Sunteți pe pagina 1din 23

SS 2007

SCRIPT OC IV
Qualitative MO Theory and its Application
to Organic Reactions, Thermal Rearrangements,
Pericyclic Reactions.



Prof. Dr. Peter Chen







Lecture assistants:

Dborah Mathis
HCI G214 tel. 24489
mathis@org.chem.ethz.ch

Jonas Hostettler
HCI G201 tel. 24493
hostettler@org.chem.ethz.ch


2
QUALITATIVE MO THEORY




I. Introduction

A. Since the mid-sixties, quantum mechanics have had an increasingly large influence
on organic chemistry. This section will serve as an introduction to qualitative mo-
lecular orbital theory at a level sufficient to understand the use of quantum me-
chanics in selection rules like the Woodward-Hoffmann rules. This is not an in-
troduction to ab initio calculations. That is a subject best left to another course.
In some senses, ab initio calculations don't tell us any chemistry -- what one gets
is an energy or a geometry, and that must then be subjected to interpretation. Mo-
lecular orbital theory at the level of the Hckel approximation is quite adequate
for the sections to follow.


B. Outline

1. From the Schrdinger equation to secular determinants.

2. The Hckel approximation and t-theory.

3. First-order perturbation theory.

4. Symmetry and group theory.

3

II. Main Body

A. From the Schrdinger equation to secular determinants.

Example of a bond formed from two orbitals.

This example will serve to introduce most of the fundamental quantum mechanical
concepts that we'll need later. It is also a derivation that everyone should see once.
Let's start with the time-independent Schrdinger equation for the electronic wave-
function (i.e. we've already made the Born-Oppenheimer assumption).

H = E

where H is the Hamiltonian operator, is the electronic wavefunction, and E is the
electronic energy. This equation, in which an operator times a function equals a scalar
times the same function is an example of an eigenvalue problem. E is the eigenvalue
of the Hamiltonian operator and is the eigenfunction. The form of the operator H is:

V
m
+ V
2
2
2
!
H

where the first term is the electronic kinetic energy term, and the second is the poten-
tial energy. V contains within it terms like:


12
2
r
Ze


which is just a Coulomb potential term for two particles with charges +Z and -1, at a
distance r
12
. The Laplacian operator in the kinetic energy term, in Cartesian coordi-
nates, is:


2
2
2
2
2
2
2
z y x c
c
+
c
c
+
c
c
V




4
For any operator, the Hamiltonian operator in particular, the most probable value of a
wavefunction , termed the expectation value for the energy, in this case, can be
found by:


} }
- -
= t t d E d H

Let's start with two normalized orbitals |
1
and |
2
, which could be hydrogen 1s func-
tions for the H
2
o-bond, or carbon 2p functions if this exercise were for the t-bond in
ethylene.

For the wavefunction , we approximate it by a linear combination of atomic orbitals.
(LCAO):


2 2 1 1
| | c c + ~

where c
1
and c
2
are scalar constants to be determined. To determine these constants,
termed orbital coefficients, we make use of the variation principle, which states:

For a normalized, well-behaved trial wavefunction that satisfies the
boundary conditions for the problem, the expectation value for the energy
will be greater than or equal to the lowest eigenvalue for the system, or:


0
d E
-
>
}
H

Application of this principle to solving for the orbital coefficients is straightforward.
One differentiates the energy with respect to the coefficients, sets the derivative equal
to zero, and solves.

Notation:


| |
|
i j i j
i j i j ij
d H
d S
| | t | |
| | t | |
-
-


}
}
H H
ij


We start with the expression below, using our LCAO wavefunction:




|
| | H
= E
5


Expanding it out explicitly in terms of the atomic wavefunctions:


2
2 12 2 1
2
1
22
2
2 12 2 1 11
2
1
2 2 1 1 2 2 1 1
2 2 1 1 2 2 1 1
2
2
|
| |
c S c c c
H c H c c H c
c c c c
c c c c
E
+ +
+ +
=
+ +
+ +
=
| | | |
| | | | H


Differentiating with respect to c
1
, and setting the derivative equal to zero to find the
minimum of the function, we get:


( )( )
( )
( )
2
2 12 2 1
2
1
12 2 1 12 2 11 1
2
2
2 12 2 1
2
1
12 2 1 22
2
2 12 2 1 11
2
1
2
2 12 2 1
2
1
12 2 11 1
1
2
2 2 2 2
2
2 2 2
2
2 2
0
c S c c c
S c c E H c H c
c S c c c
S c c H c H c c H c
c S c c c
H c H c
c
E
+ +
+ +
=
+ +
+ + +

+ +
+
=
c
c
=


which will be zero if the numerator is zero. This, and an analogous treatment for dif-
ferentiation with respect to c
2
, gives:


( )
( )
1 11 2 12 12
1 12 12 2 22
( )
( )
c H E c H ES
c H ES c H E
+ =
+ =
0
0
| |
|
\ .

