
the electronic ground state, either positive sign rotational levels are symmetric and
negative sign levels are antisymmetric or positive sign rotational levels are antisymmetric
and negative sign levels are symmetric. In both the cases, Raman selection rule is
J  0,  2 .
For the rotational Raman spectra
J  0 gives the undisplaced Rayleigh line

J  0,  2 gives the displaced Raman


lines.
The transition energy in the spectrum is
given by
  F ( J  2)  F ( J ) Stokes
Anti-Stokes
 B( J  2)( J  3)  BJ ( J  1)
 6 B  4 BJ
J"5
So in the Rotational Raman spectrum,
J" 4
the first undisplaced lines are at 6B and J"3
J" 2
J " 1
the separation between the other Stokes
" J"0
and anti-Stokes lines are 4B. 4B 6B 6B 4B

Stokes Anti-Stokes

Figure-30.7
Page-14
For the vibrational Raman spectra, let us take an example of carbon dioxide CO2.
The molecule Carbon dioxide is a triatomic molecule.
It has 3 modes of vibrations as shown in figure-30.8. They are (i) symmetric stretch, (ii)
asymmetric stretch and (iii) bending.

Symmetric stretch Asymmetric stretch

Bending
Figure-30.8

Note: For linear molecules the no. of vibrational modes is (3no. of atoms – 5) and others
are having the vibrational modes ((3no. of atoms – 6).

Symmetric stretch

Polarizability ellipsoid

Figure-30.9

The symmetric stretch vibration  1 of CO2 is shown in figure-30.9. During this vibration,
the dipole moments of C=O bonds are equal and opposite as shown by arrow, and thus
no induced dipole moment for 1 vibration. So, this  1 vibration will be inactive in IR
spectrum. But the polarizability tensor changes due to the vibration. So 1 will be active in
Raman spectrum. This 1 vibration is observed in the Raman spectrum at 1335 cm-1 from
the excitation laser line.
Page-15

The second vibration 2 is asymmetric stretch. For this vibration the net dipole moment
changes as shown by arrow in figure 3.10. So 2 will be active in IR spectrum. However,
the polarizability ellipsoid does not change due to this vibrational motion. So it will be
Raman inactive. This 2 is seen in IR spectrum at 667 cm-1.

Asymmetric stretch

Polarizability ellipsoid

+ - + -
No dipole moment
Dipole moment

Figure-30.10
Page-16
The third vibration 3 is bending. For this bending vibration the net dipole moment
changes as shown by arrow in figure 3.11. So  3 will be active in IR spectrum. However,
the polarizability ellipsoid does not change due to this vibrational motion. So it will be
Raman inactive. This 3 is seen in IR spectrum at 2350 cm-1.

Bending

Polarizability ellipsoid

No dipole moment

Dipole moment

Figure-30.11
Page-17
Let us take another example of Sulphur dioxide SO2. Figure-3012 shows the three normal
modes of vibrations. As can be seen from the figure that for all three modes the
polarizability ellipsoid changes. Thus all three modes will be Raman active. Not only
this, the dipole moment also changes due to these three vibrations. Thus all these mode
will be IR active. For SO2, 1 = 1150 cm-1, 2 = 520 cm-1, 3 = 1360 cm-1.

Symmetrical stretching 1

Bending 2

Asymmetrical stretching 3

Figure-30.12
Page-18

Note:

Comparison between Infra red and Raman spectroscopy:

For a given bond, the energy shifts observed in a Raman experiment should be identical
to the energies of its infrared absorption bands, provided that the vibrational modes
involved are active toward both infrared absorption and Raman scattering.
Infrared absorption requires that a vibrational mode of the molecule have a change in
dipole moment or charge distribution associated with it.
On the other hand, scattering involves a momentary distortion of the electrons distributed
around a bond in a molecule, followed by reemission of the radiation as the bond returns
to its normal state.
In its distorted form, the molecule is temporarily polarized; that is, it develops
momentarily an induced dipole that immediately disappears upon relaxation and
reemission.
Page-19

Recap

1. We have described spontaneous Raman effect that originates from the modulation

of the polarizability by the molecular vibrations.

2. We have shown the inadequacy of the classical picture to account for the Stokes

and anti-Stokes Raman line intensities.

3. In terms of the quantum mechanical picture, it has been described qualitatively as a

two- photon process.

4. Cause of the experimentally observed unequal intensities of the Stokes and anti-

Stokes Raman line has been traced to the difference in the population densities of

the excited and ground vibrational states of the molecule at thermal equilibrium.

5. \It has been shown that it provides access to those vibrational states of the molecule

which cannot be addressed in IR absorption spectroscopy and hence complements

the later.

You might also like