You are on page 1of 303

CH225 Inorganic Chemistry I

Course information
Instructor: Dr. Dmitri Goussev, e-mail: dgoussev@wlu.ca
Office hours: By appointment or anytime when I am in the office (N3020A).
To find the course website: type CH225 in the Search window on the WLU webpage.
Note: read lectures in advance for every class!
Marking: Midterm test (35%) (Oct. 30, in class). All midterm questions will be posted on the
CH225 website.
Assignment (20%), due before 4:00 pm, December 4. Late hand-in penalty: 10%
Final examination (45%).
Textbook: (Optional) C. E. Housecroft, A. G. Sharpe Inorganic Chemistry, Prentice Hall.
The following books have been used in the preparation of this course:
1. D. Shriver, P. Atkins Inorganic Chemistry.-3rd ed., W.H. Freeman&Co. 1999.
2. M. J. Winter Chemical Bonding, Oxford University Press, 1997.
3. R. L. Carter Molecular Symmetry and Group Theory, Wiley, 1998.
4. D. M. P. Mingos Essentials of Inorganic Chemistry 1, Oxford University Press, 1995.
5. D. M. P. Mingos Essentials of Inorganic Chemistry 2, Oxford University Press, 1998.
6. G. L. Messier, D. A. Tarr Inorganic Chemistry. 2nd ed., Prentice Hall 1998.

Course outline
In this course we are going to study fundamental concepts of Inorganic Chemistry:
structure, shape and symmetry of inorganic molecules, the nature of bonding in
inorganic compounds.
1. The fundamentals: Atomic Structure. Atomic orbitals. The Periodic Table. Properties of
elements.
2. Structure and Bonding: Lewis structures. Symmetry elements and operations.
Symmetry point groups. The VSEPR model. Valence bond theory. MO theory. Molecular
orbitals of diatomic molecules.
3. Molecular Symmetry and Group Theory: The symmetries of orbitals and application of
group theory to chemical problems. Molecular orbitals of polyatomic molecules.
4. Lewis Acids and Bases. Lewis acidity. Boron and carbon group acids. Nitrogen and
oxygen group acids. Halogen acids. Reactions. Hard and soft acids and bases.
5. Transition Metal Complexes. Representative ligands. Coordination numbers. Structures
and symmetries. Bonding and electronic structure of six-coordinate ML6 and fourcoordinate ML4 complexes.
(This list is not detailed and intended only to show the major structure of the course)

What is Inorganic Chemistry?


Organic chemistry is defined as the chemistry of hydrocarbon compounds CnHmXk (X
= nonmetal). Inorganic chemistry can be broadly described as the chemistry of
"everything else. As can be imagined, the inorganic realm is extremely broad.
Organometallic chemistry bridges both areas by considering compounds containing
carbon-metal bonds.

Contrasts with Organic Chemistry


Organic Chemistry

Inorganic and Organometallic Chemistry

Hydrogen is normally bonded to one


atom:

Hydrogen bridging several atoms:

H
H

C
H

Maximum of four atoms bonded to


carbon:

CH3

R'
R"

B
H

Carbon bonded to > 4 atoms:

R
'"R

Al
H3C

H3
C
C
H3

(CO)3
Fe

CH3
Al
CH3

(OC)3Fe
(OC)3Fe

Fe(CO)3
Fe(CO)3

Contrasts with Organic Chemistry


Organic Chemistry

Coordination geometry around carbon


is tetrahedral, trigonal planar or linear:

Inorganic and Organometallic Chemistry

Coordination numbers of 5, 6, 7 are


common. Coordination geometry is more
complex:

H
C

OC

Pri3P

Ru

PPri3
Cl

F
F

Ti

PPh3
H
H
Ir H
H
H
PPh3

3F
F

Contrasts with Organic Chemistry


Organic Chemistry
Aromatic rings are common:

Inorganic Chemistry
Aromatic rings can be bonded to metals:

Ru
Cl

Hydrocarbon cage-like compounds:

Fe

Cl
PPh3

Variety of different cluster compounds:


B
B

P
B
B

P
P

P
P4

C10H16

2B
B

B12H122-

Related to CH225.
1. Chemistry of main-group Inorganic (CH226) and Organometallic compounds
(CH328). Inorganic synthesis (laboratory component in CH226)
2. Spectroscopic determination of molecular structure (CH303)
3. Research project in Organometallic Chemistry and Catalysis (CH490)

Assigned reading for next class: Lectures 2 and 3 (prepare questions!)

CH225.2
Today:
Atomic structure
Quantum numbers and orbitals
Shapes of atomic orbitals

Hydrogen atom
The Bohr model
A model of the hydrogen atom was proposed by the Danish physicist Niels Bohr in
1913. He suggested that the atom of hydrogen consists of a negative electron
revolving around the positive nucleus - proton.

me

-central nucleus of mass mn


-electron of mass me
r -distance between the electron and nucleus

orbit

mn

The total energy E of the electron is a function of the distance between the electron
and nucleus, r :
2

E=-

e
8r

constants:
e -electron charge
-permittivity of vacuum

(Note the ve sign of E, meaning energy is released when e approaches the nucleus)
2

The principal quantum number, n


Bohr had to place an important constraint upon his model. To make it agree with the
experimental properties of atomic hydrogen, he suggested that only certain distances
r (i.e. orbits) were allowed.

h2
ao =
mee2

r = a on 2

E=

-mee4
82h2n2

where ao = 1 a.u. 0.53 and the quantum number n = 1, 2, 3.... This n is referred
to as the principal quantum number.
When r is quantized, the energy of the electron is also quantized. According to Bohr,
the electron occupies the lowest orbit (n = 1) in atomic hydrogen, at the distance of
ca. 0.53 from the nucleus.
The relative sizes of orbits:

n2

r,

0.53

-0.00022

2.12

-0.00005

4.76

-0.00002

E, J

n=3
n=2
n=1
1

9 n2

Particle or wave?
Electrons and other atomic particles should not be regarded as little hard lumps of
matter like small billiard-balls.
Electrons have properties of both particles and waves. For example, they can be
weighed (a property of a particle) and they can be diffracted (a property of a wave).
The particle-wave duality was expressed by de Broglie. He has postulated that all
matter possesses characteristics of both waves and particles; the wavelength can be
expressed as a function of mass and velocity:

The wave properties become important only for very small particles. Consider a
cricket ball, mass m = 0.15 kg moving at v =160 km/h = 44 m/s. Substituting into the
above equation gives the wavelength of the associated wave as ca. 10-34 m. For
comparison, the diameter of an atom is ca 10-10 m. A distance as short as 10-34 m is
meaningless.
4

From orbits to orbitals.


The Uncertainty Principle formulated by Heisenberg tells us that the exact position
of the electron in a hydrogen atom can not be accurately determined at any given
time. However, the probability (likelihood) of finding the electron somewhere around
the nucleus can be calculated.
If the probability at some point is high, then the electron density is said to be high at
that point.
The electron density representation describes an orbital, the term related to an orbit
in the Bohr model.
Orbital

80 % probability

Schrdinger's wave function,


The novel quantity is a wave function that can be obtained for a particle, such as
electron, by solving Schrdinger's equation. For the special case of a particle moving
in one direction (along x) it has the form of a 2nd order differential equation:

h2 d2
+ (EV) = 0
2
2
8 m dx
where E is the total energy (constant), V is the potential energy.
The probability of finding the particle in the region between x and x + dx is 2dx.
Obviously the total probability of finding the particle moving along x must be exactly 1
and can be determined by integrating over all x:

Schrdinger's wave equation


Schrdingers equation was solved for the hydrogen atom in 1926 . The mathematical
details are not easy, and Schrdinger himself needed advice from colleagues before
he could obtain the solution. The following qualitative features are important:
In the case of the hydrogen atom, is a function of three Cartesian coordinates x, y,
and z, or alternatively, three polar coordinates r, , . As a consequence, the
solution of the Schrdinger's equation for hydrogen involves 3 quantum numbers:

n, the principal quantum number, can take the values 1, 2, 3...


l, the angular momentum quantum number, can take the values 0, 1, 2... up to n 1
ml, the magnetic quantum number, can take the values -l, -(l-1),..., 0, ..., (l-1), l
The total energy depends only on n, and is given by the same equation as in the
theory of Bohr :
4

E=

-mee

82h2n2

Quantum numbers and orbital names


Any valid combination of n, l, and ml defines a unique orbital. The orbitals are named
after the quantum numbers. The first part of the name is the principal quantum
number n. The second is linked to the angular momentum quantum number l:

l=

1st shell

0,

1,

2,

name: s,

p,

d,

2nd shell

3rd shell

The principal quantum number n describes a shell. There is one orbital in the first
shell, four in the second, and nine in the third. Each set of orbitals within a shell for
which l is the same is termed a subshell.
8

Shapes of atomic orbitals


Atomic orbitals are typically shown as boundary surfaces. They include the regions
where the electron is very likely (e.g. 90% probability) to be found. Different
shading/colors can be used if the wave function changes its sign.

The 1s orbital is spherical; it is not directional:

The three 2p orbitals all have the same shape; they are directional and orthogonal:

2px

2py

2pz

CH225.3
Today:
radial distribution functions of s and p electrons
orbital shielding in many-electron atoms
For next time: read lecture CH225.4

1s orbital
Schrdingers equation can be solved only for 1-electron systems: H, He+, and Li2+.
1s, describing the 1s orbital, is an exponential function:
(e = 2.718)

1s

The value of 21s at any point is the


probability of finding the electron at that
point.

In a qualitative agreement
with Bohr, the 1s electron is
most likely ao 0.53
away from the nucleus in
the ground state.

To get a more useful picture, the radial


distribution function is defined as 4r221s.
It defines the probability of finding the
electron at any distance r from the nucleus,
that is on a spherical surface which radius
is r.
(Note, 4r2 defines the surface area of a
sphere)

Other s orbitals
2s, 3s, 4s, etc. orbitals are all spherical. Normally, they are empty in atomic
hydrogen. They are said to be excited states.
Note: ns orbitals have (n-1) radial nodes, which are spherical surfaces where 2ns =
0, and where ns changes its sign.
2s

3s

Some observations and comments:


1. All ns orbitals have the same overall shape - spherical. However, the orbitals
significantly increase in size when n increases (as n2).
2. For the 2s and 3s orbitals the greatest amount of electron density is found in the
outermost region.
3. The 2s orbital shows a spherical surface around the nucleus at which the electron
density drops to zero and the sign of ns switches from + to . This feature is
often referred to as a nodal surface or simply as a node. The 3s orbital
possesses two nodes.
4. The probability, 2ns, is not affected by the sign of ns. It is either a positive
number or zero when ns = 0.

2p and 3p orbitals
The 2p (l = 1) orbitals are directional. The 2p orbital shown below has most of its
electron density along the z-axis. It is called 2pz; the 2px and 2py orbitals are similarly
defined.

np orbitals have (n-1) nodes.

lobes

nodal plane

Electron density plots:


2p

3p

- +- +

Wave function shading/color and sign


The schematic representation of a 2p orbital contains two pieces of information. The
shape corresponds to 2np , the probability. The shading/color indicates the relative
signs of the wave function, that is, whether the value of the wave function is positive
or negative. These signs have nothing to do with charge.
2np > 0.9

The standard orientation of a Cartesian coordinate system:


z

counterclockwise

3d orbitals
There are five 3d orbitals (when l = 2). They correspond to the five possible values of
ml : -2, -1, 0, 1, 2. Four 3d orbitals (dxy dyz dzx and dx2-y2) have the same shape.
nd orbitals have (n-1) nodes.
Nodal planes

Conical
nodes

Hydrogenic ions
The solutions of Schrdingers equation for the hydrogen atom can be extended to
ions, such as He+ and Li2+, which like hydrogen have only 1 electron and are called
`hydrogenic`.
For an electron orbiting a nucleus of charge Z > 1,
all orbitals are contracted as:

r~

n2ao
Z

The electron is drawn inwards as a result of


the extra attraction. This contraction is
apparent in the figure:

Many-electron atoms
Schrdingers equation cannot be solved for He, Li, and the rest of the atoms found
in the Periodic Table. Technically, we do not exactly know their electron wave
functions, that is, their orbitals.
For most of the inorganic chemistry we rely on the orbital approximation, assuming
that the electrons in many-electron atoms occupy atomic orbitals resembling those of
the hydrogen atom.
However, there is an important difference. The 2s and 2p orbitals of hydrogen are
degenerate. In all other atoms, the 2s orbital is lower in energy than the 2p, that is
they are non-degenerate.
Consider the 3rd electron in the lithium atom. It is attracted by the nucleus (Z = +3). It
is also repelled by the inner (1s2) electrons, that is the two 1s electrons screen the
third electron. The difference between the full nuclear charge, Z, and the screening
effect is called the effective nuclear charge, Zeff:
Zeff = Z - screening constant
9

Orbital shielding (screening)


The figure below shows an important difference between 2s and 2p orbitals. A certain
part of the 2s orbital is inside the 1s orbital and is close to the nucleus. For the 2p
orbitals, penetration is less significant. As a result, the core electrons (1s2) shield an
electron in the 2s orbital less effectively than they shield an electron in a 2p orbital.
Therefore, the effective nuclear charge Zeff(2s) > Zeff(2p), and E(2s) < E(2p).
Orbital energies are proportional to 1/Zeff2

1s
2p

2s

10

Radial distribution functions of 3s, 3p, 3d


The general trend in orbital energies, ns < np < nd, is a consequence of different
repulsion from the inner shells; nd is screened most effectively, and ns least well. In
other words, 3s is most penetrating and 3d least so.

Inner shells

11

Effective nuclear charge Zeff


The values of Zeff can be significantly smaller than Z. For example, in fluorine (Z = 9)
the 2p electrons are shielded by the inner-shell electrons, and Zeff = 5.1

Increasing Zeff
Zeff increases on going from left to right in the Periodic Table because of the
increasing Z.
As Zeff increases, the orbitals contract, which is why the fluorine atom is smaller than
the boron atom.

12

CH225.4
Today:
The Periodic Table
Atomic radius
Read Lecture 5 for next class

Electron spin
Solutions of the Schrdinger equation are 'exact' in the mathematical sense. They
are not quite 'correct', however. The errors are very small for the hydrogen atom, but
become more significant for heavier atoms. Partly, this is due to the assumption that
only the electron moves, the nucleus being stationary.
In 1928, Dirac developed an equation which combined the wave theory with
relativity. For atoms with small nuclear charges, it gives results which are numerically
very similar to those from Schrdinger equation. Importantly, Dirac demonstrated that
electron has an angular momentum and resembles a tiny bar magnet. This property,
called spin, was known experimentally in 1928 but was unaccounted in the
Schrdinger equation. The spin of electron is I = , and there are two spin states
corresponding to the quantum numbers I = -1/2 and +1/2.
Therefore, the state of any electron in an atom is described by four quantum
numbers: n, l, ml, and I, which combinations should be different for all electrons in an
atom. This is the Pauli exclusion principle.

Orbital energies
The following diagram shows the order of increasing orbital energies when they are
first filled in the atoms of the Periodic Table. The ground-state electronic
configuration of an atom is obtained by filling the orbitals bottom up.

Period 7
Period 6
Period 5
Period 4
Period 3
Period 2
Period 1
3

The Periodic Table

Important features
The Hunds rule states that the lowest energy configuration is that in which as many
electrons in a subshell as possible have the same spin. For example, carbon has the
following electronic configuration:

2p2
2s2
1s2

2p2
and not

2s2
1s2

Starting at K (Z = 19, n = 4) the 4s orbital fills before any of the 3d. These orbitals
have very close energies in transition metals. For example, the valence-shell
configuration of chromium is 4s13d5 and that of copper is 4s13d10. 5s and 4d orbitals
also have very similar energies.
The electronic configurations of transition metals are difficult to predict. The
experimentally established ground-state electronic configurations for all known
elements are given on the previous slide.
5

Atomic radius
One of the useful properties of an element is the size of its atom.
By definition, the atomic radius of an element is half the distance between the atoms
in its homonuclear compound.
Example: Hydrogen (in H2)

Example: Carbon (in diamond)

0.74
H

0.77
0.37

Example: Chlorine (in Cl2)

Cl

Cl

0.99
6

Atomic radius
If the atomic radii of elements A and B are known, then a good estimate of an A-B
bond length can be obtained by combining the atomic radii.
Example: C-H bond length is predicted to be 0.77 + 0.37 = 1.14 . The experimental
values are between 1.08 and 1.10 .

Atomic radius
Atomic radii increase down a group, and within the s and p-blocks they decrease
from left to right across a period. There is the so-called lanthanide contraction
observed as a decrease in radius for elements following the f-block in Period 6.

