You are on page 1of 76

Unit – 7

Inorganic Chemistry - I

To join Test Series


Download ‘Global Online’ app

Or

@ paid WhatsApp group 8179138413


UNIT-7

1. CHEMICAL BONDING IN DIATOMICS

The idea of chemical bonding between two atoms based on sharing of electrons was given by
G.N. Lewis in 1916. The Lewis picture of bonding is still relevant today but the quantum mechanical
formulation, incorporating de-Broglie's wave particle duality, Pauli's exclusion principle and Heisenberg's
uncertainty principle, gave a deeper insight into the nature of the chemical bond. Broadly, there are two
types of bonds between atoms :
1. The ionic bond in which electrons are transferred between atoms and the bond arises
from the coulombic interactions between the resulting ions.
2. The covalent bond in which electron pairs are shared more or less equally between
neighbouring atoms. The idea of dative bond or covalent-coordinate in which one partner
donates both the electrons of the shared pair, symbolized by B  A, is now obsolete and
the term 'covalent bond' signifies that a bond has been formed by electron-sharing
irrespective of the donor of the electron pair.
Covalent bonds are also classified as single, double or triple. A single bond consists of one shared
pair of electrons; a double bond consists of two shared pairs of electrons and a triple bond consists of
three shared pairs of electrons. Quadruple bonds containing four shared pairs of electronsare rarely
encountered though they occur quite frequently in the transition metal clusters.
Modern theories of chemical bonding have largely dispensed with the distinction between the
various types of bonds. They consider bonding as a consequence of the lowering of energy that occurs
when electrons occupy molecular orbitals that spread overall the atoms in the molecule. The bonding
influence of an electron pair, according to MOT, may be distributed over many atoms. In fact, an electron-
pair is not essential to bond formation but merely represents the maximum number of electrons that can
occupy a given orbital thereby contributing to bonding.
Wave Mechanical Treatment of Covalent Bond
The Lewis concept of covalent bond based on sharing of electrons satisfactorily explains the
bonding between atoms all of which have a tendency to gain electrons in order to acquire the stable noble
gas configurations.
It cannot answer a number of questions such as
1. Why a covalent bond is formed at all ?
2. How do the electrons distribute themselves in space around the central atom in a
molecule ?
3. What are the forces of attractive interaction in a covalent molecule ?
Wave mechanics has provided satisfactory answers to most of these questions and has led
ultimately to the development of more comprehensive theories of covalent bonding. These theories are :
1. The Valence Bond Theory
2. The Molecular Orbital Theory
But, before we discuss the two theories, it would be worthwhile to elucidate the wave mechanical
principles which are involved in their treatment. These principles are :
1. à   E , then c11 is also the solution of the same
If  1 is a solution of the wave equation H
wave equation where c1 is a constant known an coefficient of the wave function. This means
that 1 and c11 give the same value of energy E.
2. à   E , then c11, c22
If  1 and 2 are separately the solutions of the wave equation H
and c11 ± c22 are also the solutions of the same wave equation. This means that 1,
2, c11, c22 and c11 ± c22 would give the same value of energy E. The squares of
the coefficients, viz, c 2 and c 2 give the contributions of  and  , respectively, to the total
1 2 1 2
wave function.
3. Several methods have been devised for finding out approximate  closest to the true  and
the corresponding approximate energy closest to the true energy of multi-electron andmulti-
nucleus system. One such method is based on the variation principle.
The Variation Principle
à   E . Multiplying each side of the wave equation by  and
Consider the wave equation H

integrating it throughout the space, we get 


à  d 
 

H E d
 

...(1)


E 2 d


( E is constant) ...(2)

where, d is a small space element.


From Eq. (2)
 
E à  d /
H  d
  

..(3)

Thus, energy E of the system can be calculated if we know true . If we do not true , even then
we can guess a wave function which is close to the true wave function  by using our chemical intuition.
If this approximate wave function is 1, then eq. (3) takes the form
 à  d /  2 d
H ...(4)


  1 1  1

where , known as energy function, will have the dimensions of energy but will not be equal
to the true energy of the system.
According to the variation principle, if we are able to guess at a number of wave functions 1,
2, 3, etc., which are thought to be close approximations of the true wave function , then we can
calculate the energy functions, 1, 2, 3, etc., with the help of eq. (4). In such a case
1. All the energy functions calculated as above would have higher values than the true
energy of the system.
2. Out of 1, 2, 3, etc., the energy function with the minimum value would be closest to the
true energy E.
3. The wave function which gives the minimum value of energy function is the closest
approximation of the true wave function .
The approximate  can also be calculated with the help of Ritz linear combination method
which too is based on the variation principle. Suppose we have reasons to believe that the wave function
 has some characteristics of a known wave function 1 and some characteristics of another known wave
function 2. The wave function  can then be written as a linear combination of 1 and 2, i.e.,as  =
c11 + c22.
Accordingly.
Energy function, 
  
 à  d /
H  2 d

 

...(5)
 à (c   c  ) d /
(c   c  )H (c   c  )2 d ...(6)


 


 1 1 2 2 1 1 22  1 1 22

The coefficient c1 and c2 are so adjusted in the total wave function  as to yield a minimum value
of . This is done by differentiating Eq. (6) with respect of c1 and also with respect to c2 and then putting
/c1 = 0 and also /c2 = 0, as these are the conditions for minimization of . By doing so we get
two equations containing , c1 and c2. From these, we can derive a quadratic equation in . The solutions
of this quadratic equation gives , c1 and c2. Similarly, if we have reasons to believe that the true wave
function  has characteristics of known wave functions 1, 2, 3, etc., the actual  can then be written as
 = c11 + c22 + c33 +... The coefficients c1, c2, c3, etc., are so adjusted as to yield minimum valueof the
energy function.
This method has an advantage that if we guess at a wrong wave function (i), its coefficient (ci)
would automatically come out to be zero.
With the help of the wave mechanical principles elucidated above, it is possible to discuss at length
the valence bond and the molecular orbital theories of covalent bonding.
2. VALENCE BOND THEORY

The valence bond theory wave first put forward by W.Heitler and F.London and extended further
by J.C.Slater and L.Pauling. The theory utilises the following wave mechanical principles.
1. If A and B are the wave functions for any two completely independent (non-interacting)
system A and B, then the total wave function  for these independent system, taken
together, would be given by :
 = A B ...(7)
and the total energy E of these non-interacting system and the total Hamiltonian can be
given as
à H
E = EA + EB and H àA H
àB

2. If the wave function for a many-electron system has characteristics of several wave
functions 1, 2, 3, ..., n, then according to the Ritz principle of linear combination, the
wave function  closest to the true wave function of the system is given by
 = c11 + c22 + c33 + cnn ...(8)
where, c1, c2, c3, ..., cn are the various coefficients. As already mentioned, the square of the
coefficient of a particular wave function gives the contribution of that wave function, generally called the
weightage of that wave function, to the total wave function. These coefficients are so adjusted as to give
a state of lowest energy, i.e., a state of maximum stability. The function  should be normalised so that
c12 + c22 + c32 + ... = 1 ...(9)
Valence bond theory can be elucidated by taking into consideration the formation of H2 molecule.
Formation of Hydrogen Molecule
Consider two hydrogen atoms lying far apart from each other so that no interaction between them
is possible. Let A and B represent the nuclei and e 1 and e2 the two electrons. The two atoms of hydrogen
may thus be represented as HA(e1) and HB(e2).
Let the wave functions for the electron (1) and (2) in the separated atoms be represented by A(1)
and B(2) respectively. According to Eq. (7) the combined wave function  for the two separated atoms of
hydrogen can be written as
 = A(1) B(2) ...(10)
When the two atoms are brought closer together, their energy changes and consequently, the
à also changes. Assuming that the combined wave function , as represented by Eq.
energy operator H
(10) still describes the wave function for H2 molecule, the energy of the hydrogen molecule, following Eq.
(3) would be given by

It has been assumed that when the two atoms are far removed from each other, i.e., when the
internuclear distance rAB is infinity, the total energy of the system is zero. As the atoms are brought closer
in a region in which the electric field due to one can influence that of the other, the energy of the system
starts decreasing till a certain minimum value is reached. This stage corresponds to the formation of H 2
molecule. The internuclear distance at this stage is designated as r0.
When we calculate the energy of the system of the two hydrogen atoms by using Eq. (11) and plot
these energies as a function of internuclear distance, the curve a is obtained. The bonding energy given
by the minimum in potential energy curve a is only 24 KJ mol–1. This is too far below the experimentalvalue
of 458 KJ mol–1 (represented by the minimum in the dotted curve). Moreover, the minimum is found at an
internuclear distance of about 0.90 # (curve a) whereas the experimental bond distance in hydrogen
molecule is 0.74 #. Evidently, the wave function A(1) B(2), as applied to H2 system, is not
a good approximation. Apparently, some other interactions between the two hydrogen atoms are required
to be taken into consideration to arrive at the correct wave function for the combined H2 system. Some
of these interactions are as follows.
(i) Exchange of Electrons
In arriving at Eq. (11), we had assumed that the electrons of the two hydrogen atoms could be
distinguished from each other and therefore, the electron around nucleus A was labelled as e 1 and that
around nucleus B was labelled as e2. This assumption, however, is not correct because when a bond
is formed, the two atomic orbitals overlap and the two electrons become indistinguishable from each
other. We can no longer be sure that electron 1 will always be near the nucleus A and electron 2 will
always be near the nucleus B. The system of two hydrogen atoms may then be represented by two
different states, I and Il, as shown below :

HA (e1 ) HB (e2 ) HA (e2 ) HB (e1 )


(I) (II)

If I and II are the wave functions representing states I and II, respectively, then, according to Eq.
(9) we can write
I = A(A) B(2) and II = A(2) B(1)
A linear combination of I and II (and not I or II alone) should be employed for calculating the
energy of H2 molecule. There are evidently two possible modes of linear combination, viz., additive and
subtractive according to which there are two wave functions, viz., + and – given by
+ = I + II = A(1) B(2) + A(2) B(1) ...(12)
and – = I – II = A(1) B(2) – A(2) B(1) ...(13)
The energy of the system, calculated by using + in the wave equation is shown by the curve
b while that calculated from – is shown by the curve c in Fig.
It is evident that – leads to increase of energy. Thus, bonding is not possible in –. We may,
therefore regard – as repulsive state. The wave function + however, leads to fall of energy. Thus,
bonding is possible in this state. We may, therefore, regard + as a bonding state.
The wave function + Eq. (12) which represents the bonding state is also referred to as a covalent
wave function. Therefore, it may be written as covalent. The reason for this terminology isthat in state
I as well as in state II, there is equal distribution of charge on both the nuclei A and B.
As mentioned above, the bonding energy calculated from Eq. (11) is only 24 kJ mol–1 and the
internuclear distance is 0.80 #. The additional energy resulting from the exchange of electrons between
the two hydrogen atoms is called the exchange energy. Thus, 303 – 24 = 279 kJ mol–1 is the exchange
energy for H2 molecule.
(ii) Screening Effect of Electrons
The wave function given by Eq. (12) is further improved by considering the screening effect of
the electrons. When the two atoms come close to each other, the electron of one atom shields the electron
of the other atom from the nucleus so that the electrons do not feel the full charge of the nucleus.The actual
charge felt by an electron is termed as its effective nuclear charge. If we use the effective nuclear charge
and improve the wave function given by Eq. (12) bonding energy of 365 kJ mol–1 is obtained.This is deplicted
by the minimum in the curved d.
(iii) Ionic Character of HóH Bond
The bond energy of 365 kJ mol–1 is still appreciably smaller than the experimental value of 458
kJ mol–1. This shows that + still does not give the true wave function for H2 molecule. An additional factor
that has been suggested is the possibility that momentarily both the electrons may simultaneously be
present close to one nucleus or the other. This gives rise to ionic configurations represented by state
IIII and IV :
e2 (e2)
(e1)HA HB HA HB(e1)
III IV
These two ionic structure may be represented as HA–HB+ and HA+HB–, respectively. However, the
coulombic repulsion between the two electrons present so close to each other would far exceed the
coulombic attraction between HA– and HB+ or between HA+ and HB–. This reduces the probability of the
existence of these additional structures to a base minimum. Nevertheless, the incorporation of these
structures as resonance structures does increase the stability of the hydrogen molecule and hence its
bond energy.
The state III and IV are, amongst themselves, equally probable and the wave functions corresponding
to state III(III) and state IV(IV) may be expressed as
III = A(1)A(2) ....................................................................................................... (14)
and Iv = B(1)B(2) ....................................................................................................... (15)
The total ionic wave function ionic may be expressed as a linear combination of III and IV, as
ionic = A(1)A(2) + B(1)B(2) ............................................................................... (16)
However, states III and IV are less probable than states I and II as the two electrons present on the
same H atom would repel each other and make state III and IV less likely compared to state I and II.
Considering the possibilities of covalent as well as ionic structures, the combined wave function
for H2 molecule is written as
 =  covalent + ionic
= [A(1)B(2) + A(2)B(1)] + [A(1)A(2) + B(1)B(2)] ...(17)
where  gives the extent to which the ionic structures contribute to the bonding. For H 2,  is
calculated to be 0.17. Knowing , we can find out the extent of the contribution of the ionic structures. The
fraction  is called the mixing coefficient.
With all these modifications, the calculated bond energy comes out to be about 388 kJ mol–1 which
can be considered as reasonably close to the experimental value of 458 kJ mol–1. The internucleardistance
between hydrogen atoms now comes out to be 0.75 # which is quite close to the experimental value of
0.74 #.
Bond energies and bond lengths for H2 molecule corresponding to various wave functions are
summed up in Table.
Table : Bond Energies and Bond Lengths for H2 Molecule
Corresponding to Various Wave Functions

