3820 Lecture Chapter - 3 - Part1 - 2004 PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Chemistry 3820 Lecture Notes Dr. M.

Gerken Page 22

3. Bonding in molecules: the molecular orbital (MO) method
There is more than one approach to the theory of the chemical bond. Within M.O.
theory, there are a variety of different approaches, as well as a vast range of levels of
approximation actually used by practicing theoreticians. We will not concern ourselves
at all with these fine points, taking the results of the wavefunctions found by others.
We will learn a few qualitative rules for the behaviour of wavefunctions under various
conditions, but that is the extent of our sophistication.
A very important and useful feature of MO theory is that it lends itself directly to
describing the bonding in solids, especially that of metals and semiconductors. With
the increasing importance of high-technology materials to modern industrial society, it
has become more important than ever that chemists have a good understanding of
solids, and thus we will consider bonds in solids (other than the ionic bond, which we covered in detail in Chemistry 2810) later
on in these lectures.
3.1 Essence of the MO method
Consider the simplest molecule, H
2
+
. Here we have a species with three
interacting particles, two protons and one electron. All three interact by
electrostatic forces, as shown at right:
What molecular orbital theory suggests is that the correct approach to this,
and every other molecule is to re-solve the Schrdinger equation for this
three-body problem. We first simplify the problem by fixing the inter-
nuclear distance at the average bond-length measured for the molecule. This assumption can always be altered later and the
problem re-solved for a different, fixed, inter-nuclear separation. The separation of electron movement from nuclear movements
is called the Born-Oppenheimer approximation; it is a very reasonable approximation since electrons are moving much faster
than nuclei. Nonetheless, this is still a more difficult problem than that for the H-atom, since the potential field, V, is now
cylindrical rather than spherical. Remember that the equation has the general form:


2
2
2
2
2
2
2
2
8

x y z
m
b
E V + + = ( )

Where the Hamiltonian operator at the left operates on the wavefunction to give a set of solutions at different energies,
which are called the Eigenvalues, and the wavefunctions are the Eigenfunctions of the system. If this mathematical system can
be solved, it leads to the derivation of the exact molecular orbitals of the simplest molecule, H
2
+
. For this molecule, the math has
been solved, and the solutions derived as closed analytical functions. However, for the vast majority of molecules, the challenge
remains to mathematicians to solve the calculus. In particular, just as for the atom, the problems start with the addition of the
second electron. Thus even the simple molecule H
2
has not been treated by exact molecular orbital theory.
The hallmarks of the MO method are that each different molecule, and each different conformation of those molecules, leads to a
unique pattern of wavefunctions delocalized over the whole molecule. As in the H-atom, the Pauli exclusion principle is valid,
which means that each orbital can accommodate a maximum of two electrons, and those electrons must have opposite spins, i.e.,
are spin-paired. Again as in the atom, the pairing of two electrons with opposite spin requires energy: the spin-pairing energy.
Again as in the atom, there is a molecular analogy to the Aufbau principle, so that the electrons fill into the new molecular
orbitals in such a way as to achieve the lowest possible overall electron configuration (most stable in energy) by obeying Hunds
rule.
3.1.1 Approximate MO theory
The difficulties in exact MO theory have largely been overcome by a wide variety of approximate methods. As a result, MO
theory is actually a collection of competing and overlapping methodologies. If you ever have a look at the computational engine
in programs such as HyperChem or Gaussian98, you will see that there a large list of "methods" to choose from, including the
Extended Hckel, CNDO, MINDO, AM1, PPP and various ab initio methods. These all refer to different levels of
approximation, with the last mentioned being the most exacting, and thus generally leading to a better approximation to the true
molecular orbitals of the molecule than the previous ones.
We will develop a qualitative bonding model in lecture that is more simplified than these computerized methods and is based
mainly on symmetry considerations, although we will from time-to-time correct our bonding schemes by comparison to what is
shown by high-level calculations, or even better, by experiment.
You were wondering
What is the name given to the
bonding theory that uses sp, sp
2
and
sp
3
hybrid orbitals, as often described
in first year texts or organic texts?
+
+
e
-
Attractive
Repulsive
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 23

3.1.2 The LCAO method
One of the most common general approaches taken in MO theory, and one that fits in with the chemist's notion of bringing
two atoms together to form a chemical bond, is to consider the molecular orbitals as being formed from constituent atomic
orbitals. One motivation for this notion is the consideration that when an electron is close to the nucleus of an atom, its behaviour
is very similar to that of equivalent electrons in free atoms of that type. This modeling of MO's () in terms of contributing
atomic orbitals (AO,
i
) is called linear combination of atomic orbitals, or LCAO. A linear combination is a sum with varying
contributions by the constituents, the extent of the contribution being indicated by weighting coefficients. (c
i
).
= c11 + c22 + c33 + c44 + ..
The basis set of AOs has to include the most important atomic orbitals from which one constructs the new molecular
orbitals of the molecule. It stands to reason, then, that if an excited state of a molecule is being calculated, more inclusion of
higher-lying atomic orbitals will be required in the basis set. Empty atomic levels close in energy to filled AO's must be included
as part of the basis set of atomic orbitals. In our discussion we will focus entirely on ground-state molecular electronic structure,
and we will operate with a minimum basis set.

