You are on page 1of 22

Acta Appl Math (2012) 118:324

DOI 10.1007/s10440-012-9675-5
Applications of Symmetry and Group Theory
for the Investigation of Molecular Vibrations
Jaan Laane Esther J. Ocola
Received: 25 April 2011 / Accepted: 16 November 2011 / Published online: 8 February 2012
Springer Science+Business Media B.V. 2012
Abstract The application of symmetry and mathematical group theory is a powerful tool for
investigating the vibrations of molecules. In this paper, we present an overview of the meth-
ods utilized. First we briey discuss the quantum mechanical nature of vibrations and the
experimental methods used. We then present the principal concepts for applying group the-
ory to molecules. The symmetry operations which are used to comprise groups are described
and then used to determine the point groups of molecules. The properties of character tables
are presented and the method for obtaining a reducible representation for all the motions of
a molecule is detailed. This can then be broken down to obtain the irreducible representa-
tion which contains the symmetry species of the individual vibrations. The determination
of symmetry adapted linear combinations is outlined and the basis for spectroscopic selec-
tion rules is presented. The paper concludes by examining how matrix algebra along with
symmetry concepts simplies calculations with molecular force constants.
Keywords Symmetry Group theory Molecular vibrations Potential energy Symmetry
adapted linear combinations Selection rules
Mathematics Subject Classication (2000) 20-01 58D19
1 Introduction
One of the notable developments in chemistry during the second half of the twentieth century
was the emergence of symmetry and group theory as a means to better understand molecular
spectroscopy, molecular bonding, and molecular reactions and dynamics. The well known
book series by Nobel laureate Gerhard Herzberg [13] published in 1945 to 1966 already
used these concepts to present the principles of molecular spectroscopy, but this was done
in a fairly complicated way. A major advance came when our late colleague F.A. Cotton
at Texas A&M University rst published his book Chemical Applications of Group Theory
J. Laane () E.J. Ocola
Department of Chemistry, Texas A&M University, College Station, TX 77843-3255, USA
e-mail: laane@mail.chem.tamu.edu
4 J. Laane, E.J. Ocola
Fig. 1 Natural symmetry of a
snowake
in 1963. The third edition [4] appeared in 1993. This text presented symmetry and group
theory in a way that the average chemist could understand. Another very good and early
book, although somewhat more advanced, was the text by Tinkham [5]. Following these,
numerous other books of varying depth have been published on this topic [620].
In the present paper we will attempt to present a condensed version of how the use of
symmetry and group theory can be applied to better understand molecular vibrations and
to simplify the calculations that are utilized for these purposes. Everyone has some kind of
conception of symmetry. Thus, when looking at the snowake in Fig. 1, it is evident that it
can be cut in half many ways leaving two identical sides. Similarly it can be rotated by 60

after which it looks just the same. These same ideas can be applied to molecules as we shall
see.
This paper is organized as follows. In Sect. 2, we present some basic concepts for the
study of molecular vibrations. In Sect. 3, we examine the applications of symmetry and
group theory to molecules. In Sect. 4, we describe the implementation of symmetry adapted
linear combinations for the representation of vibrations. Section 5 describes the origin of
spectroscopic selection rules based on symmetry, and Sect. 6 examines the effect of symme-
try on the vibrational potential energy. A brief conclusion is given in Sect. 7.
2 Molecular Vibrations
2.1 Vibrations of a Diatomic Molecule
A two-atom or diatomic molecule can have six different individual motions. It can translate
in three directions, it can rotate about two axes, or it can have a vibration where the distance
between the atoms rapidly gets larger or smaller (perhaps 10
14
times per second). There are
only two molecular rotations since one of the Cartesian axes passes through the atoms and
hence there is no moment of inertia about this axis (I =

m
i
r
2
i
where m
i
= mass of atom
and r
i
= distance from axis). Our interest is in the molecular vibration, which is generally
approximated by the harmonic oscillator model. Here, the bond is pictured to be similar to
a spring so that the potential energy function for the vibration can be written
V =
1
2
kx
2
(1)
Applications of Symmetry and Group Theory for the Investigation 5
where x = R R
e
is the bond displacement from its equilibrium position R
e
and k is a
force constant (or spring constant). Use of this potential function in the Schrdinger wave
equation


2
2
d
2

dx
2
+V =E (2)
allows the quantum energy levels E to be calculated along with their corresponding wave
functions. The constant =
h
2
where h is the well known Plancks constant, and the are
the wave functions. The reduced mass is given by
=
m
1
m
2
m
1
+m
2
(3)
where m
1
and m
2
are the masses of the two atoms in the molecule. Equation (2) can be
solved exactly to yield
E =

v +
1
2

h and
v
=N
v
H
v
(
1
2
x)e

x
2
2
(4)
where the frequency is given by
=
1
2

1
2
(5)
and the quantum number v = 0, 1, 2, . . . , N
v
is a normalization constant, H
v
is the Hermite
polynomial of degree v, and =
(k)
1
2