This set of homogenous linear equations in n unknowns can be written as a matrix
equation (eigenvalue problem):


11 12 12 1
12 12 22 2
0
0
H E H ES c
H ES H E c
| | | |
=
| |

\ . \ .

Such a system of linear equations has a non-trivial solution (all coefficients are not all
zero) only if the corresponding determinant vanishes. This determinant is called the
secular determinant, and generally has the form,


| |
j i ES H m
j i E H m
m
ij ij ij
ij ij
ij
= =
= =
=
if , and
if , with
M



6


In this case, the secular determinant gives a quadratic equation, which has two roots:


12
12 11
1 S
H H
E

= (provided that H
11
= H
22
)

The three integrals here have names which you'll run into eventually. H
ii
is called the
Coulomb integral, H
ij
is the resonance integral, and S
ij
is the overlap integral. The
Coulomb and resonance integrals are negative; the overlap integral is positive.

This expression for the energy gives the familiar splitting of two orbitals as they come
into interaction. Notice that, if the overlap integral were zero, then the splitting would
be symmetrical about the energy of the Coulomb integral. As it is, a nonzero overlap
integral raises the antibonding level more than the bonding level is lowered.

One can substitute the expression for the energy back into the expression for the en-
ergy, and solve for the orbital coefficients. In this case,


g antibondin
2
1
c ,
2
1
c
bonding
2
1
2 1
2 1
= =
= = c c


One important feature of the orbital coefficients is that they are normalized, i.e., for
= c
1
|
1
+ c
2
|
2
+ ..., the sum is equal to unity. Also, for any single coefficient, its
square is the probability that an electron occupying the LCAO MO is at that atomic
site.

2
i
c



7
B. The Hckel method.

The secular determinant can become very difficult to solve as the number of basis or-
bitals increases. In fact, the number of matrix elements goes up quadratically. Erich
Hckel proposed a series of approximations that make the secular determinant solv-
able. The approximations are quite severe, but the wavefunctions and energies that
come out are qualitatively correct. As a matter of terminology, in solid-state physics,
a Hckel calculation is called a "tight-binding" calculation. The approximations are:


1. o-t separation is good. This restricts the Hckel method to the t-systems in planar,
or at least nearly planar unsaturated compounds. The o-bonds in the molecule's skele-
ton will not interact with the t-bonds if they are of different symmetry with respect to
a plane of reflection.

2. All Coulomb integrals H
ii
have the same value, and are equal to the energy of a
carbon 2p orbital.

3. All resonance integrals H
ij
between adjacent centers have the same value, and are
equal to one-half the t-energy in ethylene.

4. All resonance integrals between nonadjacent centers are set to zero.

5. All overlap integrals S
ij
are set to zero.

In addition, the notation is typically used where the Coulomb integral is labelled o,
and the resonance integral |. The secular determinant can be rewritten using the sub-
stitution,


|
o E
x



so that our previous 2 x 2 secular determinant looks like this.


1
0
1
x
x
=

8
This gives a simple recipe for writing the secular determinant within the Hckel ap-
proximation. Follow these steps:

1. Set to zero the determinant of a square matrix of an order equal to the number of
t-centers in the molecule.

2. Set all diagonal elements equal to x.

3. Set all matrix elements between adjacent centers to unity.

4. Set all other matrix elements to zero.

The solutions of the polynomial expression will give the t-system eigenvalues; sub-
stitution back into the energy expression gives the p-orbital coefficients. Let's do two
examples of the solution of the secular determinant for t-systems. The first is allyl
radical, which is the simplest system to have non-bonding orbitals.
H
H H
H H
1
2
3

The carbon atoms are numbered as indicated on the structure. To write the secular de-
terminant within the Hckel approximation, we need only look at the connectivity.

