LCAO

Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

CHAPTER 11

MOLECULAR ORBITAL THEORY

Molecular orbital theory is a conceptual extension of the orbital model,


which was so successfully applied to atomic structure. As was once play-
fully remarked, a molecule is nothing more than an atom with more nu-
clei. This may be overly simplistic, but we do attempt, as far as posssible,
to exploit analogies with atomic structure. Our understanding of atomic or-
bitals began with the exact solutions of a prototype problemthe hydrogen
atom. We will begin our study of homonuclear diatomic molecules begin-
ning with another exactly solvable prototype, the hydrogen molecule-ion
H+2.

The Hydrogen Molecule-Ion


The simplest conceivable molecule would be made of two protons and one
electron, namely H+ 2 . This species actually has a transient existence in
electrical discharges through hydrogen gas and has been detected by mass
spectrometry. It also has been detected in outer space. The Schrodinger
equation for H+ 2 can be solved exactly within the Born-Oppenheimer ap-
proximation. For fixed internuclear distance R, this reduces to a problem of
one electron in the field of two protons, designated A and B. We can write

1 2 1 1 1
+ (r) = E (r) (1)
2 rA rB R

where rA and rB are the distances from the electron to protons A and B,
respectively. This equation was solved by Burrau (1927), after separating
the variables in prolate spheroidal coordinates. We will write down these
coordinates but give only a pictorial account of the solutions. The three
prolate spheroidal coordinates are designated , , . the first two are
defined by
rA + rB rA rB
= , = (2)
R R
while is the angle of rotation about the internuclear axis. The surfaces
of constant and are, respectively, confocal ellipsoids and hyperboloids
of revolution with foci at A and B. The two-dimensional analog should be

1
familiar from analytic geometry, an ellipse being the locus of points such
that the sum of the distances to two foci is a constant. Analogously, a
hyperbola is the locus whose difference is a constant. Fig. 1 shows several
surfaces of constant , and . The ranges of the three coordinates are
{1, } , {1, 1} , {0, 2}. The prolate-spheroidal coordinate
system conforms to the natural symmetry of the H+ 2 problem in the same
way that spherical polar coordinates were the appropriate choice for the
hydrogen atom.

=const

=const

=const
=0

Figure 1. Prolate spheroidal coordinates.

The first few solutions of the H+2 Schrodinger equation are sketched
in Fig. 2, roughly in order of increasing energy. The -dependence of the
wavefunction is contained in a factor

() = ei , = 0, 1, 2 . . . (3)

which is identical to the -dependence in atomic orbitals. In fact, the


quantum number represents the component of orbital angular momen-
tum along the internuclear axis, the only component which has a definite
value in systems with axial (cylindrical) symmetry. The quantum number
determines the basic shape of a diatomic molecular orbital, in the same
way that ` did for an atomic orbital. An analogous code is used, for
= 0, for = 1, for = 2, and so on. We are already familiar
with - and -orbitals from valence-bond theory. A second classification of
the H+2 eigenfunctions pertains to their symmetry with respect to inversion

2
through the center of the molecule, also known as parity. If (r) = +(r),
the function is classified gerade or even parity, and the orbital designation
is given a subscript g, as in g or g . If (r) = (r), the function is
classified as ungerade or odd parity, and we write instead u or u . Atomic
orbitals can also be classified by inversion symmetry. However, all s and d
AOs are g, while all p and f orbitals are u, so no further designation is nec-
essary. The MOs of a given symmetry are numbered in order of increasing
energy, for example, 1g , 2g , 3g .

Figure 2. H+
2 molecular orbitals.

The lowest-energy orbital, as we have come to expect, is nodeless. It


obviously must have cylindrical symmetry ( = 0) and inversion symmetry
(g). It is designated 1g since it is the first orbital of this classification. The
next higher orbital has a nodal plane, with = 0, perpendicular to the axis.
This function still has cylindrical symmetry () but now changes sign upon
inversion (u). It is designated 1u , as the first orbital of this type. The next
higher orbital has an inner ellipsiodal node. It has the same symmetry as
the lowest orbital and is designated 2g . Next comes the 2u orbital, with
both planar and ellipsoidal nodes. Two degenerate -orbitals come next,
each with a nodal plane containing the internuclear axis, with =const.

