Você está na página 1de 470

The Theory of the Chemical Bond: A Textbook for

Chemistry 2160
Stephen Lee
Department of Chemistry and Chemical Biology
Cornell University, Ithaca, NY 14853-1301
September 4, 2009
Chapter 1
What this book is all about
I write this book with Chemistry 2160 students, second semester rst-year honors chemistry
students, in mind. Every year around 100 of the most motivated freshman chemistry students
here at Cornell take this course (we have some 1500-2000 freshman chemistry students in
total). The students in this class are self-selected but they are among the brightest students
we have. Usually more than three-quarters of the students in the class have received a 5 on
the AP exam. When I ask the students of the class what their future plans are, typically
a third of the students say they plan on becoming engineers, a third doctors, and a third
scientists. They are an exciting group of students to teach: they represent the future of
technology in the United States.
This course generally receives bi-modal ratings from the students in their course eval-
uations: some students view the course favorably (5s and 4s) and others as an unpleasant
experience (1s and 2s). Many of the students who give favorable ratings to this course
become chemistry majors.
If you are a rst year student taking Chemistry 2160, here are some thoughts about what
will happen in this course. Things are going to be dierent. While students often feel that
Chem 2150 was an extension of AP high school chemistry, the same will not be the case this
1
semester. This course, instead of looking back to your past experiences in chemistry, looks
ahead to your future experiences. Chemistry is a big subject: three of its main elds are
organic (second year chemistry), physical (third year), and inorganic (fourth year).
Students who take these upper level chemistry classes can nd the material covered in
organic, physical and inorganic chemistry strangely disconnected. They are disconnected to
the point that it may seem that organic and physical chemists belong in disparate worlds.
The Cornell chemistry faculty would not be doing a good job of preparing you for our
discipline, if we did not try to help you unite these elds in your own minds. You, as the
student, will need to carry the responsibility here at synthesizing your knowledge, but it is
my hope that this course will provide you with some of the essential tools to do so.
If you are not planning on taking more chemistry courses do not worry. Many, in fact
most, of the students who have come back to talk to me about this course, turn out not to
be chemists at all. For the rst time, in the last year, three former students, who are now
in nance, have come by to talk about this class. All three told me the same thing: they
nd themselves thinking about the contents in this course. As one of them said to me, It
sticks. I like to think that this course has made these students happier and maybe even
more successful in their nancial careers. Whatever your nal eld may be, I hope it will
serve the same role for you.
1.1 What you will learn if you study this book
This book describes, in the plainest language I know, my understanding of the chemical
bond. The chemical bond lies at the heart of all chemistry. If you understand the chemical
bond, you will know about one theme which unites organic, physical and inorganic chemistry.
While the chemical bond is a universal and timeless truth, the human understanding of the
chemical bond changes with each generation of scientists. Linus Pauling wrote a famous
2
book (based on lectures given here at Cornell) fty years ago, The Nature of the Chemical
Bond. Much of the content of this course is not covered in this landmark book.
To give you some understanding of our uxional understanding of the chemical bond,
when I took freshman chemistry, I was taught there were three types of strong chemical
bonds: the covalent, the ionic and the metallic. In this book we will discover that the metallic
bond and the covalent bond can be productively treated in a single unied approach. So for
this course, we will need discuss only two types of chemical bonds: covalent and ionic bonds.
In this book you will learn the quantum mechanical basis of the covalent bond and the
electrostatic nature of the ionic bond. Of the two, the covalent bond is much more dicult
to understand. To understand the covalent bond you will need to learn about the electron,
quantum mechanics, the virial theorem, and molecular orbital theory. For the ionic bond,
we can rely on a classical sense of the material world. Due to the necessary ground-work
required for the former subject, the bulk of this course will be devoted to the covalent bond.
Only in the last 10-15% of this course will we consider ionic forces.
None of the ideas in this book were originally created by my research group. I can claim
no original credit for anything in this book. Roald Homann, Kenichi Fukui, David Pettifor,
and Klaus R udenberg gure among the many chemists and material scientists whose key
ideas are incorporated in this book. I do not propose, nor am I capable of, giving an accurate
scientic bibliography for each of the subjects covered in this book. Instead at the end of
each chapter a few texts or articles are listed which could help enrich your understanding of
the given material. Such references will be kept at a minimum, so as to not overwhelm the
student. Problem sets and old exams will also be given.
3
1.2 This book is not a textbook
As a student in this class, I hope to take you out of your comfort zone into places where
you will not be familiar. If we (the teaching sta and the students together) are successful,
everyone in this class will see unfold a rich and challenging view about how the material
universe around us behaves and how this material universe is based on the chemical bond.
While this book is hopefully a guidebook, it is not a textbook.
Textbooks tend to have numerous editions and are based on a collective knowledge of the
eld. They contain the promise within them, that if you just follow the textbook, you will
gain mastery of the subject. My own personal experience was that the textbook approach,
rarely worked. Followers do not become masters and teaching-learning is at its core a one-on-
one activity. When I was a student, every now and then I would nd a book where I had the
sensation the author was trying to explain to me a full subject, as they understood it to be.
The books unfolded like stories. The story could have been thermodynamics (Enrico Fermis
book with this same title), or regular polytopes (Coexters book). These books worked for
me. I didnt do very well in rst year physics, but I certainly got every exam question on
thermodynamics correct.
While this book will not come close to the clarity of the classics, I hope that its intent
can be the same. I would like you the student to get wrapped up in a story, the story of the
chemical bond. Follow the story as I tell it, but challenge the story as well. Before every
lecture read ahead the chapter or chapters in the book which the lecture will cover. Find
out which ideas make sense to you and which do not. Try to gure out the chain of logical
thought, which the book and the lectures are trying to express.
A site will be placed in the Cornell course web site, Blackboard, for you to post your
questions. To encourage you to post thought provoking questions, the best questions will be
awarded 1 or 2 extra credit points. Chemical errors in the book noted in Blackboard before
4
the relevant lecture will also receive 1 point. (The course total is 500 points). I will try to
answer the Blackboard questions in class, but if you nd the classroom lecture still does not
make sense, speak up. Please feel free to ask questions in class.
After the lecture, review what it is you heard, and what it is you understand. Start doing
the problems relevant to the material. Think of how this material relates to other material
in the course. If you can, make up your own problems.
Exams in this course are my questions to you. In the best of all possible worlds, exams
are learning experiences. I will try to design exams which test how well you see how the
course material is put together: how the course material can be used in dierent ways. On
every exam, there will be at least one question which will place two separate concepts of the
course into a single explicit problem. I will know we, the teaching sta of this course, have
succeeded as teachers, if you as a student can start thinking of such questions yourself.
5
Chapter 2
The electron
To understand the covalent bond, one must understand the electron. We therefore begin
our course by stating facts about electrons, some, but not all, of which you will have seen in
high school AP chemistry.
2.1 Orbitals
Electrons in atoms lie in orbitals. The word orbital was originally chosen for its similarity
to the word orbit. Just as planets move around the sun, electrons move around the nucleus,
see Figure 2.1. The planetary analogy for electron motion has many virtues. The electron
has a small mass compared to the nucleus and lies in the electrostatic potential well of this
nucleus. In exactly the same way a planet has a small mass when compared to the mass of
the sun, and the planet sits in the gravitational potential well of the sun. Even the relation
describing electronic and planetary potential energies are the similar.
U
m
E
m
S
r
6
Figure 2.1: Planetary vs. electronic motion
where U, m
E
, m
S
, and r are respectively the gravitational potential energy, the mass of the
earth, the mass of the sun, and the distance between the earth and the sun. Analogously,
the potential energy of an electron in the electrostatic well of a nucleus obeys the relation:
U
q

q
+
r
where q

, q
+
, and r are respectively the charge of the electron, the charge of the nucleus
and the distance between the electron and the nucleus.
The analogy between orbits and orbitals bears one main strength and one main weak-
ness. First, the strength. In both planetary orbits and electron orbitals, the concepts of
potential and kinetic energy can be exactly dened. Potential energy is the U value given
in the equations above. Kinetic energy is
p
2
2m
, where p is the momentum and m the mass.
(Classically, the momentum is dened to be p = mv, where v is the velocity.)
The Virial Theorem: On a per kilogram basis, the further away a planet is from the sun,
the less gravitational potential energy it has. Per kilogram, Mercury has the most negative
solar gravitational potential energy, while Pluto, the farthest planet from the sun, has the
7
least negative. But there is another compelling trend to planetary energies. In Table 2.1 we
list average planetary velocities. The farther a planet is from the sun, the slower it moves.
Pluto, 100 times futher from the sun than Mercury, travels at one tenth Mercurys speed.
Pluto therefore not only has a less negative potential energy than Mercury, but less kinetic
energy as well. Planetary motion obeys the Virial Theorem, a theorem which says that there
is a xed relation between potential and kinetic energy. For motion like planetary motion,
the virial theorem states:
2T = U
where T and U are respectively the kinetic and potential energies.
Table 2.1: Planetary Orbital Properties.
Planet Mean Distance from Sun (km) Mean Velocity (km/s)
Mercury 5.8 10
7
48
Venus 1.1 10
8
35
Earth 1.5 10
8
30
Mars 2.3 10
8
24
Jupiter 7.8 10
8
13
Saturn 1.4 10
9
10
Uranus 2.8 10
9
6.8
Neptune 4.5 10
9
5.4
Pluto 5.9 10
9
4.7
Electrons in stable orbitals obey the virial theorem: Electrons in stable orbitals,
just like planets in stable orbits, obey the virial theorem. The closer an electron lies to the
nucleus, the more negative is its potential energy. According to the virial theorem, the more
negative the potential energy, the faster the electron moves. In Au, for example, there are
electrons which lie very close to the nucleus. The electron orbital closest to the nucleus is
called the 1s orbital. As the Au 1s orbital lies close to the nucleus, electrons in this orbital
move fast: Au 1s orbital electrons move at 0.6 times the speed of light. In the hydrogen
atom by contrast, where the 1s orbital is much farther from the nucleus, electrons move
8
much slower, at only
1
137
times the speed of light. This latter number is so useful, it has a a
name, the Fine Structure Constant. It plays an important role in physics.
Electrons do not lie in orbits: The orbit-to-orbital analogy has one major breakdown.
Electrons are best described with quantum mechanics while planets are best described by
classical mechanics. Rather than thinking of planetary orbits, where the planets maintain a
more or less xed distance from the sun, electrons in the course of their motion vary their
distances from the nucleus.
Figure 2.2: y trajectory vs. y probability distribution
9
Comparison of electrons to ies: A qualitative picture might be useful here. A good
picture for an electron circling a nucleus might be a housey circling a plate of food. The
y (the electron) in its lowest energy state ies close to the plate of food (the nucleus) and
sometimes alights on the plate itself. In describing the ys position it is useful to envision a
probability function. This function gives the relative probability that the electron is located
at a given place, see Figure 2.2. As the Figure shows, the most probable place the y would
be located is directly on top of the plate of food. As the distance from the plate of food
increases, the probability decreases. This distribution is more or less the least energetic
prole of a non-sleeping y. We call the lowest possible energy probability distribution the
ground state distribution. For atoms which have just one or two electrons the ground state
distribution for the electron is always the 1s distribution. For historical reasons, the 1s
distribution is typically called
1s
, where is the Greek letter rho.
Excited state distributions: We can also envision excited state distributions. As the
name implies the excited state is a distribution of higher energy than the ground state
distribution. Both ies and electrons can enter excited states. Once one disturbs a y, for
instance by trying to kill it, it ies in a much more erratic and energetic manner. It ies much
farther from the plate of food and typically no longer lands on the food plate. Electrons,
like ies, also have excited states.
In the 2p orbital, one of the most important higher energy states of an electron, the
electron lies further from the nucleus than an electron in the 1s orbital. Just like in the y
excited state, if the electron enters the 2p orbital, the electron is never found on the center of
the distribution. (One major dierence between an excited y motion and an excited electron
motion is that while the y always ies faster in its excited state, the electron ies slower.)
The 2p distribution is called
2p
. Both the 1s and 2p distributions are stable distributions:
electrons can, under the right circumstances, stay in these distributions for long periods of
10
Figure 2.3: Ground state and excited state y vs. 1s and 2p electron probability distributions
11
time.
Potential energy and electron distributions: From electron distributions we can de-
duce potential energies. At this early stage in your technology career, you may only be able
to deduce qualitative potential energy trends (later, once you know integration you will be
able to get exact numerical answers). For now, compare the
1s
and
2p
distributions shown
in Figure 2.3. In the former distribution, the electron on the average lies much closer to the
nucleus than for the latter distribution.
Recalling the numerical expression for the electrostatic energy, U
q

q
+
r
, and noting that
r on the average is smaller for the 1s distribution, while q
+
(the nucleus) is a positive number,
while q

(the electron) is a negative number, we nd the value of U more negative for the
1s state than the 2p state. We can similarly compare most pairs of electron probability
distributions. This comparison is important. For example, if we know the relative potential
energies, by applying the virial theorem we know relative kinetic energies. As the total
energy is the sum of the kinetic and potential energy, we can further deduce relative total
energies.
Electrons in unstable distributions: Unfortunately, the virial theorem only applies to
stable electron orbitals. Thus while in all cases we can deduce the electrostatic potential
energy, in general, for unstable electron orbitals, knowing just the electron probability distri-
bution alone, we know nothing either about the kinetic or the total energies. In such cases,
we must go beyond the electron probability distribution.
Electrons obey quantum mechanics: Electrons obey quantum mechanics. In quantum
theory, the kinetic energy of an object can be determined by knowing its wavefunction.
Wavefunctions, generally denoted by , the Greek letter psi, are related to the probability
12
Figure 2.4: 1s and 2p wave functions and probability distributions
13
density by the relation:

2
.
The wavefunctions of the 1s and the 2p orbitals are shown in Figure 2.4.
Figure 2.5: 1s and 2p wave functions and waves with similar wavelengths
Electrons are waves: The wavefunctions
1s
and
2p
bear a qualitative similarity to
waves. The waves closest in appearence to the two wavefunctions are shown in Figure 2.5.
14
Let us consider the two waves shown in this Figure. Both waves have a wavelength. We call
these wavelengths
1s
and
2p
, where is the Greek letter lambda. In quantum mechanics
there is a xed relation between and kinetic energy. For particles moving at or close to
the speed of light
T
rel

1

while for particles moving signicantly slower than the speed of light
T
nonrel

1

2
.
The subscripts rel and non-rel refer to relativistic and non-relativisitic. (Scientists generally
use these terms to distinguish objects which are moving close to or far away from the speed
of light.)
The electrons vital to chemistry lie far away from the nucleus and hence have slow speeds,
speeds signicantly less than the speed of light. In chemistry the latter equation is therefore
much more important. For the 1s and 2p orbitals,
1s
is shorter than
2p
. By the equation
above, the 1s state therefore has more kinetic energy. Both the 1s and 2p orbitals are stable
electron states. They both therefore obey the virial theorem. We conclude that the 1s state
has a more negative potential energy than the 2p state: the same conclusion as we found
from considering the two electron probability distributions.
2.2 Appendix: Angular Momentum
We conclude this chapter with an examination of angular momentum and its connection to
the total energy of both orbits and orbitals. We begin with orbits. The results we obtain
about planetary orbits orbits proves relevant to stable atomic orbitals.
15
Figure 2.6: The position, r, velocity, v, and angular momentum,

L, vectors
Planetary picture of angular momentum: In Figure 2.6 we re-illustrate planetary
motion, where the position vector, r; the velocity vector, v; and the angular momentum
vector,

L, have been added on. The position vector, r, describes the arrow pointing from
the sun to the planet. The velocity vector, v, is an arrow describing the direction of motion
of the planet. (The direction of the arrow v gives the direction of travel at a given time;
the length of this same arrow is proportional to the speed of the planetary motion at the
same given time.) Angular momentum is also a vector. As the Figure shows, angular
momentum can be represented by an arrow which is simultaneously perpendicular to both
r and v. Furthermore, if r and v are mutually perpendicular, the length of the arrow

L is
proportional to the length of r times the length of v. These relations are expressed by the
formula

L = mr v,
where m is the mass of the planet and is the vector cross product.
16
Relation between angular momentum and stable orbitals: The angular momentum
is therefore completely determined by knowledge of r and v. In Table 2.1 we listed the
magnitudes of r and v for the nine planets. Applying the formula above to the values in this
Table, we can deduce the angular momentum for one kg of any of the listed planets. We
leave this task to the student. We note the chief nding of such a calculation: the farther a
planet lies from the sun, the greater is its angular momentum. (One of the problems at the
end of this chapter asks the student to conrm that this fact is a consequence of the virial
theorem.)
Applying the virial theorem to angular momentum: In summary, the greater the
value of r, the greater the angular momentum,

L. Conversely the greater the value of

L,
the greater the value of r. From the virial theorem, we know the greater the value of r, the
less negative is the potential energy, U, and the less the value of the kinetic energy. We
conclude, the greater the value of L, the less negative is the value of the total energy. This
result is absolutely true for planetary orbits. As our primary interest is electrons in orbitals,
we note here the principal nding for orbitals: for all atoms and ions with more than a single
electron to them, increases in the angular momentum lead to less negative total energies.
2.3 Problems
1. Based on the virial theorem prove or disprove the following statement. For an atom
with an electron, the total energy of the electron can never be positive.
2. In outer space, there is a single proton. A million miles from this single proton is a
single electron travelling at one-thousanth the speed of light. Does this situation prove
that the virial theorem is sometimes false. If not, why not.
3. One curious feature of the
1s
and
2p
probability distributions is that at distance r = 0
17
the former function has a non-zero value while the latter function is exactly zero. We
will learn later that the angular momentum of an electron in the 1s state is zero while
an electron in the 2p state is not zero. Show why the values of the angular momentum
cause
2p
to be equal to zero at r = 0.
4. The next more excited state (ie., even higher energy state) than the 2p state is the 3d
state. Draw two fundamentally dierent possible qualitative pictures of both
3d
and

3d
. For the sake of comparison draw on the same page
2p
and
2p
.
5. Angular momentum is a conserved quantity. Deduce from this fact that all planetary
motion is planar.
6. Both linear momentum and energy are functions of the mass and velocity. Compare
an adult man walking at 3 miles an hour, a baseball thrown at 50 miles per hour
and a bullet travelling at 1000 m/s. Estimate the kinetic energy and the momentum
of these three without looking up any information on the internet. Which has the
greatest momentum and which has the greatest kinetic energy? Without looking up
any information on the internet, to one signicant gure what is the conversion factor
between miles per hour and meters per second?
7. An atom is composed of a nucleus and its electrons. At room temperature a nucleus
has roughly two and a half kJ (kiloJoules) of kinetic energy per mole. Electrons are
travelling at one one-hundreths to one-tenth the speed of light. What is the order
of magnitude of the ratio of the kinetic energy for the electrons vs. the nucleus in a
carbon atom?
8. A planet of mass m is travelling in a circular orbit at a distance r from a sun of mass M.
The gravitational constant is G, ie., U = G
mM
r
. Write an expression for the angular
momentum of this planet using only the variables M, m, and r and the constant G.
18
9. In the above problem we assumed that the planet was travelling in a circular orbit
around the sun. Actually both the sun and the planet are travelling in circular orbits
around their center of mass. Let r and R be respectively the distances the planet and
the sun are from their center of mass. Further assume that the planet, the center of
mass, and the sun lie at all times in a straight line. Express R as a function of m, M,
and r. Assume the center of mass is motionless. Using the law of the conservation of
linear momentum, write an expression for the speed of the sun, V , as a function of just
m, M and the speed of the planet, v. What is the ratio of the kinetic energies of the
sun and the planet? What variables control this ratio?
10. In the above problem you calculated both the ratio of r and R as well as v and V
assuming that the sun and the planet stay in a straight line. Are your calculated
values of these two ratios compatible with this assumption?
11. If the absolute temperature is halved, the kinetic energy of the nucleus is halved, but
the kinetic energy of the electrons remain essentially the same. In what way are the
results of the previous two problems in contradiction with this fact?
19
Chapter 3
Electron orbitals and the periodic
table
In this chapter we will make a number of striking statements about electron orbitals. The
facts themselves are so peculiar that I worry that it will be hard for you, the student, to
make much sense of them. They are, however, (for the most part) facts (1) which freshman
chemistry students typically have to memorize in standard freshman chemistry courses and
(2) which allowed scientists for the rst time to understand the organization of the periodic
table (these facts are therefore important to chemistry as a whole). In this chapter we will
give as plain a rendering of these facts as possible. We reserve for subsequent chapters
explanations as to where these facts come from and how these facts can be related to the
wave-like nature of electrons.
Quantum numbers: All electrons in atoms lie in an orbital, a wave function. Such
orbitals are called atomic orbitals. All stable atomic orbitals can be described by four
dierent quantum numbers. A quantum number is a kind of physical constant. All quantum
numbers are unitless.
20
When scientists speak of constants they have two dierent constants in mind. On the one
hand there are constants like the speed of light or the rest mass of the electron. These are
constant numbers: neither the speed of light nor the rest mass of an electron ever changes.
But scientists speak of another kind of constant. In unperturbed collections of physical
objects (such collections are called systems) quantities such as the total energy or the total
momentum do not change. These latter constants, are constant only so long as the system
is unperturbed. To distinguish these constants from the xed constants (such as the speed
of light), scientists call these latter constants conserved quantities. Quantum numbers are
conserved quantities. As long as an electron in an orbital with assigned quantum numbers
is unperturbed neither by other atoms nor receives nor emits light, the quantum numbers of
the electron stay xed.
Principal quantum number: The rst quantum number is n, the principal quantum
number. The principal quantum number is always a positive whole number. The lowest
possible value for n is 1. For all atoms and all ions there is at least one atomic orbital for
each value of n. Of all the quantum numbers, it is the principal quantum number which
gives the most information about the energy of the orbital. For atoms and ions with just a
single electron (H, He
+
, Li
2+
, etc.) the relation between the principal quantum number and
the total energy is a precise one:
E
n
=
ZR
H
n
2
,
where Z is the nuclear charge (Z = 1 for H, 2 for He, and so forth) while R
H
= 2.210
18
J.
While for atoms with more than one electron, the relation describing the total energy is a
more complicated one, n remains a dominant energetic variable.
21
3.1 Orbital angular momentum
Orbital angular momentum quantum number: The orbital angular momentum quan-
tum number, l, is a whole number which ranges in value from 0 to n 1, where n is the
principal quantum number. Thus for an n value of 1, the only possible value of l is l = 0;
for an n value of 2, l = 0 or l = 1; and for an n value of 3, l = 0, l = 1 or l = 2.
From the text above, we have the sense that the principal quantum number tells us
something about the energy of an electron. The orbital angular momentum quatum number
tells us about another physical quantity, the angular momentum.
The l quantum number is suciently important in chemistry that each of the values of l
have specic names. l = 0 orbitals are called s orbitals; l = 1 are p orbitals; and l = 2 are d
orbitals. Chemists typically designate orbitals by a combination of the principal and orbital
angular momentum quantum numbers. For example, the 2p orbital refers an orbital where
n = 2 and l = 1; The 3s orbital refers to n = 3 and l = 0.
The relation between l and

L: In the case of the principal quantum number, n, one
can write an exact equation relating total energy to n, only for atoms or ions with exactly
one electron. For the case of the orbital angular momentum quantum number, l, the total
orbital angual momentum is exactly determined for all atoms and ions. The formula relating
l to L is

L
2
= h
2
l(l + 1),
where h is Plancks constant divided by 2 and where L = |

L|.
Plancks constant: Plancks constant is one of the most important constants in physical
chemistry. Its value is: h = 6.6 10
31
kJ s; if one divides h by 2 one obtains h, h =
1 10
31
kJ s. The units of Plancks constant are of interest. We derive these units by
examining the equation relating L
2
and l. As l is unitless and

L
2
is in units of angular
22
momentum squared, the equation

L
2
= h
2
l(l + 1) tells us that h is in units of angular
momentum. As we shall see later, Plancks constant will give us the order of magnitude of
many of the basic angular momentums found in chemistry.
Figure 3.1: The relation between E and l for several dierent atoms.
Relating l to the total energy: The greater the value of l, the greater the value of L.
We now make the connection between L and the total energy,E. To make this connection,
we invoke the similarity between electron orbitals and planetary orbits. For planets in stable
orbits (see the Appendix of the previous chapter), the greater the value of L, the greater is
the total energy. This statement is also true for atoms, but with one proviso: the atom or
ion in question must have more than one electron.
Thus, for all atoms with more than one electron, the greater the value of l, the greater
the value of L; and the greater the value of L, the greater is the value of E. Thus the 2p
orbital has a less negative total energy ( a less negative energy is a greater energy) than the
23
2s orbital; similarly the 3d has higher energy than 3p which is higher in energy than 3s.
Examples of these trends are shown in Figure 3.1.
Summary of the relation between n, l and the total energy: In the absence of a
magnetic eld, for atoms and ions with just a single electron to them, the total energy, E,
is independent of l and depends solely on n: the greater the value of n, the less negative is
the total energy.
In the absence in a magnetic eld, for atoms and ions with more than a single electron,
the total energy, E is solely a function of n and l. For a xed value of l, the greater the
value of n, the greater is the total energy; similarly, for a xed value of n, the greater the
value of l, the greater the total energy.
Figure 3.2: A schematic for the Aufbau principle.
24
The ordering of the atomic orbitals: The ordering of the energy of atomic orbitals
follows a straight-forward mnemonic, see Figure 3.2. For all atoms with more than one
electron, the lowest energy orbital is 1s. The energies of the orbitals then follow in the
sequence 2s, 2p, 3s, 3p, 4s, 3d, etc. As the Figure shows, the energies of the orbitals follow
a sequential pattern most easily discerned with the diagonal lines of Figure 3.2.
3.2 Directional orbital angular momentum
The third quantum number is the directional orbital angular momentum, m
l
. The value
of m
l
is restricted to being the integers which range from l to l in value. Thus for all s
orbitals, where l = 0, m
l
= 0, while for p orbitals, where l = 1, m
l
can have one of three
values: -1, 0, or 1.
Origin of m
l
: The m
l
quantum number is derived from the vector nature of the angu-
lar momentum. While the l quantum number determines the magnitude of the angular
momentum, it tells us nothing of the direction of the angular momentum. The most illumi-
nating example are the p orbitals, where l = 1. We enumerate here the

L which correspond
to the quantum numbers m
l
= 1, 0, or 1. The magnitude of

L is xed by the formula

L
2
= h
2
l(l + 1): for l = 1, L =

2 h. However, the direction of



L changes: m
l
= 1 corre-
sponds to

L = (0, 0,

2 h); m
l
= 1 to

L = (0, 0,

2 h); and the m


l
= 0 state refers to a
specic, but here unspecied, vector perpendicular to the previous two vectors.
m
l
is a conserved quantity: In the absence of any magnetic eld, m
l
is always a conserved
quantity. m
l
can furthermore remain a conserved quantity if the external magnetic eld is
aligned in an appropriate direction. For example, in the previous example of p orbitals,
whose m
l
were given with respect to the z axis, if the magnetic eld is aligned in the z
direction, the m
l
quantum number remains conserved.
25
Figure 3.3: The energies of m
l
= 1, 01 as a function of the external magnetic eld strength,
B.
Figure 3.4: Three possible orientations for an l = 1 electron.
The relation between m
l
and the energy: The quantum number m
l
aects the total
energy only in the presence of an external magnetic eld. Referring again to the previous
example where l = 1 and the magnetic eld is aligned along the +z axis, the m
l
= 1 state
is lowest in energy, the m
l
= 1 is highest in energy, while the energy of the m
l
= 0 orbital
is unchanged. These qualitative results are shown in Figure 3.3, where the energy levels of
26
the m
l
= 1, 0, 1 states are plotted as a function of the strength of the magnetic eld.
These results can be understood by thinking of a classical charge moving in a xed
circular orbit. We can imagine three such cases: the charge has an angular momentum
parallel, perpendicular and anti-parallel to the direction of the magnetic eld (which for
convenience, we place into the z direction, see Figure 3.4.
We consider now just the case where the charged particle has an angular momentum
parallel to the direction of the magnetic eld. In such a case,

L = (0, 0, L). In a classical
picture the energy of such a charged particle would lowered by the magnetic eld. We now
turn to the electron with the angular momentum (0, 0, L) as is found in the m
l
= 1 state. As
Figure 3.3 shows, just as in the classical picture, the quantum mechanical electron behaves
in exactly the same way. Similar arguments can be applied to the m
l
= 0 and 1 orbitals.
3.3 Directional spin quantum number
Electron spin quantum number Independent of an electrons spatial orbital, an electron
always has an internal angular momentum called its spin. This angular momentum carries
the spin quantum number s = 1/2. The spin quantum number while unitless, is half-integral
in value. The relation between s and the spin angular momentum,

S is the same as is found
between l and

L:

S
2
= h
2
s(s + 1).
As the s quantum number is always xed to the value of 1/2, the value of S for an electron
is also always xed at

3 h/2. S and s both being xed, chemists rarely concern themselves


with this quantum number.
Directional spin quantum number: Just as in the case of of orbital angular momentum,
where we could further consider directional orbital angular momentum, we can consider the
27
directional spin quantum number, m
s
. In just the same way that the m
l
quantum number
is limited by l, the m
s
quantum number is limited by s. m
s
can range in values from m
s
to +m
s
, where values of m
s
must dier from each other by whole numbers. As in our case
s = 1/2, the quantum number m
s
has two possible values 1/2 and +1/2. Recalling that
the magnitude of

S is

3 h/2. The value m


s
= 1/2 corresponds to the

S = (0, 0,

3 h/2) and
the value m
s
= 1/2 corresponds to a vector pointing in the opposite direction.
The number of atomic orbitals with the same value of n and l: In the absence of a
magnetic eld, neither m
l
nor m
s
play any role in either the total angular momentum or the
total energy of an electron. While m
l
and m
s
do specify the direction of the orbital and spin
angular momenta, for chemists, the most important feature of these two quantum numbers
is that they in addition specify the number of orbitals which have a given value of n and l.
Take for example the 2p or the 3p states. In both cases l = 1 and m
l
ranges over the
values of 1, 0, or -1. m
s
has the value 1/2 or -1/2. A 2p or 3p electron therefore has
simultaneously one of the possible three m
l
values and one of the two possible m
s
values: a
total of six possibilities. A d orbital, with l = 2 has simulanteously both one of ve possible
values for m
l
and one of the two possible values of m
s
: a total of ten possibilities.
3.4 Electron orbitals and the periodic table
The four quantum numbers (n, l, m
l
, and m
s
) specify completely a stable atomic orbital.
These numbers are called quantum numbers as they relate to the quantum mechanical prop-
erties of electrons. It may not be suprising therefore, that to learn about the important role
these numbers play in chemistry, the student will need to learn some additional facts about
quantum mechanics.
28
Identical particles: In quantum mechanics, a group of electrons are identical to one
another. Consider two electrons (lets call then A and B) and even let these two electrons
have dierent quantum numbers, ie., lie in dierent orbitals: once they come together in an
atom, it is none-the-less impossible to distinguish electron A from electron B. All particles
which have this sort of property are called identical particles.
Fermions and bosons: There are two types of identical particles: fermions with half-
integer spin (an electron with s = 1/2 is a fermion) and bosons with integer spins (a photon,
where s = 1 is a boson). Fermions obey the Pauli exclusion principle: only one electron can
have one complete set of four quantum numbers at a given time. Fermions with a given set of
quantum numbers exclude all other identical particles from the state which they occupy. (By
contrast, bosons draw other identical bosons into the same orbital: lasers, superconductors
and superuids are all examples of boson attraction.)
The Aufbau principle: In Figure 3.2, we gave the energy ordering of the various atomic
orbitals. We can combine this ordering with the Pauli exclusion principle. Consider for
example an atom with two electrons. In a stable arrangement of the atom, these two electrons
will occupy the lowest energy orbitals. Two electrons can occupy the 1s orbital. One of the
two electrons has m
s
= 1/2, the other has m
s
= 1/2. The atom with two electrons is He,
helium. As the arguments above demonstrate, He has two electrons in the 1s orbital.
Consider now an atom with 10 electrons. The two lowest orbitals are 1s and 2s. Each of
these orbitals can take one up-spin (m
s
= 1/2) and one down-spin (m
s
= 1/2) electron: a
total of four electrons. The next lowest orbital is 2p. In this orbital m
l
can take one of three
values and m
s
on of two values: a total of six possibilities. Thus there can be six electrons in
the 2p state. The atom Ne, neon, with ten electrons has all 1s, 2s, and 2p states lled. Both
the He and Ne examples are examples of the Aufbau principle: the principle in which the
Pauli exclusion principle is coupled with the energy requirements of the dierent orbitals.
29
Electron congurations: The electron orbital occupations play a dominant role in the
chemistry of the elements. Chemists represent these occupations in a number of dierent
ways. We illustrate two of the most important dierent ways here. The rst method is to
list the occupied orbitals in sequential order. For example for Ne the sequential ordering
is 1s
2
2s
2
2p
6
, while F, uorine, with nine electrons is 1s
2
2s
2
2p
5
, where the superscripts
indicate the number of electrons occupying a given orbital.
Figure 3.5: Three possible AO diagrams for the uorine atom. The picture on the right is
the valence atomic orbital diagram.
Atomic orbital diagrams: Even more important for this course are atomic orbital di-
agrams. Atomic orbital diagrams, AOs, list a portion of all the occupied and unoccupied
atomic orbitals. Three examples of the F AO diagram are shown in Figure 3.5, Note in the
AO diagram, we list a series of occupied and unoccupied atomic orbitals. It is not necce-
30
sary to list all the occupied and unoccupied orbitals. The third example shown is the most
commonly used. We speak more of this third example in the next two paragraphs.
Valence, core and excited-state orbitals: The electron conguration notation listed
above becomes quite cumbersome for atoms with a large number of electrons. Consider iron,
Fe, with 26 electrons. If we were to represent its full electron conguration it would be 1s
2
2s
2
2p
6
3s
2
3p
6
4s
2
3d
6
. Chemists often will abbreviate this series by stating explicitly only
those orbitals occupied after the last preceding noble gas element. In the case of Fe, the
last noble gas conguration lled is argon, Ar (Ar has 18 electrons). Chemists will therefore
write the Fe electron conguration as [Ar] 4s
2
3d
6
.
Chemists call the orbitals lled up to the last lled noble gas conguration core orbitals.
Chemists further dierentiate orbitals which lie above the last lled noble gas conguration
into one of two types. The lowest energy of these orbitals are called valence orbitals.
The nomenclature of valence orbitals is somewhat vague. However, as a good rule-of-
thumb, the valence orbitals are those orbitals which come between the last lled noble gas
conguration and the next not-completely-lled noble gas conguration. Thus is the case
of Fe, valence orbitals could be taken to be the 4s, 3d, and 4p orbitals. Orbitals which
lie above both the core and valence orbitals are called excited state orbitals. In a stable
arrangement of an atom, core orbitals are always lled; valence orbitals are sometimes lled;
and excited-state orbitals are never lled.
In Figure 3.5a, the core, valence and some of the excited state atomic orbitals were shown;
in Figure 3.5b, just the valence and some of the excited-state atomic orbitals were given;
and nally in Figure 3.5c, we show just the valence atomic orbitals. This nal diagram is
the valence atomic orbital diagram.
31
Figure 3.6: The relation between the periodic table and the Aufbau principle.
The periodic table: All chemistry students in this course have seen the periodic table.
As a student you may have wondered at its unusual format, with short rows at the top and
long rows on the bottom. We are now poised to understand the shape of the periodic table.
In Figure 3.6 we present a periodic table based on the consecutive lling of atomic orbitals.
First lled is the 1s orbital, then the 2s, and so forth. The number of elements for each
value of n and l depends on the number of possible dierent values of m
l
and m
s
. It may be
seen that the shape of the periodic table is determined by the four atomic quantum numbers
coupled with the Pauli exclusion and Aufbau principles.
3.5 Conclusions
This chapter has demonstrated that if all atoms have four quantum numbers, and if the
counting principle developed for these four quantum numbers is correct, then the quantum
numbers play a role in the periodic structure of the elements. We conclude this chapter by
32
examining this proposition more fully.
From previous chemistry courses, you have learned that elements lying in the same column
of the periodic table often have similar chemical properties. For example, examine column
14 of the periodic table. In this column, the elements carbon,C; silicon, Si; and germanium,
Ge lie one beneath the other and all have similar chemical properties. Examining Figure 3.6,
we see that for each of these three elements, the valence orbitals all have dierent n quantum
numbers, but the same l quantum numbers.
We conclude that it is the l quantum numbers which play the greatest role in periodic
trends, while the n quantutm number is of less importance. We further conclude that the m
l
and m
s
quantum numbers also play a role in atomic properites. In the case of the chemically
similar C, Si, and Ge exactly two of the six possible valence p orbitals (and all the valence
s orbitals) have been occupied. We deduce the number of occupied s and p orbitals is
important. And if this is so, it is reasonable to assume that each of occupied valence p
orbitals are based on the same combination of m
l
and m
s
states.
In the next chapter, we consider an explicit spatial represention of these m
l
and m
s
combinations. Giving a spatial shape to each of these orbitals will allow us rst to make
the bridge from quantum numbers to the shape of orbitals, second determine the connection
between the shape of orbitals and the shape of stable molecules, and nally understand the
relation between the shape of molecules and chemical reactivity.
3.6 Problems
1. Based on the Aufbau principle, draw a qualitative Atomic Orbital diagram of iron.
Include the 1s 2s, 2p, 3s, 3p, 3d, 4s and 4p orbitals in your picture.
2. Draw a quantitatively accurate atomic orbital energy diagram for Li
2+
, including the
1s 2s, 2p, 3s, 3p, 3d and 4s orbitals. Label axes and state numerically the actual
33
energies of the orbitals. Compare this diagram to the diagram of neutral lithium, see
Figure 3.1. Does Li
2+
obey the Aufbau principle as stated in Figure 3.2 ?
3. Monochromatic light has a xed frequency, see Figure 3.7. The frequency of this light,
is related to the energy of one photon of light by the formula E = h. The frequency
of light is also related to the wavelength and speed of the light, respectively and c
by the formula c = . Calculate the order of magnitude of energy of one mole of
photons of low energy ultra-violet light. A Joule is the kinetic energy of 2 kg moving
at 1 m/s. Estimate the order of magnitude of a punch by Mike Tyson. Compare these
two orders of magnitude. Why might sunlight cause skin cancer?
Figure 3.7: The names of the dierent sorts of light.
4. Atoms can have electrons in excited states. Electrons in these excited states can, by
34
emitting photons of light, lower their energy from an excited state into their ground
state. As the energies of the atomic orbitals are xed, the light emitted have xed
frequencies. For the hydrogen atom these xed frequencies typically belong to series,
see Figure 3.8. Deduce equations for the frequencies of emitted light for both the
Balmer and Lyman series.
Figure 3.8: Light emitted by excited state hydrogen atoms.
5. (a) (10 points) An electron in a Li
2+
occupies the 4p atomic orbital. What is the
shortest wavelength of light that can be emitted by this ion?
(b) (15 points) Light is absorbed by this same ion, causing an electron to y o with
speed 1.60 10
6
m s
1
. What is the frequency of this light? (Physical constants
which you might need for this problem are given on the bookends of most freshman
chemistry textbooks and on the internet.)
35
6. In the previous chapter you were asked to consider the order of magnitude of the kinetic
energy of the nucleus and the electrons in an atom at room temperature. What is the
ratio of the orders of magnitude of these energies (for a mole of atoms) compared to a
punch from Mike Tyson?
7. Draw a graph indicating how the valence atomic orbital diagram of a neutral boron
atom changes as a function of the strength of an external magnetic eld. The horizontal
axis of this graph should be the external magnetic eld strength.
8. Consider Figure 3.4. For l = 2, there are ve possible m
l
quantum numbers. Draw a
picture like that of Figure 3.4 for these ve quantum states.
9. Consider a two-dimensional world. Angular momemtum is still a conserved quantity,
but there are signicant changes in directional angular momentum. Adapt Figure 3.4
for such a 2-D world. Based on the gure you drew, deduce the 2-D law for possible
m
l
quantum numbers as a function of l. What changes would there be in a 2-D worlds
periodic table?
10. For the carbon atom, the energies of the 2s, 2p, 3s, 3p and 3d orbitals are approximately
-21, -11, -4, -3, and -1 eV. (An eV/molecule corresponds to roughly 100 kJ/mol.) Cal-
culate the energies of light which can be absorbed by a carbon atom up to a maximum
energy of -10.5 eV. Are these energies in the ultra-violet, visible, or the infra-red regions
of light? Is it possible for diamonds to be colorless?
11. In the photoelectron experiment, a photon of light travels across a chemical sample.
For low energy photons not too much happens to the sample. But at a certain photon
frequency, the sample ejects one of its electrons. Furthermore for all frequencies above
this threshold frequency, electrons continue to be ejected. Calculate the threshold
frequency for carbon. The existence of this threshold energy is part of the photoelectron
36
eect. Einstein won the Nobel prize for discovering this eect.
37
Chapter 4
The shape of atomic orbitals
In Chapter 2 we learned that electrons in atoms lie in atomic orbitals and that these atomic
orbitals can be described by wave functions, . The functions are spatial functions, ie.,
there is a value of the functions for every point in space. They give a complete description
(quantum mechanics places constraints on what a complete description is) of the electron
and from the values of the one can deduce the total energy, the potential energy, the
kinetic energy, and so forth.
In Chapter 3, you found out that all atomic orbitals can be completely described in terms
of four quantum numbers, n, l, m
l
, and m
s
. In this chapter we begin to make the connection
between the functions and the four atomic quantum numbers. At the same time, you will
learn the shape of the functions themselves.
4.1 The names of the spatial atomic wavefunctions
are spatial functions: Every pair of values of n and l, have spatial wave functions,
nl
related to them. (A spatial function is a function which has a concrete value at every point
in space.) In general, a given value of n and l can have more than one spatial wavefunction
38
associated to them. The number of these spatial wave functions is determined by the number
of m
l
allowed by the l quantum number. (m
s
as it is concerned with an internal motion of
the electron, the spin, plays no role in the shape of spatial wave functions.) For example,
consider the 3p atomic orbitals. There are three allowable values of m
l
for these values of n
and l (m
l
= 1, 0, 1) and there are three spatial 3p wavefunctions. By this same logic there
is only one 3s wavefunction and ve 3d wavefunctions.
Some notation: To faciltate the discussion in this chapter we introduce the following
notation. The symbol np will refer to an electron in a p state (with an l quantum number
of 1) and whose n quantum number is unspecied. The symbols ns, nd, (and occasionally
nf) will also be used in an analogous manner.
Table 4.1:
ns

1s
= N
1s
e

2s
= N
2s
(2 )e

3s
= N
2s
(6 6 6
2
)e

2
=
2Zr
na
0
a
0
0.5

A
The ns atomic wavefunctions: Using this notation, we note for all values of n, there
is only one ns orbital. Therefore ns designates a specic wave function,
ns
. An electron
in the ns state occupies the
ns
orbital. In Table 4.1, we explicitly list
ns
for dierent n
values where there is only one electron in the atom. The value r in this table is the distance
between the electron and the nucleus and the value Z in this formula gives the nuclear charge
in atomic units (the Z of H is one, of He two, etc.)
Also included in this formula are two constants N
ns
and the Bohr radius, a
0
. The exact
values of N
ns
may be obtained from more advanced textbooks. They will play no signicant
role in this course. The value of the Bohr radius, a
0
, is approximately 0.5

A. Note that the
39
radii for all the atoms in the periodic table range from 0.5 (for H) to 1.5

A(for the biggest
elements). These results indicate that the Bohr radius plays a pivotal role in the actual size
of all the atomic elements.
The np and nd atomic wavefunctions: For all n 2, there are three np orbitals (the
words atomic wavefunctions and orbitals are used interchangeably in this and subsequent
chapters. In general, all orbitals are wavefunctions but leaving atoms, not all wavefunctions
are orbitals). Each of these three orbitals has a distinct wave function name:
npx
,
npy
,
and
npz
.
For all n 3 there are ve nd wave functions:
ndxy
,
ndxz
,
ndyz
,
nd
x
2
y
2
, and
nd
z
2
.
In this chapter you will nd that the subscripts of both the
np
and
nd
wavefunctions have
an exact mathematical meaning with the exception of the
nd
z
2
wave function. The z
2
is a
nickname of the the true name: 2z
2
x
2
y
2
.
Origin of the np and nd wave function names: One of the most beautiful stories in
the quantum description of atoms is the precise origin of the subscript names of the np, nd,
and nf wavefunctions. We tell you the beginning of this story here (the rest of this story is
told later in this book).
Homogeneous polynomials: A homogeneous polynomial is a polynomial where every
term in the polynomial has collectively the same power to it. For example, the polynomial
x
3
+x
2
y +2zy
2
4y
3
is homogeneous as each of the four terms in the polynomial have a net
three powers of the x, y, or z coordinates. We dene the order of the polynomial to be the
sum of the x, y, and z powers of the term in a polynomial with the greatest such sum. For
example, x
3
yz + z
4
y is a homogeneous polynomial of the fth order.
40
Homogeneous polynomials of the rst order: In three-dimensional space, all homo-
geneous polynomials of the rst order can be described in the form x + y + z, where
, , and are numbers. The statement that there are three np orbitals, with names p
x
,
p
y
and p
z
is in fact determined by the fact that the three subscripts x, y, and z, taken as
polynomial terms, can be used collectively to describe all possible homogeneous polynomials
of the rst order.
Polynomials of higher orders: The ve nd orbitals all have subscript polynomials of
the second order. All seven nf orbitals have subscript polynomials of the third order. These
facts prove pertinent in the construction of the nd and nf orbitals.
4.2 The atomic wavefunction
The atomic orbital wavefunctions are suciently hard to describe that we construct them
in three separate steps. In the rst step we consider the angular part of the wavefunction,
ie., the change in the value of the function as we rotate from one point in space to another.
This part of the wavefunction is determined by the l and m
l
quantum numbers. Recall that
these two quantum numbers determine respectively the total angular momentum and the
z vector-component of the angular momentum. As angular momentum is connected with
rotation, it may not be surprising that these two quantum numbers determine the value of
the wavefunction as we rotate our position.
In the second step of the construction of the wavefunction we consider the change in the
wavefunction as we increase our distance from the nucleus. This portion, the radial part
of the wavefunction, is determined by a combination of n and l. The role of n and l is
reasonable if we recall rst that potential energy changes as a function of the distance of
the electron from the origin and second that both n and l play a role in the potential (and
kinetic) energy of the electron.
41
In the nal step we multiply the angular and radial functions to arrive at a total wave-
function. The nal step will prove to be an exercise in graphing more than anything else.
Figure 4.1: The unit sphere.
4.2.1 The angular part of the atomic wavefunction
The angular portion of the wavefunction can be thought of as the behavior of the wavefunc-
tion at a xed distance from the nucleus. In describing the angular portion it is easiest to
specify the xed distance: in this section we choose the convenient xed distance of 1 (the
unit of this distance is deliberately unspecied).
The unit sphere: Consider the unit sphere for an atom, the points in space which lie at
a unit distance away from the nucleus of the atom, see Figure 4.1.
The angular portion of
ns
: The angular portion of
ns
is Y
s
. It gives the value of the
wavefunction for all points lying on the unit sphere. Consider one of the
ns
wavefunctions
42
listed in Table 4.1. For any one of these functions, the value of the function is constant along
the unit sphere (
ns
vary solely as a function of r and r is constant on the unit sphere).
We could write these statements as Y
s
= C
s
, where C
s
is a constant. The values of these
constants are given in advanced textbooks and are not used in this course.
Note that all n
ns
orbitals have an identical angular portion of the wavefunction, and
that all these identical angular portions are equal to the same constant value.
The angular portion of
np
: The angular portion of the functions
npx
,
npy
, and
npz
(respectively Y
px
, Y
py
, and Y
pz
) are proportional to respectively the values of the functions
x, y and z on the unit sphere. Thus, Y
pz
= C
pz
z, where C
pz
is an unspecied and (for this
course) immaterial constant. Note also that for all values of n
npz
have exactly the same
angular functions, Y
pz
= C
pz
z.
In Figure 4.2a we show the relative values of Y
pz
at various points on the unit sphere.
This function is greatest at the North Pole where z = 1 and smallest at the South Pole
where z = 1. Note everywhere along the equator that z = 0.
A traditional plot for Y
pz
: Chemists traditionally plot the Y
pz
along the unit sphere by
plotting the value of the function as a distance away from the origin. Thus for the North Pole
where the relative value of Y
pz
= 1, we plot a point pointing in the direction of the North
Pole at a distance 1 away from the nucleus. At a latitude of 45
o
N (roughly the latitude of
Toronto, Canada), the value of z is

2/2, we go a distance of

2/2 away from the origin.