In the LCAO approach of MO theory, two atomic orbitals combine (interact/mix) to produce molecular orbitals with one
being higher in energy and one being lower in energy. The degree of this splitting in MO theory depends on three factors:
1. Symmetry:
Only orbitals of the same symmetry (symmetry species) can interact.
You have to apply group theory and use the symmetry species of the appropriate point group.
2. Relative Energies:
Those orbitals which are closest to each other in energy will interact the most, resulting in the largest possible
splitting.
3. Spatial Extension/Overlap:
The overlap between the AOs has to be significant to result in significant interaction.

All three factors have to be considered when combining AOs to produce MOs. These three factor do not only apply to the
LCAO approach of MO theory, but to any mixing/coupling of orbitals/wavefunctions/states.

As a consequence of the overlap requirement only orbitals within a narrow range of energy have the right properties to
undergo significant interaction. Only the valence atomic orbitals have sufficient spacial extension to produce significant
overlap between the AOs. Core orbitals are buried inside the atom and, therefore, cannot interact with the core orbitals of an
adjacent atom (assuming normal bond lengths). The valence AOs are the orbitals that mix the most in the LCAO approach and,
therefore, are essential to describe the bonding in molecules. The range of energies wherein chemical bonding can take place
may be shown using a graphic we have already considered, but which I will repeat here with a band drawn in that designates the
valence zone:
H He Li Be B C N O F Ne Na
0
480
960
1440
1920
2400
2880
3360
3840
4320
1s
1s
1s
2s
2s
2s
2s
2s
2s
3s
2p
2p
2p
2p
2p
2p
2p
VALENCE ZONE

It is this energetic consideration that is behind the concept of a valence shell. Thus, whereas the n = 1 level is the valence level
for the hydrogen, it becomes a core level for Li, and similarly the n = 2 level is the valence level for Li to F, but becomes a core
level for sodium. The valence zone that I have drawn in is not fixed in stone, but corresponds approximately to the highest
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 24

energy that can be attained by chemical ionization. In fact, the 2s levels of N and F are involved in covalent bonding, but we
must remember that they are very deeply buried atomic orbitals.
Although both, the atomic and molecular electronic structures, are derived from quantum mechanics, it helps understanding
greatly to remember the physical basis underlying atomic structure. For example, let us see if we can rationalize why the only
important levels for bonding of two elements from the second period are the 2s and 2p. At a typical chemical bonding distance,
we find that only the orbitals from the valence level are of appropriate size to allow for bonding. Remember that the radial
probability density functions we developed for the atom each have a variable radial scale depending on the size of the effective
nuclear charge that the electrons in those orbitals experience. We can pictorially represent this by drawing spheres to represent
the 90% probability density of 1, 2, and 3s levels:
(a) (b) (c)

Here we see that the very low-energy 1s orbitals (a) are so tightly bound to the nucleus that they do not reach out to the other
nucleus with appreciable electron density; the 2s valence level is optimally able to have significant overlap with high electron
density. The 3s level beyond the valence shell, however, is much larger and hence much more diffuse. Although it extends well
over the region involving both nuclei, any electrons in such an orbital would have such low electron density that they would be
useless at doing chemical bonding. As the energy level diagram indicates, the 3s orbital becomes a valence orbital in the third
period because it shrinks to the right size an energy as the nuclear charge increases.
3.1.3 One-electron theory
The next approximation that we make is to assume that the MO's we derive are unaltered by the presence of other electrons
in the same or neighbouring levels. We can then construct the possible MO's, and populate them with as many orbitals as we
need. This simplification is known as one electron theory, or as the complete neglect of differential overlap (CNDO). We will
allow for the cost of the electron-pairing energy, and will indicate this by simply adjusting the energy of doubly occupied
molecular levels up a bit higher than if they were simply 2 that of a singly occupied orbital. At the computational level, the
most common improvement on such one-electron theory is to allow for differential overlap. This is usually done by either a
variational or a perturbational approach that starts with the one-electron wavefunctions, then proceeds to modify them by
considering first the most important inter-orbital cross-terms, and working out to towards the lesser important until the right
degree of accuracy is achieved.
3.1.4 The first row diatomics as an example
We are now ready to consider the first-row diatomic molecules, i.e., those made up of H and/or He atoms. We recognize that
the minimum basis set will be the two 1s orbitals on the individual atoms. We can then see that these can be combined in an in-
phase (leading to a bonding arrangement) and an out-of-phase manner (leading to an anti-bonding arrangement).
MO's for H
2
+
Using LCAO theory, this can be written mathematically as:
)} B ( ) A ( {
2
1
s 1 s 1
b
+ = and )} B ( ) A ( {
2
1
s 1 s 1
a
=
Pictorially this can be shown by sketches showing what happens when the orbitals are
combined (both with the same size and distance separation), one case in phase, which
leads to constructive interference, and hence an increase in amplitude of the wavefunction
in the critical inter-nuclear region, and in the other case destructively, leading to
cancellation of the wavefunction and a nodal plane between the two nuclei. In MO's, as in
AO's, the square of the wavefunction is directly related to the electron density
distribution, and hence the constructive interference leads to an orbital with a net increase
in electron density between the nuclei, effectively gluing the two nuclei together by their
mutual attraction to those electron(s). On the other hand, electrons in the destructively
interferring combination will be excluded from the inter-nuclear region, and population of
this MO leads to the negation of the bond created by the constructive interference. We
refer to these two situations, respectively, as bonding molecular orbitals and anti-
+ +
- +
+