Group 1
f-block

I
Group 7

5d-metals

PRIVACY AND FREEDOM OF INFORMATION CIRCULAR #06-5


Protection of Personal Information:
Ontarios privacy law requires that we protect the personal information
we collect and use. We must not put records of personal information in
places where they may be seen by the public. According to this
legislation personal information can be as little as just a name or a
student number. The following practices, among others, therefore,
contravene this legislation:
1. Posting lists of students names on walls or doors of offices, halls
or classrooms (names with grades or other information is even worse);
2. Displaying lists of student numbers (numbers with grades or other
information is even worse);
3. Posting or passing around sign-up sheets that allow subsequent users
to see the names or numbers of previous users;
4. Any unsupervised return of assignments where students can see the
information of others (e.g., leaving piles of assignments in hallways
for pick up);
5. Unattended collection of forms or assignments, such that later
students can see material submitted those submitting material earlier.
John F. Metcalfe PhD, Director, University Information and Privacy Office
9

CH225.5
Today:
Ionization energy
Electron affinity
Electronegativity
For next time, read Lecture 6

Ionization energy
The first ionization energy, I1, is the minimum energy required to remove an electron
from an atom in the gas phase and in its ground state:

A(g)

A+(g) + e-(g)

E = I1

I1 shows an overall increase from left to right in every subshell because of the
increasing Zeff. I1 also correlates with the atomic radius; in smaller atoms the valence
electrons are more strongly attracted to the nucleus.

The trend across a period is not perfectly linear because (a) an np electron is easier
to remove than an ns electron, and (b) paired electrons repel each other and one of
them is easier to remove than an unpaired electron.
2

Ionization energy group trend


The following figure illustrates the variation of I1 for the alkali and alkaline earth
metals (Groups 1 and 2). The decrease of I1 is particularly noticeable in Group 2. The
orbitals expand on going down a group, so the valence electrons become more
distant from the nucleus and are easier to remove.

Electron affinity
The electron affinity, Ea, is the energy change observed upon addition of an electron
to an atom in the gas-phase:

A(g) + e-(g)

A-(g), E = Ea

Note: a positive value of Ea indicates that the ion A- is stable and the electron
addition is a favorable process .
Table: Electron affinities of the main-group elements (1 ev = 96.5 kJ/mol)

Electron affinity trends


An element has a high electron affinity if the additional electron enters a shell where
it can feel a strong effective nuclear charge Zeff. This is the case of the elements
toward the top right of the Periodic Table.
Electron addition is relatively unfavorable for the following two electronic
configurations :

(when the ns orbital is filled, the electron will have to enter a higher energy np orbital)

(when the p subshell is half-filled, an additional electron will have to enter a halfoccupied p orbital, and will be repelled by the electron already residing there.)

Electronegativity
The electronegativity coefficient of an element (pronounced as chi or [k]) is a
measure of the power of the atom of that element to attract electrons toward itself in
a molecule.
If an element has a strong tendency to attract electrons, it is called electronegative.
An element which has a low electronegativity is described as electropositive (which
has nothing to do with the sign of ).
There are three major definitions of electronegativity, one proposed by Pauling,
another by Mulliken, and the third by Allred and Rochow. They show the same broad
trends. The definition given by Mulliken is the simplest:

If an atom has a high ionization energy I1 and a high electron affinity Ea, then it will
acquire rather than lose electrons when it is a part of a compound, and hence is
classified as electronegative.
6

Electronegativity: Pauling scale

The electronegativity increases across the periods.


Group trends are more complicated
Variations in electronegativity across the
three series of transition metals are small.

Electronegativity: Allread-Rochow scale


A widely used scale, suggested by A. L. Allred and E. Rochow, relates to the force
exerted by the nuclear charge at the distance r = atomic radius () from the nucleus:

AR = 0.74 +

0.36 Zeff
r2

Table: AR coefficients for the s and p block elements.


Trend for
Note the large
decrease of
when n > 2

Bond covalence / ionicity


Electronegativity differences influence the polarity of chemical bonds and determine
their ionic character.
1. If two atoms A and B have identical , the A-B
bonding is covalent and the electrons are evenly
shared between the atoms. Example: H-H, 100%
covalent character.
2. If the electronegativity difference is < 2, the bond
is described as polar covalent.
Examples: C-H, C - H = 0.3; O-H, O - H = 1.3
3. If the electronegativity difference is 2, the bond
can be described as polarized ionic. Example: F-H,
F - H = 1.9. The distinction between polar
covalent and polarized ionic is more a question of
one's point of view than something that can be
measured experimentally.
4. The maximum electronegativity difference occurs
for CsF, F - Cs = 3.24, and it is assumed that in
this compound the charge separation is complete
and the bonding is 100% ionic.
9

CH225.6
Today:
Lewis structure
Resonance
For next time read Lecture 7

Part II. Structure and bonding


Chemistry begins when atoms and ions are joined together to form molecules and
molecular ions. In this chapter we focus on what holds atoms together and the
reasons why some molecular geometries are favored.
In 1916, Lewis defined a covalent bond as an interaction of two atoms which are held
together by a pair of electrons located between the two atoms. The two electrons are
shared by the two atoms.
Example: bonding in H2
The Lewis structure of H2 is very simple. Each hydrogen has one electron. In the
molecule of H2, each hydrogen attains the electronic configuration of the nearest
noble gas (helium, 1s2) by sharing the two electrons.
This is represented as H:H, or more simply HH. This representation tells us that
there are two electrons located between the two hydrogen nuclei, and that these two
electrons form the bond holding the nuclei together.
2

Lewis idea of bonding


The two positively charged protons repel each other. There is also some repulsion
between the two negatively charged electrons. However this is more than offset by
the four proton-electron attractions.

attraction

repulsion

The number of bonds x in a main-group compound other than H2 is predicted as:

x=

(2h + 8p) - number of valence e2

where h is the number of hydrogen atoms, and p is the number of p-block atoms.
The 2h + 8p is the maximum total number of electrons they can have in their valence
orbitals.
Note: that in compounds of elements from Period 3 and under, the actual number of
bonds is often greater than that predicted by the formula above.
3

Examples of Lewis structures


The following generalizations are helpful in constructing Lewis structures:
1. Hydrogen atoms are often terminally located
2. The least electronegative atom in a molecule is often located in the central
position and the more electronegative atoms occupy the outer (terminal) positions.
The great majority of inorganic compounds can be described by Lewis structures
where each atom achieves a closed shell electronic configuration.
Example: F2

x = (2 x 8 - 14)/2 = 1 bond

The resulting molecule, written F-F, demonstrates an important point. It has six pairs
of electrons which are not shared by the atoms. Such pairs of electrons are called
lone pairs.

Examples of Lewis structures


Example: O2

x = (2 x 8 - 12)/2 = 2 bonds

Oxygen needs two electrons to attain the octet structure. This places two shared
pairs between the two oxygen nuclei. Two pairs of electrons between two atoms
constitute two bonds, that is, a double bond. The bond order in the molecule is 2 and
O2 is written O=O.
Example: N2

x = (2 x 8 - 10)/2 = 3 bonds

Dinitrogen is written as NN to show the triple bond, and the bond order is 3.

Examples of Lewis structures


Example: NF3, nitrogen trifluoride

x = (4 x 8 - 26)/2 = 3 bonds

F
F

Example: ClO4, perchlorate ion

x = (5 x 8 - 32)/2 = 4 bonds

Examples of Lewis structures


Ozone

Nitrite ion

N
O

x = (3 x 8 - 18)/2 = 3
Sulfur trioxide

x = (3 x 8 - 18)/2 = 3
Sulfate ion

O
O

2-

O
O

x = (4 x 8 - 24)/2 = 4

x = (5 x 8 - 32)/2 = 4

Except in simple cases, a Lewis structure cannot predict neither the 3D shape of the
species (bond angles) nor the relative internuclear distances (bond lengths).
7

Examples of Lewis structures


Tetrafluoroborate ion

F
F

Orthophosphate ion

3-

x = (5 x 8 - 32)/2 = 4

x = (5 x 8 - 32)/2 = 4

O
N

Resonance
A single Lewis structure is often an inadequate description of a molecule. The Lewis
structure of ozone, for instance, suggests incorrectly that the two oxygen-oxygen
bonds are different. In fact, they are identical.
It is possible to draw as many as two Lewis structures to represent ozone. These
representations are called resonance structures or resonance forms.

1.28

O
O

1.28
116

The real structure is a single structure, it is just impossible to write down a single
Lewis structure to represent it.
The above diagram must be read in the following way. Ozone has two identical
oxygen-oxygen bonds. They have properties intermediate between those of typical
single O-O (1.48 ) and double O=O (1.21 ) bonds. The experimentally determined
oxygen-oxygen bond length in ozone is 1.28 .
9

Resonance
The two resonance forms of ozone are clearly degenerate. In other cases the
resonance structures may be non-degenerate. The cyanate ion, [CNO], is one such
example:

1.19
Computed APT atomic charges:

-0.89

1.22

+0.84

-0.95

The relative contributions of the resonance forms can be qualitatively established


based on electronegativity. In the case of [CNO], the structure has a slightly greater
contribution of the resonance form with a negative charge on oxygen.
10

CH225.7
Today:
Dative bond
Oxidation state
Hypervalence
For next time, read Lecture 8

Dative bond
A dative bond is a covalent bond which originates from donation of an electron pair
from one atom to another. The electron pair is usually a lone pair on the donor atom.

Example: Addition of ammonia to diborane. Note, x = (6 x 2 + 2 x 8 - 14)/2 = 7 in


H3N-BH3

Example: Grignard reagent, Et = -CH2CH3.

Oxidation state (number)


The oxidation states (numbers) are charges obtained after assigning as many
electrons as possible to the more electronegative atoms in a molecule.
Example: There are three oxygen atoms (configuration 2s22p4) in NO3. Each oxygen
can gain 2e. The oxidation state of O2 is -2. The total of 3 x 2 e = 6 e is assigned
to the three oxygen atoms. That is as if the nitrogen 'lost' 5 e. The oxidation state of
nitrogen is +5 in NO3.
O
O

1.25

-0.86
computed APT charges
+1.58

x
Oxidation numbers are formal and should not be confused with atomic charges.
3

Calculation of oxidation states


Some elements almost always have certain oxidation states.
fluorine always has an oxidation state of 1
in most compounds, oxygen has an oxidation state of 2
Exceptions: oxygen has an oxidation state of 1 in peroxides (e.g. H2O2), and of +2
in oxygen difluoride, OF2.
The alkali metals (Group I) have an oxidation state of +1 in virtually all of their
compounds.
The alkaline earth metals (Group II) have an oxidation state of +2 in virtually all of
their compounds.
The halogens (Group VII) have an oxidation state of 1 unless they are directly
combined with oxygen or with another halogen.

Calculation of oxidation states


Hydrogen is assigned an oxidation state of +1 in combination with non-metals (N, O,
P, S, halogens) and an oxidation state of -1 in combination with all metals.
There are borderline cases (read next slide). The electronegativities of H, B, C, and
Si are similar. Therefore, it is better to assume that these atoms do not oxidize/reduce
each other.

Atomic groups which exist as molecules can be treated independently when form a
compound.
Examples: H3NBH3

N(-3) and H(+1) as in NH3.

ammonia
CH2CH3
Mg
Et2O
Br
Et2O

ether

Hypervalence
When the total number of electrons around a main-group atom exceeds 8, the
species is called hypervalent.
The term 'hypervalence' may suggest that compounds which have more than 8
electrons are very unusual and are adopting an unconventional bonding mode. In
fact, there are so many examples of 'hypervalent' compounds of the main-group
elements from Periods 3 to 6 that this terminology is simply misleading.
Examples: PCl5 and SF6

Cl
Cl

F
Cl
Cl

Cl
10 e- around P

F
F

F
S

F
12 e- around S
7

CH225.8
Today:
Molecular symmetry
Symmetry elements and operations
For next time, read Lecture 9

Molecular symmetry
Molecules have shape. Chemist use names describing different shapes. For
example, the molecule of PCl5 has a trigonal-bipyramidal shape, and SF6 has an
octahedral geometry.

A systematic description of molecular shape involves consideration of molecular


symmetry.
Many molecules have at least one symmetry element and belong to distinct
symmetry point groups.
For example, PCl5 belongs to the D3h group, and SF6 belongs to the Oh group. It is
important to know what these names mean.
2

Symmetry elements and operations


There are three symmetry elements, which are imaginary geometrical entities: a line
(axis), a plane, and a point.

A symmetry operation is a prescribed movement of atoms about a symmetry element


such that the molecule looks the same (indistinguishable) before and after the
movement.
Indistinguishable means that if you turn away while someone performs a symmetry
operation on a molecule, you would not be able to tell afterwards whether the
molecule was actually manipulated in some way. Indistinguishable means looks the
same but does not necessarily mean is the same, that is identical.

Symmetry operations
According to the treatment of symmetry in the mathematical Group Theory there are
five symmetry operations:
Identity, E
Rotation, Cn
Reflection,
Inversion, i
Improper rotation, Sn
Identity
The simplest of all symmetry operations is identity, given the symbol E. It exists for
every molecule, because the molecule itself exists. When the operation of identity is
applied to a molecule, it results in an identical product. A molecule that possesses no
symmetry other than E is said to be asymmetric.

Symmetry elements and operations


Symmetry Element

Symmetry Operation

1. Line (proper axis)

Rotation by an angle
360o/n about the axis, Cn.

Examples:

C2

C3

F
N

difluorodiazene

H
H

ammonia

1.21
1.37

106

1.01

108

Symmetry elements and operations


Symmetry Element

Symmetry Operation
Reflection through (into) the plane, .

2. Mirror plane
Examples:

F
N
F

plane of the
molecule

3. Center of symmetry (point)


Examples:

Hb
Ha

Hb is a
reflection of Ha

Inversion through the center of


symmetry, i.

F
N

Rotation
This operation is referred to as n-fold rotation, where the rotational angle = 360/n; it
is designated by the symbol Cn. If a molecule has rotational symmetry, the operation
should bring every atom of the molecule into an equivalent position.
The following figure shows the effects of successive four-fold rotations of a planar
MX4 molecule. An example of this can be [PtCl4]2-

Two successive C4 rotations have an effect of a single C2 rotation. Three C4 rotations


have the effect of a single rotation in the opposite direction, C43 = C4-1. Carrying out
four C4 operations is equivalent to E.
7

Rotation
The MX4 example shows that a C4 axis is associated with only two unique symmetry
operations: C4 and C43.
C42 = C2, meaning that the molecule also has a C2 axis collinear with the C4 axis.
There are four other C2 axes in a planar MX4 molecule, as shown below:

The value of n of a Cn rotation is the order of rotation. The highest-order rotational


axis in a molecule is called the principal axis. In a planar MX4 molecule, the principal
axis is C4.
8

Reflection
The operation of reflection, , defines bilateral symmetry about a plane called the
mirror plane or reflection plane. Accordingly, if a molecule possesses a mirror plane,
it is bisected by this plane. Then for any atom at a distance r from the mirror plane
there will be an equivalent atom at a distance r.

-r

90o

Lets consider again the example of a planar compound MX4

Reflection
The figure on this slide shows five mirror planes found in a square planar
molecule MX4.
C4
Horizontal mirror plane, h, is a
symmetry plane perpendicular to the
principal rotational axis. For planar
molecules lacking rotational
symmetry, the molecular plane is h.
Vertical mirror plane, v, is defined
as containing the principal axis and
perpendicular to h (if h exists), and
containing a greater number of
atoms than d.
Dihedral mirror plane, d, has most
of the properties of v but typically
contains bond angle bisectors.
10

Reflection
This figure illustrates the effect of the five operations of reflection in MX4:

The effect of v is to transpose a pair of trans-related X atoms

The effect of d is to exchange cis-related X atoms

11

Reflection
Note that for any mirror plane, two successive reflections about the same plane bring
the molecule into its original position:

= 2 =
Thus, a mirror plane is associated with only one symmetry operation (unlike a
rotational axis).
The operation of h transforms any directional property of MX4 into the negative of
itself (consider the example of a pz orbital below).

12

CH225.9
Today:
Symmetry operations: inversion and improper rotation
For next time, read Lecture 10

Inversion
The operation of inversion is defined relative to a point which is called the inversion
center or center of symmetry. This point is usually taken as the origin of the
Cartesian coordinate system (x = 0, y = 0, z = 0).

z
i

A (x, y, z)

y
x

(-x, -y, -z) A


When inversion exists for a molecule, for every atom with coordinates (x, y, z) there
must be an equivalent atom with coordinates (-x, -y, -z). Both the element and the
operation of inversion are given the symbol i. Molecules that have inversion
symmetry are said to be centrosymmetric.
2

Inversion
In a centrosymmetric molecule, a line drawn from any atom through the inversion
center will connect it with an equivalent atom at an equal distance on the other side
of the molecule.

z
i

A (x, y, z)

y
x

(-x, -y, -z) A


Performing inversion twice should bring every atom back into itself, that is i i = i2 =
E. Like a mirror plane, an inversion center has only one operation associated with it.
Furthermore, there can be only one inversion center in a molecule.