Bond Energy Bond


Wave Function
(kJ molñ1) Length (&)
 = A(1)B(2) 24 0.90
covalent = A(1)B(2) + A(2)B(1) 303 0.80
covalent with screening effect 365 0.76
 = covalent + ionic 388 0.75
= [A(1)B(2) + A(2)B(1)]
+ [A(1)(2) + (1)(2)]
Experimental Value 458 0.74
Ex. Write down the valence bond wave function for the HF molecule (assuming that it is formed from
1s orbital of H and 2pz orbital of F) in the following three cases : (a) HF is purely covalent (b) HF
is purely ionic and (c) HF is 80% covalent and 20% ionic.
Sol. Let, the H atom be represented by A and F atom by B. Let, the electron numbered 1 be the 1s
electron of H atom and the electron numbered 2 be the pz electron of F atom.
(a) For purely covalent structure of HF, the wave function is written as
covalent = A(1)B(2) + A(2)B(1)
(b) For purely ionic structure H+F–, both the electrons are on the F atom so that
ionic = B(1)B(2)
(c) HF is a resonance hybrid shown as
HóF  H+F–
  = c + c with c 2+ c 2= 1
1 covalent 2 ionic 1 2

Since, c 12 = 0.80
and c22 = 0.20 (given),

hence c1   0.89

and c2   0.45
  = 0.89[A(1)B(2) + A(2)B(1)] + 0.45[B(1)B(2)]

electronin an atomic orbital is influenced by one positive nucleus only, an electron in a molecular
orbital is influenced by all the nuclei of atoms contained in a molecule.
If we have a simple molecular system comprising of two electrons described by two one-electron
molecular orbital wave functions 1 and 2, then the combined configurational wave function for the
system, i.e., the molecule, would be given by
 = 12 ...(18)
If, in general, there are n electrons, then the configurational wave function for the system would be
given by
 = 123 ... n ...(19)
where 1, 2, 3 etc., are one-electron molecular orbital wave functions .

The individual one-electron molecular orbital wave functions such as 1, 2, 3, ... can be
determined with the application of the wave mechanical principles especially Ritz method of linear
combination of atomic orbitals.
Linear Combination of Atomic Orbitals (LCAO). Molecular Orbital Treatment of Hydrogen Molecule
Ion, H2+
We shall apply linear combination of atomic orbitals in the molecular orbital treatment of a simple
species, viz., H2+ ion, containing only one electron and then extend the treatment to more complicated
system.
Consider a H2+ ion consisting of two H nuclei designated as A and B (Fig.). The single electron is
supposed to be moving in a molecular orbital formed by the interaction of atomic orbitals of the two
hydrogen atoms, HA and HB.
Fig. : Representation of H +
ion.

The potential energy of the H system


2
+
would evidently, be given by
2 2
e2  er  eR
V ...(20)
r
A B

where rA, rB and R are the appropriate distances, as shown in fig.


à is
As shown in previous chapter, the energy operator (i.e., the Hamiltonian operator) H for H 2+ ion
represented as

à   h  2 V
2
H ...(21)
82m

2 2 2
h2 2  e  e  e
  ...(22)
82m rA rB R

à   E becomes
Accordingly the wave equation H

e2 e2 e2
 h2 2      
 E
 
82m r r R ...(23)
 A B 
The electron in H 2+ ion is moving in molecular orbital under the influence of both the nuclei A and
B. When, at some instant, it comes nearer to nucleus A (situation I), the electron would behave as if it
were moving in the atomic orbital of HA. The motion of the electron would then be described by A (the
unmixed atomic orbital of HA). Similarly, when the electron is nearer to nucleus B (situation II), its motion
would be described by B (the unmixed atomic orbital of HB).

In between the two situations, the electron would be under the influence of both the nuclei and
its motion would be described by a wave function which is a linear combination of A and B. Thus,
 = cAA + cBB ...(24)
where cA and cB are the coefficient of the atomic orbital wave function A and B, respectively.
The squares of the coefficients, viz., c 2 and c 2 give, respectively, the contributions of  and  to the
A B A B

wave function .
The energy function  corresponding to this wave function is given by 
  
 à d /
H d 2
  

...(25)


à (c   c  ) d /
(c   c  ) H (c   c  )2 d ...(26)

 
 A A B B A A B B 
 A A B B
Differentiating Eq. (26) with respect to cA only and then with respect to cB only and applying the
conditions of minimization of energy, viz.,  /  cA = 0 and  /  cB = 0, we get a quadratic equation in
 whose solution gives two values of energy function . Since these energy functions being of minimum
energy are closest approximations of energy E, we may designate these two values of  and E+ and E–
Since, the probability of occurrence of situation I and situation II, described above, is the same,
therefore, the weightages of  and
A
 in Boverall  would be the same. Thus, c 2 = c 2.A B

In order that  is an acceptable wave function, it must be normalized i.e., it must satisfy the


 2 d  1. The atomic orbital wave functions taken for the combination are already
condition that


normalized, i.e.,

2 d  1 and  d  1
2
  A   B

The above conditions of normalization give the relation


cA2 + c 2B= 1
But, as shown above,
cA2 = c 2 B

 cA   cB  (1/ 2) ...(27)

Thus, we get two sets of values for c A and cB, viz., (1/ 2, 1 2) and (1 / 2, 1 / 2) and

hence two linear combinations of A and B corresponding to each set of the values of c A and cB.
Accordingly, we have

  (1 / 2 )A  (1 / 2 )B  (1 / 2 )(A  B ) ...(28)

  (1 / 2 )A  (1/ 2 )B  (1 / 2 )(A  B ) ...(29)

In the above equations, + corresponds to energy E+ and – corresponds to energy E–. The lower
value of energy E+ is lower that either of EA and EB (viz., the energies of isolated HA and HB). The higher
value of energy E is

higher than either of E and
A
E . Thus,
B
we see that in the case of H + ion. 2
1. A linear combination of two atomic orbitals A and B leads to the formation of two
molecular orbitals + and –.
2. The energy E+ of molecular orbital + is lower than either of EA and EB. It is, therefore,
designated as bonding molecular orbital (BMO).
3. The energy E– of molecular orbital – is higher than either of EA and EB, It is, therefore
designated as anti-bonding molecular orbital (ABMO).
4. The extent of lowering of energy of the bonding molecular orbital is equal to the extent of
increase of energy of the antibonding molecular orbital, as shown in fig.
Fig. : Energies of Bonding and Anti-Bonding Molecular Orbitals.
5. Detailed LCAO calculations show that the greater the overlap of the two combining atomic
orbitals A and B, the lower would be the energy of the bonding molecular orbital (+) formed.
6. The combining atomic orbitals A and B are completely used up during the formation of the
molecular orbitals + and –. The electron occupying originally the 1s atomic orbitalof HA
or HB, now occupies the lower level molecular orbital +, losing energy = OM and thus
gaining in stability.
7. The BMO has high electron charge density in the overlap region whereas filled ABMO has
no electron charged density in between the nuclei.
The molecular orbital treatment for H 2+ ion, discussed above, can be extended to other homonuclear
and heteronuclear systems. In general, whenever atomic orbitals A and B, of the same symmetry and of
similar energies of two atom A and B, overlap, two molecular orbitals + and – would be formed by
the linear combination of these atomic orbitals so that the equations for molecular orbitals would be
+ = N+ (A + B) ...(30)
– = N– (A – B) ...(31)
where  is called the mixing coefficient and N+ and N– are called normalising constant2s. The
values of these constants are adjusted in such a way that the normalising conditions, viz.,  d  2
 

and  2 d  1 are satisfied.




If the molecules AB is homonuclear, the weightages of A and B in both + and – would be the
same, i.e., N+ 2 = 2N +2 and N 2–2 = N 2–(weightage of a wave function is related to square of its coefficient).
It readily follows from above that 2 = 1 so that  = ± 1. In such a case, Eq. (30) and (31) for molecular
orbitals would reduce to Eq. (28) and (29) respectively because N  N  1/ 2 and  = ±1.

If, however, the molecule AB is heteronuclear, the weightages of A and B would not be same,
i.e., N 22  N 2 and N 2 2N 2 so that 2  1 and the equations for molecular orbitals would remain as
– – + +
Eqs. (30) and (31).
Ex. Write down the molecular orbital wave function for H2– molecule anion.
Sol. H2– molecule anion contains three electron which may be labelled as 1, 2 and 3. It contains two
nuclei which may be designated as A and B.
MO = 123
where, 1 = c1A(1) + c2B(1)
2 = c3A(2) + c4B(2)
3 = c5A(3) + c6B(3)
Ex. Construct a molecular orbital wave function for the bond between H and Cl in HCI assuming that
the bond is formed from the 1s electron of H atom and a 3p electron of Cl atom.
Sol. Let, H atom be designated by A and Cl atom by B. Let, the 1s electron of H atom be labelled as
electron 1 and 3p electron of CI atom be labelled as electron 2.
For notational convenience, we shall use the symbol  for MO and the symbol  for AO.
MO = 12
where, 1 = c1A(1) + c2B(1)
2 = c1A(2) + c2B(2)
 MO = 12 = [c1A(1) + c2B(1)] [c1A(2) + c2B(2)]
= c 2[ (1) (2)] + c c [ (1) (2) +  (2) (1)] + c 2[ (1) (2)]
1 A A 1 2 A B A B 2 B B
As can be seen, the first and the last terms are ionic whereas the middle terms are covalent. Also
the covalent terms have equal weight but the ionic terms have different weights from the covalent
terms and also from each other.
Ex. Using the LCAO for the wave function for H 2+, obtain the normalized wave function for the BMO
and the AMBO.
Sol. Using the LCAO-MO approximation, the wave function for H + is
2
written as
 = N(c1A + c2B) ...(i)
where, N is the normalization constant.
According to the normalization condition, for real  (using Diracís ëbraí and ëketí notation).
2d  <  |  > = 1 ...(ii)
Incorporating the value of  from Eq. (i) in Eq. (ii), we have
N2 < (c 1 +A c  )2| B(c  + 1c A ) > =2 B1
1
or c2 [  |  ]  c2 [  |  ]  2c c [  |  ]  ...(iii)
1 A A 2 B B 1 2 A B
N2
Assuming that A and B are normalized, i.e.,

  A |  A    2A d  1

Ex. Write down the wave function for the BMO for a heteronuclear diatomic molecule AB assuming that
the electron on an average spends 90% of its time on nucleus A and 10% of its time on nucleus
B.
Sol. In the LCAO-MO scheme,
MO = cAA + cBB
where the coefficients c and c are such that [c ]2 and [c ]2 determine the probability of finding
A B A B

the electron in the AOs A and B, respectively. Then, clearly


[cA]2 = 90% = 0.9
and [cB]2 = 10% = 0.1

Hence, cA    0.95

and cB   0.1  0.32

Thus, MO = 0.95A + 0.32B


Ex. Write down the normalized VB wave function and MO wave function for H2 molecule and comment
on the expression obtained.
Sol. H2 molecule contains two electron labelled 1 and 2 and two hydrogen nuclei labelled A and B. The
pairing of electrons leads to the formation of a covalent bond. In the VB approach, if electron 1
is on nucleus A, then electron 2 would be on nucleus B and vice-versa. Since, however, the two
electrons are indistinguishable,
1
therefore, the VB wave function for H2 molecule can be written as
  [ (1) (2)   (2) (1)]
AB   A B A B ...(i)
2
 

where, is the normalization constant obtained by normalizing  .