. The infrared absorption spectrum of a diatomic


molecule is generally presented on a wavenumber scale in cm
1
() where
=

c
=
1
2c

1
2
(6)
and typically the v = 0 1 quantum transition is observed at this value since all of the
energy levels are calculated to have a spacing of h.
2.2 Vibrations of Water
For larger molecules the vibrational problem becomes multi-dimensional, requiring 3N 6
coordinates for an N atom non-linear molecule. Of the 3N degrees of freedom, six represent
the translations and rotations. Vibrational motions are best described by internal coordinates
such as molecular bonds and bonding angles. As an example, Fig. 2 shows the three internal
coordinates for the two-bond stretching and one-angle bending motions of water. For this
case the quadratic potential energy function is
2V = f
R
(R
1
)
2
+f
R
(R
2
)
2
+f

()
2
+2f
RR
(R
1
)(R
2
)
+2f
R
()(R
1
) +2f
R
()(R
2
), (7)
where R
1
and R
2
represent the stretching of the two bonds and where is the bend-
ing of the HOH angle. The force constants f
R
and f

are the primary constants for the


stretching and bending motions, respectively, and f
RR
and f
R
are interactive constants.
The detailed methods for calculating the vibrational frequencies of polyatomic molecules
were nicely presented by Wilson, Decius, and Cross [21].
6 J. Laane, E.J. Ocola
Fig. 2 Internal coordinates of
water
Fig. 3 Infrared and Raman
transitions for a molecular
vibration
2.3 Experimental Methods
There are two primary methods for studying molecular vibrations which generally have their
quantum state separations in the 100 to 4000 cm
1
range. Figure 3 shows the possible tran-
sitions for a diatomic molecule. Infrared absorption experiments typically involve a direct
transition fromthe vibrational ground state (v = 0) to the rst excited quantumstate (v = 1).
Many commercial instruments are available and these utilize the concept of Fourier trans-
form infrared (FT-IR) spectrometry where two light beams are separated by a beam splitter
and then combined to give an intensity versus path difference data set. Fourier transform is
then used to convert this to an intensity versus wavenumber spectrum.
Raman spectroscopy utilizes a monochromatic laser such as an argon-ion laser at
514.5 nm wavelength or a frequency doubled Nd:YAG laser at 532 nm. The laser induces the
sample to produce a quasi-excited or virtual state which then scatters the radiation isotrop-
ically as either Rayleigh light or Raman radiation. The Rayleigh scattering, which has the
same frequency as the laser exciting line, is about a million times more intense than the
Raman scattering which is shifted in frequency. Raman spectra are generally collected as
Stokes scattering when molecules are excited from the lowest ground state vibrational level
and these have frequencies of

RAMAN
=
LASER

VIBRATION
(8)
so that the vibrational frequency can be determined. As will be discussed later, the selection
rules for which transitions can be observed are different for infrared and Raman spectra. As
an example of how the two types of spectra differ, Fig. 4 presents the observed spectra for
an organic molecule, 3-cyclopenten-1-one. This molecule has 30 vibrations and the bands
corresponding to many of them are shown in the gure. Note that both types of spectra are
needed in order to observe most of the vibrations.
Applications of Symmetry and Group Theory for the Investigation 7
Fig. 4 Infrared and Raman
spectra for the
3-cyclopenten-1-one molecule
Table 1 Symmetry elements and
operations for molecular systems
Element Symbol Operation
Identity E Nothing
Proper rotation axis C
n
Rotate by 360/n n = 2, 3, 4, . . .
Plane of symmetry Reection through plane
(
v
= vertical;
h
= horizontal;

d
= dihedral)
Center of inversion i Inversion though center
Improper rotation axis S
n
C
n
followed by
h
3 Applications of Symmetry and Group Theory
The principal concepts for applying mathematical group theory to chemical systems are the
following:
1. Symmetry operations performed on molecules make up the members of a mathematical
group.
2. All the properties and requirements for a mathematical group apply. This includes the
fact that the product of any two operations must also be a member of the group and
that any operation X must have an inverse such that XX
1
= E where E is the identity
operation.
3. A valid symmetry operation performed on a molecule produces a result where the
molecule appears to be identical to its original position.
4. Symmetry operations are carried out with respect to symmetry elements.
Table 1 presents a list of the symmetry elements and the corresponding symmetry opera-
tions. Examples of these will be shown for the benzene molecule (C
6
H
6
) which is a hexagon
of six carbon atoms with one hydrogen atom attached to each. Figure 5 rst shows the op-
erations about the C
6
proper rotation axis. A C
6
operation is a rotation of 60

, a C
2
6
= C
3
operation is a rotation by 120

, etc. Note that C


6
6
gets you back to the original arrangement
of atoms, so this is equivalent to the identity E. Figure 6 shows the two types of C
2
rotation
axes which benzene also possesses. Note that there are a total of six C
2
axes perpendicular
to the C
6
axis and this will be the basis for the molecule having D
6h
point group symmetry.
8 J. Laane, E.J. Ocola
Fig. 5 Symmetry operations about the C
6
axis for benzene
Fig. 6 The two types of C
2
axes
for benzene which are
perpendicular to the C
6
axis
Fig. 7 The three types of
reection planes for benzene
Fig. 8 The inversion (top) and
improper rotation (bottom)
operations of benzene
Figure 7 shows three different planes of symmetry,
h
(horizontal),
v
(vertical), and
d
(dihedral). Figure 8 shows the effect of the inversion and improper rotation operations. Note
that whether the atoms wind up right side up or inverted is important. In total, benzene has
Applications of Symmetry and Group Theory for the Investigation 9
T
a
b
l
e
2
C
h
a
r
a
c
t
e
r
t
a
b
l
e
a
n
d
r
e
d
u
c
i
b
l
e
r
e
p
r
e
s
e
n
t
a
t
i
o
n

f
o
r
t
h
e
b
e
n
z
e
n
e
m
o
l
e
c
u
l
e
o
f
D
6
h
s
y
m
m
e
t
r
y
D
6
h
E
2
C
6
2
C
3
C
2
3
C
2
3
C

