Vous êtes sur la page 1sur 196

William Klemperer

expanded by Pawel Swiatek

Chemistry 160:
Physical Chemistry
LECTURE NOTES
Copyright © 1999 President and Fellows of Harvard College
Table of Contents

Table of Contents:
LECTURE 1 Introduction ........................................................................................... 7
What is Chem 160? ........................................................................................................... 7
How to Use These Lecture Notes ........................................................................................ 8
Units and Conventions ....................................................................................................... 9

LECTURE 2 Classical Mechanics ........................................................................... 11


Newton’s Equations of Motion ......................................................................................... 11
Hamilton’s Equations of Motion ....................................................................................... 13
Systems with a Stable Minimum - The Simple Harmonic Oscillator ..................................... 14
Generalized Coordinates .................................................................................................. 16
Some Applications of Classical Mechanics ........................................................................ 21

LECTURE 3 Waves, Particles and Boxes ................................................................ 27


Properties of Waves and Particles...................................................................................... 27
The Schrödinger Equation................................................................................................ 30
The Particle in a Box ....................................................................................................... 31
Quantum-Based Quantities............................................................................................... 34
A Symmetric Box ........................................................................................................... 40
A Box with Finite Walls................................................................................................... 43
A Chemical Box ............................................................................................................. 45
Heisenberg Indeterminacy Principle .................................................................................. 47

LECTURE 4 More on the Schrödinger Equation.................................................... 49


Solution Properties .......................................................................................................... 49
The Schrödinger Equation in Two Dimensions ................................................................... 53
The Schrödinger Equation in Three Dimensions................................................................. 56

LECTURE 5 The Harmonic Oscillator ................................................................... 61


Solution with the Schrödinger Equation............................................................................. 61
Properties ....................................................................................................................... 65

LECTURE 6 Tunneling ............................................................................................. 67


A Divided Box................................................................................................................ 67

LECTURE 7 The Virial Theorem ............................................................................. 71


General Case .................................................................................................................. 71
Specific Examples........................................................................................................... 72

3
Table of Contents

LECTURE 8 The Boltzmann Distribution ...............................................................75


Relative Occupancy of Energy Levels ............................................................................... 75

LECTURE 9 Cylindrical Polar Coordinates ...........................................................77


The Schrödinger Equation ............................................................................................... 77
Particle on a Ring ........................................................................................................... 78

LECTURE 10 Spherical Polar Coordinates ..............................................................85


Motion in a Central Potential ........................................................................................... 85
Operators, Eigenvalues and Eigenfunctions ....................................................................... 87
The Schrödinger Equation ............................................................................................... 89

LECTURE 11 Hydrogenic Orbitals............................................................................93


Angular Solutions........................................................................................................... 93
Radial Solutions ............................................................................................................. 99

LECTURE 12 Perturbation Theory ......................................................................... 107


First and Second Order.................................................................................................. 107
Two Interacting Levels .................................................................................................. 111

LECTURE 13 Time Dependent Schrödinger Equation .......................................... 117


Solving with Time Dependence...................................................................................... 117
Time Dependent Behavior ............................................................................................. 120
Oscillation Between Interacting States ............................................................................ 124
Optical Transitions ....................................................................................................... 125

LECTURE 14 Electron Spin .................................................................................... 131


Magnetic Moment of an Electron ................................................................................... 132

LECTURE 15 The Multielectron Atom ................................................................... 135


Two Electron Systems................................................................................................... 136
Orbital Theory: Singlets and Triplets .............................................................................. 138
Angular Momentum of Two Particles ............................................................................. 143
Spin-Orbit Interaction ................................................................................................... 144
Term Symbols .............................................................................................................. 145
Hund’s Rules ............................................................................................................... 148

4
Table of Contents

LECTURE 16 The Variational Method....................................................................149


Basis Functions............................................................................................................. 149
Proof ........................................................................................................................... 151
Functions and Parameters............................................................................................... 152
Orbital Theory .............................................................................................................. 152

LECTURE 17 Models of Molecular Bonding .........................................................157


The Morse Function ...................................................................................................... 157
The Born-Oppenheimer Approximation .......................................................................... 159

LECTURE 18 Ionic Models of Bonding ..................................................................161


The Rittner Model......................................................................................................... 161
Collisions of Polarizable Spheres .................................................................................... 164
Long Range Electrostatic Interactions ............................................................................. 165
Van der Waals Forces .................................................................................................... 167

LECTURE 19 Models of the Covalent Bond ...........................................................171


Valence Bond Theory .................................................................................................... 171
Molecular Orbital Theory .............................................................................................. 176
Classification of Molecular Orbitals ................................................................................ 179
Molecular Term Symbols ............................................................................................... 180
Covalent Bond Data ...................................................................................................... 181
Polyatomic Molecular Orbital Theory ............................................................................. 183

APPENDIX I Problems............................................................................................187

APPENDIX II Physical Constants and Unit Conversions ......................................193

APPENDIX III Mathematical Reference...................................................................195

5
LECTURE 1 Introduction

Welcome to Chem 160, Physical Chemistry, affectionately referred to as p-chem.


You have probably heard a lot about this class. You have scrutinized the CUE guide,
you have talked to your friends. It seems that every time you have mentioned p-chem,
you were either countered with pity or the question “are you insane?” Most likely, you
have been told that “it’s just a lot of really abstract math that has next to nothing in
common with reality.” Chem 160 is designed to contradict these uninitiated infidels. The
key thing to note is that it is called Chem 160, not Physics 143a. This means that we will
develop topics in a fashion suitable to the chemist and with chemical applications in
mind. Presumably, this is why you are taking this class.

What is Chem 160?


We will begin in the familiar territory of Newtonian mechanics and Newton’s
equations of motion. These concepts will serve as the basis for delving into Hamilton’s
equations of motion and such concepts as generalized coordinates and phase space. We
will proceed to develop a classical description for particle collisions (which then lead to
chemical reactions), as well as analyze the classical behavior of the harmonic oscillator.
Our ultimate goal will be to describe how atoms bond. As you are probably
aware, bonding interactions are dominated by electrons, which are too small to be
described classically. We will, therefore, develop a quantum mechanical approach. First,
we will develop some quantum tools and vocabulary. The first will be the famous
Schrödinger wave equation. The other concepts will range from the familiar “wave-
length” and the ominous-sounding “eigenfunction” to the exotic “expectation value.”
Armed with these, we will attempt the first quantum problem – the particle in a one-
dimensional box. With the solutions in hand, we will then test this model on conjugated
linear polyenes. While still in one-dimension, we will recall the harmonic oscillator
problem and solve it using quantum mechanics. We will also briefly discuss the idea of
tunneling.
The next step will be to move into three dimensions, as we start thinking about
hydrogen and its atomic orbitals. In freshman and organic chemistry, you used the con-
cept of an orbital, but the real origin of these mathematical constructs was conveniently
labeled as “beyond the scope of this course.” To fill in these gaps, we will demonstrate
and explain the mathematical origin, as well as the physical significance of wavefunc-

7
Introduction: How to Use These Lecture Notes

tions and orbitals. Having developed the hydrogen atom, we will take a moment to learn quantum descriptions of inter-
acting systems. These will include Perturbation Theory and the coupling of two wavefunctions with a transition moment
dipole.
Because hydrogen is only one of the many known elements, we will then build a model to describe polyelectronic
atoms using the Hartree-Fock Approximation. We will revisit familiar ideas, such as the Pauli Exclusion Principle, Hund’s
Rules, the Aufbau Principle. Based on these, we will develop electronic state descriptions, and we will analyze allowed
and forbidden electronic transitions.
Atoms are only one step removed from molecules, our next subject. We will look at how atoms interact at different
distances, and we will develop several models of varying complexity to represent these interactions. We will start with a
simple ionic model of two charged points interacting. We will then include the polarizability of atoms and the more com-
plex model of dipole-dipole interactions. The next step will be covalent bonding and the idea of overlapping atomic
functions, which we will juxtapose with the already familiar molecular orbital model. As we become familiar with the
electronic structure of simple molecules, we will explore experimental methods of verifying our theoretical results.
At this point, you probably want to ask: “Those concepts sound reasonable enough, but isn’t it all just complicated
math in the end?” The answer is a resounding no. Doubtless, throughout the course you will have to integrate and differ-
entiate more than one equation. However, the goal of this class is neither to teach nor to test advanced calculus and dif-
ferential equations. Most of the mathematics you are likely to encounter this semester is summarized at the beginning of
this sourcebook. Become familiar with the integrals without memorizing them. Throughout the semester they will come up
with sufficient frequency that you will learn them without having to resort to brute force. The extent of differential calculus
you will be asked to reproduce on examinations will also be minimal. Once we solve the Schrödinger equation, the impor-
tant thing to know in each case will be the solutions, not the derivation.
Simultaneously, do not think that you will be able to gain a solid grasp of the concepts without devoting a solid
amount of time to the material. For most people, quantum mechanics is an epiphany-based experience, where it just
takes many tries until it “clicks.” Definitely take the time to do the problem sets. If nothing else, they will help you find
your weaknesses. Read the textbook. Read these notes. If you make an honest effort, by the end of the semester you will
have learned an impressive amount of material that will forever change the way you look at chemistry and chemical reac-
tions.

How to Use These Lecture Notes


The lecture notes have been completely reworked this year. Instead of being just a series of cryptic derivations and
mathematical expressions, they feature more detailed explanations. The lectures will follow the sourcebook very closely,
both in the order and the depth of presentation. To take full advantage of the lectures, you should skim through the
appropriate section of the notes prior to the lecture. This way, even though you may not gain a complete understanding
of the material after the first reading, you will at least be able to know what to expect. While memorizing all the equa-

8
tions might prove too daunting, you should definitely be able to recognize the important final results of derivations and
use them correctly to solve problems. You may also have noticed that a third of each page has been left blank. This is
provided for your notetaking convenience. For example, not every step of each derivation has been worked out, and you
should definitely take the time to fill in the missing ones. Should you run out of space, there is always some room at the
end of each section.

The material is presented as a series of “Lectures,” which are orgranized thematically. In almost every case, it will take
longer than ninety minutes to discuss the material in a section. Once the material has been covered, however, you should
attempt to solve the problems featured at the end of each section. They will not only be representative of what you can
expect on exams, but more importantly, they will function as a test of your understanding. Make sure to address any
questions that arise with your teaching fellow.

Units and Conventions


Physical chemists have refused to adopt SI units with as much fervor as the United States has clung to the English system.
Even journal publications continue to use cgs or atomic units. It turns out that for the quantities we will be measuring,
they are simply more convenient. If you are not familiar with either, take a quick look at Appendix II, where you will find
a summary of units and conversion factors. Throughout these notes, assume cgs values for constants, unless specified
otherwise.

All vectors are printed in italics with an arrow above, e.g. p . We feel that this notation will lend more clarity than
the traditional textbook bold-face print. Similarily, operators are shown with a “hat”, e.g. Ĥ .
LECTURE 2 Classical Mechanics

The basis for the physical understanding of chemical phenomena is primarily


through mechanics. There is today a universal acceptance of the electron structure the-
ory of matter. Thus the mechanics of electron motion in atoms, molecules, solids and
liquids is fundamental to chemistry. The understanding of the motion of electrons in the
coulombic field of the nucleii requires quantum mechanics, and in general, the dynam-
ics of atomic units can require quantum mechanics. For many problems, however, clas-
sical mechanics remains adequate. The clarity and relative simplicity of classical
mechanics, together with the determinism provided effectively by present day comput-
ing technologies has resulted in the ever increasing classical treatment of molecular
motion. Secondly, the character of the quantum mechanics is seen more clearly by com-
paring it to classical mechanics.

The task at hand is to describe the mechanics or the motion of the system in terms
of its energy. We start with Newton’s Equations in Cartesian coordinates.

Newton’s Equations of Motion


Let’s start out with the traditional definition of force, Force = Mass × Acceleration,
where both force and acceleration are vector quantities.

F = ma Eq. 2-1

Since force is a vector, we can choose to scrutinize the x-component,

2
dx dv x dx
Fx = m = m , where v x = velocity =
dt2 dt dt
dv d
F = m ------ = ----- mv since mass is a constant
dt dt

Linear momentum, p , is then defined as p = mv . It is also a vector that can be


resolved into directional components, e.g. p x = mv x . Linear momentum is also a good
way of expressing force.

11
Classical Mechanics: Newton’s Equations of Motion

dp
F = ------- Eq. 2-2
dt

Kinetic Energy. Kinetic energy, T, is a scalar quantity that measures the energy of motion. It can be expressed in a num-
ber of ways, as shown by Eq. 1-3.

1 1 1 p2 1 1
T = --- m { v x2 + v y2 + v z2 } = --- m ( v ⋅ v ) = --- mv 2 = -------- = --- { v x p x + v y p y + v z p z } = --- ( v ⋅ p ) Eq. 2-3
2 2 2 2m 2 2
1
= -------- ( p ⋅ p )
2m

In Cartesian coordinates, T is a function of the velocity, but not the position of the particle. Momentum components
∂T
can be expressed in terms of T as mv x = p x = . Using this relationship, Newton’s equation for force can now be
∂vx
written as

dp x d ∂T
Fx = = Eq. 2-4
dt d t ∂ vx

We have now successfully expressed force in terms of the kinetic energy of the system. Now for potential energy.

Potential Energy. For many molecular problems the motions may be idealized to be frictionless. If the forces are such
that F x = F x ( x, y, z ) , but does not depend on velocity, then the potential energy V is defined as V ( x, y, z ) so that


– V ( x, y, z ) = F x Eq. 2-5
∂x

or more generally

x 1, y 1, z 1 q1

∆V = V ( x 2, y 2, z 2 ) – V ( x 1, y 1, z 1 ) = ∫
x 2, y 2, z 2
F x dx + F y dy + F z dz = ∫ ( F ⋅ ds )
q2

V(x, y, z) is a scalar function of position. We can now combine Eq. 2-5 with Eq. 2-4 to completely relate the motion and
the energy of the system.

12
Classical Mechanics: Hamilton’s Equations of Motion

d ∂T ∂
= F x = – V ( x, y, z ) , with similar expressions for y and z Eq. 2-6
d t ∂ vx ∂x

Total Energy. From the kinetic and potential energy two energy related functions, the total energy E = T + V . The dif-
ference function L = T – V is also useful (see below).

Newton’s laws of motion are relations between cartesian vector components, thus they appear to depend explicitly
upon the choice of coordinate system. There is a considerable advantage in casting the equations of mechanics in terms
∂ ∂L ∂T ∂L
of derivatives of the scalar energy functions. Since F x = – V ( x, y, z ) = and p x = = , Newton’s equations of
∂x ∂x ∂ vx ∂vx
d ∂T ∂
motion are + V ( x, y, z ) = 0 , or using the difference function L = T ( v ) – V ( x, y, z )
d t∂ vx ∂ x

d ∂L ∂
– L ( x, y, z ) = 0 Eq. 2-7
d t ∂ vx ∂ x

Hamilton’s Equations of Motion


There is some advantage in using as fundamental descriptors the position and the momentum rather than position
and velocity. The energy for the particle would appear to be a function of its position and momentum, both of which are
vector quantities. We now define this function, H. It is called the Hamiltonian.

p⋅p
H ( p, q ) = ---------- + V ( q ) Eq. 2-8
2m

By differentiating, we can see that H corresponds to quantities described by Newton’s equations. Thus
∂ ∂ dp x p⋅p ∂ 2p x dx
H ( p, q ) = V ( q ) = – Fx = – , since ---------- is a constant, and H ( p, q ) = -------- = , with similar expressions for
∂x ∂x dt 2m ∂ px 2m dt
y and z. If the energy is fixed, then the Hamiltonian gives the total energy of the system, E, which is the sum of the kinetic
and potential energies.

p⋅p
H ( p, q ) = ---------- + V ( q ) = E Eq. 2-9
2m

13
Classical Mechanics: Systems with a Stable Minimum - The Simple Harmonic Oscillator

Example 2-1. To see that this approach gives us a self-consistent description, consider the motion of a particle in a sin-
p2 dx 2(E – V(x))
gle dimension for which -------- + V ( x ) = E and = ------------------------------ . Note the absence of a t term on the right side. Formal
2m dt m
t x
m m x 1
integration yields ∫ 0 dt = t = ∫x dx ----------------------------- =
2 ( E – V(x))
------ ∫ dx -------------------------- , where x 0 = x ( t = 0 ) . Of course, the result-
2E x0 V ( x )
0 1 – ----------- -
 2E 
ing expression

x
m
t = ∫x dx0
-----------------------------
2(E – V(x))
Eq. 2-10

is not very useful; it is the opposite of what is desired, namely x ( t ) and p ( t ) in terms of the energy and initial, t = 0 ,
conditions. It does, however, show that if we start with Hamilton’s equations, we can find an expression that relates time
and position.

Systems with a Stable Minimum - The Simple Harmonic Oscillator


Let us continue to examine the particle from Example 2-1. In general, systems with a stable minimum can be math-
ematically described as follows:

2
dV dV
= 0, at x = x e and > 0, at x = x e . We can expand the function V ( x ) about x e .
dx dx2
3
1 d V  1 d V 
2

V ( x ) = V ( x e ) + ---  2  ( x – x e ) 2 + ---  3  ( x – x e ) 3 + …
2 d x  x e 6 d x 

For small values of the energy and motions where the displacement from equilibrium x ( t ) – x e is small, the first two non-
zero terms of the expansion suffice. Thus,

2
1 d V  1
V ( x ) = V ( x e ) + ---  2  ( x – x e ) 2 = V ( x e ) + --- k ( x – x e ) 2 Eq. 2-11
2 d x  x e 2

14
Classical Mechanics: Systems with a Stable Minimum - The Simple Harmonic Oscillator

m x
--------------------------------------------------------------------------------
If we use Eq. 2-11 with the expression in Eq. 2-10, we obtain t = ∫ dx 
2
1 d V   . If we let
2  E – V ( x e ) – --
-  2  ( x – x e ) 2

x0
 2 d x  xe 

E – V ( x e ) = E' , we can simplify the expression further:

x dx
m -----------------------------------------------------------------
t = -------
2E' ∫  1 d V 
2

Eq. 2-12
x0 1 – ------- d x 2  ( x – x e ) 2
 2E'   xe 

We are still looking for a useful expression of x in terms of t. The next step involves using the integral
dy
∫ -----------------
1 – y2
- = sin-1 y .

( 1 – sin x )dx = ( 1 – y 2 )dx .


2
Note that if y = sin x and x = sin-1 y , then dy = cos x dx or dy =
dy dy
Separating terms ∫ dx = ∫ -----------------
1 – y2
- and integrating yields x = sin-1 y = ∫ -----------------
1 – y2
- .

This is the form of Eq. 2-12. Thus, there is a trigonometric relation between x and t, that can be expressed as
x ( t ) – x e = A sin ( ωt + δ ) , where x ( t = 0 ) – x e = A sin δ . The maximum displacement from the equilibrium x – x e is A,
dx ( t ) 2 ( E – V ( xe ) ) 2E'
where ------------- = 0 = -------------------------------- . If we use our previous notation of E – V ( x e ) = E' , then x – x e max = -------- = A ,
dt m k
the amplitude of the sine function.

The Period. The maximum displacements are called turning points, where all of the velocity vanishes, as does the
kinetic energy. Finally, since linear momentum is intrinsically tied to velocity, we can write
dx p2 1
p = m ------ = Amω cos ( ωt + δ ) . And since E′ = -------- + --- k ( x – x e ) 2 (that is the sum of kinetic and potential energies), we
dt 2m 2
can substitute the trigonometric descriptions to get the final result A 2 mω 2 cos ( ωt + δ ) + kA 2 sin ( ωt + δ ) = 2E' . Energy
2 2

for a given system is fixed in time, so the expression must be time independent. In order to satisfy this condition, the trig-
2 2
onometric functions must add up to unity according to sin x + cos x = 1 . The condition is met if A 2 mω 2 = kA 2 or

15
Classical Mechanics: Generalized Coordinates

k
ω = ----- Eq. 2-13
m

which is a general rule for simple harmonic oscillators. To determine x ( t ), p ( t ) we need to know the initial t = 0 condi-
k
tions. However since ω = ----- , the frequency of the oscillator is an intrinsic property, independent of the initial conditions
m
or the amplitude of oscillation.

The Orbit. Another way of looking at harmonic oscillation and its periodic behavior is by analyzing the relationship
between the kinetic and potential energies, where the total energy, their sum, remains constant in time. The 1-dimen-
p2 1
sional case can be described as E = -------- + V ( x ) , where V ( x ) = V ( x e ) + --- k ( x – x e ) 2 as developed earlier through a sec-
2m 2
p2 1
ond order Taylor expansion around x e . Recalling that E – V ( x e ) = E' , we write -------- + --- k ( x – x e ) 2 = E – V ( x e ) = E' or
2m 2
p 2 + km ( x – x e ) 2 = 2mE' . A rearrangement yields

p 2
 --------
- + ( k ( x – x e ) ) = 2E'
2
Eq. 2-14
 m

which is an equation of an ellipse. This tells us that there is a periodic relationship between the momentum and the posi-
tion of the particle that repeats as the particle oscillates.

Generalized Coordinates
When deriving the Hamiltonian, we introduced the generalized coordinates of position and momentum. There, we
used the momentum and position vectors ( p, q ) . These two vectors exist in 6-dimensional phase space, and have three
components each, analogous to the x-, y-, and z-components of a Cartesian vector. Let us now explore this coordinate
system using q i and p i . Note that the previously used ( p, q ) = ( ( p 1, p 2, p 3 ), ( q 1, q 2, q 3 ) )

16
Classical Mechanics: Generalized Coordinates

d ∂L ∂L ( x, y, z )
In Cartesian coordinates, the difference function L = T – V and ----- ------- – -------------------------- = 0 , with similar expressions
dt ∂v x ∂x
∂L ( x, y, z ) ∂L ∂q i dx ·
for y, z. In new coordinates, we define -------------------------- =
∂x ∑ ------
- ------- . To maintain legibility, let us use
∂q i ∂x
------ = x = v x .
dt
i

· ·
∂L ∂L ∂L ∂q i ∂q i ∂q i
Thus, ------- = ------· =
∂v x ∂x
∑ ∂q· ∂x·- , where we note that ------
------
- ------
∂x ∂x
- . We can now rewrite
·- = ------
i i

d ∂L ∂L ( x, y, z ) ∂L ∂q i ∂L ∂q i d ∂L ∂L ∂ q i
----- ------- – -------------------------- =
dt ∂v x ∂x ∑ ------
·- ------- – ∑ ------
∂q ∂x
- -------
∂q i ∂x
= ∑ ----
- ------- – -------  -------
dt ∂q· ∂q i  ∂x
= 0 Eq. 2-15
i i i i i

∂q i d ∂L ∂L
------- are coordinate relations and are unconnected to the dynamical relations, therefore ----- ------·- – ------- = 0 , just as for
∂x dt ∂q i ∂q i
Cartesian coordinates (see Eq. 2-7).

∂L
------· = p x in Cartesian coordinates, where p x is the momentum associated with the coordinate x. Similarily, in gen-
∂x
∂L
eral coordinates ------·- = p i , where p i is the momentum associated with the coordinate q i .
∂q i

Recall that for a fixed energy, the Hamiltonian is the sum of the kinetic and potential energies of the system, or in
symbols, H = T + V . This is very similar to the difference function, and we can easily derive analogous dynamical equa-
tions for the general coordinate q i with associated momentum p i . The expressions are:

∂H dp ∂H dq i
------- = – --------i and ------- = -------- Eq. 2-16
∂q i dt ∂p i dt

The explicit derivation is left as an exercise. For a shortcut, look at Eq. 2-9.
∂H ∂H dp i
Conservation Principles. If H, the total energy, is independent of q i , then ------- = 0 , but since ------- = – -------- , this implies
∂q i ∂q i dt
dp i
-------- = 0 and p i is constant. The result is easy to see for linear momentum if q i is the Cartesian coordinate; constant
dt
momentum and total energy over all possible positions imply that the total energy does not depend on the position of

17
Classical Mechanics: Generalized Coordinates

the particle. Note, though, that in this example it is the potential energy V that is independent of q i , since in Cartesian
coordinates T will always be independent of q i . The conclusion is that if the potential is independent of position, then
the particle moves freely.

Example 2-2. Consider a problem in which the potential energy depends only on the distance of a particle from a cen-
ter. One possibility is the relative motion of a light particle with respect to a very heavy particle, i.e. earth and sun or
electron and proton. The center of mass of the system is essentially at the heavy particle. If the center of mass’s motion is
ignored, it is only the motion of the light particle that we are examining.
Let V = V ( r ), r = x 2 + y 2 + z 2 and remember that H = T + V . First, describe the motion of the light particle in terms
∂H · ∂H ∂r ∂H x · ∂H y · ∂H z
of the Cartesian coordinates. The dynamic expressions are ------- = – p x = ------- ------ = ------- --- , – p y = ------- -- , and – p z = ------- --- . As
∂x ∂r ∂x ∂r r ∂r r ∂r r
∂r x
an exercise, convince yourself that ------ = --- .
∂x r
Since this is a relative motion problem, and we are assuming a stationary center of mass, the kinetic energy actually
depends on the angular momentum of the system. Angular momentum is defined as L = r × p , or alternately
l z = xp y – yp x , l x = yp z – zp y , l y = zp x – xp z . Taking the time derivative,

dl z · · · ·
------- = ( x p y – y p x ) + ( xp y – yp x ) Eq. 2-17
dt
px py ∂H – y
=  ----- p y – ----- p x +  -------  x ------ – y ------ 
–x
m m   ∂r  r r 

= 0 (always) + 0 (because V ( x, y, z ) = V ( r ))

dl z
The conclusion is that ------- = 0 and l z is a constant in time. An identical argument can be made for l x and l y . If all of
dt

dL
these components of L are constant in time, then L itself is constant in time: ------ = 0 .
dt
Note that for V = V ( r ) , r × p is a constant, but neither r nor p must be. They must vary in a coordinated manner so that

dL d
------ = ----- r × p = 0 .
dt dt

18
Classical Mechanics: Generalized Coordinates

Generalization to n particles. For a general system with n particles, Eq. 2-16 still holds.
∂H dp i ∂H dq i
------- = – -------- and ------- = -------- , except now there will be 3n values of i, to account for the 3n degrees of freedom in the system.
∂q i dt ∂p i dt

Example 2-3. Mechanics of system of two particles (a, b) interacting by central potential.
Let us now attempt to use the n particle generalization for n = 2 . Keep in mind that every quantity described in the gen-
eral coordiante system has three degrees of freedom, and therefore three coordinates. For example, the position of par-
ticle a is described by ( q xa, q ya, q za ) , or more succinctly ( q 1, q 2, q 3 ) .
6
p i2
Kinetic Energy = ∑ --------
2m i
-; m i = m a for i = 1 → 3; m i = m b for i = 4 → 6
i=1

and the six linear momenta p i correspond to p 1 = p xa , p 2 = p ya , p 3 = p za , p 4 = p xb , p 5 = p yb , p 6 = p zb .


ma xa + mb xb ma ya + mb y b ma z + mb zb
The center of mass has the coordinates R 1 = -------------------------------- , R 2 = ------------------------------- , R 3 = ----------------------------- with M = m a + m b .
M M M

To simplify the notation, we can combine the three components into one position vector R = R 1 + R 2 + R 3 .
Total linear momentum can also be combined into vector form as the sum of three components
Pj = pi + p i + 3 , j = 1 → 3 .

Any problem involving the relative motion of two bodies can be simplified by eliminating the explicit dependance
on the two individual masses. In order to achieve this, we apply the construct of a reduced mass, µ .

ma mb
reduced mass = µ = -------------------- Eq. 2-18
ma + mb

The problem consists of two parts:


d d
1. Motion of the center of mass with ----- ( m a q i + m b q i + 3 ) = ----- MR i = P j
dt dt
2. Relative motion of a “particle” mass µ moving under a potential V ( r ) , where r = r a – r b . If we define
3 2
 
q j' = q i – q i + 3 and p j' = p i – p i + 3 , then r =  ∑ q j' .
j = 1 

19
Classical Mechanics: Generalized Coordinates

To approach the first part, we note that since V depends only on r (by definition of a central potential),
∂H
H ( R, r, P, p ) does not depend upon R , the position of the center of mass. In other words, ------- = 0 for j = 1 → 3 and
∂R j

dP j dP
-------- = 0 for j = 1 → 3 , which implies that ------- = 0 , or that P , the total momentum of the center of mass, is constant in
dt dt
time and space.

To approach question of relative motion, we note that the kinetic energy of relative motion consists of two parts.
2
l 2
1. Centrifugal, described by -----------2 , where l = l x2 + l y2 + l z2 is the angular momentum.
2µr
p r2
2. Radial, described by ------ , where p r is the radial momentum.

By combining the two kinetic energy expressions, we can write down the Hamiltonian ( T + V = E ) for relative motion:

2
p r2 l
------ + -----------2 + V ( r ) = E Eq. 2-19
2µ 2µr

2
Since l is conserved, l becomes a constant and the radial motion equation can be rearranged as
l2 l2
2µ E – V ( r ) + -----------2  . Reacall that p r = µ ------ , so a simple substitution yields ------ = --- 2µ E – V ( r ) + -----------2  .
dr dr 1
pr =
  2µr  dt dt µ   2µr 
Taking further advantage of the constant angular momentum, we define the effective potential for radial motion as
l2
V eff ( r ) = V ( r ) + -----------2 to arrive at
2µr

dr 1
------ = --- 2µ ( E – V eff ( r ) ) Eq. 2-20
dt µ

This procedure has converted the complex motion of two particles (6 degrees of freedom) into a set of essentially
one dimensional problems:
1. The center of mass moves with constant linear momentum. (3 degrees of freedom)
2. The angular momentum is constant. (2 degrees of freedom — what are they?)

20
Classical Mechanics: Some Applications of Classical Mechanics

3. The radial motion is governed by the magnitude of the angular momentum only, and not the orientation of this vec-
tor. (1 degree of freedom)

2
l
Note that the term -----------2 looks like a potential energy term for the radial motion.
2µr

Some Applications of Classical Mechanics


Although classical mechanics are not applicable at the electronic level, molecular behavior
Rc can still be conveniently described using the methods developed so far. We will now present
the classical approach to molecular collisions and the closely related result, the Beer-Lambert
law for absorption.