1s
(A)
1s
(B)

b
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 25

bonding molecular orbitals. We often indicate the relative phases of the interacted wavefunctions by shading, here done in
green and red.
We now consider the relative energy of the two types of interaction. Much as we rationalized atomic energy levels in the
atom, we can think of what effect the electron experiences
in the two combinations. We first adopt a simpler orbital
labeling system, using molecular symmetry to do so. The
symmetry of both LCAO's is sigma, so we label the
bonding and antibonding MO's as and *, respectively.
The electrons in will be lowered in energy compared to
an electron in a single H atom, because it spends more
time being attracted to both nuclei in the inter-nuclear
region. On the other hand, an electron in * will
experience less nuclear attraction than in the free atom,
because it is expelled from nearly half the volume of the
free atom. Thus we diagram as dropping below the
energy of the free atom orbitals, and * as rising above the
energy of the free atom orbitals. By doing the correct
math, we can show that * rises more than drops, as shown in the more accurate energy level diagram at the far right.
Generalization for first row diatomics
This simplistic scheme, first worked out in the early days of quantum mechanics by Heitler and London, accurately accounts
for the bond energy and bond length of the simplest molecule, H
2
+
, and is adequate to treat any combination of first-row diatomic.
Thus if we take either two H 1s or two He 1s oribtals, and consider the total electron count involved in the four possible
molecular combinations of diatomics, we discover what is diagrammed in the following figure:
* * *
*
r0



1s 1s

1s 1s

1s 1s

1s 1s
272 kJ mol
-1
452 kJ mol
-1
301 kJ mol
-1
0 kJ mol
-1
B.D.E.
106 pm 76 pm
(s)
1
Configuration ()
2
()
2
(*)
1
()
2
(*)
2
H
2
+
H
2
He
2
+
He
2
0.5 Bond Order 1 0.5 0

The data below the molecules are taken from experiment, and include the bond dissociation energy (B.D.E.), the bond distance,
the electron configuration, and the bond order (dimension-less). One of the neat advantages of MO theory is that we can
postulate real bonds using a single electron, rather than the two-electron bond pairs of simple valence bond theory. In fact, H
2
+
is
a very good little molecule, which can be studied accurately by electronic spectroscopy. Its bond energy is in fact greater than
many two-electron bonds involving heavier atoms.
In order to correlate bonding in MO theory to the well-established valence-bond conventions, we define the concept of bond
order in such a way that a bond populated by two electrons has a bond order of one. However, bond orders may be any order,
unlike in VB theory:
bond order = [# of electrons in bonding MO's # of electrons in anti-bonding MO's]
*
*

1s 1s 1s 1s
E
N
E
R
G
Y
Simplified interaction diagram Correct interaction diagram
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 26