Centrosymmetric molecules
A square-planar molecule MX4 is centrosymmetric. Another example is an octahedral
MX6 molecule. The central atom M is located at the inversion center i in both.

M
X

An M2X6 molecule in the staggered conformation (left) is centrosymmetric. The


inversion center is at the midpoint between the two M atoms. Note that there is no
inversion center in the eclipsed conformation (right).

X
M
X

i
X

X X

M
X

M
X

Improper rotation
Improper rotation is the last of the symmetry operations and it is a strange one.
Another name for this operation, rotation-reflection, gives a better description of the
motion it involves.
Rotation-reflection consists of a rotation followed by a reflection in a plane
perpendicular to the rotational axis. The axis is called an improper axis. Both the
operation and the element are given the symbol Sn, and the rotation angle is 360/n.

Sn = Cn
The two parts of Sn (Cn and ) may be genuine (existing) symmetry operations of the
molecule in their own right, but often they are not. A good example is a tetrahedral
MX4 molecule (CH4, NH4+, BH4-, AlH4-) which has an improper S4 axis, but does not
possess a C4 symmetry axis.
Like proper rotations, a series of improper rotations can be performed. Two S4
operations have the effect of a single C2 operation: S4 S4 = S42 = C2. Carrying out
four S4 operations, written S44, results in the original configuration, that is S44 = E.
5

Improper rotation of a tetrahedral molecule


Example: Successive S4 operations on the tetrahedral CH4 molecule.

First S 4

S4

C4

Second S 4

C2

S4
C4

Symmetry elements of a tetrahedral molecule


In a tetrahedral MX4 molecule, there are three S4 axes, which lie along the bisectors
of the X-M-X angles, and are collinear with the three C2 axes. The three S4 axes
define the x, y, and z directions of the Cartesian coordinate system.

C2, S4
z
X

C2, S4

x
C2, S4
X

In this highly symmetric system all three directions are indistinguishable (equivalent),
that implies that the three S4 axes are also equivalent. Each has two operations
associated with it: S4 and S43. Consequently, the three S4 axes give rise to a total of
6 operations belonging to a class designated 6S4 in tabular listings.
7

Improper rotation
Example: Successive S6 operations.

1st S6

C6

5
3

1
6

2
2nd S6
3

C3

3
1

C6

6
2
5

6
3
4
2

CH225.10
Today:
Symmetry point groups

Symmetry point groups


Every molecule has a set of symmetry operations. This set of symmetry operations
defines the point group of the molecule. The word point is traditionally used and
indicates that if a molecule has more than one symmetry element, then they have at
least one common point, for example, it can be the point of intersection of two
symmetry axes.
Group theory, the mathematical treatment of the properties of groups, can be used to
analyze the molecular orbitals, vibrations, and other properties of molecules.
The chemically important symmetry point groups fall within four general categories:
(1) non-rotational, (2) single-axis rotational, (3) dihedral, and (4) cubic.
With only a few exceptions, the rules for assigning symmetry point groups are simple
and straightforward.
A table on the following slide lists families of point groups designated according to
the Schnflies notation.
2

Common Symmetry Point Groups

Systematic point group assignment


Identifying a molecule's point group by finding all symmetry elements is not a
practical approach. The best way is to look for key symmetry elements in the
following sequence.
look for a
rotational axis
'Special
groups'
Nonrotational
groups

look for a C 2 axis


perpendicular to C n

Single-axis
rotational
groups

Dihedral
groups

Examples of point group assignment


The 'special groups' are relatively easy to identify, for example
HCl has Cv symmetry (all linear non-centrosymmetric molecules have Cv symmetry)
CO2 has Dh symmetry (all linear centrosymmetric molecules have Dh symmetry)
CH4 has the tetrahedral (Td) symmetry

SF6 has the octahedral (Oh) symmetry

Examples of point group assignment


Example: SF4

C2v

The highest-order proper axis in SF4 is C2.


v'

1.68
1.58

FSF = 172

C2

FSF = 102

F
S

F
F

There is no other C2, therefore, SF4 must belong to a single-axis rotational group.
Since it has two vertical mirror planes, it belongs to the C2v group .
Example: PF5

D3h
C3

1.59
1.56
FPF = 180

FPF = 120

C2

C2

F
F

C2

There is a C3 axis, and there are 3 C2 axes perpendicular to it. There is also h the group is D3h

Examples of point group assignment


Example: H2O2

C2
0.96

HOOH =119

HOO =101

1.43

As indicated in the figure, there is a C2 axis and no other proper axis. There are no
planes of symmetry. The group is C2.
Example: Allene, C3H4
HCH =118

D2d

1.30

1.085

There is a C2 axis lying along the C-C-C axis. There are two more C2 axes
perpendicular to the first. There are two vertical mirror planes v: one containing
HaHbC and the other HcHdC. The group is D2d.
7

Examples of point group assignment


Example: C6H6

D6h

There is a C6 axis perpendicular to the plane of the ring and 6 C2 axes in the plane
of the molecule. Since there is a h, the group is D6h.
Example: N2F2

C2h

C2
F
N
h

There is a C2 axis perpendicular to the plane of the molecule. Since there is a h, the
group is C2h.
8

Non-rotational groups: C1, Cs, Ci


Group order: The number of symmetry operations found in a group defines the order
of the group, designated h.
With their low orders (h = 1 or 2) the non-rotational groups represent the lowest
symmetry point groups. C1 is the point group of asymmetric molecules (no symmetry
other than E ):

The group Cs describes bilateral molecules (only one mirror plane). The group Ci
whose only non-identity operation is inversion i is an uncommon group.
Br

Cl

i
Cl

Br

Ci
9

Single-axis rotational groups


Cn, the simplest family, consists of Cn rotations.
Example: 1,2-Dichloro-cyclobutane

Cnv groups possess one Cn


rotational axis and n vertical
mirror planes, v.

Cnh groups possess one Cn axis


and a horizontal mirror plane, h.
Since h C2 = i and h Cn = Sn,
these groups also have an n-fold
improper axis, and they are
centrosymmetric when n is even.
10

Dihedral groups
Dihedral groups have n C2 axes perpendicular to the principal n-fold axis Cn. The
C2 axes are called dihedral axes. A Dn group can be thought as a Cn group to which
n C2 operations have been added. Dn groups are rare in chemistry.

Dnd groups can be generated by addition of n C2 axes to Cnv

11

Dihedral groups
The Dnh family bears the same relationship to the Cnh
H

Cl
Cl

D3h

C2
Cl

C2

HH

C3
D3h

All linear centrosymmetric molecules belong to the Dh group.

12

Optical activity and symmetry


A chiral molecule is a molecule which cannot be superimposed on its own mirror
image. The two structures are called enantiomers. Compounds existing as
enantiomeric pairs are called dissymmetric.
Some dissymmetric molecules are asymmetric (i.e. belong to the group C1). In
general, dissymmetric molecules may have symmetry other than C1.
Example of a chiral molecule :
H
F

C1

H
Br
Cl

Br
Cl

C1

Note: a molecule containing a mirror plane or an Sn axis cannot be chiral because its
mirror image must be superimposable.

13

Optical activity and symmetry


Two types of groups are potentially chiral: Cn and Dn.
Note that a Cn or Dn molecule can be non-chiral, if it is stereochemically non-rigid.
Example: C2 symmetric chiral metal complex

C2
X
X
N

M
N

N
N

N
N

H2N

=
H2N

Ethylenediamine

C2
Example: Non-chiral molecule of H2O2 where the OH groups rotate easily about the
O-O bond.
H
O

C2

O
H

14

CH225.11
Today:
VSEPR theory
CH225 website:
Recommended problems (Problem Set A)
Solutions (Solutions to Problem Set A)
For next time, read Lecture 12

Molecular geometry: the VSEPR model


Valence Shell Electron Pair Repulsion theory (VSEPR) is a method for predicting the
shape of molecules, based on the electron pair diagrams. It was developed by
Sidgwick and Powell in 1940 and further improved by Gillespie and Nyholm in 1957.
VSEPR is a remarkably simple approach, unsurpassed as a handy predictive
method for the main group compounds (it does not work with d-block metal
complexes). The correct geometry is nearly always predicted, and the exceptions are
often rather special cases. It must be remembered, however, that VSEPR provides
information about the molecular geometry, and not about chemical bonding.
The VSEPR method is based on the following assumptions:
1. Bond electron pairs and lone pairs are negatively charged and adopt positions as
far apart from each other as possible.
2. Lone pairs occupy more space than bond electron pairs.
3. Double bonds occupy more space than single bonds.

Preferred arrangements of electron pairs


Best arrangements of electron pairs around an atom are found using simple
geometrical considerations; they are intuitively obvious.
2 e- pairs
linear

3 e- pairs
trigonal
planar

4 e- pairs
tetrahedral

trigonal
bipyramidal

octahedral
5 e- pairs

6 e- pairs
3

VSEPR: How it works


The following procedure is recommended.
1. Draw the Lewis structure.
2. Modify the Lewis structure by assigning all atoms or groups bonded to the central
atom as singly (-), doubly (=), or triply () bonded according to the following chart:
(Note: consider all double bonds as consisting of one and one bond; consider all
triple bonds as consisting of one and two bonds).

VSEPR: How it works


continued from previous slide
3. Determine the number of valence electrons in the central atom.
4. Count all bonds around the central atom and add this number to the number in
line 3 (the molecular geometry is dictated by bond electron pairs).
5. The electrons contributed by the central atom in any bonds should be
'discounted' by subtracting one electron for each bond from the electron count.
(A shortcut: Considering point 4, one can ignore any double bonds when counting
electrons in VSEPR)
7. If the molecule is charged, the corresponding number of electrons is either added
(for a ve charge) or subtracted (for a +ve charge) from the electron count.
8. Dividing the resulting number of electrons by 2 gives the number of electron pairs
around the central atom. Now a geometry can be assigned.
It is appropriate to examine some examples.

VSEPR: examples
Ozone, O3
Lewis structure:

Modified Lewis structure:


O O O
central atom: oxygen
valence electrons in oxygen:
2 O contribute 1 e- each in the s bond
Central O contributes 2 e- in the p bonds

Group: C2v

6
2
-2
6
3

total:
number of electron pairs:
3 electron pairs: trigonal arrangement

note: one can


ignore double
bonds

angular molecular geometry


Ammonia, NH3
H

Lewis structure:
H

central atom: nitrogen


valence electrons in nitrogen:
3 H each contribute 1 e-

H
H

H
N

5
3

total:
8
number of electron pairs:
4
4 electron pairs: tetrahedral arrangement

Group: C3v
H
H

trigonal pyramidal molecular geometry

VSEPR examples
Group: D3h
B-F
1.31

Group: Oh
P-F
1.63

VSEPR examples

Group: D3d
N-H 1.01

B-N 1.65

B-H 1.21

VESPR examples
Propene, H3C-CH=CH2
CH2

Lewis structure:
H3C

central atom: carbon


valence electrons in carbon:
1 CH3 group contributes 1 e1 H atom contributes 1 e-

H
H

H3C

4
1
1

H
C

Group: Cs

(the =CH2 group can be ignored)

C=C 1.33
total:
number of electron pairs:
3 electron pairs: trigonal geometry

6
3

C-C 1.50

C-H 1.09

VESPR examples
Trifluorothionitrile,

F3S N
N

Modified Lewis structure:


F

central atom: sulf ur


valence electrons in sulfur:
3 F each contribute 1 e1 N contributes 1 e- in the

6
3
bond

substract 2 e- contributed by S in the 2

F
F

Group: C3v

1
bonds -2

S-N 1.42
S-F 1.60

total:
number of electron pairs:
4 electron pairs: tetrahedral geometry

8
4

NSF 123
F-S-F 93

10

CH225.12
Today:
VSEPR: Examples of applications
For next time: Lecture 13

VESPR examples

Group: C2v

The shape of ClF3


The case of ClF3 is interesting. There are three possible ways of placing two lone
pairs in the trigonal bipyramidal arrangement of five electron pairs:
F
F

Cl

axial
site
F

Cl

Cl

equatorial
site

F
F

There are different repulsive interactions in A, B, and C:


1. <lone-pair><lone-pair>: lp-lp
(the strongest)
2. <lone-pair><bond-pair>: lp-bp
(lp-bp repulsion is stronger than bp-bp repulsion)
3. <bond-pair><bond-pair>: bp-bp
Note: Only interactions at angles ~90 are important and should be considered.
Structure B has the unfavorable lp-lp repulsion.

The shape of ClF3


There are six important interactions in every isomer of ClF3:
All important interactions
structure

repulsions

4 lp-bp
2 bp-bp

1 lp-lp
3 lp-bp
2 bp-bp

(Most unfavorable)

After cancellation of common repulsions


structure

repulsions

2 bp-bp

2 lp-bp

6 lp-bp

Considering structures A and C involves comparing


Cl-F 1.60
4 lp-bp and 2 bp-bp repulsions with 6 lp-bp
repulsions. This can be reduced to comparing 2 bpbp repulsions with 2 lp-bp repulsions, indicating that
A is more favorable than C.
The experimental structure is indeed T-shaped:

Cl-F 1.70

FClF 87.5
4

Nitrogen dioxide
When a molecule has an unpaired electron, this electron can be treated as a half
filled lone pair.

Group: C2v

NO2 vs. NO2


It is interesting to compare the structures of nitrogen dioxide, NO2, and the nitrite
anion NO2-.

1 e
ONO =134

ONO = 116
N-O = 1.18

N-O = 1.25

When the lone-pair is half filled in NO2, it demands less space than a bond pair, and
the O-N-O angle opens to 134o from the ideal trigonal angle of 120o. Addition of one
electron gives the nitrite ion, NO2-, which lone-pair occupies more space than a
bond pair and repels the two oxygen atoms. The O-N-O angle is reduced to 116.

The influence of multiple bonds


VSEPR theory considers that double and triple bonds exert stronger repulsion than
single bonds.
Examples:

Calculated DFT structures:

S-F = 1.56
S-F = 1.55

C-O = 1.17

C-F = 1.31
Group: C2v

S-N = 1.41

Group: C3v

S-O = 1.41

S-F = 1.60
Group: C2v
7

The influence of multiple bonds


The strength of repulsion between electron pairs follows the order:
lp > or ~ mbp (multiple bond) > sbp (single bond).
Examples:

Calculated DFT structures:


Se-Cl = 2.20

Group: Cs

Se-O = 1.59

Group: Cs

Group: C4v
8

How you should not count electrons in VSEPR


A different way of counting electron in VSEPR is preferred by some students. It
involves arranging ALL valence electrons in a molecule in a reasonable way to
achieve octet electron configurations around all (or most) of the atoms.
Example: Arranging 18 electrons in SO2. There
is a lone pair on sulfur the structure is angular
(bent).
Problems:

Total: 3

6 e = 18 e (9 e.p.)

1) This approach is time-consuming, because it requires handling ALL valence


electrons.
2) It is easy to make simple arithmetic mistakes when you work with large numbers.
3) The number of electrons around the central atom in a molecule can often be 8
(e.g. 6, 10, 12), that can be confusing and make you feel unsure.
4) Accurate Lewis structures can be difficult to draw for many compounds! This can
complicate making the decisions about the shape. For example, many students in
this class decided that SO2 was a linear molecule with two double bonds: O=S=O
(possibly by analogy with the linear molecule of O=C=O?)

CH225.13
Today:
Valence bond theory
Hybrid orbitals
For next time: read Lecture 14

Bond theories
The VSEPR method can be used to predict the shape of molecules. However, this
method does not tell much about the nature of bonds between the atoms.
There are two principally different ways of proceeding further. One involves regarding
bonds as localized interactions involving two electrons for every bonded pair of
atoms, while the other approach considers valence electrons delocalized over the
whole molecule in different molecular orbitals.
The localized bond approach is called the valence bond (VB) theory, which
development is associated with the names Walter Heitler, Fritz London (1927), and
Linus Pauling (in the 1930s). Both historically and conceptually, the VB model is an
extension of the Lewis idea of bond formation through sharing of electron pairs.
The delocalized bond approach is executed in the molecular orbital (MO) theory. VB
theory is less useful than the MO Theory. Modern Inorganic Chemistry texts have all
but abandoned instruction of hybridization, except as a historical footnote. The
following slides provide a brief discussion of the VB theory.
2

Valence bond theory


In the VB approach, the electron density associated with a bond is the result of an
overlap between appropriately oriented atomic orbitals. This presents an apparent
difficulty, since the conventional atomic orbitals do not always have the right
orientations. That is, they are either oriented at 90 (p, d), or they are non-directional
(s).