Collisions. Let us first assume that we are dealing with hard spheres, which touch when their
distance is R c , shown in Figure 2.1. We define the collision cross-section σ = πR c2 and the
Figure 2.1 - Two particles shown in collision frequency (or the reaction rate constant) as k = vσ = vπR c2 , where v is the velocity
contact. The area of the large circle
is the collision cross-section. of the particles. We also assume that the particles are neither pulled in nor pushed out by the
potential. In accordance with the previous derivation, we define the effective potential for
l2 dr
radial motion as V eff ( r ) = V ( r ) + -----------2 , and using Eq. 2-20, we write p r = µ ------ = 2µ ( E – V eff ( r ) ) . Note that the mini-
2µr dt
dr
mum value of r occurs at the turning point, p r = µ ------ = 0 or where E = V eff ( r ) .
dt

In order to use these results to describe a collision, we need a relation between angular momentum, energy and the
geometric factors. This relation is called the impact parameter, b. In this problem the potential V ( r ) is central, thus angu-
lar momentum is a constant. We may evaluate the angular momentum,
l = r × p or l = rp sin θ = bp at any position. At large distances, for any potential, the inter-
p⋅p ( bp ) 2
p
l2 b2E
action vanishes*, and the total energy is kinetic energy, E = ---------- , then -----------2 = ------------2- = -------- .
2m 2µr 2µr r2 b
r

Figure 2.2 - A graphical


1 representation of the impact
*. The potential of the form V ( r ) = --- is a special case and will be treated later. parameter.
r

21
Classical Mechanics: Some Applications of Classical Mechanics

The impact parameter turns out to be the distance of closest approach if the potential energy is zero. We can sum-
marize the general relationship between the potential and the distance as
V eff V eff max
V = 0, r ≥ R c
V = ∞, r < R c
b2E
V eff = --------, r ≥ Rc
r2
V eff = ∞, r < R c r

Note that for b > R c the effective potential is greater than E, so the particle is
“reflected” at distances greater than R c . Figure 2.3 - A qualitative
representation of the interaction

For the hard sphere system a collision occurs if b ≤ R c . If b c is the largest impact
parameter for which a collision takes place, then for hard spheres b c = R c . The collision cross section for hard spheres
2E
is then σ = πR c2 and the collision frequency k = σv = ------πR c2 , where v was written in terms of the relative kinetic
µ
2E
energy of the particles as v = ------ with µ equal to their reduced mass.
µ

The hard sphere model can be modified to describe ion-molecule reactions if we introduce a non-zero potential of
αe 2
V ( r ) = – --------4 , where α , the polarizability, which is roughly equivalent to the molecular volume, and e is the fundamental
2r
αe 2 l2
charge. The effective potential expression now becomes V eff = – --------4 + -----------2 , or if we use l = µbv ,
2r 2µr
αe 2 b 2 E
V eff = – --------4 + -----------2 . Clearly, the effective potential depends parametrically upon E and b. The maximum value of V eff
2r 2µr
dV eff 2αe 2 2Eb 2 2
αe 2
occurs at ----------- = 0 = ----------- 3 - , or solving for r max = --------2 . Substituting back into the effective potential equation,
5 - – -----------
dr r r Eb
( Eb 2 ) 2
V eff max = --------------- .
2αe 2

22
Classical Mechanics: Some Applications of Classical Mechanics

dr
The critical point is where ------ = 0 or where E = V eff . For the particle to reach the interior of the potential (to get
dt
( Eb 2 ) 2
over the hump), it must have E ≥ V eff max or E ≥ --------------- . In terms of the impact parameter and the collision cross-section
2αe 2
2αe 2 2αe 2
b2 = ------------ and σ = π ------------ .
E E

At low energies (room temperature) this cross section, the so-called Langevin cross section, exceeds the hard
sphere cross section by about an order of magnitude.

The rate constant, k, or constant of collision frequency is then k = vσ , where v is the velocity of the particles.
2E 2αe 2 2E 2αe 2 α
Since v = ------ and σ = π ------------ , k = ------π ------------ = 2πe --- . The key result to note is that the collision frequency
µ E µ E µ
or the rate constant is independent of energy for the ion-molecule collision. The universality of the ion-molecule collision
α
frequency is further increased when the quantity --- is examined for different species, as listed in Table 2.1.
µ

Table 2.1 - Values for various molecules showing the consistency of the ion-molecule collision model.

α
----
Mass [ g × 10 24 ] α [cm3 × 10 24 ] M
Ar 64 1.66 0.026
N2 45 1.77 0.039
C6H6 125 10.4 0.083
CO2 70 2.63 0.037
H2O 29 1.48 0.051

α
The variation of ---- in the table is less than factor of two, thus we would expect that the ion-molecule rate con-
M
αe 2
stant is k = 2π -------- = 6.3 × 4.8 ×10 × 0.2 = 7 ×10 cm3 molecule-1 sec-1, where we used the cgs value for e and 0.2
– 10 – 10
µ
α
for ---- . Results for several reactions are shown in Table 2.2.
M

23
Classical Mechanics: Some Applications of Classical Mechanics

Table 2.2 - Rate constants for some exothermic ion-molecule reactions

kobserved α µ ktheoretical
Reaction [cm s × 10 ]
3 -1 –9
[cm × 10 ]
3 24
[g × 10 – 24
] [cm3 s-1 × 10 –9 ]
H2+ + H2 → Η3+ + Η 2.1 0.82 1.6 2.1
CO+ + H2 → ΗCO+ + Η 2.0 0.82 3.0 1.6
He + CO → C + Ο + He
+ + 2.0 2.0 5.6 1.8
H3+ + CO → ΗCO+ + H2 1.4 2.0 4.3 2.1
He+ + H2 → H2+ + He < 10-4 0.82 2.1 1.9
→ H+ + H + He

d[C]
The rate of the reaction A + B → C is written as ------------ = k [ A ] [ B ] , where [ A ] is the concentration of species A, and k is
dt
the corresponding rate constant. The units of k in this case are concentration-1 seconds-1. In the above expressions, we
are using the units of concentration cm-3.

The Beer-Lambert Law. A natural extension of collisions is the Beer-Lambert Law for absorptions. Begin by defining
σ ( ν ) as the collision cross-section for the absorption of a photon per molecule in units of cm2. Since the velocity of the
photon is much larger than the velocity of the individual molecules, we neglect molecular motion. The molecular con-
centration is defined as n, or the number of molecules per cm3. Finally, the photon flux or light intensity entering a
region of thickness dL is I ( ν ) . The light intensity leaving the region will be I – dI , where the reduction is due to absorp-
dI
tion via collisions. In symbols, dI = – Inσ ( ν )dL , or equivalently ----- = – nσ ( ν )dL . If the length of the absorption cell is L
I
and the light intensity entering is I0, then the light transmitted, It, is obtained by integrating.

It L
dI
∫ ----I- = ∫ – nσ ( ν ) dL
I
0 0

It
ln --- = – nσ ( ν )L
I0

The final result is usually written as:

24
Classical Mechanics: Some Applications of Classical Mechanics

I t = I 0 e –nσ ( ν )L Eq. 2-21

Clearly, the amount of light absorbed is I 0 – I t . To make the expessions independent of the light intensity I0, we define
It I0 – I t
the transmission and absorption functions of ν , T ( ν ) = --- = e –nσ ( ν )L and A ( ν ) = ------------ = 1 – T ( ν ) = 1 – e –nσ ( ν )L , respec-
I0 I0
tively.

This reduction of intensity of a beam (of particles) is totally general. The intensity of a beam of molecules after pas-
sage through a chamber of gas with number density n and length L is then I t = I 0 e –nσL , where σ is the total scattering
cross-section forthe molecular collision.

25
Classical Mechanics: Some Applications of Classical Mechanics

26
LECTURE 3 Waves, Particles and Boxes

One of the most important concepts in quantum mechanics, and one that is perhaps
the most difficult to grasp, is the wave-particle duality of matter and energy. For example,
both photons and electrons can form interference and diffraction patterns (wave-like behav-
ior), yet their particulate nature is evidenced in the photoelectric effect. In this lecture, we
will discuss the behavior of photons as particles and waves. We will then introduce the
famous Schrödinger wave equation and apply it to the simplest case of a particle in a one
dimensional box.

Properties of Waves and Particles


De Broglie Wavelength. If the photon is treated as a wave, it is characterized by a wave-
length ( λ ) and a frequency ( ν ), which are related by the equation λν = c where c is the
velocity of light in a vacuum. The energy of such a photon is expressed as E = hν , where h
is Planck’s constant.

Einstein’s theory of special relativity shows a free particle has the energy
E = m 2 c 2 + p 2 c 2 . If the particle, for example the photon, has no mass,
m = 0 and E = pc. But the energy of a photon is also E = hν . By equating the two
energy expressions, we arrive at


p = ------ . Eq. 3-1
c

A quick rearrangement yields

h c
--- = --- = λ , Eq. 3-2
p ν

h
where λ is called the de Broglie wavelength. The equation --- = λ holds not only for pho-
p
tons, but for all matter and is usually referred to as the de Broglie hypothesis.

27
Waves, Particles and Boxes: Properties of Waves and Particles

De Broglie Hypothesis. Let us now apply the de Broglie equation to an atomic system. The velocity (and hence momentum)
distribution is described by the Boltzmann distribution since atoms constitute a thermal system. Thus, the r.m.s. momentum
h h
p = 2mE = 3mkT , where k is the Boltzmann constant, and λ = --- = ------------------- . If we insert the proper constants,
p 3mkT
6.7 ×10 2.6 ×10
–27 –7
λ = -------------------------------------------------------------------------- = --------------------cm , with mass in a.m.u.’s and T the absolute temperature in K.* A simple
3 ⋅ 1.6 ×10 ⋅ 1.4 ×10 ⋅ mT
– 24 – 16 mT

calculation at 300K yields λ electron = 65 ×10 cm, λH = 1.5 ×10 cm, λair = 0.27 ×10 cm, (m air = 29 ) .
–8 –8 –8

The calculations show clearly that the classical assumption that a particle can be treated as a point break down when the
particle’s size reaches atomic dimensions. When the linear dimension of the problem, for example, the confinement length for
2E′
the oscillator, x – x e max = --------- , is comparable to the wavelength of the particle, then quantum mechanics is required. The
k
σ
same is true for collisions where the impact parameter b c = --- is on the order of λ . Note that the wavelength of the particle
π
depends upon two quantities, its mass and its energy. When the wave associated with the particle is small compared to the linear
dimension of the problem at hand, then classical mechanics provide an adequate description.

Doppler Effect. The Doppler Effect is observed when a wave source moves towards or away from the receiver. In spectroscopic
experiments, it may be used to measure the velocity of the emitting or absorbing species along direction of photon propagation.
Its basis rests upon the principles of conservation of energy and momentum. Consider a photon of energy hν and momentum

------ striking a molecule of mass m and momentum p 1 . The absorption will cause an electron to be excited into a higher energy
c
level in agreement with the Bohr Frequency Condition: hν 0 = E 2 – E 1 , where E 1, E 2 are the respective internal energies of the
molecule before and after the photon impact, and ν 0 is the frequency of a photon absorbed if the molecule were stationary.
p 12 p 22
Thus, the initial energy of the system is hν 0 + E 1 + -------- and the final energy is E 2 + -------- . By conservation of energy
2m 2m

Initial Energy = Final Energy

p 12 p 22
hν + E 1 + -------- = E 2 + --------
2m 2m

*. Note that temperature has to be absolute to maintain a relationship with the average kinetic energy.

28
Waves, Particles and Boxes: Properties of Waves and Particles

p 22 – p 12
hν – ---------------- = E 2 – E 1
2m

p2 + p1
hν – ( p 2 – p 1 ) ---------------- = hν 0 Eq. 3-3
2m

The next useful relationship is derived via the conservation of momentum.

Initial Momentum = Final Momentum


------ + p 1 = p 2
c


------ = p 2 – p 1 Eq. 3-4
c

p2 + p 1
If we note that the average velocity of the molecule is given as v = ---------------- , then substituting into Eq. 3-3 yields
2m

v
hν – hν -- = hν 0
c

hν 1 – --  = hν 0 or hν = ----------------- hν 0 .
v 1
 c
1 – v-- 
 c

v2
For v « c , so that ----2 → 0 , we can simplify to
c

ν = ν 0 1 + --  .
v
Eq. 3-5
 c

The Doppler effect, observable as a frequency shift, can then be written ∆ν = ν – ν 0 = ν 0 1 + --  – ν 0 = ν 0 -- or succinctly
v v
 c c

∆ν v
------- = -- . Eq. 3-6
ν0 c

29
Waves, Particles and Boxes: The Schrödinger Equation

This derivation of the Doppler effect is also the “selection rule” for the center of mass translation in an optical tran-
sition. In other words, it describes the allowed shifts in the translational energy of particle after a collision with a photon.
(Optical transitions will be covered later in much detail.) The Doppler Effect uses the average velocity rather than the ini-
tial velocity, so it automatically has a symmetry between absorption and emission.

The Doppler effect is also responsible for broadening of spectral lines in emission (and also absorption) spectra. In
a thermal system, the distribution of velocities along a direction is given by
 d
– mv 2
N ( v )dv = C ⋅ exp  -------------  dv , where C is a constant. Rewriting the relationship
 2kT 
∆ν v ν – ν0
------- = -- for v , we arrive at v = c -------------- . The shape of the spectral line due to this
ν0 c ν0
distribution of velocities is then described by
 d
– mc 2 ∆ν 2 
Figure 3.1 - A schematic of intensities of a σ ( ν )dν = S ⋅ exp  ----------------------
- , S a constant
spectral line due to velocity distributions  2kTν 02 
and the Doppler effect. The maximum is
the ideal, stationary emission.
– mc 2 ( ν – ν 0 ) 2 
σ ( ν )dν = S ⋅ exp  ----------------------------------
- . Eq. 3-7
 2kTν 02 

The Schrödinger Equation


The Schrödinger Equation, due to the Austrian physicist Erwin Schrödinger, is the basis of all quantum mechanics.
It comes in two flavors: time dependent and time independent. The time dependent version is more complex, so we will
first familiarize ourselves with the time independent version and the problems it helps us solve. The key condition
behind time independence is that the energy of the system treated is conserved over time.

Because the Schrödinger Equation is strikingly similar to a standard wave equation, we call its solutions wavefunc-
tions and represent them with the Greek letter Ψ (psi). Depending on the problem, Ψ can be a function of any number
of variables. To keep things simple, we begin with Ψ ( x ) , a function of a single dimenstion. If our system is a particle of
mass m moving in a potential V ( x ) , where x is the position of this particle, and the total energy is E, then the
Schrödinger Equation is

– h– 2 ∂ 2 Ψ h
-------- ---------2- + V ( x )Ψ ( x ) = EΨ ( x ) (we use the abbreviation h– = ------ ). Eq. 3-8
2m ∂x 2π

30
Waves, Particles and Boxes: The Particle in a Box

To solve the equation, we need to find Ψ as a function of x. In terms of classical physical quantities, Ψ has no immedi-
ately intuitive meaning. Ψ 2 , however, is much easier to visualize, as it is the particle’s probability density with respect to
the given variable. In the case of a particle in one dimension, Ψ 2 ( x )dx is the probability that the particle’s position is
within the position differential xdx . Because the wavefunction is frequently a complex function, the probability density
is formally written using the complex conjugate, Ψ * ( x ) as

ρ ( x ) = Ψ ( x )Ψ * ( x ) . Eq. 3-9

The Particle in a Box


We will now solve our first quantum mechanics problem, the famous particle in a box. Consider the particle enclosed in
a one dimensional box of length L. Let’s define the potential V ( x ) = 0 when the particle is inside the box, i.e. 0 < x < L , and
V ( x ) = ∞ when the particle is outside of it, i.e. x ≥ L . Because the potential is infinite for the particle outside of the box, the
particle is always is inside the box, and that is the only part of the potential and space we need to consider here. The
Schrödinger Equation (Eq. 3-8) simplifies, therefore, to

– h– 2 ∂ 2 Ψ ( x )
- = EΨ ( x ) .
-------- ----------------- Eq. 3-10
2m ∂x 2

The conditions on the wavefunction are that it be continuous. Also, because we are dealing with probabilities, the function has
to be properly normalized. In other words, the probabilities of finding the function at any given point have to add up to unity,

or ∫ ρ ( x ) dx
–∞
= 1 . Because the particle is always inside the box, ρ ( x ) = Ψ 2 = 0 everywhere outside of the box. Thus,

formally Ψ ( x ) = 0 outside the box. The interesting part, however, is for 0 > x > L . We proceed to solve the differential
equation.

Since the problem is one dimensional, we can convert the partial derivatives to ordinary ones. The equation we want to
solve is as shown in Eq. 3-10. A simple rearrangement and a substitution of

2mE
k 2 = -----------
- yield Eq. 3-11
h– 2

d2Ψ(x)
- = –k2 Ψ ( x ) .
------------------ Eq. 3-12
dx 2

31
Waves, Particles and Boxes: The Particle in a Box

The only functions that will satisfy this condition are the sine and cosine functions. Thus, the solution must be of the form

Ψ ( x ) = A cos ( kx ) + B sin ( kx ) .* Eq. 3-13

To find the values of the constants, we apply the boundary conditions. Because the wavefunction is equal to 0 outside the box,
Ψ ( 0 ) = Ψ ( L ) = 0 as well. This implies that A = 0 , since cos ( 0 ) = 1 while sin ( 0 ) = 0 . The condition that Ψ ( L ) = 0

implies that sin ( kL ) = 0 . In order for this to be true, k must equal ------- . To summarize, the wavefunction must be of the form
L


Ψ n ( x ) = B sin ------- x , n = 1, 2, 3… . Eq. 3-14
L
Note that the solution has an explicit dependance on the number n. The number arises because the boundary conditions
can be satisfied by any number of sine waves, as long as an integral number of wavelengths fits in the box. Thus, n can take on
positive integer values. An often convenient pseudo-classical way of visualizing this phenomenon is with standing waves and
their overtones. Why is 0 not an allowed value for n?

h– 2 k 2
If we rearrange Eq. 3-11, we get an expression for the energy of the particle, E = ----------- . Substitution for k yields a
2m
very important and practical result.

h2n2
E n = -------------2 , n = 1, 2, 3… . Eq. 3-15
8mL
In other words, if we know the length of the box, the mass of the particle, and the value of n, we can calculate the energy
of the system. Notice that as n increases, so does the energy. It is a result that should be familiar from classical wave physics,
where a higher overtone (more nodes) corresponds to a higher energy. In quantum language, a number such as n that corre-
sponds to a wavefunction Ψ n ( x ) is called a quantum number. The wavefunction is often referred to as an eigenfunction and the
corresponding energy E n is called an eigenvalue. An important fact is that the lowest possible energy is not zero. Because the

d
*. For the interested reader, we present a detailed solution of the Schrödinger equation. First, define the operator D̂ ≡ ------ . Eq. 3-12
dx
can be then rewritten as ( D̂ + ik ) ( D̂ – ik )Ψ = 0 . The solution is of the form D̂Ψ = ± ikΨ . Solving for Ψ gives
Ψ = c 1 e ikx + c 2 e –ikx . Using Euler’s formula, we can expand Ψ = c 1 cos kx + c 1 i sin kx + c 2 cos kx – c 2 i sin kx . Combining like
terms yields Ψ = ( c 1 + c 2 ) cos kx + ( ic 1 – ic 2 ) sin kx . Substituting c 1 + c 2 = A and ic 1 – ic 2 = B yields equation Eq. 3-13. For
more details on solving differential equations by factoring operators, refer to Kreyszig, Advanced Engineering Mathematics, 7th.
Ed., p. 79.

32
Waves, Particles and Boxes: The Particle in a Box

h2
lowest allowed value of n is 1, the lowest energy E 1 = -------------2 > 0 . This is an example of the so-called zero point energy,
8mL
a purely quantum mechanical result. In fact, it turns out that in almost all systems there is always some kinetic energy
present. Furthermore, notice that the energy is quantized. In other words, there are only discrete energy levels avaliable
to the system. This is in direct contrast with the classical approach, where energy is a continuous function. The quantiza-
tion of energy is one of the most fundamental concepts in quantum mechanics.
The final task left is to determine the value of the constant B. To do it, we invoke the fact that the square of the wavefunc-
tion is a probability density that must be properly normalized.

L
2
B sin nπ  dx = 1
∫0  ------
L 
- x


B2 L
---------- ∫ ( 1 – cos 2y ) dy = 1
2nπ 0

(we use the identity 2 sin2 t = 1 – cos 2t )


B2 L
---------- y – --- sin 2y
1
=1
2nπ  2 
0

B2L
---------- ( nπ ) = 1
2nπ

2
B = ± --- .*
L

Thus, the normalized wavefunctions are

*. Usually, the positive root is chosen. However, recall that sin ( – x ) = – sin ( x ) . If we were to choose the negative root, the wave-
function’s phase would be shifted by half a wavelength, but the function would remain the same. Thus, technically, both values are
allowed.

33
Waves, Particles and Boxes: Quantum-Based Quantities


 2 nπ
--- sin ------- x, n = 1, 2, 3… for 0 < x < L
 L L
Ψn ( x ) =  . Eq. 3-16

 0 for x ≤ 0 or x ≥ L

The particle in a box is used for a large array of physical problems that model the containment of a small particle (for example
an electron). A thorough understanding of the solution presented above is crucial in further developments.

Quantum-Based Quantities
In a classical system, the quantities p ( t ) , q ( t ) or any function of p and q may be calculated with absolute precision. Any
and all functions of the dynamical variables can be defined and determined. The quantum approach reveals that in reality this
classical precision has a limit. The best we can do is determine average or expectation values. For example, for the particle in the
box, all we know is that the particle is somewhere inside the box, but we do not have a function available for x ( t ) . The next best
thing we can do is take an average of x over all x. Mathematically, the function’s probability density at any given interval times
the interval’s position gives the probability for that interval. To perform an average, we just need to sum over all the intervals. If

these are infinitely small, the sum becomes an integral. In symbols, x = ∫ ρ ( x )x dx . Recalling Eq. 3-9, we can write
–∞

x = ∫ Ψ ( x )Ψ * ( x )x dx , or, for reasons that will shortly become apparent,


–∞

x = ∫ Ψ * ( x )xΨ ( x ) dx .
–∞
Eq. 3-17

The equation is quite general. Thus, to compute the expectation value of, say, x 2 , we would evaluate

x2 = ∫ Ψ * ( x )x 2 Ψ ( x ) dx . To compute the expectation value of any function


–∞
f ( x ) , the expression would be

f(x) = ∫ Ψ * ( x )f ( x )Ψ ( x ) dx . The Dirac bracket notation is another way of writing f ( x )


–∞
and is frequently used in litera-

ture and some textbooks. The expectation value is written as 〈 f ( x )〉 . Regardless of notation, the integration assumes that
the probability density is properly normalized. If the wavefunctions are not normalized, the expectation value becomes

34
Waves, Particles and Boxes: Quantum-Based Quantities

∫∞ Ψ * ( x )f ( x )Ψ ( x ) dx 〈Ψ ( x )|f ( x ) |Ψ ( x )〉 *
f ( x ) = –-------------------------------------------------
- = ------------------------------------------- . Eq. 3-18
∞ 〈 Ψ ( x )| Ψ ( x ) 〉
∫ Ψ ( x )Ψ * ( x ) dx
–∞

Energy Averages. We can apply this same process to calculate the average kinetic ( T ), potential ( V ), and total ( E )
energies of a particle in the box. We already know the result, because in our treatment so far the potential is always zero.
Thus, the average kinetic energy will be the average total energy, which is a constant. Nonetheless, we follow the process
symbolically to arrive at a very important conclusion about the nature of the Hamiltonian operator in the Schrödinger
equation. We begin with a general Schrödinger equation and rearrange it slightly.

h– 2 ∂ 2 Ψ
– -------- ---------2- + V ( x )Ψ = EΨ
2m ∂x

h– 2 ∂ 2 Ψ
Ψ * – -------- ---------2-  + Ψ * V ( x )Ψ = Ψ * EΨ
 2m ∂x 

Integration produces a form very similar to Eq. 3-18.


∞ ∞ ∞
h– 2 ∂ 2 Ψ
∫ Ψ * – -------
–∞
- ----------  dx + ∫ Ψ * V ( x )Ψ dx
2m ∂x 2  –∞
= ∫ Ψ * EΨ dx
–∞
Eq. 3-19

If we invoke the fact that the total energy E is a constant independent of position, we can write
∞ ∞
E = ∫ Ψ * EΨ dx
–∞
= E ∫ Ψ * Ψ dx = E .
–∞

Eq. 3-19 becomes


∞ ∞
h– 2 ∂ 2 Ψ
*  – -------- ---------- 
∫  2m ∂x 2 
–∞
Ψ d x + ∫ Ψ * V ( x )Ψ dx = E ,
–∞

which confirms the intuitive result that

T+V = E. Eq. 3-20

*. The Dirac bracket notation implies that the first Ψ is the complex conjugate function, so there is no need to write Ψ * explicitly.

35
Waves, Particles and Boxes: Quantum-Based Quantities

p2
T turns out to be a very interesting quantity. Recall that T = -------- . Thus, we could write
2m


p2Ψ
T = ∫ Ψ * ---------
–∞ 2m
- dx

but from the treatment of the Schrödinger equation above, we know that


h– 2 ∂ 2 Ψ
*  – -------- ---------- 
T = ∫  2m ∂x 2  dx .
–∞
Ψ

A very important result follows, namely that in quantum mechanics momentum is no longer a classical function, but a differen-
tial operator. In symbols,

∂2 h– ∂
p 2 = – h– 2 --------2 , or pˆx = --- ------ .* Eq. 3-21
∂x i ∂x

Momentum is written with a “hat” over it to indicate an operator, not a function. Also, note that it requires an imaginary term.
Because the momentum operator is an integral part of the Hamiltonian expression that we have been using, there is a corre-
sponding Hamiltonian operator:

h– 2 ∂ 2
Ĥ = – -------- --------2 + V ( x ) . Eq. 3-22
2m ∂x

Thus, in short-hand operator notation, the Schrödinger equation can be written most succinctly as

ĤΨ = EΨ Eq. 3-23

This explains why the integrand in Eq. 3-17 was written as Ψ * ( x )xΨ ( x ) . If instead of x, we wanted to average momen-
tum, for example, the function would actually be a differential operator with an imaginary term, and the result would
depend on the order of the terms of the integrand.

*. The origin of the linear momentum operator can be seen in the mathematical solution of the particle in a box. See footnote on
page 32.

36
Waves, Particles and Boxes: Quantum-Based Quantities

Notice the analogy of the quantum Hamiltonian operator to the classical Hamiltonian function H ( p, q ) . Both are used to
describe the energy of the system, but the quantum operator involves wavefunctions instead of continuous variables, and the
energy eigenvalues are quantized.

Dispersion and Uncertainty. Statisticians define a quantity called the dispersion as a useful indication of the uncertainty in a
given measurement or quantity. For the position variable in the particle in the box, dispersion would be defined as
( ∆x ) 2 = ( x – x ) 2 , which can be rewritten as

( ∆x ) 2 = ( x – x ) 2 = x 2 – x 2 ≥ 0 . Eq. 3-24

The dispersion will necessarily be greater than or equal to zero, because it is an average of squares. The quantity that is most
immediately useful to us is the root-mean-square deviation, defined as

∆ * x ≡ [ ( ∆x ) 2 ] ( ∆ * x is often abbreviated as ∆x ).
1/2
Eq. 3-25

It is also a measure of uncertainty. A simple substitution yields

∆x = x2 – x2 . Eq. 3-26

An example should make this more transparent. Let us calculate the rms deviation for a particle in a box centered around
L
--- . Recalling the normalized wavefunction from Eq. 3-16, we determine x with Eq. 3-17. Eq. 3-16 tells us that the wave-
2
function is null outside of the box, so we can adjust our integration limits accordingly.

L
2 nπx 2 nπx
x = ∫0 --- sin ---------- ⋅ x ⋅ --- sin ---------- dx
L L L L

L
2 nπx
x = --- ∫ x ⋅ sin2 ---------- dx Eq. 3-27
L0 L

L
2 nπx L L L cos 2πn L sin 2πn
x = --- ∫ x ⋅ sin2 ---------- dx = --- + -------------- - – ---------------------- .
– ---------------------- Eq. 3-28
L0 L 2 4π 2 n 2 4π 2 n 2 2πn

Naturally, since n can only take on positive integer values, the sine term drops out, and the cosine term cancels the preceding
term. Thus, the final result is

37
Waves, Particles and Boxes: Quantum-Based Quantities

L
x = --- . Eq. 3-29
2

Note, however, that the brute force approach of Eq. 3-28 is completely unnecessary. In fact, we don’t even need to inte-
grate anything! If the particle is moving back and forth in the box, then it’s average position will be exactly halfway
through the box, in exact agreement with Eq. 3-29.

To calculate x 2 , we use an analogue of Eq. 3-27 with the proper x limits.

L L

x 2 = --- ∫ x 2 sin2 ---------- dx = --- ∫ x 2 1 – cos2 ----------  dx


2 nπx 1 nπx
L0 L L0  L 

L2 L2
x 2 = ---- – -------------- Eq. 3-30
3 2π 2 n 2

Substitution of Eq. 3-29 and Eq. 3-30 into Eq. 3-26 yields the final answer.

 L---- – --------------
L2  L2
2 1 1
∆x = – ---- = L ------ – -------------- Eq. 3-31
 3 2π 2 n 2  4 12 2π 2 n 2

The expression has an explicit dependence on the quantum number n, and therefore the energy of the system. The higher
energy, the higher the rms deviation, or position uncertainty. For n = 1 , ∆x = 0.1807L , for example. Note that the uncer-
1
tainty does not increase indefinitely, and in the asymptotic case of n → ∞ , ∆x = L ------ .
12

Let us now repeat the same process for the uncertainty in momentum. The integrations will seem slightly more compli-
cated, because momentum is a differential operator, as defined in Eq. 3-21, but the process is exactly analogous. First, we
find an expression for p .

L
2 nπx h– d
p = --- ∫  sin ----------   --- ------ sin ----------  dx
nπx
Eq. 3-32
L0  L   i dx L 

L
2h– nπx   nπ nπx 
p = ------- ∫  sin ---------- - cos ----------
L   ------ L  dx
Li 0  L

38
Waves, Particles and Boxes: Quantum-Based Quantities

L
2πh– n  nπx   nπx 
p = -------------
L i 0
2 - ∫ sin ----------
L 
cos ----------
L  dx

– L cos  ---------- 
2
nπx
2πh– n  L  L L cos2 ( nπ )
- ---------------------------------
p = ------------- = ---------- – -------------------------- = 0 Eq. 3-33
L2i 2πn 0
2πn 2πn

The last step is true because n can only take on integral values, which gives 1 for cos2 ( nπ ) , thus cancelling out the first
term.

The same method could be employed to find p 2 , but we will resort to a clever short cut. Recall Eq. 3-20, where we
p2
showed that T + V = E . T is simply equivalent to -------- , and since the potential in our problem is zero inside the box, a simple
2m
rearrangement yields

p 2 = 2mE Eq. 3-34

If we now take advantage of the energy level equation for the particle in the box (Eq. 3-15), we arrive at

h2n2 h2n2
p 2 = 2m -------------2 = ----------2- . Eq. 3-35
8mL 4L

Combining and Eq. 3-35, we find the uncertainty in momentum to be

hn
∆p = p 2 – p 2 = ------- . Eq. 3-36
2L

Again, notice the explicit dependance of the uncertainty on the quantum number n. As with position, the higher the energy,
the higher the uncertainty, but there is no upper asymptotic boundary.

39
Waves, Particles and Boxes: A Symmetric Box

A Symmetric Box
The coordinate system of values of x from 0 to L we imposed on the box was completely arbitrary. It is natural to expect
quantities such as energy to be independent of the coordinates. Because in more complicated problems (such as atoms) it is
L L
much easier to center the box at the origin, we now re-solve the problem for a symmetric box extending from – --- to --- , cen-
2 2
tered at x = 0 .

L
The easiest way of going about this is to let x in Eq. 3-16 equal x′ – --- , which corresponds to the coordinate shift to
2
L
the right by --- . The wavefunction
2


 2 nπ
--- sin ------- x′, n = 1, 2, 3… for 0 < x′ < L
 L L
Ψ n ( x′ ) = 

 0 for x′ ≤ 0 or x′ ≥ L

thus becomes

 2 nπ 
--- sin ------- x + --- , n = 1, 2, 3… for – --- < x < ---
 L L L
 L L  2 2 2 L
Ψn ( x ) =  since x′ = x + --- . Eq. 3-37
 L 2
L
 0 for x ≤ – --- or x ≥ ---
 2 2

If we expand the sine of a sum in the non-zero part of the wavefunction, we see that in the non-zero range

2  nπ
--- sin ------- x ⋅ cos ------- + cos ------- x ⋅ sin -------  .
nπ nπ nπ
Ψn( x ) = Eq. 3-38
L L 2 L 2 

Recall that the cosine is an even function, while the sine is an odd function. This fact greatly simplifes Eq. 3-38.