VB

In the MO theory, the MO wave function for H2 molecule is given by


MO = 12 ...(ii)
where, 1 and 2 are the normalized wave functions for MOs of H2 given by
1
 [ (1)   (1)]
1   A B ...(iii)
2
 
1
 [ (2)   (2)]
2   A B ...(iv)
2
 

Multiplying 1 and 2, we get the MO as


MO = 12[A(1) + B(1)][A(2) + B(2)] ...(v)
1 1
 [ (1) (2)   (1) (2)]  [ (1) (2)   (1) (2)] ...(vi)
A B B B A B B A
2 2
In Eq. (vi), the first two terms represent the probability of finding both the electrons on the same
atom at the same time. In other words, these terms represent the ionic structures of H2 molecule
which may be written as
   
HA HB and HA HB
The last two terms in Eq. (vi) represents the covalent structure of the H2 molecule.
Comparing VB and MO, we observe that while the wave function in VBT does not give any
weightage to ionic structures, the MOT gives equal weightage to covalent and ionic structures.
Term Symbols for a Diatomic Molecule
The hydrogen molecule (and other diatomic molecules) have a large number of excited electronic
states which are designated by molecular term symbols. As in an atom, in diatomic molecules the orbital
angular momenta of electrons couple to give a resultant orbital angular momentum L and the electron spin
momenta couple to give a resultant spin angular momentum S. The component of the orbital angular
momentum along the axis of the molecule is given by
ML = m1 + m2 + ...
where, mi = 0 for a  orbital and mi = ±1 for a  orbital, and so on. The quantum number  is defined
as the absolute value of ML and is represented by the code letters
=0123
Symbol =    
in analogy to atomic term symbols. The multiplicity of an electronic state of a diatomic molecule
is given by 2S + 1 where, S is the sum of the spins of the electrons in the molecule. The term symbol
of a molecule is represented by 2s+1.
For,  terms, a superscript of plus or minus added according to the behaviour of the wavefunction
upon reflection in the plane containing the internuclear axis. A plus sign indicates that the wavefunction is
invariant under this operation and a negative sign indicates that the wavefunction changes sign on reflection
in this plane. If a diatomic molecule, such as a homonuclear diatomic, has a centre of symmetry, a right
subscript of g or u is attached to the term symbol to denote the parity of the orbital. The parity of an orbital
is determined by observing the inversion symmetry. When a point on an orbital is inverted an equal distance
through the centre of the molecule, the orbital is said to be gerade (g) if it has the same sign at the two
points; otherwise it is said to be ungerade (u); The reader can convince himself of the fact that in an atom
an s orbital is g, a p orbital is u, a d orbital is g, and an f orbital is u. The parity of a multielectron molecule
is obtained by noting g or u for every orbital and forming products, using the relations :
g Q g = g, g Q u = u, u Q u = g
Ex. Show that the ground state term symbol of H is 1 +2. g
Sol. The H molecule has the electric configuration ( , 1s)2, i.e., both the electrons are in the 
2 s g

orbital. For the  electrons, m1 = 0, m2 = 0 so that ML = 0. Since, s = | ML |, hence  = 0. This means


that the terms symbol is . Since, S1 = 1/2 and S2 = –1/2, hence S = S1 + S2 = 0 sothat
multiplicity 2S + 1 = 1 which corresponds to a singlet state. Since, the  orbital is invariant
under reflection in the plane containing the internucler axis, we use the superscript of plus (+).
Again since the orbital is gerade (g), we use the subscript g. Thus, the complete terms symbolfor
H is 1
2
+
. g

Symmetry of Molecular Orbitals


Some of the possible combinations of atomic orbitals in the LCAO-MO scheme are shown in fig.
Those orbitals which are cylindrically symmetrical about the internuclear axis are called  orbitals,
analogous to an s orbital of the highest symmetry. If the internuclear axis lies in the nodal plane, a  bond
is formed. In  bonds (encountered in the metal-metal quadruple bond in transition-metal chemistry), the
internuclear axis lies in two mutually perpendicular nodal planes. All antibonding orbitals, identified with
an , possess an additional nodal plane perpendicular to the internuclear axis and lying between the nuclei.
Also the MOs may or may not have a centre of symmetry. Thus, p–p orbitals are ungerade and

pp orbitals are gerade.
Fig. : Combination of s, s and p, p atomic orbitals to form molecular orbitals of different symmetries. The
figure displays the contours of the atomic and molecular orbitals. The traditional shapes ofs and
p orbitals arise from the projection of the contours on the respective axes.
Molecular Orbitals for Homonuclear Diatomic Molecules
The simplest molecules are those containing two atoms of the same element; these are called
homonuclear diatomics. We shall now consider the LCAO-MO approximation for such molecules using
the basis functions of the two atoms, such as the AOs 1s, 2s, 2p, 3s, etc. The two criteria that must
be satisfied for the formation of BMOs are (i) the overlap between the AOs must be positive, (ii) the
energies of the AOs must be approximately the same in order that there is effective interaction between
the AOs of different atoms.
The molecules are built up by adding electrons to MOs in the same way as the atoms are built
up by adding electrons to AOs. The principles involved in both cases are the same. The basic principles
can be summed up as follows :
1. The MO with the lowest energy is filled first.
2. The maximum number of electrons in a MO cannot exceed two and the two electrons
must be of opposite sign.
3. If there are two degenerate MOs, pairing of electrons will occur only after each degenerate
MO has one electron.
The order of filling of molecular orbitals is
    
1s, 1s, 2s,  2s, 2p z, 2p x  2p y  2p x   2p y 2p z
(for molecules having e– > 14)
In case electrons are  14, the order is following :

1s,  1s, 2s,  2s, 2p x  2p y , 2p z,  2p x   2p y  2p z

It is useful to define the term bond order that is roughly proportional to the strength of bonding.
This is equal to one-half the difference between the number of bonding electrons and the number of anti-
bonding electrons. Thus,
1
Bond order  (N  N ) ...(32)
2 b a

where, Nb = number of electrons in bonding molecular orbitals.


Na = number of electrons in anti-bonding MOs (represented by  or  )

Bond order of 1, 2 and 3 signify single, double and triple bonds. Bond order of zero means that
no bond is formed. The bond order of a molecule is directly proportional to its bond dissociation energy.
This means that the higher the bond order, the larger the bond dissociation energy and the smaller the
bond length. e.g.,
1
(1s)2; bond order = (2  0)  1. The molecule has a single bond.
1. H2 2

The bond dissociation energy of H 2molecule has been found to be 458 kJ mol–1 and bond length
equal to 0.74 #.
2. Hydrogen molecule ion, H2 + This molecule has been detected spectroscopically when electric
discharge is passed through hydrogen gas under reduced pressure. The molecule has only one electron
and its electronic configuration is
1 1
H +  (1s)1 and its bond order = (1 0) 
2
2 2
The positive value of bond order indicates that the bond is formed and the molecule is stable.
However, the bond in H 2+ ion is weaker than that in H molecule.
2
The bond dissociation energy of H + ion
2
(269 kJ mol ) is less than that of H molecule
–1
2
(458 kJ mol –1
). The bond length in H +
ion
2
(1.04 #)is
larger than that in H molecule
2
(0.74 #). These observation support the fact that the bond in H +2 ionis
weaker than the bond in H2 molecule.
3. He2 The electronic configuration of He2 is
He 2 : (1s)2  (1s 2 )
Fig. : The Energies of Bonding and Antibonding Orbitals for the He2 Molecule.
Diatomic molecules of the second row elements In these elements, the 1s orbitals are completely
filled. In the formation of molecular orbitals, the electrons in the inner shells (i.e., 1s electrons) of each
atom remain essentially unperturbed in their respective atomic orbitals and may be kept out of consideration.
In the formulation of electronic configurations of these molecules, the letters KK are generally used for
denoting the fully filled inner shells (K shells) in the two atoms. e.g.,
4. Li2 The electronic configuration of lithium atom (Z = 3) is 1s2 2s1. There are six electrons to be
accommodated in lithium molecule. The four electrons are present in K shells and there are only two
electrons to be accommodated in molecular orbitals. These two electrons go into  (2s) BMO which has
lower energy and  ( 2s)ABMO remains empty. The electronic configuration of Li2 molecule is thus
represented as

Li2 : KK  (2s)2
1
and its bond order = (2  0)  1
2
Thus, there is one Li–Li sigma bond. The bond dissociation energy of the molecule is quite low,
being about 105 kJ mol–1. The bond length is 2.67 #.

Fig. : Molecular-orbital descriptions for Li2. In (a) the 1s electrons are regarded

as remaining in atomic orbitals. In (b) they are in g and u molecular orbitals.
5. O2 The electronic configuration of O2 molecule is

O 2 : KK( 2s)2  ( 2s)2 ( 2p z )2 ( 2p x )2 ( 2p y )2  ( 2p x )1  ( 2p y )1

Since, it contains two unpaired electrons, the oxygen molecule is paramagnetic.


1
Bond order = (8  4)  2 O molecule, thus, contains a double bond.
2
2
Because of the presence of 4 electrons in ABMOs, O 2 molecules is less stable than N2 molecule.
Its bond dissociation energy is much less, being 494.6 kJ mol–1 bond length is larger, being 1.21 #.
Molecular Orbital Energy Level
Diagrams of Heteronuclear Diatomic Molecules
The cases of heteronuclear diatomic molecules, in which the two atoms constituting a molecule
are different, may now be considered. The principles involved in the distribution of electrons are the same
as discussed before. However, the molecule orbital diagrams will not be symmetrical.
Let us discuss the electronic configurations and molecular orbital diagrams for some common
heteronuclear diatomic molecules.
1. Carbon Monoxide Molecule, CO
The electronic configuration of carbon and oxygen atoms are :
C : 1s2 2s2 2p2 and O : 1s2 2s2 2p4
There are four electrons in the outermost shell of carbon atom and six electrons in the outermost
shell of oxygen. Thus, a total of 10 electrons are to be accommodated in the molecular orbitals of CO
molecule. The MO diagram for CO molecule may be drawn similar to NO molecule as shown in Fig.
Then the molecular orbital configuration may be written as :
  2p2 
x 
CO : KK 2s 2
 2s
 2 2p z
2 
 2p y 
2


82
Bond order = 3
2
Thus, CO molecule has a triple bond. Its bond dissociation energy has been found to 1067 kJ
mol and bond length equal to 1.128 #. However, it may be noted that the molecular orbital diagram of
–1

CO is somewhat complicated than NO molecule. The above configuration is not correct.


Oxygen is more electronegative than carbon and therefore, the atomic orbitals of oxygen are

lower than those of carbon. Due to this difference, the  2s MO is slightly higher in energy than 2p , z

2px and 2py MOs. The diagram for CO molecule is shown in fig.
The better understanding of the structure is possible by comparison of CO and CO +, CO+ is formed
by the loss of one electron from CO molecule. If the electronic configuration for CO molecule similar to
NO, is correct, then electronic configuration for CO+ is
  2p2 
x 
CO : KK 2s 2
 2s
 2 2p z
2 
 2p 1

y 

72 1
Bond order = 2
2 2
1
Thus, when CO is ionized to CO+, then its bond order should be reduced 2
and the bond length
2

should be increased. However, the bond length in CO+ is 1.115 # and is less than in CO (1.128 #). This
means that the simple picture of CO molecule similar to other molecules is not correct. In fact there is m
slight increase in bond order when CO is changed to CO +. The most likely explanation for the increase fig
in bond order or decrease in bond length is that the electron must have been removed from anti-bonding

Molecular Orbitals Diagram for CO Molecule.