Lets consider the example of tetrahedral CH4 and [NH4]+ molecules which have four
equivalent bonds; the bond angles are all ca. 109.5.
H
H C H
H

Hybrid orbitals
The hybridization theory was promoted by Linus Pauling. He suggested that
formation of hybrid orbitals can be achieved by 'mixing' the conventional s and p
orbitals, through mixing the wave functions of the atomic s, px, py, and pz orbitals:

sp3 hybrid orbitals


1 = 1/2 (s + px + py + pz)

Contour diagram of a single


sp3 hybrid orbital

2 = 1/2 (s + px - py - pz)
3 = 1/2 (s - px + py - pz)

A conical node

4 = 1/2 (s - px - py + pz)

The resulting hybrid orbitals all have the same shape and they are oriented at
~109.5 from each other.

Hybrid orbitals
Different hybrid orbitals can be constructed as combinations of conventional atomic
orbitals. This mixing of orbitals has the effect of concentrating the electron density
into more specific regions of space and shifting the nodal surfaces. Hybrid orbitals
thus can better overlap with orbitals of adjacent atoms.
Example: spn hybrids.

70.5o
spn hybrids:
% orbital character

66.0o

sp3
s+p

sp2
1

/3

s + 2/ 3 p

54.5o

sp
s+p
5

Molecular shape and hybrid orbitals

The use of hybrid orbitals


For main group atoms, the predominance of the s and p valence orbitals, and the
relative unavailability of the d orbitals simplifies the choice of hybridization schemes
to sp, sp2, and sp3.
Hybrid orbitals can be used to
generate - bonds
accommodate lone pairs of electrons
act as acceptor orbitals in dative bonds
If a molecule has a lone pair, it is considered that this orbital typically receives a
greater contribution from the most stable valence atomic orbital: s - for the main
group atoms. For example, NH3 can be described in terms of four sp3 hybridized
orbitals where the three bonding orbitals have approximately 10% s character,
because of the increased s contribution to the lone pair orbital.
Hypervalent compounds such as PF5 (trigonal bipyramidal) and SF6 (octahedral)
present problems. Although sp3d and sp3d2 hybridization schemes are geometrically
feasible, they are energetically unfavorable by virtue of the relatively high energy of
3d vs. 3s, p orbitals.
7

VB Theory: Examples of application


Beryllium dihydride, BeH2

Boron trihydride, BH3

VB Theory: Examples of application


Boron trifluoride, BF3

F
B

Tetrahydridoborate, [BH4]-

Important comments
Many students (even some chemists) actually believe that molecules such as NH3,
H2O, CH4 indeed have hybridized orbitals. This is a mistake.
Orbital hybridization is merely a convenient trick that makes simple the description of
bonding in some compounds.
From a pedagogical perspective, hybridization approach tends to over-emphasize
localization of bonding electrons. In reality, electrons are delocalized in the
molecules.
Although the language and pictures arising from Valence Bond Theory remain
widespread in organic chemistry, this qualitative analysis of bonding has been largely
superseded by Molecular Orbital Theory in other branches of chemistry.
Hybridization will be used in this course, CH225, only to describe bonds which will be
relatively unimportant in the molecules.

10

Problems of the VB theory


VB theory finds its use mainly in organic chemistry and mostly concerns C, N and O
(and to a lesser extent P and S). However, it does have problems even with simple
compounds. For example, the molecule of O2 has a double bond and four lone pairs:

Valence bond interpretation of bonding:

All electrons in the model above are apparently paired, either in bonds or as lone
pairs. This representation cannot explain the experimental fact that O2 has two
unpaired electrons! Clearly, a better model is required to explain paramagnetic
molecules which have one or more unpaired electrons, like O2.

11

CH225.14
Today:
and Molecular orbitals
MO energy level diagrams of diatomic molecules (H2+, H2, H2-)

Molecular orbital (MO) theory


The fundamentals (all the math!) of the MO theory are usually treated in
Theoretical Chemistry or Quantum Chemistry. For applications, the theory is
implemented in what is called Computational Chemistry. There are several
commercial programs (Gaussian, ADF, Spartan, etc.) which allow performing
quantum-chemical calculations routinely. Computational chemistry is now an
accepted predictive tool of chemistry, used by organic, inorganic and physical
chemists.
A typical calculation in Computational chemistry results in an optimized
molecular geometry and gives the molecular energy and the energies of the
molecular orbitals in the molecule. All of the MO energy diagrams which will be
discussed in this course have been obtained with the help of quantum-chemical
calculations with the Gaussian 03 program.
Our objective is to develop a good understanding of the molecular orbitals and
MO diagrams. Students interested in theoretical and computational chemistry
are invited to take CH313 and CH444.
2

Interference of waves
Any waves, including sound, light and electron waves, can interact in a constructive
or destructive fashion.

1 + 2
1= 2

Reinforcement
of waves

1 - 2
1= -2

Destruction
of waves

s-orbital overlaps
The overlap of electron orbitals can be treated similarly. For example,
In-phase interaction of 1s orbital wave functions:
Bonding
1

MO

Out-of-phase interaction of 1s orbital wave functions:


Antibonding * MO
1

The simplest molecule: H2+


H2+ ions can be generated by shining light of an appropriate frequency on
dihydrogen, H2. The ion is made up of two protons and one electron. The electron in
H2+ occupies a molecular orbital ( MO). This MO has an ellipsoid shape, elongated
along the H-H axis. All MOs which have cylindrical symmetry about the bond are
given the name . The star * is used to distinguish bonding and antibonding,
and *, MOs.

Highest occupied
MO: HOMO

Lowest unoccupied
MO: LUMO
- bonding MO

* - antibonding MO

Bond order = (# e- in bonding MO - # e- in antibonding MO)


2

The bond order in H2+ is 1/2. The single electron cannot hold the two protons
together as well as the two electrons in H2. The bond in H2+ is weaker (61 kcal/mol)
than that in H2 (103 kcal/mol). It is also longer (1.06 ) compared to H2 (0.74 ).
5

The H2+ MO energy level diagram


The energy of the electron in H2+ has been determined by calculations as well as
experimentally. The MO has a lower energy than that of the hydrogen 1s AO.
MO energy level diagram for H2+

The calculated energy of the * MO is higher than the energy of the hydrogen 1s AO.
* MO is not occupied in H2+, but can be occupied in other molecules.
6

The MO diagram of H2
The two electrons of H2 occupy the MO and they are spin paired because the
HOMO-LUMO energy gap is greater than the electron repulsion energy in the MO.
The rules used for atomic orbitals (Pauli exclusion principle and Hund's rule) operate
for molecular orbitals.
AO

2 MO

AO

LUMO
*
1s

1s

HOMO
electronic bond energy
H atom

H2 molecule

H atom

Since there are now two electrons in the MO, the bond order of H2 is said to be
one (single bond).
7

The MO diagram of H2The H2- ion is a combination of an H atom and an H (hydride) ion. In H2, two
electrons occupy the MO and the third electron must occupy the * MO. The bond
order of H2 can be calculated as (2e- - 1e-)/2 = 0.5, and is the same as in H2+. The
calculated H-H distance in the H2 ion is 0.92 ; it is longer than the 0.74 bond in
H2.
AO

2 MO

AO

*
1s

H atom

1s

H2- ion

H- ion

Note: The bond order of H22 or He2 is (2e- - 2e-)/2 = 0, that is the energy stabilization
gained by the two electrons in the bonding MO is cancelled by the two electrons in
the antibonding MO. Since there is no chemical He-He bond, He2 is unstable.
8

MO diagram of H3+ ion


H3+ ion is the most abundant ion in the universe! H3+ exhibits trigonal planar
geometry, and forms an equilateral triangle (H-H = 0.87 ); it has two electrons which
are shared by the three protons. In H3+, the two electrons occupy 1 MO which is
produced by in-phase mixing of the three 1s AOs.
E
(Hartree)
+0.15
o

*
2

-0.19
-0.59

-1.21

1s

1
H2

H3+

1 Hartree =
627.5 kcal/mol
H+

Note: Three MOs are formed when three AOs are involved in bonding. There are
two unoccupied antibonding MOs in the H3+ ion (2 and 3 are degenerate). The
formation of H3+ from H2 and H+ is exothermic by ca. 100 kcal/mol.
9

Mixing of p orbitals
In diatomic molecules different from H2, some molecular orbitals are formed by
mixing of p orbitals. In a diatomic molecule, the internuclear axis is usually taken as
the z-axis. Thus, the pz orbital is unique and different from the px and py orbitals.

Just as two -type (cylindrical) MOs arise from the in- and out-of-phase mixing of
two s orbitals, two MOs arise from mixing of two pz orbitals.

pz

-pz

pz

pz

* MO

MO

10

px and py orbital mixing


These orbitals can overlap in a side-on fashion. The molecular orbitals formed this
way have no rotational symmetry about the bond axis and are given the name .
The in-phase combination of two px (or two py) orbitals:

The out-of-phase combination of two px (or two py) orbitals:

-p

Note: the two lobes of the MO together constitute one molecular orbital, whereas
the four lobes of the * MO also constitute one molecular orbital.
11

Possible and MOs

px px (, *)

12

Important properties of MOs


1. The electron density is low in the region of the internuclear axis, between the
nuclei, which are relatively exposed to each other.
2. The overlap of two px or py orbitals in the side-on fashion is less efficient than the
head-on overlap of two pz orbitals.
3. As a consequence, bonding is less strong than bonding. In other words,
MOs are expected to be higher in energy than MOs (which is not always the
case for reasons that we are going to have explained later).
4. In a linear molecule, the combination of two px orbitals to give a MO has the
same energy as the combination of two py orbitals. That is the two MOs are
degenerate.
5. The term antibonding is confusing and should not be taken literally. It implies an
out-of-phase combination of AOs, however it does not mean that an anti-bonding
orbital is always bad for bonding. In a molecule, the MOs are ranked according
to their energy. The lower is the energy, the better is the orbital (no matter how it
is named).

13

CH225.15
Today:
MO diagrams of diatomic molecules (F2, O2, N2, C2, B2)
Orbital mixing
Photoelectron spectroscopy
Read for next time: Lecture 16

Bonding in F2
The bond order is (8 - 6)/2 = 1. The bond is of the -type, because there is the same
number (two) of occupied and * MOs. The molecule of F2 has no unpaired
electrons. The bond energy of F2 is 37 kcal/mol (which is actually rather low) and the
F-F bond length is 1.41 .
Molecular orbitals of F2:
4
F2
F

LUMO

F
2

HOMO

2 px,y

2 px,y
1

2 pz

2 pz
3
2

2s

2s

2
1

Bonding in O2
The electronic configuration of O is 2s22p4. The degenerate * MOs are singly
occupied according to Hund's rule. There are two unpaired electrons in O2. MO
theory successfully explains the paramagnetism of O2 and also the reactivity of O2
which behaves as a diradical. The bond order is (8 - 4)/2 = 2.
O2
O
2 px,y

LUMO

*
HOMO

2 pz

The LUMO of O2
O
2 px,y

2 pz

The HOMO (SOMO) of O2


2

2s

*
2s

Removal of one electron from O2 results in O2+, which has a shorter (1.11 vs. 1.20
in O2) and stronger (154 vs. 118 kcal/mol in O2) bond. This correlates with the higher
bond order (8 - 3)/2 = 2.5 in O2+.
3

Energies of atomic orbitals


In dealing with molecules, it is important to have estimates for the energies of the
atomic orbitals of the constituent atoms. For this purpose, the energies of the
valence 2s and 2p orbitals, given in the following figure, are useful.

It is important to notice that the 2s and 2p orbitals have quite different energies in O
and F, however they become closer in the atoms of C and B.
4

Mixing of s and p orbitals


Orbitals of similar energy mix well. Mixing of orbitals is shown by dotted lines in MO
diagrams. Mixing of two 2s and two 2pz orbitals gives four molecular orbitals: 1-4.
As a result of the s/p mixing, 2 is slightly stabilized, but 3 is somewhat destabilized
and appears at a higher energy, above the MOs, in the molecule of N2.

out-of-phase combination

2s AO

2p AO

in-phase combination

2s AO

2p AO

Bonding in N2
The electronic configuration of N is 2s22p3. The bond order in N2 is (8 - 2)/2 = 3. The
molecule of N2 has a triple bond which is very short, 1.10 . It is interesting to note
that 3 has reduced electron density in the N-N bond region.

N2

The shape of 3 in N2

The shape of 3 in F2

Photoelectron spectroscopy
Photoelectron spectroscopy (PES) is an experimental technique which can be used
to determine relative energies of molecular orbitals. A sample is typically ionized by
using hard radiation at 21.2 eV (1 eV = 23.1 kcal/mol). The kinetic energy of ejected
electrons, EK, can be measured. Then, the ionization energy is calculated as I = 21.2
- EK. Electrons in different orbitals have different ionization energies. A photoelectron
spectrum of a molecule shows the MO energy levels.

Bonding in C2
The diatomic C2 is not a species that can be put in a bottle, but it can be
experimentally examined under special conditions. The electronic configuration of C
is 2s22p2, so carbon possesses four valence electrons. The bond order calculation
for C2 gives (6 - 2)/2 = 2. The double bond is composed of two bonds unsupported
by a bond. The calculated C=C bond length is 1.25 .
C

C2

The LUMO of C2
3

The HOMO of C2

Bonding in B2
The diatomic B2 is another unstable species. The electronic configuration of B is
2s22p1, therefore, there are six electrons to place into the MO diagram. The bond
order calculation gives (4 - 2)/2 = 1. The four electrons in the 1 and 2 orbitals
largely cancel each other out for the purpose of calculating the bond order. In this
case the single bond is made up from two half-filled orbitals. The B-B bond is quite
long: 1.59 .
B

B2

Bond lengths, orders, and strengths.


Higher bond order correlates with shorter bond length and greater bond strength.
The correlation of bond length and bond order

The correlation of bond strength and bond order

1.7

1000

N2

1.6

1.4

F2

1.3
1.2

O2C2
N2

O2

1.1
1.0

800
Bond energy, kJ/mol

Bond length

B2

1.5

O2 +

2
Bond order

C2

600

O2-

400

B2

200
0

O2+

O2

F2

Bond order

10

CH225.16
Today:
MO diagrams of heterodiatomic molecules (HF, CO)
Isoelectronic molecules and ions (CO, CN-, NO+)
Metal carbonyls

Heteronuclear diatomic molecules


The principles developed for homodinuclear molecules also operate for
heterodinuclear molecules. However, the AOs of different atoms have different
energies. This gives the MO diagrams a skewed appearance. Heteronuclear bonds
are polarized. The bond electrons tend to be found on the more electronegative
atom.
identical atoms

similar atoms

different atoms

antibonding MO

bonding MO

more
electronegative
atom

much more
electronegative
atom

The more electronegative element makes a greater contribution to the bonding MOs,
and the less electronegative one makes a greater contribution to the antibonding
MOs. The bonding MOs have energies similar to those of the contributing AOs of
the more electronegative atom. The antibonding MOs have energies close to those
of the contributing AOs of the less electronegative atom.
2

Energies of atomic orbitals


In dealing with heteronuclear molecules, it is important to know the energies of the
atomic orbitals of the constituent atoms.

Hydrogen fluoride
As an illustration of the general points, lets consider the example of HF. There is only
a very weak 1s - 2s orbital mixing because of the very large difference in energies of
the AOs. The H-F bond length is ca. 0.92 (in the gas phase).
Shapes of MOs in HF:
H

-13 ev
1s (H) orbital can
mix with 2pz (F)

H atom

HF molecule

F atom

strongly
antibonding

2 px,y

nonbonding

-18 ev

px, y

p
bonding

-47 ev

1
4

Carbon monoxide
C and O possess 10 valence electrons. The oxygen 2s and 2p orbitals are lower in
energy than the carbon orbitals. B.O. = (8 - 2)/2 = 3. Note: 3 has a significant nonbonding character (even slight anti-bonding character); this orbital resembles a lone
pair. The electrons of 3 can participate a dative bond.
Shapes of MOs
Photoelectron
spectrum of CO
3

-14

CO
*

2 px,y

LUMO

O
3

HOMO

-16

2 pz

2 px,y

-18

2 pz

2
-20
eV

2s

2s

2
1
5

Isoelectronic molecules and ions: CO, CN-, NO+


CO, CN-, and NO+ all have 10 valence electrons. They are said to be isoelectronic.
The same ordering in the MO diagrams is expected for isoelectronic molecules; the
only differences being in the absolute values of the orbital energies.