40
Waves, Particles and Boxes: A Symmetric Box


 2 nπ L L
 --- cos ------- x, n = an odd positive integer, for – --- < x < ---
 L L 2 2

Ψn ( x ) =  2 nπ L L
--- sin ------- x, n = an even positive integer, for – --- < x < --- Eq. 3-39
 L L 2 2

 0 L L
for x ≤ – --- or x ≥ ---
 2 2

We originally shifted the box’s coordinates under the asumption that the particle is not affected. Let us now test this
L
assumption using the wavefuntion in Eq. 3-39. When the box spanned 0 < x < L , x = --- (Eq. 3-29), which is the middle
2
of the box. Using the new wavefunction, we can write

L
---
2

x = ∫L ( Ψ n* ( x ) ⋅ x ⋅ Ψ n ( x ) ) dx = 0, for all positive integers n , Eq. 3-40


– ---
2

which is also in the exact middle of the box. In performing the integration, we have resorted to a very convenient trick: if n is
odd, then the wavefunction will be an even cosine function. A product of two even cosines with the odd x is an odd function.
The integral of an odd function about the origin is always zero. If n is even, the wavefunction is an odd sine function; but that
does not matter, because there are two of them in the integrand. The product of two odd sines and an odd x is odd overall, and
an odd integrand yields zero. Remember, though, that for the trick to work, the integration must be centered about the origin.
This calculation-free method of evaluating integrals is a great tool in quantum mechanical calculations, so let us apply it again.
To calculate the average linear momentum, write

L
---
2

Ψ * ( x ) ⋅ h- ------ ⋅ Ψ n ( x ) dx = 0, for all positive integers n .
d
p = ∫L  n --
i dx 
Eq. 3-41
– ---
2

This agrees exactly with the value laboriously derived in Eq. 3-33. Why is it true? Because for odd n, the wavefunction is
even, but its derivative with respect to x is odd, and for even n, the wavefunction is odd, but its derivative with respect
to x is even. Thus, regardless of the value of n, the integrand is odd, and the expression equals zero.

41
Waves, Particles and Boxes: A Symmetric Box

General Wavefunction Properties


The symmertic box does not affect the average momentum, and since the potential is zero inside the box the average
energy is not affected either. Thus, Eq. 3-15 still holds. Energy is a real number given by

h2n2
E n = -------------2 , n = 1, 2, 3… . Eq. 3-42
8mL

A less obvious result is that solutions with different energies are orthogonal to each other. Let us take two such
solutions Ψ n ( x ) and Ψ m ( x ) with respective energies E n and E m , where n ≠ m . The two solutions are orthogonal if
L
---
2

∫ Ψ n ( x ) ⋅ Ψ m ( x ) dx
L
= 0 . To prove the orthogonality, first take the case where n – m is odd. This implies that the inte-
– ---
2

grand Ψ n ( x ) ⋅ Ψ m ( x ) is odd, which makes the integral zero. If, on the other hand, n – m is even, then the two functions
are either both odd or both even. If they are both even, we can write

L
---
2
2 
--- ∫ cos ------- x ⋅ cos -------- x dx
nπ mπ
L L L L 
– ---
2

L
---
2
2 1 ( n + m )π ( n – m )π
= --- ∫ --- cos -----------------------x + cos ----------------------- x dx
L L2 L L
– ---
2

π π
( m – n ) sin --- ( n + m ) + ( n + m ) sin --- ( n – m )
2 2 2
= --- ------------------------------------------------------------------------------------------------------------------------ = 0, for n ≠ m, n, modd .
π (n – m)(n + m)

This proves the orthogonality for the case of two even wavefunctions (n and m are odd). The proof is similar for the case of two
odd wavefunctions.

L
---
2
2  nπ
--- ∫ sin ------- x ⋅ sin ------- x dx

L L L L 
– ---
2

42
Waves, Particles and Boxes: A Box with Finite Walls

L
---
2
2 1 ( n + m )π ( n – m )π
= --- ∫ --- cos -----------------------x – cos ----------------------- x dx
L L2 L L
– ---
2

π π
( m – n ) sin --- ( n + m ) + ( n + m ) sin --- ( n – m )
2 2 2
= --- ------------------------------------------------------------------------------------------------------------------------ = 0, for n ≠ m, n, m even .
π (n – m)(n + m)

A Box with Finite Walls


So far, we have constrained our particle in a box with infinite potenial walls. While this was convenient for calculations, it
is not a complete reflection of reality, where no potential is truly infinite. Let us now consider a system where

V ( x ) = 0 for 0 < x < L (inside the box)


V ( x ) = V 0 for all other x (outside the box)

and the particle’s energy does not exceed the V 0 wall height. Note that we have again moved the center of the box away from
the origin.

– h– 2 ∂ 2 Ψ
The Schrödinger equation -------- ---------2- + V ( x )Ψ ( x ) = EΨ ( x ) takes on two forms, correspondingly.
2m ∂x

– h– 2 ∂ 2 Ψ
-------- ---------2- = EΨ ( x ) (inside the box), E > 0 Eq. 3-43
2m ∂x

– h– 2 ∂ 2 Ψ
-------- ---------2- = ( E – V 0 )Ψ ( x ) (outside the box), ( E – V 0 ) < 0 Eq. 3-44
2m ∂x

The solution to Eq. 3-43 is similar to that derived in Eq. 3-14, although the boundary conditions no longer require
that the wavefunction be zero at the edges of the box. The general wavefunction inside the box is of the form

2mE
Ψ n ( x ) = C sin ( kx + δ n ) where k = -.
----------- Eq. 3-45
h– 2

43
Waves, Particles and Boxes: A Box with Finite Walls

The general form of the wavefunction in the regions outside of the box is readily shown to be

2m ( V 0 – E )
Ψ n ( x ) = Ae κx + Be –κx where κ = -.
--------------------------- Eq. 3-46
h– 2

As the distance from the box increases, we expect the wavefunction to decrease to zero. This is also reasonable, because the over-
all wavefuntion has to be properly normalized. Thus, in the region x < 0 , to prevent Ψ n ( x ) from increasing infinitely, B must
be set to zero. Similarily, in the region x > L , A must be set to zero. More concisely

 Ae κx for x < 0

Ψ n ( x ) =  C sin ( kx + δ n ) for 0 < x < L Eq. 3-47

 Be for x > L
– κx

The overall wavefunction is therefore a composite of the three regions. The requirements that it be a smooth, continuous and
differentiable function give rise to the boundary conditions. Thus, at x = 0 continuity requires that C sin ( δ n ) = A , and dif-
ferentiability requires that Ck cos ( δ n ) = Aκ . Correspondinlgy, at x = L , C sin ( kL + δ n ) = Be –κL and
Ck cos ( kL + δ n ) = – Bκe κL .

We expect that quantization of energy would apply to this system exactly as it did to the particle in the box with infinite
walls. This is indeed true, but the result is slightly harder to see. We can use the first two boundary conditions to write

k E 2mE 2m ( V 0 – E )
tan ( δ n ) = --- = --------------- (since k = - and κ =
----------- - ).
--------------------------- Eq. 3-48
κ Vo – E h– 2 h– 2

Analogously, the latter two boundary conditions yield

tan  -----------
- L + δ n = – --- = – --------------- .
2mE k E
Eq. 3-49
 h– 2  κ V0 – E

If we combine Eq. 3-48 and Eq. 3-49, we obtain

tan ( δ n ) = – tan  -----------


- L + δ n .
2mE
Eq. 3-50
 h– 2 

44
Waves, Particles and Boxes: A Chemical Box

For a given δ n , there are only specific energy values that satisfy Eq. 3-50 (governed by the periodicity of the tangent
function). The exact solutions can be computed numerically, but those are not as important as the fact that energy levels
remain quantized even in a finite potential well.

Qualitatively, let us note that the energy of the particle will decrease from the value obtained in the box of identical
length, but with infinite walls, because the effective wavelength is somewhat greater since Ψ ( 0 ) and Ψ ( L ) ≠ 0 . In the region
where E > V ( x ) , solutions are oscillatory, with the number of nodes increasing with energy (analogous to box with infinite
walls). In the regions where V ( x ) > E , the solutions decay exponentially.

It is useful to assess the extent of the exponential decay tail of the wavefunction. Consider a particle in a box with walls
V 0 and energy E in the region outside the box where E < V 0 . The exponential decay function is constrained by the parameter
2m ( V 0 – E )
κ = - . For a thermal particle, energy is on the order of E = 4 × 10 –14 erg . If we let V 0 – E = 5 × 10 –13 erg ,
---------------------------
h– 2
κ = 1.2 × 10 9 M cm –1 , where M is the molar mass of the particle (in grams). The function decays to 1% of the value at
the box wall where κx = – 4.6 (since e –4.6 = 0.01 ). For a H atom, this takes place at x = – 0.4Å ; for an electron, at
x = – 16Å .

A Chemical Box
A good physical example of a particle in a box is an electron in a conjugated polyene π system. For instance, take the
chain pictured below:
H H H H H
X C C C C C C C C C C X
H H H H H

h2n2
If the center of the chain is set at the origin, the energy levels are exactly as solved previously, E n = -------------2 , and the wavefunc-
8mL
nπx nπx
tions are Ψ n ( x ) = cos ---------- for odd n and Ψ n ( x ) = sin ---------- for even n. L is just the length of the chain from one termi-
L L
nal carbon to the other. We can think of the energy levels and the wavefunctions as molecular orbitals, which contain

45
Waves, Particles and Boxes: A Chemical Box

electrons. Keeping in mind that electrons have spin (so that each orbital can contain two electrons), for N electrons (N
N
even), the first --- orbitals will be filled. The highest occupied molecular orbital (HOMO) will have the energy
2
N 2
h 2  --- 
2
E HOMO = ---------------- . The corresponding lowest unoccupied molecular orbital (LUMO) will have the energy
8mL 2
2
h 2  --- + 1
N
2 
E LUMO - . Thus, we can calculate the energy involved in the lowest energy electronic transition. This would be
= -------------------------
2
8mL
h2( N + 1)
- . For the linear polyene pictured above, each carbon atom forms three σ type bonds, leaving
E LUMO – E HOMO = -----------------------
8mL 2
one π type electron, which is mobile. These occupy the molecular orbitals. The length of the box L ≈ NL C-C , where L C-C
is the average length of a carbon-carbon bond in the chain. The lowest energy electronic transition thus becomes
h2( N + 1 ) h2
E LUMO – E HOMO = ----------------------------2 ≈ ---------------------
2 .
8m ( NL C-C ) 8mNL C-C

h2
Let us insert actual values. On average, L C-C ≈ 1.4 × 10 –8 cm , m = m e = 9.11 × 10 –28 g . ----------------
2
- = 19.7eV .
8mL C-C

19.7
E LUMO – E HOMO = ---------- eV Eq. 3-51
N

hc
For a system of ten electrons, N = 10 and ∆E ≈ 2eV . If we use the relationship E = ------ , we can quickly compute that this
λ
corresponds to a wavelength λ = 620nm , or red light.

46
Waves, Particles and Boxes: Heisenberg Indeterminacy Principle

Heisenberg Indeterminacy Principle


You may have been bothered by the seemingly crude and careless rounding in solving the particle in the box. It turns out,
however, that such rounding is not completely unjustified. Take a look at Eq. 3-31 and Eq. 3-36. Notice that ∆x is propor-
1
tional to L and ∆p is proportional to --- . If we take the product of the two, we can eliminate the dependence on the box
L
length. The result is known as the Heisenberg indeterminacy principle, and for our particle in the box can be expressed
as

h– 2 ( n 2 π 2 – 6 )
∆p∆x = ------------------------------- . Eq. 3-52
12

The principle states that the linear momentum and the position of any object can simultaneously be known only to a limited
extent. The more accurate the measurement of position, the more uncertainty there is in the momentum, and vice versa.

The indeterminacy for the first quantum state ( n = 1 ) can be thus easily seen to be 0.568h– , or in its familiar form

h–
∆p∆x ≈ --- . Eq. 3-53
2

The Heisenberg principle applies for all quantum mechanical systems in its most general form

∆p∆x > 0 . Eq. 3-54

47
Waves, Particles and Boxes: Heisenberg Indeterminacy Principle

48
LECTURE 4 More on the Schrödinger
Equation

The Schrödinger equation and its solutions possess a number of interesting and useful
properties. We first examine the most important ones in one dimension, and then examine
how they can be applied in two and three dimensional spaces.

Solution Properties
Recall the Schrödinger equation for a one dimensional system.

– h– 2 d 2 Ψ
-------- ---------2- + V ( x )Ψ ( x ) = EΨ ( x ) Eq. 4-1
2m dx

To systematically examine the nature of its solutions, we consider two functions f ( x )


and g ( x ) . For these, it is very useful to define the function (called the Wronskian),

df dg
W ( x ) = ------g – ------f . Eq. 4-2
dx dx

The derivative of this new function is then given by

dW d2f df dg d 2 g dg df d2f d2g


--------- = ---------2 g + ------ ------ – --------f – ------ ------ = ---------2 g – --------f . Eq. 4-3
dx dx dx dx dx dx dx dx dx

Finally, we rewrite the Schrödinger equation as

d2Ψ
– ---------2- + V′Ψ = E′Ψ , Eq. 4-4
dx

2m 2m
- V ( x ) and E′ = -------
where we define V′ ( x ) = ------- - E (the primes are not derivatives).
h– 2 h– 2

Theorem 1. The eigenvalues E (or equivalently E′ ) are real numbers.

49
More on the Schrödinger Equation: Solution Properties

To prove this, we invoke the equation for the complex conjugate of Ψ ( x ) . If Ψ ( x ) = u ( x ) + iv ( x ) is the wavefunction
(with u ( x ) and v ( x ) both real functions), then Ψ * ( x ) = u ( x ) – iv ( x ) is its complex conjugate ( i = – 1 ). If Eq. 4-4 gives
a solution, then

d2Ψ*
- + V′Ψ * = E′ * Ψ *
– ----------- Eq. 4-5
dx 2

should produce that solution’s complex conjugate (if there indeed were an imaginary part to the eigenvalue). The potential
V ( x ) is assumed to be real for all physical problems.

Multiplying Eq. 4-4 by Ψ * and Eq. 4-5 by Ψ yields

d2Ψ
– Ψ * ---------2- + Ψ * V′Ψ = Ψ * E′Ψ Eq. 4-6
dx

d2Ψ*
– Ψ -----------
- + ΨV′Ψ * = ΨE′ * Ψ * . Eq. 4-7
dx 2

Subtracting these two equations gives

d2Ψ d2Ψ
– Ψ * ---------2- – Ψ ---------2-  = ( E′ – E′ * )ΨΨ * Eq. 4-8
 dx dx 

Note that Eq. 4-8 is analogous to Eq. 4-3. If we define f = Ψ and g = Ψ * , then

dW d2Ψ d2Ψ
--------- = Ψ * ---------2- – Ψ ---------2- = – ( E′ – E′ * )ΨΨ * . Eq. 4-9
dx dx dx

We invoke Eq. 4-2 to write

dΨ dΨ *
W = -------- Ψ * – ---------- Ψ = – ( E′ – E′ * ) ∫ ΨΨ * dx . Eq. 4-10
dx dx

We then evaluate the integral over all values of x to get

∞ ∞
 dΨ dΨ *
-------- Ψ * – ---------- Ψ = – ( E′ – E′ * ) ∫ ΨΨ * dx . Eq. 4-11
 dx dx  –∞
–∞

50
More on the Schrödinger Equation: Solution Properties


Proper normalization conditions tell us that Ψ ( ± ∞ ) = 0 , Ψ * ( ± ∞ ) = 0 , and ∫ ΨΨ * dx
–∞
= 1 . Eq. 4-11 thus simplifies to

0 = – ( E′ – E′ * ) or

E′ = E′ * . Eq. 4-12

This implies that the energy eigenvalues E must be real, since they are equal to their complex conjugates.

Theorem 2. The solutions to the Schrödinger equation are orthogonal.

While on page 42 we showed that the solutions to the one dimensional particle in the box problem are orthogonal,
here we prove the general case. Take any two unique solutions that satisfy the Schrödinger equation.

–d2 Ψ1
--------------- + V′Ψ 1 = E 1 ′Ψ 1 Eq. 4-13
dx 2

–d2 Ψ2
--------------- + V′Ψ 2 = E 2 ′Ψ 2 Eq. 4-14
dx 2

Eq. 4-14 has the complex conjugate

– d 2 Ψ 2*
- + V′Ψ 2* = E 2 ′ * Ψ 2* .
--------------- Eq. 4-15
dx 2

Multiplying Eq. 4-13 and Eq. 4-15 by the other’s eigenfunction yields

d2Ψ1
- + Ψ 2* V′Ψ 1 = Ψ 2* E 1 ′Ψ 1
– Ψ 2* ----------- Eq. 4-16
dx 2

d 2 Ψ 2*
– Ψ 1 ------------ + Ψ 1 V′Ψ 2* = Ψ 1 E 2 ′ * Ψ 2* . Eq. 4-17
dx 2

We subtract Eq. 4-16 and Eq. 4-17 to get

d2Ψ1 d 2 Ψ 2* 
– Ψ 2* -----------2- – Ψ 1 ------------ = ( E 1 ′ – E 2 ′ )Ψ 2* Ψ 1 . Eq. 4-18
 dx dx 2 

51
More on the Schrödinger Equation: Solution Properties

This is again analogous to Eq. 4-3. If we let f = Ψ 1 and g = Ψ 2* , we can evaluate the definite integral



 dΨ dΨ 2*
---------- Ψ 2* – ----------Ψ 1 = – ( E 1 ′ – E 2 ′ ) ∫ Ψ 1 Ψ 2* dx .
1
Eq. 4-19
 dx dx  –∞
–∞

We have shown previously that the left side of this equation must equal zero. Since we chose unique wavefunctions with differ-

ent eigenvalues, ( E 1 ′ ≠ E 2 ′ ) the only way Eq. 4-19 will be satisfied is that ∫ Ψ1 Ψ 2* dx
–∞
equals zero, or in other words, if the
two wavefunctions are orthogonal.

Theorem 3. Higher energy wavefunctions have more nodes.

If E 1 and E 2 are two eigenvalues of the system such that E 2 > E 1 , then Ψ 2 has more nodes than Ψ 1 . For the purposes of
the proof, we need to express this symbolically. Let Ψ 1 > 0 over the range a < x < b , and let Ψ 1 ( a ) = Ψ 1 ( b ) = 0 . Then
Ψ 2 ( ξ ) = 0 for some value ξ , a ≤ ξ ≤ b . In other words, Ψ 2 will always have a node between two adjacent nodes of Ψ 1 .

To prove this theorem, we use proof by contradiction and again resort to Eq. 4-2 and Eq. 4-3. Defining f = Ψ 1 and
g = Ψ 2 , we write

d 2 Ψ1 d2Ψ2
--------- = ( E 2 – E 1 )Ψ 1 Ψ 2 =  ------------Ψ 2 – ------------Ψ 1 ,
dW
Eq. 4-20
dx  dx dx 

dΨ 1 ( a ) dΨ 2 ( a ) dΨ 1 ( a )
W ( a ) = ------------------- Ψ 2 ( a ) – ------------------- Ψ 1 ( a ) = ------------------- Ψ 2 ( a ) and Eq. 4-21
dx dx dx

dΨ 1 ( b ) dΨ 2 ( b ) dΨ 1 ( b )
W ( b ) = ------------------Ψ 2 ( b ) – ------------------Ψ 1 ( b ) = ------------------Ψ 2 ( b ) . Eq. 4-22
dx dx dx

dΨ 1 ( a ) dΨ 1 ( b )
This implies that ------------------- > 0 and ------------------ < 0 , since Ψ 1 is positive and continuous. Now assume that Ψ 2 ( x ) > 0 in the
dx dx
dW
range a ≥ x ≥ b , i.e. that Ψ 2 has no nodes in the given range. This would make --------- = ( E 2 – E 1 )Ψ 1 Ψ 2 > 0 for a ≥ x ≥ b ,
dx
dΨ 1 ( a ) dΨ 1 ( b ) dΨ 1 ( a ) dΨ 1 ( b )
since E 2 > E 1 . But we know that W ( a ) = ------------------- Ψ 2 ( a ) > 0 and ------------------Ψ 2 ( b ) < 0 since ------------------- > 0 and ------------------ < 0 .
dx dx dx dx

52
More on the Schrödinger Equation: The Schrödinger Equation in Two Dimensions

This is the contradiction. The assumption that Ψ 2 does not change sign in the interval must be false; Ψ 2 must have a
node at some value ξ , where a ≤ ξ ≤ b . This proves the theorem that Ψ 2 has more nodes than Ψ 1 if E 2 > E 1 .

The Schrödinger Equation in Two Dimensions


The extension of the Schrödinger equation into more than one dimension is a fairly straightforward process. The general
two dimensional equation in Cartesian coordinates is

h– 2 ∂ 2 Ψ ( x, y ) ∂ 2 Ψ ( x, y ) 
– --------  ----------------------- - + V ( x, y )Ψ ( x, y ) = EΨ ( x, y ) .
- + ----------------------- Eq. 4-23
2m  ∂x 2 ∂y 2 

Note that the format remains the same as in the one dimensional case. The sum of the kinetic and potential terms gives the total
energy term. The only difference is that Ψ and V are now a functions of two variables, and the kinetic term takes this fact into
account by using partial derivatives.

Particle in a two dimensional box. As an example of a two dimensional application, let us consider a particle in a two dimen-
sional box. Let the center of the box be at the origin, and the box dimensions be x = L x and y = L y . Define the potential to
be zero inside the box and infinite everywhere outside of it:

 
 – ----
Lx Lx Lx
 – L----x > x > ----
 2 ≤ x ≤ ---- 2  2 2
V = 0 for  and V = ∞ for  .
 Ly Ly  Ly Ly
 – ---- ≤ y ≤ ----  – --- - > y > ----
2 2 2 2
 

The approach we present is contingent on the potential being separable into two components. This is referred to as the separa-
tion of variables. Thus, if we can write V ( x, y ) = V x ( x ) + V y ( y ) and Ψ ( x, y ) = X ( x )Y ( y ) , the Schrödinger equation becomes

h– 2 ∂ 2 X ( x )Y ( y ) ∂ 2 Y ( y )X ( x ) 
– --------  ---------------------------- - + V ( x, y )X ( x )Y ( y ) = EX ( x )Y ( y ) .
- + ---------------------------- Eq. 4-24
2m  ∂x 2 ∂y 2 

Since X ( x ) is a function of x only and Y ( y ) is a function of y only, the partial derivatives simplify to

53
More on the Schrödinger Equation: The Schrödinger Equation in Two Dimensions

h– 2 d 2 X ( x ) d2Y( y)
– --------  ----------------- - X ( x ) + V ( x, y )X ( x )Y ( y ) = EX ( x )Y ( y ) .
- Y ( y ) + ---------------- Eq. 4-25
2m dx  2
dy 2 

Dividing both sides by X ( x )Y ( y ) and separating the potential term into components gives

h– 2 1 d 2 X ( x ) 1 d2Y ( y )
– --------  ------------ ----------------- - + Vx ( x ) + Vy ( y ) = E .
- + ----------- ---------------- Eq. 4-26
2m  X ( x ) dx 2
Y ( y ) dy 2 

Notice that we can now split the equation into two separate ones, one solely in terms of x and the other solely in terms of y:

h– 2 1 d 2 X ( x ) 
– --------  ------------ -----------------
- + V x ( x ) = ε ( x ) , a function of x alone, Eq. 4-27
2m X ( x ) dx 2 

h– 2 1 d 2 Y ( y ) 
– --------  ----------- ----------------
- + V y ( y ) = υ ( y ) , a function of y alone. Eq. 4-28
2m  Y ( y ) dy 2 

The energy of the system is fixed. Since the sum of the two energies ε ( x ) + υ ( y ) = E is a constant, the individual energies must
be fixed as well. The two equations can now be rewritten in strikingly familiar form:

h– 2 d 2 X ( x )
- + Vx ( x ) = Ex X ( x )
– -------- ----------------- Eq. 4-29
2m dx 2

h– 2 d 2 Y ( y )
- + Vy ( y ) = Ey Y ( y ) .
– -------- ---------------- Eq. 4-30
2m dy 2

This is just a set of two one dimensional Schrödinger equations, which we have solved previously (see Eq. 3-39). The
unnormailized solutions (inside the box) are

 nx π
 cos --------- x, n x = an odd positive integer
 Lx n x2 h 2
Xn ( x ) =  , E x = -------------2 and Eq. 4-31
 nx π 8mL x
- x, n x = an even positive integer
 sin --------
Lx

54
More on the Schrödinger Equation: The Schrödinger Equation in Two Dimensions


 cos n yπ
--------- y, n y = an odd positive integer
 Ly n y2 h 2
Yn ( y ) =  , E y = -------------2 . Eq. 4-32
 nyπ 8mL y
- y, n y = an even positive integer
 sin --------
Ly

h 2 n x2 n y2
The total energy given as E = E x + E y can now be expressed as E = --------  -----2- + -----2-  . Let A = L x L y represent the area of
8m  L x L y 
the two dimensional box. If L x ≈ L y , then

n 2 = n x2 + n y2 and Eq. 4-33

h2 n2
E = -------- ----- . Eq. 4-34
8m A

Note that Eq. 4-33 imposes a limitation on the allowed values of n 2 , as summarized in the tables below:
Table 4.1 - Allowed quantum states for a particle in a two dimensional box including the degenerate cases.

number of
number of number of number of states
possible quantum possible quantum possible quantum
n2 ways numbers n2 ways numbers n2 ways numbers n g(n)
1 0 10 2 3, 1 19 0 0-1 0
2 1 1, 1 11 0 20 2 4, 2 1-2 1
3 0 12 0 21 0 2-3 3
4 0 13 2 3, 2 22 0 3-4 3
5 2 2, 1 14 0 23 0 4-5 4
6 0 15 0 24 0 5-6 7
7 0 16 0 25 2 4, 3
8 1 2, 2 17 2 4, 1 26 2 5, 1
9 0 18 1 3, 3 27 0

Clearly, as the quantum numbers increase, so does the number of available states. The table also introduces the very important
concept of degeneracy, or the idea that for a fixed total energy, the system can have more than one possible configuration. For
example, for the quantum number n 2 value of 5, there are two possibilities: either n x = 2 and n y = 1 or n x = 1 and n y = 2 .

55
More on the Schrödinger Equation: The Schrödinger Equation in Three Dimensions

This method of counting the number of states between n and n + 1 becomes inconveniently laborious as the den-
sity of states increases. A faster, though approximate method is easily derived with some trigonometry. Let n x = n cos θ
and n y = n sin θ . By the Pythagorean theorem n 2 = n x2 + n y2 , or n = n x2 + n y2 . The number of states is the area of the
circular band between n and n + 1 in the positive (first) quadrant. In short,

πn
g ( n ) ≈ ------- . Eq. 4-35
2

The Schrödinger Equation in Three Dimensions


The three dimensional Schrödinger equation is exactly analogous to the one and two dimensional versions. The general
equation in Cartesian coordinates is written

h– 2 ∂ 2 Ψ ( x, y, z ) ∂ 2 Ψ ( x, y, z ) ∂ 2 Ψ ( x, y, z ) 
– --------  -----------------------------
- + ----------------------------- - + V ( x, y, z )Ψ ( x, y, z ) = EΨ ( x, y, z ) .
- + ----------------------------- Eq. 4-36
2m  ∂x 2 ∂y 2 ∂z 2 

As in the two dimensional case, the format of the equation remains the same. The sum of the kinetic and potential terms gives
the total energy term. All functions are now three dimensional, which results in the three partial derivatives.