Empty orbitals and Spectroscopy
One important advantage of MO theory over the simpler VB method is the ease with which it can comprehend
spectroscopic processes. This is done with the same simplicity as atomic line spectra, as being
due to the promotion of an electron from a full to a higher-lying empty orbital, and/or the
decay of an excited electron from such an excited state to the ground state. The energy of the
photon, given by E = h, must match exactly the difference in energy between the two orbital
levels.
For the dihydrogen molecule, H
2
, there is an ultra-violet absorption band at 109 nm,
corresponding to an energy difference of 11.4 eV (= 1100 kJ mol
-1
) between the bonding and
antibonding levels. Most molecular absorption spectra of this type show broad lines rather
than the sharp line spectra typical of atomic spectra. Molecules, unlike atoms, are capable of
undergoing rotational and vibrational motion. It is found that such motions for small molecules
are quantized. This results in coupling of the rotational and vibrational energy levels with the
transitions due to the electron excitations. Some excitations lose a little energy to vibrational or
rotational energy-level changes. Others gain energy from theses sources. The net result is that
we no longer have sharp lines but multiplets in the spectrum, if it is measured in the gas phase.
But in solution, multiple interactions between the molecules (they collide with each other) tend
to collapse all theses energy levels together. The net result is that UV-vis absorption spectra
tend to be lumps of energy absorption rather than sharp lines. Nevertheless, the principle in
terms of the basic quantum phenomenon of an electron being excited from a lower lying
orbital to an empty higher orbital is identical to the atom and to the molecule.

1s 1s
H
2

1s 1s
H
2
*
h
109 nm
maximum
i
n
t
e
n
s
i
t
y

3.2 The second row diatomic molecules
When we turn to the second row of the periodic table, the 1s orbitals drop way down in energy. This means that (i) the
smaller orbitals overlap less, and hence the energy gap between bonding and antibonding orbitals decreases, and (ii) both
1s
and
*
1s
are occupied. The net effect is that these M.O.s no longer contribute to bonding effects between the atoms in the second row
diatomics. This is a general phenomenon, so that we always ignore the core orbitals when considering a bonding scheme.
We call the important orbitals the valence M.O.s, and these logically enough are usually derived form the valence atomic orbitals
of the constituent atoms. Theoreticians have also coined the term frontier orbitals to indicate those M.O.s which are most
important to reactivity of the molecule, and these are usually the Highest Occupied Molecular Orbital, HOMO, and the Lowest
Unoccupied Molecular Orbital, LUMO. However, sometimes orbitals slightly lower than the HOMO or higher than the LUMO
are also frontier orbitals.
3.2.1 Generalized orbital overlap
From the shapes of the orbitals we can go directly to the important concept of orbital overlap: bonds between atoms will only
occur if there is significant constructive interference between some of the waves on one atom and some on the other. This is
called orbital overlap.
You were wondering
What if we want to construct a
molecule in which H is
combined with He?

Chemistry 3820 Lecture Notes Dr. M. Gerken Page 27

Symmetry classification of orbital interactions
When we combine more than one orbital on an
atom with those on another, the geometry of the
interaction results in different symmetry classes of
orbital overlap. Orbitals can only be combined if they
belong to the same orbital class. The important
symmetry classes for the diatomic, and by inference
for any two-atom interactions (i.e. one bond in a
larger molecule), are , and . Delta bonds are
extremely rare, being encountered only for certain
heavy transition metals which can posses a
quandruple bond. Pi bonds normally form the
second and third bond in double and triple bonds.
Shown at the right is a fairly comprehensive
sample of possible orbital interactions that lead to the
three classes of overlap symmetry.
Constructive, destructive and orthogonal
overlap
Not all combinations of overlap are possible. You
will find (a) constructive and (b) destructive
interference. Orbitals having the same symmetry can
interact in a constructive and destructive way, giving
bonding and antibonding MOs, respectively. Those
interactions that lead to no net constructive or destrucive
interference are said to be orthogonal. Orthogonal
overlap occurs when atomic orbitals of different
symmetries are combined. Orthogonal interaction, which
is the absence of interactions, leads to a third class of
orbital, the non-bonding molecular orbital.
Of course, the most common form of orthogonal
interaction is among the orbitals of a single atom. Thus
the ns, np
x
, np
y
and np
y
orbitals within an atom all
occupy the same region of space, but coexist because
they are orthogonal. A common kind of orthogonal
interaction in a molecule occurs when at given atomic
orbital on one atom is of a different symmetry class, and
therefore cannot interact with any orbital on the other
atoms. Such an unhybridized AO is carried over into the
MO scheme without alteration. Bonding MOs represent
bonds in a Lewis structure, while non-bonding MOs represent lone-pairs. We will see that such a black-and-white model of MO
theory has to be refined.

3.2.2 "First-order" MO diagrams for the diatomics
We are now ready to construct at least a first attempt at the molecular orbital diagram for a second-row diatomic molecule.
We use a fragment approach, as shown below, in which we bring the two atoms with their valence orbitals (the minimum basis
set) together, and indicate which interactions lead to more stable, and which to less stable, MO's. The lines connecting the AO's
and MO's in this diagram have no real significance, and are added mostly to guide the eye. The approximate orbital shapes that
are produced by these interactions is shown in the sketches at the right hand side of the diagram.