2
i
2
S
3
2
S
6

h
3

d
3

v
A
1
g
1
1
1
1
1
1
1
1
1
1
1
1
x
2
+
y
2
,
z
2
A
2
g
1
1
1
1

1
1
1
1
1

1
R
z
B
1
g
1

1
1

1
1

1
1

1
1

1
1

1
B
2
g
1

1
1

1
1
1

1
1

1
1
E
1
g
2
1

2
0
0
2
1

2
0
0
(
R
x
,
R
y
)
(
x
z
,
y
z
)
E
2
g
2

1
2
0
0
2

1
2
0
0
(
x
2

y
2
,
x
y
)
A
1
u
1
1
1
1
1
1

1
A
2
u
1
1
1
1

1
1
1
z
B
1
u
1

1
1

1
1

1
1

1
1

1
1
B
2
u
1

1
1

1
1

1
1

1
1
1

1
E
1
u
2
1

2
0
0

1
1
2
0
0
(
x
,
y
)
E
2
u
2

1
2
0
0

2
1
1

2
0
0

3
6
0
0
0

4
0
0
0
0
1
2
0
4
10 J. Laane, E.J. Ocola
Fig. 9 Symmetry operations for
the ammonia molecule
demonstrating the product rule
Table 3 Product table for C
3v
C
3v
E C
3
C
2
3

v
(A)
v
(B)
v
(C)
E E C
3
C
2
3

v
(A)
v
(B)
v
(C)
C
3
C
3
C
2
3
E
v
(B)
v
(C)
v
(A)
C
2
3
C
2
3
E C
3

v
(C)
v
(A)
v
(B)

v
(A)
v
(A)
v
(B)
v
(C) E C
3
C
2
3

v
(B)
v
(B)
v
(C)
v
(A) C
2
3
E C
3

v
(C)
v
(C)
v
(A)
v
(B) C
3
C
2
3
E
24 symmetry operations that can be performed, and these are shown as the heading for the
D
6h
character table in Table 2. Note that the coefcients for the operations either indicate
how many there are of the same type (3C

2
, 3C

2
, 3
d
, 3
v
) or how many can be performed
about the same axis (2C
6
=C
6
, C
5
6
; 2C
3
=C
3
, C
2
3
; 2S
6
=S
6
, S
5
6
; 2S
3
=S
3
, S
2
3
). Utilization
of character tables will be discussed later.
One of the rules for mathematical groups is that the product of any two operations is also
a member of the group. Figure 9 shows the ammonia molecule (NH
3
) which belongs to the
C
3v
point group which has two C
3
operations and three reection planes. The effect of a
C
3
operation followed by a reection through plane A is shown. It is also shown that this
product is equivalent to a reection through a different plane B. In other words C
3

v
(A) =

v
(B). Table 3 presents the product table for C
3v
symmetry. Note that each product is a
member of the group and that each has an inverse.
As can be seen, it is important to identify the symmetry point group of a molecule. Table 4
summarizes the possible point groups for molecules. Scheme 1 shows in a concise way how
to determine the appropriate point group. This scheme precludes having a molecule of S
n
symmetry which is extremely rare.
Scheme 1
1. Decide if molecule is octahedral (O
h
), tetrahedral (T
d
), or linear (D
h
with center of
symmetry; C
v
with no inversion center).
2. If not, look for highest order proper axis (C
n
). If present, go to 4.
3. If no C
n
, look for which gives a C
s
point group. If no , the molecule has C
1
or the
minimal symmetry.
4. Look for nC
2
axes perpendicular to the highest order C
n
axis. If present, go to 7.
5. Look for
h
reection plane perpendicular to C
n
. If present, the symmetry is C
nh
.
Applications of Symmetry and Group Theory for the Investigation 11
Table 4 Summary of point groups
Point group Important symmetry elements Order of the group
C
1
E 1
C
i
i 2
C
s
2
C
n
C
n
n
S

n
S
n
(very rare) n
C
nv
C
n
,
v
2n
C
nh
C
n
,
h
2n
D
n
C
n
, C
2
(very rare) 2n
D
nd
C
n
, C
2
,
d
(also S
2n
) 4n
D
nh
C
n
, C
2
,
12
4n
C
v
Linear molecules without center of inversion
D
h
Linear molecules with center of inversion
T
d
Terahedral symmetry 24
O
h
Octahedral symmetry 48