( )
| o o 2 ,
2 , 0
0 2
0 2
0
1 0
1 1
0 1
2
3
=
=
=
=
=
E
x
x x
x x
x
x
x








9
From the energies, we can back-substitute to obtain the orbital coefficients.


0.707
0.707
0.707
- 0.707
- 0.5
- 0.5
0.5 0.5
o + 1.414|
o - 1.414|
o (nonbonding)
allyl radical


The orbitals are shown using the convention that we are looking at the planar t-
system from above, seeing only the phase of the top lobe.

We will consider the regularities in the Hckel solutions for linear polyenes a little
later. Another example is planar trimethylenemethane. This is a ground state triplet
molecule a biradical that has a half-filled degenerate pair of orbitals. For labelling,
the center carbon is C-1.




( ) ( )
| o o o 3 , ,
3 3
0 0
0 0
1 1 1
0 0
0 0
1 1 1
0 0
0 0
1 1 1
0 0
0 0
0 0
0 0 1
0 0 1
0 0 1
1 1 1
0
2 2 2 4 2 2 2 4
=
= = + =
+ =
=
E
x x x x x x x x
x
x
x
x
x
x
x
x
x
x
x
x
x

10

which gives a molecular orbital splitting pattern with the 4 t-electrons,


o + 1.732|
o - 1.732|
o


The picture here introduces the concept of an electronic configuration. A configura-
tion is just an assignment of electrons to available orbitals. This particular configura-
tion is for the ground-state triplet trimethylenemethane biradical. Because there are
two orbitals (nonbonding) with the same energy, and only two electrons to be distrib-
uted between them, there are four different ways to fill them in. Three of those are
singlet (spin-paired) configurations. One is a triplet with the spins parallel. By Hund's
rule, the triplet should be the lowest energy, although, in the Hckel approximation,
the four biradical configurations all have the same energy. Two orbitals, configura-
tions, or states that have the same energy are said to be degenerate. The ground con-
figurations are not the only ones. In filling the orbitals with electrons, we only have to
make sure that the Pauli principle is satisfied. A configuration does not imply that we
pick the lowest energy occupancy. If we don't, then the configuration is an electroni-
cally excited configuration. We'll see much more of this in photochemistry.

The solution for the orbital coefficients is tedious, so I won't do that explicitly. How-
ever, the results are shown pictorially above.

There are regularities in Hckel wavefunctions that are quite useful. For example, for
linear polyenes,

11
1. The eigenfunction with the lowest energy eigenvalue has no nodes.

2. Hckel orbitals are disposed such that their "center of gravity" is o.

3. For even-numbered polyenes, succeeding eigenfunctions have more nodes be-
tween atoms, and have no nonbonding orbitals.

4. Odd-numbered polyenes have one non-binding orbital.

5. Linear unbranched polyene eigenvalues can be expressed as:
N J , J
N
E
J
..., , 2 , 1
1
cos 2 =
|
.
|

\
|
+
+ =
t
| o
where E
j
is the Hckel energy of the J
th
orbital in a linear, unbranched polyene
with N t-centers.

6. In cyclic t-systems with Hckel overlap, the Hckel energies are given by the
polygon rule. Inscribing the polygon in a circle of radius 2| with a vertex at the
bottom gives a Hckel level at each vertex.

nonbonding level
2|
benzene cyclopentadienyl


We see here where the 4N-2 aromaticity rule comes from. For a cyclic array of orbi-
tals with no compulsory phase change, 4N-2 would constitute a closed shell configu-
ration in the Hckel picture. An interesting variant comes for cyclic overlap of orbi-
tals where there is one compulsory phase change. Solving the secular determinant for
this problem gives an analogue to the polygon rule, except that the polygon is in-
scribed with a side at the bottom. This manner of cyclic overlap has been termed
Mbius overlap because it has a twist in it like a Mbius strip. For a system with
Mbius overlap, an aromatic compound would be one with 4N electrons.