We furthermore plot all positive values of z in blue.
We now turn to negative values of z. At the South Pole, z = 1: we wish to plot a point
in the direction of the South Pole at a distance of one away from the nucleus. We use color
to indicate that the value of z in this direction is negative: we plot the South Pole point in
red as opposed to blue. Similarly, for latitudes of 45
o
S (where southern Chile is located)
43
Figure 4.2: The values of the function z on the unit sphere and a picture of
npz
. Blue and
red refer to respectively positive and negative values of
npz
.
44
we plot points in red a distance

2/2 away from the origin.


The complete plot of the Y
pz
function is shown in Figure 4.2b. The function has a double
balloon shape. The Northern and Southern Hemispheres are represented by respectively blue
and red balloons oriented respectively towards the North and South Poles. The function is
completely pinched in along the equator where the value of the function z is exactly zero.
4.2.2 A quick way to make traditional angular plots
Nodes: The easiest way to make angular plots is to examine the functions for those locus
of points where the function equals zero. Such locus of points are called nodes. For example,
in the case of
npz
the function z equals z along the equator: all points along the equator lie
on a node. In this example, as the nodes lie in a plane, we term this locus of nodes, a nodal
plane.
Nodes divide the unit sphere: The equator divides the unit sphere into two parts. We
may suppose that if the function equals zero at the equator, it is because in traversing the
equator, the function changes its net sign. An examination of Figure 4.2b shows that this is
true: in the Northern Hemisphere the function is positive, while in the Southern Hemisphere
it is negative.
Graphing the angular part of the
ndyz
orbital: We now turn to the angular part
of the function
ndyz
, Y
dyz
. This latter function is a constant times the function yz. This
function equals zero everywhere in the plane where y = 0 and in the plane where z = 0.
These two planes comprise two dierent nodal planes.
These planes divide the unit sphere into the four segments shown in Figure 4.3a. By the
argument used in the previous paragraph, we deduce that the function yz changes sign as it
passes from one segment into the next. A plot of yz, Figure 4.3b, shows that this deduction
45
Figure 4.3: The values of the function yz on the unit sphere and a picture of
ndyz
. Blue
and red refer to respectively positive and negative values of
ndyz
.
46
is true.
Graphing the angular part of the function
nd
z
2
: The z
2
in the function
nd
z
2
being a
nickname for 2z
2
x
2
y
2
, the angular part of this function is a constant times 2z
2
x
2
y
2
.
The crux to graphing this latter function is determining the values of x, y, and z for which
the latter function is zero.
Note that everywhere on a unit sphere x
2
+ y
2
+ z
2
= 1 and therefore x
2
+ y
2
+ z
2
1
equals zero. Adding this latter expression to 2z
2
x
2
y
2
does not change its value. We
therefore nd:
2z
2
x
2
y
2
= (2z
2
x
2
y
2
) + (x
2
+ y
2
+ z
2
1) = 3z
2
1.
Figure 4.4: The magic angle triangle.
The magic angle: The values of z in the previous paragraph relate to the so-called magic
angle right triangle (If you are interested in the magic angle look up Magic Angle on
Wikipedia.). An example of the magic angle triangle is shown in Figure 4.4. The triangle in
this picture is placed on the unit sphere. Its hypotenuse connects the origin to a point on
47
the unit sphere: it is of length 1. The vertical arm of the magic right triangle lies in the z
direction and has the value of 1/

3.
Substituting the value 1/

3 into the equation above, we nd that any point which lies on


a magic triangle has a value of 3z
2
1 = 0. Points which have such values have latitudes of
roughly 35.3
o
. (Note, the term magic angle is reserved for the other angle in the right triangle,
an angle of roughly 54.7
o
.) Cities in the Northern Hemisphere such Teheran or Knoxville,
Tennessee in the Northern Hemisphere or Buenos Aires in the Southern Hemisphere have
roughly a latitude near the magic angle.
Figure 4.5: The nodal cone and picture of
d
z
2
.
48
In Figure 4.5a we draw the nodal surfaces of the 2z
2
x
2
y
2
orbital. The nodal surfaces
have the appearence of two cones, both cones coming to a point at the origin. Applying
these nodal cones as the division between dierent-signed lobes of the 2z
2
x
2
y
2
orbital,
one arrives at the orbital shape shown in Figure 4.5b.
Figure 4.6: The three Y
np
functions, the angular part of
np
.
Figure 4.7: The ve Y
nd
functions, the angular part of
nd
.
49
Pictures of both
p
and
d
functions In an earlier section of this chapter, we mentioned
that in three dimensional space, there were three fundamental polynomials of the rst order
and ve fundamental polynomials of second order. The angular wave functions corresponding
to these polynomials are shown in Figure 4.6 and Figure 4.7.
4.2.3 The radial part of
The radial part of the wavefunction is that part of the wave function which depends solely
on the radial variable, r. Table 4.1 gave some early examples of such radial functions. The
radial part of the wave function is a single-variable function. It is typically denoted as R(r).
Square of the radial function: Just as the square of the wavefunction determines the
relative probabililties of an electron being located at dierent positions, the square of the
radial function can be used to determine the relative probabilities that an electron is found at
dierent distances from the origin. As potential energy is a function of the distance from the
origin, the function R(r) can be used to determine exactly the potential energy. For stable
orbitals, application of the virial theorem can be used to determine kinetic and therefore
total energy.
R
nl
(r): As the radial function completely determines the total energy for all stable atomic
orbitals and as the quantum numbers n and l are the sole quantum numbers which determine
total energy, it is reasonable that the values of n and l determine the radial function. This
being so the radial function requires just two quantum numbers as subscripts: it is denoted
R
nl
(r).
Table of R
nl
(r): A table of the radial parts of atomic wavefunctions is given in Table 4.2.
Inspection of this Table shows that all radial functions are composed of an exponential decay
function multiplied by a polynomial.
50
Table 4.2: R
nl
R
1s
= C
1s
e

2
R
2s
= C
2s
(2 )e

2
R
3s
= C
3s
(6 6 6
2
)e

2
R
2p
= C
2p
e

2
R
3p
= C
3p
(4 )e

2
R
3d
= C
1s

2
e

2
=
2Zr
na
0
a
0
0.5

A
The rate of the exponential decay depends on Z, n, and a
0
. The exponential decay
controls the size of the atomic orbital. The polynomials play an important role; they obey
several xed rules:
1. The order of the polynomials is always equal to n 1.
2. The number of roots of the polynomials is always equal to the order. (The root of a
polynomial gives the location of a node.)
3. The number of roots of the polynomial equal to zero is always equal to l.
4. Combining the information of the previous three points, the number of non-zero roots
is always n l 1.
The radial nodes: Just as an examination of the nodes of the angular part of the wave
function proved to be a quick way to plot the angular function, the nodes of the radial
function proves the most ecient way to plot the radial function. As the radial function is
composed of a polynomial times by an exponential decay, and as an exponential decay has
no nodes, all the nodes of the R
nl
are determined by the nodes of the polynomial. Note that
polynomial nodes are typically called roots.
51
R
ns
(r): By the above rules, the R
ns
(r) functions are the only functions which are non-zero
at the origin. R
1s
(r) has no radial nodes; R
2s
(r) has one radial node; R
3s
(r) has two radial
nodes and so forth.
R
np
(r): By the above rules, the R
np
(r) functions are zero-valued at the origin. Other than
the node at zero the R
2p
(r) function has no nodes; the R
3p
(r) has one node; and so forth.
R
nd
(r): By the above rules, the R
nd
(r) functions are zero-valued at the origin. Other than
the node at zero the R
3d
(r) function has no nodes; the R
4d
(r) has one node; and so forth.
4.2.4 Plotting R
nl
(r)
The above information provides the information necessary to make a reasonable graph of
R
nl
. We illustrate this method for R
3p
. In Figure 4.8a, we locate the nodes of R
3p
. We now
seek a smooth function which changes sign on traversing each of these nodes, whose shape
at large values of r takes on the shape of an exponential decay function.
Such a graph is illustrated in Figure 4.8b. Also plotted in this Figure is a qualitative rep-
resentation of 4r
2
R
2
3p
(r). This latter function is a weighted probability function in which we
take into account the dierent volume of space found for dierent values of r. Finally, for the
sake of comparison, we illustrate the true radial wavefunctions in Figure 4.9, Figure 4.10,and
Figure 4.11. Also shown, in Figure 4.12, are the weighted probability distributions which
correlate with these wavefunctions. As may be seen, the approximate drawing captures many
of the essential features of the true radial function (Note that the qualitative picture of
3p
is the negative of the quantitative version of
3p
.
52
Figure 4.8: R
3p
, the radial portion of the hydrogen-like
3p
wavefunctions.
53
Figure 4.9: Quantitatively correct R
np
, the radial portion of the hydrogen-like
np
wavefunc-
tions.
Figure 4.10: Quantitatively correct R
ns
, the radial portion of the hydrogen-like
ns
wave-
functions.
54
Figure 4.11: Quantitatively correct R
nd
, the radial portion of the hydrogen-like
nd
wave-
functions.
Figure 4.12: Quantitatively accurate probability distributions of the radial atomic orbitals
for hydrogen-like atoms.
55
4.3 Combining the angular and radial parts of
The wavefunctions are products of the angular and radial parts of the wavefunction. For
example,
3pz
= Y
pz
R
3p
(r).
4.3.1 Visualizing
Figure 4.13: The nodal zones of
3pz
.
Just as visualizatiion of the angular and radial parts of the wave function required visu-
alizing the nodes of these separate function, visualization of the full function is eciently
achieved by determining the nodes of the full function. That is a product of the angular
and radial parts facilitates this task. Consider again
3pz
= Y
pz
R
3p
(r). In order for
3pz
= 0
either the radial part or the angular part of the function needs to equal zero.
We consider rst the nodes of R
3p
(r). As Table 4.2 shows, this function has a node for
2Zr/3a
0
= 4. For a hydrogen atom Z = 1. Substituting the value of Z = 1 and a
0
= 0.5

A, this node lies at approximately the value of 3



A. Thus for all points where r = 3

A, the
56
radial function equals zero. The locus of such points forms a sphere of radius 3

A centered
at the origin.
We now consider the nodes of Y
pz
. As we noted earlier, this function equals zero for all
points on the plane where z = 0. We plot the nodal surfaces of both R
3p
(r) and Y
pz
in
Figure 4.13. As this Figure shows, taken together, the two nodal surfaces divide space into
four separate domains. The function
3pz
changes sign as we traverse from one of these four
domains into another.
4.3.2 Plotting
Figure 4.14: Contour maps in the xz plane for
3pz
and the xy plane for
4dxy
.
Even knowing what must look like, it is dicult to plot it in full. One method which
chemists use to convey much of the key information about is to make a contour plot. (If
you are not famililar with contour plots, go to Google images and type contour map.) In
this approach, we consider the values of for just a single plane.
For example, we can consider the y = 0 plane of
3pz
. Referring to Figure 4.13, we see
that the nodal surfaces divide this plane into four separate domains, see Figure 4.14. The
spherical radial-based node takes on the shape of a circle in this graph, while the planar
angular-based node take on the form of a sttraight line. Just as in the full three-dimensional
57
nodal plot, this planar section divides space into four domains. Also shown in this gure is
a contour plot of the
4dxy
.
At this point we can begin substituting values of the function over the y = 0 plane. Two
of the four segments must be of one phase, while the other two domains are of the other
phase. Adjacent domains must be colored dierently. In Figure 4.14 we give a contour map
which fulls these conditions.
The only unassigned feature of this plot is the relative signs of the blue and red regions.
Most conveniently, it turns out that there are essentially no experimental chemical issues
where it proves important as to whether red stands for a positive or negative number. Red
and blue must always be of opposite signs, but as to which is which, it is a matter of
convention.
Many examples of this procedure for other atomic orbitals are given in the Problems
Section of this Chapter.
4.4 Conclusion
In Chapter 2 we learned that in quantum mechanics all electrons in all atoms are described
by orbitals. These orbitals are functions. The orbitals further obey the virial theorem. They
can be used to obtain information on both the kinetic an potential energy. In Chapter 3 we
further learned that these orbitals are all described by four quantum numbers n, l, m
l
and
m
s
.
In this Chapter we have learned the exact shape of all the atomic orbitals. The angular
part of the atomic orbitals is governed by a homogeneous polynomial. The order of this
polynomial is determined by l. The radial part is an exponential decay function times by a
polynomial.
The nodes of the orbitals describe much of the shape of these functions. The nodes are
58
determined by the nodes of the angular and radial polynomials. The higher n and l are, the
greater are the number of nodes. The quantum number l controls the angular nodes, while
the quantum numbers n and l control the number of radial nodes.
4.5 Appendix: Angular plots
Angular plots have one confusing component to them. While they are plots of the value
of a function on a sphere, they themselves do not have the shape of a sphere. The key to
understanding angular plots is to recognize that in an angular plot we are plotting the value
of a function as a distance from the origin. The distances in an angular plot are the values
of the function. The following two examples might help.
Example 1: Average yearly snowfall on the surface of the Earth. Imagine the
average yearly snowfall on the Earth is as is shown in Figure 4.15. Drawan as an angular
plot such a set of values would be plotted as is shown in Figure 4.16.
Figure 4.15: Imagined average yearly snowfall on the Earth.
59
Figure 4.16: Imagined average yearly snowfall on the Earth plotted as an angular plot.
Example 2: Average temperature for dierent points on the surface of the Earth.
Imagine the average temperature on the Earth is as is shown in Figure 4.17. Drawan as an
angular plot such a set of values would be plotted as is shown in Figure 4.18. Note the
similarity of this curve to the d
z
2 angular plot.
60
Figure 4.17: Imagined average yearly temperature on the Earth during the last ice-age.
Figure 4.18: Imagined average yearly temperature on the Earth during the last ice-age
plotted as an angular plot.
61
4.6 Problems
1. Using the same conventions that are used to draw AO angular plots please draw graphs
for:
(a) The average snowfall per year. Assume that places with the same latitude have
the same snowfall. Note that, in reality, unlike the gure shown in the text of
this chapter, there is much less snowfall at the North Pole than for Toronto.
(b) The number of meteor strikes per 100,000 years. For simplicity, assume that the
Earth rotates on an axis perpendicular to the plane of its orbit around the sun
and that all meteors travel in the same plane as the plane in which the Earth
travels around the Sun.
(c) In the text for this chapter, we drew a graph of the average yearly temperature
for dierent points on the Earth. Imagine that the Earth travels in a circular
orbit around the sun but that is rotational axis was shifted by almost 90
o
, to lie
exactly in the plane of the Earths planetary orbit.
2. Please draw qualiative radial and angular pictures for = ((r
2
4r + 3)e
r
(x
2
y
2
).
Draw all 3-D nodal surfaces for this function. Please draw a contour map of this
function for points in the xy plane; for points in the yz plane.
3. Please draw a qualitative picture for a wave function where T = 0.
4. Consider the function = ((r
2
4r + 3)(x
2
y
2
) for an electron around a nucleus.
Assume the nucleus is located at the origin. What is the potential energy of this wave
function? Based on your previous answer, is this a stable orbital?
5. Does the angular part of an electron wave function give you any information about the
potential energy of an electron obeying this wavefunction? If so, how, if not, why not?
62
6. Based on the quantitative radial functions and their probability distributions shown in
this chapter, decide on the relative wavelengths for the radial part of the wave function
for the 1s, 2s, 3s, 3p and 3d states. Based on these wavelengths rank the kinetic energy
for the radial part of the wave function from lowest to highest.
7. For one-electron atom wavefunctions rank from lowest to highest the total kinetic
energy of the 1s, 2s, 3s, 3p and 3d states. Is your result for this problem potentially
compatible with your results for the preceding two problems and the virial theorem?
8. The following statements are true: Atomic orbitals obey the virial theorem. Thus
knowledge of the potential energy entirely determines the kinetic energy. Atomic or-
bitals can also be decomposed into radial and angular wavefunctions. The angular
wavefunction contains no information about the potential energy of the electron; it
only contains information about kinetic energy. One angular function looks very dif-
ferent from another. For this question, please explain away the following paradox:
How is it possible for the angular wave function to change from one orbital to another
when the kinetic energy is already determined by the virial theorem and the radial
wave function?
9. Draw the nodal zones for (a)
4d
z
2
, (b)
3dxz
and (c)
5f
x(y
2
z
2
)
. Please label your
axes in your above graphs. Draw contour maps in the xz plane for these three wave
functions.
10. Draw angular pictures which correspond to the (a) 1 + 2x, (b) 1, (c) x + y, and (d)
x 2y functions. The 1 + 2x function is called an sp hybrid wave function. It looks
neither like an s nor a p orbital, but a hybrid of these two orbitals. Is the x 2y
function a hybrid angular part of an orbital or is it a pure s, p or d state? If so, why,
if not, why not?
63
11. The fundamental wave motions on the surface of a circular drum are shown in Fig-
ure 4.19. Shown in this picture are the nodal zones, the relative frequencies, and
a pair of integers. Based on these pictures, draw contour maps of their respective
wavefunctions and their respective n and l quantum numbers. State the names of each
of the shown orbitals based on the axes given for all the orbitals in the left column of
the gure. Assume nu E. Draw the equivalent of an atomic orbital diagram for the
fundamental wave motions on the surface of a drum. On the surface of a drum there is
no potential energy just kinetic energy. Give a plausible argument why E(2s) > E(2p)
for a drum, while the converse is true for electrons bound to atoms. Apply this same
argument to the atomic vs. drum 3d vs. 3p states.
Figure 4.19: Fundamental wave motions on a drum.
12. All homogeneous polynomials of the rst order can be expressed as the weighted sum
64
of the three rst order polynomials, x, y, and z (such a weighted sum is called a
linear combination, see text). All homogeneous polynomials of the second order can
be expressed as the linear combination of six second order polynomials. State the six
simplest second order polynomials which you can think of for which this statement is
true. While, for a given n, there are three p orbitals, there are ve d orbitals. Consider
the homogeneous second order polynomial x
2
+ y
2
+ z
2
. What is the value of this
function on a unit sphere? Show that the all homogeneous second order polynomials
can be expressed as xy +xz +yz +(x
2
y
2
) +(2z
2
x
2
y
2
) +(x
2
+y
2
+z
2
),
where , , , , , and are all numbers, What statement is true about s and d
orbitals and all homogeneous second order polynomials?
13. In Figure 4.19 many of the resonant modes of a drum are shown. For each of these
resonant modes we can assign an n and and l quantum number. Develop a formula
relating n and l to the number of nodal zones. Such a formula is possible in one and
two dimensions, but not in three dimensions.
14. Consider Figure 4.20. Illustrated in this gure is a wave travelling in a circle. Draw
such a circular wave which corresponds best with:
(a) Y
npz
(b) Y
npx
(c) Y
ndxy
(d) Y
ndxz
Based on the relationship between and T, rank from lowest to highest the kinetic
energies associated with these four angular parts of the wavefunction.
65
Figure 4.20: Waves travelling in a circle.
66
Chapter 5
Particle-in-a-box, the Pauli exclusion
principle, and Hunds rule
5.1 Chapter goals
The goals of this chapter are three-fold. The rst goal is to discuss how atomic orbitals
have an analog to the so-called particle-in-a-box problem. The second goal is to consider
some of the restrictions Paulis exclusion principle places on the possible shapes of dierent
orbitals for both in-an-atom or in-a-box electrons. Finally we discuss how the Pauli exclusion
principle leads to Hunds rule (actually it leads to Hunds rst rule; there are three.)
5.2 Orbital kinetic energies
In the previous chapter, we stated that if we know the orbital an electron is in, we know
everything that we can know about that electron. (By knowing everything, we mean that
there is no further information about the electron which can be known within the bounds
of the laws of quantum mechanics.) Information contained within an orbital are its square,
66
which is proportional to the electron probability distribution, and the orbitals l and m
l
quantum numbers, which determine both the electrons total orbital angular momentum
and its directional angular momentum. This initial information can be used to derive other
useful quantities: from the probability distribution, we know the distance of the electron
from the nucleus, and hence the electrons potential energy. Due to the virial theorem, the
potential energy determines the kinetic energy, and hence determine the total energy.
In this chapter, though, we will note a dierent route to obtain a qualitative under-
standing of the kinetic energy. In Figure 5.1, we illustrate contour maps of both 2p and a 3p
orbitals. Placed beneath each picture is a cross-section of this contour map. Finally, beneath
this cross-section are 1-D waves which have wavelengths similar to the corresponding cross-
sections. As this gure shows, the 3p orbital has a longer wavelength than the 2p orbital.
Recalling that T 1/
2
, we conclude that the 3p orbital has the lower kinetic energy and
hence, by the virial theorem, the higher total energy.
We now consider a second set of orbitals. These orbitals describe electrons which are
conned to a spherical region in space. These electrons are said to be conned to a spherically
shaped box (the electrons are in-a-box rather than in-an-atom). For this second system, there
is no nucleus and no potential energy. Nonetheless, such a system will have the same types
of atomic orbitals which a true atom has. The spherically shaped box as well as contour
maps of the 2p and 3p orbitals are illustrated in Figure 5.2. Shown as well in this picture
are the cross-sections of these contour maps. As this gure shows, the wavelength of the 3p
orbital is smaller, and hence the 3p orbital has higher kinetic energy, As there is no potential
energy in this system, the 3p orbital is the higher in energy of the two.
In this example, the energy ordering of the electron-in-a-box problem is the same as the
energy ordering of the electron-in-an-atom.
67
Figure 5.1: The 2p vs. 3p state for electrons-in-an-atom.
68
Figure 5.2: The 2p vs. 3p state for electrons-in-a-spherical-box.
69
Particle in a box 2p vs. 3d: We now consider the 2p and a 3d orbitals for an electron
in-a-spherical-box. In Figure 5.3 we show contour maps of the two orbitals together with
cross-sections along both the x and y directions. As this Figure shows, while along the x
direction, the wavelengths of 2p and 3d look around the same, this is not the case for the
y direction. As the total kinetic energy is based on an average of all the wavelengths, we
conclude the 3d orbital has higher energy than the 2p. This is the same energy ordering
which is found in atoms.
Both the electron-in-an-atom and the electron-in-a-box often have the same ordering of
their total energy levels. The 2p state is lower in total energy than the 3d state in both types
of systems, even though for the former case, the kinetic energy is bigger while in the latter
case the kinetic energy is smaller. The ordering of the energy levels is often independent of
the exact model in use.
Minimal-valence basis sets: This last point is an important one in the course. In this
course we will study molecular orbitals. Molecular orbitals lie at the heart of the chemical
bond. The type of molecular orbitals we will use are minimal valence basis set molecular
orbitals. For minimal valence basis set molecular orbitals, the virial theorem is not obeyed.
For such molecular orbitals, lower energy orbitals have smaller kinetic energies. Only at the
end of this course will we explain how leaving the minimal-valence-basis-set approximation
will restore the accuracy of the virial theorem, while leaving the energy ordering of the
molecular orbital diagrams relatively unchanged.
70
Figure 5.3: The 2p vs. 3d state for elecrtons-in-a-spherical-box.
71
5.3 The Pauli exclusion principle and how it ts in to
the shapes of orbitals
5.3.1 Introduction
In Chapter 3, we learned that the Pauli exclusion principle says that two electrons can not
simultaneously enter orbitals with exactly the same quantum numbers. But in Chapter 4
we learned orbitals are also spatial wave functions. As orbitals are functions, the question
arises whether the Pauli exclusion principle, which appears at rst to be just about possible
orbital occupation, also relates to the wave functions themselves? The answer is yes, but to
understand this answer we need to generalize the way in which we view functions.
One of the central paradigms in this course is that functions can be looked at in many
dierent ways. For example, students know that the pictures of functions, graphs, are just a
dierent way in which to express the algebraic form of a function. It turns out that there is
an equally compelling connection between the the pictures of a function and another math-
ematical concept. There is a connection between functions and the mathematical concept of
the vector. (For those interested in terminology, this connection lies in the use of the Hilbert
vector space.)
In the next subsection, we will see that a function can be thought of as a sequence of
increasingly accurate vectors. More accurate vectors have more component parts to them
than less accurate vectors. But science functions are pretty much always smooth functions.
As we will nd, because of this smoothness, science functions contain less and less new real
information in every subsequent member of this sequence of vectors.
72
5.3.2 The shapes of orbitals and the Pauli exclusion principle
In Chapter 3 we discussed the Pauli exclusion principle in terms of the occupation of orbitals.
In Chapter 4 we discussed the actual shapes of these orbitals. We now consider possible
connections between the two. We start this discussion with the following scenario. Consider
an up-spin electron-in-an-atom in a 1s state. Based on its wavefunction, such an electron has
a nite probability to be located anywhere in space. Now let us consider a second up-spin
electron entering the system. By Paulis exclusion principle it can not enter the 1s state.
But the 1s state is spatially located everywhere. What kind of orbital can it enter?
Figure 5.4: The 1s, 2p
x
and 2p
y
wavefunctions evaluated on a 4-point grid.
To answer this question consider the 1s, 2p
x
and 2p
y
wavefunctions illustrated in Fig-
ure 5.4. Also shown in these pictures are four points (A, B, C, and D) located in a square
around the origin. Writing down the values of the wavefunctions at these four points
the 1s, 2p
x
, and 2p
y
orbitals are respectively

1s = (, , , ),

2p
x
= (, , , ), and

2p
y
= (, , , ).
The up-spin electron occupying the

1s state compels the the next up-spin electron to
73
enter the

2p
x
or

2p
y
states. These new electrons can enter the region of space which the
1s electron occupies. But the above formalism is set up not just to think of these states as
functions, but also as vectors:

1s,

2p
x
, and

2p
y
are all four component vectors. These three
vectors all have mutual dot-products of zero. (If you do not know what a dot product is, I
will explain it in class, or alternatively you can look up dot product wiki on the web.) Taken
as vectors, these three vectors are mutually orthogonal.
If the rst up-spin electron is in the 1s state, the Pauli exclusion principle tells us,
among other things, that the second up-spin electron can have no 1s character. To have no
1s character means that taken as a vector, the second up-spin electron is orthogonal to the
rst up-spin electron.
5.3.3 Some math about orthogonality
Reducing space to just four points is a big approximation, one which we can do away with.
In Figure 5.5, we consider a sixteen point mesh in space. Although we will not write down
the 16 component vectors here, the 1s, 2p
x
, and 2p
y
vectors all still have mutual dot products
equal to zero; they are all still orthogonal. In exactly the same way we could consider bigger
and even bigger meshes. For all symmetical choices of mesh, mutual orthogonality is still
preserved.
Mathematicians and scientists consider the limiting case where the mesh extends over all
of space in a ne even net. We write this this limit as


1s
(r)
2px
(r)dr = 0.
If the two wavefunctions are complex (rather than real), this expression is more correctly

(r)(r)dr = 0,
74
Figure 5.5: A 16-point grid.
where the superscript

refers to the complex conjugate.
5.3.4 Conclusion
The Pauli exclusion principle, that two electrons can not be in the same orbital at the same
time is compatible with the spatial picture that two dierent atomic orbitals are orthogonal
to one another. For example, if the rst elecron is in the 1s orbital, the second electron can
be in the orthogonal 2p
x
and 2p
y
orbitals. Being perfectly orthogonal is a mathematical way
of saying that two vectors are completely dierent from one another and that the 1s orbital
is entirely dierent from the 2p
x
and 2p
y
orbitals.
5.4 Hunds rule
While orthogonality is compatible with the Pauli exclusion principle, orthogonality is not
the same as the Pauli exclusion principle. To show this, we end this chapter by exploring the
connection between the Pauli exclusion principle and Hunds rule. Paulis exclusion principle
tells us that two electrons with parallel spin can not be at the same point of space at the
75
same time. (By contrast two electrons of opposite spin could be at the same point of space
at the same time.) Placing two electrons at the same point of space at the same time raises
signicantly the potential energy of the system and consequently the total energy. Hence
the Pauli exclusion principle implies it is energetically advantageous to place two electrons
into the same spin direction.
Figure 5.6: Hunds rule applied to an energy degenerate (degenerate means here equal en-
ergy) pair of orbitals.
Consider now the case of two atomic or molecular orbitals of equal energy, see Figure 5.6.
There is no energy dierence between these two orbitals and therefore no a priori energetic
advantage to place an electron in one vs. the other orbital. We place one electron into one
of these orbitals arbitrarily as an up-spin electron. Where would a second electron go into
this system if we are to minimize the energy? By placing the electron into another up-spin
orbital, by Paulis exclusion principle, it can never be at the same point in space at the same
time. This fact signicantly lowers the energy, stabilizing the system. Therefore, the second
electron enters the second orbital as a second up-spin electron (this is Hunds rule). (This
line of reasoning will be revisited more fully in the penultimate chapter of this book.)
5.5 Appendix: singlets and triplets
When two electrons have both similar m
s
orientation, the individual electron s quantum
numbers (which we will call s
1
and s
2
) can be added or subtracted to form a new combined
quantum number, S, where S = s
1
s
2
. If the two electrons both have parallel spin then
76
S = s
1
+ s
2
= 1 and S = 1. This state is a member of the triplet state. If the two electrons
can have opposite m
s
orientation, the s
1
and s
2
quantum numbers can be either added or
subtracted. In the former case S = 1, while in the latter case S = 0. The former is a member
of the triplet state, while the latter is in the singlet state. The S = 1 state is paramagnetic:
it can be attracted to a magnetic eld; the S = 0 state is diamagnetic: it is repulsed by a
magnetic eld.
5.6 Problems
1. The ground state orbital for all types of systems never has any nodes. Based on the
arguments in the text of this chapter, explain how we know that all excited states must
have at least one node.
2. In 2-D just as in 3-D, the s, p and d orbitals for both electrons-in-an-atom as well as
electrons-in-a-box are based on, respectively, homogeneous polynomials of the zeroth,
rst, and second order. However, in 2-D, these polynomials involve only the two
variables x and y. In 2-D, homogeneous polynomials of the rst degree are expressed
as the linear combination of how many dierent rst order polynomials? What are the
names of the 2-D p orbitals? What is the value of x
2
+ y
2
evaluated on a unit circle?
Using the same logic as used in the 3-D problem, nd the names of the 2-D d orbitals.
3. For a 2-D electron-in-a-circular-box, both draw pictures for and estimate the relative
energies of the 1s, 2p and 3d states. How many 2p states are there? How many 3d
states?
4. Consider a change in the circular box of the type shown in Figure 5.7. Draw a dia-
gram indicating how the change in the shape of the box changes the energies of the
orbitals given in the preceding problem. State for a system with a large circle to square
77
distortion whether a system with eight electrons is paramgnetic or diamagnetic.
Figure 5.7: A circle to more squarish shape distortion.
5. Consider the two electron occupation diagrams shown in Figure 5.8. Based on the
arguments in this chapter, which is lower in energy, and why.
Figure 5.8: Two degenerate electrons.
78
Chapter 6
Strings, drums, waves, and orbitals
6.1 The electron and waves
The previous chapters of this book have centered on the electron as it behaves in atoms (or
boxes). You have been told that while for attributes such as energy and angular momen-
tum, a planetary picture is correct, atomic electrons are best described by their probabilty
distributions and their wavefunctions. These two functions have very specic shapes, are
based on each other, and are themselves connected to four quantum numbers. While these
quantum numbers are connected to the energy and the angular momentum, they themselves
have whole number (integer) values and are unitless.
This chapter has as its goal an analysis of the above statements. We raise the question,
if electrons are waves, ie., have wavefunctions, then what might be the consequences of their
wave-like nature? For many of us, our ability to answer this question is limited by our
knowledge of waves. While some of us will know the mathematics used to describe one
dimensional waves, many students at this point in their career will not be able to similarly
describe two-dimensional or three-dimesnional waves. This is a serious weakness as electrons
in atoms are in three-dimensional waves or wavefunctions.
79
In this chapter we will observe one and two-dimensional waves. While we will nd some
overlap between one-dimensional waves and atomic orbitals, it will be for two-dimensional
waves that the connection to atomic orbitals is most clear. For two-dimensional waves, we
will nd there are radial and angular parts of the wavefunctions; s, p, and d orbitals; and
homogeneous polynomial subscripts for the various orbital types: all attributes we associate
with three-dimensional atomic orbitals.
6.2 An undulating rope
We begin this analysis by considering one-dimensional waves. In the chemistry department
at Cornell, we possess a very long spring used to make such one-dimensional waves. One can
think of it as a very long jump-rope. Just as in a jump-rope, one stretches this spring between
two people, each person holding one end of the spring. One (or both) of the people begins
to shake the spring up and down. One nds that there are several stable spring motions,
illustrated in Figure 6.1. All stable motions are wave-like, with the lowest observable energy
motion with no nodes (the ground state), the next with one node (the rst excited state),
and so forth.
Based on direct observation, energy is nor just inversely proportional to wavelength. The
wave pattern where the wavelength is a third of the original wavelength requires an energy ex-
ertion which is visibly more than three times the original energy exertion. Looking at the web
(analyzing waves on a string at http://galileo.phys.virginia.edu/classes/152.mf1i.spring02/
AnalyzingWaves.htm), one nds, that for rope waves the energy per per unit length, E,
E 1/
2
, the same relation as has been claimed for electron waves.
Encouraged by this equivalent energy relationship, we can try to nd an equivalence
between the shapes of the orbitals. We note that in the above gure, the ground, rst excited
and second excited states have respectively zero, one and two nodes. This progression in the
80
Figure 6.1: Stable wave motions of a 1-D rope.
number of nodes resembles the patterns we found for atomic orbitals, where the 1s orbital
has zero nodes, the 2s and 2p orbitals have one node, and 3s, 3p and 3d orbitals look to
have two nodes. Based on this analysis, we suspect there is a relation between the number
of nodes and the energy. We will need to consider more complicated patterns to go further
in our analysis.
6.3 String Instruments
We will do a demo involving string instruments. (String instruments convert one dimensional
string waves into sound waves). In the rst demo we will hear the change in frequency as a
function of the string length. We hear primarily the ground state. In the rst part of the
demo, we will hear the change in frequency as we shorten the length of the string. Note that
the frequency , obeys the relation: 1/. As the string is shortened, does the frequency
go higher or lower?
81
In this demonstration, the violinist will play a full octave on one string. A note one
octave higher has double the frequency. Where does the violinist place his or her nger on
the string to obtain a note one octave higher? When the violinist places his or her nger to
double the frequency, does the part of the string on the other side of the bow move or not
move? If it were to move just as much as the lower half of the string, would the string be in
its ground state or rst excited state?
In the second part of the demo we consider the pizzicato sound. Which note sounds
purer: the one created by the bow or the one created with a single nger? In Figure 6.2 we
consider a wave which is the sum of the ground state,
0
, the second excited state,
2
, and
the fourth excited state,
4
. What would a wave look like if it were the equal combination
of all three of these waves? One key in answering this question is the knowledge that waves
add. The combination wave,
comb
would therefore follow the formula:
comb
=
0
+
2
+
4
.
The combination wave is illustrated in Figure 6.2. The composite state is literally the
sum of the three component states. This composite wave has its largest value in the center
where the three compoent states are all in phase. Elsewhere, the composite function hovers
around zero in value as the three components are of varied phase and tend to cancel one
another out.
With this example in mind, we return to the question of the pizzicato sound. In Figure 6.3
we illustrate an idealized version of the pizzicato distortion of a string. As this gure shows,
the distortion is localized in one specic region of the string. As the previous discussion
illustrated, such a wave is a composite wave made up of many dierent components. To our
ears, we hear all these dierent components simulatenously. The resulting sound is less pure
in frequency than the bow-driven sound.
82
Figure 6.2: Ground, second excited, fourth excited and an example of a combination state
of a wave-like rope.
Figure 6.3: An idealization of the pizzicato string distortion.
83
6.4 Drums, plates and wineglasses
We now move to two-dimensional waves. There is a lovely Youtube presentation of two
dimensional waves located at http://www.youtube.com/watch?v=v4ELxKKT5Rw. In this
demonstration a circular membrane is placed next to a speaker. The speaker is emitting a
single frequency, the frequency itself is shown in a digital display in the background. Shown
in this demonstration are the ground, rst, second, third and fourth excited states. Draw
each of these states as a contour map. While the ground state corresponds to the 1s motion,
what do the rst four excited states correspond to?
We now move to an in-class demonstration. This demonstration is called Chladni patterns
after the rst scientist who discovered them. In this demonstration, we drive a circular metal
plate with a single frequency emission. We also place sand onto the surface of the plate.
This sand accumulates at the points on the plate which are motionless. Such points are the
nodes of the wave pattern. We will show three nodal patterns in class. Identify the names
of each of these three states.
As a nal two-dimensional demonstration we consider the Youtube presentaion at http://www.
youtube.com/watch?v=5VZcY1B8iPM. In this demonstration, we see one of the motions the
rim of a wineglass makes. Is this an s, p or d state?
6.5 Conclusion
In this chapter we have considered a number of dierent wave motions. We have found that
the classication of wave patterns based on atomic orbitals and quantum numbers provides
an excellent basis for circular waves in general. Atomic orbitals have something to do with
circles.
We deduce that the atomic orbital classication scheme is a consequence of the wave-like
nature of the electron coupled with the spherical shape of the atomic potential energy. If this
84
is true there is a signicant reduction in the number of mysteries which we need to resolve
vis-`a-vis electrons. Once we can demonstrate that electrons behave like waves, we will be
drawn to the conclusion that electrons should be found in the atomic orbitals as we have
described them.
In this course, we will go one step further. Electrons have a mass: they can follow
individual paths. We can talk about the size of an electron. Such a description may seem
uncompatible with the claim that electrons form in waves. Objects which move like planets
(or soccer balls) are said to move like particles. Physical entities which move like the surface
of a drum or the edge of a wineglass, do not follow discrete paths and are said to move as
waves. We will discover that it is possible to unite the particle description and the wave-
picture of electron motion into a single unied mechanics. The unied picture which we will
use will be based on quantum mechanics.
Finally, this chapter also introduced to us two interrelated and important concepts.
Waves add. We can generate composite waves from simpler waves by adding the simpler
waves to one another. This set of principles is so important that it bears the name: the
Superposition Principle. This principle is one of the fundamental principles of quantum
mechanics.
6.6 Problems
1. Consider functions of x where x can range from /2 to /2. Consider a string which
extends from /2 to /2 which is xed at both the /2 and the /2 end. The
ground state of this string is the wavefunction
0
= cos(x), the rst excited state is

1
= cos(2x), the second excited state is
2
= cos(3x), and so forth. If you have access
85
to a programable calculator plot the function:

10
=
10

i=0

i
.
If you can program a computer, plot

10,000
=
10,000

i=0

i
.
Please make your plot so that the maximum value of the function can be recorded on
the graph.
2. As the number of terms summed approaches innity, the function

acquires the
shape shown in Figure 46.3. Where is the

state located in space?


