You are on page 1of 8

Basics of Molecular Orbital Theory

Valence Bond theory, as we saw in the last section,


is based on the notion that electrons are localized
to specific atomic orbitals. Molecular orbital theory
asserts that atomic orbitals no longer hold
significant meaning after atoms form molecules.
Electrons no longer "belong", in a sense, to any
particular atom but to the molecule as a whole.
Molecular orbital theory holds, as its name
suggests, that electrons reside in molecular orbitals
that are distributed over the entire molecule.

Quantum mechanics specifies that we can get
molecular orbitals through a linear combination of
atomic orbitals; that is, by adding and subtracting
them. How do we add and subtract orbitals? The
best way to picture this process is to recall the
wave-like nature of electrons. Recall from physics
that two waves can interact either through
constructive interference, in which the two waves
reinforce each other, and destructive interference,
in which the two waves cancel each other out.
Mathematically, constructive interference
corresponds to addition and destructive
interference corresponds to subtraction. When
atomic orbitals interact, we can either add them to
obtain a bonding orbital or subtract them to obtain
an antibonding orbital. Antibonding orbitals are
denoted with an asterisk (*).

For instance, in the hydrogen molecule, the atomic
1 s orbitals can overlap in a (head-on) fashion to
form a -bonding molecular orbital and a -
antibonding molecular orbital. The bonding
molecular orbital is "bonding" in the sense that it is
lower in energy than its component atomic orbitals.
Forming a bond and moving electrons into the
bonding orbital lowers the total energy of the
system, which is favorable. On the other hand,
moving electrons into the antibonding orbital
raises the energy of the system, which disfavors
bond formation. The total number of orbitals is
conserved; the number of molecular orbitals equals
the number of original atomic orbitals.

Polar Bonds
When the bonding atoms significantly differ in
electronegativity, their orbital energies will also be
different. Compare the MO picture of hydrogen
with that of H-F. Because fluorine is more
electronegative than hydrogen, its 2p orbitals lie
below the 1s orbitals of hydrogen. Due to this
difference in energy, the stabilization energy of the
resulting bonding MO is not as large. Furthermore,
the bonding MO is much more like the 2p orbital in
its spatial characteristics while the antibonding MO
is much more like the 1s orbital.
Energy Considerations
One advantage of the MO Model is that it gives us
more information about the energies of the
bonding electrons. In the case of H 2 , the molecule
is stabilized by twice the E of the bonding orbital.
The larger this energy gap, the more stable the
bond. In order for this stabilization energy to be
large, several factors are important:

The atoms must be of similar electronegativity.
Atoms must be of comparable size.
Orbitals must achieve adequate spatial overlap.
For example, the fact that the O-H bond is stronger
than the S-H bond can be explained by item 2. The
2p orbital of oxygen can overlap with the 1s orbital
of hydrogen more effectively than it can with the
3p orbital of sulfur. Finally, item 3 correctly
predicts that bonds should be stronger than
bonds since they have a greater degree of direct
overlap.

Bond Order
How does MO theory describe multiple bonds? In
MO theory the bond order of a bond is the number
of bonding electron pairs minus the number of
antibonding electron pairs. Intuitively, bonding
electrons stabilize the bond while antibonding
electron destabilize the bond. The greater this
difference, the stronger the bond, and the higher
the bond order. As we'll see, in MO theory the bond
order is no longer limited to integer values.