12
Application of Hckel/Mbius concepts to strained rings.

We have been considering t-systems. However, many of the ideas that have been
developed are more generally applicable, at least in a qualitative way. An example
is the quantum mechanical description of strained small-ring compounds like
cyclopropane. There had been a number of studies aimed at describing the bond-
ing in cyclopropane, which should be quite different from that in an unstrained
hydrocarbon. A model of three sp
3
-hybridized carbons bound by ordinary single
bonds is quite far off with regards to cyclopropane reactivity, bond angles, and
CH stretching frequencies. The HCH angle is much larger (~120), and the
CH stretch (3080 cm
-1
) much higher frequency than one would expect. The
Walsh model sets each carbon up with sp
2
hybridization. Each carbon is disposed
with one hybrid orbital pointed to the center of the triangle and the other two di-
rected to form covalent bonds to hydrogen. The remaining p-orbitals lie in the
plane of the ring. It is important to notice that the sp
2
hybrids within the ring are a
Hckel system, while the p-orbitals outside the ring are a Mbius one. Inscribing
the two polygons gives this picture of the energy levels, and molecular orbitals.



H H
H
H
H
H |+
|+
|+
nonbonding level



There are six valence electrons which fill the three bonding orbitals. The lowest
energy orbital comes from the inner set of three the Hckel set of orbitals
while next lowest orbitals, a degenerate pair, come from the outer set of three
the Mbius set. The form of the orbitals is shown on the next page:

13

bonding
antibonding


There is a lot of experimental support for this model. I'll give just one example at
this point; we'll see more later.
The solvolysis of a cyclopropylmethyl tosylate, a cyclobutyl tosylate, and a 3-
butenyl-1-tosylate give the same product distribution. It is because the cyclopro-
pylcarbinyl cation is exceptionally stable, and all three solvolyses produce the
same cation. The stability for what looks like a primary cation is due to the Walsh
orbitals on the cyclopropane moiety. The cationic center, in the "bisected" con-
formation, can conjugate with one of the two high-lying, filled HOMO's of the
cyclopropane ring. This stabilizes cyclopropylcarbinyl cation to about the same
extent as allyl cation is stabilized.


H
H
H


In this picture, it is easy to see how the Walsh orbitals resemble a double bond.
Conjugation with them makes the cyclopropylcarbinyl cation resemble allyl cation
rather than a normal primary carbonium ion.

This is about all I want to say about Hckel theory and related topics for the moment.
Limitations on the Hckel method are many. It does not give quantitatively correct
energies. The assumption is made implicitly that the actual wavefunction can be well-
described by a single configuration. Electron spin is ignored altogether. Moreover,
each of the assumptions listed previously can be invalidated.
14
C. First-order perturbation theory.

We just saw from Hckel theory how one can derive orbital energies, coefficients,
and symmetries. These results apply for a molecule in isolation. When we do frontier
orbital theory in the next section, we'll need a description of what happens to orbital
energies and orbital coefficients as we bring together two molecules. For this, we use
a construct called Perturbation theory. In this theory, interaction of two moieties in-
troduces a perturbation in their wavefunctions that can be described by a correction to
their original wavefunctions (the zero-order functions). The correction we'll talk about
is only the first-order one; it contains the really large effects.

Consider the zero-order Hamiltonian H
0
. If we introduce the perturbation, the Hamil-
tonian will be different. We express the new perturbed Hamiltonian as:

H = H
0
+ H'


where H' is the Hamiltonian that expresses the perturbation, and is a disposable pa-
rameter that allows us to increase or decrease the perturbation. Similarly, the unper-
turbed eigenfunctions are
1
0
,
2
0
, etc. with eigenvalues E
1
0
, E
2
0
, etc. We assume
that there are no degenerate eigenvalues. So we start with:


s order term higher
s order term higher

1
0
0
1
0
0
0 0 0 0
+ + =
+ + =
= =
E E E
E H E H
i i
i i
i i i i i i




where the forms for the perturbed wavefunction and the perturbed energy are taken
from a truncation of a Taylor series in after the linear term. Making the substitution
for the perturbed Hamiltonian,


( )( ) ( )( )
1 1 2 0 1 1 0 0 0 0 2 0 1 0 0 0
1 0 1 0 1 0 0
' '
'
i i i i i i i i i
i i i i i i
i i i
E E E E H H H H
E E H H


+ + + = + + +
+ + = + +


15
We equate terms in like powers of to get:


order - higher ...
order - first '
order - zeroth
0 1 1 0 0 1 0
0 0 0 0
i i i i i i
i i i
E E H H
E H


+ = +
=


which still doesn't help us too much because we don't know what the correction term

i
1
to the wavefunction is. At this point we take advantage of a property of the Ham-
iltonian operator. H is a Hermitian operator, i.e.:

| | | H H =

so all of its eigenvalues are real, and its eigenfunctions form a complete orthogonal
set. What that means for us here is that any function within the boundary conditions
of the problem can be expressed as a linear combination of the eigenfunctions of H.
Another way of looking at this is to say that the eigenfunctions can serve as basis
functions in an N-dimensional space. We use this here to express the first-order cor-
rection function as a linear combination of the zeroth-order eigenfunctions.