Figure 6.4: The

state.
3. Based on the previous problems guess what the function

n=1

ns
looks like for the hydrogen atom. Imagine there is a particle which decomposes into
an electron and a proton. Imagine we measure the energy of the electron after such a
86
particle has decayed. What are the possible energy values which we might measure?
(Note that when the energy is measured for an isolated atom, only values correspond-
ing to specic atomic orbital energies of orbitals can be observed.) What would be
the measured total orbital angular momentum of the electron after such a decay has
occurred?
4. The above problems are a rst demonstration of the Heisenberg uncertainty principle.
At this point, we can express this principle in the following way: If one has complete
knowledge of the position of an electron, one has no information about the momentum
of this same electron. Explain how the above problems provide an example of the
Heisenberg uncertainty principle.
5. Imagine two circular plates come into contact with one another, see Figure 46.2. Draw
contour maps of the two lowest in energy wavefunctions. The ground state has no
nodes while the rst excited state has one node (a line or a circle). Indicate which of
thee patterns is lower in energy and why.
6. Sound waves are air density waves. Imagine there is air inside a sealed hollow metal
sphere. Draw pictures of the 2s and 3d
xy
sound waves inside the sphere. How do these
waves evolve in time?
Figure 6.5: Two circular plates which are welded to one another.
87
7. For a square Chladni plate the three lowest energy states are the one 1s and two 2p
wavefunctions. Draw contour maps of the three wavefunctions and indicate where, if
anywhere, sand would accumulate on the plate if the plate were to be resonating in
one of these three modes.
88
Chapter 7
The H
2
molecule MO diagram
In this chapter, we begin our discussion of molecular orbitals, MOs. MOs form one of the
principal themes of this course. MOs can at rst be dicult to understand. Our approach
to them will be a rule based approach. Over the next few weeks you will learn ve axiomatic
rules, the use of which will alow you to derive the MO diagram of pretty much any molecule
you are interested in.
The ve rules are stated below. It will take a few weeks for you to learn what each and
every one of these rules mean. Our goal here is just to give an informal introduction to them.
This chapter itself will actually only make use of the rst four.
7.1 Five general rules for the construction of MO dia-
grams
1. In a minimal basis set MO diagram, the number of MOs equals the number of AOs.
2. When two orbitals combine, they combine to make two new orbitals. The original
orbitals combine to make the lowest and highest energy combinations possible.
90
3. Orbitals always combine to make orthogonal combinations.
4. Orbitals which are initially close in energy interact stronger than orbitals which are ini-
tially far apart in energy. When two orbitals of dierent energy combine, the resultant
low energy combination resembles more the initially lower energy orbital; the resultant
high energy combination resembles more the initially high energy orbital. Two orbitals
of equal initial energy combine to make two new orbitals both of which resemble each
of the two starting orbitals equally.
5. If a molecule has C
2
rotations and/or mirror planes, s, which are all parallel or
perpendicular to each other, then MOs are either symmeric or antisymmetric with
respect to the C
2
, , and (if the molecule possesses one) inversion center.
7.2 Combining the AOs of two adjacent hydrogen atoms
7.2.1 Concepts needed for this chapter
This chapter is centered on the question of how the electron wavefunctions of a the hydrogen
atom change when two hydrogen atoms are placed adjacent to one another. To construct
a good answer we will use the following concepts (all of which you have seen in previous
chapters):
1. As we discussed in the last chapter, wavefunctions can add to one another.
2. As the wave functions add, the shape of the wavefunction changes.
3. If in the changed wavefunction the wavelength, , gets longer, then T must decrease.
Vice-a-versa if becomes shorter. T grows larger.
4. In the case of minimal basis set MOs, the virial theorem is not obeyed. For minimal
basis set MOs, as T increases, the total energy of the orbital becomes less negative.
91
5. Wavefunctions and the negative of the same wavefunctions are two ways of expressing
the same physical state.
7.2.2 Two hydrogen atoms approaching one another
Let us now consider two hydrogen atoms who have just begun to approach one another. In
Figure 7.1 we consider two such hydrogen atoms.
Figure 7.1: Two hydrogen 1s AOs which have begun to approach one another
We focus on the two adjacent 1s orbitals. The top picture in this gure shows the two
1s orbitals. The middle picture shows a cross-sectional view of the two 1s AOs. And in the
92
bottom picture in this gure, we bring the hydrogen atoms close enough together that the
crosssections of the individual 1s orbitals begin to overlap with one another.
In the top two diagrams of Figure 7.2, we further consider the case where the two hydrogen
AOs overlap. As the two AOs overlap, both separate AOs have non-zero values for spatial
points, and especially for those points which lie between the two hydrogen nuclei.
Figure 7.2: Two hydrogen 1s AOs approaching one another
93
As we saw in the previous chapter, two wavefunctions combine; they add. This added-
together function is shown in the second of the two top diagrams in Figure 7.2.
In the bottom three panels of this same gure we consider how this added-together
function fuurther evolves as the two hydrogen atoms come even closer to one another . As
the middle of these pictures show, as the hydrogen atoms continue to approach one another
the shape of the added-together wavefunction changes even more dramatically. We can
expect that this alteration will change the wavelength of the added-together wavefunction.
If we compare the wavelength of this bottom panel with the wavelength which corre-
sponds to the isolated hydrogen 1s orbital, see Figure 7.3, we see that once two hydrogen
atoms approach one another the wavefunctions wavelength increases. As an increased
corresponds to a smaller T, in the minimal basis set approximation, the wavefunction of two
hydrogen AOs added together is lower in energy than the initial AOs.
Figure 7.3: The of an isolated hydrogen 1s AO. Compare the here to the one shown in
the previous gure.
94
7.2.3 The case where one of the AOs arrives in its negative form
The previous section allowed us to see how two hydrogen 1s orbitals would interact if we
assumed that both orbitals were expressed in their posiitive forms. But as we discussed in
the previous chapter, a wavefunction and its negative are both equally good representations
of any given wave.
In this section we consider two hydrogen 1s which come together while one 1s orbital
is in negative form while the other is still in its positive form. As in the previous case,
as the two hydrogen atoms approach one another closely, the added-together wavefunction
changes dramatically. But, as comparison of the added-together wavefunction at the bottom
of Figure 7.4 to the isolated hydrogen atom wavefunction of Figure 7.3 shows, the wavelength
of the added-together wavefunction is this time shorter than the initial hydrogen 1s orbital.
Recalling the connection between to T, and the connection between T and total energy,
we conclude the added-together wavefunction is of higher energy than the initial AOs.
95
Figure 7.4: Two closely adjacent hydrogen 1s AOs coming together, where one of the AOs has
adopted its negative form. Please note that there is an error in this gure: the cross-section
on the upper-right should be positive, not negative.
96
7.2.4 The two dierent ways in which adjacent hydrogen 1s or-
bitals combine
In Figure 7.5 we combine the results of the two previous subsections. The top panel shows
the two cases in the cross-sectional view. The bottom panel of this same gure converts
these cross-sectional views into the more traditional angular forms of the same orbitals.
Figure 7.5: The two dierent ways in which hydrogen 1s orbitals on adjacent hydrogen atoms
can combine. The top shows a cross-sectional view of the AOs and MOs. The bottom shows
an angular picture of these same orbitals.
97
The bottom panel of Figure 7.5 is very close to the traditional H
2
MO diagram, as it
is presented in most freshman chemistry textbooks. This traditional diagram is shown in
Figure 7.6. Note in the traditional diagram one includes the total number of valence electrons
(which is two for the H
2
molecule, one from each hydrogen.) As this gure further shows,
electrons in MO diagrams follow the Aufbau principle found in AO diagrams.
Figure 7.6: The two dierent ways in which hydrogen 1s orbitals on adjacent hydrogen atoms
can combine. The outsid eof this diagram states what the various electron orbitals looked
like before the two halves came together. The middle portion shows the MOs after the two
halves of the molecule have come together. Electrons follow the Aufbau principle. Note that
there are the same number of valence electrons (two in this case) before the two parts of the
molecule come together, as after they come together.
7.3 Deriving the H
2
MO diagram from the general
rules of MO diagrams
At the beginning of this chapter we said that there were a few axiomatic rules which can be
used to derive MO diagrams of any molecule of interest. In this last section of the chapter
we will apply these general rules to the H
2
molecule. In so doing, I hope the underlying
98
structure of the H
2
MO diagram will become more clear.
Figure 7.7: The H
2
MO diagram derived from the general rules. Note that in this picture,
the electron orbitals before the two parts of the molecule come together are shown on the
left, while after they have come together are shown on the right. Compare this presentation
of the data to the more traditional picture shown in the previous gure.
On the left hand side of Figure 7.7 we draw two horizontal lines, each with one electron.
Under each horizontal line we draw a picture of an H
2
molecule. For each of these pictures,
we draw the initial two AOs.
Rule 1 states, In a minimal basis set MO diagram, the number of MOs equals the
number of AOs. As there are two initial AOs, we deduce from this rule that the nal MO
diagram will consist of two MOs.
Rule 2 states, When two orbitals combine, they combine to make two new orbitals.
The original orbitals combine to make the lowest and highest energy combinations possible,
while Rule 4 states, Two orbitals of equal initial energy combine to make two new orbitals
both of which resemble each of the two starting orbitals equally. Based on Rules 2 and 4,
we deduce the two combined orbitals will have the shapes shown on the right of the gure.
In particular, we see that the two resultant MOs are the lowest possible and highest
possible energy combinations of the initial AOs. And these combinations are both 50/50
99
combinations of the initial AOs.
We now turn to Rule 3, which states, Orbitals always combine to make orthogonal
combinations. Recall that the idea of orthogonality works best in the vector picture of
functions. In Figure 7.8, we consider the two MOs illustrated on the right of Figure 7.7.
Figure 7.8: The H
2
MOs are orthogonal to one another.
As we wish to be simple in our treatment, we consider just two points, A and B. Both
these points are illustrated in this gure. As the gure shows, one MO can be represented
by the vector (1, 1) (In Chapter 5 we would have called this same vector (, )), while the
other vector is called (1, 1) (in Chapter 5 it would have been called (, ).
The dot product of these two vectors is zero. We conclude that the two MOs are indeed
orthogonal.
7.4 Problems
1. Draw an MO diagram for the ion H

2
. Does this ion have a stronger or weaker bond
than the neutral H
2
molecule?
2. Create an MO based explanation as to why the He
2
molecule does not exist.
3. As discussed previously, atomic orbitals are orthogonal. Consider the six second order
polynomials: x
2
, y
2
, z
2
, xy, xz, and yz. By the techniques described previously,
100
these are six orthogonal orbitals. We may therefore call these six orbitals by the six
orthogonal vector names: (1, 0, 0, 0, 0, 0), (0, 1, 0, 0, 0, 0), and so forth. Express the ve
d orbitals as six-component vectors. Demonstrate that these ve vectors, so expressed,
are orthogonal. Verify that these ve vectors are also all orthogonal to the vector
(1, 1, 1, 0, 0, 0). Rationalize why all d-orbitals are orthogonal to this vector.
4. The names of the ve d-orbitals is in fact partially arbitrary. The ve d-orbitals must
(a) all be second order homogeneous polynomials, (b) all be orthogonal to one another,
and (c) all be orthogonal to the (1, 1, 1, 0, 0, 0) vector. Imagine that on the planet Oz,
freshman chemistry textbooks have decided that two of the ve d-orbitals must be the
orbitals z
2
y
2
+ xz + xy and y
2
x
2
+ xz + yz. Write down possible names for the
remaining three d-orbitals.
5. The f orbitals are all third order homogeneous polynomials which are orthogonal to
the three p orbitals x(x
2
+ y
2
+ z
2
), y(x
2
+ y
2
+ z
2
), and z(x
2
+ y
2
+ z
2
). In one
conventional set of f-orbitals there are three f orbitals named x
3
, y
3
, and z
3
. These
orbitals all have similar shape, of the type shown in Figure 46.1. Just as in the case
of the z
2
orbital, these three names are nicknames. Based on the nickname and the
provided picture, suggest a reasonable true functional name of these three f-orbitals.
Figure 7.9: The f
z
3 angular function
101
Chapter 9
The HF molecule and electronega-
tivity
We now consider the HF molecule and its molecular orbital diagram. Just as the goal of
the last chapter was to get a sense of how p orbitals on dierent atoms mix together in an
MO diagram, our goal in this chapter is to understand additional concepts pertinent to the
making of MO diagrams. Our goals are two-fold. First, we consider here an MO diagram
with two dierent elements. Second, we will introduce the concept of electronegativity. Elec-
tronegativity measures the attraction an electron has to a specic element. In this chapter
we explore the physical principles responsible for electronegativity but also we consider the
eects electronegativity has on MO diagrams in general.
9.1 A simplied HF MO diagram
Just like the MO diagrams we derived in the last chapter for N
2
and O
2
, the HF MO
diagram discussses in this chapter is really just a stepping stone. It will introduce to us the
MO General Rule 4 and it will teach us how to incorporate electronegativity in the making
103
of MO diagrams.
In Figure 9.1 we show the ve valence atomic orbitals for the HF molecule: there are four
AOs associated to the F atom, but only one emanating from the hydrogen atom. As we will
discuss in the second part of this chapter, the energies of the H 1s orbital, the F2s, and the F
2p orbitals are vastly dierent from one another. We will discover that the F 2s electrons have
the most negative potential energy, followed by the F 2p, and that the H 1s electron has the
least potential energy. And therefore by the virial theorem, E(F2s) < E(F2p) < E(H1s).
We graphically show the energies of these dierent orbitals in Figure 9.2.
Figure 9.1: The ve AOs on which the MO diagram of the HF molecule is based.
Figure 9.2: The ve original AOs in the HF molecule, ranked by energy
104
We now seek the two orbitals most likely to interact most strongly with one another. In
particular we are interested in a pair of orbitals which belong to AOs on neighboring atoms,
and which mix together, change the wavelengths of their corresponding wavefunctions, and
in so doing create bonding and antibonding combination orbitals. Recalling General Rule 4
Orbitals which are initially close in energy interact stronger than orbitals which are initially
far apart in energy. When two orbitals of dierent energy combine, the resultant low energy
combination resembles more the initially lower energy orbital; the resultant high energy
combination resembles more the initially high energy orbital., we seek two orbitals on two
two separate atoms which are initially closest in energy. Referring to Figure 9.2, the two
orbitals of initial interest to us are the H1s orbital and a F2p orbital.
105
While there is only one H1s orbital, there are three F2p orbitals; in choosing the pair of
AOs which can intera ct most strongly with one another, we will need to select just on of the
three F2p orbitals. Ideas presented in the previous chapter allow us to select the best of the
three F2p orbitals.In Figure 9.3 and Figure 9.4 we consider the three possible interactions
between the H1s and a F2p orbital. As these gures show, for only the F2p
z
orbital is there
any net interaction between the H and F AOs. The interaction for the chosen pair is a
-interaction.
Figure 9.3: The hydrogen 1s AO does not interact with the F2p
x
or F2p
y
orbital.
Figure 9.4: Two hydrogen 1s AO can combine ini a fashion with either the F2s or F2p
orbitals. But the latter interaction is between AOs which are initially closer in energy, and
therefore the latter interaction is considered rst.
106
In Figure 9.5 we illustrate this strongest HF interaction. Note that we apply General
Rule 4: we make the bonding combination resemble most the intially low energy orbital,
and the high energy combination, the originally high energy orbital. We have placed vector
labels onto Figure 9.5, showing that the two mixed orbitals are orthogonal to one another.
Figure 9.5: The hydrogen 1s and uorine 2p
z
mixing.
In Figure 9.6, we add to this picture all those F AOs which have not yet been chosen to
interact. As this diagram shows, the HF molecule appears to have one lled bonding orbitals
and three lled non-bonding orbitals. The lled bonding orbital is a -bond, while the three
non-bonding orbitals are all located on the F atom.
Figure 9.6 already has much the look of an MO diagram. Have we captured enough of
the bonding in the HF molecule that we do not have to interact furhter, ie., can we already
get a feel for the electrons in the HF molecule? I believe the best way for now to answer
this question is to compare MO diagrams with the Lewis structure of HF. (If you do not yet
107
Figure 9.6: A possible simple MO diagram for the HF molecule.
108
know how to make the Lewis structure for the HF molecule, please do not worry, you will
learn later in this course ecient methods to determine Lewis structures.)
In Figure 9.7 we show the Lewis structure for the molecule HF. The Lewis structure
shows that there are eight valence electrons in the HF molecule grouped into four pairs of
electrons. Three of the electron pairs are non-bonding and are located on the F atom; one of
the pairs lies in the H-F bond. This is exactly the information the MO diagram in Figure 9.6
conveys. We conclude for now that our HF MO diagram is suciently accurate. (Note, HF
has sp hybridization, the F2s and 2p orbitals actually miix. As our MO diagram does not
show this, we will have to revisit the HF MO diagram once we consider sp hybridization.)
Figure 9.7: The Lewis structure for HF
109
9.2 Electronegativity
The remaining topic of this chapter is why the energies of the hydrogen 1s and uorine 2s
and 2p AOs are in the order, E(F2s) < E(F2p) < E(H1s). The answer to this question
will involve introduction to Gauss law and the concept of electronengativity.
Gauss Law is a physical law which tells one about electrostatic forces. (I will explain
the law below.) Electronegativity is a concept which tries to associate to each element
a number describing its innate attraction to electrons. We wish to establish a quantitative
scale in which elements that are more electron-attractive have higher electronegativity values.
Elements with high electronegativity would be electronegative, conversely elements with low
electronegativity would be electropositive.
Such an electronegativity scale could help quantify a number of basic chemical trends.
For example, most students in this course know electrons are more innately attracted to
neutral halogen (group 17) elements and less attracted to alkali metals (group 1). The
explosive reaction of an alkali metal such as elemental sodium, Na, and a halogen such as
elemental chlorine, Cl, producing NaCl is explosive as an electron is tranferred from the less
electron-attractive sodium atom to the more electron-attractive chlorine.
(There are nice internet demonstrations of this reaction on the web. My current favorite
is at http://graysci.com. Click on Chapter 1: Experimental Cuisine. The video at this
portion of the site shows chlorine gas passing over sodium metal with popcorn placed above
it. The Na and Cl reaction is suciently exothermic to gasify the NaCl salt and thus salt
the popcorn.) The electronegativity concept, which will give us a numerical value for the
dierence in electronegativity between sodium and chlorine, helps quantify the driving force
of this reaction.
Over the years, there have been a number of dierent attempts to develop such an elec-
tronegativity scale. We will consider two such scales in this chapter. The rst, the Mulliken
110
electronegativity, relates directly to atomic orbital energies. The second, the Pauling scale,
is more widely used, is formally derived from thermochemistry, but appears to bear a rela-
tionship with approximate eective nuclear charge. We explain this latter concept later in
this chapter.
9.2.1 Ionization energy and electron anity
In Figure 9.8 we show two dierent processes relevant to electronegativity. The rst is
ionization, the process in which an electron is removed from a neutral atom; the second is
electron capture, where an electron is captured by a neutral atom. The ionization energy
(more accurately, the rst ionization energy) is the minimum energy required for ionization.
The electron anity is the minimum energy released by the capture of an electron. Both
processes are illustrated in Figure 9.8.
Figure 9.8: Ionization vs. electron capture. Note an electron with zero energy is an electron
which having escaped the electrostatic well of the atom is now a free electron.
Based on this gure, one might believe that the ionization energy and the electron anity
111
were the same. This though is not correct. The ionization energy is the energy required by
atom A for the reaction,
A A
+
+ e

,
while the electron anity is the energy released by the reaction,
A + e

.
The former reaction produces a cation, while the latter reaction creates an anion.
We will make the arbitrary decision, for now, that the height of the atomic orbital energies
in Figure 9.8 will correspond to the average of the ionization energy and the electron anity.
This average energy is the Mulliken electronegativity of the element.
The Pauling electronegativity of the elements: While the Mulliken electronegativity
scale bears a clear correlation to AO diagrams, it is not the most widely used electronegativ-
ity scale. The more widely used scale is the Pauling scale. The Pauling scale is derived from
thermochemistry. It places the most electronegative element, uorine, with an electronega-
tivity value of 4.0 and the least electronegative element, cesium with a value of 0.7, see
Figure 9.9. The numbers in the Pauling scale appear at rst arbitrary. It may be somewhat
reassuring to know that the Mulliken and Pauling electronegativites are correlated to one
another. A correlation plot of the two scales is shown in Figure 9.10.
Periodic trends: The electronegativity scales presented in the above gures look fairly
complex. But often times we wish to know the relative electronegativity of two elements.
The following mnemonics may prove useful. For main group elements (elements which belong
to columns 1,2, and 13-17) of the periodic table, the closer the element is to uorine the more
electronegative is the element. For transition and noble elements (columns 3-12), the closer
112
Figure 9.9: The Pauling electronegativity,
P
, scale.
113
Figure 9.10: Correlation of the Mulliken,
M
, and Pauling,
P
, electronegativity scales.
the element is to gold, the more electronegative it is. Finally, hydrogens electronegativity is
somewhere between boron and carbon. Thus hydrogen pulls electrons away from boron but
donates electrons to carbon atoms. In borohydrides (boron-hydrogen compounds), the hy-
drogen receives net electrons from boron, but in hydrocarbons (carbon-hydrogen compounds)
the hydrogen donates net electrons to carbon.
9.3 Z
eff
and its relation to electronegativity
The electronegativity scale is one of the basic scales used to organize our understanding of
atomic periodic trends. It, in turn, can be related to the hydrogen-like model of atomic
orbitals. To make this connection, we need to know Gauss law of electrostatics. With
Gauss law, we will be able to simplify multi-electron atoms into a type of single-electron
atom. Such a simplication is vital. In previous chapters of this book, we have been provided
with the formula for the energies of the atomic orbitals of single electron atoms, but not for
114
multi-electron atoms.
9.3.1 Gauss law
Gauss law applies to electrostatic forces caused by spherically symmetrical charge distribu-
tions. The law has two dierent parts. Consider, rst, a hollow sphere with uniform negative
charge throughout the sphere, see Figure 9.11. Imagine an electron is then placed within
this sphere. The rst part of Gauss law tells us the electron feels no net electrostatic forces
from the hollow sphere. Consider now this same hollow sphere, but the electron is located
outside the sphere. The second part of Gauss law tells us that the electron feels a force
exactly equivalent to the force it would have experienced had all the charge on the hollow
sphere been concentrated into the center of the sphere.
Figure 9.11: Electrons interior and exterior to a hollow charged shell.
9.3.2 Applications of Gauss law to atoms
Example 1: An external electron near a hydrogen atom: In the hydrogen atom the
nucleus has the charge of +1 and the electron is in a 1s orbital. The 1s electron density
function is spherical and hence Gauss law is obeyed. The hydrogen 1s radial probability
distribution is given in Figure 4.12. This gure indicates that nearly all the 1s electron
115
probability distribution lies at distances less than 5 a
0
, 2.5

A. We now consider an external
electron which comes within twenty

A of such a hydrogen atom.
The hydrogen 1s probability distribution can be thought of as a series of concentric
spherical charge shells. Each of these shells is inside with respect to the external electron.
According to Gauss law, all the inside electron density acts as if it were located at the center
of the spherical 1s orbital. As the entire 1s orbital is inside of the external electron, the
entire 1s electron probability distribution acts like one whole negative charge at the center
of the sphere, the nucleus. This whole negative charge exactly cancels out the single nuclear
positive charge. Were our assumptions exactly correct, the external electron located at 20

A
away from the hydrogen atom would feel no net electrostatic forces from the hydrogen atom!
Example 2: The helium atom: We now consider the helium atom. In this atom, there
are two nuclear protons and two 1s electrons. We consider the forces exerted on the second
of these two 1s electrons by the nucleus and the rst 1s electron, see Figure 9.12. The forces
the rst 1s electron exerts on the second 1s electron depends on whether the rst 1s electron
is interior or exterior with respect to the second electron. If the rst electron is exterior,
then it provides no net electrostatic force; if it is interior, it applies a force similar to one it
would exert if it were located at the nucleus.
As the rst and second electrons are both in the 1s orbital, the rst 1s electron is interior
with respect to the second 1s electron exactly half the time. Thus the second 1s electron
experiences a nucleus which eectively is composed of two protons and half an electron, a
net eective nuclear charge, Z
eff
, of 1.5: Z
eff
= 1.5.
In other words, we could approximate the energy of a helium 1s orbital as being the same
as a single-electron atom where the eective nuclear charge is approximately 1.5. The energy
of the 1s orbital being proportional to Z
2
for single electron, we would expect the helium 1s
orbital energy to be 2
1
4
times the hydrogen 1s energy. More accurate calculations show for
116
Figure 9.12: Applying Gauss law to the helium atom.
helium Z
eff
= 1.69, while the energy of the helium 1s orbital is approximately double the
energy of the hydrogen 1s orbital.
Example 3: The lithium atom: The analysis here is much the same as in the previous
case. The lithium electron conguration is 1s
2
2s
1
. As Figure 4.12 shows, the 1s electron
distribution is almost entirely interior with respect to a 2s electron. From the perspective of
the 2s electron, the approximate eective nuclear charge is derived from three protons and
two 1s electrons: Z
eff
= 3 2 = 1.
We can contrast the energies of the outermost electron in the hydrogen atom to that in
the lithium atom. Both electrons feel an eective nuclear charge of one. But, in the lithium
case, the outermost electron is in a 2s orbital while for hydrogen it lies in a 1s orbital.
Recalling that the 2s orbital has lower kinetic energy than the 1s orbital and that the virial
theorem tells us that if the kinetic energy is lower, the total energy is higher, we conclude
that the lithium 2s orbital has a less negative total energy than the hydrogen 1s orbital.
Lithium is a less electronegative atom than hydrogen.
Quantitatively, we might estimate the energy of the lithium 2s orbital to be one-fourth
117
of the hydrogen 1s orbital, as in the hydrogen-like energy expression E
n
1/n
2
, where n
is the principal quantum number. The true ionization energy of the lithium atom is 40% of
the hydrogen atom, not 25%: Z
eff
for the lithium 2s orbital proves to be 1.28 and not 1.
Example 4: The carbon atom: One can not as easily apply Gauss law to other second
row main group elements. First, 2p orbitals are not spherically symmetric, and, second, it is
dicult to assess the degree to which a 2s electron is interior with respect to a 2p electron.
In this example we will make two rather crude approximations. We will treat the 2s and 2p
electrons as having the same radial distribution function and we will assume that 2p angular
distributions are spherical. With this set of crude assumptions we can estimate Z
eff
for a
carbon 2p electron. The carbon electron conguration is 1s
2
2s
2
2p
2
; the carbon nucleus
contains six protons. A carbon 2p electron therefore has two 1s electrons which are almost
completely interior to it and three n = 2 electrons with radial distribution functions similar
to it. We estimate Z
eff
= 6 2 3
1
2
= 2.5. Accurate calculations show the true Z
eff
for carbon is 3.14. The Z
eff
for carbon is signicantly larger than for lithium, while the
valence electrons principal quantum number is the same: carbon is therefore much more
electronegative than lithium.
9.3.3 The Pauling electronegativity scale
Applying Gauss law to the remainder of the second row elements we can derive the approx-
imate eective nuclear charges shown in Figure 9.13. As we go across the second row of
elements in the periodic table, Z
eff
gets bigger while the valence electron principal quantum
number remains the same. We conclude that elements on the right of the row are more
electronegative than elements on the left.
Pauling developed a scale in which the electronegativity of the second row elements
roughly resembles their approximate Z
eff
. He then developed a formula based on thermo-
118
chemical data which gives values similar to the Z
eff
values. In Figure 9.13, we list the
approximate Z
eff
of the elements, accurate Z
eff
, approximate Pauling electronegativities
and accurate Pauling electronegativities.
Figure 9.13: Comparison of approximate vs. exact Z
eff
and Pauling electronegativities.
9.4 Appendix: The electronegativities of the transi-
tion elements
The electronegativity of the transition elements places the heavier elements more electroneg-
ative than lighter elements in a column of the periodic table. Gold is the most electronegative
transition or noble metal. One possible rationalization for this trend lies in the fast speed
of the Au 1s electron. As we noted earlier, the Au 1s electron is moving at roughly 2/3 the
speed of light. At such speeds, relativity begins to be obeyed. One consequence of relativity
is that the Au 1s orbital contracts.
The contraction of the Au 1s orbital aects in turn the energy of the 2s orbital. The
119
2s orbital is the lowest energy s orbital which is orthogonal to the 1s orbital. As the 1s
orbital contracts, so does the 2s orbital. A cascade phenomenon develops: where each larger
s orbital contracts because all the s orbitals smaller than it have contracted. Ultimately, the
valence Au 6s orbital is contracted. As a result of this contraction, the Z
eff
of the Au 6s
orbital increases, and, hence, the electronegativity of the Au atom also increases.
9.5 Problems
1. Which of the following molecules are paramagnetic (a) HB, (b) HC, (c) N
2
, (d) O
2
,
and F
2
? What are the bond orders of these ve molecules?
2. Let us make a Chladni plate analog for the HF problem. We will represent both the H
and the F by square plates. Should the square plate representing F be the same size,
bigger, or smaller than the square plate representing the H atom? Now consider the
two plates welded together along one of their edges to represent the HF molecule.
Consider only the ground state 1s state for the plate representing the H atom, but the
one 1s and two 2p states for the plate representing the F atom. Recalling the third rule
of MO diagrams, that wavefunctions which are initially closer in energy interact more,
how should one size the plates so that the H 1s orbital interacts most strongly with one
of the F 2p orbitals and not the F 1s orbital? Using the square plate wavefunctions as
a starting point, show what the four lowest in energy states should be for the welded
plates. State their relative energies and draw contour maps for each of the four states.
Try to draw your diagram so that it resembles somewhat an MO diagram. In what
way are we following the rst rule of MO diagrams?
3. Examine the radial probability distributions for the hydrogen-like 1s, 2s and 2p states.
Now consider the hydrogen atom with the 1s
1
electronic conguration. What is the
120
relative order of Z
eff
for the 2s vs. 2p states. Is the 2s or 2p state lower in energy?
4. Examine the radial probability distributions for the hydrogen-like 1s, 2s and 2p states.
Now consider helium where the atom has 1s
2
electronic conguration. Estimate the
relative order of Z
eff
for the 2s vs. 2p states. Is the 2s or 2p state lower in energy?
How does one use the radial probability distributions to obtain the answer to this
question?
5. In our daily lives, we feel almost no electrostatic forces. And yet we are made up
of and surrounded by both positve and negative charges. How does Gauss law help
rationalize the absence of electrostatic attraction in our daily lives?
6. Based on the notions presented in this chapter, deduce whether elements get bigger or
smaller as we go from the left to the right of a given row of the periodic table.
7. Based on the notions presented in this chapter, deduce whether elements get bigger or
smaller as we go from the top to bottom of a given row of the main group periodic
table.
8. While we tend to think of electronegativity as an invariant elemental value, electroneg-
ativites of atoms do change as a function of the circumstances of the atom. Compare
Z
eff
for C
+
, C (neutral), and C

. What would be the order of Z


eff
for these three types
of carbon atoms? How would these changes aect the electronegativity of carbon?
9. Using approximate Z
eff
values and the hydrogen-like formula of atomic orbital ener-
gies, explain why the electronegativity of hydrogen is most similar to the boron atom
amongst all the second row elements of the periodic table.
10. Using approximate Z
eff
values and the hydrogen-like formula of atomic orbital energies,
explain why the electronegativity of main group elements goes down as one goes down
the column of the alkali metals.
121
11. Explain the basis for the mnemonic, the closer a main group element lies to uorine,
the more electronegative it is.
12. Explain the basis for the mnemonic, the closer a transition or noble metal lies to gold,
the more electronegative it is.
13. The ionization energies of the elements is shown in Figure 9.14. State two trends which
this data reveal. Rationalize these two trends.
Figure 9.14: The ionization energies of the elements
14. The electron anities of the elements is shown in Figure 9.15. In this gure, in contrast
to ionization potentials, halogens and not noble gases prove to have the highest values.
Rationalize this fact.
122
Figure 9.15: The electron anities of the rst three rows of the periodic table.
123
Chapter 10
Complex numbers, lasers, and
wavefunctions
In the rst nine chapters of this book, we learned a great deal of descriptive information about
atomic and molecular wave functions. We learned that electrons in atoms and molecules are
best described by respectively atomic and molecular orbitals. In both cases the orbitals are
waves. In the minimal valence basis set approximation, molecular orbitals which have the
longest wavelength are the lowest in energy. Leaving the minimal basis set approximation,
we nd atomic orbitals with the longest wavelength are the highest in energy. For both
types of orbitals, the square of the wavefunction (or more correctly the wavefunction times
its complex conjugate) gives the relative probability that an electron is located at the various
points in space.
The description hangs together, but does not answer basic questions. What is the physical
quantity which oscillates in orbital waves? What actual physical quantity is negative in
regions where the orbital is negative? How can an electron, which is a phyically concrete
object be a wave anyway? The goal of this and the next four chapters is to answer these
questions.
124
The answers require delving into quantum mechanics. Quantum theory being fairly ab-
struse, it helps to have some macroscopic quantum mechanical objects in hand. In the lab
associated with this class, you will synthesize one such object, a high temperature supercon-
ductor. In the classroom, the simplest example is the red laser pointer.
Light emitted by laser pointers are quantum mechanical coherent plane waves. LIke all
quantum mechanical systems, these light plane waves are well described by wavefunctions.
The good news is that dierent plane wave wavefunctions are much more closely related
to one another than atomic electron wavefunctions. The bad news is that they are best
described as complex number functions (ie., complex functions).
10.1 Complex wavefunctions
You may have been taught about complex numbers in high school. And they may have seen
a bit pointless: after all, no physical quantity which we can measure is anything but a real
number. Energy, momentum, mass, charge, temperature and position are all based on real
numbers. Here in college, complex numbers will nonetheless play a useful role. Although
in quantum theory no measurement ever results in a complex number, the use of complex
numbers will make the measurements much easier to understand. Wavefunctions will turn
out to be best described as complex quantities.
To make an analogy, think of words as real things. In this analogy, words, like the
word elephant represent real entities, just as measured values represent real things. Letters
are not words, but they are useful in the understanding of words. (If you are into Chinese,
characters are words; strokes are not words, but strokes are useful in the making of characters,
which are words.) Wavefunctions, in this analogy, are like letters (or strokes), They are not
measured quantities themselves, but just as letters or sttrokes are an integral part of words,
wavefunctions are an integral part of the measurement process. In quantum mechanics,
125
wavefunctions are sometimes complex; measurements are always real.
10.1.1 Complex numbers
126
127
128
We conclude this section with the observation that a complex number times its complex
conjugate is real and positive. Therefore, it can be used to express intensity or probabiity.
10.2 Waves
10.2.1 Real waves
With this introduction to complex numbers, we turn to the red light laser pointer. As the
students in this class know, light is a wave. The simplest waves which we know of are the sine
and cosine waves. In Figure 10.1 we illustrate the two 1-D cos and sin functions, Acos(kz)
and Asin(kz). The 1-D coordinate has been chosen to be z. A is the amplitude of the waves.
As the gure shows, when z = , where is the wavelength, the cosine wave has cycled back
to its initial value. In other words, cos(k) = cos(2) = cos(0); k = 2; and therefore
k = 2/.
Are the waves given above the correct waves to be used to describe the quantum mechan-
ical wavefunctions of red laser light? The following experiment answers this question. Hold
a wall in the path of laser light. The screen shows an intense red dot. The intensity of this
dot corresponds to the proabability light has hit the wall. Let us suppose that red laser light
had a wavefunction which obeyed the law
1
(z) = Acos(kz). Then,
2
1
(z) would describe
the intensity of the red dot visible on the wall. We show this function in Figure 10.2. In this
same gure, we show the intensity function for
2
(z) = Asin(kz).
The function psi
1
= Acos(kz) or psi
2
= sin(kz), would lead, once squared, to a ickering
intensity. At some positions, intensity would vanish. As we shall see in class and indeed as
has been conrmed in countless experioments, the intensity of laser light does not nor never
has ickered. We conclude that neither the sin nor the cos function can by themselves be a
correct description for the wavefunction of laser light.
129
Figure 10.1: The Acos(kz) and Asin(kz) functions.
Figure 10.2:
1

2
1
= A
2
cos
2
(kz) and
2

2
2
= A
2
sin
2
(kz).
130
10.2.2 Complex waves
We now consider a wave function based on complex numbers. We consider the wave
= A[cos(kz) + isin(kz)],
where A is a positive real number. We are interested in the intensities which correspond to
this wavefunction.
We recall a point which we made iin the rst lecture in this course. For complex wavefunc-
tions,

. WE therefore consider

for the above stipulated complex wavefunction.

= A[cos(kz) isin(kz)]A[cos(kz) + isin(kz)] = A


2
[cos
2
(kz) + sin
2
(kz)] = A
2
. We plot this function in Figure 10.3.
As the above equation, and as the above gure show, the complex wavefunction is con-
stant in space: it does not icker. As laser light does not icker as we move the screen, we
conclude that the complex wavefunction is an appropriate wavefunction for laser light!
Figure 10.3:

for = A[cos(kz) + isin(kz).


131
Chapter 11
What is k?: The de Broglie relation
In the last chapter, we discussed how the electrons coming out of an electron gun (as modeled
by the light emitted by a red laser pointer) are in the wave function, (z) = Ae
ikz
. In this
chapter, we consider two dierent cases where this wave function is relevant. The rst is
the plane wave, where the electron (modeled by the red light) travels in a uniform front, the
second is a radiating wave, where the the wave propagates in concentric spheres away from
an intial point. These two dierent scenarios are shown in Figure 11.1. We will discuss the
wavefunctions applicable to both cases and show the connection with wave functions which
more exactly describe these states and the wave function (z) = Ae
ikz
. We then use the
more exact description of the plane-wave wavefunction to deduce a physical meaning for the
constant k, show its connection with momentum and ultimately understand the de Broglie
relation, a relation which shows that momentum is inversely proportional to wavelength.
11.1 Vectors and dot products
To carry out the program listed above, we need to make sure of our mathematical knowledge
of vectors and their dot products. As the student may already know, a vector can be thought
133
Figure 11.1: The plane and spherically radiating wave functions. Wave crests and troughs
indicated with respectively solid and dotted lines.
of as an arrow pointing in a direction. Figure 11.2 indicates such an arrow. The arrow has
both a direction and a length. In the case of arrows (or vectors) which lie in a plane, we may
express the arrow as a pair of numbers. Figure 11.2 shows how we do so: we consider the
tip of the arrow as a coordinate in space. We place the tail of the arrow at the origin and
we draw in the coordinate axes. The so-dened planar coordinates completely specify the
arrow in question. Adopting this approach, the vector shown in Figure 11.2 corresponds to
the pair of numbers 1 and 2. These numbers can be expressed in either a row or a column:
(1, 2) or

1
2

.
Vectors in planes are designated with pairs of numbers; in 3-D they are triplets, see Fig-
ure 11.2. In any case, vectors are not numbers; they are, instead, short sequences of numbers.
Important to this chapter is how we multiply such sequences. While in three dimensions
there are two types of vector multiplication, the dot product and the cross product, in other
dimensions, vector multiplication is either the dot product or a variant of the dot product.
In Figure 11.3 we show two 2-D vectors (1, 2) and (3, 1). The dot product of these two
vectors is (1 3) +(2 1) = 1. Dot products in 3-D (or even higher dimensions!) follow a
similar law. In 3-D for example, (4, 1, 2) (5, 3, 1) = (4 5) +(1 3) +(2 1) = 19.
134
Figure 11.2: The planar vector (1,2) and the 3-D vector (1,2,-4).
An important relationship between dot products and the lengths of and angle between
multiplied vectors always holds true. If v and w are vectors with respective lengths l
v
and
l
w
with an angle between them, the dot product, v w, obeys the law:
v w = l
v
l
w
cos.
Figure 11.3: The dot product. The vectors (1, 2) and (3, 1) are shown. They have respective
lengths of

5 and

10. (1, 2) (3, 1) = 1 = 5

2cos, where is the indicated angle.


One simple application of this equation is the connection between the dot product and
the length of a vector. Note that the angle between a vector and itself is by dention zero
135
degrees, ie., as is self-evident, a vector points with a deviation of zero degrees from its own
orientation. As cos0 = 1, the above equation when applied to a vectors dot product with
itself reduces to:
v v = l
v
l
v
cos0 = l
2
v
.
The dot product of a vector with itself is always the square of its length.
Physical scientists often simplify the notation used to describe the length of a vector.
Rather than calling the length of v to be l
v
, they say the length of v is v. Thus the v with
an arrow on top of it represents a vector, v without the arrow on top of it represents a single
number, the length of the vector in question:
v v = v
2
.
A second relevant application of the dot product formula is shown in Figure 11.4. In this
gure we consider a series of vectors r
i
. The tips of the vectors r
i
, form a straight line. In
this gure we show also a vector

k. The vector

k points in a direction perpendicular to the
straight line formed by the locus of arrow tips of the vectors r
i
.
Figure 11.4: Shown in this gure are a set of vectors whose tips form a straight line per-
pendicular to the direction of

k. The gure shows that r
i
cos
i
is a constant, designated
r

.
136
Placed on this gure are several right triangles. The hypotenuses of these triangles are of
the varying lengths r
i
. In every right triangle, one of the two stems of the triangle is parallel
to the direction of

k. Placed on the gure is the length of this stem, r
i
cos
i
, where
i
is the
angle between r
i
and

k. As this gure shows r
i
cos
i
is a constant for all r
i
. We call this
constant r

: it is the length of the stem of the right triangles which points in the

k direction.
In other words, for all r, whose tip lies on a straight line perpendicular to the direction of
the vector

k, the value of rcos = r

, a constant.
We now apply these ideas to the dot product between

k and all vectors r whose tips lie
in the same straight line perpendicular to

k.

k r = krcos = kr

.
This relationship will be important in the next section.
11.2 The plane wave
We now apply the idea of vectors and dot products to red laser light. Within the beam,
red laser light advances uniformly across a wavefront, see the left panel of Figure 11.1. We
call the direction that this wavefront advances the

k direction. In addition to describing the
direction in which the laser beam travels, can we also apply

k to describe the plane-wave
wavefunction itself?
To answer this question we must consider two separate steps. We consider rst only
those spatial points which lie directly in the direction

k. From the last chapter we know the
wavefunction conned to a straight line is = Ae
ikr
. We deduce that for the spatial points
lying directly in the direction

k, = Ae
ikr
= Ae
ikr

. The designation r

signies that only


those spatial points directly in the direction

k are considered.
We now turn to the full plane-wave, see Figure 11.5. Our interest is in deducing its
137
wavefunction. Applying ideas from the previous section, we note that for all r whose tips lie
in the same line perpendicular to

k,

k r = krcos = kr

. The value of

k r is a constant
across the whole perpendicular line: the dot product of

k and r is equal to the dot product
of

k and r

, where r

refers to a spatial vector which points in exactly the same direction as

k.
We now combine the expression, = Ae
ikr

(derived above) with the expression



k,

k r =
kr

. Substituting the latter expression into the former, we deduce = Ae


i

kr
. This function
describes the wave for all points in space, not just the points which lie along the

k direction:
it is the wavefunction of the wave illustrated in Figure 11.5.
Figure 11.5: The wavefunction = e
i

kr
. The central path of the planewave is indicated.
11.2.1 e
i

kr
vs. e
ikz
The student may have noted a similarity between e
i

kr
and e
ikz
. While both expressions are
exponential functions, they are respectively 3-D and 1-D in nature. The former wavefunction
is a function in space, the latter is conned to a line. We may however restrict our interest
in the spatial wavefunction e
i

kr
to its behavior along the central path of its motion. By
doing so, we have in eect, reduced the originally 3-D wave function into a 1-D one. Along
this central path the wave function e
i

kr
reduces to the 1-D expression e
ikr

. We will in
138
this course, so as to emphasize the 1-D nature of thte latter expression, substitue the 1-D
coordinate z. In such a case e
ikr

= e
ikz
.
11.3 The radiating wave
Shown in the right panel of Figure 11.1 is a second type of wave, one where the wavefunction
propagates away from a central point. It is the type of wave we associate with the ripples
caused by rain drops on a pond. In such a wave, every point which lies at the same distance
from the center of the wave has the same value. Such a wave can therefore be descibed
as e
ikr
, where r is the distance between a point in space and the center of the wave, see
Figure 11.6.
Figure 11.6: The radiating wave = e
ikr
.
The expression e
ikr
resembles the expression e
ikz
which we had previously used to de-
scribe one-dimensional wave motion. The dierence between the two is that while the latter
expression has as an input the 1-D spatial coordinate, z, the former expression uses the
distance between a spatial wave and the center of the wave, r. Thus the latter wave is a 1-D
wave, while the former wave is fully 3-D.
As Figure 11.6 shows, if we conned our interest in the radiating wave to the wavefunction
on a line segment radiating directly outward from the center of the wave, and if, in addition,
we were to call the length travelled on this line segment z, then e
ikz
would correctly describe
139
the wavefunction along the 1-D line segment.
11.4 Wave amplitudes
For the sake of completeness, we should consider the amplitude of both the plane wave and
the radiating wave. Laser beams being plane waves, do not lose intensity as they travel. But
radiating waves encompass a larger and larger surface as they radiate out. Recalling that
(1) a radiating wave propagates in a spherical manner and that the surface of a sphere is
4r
2
and (2) intensity, is proportional to
2
, we conclude:

plane
= Ae
i

kr
but that

radiating
=
A
r
e
ikr
Please note, we will talk of wave intensities more later in this course. But for the next few
chapters, our primary focus will be on the exponential portion of the wavefunctions. For this
reason, we will drop the amplitude protion of the wavefunctions for the next few chapters
except when they really needed to be included,
11.5 The de Broglie equation
We have developed two dierent 3-D wavefunctions: the planewave, e
i

kr
, and the radiating
wave, e
ikr
. The remainder of this chapter focuses on the rst of these wavefunctions, the
planewave; in the following chapter we consider again the radiating wave.
We know the direction of travel in a planewave (r) = e
i

kr
is determined by the directtion
of the vector

k. Along its central path, this wavefunction reduces to (r

) = e
ikr

= e
ikz
.
140
From the previous chapter, we know that the k used in this equation obeys the relation
k = 2/, where is the planewave wavelength.
11.5.1 What is

k?
The planewave wavefunction describes not just light waves, but also particle beams, such as
particle beams of electrons. We therefore must be able to equate the terms

k, r, and r

with
physical quantities known to play a role with particles such as electrons. In a particle sense,
the meaning of both r and r

are self-evident, the former are points in space and the latter
are the points in space which lie at the center of the particle beam path. But what is

k?
The term

k must be related to a known physical quantitity. Physical quantities used to
describe a particle are the energy, kinetic energy, potential energy, position, mass, charge,
temperature, velocity, acceleration, momentum, angular momentum and torque of the par-
ticle. We further know that

k points in the direction which the particle is travelling. By a
comparison of the two lists,

k is related to either the velocity or the momentum, ie., either

k v or

k p, where v and p are respectively the velocity and the momentum.
We need not decide between these two proportionality relationships. As p = mv, the
correctness of the one proportionality relation implies the correctness of the other. However,
it turns out that

k is more closely related to p than v. We express the relation p

k typically
not as a proportionality relation but as an equation p = C

k. The proportionality constant


is designated as h, ie., p = h

k. h/2 = h, where h is Plancks constant.