3
Their classification is 1u . The second 1u -orbital, not shown in Fig. 2,
has the same shape rotated by 90 . The 3g orbital has two hyperbolic
nodal surfaces, where = const. The 1g , again doubly-degenerate, has
two nodal planes, = 0 and =const. Finally, the 3u , the last orbital we
consider, has three nodal surfaces where =const.
An MO is classified as a bonding orbital if it promotes the bonding of
the two atoms. Generally a bonding MO has a significant accumulation
of electron charge in the region between the nuclei and thus reduces their
mutual repulsion. The 1g , 2g , 1u and 3g are evidently bonding or-
bitals. An MO which does not significantly contribute to nuclear shielding
is classified as an antibonding orbital. The 1u , 2u , 1g and 3u belong in
this category. Often an antibonding MO is designated by or .
The actual ground state of H+ 2 has the 1g orbital occupied. The
equilibrium internuclear distance Re is 2.00 bohr and the binding energy
De is 2.79 eV, which represents quite a strong chemical bond. The 1u is a
repulsive state and a transition from the ground state results in dissociation
of the molecule.

The LCAO Approximation


In Fig. 3, the 1g and 1u orbitals are plotted as functions of z, along the
internuclear axis. Both functions have cusps, discontinuities in slope, at
the positions of the two nuclei A and B. The 1s orbitals of hydrogen atoms
have the same cusps. The shape of the 1g and 1u suggests that they
can be approximated by a sum and difference, respectively, of hydrogen 1s
orbitals, such that
(1g,u ) (1sA ) (1sB ) (4)

1 g

A B

1 u

Figure 3. H+
2 orbitals along internuclear axis.

4
This linear combination of atomic orbitals is the basis of the so-called LCAO
approximation. The other orbitals pictured in Fig. 2 can likewise be ap-
proximated as follows:

(2g,u ) (2sA ) (2sB )

(3g,u ) (2pA ) (2pB )


(1u,g ) (2pA ) (2pB ) (5)
The 2p atomic orbital refers to 2pz , which has the axial symmetry of a
-bond. Likewise 2p refers to 2px or 2py , which are positioned to form
-bonds. An alternative notation for diatomic molecular orbitals which
specifies their atomic origin and bonding/antibonding character is the fol-
lowing:
1g 1u 2g 2u 3g 3u 1u 1g
1s 1s 2s 2s 2p 2p 2p 2p
Almost all applications of molecular-orbital theory are based on the LCAO
approach, since the exact H+ 2 functions are far too complicated to work
with.
The relationship between MOs and their constituent AOs can be rep-
resented in a correlation diagram, show in Fig. 4.

5
AO MO AO
3 u

1 g

2p 2p

1 u

3 g

2 u

2s 2s
2 g

Figure 4. Molecular-orbital correlation diagram.


The 1s 1g , 1u is similar to the 2s correlations.

MO Theory of Homonuclear Diatomic Molecules


A sufficient number of orbitals is available for the Aufbau of the ground
states of all homonuclear diatomic species from H2 to Ne2 . Table 1 sum-
marizes the results. The most likely order in which the MOs are filled is
given by

1g < 1u < 2g < 2u < 3g 1u < 1g < 3u


The relative order of 3g and 1u depends on which other MOs are occu-
pied, much like the situation involving the 4s and 3d atomic orbitals. The
results of photoelectron spectroscopy indicate that 1u is lower up to and
including N2 , but 3g is lower thereafter.
The term symbol , , . . ., analogous to the atomic S, P, D. . . sym-
bolizes the axial component of the total orbital angular momentum. When
a -shell is filled (4 electrons) or half-filled (2 electrons), the orbital angular

6
momentum cancels to zero and we find a term. The spin multiplicity
is completely analogous to the atomic case. The total parity is again des-
ignated by a subscript g or u. Since the many electron wavefunction is
made up of products of individual MOs, the total parity is odd only if the
molecule contains an odd number of u orbitals. Thus a u2 or a u2 subshell
transforms like g.
For terms, the superscript denotes the sign change of the wave-
function under a reflection in a plane containing the internuclear axis. This
is equivalent to a sign change in the variable . This symmetry is
needed when we deal with spectroscopic selection rules. In a spin-paired
u2 subshell the triplet spin function is symmetric so that the orbital factor
must be antisymmetric, of the form