Isoelectronic molecules and ions: CO, CN-, NO+


The bond order in CO, CN-, and NO+ molecules and ions is (8 - 2)/2 = 3. The bonds
are all very short: CO (1.13 ), CN- (1.14 ), NO+ (1.06 ). The bond strength is
not known for CN- and NO+. The triple bond in CO (256 kcal/mol) is stronger than
the triple bond in N2 (226 kcal/mol).
The molecular orbitals that will be of greatest importance for reactions between
molecules are the HOMO and LUMO, collectively known as the frontier orbitals since
they lie at the occupied-unoccupied frontier. The HOMO of CO, CN-, and NO+ is 3

LUMO (*)

HOMO (3)

The reactivity of CO, CN-, NO+


The HOMO of these molecules has a larger lobe on the less electronegative atom.
Predictably, the carbon atom in CO and CN-, and the nitrogen atom in NO+, can act
as 2e- donors.
Examples:
H
O

H
O

B
H

HH

boron has an
empty p orbital

H+ has an
empty s orbital

Hydrogen
cyanide, HCN

Metal carbonyl complexes


In 1888 Ludwig Mond observed during experiments with carbon monoxide that an
unexpected volatile substance was formed when CO had been in contact with nickel:
O
C

Ni

4 CO

O C

C
O

Ni

colorless liquid,
b.p. 34 oC
O

O
C

Ni

C
O Td

Ni(CO)4 decomposes on heating and


is used in the manufacture of high purity Ni.
There is a very large number of carbonyl
complexes of different transition metals.
The toxicity of CO is due to its strong binding
O
to the iron center of an oxygen-carrier protein
in blood. For similar reasons, the CN- ion is
also very toxic.
9

Metal carbonyl complexes


When CO binds to a transition metal it is called a ligand. It forms a bond with the
metal by donation of two electrons into an empty metal orbital. That is the CO ligand
is a donor. The LUMOs of CO are the * orbitals. These can overlap with filled
metal d orbitals. The resulting bonding leads to delocalization (donation) of electron
density from the filled d orbitals into the empty * orbitals on the CO, which acts as a
acceptor.
-bonding

MC bonding makes the C-O


bond stronger and shorter.

-bonding

MC bonding makes the C-O


bond weaker and longer.
10

CH225.17
The group theoretical approach to bonding
Character tables
Representations of point groups
Examples of applications: H2O

The group theoretical approach to bonding


It is not difficult to decide which atomic orbitals mix to form molecular orbitals in a
diatomic molecule. However, in a polyatomic molecule, even in the simplest cases
such as H2O or NH3, it is considerably more difficult to tell which atomic orbitals of
the constituent atoms have the right symmetry and therefore mix, and which do not
have the right symmetry and cannot mix.
Describing a molecule's symmetry in terms of symmetry operations conforms with
the requirements of a mathematical group. This allows us to apply mathematical
techniques of group theory to describe and analyze some of the molecule's
chemically significant properties.
In applying the group theory to chemical problems we will use vector representations
for molecular properties. For illustration, we are going to use the C2v group,
comprised of the four symmetry operations: E, C2, v, v'

Representations of point groups


Lets examine the effect of the four operations, E, C2, v, and v' on unit vectors x, y,
z. The standard orientation of the Cartesian coordinate system is chosen so that the
C2 axis coincides with the z axis, and the v plane coincides with the xz plane.

vectors

Operation
C2
v

[1] x

[-1] x [1] x

[1] y

[-1] y [-1] y [1] y

[1] z

[1] z

[1] z

v'

[-1] x
characters

[1] z
3

Representations of point groups


In addition to linear vectors x, y, z we should consider curved vectors representing
rotations about the three Cartesian axes: Rx, Ry, Rz.
For example:

All characters of the C2v group can be listed in a tabular form. We are using the
nomenclature proposed by R. S. Mulliken in 1933 (Nobel Prize in chemistry).
4classesofsymmetryoperationsinC2v

"Representations"
of the group C 2v

C 2v

C2

v'

h=4
z
Rz
x, Ry
y, Rx

Irreducible representations
The C2v group has four representations (named A1, A2, B1, B2), each composed of
four characters. These are also called irreducible representations of the group.
Character table for C2v

Irreducible representations are also referred to as symmetry species. We can say


that vector z transforms as A1 in C2v, or z belongs to the A1 species of C2v.
We will deal with non-degenerate representations A (symmetric with respect to the
principal axis Cn) and B (anti-symmetric with respect to Cn), doubly degenerate
representations E (the name should not be confused with identity E), and triply
degenerated representations T. The degeneracy is determined by the character
under E: 1 = signly, 2 = doubly, and 3 = triply degenerate.

Irreducible representations
The symbols A, B, E, and T can be modified by a subscript or/and a superscript.
1. Subscripts 1 or 2 are attached to A and B to indicate symmetry or anti-symmetry
with respect to a non-principal C2 rotation or to v, when the C2 is absent.
2. In centrosymmetric groups, a subscript g (gerade = even) indicates symmetry
and a subscript u (ungerade = uneven) indicates anti-symmetry with respect to
inversion, i.
3. Addition of a prime (') or double prime (") indicates symmetry or anti-symmetry
with respect to h.

There is a website containing more information about the symmetry point groups and
associated character tables:
http://www.phys.ncl.ac.uk/staff/njpg/symmetry/Stereographs.html

Reducible representations
Molecular properties are not always located along the axes of a Cartesian coordinate
system. Often they are oriented in general directions. An example of this is a general
vector v which has its tip at a general point with coordinates (x, y, z). It can be shown
that the transformations of v under the operations of C2v are expressed by a
reducible representation, r = {3, -1, 1, 1}.

C2
z
v'

y
x
v

Any reducible representation , r, is a combination of irreducible representations .

Decomposition of reducible representations


To decompose a reducible representation , r, we need to determine, how many
times, n(), an irreducible representations contributes to r. For groups of finite
order this number is calculated as follows:
n() =
For example, in the C2v group there are four classes of operations: E, C2, v, v,
each composed of 1 operation and the total number of the operations is the group
order h = 4. Using the example of r = {3, -1, 1, 1} we find that
n(A1) = 1/4 [(1)(1)(3) +(1)(1)(-1) + (1)(1)(1) + (1)(1)(1)] = 1
n(A2) = 1/4 [(1)(1)(3) +(1)(1)(-1) + (1)(-1)(1) + (1)(-1)(1)] = 0
n(B1) = 1/4 [(1)(1)(3) +(1)(-1)(-1) + (1)(1)(1) + (1)(-1)(1)] = 1
n(B2) = 1/4 [(1)(1)(3) +(1)(-1)(-1) + (1)(-1)(1) + (1)(1)(1)] = 1

Then, r = A1 + B1 + B2
8

The group theoretical approach


1. Find the point group (for linear molecules, substitute D2h for Dh or C2v for Cv)
2. Define a coordinate system with the origin located at the center of the molecule.
The z axis is usually collinear with the principal rotational axis. The xz plane is
chosen to contain as many atoms as possible. For a tetrahedral molecule the x, y,
and z axes are collinear with the three C2 axes.
3. Construct a reducible representation r for the orbitals of the outer atoms. In every
class, the character in r is equal to the number of orbitals unaffected by the
symmetry operations of the class, minus the number of unshifted orbitals which
show their signs reversed. = 0 for any shifted (moved) orbital.
4. Decompose r into its component species, irreducible representations. This is
equivalent to finding group orbitals, also called symmetry-adapted linear
combinations (SALCs) of the atomic orbitals.
5. Identify all orbitals of the central atom belonging to the same symmetry species.
6. The orbitals of the central atom and the SALCs with the same symmetry and
similar energy can be combined to form molecular orbitals. Several computer
packages are available that can further calculate the relative energies and
generate the shapes of the MOs.
9

Example of H2O
Example 1. H2O, water.
1. The point group of H2O is C2v
2. The C2 axis is chosen as the z axis. The xz plane contains all atoms of H2O.
3. The outer atoms, two hydrogens, possess 1s orbitals. These are involved in
bonding and are represented by HO vectors.
Lets consider transformations (symmetry properties) of the hydrogen orbitals under
the operations of C2v
E leaves both hydrogen orbitals (bond vectors) unchanged, = 2
C2 moves both orbitals, = 0
v leaves both orbitals unchanged, = 2
v' moves both orbitals , = 0
The reducible representation r:
C2v
r

C2
v
v'

z y

H
O

C2

v'

10

Example of H2O
4. r can be decomposed into the irreducible representations:

The molecule of H2O has two symmetry-adapted linear combinations (called group
orbitals) of the two 1s orbitals of the hydrogen atoms: A1 = (Ha) + (Hb) and B1 =
(Ha) - (Hb)
C2
C2

v
A1

v
B1

11

Example of H2O
5. Symmetries of the orbitals of the central atoms (e.g. oxygen in H2O) are know for
different groups and can be determined from the table:

6. Three oxygen orbitals (s, pz and px) belong to the A1 and B1 species of C2v. That is
they have symmetry matching that of the group orbitals of the two hydrogen atoms.
12

Molecular orbitals of H2O


2b1

2b1

2H

3a1

3a1
B1 1s
2pz 2py 2px

A 1 B2 B1

A1
1b2

A1
2py

B1

1b2

2pz

2a1

A1

2a1

2px

B1

1b1
2s

1b1

A1
1a1

2s

A1

1a1
13

Summary for H2O


1. The combination of A1 + 2s + 2pz gives three MOs: 1a1 (which is mostly the 2s
orbital), 2a1 (slightly bonding, mostly 2pz), and 3a1 (antibonding).
2. The combination of B1 + 2px gives two MOs: 1b1 (bonding) and 2b1(antibonding).
3. The oxygen 2py orbital belongs to the B2 species in C2v, and is a nonbonding
orbital.
4. The molecular orbital picture differs from the common conception of the water
molecule having two identical lone pairs and two identical bond pairs. In the MO
picture, the HOMO (1b2) is truly nonbonding, occupying the 2py orbital, which is
perpendicular to the plane of the molecule. The next two pairs (2a1 and 1b1) are
bonding pairs, resulting from the overlap of the 2pz and 2px with the 1s orbitals of
the hydrogens. All four occupied molecular orbitals are different in the molecule of
H2O!

14

CH225.18
Molecular orbitals in NH3 and CH4 molecules.
For next time, read Lecture 19

Example 2. NH3, ammonia.

Example of NH3

1. The point group of NH3 is C3v which contains three classes of symmetry
operations: E, C3, and v. The following is the character table for C3v.

C3
z
H
H

h=6

y
H

x
v

2. The C3 axis is chosen as the z axis. The xz plane contains two atoms: N and H.
3. The outer atoms, 3 H, possess three 1s orbitals. These are involved in
bonding and are represented by HN vectors.
E leaves 3 vectors (hydrogen orbitals) unchanged, = 3
C3 shifts all vectors, = 0
v leaves one vector unchanged, = 1
Reducible representation r:
2

Hydrogen group orbitals


4. r can be decomposed into the irreducible representations:
3v
E
2C3
n(A1) = 1/6 [113 +210 + 311] = 1
n(A2) = 1/6 [113 +210 + 3(-1)1] = 0
n(E) = 1/6 [123 +2(-1)0 + 301] = 1

The symmetry-adapted linear combinations (LCAOs) of the three 1s orbitals: A1 =


(Ha) + (Hb) + (Hc), and E = 2(Ha) - (Hb) - (Hc) and (Hb) - (Hc). Note: Two
group orbitals transform as E in C3v. Despite the different appearance, they have the
same energy. LCAOs are generated in group theory using the projection operator
approach. Their pictorial representations are shown below:

A1

E
3

Symmetry of 2s and 2p AOs in C3v


5. The nitrogen orbitals 2s and 2pz belong to the A1 species, while 2px and 2py
belong to the doubly degenerate species E in C3v.

6. The orbitals which have matching symmetry can mix to produce molecular orbitals.

The MO diagram for NH3


2e

2e
3a1

3H
E

2px,y 2pz

A1

2a1

3a1

A1

2a1

1e
2s

2e

1e

A1

1e

1a1

1a1

Summary for NH3


1. The nitrogen s and pz orbitals combine with the hydrogen A1 group orbital to give
three a1 orbitals. The nitrogen px and py orbitals combine with the E group orbitals
of the hydrogen atoms to form four molecular orbitals e : two bonding and two
antibonding.
2. The nitrogen 2s orbital has a low energy and its interaction with the hydrogen A1
orbital is quite weak. Therefore, the 1a1 orbital has nearly the same energy as the
nitrogen 2s orbital. The HOMO of NH3 (2a1) is the lone pair of the Lewis and
VSEPR models. This is a slightly bonding molecular orbital. Some texts describe
it as essentially non-bonding.

Example of a tetrahedral molecule, CH4


Example 3. CH4
1. The point group of CH4 is Td, which contains five classes of symmetry operations:
E, C3, C2, S4 and d. The following is the character table for Td.

h = 24

Transformations of CH4 in Td
2. For a tetrahedral molecule, the three C2 axes are chosen as the x, y, and z axes
of the Cartesian system.
3. The outer atoms, 4 H, possess four 1s orbitals. These are involved in bonding
and are represented by four vectors.
E leaves 4 vectors (hydrogen orbitals) unaffected, = 4
C3 does not move 1 vector, = 1

C3

C2 moves all vectors, = 0

d H

S4 moves all vectors, = 0

C2 (S4)

d does not move two vectors, = 2

C
r

8C3

3C2

3S4 6d

y
H

Transformations of CH4 in Td
Effect of the symmetry operations of Td on the four-vector basis of the representation r:

Hydrogen group orbitals


4.

r can be reduced to the irreducible representations A1 + T2.


n(A1) = 1/24 [114 + 811 + 310 + 610 + 612 ] = 1
n(A2) = 1/24 [114 + 811 + 310 + 6(-1)0 + 6(-1)2] = 0
n(E) = 1/24 [124 +8(-1)1 + 320 + 600 + 602] = 0
n(T1) = 1/24 [134 +801 + 3(-1)0 + 610 + 6(-1)2] = 0
n(T2) = 1/24 [134 +801 + 3(-1)0 + 6(-1)0 + 612] = 1

r = A1 + T2
5. The symmetry-adapted linear combinations of the four 1s orbitals have the
following composition:
A1 = (Ha) + (Hb) + (Hc) + (Hd)
T2 = (Ha) - (Hb) - (Hc) + (Hd)
T2 = (Ha) - (Hb) + (Hc) - (Hd)
T2 = (Ha) + (Hb) - (Hc) - (Hd)
10

Symmetry of 2s and 2p AOs in Td


Group orbitals
(LCAOs)

A1

T2

6. One atomic orbital of carbon (2s) belongs to the A1 species, while 2px, 2py and
2pz belong to the triply degenerate T2 species in Td.

s
A1

pz

py

px

T2

11

The MO diagram for CH4

2t2

4H
2t2

px,y,z

2a1

T2
s

A1

1t2

A1 + T2

2a1

1t2

1a1

1a1
12

Summary for CH4


1. The 2s orbital combines with the hydrogen A1 group orbital to give two a1
molecular orbitals (bonding and antibonding). The 2px, 2py, and 2pz orbitals
combine with the T2 group orbitals to form six t2 MOs: three bonding and three
antibonding.
2. This model contrasts noticeably with what might be expected from either Lewis or
VSEPR models. Instead of four equal-energy electron pairs confined to four
identical bonds, we see three pairs of electrons at one energy level and a single
pair at a lower level. This is consistent with the observed photoelectron spectrum
for CH4 that shows two bands corresponding to ejection of electrons from MOs of
two different energies (1a1 and 1t2).
PE spectrum of CH4

15

20

25 (eV)

13

CH225.19
Molecular orbitals of BF3

Example of BF3
Example 4. BF3, boron trifluoride.
1. The point group of BF3 is D3h, h = 12.

C3 S3
z
y
B

C2

2. The C3 axis is chosen as the z axis. The xz plane contains two atoms: B and F
3. The outer atoms, 3 F, possess 2px,y,z orbitals. These are represented by 9
vectors. We can neglect the 2s orbitals of fluorine which have a very low energy
and are not significantly involved in B-F bonding.