n must be even, or else S


n
=C
nh
Fig. 10 Examples of molecules
with octahedral, tetrahedral, and
linear symmetries
6. If no
h
, look for n
v
planes. If present, the symmetry is C
nv
. If not, the symmetry point
group is C
n
.
7. Look for
h
. If present, the symmetry is D
nh
.
8. If no
h
, look for n dihedral planes bisecting C
2
axes. If present, the symmetry is D
nd
. If
not, it is D
n
.
Figure 10 shows O
h
, T
d
, D
h
, and C
v
molecules and lists some of their operations.
Figures 11 and 12 show C
2v
, C
3v
, D
3h
, and D
2h
molecules along with the steps which lead
to their identication.
We now turn to the use of character tables. Based on Schurs lemma in the representation
theory of nite groups [22, 23], matrix elements of irreducible representations of a nite
group satisfy certain orthogonality relations. Here the great orthogonality theorem applies.
As given by Nowick [24] it can be stated as follows:
Theorem 1 (The Great Orthogonality Theorem [24]) Express the th irrep of a group by
the set of matrices [D
()
(R)]
ij
where R is one of the operations of the group G, and take the
12 J. Laane, E.J. Ocola
Fig. 11 Examples of C
2v
and C
3v
molecules
Fig. 12 Examples of D
3h
and D
2h
molecules
dimensionality of this irrep to be l

. Then the components of the various irreps of group G


obey the equation

R
[D
()
(R)]

ij
[D
()
(R)]
pq
=(h/l

ip

jq
where denotes the complex conjugate, h is the order of the group G and the s are all
Kronecker s.
This rather complicated equation may be interpreted as follows. The fact that the order
of the group is h means that each irrep has h matrices. Suppose we treat the corresponding
elements of these matrices, for example [D
()
(R)]
ij
for given i and j, as we vary the group
operation R, as a vector in h-dimensional space. There are l
2

such vectors for irrep and

l
2

in total. But group theory also tells us that

l
2

=h, so that there are h such vectors


in all. All members of this collection of h vectors are orthogonal to each other (dened so
that complex conjugates are used in the dot products, where necessary). Further it tells us
that the magnitude of each other vector (the dot product with itself) equals h/l

.
Applications of Symmetry and Group Theory for the Investigation 13
Fig. 13 Effect of a C
2
operation
on the translational motions of
water
Table 5 The character table for the C
2v
symmetry point group. The reducible representation for water is
also given
C
2v
E C
2

v
(xz)

v
(yz)
A
1
1 1 1 1 z x
2
, y
2
, z
2
A
2
1 1 1 1 Rz xy
B
1
1 1 1 1 x, Ry xz
B
2
1 1 1 1 y, Rx yz
9 1 3 1
Next we turn to the characters
()
(R) of the various irrep matrices. From the denition
of the character, it readily follows that the characters obey their own orthogonality relation

()
(R)

()
(R) =h

Again, we may think of the characters of each irrep as forming an h-dimensional vector.
These vectors are all orthogonal to each other and that the magnitude of each is equal to h,
the order of the group.
The sum of the diagonal elements of a matrix representation is called a character. The
number of irreducible representations of a nite group is equal to the number of conjugacy
classes of that group. Each character takes a constant value on a conjugacy class. We will
use the water molecule (Figs. 2 and 13) of C
2v
symmetry as an example. Table 5 shows
the character table for any C
2v
molecule which has C
2
,
v
, and

v
operations in addition
to the identity. The character table lists four symmetry species and each has either a +1 or
1 value for the operations shown. A positive value means that the motion (or orbital) of
interest is symmetric with respect to that operation while a 1 means it is antisymmetric.
This is readily understood if we examine the translational and rotational motions of water.
Figure 13 shows the effect of a C
2
operation on the three translational motions. The C
2
axis is taken to be the z direction and the molecule is in the xz plane. It is evident that the
C
2
operation will leave the direction of T
z
unchanged, but the vectors of T
x
and T
y
change
direction. Hence, we can place a +1 character under C
2
for the T
z
operation while T
x
and
14 J. Laane, E.J. Ocola
Table 6 Characters for molecular motions of water
C
2v
E C
2

v
(xz)

v
(yz)
T
z
1 1 1 1 A
1
T
x
1 1 1 1 B
1
T
y
1 1 1 1 B
2
R
z
1 1 1 1 A
2
Vibrations:
Symmetric stretch 1 1 1 1 A
1
Antisymmetric stretch 1 1 1 1 B
1
Bend 1 1 1 1 A
1
Fig. 14 Effect of C
2
and
v
operations on a rotation of the
water molecule
Fig. 15 The three vibrations of
water: symmetric stretching
(A
1
), antisymmetric stretching
(B
1
), and angle bending (A
1
),
respectively
T
y
have 1 characters. The
v
(xz) reection leaves T
z
and T
x
in the same direction (+1
characters) while T
y
switches direction. Similarly

v
(yz) leaves T
z
and T
y
in their original
direction but the direction of T
x
is changed. Table 6 summarizes these results and shows that
the T
z
, T
x
, and T
y
characters match symmetry species A
1
, B
1
, and B
2
, respectively. Note
that in Table 5 this fact is already identied by the location of the x, y, and z to the right of
the character table. Figure 14 shows the effect of the C
2
and
v
operations on the rotation
about the z axis and Table 6 presents the characters for each of the operations. These show
R
z
to be of symmetry species A
2
, and Table 5 can be seen to identify each of the rotational
symmetries.
When a vibrational description is known, we can use this + or method to also deter-
mine the symmetry species for vibrations. Figure 15 shows the three vibrations of water. The
behavior of each motion (symmetric or antisymmetric) upon each operation is also summa-
Applications of Symmetry and Group Theory for the Investigation 15
Fig. 16 The effect of a C
4
operation on the Cartesian
coordinates of PtCl
2
4
rized in Table 6. For example, the antisymmetric stretch is antisymmetric with respect to the
C
2
and