Chemistry 3820 Lecture Notes Dr. M. Gerken Page 28

2s
2s
2p 2p
E
N
E
R
G
Y
bring together
MO description of 2nd row homonuclear diatomic molecules
Energy level diagram
Sigma and pi overlap leading to bonding and antibonding MO's
1
g
(2s)
1
u
(2s)*
2
u
(2p)*
2
g
(2p)
1
g
(2p)*
1
u
(2p)

Note that the total number of nodes are preserved between the incident atomic orbitals and the resulting MO's. Also, it is often
difficult to render the resultant MO's accurately, unless you happen to be a good artist. For most purposes, topologically correct
diagrams will suffice, i.e., just draw the overlapping AO's in the position they hold in the molecule, but shade them as to show the
constructive and destructive overlap. Nonetheless, it is very important to study the final forms of these MO's on the right because
that is closer to the actual electron distributions that our topologically correct diagrams represent. For an even better
visualization, run an empirical MO calculation on a computer, and rotate the resulting wire-frame orbitals in space to appreciate
their 3-dimensional nature better.
versus overlap
A very important conclusion has been included on the diagram above, and that is that the end-on overlap of the 2p orbitals
leads to greater bonding and antibonding interaction than the side-on overlap. Why is this the case? This can better be
appreciated by considering some scale diagrams of orbitals drawn at typical single and triple-bonding distances.
First of all, a single bond is longer than a double bond. This is because the optimal overlap is achieved in bonding when the
nuclei are further apart than in is optimal for
overlap. Indeed, at the distance shown in
the diagram at the left, the sideways overlap
of the same 3p orbital is negligible at best.
Thus on forming a double, the bond distance
gets shorter. A triple ( + 2) bond is even
shorter, because there are two bonds
requiring a shorter distance for optimal
overlap. What is important to realize, as
shown on the next diagram, is that even at
the shorter distance found in a multiple
bond, the overlap efficiency of the
interaction is still greater than that of the
interaction. This should be seen as strictly a
consequence of geometry.
Two 3p orbitals in bonding orientation at the distance of a single bond
Location of nuclei
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 29

Location of nuclei
Two 3p orbitals in bonding orientation at the distance of a triple bond
Two 3p orbitals in bonding orientation at the distance of a triple bond
Location of nuclei

Since the overlap is greater, the electrons experience a stronger attraction to the nuclei (or repulsion in the antibonding case),
and thus the energy of the orbitals drops lower than that of the orbitals when equivalent np orbitals overlap in this fashion.
3.2.3 Orbital mixing and a "second order" treatment of the diatomics
An important reason to use symmetry labels in MO theory is that symmetry controls the types of interactions that are
possible. We saw that AOs (wavefunctions localized on atoms) of the same symmetry can interact; the same is true for the
interaction between molecular orbitals (wavefunctions delocalized over the whole molecule). This leads first to another of our
basic rules of LCAO theory;
Only orbitals of the same symmetry can interact constructively or destructively to form bonding and antibonding
orbitals.
But a second consequence of symmetry that is almost equally important is the so-called non-crossing rule. This states that:
When two energy levels of the same symmetry class approach each other in energy, they will interact (mix) in such a way
that the lower level will drop in energy and the higher level will rise in energy.
In the diatomic molecules, this effect is most dramatic for the early molecules such as Li
2
and Be
2
. We have already noted that in
these elements, the atomic 2s and 2p levels are very close to one another. A consequence of this is that the energy of the resulting
diatomic 1
g
and 2
g
orbitals come very close to one another. They therefore undergo extensive s-p mixing, with the result that
the lower level gains p character, and is lowered in energy, while the higher level gains s character in order to become less
strongly bonding. A consequence of this is that the strict division of the MO's to individual atomic parentage becomes less clear-
cut, and from now on we will not be using this labeling method. Instead, we will simply number orbitals with the same symmetry
labels from 1 to n up the MO diagram starting from the bottom. The non-crossing rule can be diagrammed as follows:

The "g" orbitals will end up being much closer to each other as a consequence of how the diagram is constructed, and it is clear
from (a) in the figure that these two levels would have crossed if linear correlation had occurred. The situation for the "u"
orbitals is less dramatic, but the effect is nonetheless the same, so that they two levels mix and thereby pull apart from one
another as the nuclei are brought closer together. You may wonder as to how this mixing is accomplished, and this is best
understood by considering the topological sketches of these two orbital types.
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 30