=
j
j ij i
a
0 1


which, after substitution into the first order term, gives,



+ = +
j j
j ij i i i j ij i
a E E H a H
0 0 0 1 0 0 0
'

where, now, everything is expressed in terms of zero-order eigenfunctions. This
means that our knowledge of the perturbation depends only on the perturbation opera-
tor, which we can often simply write out. Next, one multiplies through by and
integrates over all space.
- 0
k





+ = +
+ = +
j
jk ij i i k i
j
jk j ij i k
j
j k ij i i k i
j
j k ij i k
a E E E a H
a E E H a H
o o

0 0 0 1 0 0 0
0 0 0 0 0 1 0 0 0 0 0
| | ' |
| | | | | ' |


16
where much use has been made of the orthogonality of the zero-order eigenfunctions,
i.e. <
i
|
j
> = o
ij
. Let's look at this expression for a number of cases. For the case
where i = k, the expression reduces to,


0 0 1
| ' |
i i
H E
i
=

which is not too surprising. When the perturbation is small, then a good first-order
correction to the energy is the perturbation operator applied to the zero-order eigen-
functions. For i = j = k, the expression reduces to,

( ) 0 | ' |
0 0 0 0
= +
i k ik i k
E E a H

which rearranged, becomes an expression for the coefficients a
ik
.


0 0
0 0
| ' |
k i
i k
ik
E E
H
a

=



This, then, gives the first-order perturbed wavefunction as,

+ =
k
k
k i
i k
i i
E E
H
0
0 0
0 0
0
| ' |




The second-order terms in the perturbation expansion can be evaluated, but I won't do
that here. What I will do is give the result for the energy, correct to second-order.

+ + =
k k i
i k
i i i i
E E
H
H E E
0 0
2
0 0
0 0 0
| ' |
| ' |



To summarize perturbation theory, the effect of a perturbing term H' in the Hamilto-
nian is to mix the zero-order states according to an energy-weighted average, causing
a shift in the eigenstate energies. We'll use this result later when we come to frontier
orbital theory.


17
D. Symmetry and group theory.

Group theory is an odd topic because it contains no fundamentally new information.
It is simply a book-keeping tool which makes obvious thing we should have seen
anyway. Objects, or molecules, have symmetry operations. These are operations that
map the object onto itself, such as a rotation about an axis, or reflection through a
mirror plane. We will go through some terminology that makes it much easier to use
group theory.

1. Symmetry operations.

A more rigorous definition is: A symmetry operation is a transformation of a body
such that the final position is physically indistinguishable from the initial position,
and such that the distances between all pairs of points in the body are preserved.

A symmetry element is a geometrical entity with respect to which a symmetry op-
eration is carried out. It could be a point, line, or plane.

Let's start by considering the various symmetry elements that we will use in chem-
istry. I'll give an example of each one.

E Identity

C
n
n-fold (proper) rotation axis.

n is the order of the axis. A rotation of 2t/n radians results in a configuration
indistinguishable from the initial configuration. The operation itself is also de-
noted C
n
, with rotation through 2mt/n radians denoted C
n
m
.
F B
F
F
C
3
C
2
o
v
o
h

18

For molecules with multiple proper axes of rotation, like BF
3
, the axis of
highest order is the principal axis.

o Reflection through a plane of symmetry.

o comes from German Spiegel, which means mirror. A plane to which a prin-
cipal axis of rotation is perpendicular is termed o
h
, for horizontal, and a plane
containing the principal axis is termed o
v
, for vertical. In cases, such as a
square planar molecule, where there are two kinds of vertical planes, one is la-
belled o
d
, for dihedral.

i Inversion through a center of symmetry.

Inversion through a center of symmetry corresponds to the operation whereby
all points at coordinates (x,y,z) are sent to (-x,-y,-z) with the origin of the co-
ordinate system at the center of symmetry. BF
3
does not have a center of
symmetry, but SF
6
does.