Constructing a reducible representation for BF3


E leaves 9 vectors (orbitals) unshifted, = 9
C3 affects all vectors, = 0
C2 leaves 1 orbital unchanged and 2 reversed, = -1
h leaves 6 orbitals unchanged and 3 reversed, = 3
S3 affects all vectors, = 0
v leaves two orbitals unaffected and one reversed, = 1

2C 3 3C 2

2S 3 3

4.

r can be reduced to the irreducible representations: A'1 + A'2 + 2E' + A"2 + E"

Group orbitals of BF3


The symmetry adapted combinations of the 2p orbitals in a D3h symmetric molecule
are shown on this slide:

Symmetry of 2s and 2p AOs in D3h


5. The boron 2s and 2pz orbitals belong to the A'1 and A"2 species, respectively. The
2px and 2py orbitals of boron belong to the doubly degenerate species E'.

6. The orbitals which have matching symmetry can mix to produce molecular orbitals.

The MO diagram for BF3


3e'
2a'1

2p

2a"2

3F

A''2 + E'

2s

non-bonding
a'2 e"

A1

2e'

2px,y, z
15eA1 + A2 + A"2 + E" + 2E

1a"2
1e'
1a'1

2s
6

Non-bonding MOs

Upper occupied MOs in BF3


(orbital energies are approximate,
1 Hartree = 627.5 kcal/mol)

e" (-0.478)

e" (-0.478)

a2 (-0.481)

Bonding MOs

2e (-0.48)

2e (-0.48)

1a"2 (-0.61)

1e (-0.64)

1e (-0.64)
1a1 (-0.68)

-bonding in BF3
The molecule of BF3 has three bonding orbitals (1a'1 and
1e') and one bonding orbital (1a"2). Three resonance
structures can be drawn for BF3 which demonstrate that the
B-F bonds has some double bond character; total number of
bonds = (32 24)/2 = 4

SO3, NO3-, and CO32- are isoelectronic with BF3 (24 valence
electrons). These molecules and ions are isostructural and
all have three bonds (1a'1 and 1e') and one bonding
orbital (1a"2).

1a"2

Summary for BF3


1. The 2s orbital of boron combines with the fluorine A'1 group orbital to give two a'1
MOs (bonding 1a1 and antibonding 2a1). The boron 2px and 2py orbitals
combine with the two E' group orbitals to form six e' MOs: two bonding, two
nonbonding, and two antibonding. The boron 2pz orbital combines with the A"2
group orbital to give a bonding and an antibonding molecular orbital.
2. The LUMO of BF3 is 2a2 which main contributor is the 2pzorbital of boron. The
LUMO of BF3 can accept electrons in a dative bond formation. In other words,
BF3 is a Lewis acid.

LUMO: 2a"2

CH225.20
Molecular orbitals of CO2

Example of CO2
Example 5. CO2, carbon dioxide.
1. The point group of CO2 is Dh but the simpler D2h can be used instead.
h=8

2. The molecular axis O=C=O is chosen as the z axis.


3. The outer atoms, 2 O, possess 2px,y,z orbitals. These are represented by 6
vectors. To a first approximation, we can neglect the 2s orbitals of oxygen which
have low energy and are not significantly involved in the C-O bonding.
2

Constructing a reducible representation for CO2


E leaves 6 vectors (orbitals) unshifted, = 6
C2(z) leaves 2 orbitals unchanged and 4 reversed = -2
C2(y) moves all vectors, = 0
C2(x) moves all vectors, = 0
i moves all vectors, = 0
(xy) moves all vectors, = 0
(xz) leaves 4 orbitals unchanged and 2 reversed, = 2
(yz) leaves 4 orbitals unchanged and 2 reversed, = 2

E
r
4.

C2(z) C2(y) C2(x)


-2

i
0

(xy) (xz) (yz)


0

r can be reduced to the irreducible representations:


n(Ag) = 1/8 [(1)(1)(6) +(1)(1)(-2) + (1)(1)(0) + (1)(1)(0) + (1)(1)(0) + (1)(1)(0) + (1)(1)(2) + (1)(1)(2) ] = 1
n(B1g) = 1/8 [(1)(1)(6) +(1)(1)(-2) + (1)(-1)(0) + (1)(-1)(0) + (1)(1)(0) + (1)(1)(0) + (1)(-1)(2) + (1)(-1)(2) ] = 0
n(B2g) = 1/8 [(1)(1)(6) +(1)(-1)(-2) + (1)(1)(0) + (1)(-1)(0) + (1)(1)(0) + (1)(-1)(0) + (1)(1)(2) + (1)(-1)(2) ] = 1
n(B3g) = 1/8 [(1)(1)(6) +(1)(-1)(-2) + (1)(-1)(0) + (1)(1)(0) + (1)(1)(0) + (1)(-1)(0) + (1)(-1)(2) + (1)(1)(2) ] = 1
n(Au) = 1/8 [(1)(1)(6) +(1)(1)(-2) + (1)(1)(0) + (1)(1)(0) + (1)(-1)(0) + (1)(-1)(0) + (1)(-1)(2) + (1)(-1)(2) ] = 0
n(B1u) = 1/8 [(1)(1)(6) +(1)(1)(-2) + (1)(-1)(0) + (1)(-1)(0) + (1)(-1)(0) + (1)(-1)(0) + (1)(1)(2) + (1)(1)(2) ] = 1
n(B2u) = 1/8 [(1)(1)(6) +(1)(-1)(-2) + (1)(1)(0) + (1)(-1)(0) + (1)(-1)(0) + (1)(1)(0) + (1)(-1)(2) + (1)(1)(2) ] = 1
n(B3u) = 1/8 [(1)(1)(6) +(1)(-1)(-2) + (1)(-1)(0) + (1)(1)(0) + (1)(-1)(0) + (1)(1)(0) + (1)(1)(2) + (1)(-1)(2) ] = 1

Group orbitals of CO2


r = Ag + B2g + B3g + B1u + B2u + B3u
The symmetry adapted combinations of the 2p orbitals in a D2h symmetric molecule
are shown on this slide (red dots stand for O nuclei and the black ones are for C):
x
px +/- px
z

3u

2g

2u

3g

1u

y
py +/- py

pz +/- pz

Symmetry of 2s and 2p AOs in D2h


5. The carbon 2s and 2pz, 2px and 2py orbitals belong to the Ag and B1u, B3u, and B2u
species, respectively.

6. The orbitals of matching symmetry can mix to produce molecular orbitals.

Oxygen group orbitals: Ag + B2g + B3g + B1u + B2u + B3u


non-bonding

The MO diagram for CO2


2b1u

2ag

2b2u 2b3u

B1u B2u B3u

b2g b3g

Ag B2g B3g B1u B2u B3u


1b2u1b3u

Ag

non-bonding

1b1u

1ag

Bonding MOs

Non-bonding MOs

Upper occupied MOs in CO2

b2g

b3g

1b2u

1b3u

1b1u

1ag

Summary for CO2


1. The 16 valence electrons in CO2 occupy, from the bottom, two essentially nonbonding orbitals (we have ignored these), two bonding orbitals (1ag and 1b1u),
and two bonding orbitals (1b2u and 1b3u). There are also two occupied nonbonding MOs (b2g and b3g). In other words, there are two and two bonds in
the molecule, as expected.
2. The molecular orbitals of the isoelectronic linear triatomic molecules and ions
such as N3, CS2, NCO, and CNO ( all have 16 valence electrons) are very
similar.

N
azide

C
carbon
disulfide

S O

C
cyanate

N O

fulminate

CH225.21
Molecular orbitals in a hypervalent compound: SF6
For next time: read Lecture 22

Example of SF6
Example 6. SF6, sulfur hexafluoride.
1. The point group of SF6 is Oh

zF
F

F y
Fx

F
F

2. The Cartesian system is easy to define: its origin is at the S atom and the three
axes coincide with the S-F bonds.
3. For simplicity, we can ignore bonding in SF6. The outer atoms, 6 F, each use
only one 2p orbital in bonds to sulfur, represented by 6 FS vectors. We can
completely ignore the 2s orbitals of fluorine because of their very low energy.
2

Constructing a reducible representation for SF6


C 4, C 2, S 4
d

C2

F
S

C3 affects all vectors, = 0


C2 affects all vectors, = 0
C4 does not move two orbitals, = 2

F
h F

zF
F

E leaves 6 vectors (orbitals) unshifted, = 6

C 3, S 6

C2

C2=C42 does not move two orbitals, = 2


i affects all vectors, = 0
S4 affects all vectors, = 0

F y

S6 affects all vectors, = 0

Fx

h leaves 4 orbitals unchanged, = 4


d leaves 2 orbitals unchanged, = 2

F
F

4.

8C3

6C2

6C4 3C2

6S4 8S6 3h 6d

r can be reduced to the irreducible representations: A1g + Eg + T1u


3

Group orbitals of SF6


r = A1g + Eg + T1u
Thus we have six group orbitals with three different symmetries, which can form
bonding and antibonding combinations with like symmetry AOs on sulfur.
For simplicity, only the lobe directed to sulfur is shown for each 2p orbital of the F
atoms:
or
Group orbitals of 6 F atoms in SF6:

A1g

Eg

T1u

Symmetry of 2s and 2p AOs of sulfur


5. The orbitals of sulfur belong to the A1g and T1u species in Oh:

3s, A1g

3pz, T1u

3py, T1u

3px, T1u

6. The orbitals of matching symmetry can mix to produce molecular orbitals.

Fluorine group orbitals: A1g + Eg + T1u

non-bonding

A partial MO diagram for SF6

T1u

A1g + Eg + T1u
A1g

Bonding MOs

Non-bonding MOs

Important occupied MOs in SF6

eg

t1u

eg

t1u

a1g

t1u

Summary for SF6


A satisfactory qualitative MO description of bonding in hypervalent molecules can
be developed using valence s and p orbitals on the central atom.
Considering bonding in SF6 involving 6 2p orbitals of fluorine atoms and four
orbitals of sulfur, we find that they form four bonding, two nonbonding, and four
antibonding orbitals. 12 valence electrons of SF6 occupy 4 bonding and 2
nonbonding MOs. The antibonding orbitals are all empty.
Sulfur hexafluoride is chemically unreactive. It is both the most inert sulfur
compound and the most inert covalent fluoride. As a result of its inertness and
high stability, SF6 finds application as an insulating gas in high-voltage electrical
systems. Because the velocity of sound is low in SF6, it is also used to fill the
space between the glass sheets in double-glazing units found in airport
concourses.

CH225.22
Part 4. Lewis acids and bases
Important types of Lewis acids
Boron group acids
For next time, read Lecture 23

Lewis concept of acidity and basicity


Definition: A Lewis base is a 2e- donor and a Lewis acid is a 2e- acceptor. We can
denote a Lewis acid as A and a Lewis base as :B. The fundamental reaction of Lewis
acids and bases is the formation of complexes (or adducts), A-B, where A and :B
share the electron pair supplied by the base.
Example 1: A proton, H+ is a Lewis acid because it can attach to an electron pair of a
base.

Example 2: A main-group atom with an incomplete octet in a molecule can complete


its octet by accepting an electron pair.

trimethylborane

trimethylborane-ammonia complex
2

MO description of Lewis acid-base reactions


MO description of Lewis acid-base reactions involves the frontier molecular orbitals.
For Lewis acid-base complex formation, the base should have 2e in the HOMO of
suitable symmetry to interact with the LUMO of the acid. This can be illustrated by
reaction of H+ with NH3.

Antibonding MO

Unoccupied 1s orbital

Bonding MO

HOMO of NH3

Important types of Lewis acids


A useful reagent in chemical synthesis is solution of borane in diethyl ether where
complex formation takes place:

The LUMO of BH3 is the 2pz orbital of boron. This empty orbital is responsible for the
Lewis acidity of the molecule.

Important types of Lewis acids


Example 3: An atom with a complete octet can rearrange its valence electrons and
accept an additional electron pair.
O
C
O

OH
O

OH

The LUMO of CO2 is concentrated on carbon. This empty orbital is responsible for
the Lewis acidity of the molecule.

Important types of Lewis acids


Example 4: A particularly subtle example of Lewis acidity is one when an atom in a
molecule or ion is able to expand its valence shell and accept an electron pair to
form a hypervalent compound.

C-N 1.32

+
2

Si-F 1.68

SiF62 has been structurally characterized by


X-ray diffraction. The crystal structures of
diguanidinium and bis(methylammonium)
C-N 1.47
hexafluorosilicate are shown here.
Question: why [C(NH2)3]+ is planar and has short C-N bonds?

Important types of Lewis acids


A satisfactory MO description of bonding in SiF62 can be developed using the 3s and
3p orbitals of Si interacting with a single 2p orbital on each of the six fluorine atoms:

Si

SiF4

3p

4F
2p

+2

:F

SiF62-

Si
3p

6F
2p

eg
3s

t2
a1

Tetrahedral (Td)

3s

t1u
a1g

Octahedral (Oh)

Two electron pairs in SiF62 are accommodated in nonbonding MOs. The electronic
structure of SiF62 is analogous to that of SF6 (see Lecture 21).
7

Important types of Lewis acids


Example 5: A transition metal ion or compound can accept an electron pair supplied
by a Lewis base to form a coordination compound. Coordination compounds will be
treated in the last part of this course.

The crystal structure of the


trinitrophenolate salt of the diammine silver
complex:

Ag-N 2.11

Trinitrophenolate ion
8

Boron group acids


Boron and Aluminum have incomplete octets in the planar molecules BX3 and AlX3.
The central atom pz orbital is either vacant (when X is not a -donor) or involved in
weak -bonding (e.g. when X = halide).
CH3

H3C

CH3

B(CH3)3

BF3

In either case boron can accept 2e- from a Lewis base, for example:

boron trifluoride
CH3
B

tetrafluoroborate ion
H

H3C CH3

trimethylborane

H3C
B

C
H

f rom LiCH3

H3C

CH3

CH3

tetramethylborate ion

Relative Lewis acidity


For Lewis acids such as BX3 (X = F, Cl, Br, I) the stability constants of the
compounds which are formed with Lewis bases have been measured and the trend
have been established which reflects the better -donor ability of the smaller halides.

stronger

Lewis acids

weaker

Although BF3 is the least reactive among the boron trihalides, it is still strongly Lewis
acidic. For example, BF3 (gas, b. p. = 100 C) dissolves in diethyl ether (b. p. = 35
C) to give a product liquid F3BOEt2/Et2O that boils at 125 C and is widely used as
an industrial catalyst.
The role of BF3 in organic chemistry is to abstract a halide bound to carbon and
generate a reactive carbocation:

10

CH225.23
Boron group acids
Carbon group acids
Nitrogen group acids

Formation of dimers
Under normal conditions, some of the boron and aluminum Lewis acids do not exist
in the monomeric form but produce dimers or even polymers. For example, AlCl3 is a
dimer Al2Cl6 in the gas phase. The electron unsaturation of one metal center is
relieved by 2e donation from a chloride of the other AlCl3 unit in the dimer:
2.21

2.06

D2h
The molecule of Al2Br6 has been
structurally characterized by X-ray
diffraction. The structure is shown on
this slide.

Dimers of borane and trimethylaluminum


Even more interesting are boron and aluminum dimers with hydrogen or carbon
atoms in the bridging positions:

Diborane

Trimethylaluminum dimer

D2h

C2h

1.19

1.94

1.32

The structure of B2H6 (X-ray)

2.14
The structure of B2Me6 (neutron diffraction)
3

Electronic structure of B2H6


In diborane, the octet electronic configuration is
achieved by forming the so-called three-center
two-electron bonds (3c, 2e bonds):
MO description of 3c, 2e bonding in diborane:
Group orbitals
of Boron:

energy

2px

2px

Group orbitals
Molecular orbitals: of Hydrogen:
Antibonding MOs

H
b3u

2pz

D2h

b3u

Bonding MOs

-2pz
ag

ag
4

Electronic structure of Al2Me6


For Al2Me6, MOs can be constructed using the lone pairs of the CH3 groups. Note:
this lone pair is the a1 molecular orbital (HOMO) of the CH3 anion (note that CH3
and NH3 are isoelectronic and isostructural, and their MO descriptions are
qualitatively the same).

Only the bonding MOs are occupied in B2H6 and Al2Me6, and this results in formation
of reasonably stable B-H-B and B-C-B bonds.