1
u
(2s)
1
u
(2p) 1
u

1
g

2
u

2
g

2
u
(2p)*
2
g
(2p)*
+
+
+
+
30%
30%
30%
30%
=
=
=
=

I have arbitrarily mixed in 30% of the same-symmetry orbital in each case: in fact the extent of mixing will change directly with
the atomic sp separation in energy. Further, it will be unusual that the degrees of mixing of
g
and
u
are the same in a given
molecule. But the principle operates this way in all cases. Note that the resulting MO's are no longer sub-labeled "2s" and "2p",
because that strict distinction has been lost. In fact, from now on we will use symmetry labels directly created for the MO's, and
not worry too much as to which AO's they originally came from. Note also that we have created orbitals of partial bonding and
anti-bonding character! This will also complicate our calculations of bond order, which in most MO diagrams will not be
completely cut and dried.
Consider carefully what has happened to the MO's resulting from the mixing: 1
g
has gained increased electron density in
the internuclear region. We would expect it therefore to be lowered in energy. 1
u
has decreased electron density in the inter-
nuclear region, but this was anti-bonding electron density, so this orbital has become less anti-bonding than previously. Thus we
would also expect this orbital to be lowered in energy. On the other hand, 2
g
has lost electron density in the internuclear region,
which was bonding, and has become polarized to the left- and right-sides. It is well on its way to be a non-bonding MO, one that
corresponds to the VB notion of a lone pair. However, note that in MO theory, such lone pairs are described as delocalized
rather than as localized regions of electron density. Which is more accurate? We will look for evidence on this issue a little later
in our discussion. For the moment, we recognize that this orbital will be raised in energy. In fact, it is the raising of the energy
of 2
g
which is the most noticeable effect of second-order mixing among the second-row diatomic molecules, because it is
possible for this level to be raised so much that it ends up above the
u
level, thus causing a level-inversion in the MO diagrams
of the second-row diatomics. Finally we recognize that 2
u
"paid the penalty" of all this re-mixing by becoming a highly-
unfavourable strongly anti-bonding orbital, and thus we expect this orbital to be raised in energy significantly where strong
second-order mixing occurs.
We are now ready to see the effect of all these hybridizations on the energy level of the molecule. Note that the levels are
unaffected, because they have nothing that they can mix with in the homonuclear diatomics.
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 31

2s
2s
2p 2p
MO description of 2nd row homonuclear diatomic molecules
Energy levels in original diatomic construction
1
g
(2s)
1
u
(2s)*
2
u
(2p)*
2
g
(2p)
1
g
(2p)*
1
u
(2p)
Varying
degree
of s-p
separation
2s
2s
2p 2p
1
g
1
u
2
u
2
g
1
g
(2p)*
1
u
(2p)
Less s-p
separation
Relative order may vary here
Energy levels in re-mixed diatomic
E
N
E
R
G
Y

Now that the theory is in place, we are ready to apply it to the actual molecules, and the results of detailed calculations and
measurements on all the second-row diatomics are shown in the following diagram:

To see how this picture fits with the experimental results on such molecules, consider the following table of data:
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 32