v
(yz) operations, but symmetric to reection in the
v
(xz) plane and thus has B
1
symmetry. The other two vibrations have A
1
symmetry species.
We now address how symmetry operations are represented mathematically. Figure 16
shows the Pt Cl
2
4
molecular ion of D
4h
symmetry. The fteen Cartesian coordinate vectors
needed to represent all the motions for the ve atoms are also shown, and the effect of a C
4
operation on the positions of all these vectors is presented. A matrix for the C
4
operation
can be used to transform a column matrix of the original coordinates to the new coordinate
positions, and this is shown below.

0 0 0 0 1 0 0 0 0 0 0 0 0 0 0
0 0 0 1 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 1 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 1 0 0 0 0 0 0 0
0 0 0 0 0 0 1 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 1 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 1 0 0 0 0
0 0 0 0 0 0 0 0 0 1 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 1 0 0 0
0 1 0 0 0 0 0 0 0 0 0 0 0 0 0
1 0 0 0 0 0 0 0 0 0 0 0 0 0 0
0 0 1 0 0 0 0 0 0 0 0 0 0 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 1 0
0 0 0 0 0 0 0 0 0 0 0 0 1 0 0
0 0 0 0 0 0 0 0 0 0 0 0 0 0 1

x
1
y
1
z
1
x
2
y
2
z
2
x
3
y
3
z
3
x
4
y
4
z
4
x
5
y
5
z
5

y
2
x
2
z
2
y
3
x
3
z
3
y
4
x
4
z
4
y
1
x
1
z
1
y
5
x
5
z
5

(9)
This shows that x
1
is moved to where y
2
was previously, y
1
goes to x
2
, and so on as
a result of the C
4
operation. Fortunately, it is only necessary to have the character or trace
(sum of the diagonal elements) of the C
4
matrix in order to analyze molecular motions. Note
that for PtCl
2
4
the character for the C
4
operation is +1 and this arises from the fact that the
z
5
vector on the Pt atom has not moved. In all cases, contributions to the character only
come from unshifted atoms when an operation is performed. Although vectors x
5
and y
5
are
also on the Pt atom, these do not contribute to the character since they are turned 90

from
their original direction. What simplies matters a great deal is that each type of operation
produces the same contribution to the character per unshifted atom. For a C
4
operation this
is always +1 since one vector goes into itself and the other two do not contribute. Figure 17
shows the effect of E, , and C
2
operations on any unshifted atom and why they contribute
+3, +1, and 1, respectively, to the character. Table 7 shows the contribution per unshifted
atom for all types of the operations.
We will now apply this approach to the motions of water using Table 5. For the C
2
and

(yz) operations only the oxygen atom is unshifted while all three atoms are unshifted
16 J. Laane, E.J. Ocola
Fig. 17 Effect of E, , and C
2
operations on the Cartesian
coordinates of an unshifted atom
Table 7 Contribution per unshifted atom (for atom vectors)
Operation Contribution Operation Contribution
E 3 (any) 1
C
2
1 i 3
C
3
0 S
3
2
C
4
1 S
4
1
C
5
1 +2cos(72

) S
5
1 +2cos(72

)
C
6
2 S
6
0
C
L
1 +2cos(360

/L) S
L
1 +2cos(360

/L)
for the (xz) reection. Utilizing the contribution per unshifted atom information in Table 7
then results in the reducible representation in Table 5. This can then be broken down to an
irreducible representation consisting of the symmetry species as components. The following
formula is used:
n

1
g

j
g
j
x

j
x
j
g
, (10)
where
= particular symmetry species such as A
1
or A
2
,
n

= number of species represented in reducible representation ,


g =

g
j
= total number of symmetry operations (add coefcients in heading),
g
j
= number of operations in class j (coefcient for particular column);
this is often one (implied),
x

j
= number in character table for species for operation class j;
this is +1 or 1 for non-degenerate species,
x
j
g
= character in reducible representation for operation j.
Applications of Symmetry and Group Theory for the Investigation 17
As an example, to calculate how many of the A
2
symmetry species are in , we see that
each g
j
is one and g = 4. Then
n
A
2
=
1
4
[1 1 9 +1 1 1 +1 1 3 +1 1 1] = 1. (11)
The total result is that the 3N = 9 motions of water can be described as
= 3A
1
+A
2
+3B
1
+2B
2
. (12)
Table 5 shows the symmetry species for the translations and rotations are

TRANS
=A
1
+B
1
+B
2
. (13)
and

ROT
=A
2
+B
1
+B
2
. (14)
When these are subtracted from above, we are left with the symmetry species for the
vibrations as