Particle in a three dimensional box. TobecomefamiliarwiththethreedimensionalSchrödingerequation,weresortyetagain


to the particle in the box problem. The box is centered at the origin, with dimensions x = L x , y = L y , and z = L z . The
potential is zero inside the box and infinite everywhere outside of it:

 Lx Lx  Lx Lx
 – ---- ≤ x ≤ ----  – ---- > x > ----
 2 2  2 2
 Ly L  Ly Ly
V = 0 for  – ---- ≤ y ≤ ----y and V = ∞ for  – --- - > y > ---- .
 2 2  2 2
 L L  L L
 – ----z ≤ z ≤ ----z  – ----z > z > ----z
 2 2  2 2

As in the two dimensional case, we assume that the potential and the wavefunction can be separated into three orthogonal com-
ponents. A separation of variables and a derivation exactly analogous to that presented on page 53 yields:

56
More on the Schrödinger Equation: The Schrödinger Equation in Three Dimensions

h– 2 ∂ 2 X ( x )Y ( y )Z ( z ) ∂ 2 Y ( y )X ( x )Z ( z ) ∂ 2 Z ( z )X ( x )Y ( y ) 
– --------  ----------------------------------------
- + ---------------------------------------- - + V ( x, y, z )X ( x )Y ( y )Z ( z ) = EX ( x )Y ( y )Z ( z )
- + ---------------------------------------- Eq. 4-37
2m  ∂x 2 ∂y 2 ∂z 2 

h– 2 ∂ 2 X ( x ) ∂2Y ( y ) ∂2Z( z )
– --------  ----------------
-Y ( y )Z ( z ) + ----------------X ( x )Z ( z ) + - X ( x )Y ( y ) + V ( x, y, z )X ( x )Y ( y )Z ( z ) = EX ( x )Y ( y )Z ( z )
---------------- Eq. 4-38
2m  ∂x 2 ∂y 2 ∂z 2 

h– 2 1 d2X ( x ) 1 d2Y ( y ) 1 d2 Z ( z )
– --------  ------------ -----------------
- + ----------- ---------------- - + V x ( x ) + Vy ( y ) + Vz ( z ) = E .
- + ----------- ---------------- Eq. 4-39
2m  X ( x ) dx 2
Y ( y ) dy 2
Z ( z ) dz 2 

Exactly as in the two dimensional case, the Schrödinger equation can now be split into one dimensional components.

h– 2 1 d 2 X ( x ) 
– --------  ------------ -----------------
- + V x ( x ) = ε ( x ) , a function of x alone, Eq. 4-40
2m  X ( x ) dx 2 

h– 2 1 d 2 Y ( y ) 
– --------  ----------- ----------------
- + V y ( y ) = υ ( y ) , a function of y alone, Eq. 4-41
2m  Y ( y ) dy 2 

h– 2 1 d 2 Z ( z ) 
– --------  ----------- ----------------
- + V z ( z ) = ζ ( z ) , a function of z alone. Eq. 4-42
2m  Z ( z ) dz 2 

The energy of the system is fixed, so ε ( x ) + υ ( y ) + ζ ( z ) = E is a constant. The three equations are easily rewritten in the usual
form of the one dimensional Schrödinger equation:

h– 2 d 2 X ( x )
- + Vx ( x ) = Ex X ( x ) ,
– -------- ----------------- Eq. 4-43
2m dx 2

h– 2 d 2 Y ( y )
- + Vy ( y ) = Ey Y ( y ) ,
– -------- ---------------- Eq. 4-44
2m dy 2

h– 2 d 2 Z ( z )
- + Vz ( z ) = Ez Z ( z ) .
– -------- ---------------- Eq. 4-45
2m dz 2

57
More on the Schrödinger Equation: The Schrödinger Equation in Three Dimensions

The solutions are of the very familiar form:


 cos n xπ
 --------- x, n x = an odd positive integer n x2 h 2
Lx
Xn ( x ) =  , E x = -------------2 , Eq. 4-46
 nπ 8mL x
 sin ------
- x , n = an even positive integer
Lx


 cos n yπ
--------- y, n y = an odd positive integer
 Ly n y2 h 2
Yn ( y ) =  , E y = -------------2 and Eq. 4-47
 ny π 8mL y
- y, n y = an even positive integer
 sin --------
Ly

 nzπ
 cos --------- z, n z = an odd positive integer
 Lz n z2 h 2
Zn( z ) =  , E z = -------------2 . Eq. 4-48
 nz π 8mL z
- z, n z = an even positive integer
 sin --------
Lz

h 2 n x2 n y2 n z2
The total energy given as E = E x + E y + E z can be expressed as E = --------  -----2- + -----2- + ------  . Let V = L x L y L z represent the volume
8m  L x L y L z 
of the box. If L x ≈ L y ≈ L z , then n 2 = n x2 + n y2 + n z2 and

h2 n2
E n = -------- --------
-. Eq. 4-49
8m V 2 / 3

58
More on the Schrödinger Equation: The Schrödinger Equation in Three Dimensions

Again, there is a limitation on the allowed values of n 2 , which we summarize in tables:


Table 4.2 - There seem to be fewer allowed quantum states for the particle in a three dimensional box, but the level of degeneracy quickly increases as the
energy of the system is increased.

number of
number of number of number of states
possible quantum possible quantum possible quantum
n2 ways numbers n2 ways numbers n2 ways numbers n g(n)
1 0 6 3 2, 1, 1 11 3 3, 1, 1 1-2 1
2 0 7 0 12 1 2, 2, 2 2-3 3
3 1 1, 1, 1 8 0 13 0 3-4 13
4 0 9 3 2, 2, 1 14 6 3, 2, 1 4-5 21
5 0 10 0 15 0

The number of degenerate states is increased over the two dimensional case. Counting individual states available for each
value of n involves a fair amount of effort, especially as n becomes large. To circumvent this difficulty, let us note that the
quantum numbers n x, n y, n z can be thought of as three dimensional coordinates, each set of three representing a point
in space. For sufficiently large n, g ( n ) , the number of states in a unit increment of n, is approximately equal to the vol-
ume of a shell of unit thickness with radius n (in the first quadrant, where n x, n y, and n z are all positive), or

4πn 2 πn 2
g ( n )dn ≈ ------------dn = --------- dn . Eq. 4-50
8 2

Air molecules in a three dimensional box. Let us see how this approach to the number of available states works for
physical examples. Consider a 1 liter cube filled with air at 1 atm pressure. The molecular mass of air is about
29 g ⋅ mol –1 , which makes the mass of a single “air” particle m = 46 × 10 –24 g . Invoking Eq. 4-49, and using
V 2 / 3 = 100 cm 2 , we compute that E n = 1.2 × 10 –33 ⋅ n 2 erg. At 1 atm pressure, there are approximately 3 × 10 22 mole-
cules inside the cube. Recalling Eq. 4-50, we can note that the total number of states with quantum number less than
1 4πN 3
some arbitrary limit N is --- ⋅ ------------- . In order for each molecule to be in a different quantum state, the number of available
8 3
1 4πN 3
states must be greater than the number of molecules, --- ⋅ ------------- > 3 × 10 22 , or N > 3 × 10 7 . The energy of the highest occu-
8 3
pied state E N = 1.2 × 10 –18 erg , which corresponds roughly to a temperature of 0.01 K (taking the average energy to be
3
--- kT ).
2

59
More on the Schrödinger Equation: The Schrödinger Equation in Three Dimensions

It is interesting to calculate how closely the individual states are spaced apart, on average, at room temperature.
3
Thus, taking the temperature to be 300 K , E = --- kT ≈ 6 × 10 –14 erg . Using E n = 1.2 × 10 –33 ⋅ n 2 erg , we solve
2
6 × 10 –14 erg = 1.2 × 10 –33 ⋅ n 2 erg for n, to obtain n = 5.5 × 10 9 . This is the number of occupied states at room temper-
ature. Dividing the energy of the highest occupied state by the number of occupied states yields the average spacing
6 × 10 –14 erg
- = 1 × 10 –23 erg apart.
between adjacent states. Thus, at room temperature individual energy levels are ------------------------------
5.5 × 10 9
Because this amount of energy is very small, to a good approximation the energy levels represent a continuum, and air at
room temperature can be readily described using classical approximations.

Electrons in a three dimensional box. Electrons in a sample of metal are also often treated as particles in a three
dimensional box. Consider, for example, a piece of lithium metal. A good question to ask is what is the energy of the
highest filled orbital in lithium’s conduction band? With a density of 0.5 g ⋅ cm –3 , a mole of lithium (molar mass
6.9 g ⋅ mol –1 ) has a volume of 13.8 cm 3 . There are thus 6 × 10 23 free unpaired electrons from the 2s orbitals of each lith-
ium atom. Since two electrons of opposite spin can occupy a single orbital, this requires that 3 × 10 23 orbitals be avail-
1 4πN 3
able. This number is given by --- ⋅ ------------- . Solving for N, we obtain N = 8.3 × 10 7 . Using Eq. 4-49, we find that
8 3
E N = 7.2 × 10 –12 erg = 4.5 eV .

60
LECTURE 5 The Harmonic Oscillator

The quantum mechanical description of the harmonic oscillator is at the heart of phys-
ical chemistry. It is the basis for quantifying all forms of molecular vibrations, which are
truly omnipresent. Vibrations are a means of storing energy and thus a key topic in thermo-
dynamics at the molecular level, they play an important role in spectroscopy. Measuring
vibrations is a method for comparing the stiffness of bonds. We therefore devote an entire
section to a derivation of harmonic oscillator wavefunctions and their properties.

Solution with the Schrödinger Equation


We approach the harmonic oscillator problem with the one dimensional Schrödinger
equation. Unlike the particle in the box where the potential was either zero or infinite,
however, the potential for a harmonic oscillator is a specific function, as given by Hooke’s
law:

1
V ( x ) = --- kx 2 . Eq. 5-1
2

When this potential is included in the Schrödinger equation, we obtain

h– 2 d 2 Ψ 1
– -------- ---------2- + --- kx 2 Ψ = EΨ , where Ψ is a function of x. Eq. 5-2
2m dx 2

Recall from classical mechanics that the classical vibration frequency (in radians per second)
k
ω = ----- , where k is the spring constant and m is the particle’s mass. This allows us to
m
rewrite Eq. 5-2 as

d2Ψ m2ω2 2 2m
-x Ψ = -------
– ---------2- + ------------ - EΨ . Eq. 5-3
dx h– 2 h– 2

mω 2mE
Let α = --------
- and λ = -----------
- to write
h– h– 2

61
The Harmonic Oscillator: Solution with the Schrödinger Equation

d2Ψ
– ---------2- + α 2 x 2 Ψ = λΨ . Eq. 5-4
dx

For very large values of x (so that α 2 x 2 » λ ), Eq. 5-4 reduces to

d2Ψ
– ---------2- + α 2 x 2 Ψ = 0 . Eq. 5-5
dx

Therefore, the asymptotic solution is

– αx 2 dΨ – αx 2
Ψ = exp  ------------  ; -------- = – α exp  ------------  . Eq. 5-6
 2  dx  2 

αx 2
Note that Ψ = exp  ---------  also satisfies the differential equation, but diverges for large x. The solution can be easily verified:
 2 

d2Ψ – αx 2 – αx 2 – αx 2
---------2- = – α exp  ------------ + α 2 x 2 exp  ------------ ≈ α 2 x 2 exp  ------------ for large x.
dx  2   2   2 

– αx 2
Note that the function Ψ = exp  ------------  will solve Eq. 5-4 if α = λ . Since α = --------
mω 2mE
- and λ = -----------
- , the energy associ-
 2  h– h– 2
ated with this particular wave function is

h– ω hν ω
E = -------- = ------ , where ν = ------ . Eq. 5-7
2 2 2π

– αx 2
Ψ = exp  ------------  has no nodes. It is the lowest energy solution.
 2 

The general solution set will be of the form

– αx 2
Ψ = exp  ------------  f ( x ) , Eq. 5-8
 2 

62
The Harmonic Oscillator: Solution with the Schrödinger Equation

which has the proper asymptotic form for almost any “reasonable” function f ( x ) at large x. f ( x ) will be an oscillatory function
1
in the region where E > V , i.e. where E – --- kx 2 > 0 , or between the classical turning points. The classical turning points are the
2
points where E = V :

2E 1 2E
x tp = ± ------ = ± ---- ------ . Eq. 5-9
k ω m

– αx 2
If the general solution is of the form Ψ = exp  ------------  f ( x ) , we can write
 2 

– αx 2 – αx 2
-------- = – αx exp  ------------  f ( x ) + exp  ------------  f′ ( x ) and

dx  2   2 

d2Ψ – αx 2 – αx 2 – αx 2 – αx 2
-------2- =  – α exp  ------------  + α 2 x 2 exp  ------------   f ( x ) – 2αx exp  ------------  f′ ( x ) + exp  ------------  f′′ ( x. Eq. 5-10
dx   2   2   2   2 

Introducing Eq. 5-10 into Eq. 5-4 and cancelling Ψ yields

αf ( x ) + 2αxf ′ ( x ) – f ′′ ( x ) = λf ( x ) or

( α – λ )f ( x ) + 2αxf ′ ( x ) – f ′′ ( x ) = 0. Eq. 5-11

Eq. 5-11 solves if f ( x ) is a polynomial.* The order of the polynomial corresponds to the number of nodes in the wave-
function. In other words, let
v

f(x) = ∑ an x n, which makes


n=0
Eq. 5-12

f ′( x ) = ∑ nan x n – 1, n ≠ 0, and
n=0
Eq. 5-13

*. See Levine, Section 4.2 for a more general proof of the polynomial solution.

63
The Harmonic Oscillator: Solution with the Schrödinger Equation

f ′′ ( x ) = ∑ n ( n – 1 )an x n – 2, n ≠ 0, 1.
n=0
Eq. 5-14

Substituting into Eq. 5-11, we obtain

α – λ ) ∑ a n x n + 2αx ∑ na n x n – 1 – ∑ n ( n – 1 )a n x n – 2 = 0or

α – λ ) ∑ a n x n + 2α ∑ na n x n – ∑ n ( n – 1 )a n x n – 2 = 0, which can be rearranged to

∑ [ ( 2n + 1 )α – λ ]a n x n – ∑ n ( n – 1 )a n x n – 2 = 0. Eq. 5-15

will only hold true if all powers of x from all the sums cancel out. Thus, we can equate all like powers of x with the
expression

[ λ – ( 2n + 1 )α ]a n = – ( n + 2 ) ( n + 1 )a n + 2. Eq. 5-16

Note that the index on the right side has to be two values of n higher in order to correspond to the correct power of x. If we let
the solution polynomial be of order v, then a v ≠ 0 and a v + 2 = 0 (since there are no terms in higher powers of x than the
degree of the polynomial itself). Eq. 5-16 can be thus restated

[ λ – ( 2v + 1 )α ]a v = 0 and solved for λ : Eq. 5-17

λ = ( 2v + 1 )α , Eq. 5-18

mω 2mE
with α = --------
- and λ = -----------
- as specified previously. Combining these equations with Eq. 5-18 produces an explicit
h– h– 2
expression for the energy levels of the harmonic oscillator:

h– 2 mω
- ( 2v + 1 ) = h– ω  v + --- = hν  v + --- .
1 1
E v = -------- -------- Eq. 5-19
2m h–  2   2

The (unnormalized) wavefunctions can be obtained from the recursion relation given in Eq. 5-16 with the expression for
λ from Eq. 5-18. Substituting and solving, we obtain

2(v – n)
a n + 2 = ------------------------------------ αa n. Eq. 5-20
(n + 2)(n + 1)

64
The Harmonic Oscillator: Properties

– αx 2
Recalling from Eq. 5-8 that the general form of the solutions is Ψ v ( x ) = exp  ------------  f v ( x ) , we use the recursion to find
 2 
that

f0 = 1
f1 = x . Eq. 5-21
2
f 2 = 1 – 2αx

Note that f v contains only even powers of x if v is even and only odd powers of x if v is odd. For the even functions, the follow-
ing integrals may be useful when evaluating expectation functions:
∞ ∞
1 π
∫ exp ( – αx 2 ) dx = -------α –∫∞ exp ( –x 2 ) dx =
–∞
---
α
Eq. 5-22

∞ ∞
∂ ∂ π
------- ∫ exp ( – αx 2 ) dx = ∫ – x 2 exp ( – αx 2 ) dx = ------- ---, while Eq. 5-23
∂α –∞ –∞ ∂α α

∫ –x n exp ( – αx 2 ) dx
–∞
= 0 for odd n. Eq. 5-24

Properties
We now briefly summarize the properties of the harmonic oscillator. The first is the symmetry of the eigenfunctions:

Ψv ( –x ) = ( – 1 ) v Ψv ( x ) . Eq. 5-25

In other words, if v is an odd integer, then the wavefuntion is odd to inversion. If v is even, then the function is even. Note that
the vibrational quantum number (v) corresponds to the number of nodes of the wavefunction, since it is also the degree of the
polynomial.

A wavefunction allows us to compute the probability density of the harmonic oscillator at any point. For example, we
may wish to look at the extremes of the classically allowed region, i.e. the classical turning points where E = V . The position
of the turning points is derived from Eq. 5-9 and Eq. 5-19:

65
The Harmonic Oscillator: Properties

2E ( 2v + 1 )h– ω ( 2v + 1 )h– ( 2v + 1 )
x tp = ± ------ = ---------------------------- = ------------------------ = -------------------- . Eq. 5-26
k mω 2 mω α

– αx 2
For v = 0 , x tp2 = --- , and Ψ 0 ( x ) = N 0 exp  ------------ , where N 0 is the normalization factor. Thus,
1
α  2 

Ψ 0 ( x tp ) = N 0 exp  – --- = 0.607N 0 . The square of the function is the probability density: Ψ 02 ( x tp ) = N 02 exp ( – 1 ) = 0.37N 02 . It
1
 2
is important to note that beyond the turning point, the probability density is far from zero, a result expected in a classical
approach. For example, at the point 2x tp , the square of the wavefuntion is Ψ 02 ( 2x tp ) = N 02 exp ( – 4 ) = 0.02N 02 , a non-negligible
value. Clearly, the quantum character of the harmonic oscillator allows it to occupy positions completely forbidden as unphys-
ical in the classical approach.

The functional form of a probability density is useful in computing the total energy stored in a system of many
oscillators, such as molecules. Consider, for example, one mole of harmonic oscillators with a vibrating frequency of
3000 cm –1 . This is the frequency of CH bond stretching or HCl vibration. The question to ask is how much energy is
stored in such a system? The probability of excitation of a single oscillator is given by the Boltzmann law, or

P = exp  ---------- . Taking care to convert wavenumbers to Hz, we arrive at


– hν
 kT 
6.7 × 10 –27 ⋅ 3 × 10 10 ⋅ 3000
P ν = exp  ----------------------------------------------------------------- = 6.17 × 10 –7. The energy contained in a mole of excited oscillators is
 1.4 × 10 –16 ⋅ 300 
E = N a hν = 6 × 10 23 ⋅ 6.7 × 10 –27 ⋅ 3 × 10 10 ⋅ 3000 = 3.6 × 10 11 erg. The energy of the system (in units of RT ) is
hν 6.7 × 10 –27 ⋅ 3 × 10 10 ⋅ 3000
E = hνP ν N a = ------ ⋅ P ν ⋅ RT = ----------------------------------------------------------------- ⋅ 6.17 × 10 –7 ⋅ RT = 9 × 10 –6 ⋅ RT (here, we use the fact that
kT 1.4 × 10 –16 ⋅ 300
R
k = ------ to eliminate Avogadro’s number from the equation). The odd-looking energy units of RT lend themselves to a
Na
direct comparison with a classical system, where the energy is E = RT per mole. Quantum mechanics brings about a
correction of five orders of magnitude.

66
LECTURE 6 Tunneling

Quantum treatment of such problems as the particle in the box and the harmonic
oscillator demonstrated that the particle or oscillator has a non-zero probability of occupy-
ing space classically forbidden as unphysical, where the potential is greater than the energy
of the system, or V > E . This kind of behavior is known as tunneling and is the basis of
modern atomic-scale microscopic techniques, such as Scanning Tunneling Microscopy
(STM). The quantitative and rigorous mathematical treatments of general cases of tunneling
are rather involved and are best done on a computer. Below, we present a simple, yet illus-
trative example of a particle confined to a box with infinite potential walls that is divided by
a barrier of finite width and potential.

A Divided Box
In Lecture 3 we derived the
solutions for a particle confined in a
box with walls of first infinite and
 
then finite potential. In the case of
the latter, we found that the particle
is capable of penetrating finite poten-
tial barriers and that the probability
of finding it outside of the confining
V0
box is non-zero.

Consider a slightly more com-


plicated situation, where a box of
infinite walls contains a barrier of I II III IV V
finite height (or finite potential
energy). The potential in the rest of L L
the box is zero. A schematic drawing
is shown in Figure 6.1. The solution
to the Schrödinger equation given
this piecewise potential is a piecewise
function. The froms of the pieces -L - a - a a L + a
2 2 2 2
arise from our previous experience
with particles in boxes. Figure 6.1 - A box with walls of infinite potential and a finite-
potential barrier. The potential in regions II and IV is zero.

67
Tunneling: A Divided Box

Table 6.1 - Summary of the particle in two square wells separated by a finite barrier.

Region I II III IV V
Potential ∞ 0 V0 0 ∞
Domain a a
x < – L + ---  – L + ---  ≤ x ≤ – --- + --- ≤ x ≤ L + ---  x > L + --- 
a a a a a a
– --- ≤ x ≤ + ---
 2  2  2 2 2 2  2  2

Ψ(x) 0 A 1 sin kx + B 1 cos kx Ce –κx + De +κx A 2 sin kx + B 2 cos kx 0

The box can be divided into five regions, as shown. In regions I and V, the potential is infinite and there is no probability
of finding the particle there. In regions II and IV, the potential is zero, and we expect the wavefuntion to be of the trigonometric
form we derived in Eq. 3-39 on page 39. In region III, there is a non-zero finite potential, where we expect an exponen-
tial decay of probability, and a functional form analogous to that seen in Eq. 3-46 on page 42. All of this is summarized in
Table 6.1.

The potential is symmetric about x = 0 , and therefore the solutions will be odd and even functions. The number of
nodes will be a monotonically increasing function of energy, the first eigenfunction Ψ 0 with the lowest energy E 0 will
have no nodes and thus be even. The second, Ψ 1 , with one node and energy E 1 , will be odd, the third even again, etc.

The dominant form of the wavefunctions is determined in regions II and IV, where the functions are large. In these

regions the functions are trigonometric, terminating at ± L + ---  . The length of each region, L , effectively determines the
a
 2
energy for those states with energy appreciably less than the barrier height, or E < V 0 .

h2n2
The first two eigenfunctions are degenerate in energy. If we were to use Eq. 3-42 ( E n = -------------2 ), both have the same
8mL
h2
value of n = 1 . Thus, E 0, E 1 ≈ -------------2 . The two functions are just the even and odd combinations of two sine functions.
8mL
The curly braces indicate the domain of each “piece” of the function.

a a
 x + ---   x – --- 
2   2  a a 
Ψ 0 = – sin π -------------   – L + ---  ≤ x ≤ – ---  + sin π ------------   + --- ≤ x ≤ L + ---  
a a
Eq. 6-1
 L   2  2  L  2  2 
   

68
Tunneling: A Divided Box

a a
 x + ---   x – --- 
  2  a a 
Ψ 1 = – sin π -------------   – L + ---  ≤ x ≤ – --- – sin π ------------   + --- ≤ x ≤ L + ---  
2 a a
Eq. 6-2
 L   2 2  L  2  2 
   

The only difference between the two is that Eq. 6-2 is an odd function overall, and it forces a node to exist in region III.
The second set of functions is exactly analogous. There are two of them, one odd and one even, where n = 2 for both.
h2
E 2, E 3 ≈ -------------2 .
2mL

a a
 x + ---   x – --- 
2   2  a a 
Ψ 2 = – sin 2π -------------   – L + ---  ≤ x ≤ – ---  + sin 2π ------------   + --- ≤ x ≤ L + ---  
a a
Eq. 6-3
 L   2 2  L  2  2 
   

a a
 x + ---   x – --- 
  2  a a 
Ψ 3 = – sin 2π -------------   – L + ---  ≤ x ≤ – --- – sin 2π ------------   + --- ≤ x ≤ L + ---  
2 a a
Eq. 6-4
 L   2 2  L  2  2 
   

2m ( V 0 – E )
Region III, where E – V 0 < 0 is computationally more difficult. Recall that κ = - . Thus, we expect that the
---------------------------
h– 2
a
wavefunction will decay exponentially as we move into the wall, from ± --- towards x = 0 . However, since the potential is
2
finite, regardless of the thickness of the barrier, the probability never reaches zero. Furthermore, if the overall wavefunction is
even ( Ψ 0, Ψ 2 ), the exponentially decaying function from region II adds to the exponentially decaying function from region IV,
which actually raises the probability that the particle will tunnel through the barrier. Note that if the overall function is odd
( Ψ 1, Ψ 3 ), a node must exist at x = 0 and tunneling is forbidden. There is zero probability of the particle existing at x = 0 .

For E i > V 0 , the wavefunction is trigonometric everywhere, because the energy of the particle surpasses the barrier. The

problem is then soluble as a particle in a box with infinite walls and box length 2 L + ---  .
a
 2

The mathematics of tunneling make presenting the topic on paper a considerable challenge. Take some time to use the Math-
ematica notebook that will be provided to you to see a complete plot of the solutions and to experiment by varying parameters.

69
LECTURE 7 The Virial Theorem

The virial theorem connects the average kinetic, average potential and total energies.
The exact form of the theorem depends on the specified potential. Here, we first derive the
general case and follow up with some examples from classical mechanics. The validity of the
virial theorem in quantum mechanics will be discussed later.

General Case
In Lecture 2, we developed a set of general coordinates, ( p, q ) . For the purposes of the
n

virial theorem, we define a quantity p i ⋅ q i (in case of n dimensions, we use ∑ p i ⋅ qi ). We


i=1

are interested in relating the time average of the potential to the time average of the kinetic
energy. To do this, we first average the time derivative of p i ⋅ q i :

τ
1 d 1
--- ∫ ( p i ⋅ q i ) dt = --- [ p i ⋅ q i ( τ ) – p i ⋅ q i ( 0 ) ]. Eq. 7-1
τ 0 dt τ

For bound states, this goes to 0 as τ increases to infinity, because both p i ( τ ) and q i ( τ ) are
always finite.

· dp ∂H · dq i ∂H
Recall from Eq. 2-16 on page 15 that p i = --------i = – ------- and q i = -------- = ------- . The
dt ∂q i dt ∂p i
d
expression ( p i ⋅ q i ) can therefore be rewritten as
dt

d · · ∂H ∂H
( p ⋅ q ) = p i ⋅ q i + p i ⋅ q i = – ------- q i + ------- p i . Eq. 7-2
dt i i ∂q i ∂p i

The Hamiltonian is the sum of the kinetic and potential energies, H = T + V , or as stated
p⋅p
in Eq. 2-8, H ( p, q ) = ---------- + V ( q ) . Clearly,
2m

71
The Virial Theorem: Specific Examples

∂H ∂V ∂H ·
------- q i = -------q i and ------- p i = q i p i = 2T . Eq. 7-3
∂q i ∂q i ∂p i

Through some simple substitutions, the time average in Eq. 7-1 can be rewritten as
τ
1  ∂H ∂H · ∂V
--- ∫ – ------- q i + ------- q i dt = – -------q i + 2T Eq. 7-4
τ 0  ∂q i ∂p i  ∂q i

for τ » 0 , or a long time average. As pointed out earlier, as τ → ∞ , the integral in Eq. 7-4 approaches zero. Using a long time
average, we can write the fundamental relation of the virial theorem:

∂V
-------q = 2T . Eq. 7-5
∂q i i

If the potential is an n-th order polynomial function of the coordinate, V ( q i ) = Cq in , then Eq. 7-5 becomes

nV = 2T . Eq. 7-6

Specific Examples
The relationship shown in Eq. 7-6 is quite general. The precise relationship depends on the functional form of the poten-
tial. In the case of the harmonic oscillator, for example, the potential has a second degree dependence on the coordinate
1
( V ( x ) = --- kx 2 ). Thus, n = 2 , and the virial theorem for the harmonic oscillator states that V = T , or that the average poten-
2
tial energy is exactly equal to the average kinetic energy.

1
Another example we can look at is the central Coulomb potential, where V ( r ) ∝ --- . The potential has an inverse depen-
r
dence on the coordinate, which makes n = – 1 . Thus, the virial theorem for a spherically symmetric Coulomb potential is
– V = 2T .

The virial theorem also applies to quantum systems, which we will prove a little later. However, let us apply the result to
the quantum oscillator to obtain the mean displacement or the mean oscillation amplitude. The potential is the same as for the

72
The Virial Theorem: Specific Examples

1
classical harmonic oscillator, so V = T = --- E . Recall from Eq. 5-19 that the enegy of a quantum oscillator in a quantum state v
2

is given by E v = h– ω v + ---  . The average potential energy is equal to half the total energy:
1
 2

1
V = --- E . Eq. 7-7
2

--- kx = --- h– ω v + ---  ,


1 2 1 1
Eq. 7-8
2 2  2

where k = mω 2 (the spring constant). Substituting and solving for x 2 , we obtain

v + 1 ---  h– v + 1 ---  h
 2  2
x 2 = --------------------- = -------------------- . Eq. 7-9
mω 4πmν

The final result, the main displacement, is just a square root of x 2 :

v + 1 ---  h
 2
∆x = x2 = -------------------- . Eq. 7-10
4πmν

73
LECTURE 8 The Boltzmann Distribution

The Boltzmann distribution law relates the relative number of molecules in a quan-
tum state to the energy of the state and the temperature of the system. It provides a relation
between the microscopic components and the macroscopic system at a given temperature
(thermostat). The Boltzmann law is used so frequently that we introduce the concept here
without proof, only as a tool to be invoked as needed.

Relative Occupancy of Energy Levels


The energy levels of a system are E i , with E 0 being the lowest level. The number of
molecules in an energy level E i with a degeneracy g i is symbolized by N i . The number is
only meaningful as a quantity relative to the number of molecules in the lowest energy level.
Thus, we write

Ei
g i exp – -------- 
Ni  kb T
----- = -------------------------------- , Eq. 8-1
N0 E0
g 0 exp – -------- 
 kb T

where k b is Boltzmann’s constant (often abbreviated as k when there is no risk of confusion).


The relationship is most often expressed as

Ei
g i exp – -------- 
 k b T
N i = N 0 -------------------------------- . Eq. 8-2
E0
g 0 exp – -------- 
 k b T

The expression may be converted into the fraction of molecules in the energy state E i
by adding up all the occupancies of all the available quantum states. Keeping in mind that

∑ Ni
i=0
= N , the total number of molecules, we can state

75
The Boltzmann Distribution: Relative Occupancy of Energy Levels

Ei
g i exp – -------- 
∞ ∞
 k b T N0 Ei
N = ∑ N 0 -------------------------------- = -------------------------------- ∑ g i exp – --------  . Eq. 8-3
E E  k bT
g 0 exp – --------  g 0 exp – --------  i = 0
i=0 0 0
 k b T  kb T

Taking the quotient of Eq. 8-2 and Eq. 8-3, we obtain

Ei
g i exp – -------- 
Ni  kb T

-.
----- = --------------------------------------- Eq. 8-4
N
E
∑ g i exp – k------- -
i

i=0 bT


Ei
Making the simple substitution q ( T ) = ∑ g i exp – k-------
i=0 bT
- yields a compact result

Ei
g i exp – -------- 
Ni  kb T
----- = -------------------------------- . Eq. 8-5
N q( T)

For many problems, q ( T ) or the partition function can be easily evaluated.

76
LECTURE 9 Cylindrical Polar Coordinates

Cartesian coordinates are very well suited for particles in boxes, but become unwieldy
and cumbersome in treatments of problems involving central potentials. In those instances,
the cylindrical polar coordinate system proves far more convenient. The cylindrical
polars are defined as follows:

x = ρ cos φ
y = ρ sin φ . Eq. 9-1
z=z

ρ is the radius from the z-axis to the given point ( ρ ≥ 0 ), and φ is the polar angle
( 0 ≥ φ ≥ 2π ). When integrating over all space in cylindrical polar coordinates, the vol-
ume element dV becomes

dV = dxdydz = ρdρdφdz . Eq. 9-2

In case of two dimensions only, dxdy = ρdρdφ .

The Schrödinger Equation


The transformation of the Schrödinger equation into cylindrical polar coordinates
is essentially just a long exercise in the chain rule for differentiation. Some very useful
relationships to keep in mind are:

∂ρ ∂ρ
------ = cos φ ------ = sin φ
∂x ∂y
Eq. 9-3
∂φ sin φ ∂φ cos φ
------ = – ----------- ------ = ------------
∂x ρ ∂y ρ

77
Cylindrical Polar Coordinates: Particle on a Ring

x 2 + y 2 and φ = ----- ln  --------------  , since e iφ = cos φ + i sin φ = -------------- and


1 x + iy x + iy
These are all easily obtained from ρ =
2i  x – iy  ρ
x – iy
e –iφ = -------------- . It is also possible to show that
ρ

∂ 2 f 1 ∂f 1 ∂ 2 f ∂ 2 f
∇2f = --------2 + --- ------ + -----2 --------2 + --------2 . Eq. 9-4
∂ρ ρ ∂ρ ρ ∂φ ∂z

h– 2
The Schrödinger equation is – -------- ∇2Ψ + VΨ = EΨ . If we consider just the two dimensional problem with the coordinates
2m
( x, y ) or ( ρ, φ ) , the equation becomes

h– 2 ∂ 2 Ψ 1 ∂Ψ 1 ∂ 2 Ψ
– --------  ---------2- + --- -------- + -----2 ---------2-  + V ( ρ, φ )Ψ = EΨ . Eq. 9-5
2m  ∂ρ ρ ∂ρ ρ ∂φ 

Particle on a Ring
A practical application for equation Eq. 9-5 is the case of a particle confined to a ring or a circular track. In general,
the potential involved in such a case is only a function of radius from the origin, since there is no preferred polar angle
for the particle; V ( ρ, φ ) = V ( ρ ) . The wavefunction is then a product function Ψ ( ρ, φ ) = P ( ρ )Φ ( φ ) and the Schrödinger
equation becomes

h– 2 ∂ 2 P ( ρ ) 1 ∂P ( ρ ) 1 ∂2Φ( φ )
– --------  ---------------- - P ( ρ ) + V ( ρ )P ( ρ )Φ ( φ ) = EP ( ρ )Φ ( φ ) and
- Φ ( φ ) + --- --------------- Φ ( φ ) + -----2 ----------------- Eq. 9-6
2m ∂ρ  2 ρ ∂ρ ρ ∂φ 2 

h– 2 ∂ 2 P ( ρ ) 1 ∂P ( ρ )  1 h– 2 1 ∂ 2 Φ ( φ ) 1
– --------  ----------------
- + --
- --------------
- ----------- + V ( ρ ) – -------
- ----- ------------------ ------------ = E . Eq. 9-7
2m  ∂ρ 2 ρ ∂ρ  P ( ρ ) 2m ρ 2 ∂φ 2 Φ ( φ )

The equation consists of two parts, separated by variable. In other words, f ( ρ ) + g ( φ ) = E . Since the total energy is a
constant, Eq. 9-7 can be broken up into two separate equations, which are easier to solve; f ( ρ ) = E ρ would be used to
solve the radial part of the overall equation, and g ( φ ) = E φ would be used for the angular part.