S
n
n-fold improper axis of rotation.

Rotation about an improper axis corresponds to a C
n
operation followed by re-
flection through a plane normal to the axis. That plane does not necessarily
have to be a symmetry element by itself.


H
H
H
H
C
S
4




19
2. Point groups.

We can classify an object or a molecule by the symmetry operations that may be
performed on it. The category is called a point group. There are further classifica-
tions based on other symmetry properties, e.g. space groups, but we won't be con-
cerned about them. Each one is given a name, such as C
2v
or D
6h
, which describes
some of its most important elements. An instructive table of the chemically rele-
vant point groups, their names, and the operations they include, can be found in
most chemistry books.

In general, the symmetry operations can be expressed as matrix operators on the
coordinates of a representative point. For example, the identity matrix is:


|
|
|
.
|

\
|
1 0 0
0 1 0
0 0 1

which, multiplied by a column vector (x,y,z), returns the same vector. Similarly,
counterclockwise rotation through an angle u about the z-axis is given by a matrix,


|
|
|
.
|

\
|
1 0 0
0 cos sin
0 sin cos
u u
u u

A product of symmetry operations gives a product of matrices. Now, a representa-
tion of a group is defined as a set of matrices (or any other objects), each corre-
sponding of a single operation in the group, that can be combined among them-
selves in a manner parallel to the way in which the group elements combine. There
are an infinite number of representations of a group, but a finite set of irreducible
representations. All others can be block-factored into a combination of irreducible
representations.

For example, using the 3 x 3 matrices for three operations in the C
2v
point group,
C
2
(z), o
v
(xz), and o
v
(yz), we can show how they are related by multiplication:


|
|
|
.
|

\
|
=
|
|
|
.
|

\
|

=
1 0 0
0 1 0
0 0 1
) (
1 0 0
0 1 0
0 0 1
) (
2
xz z C
v
o
20

so the operation o
v
(xz), followed by C
2
(z), applied to a representative point, can be
written as:

) (
1 0 0
0 1 0
0 0 1
1 0 0
0 1 0
0 0 1
1 0 0
0 1 0
0 0 1
) ( ) (
2
yz xz z C
v v
o o
|
|
|
.
|

\
|
=
|
|
|
.
|

\
|

|
|
|
.
|

\
|

=

The 3 x 3 matrices for C
2
(z), o
v
(xz), and o
v
(yz), along with the unit matrix corre-
sponding to E, form a representation of the C
2v
point group because their multipli-
cation table is the same as that for the symmetry operations in C
2v
. Any group of
objects with that same multiplication table is a representation of C
2v
. A simple ex-
ample for the point group C
2v
can be done by assigning a +1 or 1 to each symme-
try operation. It only matters that the product of two of the +1's or 1's is the sign
assigned to the product of the corresponding symmetry operations.

E C
2
o
v
o
v
'


+1 +1 +1 +1
+1 1 +1 1
+1 1 1 +1
+1 +1 1 1

These numbers can be considered as 1 x 1 square matrices that transform according
to an irreducible representation of the C
2v
point group. Tables of the sum of the di-
agonals of the transformation matrices are called character tables. The character of
a one-dimensional matrix, i.e. a number, is just that number. In this case, +1 means
that the matrix is symmetric with respect to the operation, while 1 means that it is
antisymmetric. The 3 x 3 matrices that correspond to the symmetry operations also
form a basis for the same irreducible representations of C
2v
. They can be assigned
to the irreducible representations above based on their transformation properties.

Each irreducible representation has a name. The names are given according to
some conventions.

1. All one-dimensional representations are either A or B; two-dimensional repre-
sentations are E; three-dimensional representations are T.

21
2. One-dimensional representations that are symmetric with respect to a rotation
by 2t/n about the principal C
n
axis are designated A, those antisymmetric are
designated B.

3. Subscripts 1 and 2 are attached to A or B according to whether the representa-
tion is symmetric or antisymmetric, respectively, with respect to a C
2
axis per-
pendicular to the principal axis, or, in the absence of such an axis, to o
v
.

4. Primes or double primes are attached to all labels according to whether the
representation is symmetric or antisymmetric with respect to o
h
.

5. Where there is a center of inversion, the representation is labelled with a sub-
script g or u for gerade or ungerade, meaning even or odd.