11.5.2 The relation between p, and
The vector relation p = h

k implies that the length of p and the length of



k are similarly
related, p = hk. We already know the relation between k and the wavelength , k = 2/.
Combining the two equalities we nd p = hk = h2/ = h/. This last equation is the
famous de Broglie equation which relates the magnitude of the momentum to the wavelength
141
of the particle. This relation is one of the cornerstones of quantum theory and is a relation
which we will frequently invoke in this course.
142
Chapter 12
The Heisenberg Uncertainty Principle
Knowing that plane waves can be described by the wavefunction
plane
= A
1
e
i

kr
while
radiating waves can be described by
radiating
= A
2
e
ikr
/r can be a powerful aid in the
understanding some of the intrinsic laws which govern the universe. I hope to show you one
such connection today.
We begin this chapter by considering a plane wave which passes through a small hole in
a screen before arriving at a far wall. There is a clear Youtube demonstration of this set up
for water waves, at http://www.youtube.com/watch?v=4EDr2YY9IyA. We illustrate this
situation in Figure 12.1. As this gure shows, while initially the wave is a plane wave, after
passing through the small hole, the wave becomes a radiating wave.
Let us analyze this situation with the equations developed in the last chapter. In Fig-
ure 12.2, we follow a plane wave as it leaves the initial source, S, passes through a hole H,
and nally arrives at two dierent points, B and C. In the rst part of its journey, as the
light travels from S to H, the wave is a planewave and there is no diminuation in the light
intensity. In the second part of its journey, as the light travels in a radiating wave from H
to either B or C, the intensity of light decreases as the distance travelled increases.
143
Figure 12.1: A plane wave which passes through a hole before reaching a wall.
We are interested in calculating the intensity of light as it travels along the above paths.
We ignore for now the journey from S to H, as the the light intensity is constant throughout
this part of the journey. Our focus turns to the latter part of the light path, where the light
is in a radiating wavefunction.
Recalling that intensity, or probability distribution, is proportional to

, we note:

radiating

radiating
= A
2
e
ikr
/r A
2
e
ikr
/r = A
2
2
/r
2
,
where r is the distance travelled. The relative intensities of light observed at points B and
C are therefore respectively A
2
2
/r
2
HB
and A
2
2
/r
2
HC
. C, lying signicantly further away from
the hole in the screen than B (ie., r
HC
> r
HB
), has a smaller intensity of light associated
to it. In class we hope to give a demonstration of this with a hand held laser pointer. We
illustrate this result in Figure 12.3, where we use a qualitative graph to indicate the relative
intensity of the light received on the nal wall.
144
Figure 12.2: A plane wave which passes through a hole before reaching a wall.
Figure 12.3: A

k-plane wave which passes through a hole before reaching a wall. The
intensity of the light on the far wall is illustrated with a qualitative graph.
145
12.0.3 Probability Amplitudes
In Figure 12.4, we illustrate this same experiment using three dierent notations. The
middle panel introduces the term probability amplitude. The values of a wavefunction are
the probability amplitudes that a given particle in a given state can be found at dierent
points in space. All wavefunctions are probability amplitudes (but as we will learn later in
this course, not all probability amplitudes are wavefunctions). We will give the basic laws
of probability amplitudes in this and the next chapter.
And nally the last panel refers to the bra and ket notation. You will learn more about
this notation later. For now you can think of this third panel as a heads-up. Our focus
now is the connection between the top two panels of this gure, the connection between
wavefunctions and probability amplitudes.
In particular, the probability amplitude for a photon of light to travel from the source S,
to the hole H is PA
SH
. In the same way PA
HB
refers to the probability amplitude of a photon
of light, having reached the point H, further travelling to the point B. As wavefunctions are
a type of probability amplitude,
PA
SH
=
plane
(r
SH
) = A
1
e
i

kr
SH
and
PA
HB
=
radiating
(r
HB
) =
A
2
e
ikr
HB
r
HB
.
As wavefunctions are probability amplitudes, and as wavefunctions times their complex
conjugates are probabilities, we conclude that probabilities, P, follow the law
P = PA

PA,
the same law which wavefunctions follow. The above is one of the fundamental laws of
146
Figure 12.4: Dierent physical notation used to describe a

k-plane wave which leaves a point
S, passes through a hole, H before reaching a wall at points B or C. The third panel
introduces the reader to the bra and ket notation. As the bra and ket notation is the one
most widely used in quantum theory, you will need to learn how this notation works. We
will introduce this notation slowly over the next two weeks, especially becauuse this is the
notation used in Feynman Volume 3. For now note as one example of this notation that
< H | = A
1
e
i

kr
> denotes the probability amplitude a particle which is in a plane wave
travelling in the

k direction starting from the point S can reach the point H.
147
quantum theory.
The probability amplitude that light travels from S to H is A
1
e
i

kr
SH
and therefore
P
SH
= PA

SH
PA
SH
= A
1
e
i

kr
SH
A
1
e
i

kr
SH
= A
2
1
.
Similarly the probability amplitude that light travels from H to B is r A
2
e
ikr
/r and
therefore
P
HB
= PA

HB
PA
HB
= A
2
e
ikr
HB
A
2
e
ikr
HB
=
A
2
2
r
HB
.
We further note that the probability amplitude that the probability amplitude that a
particle starts at a point S and reaches the point B is the product of the two component
probability amplitudes:
PA
SB
= PA
SH
PA
HB
and therefore the probability the light travels form S to H, P
SB
, follows the equation
P
SB
= PA

SB
PA
SB
= PA

SH
PA

HB
PA
SH
PA
HB
= A
2
1
A
2
2
r
HB
.
The basic law is that the probability amplitude of a process composed of two sequential events
equals the probability amplitude of the two sequential events multiplied by each other.
12.1 The Heisenberg Uncertainty Principle
One might think we have said everything that could be said about the experiment of light
travelling through a hole before reaching a wall. But it turns out there is much more to
be said. First, it turns out that all particles behave in exactly the same manner. Most
importantly for us, an electron beam, passing through a small hole spreads out exactly as
the laser beam does.
148
But even more interesting was Heisenbergs exposition on this and related experiments.
Heisenberg noted two important points. First, he recognized that passing light or electrons
through a hole is a measurement of where the light or electron is located. And that measuring
the light or electron position changed the state of the particle. While initially the light or
electron was in a plane-wave, after the position measurement, the light or electron had turned
into a radiating wave. Measurements change the state of the system.
Second, he recognized that in the plane wave state, the momentum of the particle was
exactly known. But once position was exactly measured, knowledge of the momentuum was
entirely lost. We hope to illustrate this with some in-class experiments. We hope to nd
results like those illustrated in Figure 12.5.
As the width of the hole increases, one simultaneously learns less and less about the
position and correspondingly retain more and more knowledge of the nal momentum. This
result can be quantied:
xp
h
2
.
The above is a statement of the Heisenberg Uncertainty Principle. In fact the incorporation
of Plancks constant is entirely due to the relation p = h

k.
149
Figure 12.5: A

k-plane wave which passes through holes of dierent width before reaching a
wall. The intensity of the light on the far wall is illustrated as qualitative graphs. The wider
the hole, the less spread is observed on the far wall.
150
12.2 Parallel probability amplitudes
We end this chapter with a brief look at the two hole (or two slit) experiment. In the next
chapter we will handle the actual math associated with this sort of experiment. For this
chapter, we wish to illustrate that there is something very interesting in this experiment.
And as we shall see in the next chappter , this interesting thing is very well managed with
probability amplitudes.
The question we look at for now is what the intensity of light looks like if laser light
passes through two side-by-side holes rather than a single hole.As illustrated in the next three
gures, we might expect that the side-by-side experiment wouuld look like a superposition
of two dierent single hole experiments. The reality of the situation is quite dierent. When
light (or electrons) pass through two holes the cumulative result is not the sum of what
would happen if light passed through either one hole or the other hole separately.
This result violates our intuition about probabilities.
Figure 12.6: What do we expect to happen if laser light passes through two side-by-side
holes?
151
Figure 12.7: What we might expect to happen if laser light passes through two side-by-side
holes.
152
Figure 12.8: What actually happens when laser light passes through two side-by-side holes.
153
Chapter 13
Probability amplitudes
The experiments described in the last chapter suggest that we need a new way of guring
out probabilities. This new way is called quantum mechanics, QM. In this course, we will
introduce some of the key ideas of quantum theory, but not all of them. (we will not, for
example, introduce concepts related to time.) We begin our exposition here.
1. In quantum theory one separates the system from the observer. Light, electrons,
electrons and protons in an atom, all three are examples of systems. Observers make
measurements on these systems. (Observers are the people making the measurements.)
Observers measure the energy, or the position, or the momentum of the light, the
electron or the atom. Observers measure the systems.
Situations which can not be so divided are outside the purview of ordinary non-
relativistic quantum theory. (The paradox of Schroedingers cat, which you can read
up on in Wikipedia if you are interested, is an example of a situation which can not
be so divided.) The good news is that, for physical scientists, every non-relativistic
situation which needs to be understood, is describable by quantum theory.
2. Systems are in states; states are describable by wavefunctions, . In the bra or ket
154
notation, states can be represented as | > or < |. For laser light = A
1
e
i

kr
; for
a radiating wave = A
2
e
ikr
/r; and for an electron in the 1s orbital, = N
1s
e
Cr
.
The wavefunctions, , are not probabilities, they are probability amplitudes. (Wave-
functions, being neither always real nor always positive, could never be probabilities.)
Wavefunctions are position probability amplitudes, ie., the probability amplitude, PA,
an electron or a photon is located at each and any given position.
3. Here are how probability amplitudes work. Consider light is a planewave. This light
travels to a hole in a screen, see the top panel of Figure 13.1. Let source be located at
(0, 3, 0) and the hole be located at the point (0, 3, 2). If the plane wave was in the
state = A
1
e
i

kr
, the probability amplitude that the photon reached the hole in the
screen is = A
1
e
i

k(0,3,2)
and the relative probability that the light has reached this
same point is =

= A
2
1
.
4. If we now ask ourselves what wavefunction the photon is in after it passes through
the slit, we nd the wavefunction has changed into the state A
2
r
1
e
ikr
, where r is
now the distance a point is from the hole located at (0, 3, 2). To nd the probability
amplitude that the photon reaches a point on a wall, W, where W = (0, 0, 6) we need
to calculate the distance between (0, 0, 6) and (0, 3, 2), which is 5. We therefore nd
the probability amplitude that the photon travels from the hole at (0, 3, 2) to the
point r = (0, 0, 6) on the wall is = A
2
5
1
e
i5k
.
The probability amplitude that the laser light has travelled from the source all the way
to the point on the far wall is the product of the two separate probability amplitudes,
ie., PA
SW
= PA
SH
PA
HW
, where S, H, and W refer to respectively the source, the
hole, and the wall. From the above we deduce PA
SW
= A
1
e
i

k(0,3,2)
A
2
5
1
e
i5k
.
5. Probabilities in ordinary, non-quantum mechanical situations follow two simple laws.
155
Figure 13.1: Single pin-hole and double pin-hole experiments.
156
The rst is the multiplicative law: if the probablity of wearing a pair of shoes with
no holes in them is 2/3, and the probability of wearing socks with no holes in them
is 1/5, the probability of wearing both shoes and socks with no holes in them is the
multiplication of the two probabilities: 2/15. The second law is an additive law. If the
probability of rolling a 6 with a die is 1/6 and the probability of rolling a 5 is 1/6, the
probability of rolling a 5 or a 6 is the sum of the two separate probabilites, 1/3.
In quantum theory, probabilites do not obey these two laws: probability amplitudes,
however, do. Consider the panel on the right of Figure 13.1. The probability amplitude
that the photon makes it through the hole and then travels to the point A is the product
of the two separate PAs: PA = A
1
e
i

k(0,3,2)
5
1
A
2
e
i5k
.
6. In the last paragraph,we gave an example of the multiplicative law of probability
amplitudes. Now we consider an example of the additivity law. In Figure 13.1, we
show a similar set-up to that given in the previous section, but where a second hole
located at (0, 2, 6) has been added to the rst hole. For each separate pathway,we
can calculate a separate PA. The PA for the top pathway is the same as before:
PA
top
= A
1
e
i

k(0,3,2)
5
3
A
2
e
i5k
. The PA for the bottom pathway is PA
bottom
=
A
1
e
i

k(0,2,6)
2
1
A
2
e
i2k
.
The PA
comb
, the PAthe photon travelled either through the top or the bottom pathway
before nally reaching the point W is: PA
comb
= PA
top
+ PA
bottom
. The relative
probability,P, is the probability amplitude times its complex conjugate.
P
comb
= PA

comb
PA
comb
= (PA
top
+PA
bottom
)

(PA
top
+PA
bottom
)
P
comb
= PA

top
PA
top
+PA

top
PA
bottom
+PA

bottom
PA
top
+PA

bottom
PA
bottom
7. Lastly, and most importantly, probabilities can be calculated from probability ampli-
157
tudes. A probability is a probability amplitude times its complex conjugate. Given the
math of complex conjugates, in QM, probabilities are always real and non-negative.
13.1 The double slit experiment
In this section we work out the consequences of these principles. We will slightly simplify
our math. Both radiating and planewave wavefunctions are of the form f(r)e
ikr
, the rst
part gives the amplitude, the second the phase of the various wavefunctions. It turns out the
most interesting things happening in the double pin-hole experiment have only to do with
the phase part of the wave functions.
For the remainder of this chapter we will herefore ignore chnages in ampliude amplitude
and concentrate only on the phase part of the wavefunctions. As we saw in an earlier chapter,
and as we illustrate in Figure 13.2, if we ignore the amplitude component, then both plane
wave and radiating waves just depend on the distance travelled. We shall call these distances
r
1
, r
2
, ...
13.1.1 Two single hole experiments
Before moving onto the double slit experiment, we revisit the single hole experiment. As
discussed above, we consider only the distance travelled, see Figure 13.3. In the rst ex-
periment, a photon of red light starts at a source, passes through the pin-hole on the top
and nally arrives at a wall. As this gure shows, the light travels from crest to trough to
crest as it progresses on this journey. And the evolution of the phase component of the wave
function depends only on the distance travelled. The distance travelled from the source to
the pin-hole being r
1
and the distance travelled from the pin-hole to the wall being r
2
, the
total distance travelled is r
1
+ r
2
.
158
Figure 13.2: Phase portion of the plane and radiating waves depends only on the distance
travelled. The distance travelled to points in both the upper and lower panels is r
1
.
159
Figure 13.3: Two single hole experiments.
160
The nal phase of the wave is determined by this distance travelled. The nal phase is
therefore e
ik(r
1
+r
2
)
. Expressed in probability amplitudes:
PA
top
= PA
SH
T
PA
H
T
W
= e
ikr
1
e
ikr
2
= e
ik(r
1
+r
2
)
,
where S, H
T
, and W refer to respectively the source, top hole, and a point on the wall.
In exactly the same way we can consider the second single-hole experiment, where the only
open hole is the bottom hole.
PA
bottom
= PA
SH
B
PA
H
B
W
= e
ikr
3
e
ikr
4
= e
ik(r
3
+r
4
)
,
and where H
B
refers to the bottom hole.
P
top
= e
ik(r
1
+r
2
)
e
ik(r
1
+r
2
)
= 1,
and similarly,
P
bottom
= e
ik(r
3
+r
4
)
e
ik(r
3
+r
4
)
= 1.
The combined intensity for the top pls the bottom path is 1 + 1 = 2.
13.1.2 The double hole experiemnt
Now consider Figure 13.4. In this experiment, there are two side-by-side pin-holes which the
red photon can travel through. (As we shall later see, it is important we do not measure
which pin-hole is the actual pin-hole traversed by the photon.) Referring to the distances
r
1
, r
2
, r
3
, and r
4
in this gure, the probability amplitude associated with travelling through
the top and bottom paths are the same as in the previous section: they are respectively
(PA)
top
= e
ik(r
1
+r
2
)
and (PA)
bottom
= e
ik(r
3
+r
4
)
.
161
Figure 13.4: A double pin-hole and wall experiment.
In such a case probability amplitudes are summed: (PA)
comb
= (PA)
top
+ (PA)
bottom
=
e
ik(r
1
+r
2
)
+ e
ik(r
3
+r
4
)
. The intensity of light seen at the wall equals
P
comb
= (PA)

comb
(PA)
comb
.In this case the net intensity would be
P
comb
= PA

comb
PA
comb
= (PA
top
+ PA
bottom
)

(PA
top
+ PA
bottom
)
P
comb
= (e
ik(r
1
+r
2
)
+ e
ik(r
3
+r
4
)
)(e
ik(r
1
+r
2
)
+ e
ik(r
3
+r
4
)
)
13.1.3 Diraction
We now consider the special cases where the top and bottom path lengths exactly dier by
/2. Lets say, for the sake of discussion, that the top path length is r
T
, ie., r
T
= r
1
+ r
2
.
162
Then the bottom path length would be r
B
= r
T
+/2.
P
comb
= (e
ikr
T
+ e
ik(r
T
+/2)
)(e
ikr
T
+ e
ik(r
T
+/2)
)
= e
ikr
T
(1 + e
ik/2
)e
ikr
T
(1 + e
ik/2
)
= (1 1)(1 1) = 0.
We can just as well consider the case where the top and bottom path lengths dier by ,
ie., r
B
= r
T
+/2. In this later case
P
comb
= (e
ikr
T
+ e
ik(r
T
+)
)(e
ikr
T
+ e
ik(r
T
+)
)
= e
ikr
T
(1 + e
ik
)e
ikr
T
(1 + e
ik
)
= (1 + 1)(1 + 1) = 4.
In just the same way we could consider two path lengths which dier by 3/2 or 2. The
intensity for the former case would be zero and for the latter case it would be 4.
In Figure 13.5 we consider a series of points on the wall. As we progress from W
0
to W
1
to W
2
etc.. the top path length grows shorter while the bottom path length grows longer.
Thus we will pass from points where the dierence in path lengths progresses from 0 to /2
to and so forth. The overall intensity would therefore oscillate between 4 and 0. This is
the oscillating image which we discussed at the snd of the predeing chapter. This oscillating
image is termed diraction. Diraction is observed whenever we have pairs of waves, each
with a diering path length.
We will show a Youtube demonstration of diraction.
163
Figure 13.5: Diraction as the result of the double hole experiment.
Figure 13.6: The double hole experiment for an electron beam where the path chosen by the
electron is detected by a small magnet does not show diraction.
164
13.2 Diraction, electrons, and measurement
All waves exhibit diraction. But quantum systems have a special character all Their own.
To understttand tthe special character of quantum systems we nconsider the same set of
experiments as the preceding experiments but with an electron beam rather than a photon
beam. Assume the electrons travel throughout with an initial de Broglie wavelength of 700
nm just as before.
In such a case the electron is initially described by the same wavefunction as before.
We associate a probability amplitude with the top path, (PA)
top
and the bottom path,
(PA)
bottom
. As the set-up is the same, (PA)
comb
= (PA)
top
+ (PA)
bottom
and P
comb
=
(PA)

comb
(PA)
comb
, where P
comb
is the nal measured intensity.
This is a remarkable statement. Consider the diraction experiment with electrons. Lets
slow down the nuumber of electrons travelling in our electron beam until just one electron is
travelling throuugh the experimental set-up at a time. Our equations say diraction occurs.
But doesnt this mean the electron travels through both holes at the same time? How can
an electron pass throuugh two holes at the same time?
13.2.1 Something really amazing
And if that was not incredible enough , here is something that almost dees belief. Electrons,
as they move, create a small magnetic eld. In principle, this magnetic eld can be measured.
Imagine in the last experiment, we measure the magnetic eld near one of the pin-holes. We
illustrate his situation in Figure 13.6 In this experiment we therefore determine which hole
the electron passed through. In such a case, no matter how small we make the measuring
magnet, the nal probability distribution changes. The nal probability is no longer
P
comb
= PA

comb
PA
comb
= (PA
top
+PA
bottom
)

(PA
top
+PA
bottom
).
165
Instead,
P
comb
= PA

top
PA
top
+PA

bottom
PA
bottom
= 1 + 1 = 2.
This result is illustrated in Figure 13.6. Once, the electron path is determined, electron
diraction entirely disappears.
Its mind-boggling. No matter how small one makes the measuring device, the resultant
observed electron intensity changes. Just the act of measurement, by its very nature, changes
everything.
The Feynman lectures provided on Blackboard iin the Course Document section makes
for an interesting reading on this subject. I advise reading this Feynman text. But if you
do read it, please note this text uses the bra and ket notation.
166
Chapter 14
What are orbitals?
In the rst nine chapters we considered atomic and molecular orbitals. In the last four chap-
ters, we treated plane waves detected by a wall or a screen. In this chapter we apply our new
understanding of detected plane waves to understand what atomic and molecular orbitals re-
ally are. We shall discuss dierent ways of thinking of orbitals. First we will extend our ideas
based on the double pin-hole experiment. Second, we will discuss the realtionship between
orbitals and measuurement. Orbitals are wavefunctions and wavefunctions are probability
amplitudes for position measurement. At the end of this chapter we will contrast posiiton
measurement with momentum or energy measurement.
14.1 Dierent ways of reaching a wall
In the last chapter we studied a laser beam travelling through a one and two-hole screen on
its way to the wall. Now let us imagine a very small wall, W. Looking at the rst panel in
Figure 14.1, we nd the probability amplitude the photon travels from the source, S, to the
wall, W, is
PA
SW
= PA
1
PA
2
+PA
3
PA
4
.
167
Figure 14.1: Dierent ways a photon can traverse screens with holes to reach a nal small
wall.
168
In the second panel we introduce one more hole. In this case,
PA
SW
= PA
1
PA
2
+PA
3
PA
4
+PA
5
PA
6
.
For the third panel we nd:
PA
SW
= PA
1
PA
2
+PA
3
PA
4
+PA
5
PA
6
+PA
P
A
8
+PA
9
PA
10
+PA
11
PA
12
.
But we could as well envision two screens. The fourth picture in Figure 14.1 has
PA
SW
= PA
1
(PA
2
PA
7
+PA
5
PA
8
) +PA
2
(PA
4
PA
8
+PA
6
PA
7
).
And we could envision even more screens with even more holes. The fth panel illustrates
such a scenario, where we omit drawing the screens. And nally, in the last panel we
consider what wouuld happen if there were more than one source. The photon could be
starting anywhere and it could be travelling through any path.
We are constructing a very general situation. By this method, there is a probability am-
plitude associated with the photon starting anywhere, going through any path, before nally
arriving at the wall W. And furthermore we can move the wall itself can change position.
The wall can be located at any posititon r. For every such position there is a probability
amplitude which is the sum of all possible probability amplitudes for all the possible paths.
Lets call this probability amplitude, PA

(r). We can now dene a wavefunction:


(r) = PA

(r).
169
Figure 14.2: A 2p
z
orbital and a small detection device. The small detection device can be
thought of as a wall.
14.2 Orbitals
In this section we consider electrons. These electrons pass through screens before reaching
a wall. Imagine now we have an electron not in a plane wave, but in an orbital. The small
wall detects the electron only when the electron is located exactly where the innitely small
wall is placed. The electron coould howvever have reached this wall by travelling over any
possible path.
For the sake of visualization, imagine the electron is in the 2p
z
orbital and the detection
device (the wall) is placed as indicated in Figure 14.2. From the previous discussion we know
that there is a probability amplitude (which itself is the sum of many probability amplitudes)
associated with this measurement. We could call this probability amplitude, (PA)
2pz
. But
this probaility amplitude is just the wavefunction
2pz
ie.,

2pz
= (PA)
2pz
.
14.3 Other types of measurement
Walls measure position. But position is not the only type of measurement. We can measure
any physical quantity. In Figure 14.3 and Figure 14.4 we consider devices which measures
170
momentum and energy.
Figure 14.3: A measuring device which measures the momentum of an electron.
The rst device measures momentum. It consists of two doors which open and close very
fast. An example of a door which opens and closes very fast are camera shutters, as found
in old-fashioned cameras. The crux of this measuring device is that there are two sequential
doors which are a xed distance d apart. Furthermore, the opening and shutting of the
doors is times to be a time t apart. Only an electron moving at the speed v = d/t can pass
through both doors. As momentum is p = mv, and as the m and as the mass of an electron
is known, this device measures the momentum of the electron
The second measuring device measures energy of the electron in a hydrogen atom. It
consists of an optical sensor. An eye is an example of an optical sensor. We measure the
frequency of the light emitted by the electron as the electron transits into the 1s state. As
the energy of the 1s state is known (it is 13.6 eV), this frequency determines the initial
energy of the electron. Note that for light E = h.
171
Figure 14.4: A measuring device which measures the total energy of an electron in a hydrogen
atom.
14.3.1 States
Imagine now that we have placed the momentum measuring device so that it will detect
the momentum p = (0, 6, 0). Imagine the electron is initially in the state . There will
then be a probability amplitude that such an electron will have the momentum Figure 14.3.
But we could also have another kind of detector, a position detector. This position detector
could be a small wall at the position r = (0, 4, 2). There is a probability amplitude that
the electron is detected at r = (0, 4, 2). And nally we could have an energy detector, an
optical sensor which only detects green light. Green light has the energy of 2.3 eV. There
would be a probability amplitude associated to this measurement.
We need a way to express these three dierent probability amplitudes. The bra and ket
notation is the easiest way to express probability amplitudes. We say the state the system
is in is | >. The three probabiliy amplitudes are then respectively, < p = (0, 6, 0) | >,
< r = (0, 4, 2) | > < E = 2.3eV | >.
We call the state in which the system is in | >. And as each of these probability
amplitudes are complex numbers, they each have a complex conjugate: < p = (0, 6, 0) |
>

=< | p = (0, 6, 0) >, < r = (0, 4, 2) | >

=< | r = (0, 4, 2) >, and


< E = 2.3eV | >

=< | E = 2.3eV >.


172
14.3.2 What happens to the state after measurement?
Measurements change the system. We have seen only one exaple of this. Right after we
determine the spatial coordinate of an photon in a laser beam it turns from a planewave into
a radiating wave.
But we could imagine other transformations. In Figure 14.5, we show what happens to
the wavefunction after respectively the momentum and the energy are measured.
Figure 14.5: Momentum and energy measurement changes the state of the system.
How should we name these new states? For example, initially a particle is in the state
| >. What would be a good name after the posititon is located?
One possibility would be to name the new state | r >. The state | r > represents a state
where the particle is located at the position r. Similarly | p > represents a state where the
particle whose momentum is p, and | E > represents a state where the particle whose energy
is E
Here are a list of questions we can answer together.
173
1. For a hydrogen atom in the 1s state, what is < E |
1s
>?
2. For a hydrogen atom in the 1s state, what is <
1s
| E >?
3. Right after an electron in a 1s state has its position measured at r, what state is the
electron in?
4. What does the wave function corresponding to | r > look like?
5. A 1s electron is measured to have momentum p. What state has the electron turned
into after this measurement?
6. What does the wave function corresponding to | p > look like?
7. Express as a function < r | p >?
8. Give an argument which demonstrates that the probability amplitude that a state | r >
after detection of its position turns into the state | r > is < r | r >= 1.
174
Chapter 15
Measurements and Matrices
Up to now in this course, we have lots of math to describe states and systems. Systems can
be an electron in an s, p or d orbital, or a photon in a radiating or a plane wave. We know
the wavefunctions functions for each and every one of these systems. By contrast we know
very little about the math of measurements. We know measurements change the system and
we know what a few diferent measuring devices look like physically, we have no math yet
for them. As measurement is an integral part of quantum theory, we need to develop this
math.
What kind of math will this be? A rst clue is contained in the following pair of exper-
iments shown in Figure 15.1. In the rst experiment, shown in the top panel of the gure,
an electron passes through a position measurement followed subsequently by a momentum
measurement. In the second experiment, shown in the bottom panel, an electron experiences
the momentum measurement followed by the position measurement.
As the gure notes, the order of the measurement matters. For the rst experiment, the
nal outcome is a planewave, but for the second experiment the resultant wavefunction is a
radiating one. Let us call the position measurement r and the momentum measurement p.
182
We note:
rp = pr.
Figure 15.1: Two dierent experiments. In the rst experiment, the electron passes through
a position measurement followed by a momentum measurement. In the second experiment
in travels through the momentum measurement followed by the position measurement. Note
that the order of the measurement changes the nal outcome of the system.
If rp = pr, we say the commutative law is not obeyed. In regular math, there are two
types of entities where the commutative law is not obeyed: composite functions and matrices.
Both can be used to formulate the laws of quantum theory. In this course we shall use the
latter. (Much of this chaper is devoted to telling you how the arithmetic of matrices works.)
183
15.1 Polarized light
It turns out that the position and momentum operators are not the simplest mathematically
of all measurements. Among the simplest is the measurement of the polarization of light.
(And as an added bonus, the math used for polarization is the same math needed to nd the
molecular orbitals used in this course.) The basic experiment is illustrated in Figure 15.2.
Light has two dierent components, one which is up-down polarized and the other which
is left-right. As the gure suggests polarizing screens let through only light with with one
polarization.
After passing through the screen, a measurement has been made. The light has entered
a new quantum state. And as polarized light has only two components, these states can be
easiest described as a two-component vector. We call vertical and horizontal polarized light
respectively

1
0

and

0
1

. Alternatively we can call vertical polarized light v or even


| v > and horizontal light

h or | h >.
Figure 15.2: Vertical and horizontal polarizing screens.
184
We now must check that such polarizing screens are suciently complicated that the
commutative law is not obeyed. We do so with a demonstration in class. This demonstration
is illustrated in Figure 15.3.
Figure 15.3: Do polarizing screens commute?
Let us call the initial state w (or alternatively, in the bra and ket notation, | w >.
The top experiment in this gure could be written, HV w = 0 (similarly we can write
HV | w >= 0). Using this notaion, we can summarize the result in this class as D
R
HV w = 0
and VD
R
H w = 0. We conclude D
R
HV = VD
R
H. Polarizing screens, like the momentum
and position measurements, do not commute.
185
15.2 Matrices
One of the simplest mathematical objects which do not commute are matrices. And we will
use matrices to describe the various polarization measurements. I know I was not introduced
to matrices until college, and they seemed quite abstract at rst. But I think it is useful
to recognize that things which dont commute are quite common in our daily life. Consider
the process of opening a door and then walking through the door. This is quite dierent
from walking through a door and then opening the door. Putting food on a plate followed
by eating the food is dierent from eating the food followed by putting it on the plate. So
in some ways our learning about matrices is an example of our math catching up with our
life experiences.
In any case, matrices are rectangular arrays of numbers, as illustrated in Figure 15.4.
Figure 15.4: Four examples of matrices
186
187
188
189
190
191
15.3 New notatiton for the dot product
In the last section we learned the rules of matrix multiplication. Our goal in this section
is to revise our vector notation in view of these rules. Previously we considered vector
dot products of real numbers, u w. Such vector dot products are one example of matrix
multiplication.
u w =

u
1
u
2

w
1
w
2

= u
1
w
1
+ u
2
w
2
and
u u =

u
1
u
2

u
1
u
2

= u
1
u
1
+ u
2
u
2
= l
2
u
.
These formulas break down if the components of the vectors are complex numbers. In
partucular, by the above denitions, if the components of u are complex, there is no reason
for l
2
u
to be a real number, let alone a positive number. As l
u
is suppossed to be a length,
we wish it to be both real and positive.
For complex numbers we change the rules of dot products to:
u w =

1
u

w
1
w
2

= u

1
w
1
+ u

2
w
2
.
This notation has the advantage that when
u u =

1
u

u
1
u
2

= u

1
u
1
+ u

2
u
2
= l
2
u
,
l
u
can always be a real positive number, the length.
We introduce the following new names: If u =

u
1
u
2

, then u

1
u

. We can
192
refer to u as | u >. We will sometimes call u

by the alternate name < u |. The dot product


of u and w would then be:
u w = u

w =< u | w > .
193
Chapter 16
Polarizing screens and PAs
In the last chapter we began to analyze the math of polarizing screens. As there are only two
components to light polarization, we hypothesized that light polarization can be repesented
as a two-component vector. Vertical polarized light was chosen to be represented as

1
0

,
v, or | v > and horizontal polarized light as

0
1

,

h, or | h >. (In this representation, light
polarized in any direction corresponds to some two-component vector, diagonal polarized
light would be

d
1
d
2

, where d
1
and d
2
are numbers. It would be denoted as

d or | d >.)
Vertical, diagonal, and horizontal polarizing screens are measurements and as we shall see
in this chapter will be represented respectively as the matrices V, D, and H.
Consider the experiment shown in Figure 16.1. As the top panel in this gure shows,
diagonal polarized light reaches a vertical polarization screen and turns into vertical polarized
light. In the bottom panel vertical polarized light reaches a diagonal polarization screen and
the light turns into vertical diagonal light. We will demonstrate both processes in class.
Both of the above experiments illustrate that measurements change the system. Be-
fore the polarization measurement, a photons light polarization is expressed as one two-
194
Figure 16.1: Examples of the polarization measurement
component vector. After polarization measurement, the photon has either been extinguished
or it has become another two-component vector. In the upper panel experiment the vertical
polarizing screen turns some fraction of the photons into vertical polarized light. We write
this statement mathematically as:
V

d =
1
v,
where
1
is a number whose magnitude is less than one (it is less than one as some of the
light has been extinguished). This statement may also be written as,
V | d >=
1
| v > .
Simililarly for the bottom experiment we write
Dv =
2

d,
195
where now
2
is a number whose magnitude is also less than one.
D | v >=
2
| d > .
16.1 Matrices for polarization screens
The polarization screen turns one two component vector into another two component vector.
What sort of mathematical object does this? We know from the last chapter that a good
answer is a 2 2 square-matrix. If we were to call,
V =

V
11
V
12
V
21
V
22

,
where V
11
, V
12
, V
21
, and V
22
are all numbers, then the statement V

d =
1
v can be written:

V
11
V
12
V
21
V
22

d
1
d
2

V
11
d
1
+ V
12
d
2
V
21
d
1
+ V
22
d
2

=
1

1
0

1
0

.
22 square-matrices always convert one two-component vector into another two component
vector.
The remainder of the chapter will be devoted to nding explicit real values for the com-
ponents of the V, D, and H matrices. We begiin with the two-by-two V matrix. We will
distinguish in this and the next chapter two kinds of diagonal polarization screens, D
R
and
D
L
, which correspond respectively to a 45
o
clock-wise or counter-clock-wise rotation of the
vertical polarized screen.
196
16.1.1 V
In Figure 16.2 we show the two experiments needed to determine the four terms in V:
V
11
, V
12
, V
21
, and V
22
. In the top panel, we show that a vertical polarized photon passes
through a vertical polarization screen unchanged. The bootom panel shows that a horizontal
polarized photon is extinguished by the vertical polarizing screen. We conclude Vv = v (or
equivalently that V | v >=| v >) and that V

h = 0 (or equivalently that V | h >= 0). These


expressions can be rewrittten

V
11
V
12
V
21
V
22

1
0

V
11
V
21

1
0

,
and similarly

V
11
V
12
V
21
V
22

0
1

V
12
V
22

0
0

.
We conclude V
11
= 1, V
12
= 0, V
21
= 0, and V
22
= 0.
Figure 16.2: Two polarization measurements whose results can be used to determine V.
197
V =

V
11
V
12
V
21
V
22

1 0
0 0

.
By exactly the same procedure we can nd
H =

H
11
H
12
H
21
H
22

0 0
0 1

.
16.1.2 V
ij
are probability amplitudes
In the previous subsection we found the specic values for V
ij
: V
11
= 1, V
12
= 0, V
21
= 0,
and V
22
= 0. In this section we begin to assign a meaning for each of these four terms.
We will rst express our result in vector notation. First we recall that v

1 0

and

0 1

. We consider v

Vv, which can also be rewritten as < v | V | v >. We note:


< v | V | v >= v

Vv =

1 0

V
11
V
12
V
21
V
22

1
0

= V
11
= 1.
Similarly
< v | V | h >= v

h =

1 0

V
11
V
12
V
21
V
22

0
1

= V
12
= 0.
< h | V | v >=

h

Vv =

0 1

V
11
V
12
V
21
V
22

1
0

= V
21
= 0.
< h | V | h >=

h

h =

0 1

V
11
V
12
V
21
V
22

0
1

= V
22
= 0.
198
We may express all these equations with the following words: < v | V | v > is the PA
that the state V | v > has become the state < v |. < h | V | v > is the PA that the state
V | v > has become the state < h | and so forth. We will learn about the meaning of these
statements later. But note all the numbers in the V matrix refewr to specic PAs.
16.2

d
R
In the last section of this chapter we derive a possible set of values for

d
R
. We note

d
R
=

d
R1
d
R2

= d
R1

1
0

+d
R2

0
1

= d
R1
v +d
R2

h+
As diagonal polarized light is half-way imbetween v and

h, we expect d
R1
= d
R2
. We
conclude

d
R
=

d
R1
d
R1

.
We can determine the value of d
R1
if we recall

d
R
=< d
R
| d
R
>=

d
R1
d
R1

d
R1
d
R1

= 2d
2
R1
.
But < d
R
| d
R
> is just the PA that one state is itself. If there is a single photon with a
100% chance of existing then this PA=1. We conclude 2d
2
R1
= 1, d
R1
=
1

2
, and

d
R
=

2
1

.
199
200
201
202
204
16.3 A fundamental law about Probability Amplitudes
We can use the math we have developed for polarizing screens to express one of the funda-
mental laws of QM.
If the state w is of length one, the probability amplitude that a system in the state u
behaves as if it were in the state w is w

u. Alternatively, if | w > is of length one, the


probability amplitude that a system in the state | u > behaves as if it were in the state | w >
is < w | u >.
As an example consider diagonal left polarized light. What is the PA that such light
behaves as vertical polarized light?
We calculate
v

d
L
= v

[
1

2
v
1

h] =
1

2
or
< v | d
L
>=< v | [
1

2
| v >
1

2
| h >] =
1

2
,
the PA that

d
L
light behaves as vertical polarized light is 1/

2.
Similarly we ask what is the PA that diagonal left polarized light behaves as horizontal
polarized light. We calculate

d
L
=

h

[
1

2
v
1

h] =
1

2
or
< h | d
L
>=< h | [
1

2
| v >
1

2
| h >] =
1

2
,
the PA that

d
L
light behaves as horizontal polarized light is 1/

2.
205
Chapter 17
Eigenvectors, eigenstates, and
eigenvalues
In this chapter we discuss eigenstates, eigenvectors, and eigenvalues. For our purposes
eigenstates and eigenvectors are synonomous: they are both states of the system. The
measured value of an eigenvector is its eigenvalue. We will learn that systems, at the exact
time of the measurement, become an eigenstate or eigenvector of the measurement with a
measured value for the measurement which is the eigenvalue corresponding to this eigenstate.
If the system is already in an eigenvector of the measurement then the measured value
is the eigenvalue, 100% of the time. If the system is not in an eigenstate or eigenvector of
the measurement then the measured value will be one of at least two dierent eigenvalues,
and neither eigenvalue will be measured 100% of the time.
17.1 Polarizing screens
In the previous two chapters we have expressed polarizing screens and polarized light as the
following:
206
V =

1 0
0 0

with | v >= v =

1
0

.
H =

0 0
0 1

with | h >=

h =

0
1

.
D
R
=

1/2 1/2
1/2 1/2

with | d
R
>=

d
R
=

2
1

.
D
L
=

1/2 1/2
1/2 1/2

with | d
L
>=

d
L
=

2
1

.
After the polarizing screens V has measured a photon of light, the photon of light is
either extinguished or is now a v or | v > polarized photon and so forth.
17.1.1 Eigenvectors and eigenvalues of a measurement
A measurement M has an eigenvector (or eigenstate) w or | w > with eigenvalue if
M w = w
or alternatively
M | w >= | w > .
207
An eigenvector when measured stays in the same vector state; its measured value is its
eigenvalue.
For example, we observe
Vv = 1v
or alternatively
V | v >= 1 | v >
Based on the denitions of eigenvectors and eigenvalues above we see that v is an eigenvector
of V with an eigenvalue of one. When we measure a vertical photon in a vertical polarizing
screen, we are measuring to what probability a photon is a vertical polarized photon. If the
state is v, 100% of the time (ie., with probability 1) it is in in the vertical polarized state. An
eigenvector, when measured by its corresponding measurement results in a measured value
which is the eigenvalue of this same eigenvector 100% of the time.
17.1.2 V has a second eigenvector
Note that
V

h = 0 = 0

h
or alternatively
V | h >= 0 = 0 | h > .
The state

h must also be an eigenvector of V, this time with eigenvalue 0. 100% of the time
when horizontal polarized light is measured as to whether it is vertical polarized light will
respond that it is not vertical polarized light.
208
17.1.3 The eigenvectors of D
R
D
R

d
R
= 1

d
R
and
D
R

d
L
= 0 = 0

d
L
.
Alternatively
D
R
| d
R
>= 1 | d
R
>
and
D
R
| d
L
>= 0 = 0 | d
L
> .
The eigenvectors of the diagonal right polarizing screen are diagonal right and diagonal left
polarized light. 100% of the time diagonal right polarized light behaves like diagonal right
polarized light and 100% of the time digonal left polarized light does not act like diagonal
right polarized light.
Eigenvectors are states which respond to measurements with 100% certainty.
17.1.4

d
L
is not an eigenvector of V
We note
V

d
L
= 1/

2v
or alternatively
V | d
L
>= 1/

2 | v > .
209
We conclude

d
L
or | d
L
> is not an eigenvector of V. We nd

d
L
=
1

2
v
1

h
or
| d
L
>=
1

2
| v >
1

2
| h > .
The PA that diagonal left polarized light behaves as vertical polarized light is 1/

2 and the
PA that diagonal left polarized light behaves as horizontal polarized light is 1/

2. The
probability that diagonal left polarized light behaves as vertical polarized light is 50% and
the PA that diagonal left polarized light behaves as horizontal polarized light 50%.

d
L
or
| d
L
> light is not an eigenvector of the V measurement. But after the V measurement is
made, the system has become an eigenvector of the V measurement.
Note that in the above expression, we have expressed the state

d
L
or | d
L
> as the
weighted sum of the actual eigenvectors of the V measurement. Such a weighted sum
is called a linear combination. The coecient attached to any given eigenvector in this
expression is the PA that the measured value for the measurement is the eigenvalue of that
given eigenvector.
17.2 The position and energy measurements
Let us now consider the position measurement r. The eigenstates of the position measure-
ment are those states which give exactly the smae position 100% of the time. If we place a
detector at a hole in a screen, whenever we detect a particle pass through the hole, we know
exactly where the particle is located 100% of the time. Recalling that we call this state | r >,
210
we conclude | r > is an eigenstate of the position measurement.
r | r >= r | r > .
The position of | r > when measured is r 100% of the time. After this position measurement
the particle is in the state | r >.
The electron of a hydrogen atom in a 1s orbital is not in a position eigenstate. Measuring
the position leads to many dierent measured positions. After the position is measured on
a 1s electron, however, the system changes: it turns into an eigenstate of the position
measurement. If the position measured is r, the system turns into | r >. By contrast the 1s
state is an eigenstate of the energy measurement. 100% of the time a hydrogen 1s orbital
has its energy measured, its energy is -13.6 eV. After a hydrogen electron has had its energy
measured at -13.6 eV, the state of that electron has turned into the 1s state. The energy
measurement is called the Hamiltonian: it is abbreviated H.
H |
1s
>= 13.6 |
1s
> .
17.3 The momentum measurement
Let us try our understanding of the above material with the following questions:
1. The momentum measurement is called p. What are eigenstates of the momentum
measurement?
2. What happens after a plane wave has had its momentum measured?
3. Is the meaured momentum of an electron in a plane wave always the same?
4. What happens after a plane wave has had its position measured?
211
5. Is the meaured position of an electron in a plane wave always the same?
6. What interpretation can be given to the Heisenberg Uncertainty Principle based on
the eigenvector and eigenvalue concept?
212
Chapter 18
The laws of quantum mechanics and
average energy
A portion of this chapter is written in the bra and ket notation. You may still feel unfamiliar
with this notation. If so, please remember that | > is just another way of writing

, while
< | could be alternatively written as

. It is important also to recognize that the bra and


ket notation is the accepted notation in quantum mechanics, while the vector-arrow notation
is not very widely used. An Appendix has been added to this chapter which discusses a bit
ideas behind the bra and ket notation.
18.1 A rst two laws of quantum mechanics
Law 1: Systems are in states. These states can be thought of as vectors, u or u

, (as
treated in polarization experiments) or as kets or bras | > or < |.
Law 2: The probability amplitude that a system in the state u will behave as if it is in the
state w is w

u, if the vector w is of unit length. In bra and ket notation, this statement is
expressed: the probabilty amplitude the state | > will behave as if it is in the state | >
213
is < | >, if the ket | > is of unit length.
Example 1: Consider the dot product of a diagonal-left polarized photon with a vertical
polarized photon.
v

d
L
= (
1 0
)

1/

2
1/

= 1/

2.
By Law 2, the above dot product is the probability amplitude that a diagonal-left polar-
ized photon will behave as a vertical polarized photon. As a test of this law, consider the
experiment shown in Figure 18.1. This gure shows that a diagonal-left photon passing
through a vertical polarized screen, has the probability amplitude of behaving as if it were a
vertical-polarized photon of 1/

2. The probability amplitude predicted by Law 2 is correct.


Figure 18.1: A

d
L
photon travelling through a vertical polarization screen.
Example 2: In exactly the same manner, consider the dot product of a horizontal polarized
photon with a vertical polarized photon.

v = (
0 1
)

1
0

= 0.
214
By Law 2, the above dot product is the probability amplitude that a vertical polarized
photon will behave as a horizontal polarized photon. As a conrmation of this law we
consider the experiment Figure 18.2. This gure shows that a vertical photon passing through
a horizontal polarized screen, has zero probability amplitude of behaving as if it were a
horizontal-polarized photon. The probability amplitude predicted by Law 2 is correct.
Figure 18.2: A v photon encountering a horizontal polarization screen.
Example 3: Consider the slit experiment in which a photon in the | 2p
z
> travels to a
wall, where its position is measured to be at the point r
1
, see Figure 18.3. Immediately
upon hitting the wall the particle state changes to the state | r
1
>. In the state | r
1
>, the
particle is exactly located at the point r
1
. According to Law 2, the probability amplitude
the state | 2p
z
> behaves as if it is in the state | r
1
> is < r
1
| 2p
z
>. From the denition
of wavefunctions we know
2pz
(r
1
) =< r
1
| 2p
z
>. As wavefunctions are the probability
amplitude that a particle is located at a given point in space, we see the probability amplitude
calculated by Law 2 is correct.
215
Figure 18.3: A 2p
z
orbital and a small location detection device located at r
1
. The small
detection device can be thought of as a wall.
18.2 A third and fourth law of quantum mechanics
Law 3: A measurement will change the state of the system into an eigenstate (or eigen-
vector) of the measurement. An eigenstate or an eigenvector of a given measurement is a
state which upon this given measurement always gives exactly the same measured value.
Law 4: When a a measurement is made, the measured value is always an eigenvalue of the
measurement
Example 1: The diagonal-left polarizing screen is a measuring device corresponding to
the matrix
D
L
=

1/2 1/2
1/2 1/2

.
This matrix has two eigenvectors

d
R
and

d
L
with respective eigenvectors of 1 and 0. Upon
measurement, the photon responds either by declaring it is a diagonal-right polarized photon,
a state which it then stays in until subsequent measurement or a diagonal-left polarized
photon, which then gets extinguished.
Example 2: An electron in the wavefunction | 2p
z
> hits a wall, where its position is
measured to be at the point r
1
, The electron state turns into an eignestate of position.
216
The state | r
1
> is a state which upon spatial measurement is always located at the same
meqasured position, r
1
.
Example 3: An electron is in the wavefunction = e
i pr/h
. This wavefunction must be
an eigenstate of the momentum measurement as every time its momentum is measured its
measured momentum is the same, p. We therefore conclude
= e
i pr/h
=< r | p > .
If an electron is in the state | p > and its position is measured to be at the point r
1
, the
state of the electron changes to | r
1
>.
18.3 A fth and sixth law of quantum mechanics
Law 5: Just as all states can be thought of as vectors, all measurements can be thought
of as matrices. M w is the state w which has had the measurement M made upon it. In the
bra and ket notation this can be written as M | w >.
Law 6: All measurement matrices are Hermetian. A Hermetian matrix, M, is one where
M

ij
= M
ji
. Using the dagger convention, if M is Hermetian, then M

= M.
Law 6 has a number of important consequences. Among them are:
1. All measurement eigenvalues are real numbers.
2. Two eigenvectors of the same measurement which have dierent eigenvalues are always
orthogonal to each other.
3. I suspect, but am not sure, that this assumption ties in with Newtons third law.
217
Example 1: The V, H, D
R
, and D
L
matrices are all Hermetian. For any one of these
measurements, the eigenvalues are either 1 or 0, both real numbers. For any one of the mea-
surements, the eigenvectors with dierent eigenvalues, corresponding to this measurement
are orthogonal to one another. For example, the D
L
matirx shown on the previous page has
two eigenvectors with dierent eigenvalues:

2
1

and

2
1

.
These two eigenvectors have a mutual dot-product of zero and are therefore orthogonal to
one another.
Example 2: For this example, recall all atomic orbitals are eigenvectors of both the en-
ergy and the total angular momentum measurements. If two atomic orbitals have dierent
eigenvalues with respect to either of these two measurements, they will be orthogonal. For
example, for the hydrogen atom, the states
2s
and
2px
have the same energy but dierent
angular momentum eigenvalues. These two states are therefore orthogonal to each other. In
bra and ket notation, we write:
<
2s
|
2px
>= 0.
In a similar manner, all atomic orbitals with either dierent energies or dierent angular
momenta are orthogonal to each other.
18.4 The completeness theorem
In order for the above laws to be logically self-consistent, the completeness theorem must
hold true. Note that the completeness theorem is actually a mathematical consequence of
the Hermetian nature of all measurements.
218
The completeness theorem: Any quantum state can be expressed as the sum of the
various eigenstates corresponding to any one measurement.
Example 1: The V screen has two eigenvectors v and

h. Consider the

d
L
state:

d
L
=

2
1

=
1

1
0

0
1

=
1

2
v
1

h
= (v

d
L
)v + (

d
L
)

h.
The

d
L
state can be written as a linear combination of the probability amplitudes that the
photon is in the v and

h states times their respective states.
Example 2: An electron is in as sp
3
hybrid orbital:
| sp
3
>=
1
2
| 2s > +

3
2
| 2p
z
> .
We measure the total angular momentum. The probability amplitude that the total angular
momentum is zero is 1/2, while the probability amplitude the electron has l = 1 is

3/2.
Note the | sp
3
> state is a linear combination of the | 2s > and | 2p
z
> states.
18.5 The meaning of < | H | >
If the bra and ket notation still makes you uncomfortable, please remember that | > and
< | can be written respectively as

and

: the bra and ket notation is just a vector


notation.
A model system: Consider a system which has at least two energy eigenstates, |
1
> and
|
2
>, both of which are kets of length 1, ie., <
1
|
1
>=<
2
|
2
>= 1 and assume these
219
energy eigenstates have dierent energy eigenvalues, respectively E
1
and E
2
. Now consider
a state | >. Assume < | >= 1 and that | >=
1
|
1
> +
2
|
2
>, ie., | > is a
linear combination of the the two energy eigenstates |
1
> and |
2
>. We now show:
Theorem: < | H | >= E
avg
, where H is the energy measurement operator, the
Hamiltonian, and E
avg
is the average energy of an electron in the state | >.
Proof: To prove this theorem, we note, rst, that |
1
> and |
2
> are both of length one
and, second, according to Law 6.2 the two states are orthogonal to one another, In other
words, <
1
|
1
>= 1 and <
1
|
2
>= 0 . Consider the expression for | >,
| >=
1
|
1
> +
2
|
2
>
. We apply <
1
| to both the left and right of this equality. We nd:
<
1
| >=
1
<
1
|
1
> +
2
<
1
|
2
>=
1
,
and similarly :
<
2
| >=
2
.