The angular part of the solution of Eq. 9-7 is

78
Cylindrical Polar Coordinates: Particle on a Ring

Φ ( φ ) = e iMφ , Eq. 9-8

with M = 0, ±1, …, ± n since Φ ( φ + 2π ) = Φ ( φ ) . The form of Φ ( φ ) is based on its periodicity. Since it is the polar angle, all
of its functions must have the same period of 2π .

h– 2 1 ∂ 2 Φ ( φ ) 1
The differential equation separates into two independent parts if – -------- -----2 -----------------
- ------------ can be approximated by replacing
2m ρ ∂φ 2 Φ ( φ )
1 1 1
-----2 with the constant --------------
- , or -----2 . The expression for the energy of the M-th quantum state is
ρ ρ average
2 ρ0

h– 2 1 2 ( h– M ) 2
E M = -------- -----M = --------------2- . Eq. 9-9
2m ρ 02 2mρ0

Eq. 9-8 and Eq. 9-9 are given as results. Let us take a moment to understand their physical meaning by examining
the angular momentum. Since the motion is confined to the xy plane, the only component of angular momentum is L z .
From Example 2-2 on page 16 recall the classical definition of L z = xp y – yp x . In quantum mechanics, operators (indi-
cated by “hats”) replace functions such as angular momentum. The correspondence is as follows:

x̂ = x and ŷ = y , Eq. 9-10

∂ ∂
p̂ x = – h– ------ and p̂ y = – h– ------ , Eq. 9-11
∂x ∂y

∂ ∂
L̂ z = – ih– x ------ – y ------  . Eq. 9-12
 ∂y ∂x 

Angular momentum is therefore not a function itself, but an operator which can be applied to any function. Thus,

∂ ∂
L̂ z f = – ih– x ------ – y ------  f . Eq. 9-13
 ∂y ∂x 

∂f ∂φ ∂f ∂ρ ∂f ∂f ∂φ ∂f ∂ρ ∂f
The chain rule for differentiation states that ------ = ------ ------ + ------ ------ and ------ = ------ ------ + ------ ------ . From Eq. 9-3, we know that
∂x ∂x ∂φ ∂x ∂ρ ∂y ∂y ∂φ ∂y ∂ρ

79
Cylindrical Polar Coordinates: Particle on a Ring

∂ρ ∂ρ
------ = cos φ ------ = sin φ
∂x ∂y
.
∂φ sin φ ∂φ cos φ
------ = – ----------- ------ = ------------
∂x ρ ∂y ρ

Combining all of this information with the angular momentum operator yields

∂φ ∂f ∂ρ ∂f ∂φ ∂f ∂ρ ∂f
L̂ z f = – ih– ρ cos φ  ------ ------ + ------ ------  – ρ sin φ  ------ ------ + ------ ------  Eq. 9-14
 ∂y ∂φ ∂y ∂ρ   ∂x ∂φ ∂x ∂ρ 

cos φ ∂f ∂f sin φ ∂f ∂f
L̂ z f = – ih– ρ cos φ  ------------ ------ + sin φ ------  – ρ sin φ – ----------- ------ + cos φ ------  Eq. 9-15
ρ ∂φ ∂ρ ρ ∂φ ∂ρ

cos φ ∂f sin φ ∂f
L̂ z f = – ih– ρ cos φ  ------------ ------  + ρ sin φ  ----------- ------  Eq. 9-16
 ρ ∂φ   ρ ∂φ 

∂f ∂f
L̂ z f = – ih–  ------  +  ------  Eq. 9-17
 ∂φ   ∂φ 


L̂ z = – ih– ------ . Eq. 9-18
∂φ

In other words, the angular momentum operator turns out to operate only on a single variable, the polar angle φ , a fact that
was not readily transparent in the original definition in Eq. 9-12.

Using Eq. 9-18 and the solution provided in Eq. 9-8, we can write a general expression for angular momentum.


L̂ z Φ = – ih– ------ e iMφ Eq. 9-19
∂φ

Upon differentiation, the expression simplifies to a very important result:

L̂ z Φ = Mh– Φ . Eq. 9-20

In other words, the value of the angular momentum

80
Cylindrical Polar Coordinates: Particle on a Ring

L z = Mh– . Eq. 9-21

Eq. 9-8 is an eigenfunction of the angular momentum operator and the resulting eigenvalue is a number, not a function;
it is independent of the angle φ . Note that Eq. 9-20 is only true for the functional form Φ ( φ ) = e iMφ . If a linear combina-
tion of two solutions with the same value of M (which we will show to be degenerate) were used, such as e iMφ ±e –iMφ ,
the angular momentum would no longer be independent of the polar angle.

e iMφ ± e –iMφ = λ cos ( φM ) or λ sin ( φM ) , Eq. 9-22

depending on the ± , where λ is a coefficient. If L̂ z operates on Eq. 9-22, the dependence on φ is not eliminated. Note,
however, that if a closed shell is considered, which fills both e iMφ + e –iMφ and e iMφ – e –iMφ , the overall angular momentum
becomes independent of φ . This is equivalent to setting Φ ( φ ) = ( e iMφ + e –iMφ ) + ( e iMφ – e –iMφ ) = 2e iMφ, or a function of the
form given in Eq. 9-8.

h– 2 ∂ 2 P ( ρ ) 1 ∂P ( ρ )  1 h– 2 1 ∂ 2 Φ ( φ ) 1
If we return to the original Schrödinger equation – --------  ----------------
- + --- --------------- ----------- + V ( ρ ) – -------- -----2 -----------------
- ------------ = E , we
2m  ∂ρ 2 ρ ∂ρ  P ( ρ ) 2m ρ ∂φ 2 Φ ( φ )
∂2Φ( φ ) 1
note that given the angular solution Φ ( φ ) = e iMφ , the term -----------------
- ------------ becomes – M 2 , where M is an integer as defined
∂φ 2 Φ ( φ )
above. The radial equation thus becomes

h– 2 ∂ 2 P ( ρ ) 1 ∂P ( ρ )  1 h– 2 1
– --------  ----------------
- + --- --------------- ----------- + V ( ρ ) + -------- -----2 M 2 = E or Eq. 9-23
2m  ∂ρ 2 ρ ∂ρ  P ( ρ ) 2m ρ

h– 2 ∂ 2 P ( ρ ) 1 ∂P ( ρ )   h– 2 1
– --------  ----------------
- + --- --------------- + V ( ρ ) + -------- -----2 M 2 P ( ρ ) = EP ( ρ ) . Eq. 9-24
2m  ∂ρ 2 ρ ∂ρ   2m ρ 

Eq. 9-24 can be somewhat simplified by the substitution F ( ρ ) = ρP ( ρ ) , which gives

1
M 2 + ---
h– 2 d 2 F ( ρ ) h– 2 4
- F ( ρ ) = EF ( ρ ) (steps omitted).
– -------- ----------------- + V ( ρ ) + -------- --------------- Eq. 9-25
2m dρ 2m ρ 2

81
Cylindrical Polar Coordinates: Particle on a Ring

1 1
M 2 + --- M 2 + ---
h– 2 4 h– 2 4
One can consider V ( ρ ) + -------- ---------------
- the effective potential for radial motion, or V eff ( ρ ) = V ( ρ ) + -------- ---------------
- . The
2m ρ 2 2m ρ 2
Schrödinger equation ths becomes

h– 2 d 2 F ( ρ )
– -------- ----------------- + V eff ( ρ )F ( ρ ) = EF ( ρ ) . Eq. 9-26
2m dρ

Notice that we have reduced the problem to an equation of one coordinate variable, ρ .

If we label the radial functions with n, then for every value of M , there is a set of orthogonal functions and energies,
P n M ( ρ ) and E n M . The complete wavefunction is of the form

Ψ ( ρ, φ ) = P n M ( ρ )e iMφ . Eq. 9-27

For M = 0 , the set of energy levels E n0 is twofold degenerate.

Every wavefunction can be used to calculate a probabilty density. In this case, the probability density is

Ψ * ( ρ, φ )Ψ ( ρ, φ )dV = Ψ * ( ρ, φ )Ψ ( ρ, φ )dρdφ . Eq. 9-28

Using the wavefunctions from above

Ψ * ( ρ, φ )Ψ ( ρ, φ )dV = P n* M ( ρ )P n M ( ρ ) ( e iMφ ) * e iMφ ρdρdφ Eq. 9-29

F( ρ)
or using the fact that Ψ ( ρ, φ ) = ----------- e i M φ ,
ρ

Ψ * ( ρ, φ )Ψ ( ρ, φ )dV = F n* M ( ρ )F n M ( ρ ) ( e iMφ ) * e iMφ dρdφ . Eq. 9-30

If the angular functions e iMφ are used, the probability density, much like the angular momentum, is independent of angle. This is
because ( e iMφ ) * e iMφ = e –i Mφ e iMφ = 1 , which eliminates all functions of φ . If, on the other hand, the angular functions
e iMφ ± e –iMφ = λ cos Mφ or λ sin Mφ are used instead ( λ being a constant), the probability density is a function of
cos2 Mφ or sin2 Mφ . This is the general problematic behavior of degenerate functions. Note that the closed shell corresponding
to filling both cos Mφ and sin Mφ is also independent of angle.

82
Cylindrical Polar Coordinates: Particle on a Ring

h– 2 ∂ 2 P ( ρ ) 1 ∂P ( ρ )  1 h– 2 1 2
Let us consider the solutions Ψ ( ρ, φ ) = P n M ( ρ )e iMφ to – --------  ----------------
- + --
- --------------
- ----------- + V ( ρ ) + -------
- ----- M = E with n
2m  ∂ρ 2 ρ ∂ρ  P ( ρ ) 2m ρ 2
h– 2 1
constant. The term -------- -----2 M 2 (initially introduced as Eq. 9-9) is the centrifugal energy of a particle of mass m with angular
2m ρ
momentum Mh– (as shown in Eq. 9-21), at a distance ρ from the origin. If the radial wavefunction P n M ( ρ ) is sharply
peaked at the value ρ = ρ 0 , then the solutions P n M ( ρ ) will be relatively independent of M. The energy levels for the
particle on a ring are therefore:

h– 2 1 2
E M = -------- -----M , with M = 0, ± 1, ± 2, … . Eq. 9-31
2m ρ 02

Aromatic Systems. A good example of a particle in a ring as an electron occupying one of the delocalized π system of an aro-
matic molecule. Using Eq. 9-31, it is possible to estimate the energy of optical transitions in such aromatic systems. Table
9-1 contains some sample data.
Table 9-1 - Examples of aromatic systems.

Number of M of the M of the


System π electrons ρ0 HOMO LUMO
C 3 H 3+ 2 L CC 0 1
-------
3
C6 H 6 6 L CC 1 2

C 10 H 8 10 ≈ 2L CC 2 3

Just as in the case of the linear polyene, the energy of the first optical transition is E ( LUMO ) – E ( HOMO ) .

83
LECTURE 10 Spherical Polar Coordinates

Spherical polar coordinates are best suited for problems involving spherically symmet-
ric central potentials. The coordinates are defined as follows:

x = r sin θ cos φ
y = r sin θ sin φ . Eq. 10-1
z = r cos θ

As in the case of cylindrical polar coordinates, φ is the polar angle ( 0 ≥ φ ≥ 2π ). The second
π
angle, θ , is the azimuthal angle ( 0 ≥ θ ≥ --- ), and r is the radius from the origin ( r ≥ 0 ). The
2
volume differential dV is

dV = r 2 sin θdrdθdφ . Eq. 10-2

It is useful to be able to convert between the spherical polar and cylindrical polar coor-
dinates. The two sets are related as follows:

z = r cos θ
ρ = r sin θ . Eq. 10-3
φ=φ

Motion in a Central Potential


A central potential is most often encountered in problems where one body orbits
another. Below, we give two examples, and show that the best way of treating such prob-
lems is by using spherical polar coordinates.

The Planetary Model of the Atom. This is the atomic model where the electron is taken as
orbiting the nucleus in concentric spherical orbits. If M is the mass of the nucleus and
m e the electron’s mass, the reduced mass of the whole system is

85
Spherical Polar Coordinates: Motion in a Central Potential

Mm e me
µ = ------------------ ≈ m e 1 – ------  . Eq. 10-4
M + me  M

In such a case, the potential is always of the form V = V ( r ) , where r is the distance between the electron and the
nucleus.

Diatomic Molecule. For a diatomic molecule, the two nuclei orbit each other. They can be labeled as M a and M b . The
reduced mass is

MaMb
µ = -------------------- . Eq. 10-5
Ma + Mb

In such a case, r is the internuclear distance, and the potential is again of the form V = V ( r ) .

Center of Mass Motion. Both, the atom and the diatomic molecule are unlikely to remain fixed in space. The center of mass of
either the nucleus and electron or the diatomic molecule may very well be moving in a specified direction. That motion, how-
ever, is completely independent of the the internal potential experienced by the two particles. If the motion in relation with
one another is confined to a large box, then the states of the system are bound. The motion can also be completely free. In
either case, the internal potential remains unaffected.

Relative Motion. Because the center of mass motion does not figure into the two particle interaction, it is possible to separate
and neglect it. What is important in the problem is the relative motion of the two bodies, and in particular, the distance r
between them, which can be expressed using Cartesian coordinates:

r = ( xa – xb )2 + ( ya – yb ) 2 + ( za – zb ) 2 . Eq. 10-6

We can easily write the Schrödinger equation, since the potential due to either particle is spherically symmetric:

h– 2 ∂ 2 Ψ ∂ 2 Ψ ∂ 2 Ψ
– ------  ---------2- + ---------2- + ---------2-  + V ( ( x a – x b ) 2 + ( y a – y b ) 2 + ( z a – z b ) 2 )Ψ ( x, y, z ) = EΨ ( x, y, z ) , Eq. 10-7
2µ  ∂x ∂y ∂z 

where we define x = x a – x b , y = y a – y b , and z = z a – z b . Note that despite its daunting form in Cartesian coordinates,
the potential is central and depends only on one variable:

V ( ( xa – xb ) 2 + ( ya – yb ) 2 + ( z a – zb ) 2 ) = V ( r ) . Eq. 10-8

86
Spherical Polar Coordinates: Operators, Eigenvalues and Eigenfunctions

The Schrödinger equation can therefore be expressed entirely in spherical polar coordinates. The kinetic part of the Hamilto-
nian can be shown to have the following form (we use the cylindrical polar equivalent as the first step):

h– 2 ∂ 2 Ψ ∂ 2 Ψ ∂ 2 Ψ h– 2 ∂ 2 Ψ 1 ∂Ψ 1 ∂ 2 Ψ ∂ 2 Ψ
– ------  ---------2- + ---------2- + ---------2- = – ------  ---------2- + --- -------- + -----2 ---------2- + ---------2- . Eq. 10-9
2µ  ∂x ∂y ∂z  2µ  ∂ρ ρ ∂ρ ρ ∂φ ∂z 
–2 1 ∂ ∂Ψ 1 ∂ ∂Ψ 1 ∂2Ψ
= – ------ ----2 -----  r 2 -------- + --------------- - ------ sin θ -------- + -----------------
h
- ----------
2µ r ∂r  ∂r  r sin θ ∂θ
2  ∂θ  r sin2 θ ∂φ 2
2

Thus, given a central potential, the complete Schrödinger equation reads

h– 2 1 ∂ ∂Ψ 1 ∂ ∂Ψ 1 ∂2Ψ
– ------ ----2 ----- r 2 --------  + ---------------
- ------ sin θ --------  + -----------------
- ---------- + V ( r )Ψ = EΨ , where Ψ = Ψ ( r, θ, φ ) . Eq. 10-10
2µ r ∂r ∂r r sin θ ∂θ 
2 ∂θ  r sin2 θ ∂φ 2
2

Operators, Eigenvalues and Eigenfunctions


We have already encountered operators, eigenvalues and eigenfunctions, but it may be useful to formalize their relation-
ship. In classical mechanics, the Hamiltonian was simply the sum of all the energy functions:

p r2 L2
------ + -----------2 + V ( r ) = E Eq. 10-11
2µ 2µr

In quantum mechanics, the Hamiltonian is an operator that operates on wavefunctions:

h– 2 2
– -------
- ∇ + V Ψ = EΨ . Eq. 10-12
 2m 

The operator itself,

h– 2
Ĥ = – -------- ∇2 + V . Eq. 10-13
2m

Thus, the Schrödinger equation, in its most succinct form is

ĤΨ = EΨ . Eq. 10-14

87
Spherical Polar Coordinates: Operators, Eigenvalues and Eigenfunctions

If Ψ is a solution to the Schrödinger equation, it is called an eigenfunction. What we understand by “solution” is that if a func-
tion is subjected to the operator, then the result is a multiple of the same function. The number by which the function is multi-
plied is called the eigenvalue. In case of the Schrödinger equation, the total energy of the system is the eigenvalue.


An important operator that we have already encountered is the angular momentum operator, L̂ z = – ih– ------ . In Eq. 9-
∂φ
20 on page 80, we saw that in the case of two dimensions and a central potential V = V ( r ) ,

– ih– ------ Ψ = L̂ z Ψ = L z Ψ = Mh– Ψ , where M is an integer. Here, Ψ is the eigenfunction and L z is the eigenvalue (note
∂φ
that without the “hat,” L z represents a number, not an operator).

The two dimensional example can be generalized into three dimensions. Convince yourself that

Lˆ2 = Lˆx2 + Lˆy2 + Lˆz2 , where Eq. 10-15

Lˆ2 Ψ = L 2 Ψ . Eq. 10-16

It can be shown that in spherical polar coordinates*

1 ∂ ∂ 1 ∂2
Lˆ2 = – h– 2 ----------- ------  sin θ ------ + ------------
- -------- . Eq. 10-17
sin θ ∂θ  ∂θ sin2 θ ∂φ 2
∂2 ∂ 1 ∂2 
= – h– 2  --------2 + cot θ ------ + ------------
- --------
 ∂θ ∂θ sin2 θ ∂φ 2

We will later prove that if Ψ is an eigenfunction of angular momentum, then Lˆ2 Ψ = L 2 Ψ , with

L 2 = l ( l + 1 )h– 2, where l = 0, 1, 2, … . Eq. 10-18

Also, since all the magnitudes of the angular momentum components are positive terms

L 2 = L x2 + L y2 + L z2 > L z2 . Eq. 10-19

This means that

*. See Levine, Section 5.3.

88
Spherical Polar Coordinates: The Schrödinger Equation

l ( l + 1 )h– 2 > ( Mh– ) 2 and l ( l + 1 ) > M 2 . Eq. 10-20

The Schrödinger Equation


The spherical polar version of the Schrödinger equation presented in Eq. 10-10 may seem a little daunting. How-
ever, a slight rearrangement and an application of our knowledge of angular momentum simplify things a great deal.

 h– 2 1 ∂  2 ∂  h– 2 1 1 ∂  ∂ 1 ∂2 
 – ------ ----2 ----- r ----- – ------ ----2 ----------- ------ sin θ ------  + ------------
- -------- Ψ + V ( r )Ψ = EΨ
sin2 θ ∂φ 2 
Eq. 10-21
   
 2µ r ∂r ∂r 2µ r sin θ ∂θ ∂θ

Substituting in Eq. 10-17,

h– 2 1 ∂ ∂ Lˆ2
– ------ ----2 ----- r 2 -----  + -----------2 + V ( r ) Ψ = EΨ . Eq. 10-22
2µ r ∂r ∂r   2µr

Φ
A further simplification can be made if we substitute Ψ = ---- , or rΨ = Φ .
r

h– 2 ∂ 2 Lˆ2
– ------ -------2 + -----------2 + V ( r ) Φ = EΦ Eq. 10-23
2µ ∂r 2µr

or using the fact that Lˆ2 Ψ = L 2 Ψ ,

h– 2 ∂ 2 l ( l + 1 )h– 2
– ------ -------2 + ----------------------- + V ( r ) Φ = EΦ . Eq. 10-24
2µ ∂r 2µr 2

To solve this Schrödinger equation, we try to achieve a separation of variables, where Φ is separable into radial and angu-
lar parts, Φ = R ( r )Q ( θ, φ ) . Furthermore, since the dependence on φ has to be periodic (see Eq. 9-8 on page 79), we
can specify that the function has to be of the form

Φ = R ( r )P ( θ )e imφ . Eq. 10-25

If we try using this function in Eq. 10-24, a nice simplification takes place.

89
Spherical Polar Coordinates: The Schrödinger Equation

h– 2 ∂ 2 l ( l + 1 )h– 2
- + V ( r ) R ( r )P ( θ )e imφ = ER ( r )P ( θ )e imφ
– ------ -------2 + ---------------------- Eq. 10-26
2µ ∂r 2µr 2

If we multiply all the terms out, we see can separate two diffrential equations, each one operating on different variables. Since
the kinetic operator affects only functions of r, the radial equation can be stated as

h– 2 ∂ 2 l ( l + 1 )h– 2
– ------ -------2 + ----------------------- + V ( r ) R nl ( r ) = E nl R nl ( r ) . Eq. 10-27
2µ ∂r 2µr 2

Correspondingly, the angular momentum operator affects only the angular functions. Thus, the angular momentum
equation can be written as

∂2 ∂ 1 ∂2 
– h– 2  --------2 + cot θ ------ + ------------
- -------- P ( θ )e imφ = L 2 P lm ( θ )e imφ . Eq. 10-28
 ∂θ ∂θ sin2 θ ∂φ 2 lm

= l ( l + 1 )h– 2 P lm ( θ )e imφ

Simplifying to eliminate derivatives in φ gives

∂ ∂2
– h– 2  cot θ ------ + --------2 – ------------
m2 
- P ( θ )e imφ = l ( l + 1 )h– 2 P lm ( θ )e imφ . Eq. 10-29
 ∂θ ∂θ sin2 θ  lm

By cancelling the exponential, we arrive at the final form of the orbital angular momentum, a general equation for all central
force problems:

∂ ∂2 m2 
– h– 2  cot θ ------ + --------2 – ------------
- P ( θ ) = l ( l + 1 )h– 2 P lm ( θ ) Eq. 10-30
 ∂θ ∂θ sin2 θ  lm

Eq. 10-27 is the second equation we need to solve. It depends upon V ( r ) and the value of l ( l + 1 ) . Note that the
form of the equation is that of a one dimensional Schrödinger equation with an effective potential for radial motion

l ( l + 1 )h– 2
V eff ( r ) = V ( r ) + ----------------------- . Eq. 10-31
2µr 2

If V ( r ) has a deep, sharp minimum and we are in a low energy state, we can approximate the potential by assuming that r is
never very different from the equilibrium position r 0 . Thus, Eq. 10-27 can be rewritten with r = r 0 as

90
Spherical Polar Coordinates: The Schrödinger Equation

h– 2 ∂ 2 l ( l + 1 )h– 2
– ------ -------2 + ----------------------- + V ( r ) R nl ( r ) = E nl R nl ( r ) . Eq. 10-32
2µ ∂r 2µr 02

l ( l + 1 )h– 2
The term ----------------------- is the rotational energy of the system, and since it is a constant, it can be moved to the right side of the
2µr 02
equation.

h– 2 ∂ 2 l ( l + 1 )h– 2
– ------ -------2 + V ( r ) R nl ( r ) = E nl – ----------------------- R nl ( r ) Eq. 10-33
2µ ∂r 2µr 02

The original equation’s eigenvalue, E nl , was the rotational and vibrational energy of the system. Eq. 10-33 as written
above gives the vibrational energy only (since the rotational is being subtracted explicitly).

91
LECTURE 11 Hydrogenic Orbitals

In the previous section, we began solving the Schrödinger equation in spherical


polar coordinates given a central potential. We separated the variables into two differ-
ential equations, one involving the radial coordinate ( r ), and another, which treated the
angular variables ( φ, θ ). We now proceed to develop the functional solutions for the
orbitals of a hydrogen atom. We derive the angular solutions first and then follow with
the radial ones. We will make no attempt to develop the solutions a priori. With the
functional forms of the solutions given at the outset, we try to explain why they work.

Angular Solutions
The solution to the angular equation is a function of the two angular variables,
most often written as Y lm ( θ, φ ) , where m and l are quantum numbers as defined previ-
ously ( l = 0, 1, 2, … and m = 0, ± 1, ± 2, … ).

As shown in Eq. 10-28 on page 90, the angular equation results from applying the
angular momentum operator (Eq. 10-17)

∂2 ∂ 1 ∂2 
Lˆ2 = – h– 2  --------2 + cot θ ------ + ------------
- -------- Eq. 11-1
 ∂θ ∂θ sin2 θ ∂φ 2 

to the angular function:

∂2 ∂ 1 ∂2  m
Lˆ2 Y lm ( θ, φ ) = – h– 2  --------2 + cot θ ------ + ------------
- -------- Y ( θ, φ ) = λY lm ( θ, φ ) . Eq. 11-2
∂θ ∂θ sin2 θ ∂φ 2  l

In Eq. 10-25, we showed that the dependence on φ must be of the form e imφ . We
will now show that the part of the equation dependant on θ is of the form
sin m θ P lm ( cos θ ) . The symbols mean that the solution is a product of the m -th power
of the sine function and an m-th degree polynomial of the cosine function. The com-
plete angular function is therefore of the form

Y lm ( θ, φ ) = sin m θ P lm ( cos θ ) ⋅ e imφ . Eq. 11-3

93
Hydrogenic Orbitals: Angular Solutions

We can eliminate φ from Eq. 11-2 by performing the differentiation with respect to φ . This yields

∂2 ∂ m2
– h– 2  --------2 + cot θ ------  sin m θ P lm ( cos θ ) – ------------
- sin m θ P lm ( cos θ ) = λ sin m θ P lm ( cos θ ) . Eq. 11-4
 ∂θ ∂θ  sin2 θ

Note that this equation has singularities at θ = 0, π . This explains the need for the sine term in the solution. Rewriting
the angular function as

Y lm ( θ, φ ) = e imφ ⋅ sinm′ θ ∑ a n cosn θ Eq. 11-5

shows clearly that the solution vanishes at θ = 0, π ; the meaning of the polynomial in cosine is also specified more
clearly.

We can readily show that m′ = m . For clarity, let us perform one differentiation at a time.


------ sinm′ θ ∑ a n cosn θ = m sinm′ – 1 θ cos θ ∑ a n cosn θ + sinm′ θ ∑ – na n cosn – 1 θ sin θ Eq. 11-6
∂θ

∂2
--------2 sinm′ θ ∑ a n cosn θ = [ – m′ sinm′ θ + m′ ( m′ – 1 ) cos2 θ sinm′ – 2 θ ] ∑ a n cosn θ Eq. 11-7
∂θ
– 2m′ cos θ sinm′ – 1 θ ∑ na n cosn – 1 θ sin θ

+ sinm′ θ [ ∑ n ( n – 1 )a n cosn – 2 θ sin2 θ – ∑ na n cosn θ ]

Collecting terms and using the relation sin2 θ + cos2 θ = 1 , turns Eq. 11-4 into


λ sin m θ ∑ a n cosn θ = – h– 2  [ – m′ sinm′ θ + m′ ( m′ – 1 ) sinm′ – 2 θ – m′ ( m′ – 1 ) sinm′ θ ] ∑ a n cosn θ Eq. 11-8

– 2m′ cos θ sinm′ – 1 θ ∑ na n cosn – 1 θ sin θ + sinm′ θ [ ∑ n ( n – 1 )a n cosn – 2 θ sin2 θ – ∑ na n cosn θ ]


+ cot θ [ m′ sinm′ – 1 θ cos θ ∑ a n cosn θ + sinm′ θ ∑ – na n cosn – 1 θ sin θ ] – m 2 sinm′ – 2 θ ∑ a n cosn θ 

Further simplifications give

94
Hydrogenic Orbitals: Angular Solutions


λ sin m θ ∑ a n cosn θ = – h– 2  [ – m′ sinm′ θ + m′ ( m′ – 1 ) sinm′ – 2 θ – m′ ( m′ – 1 ) sinm′ θ ] ∑ a n cosn θ Eq. 11-9

– 2m′ cos θ sinm′ – 1 θ ∑ na n cosn – 1 θ sin θ + sinm′ θ [ ∑ n ( n – 1 )a n cosn – 2 θ sin2 θ – ∑ na n cosn θ ]

+ [ m′ sinm′ – 2 θ ( 1 – sin2 θ ) ∑ a n cosn θ + sinm′ – 1 θ cos θ ∑ – na n cosn – 1 θ sin θ ]


– m 2 sinm′ – 2 θ ∑ a n cosn θ 

and then


λ sin m θ ∑ a n cosn θ = – h– 2  [ – m′ sinm′ θ + m′ 2 sinm′ – 2 θ – m′ ( m′ – 1 ) sinm′ θ ] ∑ a n cosn θ Eq. 11-10

– 2m′ cos θ sinm′ – 1 θ ∑ na n cosn – 1 θ sin θ + sinm′ θ [ ∑ n ( n – 1 )a n cosn – 2 θ sin2 θ – ∑ na n cosn θ ]

+ [ – m ′ sinm′ θ ∑ a n cosn θ + sinm′ – 1 θ cos θ ∑ – na n cosn – 1 θ sin θ ]


– m 2 sinm′ – 2 θ ∑ a n cosn θ 

The terms involving sinm′ – 1 θ multiply out to become terms in sinm′ θ . The differential equation allows for only one
power of sin θ . Eq. 11-10 has two powers of sin θ , so we want to remove the lowest power one. The lowest order of
sin θ is present in terms involving sinm′ – 2 θ . We note that if we collect like terms, the coefficient of sinm′ – 2 θ is m′ 2 – m 2 .
The only way terms in sinm′ – 2 θ will cancel out is if m′ = m .