6. E and T representations have numerical subscripts that are assigned by a
mathematical convention, but best treated as arbitrary labels.

These character tables are very useful in quantum mechanics. Molecular orbitals
can be assigned to a representation according to their phase properties. This is
why Hckel calculations are so useful. Although the absolute energies are just ter-
rible, the orbitals come out in the right order with the right symmetry properties.
As an example, the four t orbitals of cis-butadiene are shown below. The coordi-
nate system is included you should always specify your coordinate system so
that there is no ambiguity as to which axis or plane is being used. It can swap the
symmetry labels of different orbitals, which is confusing however, one gets the
right answer in the end as long as the choice of coordinate system is used consis-
tently.

The character table that is relevant is that for C
2v
. Note that, if it were trans-
butadiene that we were looking at, the point group would be D
2h
instead. The
characters are:

C
2v
E C
2
(z) o
v
(xz) o
v
(yz)
A
1
1 1 1 1
A
2
1 1 1 1
B
1
1 1 1 1
22
B
2
1 1 1 1

which can be used to assign symmetry labels to each of the butadiene molecular
orbitals based on their behaviour with respect to each symmetry operation.


z
x
y

4
|+
|+
b
1
a
2
b
1
a
2


The total symmetry of an electronic configuration is obtained by taking each elec-
tron, assigning it to an orbital, and giving that electron the symmetry of its orbital.
The symmetries of all the electrons are multiplied together to get the overall sym-
metry of the configuration. For example, for cis-butadiene, four electrons in the
two bonding orbitals gives (b
1
)(b
1
)(a
2
)(a
2
) = a
1
. For the first excited configura-
tion, the symmetry is (b
1
)(b
1
)(a
2
)(b
1
) = b
2
. Notice that for a closed-shell electron
configuration, the symmetry is always the totally symmetric representation.

Orbitals, configurations, and states can be assigned to a particular representation
of a point group. Vibrations more specifically, normal modes can also be as-
signed in a similar fashion. We'll just look at one example in the same C
2v
point
group. For H
2
O, there are three vibrations. Operating on each mode by the sym-
metry operations of C
2v
produces either the same vibration or its inverse. Exami-
nation of the character table then gives the representation.


a
1
b
2
a
1
H
O
H H
O
H
H
O
H

23

The immediate use that we're going to make of point groups and character tables
is in the evaluation of matrix elements. I don't mean the numerical evaluation, but
rather, a judgement as to whether the integral is zero or not. The criterion is that,
when one integrates over all space, the integral is non-zero only if the integrand
contains the totally symmetric representation of the molecular point group. This is
quite analogous to the situation where one integrates something like sin(x) from
t to +t.

To make use of this in the matrix elements <
i
|H|
j
> and <
i
|
j
>, we need to
know the symmetry of the Hamiltonian operator. It is totally symmetric, so the
two matrix elements are nonzero only when the wavefunctions belong to the same
representation. Other operators, such as the electric dipole operator or the spin-
orbit operator are not totally symmetric, so they connect states of different sym-
metry.

The best way to see the use to which symmetry properties of electronic wavefunc-
tions can be put to use is to solve a problem using them. In this first example, let's
look at the benzyl radical. This is a t-system for which Hckel theory gives a
qualitatively correct answer. However, we can arrive at the same answer without
solving any secular determinant at all. Benzyl can be considered to be a benzene
ring onto which a radical center has been attached. If there were a way to "turn
off" the interaction between the ring and the exocyclic carbon 2p orbital, we
would have a problem that we have solved before. This sounds like the start of a
perturbation theory approach. We start with the benzene orbitals, classified ac-
cording to D
6h
. Benzyl is only C
2v
so we reduce the symmetry according the ta-
bles from Herzberg, or by inspection of the character tables themselves. Those
benzene orbitals of the same symmetry as the exocyclic orbital interact with it
they split apart in energy and mix in character. The closer the orbitals are in en-
ergy, the greater the interaction. The four benzene orbitals that reduce to b
1
sym-
metry interact with the exocyclic methylene orbital (also b
1
). The two a
2
orbitals
are unaffected. The resulting orbital energy diagram (with symmetry labels) is the
same one that we would have gotten had we solved the secular determinant, but
this one is derived from perturbation and symmetry arguments.

S-ar putea să vă placă și