Electron Delocalization
One of the greatest successes of MO theory is that
it accounts for electron delocalization in a natural
way. We have seen that some molecules require
resonance structures to be represented accurately.
In all such cases, electrons are delocalized over
several bonds/atoms. One main drawback of the VB
model is that it assigns electrons to specific
atoms/bonds and therefore breaks down when it
comes to explaining delocalized electrons. The MO
model has no such problem; it offers a clean
approach to describing delocalization that is
superior to writing a bunch of awkward resonance
structures.
Application of MO Theory to Extended -systems
Unfortunately, the complexity of the full MO model
increases exponentially with the size of the
molecule. In order for MO theory to be useful in
practice, we limit its application to portions of a
molecule that are extensively delocalized. This
often occurs when electrons and lone pairs
overlap over several contiguous atoms.
MOLECULAR ORBITAL THEORY
1. Molecular orbitals (MOs) are made of
fractions of atomic orbitals. All atoms in the
molecule provide their atomic orbitals for
construction of MOs, but not all atomic orbitals
must participate in all MOs. The number of MOs
is equal to the number of atomic orbitals used to
generate them.
Instead of making bonds one at a time by
overlapping pairs of atomic or hybridized
orbitals, in the the MO procedure all available
atomic orbitals are mixed into multiple
combinations (MOs). This mixing procedure is
called the Linear Combination of Atomic
Orbitals (LCAO), and it simply means that we
"add" and "subtract" fractions of atomic orbitals
(wavefunctions) to make new molcular orbitals.
We have to use each atomic orbital completely,
we have to generate normalized molecular
orbitals (MOs), and the number of MOs must be
equal to the number of atomic orbitals that we
have started with. Again, it is a bit like making
mixed drinks (except for the "subtracting"
part), but now we mix orbitals of different
atoms all at once, instead of just premixing
individual atom's orbital as we din
inhybridization.
2. The MO are delocalized over many atoms.
In general, they do not directly correspond to
specific bonds (the exceptions include simple
diatomic molecules or some isolated bonds,
such as one in ethylene).