We can therefore rewrite Eq. 11-10 in a simplified form and finally arrive at the exact solutions to the angular equa-
tion. The equation reads

95
Hydrogenic Orbitals: Angular Solutions


λ sin m θ ∑ a n cosn θ = – h– 2  ( – m ) ( m + 1 ) sin m θ ∑ a n cosn θ Eq. 11-11


– 2 ( m + 1 ) sin m θ ∑ na n cosn θ + sin m θ [ n ( n – 1 ) ∑ a n cosn – 2 θ ( 1 – cos2 θ ) ] 

or without the sine terms


λ ∑ a n cosn θ = – h– 2  ( – m ) ( m + 1 ) ∑ a n cosn θ . Eq. 11-12


– 2 ( m + 1 ) ∑ na n cosn θ + ∑ n ( n – 1 )a n cosn – 2 θ – ∑ n ( n – 1 )a n cosn θ 

Rearranging slightly, we can separate the powers of the cosine term:

λ
∑ ( an cosn θ ) m ( m + 1 ) + 2n ( m + 1 ) + n ( n – 1 ) – ----
- =
h– 2 ∑ n ( n – 1 )an cosn – 2 θ . Eq. 11-13

The equation is similar to the equation we encountered when solving the harmonic oscillator (Eq. 5-15 on page 62). We
also use a similar argument – all the powers of cosine must cancel – to write

λ
∑ ( an cosn θ ) m ( m + 1 ) + 2n ( m + 1 ) + n ( n – 1 ) – ----
- =
h– 2 ∑ ( n + 2 ) ( n + 1 )an + 2 cosn θ . Eq. 11-14

If we cancel the cosine terms on either side, we obtain a polynomial relation:

λ
m ( m + 1 ) + 2n ( m + 1 ) + n ( n – 1 ) – ----
- a = ( n + 2 ) ( n + 1 )a n + 2 . Eq. 11-15
h– 2 n

If we choose the k-th order solution, the polynomial will have only k terms, and a k + 2 = 0 , since a non-zero a k + 2
would imply a ( k + 2 )-th order polynomial. The allowed values of m are specifically given by the equation

λ
m ( m + 1 ) + 2k ( m + 1 ) + k ( k – 1 ) – ----
- = 0, Eq. 11-16
h– 2

96
Hydrogenic Orbitals: Angular Solutions

or more succinctly,

λ
m ( m + 1 ) + 2 m k + k ( k + 1 ) – ----
- = 0. Eq. 11-17
h– 2

ak + 2 k(k + 1)
Note that k has to be finite, because otherwise the series above will diverge ( lim ----------- = ---------------------------------- = 1 ).
k → ∞ ak (k + 2)(k + 1)

To simplify matters, we can let l = m + k . The product

l(l + 1) = ( m + k)( m + k + 1) = m ( m + 1) + 2 m k + k(k + 1) , Eq. 11-18

when substituted into Eq. 11-17 gives

λ = l ( l + 1 )h– 2 . Eq. 11-19

λ is the eigenvalue of the angular momentum equation (Eq. 11-2). In other words, Eq. 11-19 specifies the allowed values
of the total angular momentum, L 2 , for a given set of quantum numbers m and l. We can obtain a recursion relation sim-
λ
ilar to that in Eq. 5-20 on page 62 by eliminating ----
- from Eq. 11-15 and solving for a n + 2 .
h– 2

m ( m + 1 ) + 2n ( m + 1 ) + n ( n – 1 ) – l ( l + 1 )
a n + 2 = ---------------------------------------------------------------------------------------------------------------------- a n Eq. 11-20
(n + 2)(n + 1)

If we know the coefficient of the n-th degree solution, we can use this relation to generate the ( n + 2 )-th solution. Note
that a n is just a coefficient of the cosine term of the n-th degree. Thus, the series are either odd or even with respect to

Table 11.1 - The first nine solutions to the radial equation; in chemical terms, these are the radial equations for the s, p, and d orbitals.

l m Y lm ( θ, φ ) l m Y lm ( θ, φ ) l m Y lm ( θ, φ )
0 0 1 1 –1 sin θ e –iφ 2 0 ( 3 cos2 θ – 1 )/2
1 1 sin θ e iφ 2 2 sin2 θ e i2φ 2 –1 sin θ cos θ e –i φ
1 0 cos θ 2 1 sin θ cos θ e iφ 2 –2 sin2 θ e –i 2φ

97
Hydrogenic Orbitals: Angular Solutions

cos θ . The solution functions themselves are easily generated, following the form specified in Eq. 11-5,
Y lm ( θ, φ ) = e imφ ⋅ sinm θ ∑ a n cosn θ .

Polar plots of the functions listed in Table 11.1 clearly show that the radial functions have specific directionality if plotted
in a Cartesian coordinate system (see Karplus and Porter, pp. 132-135). It is customary to assign the Cartesian polariza-
tions to these functions. These are summarized in Table 11.2 below.

Table 11.2 - The Cartesian polarizations of the s, p, and d orbitals.

Polarization
l = 0 1
x
l = 1 y
z
xy
x2 – y2
l = 2 xz
yz
3z 2 – 1

Symmetry of Angular Functions


The angular functions of the form Y lm ( θ, φ ) = sin m θ P lm ( cos θ ) ⋅ e imφ (as shown in Eq. 11-3) have distinct inversion
properties depending on the quantutm number l. To see these, let us recall the relationships between the Cartesian and
sherical polar coordinates (Eq. 10-1), namely x = r sin θ cos φ , y = r sin θ sin φ , z = r cos θ , where 0 ≤ θ ≤ π and
0 ≤ φ ≤ 2π . The inversion operation maps x → –x , y → – y , and z → – z .

In terms of spherical polar coordinates, the z inversion can be expressed as θ → π – θ , since cos ( π – θ ) = – cos θ
and sin ( π – θ ) = sin θ . The x and y inversions, taken together, give φ → π + φ , since cos ( π + φ ) = – cos ( φ ) and
sin ( π + φ ) = – sin φ . If we apply the effects of all these transformations to the general angular equation we see that

98
Hydrogenic Orbitals: Radial Solutions

Y lm ( π – θ, π + φ ) = e imπ ⋅ e imφ ⋅ sin θ m P lm ( – cos θ ) . Eq. 11-21


= ( – 1 ) m ⋅ ( – 1 ) l – m ⋅ Y lm ( θ, φ )
= ( – 1 ) l ⋅ Y lm ( θ, φ )

Thus, the symmetry of the angular solutions under inversion depends only on the quantum number l and is independent
of m. If l is odd, the overall function is odd under inversion, in l is even, the function is even.

Radial Solutions
Given a central potential such as in the case of a hydrogen atom and a wavefunction with separable radial and
angular components such as Φ = rψ , the radial equation is

h– 2 ∂ 2 Φ
– ------ ---------2- + V eff ( r )Φ = EΦ , Eq. 11-22
2µ ∂r

l ( l + 1 )h– 2
where V eff ( r ) = V ( r ) + ----------------------- as developed in Eq. 10-31. Note that the effective potential is an equation in r alone. If
2µr 2
we write the wave function Φ = R ( r )Y lm ( θ, φ ) , the radial equation simplifies to

h– 2
– ------R′′ + V eff ( r )R = ER . Eq. 11-23

e c2
We will take the potential in the hydrogen atom to be the simple Coulombic interaction V ( r ) = – ---- *, where e c is the
r
electronic charge. Substituting in all the parts of the effective potential and rearranging slightly, Eq. 11-23 becomes

2µ 2µ e c2 l ( l + 1 )
R′′ + -----
– 2-E + ----- - R = 0.
- ---- – ---------------- Eq. 11-24
h h– 2 r r2

*. If we use atomic units, we can avoid clutter such as the 1/4πε 0 term in Coulomb’s law.

99
Hydrogenic Orbitals: Radial Solutions

For the bound states of the hydrogen atom, that is where the electrons remain around the nucleus , E < 0 . We can prove
this somewhat counterintuitive fact with the virial theorem. Recall from Eq. 7-6 that 2T = nV , where n depends on the
functional form of the potential. For a central potential, n = – 1 and

e c2 1
E = T + V = --- V = – ----  ---  .
1
Eq. 11-25
2 2 r

Since r is never negative, the total energy for a bound state always is.

In order to simplify the ensuing equations, let us state

2µ 2µe c2
α 2 = – -----
– 2- E and λ = -----------
-. Eq. 11-26
h h– 2

This turns Eq. 11-24 into

λ l(l + 1)
R′′ + – α 2 + --- – ----------------
- R = 0. Eq. 11-27
r r2

For any realistic atomic (or molecular) potential, V ( r ) → 0 as r → ∞ , so in its asymptotic form Eq. 11-27 becomes

R′′ – α 2 R = 0 , Eq. 11-28

where α is as defined above, and E is again negative in the bound states. The solution to the differential equation in Eq.
11-28 can be obtained almost by inspection alone. It is the general asymptotic form for the electronic wave equation in
atoms, ions and molecules,

R = e –αr . Eq. 11-29

There is just one problem with this solution. If we apply it to Eq. 11-27, a singularity occurs at r = 0 . In order to avoid
this problem, let

R = r s e –αr F , Eq. 11-30

where F is a function, s is a constant, and the r s term ensures that R → 0 and r → 0 . The condition from Eq. 11-28 still
has to be met, so we need to find the second derivative of this new R.

100
Hydrogenic Orbitals: Radial Solutions

R′ = sr s – 1 e –αr F – αr s e –αr F + r s e –αr F′ ,

R′′ = s ( s – 1 )r s – 2 e –αr F – 2sαr s – 1 e –αr F + α 2 r s e –αr F + 2sr s – 1 e –αr F′ – 2αr s e –αr F′ + r s e –αr F′′ .

Substituting this rather lengthy expression into Eq. 11-27 gives the differential equation of interest:

s ( s – 1 )r s – 2 e –αr F – 2sαr s – 1 e –αr F + α 2 r s e –αr F + 2sr s – 1 e –αr F′ – 2αr s e –αr F′ + r s e –αr F′′ + λr s – 1 e –αr F – l ( l + 1 )r s – 2 e –αr F = 0 .

Cancelling out the r s – 2 e –αr term gives a simpler

s ( s – 1 )F – 2sαrF + α 2 r 2 F + 2srF′ – 2αr 2 F′ + r 2 F′′ + λrF – l ( l + 1 )F = 0 . Eq. 11-31

At r = 0 , the function becomes

s ( s – 1 )F – l ( l + 1 )F = 0 or s = l + 1 . Eq. 11-32

Thus,

– 2sαrF + 2srF′ – 2αr 2 F′ + r 2 F′′ + λrF = 0 or

rF′′ + 2 ( l + 1 – αr )F′ + [ λ – 2 ( l + 1 )α ]F = 0 . Eq. 11-33

The function F can be expanded using a MacLaurin series, or

F= ∑ an r n
F′ = ∑ nan r n – 1 Eq. 11-34

F′′ = ∑ n ( n – 1 )an r n – 2
Substituting into Eq. 11-33, we get

∑ n ( n – 1 )a n r n – 1 + 2 ( l + 1 – αr ) ∑ nan r n – 1 + [ λ – 2 ( l + 1 )α ] ∑ an r n = 0, Eq. 11-35

and collecting all the terms in r of order n,

[ n ( n + 1 ) + 2 ( n + 1 ) ( l + 1 ) ]a n + 1 – [ 2nα – λ + 2 ( l + 1 )α ]a n = 0 . Eq. 11-36

101
Hydrogenic Orbitals: Radial Solutions

This allows us to write a recursion formula for the coefficient a n + 1 :

λ
2 ( l + 1 + n )α – ---
2
an + 1 = -----------------------------------------------------------------a n . Eq. 11-37
n( n + 1) + 2( l + 1 )( n + 1)

a n + 1 2α
Since for large n, ----------- ≈ ------- , the series is like e αr .
an n

The series will diverge at r = ∞ unless a k + 1 = 0 with a k ≠ 0 , which retains convergence and solutions of a poly-
nomial form. If a k + 1 = 0 , then

λ
( l + 1 + k )α = --- . Eq. 11-38
2

2µ 2µe c2
Recalling the definitions α 2 = – ------
– 2 E and λ = -----------
- from Eq. 11-26, we can now write that
h h– 2

 µe c-  or
2 2

– ( l + 1 + k ) 2 -----
– 2- E =  -------
–2  Eq. 11-39
h h

µe c4 µe c4
E = – ----------------------------------- -,
-2 = – -------------- Eq. 11-40
2h ( l + 1 + k )
– 2 2h– 2 N 2

where N = l + k + 1 .

We can now write a complete solution to the radial equation. The solutions are

k
R nl ( r ) = e –αr r l + 1 ∑ a n r n . Eq. 11-41
0

Note that the highest order term is r k + l + 1 = r N .

The lowest energy state, where k = l = 0 , is

102
Hydrogenic Orbitals: Radial Solutions

R ( r ) = e –α r r . Eq. 11-42

We can use the first derivative to find the maximum.

dR ( r )
-------------- = e –αr – αe –αr r ,
dr

dR ( r ) 1
and -------------- = 0 at r = --- .
dr α

µe c2 * 1
Recall that α 0 = -------- . ----- = a 0 , which is called the Bohr radius and is numerically equivalent to 0.53 × 10 –8 cm (taking
h– 2 α 0
h– 2 α 2 µe c4 e c2
µ ≈ m e ). The energy of the first Bohr orbital is also readily calculated. E = – ------------ = – --------
– -2 = – ----- = – 13.61eV .
2µ 2h a0

In general, the energy is

µe c4 µe c4
E = – --------------------------------
– 2 ( + 1 + )-2 = – ----------- – 2 -2 . Eq. 11-43
h l k h N

Note that for each value of N there are N different possible values of l with the same energy. The values of l are from
l = 0 to l = N – 1 . For each value of l there are 2l + 1 different values of m l , with the same magnitude of angular
momentum (and the same energy). The total number of states at a fixed value of N is

N–1

∑ 2l + 1
l=0
= N2 . Eq. 11-44

Normalization of wave functions


To obtain a complete set of specific solutions, the wavefunctions need to be normalized. In spherical polar coordi-
nates, the volume element dV = r 2 sin θ drdθdφ . The complete set of wavefuntions Ψ would then be normalized as

*. Again, remember that all constants must be in atomic units. If you wish to use SI units, you must include the 1/4πε 0 constant.

103
Hydrogenic Orbitals: Radial Solutions

∞ π 2π

∫ Ψ * Ψ dV = ∫ r 2 dr ∫ sin θ dθ ∫ Ψ * ( r, θ, φ )Ψ ( r, θ, φ ) dφ .
0 0 0
Eq. 11-45

Since Φ = rΨ ,

∞ π 2π

∫ Ψ * Ψ dV = ∫ dr ∫ sin θ dθ ∫ Φ * ( r, θ, φ )Φ ( r, θ, φ ) dφ .
0 0 0
Eq. 11-46

4πr 3
The volume of a sphere centered at the nucleus with radius r is ----------- . The nuclear radius is between 10 –13 and
3
10 –12 cm . Using this as the value of r, we note that the density of an s type function ( l = 0 ) compared to a function with
angular momentum l will be ≈ 10 –8l , since the rations of their densities scale as ( r l ) 2 .

Construction of radial wave functions for the H atom


Let us now use Eq. 11-41 and the recursion formula from Eq. 11-37 to actually write out several of the hydrogenic
radial wave functions. All the equations we need are summarized below:

2µ – 2µE 2µe c2 µe c2 1 λ λ
α 2 = – ------
– 2 E , α = -----------------
– , λ = ----------- - = ----- = --- , ( l + 1 + k )α = ---
- , ------- Eq. 11-47
h h h– 2 h– 2 a0 2 2

k
R nl ( r ) = e –αr r l + 1 ∑ a n r n Eq. 11-48
0

Recall that k is the index of the highest-degree non-vanishing coefficient.

λ
2 ( l + 1 + n )α – ---
2
an + 1 = -----------------------------------------------------------------a n Eq. 11-49
n( n + 1 ) + 2( l + 1 )( n + 1)

λ
Also recall that l + k + 1 = N , which allows us to write Nα = --- , or that
2

1
α = --------- . Eq. 11-50
Na 0

104
Hydrogenic Orbitals: Radial Solutions

Using these equations, we can restate the recursion relation as

2 [ ( l + 1 + n ) – N ]α
a n + 1 = -----------------------------------------------------------------a n , Eq. 11-51
n( n + 1) + 2( l + 1 )( n + 1)

and to explicitly write out the coefficients for the first several radial wave functions.

Table 11.3 - The first several coefficients of the hydrogenic radial wave functions.

N l k b0 b1 b2 b3
1 0 0 1 — — —
1
2 0 1 1 – -------- — —
2a 0
2 1 0 1 — — —
2 2
3 0 2 1 – -------- ----------- —
3a 0 27a 0
1
3 1 1 1 – -------- — —
6a 0
3 2 0 1 — — —

The coefficients above generate unnormalized radial wave functions Normalization is obtained as described above from
the realation

∞ ∞ 2 ∞ 2
 k  k 
∫0 R dr = ∫0 e –αr r l + 1 ∑0 a n r n dr =
2
∫0 e – 2αr 2l + 2
r  ∑ a n r n  dr = 1 . Eq. 11-52
  0 

105
Hydrogenic Orbitals: Radial Solutions

106
LECTURE 12 Perturbation Theory

So far we have only considered problems with exact solutions. In each instance,
the differential equations eventually yielded an precise eigenfunction. However, most
systems are not so easily soluble. For example, an atom with just two electrons does
not yield exact solutions. Instead, we have to be content with approximating them with
so called perturbation theory.

Consider the insoluble Hamiltonian Ĥ . Suppose we can generate an approximate


one, Ĥ ( 0 ) , that is soluble with the equation Ĥ ( 0 ) Ψ n( 0 ) = E n( 0 ) Ψ n( 0 ) and that the resulting
energy levels are non-degenerate. The real problem is

Ĥ = Ĥ ( 0 ) + Ĥ′ Eq. 12-1

where Ĥ′ is called the perturbation. If the perturbation is small, then the eigenfunc-
tions will presumably be close to the solutions of the unperturbed system. These can be
approximated as

Ψ n = Ψ n( 0 ) + Ψ n( 1 ) + Ψ n( 2 ) + … , Eq. 12-2

where each Ψ n( k ) is a successive correction in the approximation of Ψ n . Analogously,


the energy eigenvalues of the system can be approximated as

E n = E n( 0 ) + E n( 1 ) + E n( 2 ) + … , Eq. 12-3

where each E n( k ) is a successive correction in the approximation of the true eigenvalue.

First and Second Order


The consecutive terms in the eigenvalue approximation become smaller and
smaller and also less and less important. For many purposes, it is acceptable to drop all
of the terms except for the unperturbed eigenvalue and the first order correction. The
value of E n( 1 ) can be computed by averaging Ĥ′ over the unperturbed eigenfunctions,
or

107
Perturbation Theory: First and Second Order

E n( 1 ) = ∫ Ψ n( 0 )* Ĥ′Ψ n( 0 ) dτ . Eq. 12-4

Thus, the first order approximation of the energy eigenvalue becomes

E n ≈ E n( 0 ) + ∫ Ψ n( 0 )* Ĥ′Ψ n( 0 ) dτ . Eq. 12-5

A more accurate result can be obtained by including the second order energy term. This is sometimes useful when
(1)
E turns out to be zero, but usually introduces a much higher level of complexity into the calculations. The second
i

order term is

( ∫ Ψ m( 0 )* Ĥ′Ψ n( 0 ) dτ )
2

E n( 2 ) = ∑
m≠n
-,
-------------------------------------------
E n( 0 ) – E m( 0 )
Eq. 12-6

where the sum is over all the unperturbed states, Ψ m( 0 ) , where m ≠ n . The exact evaluation of the second order term can
be difficult or even impossible and is most frequently performed by computational methods.*

Example 12-1. To demonstrate how perturbation theory is used, we will attempt to approximate the energies of the 2s
and 2p electrons of a lithium atom. The approach is highly approximate, although in this case it yields very good results.

Our unperturbed system will be the Li 2+ ion (with the nuclear charge Z = 3 ). The equation we need to solve is

λ l( l + 1 )
R′′ + – α 2 + --- – ----------------
- R = 0, Eq. 12-7
r r2


which is similar to Eq. 11-27. α 2 = – ------ E remains the same, but
h– 2

2µe c2
λ = Z -----------
-. Eq. 12-8
h– 2

λ
Recall from Eq. 11-38 that ( l + 1 + k )α = Nα = --- . Combining this with Eq. 12-8 and Eq. 11-50, we get
2

*. For a derivation and a more complete discussion of perturbation theory, see Levine, Section 9.2.

108
Perturbation Theory: First and Second Order

2µe c2
Z ----------- -
h– 2 Z
α Z = ---------------- = ----- . Eq. 12-9
N a0

The radial solution for the 1s electron of the Li 2+ cation is just like that of the H atom, except it takes the higher nuclear
charge into account. The function is

3
– ------ r
R 1s ( r ) = re a0
. Eq. 12-10

a0
Its maximum occurs at r = ----- . We can now perturb the system by adding the other two electrons. This addition causes
3
a change in the potential function of the system. To avoid being distracted by the details of inter-electronic repulsions,
we can make the assumption that the potential function can be broken into two parts. The first is inside of the 1s wave-
function maximum, where we assume an absence of shielding, and the second is outside the 1s wavefunction maximum,
where we assume that the Z eff = 1 due to shielding by the two 1s electrons. The potentials thus break down into

 2
 – 3e c
-------
-
aa
, 0 < r < -----
 r 3
V(r) =  . Eq. 12-11
 ec a 2 a

 – ---- , ----- < r < ∞


r 3

The perturbation itself is the change in the potential.* Thus, the operators would be

e c2 e c2
Ĥ ( 0 ) = – ---- , Ĥ′ = V ( r ) – – ----  . Eq. 12-12
r  r

Note that the perturbation turns out to be zero outside of the 1s orbital maximum, but is clearly non-zero inside of it. This
a0
creates a discontinuinty at ----- . However, perturbation theory calls for a continuous function. We will therefore make
3
another adjustment by fitting a function which approximates the two pieces of the perturbation at their respective range
limits. A suitable function to use is
( Z – 1 )e c2
Ĥ′ = – ----------------------e –βr , Eq. 12-13
r

*. Notice, however, that the perturbation to the radial potential does not affect any of the angular components of the eigenfunctions.

109
Perturbation Theory: First and Second Order

where β is a parameter that we can adjust to sculpt the function. This


process is pictured schematically in Figure 12.1. For the case of lith- V(r) = 0
3
ium, Z = 3 and β ≈ ----- . Finally, we assume that the Li 2s and 2p V(r) = -2e
2
c -βr
a0 r e
functions can be approximated with the corresponding hydrogenic
wavefunctions.

Using Eq. 12-4 and Eq. 12-13, we can write a general expression
for the change from the H atom. Ψ α stands for a hydrogenic wave-
function, and since the perturbation affects only the radial part of the 2
V(r) = -2ec

solutions, we can replace Ψ α with R α : r

Figure 12.1 - A schematic picture of the two pieces of the


perturbation (dashed) and the “averaged” function used for

( Z – 1 )e 2

∫ Ψ α* Ĥ′Ψ α dτ ∫0 R α2 – ---------------------  dr = H ′ . integration (solid).


c – βr
= -e Eq. 12-14
r  αα

In this case, we are most concerned with the effect on the electron-containing orbitals, namely 2s and 2p. We
already know the hydrogenic radial functions as

R 2s = N s r 2 – -----  exp – --------  and


r r
Eq. 12-15
 a 0  2a 0 

r2
R 2p = N p ----- exp – --------  .
r
Eq. 12-16
a0  2a 0 

1 1/2 1 1/2
N s and N p are normalization constants equal to  ---------3  and  -----------3  respectively. The two orbitals remain orthogonal
 4a 0   4!a 0 
due to the angular terms, which can be shown directly with

∫ Ψ2s* Ĥ′Ψ 2p dτ = 0. Eq. 12-17

′ , is
The perturbation to the 2s orbital, H 2s2s


r 2
′ = N s2 ( – 2e c2 ) ∫ r 2 – -----  exp – -----  e –βr dr .
r
H 2s2s Eq. 12-18
0  a0  a0

110
Perturbation Theory: Two Interacting Levels

3
Recall, however, that β ≈ ----- based on the optimization of the potential. Thus, the above equation can be rewritten (with
a0
a very rough approximation) as


r 2 a 02 – e c2 IP H
′ = N s2 ( – 2e c2 ) ∫ r 2 – -----  exp – -----  dr ≈ N s2 ( – 2e c2 ) ----- = ----------- = – -------- .
4r
H 2s2s Eq. 12-19
0  a 0   a 0  4 16a 0 8

In other words, the perturbation is equivalent to the negative of an eighth of the first hydrogenic ionization potential. A
similar process yields the perturbation to the 2p orbital, again, with a rough approximation:

∞ 3 ∞ 3
a 04 – e c2 IP H
′ = N p2 ( – 2e c2 ) ∫ -----2 exp – -----  e –βr dr = N p2 ( – 2e c2 ) ∫ -----2 exp – -----  dr ≈ N p2 ( – 2e c2 ) ----4-3! = -------------- = – --------- .
r r r 4r
H 2p2p Eq. 12-20
a
0 0  a 0  a
0 0  a 0  4 512a 0 256

13.595
The IP for H at the n = 2 level is – ---------------- = 3.40 eV . Thus, according to our calculations, the energies of a Li 2s and 2p
4
orbitals should be

E 2s = – 3.40 – 1.70 = – 5.10 eV , Eq. 12-21

E 2p = – 3.40 – 0.05 = – 3.45 eV . Eq. 12-22

This compares very well with the experimental values of – 5.390 eV and – 3.542 eV . Note that the approach directly
explains why the s and p orbitals in any species with more than one electron cease to be energetically degenerate.

Two Interacting Levels


Imagine two states or energy levels which interact with each other. Perturbation theory can be used to analyze the
extent of the mixing between the two and the effect that the two mutually exert on each other. Consider an approximate
Hamiltonian H 0 with two orthogonal solutions ψ a and ψ b .

Ĥ 0 ψ a = E 0a ψ a Eq. 12-23

Ĥ 0 ψ b = E 0b ψ b Eq. 12-24

111
Perturbation Theory: Two Interacting Levels

The real problem is again

Ĥ = Ĥ 0 + Ĥ′ . Eq. 12-25

We next choose combinations of the starting functions ψ a and ψ b , such that

ψ + = cos ω ψ a – sin ω ψ b and Eq. 12-26

ψ - = sin ω ψ a + cos ω ψ b . Eq. 12-27

A particular value of ω is chosen so that

Ĥψ + = E + ψ + and Eq. 12-28

Ĥψ - = E - ψ - . Eq. 12-29

The solutions remain orthogonal, so that

∫ ψ+ Ĥψ- dv = ∫ ψ - Ĥψ + dv = 0 . Eq. 12-30

In order to abbreviate the ensuing equations, we define

H aa = ∫ ψ a* Ĥψ a dv , H bb = ∫ ψ b* Ĥψ b dv Eq. 12-31

H ab = ∫ ψ a* H′
ˆ ψ b dv , H ba = ∫ ψ b* H′
ˆ ψ a dv . Eq. 12-32

Note that

H ba = ∫ ψ b* H′
ˆ ψ a dv = [ ∫ ψ a* Ĥψ b dv ] * Eq. 12-33

since

∫ ψ a* Ĥψb dv = ∫ ψ a* ( Ĥ 0 + Ĥ′ )ψ b dv = ∫ ψ a* Ĥ 0 ψ b dv + ∫ ψ a* Ĥ′ψ b dv Eq. 12-34

and according to Eq. 12-24,

112
Perturbation Theory: Two Interacting Levels

∫ ψa* Ĥ0 ψb dv = ∫ ψ a* E 0b ψ b dv = 0 . Eq. 12-35

Furthermore, since Ĥ′ is real,

[ ∫ ψ a* Ĥ′ψ b dv ] * = ∫ ψ b* Ĥ′ψ a dv . Eq. 12-36

Given all of the above, we can obtain the energy eigenvalues of the perturbed system. First, we solve for ω . Substitute
Eq. 12-26 into Eq. 12-28. This yields

Ĥ ( cos ω ψ a – sin ω ψ b ) = E + ( cos ω ψ a – sin ω ψ b ) . Eq. 12-37

After multiplying each side by ψ a* , we can integrate

∫ ψa* Ĥ ( cos ω ψa – sin ω ψ b ) dv = ∫ ψ a* E + ( cos ω ψ a – sin ω ψ b ) dv . Eq. 12-38

Recalling Eq. 12-25, we expand into

∫ ψa* Ĥ cos ω ψa dv – ∫ ψa* Ĥ sin ω ψ b dv = ∫ ψ a* E + cos ω ψ a dv – ∫ ψ a* E + sin ω ψ b dv . Eq. 12-39

cos ωH aa – ∫ ψ a* Hˆ 0 sin ω ψ b dv – ∫ ψ a* Ĥ′ sin ω ψ b dv = E + cos ω – 0 Eq. 12-40

cos ωH aa – 0 – sin ω H ab = E + cos ω . Eq. 12-41

A similar process using Eq. 12-27 and multiplying by ψ b instead of ψ a yields

cos ωH ab – sin ω H bb = – E + cos ω . Eq. 12-42

The same steps are then repeated using Eq. 12-29 to complete the set of four algebraic equations:

sin ω H aa + cos ω H ab = E - sin ω Eq. 12-43

sin ω H ab + cos ω H bb = E - cos ω Eq. 12-44

Eq. 12-41 and Eq. 12-42 can be solved for the energy eigenvalue.

113
Perturbation Theory: Two Interacting Levels

E + = H aa – H ab tan ω and Eq. 12-45

H ab
– E + = ------------- – H bb . Eq. 12-46
tan ω

If we put the above equations together (or the corresponding ones involving E - ) to eliminate the energy term, we get

sin ω cos ω sin2 ω – cos2 ω


H aa – H bb = H ab  tan ω – -------------  = H ab  ------------- – -------------  = H ab  ----------------------------------- =
1
 tan ω   cos ω sin ω   cos ω sin ω 
 
– cos 2ω 1
H ab  --------------------  = – 2H ab ---------------- , Eq. 12-47
1  tan 2ω
 --
2
- sin 2ω 

or solving for ω ,

H ab tan 2ω
----------------------- = ---------------- Eq. 12-48
H bb – H aa 2

1 H ab
ω = --- tan–1 2 -----------------------  . Eq. 12-49
2  H bb – H aa 

We can now use this result to find expressions fot E + and E - . From Eq. 12-45 and Eq. 12-46 we know that

– H ab tan ω = E + – H aa and Eq. 12-50

H ab
– ------------- = E + – H bb . Eq. 12-51
tan ω

We can eliminate the tan ω term from the system to arrive at the following quadratic equation:

2
H ab = ( E + – H aa ) ( E + – H bb ) . Eq. 12-52

This (or its E - analogue) can be easily treated with the quadratic formula to give

114
Perturbation Theory: Two Interacting Levels

H bb + H aa – ( H bb – H aa ) 2 + 4H ab 2
E + = ----------------------------------------------------------------------------------- and Eq. 12-53
2

H bb + H aa + ( H bb – H aa ) 2 + 4H ab 2
E - = ------------------------------------------------------------------------------------ . Eq. 12-54
2

Think of H ab as a measure of how strongly the two levels interact. Depending on the relative magnitude of H ab ,
there are two possible limiting cases for the above solutions. If there is very little of interaction, then

H ab « H bb – H aa Eq. 12-55

and the radical can be expanded to*

H ab 2
( H ab ′ )2
′ – ------------------------------------------------------ .
E + = H aa – ----------------------- = E 0a + H aa Eq. 12-56
H bb – H aa ( E 0b – E 0a ) ( H bb ′ – H aa ′ )

For small ω , tan ω = ω and thus

H ab
ω = ----------------------- , Eq. 12-57
H bb – H aa

which (using Eq. 12-26) makes

H ab
ψ + = ψ a – -----------------------ψ b . Eq. 12-58
H bb – H aa

The same process, repeated for E - gives

2
H ab
E - = H bb + ----------------------- and Eq. 12-59
H bb – H aa

H ab
ψ - = ψ b + -----------------------ψ a . Eq. 12-60
H bb – H aa

2H ab
*. The expansion is of the form 1 + x 2 , where x = ----------------------- .
H bb – H aa

115
Perturbation Theory: Two Interacting Levels

Note that as H ab → 0 (or the level mixing approaches 0), the wave functions tend towards their unperturbed forms, or
ψ + → ψ a and ψ - → ψ b .