Carbon group acids


Two Lewis acidic compounds have been already mentioned in the preceding
discussion: CO2 and SiF4. Germanium tetrahalides GeX4 are Lewis acids like SiF4.
The preference for +2 over +4 oxidation state increases down Group IV (the inert
pair effect). Tin (II) dichloride is a Lewis acid. It has the same number of electrons as
BCl3 (6 e), and the two molecules have qualitatively similar electronic structure:
2

Tin dichloride, SnCl2


Cl

Sn

Cl

tin

Cl
Sn
Cl

97

total:

C2v

trigonal
6

Carbon group acids


SnCl2 combines with H2O to form the hydrate SnCl22H2O, which is commercially
available and is used as a reducing agent. SnCl2 accepts Cl to give trigonal
pyramidal SnCl3

Sn
Cl

Cl-

Cl
Cl

Cl

Sn
Cl

C3v
Crystal structure of [HNMe3]SnCl3

SnCl3 is isoelectronic and isostructural with NR3, i.e. it is formally a Lewis base.
Sn(IV) halides are Lewis acids and their strength follows the order SnF4 > SnCl4 >
SnBr4 > SnI4
Oh
2KCl + SnCl4

K2[SnCl6]
Potassium
hexachlorostannate
Crystal structure of [HNMe3]SnCl6

Nitrogen group acids


The heavy element antimony, Sb, forms an important strong Lewis acid SbF5, which
can be used to produce some of the strongest known Brnsted acids by reaction
with hydrogen fluoride:

Brnsted acid solutions more acidic that H2SO4 are called superacids. A superacid
can protonate almost any organic compound. Another superacid can be prepared by
reacting SbF5 with fluorosulfonic acid, FSO3H.

Lewis acid

Lewis base
Brnsted acid

Lewis acid&base
complex

Brnsted acid
8

CH225.24
Oxygen group acids
Halogen acids

Electronic structure of SO2


SO2 is an angular molecule with and bonding between the sulfur and oxygen
atoms. The central atom possesses a lone pair, i.e. SO2 can function as a Lewis
base. SO2 has a low-lying LUMO and can also react as a Lewis acid.
Lewis base

1.43

Lewis acid

O-S-O 119

C2v
LUMO can accept electrons

LUMO

HOMO can donate electrons

HOMO
2

Lewis acidity of SO2


The Lewis acidity of sulfur dioxide is illustrated by formation of complexes with the
Lewis bases, trimethylamine and F ion:

OSF 103

OSO 107

Crystal structure of [Me4N]SO2F

OSN 99

Cs

OSO 114

Cs

Crystal structure of Me3N-SO2


3

Lewis basicity of SO2


SO2 can donate an electron pair in complex formation with transition metals. The
crystal structures of two octahedral 18-e- complexes of Cr(0) and Ru(II):
OSO = 115

Cr(CO)5(SO2)

[Ru(NH3)4(TFA)(SO2)]+

Electron count:
6+52+2 = 18 e-

Electron count:
8+42+1+2-1 = 18 e4

Lewis acidity of SO3


Sulfur trioxide, SO3 (isoelectronic with AlF3), is a strong Lewis acid.
LUMO in SO3 (*) can
accept electrons

D3h

The crystal structures of O3Spyridine and O3SNH2CH3 complexes:

Cs

Cs

Sulfuric acid from SO3


In the two-stage industrial preparation of sulfuric acid, SO3 is first dissolved in
H2SO4 to give oleum:

The crystal structure of the [S2O7]2 ion:


C2
The resulting H2S2O7 is diluted with water and is hydrolyzed:

Halogen acids
The molecules of bromine, Br2, and iodine, I2, are mild Lewis acids. The well known
reaction between I2 (acid) and I (base) gives the triiodide ion, I3
2

The crystal structure


of [H2NMe2]I3

Dh
In this reaction an electron pair is donated by I to the LUMO of I2 (antibonding *
MO).
I

antibonding * MO of I2
7

Halogen acids
Examples of crystal structures of complexes of I2 and Br2 acting as Lewis acids:

Me3As-I2

Me2Se-I2

C3

Cs
Me2S-Br2
Ph3P-Br2

[Me4P]Br3
8

CH225.25
Lewis base and acid strength
Hard and soft acids and bases
For next time, read Lecture 26

Lewis base and acid strength


Acids and bases can be weak or strong. When two Lewis bases are combined with
an acid, the stronger base will bind more strongly and form a more stable product.
Periodic trend: Lewis basicity generally decreases down a group.
Example:

Me3E + BMe3 Me3E-BMe3

Base

G = H - TS

Thermodynamics of reaction with BMe3

trimethylamine

Me3N

trimethylphosphine
trimethylarsane

Me3P
Me3As

-16 kcal/mol
favorable at -80 oC

trimethylstibane

Me3Sb

practically no product formed

-17.6 kcal/mol

The group trend is probably the only reliable predictive tool. The influence of other
factors, such as inductive, steric, and solvent effects is often difficult to predict.
2

Inductive effect
Substitution of electronegative atoms, such as fluorine or chlorine, in place of
hydrogen results in weaker bases. For example, PF3 is a much weaker base than
PCl3 and PH3. This is explained by the inductive effect, that is by reduced electron
density on the donor atom.

Proton affinity: 160 kcal/mol


Atomic charge on P: +1.93

Proton affinity: n/a


Atomic charge on P: +1.54

Proton affinity: 187 kcal/mol


Atomic charge on P: +0.31

Stronger base
There are exceptions to this rule. For example,
P(OMe)3 (trimethylphosphite) is more basic than
PH3, although phosphorus is certainly more +ve in
P(OMe)3 than in PH3.

Proton affinity: 218 kcal/mol


Atomic charge on P: +1.64
3

Inductive, solvent and steric effects


Gas phase proton (H+) affinity of simple amines follows the order:
Me3N > Me2NH > MeNH2 > NH3
as predicted on the basis of electron donation (induction) by the methyl groups and
resulting increased electron density on the nitrogen.
However, the relative basicity is different toward BMe3:
Me2NH > MeNH2 > Me3N > NH3
This is due sterics. Repulsion between the methyl groups is greater than repulsion
between methyl groups and hydrogen atoms. The larger molecule of triethylamine,
Et3N, does not form a stable product with BMe3.
The basicity is greatly affected by solvation. For example, the proton affinity of
methylamines in water follows the order: Me2NH > MeNH2 > Me3N > NH3.
For phosphines, gas phase proton affinity is in the order:
PPh3 > PMePh2 > PMe2Ph > PMe3
The order is completely reversed in organic solvents:
PMe3 > PMe2Ph > PMePh2 > PPh3

Solubilities of metal halides


In addition to the intrinsic strength, Lewis acids and bases have other properties,
called chemical softness and hardness, that determine the outcome of their
reactions. For example, silver halides have a range of solubilities in water.
H2O
Ag+(aq) + X-(aq),

AgX(s)

205

Cl

1.8 x 10-10

Br

5.2 x 10-13

8.3 x 10-17

K = [X-][[Ag+]
[AgX]

more soluble

The trend is due to systematic differences in the strength of bonding between the
halides and silver ion. Ag+ is strongly bonded to I but is more weakly bonded to F.
The trend of solubilities is reversed for lithium halides, where LiF shows the strongest
bonding:

Hard and soft acids and bases


Reactions on the previous slide illustrate two general trends.
[A] Chemists established that late transition metals and the heavier post-transition
metals in low oxidation states (Pd2+, Pt2+, Cu+, Ag+, Au+, Cd2+, Hg2+) form more stable
complexes with the donor atoms P, S, and I.
[B] Electropositive main-group elements, early transition metals in relatively high
oxidation states (Ti4+, Fe3+, Co3+, Al3+) form more stable complexes with N, O donors
and F.
R. G. Pearson named class A cases soft acids/bases and class B cases hard
acids/bases.
A hard acid prefers to combine with a hard base. Hard acid-base interactions are
often primarily electrostatic. A soft acid prefers to combine with a soft base. Soft acidbase bonds are more covalent.

The Periodic Table

Late transition and posttransition metals


Early transition metals

Hard acids/bases
Relatively small, electropositive and highly oxidized species are hard acids.
Examples:

Be2+, Al3+, H+, Mn3+, Sc3+, Cr3+, Fe3+, Ti4+


Hard bases have strongly electronegative donor atoms (N, O, F)

Examples:

R2O, ROH, NR3, OH-, F-, Cl-, SO42-, CH3CO2-, SCN-

Soft acids/bases
The larger and less electropositive metals form soft Lewis acids. The acids become
softer as the oxidation state is reduced.
Examples:

Ga3+, Cu+, Ag+, Au+, Hg+/Hg2+, Pd2+, Pt2+/Pt4+, Rh+, Ir+


Soft bases have less electronegative non-metal donor atoms (C, S, P, As, I).
Examples:

R2S, PR3, AsR3, RNC, C2H4, CO, CN-, I-, SCNBorderline acids and bases:

Fe2+, Co2+, Ni2+, Cu2+

Br-, py

Examples of acid-base reactivity


Oxygen and sulfur are hard and soft bases, respectively, in the crown ether ligands
C12H24O6 (O6) and C12H24O4S2 (O4S2). The following equilibrium constants (log K)
for K+ and Ag+ illustrate the relative stability of their complexes:

O
O

18-crown-6 (O6)

O
M

O
O

O
O

[M(18-crown-6)]+

Ag+ forms a more stable complex with the sulfur containing ligand O4S2, whereas
K+ forms a stronger complex with O6.
10

CH225.26
Part 5. d-Metal Complexes
Ligands
Nomenclature
Electron counting
For next time, read Lecture 27

d-Metals and Transition metals


A transition metal is "an element whose atom has an incomplete d subshell, or which
can give rise to cations with an incomplete d subshell."

3d
4d
5d

d-Metal complexes
Complexes or coordination compounds are composed of a metal atom or metal ion
and ligands, the latter can be thought of as donating electrons to the metal. A ligand
can be an ion (Cl), a molecule (NH3), or a group of atoms. In other words, a complex
is a combination of a Lewis acid (the metal) with Lewis bases (the ligands).

-NH2

Trichloro-p-tolylamino-ptolylimino-triphenylphosphinerhenium(V)

cis-Dichloro-(dihydrogensulfide)((2-(N,N-dimethylamino)phenyl)
diphenylphosphine)-(triphenylphosphine)ruthenium(II)

Coordinate bonds L:M are covalent bonds; only the formal electron counting
distinguishes them.
3

Important ligands
Monodentate ligands occupy one coordination site in metal complexes, i.e. they have
one point of attachment to metal.
Monodentate ligands:

Hydride: H- (hydrido)
CN (cyano)

CO (carbonyl)
NH3 (ammine)

(pyridine, py) NO+ (nitrosyl) NO2- (nitro)

PR3 (phosphine): e.g. PMe3, PPh3


O2- (oxo)

OH- (hydroxo)

SCN- (thiocyano)

OH2 (aqua)

CNS- (isothiocyano)

Halides: F- (fluoro), Cl- (chloro), Br- (bromo), I- (iodo)


Ligands are frequently named with their older trivial names rather than with the
IUPAC names.

Important ligands
Polydentate ligands have two or more points of attachment to metal. Most common
are bidentate (which are called chelating ligands) and tridentate ligands.
Bidentate ligands:

H2N

NH2

Ph2P

PPh2

Important ligands
Tridentate ligands:

Nomenclature
Complexes are named with the ligands in the alphabetical order, followed by the
name of the metal with its oxidation number in parentheses. The suffix -ate is added
if the complex is an anion. The systematic names, if complicated, are rarely used.
The Greek prefixes are used to tell the number of each type of ligand in a complex:
mono-, di- (bis-), tri- (tris-), tetra- (tetrakis-), penta-, hexaNi(CO)4 tetracarbonylnickel(0)
RuCl2(PPh3)3 di-chloro-tris-(triphenylphosphine)-ruthenium(II)
[Co(NH3)6]3+ hexaamminecobalt(III)
[CoCl2(NH3)4]+ tetra-ammine-di-chloro-cobalt(III)
[CoCl2(en)2]+ di-chloro-bis-(ethylenediamine)-cobalt(III)
[Fe(bipy)3]2+ tris-(bipyridine)-iron(II)
[Fe(CN)6]4- hexacyanoferrate(II)
[Fe(OH2)6]2+ hexaaquairon(II)
[PtCl4]2- tetrachloroplatinate(II)

Nomenclature
The prefixes cis- and trans- designate adjacent and opposite coordination sites:

Chemotherapy drug Cisplatin

The trans isomer does not have the


pharmacological effect of Cisplatin

Nomenclature
Octahedral complexes with at least three identical ligands exist as meridianal (mer-)
and facial (fac-) geometrical isomers. In fac- isomers, the three identical ligands
occupy one face of the octahedron.

R3P
R3P

PR3
Ir

Cl

Cl

Cl

L
fac-tri-chloro-tris-(dimethylphenylphosphine)iridium(III)
fac-IrCl3(PPhMe2)3

R3P
Cl

PR3
Ir
PR3

Cl
Cl

mer-tri-chloro-tris-(dimethylphenylphosphine)iridium(III)
mer-IrCl3(PPhMe2)3

Nomenclature
Bridging ligands between two metal atoms have the prefix -

CMe MeC

MeC
N
MeC

N
O

Mo
N

MeC

N
Mo

2+

CMe

CMe

N
CMe

CMe MeC

-oxo-bis(penta(acetonitrile)molibdenum(II))

H3N

H3N
H3N

NH3
Cr

NH3

NH3
NH3
Cr
NH3
NH3

NH3

-oxo-bis(pentaamminechromium(I))

10

Electron counting
Transition metal complexes typically have 18 electrons around the metal. The
following chart shows numbers of electrons donated by common ligands and how
they affect the metal oxidation state.
Metal O.N. change

M H

hydride, 1e-

+1

M Cl

halide, 1e-

+1

M Cl M

-halide, 3e/2M

+1/2 per M

M O-R

alkoxide, 1e-

+1

M NR3

amine, 2e-

no change

M PR3

phosphine, 2e-

no change

M CO

carbonyl, 2e-

no change

M CN

cyanide, 1e-

+1
11

CH225.27
Coordination numbers 2, 3, 4, 5, and 6

Coordination numbers
The number of coordination sites occupied in a complex by the ligands is called the
coordination number. Common coordination numbers of d-metal complexes range
from four (CN = 4) to eight (CN = 8). Other coordination numbers are known.
The following factors determine the coordination number of a complex.
1. The number of valence electrons on the metal
2. The atomic radius of the metal
3. The size of the ligands
High coordination numbers, CN > 6, are most common for the relatively large 4d and
5d transition metals (i.e., from Periods 5 and 6) with 4 to 7 valence electrons (d4 to d7
electron configurations)
Low coordination numbers are found on the very right of the transition series, where
the metals have a relatively small number of empty orbitals in the valence shell: in
nickel, copper, and zinc groups.

Coordination number 2
In the following examples, the electronic configuration of gold and mercury is d10,
which is calculated as group # - O.N. (note that all valence metal electrons in
complexes are treated as d).

NC Hg

Electron count:
11(Ag) + 22 (:NH3) 1 = 14 e
This is a d10 14e complex

CN

Electron count:
12(Hg) + 21 (CN) = 14 e
This is a d10 14e complex
3

Coordination number 3
Coordination number 3 is rare; it is more likely with d10 metal centers. Threecoordinate trigonal-planar complexes of Au(I) are known:
Ph3P

Au

PPh3

PPh3

tris(triphenylphosphine)gold(I)

The low coordination number is, in part, due to ligand crowding:

Electron count:
11(Au) + 32 (:PPh3) 1 = 16 e
This is a d10 16e complex

Electron count:
11(Au) + 22 (:PPh3) + 1(Cl) = 16 e
This is a d10 16e complex
4

Coordination number 4
Four-coordination is found in a large number of compounds. Tetrahedral geometry is
favored when the metal is small and/or the ligands are large. Halide complexes of
the late 3d metals Fe, Co, and Ni are often tetrahedral. Those of the 4d and 5d
metals Rh, Pd, Ir, and Pt are typically square-planar.

tetrachloroferrate(II) tetrachlocobaltate(II)

10(Ni) + 42 (:CO) = 18 e
This is a d10 18e complex

11(Cu) + 42 (:py) 1 = 18 e
This is a d10 18e complex

9(Co) + 41 (Cl) + 2 = 15 e
This is a d7 15e complex
5

Coordination number 4
Square-planar complexes are common with the larger 4d and 5d d8-metal centers :

10(Pd) + 41(Cl) + 2 = 16 e
This is a d8 16e complex

10(Pt) + 42(NH3) - 2 = 16 e
This is a d8 16e complex

Coordination number 5
The geometries possible for coordination number 5 are the trigonal bipyramid and
the square pyramid. Coordination compounds can switch easily from one to the other
shape because the energy difference between the two is small for d-metal
complexes.

Lax
Leq

M
Lax

D3h

Leq
Leq

Lax
Lb

Lb
Lb

Lb

C 4v

Coordination number 5
Two rhenium complexes on this slide are examples of trigonal-bipyramidal and
square-pyramidal coordination geometries.