Since many atomic orbitals participate in
the "mixture" to form the molecular orbitals,
the volumes of these new orbitals encompass
many atoms, sometimes even the whole
molecule. The so formed MOs do not any
longer correspond to specific bonds (as they
were in VB theory). In fact, they can be
bonding between some pairs of atoms (where
"addition" of wavefunctions took place) and
antibonding between other pairs of atoms
(where "subtraction" of wavefunctions
happend), and sometimes they will have a node
at a given atom (like in the allyl system). On
some occasions the MO and VB orbitals will
"look" exactly the same. For example, we may
find such cases in isolated bonds, or orbitals
containing lone pairs that are not adjacent to
systems.
3. The (or *) type MOs are usually
separated from (or *) type MOs. (One
exception that will be discussed by us is called
hyperconjugation).
This separation is the consequence of
symmetry. bonds are usually perpendicular
to bonds, i.e. they cannot mix (because
the overlap is zero). In many cases this
arrangement simply means that and
networks do not interact and can be treated
separately. This situation simplifies the
analysis. For example, look at benzene: its
and networks are essentially independent;
when we analyze benzene (and aromaticity
later on) we talk exclusively about electrons.
Of course, there are exceptions. There are
many situations where bonds are in geometry
that allows for overlap with a system. One
such exception is called hyperconjugation. In
fact, the concept is (again) derived from the VB
theory to account for the delocalization of
electrons from the bond to the system. And
likeresonance (= conjugation) it is a fix of our
model, and not the problem of molecular
structure. In MO theory, the hyperconjugation
shows naturally: the appropriate, mostly -
type MO show contributions from some atomic
orbitals (s or p) of adjacent atoms that are
properly aligned with the system (but would
not be, in the VB language, considered a part of
it).
4. The MO are filled by all available electrons
(no more than two per orbital), starting from the
lowest energy MO orbital.
Each MO has energy associated with it (see
above). All available electrons (from all
participating atoms) are placed (two per orbital)
in the molecular orbitals, starting at the bottom
of the energy scale and moving up, until no
more electrons are left. What counts is not
whether the orbital is bonding or antibonding
between specific atoms within the molecule,
but what is the energy of the orbital.
5. All electrons in all MOs determine the
structure of the molecule, but the Highest (in
energy) Occupied MO (HOMO) and the Lowest (in
energy) Unoccupied MO (LUMO) are the most
important from the point of view of reactivity.
The energy of the specific molecular structure
depends on energy of its electrons in occupied
molecular orbitals. Different structures (i.e.
molecular geometries) will have different
energies of their molecular orbitals. Thus, all
electrons will influence the structure
(remember the compromises discussed above).
But from the point of view of reactivity some
electrons and some orbitals are more important
than others. The electrons of the highest
energy are the ones that the molecule would
like to "dump", and empty orbitals of the lowest
energy (in the reaction partner) are the best
"dumping grounds". In some chemical
reactions (for example electron-transfer
reactions or Lewis acid-base coplex
formations) this is exactly what takes place, in
others the interactions between the HOMO
(occupied) and the LUMO (unoccupied) "starts"
the reorganization of bonding of both reacting
partners.
6. The usual energy ordering of MOs is as
follows: -type orbitals (the lowest energy), -
type orbitals, nonbonding orbitals (atomic
orbitals, lone pair orbitals, or non-bonding -
type orbitals), *-type orbitals and *-type
orbitals (the highest in energy). The exceptions
are known (for example, CO molecule). This
ordering may be used to rapidly identify the
HOMO and the LUMO in organic molecules
7. In general for each type ( or ) the energy
of the MO increases with the increasing number of
nodes (in the bonding sense).
Predicting the ordering of the energy levels of
the orbitals that are farther removed from the
line dividing the occupied and unoccupied
orbitals is more difficult, short of performing
calculations. But we rarely need it anyway. On
the other hand, it is useful to know that the
energy of each type of orbital ( and especially
) increases with the number of nodes. Here,
we want to count the nodes that result in
antibonding interaction between atoms that are
bonded in the molecule. The more nodes of
this type, the higher the energy of the orbital.
So, although we cannot order the and
type orbitals relative to each other, we can
arrange the orbitals according to their energy
and decide easily (for example) which is the
highest occupied.
8. The electronegativity of atoms is reflected
in the size of their lobes within the MO. Usually,
in bonding orbitals there is more participation (the
lobes are larger) by the more electronegative
atoms. In the corresponding antibonding orbitals
the lobe sizes are reversed.
The situation here is slightly complicated. The
size of the lobe of the atomic orbital
participating in a given molecular orbital
depends on the energy of that orbital and its
size (these two are of course related, see
above). The bonding molecular orbital will
have larger contributions from the lower energy
(i.e. more electronegative) atoms, and the
antibonding orbitals will have larger
contributions from the higher energy (i.e. less
electronegative) atoms. This usually works well
for atoms from the same row of periodic table,
but deviations from this pattern can be
expected if the orbital lobes are contributed by
atoms belonging to different rows (compare for
example the size of the p orbitals of oxygen
and sulfur in the table in Part I).
9. Most often, the HOMO corresponds to a
filled -type orbital or a lone pair (nonbonding
electrons), and the LUMO corresponds to an empty
atomic p, *-type orbital or (if there is no
system) to an empty * orbital.
This is just the consequence of orbital
ordering discussed in p. 6. A minor
complication (not really) is when a lone pair
orbital is part of the system (as happens
when resonance is present). Well... then the
whole system: needs to be analyzed.
Typically, the HOMO in such situations is an
occupied nonbonding type orbital.
10. The chemical reactions between
molecules are largely governed by HOMO-LUMO
interactions (the highest-energy electrons (HOMO)
of one molecule "looking for" the lowest-energy
unfilled space (LUMO) in the other molecule). The
electron-rich component's (see: Brnsted base,
Lewis base, nucleophile, electron donor) HOMO
will interact strongly with the electron-deficient
component's (see: Brnsted acid, Lewis acid,
electrophile, electron acceptor) LUMO. The
difference in energy between these orbitals and
the overlap between them (orbital lobe size) will
largely determine the facility of the reaction and
the site of attack (bond formation).

Since interaction between filled (occupied)
orbitals does not result in net bonding (electron-
electron repulsion), and interaction between
empty (unoccupied) orbitals cannot contribute to
bonding (no electrons to be shared) only the
interaction between occupied and unoccupied
orbitals may provide the initial impetus for the
reorganization of existing bonding. Of course the
highest energy electrons (HOMO) and the lowest
energy empty orbitals (LUMO) will interact the
strongest (they are closest in energy, see above).
Essentially all chemical reactions are dependent
on these HOMO-LUMO interactions.
The strength of interaction between orbitals is
proportional to their overlap (here assumed the
same for both pairs of interaction) and inversely
proportional to their energy separation. HOMOB
LUMOA interaction will control the reactivity
between A (Lewis acid or electrophile) and B (Lewis
base or nucleophile).

You might also like