There are some important features of these molecules that bear consideration, partly because of their intrinsic interest, and also
because similar situations will appear for many of the more complex molecules we will encounter later. Note first of all that 6 of
the 8 possible second-row diatomic molecules can exist (though only N
2
, O
2
and F
2
are the most stable forms of those elements).
Note that strong evidence for the orbital mixing we have been considering is provided by the magnetic properties of B
2
and C
2
.
Thus the former is paramagnetic (two unpaired electrons in
u
) and the later diamagnetic (all electrons paired), consistent with the
diagram. If a level inversion did not exist, then the opposite should be true, i.e. B
2
diamagnetic and C
2
paramagnetic. The
bonding in B
2
is also quite unique, with the net single bond being a -bond, but without a bond to support it! Similarly C
2
has
two -bonds but no net bond!
Note that N
2
would be diamagnetic with or without a level inversion, and so we will need additional experimental evidence
to support the order of the MO's in this molecule. Note that MO theory clearly predicts that O
2
will be paramagnetic with two
unpaired electrons. This is not obvious from the Lewis structure of dioxygen, which has a double bond and 4 lone pairs. Note
also that the "double bond" in this molecule is the net result of having a "triple bond" and a single anti-bond. This situation leads
to weaker over-all bonding than a bond order of 2 when there are no anti-bonds, e.g. compare O
2
with C
2
. Similarly F
2
has a
single bond that is the net of a "triple bond" and a "double anti-bond". Consequently F
2
has a very low bond energy. We are not
surprised to learn that much of the chemistry of this diatomic molecule occurs via radical processes, in which the weak FF bond
is first broken to form two reactive F atoms. Finally note that bond distances go down from left to right across the PT for bonds
of similar bond order. In part this reflects the reduction in size of the atoms themselves in the same direction.
3.2.4 Using CACAO and HyperChem
A quantum-chemistry computer program will calculate a "best" MO description of a molecule in one go, the degree of
accuracy depending on the sophistication of the method being used. A well-known approximate MO method is the Extended
Hckel method developed originally by Nobel laureate Roald Hoffman. This is a low-power MO method that uses the one-
electron approximation in conjunction with experimentally derived atomic energy levels. It is therefore often called and
"empirical MO method". The Extended Hckel method (EHMO) is incorporated into several popular software packages, of
which two are ones you are likely to have a chance to use. The first program is CACAO, a wonderful implementation of
Hoffman's original code into a Window's compatible software package that is provided free by an engaging pair of Italian
inorganic chemists. Please consult your instructor if you wish to obtain your own copy of CACAO.
HyperChem is a commercial product, and is quite expensive. If you wish to own a copy of this program, an inexpensive version
called HyperChem Lite, which incorporates the EHMO method, is available for about US$50 by downloading off the internet.
The URL of the company is www.hypercube.com. The full versions of HyperChem also contain more sophisticated "semi-
empirical" and "ab initio" molecular orbital methods. The interface in HyperChem is entirely interactive on the computer screen,
and uses a mouse to create and manipulate the structure. CACAO, by contrast, is a command-driven Fortran computer program
that was developed from code first written in the 1960's. However, despite the antiquated interface, the program does incorporate
some nifty features that directly reflect Hoffman's insightful approach to MO theory. In particular, the fragment orbital analysis
used by the program most directly resembles the approach taken in these lecture notes. I would therefore recommend that you
invest the time required to become an adequate CACAO user. Here follow some extracts of the input code and the output from
CACAO for dinitrogen.
Input file for N2: TITL nitrogen
PAR 0 DIST
KEYW EL WF CM OV OP RO NC
ORIG DU
INT N -1 1 0.564 180 0
INT N -1 2 0.564 0 0
FMO
FRAG 2 1
END
(consult the manual for explanation of the arcane command language!)
Here is the energy level diagram:
Property Li
2
Be
2
B
2
C
2
N
2
O
2
F
2

Bond length
()
2.67 1.59 1.24 1.10 1.21 1.42
Bond
energy
(kJ mol
1
)
110 272 602 941 493 138
Bond order

1 0 1 2 3 2 1
Chemistry 3820 Lecture Notes Dr. M. Gerken Page 33


Here are the plots of the MO's in wire-frame representation:

Chemistry 3820 Lecture Notes Dr. M. Gerken Page 34

3.2.5 Ultra-violet photoelectron spectroscopy
Although ultraviolet absorption spectra
have proved very useful for the analysis of the
electronic structure of diatomic molecules more
complex than H
2
, a more direct portrayal of the
molecular orbital energy levels can often be
achieved by using ultraviolet photoelectron
spectroscope (UV-PES) in which electrons are
ejected from the orbitals they occupy in
molecules. In this technique, a sample is
irradiated with 'hard' (high frequency) ultraviolet
radiation (typically the He(I) line of 21.22 eV
radiation from excited He atoms, or the He(II)
line at 40.8 eV) and the kinetic energies of the
photoelectrons - the ejected electrons - are
measured. Because a photon of frequency has
energy h, if it expels an electron with
ionization energy IE from the molecule, the
kinetic energy E
K
of the photoelectron will be:
E h IE
K
=
The lower in energy the electron lies initially
(that is, the more tightly it is bound in the
molecule), the greater its ionization energy and
hence the lower its kinetic energy after it is
ejected. The instrument used to measure UV-
PES shares many components with a mass
spectrometer. A schematic of the device is
shown at right. Since all electrons have the
same mass, instead of selecting out the mass in
the magnetic analyzer (as done in a mass
spectrometer), the kinetic energy can be measured. The detector circuitry subtracts the kinetic energy from the known source
energy, and thus we get a readout of intensity (y axis) against IE - a photoelectron spectrum.
Because the peaks in a photoelectron spectrum correspond to the various kinetic energies of photoelectrons ejected from
different energy levels of the molecule, the spectrum gives a vivid portrayal of the molecular orbital energy levels of a molecule.
To analyze a photoelectron spectrum we use the simplifying assumption called Koopman's theorem:
IE = E
orbital

In other words, we assume that the ionization energies are the exact inverse of the orbital energies of the M.O.s in the molecule.
Thus UV-PES gives us a literal "snapshot" of the M.O.s of a molecule! Though an approximation, it holds remarkably well for
small molecules and proved to be one of the biggest successes of MO theory over the competing VB (valence bond) theory. A
typical UV-PES of a small molecule is shown in schematic form below, keyed to the vibrational changes that occur in the
molecule during ionization.
Ground state
1st ionized state
2nd ionized state
3rd ionized state
Vertical ionization
populates many
vibrational levels
Vertical ionization
populates mostly
groung vibrational
level
Internuclear distance
1st ionized state
2nd ionized state
3rd ionized state
Schematic diagram of the UV-PES experiment and the output