1
x (1)y (2) y (1)x (2) (6)
2

This will change sign under the reflection, since x x but y y. We


need only remember that a u2 subshell will give the term symbol 3 g .
The net bonding effect of the occupied MOs is determined by the bond
order, half the excess of the number bonding minus the number antibond-
ing. This definition brings the MO results into correspondence with the
Lewis (or valence-bond) concept of single, double and triple bonds. It is
also possible in MO theory to have a bond order of 1/2, for example, in H+ 2
which is held together by a single bonding orbital. A bond order of zero gen-
erally indicates no stable chemical bond, although helium and neon atoms
can still form clusters held together by much weaker van der Waals forces.
Molecular-orbital theory successfully accounts for the transient stability of
a 3 +
u excited state of He2 , in which one of the antibonding electrons is pro-
moted to an excited bonding orbital. This species has a lifetime of about
104 sec, until it emits a photon and falls back into the unstable ground
state. Another successful prediction of MO theory concerns the relative
binding energy of the positive ions N+ +
2 and O2 , compared to the neutral
molecules. Ionization weakens the NN bond since a bonding electron is
lost, but it strengthens the OO bond since an antibonding electron is lost.
One of the earliest triumphs of molecular orbital theory was the pre-
diction that the oxygen molecule is paramagnetic. Fig. 5 shows that liquid
O2 is a magnetic substance, attracted to the region between the poles of a

7
permanent magnet. The paramagnetism arises from the half-filled 1g2 sub-
shell. According to Hunds rules the two electrons singly occupy the two
degenerate 1g orbitals with their spins aligned parallel. The term sym-
bol is 3
g and the molecule thus has a nonzero spin angular momentum
and a net magnetic moment, which is attracted to an external magnetic
field. Linus Pauling invented the paramagnetic oxygen analyzer, which is
extensively used in medical technology.

Figure 5. Demonstration showing blue liquid O2


attracted to the poles of a permanent magnet. From
http://jchemed.chem.wisc.edu/jcesoft/cca/CCA2/
STHTM/PARANIO/9.HTM

Variational Computation of Molecular Orbitals


Thus far we have approached MO theory from a mainly descriptive point
of view. To begin a more quantitative treatment, recall the LCAO approx-
imation to the H+
2 ground state, Eq (4), which can be written

= cA A + cB B (7)

Using this as a trial function in the variational principle (4.53), we have


R
H d
E(cA , cB ) = R 2 (8)
d

where H is the Hamiltonian from Eq (1). In fact, these equations can be


applied more generally to construct any molecular orbital, not just solu-
tions for H+
2 . In the general case, H will represent an effective one-electron

8
Hamiltonian determined by the molecular environment of a given orbital.
The energy expression involves some complicated integrals, but can be sim-
plified somewhat by expressing it in a standard form. Hamiltonian matrix
elements are defined by
Z
HAA = A H A d
Z
HBB = B H B d
Z
HAB = HBA = A H B d (9)

while the overlap integral is given by


Z
SAB = A B d (10)

Presuming the functions A and B to be normalized, the variational energy


(8) reduces to

c2A HAA + 2cA cB HAB + c2B HBB


E(cA , cB ) = (11)
c2A + 2cA cB SAB + c2B
To optimize the MO, we find the minimum of E wrt variation in cA and
cB , as determined by the two conditions
E E
= 0, =0 (12)
cA cB
The result is a secular equation determining two values of the energy:

HAA E H ES
AB AB
HAB ESAB HBB E =0 (13)

For the case of a homonuclear diatomic molecule, for example H+2,


the two Hamiltonian matrix elements HAA and HBB are equal, say to .
Setting HAB = and SAB = S, the secular equation reduces to

E ES
2 2
ES E = ( E) ( ES) = 0 (14)

9
with the two roots

E = (15)
1S
The calculated integrals and are usually negative, thus for the bonding
orbital
+
E+ = (bonding) (16)
1+S
while for the antibonding orbital


E = (antibonding) (17)
1S

Note that (E ) > ( E + ), thus the energy increase associated with


antibonding is slightly greater than the energy decrease for bonding. For
historical reasons, is called a Coulomb integral and , a resonance integral.

Heteronuclear Molecules
The variational computation leading to Eq (13) can be applied as well to the
heteronuclear case in which the orbitals A and B are not equivalent. The
matrix elements HAA and HBB are approximately equal to the energies of
the atomic orbitals A and B , respectively, say EA and EB with EA > EB .
It is generally true that |EA |, |EB | |HAB |. With these simplifications,
secular equation can be written

EA E H ES
AB AB =
HAB ESAB EB E

(EA E)(EB E) (HAB ESAB )2 = 0 (18)


This can be rearranged to

(HAB ESAB )2
E EA = (19)
E EB
To estimate the root closest to EA , we can replace E by EA on the right
hand side of the equation. This leads to