The other limiting case is where the two levels mix completely, and

H ab » H bb – H aa . Eq. 12-61

π
Based on Eq. 12-49, we can state that ω → --- , and the energy solutions approach
4

H bb + H aa
E + = ------------------------ – H ab and Eq. 12-62
2

H bb – H aa
E - = ----------------------- + H ab . Eq. 12-63
2

π
For ω = --- , the two original wavefunctions are mixed into two normalized linear combinations:
4

( ψa + ψ b )
ψ + = ----------------------- and Eq. 12-64
2

( ψa – ψb )
ψ - = ----------------------- . Eq. 12-65
2

Note that when E 0a = E 0b , the system is unstable with respect to perturbation. In other words, an infinitesimal perturba-
tion can produce a finite change in the wave function and first order perturbation theory fails, unless the proper wave-
functions (Eq. 12-64 and Eq. 12-65) are used.*

*. For further details on the treatment of many interacting levels, see Levine, Chapter 9.

116
LECTURE 13 Time Dependent Schrödinger
Equation

Until now, all the potentials we have treated have been constant in time and all the
wavefunctions have been fixed, standing waves. Naturally, many problems in quantum
mechanics involve the fourth dimension of time in addition to the three Cartesian coor-
dinates. The time-dependent Schrödinger equation used for treating such problems is *

∂Ψ ( x, y, z, t )
ih– --------------------------------- = Eq. 13-1
∂t

h– 2 ∂ 2 Ψ ( x, y, z, t ) ∂ 2 Ψ ( x, y, z, t ) ∂ 2 Ψ ( x, y, z, t )
= – -------- ----------------------------------
- + ---------------------------------- - + V ( x, y, z, t )Ψ ( x, y, z, t ) .
- + ----------------------------------
2m ∂x 2 ∂y 2 ∂z 2

Solving with Time Dependence


Consider a system with no varying fields, or where V ( x, y, z, t ) = V ( x, y, z ) . This
is a system that in classical mechanics would have a time independent energy. The cor-
responding quantum mechanical problem is:

∂Ψ ( x, y, z, t )
ih– --------------------------------- = Eq. 13-2
∂t

h– 2 ∂ 2 Ψ ( x, y, z, t ) ∂ 2 Ψ ( x, y, z, t ) ∂ 2 Ψ ( x, y, z, t )
= – -------- ----------------------------------
- + ---------------------------------- - + V ( x, y, z )Ψ ( x, y, z, t ) .
- + ----------------------------------
2m ∂x 2 ∂y 2 ∂z 2

Since the differentiation on the left side of the above equation is with respect to time,
and the differentiations and multiplications on the right side are with respect to spatial
quantities, we can attempt a separation of variables by defining:

Ψ ( x, y, z, t ) = ψ ( x, y, z )φ ( t ) . Eq. 13-3

A simple substitution allows us to rewrite Eq. 13-2 as:

*. See Karplus and Porter, p. 140ff for the details of the derivation.

117
Time Dependent Schrödinger Equation: Solving with Time Dependence

∂φ ( t )
ih– ------------- ψ ( x, y, z ) = Eq. 13-4
∂t

 h– 2 ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) 
=  – -------- ----------------------------- + ----------------------------
- + ----------------------------- + V ( x, y, z ) ψ ( x, y, z ) φ ( t ) ;
 2m ∂x 2 ∂y 2 ∂z 2

completing the separation, we state

∂φ ( t )
-------------
∂t
– ------------
ih - = Eq. 13-5
φ( t)

 h– 2 ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) 
 – ------- - ----------------------------- + ---------------------------- - + ---------------------------- - + V ( x, y, z ) ψ ( x, y, z ) 
 2m ∂x 2 ∂y 2 ∂z 2

= --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- .
ψ ( x, y , z )

One the left, we have a function of time only, and on the right we have a function of space only. Because the overall
energy of the system is fixed, either one of the two sides is equal to E:

∂φ ( t )
-------------
∂t
ih– ------------- = E and Eq. 13-6
φ( t)

 h– 2 ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) 
 – ------- - ----------------------------- + ---------------------------- - + ---------------------------- - + V ( x, y, z ) ψ ( x, y, z ) 
 2m ∂x 2 ∂y 2 ∂z 2

--------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- = E . Eq. 13-7
ψ ( x, y, z )

Rewriting the space-only fuction yields the familiar time independent Schrödinger equation:

h– 2 ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z ) ∂ 2 ψ ( x, y, z )
– -------- ----------------------------
- + ----------------------------
- + ----------------------------- + V ( x, y, z ) ψ ( x, y, z ) = Eψ ( x, y, z ) . Eq. 13-8
2m ∂x 2 ∂y 2 ∂z 2

The new aspect lies in the time-only function. Eq. 13-6 can be solved to give

118
Time Dependent Schrödinger Equation: Solving with Time Dependence

φ ( t ) = exp  ----–-  = exp – ------


Et iEt 
- . Eq. 13-9
 ih   h– 

We now have enough information to write the time dependent wavefunction. It turns out to be the solution to the
time independent Schrödinger equation multiplied by the additional time dependent factor.

Ψ ( x, y, z, t ) = ψ ( x, y, z )φ ( t ) = ψ ( x, y, z ) exp – ------
iEt 
- Eq. 13-10
 h– 

As with time independent solutions, the wavefunction itself is not as physically significant as the probability density, or
the product of the wavefunction and its complex conjugate.

ρ ( x, y, z, t ) = Ψ ( x, y, z, t )Ψ * ( x, y, z, t ) Eq. 13-11

= ψ ( x, y, z ) exp – ------
iEt  *
- ψ ( x, y, z ) exp  ------
iEt 
- = ψ ( x, y, z )ψ * ( x, y, z )
 h–   h– 

Therefore, the probability density of a time dependent equation wavefunction with a time independent potential is itself
time independent.

ρ ( x, y, z, t ) = ρ ( x, y, z ) Eq. 13-12

Even if the energy level happens to be degenerate, and the wave function is a linear combination of orthogonal basis
functions,

Ψ ( x, y, z, t ) = c 1 ψ 1 ( x, y, z )φ 1 ( t ) + c 2 ψ 2 ( x, y, z )φ 2 ( t ) , where c 12 + c 22 = 1 , Eq. 13-13

the time dependent functions are equivalent ( φ 1 ( t ) = φ 2 ( t ) = exp – ------


iEt 
- ) and the probability density remains time inde-
 h– 
pendent. In order to produce a time dependent probability density and wavefunctions affected by the time dependence,
we need a time dependent potential function V ( x, y, z, t ) of some form It should not be surprising that problems involv-
ing a time varying potential are very challenging to solve.

119
Time Dependent Schrödinger Equation: Time Dependent Behavior

Time Dependent Behavior


Consider a problem in which two states are almost degenerate The energies of the two states, labeled “+” and “-”,
are

E + = E 0 – δ , and Eq. 13-14

E- = E 0 + δ Eq. 13-15

The corresponding wave functions for the two states are

1 iE + t 
Ψ + ( x, y, z, t ) = ------- [ ψ 1 ( x, y, z ) + ψ 2 ( x, y, z ) ] exp – --------
- and Eq. 13-16
2  h– 

1 iE - t 
Ψ - ( x, y, z, t ) = ------- [ ψ 1 ( x, y, z ) – ψ 2 ( x, y, z ) ] exp – -------- . Eq. 13-17
2  h– 

These are solutions of the Schrödinger equation with the time independent potential,

ĤΨ + ( x, y, z, t ) = E + Ψ + ( x, y, z, t ) and Eq. 13-18

ĤΨ - ( x, y, z, t ) = E - Ψ - ( x, y, z, t ) . Eq. 13-19

The wave functions ψ 1 ( x, y, z ) and ψ 2 ( x, y, z ) with the energy E 0 are solutions to a simpler Schrödinger equation. Note
their degeneracy.

Suppose we prepare the system at t = 0 in the state ψ 1 and then let it go. At t = t 0 ,

Ψ ( x, y, z, t 0 ) = c + Ψ + ( x, y, z, t0 ) + c - Ψ - ( x, y, z, t 0 ) . Eq. 13-20

Since there are no time varying forces, Ĥ is independent of time,

∂Ψ ( x, y, z, t )
ih– --------------------------------- = ĤΨ ( x, y, z, t ) = Ĥ [ c + Ψ + ( x, y, z, t ) + c - Ψ - ( x, y, z, t ) ] , Eq. 13-21
∂t

and for all time

120
Time Dependent Schrödinger Equation: Time Dependent Behavior

Ψ ( x, y, z, t ) = c + Ψ + ( x, y, z, t ) + c - Ψ - ( x, y, z, t ) . Eq. 13-22

Note that c + and c - are independent of time. The proof follows.

∂Ψ ( x, y, z, t )
ih– --------------------------------- = ĤΨ ( x, y, z, t ) = Ĥ [ c + Ψ + ( x, y, z, t ) + c - Ψ - ( x, y, z, t ) ] Eq. 13-23
∂t


= ih– ----- [ c + Ψ + ( x, y, z, t ) + c - Ψ - ( x, y, z, t ) ] Eq. 13-24
∂t

∂c + ∂ ∂c - ∂
= ih– -------- Ψ + ( x, y, z, t ) + c + ----- Ψ + ( x, y, z, t ) + ------- Ψ - ( x, y, z, t ) + c - ----- Ψ - ( x, y, z, t ) Eq. 13-25
∂t ∂t ∂t ∂t

∂c + ∂c -
= ih– -------- Ψ + ( x, y, z, t ) + c + E + Ψ + ( x, y, z, t ) + ------- Ψ - ( x, y, z, t ) + c - E - Ψ - ( x, y, z, t ) Eq. 13-26
∂t ∂t

Substituting Eq. 13-18 and Eq. 13-19 gives

∂c + ∂c -
ĤΨ ( x, y, z, t ) = ih– -------- Ψ + ( x, y, z, t ) + c + ĤΨ + ( x, y, z, t ) + ih– ------- Ψ - ( x, y, z, t ) + c - ĤΨ - ( x, y, z, t ) Eq. 13-27
∂t ∂t

∂c + ∂c -
= Ĥ [ c + Ψ + ( x, y, z, t ) + c - Ψ - ( x, y, z, t ) ] + ih– -------- Ψ + ( x, y, z, t ) + ------- Ψ - ( x, y, z, t ) Eq. 13-28
∂t ∂t

∂c + ∂c -
= ĤΨ ( x, y, z, t ) + ih– -------- Ψ + ( x, y, z, t ) + ------- Ψ - ( x, y, z, t ) Eq. 13-29
∂t ∂t

Thus, in order for the equation to be true, the following must hold:

∂c + ∂c -
ih– -------- Ψ + ( x, y, z, t ) + ------- Ψ - ( x, y, z, t ) = 0 . Eq. 13-30
∂t ∂t

Since the wavefunctions are non-zero, the time derivatives of c + and c - must be nil.

Let us calculate the probability density of such a system. In order to do so, we need to find a better expression for
the wavefunction Ψ ( x, y, z, t ) . Recall that at t = 0 the system is fixed in the ψ 1 state. Invoking Eq. 13-22,

121
Time Dependent Schrödinger Equation: Time Dependent Behavior

Ψ ( x, y, z, 0 ) = ψ 1 ( x, y, z ) = c + Ψ + ( x, y, z, 0 ) + c - Ψ - ( x, y, z, 0 ) , Eq. 13-31

which using the definitions from Eq. 13-16 and Eq. 13-17 can be restated as

ψ 1 ( x, y, z ) ψ 2 ( x, y, z ) ψ 1 ( x, y, z ) ψ 2 ( x, y, z )
ψ 1 ( x, y, z ) = c + -------------------------- + c + -------------------------- + c - -------------------------- – c - -------------------------- . Eq. 13-32
2 2 2 2

The terms in ψ 2 must cancel, so c + = c - must be true. This leaves us with

ψ 1 ( x, y, z ) ψ 1 ( x, y, z ) 2ψ 1 ( x, y, z )
ψ 1 ( x, y, z ) = c + -------------------------- + c - -------------------------- = c ----------------------------- = c 2ψ 1 ( x, y, z ) . Eq. 13-33
2 2 2

Since c 2 = 1 ,

1
c = c + = c - = ------- . Eq. 13-34
2

Having proved above that c + and c - do not change over time, we can state more generally that

1
Ψ ( x, y, z, t ) = ------- [ Ψ + ( x, y, z, t ) + Ψ - ( x, y, z, t ) ] . Eq. 13-35
2

Using the results from Eq. 13-16 and Eq. 13-17 yet again, we can write

1  1 iE + t  1  iE - t   ,
Ψ ( x, y, z, t ) = -------  ------- [ ψ 1 ( x, y, z ) + ψ 2 ( x, y, z ) ] exp – --------
– -  + ------- [ ψ 1 ( x, y, z ) – ψ 2 ( x, y, z ) ] exp  – --------
h–  
Eq. 13-36
2 2  h 2

or after combining terms,

iE + t  iE - t  1 iE + t  iE - t 
Ψ ( x, y, z, t ) = --- ψ 1 ( x, y, z ) exp – --------
- + exp – -------- + --- ψ ( x, y, z ) exp – --------
- – exp – --------
1
. Eq. 13-37
2  h–   h–  2 2  h–   h– 

If we substitute in the definitions from Eq. 13-14 and Eq. 13-15, the expression becomes

122
Time Dependent Schrödinger Equation: Time Dependent Behavior

1 i ( E 0 – δ )t i ( E 0 + δ )t
Ψ ( x, y, z, t ) = --- ψ 1 ( x, y, z ) exp  – ----------------------- + exp  – ----------------------- . Eq. 13-38
2  –
h   h– 
i ( E 0 – δ )t i ( E 0 + δ )t
+ --- ψ 2 ( x, y, z ) exp  – ----------------------
- – exp  – ----------------------
1
-
2  h–   h– 

After factoring out the E 0 term, we have

1
Ψ ( x, y, z, t ) =  --- ψ 1 ( x, y, z ) exp  ------
iδt
- + exp  – ------
iδt
- . Eq. 13-39
2  h–   h– 

iδt  E 0 t
+ --- ψ 2 ( x, y, z ) exp  ------
iδt
- – exp  – ------ exp  – ------
1
 h–  
- -
2  h–   h– 

The exponential functions readily convert into trigonometric ones:

δt δt E0 t
Ψ ( x, y, z, t ) = ψ 1 ( x, y, z ) cos ---- - exp – ------
- + iψ 2 ( x, y, z ) sin ---- - . Eq. 13-40
h– h–  h– 

We can now write a meaningful probability density function for this system.

ρ ( x, y, z, t ) = Ψ ( x, y, z, t )Ψ * ( x, y, z, t ) Eq. 13-41
δt δt
= ψ 1 ( x, y, z )ψ 1* ( x, y, z ) cos2 ----
- + ψ 2 ( x, y, z )ψ 2* ( x, y, z ) sin2 ----
-
h– h–

h h
This means that the system oscillates between the two “non-stationary” states in a time --- , or with a period ------ , or with a
δ 2δ

frequency ------ .
h

123
Time Dependent Schrödinger Equation: Oscillation Between Interacting States

Oscillation Between Interacting States


There are many examples of oscillation between two interacting states. We will look at two examples to find out
what sorts of δ s exist when two initially degenerate states are allowed to act on one another.

Example 13-1. NH 3 molecules are known to invert back and forth between their right- and left-handed forms, much
like an umbrella turning inside out, and then back again. What is the frequency of this inversion?

The two conformations (or states) can be thought of as oscillators. If the trigonal planar transition state is at x = 0 ,
then the left-handed oscillator (with an eigenfunction L ( x ) ) is centered at x = – l 0 and the right-handed one (with an
eigenfunction R ( x ) ) at x = l 0 . If the two states are allowed to interact, then two new real-time independent states result.
These are

1
------- ( L ( x ) + R ( x ) ) = ψ + , Eq. 13-42
2

with a corresponding energy eigenvalue E + , and

1
------- ( L ( x ) – R ( x ) ) = ψ - , Eq. 13-43
2

with the energy eigenvalue E - . For the lowest energy states of NH 3 , the oscillator states are

L ( x ) = exp [ – α ( x + l 0 ) 2 ] and R ( x ) = exp [ – α ( x – l 0 ) 2 ] . Eq. 13-44

Note that the function ψ + has no nodes, while ψ - has a node at x = 0 . The energy E + is therefore lower than E - Thus,
if we start with the left handed state at t = 0 , the molecule will oscillate between the two conformations at the frequency

E- – E+
ν = ---------------- . Eq. 13-45
h

For NH 3 this frequency is approximately 2 × 10 10 Hz or 0.67 cm –1 .

124
Time Dependent Schrödinger Equation: Optical Transitions

Optical Transitions
Optical transitions, or electron shifts caused by electromagnetic radiation are the basis of optical spectroscopy. Fun-
damentally, the interaction of light with electrons can be described as the interaction between two electronic states which
interact with each other by means of radiation.

Example 13-2. Consider a system with two states. The first state consists of an atom in its ground state (with energy E g )
and N photons of frequency ν 0 . The second state is the atom in an excited state (with energy E g + hν o ) and N – 1 pho-
tons with frequency ν 0 . How fast does the system convert between the two states?

In the absence of the interaction between matter and radiaton, the two states are exactly degenerate. They have the
same energy because E g + Nhν 0 = E g + hν 0 + ( N – 1 )hν 0 . Let E be the electric field of the radiation, and er the dipole
moment of the atom. The energy of interaction between the two is therefore

E ⋅ ec r = e c E ⋅ r . Eq. 13-46

The Hamiltonian of the interacton between the two states mediated by radiation is


H 12 = – e c ∫ ψ 1* E ⋅ rψ 2 dx dy dz , Eq. 13-47
–∞

which can be broken up into directional components:

 ∞ ∞ ∞

H 12 = – e c E x ∫ ψ 1* xψ 2 dx dy dz + E y ∫ ψ 1* yψ 2 dx dy dz + E z ∫ ψ 1* zψ 2 dx dy dz Eq. 13-48
 –∞ –∞ –∞ 

The wavelength of light is on the order of 5000 Å , while the dimension of an atom is about 1 Å . Therefore, the electric
field can be considered spatially uniform across the atom, and in general only one component of E is non-vanishing.
Thus, if E z ≠ 0 and E x = E y = 0 , then the interaction Hamiltonian reduces to


H 12 = – e c E z ∫ ψ 1* zψ 2 dx dy dz = – e c E z µ 12 , Eq. 13-49
–∞

where µ 12 is called the transition dipole moment.

125
Time Dependent Schrödinger Equation: Optical Transitions

The two states are no longer degenerate, being split by 2 H 12 . We can describe the time dependent behavior of this
system with the approach elaborated in Example 13-1. Using Eq. 13-41 with δ = 2 H 12 , the probability density can be
expressed as

ρ ( x, y, z, t ) = Ψ ( x, y, z, t )Ψ * ( x, y, z, t ) . Eq. 13-50

H 12 t H t
= ψ 1 ( x, y, z )ψ 1* ( x, y, z ) cos2 ---------
– - + ψ 2 ( x, y, z )ψ 2 ( x, y, z ) sin ---------
* 2 12-
h h–

The system undergoes transitions between the two states, as long as H 12 ≠ 0 , that is as long as


µ 12 = ∫ ψ 1* zψ2 dx dy dz ≠ 0 .
–∞
Eq. 13-51

In order for the transition dipole integral not to vanish, the two wavefunctions have to be of opposite parity, that is one
must be odd and the other even with respect to inversion. This is known as an optical selection rule, and can be rigorized
by looking at the actual wavefunctions.

For an atom (for instance, the H atom), we have shown that ψ ( – x, – y, – z ) = ( – 1 ) l ψ ( x, y, z ) (see Eq. 11-21). Thus,
µ 12 = 0 unless l 1 + l 2 + 1 is even. With two generic wavefunctions

ψ 1 = R nl ( r )P lm ( cos θ )e imφ and Eq. 13-52

ψ 2 = R n′l′ ( r )P l′m′ ( cos θ )e im′φ , Eq. 13-53

we can re-express the transition dipole moment more precisely in polar coordinates:

∫ ψ1* zψ2 dx dy dz = ∫ Rnl ( r )Plm ( cos θ )e –imφ ( r cos θ )Rn′l′ ( r )Pl′m′ ( cos θ )e im′φ r 2 dr sin θ dθdφ .
–∞
Eq. 13-54

By separating the variables into discrete integrals, we can get a better picture of when the integral vanishes.

∞ π 2π
µ 12 = ∫ R nl ( r )R n′l′ ( r )r 3 dr ∫ P lm ( cos θ )P l′m′ ( cos θ ) cos θ sin θ dθ ∫0 e –i mφ e im′φ dφ ≠ 0 Eq. 13-55
0 0

The above equation is satisfied as long as

126
Time Dependent Schrödinger Equation: Optical Transitions

∫0 e –i mφ e im′φ dφ ≠ 0 Eq. 13-56

which holds true if

m = m′ Eq. 13-57

and as long as

∫0 Plm ( cos θ )P l′m′ ( cos θ ) cos θ sin θ dθ ≠ 0 Eq. 13-58

or

l = l′ ± 1 Eq. 13-59

Note that the inversion symmetry of cos θ is – 1 . That of P lm is ( – 1 ) l , and that of P l′m′ is ( – 1 ) l′ . Thus, the inversion sym-
metry of the integrand in Eq. 13-58 is ( – 1 ) ( l + l′ + 1 ) . For the integral not to vanish, l + l′ + 1 must be even. Recall that
P lm ( cos θ ) is a k-th order polynomial in cos θ , and that k = l – m . If we take a concrete example, where l ≥ l′ , then P lm
is

P lm = ( sinm θ )P ( l – m ) ( cos θ ) Eq. 13-60

or the product of sinm θ and a polynomial of order ( l – m ) in cos θ , and correspondingly,

P l′m = ( sinm θ )P ( l′ – m ) ( cos θ ) Eq. 13-61

or P l′m is the product of sinm θ and a polynomial of order ( l′ – m ) in cos θ . If we include the additional cos θ term with
the last expression, we can write it as

P l′m cos θ = ( sinm θ )P ( l′ + 1 – m ) ( cos θ ) . Eq. 13-62

In order for the integral not to vanish, the orders of the two polynomials (Eq. 13-60 and Eq. 13-62) have to be equal,
l – m = l′ + 1 – m , and l = l′ + 1 .

Note that light must not necessarily be polarized in the z direction, which in turn affects the selection rules. For
example, if E x ≠ 0 , then the transition moment dipole is

127
Time Dependent Schrödinger Equation: Optical Transitions


µ 12 = ∫ ψ1* xψ2 dx dy dz ,
–∞
Eq. 13-63

which when expanded and converted into polar coordinates becomes

∫ ψ1* xψ 2 dx dy dz = ∫ Rnl ( r )Plm ( cos θ )e –i mφ ( r sin θ cos φ )R n′l′ ( r )P l′m′ ( cos θ )e im′φ r 2 dr sin θ dθdφ .
–∞
Eq. 13-64

Separating the variables into separate integrals for clarity, we obtain

∞ π 2π
µ 12 = ∫ R nl ( r )R n′l′ ( r )r 3 dr ∫ P lm ( cos θ )P l′m′ ( cos θ ) sin2 θ dθ ∫0 e –imφ e im′φ cos φ dφ ≠ 0 . Eq. 13-65
0 0

Again, the forbidden transitions are between states that make any one of the above integrals vanish. For

π 2π

∫ Plm ( cos θ )P l′m′ ( cos θ ) sin2 θ dθ


0
∫0 e –i mφ e im′φ cos φ dφ ≠ 0 Eq. 13-66

the same rule holds as above, namely that l = l′ ± 1 . The difference lies in the integral in φ , where

∫0 e –i mφ e im′φ cos φ dφ ≠ 0 Eq. 13-67

only when

m = m′ ± 1 . Eq. 13-68

The dependence of the radial integral on ( n, n′ ) and l is somewhat complex. The selection rule, however, is that the
change in n can be somewhat arbitrary.

128
Time Dependent Schrödinger Equation: Optical Transitions

To summarize, optical transitions for atoms are allowed for

∆n = arbitrary Eq. 13-69

∆l = ± 1 Eq. 13-70

∆m = 0, ± 1 . Eq. 13-71

The results of these selection rules are quite dramatic. For example, in the transition from the n = 2 state to the 1s
orbital of hydrogen, only the 2p-1s transition has ∆l = – 1 . The lifetime of the 2p state is approximately 10 –9 sec , while
that of the 2s (for which there is no allowed electronic transition to the 1s state) is approximately 0.1 sec .

129
Time Dependent Schrödinger Equation: Optical Transitions

130
LECTURE 14 Electron Spin

So far, we have treated electrons as being placed in orbitals described by three


quantum numbers only. We will now introduce a fourth, which should be familiar from
general chemistry. Electron spin is not an obvious concept. It is not a part of any of the
wave equations we have tackled so far, nor does it result from the Schrödinger equa-
tion. It can be derived rigorously through a relativistic treatment of electron motion, but
we will simply accept it here as an empirical fact. The empirical evidence for the exist-
ence of electron spin comes from the so-called fine structures of electronic configura-
tions of atoms.

Before we discuss these, let us define some nomenclature used in the context of
electron spins. The spin function, s ( σ ) , is a vector function of the spin variable σ . The
form of the function is of no concern. The important thing is that the spin function
1
always takes the value of --- , which can be directed either up or down. The spin of the
2
electron results in a certain amount of angular momentum. The z-component of the
1
angular momentum can also have only two values, namely S z = ± --- h– . The quantum
2
1
number associated with spin, m s , is given the values ± --- . Other conventions for
2
describing the spin state of an electron include α, β and ↑, ↓ . Without proof, we will
state that the spin functions are orthogonal. This means that

α⋅α = 1
β⋅β = 1 Eq. 14-1
α⋅β = 0

Thus, the complete wavefunctions including spin should be written as Ψ ( r, s ) ,


where r describes the spatial components and s the spin of the electron’s wavefunc-
tion. We can now specify Ψ entirely in terms of four quantum numbers. There are
three spatial designations, N , the radial quantum number, l , or the angular momentum
quantum number, where L 2 = l ( l + 1 )h– 2 , and m l , or the z component of the angular
momentum, where L z = m l h– , and the fourth quantum number, m s , which is a spin
1
designation with values of m s = ± --- .
2

131
Electron Spin: Magnetic Moment of an Electron

Magnetic Moment of an Electron


An electron in an orbital is a moving charge, which possesses a magnetic moment. There are two sources of the
magnetic moment of a moving electron: the magnetic moment due to the electron’s motion in the orbital, and the mag-
netic moment due to its intrinsic spin.

Orbital Magnetic Moment. For a classical current in a loop, the magnetic moment (in cgs units) is given by

IA
µ = ------ Eq. 14-2
c

qv
where I is the current and A is the area of the current loop. The current is expressed as I = --------- , and A = πr 2 . µ can be
2πr
rewritten as

qvr
µ = -------- Eq. 14-3
2c

Since the velocity and the radius r are always perpendicular in circular motion, v × r = vr , and by definition of angular
momentum ( l = mv × r ),

– ec
µ = ------------- L Eq. 14-4
2cm e

We can express the magnitude of the magnetic moment vector by substituting L = h– l ( l + 1 ) , to obtain

–e c
µ = ------------- h– l ( l + 1 ) Eq. 14-5
2cm e

Suppose that the an external uniform magnetic field is applied to the electron. Let the direction of the magnetic field
define the z direction. The energy of interaction of the electron’s magnetic dipole with the magnetic field H would be
µ ⋅ H . However, since L z = m l h– , we can write

–ec
µ ⋅ H = ------------- h– m l H z Eq. 14-6
2cm e

where H = H z .

132
Electron Spin: Magnetic Moment of an Electron

Spin Magnetic Moment. Such an external magnetic field is created by the intrinsic spin of the moving electron. Stated
here without a derivation, the magnetic moment due to electron’s spin is

–ec
µ s = ---------- S Eq. 14-7
cm e

1
and its magnitude (with s = --- ) is
2

ec
µ s = ---------- h– s ( s + 1 ) Eq. 14-8
cm e

This is responsible for the Zeeman effect, or the splitting of electrons in a strong magnetic field due to the interaction
given by

– ec
µ s ⋅ H = ---------- m s H z Eq. 14-9
cm e

133
Electron Spin: Magnetic Moment of an Electron

134
LECTURE 15 The Multielectron Atom

The next task at hand is to develop a model for atoms with more than one elec-
tron. Already a single additional electron complicates the Hamiltonian tremendously, as
the two electrons interact not only with the nucleus, but also with each other. The
Schrödinger equation for a two electron atom is

h– 2 e c2
ĤΨ = – ---------- ( ∇ 12 + ∇ 22 ) – Ze c2  ---- + ----  – ------ Ψ ( ξ 1, ξ 2 ) = EΨ ( ξ 1, ξ 2 ) .
1 1
Eq. 15-1
2m e  r 1 r 2  r 12

The ∇ i2 ‘s are the Laplacian operators for the two electrons. The potential term includes
the interaction of the two electrons with the nucleus, where r i is the electronic distance
from the nucleus, as well as with each other, where r 12 is the interelectronic distance.
The wavefunction Ψ is a function of two ξ i ‘s. Each ξ i is a four dimensional vector that
combines the three space ( x, y, z ) and single spin ( s ) coordinates.

The same type of Hamiltonian can be generalized to n electrons:

n
h– 2 n
e c2
∑ – ---------
- ∇ 2 – Ze c2 ---  + ∑ ----
1
Ψ ( ξ 1, ξ 2, …, ξ n ) = EΨ ( ξ 1, ξ 2, …, ξ n ) Eq. 15-2
i=1 2m e i r i  i > j = 1 r ij

The Hamiltonian does not include the spin-orbit interaction. This is reasonably correct
for light atoms, and will be added and discussed later.

n
h– 2 2
– --------- 1
2 ---  +
n
e c2
A very important aspect of the ∑
i=1
 2m e
- ∇ i – Ze c
r i  ∑ ---- operator is that it is
i > j = 1 r ij

symmetric with respect to electron interchange. In other words, the result is the same,
irrespective of the order in which the electrons are numbered. Another important char-
acteristic of the multielectron operator is that the symmetry of its eigenfunctions does
not change in time. The general time-dependent Schrödinger equation is

∂Ψ ( t, ξ 1, ξ 2, …, ξ n )
ĤΨ ( t, ξ 1, ξ 2, …, ξ n ) = – ih– ------------------------------------------------ . Eq. 15-3
∂t

135
The Multielectron Atom: Two Electron Systems

If the wavefunction has a particular symmetry with respect to particle interchange at a time designated t = 0 , then it will
retain that symetry for all times, since the Hamiltonian will not affect it. Nonetheless, the order of particles at time t = 0
does matter. One of the fundamental postulates of quantum mechanics is that the wavefunction for electrons must be
antisymmetric with respect to particle interchange.

Ψ ( ξ 1, ξ 2, …, ξ n ) = – Ψ ( ξ 2, ξ 1, …, ξ n ) Eq. 15-4

This is known as the Pauli Exclusion Principle, and it implies that no two electrons can have the same set of four quan-
tum numbers ( ξ i ≠ ξ j for i ≠ j ), because such a configuration would allow for an exchange of electrons without a change
in the sign of the wavefunction.