The above figure shows that  2s MO is higher in energy than 2p ,z 2p xand 2p yMOís. Now

when an electron is lost, it is removed from  2sMO and therefore, bond order increases. As a result,

CO+ has slightly strong bond than does CO and bond length in CO + is less than that in CO molecule.
The molecular orbital configuration may be written as
 2p2  
CO : KK 2s2 2p2 x
 2s2 
z
 2p2y 
2. HF
The electronic configurations of hydrogen and fluorine are 1s 2 and 1s2 2s2 p5, respectively. In the
formation of HF molecule, only the 2p electrons of fluorine atom would combine effectively with the solitary
electron of hydrogen atom. As has been already explained, only a p z orbital is able to combine with an s
orbital. The combination of a px or py orbital with an s orbital is ruled out on symmetry
considerations. The 2pz orbital of fluorine gives a highly effective overlap with the s orbital of hydrogen to

form (sp) BMO and (sp) ABMO .

Unsymmetrical shape of molecular orbital formed by the combinationof 1s orbital of


hydrogen with 2pz orbital of fluorine.
The bonding molecular orbital is fully filled with two electrons. The rest of the electrons remain
in their atomic orbitals. The MO formed in the case of HF molecule will not be symmetrical. The
asymmetry occurs because the energies of H(1s) and F(2pz) atomic orbitals are not the same. The MO
diagram for HF molecule is shown in fig.
As can be seen, the bonding Mo has an energy which is less closer to that of 2p z orbital of fluorine
than to that of higher energy 1s orbital of hydrogen.
3. HCl
The electronic configurations of H and Cl atoms are 1s1 and 1s2, 2s2, p6, 3s2, p5, respectively.
In this case also, there is only one bonding MO and one anti-bonding MO formed by the combination of
1s orbital of hydrogen with 3pz orbital of chlorine. The rest of the electrons of the chlorine atom remain
in their respective AOs.
The electronic configuration of HCl may thus be represented as
HCl : KL(3s)2, (sp)2, (3px)2, (3py)2
Here, too, the MO formed will not be symmetrical. The MO energy level diagram of HCl molecule
will be exactly similar to the one shown in fig. for HF molecule.
Table gives the ground state electronic configuration of homonuclear diatomics (along with other
relevant data), in a slightly different notation.
Table : Ground State Electronic Configurations of Homonuclear Diatomics
Number of Electronic Term Bond 4.
Molecule Re (pm) De (eV) De (kJ molñ1)
Electron Configuration Symbol Order
2 
1
H2+ 1 (1g)1  106.0 2.793 269.5
2
1 
H2 2 (1g)2  1 74.1 4.748 458.1
1
He2+ 3 (1g)2 (1u) 2
 108.0 2.5 238.0
2
He2 4 (1g)2 (1u)2 1
 - - -
1 
Li2 6 [He2](2g) 2
 1 267.3 1.14 110.0
Be2 8 [He2](2g)2 (2u)2 1
 0 - - -
3 
B2 10 [Be2](1u) 2
 1 158.9 -3.0 -290
C2 12 [Be2](1u) 4 1
 2 124.2 6.36 613.8
1 
N2 14 [Be2](1u)4 (3u)2  3 109.4 9.902 955.4

O2+ 15 [N2](1g)2 2
 21 112.27 6.77 653.1
2
O2 16 [N2](1g)2 3
 2 120.74 5.213 502.9
1 
F2 18 [N2](1g) 4
 1 143.5 1.34 118.8
Ne2 20 [N2](1g)4 (3u)2 1
 0 - - -
The Water Molecule
The procedures for constructing molecular orbitals for more complicated molecules may be
exemplified by a brief treatment of the water molecule, the symmetry elements for which are shown in Fig.
The oxygen atom lies on all three of these symmetry elements, and in Fig(a). We show orbitals of the
oxygen atom in relation to these elements. Since, the 1s and 2s orbitals are spherically symmetrical, none
of the symmetry operations brings about any change in them and both are therefore of A 1 symmetry.
The 2pz orbital has a positive lobe in one direction along the Z axis and a negative lobe in the other. The
C2, , and ' operations bring about no change in this orbital, which is therefore also of A1 symmetry.
The 2px and 2py orbitals, however, behave differently. Reflection of the 2px orbital of the ' plane
interchanges the + and – lobes and therefore changes the sign, and the same happens with a C 2
rotation. Reflection in the  plane brings about no change. The 2px orbital is therefore of B1 symmetry.
By similar arguments we find that the 2py orbital is of B2 symmetry.
The situation with the orbitals on the hydrogen atoms in the water molecule is a little more
complicated, since these atoms do not lie on all the symmetry elements of the molecule; they only lie
in the ' plane. The 1s orbital of the HA atom, which we will write as 1sA , therefore does not correspond
to the symmetry of the molecule, and the same is true of the 1s B orbital. However, from these two orbitals
we can construct orbitals that do have the right symmetry. Thus the sum of the atomic orbitals, 1s A +
1sB, is unchanged when any of the symmetry operations is performed, and it is therefore of A1 symmetry.
The difference, 1sA – 1sB, changes its sign when the C2 operation is performed and when the 
operation is performed, but it remains unchanged when í is performed. It is therefore of B2 symmetry.
The symmetry properties of the various orbitals on the H and O atoms are summarized in Fig.(b).
The complete molecular orbital must correspond in symmetry to one of the symmetry species of the
molecule, and this is impossible if we combine atomic orbitals of different symmetries. We can, however,
obtain a molecular orbital of symmetry A1 if we make a linear combination of 2s, 2p z, and 1sA + 1sB orbitals,
all of which are of A1 symmetry. Such an orbital is designated a 1 (the lower case letter is used) and is of
the form
a1 = 1sA + 1sB + 12s + 22pz ...(33)
where 1 and 2 are coefficients which could be determined by a variation procedure. This orbital
is represented schematically in fig.
Fig. : Orbitals for the water molecule. (a) The atomic orbital and their symmetry species. The ív plane
is the plane of the molecule.(b) The C2v character table showing the symmetries of the atomic
orbitals and symmetry-adapted orbitals. (c) and (d) The bonding molecular orbitals a1 and b2.

We can also construct the orbital


b2 = 1sA – 1sB + 2py ...(34)
in which orbitals of B2 symmetry are combined. This orbital is represented in fig.(d).
Both of these orbitals are bonding, since the atomic orbitals have been added together and there
is a piling up of charge between the nuclei. For each bonding molecular orbital there is an anti-bonding
one, of the same symmetry, formed by subtracting the atomic orbitals. These are indicated by an asterisk.
The anti-bonding molecular orbital corresponding to a1 is
a  1s  1s   ' 2s   ' 2p ...(35)
1 A B 1 2 z
That corresponding to b2 is
b  1s  1s   ' 2p ...(36)
2 A B y

There is also an anti-bonding orbital of B1 symmetry, which consists only of 2px(O) itself; this cannot
combine in any way with the hydrogen atom orbitals and, therefore, gives no bonding.
These conclusions are shown in Fig. in the form of an energy diagram. The order of the energy
levels is that predicted by self-consistent field calculations. The Is atomic orbital of the oxygen atom has
a much lower energy than the other atomic orbitals, and it is therefore carried over as the lowest MO;
electrons in this orbital are non-bonding. The 2s atomic orbital of the oxygen atom also forms an MO,
designated aí1, and electrons in this level are also non-bonding.
The assignment of the 10 electrons in the water molecule is shown in Fig. The 2 electrons in the
1s orbital of oxygen make no contribution to bonding. Two electrons are in each of the a 1 and b2 orbitals,
and these four electrons are responsible for the two O H bonds. The remaining four electrons are in the
aí1 and b1 orbitals and constitute the two lone pairs.

Fig. : The energy levels corresponding to the molecular orbitals for water, showing
how they are derived from the atomic orbitals of oxygen and hydrogen (not to scale).
Ex. Of the following species, which has the shortest bond length : NO, NO +, NO2+, NO– ?
Sol. We shall first calculate the bond orders for the given species from their electronic configurations.

NO KK ( 2s)2  ( 2s) 2 ( 2p z )2 ( 2p x )2 ( 2p y )2  ( 2p x)3


1 1
Bond order  (8  3)  2
2 2
NO+ One electron is removed from the anti-bonding  ( 2p x ) MO.
1
Bond order  (8  2)  3
2

N O 2+ One electron is removed from the anti-bonding  ( 2p )x MO and the other electron is removed
from the bonding (2py)MO.
1 1
Bond order  (7  2)  2
2 2
NOñ One electron is added to the anti-bonding  ( 2p x ) MO.
1
Bond order  (8  4)  2
2
Evidently, the species NO+ has the highest bond order and thus has the shortest bond length.
4. HUCKEL THEORY FOR CONJUGATED -ELECTRON SYSTEMS

This theory was originally introduced to permit qualitative study of the -electron system in planar,
conjugated hydrocarbon molecules. It is thus most appropriate for molecules such as benzene or
butadiene, but the approach and concepts have wider applicability.
Basic Assumptions
1. The atomic orbitals contributing to the -bonding in a planar molecule (e.g., the so-called
p orbitals in a molecule such as benzene) are anti-symmetric with respect to reflection
in the molecular plane; they are therefore of a different symmetry to the atomic orbitals
contributing to the -bonding and may be treated independent

The basic form of the secular determinant for the bonding arising from the overlap of two orbitals
is reproduced below.

1  E 12
12 2  E

For three overlapping orbitals the approach leads to a secular determinant of the form :
1  E 12 13
12 2  E 13
13 23 3  E

From a comparison of the two secular determinants given above, it is becoming clear that all such
secular determinants have a characteristics structure
1. Each row and column may be associated with one of the atomic orbitals; thus the first row
and first column contain information about the nature of orbital 1 and its interactions with
the other orbitals, the second row and the second column contain information about the
nature of orbital 2 and its interactions with the other orbitals.
2. The diagonal set of elements (comprised of those elements where row 1 intersects column
1, row 2 intersects column 2,... and so on) include the values of the relevant Coulomb
integrals (1, 2 etc.).
3. The off-diagonal elements (comprised of those elements having different row numbers and
column numbers) are equal to the relevant resonance integrals (e.g., 12 at the intersection
of row 1 and column 2)
This structure is summarised below, where the rows and columns have been labelled with the
numbers identifying the associated atomic orbital.

x 10
1 x 1 0

0 1 x

(  E)
where x 


 x(x.x – 1.1) – 1(1.x – 1.0) + 0(1.1 – x.0) = 0


 x3 – x – x = 0
 x3 – 2x = 0
 x(x2 – 2) = 0

SHAPE \* MERGEFORMAT  x  0 or x  
5. GROUP THEORY

The relationships among the symmetry elements can be treated in terms of the concepts of
group theory which can be stated as four rules.
(1) The combination of any two members A and B results in a third member C, which also
belongs to the same group.
The order of combination is very important as AB is not necessarily equal to BA.
If AB = BA the members A and B are said to be commutative.
(2) There must be a member E in the group such that
AE = EA = A, BE = EB = B, etc.
The member E is called identity.
(3) When more than two members of a group combine, they do so in an associative manner
i.e.,
(AB)C = A(BC)
(4) Every member A must have its inverse (or reciprocal)A–1, which is also  member of the
group i.e.,
AA–1 = A–1A = E
Main Application of Group Theory
(1) Construction of hybrid orbitals.
(2) Construction of SALCs (symmetry adapted linear combinations of atomic orbitals)
(3) Determination of the irreps to which the vibrational modes of molecules belong.
(4) Determination which spectral transitions in infrared and Roman spectra are allowed
forbidden.
(5) Determining the selection rules for n   and    transitions in carboxyl compounds.