1
and
2
are therefore the probability amplitudes that | > behaves as if it were in
respectively the states |
1
> and |
2
>.

1
and

2
are therefore the probabilities that
| > behaves as if it were in respectively the states |
1
> and |
2
>. We now consider the
expression < | H | >. Noting | >=
1
|
1
> +
2
|
2
>, we nd:
< | H | >= (

1
<
1
| +

2
<
2
|) H(
1
|
1
> +
2
|
2
>)
=

1
<
1
| H |
1
> +

2
<
1
| H |
2
> +

1
<
2
| H |
1
> +

2
<
2
| H |
2
>
220
=

1
<
1
| H |
1
> +

2
<
2
| H |
2
>
=

1
E
1
<
1
|
1
> +

2
E
2
<
2
|
2
>
=

1
E
1
+

2
E
2
Recalling that

i
is the probability that the measured energy is E
i
, the last expression
is the sum of the two possible measured energies weighted by their individual probabilities. In
other words, the last expression is just the probability weighted average energy, the assertion
which we set out to demonstrate.
The completeness theorem and < | H | >: In the last example, we worked out that
if a state | > is the linear combination of two energy eigenstates, then < | H | >= E
avg
.
The exactly same argument holds were | > to be the linear combination of three or even
more energy eigenstates. Some reection shows, that as long as | > can be expressed as a
linear combination of energy eigenstates, then < | H | >= E
avg
. But the completeness
theorem tells us that it is always possible to express any state as a linear combination of the
eigenstates of any measurement. Thus the result given in the last example is a general one.
18.6 Summary
With this chapter, we complete our discussion of quantum theory. The material being
both new and unexpected can be confusing for the student. But please bear with these
complexities. You have been presented with an entirely dierent way of viewing the world.
Not only do we generally not want to give up our old views, this new world-view is an
inherently perplexing one.
One suggestion on how to master this material is the following. Take a look at the lecture
material from the beginning and ask yourself if can you see how the language developed in the
221
last few chapters can be used to clarify earlier lectures in this course. Reexamine each of the
three quantum mechanical systems we have examined and one classical system, laser light,
polarized light, electrons in orbitals and musical instruments. Find where measurements,
eigenvalues, eigenstates, and PAs come to play in each of these topics (for the musical
instruments look only for the eigenvectors and eigenvalues). Puzzle you way through each
of these topics trying to see that all this material is one cohesive whole. And please come to
talk to a TA, John Terry, or me as you reason your way through this material.
Here is a brief summary:
1. Quantum systems are in states. States can be thought of as vectors.
2. Measurements correspond to matrices.
3. A measurement on a system always results in an eigenvalue as the measured value.
4. The probability amplitude for a state | > that the measured value is t, where t is
the eigenvalue corresponding to the eigenstate | > is < | >, assuming | > is of
length one.
5. The probability is the probability amplitude times its complex conjugate.
6. Measurement changes the system. Directly after measurement, the system is in the
eigenstate which corresponds to the measured eigenvalue.
18.7 Appendix: Vectors and the bra and ket notation
The bra and ket notation is abstract. You may, as a student, nd it uncomfortable. In this
section, we will try to show you the need for and the utility of this notation.
Bras and kets come about because quantum states are vectors. Vectors are inherently
more abstract than numbers. Consider the vector illustrated in Figure 18.4. Would it be
222
more correct to call this vector
(1, 0) or

1
1

or (1, 0, 1) ?
Figure 18.4: A vector.
Unless we have chosen a set of coordinate axes, it is impossible to answer the question.
In Figure 18.5, for example, we illustrate two dierent choices for coordinate axes. The
rst choice makes (1, 0) a resonable descriptor of the vector in question. The second set
of coordinate axes makes the vector (1, 0, 1) the more appropriate choice. Vectors, which
can be thought of as arrows, do not by themselves, specify the individual components of the
vector. The individual components, and even the number of individual components, of a
vector are specied by the arrow with respect to a set of coordinate axes.
In quantum mechanics, coordinate axes are chosen by the measurement which we are
interested in. Any measurement, whether it is energy, or position, or momentum, has a
complete set of orthogonal eigenvectors. Once a measurement is decided upon, the reasonable
thing to do is to choose the eigenvectors of the chosen measurement as ones coordinate axes.
The completeness theorem given in the main body of this chapter ensures that it is always
possible to create a full set of coordinate axes out of a measurements eigenvectors.
As an example of completeness and how it aects the choosing of a set of coordinate axes,
consider an electron in a hydrogen atom. Furthermore, consider the energy measurement.
223
Figure 18.5: The same vector shown in the previous gure, but which now because of a
chosen set of coordinate axes can now be represented as a short sequence of numbers.
In creating a coordinate set of axes, we therefore need to consider the energy eigenstates of
the hydrogen atom. These eigenstates should by now be well known to the student: they
are | 1s >, | 2s >, | 2p
x
>, | 2p
y
>, | 2p
z
>, | 3s >, and so forth. The completeness theorem
states:
| >=
1s
| 1s > +
2s
| 2s > +
2px
| 2p
x
> +
2py
| 2p
y
> +
2pz
| 2p
z
> +...,
where | > is any possible state.
The student can show as an exercise that the value of each of the s corresponds to the
dot product between a bra and a ket.
Exercise: Show that
2pz
=< 2p
z
| >. In order to carry out the exercise recall that all
atomic orbitals can have unit length and that all dierent atomic orbitals are orthogonal to
one another.
In the above example we see that the sequence of numbers (
1s
,
2s
,
2px
,
2py
,
2pz
,
3s
...
224
completely determines | >. This sequence of numbers is one choice for the numbers
associated with a vector. We can write this sequence of numbers as a row vector or as a
column vector:

1s
,
2s
,
2px
,
2py
,
2pz
,
3s
, ...

or

1s

2s

2px

2py

2pz

3s
...

The above example shows how a quantum state | > is converted into a sequence-
of-numbers vector by the choice of a measurement and a set of coordinate axes which are
eigenstates of the chosen measurement.
18.7.1 Another set of coordinate axes for the state | >:
We now consider another set of coordinate axes to describe the | > state of an electron
in a hydrogen atom. A new set of coordinate axes comes about by the choice of a new
type of measurement. In this subsection, the new measurement is position. The student
may have more trouble recognizing position eigenvectors. The following clue may help: An
eigenvector of a measurement is a state gives the same measured value every time such a
measurement is made. An electron in a position eigenvector would, were position to be
measured, always declare itself to be at an exactly specied point in space. Consider the
state | r = (2, 3, 1) >. Let us dene such a state as the state where if position were
measured, the electron would always be found at the position (2, 3, 1).
We now can dene a set of position eigenvectors: | r = (0, 0, 0) >, | r = (dx, 0, 0) >, | r =
(0, dx, 0) >, | r = (0, 0, dx) >, | r = (2dx, 0, 0) >, | r = (dx, dx, 0) >, | r = (dx, 0, dx) >, ...
225
We could in such a way indicate to reasonable accuracy every point in space and therefore
every position eigenvector. The state | > could be expressed as:
| >=
1
| r = (0, 0, 0) > +
2
| r = (dx, 0, 0) > +
3
| r = (0, dx, 0) > +
4
| r = (0, 0, dx) > +...
In exactly the same way as before, we nd that the coecients are the dot products
of a bra times a ket. For example,
2
=< r = (dx, 0, 0) | >. Just as with the energy
eigenvectors, we can create row or column vectors:
(
1
,
2
,
3
,
4
, ...) or

4
...

,
where we use the position eigenstates as the coordinate axes.
Coecients such as
2
equal the dot product, < r = (dx, 0, 0) | > which are tra-
ditionally expressed as the value of the wavefunction at the position (dx, 0, 0). In other
words:

2
=< r = (dx, 0, 0) | >= (dx, 0, 0).
The vectors written above are just a sequential list of the values of the wavefunction:

4
...

(0, 0, 0)
(dx, 0, 0)
(0, dx, 0)
(0, 0, dx)
...

.
226
This notation is however voluminous and unwieldy. Better is just to think of the function
(r), and that were we interested, we could express this wavefunction as either a row or a
column vector.
18.8 Problems
1. Please state the wavefunction for a photon of red-light in a laser-beam travelling in the
(1, 1, 0) direction.
2. Imagine an electron is in the state | r
1
>. Its position is measured. What is the mea-
sured position value? What state is the electron in immediately after measurement?
3. Recall that < p | r >=< r | p >

. Imagine an electron is in the state | r


1
>. Its
momentum is measured. What are the probability amplitudes for dierent momentum
measurements? What is the relative probability of the momentum measurement being
p
1
vs. p
2
?
4. Please state the function A( p) for an electron located at the position r
1
.
5. This problem illustrates one reason as to why measurement matrices are Hermetian.
Consider the matrix:
G =

1 0
1 0

with eigenvectors w
1
=

2
1

and w
2
=

0
1

.
Consider the probability amplitudes w

1
w
2
and w

2
w
2
. Can G be a measurement?
6. This problem illustrates another reason as to why measurement matrices are Herme-
227
tian. Consider the matrix:
J =

0 0
1 0

.
Find the eigenvectors of this matrix. Now consider the state
u =

1
0

. Can J be a measurement?
7. Do the above two examples show that the eigenvectors (with dierent eigenvalues) of
a measurement have to be orthogonal? Do the above examples show that eigenvectors
with the same eigenvalue have to be orthogonal? Do the above examples show that
eigenvectors have to be complete? A mathematical theorem states that the eigenvec-
tors of a Hermetian matrix are always complete and that eigenvectors with dierent
eigenvalues are always orthogonal. Does is seem reasonable that all measurements are
represented by Hermetian matrices?
228
Chapter 20
The H
2
molecule
derives the Hamiltonian and MO diagram of H
2
. Show rules 1-3 of general rules of making
MO digrams are obeyed. Handwritten notes given in document 6-1.
254
Chapter 21
The N
2
molecule
derives the Hamiltonian and MO diagram of N
2
Show rules 1-3 of general rules of making
MO digrams are obeyed.. Handwritten notes given in document 6-2.
255
Chapter 22
The HF molecule
derives the Hamiltonian and MO diagram of HF Show rules 1-4 of general rules of making
MO digrams are obeyed. Handwritten notes given in document 6-3.
256
Chapter 23
The rst four rules used in making
MO diagrams
23.1 Introduction
We have examined the H
2
, N
2
and HF MO diagrams twice in this course. Early on in the
course we derived these three MO diagrams using the general rules of all MO diagrams.
In the last three lectures, we have derived these same three MO diagrams based directly
on quantum theory itself. While in principle, we can always use either approach in the
generation of MO diagrams, in practice, for molecules of ordinary complexity, only the former
approach will allow the student to derive chemically useful MO diagrams. (Computers may
always be used to derive full MO diagrams: for students going on in chemistry, you may
use computer calculations to obtain accurate electronic energies, and then rationalize your
computer results with the general rules.)
For many of the molecules we will study in this course, by necessity we will use the general
rules to derive their corresponding MO diagrams. It is therefore important to understand
the connection between the general rules and quantum mechanics itself. In this chapter, we
257
show the connection between the rst four rules and quantum theory. In the next chapter,
we will expound on the remaining fth rule, a rule which we have hitherto not applied.
We begin by recalling all ve rules. Stated in parantheses after each rule are the physical
principles connected with the rule.
23.1.1 The ve general rules used in making MO diagrams
1. The number of molecular orbitals equals the number of atomic orbitals. (Hermeticity)
2. The lowest energy MO is the lowest possible (energy) orbital one can make. The
highest energy MO is the highest possible (energy) orbital one can devise. (Variational
theorem)
3. When two orbitals mix, the amount the orbitals mix is inversely proportional to the
initial dierence in energy of the orbitals. Therefore, orbitals which are initially close
in energy mix (ie. interact) more than orbitals which are initially far apart in energy.
When two orbitals of dierent energy can mix (ie. interact), the low energy combination
resembles more the initial low energy orbital; the higher energy combination resembles
more the initial high energy orbital. When two orbitals can interact and they are of
the same initial energy, then the two resultant combination orbitals are derived equally
from the two initial orbitals. (Second order perturbation theory)
4. Molecular orbitals in an MO diagram are always orthogonal to one another. (Her-
meticity)
5. To be discussed in the next chapter: If there are a set of perpendicular and/or parallel
mirror planes and C
2
-axes, then the MOs can always be written as symmetric or
antisymmetric orbitals with reespect to each of these operations. (Conservation laws)
258
23.2 Rule 1: Number of MOs = number of AOs
In all three MO diagrams so far seen in this course, the number of MOs are equal to the
number of AOs, see Table 23.1. The source for these equal numbers lies in the Hamiltonian
we use. This Hamiltonian is a Hermetian square matrix where each row or each column
represents one of the atomic orbital. Thus the number of atomic orbitals equals the dimension
of the square matrix. This is important as a mathematical theorem tells us the number of
orthogonal eigenvectors exactly equals the dimension of square Hermetian matrices like the
Hamiltonian (more on this point later). As the orthogonal eigenvectors of the Hamiltonian
are the MOs of the MO diagram, the number of MOs exactly equals to the number of AOs.
Table 23.1: Number of Orbitals.
Molecule Number AOs Number MOs
H
2
2 2
N
2
8 8
HF 2 2
Rule 1 is therefore derived from the quantum mechanical treatment developed in this
course. It works best in conjunction with the minimal valence basis set of atomic orbitals
assumed in this course.
23.3 Rule 2: The variational theorem
Rule 2 is a restatement of the variational theorem of quantum mechanics. This theorem states
that the lowest energy MO is the lowest possible energy orbital. Equally, when dealing with
a nite number of MOs, as we always do in the minimal valence basis set approach, the
variational theorem also states the highest energy MO is the highest energy orbital possible.
To understand Rule 2, we must therefore understand the variational theorem.
This theorem is a consequence of one the fundamental principles of quantum mechanics,
259
the principle which states that when one makes a measurement, the measured value is always
an eigenvalue of the measurement. Thus, if one were to measure the energy, the measured
energy would always be an eigenvalue of the Hamiltonian. The energy of any orbital is a
probability weighted average of the dierent eigenvalues.
Expressed as equations, these thoughts state that if |
i
> are the dierent MOs (all
normalized to be of length one) and each with respective eigenvalues E
i
, then an arbitary
orbital | w > (also normalized to length) one can always be expressed as:
| w >=
1
|
1
> +
2
|
2
> +... +
3
|
N
>,
where N is the number of orthogonal molecular orbitals, and
i
=<
i
| w >, the probability
amplitude that the | w > state behaves like the |
i
> state. The energy of the arbitrary
orbital | w >, E
w
is
E
w
=

1
E
1
+

2
E
2
+... +

N
E
N
.
This last expression states mathematically that the energy of the state | w > is the average
of all the energy eigenvalues, E
i
weighted by the probabilities that each individual energy
eigenvalue will be measured. E
w
is the weighted average of the energy eigenvalues.
A consequence of the above statements is Rule 2, the variational theorem. The proof is
by contradiction.
Variational theorem: The lowest possible energy orbital is the lowest energy MO.
Proof: Assume that the state | w > has an energy lower than the lowest energy eigenvalue,
which we will call E
1
. By the above, the energy of the | w > state, E
w
is:
E
w
=

1
E
1
+

2
E
2
+... +

N
E
N
,
260
where

i
is the probability that | w > is in the |
i
> state. As all E
i
are greater or equal
to E
1
the below holds true:
E
w
=

1
E
1
+

1
E
2
+... +

N
E
N

1
E
1
+

2
E
1
+... +

N
E
1
= (

1
+

2
+... +

N
)E
1
= E
1
.
Thus E
w
is greater or equal to E
1
. But by assumption E
w
is less than E
1
, a clear
contradiction.
By a similar argument we can show that the highest possible energy orbital is the highest
energy MO.
23.3.1 Loaded dice
A simple analogy to the variational theorem can be found in a loaded die. In this analogy
a loaded die is the state | w >, the six sides of the die are the six eigenvectors, and the
corresponding numbers 1 through 6 are the six eigenvalues. While an unloaded die has,
when rolled, an average value of 3.5, a loaded die can have an average value anywhere from
1 to 6. However, the loaded die can never have an average value below 1 or above 6. In just
the same way, the state | w > can never have an average energy below the lowest eigenvalue
or above the highest eigenvalue.
23.4 Rule 3: The mixing of orbitals with dierent ini-
tial energy
This rule is a key rule. It applies to the mixing of just two orbitals. Thus this rule is about
the eigenvectors of two-by-two square Hamiltonian matrices. Let us call the two initial
261
orbitals |
1
> and |
2
>. The two-by-two Hamiltonian matrix is made up of the terms,
H
ij
=<
i
| H |
j
>. We can envision two dierent classes of two-by-two Hamiltonian
matrices. In the rst case, H
11
= H
22
= . The Hamiltonian is:
H =

.
In both the H
2
and N
2
MO diagrams, we have dealt with such a Hamiltonian. As Figure 23.1
shows, for such systems, both MOs have equal magnitude contributions of the two starting
states.
Figure 23.1: Two initially equal orbitals mixing to make two equally weighted MOs.
We have seen one example of the second type of two-by-two Hamiltonian matrix, where
the rst orbital has an initial energy and the second orbital has an initial energy , the
262
HF MO diagram. The Hamiltonian is:
H =

.
In Figure 23.2 we show the MO diagram for this Hamiltonian.
Figure 23.2: Two initially not equal energy orbitals mixing to make two not equally weighted
MOs.
As we see in this gure, the lower and higher energy MO are weighted more on respectively
the initially lower and higher energy states. The amount of the initially higher energy state
which mixes into the lower energy MO, or conversely the amount the initially lower energy
state mixes into the higher energy MO are both inversely proportional to the inital dierence
of energy of the two states, 2.
Note however that there is a dierence between the rst type of Hamiltonian, where
H
11
= H
22
, and the second case, where H
11
= H
22
. In the former case, we solved exactly
263
for the eigenvectors and eigenvalues. In the latter case, we made an approximation based
on the assumption 2 >> . The latter results are therefore only approximately true.
The approximation used to obtain these results in quantum theory is called second order
perturbation theory.
23.5 Rule 4: Orthogonality of MOs
As both Figures 23.1 and 23.2 show, MOs are orthogonal to one another. (Orthogonality
between two vectors is most easily seen by taking the dot product of the vectors. If this dot
product is zero, the vectors are orthogonal.)
Both the orthogonality of MOs with dierent energies and that the number of orthogonal
MOs equals the number of initial AOs are consequences of the Hermeticity of the Hamiltonian
matrix. A square matrix, H, is Hermitian (ie., exhibits Hermiticity) if for all i and j,
H
ij
= H

ji
. Note that in the previous section both Hamiltonians were Hermetian.
A theorem in linear algebra, which we state without proof, is that the all eigenvectors of
a Hermitian matrix with dierent eigenvalues are orthogonal to one another. Furthermore,
the number of orthogonal eigenvectors equals the dimension of the square Hermetian matrix.
As MOs are eigenvectors of the Hamiltonian, and as the Hamiltonian is Hermetian, both
Rules 1 and 4 are always true.
23.6 Conclusion
This chapter provides some of the key underpinning for results derived in future chapters.
Together with Rule 5, the subject of the folowing chapter, we will use these general rules to
derive a number of MO diagrams. We will begin with organic planar molecules, move on to
simple main group compounds, transition metal complexes, and nally some simple metal
264
systems. In every case, the general rules will prove to be all we need to derive chemically
useful MO diagrams. This and the next chapter show that these rules are themselves derived
from quantum theory.
265
Chapter 24
Rule 5 of MO diagrams
derives rule 5 of MO diagrams include the example of trans-butadiene. Handwritten notes
given in documents 6-5 and 7-1
266
Chapter 25
Butadiene, H
+
3
, and benzene
derive MO diagrams of butadiene, H
+
3
, and benzene. Handwritten notes given in document
7-1.
267
Chapter 26
Classical arrow convention: electron
donors and acceptors
Two examples. One is amide. The -NHR group is an electron donor. Show how this works
in an MO picture with amide and hydride reaction. -OH is an electron donor. -COR is
an electron acceptor. Go through classic organic pictures, with arrows, of hydride addition,
-NHR being an electron donor. We can go over -COR and -OH with classic arrow pictures
of electrophilic aromatic substitution. Handwritten notes given in document 7-2 and 7-3.
268
Chapter 27
Frontier orbitals and hydride addition
derives frontier orbital picture, that the HOMO and LUMO control the MO interactions
between two molecules. Introduces the key concepts of electrophiles (Lewis acid character,
LUMOs and nucleophiles (Lewis base character, HOMOs) Shows how we can predict ease
of hydride reactions based on MO diagrams giving HOMOS and LUMOS. Hydrides are nu-
cleophiles; organic molecule electrophiles. We examine organic molecule LUMOS. Examine
alkenes vs. C=O. Handwritten notes given in document 7-3.
269
Chapter 28
Electrophilic aromatic substitution
EAS MO theory explained. Handwritten notes given in document 8-1.
270
Chapter 29
Naphthalene
The MO diagram of naphthalene derived and its EAS preferences explained. Handwritten
notes given in document 8-2. Figure 29.1
271
Figure 29.1: A comparison with the accurate calculation of the key frontier orbitals of aro-
matic molecules with the experimentally observed sites of electrophilc aromatic substitution
(EAS), and especially nitration. Although many of these systems are of low symmetry and
therefore have dicult -MO diagrams to derive by hand, to my knowledge, application of
the general rules of making MO diagrams obtains the correct site of EAS in every case.
272
Chapter 30
The Woodward-Homann rules
An example of the Woodward-Homann rules given. Either Diels Alder or butadiene cy-
clization. Handwritten notes given in document 8-3
273
Chapter 32
VSEPR
Of all the subjects presented in freshman chemistry textbooks, there may be none more
controversial than the V alence Electron Pair Repulsion theory or VSEPR. It is a theory
taught to almost all freshman chemistry students and is used to rationalize the shapes of
molecules such as water, ammonia, and carbon dioxide. The controversy surrounding VSEPR
is not with the rules themselves (the rules work) but rather the physical basis of the rules.
Most freshman chemistry textbooks would have the student believe that the source for
the rules, as the name implies, is valence shell electron pair repulsion. Work done in the 1960s
and 1970s by chemists such as Larry Bartell however have led many chemists to believe that
the VSEPR rules are not rooted in valence shell electron pair repulsion. Instead, MO theory,
of the type taught in this course and used to derive the Woodward-Homann rules, can be
used to account for the VSEPR rules. The VSEPR rules appear to be based in the same
electronic behavior which leads to electrophiles, nucleophiles, atomic orbitals and periodic
trends in general.
In this chapter, I will follow the exposition of this subject as taught to me by my Ph.D.
advisor, Jeremy Burdett when I was in graduate school. Jeremy himself began his Ph.D.
studies at the University of Michigan, while there studied under Larry Bartell and later was
288
a postdoctoral fellow of Roald Homann.
32.1 The VSEPR rules
These rules are clearly stated in any freshman chemsitry textbook. Please consult your 2150
textbook. You will be responsible for knowing the rules; theyre not too hard to learn. I will
go over them in class.
32.2 Central atom molecules and MO diagrams
Main group molecules and ions as diverse as H
2
O, NH
3
, CO
2
, CCl
4
, SiF
2
6
all contain central
atoms to which all the other atoms are bonded. The VSEPR rules determine the structure
and shape of such central atom molecules and ions. In this chapter we will study just one
example of the VSEPR rules, that for a central main group atom bonded by -bonds to two
peripheral atoms. Examples of such systems are listed in Table 32.1
Table 32.1: Geometries of some AX
2
molecules and ions.
Number of Number of Geometry Example
valence e

lled orbitals
4 e

2 lled orbitals linear BeF


2
6 e

3 lled orbitals bent CH


2
CF
2
8 e

4 lled orbitals bent OH


2
OF
2
10 e

5 lled orbitals linear IF

2
As this table shows, AX
2
molecules are either linear or bent, geometries which are illus-
trated in Figure 32.1.
289
Figure 32.1: Two possible geometries of AX
2
molecules, where A is the central atom.
32.3 The MO theory for AX
2
molecules and ions
MO analyses of the VSEPR rules are carried out one stoichiometry at a time. In this section
we examine the linear and bent AX
2
geometries. But the MO procedure used to analyze
AX
2
systems is much the same as the one used for other AX
n
geometries: it is a procedure
with three separate steps.
In the rst step we examine all the geometries compatible with the given stoichiometry.
We determine the geometry which is the most symmetrical, ie., contains the greatest number
of mutually perpendicular mirror planes and C
2
axes. Second (using the general rules of MO
diagram construction) we determine the MO diagram for the most symmetrical geometry.
Finally, we consider how this MO diagram changes as this highest symmetry geometry
gradually transforms into a second less symmetrical geometry. This nal step is done with
a graph called a Walsh Diagram.
As Table 32.1 indicates, there are two possible geometries for AX
2
molecules, bent and
linear, illustrated in Figure 32.1. The more symmetric of these geometries is the linear
geometry. Applying the procedure indicated above, we therefore begin by determining the
linear AX
2
MO diagram.
290
32.3.1 Constructing the more symmetrical geometrys MO
We start by constructing the AX
2
MO diagram. This diagram diers from the organic MO
diagrams which we have constructed in the previous chapters in two important regards.
First, as we are interested in a very wide range of electron counts (ranging from 4 to 10
valence electrons), we will need to examine both the - as well as the -MOs. Secondly,
we will need to construct an MO diagram applicable to a wide range of A as well as X
atoms. Some simplications will be needed. Let us consider that the X atoms are hydrogen
atoms, and that the A atom is a second row main group element. The energies of the valence
attomic orbitals of these atoms is shown in Table 32.2
Table 32.2: Energies of hydrogen and second row main group AOs.
Element Valence s energy Valence p energy
H -13.6 eV
Li -5.4 eV -3.5 eV
Be -10 eV -6 eV
B -15.2 eV -8.5 eV
C -21.4 eV -11.4 eV
N -26.0 eV -13.4 eV
O -23.3 eV -14.8 eV
F -40.0 eV -18.1 eV
Ne -41.0 eV -21.6 eV
We wish to draw a generic MO diagram for the AH
2
system where A is a second main
group atom. As C is in the middle of this series, we choose a carbon atom to represent the
second row main group atom. According to the Table above, the energy of the H 1s orbital
is therefore initermediate between the C 2s and the C 2p energies.
32.3.2 The linear AX
2
MO diagram
We proceed to construct the higher symmetry linear AX
2
MO diagram. We assume that the
X atotms are hydrogen-like and that these hydrogen-like AOs start at an energy between
291
the A atom 2s and 2p energies. The linear AX
2
MO diagram is derived in Figure 32.2. As
this gure shows, in linear AX
2
, there are six MOs: two bonding, two non-bonding and two
antibonding.
32.3.3 The AX
2
Walsh Diagram
We now consider how these six MOs change as we gradually transform the geometry from
linear to bent. It is easiest to consider these changes in two separate steps. In the rst step
we consider the MO energy change assuming that there is no remixing of thte MOs. If due
to the gemetry change, the orbital looks more bonding, then the orbital goes down in energy,
and if it looks less bonding than before it goes up in energy. This rst step is shown in the
middle panel of Figure 32.3.
But the middle panel of Figure 32.3 shows that there is another signicant change in the
MO diagram in passing from the linear to the bent geometry. As this panel shows, one of
the mutually perpendicular mirror planes in the linear AX
2
molecule has been lost in the
bent shape. Accordingly, there are MOs which previously were of dierent symmetry labels
and which therefore could not interact which are now of the same label and can interact.
This mixing between orbitals which were previously of the wrong symmetry to mix but
are now of the correct symmetry to mix is generally quite important. In this chapter, we
focus on the most important of such mixings. This most important mixing will result from
two orbitals which previously could not mix, can under the new symmetry mix, and which
before mixing are close in energy and therefore can interact strongly. A study of the middle
panel of Figure 32.3 shows that the most important mixing is the one shown in Figure 32.4.
This energy change is the largest of all the energy changes in the MO diagram.
The resultant change to the bent AX
2
MO diagram is shown in the right panel of Fig-
ure 32.3. Our MO analysis of the AX
2
system is now complete.
292
Figure 32.2: Construction of the AX
2
MO diagram where A is a second row main group
element and X is hydrogen-like.
293
Figure 32.3: The change in the AX
2
MO diagram as the molecule transforms from thte linear
to the bent geometry.
294
Figure 32.4: The most important orbital mixing allowed by the symmettry change from the
linear to tthe bent AX
2
geometry.
32.3.4 The MO basis for the AX
2
VSEPR rules
An examination of Figure 32.3 will allow us to predict the MO-rationalized most stable
structure as a function of electron count. If the summed energies of the electrons increases
in going from the linear to the bent geometry, then the linear geometry is most stable. ANd
if it decreases, then the bent geometry is preferred.
We start with four valence electrons. For this number of valenece electrons, only the
lowest two MOs are lled. Turning to Figure 32.3, we see while there is no appreciable
change in the energy of the lowest energy MO, the second lowest goes up in energy in going
from the linear to bent geometries. Therefore for 4 valenece electrons, we would expect the
linear geometry to be preferred. This accords with the experimental results presented in
Table 32.1.
We now move to 6 valence electrons. For 6 valence electrons, the third lowest MO is also
lled. This third lowest MO is one of the two MOs which change most in energy. And as
this MO is lower in energy in the bent geometry, for 6 valence electrons, the preferred AX
2
295
geometry is bent. Again, the experimental ndings of Table 32.1 are in agreement with the
MO analysis.
The 8 valence electron picture is the same is the same as the 6 electron picture. The
fourth lled MO does not change in energy and therefore does not change the overall electron
energetics. 8 valenece electron AX
2
systems are therefore bent.
And nally we consider 10 valence electron systems. Under the linear to bent geometry
change, the fth from the bottom MO raises up in energy as much as the third MO lowers
in energy. The net eect of all ve MOs being lled is that the linear geometry is preferred.
Again, this is MO-based rationalization is in agreement with experiment.
32.4 The MO basis of other VSEPR systems
MO analyses exactly like the one shown above for the AX
2
geometries can be realized for all
other AX
n
systems. I will give you a problem in the next problem set which will allow you
to try your hand at deriving such MO-based rationalizations. The problem you will be given
is for the AX
3
stoichiometry. In doing this problem, please follow the procedure developed
for the AX
2
geometries. If you do so, you will be able to derive the relation between valence
electron count and AX
3
geometry.
296
Chapter 33
Hypervalent systems
A worthwhile test for a bonding model is that the model makes correct but otherwise unex-
pected predictions of molecular reactivity or shape. In this chapter we will examine a type
of inorganic compound where such predictions have been made and subsequently tested ex-
perimentally: hypervalent heavy main group molecules (molecules containing I, Br, Te, Se,
or Bi with supernumerary bonds). Such hypervalent systems have a surfeit of bonds, beyond
the number we normally expect main group compounds to have. We will discover that the
extra bonds in hypervalent systems are based on a Lewis acid - Lewis base interaction, these
Lewis acid and base interactions can be understood with ordinary MO theory, and this MO
theory has proven to be predictive.
We begin this chapter by reviewing some essential facts about main group compounds.
33.1 Lewis acids and bases
A Lewis base is a molecule with a high lying lled MO which is ready to share its electrons
with an empty orbital on a neighboring molecule. A Lewis acid is a molecule with a low lying
empty MO which is ready to share its empty orbital with a pair of electrons on a neighboring
297
molecule. Lewis acids naturally bind to Lewis bases. A classic example of a Lewis acid-Lewis
base interaction is that between BH
3
and NH
3
. The MOs of these two separate molecules is
shown in Figure 33.1. As is shown in Figure 33.2 the HOMO on the Lewis base NH
3
reacts
strongly with the LUMO on BH
3
in a Lewis acid-Lewis base manner to create the NH
3
BH
3
molecule.
Figure 33.1: BH
3
and NH
3
MO diagrams
In general the Lewis acid and base properties are governed by respectively the LUMO
298
Figure 33.2: BH
3
and NH
3
LUMO-HOMO Lewis acid-Lewis base interaction
and HOMO orbitals. Just as we saw for electrophiles, Lewis acids tend to have low energy
outward facing LUMO orbitals. Lewis bases, like nucleophiles, have high energy outward
pointing HOMOS. If we were to make a distinction between Lewis acid and bases vs. elec-
trophiles and nucleophiles, the former concepts are thermodynamic concepts, while the latter
are kinetic ones.
33.2 The octet rule
The next concept we need to refamilarize ourselves with is the octet rule. The rule is a
simple one. Among other things, it tells us about main group atom bonding. We turn rst
to atoms with a zero formal charge (if you do not know the formal charge concept please read
up on this concept in your 2150 textbook). The rule tells us the number of bonds bonded
to a given main group neutral atoms (in most stable compounds) depends on the column
299
of the periodic table to which the neutral element belongs. For elements in the 18th, 17th,
16th, 15th and 14th column of the periodic table the number of bonds is respectively zero,
one, two, three, or four.
If the formal charge is +1, +2, +3, then the atom has the same number of bonds as a
neutral atom lying in the column respectively one, two, or three columns to the left of it in
the periodic table. If the formal charge is -1, -2, -3, then the atom has the same number of
bonds as a neutral atom lying in the column respectively one, two, or three columns to the
right of it in the periodic table.
We can practice our ability to use the octet rule with the following questions.
1. How many bonds does a neutral carbon atom have in a stable compound?
2. How many bonds does a neutral uorine atom have in a stable compound?
3. How many bonds does a Si
2
atom have in a stable compound?
4. How many bonds does a neutral Bi atom have in a stable compound?
5. How many bonds does an O

atom have in a stable compound?


6. How many bonds does a N
+
have in a stable compound?
7. How many bonds does a Xe
2+
have in a stable compound?
8. How many bonds does a neutral tellurium atom have in a stable compound?
9. How many bonds does a neutral iodine atom have in a stable compound?
10. How many bonds does a I

have in a stable compound?


11. How many bonds does a neutral bismuth atom have in a stable compound?
300
The octet rule applies to compounds which are not metals, ie., do not conduct electricity.
Turning to the main group elements, see Figure 33.3, the octet rule applies most often to
those insulating main group elements in the upper right of the periodic table, and least
to those metallic elements on the lower left. For metals, ie., compounds which conduct
electricity well, the octet rule is rarely obeyed. Metals typically have more bonds in them
than would have been expected by the octet rule. Semiconducting elements, those lying
between the metals and the insulators vascillate between obeying and not obeying the octet
rule. In the latter case, extra bonds can appear.
Figure 33.3: The periodic table for the main group elements coded by the type of conductivity
the element exhibits. Metals are shown in orange, semiconductors in yellow, and insulators
in purple. Note the octet rule applies most often to the insulating main group elements.
301
33.3 Bonded vs. non-bonded atoms
The third concept we wish to familiarize ourselves with is the relation between chemical bonds
and interatomic distances. The two concepts are intrinsically related. Chemists can often
determine whether atoms are bonded to one another by examining the distances between
neighboring atoms. (Such data is typically derived from X-ray crystallographic studies.) For
example, in crystals, the observed distances between pairs of carbon atoms is as shown in
Figure 33.4.
Figure 33.4: The distances found between carbon atoms in the Cambridge Structure
Database, a compendium of all organic crystal structures
As this gure shows, all carbon-carbon distances are either less than 1.7

A or longer than
1.9

A (an

Angstrom, or

A, is 10
8
cm long). This divide can be used to separate pairs of
carbon atoms which are bonded to one another from those which are not. Only the former
are in C-to-C bonds. All main group atoms which lie in the upper right of the periodic
table have such clear separations between bonded vs. non-bonded distances. (In the case
of carbon-to-carbon bonds, the quantitative C-C distances can even be used to dierentiate
302
between C-C single, double, and triple bonds. These bond lengths are typically respectively
near 1.54, 1.34, and 1.22

A long.)
All insulating main group elements in the upper right of the periodic table show this
characteristic divide. Interatomic distances shorter than this divide always correspond to
bonds, distances longer than this value do not represent bonds. The situation however
becomes less clear as we approach the semiconducting elements. In Figure 33.5 we show
crystallographic data on Br-Br and I-I interatomic distances.
Bromine, an element already somewhat removed from the upper right of the periodic
table, shows the divide between bonded and non-bonded distances. However, close exami-
nation of the gure shows that there are a few Br-Br distances which lie in the divide itself.
Iodine, an element one step further removed from the upper right hand corner shows that
the divide lls in even further.
We know to call pairs of atoms below the divide as bonded atoms, and those longer than
thte divide non-bonded atoms, but how do we term those distance lying in the divide itself?
This chapter is devoted to answering this question: the answer, we will nd, is related to
the concept of hypervalency.
We leave this section noting that when we cross to the metallic main group elements, the
divide between bonded and non-bonded atoms disappears almost completely, see the data
for Ga-Ga distances in Figure 33.6. Metallic elements do not obey the octet rule; it proves
dicult for such elements even to determine which atoms are bonded to which.
33.4 The inert pair eect
We will also make brief reference to the inert pair eect, a well established rule of heavy
main group elements that their valence s orbitals are almost core-like in behavior. The rule
applies most clearly to the bottom row main group elements, Tl, Pb, Bi, and Po and to a
303
Figure 33.5: The distances found between (a) bromine and (b) iodine atoms in the Cambridge
Structure Database, a compendium of all organic crystal structures
304
Figure 33.6: The distances found between gallium atoms in the Cambridge Structure
Database, a compendium of all organic crystal structures
lesser extent to thesecond to bottom row elements In, Sn, Sb, Te and I. According to this
rule, these heavy elements highest occupied s orbital are nearly inert in character. The only
valence electrons belong to their partially lled p band Accordingly, in this chapter we will
consider the electronic conguration of neutral iodine as 5p
5
.
305
33.5 Hypervalent systems
33.5.1 I

3
The two most common iodine anions are I

and I

3
, see Figure 33.7. As this gure shows,
I

has no bonds, a result anticipated from the octet rule. Two dierent types of resonance
forms for I

3
are shown. The traditional Lewis structure for the ion is shown on the left. In
this resonance structure, the net bond order of the ion appears to be two, the central iodine
atom has a -1 charge, and the molecule is hypervalent. The two resonance forms on the
right portray this same ion as having a net bond order of one, the negative charge having
accumulated on the peripheral iodine atoms, and the octet rule being obeyed.
Figure 33.7: The I

and I

3
ions. Following the octet rule. I

, having the same number of


electrons as the noble gas Xe, is expected to have no bonds. I

3
is more complicated. Three
resonance forms of I

3
are drawn. The traditional Lewis structure for the ion is shown on
the left. In this resonance structure, the net bond order of the ion appears to be two, the
central iodine atom has a -1 charge, and the molecule is hypervalent. The two resonance
forms on the right portray this same ion as having a net bond order of one, the negative
charge having accumulated on the peripheral iodine atoms, and the octet rule being obeyed.
Although tradition favors the former resonance structure as being the most important, MO
theory conrms that the latter two forms are more dominant.
These two types of resonance structures portray the I

3
ion in very dierent light. In this
chapter we study the MO theory of this ion. Our MO theory will show us that the resonance
forms shown on the right of Figure 33.7 are the most signicant.
306
We start our MO study by considering the MO diagram of I

and I

3
separately, see
Figure 33.8. The former sysetm being just a monoatomic ion has an AO diagram for its
molecular orbitals; the latter system is the classic homoatomic diatomic MO diagram with
the further simplication that the only valence orbitals are 5p in character.
Figure 33.8: The MO diagrams for the isolated I

and I

3
ions. For both MO diagrams, we
assume that the valence electron conguration is comprosed solely of 5p electrons, see the
section on the inert pair eect. The former MO diagram is actually just an AO diagram
where all valence orbitals are lled. The I
2
MO diagram is a classic diatomic MO diagram,
The HOMO orbitals are

and the LUMO the

. Note that as I-I -bonds are much


stronger than I-I -bonds, there is a much smaller

gap than

one.
We now consider the reaction
I
2
+ I

3
.
Such a reaction is governed by its HOMO-LUMO interactions. The I

has no valence orbital


LUMO. Therefore the interaction between I

and I
2
is governed by the I

HOMOs and
the I
2
LUMO. In Figure 33.9 we consider these HOMO-LUMO interactions where the angle
307
formed between the I
2
and I

ion varies from 90


o
to 135
o
and nally to 180
o
. As seen in this
Figure, the HOMO-LUMO interaction is most favorable when the interatomic angle is 180
o
.
The crystallographic data supports this MO picture. In Figure 33.10 we plot all crystal-
lographic data for crystals which contain I

and I
2
in close proximity to one another. As this
gure illustrates, we will call the angle between the I

and I
2
units . BAsed on the crys-
tallographic data, this angle always proves to be equal or near 180
o
. WE can understand
this experimental result with our electronic structure theory: at 180
o
the HOMO-LUMO
interactions are greatest.
The crystallographic data tells a vivid story of the I
2
and I

interaction. In Figure 33.11


we plot the distance between the bonded pair of iodine atoms vs. the distance between the
I

and I
2
units As this gure shows as the distance between the two groups grows shorter,
the I
2
bond distance grows longer. The locus of points shown in Figure 33.11 describes the
reaction coordinate, the geometrical parameter which plots the course of a reaction. There
is a homework problem which explores this data further.
Note that in the I
2
and I

interaction, the former acts as a Lewis acid and the former as


a Lewis base.
308
Figure 33.9: The interaction between the LUMO of I
2
and the HOMO of I

for interatomic
angle of 90, 135 and 180
o
.
309
Figure 33.10: The Cambridge Structure Database of I
2
and I

interactions.
310
Figure 33.11: The reaction coordinate of the I
2
+ I

3
as revealed from crystal data.
311
33.6 The crystal structure of I
2
In comparing the elemental halogens chlorine, bromine, and iodine, the rst is a gas, the
second a liquid, and the third a solid. The dierence in their physical properties is a man-
ifestation of the dierent bonding propensities of these three elements. Chlorine belonging
to the upper right of the periodic table follows the octet rule closely, has just a single bond,
forms in the light weight molecule Cl
2
, and consequently at STP is a gas. By contrast, iodine
is much lower in the periodic table, follows only partially the octet rule, forms intermolecular
interactions in excess to those expected from the octet rule, and is therefore a solid.
The crystal structure of iodine is shown in Figure 33.12. As this gure illustrates, iodine
atoms are found at two diferent short distances with respect to one another. The shortest
I-I distances are 2.7