N
Cl
Cl

Re

Cl
Cl

tetrachloro-nitrido-rhenate(VI)

7(Re) + 31(Cl) + 22(NBut) = 14 e


This is a d0 14e complex

7(Re) + 41(Cl) + 3(N) + 1 = 15 e


This is a d1 15e complex
8

Coordination number 6
Six is the most common coordination number. Six-coordinate complexes are usually
octahedral.
H
OC

Ir

O
CO

Cl

Cl

Mo

Cl

Cl

Cl

Cl

OH2

dicarbonyl-f ac-triiodohydrido-iridate(III)

aqua-tetrachlorooxo-molibdate(V)

9(Ir) + 31(I) + 22(CO) +


1(H) + 1 = 18 e
This is a d6 18e complex

6(Mo) + 41(Cl) + 2(OH2) +


2(O) + 1 = 15 e
This is a d1 15e complex

OEt2
Cl
Zr
Cl
OEt2

tetrachloro-bis(diethylether)zirconium(IV)

4(Zr) + 41(Cl) + 22(OEt2) = 12 e


This is a d0 12e complex
9

CH225.28
Bonding in square-planar complexes

Bonding in square-planar complexes


Square-planar complexes, ML4, form an important group of coordination compounds.
To understand bonding in square-planar complexes we are going to develop a
qualitative molecular orbital diagram and examine the molecular orbitals. We shall
consider a simple compound that has only metal-ligand -bonds, [PdH4]2.

z C4, C2, S4
C2"
x

C2' y

D4h

Tetrahydridopalladate(II)
16-e, d8-complex
2

Character table for D4h

z C4, C2, S4

1. The point group of [PdH4]2 is D4h, which contains ten classes of


symmetry operations. The following is the character table for D4h.
2. The C2 and C4 axes are chosen as the x, y, and z axes.

D4h

E 2C4

C2

2C2

2C2

2S4

A1g

A2g

-1

-1

-1

-1

B1g
B2g
Eg

1
1
2

-1
-1
0

1
1
-2

1
-1
0

-1
1
0

1
1
2

-1
-1
0

1
1
-2

1
-1
0

-1
1
0

A1u

-1

-1

-1

-1

-1

A2u

-1

-1

-1

-1

-1

B1u

-1

-1

-1

-1

-1

B2u

-1

-1

-1

-1

-1

Eu

-2

-2

C2"
C2' y

2v 2d h = 16

Pd orbitals

x2+y2, z2

s, dz2

x2-y2
xy
(yz, zx)

dx2-y2

dxy
dxz, dyz

pz

(x, y)

px,py
3

Group orbitals of [PdH4]


3. The outer atoms, 4 H, possess four 1s orbitals. These are involved in bonding
and are represented by four vectors.
E leaves 4 vectors (hydrogen orbitals) unaffected, = 4
C4 moves all vectors, = 0
C2 moves all vectors, = 0

z C4, C2, S4

C2 does not move two vectors, = 2


C2 moves all vectors, = 0

C2"

i moves all vectors, = 0

C2' y

S4 moves all vectors, = 0


h leaves all vectors unaffected, = 4
v does not move two vectors, = 2
d moves all vectors, = 0
D4h E 2C4
r

C2

2C2

2C2

2S4

2v 2d
2

0
4

Group orbitals of [PdH4]2


4.

r can be reduced to the irreducible representations A1g + B1g + Eu.

n(A1g) = (1/16)[114 + 210 + 110 + 212 + 210 + 110 + 210 + 114 + 212 + 210] = 1
n(A2g) = (1/16)[114 + 210 + 110 + 2(-1)2 + 2(-1)0 + 110 + 210 + 114 + 2(-1)2 + 2(-1)0] = 0
n(B1g) = (1/16)[114 + 2(-1)0 + 110 + 212 + 2(-1)0 + 110 + 2(-1)0 + 114 + 212 + 2(-1)0] = 1
n(B2g) = (1/16)[114 + 2(-1)0 + 110 + 2(-1)2 + 210 + 110 + 2(-1)0 + 114 + 2(-1)2 + 210] = 0
n(Eg) = (1/16)[124 + 200 + 1(-2)0 + 202 + 200 + 120 + 200 + 1(-2)4 + 202 + 200] = 0
n(A1u) = (1/16)[114 + 210 + 110 + 212 + 210 + 1(-1)0 + 2(-1)0 + 1(-1)4 + 2(-1)2 + 2(-1)0] = 0
n(A2u) = (1/16)[114 + 210 + 110 + 2(-1)2 + 2(-1)0 + 1(-1)0 + 2(-1)0 + 1(-1)4 + 212 + 210] = 0
n(B1u) = (1/16)[114 + 2(-1)0 + 110 + 212 + 2(-1)0 + 1(-1)0 + 210 + 1(-1)4 + 2(-1)2 + 210] = 0
n(B2u) = (1/16)[114 + 2(-1)0 + 110 + 2(-1)2 + 210 + 1(-1)0 + 210 + 1(-1)4 + 212 + 2(-1)0] = 0
n(Eu) = (1/16)[124 + 200 + 1(-2)0 + 202 + 200 + 1(-2)0 + 200 + 124 + 202 + 200] = 1

r = A1g + B1g + Eu
5

Group orbitals of [PdH4]2Four 1s orbitals of the four hydride ligands form four group orbitals in [PdH4]2 :

The group orbitals have the right symmetry


to mix with the following metal orbitals: 5s
(A1g), 5px,y (Eu), 4dx2-y2 (B1g), and 4dz2 (B1g)
to form nine MOs in [PdH4]2.
Metal pz, dxy, dyz, dxz orbitals form four
nonbonding MOs in [PdH4]2, because there
are no matching symmetry hydrogen
orbitals in the complex. This is a
consequence of the orientation of these
metal orbitals relative to the ligands.
6

MO diagram for a square-planar complex


z
LUMO
nonbonding MOs

a2u (pz)
HOMO

b2g (dxy)

a1g (dz2)

eg (dyz, dxz)

-bonding MOs

b1g
eu

eu
a1g

Summary for square-planar complexes


1. There are four -bonding MOs in [PdH4]2, corresponding to four single bonds in
the complex.
2. Square-planar complexes have four occupied non-bonding orbitals which are the
metal dxy, dxz, dyz, and dz2. Therefore, square-planar complexes have 16 electrons
and d8 electron count.
3. The LUMO of a square-planar complex is the metal pz orbital. Despite being
unsaturated, 16-electron square-planar complexes are reasonably stable. There
are two reasons for this: (a) the LUMO has a relatively high energy and (b) the
LUMO and HOMO have the same orientation, therefore addition of electrons to
the LUMO is relatively unfavorable.
4. The occupied non-bonding orbitals dxy, dxz, dyz can be involved in -bonding in
square-planar complexes containing -acceptor ligands such as CO, CN, NO+

2e
dxz or dyz

LUMO (CO)

Reactivity of square-planar complexes


16-electron square-planar complexes oxidatively add molecules such as H2, HCl,
CH3I. The best example of this reactivity is shown by the Vaskas complex, transchlorocarbonylbis(triphenylphosphine)iridium(I): trans-IrCl(CO)(PPh3)2 (L. Vaska and J. W.
DiLuzio J. Am. Chem. Soc. 1961, 83, 2784).

Vaskas complex

Bonding in square-pyramidal complexes

Bonding in square-pyramidal complexes


Square-pyramidal complexes, ML5, form an important group of coordination
compounds. To understand bonding in square-planar complexes we are going to
develop a qualitative molecular orbital diagram and examine the molecular orbitals.
We shall consider a simple compound that has only metal-ligand -bonds, like
[RhH5]2.

C4, C2
H
H

H
Rh

v
H

C4v
Pentahydridorhodate(III)
16-e, d6-complex

Character table for C4v


1. The point group of [RhH5]2 is C4v, which contains 5 classes of
symmetry operations. The following is the character table for C4v.

C4v

E 2C4

C2

2v 2d

A1

A2

-1

-1

B1
B2
E

1
1
2

-1
-1
0

1
1
-2

1
-1
0

-1
1
0

h=8
Metal orbitals
s, pz, dz2
dx2-y2

dxy
(px, py), (dxz, dyz)

2. The Rh-H bonds are chosen as the x, y, and z axes.

Group orbitals of [RhH5]2


3. The outer atoms, 5 H, possess four 1s orbitals. These are involved in bonding
and are represented by four vectors.
E leaves 5 vectors (hydrogen orbitals) unaffected, = 5
C4 does not move 1 vector, = 1
C2 does not move 1 vector, = 1

v does not move 3 vectors, = 3


d does not move 1 vectors, = 1
C4v E 2C4
r

C2
1

2v 2d
3

C4, C2

H
Rh

v
H
H

d
y

C4v

r = 2A1 + B1 + E

Group orbitals of [RhH5]2Four 1s orbitals of the four hydride ligands form four group orbitals in [RhH5]2 :

A1

A1

B1

The group orbitals mix with the following


metal orbitals: 5s + 5pz + 4dz2 (A1), 5px,y
(Eu), 4dx2-y2 (B1) to form the total of eleven
bonding and antibonding MOs.
Metal 4dxy, 4dyz, 4dxz orbitals form three
nonbonding MOs. Note: the dyz, dxz orbitals
become weakly bonding in complexes
where the metal is removed from the basal
plane.

E
C4v

Metal orbitals

A1

s, pz, dz2

A2
B1
B2
E

dx2-y2

dxy
(px, py), (dxz, dyz)

MO diagram for a square-pyramidal complex (16-e)


5a1
Rh

3e
A1

5p

antibonding MOs

4a1

5s

5H

A1

nonbonding MOs

2b1

1s

3a1
A1

4d

B1
B2
E

1b2

2e

B1
E
A1

1e
1b1

-bonding MOs
2a1

1a1
6

Summary for square-pyramidal complexes


1. There are five -bonding MOs corresponding to five single bonds.
2. Square-pyramidal complexes have three non-bonding d orbitals: dxy (1b2), dxz,
dyz (2e).
3. Immediatelly above the non-bonding d orbitals, there are two anti-bonding MOs:
3a1 and 2b1.
4. The d orbitals and 3a1 + 2b1 orbitals of square-pyramidal compexes correspond
to the t2g and eg orbitals, respectively, of octahedral complexes. Therefore, highspin configurations can be expected for all ML5 complexes which ML6 analogues
are high-spin. The relatively low energy of 3a1 should make high-spin
configurations very likely for d6 square-pyramidal complexes. d8 configurations
can be diamagnetic when low-spin.
5. Occupied non-bonding orbitals dxy, dxz, dyz can be involved in -bonding in
square-planar complexes containing -acceptor ligands such as CO, CN, NO+

CH225.29
Bonding in octahedral complexes

Bonding in octahedral complexes


Octahedral complexes, ML6, represent one of the most important types of
coordination compounds. To understand bonding in octahedral complexes we are
going to develop a qualitative molecular orbital diagram and examine the molecular
orbitals. We shall first consider ML6 compounds that have only metal-ligand -bonds,
such as [Co(NH3)6]3+ or [FeH6]4 complexes.

Hexamminecobalt(III)

Hexahydridoferrate(II)

18-e, d6-complex

18-e, d6-complex
2

MO diagram for [FeH6]4

Orbitals of an octahedral complex, [FeH6]4Group orbitals of the hydrogen atoms in [FeH6]4 are qualitatively the same as the
groups orbitals of the main-group octahedral compound, SF6, which we have
considered before (in CH225.21).

d-Metal complexes have s, p and d orbitals in the valence shell. The s and p orbitals
in [FeH6]4 have the same symmetries as the s and p orbitals of sulfur in SF6:
4s (A1g)

4px,y,z (T1u)

The d orbitals in Oh have the following two symmetries:


3dz2 and 3dx2y2 (Eg)

3dxy, 3dyz, and 3dxz (T2g)

The T2g orbitals dxy, dyz, dxz have no matching symmetry hydrogen orbitals and
remain nonbonding. This is a consequence of the orientation of these orbitals relative
to the ligands.
4

Ligand and metal orbitals in [FeH6]4Fe

Fe

A1g

T1u
Eg

Occupied -bonding MOs in octahedral complexes

t1u

t1u

t1u

eg

eg

a1g

Important non-bonding and antibonding MOs

eg (*)

eg (*)

z
3dyz

3dxz

3dxy

t2g
x

t2g

t2g
7

MO diagram for [FeH6]4


The following is a qualitative MO
diagram applicable for any
octahedral complex with only
interactions between the metal
and the ligands.
Note that 12 electrons fill six
bonding MOs. The other 6
electrons of the 18-e- [FeH6]4
complex occupy the nonbonding
t2g set. The energy gap between
the t2g and eg orbitals is called the
ligand-field splitting parameter,
o.
This energy gap can be very
small in some 3d metal
complexes.

Fe
[FeH6]4-

p
1u

eg

1g

6H

d
g

2g

t2g

1u
g
1g

t1u
eg
a1g
8

High and low spin octahedral complexes


For the configurations d1 (13 valence e), d2 (14 valence e), and d3 (15 valence e),
the electronic configurations are unambiguous (the quantum number S = (# of
unpaired e-)/2)

The splitting energy o depends on the ligands L in ML6. Therefore, for configurations
d4 through d7 both high spin and low spin configurations are possible. For example:

CH225.30
High and low-spin octahedral complexes
-Bonding in metal complexes

High- vs. low-spin configurations


Two electrons in different orbitals experience less repulsion than when they are
paired. In the d4, d5, d6, and d7 cases, the number of unpaired electrons (and the
total spin S) is determined by the magnitudes of o and the electron-pairing energy
P.

When o < P, the electrons are spread across the t2g and eg levels, resulting in a
high-spin configuration.
When o > P, all electrons are in the t2g orbitals, resulting in a low-spin configuration.
2

High- vs. low-spin configurations

P increases in the order 5d < 4d < 3d, as the d orbitals become smaller.

o shows the opposite trend: 5d > 4d > 3d (also o for M(o.n) < o for M(o.n.+1)).
Consequently, o > P in the ML6 complexes of 4d and 5d metals, which are
typically low spin compounds. Many of them have18 electrons which are all
paired (d6 or t2g6 configurations).

3d
4d
5d

A relatively small number of metals may have O < P in their complexes.


They are typically 3d metals: V, Cr, Mn, Fe, Co, Ni, and Cu.

High- vs. low-spin configurations of 3d metals


Even for 3d metals, O is strongly dependent on the nature of the coordinated ligands.
The orbital splitting O decreases for common ligands according to the spectrochemical
series (experimentally established) :
NO+,CO, CN > NO2 > en > NH3 > NCS > H2O > F, RCOO > OH > Cl > Br > I
-acceptors
Large O
Low-spin

-donors only

-donors
Small O
High-spin

Examples of low-spin 3d metal complexes: [Mn(CN)6]3, [Fe(CN)6]3, [Co(CN)6]3

Mn(III), t2g4

16 e

Fe(III), t2g5

17 e

Co(III), t2g6

18 e

High-spin 3d metal complexes


High-spin complexes are common for ligands that are low in the series (appear on
the right hand side) in combination with 3d metal ions. Examples: [Mn(H2O)6]2+,
[FeF6]3, [Fe(C2O4)3]3,
2+

OH2
H2O
H2O

Mn
OH2

OH2
OH2

HOOC-COOH
Oxalic acid

t2g3eg2

17-e, d5 Mn(II)

17-e, d5 Fe(III)

17-e, d5 Fe(III)
5

-Bonding in d-metal complexes


So far we have considered only ML interactions. However, many common ligands
in metal complexes are either -donors or -acceptors. They have atomic or
molecular orbitals of -symmetry with respect to the LM bond axis.

In an octahedral complex with -donor/acceptor ligands, the t2g orbitals (dxy, dyz, and
dxz) are involved in -bonding with the ligands and are no longer non-bonding.

HOMOs in [FeH6]4 and [Fe(CO)H5]3


-bonding
MCO

[Fe(CO)H5]3

3dyz

3dxz

3dxy

3dyz

3dxz

3dxy

[FeH6]4

t2g
x

t2g

t2g
7

-Acceptor ligands
The effect of -acceptors on the orbital splitting parameter, o

t2g
eg
o (no -bonding)
o ( -bonding)

d zx d yz d xy

t2g
Ligands that act as -acceptors increase the orbital splitting o. These ligands are
found on the left-hand side in the spectrochemical series and give low-spin
complexes.
8

-Donor ligands
We shall now consider the effect of -bonding on the orbital splitting parameter, o.
The effect of -donation:

eg
o ( -bonding)

o (no -bonding)

d xy
d yz
d zx

t2g

t2g
Ligands that act as -donors decrease the orbital splitting o. These ligands are
found on the right-hand side in the spectrochemical series and often give high-spin
complexes with 3d metals.

You might also like