VIB
=
TRANS

ROT
= 2A
1
+B
1
. (15)
These vibrations were shown in Fig. 15.
We will now apply this approach to the 12 atom benzene molecule using Table 2 where
the reducible representation is also shown. results from examining the number of un-
shifted atoms for each operation and then using Table 7 to get the total contribution. The
result for the 36 motions of benzene is
=2A
1g
+2A
2g
+2B
2g
+2E
1g
+4E
2g
+2A
2u
+2B
1u
+2B
2u
+4E
1u
+2E
2u
. (16)
Note that E type symmetry species represent two motions. Table 2 shows the translations
to be of symmetry A
2u
+E
1u
and rotations to be A
2g
+E
1g
. Thus the symmetry species for
the 30 vibrations of benzene are

VIB
=
TRANS

ROT
= 2A
1g
+A
2g
+2B
2g
+E
1g
+4E
2g
+A
2u
+2B
1u
+2B
2u
+3E
1u
+2E
2u
. (17)
Previously in Fig. 4 we presented the spectra of an organic molecule of C
2v
symmetry.
When symmetry considerations are applied here, we nd the vibrational symmetry species
are

VIB
= 10A
1
+5A
2
+9B
1
+6B
2
. (18)
When the vibrations are numbered, as they are in Fig. 4, we start with the rst symmetry
species and label the vibrations from highest frequency down. The rst vibration (
1
) is
then the highest frequency A
1
mode which happens to be a CH bond stretching vibration
with frequency 3079 cm
1
. Vibration
10
is a bending motion of the ve-membered ring
of A
1
symmetry at 623 cm
1
. The next vibration numbered is
11
which is the highest
frequency vibration with A
2
symmetry and so on. The last numbered vibration
30
is the
lowest frequency B
2
motion at 83 cm
1
.
18 J. Laane, E.J. Ocola
4 Symmetry Adapted Linear Combinations (SALCs)
In our discussion of water we pointed out that vibrations are best represented by internal
coordinates as shown in Fig. 2. We now show how the application of symmetry enables us to
provide better descriptive pictures of the vibrations. During a molecular vibration equivalent
bonds or angles stretch or bend simultaneously, but not necessarily in the same direction.
Thus, for water, the two oxygen-hydrogen molecular bonds R
1
and R
2
will either stretch
at the same time or one will stretch while the other contracts. This is shown in Fig. 15.
The method of symmetry adapted linear combinations (SALCs) allows us to calculate how
equivalent internal coordinates combine. The formula is
S

Op

Op
Op(In
1
), (19)
where S

is the SALC with symmetry species ,

Op
is the value in the character table
for species and operation Op, and Op(In
1
) is the resulting internal coordinate when the
operation Op is performed on the initial internal coordinate of choice In
1
, and is the
normalization constant. This process is shown below when the bond stretching coordinate
R
1
of water is selected as In
1
:
C
2v
E C
2

v
(xz)

v
(yz)
Op(R
1
) R
1
R
2
R
1
R
2
(20)
Thus R
1
stays in its own position upon E and
v
(xz) operations but moves to where R
2
was
upon C
2
and

v
(yz) operations. The result above in combination with (19) yields
S
A
1
= (2R
1
+2R
2
), (21)
S
B
1
= (2R
1
2R
2
), (22)
and
S
A
2
=S
B
2
=0. (23)
The normalization condition requires that

c
2
i
= 1, (24)
where the c
i
are the coefcients for the internal coordinates. Thus, (21) and (22) become
S
A
1
1
=2

1
2
(R
1
+R
2
) (25)
and
S
B
1
3
=2

1
2
(R
1
R
2
) (26)
The have been inserted to indicate the vibrations actually involve the change in the bond
distances R
1
and R
2
. The B
1
SALC is labeled as S
3
since
S
A
1
2
= (27)
is numbered rst since it belongs to A
1
symmetry. In this case, there is only one angle to
bend and has nothing to combine with so the SALC is the same as the internal coordinate.
Applications of Symmetry and Group Theory for the Investigation 19
Fig. 18 The effect of symmetry
operations on the internal
coordinates of ethylene
Fig. 19 SALCs representing the CH bond stretching and angle bending vibrations of ethylene
To further demonstrate SALCs we examine the CH bond stretching and HCH angle
bending internal coordinates of ethylene shown in Fig. 18. The effect of each operation is
shown. When used together with a D
2h
character table, the SALCs shown in Fig. 19 result.
For this molecule and in many cases, the SALCs shown are actually very good representa-
tions for the so-called normal coordinates which are the precise descriptions of the vibra-
tional motions. Mixing or vibrational coupling of motions of the same symmetry species
is allowed, but no interactions between different symmetry species may occur.
5 Spectroscopic Selection Rules
The vibrational spectra of molecules involve transitions between quantum states and these
were shown for infrared absorption and Raman scattering transitions in Fig. 3. For an in-
frared transition to occur the molecule must have a change in dipole moment (a change
20 J. Laane, E.J. Ocola
in charge distribution) during the vibration, and this can be determined by symmetry con-
siderations. Namely, the following transition moment integral R
if
must not be zero for an
allowed transition:
R
if
=

f
M

i
d = f ||i. (28)
The expression on the right is in the more concise Dirac notation. Here,
i
and
f
are
the wave functions for the initial and nal states of the transition, and