A long. Comparing this value with the interatomic I-I distances of
Figure 33.6, we conclude that these shortest I-I distances are bonds. But, as Figure 33.12
also shows, in the iodine crystal structure, pairs of iodine atoms are 3.5

A apart, a distance
which falls in the divide between bonded and non-bonded iodine atoms. As the gure shows,
it is these longer interatomic interactions which turns iodine from a relatively light weight I
2
molecules, which by itself would have been a gas, into an innite sheet of partially bonded
iodine atoms which is a fairly robust solid.
We examine the MO basis of these 3.5

A I-I distances. Figure 33.13 shows one such
interaction. It can be seen that this interaction connects two I
2
molecules in an L shape,
with two I-II angles, of 180 and 90
o
. This same gure also shows that this L shape optimizes
the HOMO-LUMO interaction between adjacent iodine molecules. Turning back to the atual
iodine crystal structure, we see tthe actual crystal structure optimizes the number of such
L-shaped units. Recalling that LUMOS correspond to Lewis acids and HOMOS to Lewis
bases, the interactions between iodine atoms are all of the Lewis acid-Lewis base type.
312
Figure 33.12: The crystal structure of I
2
.
Figure 33.13: The optimal HOMO-LUMO interaction between two I
2
molecules.
313
Chapter 34
Hard and soft acids and bases
34.1 Introduction
Over the past two weeks we have illustrated the ability of qualitative molecular orbital
theory to rationalize both the shapes of molecules and the course of reactions. From hydride
additions and electrophilic aromatic substitutions in organic chemistry, to the VSEPR rules
and hypervalent main group moities of inorganic chemistry, the examples have been chosen to
show molecular orbital theory at its most persuasive. In particular, in each case mentioned,
the covalent rather than the ionic bond has dominated the chemical bonds in question.
In this chapter, we will lay down a framework which goes beyond the covalent bond and
allows both covalent and ionic systems to be studied: the hard-soft theory of acids and bases.
Hard acids and bases are molecules and ions which are governed primarily by ionic forces.
Soft acids and bases, by contrast, have a more subtle bonding where the covalent bond and,
hence, molecular orbital theory plays a greater role. By the end of the chapter within the
framework of hard vs. soft acids and bases, we will have laid down the ground work for a
more complete picture of the chemical bond where both the ionic and covalent terms can
play a simultaneous role.
314
34.2 Hard vs. soft
The basic rule of hard-soft theory is that hard Lewis acids make strong bonds with hard Lewis
bases while soft Lewis acids make strong bonds with soft Lewis bases. By contrast hard acids
and soft bases or, alternatively, soft acids and hard bases make weak bonds.
Metal cations typically are Lewis acids while main group anions are generally Lewis bases.
In this chapter you will learn that Be
2+
and Hg
2+
are respectively hard and soft acids, while
F

and I

are hard and soft bases. The hard-soft principle will therefore tell you that Hg
2+
and Be
2+
make strong bonds with respectively I

and F

. It is therefore not a surprise that


for :
HgF
2
+ BeI
2
HgI
2
+ BeF
2
, H 400kJ.
We will delve more into the applications of hard-soft theory at the end of this chapter. First
however, we will learn which Lewis acids and bases are hard and which are soft.
34.3 Hard and soft bases
Fairly complete tables of hard and soft acids and bases are available in a number of inorganic
textbooks. Figure 34.1 shows one such table for bases.
Tables like the one shown in this gure are compendiums of a large amount of data. As
one goes on in chemistry, one sees more and more of such tables. I suspect that the initial
reaction of many students might be to gloss over the tables themselves or worse yet try to
memorize the tables contents. In my opinion, the best way of dealing with such tables is for
the student to examine them carefully and try to develop a scheme which makes sense of the
data contained within. Such a scheme may be dicult to obtain: it may involve guesswork,
an activity which many students do not enjoy.
Examining Figure 34.1, the rst question one might have is how the molecules and ions in
315
Figure 34.1: Hard, borderline, and soft Lewis bases.
the table are Lewis bases. A Lewis base is a chemical system with a pair of electrons which
can be shared with another molecule. In terms of MO theory, Lewis bases are frequently
governed by the HOMO, as long as this HOMO has readily accessible orbital lobes (in other
words, if the orbital lobes of the HOMO are either tiny or pointed in a way that an acid can
not reach them, such a HOMO would make a poor Lewis base.) In Figure 34.2 we draw out
a number of these Lewis bases highlighting the key basic electrons in red.
Examine Figure 34.1 and Figure 34.2 and try to develop a scheme which accounts for the
hardness and softness of bases.
316
Figure 34.2: Hard, borderline, and soft Lewis bases. The electron pairs responsible for the
basic character of the ion or molecule are highlighted in red. Before turning the page, try to
establish in your mind which periodic trends help classify the dierent types of Lewis bases.
317
In making sense of Figure 34.1 and Figure 34.2, we review the periodic trends of the
elements while searching for criteria which correlate well with the tables. My eye turns
to the electronegativities of the elements. The four most electronegative elements are, in
order, F, O, Cl, and N. Base pairs on these four very electronegative elements are found in
hard bases. From ammonia, NH
3
, to ethers, ROR, to phosphates (PO
3
4
), sulfates (SO
2
4
),
perchlorates (ClO

4
), uorides (F

), and chlorides (Cl

), hard bases all have their base


character located on one of the four most electronegative elements.
Similarly we can examine the soft bases. A review of the soft bases in these two gures
shows that it is the less electronegative elements, H, C, I, and S, which play a dominant role.
Hydrides (H

), benzene, cyanides (CN

), thioethers (RSR), and iodides (I

) all have their


basic character located on one of the less electronegative atoms.
Is electronegativity enough of a concept to make sense of hard and soft bases? After all,
there are many other elements in the periodic table. Bases centered on less electronegative
elements such as selenium, tellurium and phosphorus do prove to be soft, while even less
electronegative elements, boron, silicon, germanium, and the metallic elements are primarily
involved in Lewis acids and are not relevant to this question.
An examination of the borderline bases is instructive. As an exercise, please look at the
list of borderline bases in Figure 34.1 and Figure 34.2 and see what thoughts you can gather
from this list. Only after you have gathered your own thoughts would it be best for the
student to turn to the next page. The goal here is for the student to practice thinking their
way through data rather than to memorizing spoon-fed conclusions.
318
Lets compare answers: looking at the list of borderline bases, I nd only nitrogen-,
sulfur- and bromine-centered bases. Among our list of elements involved in hard bases, F,
O, Cl, and N, nitrogen is the least electronegative element. Among our list of elements
involved in soft bases, H, I, S, and C, sulfur is the most electronegative. Finally, bromine is
intermediate in electronegativity between the hard elements, Cl and N vs. the soft elements,
S and C. Electronegativity therefore does appear to play a signicant role in the selection of
borderline bases.
But electronegativity alone is not enough to sort out the borderline bases. Nitrogen-
centered bases such as NH
3
and N
2
H
4
are hard, while N

3
and NC
5
H
5
are borderline. Among
sulfur-centered bases, S
2
O
2
3
and SCN

are soft while SO


2
3
is borderline. As for both of
these groups of bases, it is the same two elements involved, it can not be electronegativity,
which is a constant, that solely plays a role in determining type-of-base character. What
factors are responsible for dierentiation amongst the sulfur- and nitrogen-centered bases?
If your analysis got you to a point such as this one, I think you did pretty well. Gathering
your thoughts of data into concrete form is a dicult and important scientic skill. The next
step of our exposition, however, involves a concept new to this course, polarizability. As it is
a new concept, I suspect, the material will go down clearest in a more traditional narrative
form.
34.3.1 Polarizability
Polarizability is both an atomic and molecular concept. Here, we will concern ourselves
with atomic polarizability. If an atom is placed in a static electric eld, the electron charge
distribution of the atom responds by shifting and developing a static electric dipole moment.
The plus end of the induced static dipole points towards the negative direction of the electric
eld, while the negative end of the induced static dipole points to the positive direction of
319
the electric eld. The polarizability of individual elements is both known and tabulated. In
Figure 34.3 we present atomic polarizabilty in graphical form.
Figure 34.3: The atomic ionization potentials, IP, in eV and in red, and the atomic polariz-
ability, , in 10
1
au and in black, as a function of atomic number.
In this gure, atomic polarizability, , is presented as a function of atomic number. A
larger value of implies that a larger dipole moment is induced in the element. The vertical
dotted lines in the gure correspond to the noble gases (ignore the capital Roman letters at
the top of the gure). The gure shows a number of periodic trends. In each period, the noble
gas and the alkali metal elements have respectively the lowest and highest polarizability. As
this gure demonstrates, this is the exact reverse of what is found for ionization potentials.
For any given column of the periodic table, the gure also shows that the heavier the
element, the greater is the polarizability. This trend is again the exact reverse of what is
found for ionization potentials. Combining both this and the previous periodic trend, we
nd the elements with the greatest polarizabilities are the heaviest alkali earths, Rb (element
37) and Cs (element 55).
320
We can further relate polarizability to both Z
eff
and electronegativity. If Z
eff
is low,
the electron is not tightly bound, orbital shapes distort more readily, and polarizability
is therefore high. Elements with low Z
eff
are less electronegative. Among the elements
involved in soft bases, iodine and tellurium, with atomic numbers 52 and 53 are the most
polarizable and are the least electronegative, while among hard bases, the elements F, O, N
and Cl, respectively atomic numbers 9,8,7, and 17, have the smallest polarizabilities and are
the most electronegative.
Polarizability diers from electronegativity in an important regard. The induced static
electric dipole is a measurable physical quantity not just for atoms but for molecules as
well: we can therefore measure molecular polarizabiliites and deduce how elemental polar-
izabilites are altered by the molecular environment in which the elements are placed. While
electronegativity is a specic number for a specic element, polarizabilities vary and can
delineate the roles of molecular bonding, oxidation states and molecular shape in the hard
vs. soft concept.
34.3.2 Polarizability and hard vs. soft bases
Highly polarizable and therefore low Z
eff
and low electronegativity elements make for soft
bases, while low polarizability, high Z
eff
and high electronegativity elements turn into hard
bases. The polarizability of an element changes depending on the molecule in which it is
placed. Three important changes are the following:
1. If an element is in a high positive oxidation state it has low polarizability. An ele-
ment with high positive oxidation state has high Z
eff
and hence a low polarizability.
Elements with negative oxidation states, by contrast, are more highly polarizable.
Oxidation of an element lowers polarizability, reduction of the same element raises
polarizability.
321
2. If an element is in a metal, it is essentially innitely polarizable. A metal placed in an
electric eld conducts electrons from the negative side of the electric eld to the positive
side. The electrons travel macroscopic distances in metals, irregardless of electric eld
strength. Conductive metals therefore have very large induced dipoles and very large
polarizabilities.
3. Elements involved in -bonding are more polarizable than elements with only -
bonding. Carbon in its two most stable forms, diamond and graphite, have markedly
dierent polarizabilities. Diamond, with only -bonds, is not very polarizable; graphite,
with -bonds, is a metal and is highly polarizability. The high polarizability of -
bonded systems is one of the important features of this bond type, playing a role in
reactivity, bond dipole moments, and solubilities.
We now apply these principles to sulfur- and nitrogen-centered bases. Among the sulfur
bases in these tables, S
2
O
2
3
and SCN

are soft while SO


2
3
is borderline. The oxidation
states of the sulfur atoms in S
2
O
2
3
, SCN

, and SO
2
3
are respectively +I, -I (assuming S
is more electronegative than C), and +IV. The sulfur atom in SO
2
3
is considerably more
oxidized than the other two sulfur atoms. This latter sulfur is therefore less polarizable and
its sulfur-centered base is therefore borderline rather than soft.
Among the nitrogen-centered bases, NH
3
and N
2
H
4
, are hard while N

3
and NC
5
H
5
are
borderline. The nitrogen oxidation states in these four molecules and ions are respectively -
III, -III, -I, and -III. Oxidation state therefore does not appear to play a role in the hardness or
softness of nitrogen bases. We note however, that the latter two nitrogen bases have nitrogen
-bonds. The latter bases are therefore more polarizable. The former purely -bonded
nitrogen bases are hard while the latter -bonded nitrogen-centered bases are borderline.
We see, in both the sulfur and nitrogen-centered cases, that the bases where the key atoms
are more polarizable are the softer bases.
322
34.4 Hard and soft acids
We now turn to hard and soft Lewis acids. As hard acids make strong bonds with hard bases
and soft acids combine well with soft bases, we may use our knowledge of hard vs. soft bases
to dierentiate hard from soft acids. In Figure 34.4 we compare the heats of formation of a
given metal with oxygen versus sulfur. O
2
and RO

are hard while S


2
and RS

are soft.
Hard acids will make stronger bonds to oxygen and soft acids bind better to sulfur.
Figure 34.4 plots the dierence between oxide and sulde heats of formation. The gure
convention is that positive values correspond to the oxide having a more negative heat of
formation, ie, the metal is hard; negative values in the gure indicate the metal is soft.
Figure 34.4: H of a metal oxide compared to the same metal but sulde. Plotted are the
dierences of these two H. If the plotted value is positive, the oxide has a more negative
heat of formation.
The four most oxophilic, ie., most oxygen favoring, are in order Al
3+
, Be
2+
, Si
4+
and B
3+
followed by Ti
4+
, H
+
, Mg
2+
and C
4+
. Hard cations are therefore highly oxidized light atoms.
323
Both high oxidation and small atomic number correlate with low polarizability. Figure 34.4
tells us that these eight low polarizability cations are all hard.
Among this list of eight hard cations, two deserve further comment. C
4+
, is lighter and
and has a higher oxidation state than many of the other cations on the list, but is the
least hard of the listed eight. Perhaps it is relevant that the stable carbon oxide is CO
2
,
a molecule with two C-O -bonds; such -bonds must lessen CO
2
s overall hardness. Also
worth mentioning is H
+
, the only hard cation on the list with only a single positive charge.
H
+
is an exceptional ion with no electrons associated to it; a true H
+
is just a proton with
essentially zero polarizability.
The most thiophilic, ie, sulfur favoring, of the metal cations are the heaviest main group
elements, Cs, Rb, and K as well as Au and the metal atoms nearest to it on the periodic
table, Ag
+
, Pt
2+
, and Hg
2+
. (Gold itself is not tabulated on this list as there is no known
gold-oxide and therefore no determinable dierence in oxide vs. sulde heat of formation.)
Each of these elements are highly polarizable. Recall that Au is the most electronegative of
all metal atoms. High electronegativity but especially high polarizability metal cations are
soft.
34.4.1 Fluorides vs. chlorides
As a test of the applicability of the hard-soft concept, we compare in Figure 34.5 the metal
heats of formation of the uorides versus the chlorides. Both F

and Cl

are hard, but


uorine, being the most electronegative of the elements, is the signicantly harder of the
two. The hard-soft concept would lead us to expect the dierences of uoride and chloride
heats of formation would mirror the previous oxide vs. sulde results. A comparison of
two gures conrms these expectations. The two graphs resemble one another. Fluorophilic
metal cations are the same as the previous oxophilic list; chloro-preferring metal cations are
324
the Au and its near neighbors, Ag and Hg as well as other Au neighbors, Tl(I), In(I) and
Cu(I).
Three further points are worth noting.
1. The highly polarizable K, Rb and Cs cations are chloro-preferring, but less so than
they were sulfur-preferring amongst oxide and sulde heats of formation. This fact
shows that the dierences of heats of formation between uorides to chlorides does not
mirror exactly the same as the heats of formation of oxides and suldes.
This suggests that the hard-soft concept sets a general order, not exact dierences, in
heats of formation. Hard vs. soft is best used as a qualitative rather than an exact
quantitative concept. Electronegativity, for example, appears to play a greater role in
determining softness for uorides vs. chloride preferences than for oxide vs. sulde
preferences. Electropositive K, Rb, and Cs are less chloro-preferring than they are
sulfur-preferring.
2. In Figure 34.5 a number of metal cations in two dierent oxidation states are listed:
Cu(I) vs. Cu(II); In(I) vs. In(III); and Fe(II) vs. Fe(III). In every case, the more
highly oxidized state of the given pair prefers the uoride more over the chloride. This
result agrees with the smaller polarizability and greater hardness of the more oxidized
state.
3. A few non-metallic cations are also included in this table, for example I
+
, Br
+
, and
Cl
+
. The more electronegative the non-metal cation is, the softer it is. Cl
+
has a
softness rivalling that of Au itself. This result corroborates the suggestion in the rst
item of the list: that elemental electronegativity plays a signicant role in uoride vs.
chloride softness.
325
Figure 34.5: H of a metal uoride compared to the same metal but chloride. Plotted are
the dierences of these two H. If the plotted value is positive, the uoride has a more
negative heat of formation.
326
34.5 Applications of the hard-soft concept
I give here a short list of examples of the utility of the hard-soft concept.
Benzene and water: Benzene is a soft base. Water is both a hard base (oxygen-centered)
or a hard acid (H(+I) centered). The benzene-water acid-base interaction is therefore weak.
Water and benzene are immiscible. By contrast, Ag
+
, one of the softest cations, readily
forms complexes with benzene and other -bonded carbon systems.
Metal salt solubilities: In Table 34.1 we list the solubilities of various metal salts in
water. We will come back to this table later in the course. For now, however, we show how
a number of the facts in this table can be understood with the hard-soft concept.
1. F(-I) is more electronegative and less polarizable than O(-II). Fluorides are therefore
harder than oxides. Ag(I) is a soft cation. It would therefore form stronger bonds with
water than with F

. As Table 34.1 shows, AgF is very soluble in water.


2. Iodide is a soft base. As Ag(I) is a soft acid, and water is a hard base, AgI is essentially
insoluble in water.
3. Li(I) is hard, I(-I) is soft, and water is hard. LiI is therefore highly soluble in water.
4. Cs is much softer than Li. We therefore expect the solubility of CsI to be smaller than
the solubility of CsCl.
Reactivity with hydrides In an earlier chapter in this book, we contrasted the reaction
rates of H

with a variety of organic electrophiles. MO theory successfully rationalized


the relative reaction rates of H

with C=C, C=O, and CN. In each of these cases, the


nucleophile, the Lewis base, is H

, while the electrophile, the Lewis acid, is centered on


327
Table 34.1: Solubility of metal salts in water at room temperature
compound solubility compound solubility compound solubility
AlCl
3
70 g AlF
3
0.6 g Al(SO
4
)
3
31 g
Al(OH)
3
ins AlI
3
6H
2
O very sol (NH
4
)
2
SO
4
71 g
(NH
4
)NO
3
118 g (NH
4
)
2
CO
3
H
2
O 100 g (NH
4
)CN very sol
(NH
4
)OH sol (NH
4
)I 154 g NH
4
F 100 g
BaBr
2
104 g BaCl
2
38 g BaF
2
0.1 g
BaCO
3
0.002 g Ba(NO
3
)
2
9 g Ba
3
(PO
4
)
2
ins
BaSO
4
0.0002 g Ba(OH)
2
8H
2
O 6 g Ba(CN)
2
80 g
BeCO
3
in base ins Be(NO
3
)
2
3H
2
O very sol BeSO
4
ins
BeF
2
CaI
2
209 g Ca(NO
3
)
2
121 g
CaCl
2
75 g CaF
2
0.002 g CaBr
2
142 g
CaSO
4
0.2 g Ca(OH)
2
0.2 g CaCO
3
0.001 g
CsF 367 g CsCl 162 g Cs
2
SO
4
167 g
CsI 44 g CsOH 396 g CsNO
3
9 g
CuF ins CuF
2
5 g Cu(NO
3
)
2
6H
2
O 244 g
CuCN ins Cu(CN)
2
ins Cu(OH)
2
ins
PbCl
2
1 g Pb(OH)
2
0.02 g PbF
2
0.06 g
PbCO
3
0.0001 g PbSO
4
0.004 g PbI
2
0.04 g
Pb(CN)
2
sl. sol Pb
3
(PO
4
)
2
0.00001 g Pb(NO
3
)
2
38 g
LiF 0.3 LiCl 64 g Li
2
SO
4
26 g
LiI 165 g Li
3
PO
4
0.04 g LiNO
3
90 g
AgF 182 g AgCl 0.00009 g Li
2
SO
4
26 g
AgI 10
7
AgCN 0.000024 g AgNO
3
122 g
Ag
2
CO
3
0.003 g Ag
2
SO
4
0.6 g TiCl
4
sol
Ti(SO
4
)
2
ins Sc
2
(SO
4
)
3
10 g Sc(NO
3
)
3
very sol
328
a -bonded carbon atom and is soft. Both hydrides and -bonded carbon atoms are soft.
Their interaction is a favored one.
The MO model we have developed in this course accounts for trends amongst soft acids
and bases. It did not account for the low reactivity of the RNO
2
group with the hydride
ion. The hard-soft concept suggests one reason why hydrides have a low rate of reactivity
with RNO
2
groups. Hydrides are soft while RNO
2
groups are borderline. These two groups
therefore have a lower anity for each other than would have been otherwise expected.
34.6 MO theory and the hard-soft concept
In applying MO theory to chemical systems, we focused our attention on softer systems.
1. In organic chemistry, we considered the reactivity of the soft bases: benzene, toluene,
and naphthelene.
2. Later this week, Roald Homann will further discuss the MO theory of soft -bonded
hydrocarbons.
3. We considered the crystal-packing preferences of the soft acid or base, iodine.
The qualititative MO theory we have developed in this course is particularily apt at
understanding soft base to soft acid interactions.
34.6.1 MO theory and hard-soft theory
In Figure 34.6 we contrast two pairs of HOMO-LUMO interactions. In the top portion of
this gure, MO theory makes a big deal of the of the HOMO-LUMO interaction (between
the H

HOMO and the C=O based LUMO). This rst HOMO-LUMO interaction involves
a high energy HOMO mixing with a low energy LUMO: the MO energy mixing term, E, is
329
large. MO theory, which studies the eect of the HOMO-LUMO interaction, has something
important to say about the mixing of these two chemical systems.
Figure 34.6: Strong and weak HOMO-LUMO interactions
By contrast, the second pair of HOMO-LUMO interactions, shown at the bottom of
Figure 34.6, is between a low energy HOMO and a high energy LUMO (the HOMO is
based on an F

ion, the LUMO on a Li


+
cation). As this gure shows, the HOMO-LUMO
330
interaction term, E, is small. The qualtitative MO theory taught in this course does not
speak to the Li
+
-F

interaction.
Hard-soft theory tells us that the soft H

and soft C=O interaction is energetically


signicant but so is the hard F

and hard Li
+
interaction. The former pairs interaction
involves the covalent bond; the second pairs interaction is ionic in nature. MO theory has
a great deal to say about the soft interaction, the covalent bond; Coulombic electrostatics
informs us about hard interactions, the ionic bond.
A full theory of the chemical bond will join MO theory, useful for soft systems, to the
ionic model, pertinent to hard systems.
34.7 Problems
A weekly problem set corresponds to solving the starred problems given in this section.
34.7.1 Solubility
1. Rationalize the relative solubility of ammonium sulfate and ammonium nitrate.
2. Rationalize the relative solubilty of aluminum uoride and aluminum chloride. Why
does aluminum iodide typically have water molecules in its stable crystal percipitate?
3. Rationalize the relative solubilty of lithium nitrate, lithium sulfate and lithium phos-
phate.
4. Without examining the table, predict the relative solubility of titanium sulfate, scan-
dium sulfate, ammonium sulfate and lithium sulfate. Compare your answer with the
actual data. Can you rationalize the actual data?
5. Without examining the table, predict the relative solubility of lead sulfate, lead nitrate,
and lead phosphate. Compare your answer with the actual data. Can you rationalize
the actual data?
6. Without examining the table, predict the relative solubility of silver nitrate and silver
carbonate. Compare your answer with the actual data. Can you rationalize the actual
data?
7. *Without examining the table, predict the relative solubility of silver uoride, silver
chloride, silver iodide, silver nitrate, and silver carbonate. Compare your answer with
the actual data. Can you rationalize the actual data?
331
8. Based on the silver cyanide data, is cyanide hard or soft? Can you use the results on
bond dipoles (discussed in the next section) to rationalize this observation?
9. Without examining the table, predict the relative solubility of lithium uoride, lithium
chloride and lithium iodide. Compare your answer with the actual data. Can you
rationalize the actual data?
10. Without examining the table, predict the relative solubility of cesium uoride, cesium
chloride and cesium iodide. Compare your answer with the actual data. Can you
rationalize the actual data?
11. *Without examining the table, predict the relative solubility of lithium uoride, cesium
uoride and silver uoride. Compare your answer with the actual data. Can you
rationalize the actual data?
12. Why might you think that the data for CuF solubility is incorrect?
13. Based on the data, do you think Ca
2+
should be considered hard or soft?
34.7.2 Dipole Moments
14. *List in order the ve largest bond dipole moments. What chemical feature do these
ve bonds have in common? Rationalize why these ve bonds have the largest bond
dipole moments.
15. Which bond has the larger bond dipole moment: C-S or C-O? Rationalize this fact.
16. Which bond has a larger dipole moment: C-N, C-O or C-F? Rationalize this fact.
17. *Rationalize the order of the dipole moments for C-N, C=N, and CN.
18. Rationalize the order of the dipole moments for N-O, P-S and P=S. Compare your
answer with the actual data.
19. *List six bond types which are among the hardest of the listed bond types. Does elec-
tronegativity appear to play a dominant role in the order of the bond dipole momentsn
(Ignore the B-O bond in answering this question)? Based on your analysis, what are
your thoughts on the accuracy of the B-O data? Is there any indication in the data
which suggests that the B-O data might be suspect?
20. *List ve bond types which are among the softest of the listed bond types. Does elec-
tronegativity appear to play a dominant role in the order of the bond dipole moments?
Does polarizability appear to play a dominant role?
21. *Are multiple bonded systems hard or soft?
332
Figure 34.7: Bond dipole moments. Bond dipole moments are given in Debye (D). 1 D is
the dipole moment of a proton and an electron separated by 1

A.
333
34.7.3 Boiling points
Table 34.2: boiling points
1
of various compounds and elements
compound boiling point compound boiling point compound boiling point
He 4 Ar 87 Kr 120
Ne 27 Xe 165 F 85
Cl 239 Br 332 Se 958
Te 1261 O 90 I 457
S 600 N 77 P 553
Cl 239 Br 332 water 370
As 887 Sb 1860 Si 3500
Ge 3100 C 5100 acetone 330
aluminum sulfate 2800 barium chloride 1800 cadmium nitrate 400
sodium chloride 1100 CO
2
200 Mg chloride 1400
aluminum chloride 500 barium uoride 2600 sodium uoride 2000
Mg uoride 2500 Mg oxide 4200 sodium oxide 2300
aluminum oxide 3300 hexane 342 decane 447
ethanol 352 methane 110 propane 230
butane 280 dodecane 490 HF 293
HCl 188 HBr 207 HI 238
ammonia 240 PH
3
180 BiH
3
290
SbH
3
250 H
2
S 213 H
2
Se 230
AsH
3
210 cyanogen 250 SnH
4
220
CF
4
140 CCl
4
350 SiH
4
150
CF
2
Cl
2
240 CF
3
Cl 190 CCl
3
F 297
GeH
4
180 CS
2
319 CSe
2
400
o-F
2
-benzene 360 m-F
2
-benzene 350 p-F
2
-benzene 360
o-Br
2
-benzene 490 m-Br
2
-benzene 490 p-Br
2
-benzene 490
o-(OH)
2
-benzene 510 m-(OH)
2
-benzene 450 p-(OH)
2
-benzene 600
butadiene 270 cyclobutene 270 1-butene 270
2-butene 280 formaldehyde 250 acetonitrile 350
1
Boiling points are given for 1 atm pressure and are given in degrees Kelvin.
22. *Look up the boiling points of He, Ne, Ar, Kr, and Xe in the provided table. Rationalize
the observed trend.
23. *Examine the boiling points of magnesium uoride, magnesium oxide, sodium oxide,
and sodium uoride. Rationalize the order of their boiling points.
24. The letters o-, m-, and p- in the rable refer to the ortho, meta and para arrangements
of atoms around a benzene ring. The o- arrangement is sometimes referred to as 1,2-;
334
m- is sometimes 1,3-; and p- is sometimes 1,4-. Please use the web to determine the
structures of o-diuorubenezne, m-diuorobenzene, and p-diuorobenzene. Which of
these molecules has no molecular dipole moment? Which of these molecules is dened
to be non-polar? For this data set, does the polarity of the molecule aect boiling
points?
25. Consider in the same way dichlorbenzene and dihydroxybenzene data. Does the po-
larity of the molecule aect boiling points?
26. Look up the structure of 1-butene and 2-butene. Which of these molecules might be
non-polar? Does molecular polarity aect boiling points?
27. Look up the structure of butadiene and cyclobutene. Rate in terms of softness butane,
cyclobutene, 1-butene, and butadiene. Does softness aect the boiling point? It turns
out that for second row compounds softness aects reactivity but not boiling points.
Our colleague, Roald Homann won the Nobel Prize for determining the role of softenss
in organic chemistry. But for the second row, the empirical fact is that softness does
not aect boiling points. By contrast in 3rd and higher row main group compounds,
softness aects reactivity and boiling points.
28. Do the boiling points of carbon dioxide and cyanogen suggest that for the second row,
softness does not control boiling points?
29. *Examine the boiling points of compounds CF
2
Cl
2
(freon), CF
3
Cl, CF
4
, CFCl
3
and
CCl
4
. Which of these molecules has no moleuclar dipole moment? Rationalize the fac-
tors controling boiling points. Does molecular polarity form part of your explanation?
30. *Rationalize the boiling points of HF, HCl, HBr, and HI.
31. *Rationalize the boiling points of uorine, chlorine, bromine, and iodine.
32. Rationalize the boiling points of uorine, oxygen, nitrogen, and carbon.
33. Rationalize the boiling points of water, H
2
S, and H
2
Se.
34. Rationalize the boiling points of oxygen, carbon dioxide, and aluminum oxide.
35. The molecule P
4
is found in white phosphorus. Please deduce its structure. (Remember
multiple bonds require 2nd row main group atoms).
36. Rationalize the boiling points of nitrogen, phosphorus, arsenic, and antimony. Hint:
which of these structures is most likely not to obey the octet rule?
37. *In comparing the boiling points of nitrogen to phosphorus versus uorine vs. bromine,
there is a much more dramatic increase in boiling points. Yet bromine is probably softer
than phosphorus. Rationalize this fact. Hint: Which of these elements form multiple
bonds?
335
38. *Rationalize the boiling points of ammonia, PH
3
and SbH
3
. Why are the trends in
boiling points so much less dramatic than the boiling points of P and Sb?
39. *The molecule aluminum chloride has a suprisingly low boiling point when compared
with magnesium chloride and sodium chloride. Two of these metals are coordinated
to six oxygen atoms, while the third can be as little as three-coordinate. Which metal
atom might be three-coordinate? (Hint: which compound can most easily make a
partially covalent bond?) Draw a possible structure of aluminum chloride. Rationalize
the low boiling point of aluminum chloride.
40. *Cadmium nitrate has an unusually low melting point for an ionic solid. Rationalize
this fact.
41. Rationalize why sulfur has a much higher boiling point than oxygen.
42. Rationalize the boiling points of iodine, tellurium and antimony.
43. Rationalize the boiling points of bromine, selenium, and arsenic. Guess what the
gaseous moelecular structure of arsenic is.
34.7.4 Comparing K
sp
44. *Without looking at a table, guess which K
sp
is smaller: CuI or CuBr?
45. Without looking at a table, guess which K
sp
is smaller: CuI or AgI?
46. Without looking at a table, guess which K
sp
is smaller: AgF or NaF?
47. Without looking at a table, guess which K
sp
is smaller: CuI or CuBr?
48. *Without looking at a table, guess which K
sp
is smaller: NiS or CoS?
49. *Why does the hard-soft concept tell us relatively little about metal-hydroxide K
sp
?
336
34.7.5 pK
a
Table 34.3: pK
a
1
of some common acids
compound pK
a
compound pK
a
compound pK
a
HI -10 HCl -8 H
2
SO
4
-3
HNO
3
-1.5 CH
3
CO
2
H 4.8 H
2
CO
3
6.4
H
2
S 7 HCO

3
10.3 H
2
O 15.7
HCCH 25 NH
3
36 CH
4
60
HCN 9.2 HF 3.2 H
3
PO
4
2.2
1
The smaller the value pK
a
is the stronger os the acid.
Negative values of pK
a
correspond to very strong acids.
50. *Rationalize the order of the pK
a
of HI, HBr and HF. Note these pk
a
refer to dissoci-
ation in an aqueous medium.
51. *Rationalize the order of the pK
a
of H
2
S and H
2
O.
52. *Rationalize the order of the pK
a
of HCCH and CH
4
.
34.7.6 Bond energies
Table 34.4: Bond energies
bond energy (kJ/mole) bond energy (kJ/mole) bond energy (kJ/mole)
Li-F -570 Li-Cl -460 Li-I -347
Na-F -477 Na-Cl -408 Na-I -204
Be-F -632 Be-Cl -461 Be-I -289
Ba-F -553 Ba-Cl -469 Ba-I -360
Ag-F -300 Ag-Cl -310 Ag-I -260
Au-F -305 Au-Cl -289 Au-I -230
Hg-F -268 Hg-Cl -225 Hg-I -146
1
Bond energies are taken from J. E. Huheey Inorganic Chemistry, second edition.
337
Table 34.5: Bond energies II
bond energy (kJ/mole) bond energy (kJ/mole) bond energy (kJ/mole)
Li-F -570 Li-Cl -460 Li-I -350
Be-F -632 Be-Cl -461 Be-I -289
Na-F -477 Na-Cl -408 Na-I -304
Mg-F -513 Mg-Cl -406 Mg-I -264
Al-F -580 Al-Cl -420 Al-I
K-F -490 K-Cl -423 K-I -330
Ca-F -550 Ca-Cl -429 Ca-I -330
Ti-F -586 Ti-Cl -440 Ti-I -296
1
Bond energies are taken from J. E. Huheey Inorganic Chemistry, second edition.
53. *For the reaction HgF
2
+ BeI
2
HgI
2
+ BeF
2
use the table of bond energies to cal-
culate the enthalpy of reaction. For want of better data assume the Be compounds
both have 4-coordinate Be, and the Hg compounds both have 2-coordinate Hg.
54. *Which bond is stronger Hg-F or Hg-I?
55. *How does the data in the bond energy table demonstrate the validity of the hard-soft
concept?
338
Chapter 35
Co(NH
3
)
3+
339
340
341
342
343
344
345
346
35.1 The d-orbital portion of the full MO diagram
In the previous sections of this chapter, we derived the MO diagram of Co(NH
3
)
3+
. The
HOMO were the three d non-bonding AOs, while the LUMO were the two metal-to-ligand
antibonding orbitals. In this model the HOMO are entirely located on the transition metal
atoms. But what about the LUMO: are they more ligand- or more metal-based? The LUMO
are antibonding metal-ligand orbitals. As the transition metal element is more electropositive
than the ligand main group elements, the metal-ligand antibonding orbital is primarily metal-
rather than ligand-based.
Both the HOMO and the LUMO are either entirely or principally located on the transition
metal. That the HOMO and LUMO are metal d-based is not just true for Co(NH
3
)
3+
, but
for many transition metal complexes. The d-electron counts, themselves, help demonstrate
this fact. Many transition metal complexes have a d
n
electron count where n is neither 0
or 10. Therefore many transition metal complexes have both lled and unlled d-orbitals.
Amongst the lled and unlled d-orbitals are often respecttively the HOMO and the LUMO.
Partially lled sets of d-orbitals (the set of d-orbitals is sometimes called the d-band) indicate
that d-orbital energies are centered on the HOMO and LUMO.
This situation is somewhat analogous to what we saw in planar organic molecule MO dia-
grams. While planar organic MO diagrams contain both - and -MOs, both the HOMO and
the LUMO are typically -type. In considering organic planar MO diagrams, we therefore
only considered the subset of the full MO diagram.
Transition metal MO diagrams follow a similar scenario. Often, instead of giving a full
MO diagram, chemists report only on the shape and energies of the MOs whose principal
components are d-orbitals. They do so with the understanding that they are reporting on
the HOMO and/or LUMO of the corresponding complex.
347
35.1.1 Deriving the d-orbital diagram
d-orbital diagrams bear, however, a signicant complication not found in organic -MO
diagrams. While planar molecules never mix - and -MOs, principally d-orbital MOs are
just that, they are principally of d-character, ie., they may, and often do, contain ligand
character as well. In deriving the principally d-character MOs, we need to consider the
ligand orbitals as well.
Luckily, a great deal of information about what this ligand character should be is con-
tained in the d-orbitals themselves. As we shall see in the next subsection, if we nd the
combination of ligand orbitals which interact best with a given d-orbital, not only will we
have found a combination of ligand orbitals which are of the right symmetry to interact with
the transition element, but we will have found the optimal combination. The variational
theorem suggests that this best seeming mixture will be an MO itself.
This approach only works, however, if the ve orbitals, so derived, all prove orthogonal
to one another. If they are orthogonal, they are typically the ve principally d-MOs (we
assume here that the only metal AOs are the ve d-orbitals).
35.1.2 Co(NH
3
)
3+
To illustrate the above approach, we consider Co(NH
3
)
3+
. We begin by drawing the ve
d-orbitals, indicating on the drawing the locations of the N atom in this complex. This
diagram is shown in Figure 35.1
We see that the xy, xz, and yz AOs can not -interact with the ammonia molecules.
These three orbitals are non-bonding. We now turn to the x
2
y
2
orbital. Four ammoia
molecules can interact in a -fashion with the x
2
y
2
orbital. As we are interested in the
principally d-MOs, and as N is more electronegative than Co, the principally x
2
y
2
MO
is antibonding in nature. It is shown on the left of Figure 35.2. Note that by maximizing
348
Figure 35.1: Co d-AOs in Co(NH
3
)
3+
with the location of the N atoms indicated with dots.
the ligand-metal antibonding character, we have successfully symmetry-adapted the ligand
orbital phases.
Figure 35.2: Principally x
2
y
2
and the 2z
2
x
2
y
2
MOs.
In exactly, the same way, we nd the principally 2z
2
x
2
y
2
orbital. It is shown on the
right panel of Figure 35.2. Note that we make the top and bottom ligand orbitals larger than
the other four ligand orbitals. We do so as the 2z
2
x
2
y
2
orbital has bigger lobes pointing
in these two directions. The most antibonding MO will therefore have bigger coecients on
these two orbitals.
We now examine if the x
2
y
2
and the 2z
2
x
2
y
2
MOs so generated are orthogonal to
one another. AOs being always orthogonal to one another, we need not consider the d-orbital
349
portion of these two MOs: we need only examine the ligand portions. In Figure 35.3, we list
the ligand portions of these two MOs as vectors. The 2z
2
x
2
y
2
top and bottom ligand
orbitals prove to be twice as big as the other four; the x
2
y
2
and 2z
2
x
2
y
2
MOs are
orthogonal to one another. These x
2
y
2
and the 2z
2
x
2
y
2
orbitals are therefore suitable
MOs. If so, the d-orbital portion of the MO diagram would be as shown in Figure 35.4.
Comparison to the full MO diagram, found previously, shows that this partial MO diagram
is correct.
Figure 35.3: Ligand portion of the principally x
2
y
2
and the 2z
2
x
2
y
2
MOs viewed as
vectors.
Figure 35.4: Principally d-MOs in Co(NH
3
)
3+
.
350
35.2 Advice for exam questions
This chapter has shown two ways to derive the principally d-orbital MO diagram. One can
either derive the full MO diagram and then extract from this full diagram the ve MOs
which are principally d-orbital in character, or one can adopt the much faster approach of
directly writing down the ve relevant MOs based on the individual d-orbital shapes. In
examinations, where time is limited, you are advised to use the latter approach (if asked to
derive the d-orbital portion of an MO diagram). If you do use this latter approach, make
sure that the MOs so derived are orthogonal. If they are not, you will have to revert to the
much longer rst way of deriving the diagram.
351
Chapter 36
The Cr(CO)
6
MO diagram
352
353
354
355
356
357
358
359
360
36.1 The d-only MO diagram
We apply the section given in the last section of the previous chapter to derive the d-only
portion of the Cr(CO)
6
MO diagram. Just as for the Co(NH
3
)
3+
MO diagram, we begin with
the ve d-orbitals, placed with the six ligands around them, Figure 35.1. We now consider
CO rather than NH
3
. The orbitals which are outward pointing and which are located on the
carbon atom are the lled -nonbonding and unlled -antibonding orbitals.
We now nd the ligand orbitals which optimally mix with the d-orbitals. -nonbonding
and -antibonding orbitals are respectively lower and higher in energy than the Cr d-AOs.
Therefore, the principally d MOs will be
respectively antibonding and bonding to the -nonbonding and -antibonding ligand
orbitals. This is shown in Figure 36.1. All ve orbtials so obtained are orthogonal to one
another. Comparison to the full MO diagram, shows that these ve orbitals are indeed MOs.
Figure 36.1: Principally d-MOs in Cr(CO)
6
.
361
Chapter 37
High-spin and low-spin compounds;
-donors, -donors, and -acceptors
The last two chapters have developed the MO theory of transition metal octahedral com-
pounds. For Co(NH
3
)
3+
and Cr(CO)
6
, we considered respectively a -bonding model and
a - and -bonding model. The two cases are contrasted in the left and middle panel of
Figure 37.1.
In both cases, the ve d-orbitals split into a high-lying set of two orbitals, the e
g
set, and
a low lying set of three orbitals, the t
2g
set. In both cases, low-lying lled ligand orbitals
interact in a -fashion with the d-based e
g
orbitals, raising the energy of these e
g
-orbitals.
We call both ammonia, NH
3
, and cabon monoxide, CO, -donors as both these ligands
have low-lying lled orbitals of the correct orientation and symmetry to share, ie, donate,
electrons in these orbitals with the transition metal in a -fashion (these -orientations are
illustrated in the gures of the previous two chapters).
Co(NH
3
)
3+
and Cr(CO)
6
dier, however, in their t
2g
orbitals. In the former case, their
t
2g
orbitals are non-bonding while in the latter case, the t
2g
orbitals interacted in a -fashion
with high-lying empty ligand orbitals. Such empty high-lying -interacting ligand orbitals
362
Figure 37.1: d orbital diagrams for Co(NH
3
)
3+
, Cr(CO)
6
and CoI
3
6
.
363
are the hallmark of a -acceptor. A -acceptor uses its high-lying empty -orbitals ligand
orbitals to bond with transition metal d-orbitals, and in so doing accepts electrons into these
previously empty ligand orbitals. As Cr(CO)
6
shows, carbon monoxide is a -acceptor. It
is also a -donor. Ammonia, as Co(NH
3
)
3+
shows, is just a -donor.
In the right panel of Figure 37.1, we show a third type of ligand-metal interaction, the
type found in CrI
3
6
. As this MO diagram shows, the transition metal d-orbitals interact
both with low lying lled iodine and orbitals. Iodine is therefore both a -donor and
a -donor. The -donating orbitals are just the same as those found in the Co(NH
3
)
3+
and Cr(CO)
6
cases. The -ligand orbitals have the same -overlaps to the transition metal
d-orbitals as was found for Cr(CO)
6
, but as the -ligand orbitals in the iodine ion, I

, are
low-lying in energy and are initially lled, the iodine ion ligand orbitals are -donating rather
tham -accepting.
37.1 Spectrochemical series
The three cases of Co(NH
3
)
3+
, Cr(CO)
6
, and CrI
3
6
exemplify the three types of common
transition metal-ligand interactions. From the perspective of the ligand, the ammonia, car-
bon monoxide and iodine ions are respectively a -donor; a -donor and a -acceptor; and
a - and donor. The other common ligands have also been assessed as to their donating
and accepting ability.
These results of this assessment are traditionally contained in the spectrochemical series.
This series is an empirical one which rates the ability of a ligand to create a large energy
splitting between the d-orbital t
2g
and e
g
sets. The ligands with the largest t
2g
to e
g
splittings
are -donors and -acceptors, those with the smallest t
2g
to e
g
splittings are ligands which
are -donors and -donors, and those with intermediate splittings are just -donors lie with
little or no -interactions. In the spectrochemcial series, carbon monoxide therefore has the
364
largest splitting, iodine the smallest, and ammonia lies somewhere imbetween.
A more complete rendition of the spectrochemical series is
CO CN

> NO

2
> en > NH
3
> H
2
O > OH

> F

> Cl

> Br

> I

Note the cyanide ion is just as strong a -acceptor as carbon monoxide, while the next-
strongest -donor is Br