(6) Simplifying the so-called secular equation (quantum mechanical calculations in both VBT
and MOT).
(7) Determining optical activity of molecules.
(8) Determining which molecules are polar or non-polar.
(9) Classification of elementary particles into fermions and boson in quantum mechanics.
Q. Assuming that a molecule AB6 belongs to On point group, determine the point groups that result
if it is changed into :
(a) AB5C
(b) Trans-AB4C2
(c) Cis AB4C2
Sol. (a) C4V
(b) D4h
(c) C 2v
Q. In the table given below of symmetry elements and operations, fill the blanks with proper element
or symbol or operation.
Operation Element Symbol
1. To leave the molecule as such. ó E
2. Rotation about an axis through n-fold ëproperí ó
2
axis of rotation.

angle, n being 1, 2, 3, ...


n  
3. Reflection in a plane Plane of symmetry ó
4. ó Centre of inversion. i
5. Rotation about an axis through ó Sn
2
angle followed by reflection
n
in a plane perpendicular to that
axis.
Sol. 1. Identity
2. Cn
3. 
4. Inversion through a point.
5. n fold improper axis of rotation.
  (R)
R
i mn  j (R)m n  0 , if i  j ...(2)

i(R)mn j(R)m'n'  0 , if m  mí
R

and/or n = ní ...(3)
h
 (R)  (R) 
i mn j mn ...(4)
R
li

Equation (2) indicates that the vectors differ only in the fact that they are chosen from matrices
of different representations and are orthogonal. Equation (3) indicates that the vectors from the same
representation but from different sets of elements in the matrices of this representation are orthogonal.
h
Equation (4) indicates the fact that the square of the length of any such vector equals l .
i
à 2 , à v , à 'v and E
The Operators C à form A Group.

Taking water molecules as the object of operations, let v be the reflection in the molecular plane
and ív the reflection in the plane bisecting the HóOóH angle. Then, constructing the multiplication table
by investigating the effect of two successive operation.

The effect of à followed by another C


C à is to bring the water molecule back to the original
2 2

position.

 à 2C
C à 2  Eà

The result of à v followed by à 'v is the same as the rotation by 180c. Hence,

à2
à v à ' v  C
By proceeding along these lines, we get the following group multiplication table :
Let us denote the three basis functions as f1, f2 and f3 and the SAFS by (a), (e) and í(e).
From the projection operator theorem
(a)  1Eàf  1C
à f  1C
à 2f
1 3 1 3 1

= f1 + f2 + f3
The E representation is doubly degenerate. The functions belonging to it are
( e)  2Eàf  1C
à f  1C
à 2f
1 2 1 2 2

= 2f1 – f2 – f3

r
o
m
t
h
i
s
t
a
b
l
e
w
e
s
e
e
t
h
a
By starting with the function f2, we have
í(e) = 2f2 – f3 – f1
By starting with f3, we obtain (2f3 – f1 – f2).
However, this function is not independent of the other two functions, as may be seen by adding
them.
Instead of the above set we may work with
1 = (e) + í(e) = 2f3 – f1 – f2
and 2 = (e) – í(e) = f1 – f2
Since, any linear combination of degenerate eigen functions is also an eigen function.
6. MOLECULAR SPECTROSCOPY

Spectroscopy deals with the transitions that a molecule undergoes between its energy levels upon
absorption of suitable radiations determined by quantum mechanical selection rules.
Let us consider how a spectrum arises. Consider two molecular energy levels E n and Em, as shown
in Fig.

Fig. : Spectroscopy Transitions between Molecular Energy Levels.


If a photon of frequency v falls on a molecule in the ground state and its energy hv is exactly equal
to the energy difference E (= Em – En) between the two molecular energy levels, then the molecule
undergoes a transition from the lower energy level to the higher energy level with the absorption of a
photon of energy hv. The spectrum thus, obtained is called the absorption spectrum. If the molecule falls
from the excited state to the ground state with the emission of a photon of energy hv, the spectrum
obtained is called the emission spectrum.
Basic Features of Different Spectrometers
The schematic diagram of the apparatus for absorption spectroscopy is given in Fig.

Fig. : Apparatus for Absorption Spectroscopy.


Source
The spectrometer is used when absorption occurs in infrared and ultraviolet regions of the
electromagnetic spectrum. The source in a spectrometer produces radiation spanning a range of
frequencies, but in a few cases (such as lasers), it is almost a monochromatic radiation. The radiation
source in an absorption spectrometer is a heated ceramic filament coated with rare-earth oxides (a Nernst
emitter or filament) for the infrared region. The source for the visible region of the spectrum isa tungsten
filament which gives out intense white light : for the ultraviolet region the source is a hydrogendischarge
lamp. A klystron (which is also used in radar installations and microwave ovens) or more commonly, a
semiconductor device called Gunn diode, is used to generate microwaves. The radio- frequency radiation
is generated by causing an electric current to oscillate in a coil of wire.
Dispersing Elements or Analyser
The variation of absorption with frequency is determined, traditionally, by analyzing the spectral
radiation by means of a dispersing element which separates different frequencies into rays that travel
in different directions. The simplest dispersing element is a glass or quartz prism but a diffraction grating
is more widely used.
Detector
The third component of spectrometers is the detector, a device that converts the spectral radiation
into an electrical signal that is passed on to a recording device operating synchronously with the analyzer,
thus producing either a trace on a chart recorder or a computer record of the spectrum. Common detectors
are the radiation-sensitive semiconductors. The radiation is chopped by a shutter that rotates
in the beam so that an alternating signal is obtained from the detector (an oscillating signal is easier to
amplify than a steady signal). A modulator is introduced to convert the signal to an alternating character.
The procedure enables more AC electronics to be employed in the recording stages. In the microwave
region the source frequency is varied and the analyser is not necessary.
Sample
The highest resolution is obtained when the sample is gaseous and at such low pressure that
collisions between molecules are infrequent. Gaseous samples are essential for microwave (pure rotational)
spectroscopy for molecules can freely rotate only in the gaseous state.
In order to achieve sufficient absorption, the path lengths of gaseous sample must be very long, of
the order of metres. Long path lengths are achieved by multiple passage of the beam between two parallel
mirrors at each end of the sample cavity. For infrared spectroscopy, the sample is typically a liquid held
between windows of sodium chloride (which is transparent down to 700 cm–1) or potassium bromide (down
to 400 cm–1). Other ways of preparing the sample include grinding it into a paste with ëNujolí, a hydrocarbon
oil, or passing it into a solid disk, with powdered potassium bromide.
Fourier Transform Technique
Now-a-days, it is a common practice to use Fourier transform technique in spectroscopy,
particularly with infrared (IR) and nuclear magnetic resonance (NMR) spectroscopies. In Fourier-transform
infrared (FTIR) spectroscopy, a Michelson-type interferometer is used to analyse the spectrum. It functions
by producing an interferogram which is the superposition of a series of waves, each of which represents
a component in the spectrum in terms of intensity and wave number. A Fourier transformation of the
interferogram then produces a well-resolved absorption spectrum of the species, with a good signal-to-
noise (S/N) ratio, also abbreviated as SNR.
Selection Rules
The molecular spectra are governed by the so-called selection rules which specify the changes
in the quantum numbers accompanying a particular transition. The chemist is lucky that selection rules
exist which determine a spectrum. It there were no selection rules. The resulting spectrum would by very
chaotic, indeed ! The selection rules are, in fact, the 'backbone' of spectroscopy and are obtained from
the quantum theory of interaction of radiation with matter. Let us enumerate a few examples which would
later be elucidated. For a diatomic molecule, such as H 2, NO, CO, etc., the selection rule for a pure
rotational transition is J = ± 1, where, J is the rotational quantum number. The selection rule for a pure
vibrational transition is V = ± 1, where, V is the vibrational quantum number.
The selection rules, however, are not always obeyed strictly. This is because certain approximations
which have been used in the derivation of the selection rules are not valid strictly. The spectral transitions
which obey a given selection rule are called allowed transitions whereas those which violate a selection
rule are called forbidden transitions. In general, the allowed transitions are more intense (stronger) than
the forbidden transitions which are weak.
The natural line width (or line-time broadening) of a spectral line is determined by the
Heisenberg uncertainty principle, Et  h / 4, where, E is the uncertainty in the energy and t is the
uncertainty in the life-time of the energy level. Since, for a photon E = hv so that E = hv, hence the
natural line width, v, is given by
v  (4t)–1 ...(1)
Width and Intensity of Spectral Lines
When we analyse the spectrum of a molecule, the first thing we wish to know is how sharp and
how intense (strong) is the spectral line. These two quantities are common to all branches of spectroscopy.
fig. shows sharp spectral line having no width while fig. shows a spectral line having a width E at half-
height. The chemist would, indeed, be a happy person if the spectral lines were all very sharp and very
intense. In practice, this is not so.
Fig. : (a) Sharp Spectral Line, (b) Spectral Line having a Width.
Two factors contribute to broadening of a spectral line : (i) The collision broadening and (ii) the
Doppler broadening. The collision broadening is largely responsible for the width of spectral lines in the
ultraviolet (UV) and visible regions. These transitions mostly take place between electrons in the outer
shells in a molecule. When molecules in the gaseous or liquid phase collide with one another, they deform
the charge clouds of the outer electrons thereby slightly perturbing the energy levels of these electrons.
Hence, the spectral transitions between these perturbed energy levels are broadened.
The Doppler broadening arises when the molecule under investigation has a velocity relative
to the observer or observing instrument. This is generally, the case with gaseous samples where the
molecules are undergoing random motion according to the postulates of the kinetic theory of gases. If
the molecule is moving towards the measuring instrument with velocity u, then the frequency v' of radiation
'seen' by the molecule is given by
ví = v(1 + u/c) ...(2a)
where, v is the radiation frequency and c is the velocity of light. If, on the other hand, the molecule
is moving away from the measuring instrument, the frequency of radiation 'seen' by the molecule is given
by
ví = v(1– u/c) ...(2b)
Rearrange Eq. (2a), we obtain
(v – v')/v = v/v = – u/c ...(3a)
Similarly, rearranging Eq. (2b), we obtain
(v – v')/v = v/v = u/c ...(3b)
The quantity v is the Doppler broadening. From the kinetic theory of gases, it can be shown that
the Doppler broadening of the spectral line of a molecule of mass m is given by
v/v = (2/c) (2kT In 2/m)1/2 ...(4)
Since, v/v is directly proportional to T , Doppler broadening can be reduced (and spectral lines
1/2

of maximum sharpness can be obtained) by working with cold gaseous samples.


Intensity of Spectral Lines
The intensity of a spectral line is determined by (i) the Boltzmann population of the energy levels
and (ii) the transition probability between the energy levels.
According to Boltzmann, if, at temperature T, N0 is the number of molecules in the ground state,
then the number of molecules, N, in the excited state is given by
N = N0 e–E/kT ...(5)
where, E is the energy difference between the ground and excited states and k is the Boltzmann
constant. The relative population at equilibrium is, thus, given by
N/N0 = e–E/kT ...(6)

Evidently, if E is large, N/N0 is small, i.e., the number of molecules in the excited state is less than
that in the ground state. In fact, at room temperature, most of the molecules are in the ground state.
Hence, the spectral lines originating from transitions from the ground state to a higher, say, third excited
state would be more intense than those originating from transitions from the first excited state to the third
excited state.
7. PURE ROTATION (MICROWAVE) SPECTRA

The investigation of pure rotation spectra provides a powerful method of finding molecular
parameters of simple molecules that may be obtained in the gas phase. A rotating diatomic molecule
whose nuclei are considered as being separated by a definite mean distance may be treated as a rigid
rotator with free axis. As a suitable model for the interpretation of pure rotation spectra, we may consider
two atoms of masses m and M, respectively at a fixed distances of r1 and r2 from the centre of gravity
as shown in fig.

Fig. : Rotation of A Diatomic Molecule about Its Centre of Gravity.