f
is the complex
conjugate of
f
. Also here M

is the dipole moment expression for the direction ( =


x, y, or z) in terms of the vibrational coordinate Q:
M

=M
0

0
Q+

2
M

Q
2

0
Q
2
+ . (29)
Then
f ||i =M
0

f |i +

2
M

Q
2

0
f |Q|i + . (30)
The rst term is zero due to the fact that the wave functions are orthogonal. It can readily
be shown that R
if
is non-zero only if the direct product of the symmetries of
i
, Q, and

f
is totally symmetric where all the characters for the symmetry species are +1 (A
1
for
C
2v
symmetry, for example). For vibrations the initial state is totally symmetric and the
symmetry species for Q is the same as T
x
, T
y
, or T
z
. Thus for R
if
to be non-zero,
f
must have the same symmetry species as T
x
, T
y
, or T
z
. This is because the direct product
of a symmetry species times itself is totally symmetric (e.g., B
1
B
1
= A
1
for C
2v
). This
provides a simple way to identify infrared active vibrations since they are those that have an
x, y, z at the right hand side of a character table. For C
2v
these are A
1
, B
1
, and B
2
. For D
6h
they are A
2u
and E
1u
.
For Raman transitions the polarizability of a molecule must change during a vibration.
This essentially means that the shape of the molecule must change. This is generally true
for more symmetric vibrations but often not true for antisymmetric motions such as the
antisymmetric CO bond stretching of carbon dioxide. Mathematically, the product of two
integrals must be non-zero:
P
if
= f ||jj||i, (31)
where i and f are the initial and nal states and j is the virtual state shown in Fig. 3.
For the second integral to be non-zero j must have the same symmetry as since i is
totally symmetric. Then in the rst integral the direct product of and j must match that
of the vibrational level f . Hence, the vibration must have the same symmetry as the tensor
element to be Raman active. Each character table lists these at the right side of the
table. For C
2v
all of the symmetry species correspond to Raman active vibrations. For D
2h
symmetry only the A
1g
, E
1g
, and E
2g
vibrations are Raman active. In Raman experiments
different vibrations can also be determined to be either Raman polarized or depolarized
depending on how the intensities vary with laser beam polarization. As it turns out, only the
totally symmetric vibrations have polarized Raman bands. There are also other spectroscopic
processes such as two-photon absorption and hyper Raman for which the selection rules can
be determined in a similar way.
To demonstrate the application of selection rules we will examine the predicted spectral
features of two organic molecules with the same formula but belonging to different symme-
try point groups. The cis- and trans-dichloroethylene molecules of C
2v
and C
2h
symmetry,
Applications of Symmetry and Group Theory for the Investigation 21
Fig. 20 Cis- and
trans-dichloroethylene molecules
Table 8 Infrared and Raman bands expected for isomers of 1,2-dichloroethylene
Isomer Symmetry Raman bands Raman polarized Infrared bands
cis C
2v
12 5 10
trans C
2h
6 5 6
respectively, are shown in Fig. 20. The cis molecule has vibrations

cis
(C
2
) = 5A
1
+2A
2
+4B
1
+B
2
, (32)
IR, R(P) R IR,R IR,R
while the trans has

trans
(C
2h
) = 5A
g
+B
g
+2A
u
+4B
u
. (33)
R(P) R IR IR
For each symmetry species, the infrared (IR) and/or Raman (R) activity are shown and
the polarized Raman bands [R(P)] are also indicated. Table 8 summarizes the number of
bands expected for the infrared and Raman spectra. As can be seen, the two molecules are
easy to distinguish based on symmetry selection rules. The experimental spectra for these
molecules can be found elsewhere [25].
6 Effect of Symmetry on Potential Energy
We now wish to show how symmetry helps to simplify the computation of vibrational fre-
quencies and force constants. We will use the water molecule with its internal coordinates
shown in Fig. 2 as an example. First we note that the internal coordinates can be represented
by a column matrix R which is dened by a transformation from Cartesian coordinates X:
R =BX. (34)
For water X is a column matrix of the nine Cartesian coordinates required to dene the
positions of the three atoms. B is the 3 9 transformation matrix which will not be shown
here. For water,
R =

R
1
R
2

. (35)
22 J. Laane, E.J. Ocola
These internal coordinates can be transformed to symmetry coordinates using
S =UR, (36)
where for water
S =

S
1
S
2
S
3

(37)
is based on the denitions in (25)(27) so that
U=

1
2
2

1
2
0
0 0 1
2

1
2
2

1
2
0

. (38)
The potential energy is given by
2V = R

FR. (39)
For water the potential function in terms of internal coordinates is given in (7) so that
F =

f
R
f
RR
f
R
f
RR
f
R
f
R
f
R
f
R
f

. (40)
This can be converted to a symmetric set of force constants using the similarity transforma-
tion
F
sym
= UFU

. (41)
For water this becomes
F
sym
=

F
11
F
12
0
F
21
F
22
0
0 0 F
33

(f
R
+f
RR
) 2
1
2
f
R
0
2
1
2
f
R
f

0
0 0 (f
R
f
RR
)

. (42)
Then in terms of symmetry coordinates we have
2V = S

F
sym
S. (43)
The solution of the vibrational problem also requires the consideration of the kinetic energy
which can be written as
2T = R

G
1
R. (44)
where G
1
is the reciprocal of the G matrix which is determined from the geometry of the
molecule and the masses of the individual atoms. Details for its calculation can be found
elsewhere [21]. The same symmetry considerations apply to the G matrix as they do for F.
Namely
G
sym
= UGU