. Much of the spectrochemical series can be undertood with MO


theory. But a nuanced understanding of the series requires the combined application of MO
theory, the ionic bond, and the hard-soft concept. This more nuanced treatment will be given
in the next chapter on Crystal Field Theory. In this chapter our focus will be in applying
the results of the spectrochemical seies to understand the physical properties of transition
metal complexes.
37.2 Transition metal oxidation states
Before going on to these physical properties of transition metal complexes, we need to briefy
discuss the transition metal atoms themselves. Transition metal atoms dier from the main
group elements (with which you are probably more familiar) in that, in stable complexes,
they can have a varied number of electrons associated to them. While the carbon in stable
organic compounds always obeys the octet rule, transition metal atoms, often do not follow
any xed electron count (please note however, in organometallic chemistry chemists study
a subset of transition metal compounds where transition elements do follow a xed electron
count rule).
In order to determine the number of electrons associated to the transition metal, it is
best to know the oxidation state of the metal atom. A table of some practical importance is
the most common oxidation states for the rst row transition metals, see Table 37.1. Also
365
included in this table are the d-electron counts associated to these oxidation states as well
as the typical complexes which surround each rst row transition element in water. Besides
thte oxidatiton states listed in this table, the three most important addititonal oxidation
states are Fe(II), Co(III), and Cu(II).
Table 37.1: Common oxidation states for 1
st
row transi-
tion elements
Common oxidation state Aqueous complex d-count
Sc(III) Sc
3+
(H
2
O)
6
d
0
Ti(IV) Ti
4+
(H
2
O)
6
d
0
V(V) V
5+
(H
2
O)
6
d
0
Cr(III) Cr
3+
(H
2
O)
6
d
3
Mn(II) Mn
2+
(H
2
O)
6
d
5
Fe(III) Fe
3+
(H
2
O)
6
d
5
Co(II) Co
2+
(H
2
O)
6
d
7
Ni(II) Ni
2+
(H
2
O)
6
d
8
Cu(II) Cu
2+
(H
2
O)
6
d
9
Zn(II) Zn
2+
(H
2
O)
6
d
10
First row transition elements are much more common than their second and third row
equivalents. As a pracitical matter, chemists tend to learn the common oxidation states of
this row rst. And while you as a student may want to learn these oxidation states, you
should note there are two very dierent sources for these common oxidation states. On
the one hand, metal atoms are inately more stable in some oxidation states over another.
Scandium is genuinely more stable in the 3+ state than in the 2+ or 1+ states. But there
is a second reason behind the common oxidation states of transition metal elements.
While the Earth is not in chemical equilibrium, it is not in tremendous disequilibrium
366
either. Transition metal atoms found in geological samples tend therefore to be in pseudo-
chemical equilibrium with one another. Most iron, cobalt or manganese atoms found in the
Earth tend to be in respectively the 3+, 2+, and 2+ oxidation states because these elements
in these oxidation states tend to neither oxidize or reduce each other. Geologically speaking,
titanium can not be in the 3+ oxidation state while iron is in the 3+ state as it would be
likely that over time that these Ti(III) ions would transfer electrons to the Fe(III) ions.
If we were on another planet and you were taking a freshman chemistry course there, you
might very well be told that the most common oxidation state of scandium is Sc(II), but the
iron might be Fe(II) or alternatively the Co might be Co(III). (One of the elements on Earth
found in the greatest oxidation state disequilibrium is oxygen itself. WHile most oxygen on
Earth is in the O(-II) state, atmospheric oxygen is neutral. In fact were the conditions of the
Earth just determined by geological forces, probably there would be no atmospheric oxygen.
Such oxygen is a by-product of life. Life is similarly responsible for much of the terrestial
reduced nitrogen and carbon.)
The important point for the remainder of this chapter is the right-most column of Ta-
ble 37.1. Dierent rst row transition metal atoms have genuinely dierent d-electron counts.
These counts range all the way from d
0
to d
10
. It should be further noted that chemists are
not bound by terrestial oxidation requirements. Each of the transition elements have been
incorporated into compounds with an extremely wide range of oxidation states. Even an ele-
ment such as Sc has been prepared in oxidation states other than its most common oxidation
state.
37.3 High vs. low spin compounds
We now combine our knowledge of the spectrochemical series with that of transition metal
oxidation states. Our interest will be in octahedral complexes and the splitting between the
367
t
2g
and e
g
orbitals. Ligands on the left of the specrtochemical series such as carbon monoxide
and cyanide cause large t
2g
-to-e
g
splittings, while those on the right, ions such as iodide and
bromide, cause small splittings.
Dierent transition metal ions cause dierent splitttings. A metal atom in a higher
oxidation state tends to cause a bigger splitting than the same element in a lower oxidation
state. Second and third row transition metal atoms have much bigger splittings than their
rst row equivalents.
In Figure 37.2 we examine the consequence of a large vs. a small t
2g
-to-e
g
splitting. We
choose for illustrative purposes a Mn(III) ion. Mn(III) has a d
4
electron count. As this
gure shows, when the splitting is zero there are four unpaired electrons As the splitting
becomes very large, there are only two unpaired electrons. We see there are two opposing
forces behinfd the electron-lling. In the formwe case, there is Hunds rule which favors the
maximum number of unpaired spins. In the latter case, there is thte Aufbau principle which
favors the occupation of the lowest energy orbitals.
Figure 37.2: Mn(III), d
4
, in an octahedral complex for dierent E splittings.
The central region of Figure 37.2 shows two cases. On the left side of the central region,
the t
2g
-to-e
g
splitting is still small enough that the Hunds rule forces still dominate. On the
368
right side of the center of this gure, the splitting is large enough to dominate over Hunds
rule. The former case is called high spin, while the latter is low spin.
369
37.4 Physical properties
370
371
372
373
Chapter 38
Crystal Field Theory
In the previous three chapters, we have given an MO-based account of transition metal
compounds. We have seen that MO theory can be used to explain many of the properties of
these compounds. But at the same time, we know from hard-soft theory that such a theory
is not ideally suited for all transition metal compounds equally. MO theory, of the type
considered in this course, works best with low oxidation state, electronegative metal atoms.
It is not particularily well suited for electropositive, high oxidation state transition metals.
Qualitative MO theory works best for soft rather than hard metal atoms.
This is important as many of our most important transition metals, for example Ti
4+
,
V
3+
, or Cr
3+
, are hard as are many of our most important ligands. H
2
O, for example,
is distinctly hard. In hard systems, such as Ti(H
2
O)
4+
6
or Na
+
(aq) MO theory, which is
based on the overlap and mixing of atomic orbitals, does not play a large energetic role. An
electrostatic model would appear to be more appropriate.
In this chapter, we develop a simple electrostatic picture of hard metal to hard ligand
interactions. We concentrate on octahedral complexes. Not only is the octahedral complex
the most common one, electrostatic or ionic forces, themselves, are often responsible for the
low energy of the octahedral geometry.
374
38.1 The ionic model
In Figure 38.1 we illustrate M(H
2
O)
n+
6
. The metal ion is coordinated to six water molecules.
We have discussed in previous chapters the MO theory of this type of complex. In the limit
of a very electropositive metal and very electronegative ligands such MO energetics need not
dominate (see the nal section of the chapter on hard-soft theory for an explanation of this
point).
Figure 38.1: An octahedral ionic model to describe a metal-octahedral complex.
That the MO energetics are not all important does not mean that the bonds are weak.
Instead it means that other bonding forces come to a fore. The chief of these are ionic
forces. In the ionic model we associate with the metal atom a net positive charge and with
the electronegative ligands a net negative charge. These charges are governed by the simple
electrostatic equation
E
q
+
q

r
,
where r is the distance between the cation and the anion.
This equation tells us:
1. The higher the cation or anion charge, the larger the ionic attraction.
375
2. The smaller the cation or anion, the smaller the distance between the two, and the
stronger the ionic bond.
3. Variations in charge, being generally bigger than variations in bond length, tend to
play a stronger role in the variation of ionic bond strength.
38.2 Verication of the electrostatic model
How useful is this ionic model? Can it be used to understand features of transition metal
energetics? To answer these questions let us look at metals to the upper left of the periodic
table. Such metal atoms are the hardest and therefore should accord best with the ionic
model.
The best data would be data which explicitly told us metal-anion bond strength. We saw
some data of this type in the chapter on hard-soft theory. But it turns out while reference
books carry tables of metal-halide bond strengths, there is suprisingly little tabulated data
on metal-oxygen or metal-nitrogen bond energies.
In this chapter we therfore consider data which would logically be governed by the
strength of the metal-anion bond, but which does not give us a numerical value for these
bond stregnths. We consider two types of data. The rst is the melting points of compounds
of these early metal atoms together with one of the four most electronegative main group
elements: F, O, Cl and N. The second is the rate of exchange of water molecules bound in
aqueous solution to a metal. Both the melting point and water exchange data speaks to the
strength of the metal-ligand bond.
376
38.2.1 Melting point data
Melting point data is useful, but at times convoluted. We start our discussion with its
problems.
1. While melting does generally require the rupturing of bonds, and hence melting points
do provide a measure of bond strength, in some cases molecular species can form. Such
molecular species can melt without the rupturing of bonds. The coordination number
of the metal is key in rationalizing the forming of molecular metal-ligand complexes:
if the coordination number of the metal can match the metal to ligand stoichiometry,
molecular species can form.
For a (non-metallic) example consider CO
2
. The carbon atom can be two coordiante
and the stoichiometry is 1:2. A stable molecule, O=C=O results. CO
2
when it sublimes
does not rupture any of its primary bonds: CO
2
sublimes at low temperature. The
sublimation temperature, or its melting point (under pressure) are not measures as to
the strength of the C=O bond.
Among electropositive metal atoms there is a tremendous variation in coordination
numbers. For uorides, oxides and nitrides, while most metal atoms are six coordinate,
the smallest metal atoms, Be, Zn, Cu, Ga, and Al are, dependinig on the metal, either
often or occasionally four coordinate. Similarly the largest metal atoms, K,Ca, Rb,
Sr, Cs, Ba, La (the lanthanides in general as well as the elements near them), U, and
Pb are often more than six coordinate. For chlorides, lower coordination numbers are
frequently found: chlorine is signicantly larger than uorine.
2. Not all compounds melt. Many decompose or sublime. Decomposition and submli-
mation temperatures can be signicantly below what the melting temperature might
have been had such a process not been stopped by these alternate processes.
377
3. Not all melting point data are tabulated in common data sources.
With these provisos, we can now examine melting points of early metal uorides, chlo-
rides, oxides, and nitrides. These melting points are listed in Table 38.1.
Table 38.1: Melting points of various compounds in degrees Celsius
ele- uoride chloride oxide nitride
ment melting point melting point melting point melting point
Li 845 605 1700 850(dec)
1
Be 800(sub)
2
405 2530 2200
B -127 -107 450 3000
Na 993 801 1275(sub) 300(dec)
Mg 1261 714 2852 800 (dec)
Al 1291(sub) 190 2015 2200(in N
2
)
K 858 770 350(dec) (dec)
Ca 1423 782 2614 1195
Sc
Ti(IV) <300 (sub) -25 1840
Ga 800(sub in N
2
) 164 >660 800(sub)
1
(dec)=decomposes
2
(sub)=sublimes
In this table, we see that a number of the melting points are extremely low: BF
3
, BCl
3
,
AlCl
3
, GaCl
3
and TiCl
4
. In each of these cases, and at these stoichiometries, a stable molecule
exists. The low melting point is therefore an indication of the molecularity of the species, not
the weakness of the metal-ligand bond. Boron being capable of three- and four-coordinate
geometries, I suspect that for B
2
O
3
a molecular species may also arise at slightly elevated
temperatures.
Eliminating this data from consideration, we see for the same metal that oxides and
nitrides generally have decidely higher melting points than uorides or chlorides. The ionic
model accounts for this as oxygen and nitrogen have respectively twice and thrice the negative
charge of halogens. Nitrides however do not always have higher melting points than oxides.
Nitride melting points are complicated by the redox reaction in which metal plus N
2
is
378
produced at high temperature.
For uorides and oxides, +2 cations have higher melting points than +1 cations. We do
not have enough data for nitrides to draw any conclusion, while for chlorides, the ionic trend
is not observed. Perhaps chlorine, a neighbor of sulfur in the periodic table, is soft enough
that the ionic model does not hold sway. (In the table of hard bases, uorine and oxygen
are denitely hard, while chlorine and nitrogen are only partially placed in the hard column.
Perhaps of interest here, chlorine and bromine have the same crystal structure as soft iodine,
but crystalline uorine has a dierent structure.)
In Table 38.2 we give an abbreviated table in which all decomposition temperatures have
been removed, sublimation temperatures are listed as lower bounds of melting points, all
chloride data has been removed, and molecular melting temperatures are not listed. The
ionic model appears to have relevance to this ltered data. Oxides melt at considerably
higher temperature than uorides. Interestingly, oxides melt at roughly double the uorides
temperatures, a qualitiative prediction of the ionic model (oxides are in the -II oxidation
state while uorine is in the -I state.). Nitrides, which contain nitrogen in the -III state, are
high melting, but they do not quantitatively t in the same way.
Table 38.2: Melting points of uorides, oxides and ni-
trides, in degrees Celsius
ele- uoride oxide nitride
ment melting point melting point melting point
Li 845 1700
Be >800 2530 2200
Na 993 >1275
Mg 1261 2852
Al >1291 2015 2200(in N
2
)
K 858
Ca 1423 2614 1195
Ti(IV) 1840
379
Transition metal uorides, oxides and nitrides
In Table 38.3 we list the melting points of the transition metal ions uorides, oxides and
nitrides in the +II and +III oxidation state.
Table 38.3: Transiton metal compound melting points,
o
C
ele- uoride oxide nitride
ment melting point melting point melting point
Ti(II) 1750
Ti(III) 1200 2130 (dec) 2930
V(II)
V(III) >800 1970 2320
Cr(II) 1100
Cr(III) >1000 2270 1700 (dec)
Mn(II) 856 1564
Mn(III)
Fe(II) >1000 1369
Fe(III) >1000 1565
Co(II) 1200 1795
Co(III) 895(dec)
Oxides melt at higher temperatures than uorides. Among the earlier more electropos-
itive metal atoms, oxide melting points are double that of uorides. The later transition
metals, iron and cobalt, have oxide melting points which are higher but are not double the
uoride melting temperatures. Nitrides, where they are stable at high temperature, are high
melting. Only titanium(III) nitride, however appears to semi-quantitatively t the ionic
model.
Early electropositive transition metal oxides, uorides and perhaps nitrides appear to
be goverened by the ionic model; later transition metals might well be responding to other
bonding forces. Such a picture ts in with the hard-soft theory.
380
38.2.2 Water exchange data
In this section we consider Manfred Eigens water exchange rate constant data, experimental
work which won him the Nobel prize. While the prize was won, in part, for the development
of new rate measuring techniques, we concentrate here on the data itself.
The rate constant in question is the rate constant for the reaction:
M
n+
(H
2
O)
m
+ H
2
O

M
n+
(H
2
O)
m1
(H
2
O

) + H
2
O
This rate constant is an experimental measure of the strength of the metal-to-water bond.
Before presenting the data, we list diculties with data interpretation.
1. We can imagine two dierent limiting transition states or intermediates. In the rst
limiting case, a water molecule leaves before the new water molecule joins the complex.
In the second extreme, the new water molecule joins before the vacating water molecule
leaves the complex. In either limit, the rate constant is a measure of the metal-water
bond strength. If all metal atoms were to follow the same reaction pathway, we would
have a reasonable qualitative measure of bond strengths. But if dierent metal atoms
adopt dierent reaction pathways, we will be adding to our picture the dierence in
pathway energies.
2. The number of water molecules in the complex varies. While typically m=6, the
smallest metal atoms, Be, Zn, Cu, Ga, and Al are often four coordinate while the
largest metal atoms, K,Ca, Rb, Sr, Cs, Ba, La (the lanthanides in general as well
as the elements near them), U, and Pb are often more than six coordinate. (These
coordination number rules are quite general and were also discussed in the section on
melting point data.)
The diculty is that as the metal-ligand coordination number becomes smaller, metal-
381
ligand bond distances shrink, and metal-ligand bonds become stronger. Four-coordinate
metal complexes will have smaller exchange rate constants than six-coordinate ones.
It is therefore generally best, when examining water exchange data, to try to compare
two metal atoms which have roughly the same coodination number.
3. Included in the water exchange data is a great deal of transition metal data. Tran-
sition metal exchange rates can, as we will discuss below, depend on the number of
d-electrons. As we are examining the ecacity of the ionic model, where up to now the
d-orbitals have not been considered, it is best for now to limit our examination to d
0
,
high-spin d
5
, and d
10
species. In all three of these cases d orbital eects are minimized.
With these provisos, we examine the data, Figure 38.2. We consider here the alkaline
metal, alkaline earth, rare earth, main group, Zn
2+
, and Mn
2+
. These last two metals are
respectively d
10
, and d
5
. (We omit Cd
2+
, In
3+
, and Hg
2+
being all soft metals.)
We observe:
1. We examine rst the octahedral complexes: Na
+
, Mg
2+
, Mn
2+
, and possibly Al
3+
,
Ga
3+
, and Zn
2+
are octahedrally coordinated. The table shows clearly that monovalent
metal atoms exchange fastest, while trivalent cations exchange slowest. The dierence
in exchange rates between Na
+
and Al
3+
is 10
9
, an impressively large dierence. This
is in keeping with the ionic model.
2. In any column of the periodic table, the smallest metals exchange slowest while the
largest metals exchange fastest. The rst two rows of data in the gure give the full
alkali metal and alkaline earth series of rate constants. This is in keeping with the
ionic model.
3. Amongst the lanthanides, where there is little chemical dierence between one metal
and the next, the heavier lanthanide, which is generally the smaller ion, has the slowest
382
Figure 38.2: Manfred Eigens water exchange data.
383
exchange rate. This is in keeping with the ionic model.
4. Based on points 1-3 above and the data itself, size eects play a smaller role than
charge eects, again this is in keeping with the ionic model,
We conclude that the ionic model rationalizes hard-metal water exchange constants.
38.3 The ionic model of crystal eld theory
In this section you will be presented with Crystal Field Theory, an ionic model of the tran-
sition metal-ligand interaction rst developed by Hans Bethe here at Cornell. Bethes work
predates the development of MO theory and relies solely on the ionic interaction.
38.3.1 A point charge model for ligands
From the previous sections, we suspect the ionic model might rationalize the framework of
hard-metal uorine, oxygen, and perhaps nitrogen bonds. We now amplify this model to
transition metals in general. We return to Figure 38.1, the M(H
2
O)
n+
6
complex. In this
complex the transition metal is positively charged while the oxygen atoms bound to the
water molecule are negatively charged. This complex, viewed as an ionic entitiy, adopts the
form shown on the right of Figure 38.1.
Consider now Mn(H
2
O)
2+
6
, where each of the d-orbitals is singly occupied. The transition
element in the center of the complex continues to have valence d-orbitals and d-electrons.
Each of these d-electrons will experience the electrostatic eld of the ligand ions. Thus each
of these d-electrons has its own potentially dierent electrostatic energy. For us to estimate
the dierent d-orbital energies, we need to calculate the dierences in electrostatic energies,
we need to compare the d
xy
, d
x
2
y
2 and d
2z
2
x
2
y
2 environments, see Figure 38.3.
384
Figure 38.3: d-orbitals and , their electron densities. Red dots indicate ligand lobes repre-
sented as point charges.
385
In Figure 38.3 we illustrate the electron densities of electrons in each of these orbitals.
The electron densities shown are the square of the wave functions. Also illustrated in this
picture are the six ligands, each represented as a point charge along one of the Cartesian
axes. The electrostatic energy of an electron in any of these d-orbitals is the sum of the
electrostatic energy of the electron density at each point in space interacting with the six
ligand charges. This sum is a 3-D integral, of a type which we will not solve in this course.
Instead, we will estimate the values of these integrals based on pictures. For example,
in Figure 38.3, we see that the d
x
2
y
2 and d
2z
2
x
2
y
2 would have higher electrostatic energy
than the d
xy
as these two orbitals have lobes directly pointed to the six ligands while the
d
xy
orbital has lobes pointing between the ligands. The lobes being closer to the ligands
in the former two orbitals, their electrostatic energy is higher. Furthermore, the d
xz
, and
d
yz
orbitals, by symmetry, have the same point charge environment as the d
xy
: therefore
electrons in these two orbitals have the same electrostatic energy as one in the d
xy
orbital.
We are close to the nal answer as to the electrostatic energies of electrons in the ve d-
orbitals. To complete this answer we need to compare the electrostatic energy of a d
2z
2
x
2
y
2
and a d
x
2
y
2 electron. To do so note that 2z
2
x
2
y
2
= (z
2
x
2
) +(z
2
y
2
). This equality
is illustrated in Figure 38.4. Let us make the approximation that there is no overlap between
dierently oriented lobes in the (z
2
x
2
) and (z
2
y
2
) orbitals. If this is so then the electron
density of the 2z
2
x
2
y
2
can be approximated as shown in Figure 38.4, four smaller lobes
and two larger lobes. As the wave function of these larger lobes is two times that of the
origianl clover-leaf orbitals, the electron density of these larger lobes is four times that of
the clover-leaf orbital.
We now calculate the energies of each of these lobes with the six ligand point charges.
As Figure 38.5, we can associate with the six lobes, six separate charges. These point
charges are just like the point charges used to describe the six ligands (ie., points are used to
represent orbital lobes). In Figure 38.5 we normalize the six lobes so that the overall charge
386
Figure 38.4: A simplied picture of the electron density in the d
2z
2
x
2
y
2 orbital
is one.
In Figure 38.6 we calculate the electrostatic energies of the d
x
2
y
2 and d
2z
2
x
2
y
2 electrons.
They prove to have the same electrostatic energy. Our approximate answer ia actually exactly
right, electrons in the d
x
2
y
2 and d
2z
2
x
2
y
2 have the same electrostatic energy.
In Figure 38.7 we show the d-orbital energies. We see that the electrostatic model gives
the same splitting pattern as does MO theory. This is important. As transition metal-
ligand bonds are a combination of electrostatic and MO forces, the true bond energies are a
combination of both the electrostatic and MO terms. In this case, both MO theory and the
ionic model give the same splitting pattern. We therefore expect transition metals to adopt
the calculated d-orbital diagrams.
387
Figure 38.5: A simplied picture of the electron density in the d
2z
2
x
2
y
2 and d
x
2
y
2 orbitals.
Point charges represent similar-shaped orbital lobes.
Figure 38.6: Calculation suggesting that the d
z
2 and d
x
2
y
2 orbitals have the same electro-
static energy. Blue dots are the d-electron point charge description. The red dots indicate
ligand charges.
388
Figure 38.7: The electrostatic model of the d-orbitals in an octahedral complex. This is
generally called the crystal eld model. The splitting between the top two and the lower
three orbitals is typically called 10D
q
.
38.4 The spectrochemical series
In the MO model, we rationalized that the spectrochemical series should proceed from
strongest to weakest eld in the order:
CO PR
3
CN

>> NH
3
> H
2
O > F

.
For the ionic model, from our work on melting point data, we expect that the spectrochemical
series goes in the order:
N(III) > O(II) > F(I).
We apply both models to rationalize the full spectrochemical series:
CO PR
3
CN

>> NH
3
> H
2
O > F

> Cl

> Br

> I

.
Starting from the left, we look at the three strongest splitting ligands: CO, PR
3
and CN

.
All three are soft, MO theory is therefore relevant: they are all -donors and -acceptors.
The ligand -donating orbitals push up the z
2
and the x
2
y
2
energies, while the ligand
389
-accepting orbitals push down the energy of the xy, xz, and yz orbitals. (The former two
orbitals are often called the e
g
-set while the latter three are the t
2g
-set.) CO, PR
3
and CN

therefore have large splittings associated with them.


In the middle of the spectrochemical series, we have NH
3
, H
2
O and F

. Both MO theory
and the ionic model can be applied to these ligands, although as these ligands are hard, we
might expect an ionic model is best suited in their treatment. MO theory places a weaker
interaction between the electropositive metal and electronegative F orbitals and hence a
weaker splitting. Nitrides, being higher melting than oxides, which in turn are higher melting
than uorides, the ionic model suggests that NH
3
has the strongest interaction, followed by
H
2
O and nally F

. The electrostatic splitting is therefore in the order NH


3
> H
2
O > F

.
Both the MO picture and the ionic picture have the same ordering of these three ligands;
both pictures are correct.
We have not as yet discussed the last portion of the spectrochemical series. F

> Cl

> Br

> I

. In many ways, it pays to have held o our discussion of these dierent halide ligands
until we had nished developing both the MO and ionic models of transition metal d-orbitals.
Our discussion will combine elements from both theories. F

is a hard ligand driven by its


ionic nature. The uorine being so electronegative, we can imagine the metal-to-uorine
interaction is primarily an electrostatic one. But as we go to Br

and I

, the halides become


quite soft. We should look at the orbitals of the latter two ions
I

has no valence LUMOS, only HOMOs. The HOMO orbitals are the three lled p-
atomic orbitals. For each I

, one of its p-AOs can form a bond while the other two can
make bonds with a transition metal. The p-AOs are therefore both -donors and -donors.
Unlike -acceptors which widen the gap between d-orbitals, -donors narrow the gap. I

,
the softest of the halides has the greatest -donation, the narrowest d-orbital splitting, and
is therefore the last member of the spectrochemical series.
It is really a combined MO and electrostatic model which speaks clearest to the metal-
390
ligand coordination bond.
391
Chapter 39
The Jahn-Teller eect and the 18
electron-rule
This chapter will wrap up the discussion of transition metal complexes. This chapter is not
needed for the later chapters in this book. Based on a preliminary count, we will skip thtis
chapter in the 2011 version of Chem 2160.
392
Chapter 40
Hybridization and bands
In the nal two weeks of the course we move to bonding in crystals. The good news here
is that the MO theory and the ionic model which has been developed in this course applies
equally well to crystals as to molecules. The bad news is that crystals are very large chemical
systems. Consider a crystal such as diamond. Diamond which is pure carbon contains
only carbon-carbon single () bonds. Its crystal structure is illustrated in Figure 40.1. As
diamond bonds are covalent, qualitative MO theory is the appropriate way to understand
the diamond system.
There is one complication. All the carbon atoms are bonded together into a single unit.
If we were to think of a diamond molecule, it would be the entire diamond crystal viewed
as a single whole. The Hope diamond is a macroscopic molecule; it contains Avogadros
number of carbon atoms within it.
The MO diagram of diamond must reect this fact. The number of molecular orbitals
equals the number of atomic orbitals; diamond posesses Avogadros number of atomic orbitals
and hence Avogadros number of MOs. While today we do in fact know how to calculate all
these diamond MOs, for many purposes we do not need the full diamond MO diagram. In
this chapter, we will use the concept of hybridized orbitals to develop a simplied picture
393
Figure 40.1: The crystal structure of diamond. On the right a cube is shown. This is the
unit cell of diamond, the fundamental building block on which the structure is based. We
will discuss unit cells in the last week of this course.
394
of the diamond MO diagram. Students in this course may have heard of hybrid orbitals
already: they are the sp, sp
2
, and sp
3
hybrid orbitals.
40.1 Ethylene
We begin our discussion of hybrid orbitals by discussing the MO diagram of ethylene, C
2
H
4
.
Ethylene is a planar molecule, as is shown in Figure 40.2. (As the goal of this lecture is
to obtain a band picture of -bonded diamond, the more direct molecule to consider here
would be the purely -bonded ethane molecule, C
2
H
6
. We choose this example as our
previous experience in this course has been wiith systems and ethylene, unlike ethane,
contains both - and -MOs.)
Figure 40.2: Ethylene.
Ethylenes MO diagram is most easily derived from the mixing of the four peripheral H
AOs with a central C=C moiety. Ethylene has three perpendicular mirror planes and it is
therefore straightforward to derive its MO diagram. The derivation of the MO diagram is
shown in Figure 40.3. The key to obtaining ethylenes MO diagram proves to be using the
C=C moiety as part of the starting orbitals. In C=C the most bonding and antibonding
orbitals are inward pointed and can therefore mix very little with the peripheral hydrogen
atoms.
On the right of Figure 40.3 we further introduce the band picture of MO diagrams. In
the band picture, rather than indicate the indivdual MOs and their energies, we indicate the
395
Figure 40.3: The ethylene MO diagram.
396
approximate energies of the dierent MOs. As the band picture summarizes, ethylene has
ve -bonding, ve -antibonding, and two MOs.
40.1.1 sp
2
hybridization
We now introduce an electronic structure scheme which will obtain the number of -bonding,
-antibonding orbitals without the deriving of the MO diagram. In other words, we will de-
velop a scheme which gets the band diagram on the right of Figure 40.3, without determining
the actual MOs. This scheme has one weakness: while it calculates correctly the number of
gets correct the number of -bonding vs. -antibonding orbitals, for many ring systems it
gets the wrong number of -bonding vs. -antibonding orbitals. The hybridization scheme
always gets right the total number of -MOs, but not how they are apportioned between
bonding and antibonding states.
The scheme works well in conjunction with the VSEPR model. Apply VSEPR to an
atom in a molecule. For main group atoms, very oftern the atom has two, three, or four
electrton groups around it. The two groups are pointed in a line, the three groups make a
triangle, and the four groups form a tetrahedron. We therefore mix the AOs in order to get
the orbitals pointed either in a line, sp-hybridization, a triangle, sp
2
hybridization, or in a
tetrahedron, sp
3
hybridization, see Figure 40.4.
Note that these schemes break the basic rules of mixing, we mix orbitals to get them
closer in energy rather than farther apart. Hybridizatiion schemes require a suspension in
the application of basic quantum mechanical principles.
Applying the sp
2
hybridization scheme to the central carbon atoms, and mixing these
orbitals with each other and the peripheral hydrogen atoms we obtain Figure 40.5. Just as
in thte full ethylene MO diagram, there are ve -bonding, ve -antibonding, and two
MOs. The hybridization scheme gets right the ethylene band picture. But note that as the
397
Figure 40.4: The sp, sp
2
, and sp
3
hybridization schemes.
398
hybridized scheme does not follow the basic quantum mechanical symmetry of ethylene, the
individual hybridized MOs can not, and are not, generally correct.
Figure 40.5: The ethylene hybridized orbitals.
40.2 Ethane
As Figure 40.4 shows, it is possible to mix the s and p atomic orbitals to make four sp
3
hybridized orbitals, each pointing to the four corners of a tetrahedron around each atom. It
is a scheme ideally set up to deal with tetrahedrally coordinated carbon atoms.
In the problem set, the student is asked to derive the MO diagram of ethane, C
2
H
6
, see
Figure 40.6. In working out this problem the student will discover that ethane contains seven
-bonding MOs and seven -antibonding MOs.
We can obtain these same numbers of -bonding and -antibonding MOs directly from
a hybridization scheme. In ethane, each carbon atom is tetrahedrally coordinated to atoms.
Similarly in the sp
3
-hybridization scheme the four AOs around a main group atom are
399
Figure 40.6: Ethane.
converted into four orbitals, each with a single main lobe pointing in one of the four directions
of a tetrahedron.
We therefore sp
3
hybridize both of the carbon atoms in ethane. As Figure 40.7 shows
these sp
3
hybrids can mix with either another carbon sp
3
hybrid or a hydrogen s orbital to
make seven -bonding and seven -antibonding orbitals. These numbers are the same as
produced by the full MO diagram.
As the problem set shows, however, the correct orbitals are the MOs. Only the MOs can
be used to explain why, when ethane absorbs light, the C-C bond becomes shorter, while
the C-H bonds become longer.
40.3 Diamond
We now turn to the diamond crystal. In this crystal, every carbon atom is surrounded in a
tetrahedron by four other carbon atoms. We hybridize all the carbon AOs into sp
3
hybrids
pointing in the four directions of a tetrahedron. As Figure 40.8 and Figure 40.9 show, these
sp
3
hybrids mix, resulting in an equal number of bonding and -antibonding orbitals. Let
there be Avogadros number, N, of carbon atoms in a diamond crystal. There would then
be both 2N -bonding and 2N -antibonding orbitals in the crystal. This is represented in
400
Figure 40.7: Ethane hybridized orbitals.
401
Figure 40.10.
Figure 40.8: Diamond bonding hybridized orbitals
40.4 Hybrid vs. molecular orbitals
In this chapter, we have compared two dierent orbital schemes, the MO and the hybrid
orbital. Of the two, the MO scheme is much more dicult to apply. Deriving an MO diagram
requires the use of symmetry and the application of multiple mixing. The hybrid scheme,
by contrast, as our ethylene and ethane examples show, closely tracks the molecules Lewis
structure.
It is easy to fall in the trap to believe that both MO and hybridization models are equally
correct, or worse yet, that the hybrid scheme, being simpler, should be used in preferenceto
the MO approach. This would be a mistake. MOs are the eigenvectors of the Hamiltonian,
the energy measurement operator. The eigenvalues corresponding to the MOs are the only
402
Figure 40.9: Diamond antibonding hybridized orbitals
Figure 40.10: The band diagram for diamond. All -orbitals are lled and all

-orbitals are
empty.
403
possible measured electron energy values. MOs correspond to a measurable quantity, the
energy.
Let us compare the MOs and the hybridized orbitals of the molecule H
2
O, see Fig-
ure 40.11. The MO diagram has four MOs, each with a dierent energy. The hybridized
scheme has two types of lled MOs: one type is bonding, the other non-bonding.
Figure 40.11: Comparison of MO vs. hybridized orbitals for H
2
O.
In photoelectron spectroscopy, we measure the energy of the electron. In this spec-
troscopy, we irradiate a sample with light of a known energy. This energy is sucient to
expel electrons from the sample. We measure the energy of these electrons. By subtraction,
we deduce the electrons initial energy. The low energy photoelectron spectrum of water is
shown in Figure 40.12. As can be seen there are three dierent energy levels. Not shown, at
slightly higher energies, is a fourth level. The number of levels corresponds to the number
of MOs.
Hybrid orbitals are easy to work with. But they are not the eigenvectors of any known
measurement. Unlike MOs, hybrid orbitals do not have measurable eignevalues. Hybridiza-
tion schemes can be useful. For example, they can be used in counting schemes, when we
404
Figure 40.12: The photoelectron spectrum of water.
405
wish to know the number and basic types of lled and unlled MOs. For example in both
ethane and diamond, the hybridization scheme tells us that the HOMO is a -bonding or-
bital, while the LUMO is -antibonding. The dierences in HOMO-LUMO energies are
large and hence both these systems absorb only in the UV. We have no need to go to a
full MO treatment to understand that ethane and diamond are UV absorbers. And nally,
and perhaps most importantly, hybridization is relevant as many chemists vastly prefer the
hybridized scheme over the full MO picture when giving chemistry lectures.
406
Chapter 41
Bi, Se, and Te
407
408
409
410
411
412
413
414
415
416
417
418
419
420
421
422
423
424
425
426
427
428
429
430
431
Figure 41.1: Color lost in transcription. These are supposed to be in-phase bonding orbitals.
432
Figure 41.2: Color lost in transcription. These are supposed to be out of phase antibonding
orbitals.
433
Chapter 42
Metals
Most of the elements in the periodic table conduct electricity well: they are metals. And
with very very few exceptions, when one mixes two metallic elements, their mixture is still
a conducting metal. Metals are therefore common.
In this chapter we will learn why most elements are metals. At the same time, we will
learn the three most common metallic crystal structures, face-centered-cubic (fcc), hexagonal-
closest-packed (hcp), and body-centered-cubic (bcc). Finally we will relate these three struc-
tures to the MO needs of the metallic elements in general and thus understand why these
structures are the most common.
42.1 Electron decient systems
Metallic elements tend to be elements on the left-hand side of the periodic table. Being on
the left side of the table, they are elements which have fewer lled MOs than empty MOs.
Noting that many of the most stable chemical systems have their bonding and nonbonding
orbitals lled but their antibonding orbitals empty, we conclude that metal elements should
adopt structures whose MO diagrams have fewer bonding and nonbonding orbitals than
434
antibonding ones.
In Figure 42.1, we see how these general considerations fare in practice. We consider the
element Li. To simplify matters, we assume that the only valence Li orbital is its 2s orbital.
Using the techniques established in this course, we derive the MO diagram for a variety of
Li clusters. Although we do not explicitly show the derivations here, the interested student
can, if they wish, verify that the MO diagrams shown in this gure are correct.
In each of these MO diagrams, we occupy all the bonding and non-bonding orbitals and
none of the antibonding orbitals. We give in this gure two ratios. The rst ratio equals the
(number of bonding + nonbonding orbitals)/(total number of orbitals). The second ratio is
the ratio between the number of electrons in the system to the number of bonded pairs of
atoms.
Let us rst turn to the ratio of bonding and nonbonding orbitals to the total number of
orbitals. In Figure 42.1 we see that the rst of the two structures have ratios which are less
than one-half. The rst two structures are therefore good candidates for metal structures;
the second two structures are not.
Chemists know, just by looking at the structure, to tell if a molecules MO diagram
contains fewer bonding than antibonding orbitals. Molecules composed of mutually bonded
triangles of atoms have such MO diagrams. It is the triangles of bonded atoms which is
reponsible for the ratio of bonding and non bonding orbitals to total orbitals to be less than
one-half. We conclude the structtures of from the left-side of the periodic table, which all
have comparitavely few valence electrons, should all contain triangles of bonded atoms.
We now turn to the ratio of electrons to pairs of bonded atoms. As Figure 42.1 shows,
only structures wiith very few electrons per pairs of bonded atoms have the desired MO band
structure. Metals from the left-side of the perioidc table therefore adopt structures where
there are lots of bonded pairs of atoms with very few electrons to spread between them. For
435
Figure 42.1: MO diagrams with all bonding and nonbonding orbitals lled for a variety of
Li clusters.
436
metals, the concept of two electrons per bonded pair of atoms is not a feasible one.
This short MO exposition allows us to see the two principles which govern metallic
structures. Metal structures contain triangle of mutually bonded atoms. Metal structures
have a large number of bonded pairs of atoms compared with the total number of valence
electrons. Chemical systems which obey these two principles are called electron decient
systems. And metals, except those from the far right of the perioidc tabel, are all electron
decient.
437
42.2 fcc, hcp and bcc
The three most common metal structures are face-centered-cubic (fcc), hexagonal-closest-
packed (hcp), and body-centered-cubic (bcc). All three of the stuctures contain lots of
bonded pairs of atoms and lots of mutually bonded triangles. They are illustrated in the
attached slides. In fact, if the requirement was to optimize the number of bonds and the
number of mutually bonded triangles while at the same time keeping the structures as simple
as possible, ie., keeping the crystal unit cells as small as possible, we would probably arrive
at the fcc, hcp, and bcc structures as the most likely candidates for metal structures.
438
The last slide shows just how common the fcc, hcp, and bcc structures are. As this gure
shows, the hcp structure is found for elements with 3-4 valence electrons per atom as well as
7-8 electrons, bcc structures tend to have 5-6 valence electrons per atoms, and fcc structures
have 9-11 valence elctrons per atom. As in fcc, hcp and bcc each atom has 12 to 14 bonds
to it, for none of these systems are there enough electrons for two full electrons per bond.
The reader might wonder how it is possible that with all the bonding orbitals lled and all
the antibonding orbitals unlled, that we can have these structures appearing at a multitude
439
440
441
442
443
444
of electron counts. Analysis of the MO diagrams of metals shows that while metal structures
tend to ll all the bonding orbitals and not ll all the antibonding orbitals, they do not
do so in a perfect manner. Sometimes these imperfections lead to additional metal crystal
structures. These additional crystal structures can be quite complicated.
42.3 The Hume-Rothery metals
Metal structures whose structures are controlled by optimizing the number of lled bonding
orbitals and unlled antibonding orbitals are called Hume-Rothery metals. They are found
most often for metals with 6-8 and 11-12 valence electrons per atom. We will show only one
of these structures here, that of -brass. Brass is a metal composed of copper and zinc. -
brass is CuZn and -brass is Cu
5
Zn
8
. The -brass crystal structure is shown in Figure 42.2.
The structure is complicated and goes beyond the structural complexity which a freshman
chemistry student needs to learn. We illustrate it here as this structure for what it can teach
us about the accuracy of the MO methods introduced in this course. See the next section.
42.4 Detailed metal MOs
42.4.1 -brass
We give two examples here of metal MO diagrams, both for dierent types of brass. Both
examples follow the two literature convention of metal MO band diagrams. The literature
format contains more information than the band diagrams given previously in this book. In
particular, literature MO diagrams show the MO energies as curves. Every value in every
curve of these diagrams corresponds to a dierent MO. As this convention makes the MO
energies look a lot like spaghetti such diagrams are called spaghetti diagrams.
The rst literature band structure we show is for -brass, CuZn, which is in a bcc-type
445
Figure 42.2: The structure of -brass, Cu
5
Zn
8
. The structure is an example of the com-
plicated Hume-Rothery phases. We introduce it here as this structure can teach us a lot
about the accuracy of the MO methods introduced in this course. See the next section. This
section contains as well as a discussion of the IT, OT, OH, and CO naming convention
446
structure (with Cu atoms at the corner of the cubic unit cell and Zn atoms in the cube body
center). In Figure 42.3 we show two dierent pictures of the band structure of -brass. On
the left we show the spaghetti diagram, where the energies of detailed valence molecular
orbitals are shown. (The horizontal axis contains information of the phase relationship of
atomic orbitals in one unit cell compared to atomic orbitals in a neighboring cell. The
nomenclature of , X. M. R, and is complicated and I believe due to the Cornell scientist,
Hans Bethe.)
On the right of this gure we show the summed band diagram, where the number of
orbitals at dierent energies are shown by the height of the curve. The function labeled
g(E), the electron density of states, gives the relative number of energy levels as a function
of the MO eigenvalues. Note the electron density of states has two spikes. The lower and
upper spikes correspond to respectively the Zn and Cu d-orbital energies. The gure also
shows a dotted line. This dotted line is the Fermi energy, the energy which separates lled
from empty orbitals.
The calculations shown are local density approximation density functional calculations,
LDA-DFT - a highly accurate calculation of the energy eigenvalues. In Figure 42.4 we
compare the -brass LDA-DFT electron density of states with a density of states using the
method taught in this course, the extended H uckel, eH, method. As can be seen the eH
method captures the important energy features of the accurate LDA-DFT calculation.
42.4.2 -brass
Cu
5
Zn
8
, -brass, has a much more complicated structure than -brass, with 52 atoms in a
cubic unit cell. Its crystal structures is again illustrated in Figure 42.2. The notation IT, OT,
OH, and CO is a standard one. They stand for Inner Tetrahedron, Outer Tetrahedron,
OctaHedron, and CuboOctahedron. As the gure shows, the atoms in Cu
5
Zn
8
occupy
447
Figure 42.3: The band structure of -brass.
448
Figure 42.4: Comparison of LDA-DFT and EH spaghetti diagrams for -brass.
449
respectively yellow, orange, red, and purple polyhedra. In order these four polyhedra are
a small tetrahedron, a large tetrahedron, an octahedron, and a cubooctahedron. The Cu
atoms occupy all the IT and CO sites while the Zn atoms occupy the OT and OH sites.
The point of all this is not to learn the -brass structure, but to see that eH theory
captures correctly the full LDA-DFT result, see Figure 42.5 and Figure 42.6. In other
words, as these gures show, the MO techniques which you have learned in this course get
the correct energy eigenvalues for metal electrons. Metals follow the same bonding model as
other covalently bonded systems. The covalent bond and thte metallic bond are caused by
the exact same physical forces, the drive to lengthen the de Broglie wavelength.
Figure 42.5: The band structure of -brass.
450
Figure 42.6: Comparison of LDA-DFT and EH spaghetti diagrams for -brass.
451
The results in this section, on both - and - brass, were obtained by Peter Walters, a
freshman in this course in 2004. Pete carried out this work as part of his Cornell undergrad-
uate honors thesis. Roald and I submitted Petes work as part of a larger piece of research.
It was published this year in Chemical Reviews, a journal with a 2009 Impact Factor of 36.0,
a very high number). Roald remarked in the writing of this paper, that Peters results came
thirty years too late. I take this to be a comment on many of the current generation of
chemists. Many of the new generation have lost their ability to reason through molecular
orbitals: the current generation of chemists were told, quite incorrectly, that eH MOs were
not accurate. Roald believes that it will be future generations of chemists, chemists such as
yourselves, who will have to rediscover the power of this technique.
452
Chapter 43
Ionic structures
In this chapter we will discuss the common one-to-one or AX crystal structures. These are
structures composed of an equal number of two dierent elements, A and X. We examine
rst the three common AX structures favored by the ionic bond: the NaCl, the sphalerite
(ZnS), and the CsCl structure types. Of these, it is the NaCl structure which will prove to
have the best ionic bonding of all.
We contrast these structures with those compatible with the covalent and metallic bonds.
Structures discussed here extend from the sphalerite and CsCl structures to the AuCu and
AuCd structures, two common metallic AX crystal structures. Finally we will study the
Mooser-Pearson diagram, a structural model developed in the 1950s which examines the
interplay between the covalent, metallic, and ionic bond in the experimentally observed AX
crystal structure types.
43.1 The crystal structure of NaCl
We start this chapter by considering table salt, sodium chloride. As sodium belongs to the
alkali metal elements and chlorine to the halogens, a stable sodium chloride compound has
453
a one-to-one ratio. Sodium chloride is therefore NaCl and is comprised of an equal number
of mono-cations and mono-anions.
Can we deduce the crystal structure of sodium chloride from its stoichiometry? To
construct an answer, we rst note that as sodium is very electropositive and chlorine very
electronegative, the MO interaction between sodium and chlorine valence AOs is slight.
What interaction there is is electrostatic in origin. The NaCl bond is ionic, not covalent.
For the ionic bond, E q
+
q

/r, where q
+
, q

, and r are respectively the charge of the Na


+
,
the Cl

, and the distance between adjacent ions.