If this molecule rotates about an axis perpendicular to the internuclear axis through the centre of
gravity O, then moment of inertia is given by
I = mr1 2 + Mr22 ...(7)
or I mr 2
i i i

where, ri is the distance of ith particle from the centre of gravity of the system.
For a diatomic molecule, for instance, the masses and distances of the above fig. lead to the
centre of gravity being located so that,
mr1 = Mr2 (By law of momentum)
or since, r1 + r2 = r, to the relations,

M
r1  r
m M
 M
and r2  r
m M

Substituting these values in Eq. (7) we obtain

mM
2
m2M
I r 2
 r2
(m  M) 2 (m  M)2

mM
 r2
mM
 I = r2
where,  the reduced mass is defined as,

mM

mM
The kinetic energy of a system of moving particles is given by
1
T(KE)   m v2
i 2 i i
Fig. : Rotating Dumbell.
If the system performs a rotational motion then the particle velocities are more conveniently
treated in terms of the angular velocity  of the system, so that we can write the kinetic energy T, as
1
T m v2
i i
2
1
or T m 2r2
i i
2
1
 2m r2
i i
2

1 2
T l ...(8)
2

Thus, the rotation of any system is most conveniently treated in terms of the angular velocity, 
and the moment of inertia, I.
The above equation is the classical expression for the rotational kinetic energy of the rigid rotator
model.
By analogy with linear momentum, which is defined as the product of the linear velocity and
mass, the angular momentum is defined as the product of angular velocity and moment of inertia.
Denoting the angular momentum by P we can write
P=Ihw
and so the expression for KE becomes
KE = P2/2l ...(9)
In Eq. (9), the allowed amount of angular momentum are multiples of the quantity h/2 where,
h is Planck's constant. Thus, the angular momentum will be given by,

P  (J(J  1))  h / 2 ...(10)


in which J is a rotational quantum number having only integral values, 0, 1, 2, ... The expression
for allowed rotational energies for a diatomic molecule is then obtained by substituting (10) in (9).
Denoting the rotational energy for a molecule by E R, the expression becomes
h2
ER  J(J  1) ...(11)
82l
where, I is the moment of inertia which is the characteristic of the molecule.

If a rotational transition occurs, from an upper level with quantum number Jí to lower level of
quantum number J”, the energy emitted is given by

h2
E'R  E''R  {J'(J' 1)  J"(J" 1)} ...(12)
82I
The frequency of corresponding spectral line should be

E'R  E"R
v
h

E'R  E"R
or (cm1 )  , in wave number.
hc
Substituting the value of EíR – E”R in the above equation we obtain
1 h2
 1
(cm )   {J'(J' 1)  J"(J" 1)}
hc 82I

h2 {J'(J' 1)  J"(J" 1)}


 ...(13)
82Ic

The selection rule for rotational transition of linear molecules that occur with absorption or emission
of electromagnetic radiation is J = ± 1, where, J is equal to Jí – J”.
For studies of the absorption of rotation, the experiment that is invariably done when rotational
spectra are obtained, the appropriate part of the selection rule is J = + 1 or Jí – J” = + 1 or Jí = J” +
1. Now, substituting the value of Jí in Eq. (13), we get

h
(cm–1)  {(J" 1)(J" 1 1)  J"(J" 1)}
82Ic

2h
 (J" 1)
82Ic
For the sake of simplicity, writing J instead of J” we have,

h
(cm1 )  (J  1)
82Ic

or (cm–1) = 2B(J + 1)
h
where, B  cm1
82Ic

and this B is called the rotational constant of the molecule and is a characteristic property of the
molecule. We can expect the absorptions of radiation at frequencies of 2B, 4B, 6B, ... etc., for the levels
with J = 0, 1, 2, ... etc. Therefore, the lines are spaced equally and the constant frequency separation
h
between successive line is equal to 2B, i.e., 42Ic .

According, to classical electrodynamics, an intermolecular motion leads to a radiation of light if


a changing dipole moment is associated with it. For a rigid rotator this can be caused by the rotating mass
having a permanent dipole moment, that lies in the direction perpendicular from the mass point to the
axis of rotation. This applies to all diatomic molecules that consist of unlike atoms like CO, NO, HCI, etc.
For such molecules the centres of positive and negative charges do not coincide, i.e., such molecules
have a permanent dipole moment. During rotation, the component of dipole changes in a fixeddirection
with rotational frequency, i.e., classically light of frequency, vrot should be emitted.
For molecules consisting of two like atoms (homonuclear) such as H2, O2, N2, no dipole moment
arises and, therefore, no light is emitted. Conversely, only if a permanent dipole moment is present, then
an infrared frequency can be absorbed and thereby a rotation of the system be produced.
Thus, a rigid rotator is assumed by making all these derivations and the expression for a rigid
planar rotator can also be written as :
ER h ER
  J(J  1), where values are called rotational terms or sometimes the spectroscopic

hc 82Ic hc
terms of the molecules.
Ex. The pure rotational (microwave) spectrum of gaseous HCl consists of a series of equally spaced lines
separated by 20.80 cm–1. Calculate the internuclear distance of the molecule. The atomic masses
are :
1
H = 1.673 Q 10–28 kg;
35
Cl = 58.06 Q 10–27 kg.
Sol. The spacing between the lines
= 2B = 20.80 cm–1
20.80
 B=  10.40 cm–1
2

h
B cm1
82Ic
h
I

82Bc

 6.626  1034 J s

(82 )(10.40 cm1)(3  1010 cm s1)

= 0.2689 Q 10–46 kg m2 (J = kg m2 s–2)


The reduced mass is given by

m1m2

m1  m2
where, m1 and m2 are the atomic masses of H and Cl, respectively. Thus,

(1.673  1027 kg)(58.06  1027 kg)



(1.673  58.06)  1027 kg

(1.673  58.06)1027
 kg  1.626  1027 kg
59.73

Since, I = r2
1/ 2
I
hence, r   
  

 0.2689 1046 kg m2 
 1.626 1027 kg
 
= 1.29 Q 10–10 m = 129 pm.
Ex. Compare the first excited state rotational energy of CO molecule obtained in the last example with
its thermal energy at room temperature and comment on the result.
Sol. Thermal energy of CO molecule at room temperature, i.e., 25cC = kT
1.38  1023 JK–1(298 K) 2.6 102 eV
  
1.602  1019 J(eV)1
Comparing it with the first excited state rotational energy of CO (viz., 4.76 Q 10–4 eV), obtained
in the last example, we find that the thermal energy is far greater. Hence, we conclude that at room
temperature, all the molecules in the gaseous sample of carbon monoxide are in the excited
rotational energy levels.
Relative Intensities of Rotational Spectral Lines
The relative intensities of spectral lines depend upon the relative populations of the energy levels.
Since, the energy level population is given by the Boltzmann distribution, the intensity of rotational lines
is evidently proportional to the Boltzmann distribution of molecules in the rotational energy levels, i.e.,
Intensity  N J/N =
0
e–EJ/kT
Rotational energy levels are, however, degenerate, their degeneracy (gJ) for a diatomic molecule
being given by
gJ = 2J + 1 ...(15)
In other words, for a given value of J, the energy level is (2J + 1) fold degenerate. For J = 0, g J
= 1; for J = 1, gJ = 3; for J = 2, gJ = 5; and so on.
Hence, the intensity of the rotational spectra is determined by the product of the degeneracy factor
and the Boltzmann exponential factor. Thus,
Intensity  N J/N 0 = (2J + 1)e–EJ/kT ...(16)
Since, EJ = hcF(J) ...(17)
and F(J) = BJ(J + 1) ...(18)
we have, NJ/N 0 = (2J + 1)e –BJ(J + 1)hc/kT
...(19)
The quantity NJ/N0 is ploteed versus J for a rigid diatomic molecule at room temperature in fig.

Plot of The Relative Boltzmann Population Versus J for A Rigid Diatomic Molecules
We see that the relative intensity passes through a maximum. It can be shown that the value of
J corresponding to the maximum in population is given by
 kT 1/2 1
Jmax     (20)
 2hc B  2 .........................................................................................................

The Jmax should be rounded off to the nearest integral value.


Ex. Calculate Jmax for a rigid diatomic molecule for which at 300 K, the rotational constant is 1.566
cm–1.
 kT 1/2 1
Jmax   
Sol. 2hc B  2
[From Eq. (20)]

 (1.38  1023 JK1 )(300 K) 


1/ 2
1
 34 1 1 
 = 7.56 = 8
 2(6.626 10 Js)(3  10 cms )(1.566 cm ) 
10
2
Rotational Spectra of Polyatomic Molecules
The diatomics have only one bond distance and hence only one moment of inertia, polyatomic
molecules have more than one bond distance and hence several moment of inertia. In fact, the Cartesian
coordinate system (x, y, z) there are nine components of the moment of inertia.
However, in the so-called principal axis system (a, b, c), these components reduce to three
moments of inertia Ia, Ib, Ic, where Ia, Ib and Ic are, respectively, the principal moments of inertia about the
a, b and c axes. Thus, polyatomic molecules in microwave spectroscopy are classified into different types
of rotators (rotors) as shown in Table.
Table : Classification of Polyatomic Molecules
Moment of Inertia Type of Rotor Example
Ia = Ib, Ic = 0 Linear HX, O=C=S, H–CN
Ia = I b = I c Spherical top CH4, SF6, UF6
Ia < I b = I c Prolate symmetric top NH3, CHCl3, CH6Cl
Ia = I b < I c Oblate symmetric top
Ia  I b  I c Asymmetric top H2O

n
c
a
s
e
o
f
n
o
n
-
li
n
e
8. VIBRATIONAL (INFRARED) SPECTRA OF DIATOMIC MOLECULES

A diatomic molecule with atomic masses m1 and m2 joined by a chemical bond vibrates as a one-
dimensional simple harmonic oscillator (SHO). Classically, the vibrational frequency of a mass point m
connected by a spring of force constant k is given by
1/ 2
1  k 
v    ...(21)
2  m 

In the case of a diatomic molecule, the masses m1 and m2 vibrate back and forth relative to their
centre of mass in opposite directions fig.

Fig. : Vibration of A Diatomic Molecule.


The two masses reach the extreme of their respective motions at the same time. The vibrational
frequency of the molecule is given by a relation analogous to that of Eq. (21) with m replaced by the
reduced mass 

v  1  k  s1
1/2

...(22)
2   

Two convert the frequency v from s–1 (hertz) to cm–1, we divide by c, the velocity of light. Thus,

1/2
1  k 
v   cm1 ...(23)
2c  

The units of the force constant are dyn cm–1 in the c.g.s. system and Nm–1 in SI system. The
potential energy of the simple harmonic oscillator (SHO) as a function of displacement from the equilibrium
configuration is given by the parabolic Hookeís law equation, viz.,
1 1
V(x)  k(r  r )2  kx2 ...(24)
e
2 2
where x(= r – re) is the displacement, re being the equilibrium bond length. The Hookeís law
potential energy is shown in fig.
Fig. : The Hookeís Law Potential Energy for A Diatomic Vibrator.
The solution of the Schrodinger equation for a simple harmonic oscillator gives the quantized
vibrational energy levels,
 1 
E  v  hv '; v  0, 1, 2, 3... ...(25)
v  
2
 

where, v is the vibrational quantum number and ví is the vibrational frequency given by Eq. (22).
1
Notice that the energy of the lowest vibrational level of the oscillator is not zero but is equal to hv ; this
2
is called zero-point energy (Z.P.E.). The energy levels of the SHO are equally spaced, the spacing being
equal to hv. To convert the energy from joules to cm –1 we divide by hc, obtaining
E  1v  1
G(v)  v  v   v  ...(26)

 

hc 2c  2 e
   

where, e is the equilibrium vibrational frequency. G(v) is called the vibrational term.
The selection rule for a vibrational transition in the simple harmonic oscillator is
v = ±1 ...(27)
The operative part of the selection rule for the absorption spectrum is v = + 1, i.e., the vibrational
quantum number changes by unity. Using the selection rule, the frequency of the vibrational transition is
given by
 1  1
v  G(v  v  1)  v  1   v   ...(28)
 2 e  2 e e

   