. (45)
Applications of Symmetry and Group Theory for the Investigation 23
Once the F and G matrices are available, the vibrational eigenvalues can be calculated and
then related to the vibrational frequencies :
=4
2

2
. (46)
In internal coordinates,
|GF E| =0, (47)
where E is the identity matrix. For water this would be a 3 3 determinant that needs
to be solved (or matrix diagonalization methods can be used). However, use of symmetry
coordinates results in
|G
sym
F
sym
E| = 0, (48)
and the resulting determinant can be broken up in terms of the individual symmetry species.
For water a 22 determinant for A
1
results and a simple 11 determinant for the B
1
block.
For larger molecules this symmetry factoring is even more signicant. For example, for
benzene and its vibrations given in (17), instead of having a 30 30 matrix or determinant
to deal with, symmetry factoring results in ten separate symmetry blocks with the largest
being a 4 4 matrix.
7 Conclusions
In this paper we have tried to convey how powerful the application of symmetry and group
theory is to the analysis of molecular vibrations. However, we have only touched the surface
as the understanding of electronic and rotational spectroscopies is also greatly facilitated by
similar methods. We have also not delved deeply into demonstrating how the computation
of vibrational frequencies and force constants is simplied to a great extent by symmetry
considerations. Symmetry and group theory are also invaluable for understanding molecular
bonding based on molecular orbital theory where molecular orbitals are generated from the
linear combination of atomic orbitals (MO-LCAO theory).
Acknowledgements The authors wish to thank the Robert A. Welch Foundation (Grant A-0396) for nan-
cial support. Yi-Ching Wang assisted with adapting the manuscript to LaTeX format.
References
1. Herzberg, G.: Molecular Spectra and Molecular Structure: II. Infrared and Raman Spectra of Polyatomic
Molecules. Van Nostrand, Princeton (1945)
2. Herzberg, G.: Molecular Spectra and Molecular Structure: I. Spectra of Diatomic Molecules. Van Nos-
trand, Princeton (1950)
3. Herzberg, G.: Molecular Spectra and Molecular Structure: III. Electronic Spectra and Electronic Struc-
ture of Polyatomic Molecules. Van Nostrand Reinhold, New York (1966)
4. Cotton, F.A.: Chemical Applications of Group Theory, 3rd edn. Wiley, New York (1990)
5. Tinkham, M.: Group Theory and Quantum Mechanics. McGraw-Hill, New York (1964)
6. Kettle, S.F.A.: Symmetry and Structure: Readable Group Theory for Chemists, 3rd edn. Wiley, Sussex
(2007)
7. Bunker, P.R., Jensen, P.: Fundamentals of Molecular Symmetry. Institute of Physics, Bristol (2005)
8. Molloy, K.C.: Group Theory for Chemists: Fundamental Theory and Applications. Horwood, Chichester
(2004)
9. Ogden, J.S.: Introduction to Molecular Symmetry. Oxford University Press, London (2001)
24 J. Laane, E.J. Ocola
10. Kim, S.K.: Group Theoretical Methods and Applications to Molecules and Crystals. Cambridge Univer-
sity Press, Cambridge (1999)
11. Walton, P.: Beginning Group Theory for Chemistry. Oxford University Press, London (1998)
12. Carter, R.L.: Molecular Symmetry and Group Theory. Wiley, New York (1997)
13. Narang, K.: Group Theory in Chemistry and Physics. Prints India, New Delhi (1994)
14. Tsukerbiat, B.S.: Group Theory in Chemistry and Spectroscopy: A Simple Guide to Advanced Usage.
Academic, London (1994)
15. Harris, D.C., Bertolucci, M.D.: Symmetry and Spectroscopy: An Introduction to Vibrational and Elec-
tronic Spectroscopy. Dover, New York (1989)
16. Flurry, R.L. Jr.: Symmetry Groups: Theory and Chemical Applications. Prentice-Hall, Englewood Cliffs
(1980)
17. Jaffe, H.H., Orchin, M.: Symmetry in Chemistry. Krieger, Huntington (1977)
18. Vincent, A.: Molecular Symmetry and, Group Theory: A Programmed Introduction to Chemical Appli-
cation. Wiley, New York (1977)
19. Nakazaki, M.: Chemistry and Symmetry: Introduction to Group Theory. Nankodo, Tokyo (1976)
20. Bishop, D.M.: Group Theory and Chemistry. Oxford University Press, London (1973)
21. Wilson, E.B., Decius, J.C., Cross, P.C.: Molecular Vibrations: The Theory of Infrared and Raman Vibra-
tional Spectra. McGraw-Hill, New York (1955)
22. Curtis, C.W., Reiner, I.: Representations of Finite Groups and Associate Algebras. Wiley-Interscience,
New York (1962)
23. Serre, J.P.: Linear Representations of Finite Groups. Springer, Heidelberg (1977)
24. Nowick, A.S.: DThe Great Orthogonality Theorem, pp. 204205. Cambridge University Press,
Columbia University, New York (1995)
25. Chen, G., Laane, J., Wheeler, S.E., Zhang, Z.: Greenhouse gas molecules: a mathematical perspective.
Not. Am. Math. Soc. 58, 14211434 (2011)

You might also like