We will presuppose that the r in the above equation is a distance controlled by the HOMO
electrons of both ions. (While there are only weak MO interactions between valence AOs, if
the distance between the Na
+
and Cl

becomes so slight that the core Na electrons begin


overlapping with valence Cl

, there is a signicant core Na HOMO to valence Cl HOMO


repulsion. This mimimum Na
+
-Cl

distance, r
NaCl
, is generally held to be equal to the
sum of the size of the core electrons in Na
+
and the valence electrons in Cl

. These sizes are


usually expressed in terms of the radii of spheres, the ionic radii. Thus r
NaCl
= r
Na
+ +r
Cl
,
where r
Na
+ and r
Cl
are respectively the ionic radii of Na
+
and Cl

.)
With this preamble it is clear a single Na
+
and Cl

would place themselves optimally


as shown in Figure 43.1. This monomer unit places the Na cation and Cl anion as close
as they can possibly be. We now consider the 1-D Na-Cl structure. The optimal structure
will be one which maximizes the number of cation-anion contacts, while at the same time
minimizing cation-cation and anion-anion contacts. From the above considerations we see
the ideal 1-D NaCl structure is the one shown in this same gure.
Just as we assembled the NaCl monomer units to construct the 1-D Na-Cl crystal, we can
assemble distinct 1-D chains to form the 2-D sheet structure, see Figure 43.2. As this gure
shows, the 2-D sheet assembles into a square motif, with each Na
+
with four Cl

nearest
neighbors and vice-versa.
454
Figure 43.1: The optimal Na-Cl monomer and 1-D crystal structures.
455
Figure 43.2: Assembling the optimal NaCl 2-D crystal from the optimal 1-D crystal structure.
456
We can now see our way to the optimal 3-D crystal structure. We make this structure
by stacking 2-D NaCl sheets in such a way as to optimize Na-Cl contacts. This structure
is shown in Figures 43.3 and 43.4. Figure 43.4 shows the true 3D NaCl crystal structure.
Shown in this gure are several alternate views. The ball view shows that each Na ion is
surrounded by six Cl ions and vice-versa. The stick view shows the cubic nature of the
NaCl structure. Finally, one common literature view is to show the shape of the polyhedron
formed by all ita anion nearest neighbors. For NaCl this polyhedron is an octahdedron.
The NaCl structure is adopted by crystals other than sodium chloride. Other examples
include AgCl, MgO, PbS, TiC, CrN, LiF, LiI, NaF, NaBr, CsF, RbI, InAs, BaTe, CaS, VN,
SrS, ZnTe, and UN. This list extends to several hundred dierent AX crystal structures. As
the above discussion shows, the NaCl structure is the ultimate ionic AX crystal structure.
43.2 Other AX crystal structures
Besides the analysis given for the NaCl structure in the section above, this same structure
can be used to illustrate an even more productive way of understanding common AX crystal
structures. This second analysis is illustrated in Figure 43.5. In this gure we compare NaCl
with the crystal structure which would have been formed if the Na and Cl ions were replaced
by the same element.
As this gure shows, the NaCl structure is related to the elemental polonium structure
which we examined two chapters ago. The idea here then is to think of the AX crystal
structure which would come about from populating an elemental structure with two dierent
atoms. We suppose that one of these two atoms will be a cation, while the other an anion.
We further suppose that these cations and anions will have at least some component of
an ionic bond. Therefore, we will choose cation and anion sites so as to maximize the
number of cation-anion bonds, while simultaneously minimizing cation-cation and anion-
457
Figure 43.3: Assembling the optimal NaCl 3-D crystal from the optimal 2-D crystal structure.
458
Figure 43.4: The 3-D NaCl crystal structure.
459
Figure 43.5: The relation between the NaCl and elemental polonium crystal structures.
460
anion interactions.
In Figures 43.6, 43.7, 43.8, and 43.9, we show the AX structures so generated from
respectively the cubic diamond, the fcc, hcp, and bcc elemental structures. The four respec-
tive AX structures so generated are sphalerite (ZnS), AuCu, AuCd, and the CsCl structures.
As these gures show, the diamond and bcc structures convert to excellent ionic structures
while the fcc and hcp do not. We can expect partially ionic crystals form in the sphalerite
and CsCl structures, but the AuCu and AuCd structures will be found in for less ionic AX
stoichiometries.
Figure 43.6: The relation between the diamond and sphalerite crystal structures.
461
Figure 43.7: The relation between the fcc and AuCu crystal structures.
462
Figure 43.8: The relation between the hcp and AuCd crystal structures.
463
Figure 43.9: The relation between the bcc and CsCl crystal structures.
464
The data listed in these gures support this conjecture. AX metals where the A and X
are very dierent in electronegativity are often found in the CsCl structure but those where
the A and X electronegativities adopt the AuCd and AuCu structures as well.
43.3 The Mooser-Pearson Diagram
We nish this chapter examining AX crystal structures, where A and X are main group
atoms and where the number of A plus X valence electrons adds up to 8 electrons per AX
formula unit. There are roughly 50 such compounds. In the 1950s two Canadian scientists,
Mooser and Pearson constructed a diagram which allows one to understand why dierent
AX structures adopt the structures which they do. They considered two dierent elemental
parameters, the dierence between elemental electronegativity and the elements average
principal quantum number.
Mooser and Pearson decided that is the AX electronegativity dierence is large, the ionic
bond will dominate, while if the dierence is small, the covalent and metallic bond will be
important. For large dierences, one would expect the NaCl structure to dominate. They
considered next the average principal quantum number. Recalling that heavier main group
elements are either hypervalent or metallic like in character, they expected heavier elements
would adopt either the Po type hypervalent structures (which as an AX compound would
be the NaCl structure) or a metallic structure like fcc, bcc, or hcp. Of these three metallic
structures, the bcc-based CsCl structure would be found for the more ionic of the metal
structures. Finally covalent structures will obey the octet rule. For eight valence electrons
per AX unit, the diamond or sphalerite structure should dominate.
As Figure 43.10 shows, the actual observed crystal structures follows the above analysis.
For high electronegativiy dierences, the NaCl structure is always observed. For large average
principal quantum numbers the NaCl and CsCl structures are obserevd. Finally at low
465
electronegativity dierencs, for lighter elements, the octet rule is obeyed and the diamond
or sphalerite structure is the main observed structure type.
Figure 43.10: The Mooser-Pearson diagram for AX main group compounds.
43.4 The radius-ratio rules
Most inorganic textbooks base their discussion of the AX structures on an entirely dierent
analysis. This analysis is called the radius ratio rules. The radius ratio rules are purely
geometric rules which putatively sort out the CsCl, NaCl or sphalerite structures. Textbooks
rarely discuss the data on which the radius ratio rules are based. Analysis of the data shows
the rules are correct slightly less than 50% of the time. As the radius-ratio rules are being
used to sort out only three structures, the data supporting this model is not very convincing.
The Mooser-Pearson diagram suggests what is missing in the radius-ratio rules. It is
always important when thinking of crystal structure to consider the nature of the chemical
466
bonds in question.
43.5 Summary
This chapter has explored the relation between the ionic bond and crystal structures. In
ionic structures one maximizes cation-anion interactions while simultaneously minimizing
cation-cation and anion-anion interactions. Among the octet rule crystal structures, the
diamond structure is compatible with the ionic bond, while among the metal structures,
bcc is the best. Ultimately, in understanding crystal structures, it is best to think of the
interplay of the covalent, metallic, and ionic bonds in deducing optimal structures. We study
this last idea in detail in the next chapter.
467
Chapter 44
Identifying crystal structures
A problem of the following type will be on the nal. The student will be given a crystal
structure. The crystal structure typically has more than one element type (dierent elements
are represented with dierent sized and colored spheres). The student is then asked to suggest
a chemical compound which could most plausibly form in the illustrated structure.
To solve these problems, the student must put together lots of disparate information
about many dierent types of bonding. The student must also have mastered skills associated
with looking at unit cells. In this chapter, I will restate some of the most useful facts. I will
then make some suggestions about things to think about when solving this type of problem.
The majority of this chapter will be devoted to giving you examples for you to work on.
44.1 Some key facts
1. In ionic compounds, the coordination number of the cation (the coordination number ot
the metal atom is essentially the number of atoms the metal atom is bonded to) depends
much more on the type of anion rather than the stoichiometry of the compound. As O
and F are the two principal ionic compound anions, we focus on these. In Figure 44.1
468
we show the most typical coordination numbers of metals with second row atoms such
as F or O.
Figure 44.1: TYpical metal coordination numbers with oxygen or uorine.
As we saw in the chapters on transition metal complexes, the most typical coordination
number is six: the octahedral coordination. In TiO
2
as well as Ti(H
2
O)
4+
6
, the titanium
atom is octahedrally coordinated to oxygen atoms.
2. Figure 44.1 is just a general outline of metal-F and metal-O coordination numbers.
Metal atoms to the right of the periodic table: Ni, Cu, Zn, and Al range from 4 to
6 coordinate geometries. Other metallic elements can have 4 coordinate geometries.
For example, highly oxidized metal cations, which are very small, are also often 4-
coordinate. The alkali metal atoms show a pronounced dierence in coordination
numbers: Li is often 4-coordinate to oxygen, Na would more likely be 6-coordinate,
and K higher than 6 coordinate.
3. Large metal atoms oftwn have very high coordination numbers to O and F. Large
alkali metal atoms range often from 8-12 coordinate. B is either 4 or 3 coordinate to
oxygen. Carbon, which obeys the octet rule, can make carbon monoxide, but 2 and 3
coordinate carbon to oxygen is more common.
469
4. Figure 44.1 follows roughly the size of the metal atom. Elements to the left and to
the bottom of the periodic table are larger. When the anion gets bigger, as we pass
from F to Cl, metal coordination numbers get smaller: for large anions it is sometimes
not possible to place six anions around the same cation. In TiF
4
, the titanium atom
is 6-coordinate to F; in TiI
4
it is 4-coordinate.
5. The anion coordination number is often determined by the metal coordination number
and the metal oxidation state. Recall that while left-side rst row transition metal
oxidation states typically are governed by their column in the periodic table: Sc
3+
,
Ti
4+
, and V5+, by the middle of the transition elements, lower oxidation states are
more typically achieved: we nd Cr
3+
, Mn
2+
, Fe
3+
, Co
2+
, Ni
2+
, and Zn
2+
. If we wish
to determine the oxygen coordination number in chromium oxide, we note that the
chromium metal oxidation state results in Cr
2
O
3
being a common chromium oxide. Cr
is 6-coordinate to oxygen. We deduce oxygen is most probably coordinated to four Cr
atoms.
6. Metal atoms when they mix together make another metal. Metal crystal structures are
high coordinate and are often related to fcc, hcp, and bcc. Most metals are electron
decient: they do not have two electrons a bond. Triangles and tetrahedra often result
due to this electron decient nature.
7. Ionic structures have cations only bonded to anions. Some elemental structures can
readily be turned into ionic structures (simple cubic, diamond, and bcc) but others
can not (fcc and hcp).
8. Light main group atoms typically obey the octet rule. Carbon and Si with formal
charge zero have four bonds, N and P have three bonds, etc... These atoms can obey
standard hybridization schemes: sp, sp
2
, and sp
3
.
470
9. Heavy main group elements have an inert pair. Their most important valence orbitals
are p orbitals. Heavy main group elements are often near-metals. They are often
hypervalent. The hypervalent bond is a Lewis acid-Lewis base interaction. Electron
rich main group metals and near-metals often have structures with square-motifs.
44.2 Notions to follow in solving the crystal identica-
tion problem
It might be possible to make a ow chart telling you the best way to deduce possible crystal
chemical formulas. But this is really counter to the point of the exercise. The exercise is to
take the enormous amount of information which you are assembling over chemical compounds
and to organize it in a fashion which can answer many questions. Crystal identication is
just an example of how this inofrmation can be applied. Nonetheless, I would like to list
here some things to think about when you determine possible chemical formulae.
1. Look at the structure. Is it metallic-like (have lots of bonds), ionic-like (only dierent
elements are bonded to one another), or covalent-like (possibly sp-hybridized, sp
2
, or
sp
3
hybridized)?
2. If it is covalent, the electron count could follow the octet rule. If it is ionic, con-
sider possible metal oxidation states. If it is a metal, we have no understanding of
stoichiometry, and you have a lot of freedom in suggesting possible chemical formulas.
3. Establish the stoichiometry of the unit cell. There are two ways of doing this. One is by
establishing the number of each sphere type in the unit cell. The other is by examining
the dierent coordination numbers. Sometimes one method is quicker, sometimes, the
other.
471
4. Remember common crystal structure types: fcc, hcp, bcc, diamond, NaCl, bcc, and
simple cubic. Is there a connection between these structures and the structure at hand?
Remember how the Mooser-Pearson diagram organizes these structures.
44.3 Examples of problems
The next few pages are devoted to dierent problems. Each identication problem is one
page long.
472
473
474
475
476
477
478
479
480
481
Chapter 45
The virial theorem
We have come to our last lecture. While for the other lectures in this course, I have con-
centrated on what I believe the student should know about the chemical bond, this last
lecture presents material better suited to junior level physical chemistry. To my knowledge,
however, the second half of the lecture material is not given in physical chemistry textbooks
and therefore is, up to now, not in a format which undergraduate students might readily see.
The goal of this lecture is to describe what happens to the kinetic and potential energy
when a chemical bond forms. This lecture material is motivated by a need to unravel what
chemical bonds really are rather than teach specic student skills. Still, I have tried to write
this lecture in a form which you, the student, can follow.
At the outset let me warn you, this chapter will use one-dimensional calculus to derive
the optimal size of the hydrogen atom and the energy of the hydrogen molecule. Calculus is
important as optimization (ie., minimization) of both the kinetic and the potential energy
plays a crucial role in both atomic orbitals and the covalent bond. We shall rst examine
the individual atom, nding a connection between atomic energy optimization and the virial
theorem. This initial material is contained in third year physical chemistry textbooks.
482
45.1 The hydrogen-like atom
We consider a one electron atom. The energy of the electron in this atom is the sum of its
kinetic and potential energy:
E
Total
= E
Kinetic
+ E
Potential
= E
K
+ E
P
=
p
2
2m

Ze
2
r
,
where p is the momentum, Z is the atomic number, and r is the distance between the electron
and the nucleus.
We apply now the concept of the de Broglie wavelength. In its ground state the electron
will, on the average, be located on the order of magnitude of the de Broglie wavelength, ,
away from the nucleus. Further, the momentum follows the de Broglie relation, p = h/.
Combining these ideas with the above energy equation we nd:
E
h
2
2m
2

Ze
2

.
We now search for a value of which minimizes the total energy. This is a type of problem
which you have seen in rst semester calculus. At the minimal value of the energy, dE/d =
0. We therefore nd:
0 =
dE
d
=
h
2
m
3
+
Ze
2

2
=
1

2
(
h
2
m
+ Ze
2
).
=
h
2
mZe
2
.
We dene the Bohr radius, a
o
, to be the de Broglie wave length for the hydrogen atom where
Z = 1.
a
o
=
h
2
me
2
0.5

A
483
The size of the hydrogen atom is determined by the above relationship. As verication of
the accuracy of the above equations, note that:
E
K
=
mZ
2
e
4
2 h
2
,
E
P
=
mZ
2
e
4
h
2
.
The virial theorem, 2E
K
= E
P
, is obeyed by our model.
Finally we compare our expression for the de Broglie wavelength of the 1s electron with

1s
given in the fourth chapter of this textbook We nd that
1s
(r) = N
1s
e
r/
, where N
1s
is an unspecied constant. The 1s wavefunction is an exponentially decaying function whose
decay length is the de Broglie wavelength itself.
45.1.1 Translating the above equations into pictures
The previous section shows, on the one hand, that the radius of the hydrogen atom is
governed by the Bohr radius and, on the other hand, that the size of the atom comes
from the simultaneous optimization of both the kinetic and the potential energies. Lets
translate this story into pictures. For pictures, we consider two asymptotes, where the de
Broglie wavelength is either set to minimize only the kinetic or only the potential energy.
In Figure 45.1 we show the optimal wavefunctions for these two scenarios. In the former
case, where we wish to minimize solely the kinetic energy, we note E
K
= p
2
/2m = h
2
/2m
2
.
The longer is, the smaller the kinetic energy. E
K
is therefore minimized when = , see
Figure 45.1.
If, instead, we were to minimize just the potential energy, the electron would be located
entirely at the nucleus. The wavefunction would be like the one shown in the middle panel
of Figure 45.1; its de Broglie wavelength would be zero. The actual 1s wavefunction, which
484
Figure 45.1: which minimize respectively the kinetic, potential, and total energies.
optimizes both the kinetic and the potential energy is
1s
= N
1s
e
r/
, shown in the right
panel of the gure. The true 1s wave function is an exponential decay function which retains
a spike at r = 0 to help optimize the potential energy, but also keeps an exponential tail at
larger values of r to help minimize the kinetic energy portion of the total energy.
An example will help clarify the situation. In the previous section, we found that the de
Broglie wavelength was inversely proportional to Z, ie., 1/Z. As Z increases in size,
the potential energy term gets bigger, becomes smaller, N
1s
e
r/
becomes a more swiftly
decaying exponential function, the tail of the wavefunction gets smaller, and the spike grows
larger. An increase in the value of Z shifts the balance between the spike and the tail
component of the wave function so that the spike plays a greater role. This is in accord with
the greater role that potential energy would be playing.
To make an analogy, the kinetic energy and the potential energy are like two dierent
reins which steer a horse (the horse being the total energy). As we increase the needs of one
of the reins, the corresponding rein tightens up and that term dominates more. Conversely,
if one of the two reins loosens up, it allows the other rein to tighten and exert greater control.
The idea that the shape of the wavefunction changes as we disturb the balance between
kinetic and potential energies will prove to lie at the heart of the covalent bond.
485
45.2 The covalent bond
45.2.1 The H
2
molecule
We now consider the H
2
covalent bond. In this course, we have adopted a model where
we take the two exponential decay functions,
1s
= N
1s
e
r/
, one function centered at one
nucleus and the other at the other nucleus, and place these two wavefunctions in phase with
one another. As we have seen, the kinetic energy decreases as we do so. But in making this
model we made a perhaps unwarranted assumption, that the specic atomic orbitals which
make up the H
2
molecular orbital remain themselves unchanged. (This assumption was part
of the minimal valence basis set approximation.)
The assumption is unwarranted as the H
2
covalent bond has changed the balance between
the atomic kinetic and potential energy in their roles of optimizing the atomic electronic
energy. By introducing a second hydrogen atom, wavefunctions have become longer. Going
back to our analogy of reins, the kinetic energy rein has loosened up. The potential energy
rein, therefore, has tightened, and plays a more dominant role in determining the size of the
atomic de Broglie wavelength. The atomic orbtial wavelength given in the expression,
1s
=
N
1s
e
r/a
, called
a
here, to distinguish it from the full de Broglie wavelength, contracts,
lowering the total potential energy.
We use Figure 45.2 to illustrate the situation. In this gure we show the total molecular
energy as a function of the
a
decay constant, taken from the expression
1s
= N
1s
e
r/a
.
Note that there are two such 1s orbitals in the H
2
problem. Note further that for each value
of
a
, the interatomic distance between the two nuclei, R, adjusts.
We know there will be one optimal value of the atomic de Broglie wavelength,
a
which
minimizes the total energy of the H
2
molecule. We will call this value
a2
. This value is
shown in the gure. At this equilibrium geometry; the virial theorem is obeyed: E
K
= E
T
and E
P
= 2E
T
. We use these equivalences to place the values of both the kinetic and
486
Figure 45.2: The H
2
energy as a function of the atomic exponential decay parameter
a
. a1
and a2 correspond to respectively the otimal values of the lone hydrogen atom, H
1
, and the
hydrogen molecule, H
2
.
potential energies on the gure.
As
a
increases, the kinetic energy decreases; conversely as decreases, the total kinetic
energy increases. Further as
a
increases, the potential energy decreases; as decreases,
the potential energy increases. These relations are in accord with the requirement that
E
T
= E
K
+ E
P
. Also placed on the gure is the value
a1
.
a1
is the initial atomic de
Broglie wavelength, a length larger than
a2
45.2.2 A truncated Taylor series
We now apply one-dimensional calculus. We know that any function can be expressed as a
Taylor series around a specic value of the function, ie.,
E
T
(
a
) = E
T
(
a2
) +
dE(
a2
)
d
a
(
a

a2
) +
d
2
E(
a2
)
d
2
a
(
a

a2
)
2
2
+ ...
487
For values near
a2
we can truncate this expression to :
E
T
(
a
) E
T
(
a2
) +
dE(
a2
)
d
a
(
a

a2
).
Recalling that
a2
minimizes the total energy, this expression further simplies to:
E
T
(
a
) E
T
(
a2
)
We show this relation in Figure 45.2 by making the slopes of E
K
and E
P
equal in magnitude
and opposite in sign.
We can summarize the situation in Figure 45.3, a blow-up of the central portion of the
previous gure. We use this gure to follow the story of the covalent bond. In this course,
we considered a model of atomic orbitals xed in size, where
a
was taken to be a constant.
In this model, covalent bonds appeared to lower the kinetic energy, and, if we had thought
about it, raised the potential energy. But
a
is not a constant. Once the covalent bond
forms, the atomic orbital decay constant,
a
, becomes smaller. It becomes smaller as the
urgency of optimizing the kinetic energy has been partially relieved by the formation of the
in-phase MO.
In making the covalent bond,
a
shrinks in value from
a1
to
a2
. Importantly, the
shrinking barely changes the total energy. By the time
a
has shrunk to its nal value
a2
,
the kinetic energy has increased, and the potential energy has decreased to the point that
the virial theorem is obeyed, but the total energy has barely changed.
In summary, for the covalent bond, ultimately, kinetic energy increases and potential
energy decreases. But, we can understand the covalent bond perfectly well with a model
where the atomic orbitals have not contracted, the kinetic energy has decreasesd and the
potential energy increased. The model used in the course gives valid energies for while the
488
Figure 45.3: Blow-up of the central portion of the H
2
energy as a function of the atomic
exponential decay parameter
a
. a1 and a2 correspond to respectively the otimal values of
the lone hydrogen atom, H
1
, and the hydrogen molecule, H
2
. A Taylor series approximation
has been used.
indiividual kinetic and potential energies are both wrongs, the total energy is close to correct.
45.3 The ionic bond
The ionic bond, as always, is simpler. It will take us one paragraph and one gure. Consider
an Na
+
cation interacting with a F

anion. As these two ions approach, the electrostatic


energy decreases. Were it just up to the electrostatic energy, the cation and anion would fuse
into one. At some point however, the size of the atoms becomes involved. The size of the
atoms, as we showed in the beginning of this chapter, is another way of saying, the size of the
relevant atomic orbitals. As the lled atomic orbitals start interfering, the Pauli exclusion
principle requires that higher, unlled atomic orbitals be involved, THe kinetic energy of the
combined Na-F system increases. the equilibrium Na-F distance is reached when the virial
theorem is obeyed. Figure 45.4 illustates this story.
489
Figure 45.4: Blow-up of the central portion of the H
2
energy as a function of the atomic
exponential decay parameter
a
. a1 and a2 correspond to respectively the otimal values of
the lone hydrogen atom, H
1
, and the hydrogen molecule, H
2
. A Taylor series approximation
has been used.
45.4 Bonding across the periodic table
45.4.1 Second row main group elements
This chapter explains a great deal about the covalent bond. It suggests even more. Why is
it that C-C bonds are so much stronger than Sn-Sn bonds? Diamond melts at over 4000
o
C
while Sn melts at 200
o
C. Why does carbon make double and triiple bonds while Sn does not?
This chapter suggests a solution. The carbon valence orbitals are primarily 2p orbitals. 2p
orbitals are special, as there are no 1p orbitals. Carbon 2p orbitals have therefore no lower
in energy p orbitals to stop their contraction process. They contract enough that we can no
longer contract the Taylor series to the rst derivative term. The total energy term becomes
genuinely smaller.
If this is correct, and I only suspect that it is, there would be two consequences. First
C-C covalent bonds would become stronger. Second, the -orbitals would become so small
that the orbitals would become bigger and overlaps would increase. Carbon could make
490
good bonds, but tin would not.
45.4.2 Magnetic metal elements
This same story has been proposed (one of the proposers is Greg Landrum, a student of Roald
Homann) to be responsible for the magnetism of metallic elements. In Landrums model,
the 3d and the 4f atomic orbitals are both the rst of their kind. Both can therefore contract.
Unlike the 2p elements, what happens is that atomic orbitals with one spin contract. As a
consequence atomic orbitals with the opposite spin expand. The occupation of the up vs.
down spin orbitals is no longer equal, and among other things, a ferromagnet can result.
The metallic elements involved in ferromagentism are indeed 3d and 4f elements. On
the transition metal side, they are Fe, Co and Ni; for the rare earths Pr, Nd, Sm, Gd, and
Dy are often found: 3d and 4f elements all.
I wonder too about the hydrogen bond. Could the absence of a 0s orbital play a role
in the hydrogen bond? Maybe a student in this course can solve this last problem through
their own research.
45.4.3 Hunds rule
The discussion on the virial theorem also might play a apart in Hunds rule. Early in
this book we suggested that triplets are lower in energy than singlets as triplets minimize
the potentitial energy. If the analogy of kinetic and potential energy being separate reins
controlling the total energy is true, then, in the triplet state, the rein on the potential energy
has been loosened, allowing kinetic energy, ie., de Broglie wavelengths, to lengthen. I have
not yet checked the literature to see if this idea is correct.
I hope the student can see that there is still so much to understand about the chemical
bond. There is so much more work to read, think about, and understand. Perhaps, a student
491
reading this book could become suciently interested not just to learn more but to actually
join in further research on the theory of the chemical bond. I know I speak for many, when
I tell you that your help would be most welcome.
492
Chapter 42
Identifying crystal structures
A problem of the following type will be on the nal. The student will be given a crystal
structure. The crystal structure typically has more than one element type (dierent elements
are represented with dierent sized and colored spheres). The student is then asked to suggest
a chemical compound which could most plausibly form in the illustrated structure.
To solve these problems, the student must put together lots of disparate information
about many dierent types of bonding. The student must also have mastered skills associated
with looking at unit cells. In this chapter, I will restate some of the most useful facts. I will
then make some suggestions about things to think about when solving this type of problem.
The majority of this chapter will be devoted to giving you examples for you to work on.
42.1 Some key facts
1. In ionic compounds, the coordination number of the cation (the coordination number ot
the metal atom is essentially the number of atoms the metal atom is bonded to) depends
much more on the type of anion rather than the stoichiometry of the compound. As O
and F are the two principal ionic compound anions, we focus on these. In Figure 42.1
235
we show the most typical coordination numbers of metals with second row atoms such
as F or O.
Figure 42.1: TYpical metal coordination numbers with oxygen or uorine.
As we saw in the chapters on transition metal complexes, the most typical coordination
number is six: the octahedral coordination. In TiO
2
as well as Ti(H
2
O)
4+
6
, the titanium
atom is octahedrally coordinated to oxygen atoms.
2. Figure 42.1 is just a general outline of metal-F and metal-O coordination numbers.
Metal atoms to the right of the periodic table: Ni, Cu, Zn, and Al range from 4 to
6 coordinate geometries. Other metallic elements can have 4 coordinate geometries.
For example, highly oxidized metal cations, which are very small, are also often 4-
coordinate. The alkali metal atoms show a pronounced dierence in coordination
numbers: Li is often 4-coordinate to oxygen, Na would more likely be 6-coordinate,
and K higher than 6 coordinate.
3. Large metal atoms oftwn have very high coordination numbers to O and F. Large
alkali metal atoms range often from 8-12 coordinate. B is either 4 or 3 coordinate to
oxygen. Carbon, which obeys the octet rule, can make carbon monoxide, but 2 and 3
coordinate carbon to oxygen is more common.
236
4. Figure 42.1 follows roughly the size of the metal atom. Elements to the left and to
the bottom of the periodic table are larger. When the anion gets bigger, as we pass
from F to Cl, metal coordination numbers get smaller: for large anions it is sometimes
not possible to place six anions around the same cation. In TiF
4
, the titanium atom
is 6-coordinate to F; in TiI
4
it is 4-coordinate.
5. The anion coordination number is often determined by the metal coordination number
and the metal oxidation state. Recall that while left-side rst row transition metal
oxidation states typically are governed by their column in the periodic table: Sc
3+
,
Ti
4+
, and V5+, by the middle of the transition elements, lower oxidation states are
more typically achieved: we nd Cr
3+
, Mn
2+
, Fe
3+
, Co
2+
, Ni
2+
, and Zn
2+
. If we wish
to determine the oxygen coordination number in chromium oxide, we note that the
chromium metal oxidation state results in Cr
2
O
3
being a common chromium oxide. Cr
is 6-coordinate to oxygen. We deduce oxygen is most probably coordinated to four Cr
atoms.
6. Metal atoms when they mix together make another metal. Metal crystal structures are
high coordinate and are often related to fcc, hcp, and bcc. Most metals are electron
decient: they do not have two electrons a bond. Triangles and tetrahedra often result
due to this electron decient nature.
7. Ionic structures have cations only bonded to anions. Some elemental structures can
readily be turned into ionic structures (simple cubic, diamond, and bcc) but others
can not (fcc and hcp).
8. Light main group atoms typically obey the octet rule. Carbon and Si with formal
charge zero have four bonds, N and P have three bonds, etc... These atoms can obey
standard hybridization schemes: sp, sp
2
, and sp
3
.
237
9. Heavy main group elements have an inert pair. Their most important valence orbitals
are p orbitals. Heavy main group elements are often near-metals. They are often
hypervalent. The hypervalent bond is a Lewis acid-Lewis base interaction. Electron
rich main group metals and near-metals often have structures with square-motifs.
42.2 Notions to follow in solving the crystal identica-
tion problem
It might be possible to make a ow chart telling you the best way to deduce possible crystal
chemical formulas. But this is really counter to the point of the exercise. The exercise is to
take the enormous amount of information which you are assembling over chemical compounds
and to organize it in a fashion which can answer many questions. Crystal identication is
just an example of how this inofrmation can be applied. Nonetheless, I would like to list
here some things to think about when you determine possible chemical formulae.
1. Look at the structure. Is it metallic-like (have lots of bonds), ionic-like (only dierent
elements are bonded to one another), or covalent-like (possibly sp-hybridized, sp
2
, or
sp
3
hybridized)?
2. If it is covalent, the electron count could follow the octet rule. If it is ionic, con-
sider possible metal oxidation states. If it is a metal, we have no understanding of
stoichiometry, and you have a lot of freedom in suggesting possible chemical formulas.
3. Establish the stoichiometry of the unit cell. There are two ways of doing this. One is by
establishing the number of each sphere type in the unit cell. The other is by examining
the dierent coordination numbers. Sometimes one method is quicker, sometimes, the
other.
238
4. Remember common crystal structure types: fcc, hcp, bcc, diamond, NaCl, bcc, and
simple cubic. Is there a connection between these structures and the structure at hand?
Remember how the Mooser-Pearson diagram organizes these structures.
42.3 Examples of problems
The next few pages are devoted to dierent problems. Each identication problem is one
page long.
239
240
241
242
243
244
245
246
247
248
Chapter 42
The virial theorem
We have come to our last lecture. While for the other lectures in this course, I have con-
centrated on what I believe the student should know about the chemical bond, this last
lecture presents material better suited to junior level physical chemistry. To my knowledge,
however, the second half of the lecture material is not given in physical chemistry textbooks
and therefore is, up to now, not in a format which undergraduate students might readily see.
The goal of this lecture is to describe what happens to the kinetic and potential energy
when a chemical bond forms. This lecture material is motivated by a need to unravel what
chemical bonds really are rather than teach specic student skills. Still, I have tried to write
this lecture in a form which you, the student, can follow.
At the outset let me warn you, this chapter will use one-dimensional calculus to derive
the optimal size of the hydrogen atom and the energy of the hydrogen molecule. Calculus is
important as optimization (ie., minimization) of both the kinetic and the potential energy
plays a crucial role in both atomic orbitals and the covalent bond. We shall rst examine
the individual atom, nding a connection between atomic energy optimization and the virial
theorem. This initial material is contained in third year physical chemistry textbooks.
230
42.1 The hydrogen-like atom
We consider a one electron atom. The energy of the electron in this atom is the sum of its
kinetic and potential energy:
E
Total
= E
Kinetic
+ E
Potential
= E
K
+ E
P
=
p
2
2m

Ze
2
r
,
where p is the momentum, Z is the atomic number, and r is the distance between the electron
and the nucleus.
We apply now the concept of the de Broglie wavelength. In its ground state the electron
will, on the average, be located on the order of magnitude of the de Broglie wavelength, ,
away from the nucleus. Further, the momentum follows the de Broglie relation, p = h/.
Combining these ideas with the above energy equation we nd:
E
h
2
2m
2

Ze
2

.
We now search for a value of which minimizes the total energy. This is a type of problem
which you have seen in rst semester calculus. At the minimal value of the energy, dE/d =
0. We therefore nd:
0 =
dE
d
=
h
2
m
3
+
Ze
2

2
=
1

2
(
h
2
m
+ Ze
2
).
=
h
2
mZe
2
.
We dene the Bohr radius, a
o
, to be the de Broglie wave length for the hydrogen atom where
Z = 1.
a
o
=
h
2
me
2
0.5

A
231
The size of the hydrogen atom is determined by the above relationship. As verication of
the accuracy of the above equations, note that:
E
K
=
mZ
2
e
4
2 h
2
,
E
P
=
mZ
2
e
4
h
2
.
The virial theorem, 2E
K
= E
P
, is obeyed by our model.
Finally we compare our expression for the de Broglie wavelength of the 1s electron with

1s
given in the fourth chapter of this textbook We nd that
1s
(r) = N
1s
e
r/
, where N
1s
is an unspecied constant. The 1s wavefunction is an exponentially decaying function whose
decay length is the de Broglie wavelength itself.
42.1.1 Translating the above equations into pictures
The previous section shows, on the one hand, that the radius of the hydrogen atom is
governed by the Bohr radius and, on the other hand, that the size of the atom comes
from the simultaneous optimization of both the kinetic and the potential energies. Lets
translate this story into pictures. For pictures, we consider two asymptotes, where the de
Broglie wavelength is either set to minimize only the kinetic or only the potential energy.
In Figure 42.1 we show the optimal wavefunctions for these two scenarios. In the former
case, where we wish to minimize solely the kinetic energy, we note E
K
= p
2
/2m = h
2
/2m
2
.
The longer is, the smaller the kinetic energy. E
K
is therefore minimized when = , see
Figure 42.1.
If, instead, we were to minimize just the potential energy, the electron would be located
entirely at the nucleus. The wavefunction would be like the one shown in the middle panel
of Figure 42.1; its de Broglie wavelength would be zero. The actual 1s wavefunction, which
232
Figure 42.1: which minimize respectively the kinetic, potential, and total energies.
optimizes both the kinetic and the potential energy is
1s
= N
1s
e
r/
, shown in the right
panel of the gure. The true 1s wave function is an exponential decay function which retains
a spike at r = 0 to help optimize the potential energy, but also keeps an exponential tail at
larger values of r to help minimize the kinetic energy portion of the total energy.
An example will help clarify the situation. In the previous section, we found that the de
Broglie wavelength was inversely proportional to Z, ie., 1/Z. As Z increases in size,
the potential energy term gets bigger, becomes smaller, N
1s
e
r/
becomes a more swiftly
decaying exponential function, the tail of the wavefunction gets smaller, and the spike grows
larger. An increase in the value of Z shifts the balance between the spike and the tail
component of the wave function so that the spike plays a greater role. This is in accord with
the greater role that potential energy would be playing.
To make an analogy, the kinetic energy and the potential energy are like two dierent
reins which steer a horse (the horse being the total energy). As we increase the needs of one
of the reins, the corresponding rein tightens up and that term dominates more. Conversely,
if one of the two reins loosens up, it allows the other rein to tighten and exert greater control.
The idea that the shape of the wavefunction changes as we disturb the balance between
kinetic and potential energies will prove to lie at the heart of the covalent bond.
233
42.2 The covalent bond
42.2.1 The H
2
molecule
We now consider the H
2
covalent bond. In this course, we have adopted a model where
we take the two exponential decay functions,
1s
= N
1s
e
r/
, one function centered at one
nucleus and the other at the other nucleus, and place these two wavefunctions in phase with
one another. As we have seen, the kinetic energy decreases as we do so. But in making this
model we made a perhaps unwarranted assumption, that the specic atomic orbitals which
make up the H
2
molecular orbital remain themselves unchanged. (This assumption was part
of the minimal valence basis set approximation.)
The assumption is unwarranted as the H
2
covalent bond has changed the balance between
the atomic kinetic and potential energy in their roles of optimizing the atomic electronic
energy. By introducing a second hydrogen atom, wavefunctions have become longer. Going
back to our analogy of reins, the kinetic energy rein has loosened up. The potential energy
rein, therefore, has tightened, and plays a more dominant role in determining the size of the
atomic de Broglie wavelength. The atomic orbtial wavelength given in the expression,
1s
=
N
1s
e
r/a
, called
a
here, to distinguish it from the full de Broglie wavelength, contracts,
lowering the total potential energy.
We use Figure 42.2 to illustrate the situation. In this gure we show the total molecular
energy as a function of the
a
decay constant, taken from the expression
1s
= N
1s
e
r/a
.
Note that there are two such 1s orbitals in the H
2
problem. Note further that for each value
of
a
, the interatomic distance between the two nuclei, R, adjusts.
We know there will be one optimal value of the atomic de Broglie wavelength,
a
which
minimizes the total energy of the H
2
molecule. We will call this value
a2
. This value is
shown in the gure. At this equilibrium geometry; the virial theorem is obeyed: E
K
= E
T
and E
P
= 2E
T
. We use these equivalences to place the values of both the kinetic and
234
Figure 42.2: The H
2
energy as a function of the atomic exponential decay parameter
a
. a1
and a2 correspond to respectively the otimal values of the lone hydrogen atom, H
1
, and the
hydrogen molecule, H
2
.
potential energies on the gure.
As
a
increases, the kinetic energy decreases; conversely as decreases, the total kinetic
energy increases. Further as
a
increases, the potential energy decreases; as decreases,
the potential energy increases. These relations are in accord with the requirement that
E
T
= E
K
+ E
P
. Also placed on the gure is the value
a1
.
a1
is the initial atomic de
Broglie wavelength, a length larger than
a2
42.2.2 A truncated Taylor series
We now apply one-dimensional calculus. We know that any function can be expressed as a
Taylor series around a specic value of the function, ie.,
E
T
(
a
) = E
T
(
a2
) +
dE(
a2
)
d
a
(
a

a2
) +
d
2
E(
a2
)
d
2
a
(
a

a2
)
2
2
+ ...
235
For values near
a2
we can truncate this expression to :
E
T
(
a
) E
T
(
a2
) +
dE(
a2
)
d
a
(
a

a2
).
Recalling that
a2
minimizes the total energy, this expression further simplies to:
E
T
(
a
) E
T
(
a2
)
We show this relation in Figure 42.2 by making the slopes of E
K
and E
P
equal in magnitude
and opposite in sign.
We can summarize the situation in Figure 42.3, a blow-up of the central portion of the
previous gure. We use this gure to follow the story of the covalent bond. In this course,
we considered a model of atomic orbitals xed in size, where
a
was taken to be a constant.
In this model, covalent bonds appeared to lower the kinetic energy, and, if we had thought
about it, raised the potential energy. But
a
is not a constant. Once the covalent bond
forms, the atomic orbital decay constant,
a
, becomes smaller. It becomes smaller as the
urgency of optimizing the kinetic energy has been partially relieved by the formation of the
in-phase MO.
In making the covalent bond,
a
shrinks in value from
a1
to
a2
. Importantly, the
shrinking barely changes the total energy. By the time
a
has shrunk to its nal value
a2
,
the kinetic energy has increased, and the potential energy has decreased to the point that
the virial theorem is obeyed, but the total energy has barely changed.
In summary, for the covalent bond, ultimately, kinetic energy increases and potential
energy decreases. But, we can understand the covalent bond perfectly well with a model
where the atomic orbitals have not contracted, the kinetic energy has decreasesd and the
potential energy increased. The model used in the course gives valid energies for while the
236
Figure 42.3: Blow-up of the central portion of the H
2
energy as a function of the atomic
exponential decay parameter
a
. a1 and a2 correspond to respectively the otimal values of
the lone hydrogen atom, H
1
, and the hydrogen molecule, H
2
. A Taylor series approximation
has been used.
indiividual kinetic and potential energies are both wrongs, the total energy is close to correct.
42.3 The ionic bond
The ionic bond, as always, is simpler. It will take us one paragraph and one gure. Consider
an Na
+
cation interacting with a F

anion. As these two ions approach, the electrostatic


energy decreases. Were it just up to the electrostatic energy, the cation and anion would fuse
into one. At some point however, the size of the atoms becomes involved. The size of the
atoms, as we showed in the beginning of this chapter, is another way of saying, the size of the
relevant atomic orbitals. As the lled atomic orbitals start interfering, the Pauli exclusion
principle requires that higher, unlled atomic orbitals be involved, THe kinetic energy of the
combined Na-F system increases. the equilibrium Na-F distance is reached when the virial
theorem is obeyed. Figure 42.4 illustates this story.
237
Figure 42.4: Blow-up of the central portion of the H
2
energy as a function of the atomic
exponential decay parameter
a
. a1 and a2 correspond to respectively the otimal values of
the lone hydrogen atom, H
1
, and the hydrogen molecule, H
2
. A Taylor series approximation
has been used.
42.4 Bonding across the periodic table
42.4.1 Second row main group elements
This chapter explains a great deal about the covalent bond. It suggests even more. Why is
it that C-C bonds are so much stronger than Sn-Sn bonds? Diamond melts at over 4000
o
C
while Sn melts at 200
o
C. Why does carbon make double and triiple bonds while Sn does not?
This chapter suggests a solution. The carbon valence orbitals are primarily 2p orbitals. 2p
orbitals are special, as there are no 1p orbitals. Carbon 2p orbitals have therefore no lower
in energy p orbitals to stop their contraction process. They contract enough that we can no
longer contract the Taylor series to the rst derivative term. The total energy term becomes
genuinely smaller.
If this is correct, and I only suspect that it is, there would be two consequences. First
C-C covalent bonds would become stronger. Second, the -orbitals would become so small
that the orbitals would become bigger and overlaps would increase. Carbon could make
238
good bonds, but tin would not.
42.4.2 Magnetic metal elements
This same story has been proposed (one of the proposers is Greg Landrum, a student of Roald
Homann) to be responsible for the magnetism of metallic elements. In Landrums model,
the 3d and the 4f atomic orbitals are both the rst of their kind. Both can therefore contract.
Unlike the 2p elements, what happens is that atomic orbitals with one spin contract. As a
consequence atomic orbitals with the opposite spin expand. The occupation of the up vs.
down spin orbitals is no longer equal, and among other things, a ferromagnet can result.
The metallic elements involved in ferromagentism are indeed 3d and 4f elements. On
the transition metal side, they are Fe, Co and Ni; for the rare earths Pr, Nd, Sm, Gd, and
Dy are often found: 3d and 4f elements all.
I wonder too about the hydrogen bond. Could the absence of a 0s orbital play a role
in the hydrogen bond? Maybe a student in this course can solve this last problem through
their own research. With that note of hope, we end the corse.
239
A brief reminder on how determine probability amplitudes and probabilities of
states under measurement. Consider a Hermitian matrix H with eigenvectors
v
1
, v
2
, v
3
and eigenvalues
1
,
2
,
3
. If H is a measuerment, then the possible
measured values are
1
,
2
, and
3
. Immediately after measurement, the system
will be in one of the states v
1
, v
2
, or v
3
.
Now consider a general state u. We already know the possible outcomes of
measurement Hon u, but with what probability would each value occur? There
are a couple equivalent ways of thinking about this. First, you could remember
that u is a combination of v
1
, v
2
, and v
3
:
u = a v
1
+ b v
2
+ c v
3
| u = a | v
1
+ b | v
2
+ c | v
3

Both lines say the same thing, but the second one is in braket notation. Here,
youd have to solve for a, b, and c which are the probability amplitudes u behaves
like v
1
after measurement (now only in braket notation):
| u = a | v
1
+ b | v
2
+ c | v
3

v
1
| u = av
1
| v
1
+ bv
1
| v
2
+ cv
1
| v
3

= a
Similarly, v
2
| u = b and v
3
| u = c. Now we know that if these are
the probability amplitudes, the probabilities should be the squares, and they
should add up to one. Heres how we check: the probability that | u behaves
like itself should be 1. That is, u | u
2
= u | u = 1. So,
u | u = 1
u | u = (a

v
1
| +b

v
2
| +c

v
3
|)(a | v
1
+ b | v
2
+ c | v
3
)
= a

av
1
| v
1
+ b

bv
2
| v
2
+ c

cv
3
| v
3

= a
2
+ b
2
+ c
2
= 1
You may be wondering how we did this without once using the measurement
matrix H its because we were working with eigenvectors of H. One can do
the meausurement explicitly, but its a lot more work. Well go through it here
for completeness sake (in regular and braket notation again). If you make the
measurement H, you get the new state:
Hu = aH v
1
+ bH v
2
+ cH v
3
H | u = aH | v
1
+ bH | v
2
+ cH | v
3

But remember that because v


1
, v
2
, v
3
are eigenvectors, H v
i
=
i
v
i
, which means:
Hu = a
1
v
1
+ b
2
v
2
+ c
3
v
3
H | u = a
1
| v
1
+ b
2
| v
2
+ c
3
| v
3

1
So what if we wanted to determine a, the probability amplitude that Hu behaves
like v
1
?
v
1

Hu = a
1
v
1

v
1
+b
2
v
1

v
2
+c
3
v
1

v
3
v
1
| H | u = a
1
v
1
| v
1
+b
2
v
1
| v
2
+c
3
v
1
| v
3

Because the eigenvectors are orthogonal and of unit length, this reduces to:
v
1

Hu

1
= a
v
1
| H | u

1
= a
2

Você também pode gostar