At room temperature, most of the molecule are in the ground vibrational state (v = 0) so that the
only transition of interest is that which takes place from v = 0 to v = 1. The vibrational frequency
corresponding to this transition is called fundamental vibrational frequency. Diatomic molecules have
only one vibrational frequency. This is a stretching vibrational frequency. Diatomic molecules do not have
a bending vibrational frequency.
Vibrational spectra are observed in the infrared (IR) region. There is a very important requirement
for a molecule to show an infrared spectrum. It states that the dipole moment of the molecule must change
during the vibration. Homonuclear diatomic molecules such H2,O2, N2, etc., do not have a
permanent dipole moment nor does the stretching of the bond between the two atoms change the dipole
moment from zero. Hence, homonuclear diatomics do not show IR spectra. On the other hand,
heteronuclear diatomic molecules such as CO, NO, CN, HCI, do possess a dipole moment which changes
when the bond length changes. Hence, they show IR spectra. Thus, homonuclear diatomics are IR-
inactive while heteronuclear diatomics are IR-active.
In 1929, P.M. Morse suggested an empirical expression for the potential energy of an harmonic
diatomic oscillator, given by
V(r) = De [1 – exp a{r – re}]2 ...(29)
Here, a is a constant and De is the dissociation energy of the molecule. The Morse potential energy
is sketched in fig.
Fig. : Morse Potential Energy for An Anharmonic Diatomic Oscillator.
It may be pointed out that the dissociation energy De in the Morse potential energy curve is
measured from the bottom of the potential well. The experimentally measured dissociation energy D 0 is
measured from the ground level (v = 0) to the top. Thus,
De = D0 + hv/2 ...(30)
When the Schrodinger equation for an anharmonic oscillator is solved using the Morse potential
energy, the energy levels are given by
1 
2
 1 
G(v)   v      v   e xe ; (v = 0, 1, 2, 3, ...) ...(31)
 2 e  2 

where, exe is called the anharmonicity constant. As expected exe << e. The consequence of
an harmonicity is that the vibrational energy levels of the SHO are all slightly lowered and the spacing
between them is no longer constant but goes on steadily decreasing with increase in vibrational quantum
number. Also, because of anharmonicity, the rule for vibrational transitions is no longer v = ± 1. Instead,
transitions corresponding to v = ± 2, ± 3, etc., are also observed in the IR spectra. These are called
the first overtone, the second overtone, etc., respectively. The intensity of an overtone is dependent on
the anharmonicity of the vibration. Compared with the highly intense fundamental vibrational frequency,
the overtones are very weak, i.e., of considerably low intensity.
We see that kT < E. Hence, we conclude that most the molecules are in the ground vibrational
state at room temperature. This situation is in contrast with the rotational states where, kT > E and
consequently most of the molecules are in the excited rotational states at room temperature.
Ex. The force constant of CO is 1840 N m–1. Calculate the vibrational frequency in cm–1 and the spacing
between the vibrational energy levels in eV. Compare this spacing with the thermal energy at room
temperature and comment on your result. The atomic masses are 12C = 19.9 Q 10–27 kg; 16O = 26.6
Q 10–27 kg, 1 eV = 8066 cm–1.
Sol. The reduced mass of CO is given by

m1m2

m1  m2

(19.9  1027 kg)  (26.6  1027 kg)



(19.9  26.6)  1027 kg

= 11.4 Q 10–27 kg
1/ 2
1  k  1
v   
2c    2(3.1416)(3  1010 cms1)

 1840 kg s2   
 27  2140 cm1
 11.4 10 kg 

The spacing between energy levels,


E = Ev+1 – Ev = hv joules

hv v
   v  2140 cm1
hc c
Since, 1 eV = 8066 cm–1
2140 cm1
 E   0.265 eV
8066 cm1 / eV
(1.38  1023 JK1)(300 K)

Thermal energy  kT  

1.602  1019 J / eV
= 2.6 Q 10–2 eV = 0.026 eV.
Ex. For the molecule BH,  = 2368 cm–1 and the anharmonicity  x = 49 cm–1. Calculate the
e e e
vibrational terms of the first four vibrational levels and determine the spacing between them.
Sol. From Eq. 31
2
 1  1
G(v)   v   e   v   e xe
 2   2 

2
 1   1
G(0)   0    2368   0    49
 2   2

= 1184 – 12 = 1172 cm–1

2
 1   1

G(1)  1   2368  1    49


 2  2 

= 3552 – 110 = 3442 cm–1

 1  1
G(2)  2   2368  2   49
 2   2 
   
= 5920 – 306 = 5614 cm–1
 1  1
G(3)  3   2368  3   49
 2   2 
   
= 8288 – 600 = 7688 cm–1
The spacing between the energy levels are :
G(1) – G(0) = 3442 – 1172 = 2770 cm–1
G(2) – G(1) = 5614 – 3442 = 2172 cm–1
G(3) – G(2) = 7688 – 5616 = 2074 cm–1
Notice that the energy levels are not equally spaced and the spacing goes on steadily decreasing
with increase in the vibrational quantum number.
9. ELECTRONIC SPECTRA

The electronic band spectra of molecules are observed in the ultraviolet and visible regions of the
electromagnetic spectrum. Their complexity arises from the fact that a transition between two electronic
states is almost invariably accompanied by simultaneous transitions between the vibrational and rotational
energy levels as well. This is expressed by saying that electronic spectra have vibrational fine structure
and rotational fine structure. According to the Born-Oppenheimer approximation, the total energy of
molecule in the lower (ground) state is given by
E” = E”el + E"vib + E"rot ...(47)
neglecting the translational energy, E”tr, which is not quantized. Here, E"el, E"vib and E''rot are
respectively, the electronic, vibrational and rotational energies. Assuming that the Born-Oppenheimer
approximation is valid in the upper (excited) state as well, the excited state energy E' is given by
E' = E'el + Eívib + E'rot ...(48)
The energy change for an electronic transition is given by
E = Eí – E” = (Eíel – E”el) + (Eívib + E”vib) + (Eírot – E”rot) ...(49)
= Eel + Evib + Erot ...(50)
Considerable simplification of spectra results by recognizing that
Eel >> Evib >> Erot ...(51)
The frequency for the electronic transition is given by the Bohr frequency condition, viz.,

E Eel  Evib  Erot 1


v  cm ...(52)
hc hc

Eq. 52 shows how an electronic transition possesses the vibrational and rotational fine structure.
Franck-Condon Principle
A very useful guiding principle for investigating the vibrational structure of electronic spectra is
provided by the well known Franck-Condon principle which states that an electronic transition takes place
so rapidly that a vibrating molecule does not change its internuclear distance appreciably during the
transition. This principle is, to a first approximation, true since the electrons move so much faster than
the nuclei that during the electronic transition the nuclei do not change their position. Hence, an electronic
transition may be represented by a vertical line on a plot of potential energy versus the internuclear
distance.
Let us demonstrate the Franck-Condon principle for the electronic transition of a diatomic molecule.
Consider fig. where we have shown two potential energy curves for the molecule in the ground electronic
state (E0) and in the first excited electronic state (E1). Since, the bonding in the excited state is weaker
than in the ground state, the minimum in the potential energy curve for the excited state occurs at a
slightly greater inter-nuclear distance than the corresponding minimum in the ground electronic state. Also,
quantum mechanically it is known that the molecule is in the centre of the ground vibrational levelof the
ground electronic state. Thus, when a photon falls on the molecule, the most probable electronic transition,
according to the Franck-Condon principle, takes place from v" = 0 to v' = 2 (written schematicallyas 0  2).
Transitions to other vibrational levels of the excited electronic state occur with lower probabilitiesso that their
relative intensities are smaller than the intensity of the 0  2 transition, as shown in fig.
Fig. : Electronic Spectrum of A Diatomic Molecule.
Consider now a slightly different case of a diatomic molecule which has a potential energy curve
E0 in ground state and two potential energy curves E1 and E2 in excited states.
The potential energy curve E2 does not have a minimum, as shown in fig.
The equilibrium internuclear distance is longer in E 1 than in E0 since the bond is weaker in the
excited state. The transition indicated by arrow a has, according to the Franck-Condon principle, maximum
intensity. Other possible transitions are from the first excited vibrational level of the ground electronic state;
they are indicated by the duplication of the vertical arrows b 1 and b2. The transition b1 ends in the lowest
vibrational level of the first excited electronic state where the molecule is still held firmly. However,the
transition b2 promotes the molecule to a point where its energy is far above the potential energy plateau
of state E1 at large distance with the result that the molecule dissociates.
The molecule will also dissociate if it is excited to a level below the plateau in E 1 but above point
c where the potential energy curves for E1 and E2 cross. At an inter-nuclear distance corresponding to
the point c, the molecule may undergo a crossover from one state to the other without energy change and
continue to dissociate.
Types of Electronic Transitions
Organic compounds, particularly those containing groups like C C, C O, N N
extensively conjugated systems, form a special class of polyatomic molecules whose electronic spectra
are amenable to simple interpretation even though the investigation of their detailed spectral features may
require knowledge of quantum mechanics and group theory. On the basis of the molecular orbital theory
(MOT), the electrons can be classified as  ,  or n (non-bonding) depending upon the MOs they occupy.
For organic carbonyl compounds, the electronic transitions involve promotion of the electrons in n,  n,
 and  orbitals in the ground state to  and  ABMOs in the excited state (Fig.). In other words, only the

  
transitions of the type   ,    and   n are allowed. Since electrons in the n orbitals are not

involved in bond formation, there are no ABMOs associated with them.

The    transitions occuring in saturated hydrocarbons and other types of compounds in

which all valence shell electrons are involved in single bonds, are found in ultraviolet region since they
involve very high energy.
The    and the   n transitions, on the other hand, are found either in the UV or visible

regions. The unsaturated molecules containing C C and C O groups (such as aldehydes and ketones)
show   n and   n transitions. For aldehydes and ketones the more intense band near

1800 # (180 nm) is due to    transition and the weaker band around 2850 # (285 nm) is due to

  n transition. Olefinic hydrocarbons show a    transition in the wavelength 160-170 nm. Acetylene
shows an absorption near 180 nm.
Functional groups such as C C, N N which absorb at wavelengths longers than 180 nm,
are called chromophores. Other chromophores are nitro, nitroso, carbonyl, thiocarbonyl, sulphoxide
groups as well as aromatic rings.
Molecules such as methyl amine and methyle iodide which do not contain a  orbital, show
  n transitions.
In extensively conjugated systems, the -electrons are delocalized over the entire skeletal frame-
work. Such systems are treated in terms of the ëfree-electron modelí and it is found that the absorption
band shift to longer wavelengths as the extent of conjugation increases. Thus, in the compound

C6H5 (CH CH)n C6H5 ,    transitions lies in the UV region when n = 1 or 2. As n increases,
the electronic transition shifts to the visible region. This phenomenon is shown schematically in fig.

Fig. : Energy level diagram showing that in a conjugated -electron system, the wavelength of
absorption maximum is directly proportional to the extent of conjugation.
The intensity of an electronic band is determined by the extent of overlap of the wave functions
in the group and excited states. Since, there is very poor overlap of wave functions of the ground and
excited states in the   n transition and there is considerable overlap of the corresponding wave

= 2.6 Q 10–2 eV = 0.026 eV.


Ex. For the molecule BH,  = 2368 cm–1 and the anharmonicity  x = 49 cm–1. Calculate the
e e e
vibrational terms of the first four vibrational levels and determine the spacing between them.
Sol. From Eq. 31
2
 1  1
G(v)   v   e   v   e xe
 2   2 

2
 1   1
G(0)   0    2368   0    49
 2   2 

= 1184 – 12 = 1172 cm–1

2
 1   1

The term auxochrome refers to an atom or a group of atoms which does not give rise to an
absorption band on its own but, when in conjugation with a chromophore, causes a bathochromic shift

and a hyperchromic effect. For instance, C C group is a chromophore in ethylene; when one of

the hydrogens is replaced by a halogen atom, a bathochromic shift and a hyperchromic effect are
produced. This is because the lone pair on the halogen atom conjugates with the alkene double bond. We
conclude that the halogen atom acts as an auxochrome.
As said earlier. solvent effects are important in identifying    and   n tansitions. Polar

solvent often stabilize the ground state   n transition more than the excited state, causing blue shifts
   transition the excited state is stabilized more than the ground
(fig.(a)). On the other hands, for
state, causing red shifts (fig.(b)).
Fig. : Solvent effects on (a)   n and (b)    transitions.