Chemistry 3820 Lecture Notes Dr. M. Gerken Page 35

In the un-ionized molecule, most of the molecules will be in the ground vibrational state, and have the average inter-nuclear
separation that chemists call the "bond length". If ionization results in a 1+ molecule with very similar bond strength to the
original, the vibrational levels of the produced ion will be very similar to that of the neutral molecule, and most of the ions will be
in the ground vibrational state. In this case, the ionization will result in a single peak corresponding to the transition from the
ground vibrational state of the molecule to the ground vibrational state of the ion. There will be a few small daughter peaks
resulting from that fraction of both molecule and ion that populate higher vibrational states, but these will be minimal.
On the other hand, if the ionization causes a significant change in bond-strength, either by removal of an anti-bonding
electron (bond strengthening), or removal of a bonding electron (bond weakening), the bond length will alter significantly. My
sketch shows the case of bond weakening, with the average inter-nuclear separation increasing. Vertical ionization will then
automatically push the ion into a large number of vibrational excited states, and the result is an envelope of evenly spaced peaks
instead of a single peak. This is shown in the spectrum by the middle band, which has lots of vibrational fine structure,
whereas the 1
st
and 3
rd
peaks are sharp.
The interpretation of this in terms of molecular orbitals is that the 1
st
and 3
rd
peaks are from orbitals that are mostly or
completely non-bonding, whereas the 2
nd
peak is from an orbital that is either strongly bonding or strongly anti-bonding. The
way I have drawn the potential wells, in this case it is from a strongly bonding orbital (though this choice is not obvious from the
spectrum alone.) To summarize our discussion:
The smallest IE represents electrons from the HOMO
If the MO is strongly bonding or anti-bonding, the ionization will have a vibrational fine structure. This can actually be
used in small molecules such as N
2
to measure the NN stretching frequency, which does not show up in the IR
spectrum.
If the peak is sharp, the MO is essentially non-bonding.
For larger molecules, the lower orbitals tend to be closely spaced in energy, so that a virtually continuous ionization band
is usually seen at higher energies for larger molecules.
Ionizations requiring >21.22 eV require the use of He-II photons, and such ionizations are always broad, so that little can
be said about bonding/anti-bonding vs. non-bonding for such high-energy processes. Often only He-I data is reported, so
that as a general rule, only the higher-lying MO's in an MO diagram will be interpreted by UV-PES.
Let us now apply our ideas to a real-life example: dinitrogen. This vitally important diatomic molecular element has been
studied by UV-PES. In the diagram below, the UV-PES spectrum has been turned on its side, and the ionization energies keyed
to the energy levels in the MO diagram. Topologically correct sketches of the orbitals are also provided.

1
u

1
g

2
u

2
g

2s
2s
2p 2p
1
g
1
u
2
u
2
g
Energy levels of N
2
1
u

1
g
1
u

1
g

Using a UV-PES to scale the MO diagram of N


2

Chemistry 3820 Lecture Notes Dr. M. Gerken Page 36


Here are some of the important features to pay attention to:
The first IE has little bonding character, consistent with ionization of electrons from the 2
g
level, which through second-
order mixing has lost most of its bonding character, and become non-bonding.
The second IE is an envelope of peaks due to a lot of vibrational fine-structure. This is fully consistent with ionization
out of the strongly bonding
u
MO.
The third IE is again from a non-bonding level, and this is consistent with ionization out of the 1
u
level in this molecule.
Note that the labelling of the UV-PES, taken from the textbook, is different, because they have included the
g
and
u

formally derived from the 1s level in their count. We will not do this in our own diagrams, although it is never wrong to
do so - just unecessary.
We cannot see the fourth IE from the 1
g
level using He-I radiation. However, it has been measured with the He-II
source, and is found as a broad lump at about 40 eV!
The UV-PES data therefore provide experimental evidence for the level inversion shown in the previous diagram of all
the second-row diatomics. In the case of N
2
, both alternative electron arrangements lead to diamagnetic structures, so
that we need this additional evidence to prove the structure.
Note also that in comparison to the Lewis structure of N
2
, :NN:, the MO description has similarities and differences.
The bond order is indeed 3 by both descriptions (6 electrons in the bonding MO's 1
g
and
u
). There are two lone-pair
type orbitals, 1
u
and 2
g
. However, the lone pairs are not equivalent as the Lewis structure would suggest. One is the
HOMO, and will be used as the donor atom when N
2
acts as a (weak) Lewis base, the other is deeply buried and will not
be involved in reactivity (its not a frontier orbital).

You might also like