(HAB EA SAB )2
E EA + (20)
EA EB

10
and analogously for the other root,

+ (HAB EB SAB )2
E EB (21)
EA EB

The following correlation diagram represents the relative energies of these


AOs and MOs:
E-
EA

EB
E+

A simple analysis of Eqs (18) implies that, in order for two atomic
orbitals A and B to form effective molecular orbitals the following con-
ditions must be met:
(i) The AOs must have compatible symmetry.
For example, A and B can be either s or p orbitals to form a -bond
or both can be p (with the same orientation) to form a -bond.
(ii) The charge clouds of A and B should overlap as much as possible.
This was the rationale for hybridizing the s and p orbitals in carbon. A
larger value of SAB implies a larger value for HAB .
(iii) The energies EA and EB must be of comparable magnitude.
Otherwise, the denominator in (20) and (21) will be too large and the MOs
will not differ significantly from the original AOs. A rough criterion is that
EA and EB should be within about 0.2 hartree or 5 eV. For example, the
chlorine 3p orbital has an energy of 13.0 eV, comfortably within range of
the hydrogen 1s, with energy 13.6 eV. Thus these can interact to form a
strong bonding (plus an antibonding) MO in HCl. The chlorine 3s has an
energy of 24.5 eV, thus it could not form an effective bond with hydrogen
even if it were available.

11
Huckel Molecular Orbital Theory
Molecular orbital theory has been very successfully applied to large conju-
gated systems, especially those containing chains of carbon atoms with al-
ternating single and double bonds. An approximation introduced by Huckel
in 1931 considers only the delocalized p electrons moving in a framework of
-bonds. This is, in fact, a more sophisticated version of the free-electron
model introduced in Chap. 3. We again illustrate the model using buta-
diene CH2 =CHCH=CH2 . From four p atomic orbitals with nodes in the
plane of the carbon skeleton, one can construct four molecular orbitals
by an extension of the LCAO approach:

= c1 1 + c2 2 + c3 3 + c4 4 (22)

Applying the linear variational method, the energies of the MOs are the
roots of the 4 4 secular equation

H11 E H ES . . .
12 12
H12 ES12 H E . . . =0 (23)
22
... ... ...

Four simplifying assumptions are now made


(i) All overlap integrals Sij are set equal to zero.
This is quite reasonable since the p-orbitals are directed perpendicular to
the direction of their bonds.
(ii) All resonance integrals Hij between non-neighboring atoms are set equal
to zero.
(iii) All resonance integrals Hij between neighboring atoms are set equal
to .
(iv) All coulomb integrals Hii are set equal to .
The secular equation thus reduces to

E 0 0

E 0
=0 (24)
0 E

0 0 E

12
Dividing by 4 and defining
E
x= (25)

the equation simplifies further to

x 1 0 0

1 x 1 0
=0 (26)
0 1 x 1

0 0 1 x

This is essentially the connection matrix for the molecule. Each pair of
connected atoms is represented by 1, each non-connected pair by 0 and
each diagonal element by x. Expansion of the determinant gives the 4th
order polynomial equation

x4 3x2 + 1 = 0 (27)

Noting that
this is a quadratic equation in x2 , the roots are found to be
x2 = (3 5)/2, so that x = 0.618, 1.618. This corresponds to the four
MO energy levels

E = 1.618, 0.618 (28)

Since and are negative, the lowest MOs have

E(1) = + 1.618

and
E(2) = + 0.618
and the total -electron energy of the 12 2 2 configuration equals

E = 2( + 1.618) + 2( + 0.618) = 4 + 4.472 (29)

The simplest application of Huckel theory, to the ethylene molecule


CH2 =CH2 gives the secular equation

x 1

1 x = 0 (30)

13
This is easily solved for the energies E = . The lowest orbital has
E(1) = + and the 1 2 ground state has E = 2( + ). If butadiene
had two localized double bonds, as in its dominant valence-bond structure,
its -electron energy would be given by E = 4( + ). Comparing this
with the Huckel result (29), we see that the energy lies lower than the that
of two double bonds by 0.48. The thermochemical value is approximately
17 kJmol1 . This stabilization of a conjugated system is known as the
delocalization energy. It corresponds to the resonance-stabilization energy
in valence-bond theory.
Aromatic systems provide the most significant applications of Huckel
theory. For benzene, we find the secular equation

x 1 0 0 0 1

1 x 1 0 0 0

0 1 x 1 0 0
=0 (31)
0 0 1 x 1 0

0 0 0 1 x 1

1 0 0 0 1 x

with the six roots x = 2, 1, 1. The energy levels are E = 2 and


two-fold degenerate E = . With the three lowest MOs occupied, we
have
E = 2( + 2) + 4( + ) = 6 + 8 (32)
Since the energy of three localized double bonds is 6+6, the delocalization
energy equals 2. The thermochemical value is 152 kJmol1 .

14

You might also like