One of the most important applications of UV spectroscopy, based on the Lambert-Beer law, is
in quantitative analysis. According to the Labert-Beer law, the absorbance A of a compound of concentration
c is given by A = bc where, b is the path length and E is the molar extinction coefficient. Consider the
following equilibrium X Y . If X and Y have absorption bands that overlap and if at some wavelength
over the overlapped region both X and Y have equal absorbance, then there will be no change in the
absorbance at this wavelength as the ratio [Y]/[X] is varied. This gives rise to the so-called isosbestic point.
When there is more than one isosbestic point present, we have
A  b((1)c  (1)c )
1 X X YY

A  b( c   c )
(2) (2)
2 X X Y Y

where, 1 and 2 indicate the two wavelengths. Since, A1 and A2 can be measured, the concentrations
of X and Y, viz., cX and cY (and hence the equilibrium constant K = [Y]/[X] can be calculated if the various
extinction coefficients (1), (2), (1) and (2) are known.
X X Y Y

Finally, we may mention that UV spectroscopy has an advantage over IR spectroscopy as far as
applications to biological systems are concerned. Most biological systems have high water contents and
since water absorbs strongly in the IR, its use is severely limited. On the other hand, aqueous solutions
can be conveniently investigated by UV spectroscopy. UV spectra have also better resolution than IR
spectra of biochemical systems.
Electronic Spectra of Conjugated Molecules
The electronic spectra of conjugated systems using free-electron molecular orbital (FEMO) theory
is based on the theory proposed in the 1930s by Pauling and, independently, by Kathleen Lonsdale and
H. Kuhn, that the pi-electrons are delocalized over the entire carbon skeletal framework Thus, there is no
true carbon-carbon single bond or carbon-carbon double bond and we can use the particle-in-a-one
dimensional box-approach for such systems. We know that for an electron in a one-dimensional box with
width equal to a and infinite height, with potential energy equal to zero everywhere inside the box,
En = n2h2/(8m ea2) (n = 1, 2, 3...) ... (53)
The electrons in the MOs are filled according to the Pauli principle and Hund's rule. Thus, if N
is the number of -orbitals, only N/2 -orbitals are filled in the ground state.
The longest wavelength transition involves the excitation of the -electron from the N/2 HOMO
to (N/2) + 1 LUMO. Hence, the frequency of the electronic transition is given by

E E h  N 
2
 N  
2

 1    
v (N/2)1 N/2
 

 

h 8m e a2   2   2  
= h(N + 1)/8m a2 ...(54)
e

 v  v / c  h(N  1) / (8me a2c) ...(55)


Let us now estimate the value of the chain length a of the conjugated system (i.e., the width of the
box). Since, there is a residual free valence at each of the two terminal carbon atoms, hence
a = number of actual CóC bond between N carbon atoms + 2
= (N – 1 + 2)l = (N + 1)l ...(56)
where, l is the CóC bond length, assumed to be equal to 139 pm. Here, we have made
allowance for the molecular orbitals to extend on bond length l past each terminal carbon atom which
is a very plausible assumption. Substituting for a Eq. 56, we have
Table : Pascalís Triangle giving Intensities of NMR Signals

Number of protons, n Number of NMR lines Relative Intensities


0 1 1
1 2 11
2 3 121
3 4 1331
4 5 14641
5 6 1 5 10 10 5 1
6 7 1 6 15 20 15 6 1
It may be noted that the spin-spin interaction is independent of the applied magnetic field strength
whereas the chemical shift depends on the field strength.
The proton NMR spectrum of a molecule thus gives information about
(a) The number of peaks which enables us to know about the kinds of protons present in a
molecule.
(b) The positions of the peaks which tell us about the electronic environment of each kind of
proton.
(c) The intensities of the peaks which tell us about the number of protons of each kind that
are present and
(d) The splitting of a peak into several peaks which tells us about the environment of a proton
with respect to other nearby protons in the molecule.
Ex. Predict the proton NMR spectrum of ethyl acetate using the three rules mentioned earlier.
Sol. The molecule has the structure
CH3óCOóOóCH2óCH3
(a) From rule 1, there is no splitting due to spin-spin interaction of the protons in either of the
two methyl groups or in the methylene group.
(b) From rule 2, the CH2 protons split the CH3 peak into three lines (i.e., a triplet) and the CH3
protons split the CH2 peak into four lines (i.e., a quartet).
(c) From rule 3, the intensities of the CH3 triplet are given by the coefficients of the expansion
(1 + x)2 = 1 + 2x + x2, i.e., by the ratio of 1 : 2 : 1 and the intensities of the CH quartet
3 are
given by the coefficients of the expansion (1+x)3 = 1 + 3x + 3x2 + x3, i.e., by the ratioof 1 :
3 : 3 : 1.
Chemical Shift-Equivalent and Magnetically Equivalent Nuclei
If nucleus have exactly the same chemical shift, they are called chemical shift-equivalent nuclei
and designated by the same letter. Chemical shift-equivalent nuclei can be interchanged by one of the
symmetry operations of a molecule. Thus, all the protons in benzene are chemical shift equivalent since
they can be interchanged by the six-fold symmetry axis of benzene. If nuclei are coupled in exactly the
same way to every other nucleus in the molecule, they are called magnetically equivalent nuclei. Thus,
in para-fluoronitrobenzene, the protons (a) and (b) are chemical shift-equivalent since they are related
by the mirror plane which bisects the molecule.
1
The nuclear spin of F is I  . Since, F has identical bond distances and angles with respect to
2
protons (a) and (b), these protons are identically coupled to F. However, protons (a) and (b) are not
identically coupled to protons (c) and (d). Hence, they are not magnetically equivalent. Similarly, protons
(c) and (d) are not magnetically equivalent.
Another way to decided the equivalence of two protons is by replacing each alternately by a group
X. If the X derivatives are the same, the protons are equivalent. If the replacement of each H in a CH 2
gives 2 diastereomers, the protons are not equivalent.
Ex. Determine which of the following molecules will show spin-spin coupling in their NMR spectra.If
the coupling or splitting is observed, give the multiplicity of each kind of proton, i.e., give the number
of lines arising form each proton :
(a) ClCH2CH2Cl (b) ClCH2CH2I

H H H Cl
(c) C C (d) C C
Br Br Br H

is split into a doublet and the –CH=O resonance is split into a 1 : 3 : 3 : 1 quartet.
Spin-Spin Coupling Constant
The magnitude of the coupling J is also intimately related to the geometrical relationship of the
bonds over which the coupling is transmitted. Thus, in the system HóCóCóH, the HóH coupling
constant, known as the vicinal coupling constant, depends upon the dihedral angle  between the CóH
bonds. The angle  is the angle between the projections of the two CóH bonds viewed along the CóC
bonds axis. It is found that
Jvic = 4.22 – 0.5 cos  + 4.5 cos 2 ...(76)
The is called the Karplus relation.
The variation of the vicinal coupling constant, Jvic, with the dihedral angle  between the two
vicinal CóH bonds, is shown in fig.
Fig. : Variation of vicinal coupling constant and dihedral angle.
Some typical J-values are given in table.
Table : Spin-Spin Coupling Constants of Some Systems

Largest vicinal couplings arise with the protons in the trans coplanar position ( = 180c). Vicinal
coupling for cis coplanar protons are almost are large ( = 0). Very small coupling arise between protons
which are 90c to each other.
NMR Spectra of Other Nuclei
1
Besides protons, other nuclei having nuclear spin I  m, such as 13C, 19F, 29Si and 31P, have also
2
been investigated by NMR spectroscopy. The NMR spectra of 1H and 19F are most easily obtained
because of their high natural abundance and large magnetogyric ratios. Spectra of other nuclei provide
structural information about compounds in the same way as the proton magnetic resonance (p.m.r.)
spectra do. While 1H-chemical shifts extend only over a small range of the order of 15 ppm, 31P-chemical
shifts have a span of about 400 ppm and 19F-chemical shifts have a still wider span of 600 ppm. This wider
range can be attributed to the greater number and mobility of extranuclear electrons which producegreater
variation in diamagnetic shielding.
In recent years, the 13C-NMR spectroscopy has emerged as good structural tool although there
is considerable difficulty in recording the 13C-NMR spectra. This is due to the fact that has a very small
natural abundance and a small value of the magnetogyric ratio.
(d) HI
1. The MOT
(a) puts equal importance on both ionic and 7. In momomeric BH3, let an axis
covalent structures definition place the molecule in the xz
(b) overestimates the importance of ionic plane. Which atomic orbital on B is non-
structures bonding ?
(c) underestimates the importance of (a) 2s
covalent structures (b) 2px
(d) none of the above (c) 2py
(d) 2pz
2. The VBT
(a) underestimated the importance of 8. How many molecular orbitals may be
covalent structures constructed from the valence shell orbitals
(b) overestimated the importance of of the constituent atoms in CH4 ?
covalent structure (a) 7
(c) underestimates the importance of ionic (b) 8
structure (c) 4
(d) puts equal emphasis on the both ionic (d) 6
and covalent structures
9. The Pauling electronegativity of
3. If electronegativity difference between hydrogen is 2.1. Use the definition of
atoms A and B is 1.5, percentage ionic Pauling electronegativity to estimate the
character of A – B bond is electronegativity of chlorine given that the
(a) 8 dissociation energy of an H –CL bond is
(b) 24.22 427 Kj mol-1 , for a H –H bond is 432 Kj
(c) 28.22 mol-1 and for a Cl –Cl bond is 239 Kj mol-1
(d) 31.88 (a) 3.0
(b) 1.1
4. Hydrogen bonding is maximum in (c) 2.1
(a) (CH3)3N (d) 4.2
(b) CH3Cl
(c) (C2H5)2O 10. Use molocular orbital theory to
(d) C2H5OH determine the bond order for the O+2 ion.
(a) 3
5. The non-polar molecule is (b) 1 ½
(a) cis 1,2-dichloroethane (c) 2 ½
(b) 1,3-dichlorobenzene (d) 2
(c) NH3
(d) tans 1,2-dichloroethene 11. The number of ionic terms present in
the molecular orbital wave function for the
6. The dipole moment is maximum for bond between H and Cl in HCl (assuming
(a) HF that the bond is formed from the 1s
(b) HCL electron of H atom and a 3p electron fo Cl
(c) HBr atom) is
(a) 0 (d) C2
(b) 1
(c) 2 16. For a planar unsaturated hydrogen
(d) 3 having formula CxHy, where, all the carbon
are part of the usaturated frame work, how
12. For n-chain atoms, as the n increases many pi MOs are there ?
the HOMO- LUMO separation (a) 1 – y
(a) increases (b) x
(b) decreases (c) y
(c) first increase then decreases (d) 2
(d) remains the same
17. The paramagnetic molecule is
13. Which of the following statements is (a) O2
correct ? (b) N2
(a) Bonding molecular orbital have higher (c) Cl2
energy than the combining atomic orbitals (d) ClO3
(b) bonding molecular orbital has high
electron orbital density than the anti 18. The molecule containing one unpaired
bonding molecular orbitals electron is
(c) bonding molecular orbitals are the (a) CO
result of substractive interaction (b)CN-
(d) all of the above (c) NO
(d) O2
14. The number of p-function
perpendicular to the molecular plane in 19. The compound which gas zero dipole-
case of butadiene is moment is/are
(a) 2 (a) BF3
(b) 3 (b) NF3
(c) 4 (c) PF2Cl2
(d) 6 (d) CH2Cl2

15. According to the LCAO-MO model, 20. Which of the following compounds
which one of the following second period possesses zero dipole moment ?
diatomic molecules has a double bond in (a) benzene
the ground electronic state ? (b) carbon tetrachloride
(a) Li2 (c) boron trifluoride
(b) Be2 (d) water
(c) B2
Answer
1.b 2.c 3.d 4.d

5.d 6.a 7.b 8.b

9.a 10.c 11.c 12.b

13.b 14.c 15.d 16.b

17.a 18.c 19.a,c 20.a,b

You might also like