Two Electron Systems


Two electron systems (or in chemical language, electron pairs) are very useful, because it is possible to sparate their
eigenfunctions into spatial and spin parts. Ψ can be separated into two functions:

Ψ ( ξ 1, ξ 2 ) = φ ( r 1, r 2 )χ ( s 1, s 2 ) , Eq. 15-5

and since Hamiltonian from Eq. 15-1 does not operate on spin, the expression

h– 2 e c2
– ---------- ( ∇ 12 + ∇ 22 ) – Ze c2  ---- + ----  – ------ φ ( r 1, r 2 )χ ( s 1, s 2 ) = Eφ ( r 1, r 2 )χ ( s 1, s 2 )
1 1
Eq. 15-6
2m e  r 1 r 2  r 12

can be written entirely without the spin function:

h– 2 e c2
– ---------- ( ∇ 12 + ∇ 22 ) – Ze c2  ---- + ----  – ------ φ ( r 1, r 2 ) = Eφ ( r 1, r 2 ) .
1 1
Eq. 15-7
2m e  r 1 r 2  r 12

However, since the overall eigenfunctions must be antisymmetric with respect to electron interchange, two possibilities
occur. If the spin is symmetric, χ ( s 2, s 1 ) = χ ( s1, s 2 ) , then the spatial function must be antisymmetric,
φ ( r 2, r 1 ) = – φ ( r 1, r 2 ) . Alternately, if the spin function is antisymmetric, χ ( s 2, s 1 ) = – χ ( s 1, s 2 ) , then the space must be
symmetric, φ ( r 2, r 1 ) = φ ( r 1, r 2 ) . We are interested in writing at least approximate expressions for these functions.

136
The Multielectron Atom: Two Electron Systems

Indistinguishability of Electrons. Before we can treat the approximations, we must pause to note that electrons are
indistinguishible. In the expressions above, we have given number labels to individual electrons, but in reality it is
impossible to label them. Thus, in case of two electrons in the same orbital with opposite spin, we cannot distinguish
between α ( 1 )β ( 2 ) and α ( 2 )β ( 1 ) as two separate functions. Instead, indistinguishablity requires that we write out all the
possible linear combinations of the two functions. In this case, the possibilities are

1
χ 1 = ------- [ α ( 1 )β ( 2 ) + β ( 1 )α ( 2 ) ] and Eq. 15-8
2

1
χ 2 = ------- [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ] . Eq. 15-9
2

1
The functions are normalized with the ------- . Note that once we begin to write out linear combinations of functions, the
2
neat picture of individual electrons being allocated to orbitals no longer holds, and the arrow notation ( ↑, ↓ ) loses its
meaning.

Singlet and Triplet States. As discussed above, a two electron system can have a spin function that is either symmetric
or antisymmetric with respect to electron interchange. If the two electrons have opposite spins, then two possible spin
functions result (Eq. 15-8 and Eq. 15-9). If the two electrons have parallel spins, then the two possible spin functions are

χ 3 = α ( 1 )α ( 2 ) and Eq. 15-10

χ 4 = β ( 1 )β ( 2 ) . Eq. 15-11

These two functions are not a basis for linear combinations, because it is possible to distinguish them based on the direc-
tion of the spin. Out of the four functions, only one is antisymmetric when the electron labels are changed. A state char-
acterized by this spin function is called a singlet state.

1
χ 2 = ------- [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ] Eq. 15-12
2

1 1
------- [ α ( 2 )β ( 1 ) – β ( 2 )α ( 1 ) ] = – ------- [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ] = – χ 2 Eq. 15-13
2 2

The other three spin functions are symmetric with respect to electron exchange. They are called the triplet spin functions.

137
The Multielectron Atom: Orbital Theory: Singlets and Triplets

The singlet and the triplet spin functions are orthogonal. For example,

1
χ 2 ⋅ χ 3 = ------- [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ]α ( 1 )α ( 2 ) Eq. 15-14
2
1
= ------- [ α ( 1 )α ( 1 )β ( 2 )α ( 2 ) – α ( 1 )β ( 1 )α ( 2 )α ( 2 ) ]
2
1
= ------- [ β ( 2 )α ( 2 ) – α ( 1 )β ( 1 ) ]
2
= 0

The same can be shown for the other two combinations.

Optical Selection Rules. Optical selection rules depend upon the interaction of the electric field of the radiation with
the transition dipole moment between the states. This is a spatial quantity, which does not affect spin in any way. The
triplet and singlet states cannot interact optically (or electrostatically). They are orthogonal due to their spin functions,
which cannot be coupled by spatial quantities.

Orbital Theory: Singlets and Triplets


n
e c2
In the multielectron Hamiltonian (see Eq. 15-2), it is the ∑ ----
i > j = 1 r ij
term that prevents a separation of φ ( r 1, r 2, …, r n )
n
into a product of single electron functions ∏ u ( r i ) . An approximate set of single electron solutions can be achieved if
i=1
the interelectronic repulsion term is averaged for all the electrons except for the one being examined. The approximate
Hamiltonian thus becomes

h– 2 1 n–1
e c2
= – ---------- ∇ i2 – Ze c2 --- + ∑ ∫ ----ρ ( j ) dr j ,
eff
Ĥ i Eq. 15-15
2m e r i j ≠ i, j = 1 r ij

where ρ ( j ) = ∫ u * ( r j )u ( r j ) dr j , which is the electron density of the j-th electron. Using the approximation, we are faced
with n single electron equations of the form

Ĥ ieff ( r i )u ( r i ) = εu ( r i ) . Eq. 15-16

138
The Multielectron Atom: Orbital Theory: Singlets and Triplets

There is an assortment of notations for describing the orbitals. We can label the u’s with the three spatial quantum num-
bers as u N, l, ml ( r i ) , or to keep things even simpler, we can use orbital notation. Instead of u 1, 0, 0 , we can write 1s and
instead of u 2, 1, +1 , 2p ml = +1 .

The Two Electron Atom. The SCF method is a valid approximation for any number of electrons, including two. We use
it as a convenient example.

As discussed above, the two electron spatial function φ ( r 1, r 2 ) can be either symmetric or antisymmetric. If the elec-
trons have opposite spins, the symmetric spatial function is

φ ( r 1, r 2 ) = u N1, l1, ml1 ( r 1 )u N2, l2, ml2 ( r 2 ) + u N2, l2, ml2 ( r 2 ) ( r 1 )u N1, l1, ml1 ( r 2 ) . Eq. 15-17

Note that spatial indistinguishibility is retained. The corresponding antisymmetric function is

φ ( r 1, r 2 ) = u N1, l1, ml1 ( r 1 )u N2, l2, ml 2 ( r 2 ) – u N2, l2, ml2 ( r 2 ) ( r 1 )u N1, l1, ml1 ( r 2 ) . Eq. 15-18

If we use the 1s and the 2p ml = 0 functions for u ( r 1 ) and u ( r 2 ) , respectively, we can abbreviate the notation to

φ ( r 1, r 2 ) = 1s ( 1 )2p ml = 0 ( 2 ) + 2p ml = 0 ( 2 )1s ( 2 ) Eq. 15-19

for the symmetric function, and

φ ( r 1, r 2 ) = 1s ( 1 )2p ml = 0 ( 2 ) – 2p ml = 0 ( 2 )1s ( 2 ) Eq. 15-20

for the antisymmetric one. If we wish to include spin in this notation, we simply append it at the end as α or β . For
example, 1sα, 2p ml = 1 α, 2sβ .

The Slater Determinant. The Slater Determinant is the simplest many electron antisymmetric function. It is simply the
determinant of a possible combinations of electrons and space functions. In case of a two electron system, the two spin
orbitals could be u and v. The properly normalized wavefunction can be expressed as

1 1 u( 1) u(2 ) u ( 1 )v ( 2 ) – v ( 1 )u ( 2 )
Ψ ( 1, 2 ) = ---------- Det u v = ------- = ---------------------------------------------------- Eq. 15-21
N! 2 v(1) v(2 ) 2

139
The Multielectron Atom: Orbital Theory: Singlets and Triplets

In case of a two-electron system, there are two contributions to overall electron density.

ρ( r) = ρ( 1) + ρ( 2) . Eq. 15-22

We can obtain an expression for the electron density due to the first electron by taking the probability density of the
wavefunction over the coordinates of the second electron.

ρ ( 1 ) = ∫ Ψ * Ψ dr 2 Eq. 15-23

Substituting Ψ from Eq. 15-21, we write

1 1
ρ ( 1 ) = --- ∫ [ u 2 ( 1 )v 2 ( 2 ) + v 2 ( 1 )u 2 ( 2 ) ] dr 2 – --- ∫ u ( 1 )v ( 1 )u ( 2 )v ( 2 ) dr 2 . Eq. 15-24
2 2

Since u and v are orthogonal, the second integral disappears. In the first, the normalized functions of the second electron
become unity when integrated over r 2 . The resulting expression is

1
ρ ( 1 ) = --- [ u 2 ( 1 ) + v 2 ( 1 ) ] . Eq. 15-25
2

An analogous equation can be derived for the second electron,

1
ρ ( 2 ) = --- [ u 2 ( 2 ) + v 2 ( 2 ) ] . Eq. 15-26
2

The overall electron density is the sum of the individual electron densities.

1 1
ρ ( r ) = --- [ u 2 ( 1 ) + v 2 ( 1 ) ] + --- [ u 2 ( 2 ) + v 2 ( 2 ) ] Eq. 15-27
2 2

If we remove the labels and use a generalized coordinate vector r , the overall density becomes the sum of the electron
densities due to the original two orbital functions, exactly as expected.

1 1
ρ ( r ) = --- [ u 2 ( r ) + v 2 ( r ) ] + --- [ u 2 ( r ) + v 2 ( r ) ] = u 2 ( r ) + v 2 ( r ) Eq. 15-28
2 2

The effective charge density is the sum of the charge density of an electron in each orbital.

140
The Multielectron Atom: Orbital Theory: Singlets and Triplets

The Helium Atom. We can use the SCF model to describe the ground state of helium. The two basis functions are the
u = 1sα and v = 1sβ . The overall wavefunction has to be antisymmetric. Since both of the electrons are in the same 1s
orbital, the spin function has to be the antisymmetric part. This gives makes

Ψ = [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ]1s ( 1 )1s ( 2 ) . Eq. 15-29

There are two possible forms of an excited state of helium. If the spins are parallel, it is a triplet state, where u = 1sα
and v = nlα . The antisymmetrized wavefunction is then

Ψ = 1sα ( 1 )nlα ( 2 ) – 1nlα ( 1 )1sα ( 2 ) = [ 1s ( 1 )nl ( 2 ) – nl ( 1 )1s ( 2 ) ]α ( 1 )α ( 2 ) . Eq. 15-30

Alternately, if the spins are antiparallel, the helium atom is in an excited singlet state with u = 1sα and v = nlβ . How-
ever, while the ground state and the excited triplet state functions could be expressed as a single Slater determinant, the
excited singlet is more complicated. It is a sum of two Slater determinants, which ensures that both the space and the
spin functions allow for indistinguishability of electrons.

Ψ = [ 1s ( 1 )nl ( 2 ) + nl ( 1 )1s ( 2 ) ] [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ] Eq. 15-31

As an exercise, write out the two Slater determinants that give rise to this wavefunction.

N Electron System. Slater determinants are especially useful as a approximate wavefunctions for atoms or molecules
with more than two electrons. For example, the ground state of Be ( 1s 2 2s 2 ) can be written as

1sα ( 1 ) 1sα ( 2 ) 1sα ( 3 ) 1sα ( 4 )


1 1sβ ( 1 ) 1sβ ( 2 ) 1sβ ( 3 ) 1sβ ( 4 ) 1
Ψ = --------- = --------- Det 1sα 1sβ 2sα 2sβ . Eq. 15-32
4! 2sα ( 1 ) 2sα ( 2 ) 2sα ( 3 ) 2sα ( 4 ) 4!
2sβ ( 1 ) 2sβ ( 2 ) 2sβ ( 3 ) 2sβ ( 4 )

The rows of a Slater determinant are indexed by spin orbitals and the columns by the electrons. The wavefunction is nor-
1
malized by a factor of ---------- , where N is the number of electrons involved.
N!

Singlets and Triplets. Recall the Hamiltonian for a two electron atom

h– 2 e c2
Ĥ = – ---------- ( ∇ 12 + ∇ 22 ) – Ze c2  ---- + ----  – ------
1 1
Eq. 15-33
2m e  r 1 r 2 r 12

141
The Multielectron Atom: Orbital Theory: Singlets and Triplets

The singlet and triplet states of the two electron systems are non-degenerate. Consider, for example, the two states
1
in the He configuration 1s2s. For comparison, we use the triplet spin function ------- [ α ( 1 )β ( 2 ) + β ( 1 )α ( 2 ) ] and the singlet
2
1 1
spin function ------- [ α ( 1 )β ( 2 ) – β ( 1 )α ( 2 ) ] . The associated spatial functions are ------- [ 1s ( 1 )2s ( 2 ) – 2s ( 1 )1s ( 2 ) ] and
2 2
1
------- [ 1s ( 1 )2s ( 2 ) + 2s ( 1 )1s ( 2 ) ] , respectively. The orbital functions are assumed to be those appropriate to an “average”
2
e c2
value of ------ appropriate to the orbital involved, i.e. screened potentials. The total energy eigenvalue for the system is
r 12
thus approximated as a sum of the eigenvalues for the orbital functions for individual electrons.

h– 2 2
– --------- 1 e c2
- ∇ – Ze c2 ---- – 〈 ------〉  1s ( 1 ) = ε 1s 1s ( 1 ) Eq. 15-34
 2m e 1 r1 r 12 

h– 2 2
– --------- e c2
- ∇ 1 – Ze c2 ---- – 〈 ------〉  2s ( 1 ) = ε 2s 2s ( 1 )
1
Eq. 15-35
 2m e r1 r 12 

h– 2 2
– --------- 1 e c2
- ∇ – Ze c2 ---- – 〈 ------〉  1s ( 2 ) = ε 1s 1s ( 2 ) Eq. 15-36
 2m e 2 r2 r 12 

h– 2 2
– --------- e c2
- ∇ – Ze c2 ---- – 〈 ------〉  2s ( 2 ) = ε 2s 2s ( 2 )
1
Eq. 15-37
 2m e 2 r2 r 12 

The difference in energies between singlet and triplet states of the same configuration arise from the electrostatic
e c2
repulsion term, ------ , which when averaged over the triplet and singlet states gives a different value. It can be shown that
r 12

1 e c2 1
∫ ------2- [ 1s ( 1 )2s ( 2 ) – 2s ( 1 )1s ( 2 ) ] -----
- ------- [ 1s ( 1 )2s ( 2 ) – 2s ( 1 )1s ( 2 ) ] dr 1 dr 2 ≠
r 12 2
Eq. 15-38

1 e c2 1
∫ ------2- [ 1s ( 1 )2s ( 2 ) + 2s ( 1 )1s ( 2 ) ] -----
- ------- [ 1s ( 1 )2s ( 2 ) + 2s ( 1 )1s ( 2 ) ] dr 1 dr 2
r 12 2

and that the difference between the two is

142
The Multielectron Atom: Angular Momentum of Two Particles

e c2
2 ∫ 2s ( 1 )1s ( 2 ) ------ 1s ( 1 )2s ( 2 ) dr 1 dr 2 > 0 . Eq. 15-39
r 12

Thus, the triplet state with an antisymmetric spatial function is lower in energy than the singlet state. This is true for all
configurations. The integral in Eq. 15-39 is called the exchange integral, and it expresses the energy diffrence between
5 e c2
the triplet and the singlet states. For a helium atom, the exchange integral can be evaluated to be ---  --------  .*
2 2a 0

Angular Momentum of Two Particles


A system of two particles, such as two electrons, has a total angular momentum. This momentum is a vector sum of
the angular momenta of the two particles. Thus

L = l1 + l2 Eq. 15-40

Another way of measuring total angular momentum is to add the z projections of the angular momentum vectors to get
the magnitude of the angular momentum along the z direction. This is scalar addition.

L = l z1 + l z2 Eq. 15-41

Note that l z is frequently denoted as m l .

In spherical polar coordinates, the operator

∂ ∂
Lˆz = – ih–  -------- + --------  Eq. 15-42
 ∂φ 1 ∂φ 2 

gives the z-projection of total angular momentum. Another way of looking at the angular momentum is

2
L = L ⋅ L = ( l 1 + l 2 ) ⋅ ( l 1 + l 2 ) = l 1 ⋅ l 1 + l 2 ⋅ l 2 + 2l 1 ⋅ l 2 Eq. 15-43

*. Refer to Karplus and Porter, pp. 174-5 for complete details.

143
The Multielectron Atom: Spin-Orbit Interaction

2
The allowed values of L are

2
L = h– 2 [ l 1 ( l 1 + 1 ) + l 2 ( l 2 + 1 ) + 2 cos χ l 1 ( l 1 + 1 )l 2 ( l 2 + 1 ) ] Eq. 15-44

where 1 > cos χ > – 1 , and χ is the angle between the two angular momentum vectors. The allowed values of L z are

L z = h– ( m l1 + m l2 ) Eq. 15-45

The maximum occurs when both angular momentum vectors are aligned with the z-axis, and it is equal to

L z = h– ( l 1 + l 2 ) Eq. 15-46

The minimum value occurs when both angular momentum vectors are aligned antiparallel to the z-axis, in which case

L z = – h– ( l 1 + l 2 ) Eq. 15-47

Angular momentum is quantized in steps of h– . Thus, the intermediate values of L z are

L z = h– ( l 1 + l 2 – n ) , with n = 1, 2, …, 2 ( l 1 + l 2 ) Eq. 15-48

This introduces a large amount of degeneracy. For example, while there is only one vector arrangement that will yield
L z = h– ( l 1 + l 2 ) or L z = – h– ( l 1 + l 2 ) , there are two for L z = h– ( l 1 + l 2 – 1 ) and L z = – h– ( l 1 + l 2 – 1 ) , and so on.

Thus, given the same system, there are several states of different total angular momentum. The total number of
states is given by

( 2l 1 + 1 ) ( 2l 2 + 1 ) Eq. 15-49

Spin-Orbit Interaction
Spin-orbit interaction results from the interaction of the electron’s spin magnetic moment with the effective mag-
netic field produced by its motion relative to the nucleus. The energy of the spin orbit interaction for one electron can be
shown to be

144
The Multielectron Atom: Term Symbols

e c2 h– 2 Z 4
E S.O. = -------------------------------------------------------------- L ⋅ S Eq. 15-50
2m e c 2 a 03 n 3 l l + ---  ( l + 1 )
1
 2

We evaluate L ⋅ S . To do this, we express the total angular momentum as

J = L+S Eq. 15-51

L ⋅ S can now be evaluated by noting that

2 2 2
J = J ⋅ J = ( L + S ) ( L + S ) = L + S + 2L ⋅ S Eq. 15-52

or that

2 2 2
J – (L + S ) h– 2
L ⋅ S = ------------------------------ = ----- [ J ( J + 1 ) – L ( L + 1 ) – S ( S + 1 ) ] Eq. 15-53
2 2

2
Note that this expression assumes that the magnitude of L is unchanged, or that L = L ( L + 1 )h– 2 . See Table 15.1 at the
end of this lecture for some experimental spin-orbit coupling data.

Term Symbols
The spin-orbit interaction gives rise to the fine structure of the electronic states of the elements and accounts for the
energy splittings observed in a single electron configuration. To appreciate the subtle differences in electronic configura-
tions, we need to resort to term symbols, a systematic manner of describing them. The term symbol is a shorthand way of
condensing a great deal of information. Although we have not yet formally discussed many electron systems, term sym-
bols are used primarily in that context. They will prove useful in the rigorous development in the ensuing lectures. The
general form of a term symbol is

2S + 1
LJ

L is the total orbital angular momentum number, S is the total spin quantum number, and J is the total angular momentum
number. The correspondence between the value of L and the symbol is similar to that of the familiar correspondence
between the quantum number l and the orbital designations, namely

145
The Multielectron Atom: Term Symbols

L= 0 1 2 3 4 5 …
S PD F G H …

1 3
For example, if we ignore the J subscript, our two electron systems can be described either as a S and P (read: singlet
S and triplet P).

The total orbital angular momentum and the total spin angular momentum add as vectors, or

L = ∑i l i and Eq. 15-54

S = ∑i si Eq. 15-55

However, their z-components add as scalar sums, where

Lz = ∑i ( lz )i = ∑i ( ml )i = M L and Eq. 15-56

Sz = ∑i ( sz )i = ∑i ( ms )i = MS Eq. 15-57

Just like the z-component of l can assume 2l + 1 values m l = l, l – 1, …, 0, …, – l , the z-component of L can assume
2L + 1 values M L = L, L – 1, …, 0, …, – L . The sum of the spins, M S , can be treated analogously.

As an example, let us consider a boron atom with a 1s 2 2s 2 2p 1 electron configuration. For the two closed electron
1
shells, both L and S are necessarily 0. For the 2p orbital, m l = 1, 0, – 1 , and therefore L = 1 , and S = --- . Thus, the term
2
2
symbol would be P (read: doublet P).

As defined in Eq. 15-51, J is a vector sum. J, or the vector’s magnitude, can take on a whole array of values,
depending on the mutual alignment of L and S . If the two vectors are parallel, J is at a maximum. Because of quantiza-
tion, J can only take on distinct values in the range

146
The Multielectron Atom: Term Symbols

J = L + S, L + S + – 1, L + S + – 2, …, L – S Eq. 15-58

1 3 1
In case of the boron atom mentioned above, where L = 1 and S = --- , J = ---, --- . Thus, if the orbital angular momentum
2 2 2
2
and the spin angular momentum are aligned, the complete term symbol is P3/2 . If the two are directed in an antiparallel
2
fashion, the term symbol is P1/2 .*

Table 15.1 - A summary of the energies involved in the spin-orbit effect for atoms with a single electron or a single hole. Overall, the effect is small
compared to the energies of different configurations for low-Z atoms. Notice the sharp inverse cube dependance on n in the excited states of Ga, as
predicted by Eq. 15-50.

Element Valence Term ES.O. Element Valence Term ES.O.


(Z) Orbital Symbol (cm-1) (Z) Orbital Symbol (cm-1)
B 2p 2P 0 F 2p5 2P 0
1/2 3/2
Z=5 2p 2 16 Z=9 404
P3/2 2p5 2
P1/2
Al 3p 2
P1/2 0 Cl 3p5 2
P3/2 0
Z = 13 3p 2P 112 Z = 17 881
3/2 3p5 2P
1/2

Ga 4p 2
P1/2 0 Br 4p5 2
P3/2 0
Z = 31 4p 2 826 Z = 35 3685
P3/2 4p5 2
P1/2
5s 2S 24788
1/2

5p 2
P1/2 33044

5p 2
P3/2 33155

In 5p 2P 0 I 5p5 2P 0
1/2 3/2
Z = 49 5p 2 2212 Z = 53 7603
P3/2 5p5 2
P1/2
Tl 6p 2
P1/2 0
Z = 81 6p 2P 7792
3/2

*. For a thorough treatment of term symbols, refer to McQuarrie and Simon, p. 292ff.

147
The Multielectron Atom: Hund’s Rules

Hund’s Rules
We mentioned that the differences in electronic configuration indicated by the different term symbols lead to subtle
differences in energy due to spin-orbit coupling. While the precise energies can be computed using Eq. 15-50, frequently
only a relative order is necessary. A set of empirical rules called Hund’s rules provides a systematic approach.
1. The state with the lowest energy has the largest S value.
2. Among states with the same S value, the state with the largest L value has the lowest energy.

If states have the same S and L, then for a subshell that is less than half-filled, a smaller J corresponds to lower energy; for
a subshell that is more than half-filled, a higher J corresponds to lower energy.

148
LECTURE 16 The Variational Method

When tackling a problem involving the Schrödinger equation, very frequently we


have a general idea of what the solution should be, but do not know its precise func-
tional form. An alternative method to perturbation theory, the Variational Method
serves as a good starting point or a reasonable approximation.

Basis Functions
We begin with a set of orthogonal, normalized basis functions. No matter whether
approximate or exact, the solution to the problem can be expressed by some linear
combination of the basis functions. If the problem at hand is the usual ĤΨ = EΨ , then
we write

Ψ(x) = ∑ cnφn( x )
n=0
Eq. 16-1

to express the expansion of the wavefunction in terms of the set of basis functions,
{ φ n ( x ) } . Note that c n = ∫ Ψ ( x )φ n* ( x ) dx and ∫ φ n ( x )φ m* ( x ) dx = δ mn . The functions
may be chosen, for example, from the solution of a particular Schrödinger equation.
The solution set of any Schrödinger equation will work.

The problem can be reexpressed as

ĤΨ = Ĥ ∑ c n φ n ( x ) = EΨ = E ∑ c n φ n ( x ) Eq. 16-2


n=0 n=0

When multiplied through by φ m* and integrated, it becomes

∑ cn ∫ φ m* ( x ) Ĥφn ( x )dx
n=0
= E ∑ c n ∫ φ m* ( x ) φ n ( x )dx = Ec m
n=0
Eq. 16-3

We can use the notation developed in the context of perturbation theory in Eq. 12-31
and Eq. 12-32 to abbreviate this very important result:

∑ c n Hmn
n=0
= Ec m Eq. 16-4

149
The Variational Method: Basis Functions

where H mn = ∫ φ m* ( x ) Ĥφ n ( x )dx . This becomes especially useful if we do not need an infinite number of terms. For those
eigenfunctions Ψ such that

N
Ψ(x) ≈ ∑ cn φn ( x ) Eq. 16-5
n=0

we can still state that

∑ c n Hmn = Ec m Eq. 16-6


n=0

and

∑ c n ( Hmn – Eδ mn ) = 0 Eq. 16-7


n=0

In such a case, we are left with N homogeneous equations of this type with N unknowns. The best way of tackling this
problem is by solving the determinant

H – EI = 0 Eq. 16-8

where H is a matrix with the elements H mn , I is the unit matrix and E the eigenvalues of H.

Note that this is only an approximation. If we start with the exact expressions

Ψ(x) = ∑ cnφn ( x )
n=0
and Ψ * ( x ) =
n=0
∑ cn* φ n* ( x ) Eq. 16-9

and the function is normalized so that

∫ Ψ * ( x )Ψ ( x ) dx = ∫ n∑= 0 cn* φ n* ( x ) n∑= 0 cn φn ( x ) dx = ∑ cn* cn


n=0
= 1 Eq. 16-10

the finite expression

N
Ψ(x) ≈ ∑ cnφn( x ) Eq. 16-11
n=0

150
The Variational Method: Proof

will only be accurate if the remaining terms are small, i.e.

∑c*
m
m = N+1
cm ≈ 0 Eq. 16-12

Note that c m* c m ≥ 0 .

Proof
The theorem behind the variational method is that the energy of an approximate solution based on a set of func-
tions will always be greater than, or at best, equal to the ground state energy of the system. For an arbitrary function f ,
we define the variational function as

∫ φk* Ĥφk
W k = ------------------ . Eq. 16-13
∫ φk* φk
The variational principle states that

Wk ≥ E0 , Eq. 16-14

where E 0 is the lowest energy solution to ĤΨ n = E n Ψ n . Here’s the proof. We start with the formal expansion

φk = ∑ c kn Ψn ( x ) .
n=0
Eq. 16-15

Even though we do not know what Ψ is, we know the expression to be true. We can then rewrite Eq. 16-13.

∫ n∑= 0 ckn* Ψn* ( x ) Ĥ n∑= 0 c kn Ψn ( x ) dx


W k = -------------------------------------------------------------------------------- Eq. 16-16
∫ ∑ ckn* Ψn* ( x ) ∑ c kn Ψn ( x ) dx
n=0 n=0

151
The Variational Method: Functions and Parameters

∑ c kn* ckn E n n∑= 0 c kn* ckn E 0


n=0
W k = ------------------------------ ≥ ----------------------------- = E 0 Eq. 16-17
∑ c kn* ckn
n=0
∑ c kn* c kn
n=0

Since E 0 is the ground state energy, the only way that the expressions can be equal is if φ k = Ψ 0 and E n = E 0 .

Functions and Parameters


In addition to the weighing constant c i , the basis functions can include a parameter that can be further adjusted in
order to improve the approximation. This is expressed formally in the variational function by including the parameter as
a variable.

∫ f * ( x, α i )Ĥf ( x, α i ) dx
W k = ----------------------------------------------------- Eq. 16-18
∫ f * ( x, α i )f ( x, α i ) dx
A function may have only one or many more parameters, as indicated by the index i. Regardless of their number, in order
to provide the best approximation, the variational function must be minimized with respect to each parameter. In case of
Eq. 16-18, the best function f ( x, α i ) will be that which satisfies

∂W trial ( α i )
------------------------- = 0 Eq. 16-19
∂α i

for every α i . In other words, the best fuction f ( x, α i ) will give the minimum W trial ( α i ) .

Orbital Theory
The variational method is at the heart of orbital theory. Before we proceed, let us review some basics. For an n
electron system, the Schrödinger equation is

152
The Variational Method: Orbital Theory

n
h– 2 1 n
e c2
∑ – ---------
i=1 2m e i
- ∇ 2 – Ze c2 ---  + ∑ ----
r i  i > j = 1 r ij
Ψ ( ξ 1, ξ 2, …, ξ n ) = EΨ ( ξ 1, ξ 2, …, ξ n ) Eq. 16-20

and in spherical polar coordinates, the Laplacian operator for each electron is

1 ∂ ∂ 1 1 ∂ ∂ 1 ∂2 1 ∂ ∂
2
∇ i2 = ----2 ------ r i2 ------  + ----2 ------------ -------  sin θ i -------  + -------------
- -------- = ----2 ------ r i2 ------  + -----------2
li
Eq. 16-21
r i ∂r i  ∂r i  r i sin θ i ∂θ i  ∂θ i  sin2 θ i ∂φ i2 r i ∂r i  ∂r i  2µr i

The square of the angular momentum vector is l i = l i ⋅ l i = l ix2 + l iy2 + l iz2 . The Hamiltonian can be arbitrarily separated
2

into those operators acting on a single electron and those acting on two electrons.

n
Ĥ = ∑ ĥ i + ∑
i=1 i, j
K̂ ij Eq. 16-22

The solutions to ĥ i φ a = ε a φ a are the space orbitals. Note that the partition of the Hamiltonian is completely arbitrary. For
example,

n n

∑ ( ĥ i + ĝ i ) + ∑
i=1 i, j
( K̂ ij – ĝ i ) = ∑ hˆ i′ + ∑
i=1 i, j
K̂ ij′ = Ĥ Eq. 16-23

The solutions to ĥ i′ φ a′ = ε a φ a′ are still the space orbitals.

The overall wave function (the orbital approximation) is the Slater determinant of the n spin orbitals (for n elec-
trons). The normalized trial function is

1
Ψ ( ξ 1, ξ 2, …, ξ n ) = ---------Det φ 1′α φ 1′β φ 2′α φ 2′β … φ n--′α
-
φ n′β
---
Eq. 16-24
n! 2 2

The function is constructed based on n electrons, n spin orbitals, and n/2 spatial functions. It neglects spin-orbit interac-
tion, which is reasonably accurate for light atoms. The orbital approximation of the energy is

153
The Variational Method: Orbital Theory

E = ∫------------------------------------------------------------------------------------------------
Ψ * ( ξ 1, ξ 2, …, ξ n )ĤΨ ( ξ 1, ξ 2, …, ξ n ) dτ
Eq. 16-25
∫ Ψ * ( ξ1, ξ2, …, ξn )Ψ ( ξ1, ξ2, …, ξn ) dτ
*

∫ Det φ ′α
1 φ ′β
1 φ ′α
2 φ ′β
2 … φ n′α φ n′β ĤDet φ ′α 1 φ ′β
1 φ ′α
2 φ ′β
2 … φ n′α φ n′β dτ
--- --- --- ---
2 2 2 2
= -------------------------------------------------------------------------------------------------------------------------------
* -----------------------------------------------------------------------
∫ Det φ′α 1 φ ′β1 φ ′α
2 φ ′β2 … φ n′α φ n′β Det φ ′α
---
2
---
2
1 φ ′β
1 φ ′α2 φ ′β
2 … φ n′α φ n′β dτ
---
2
---
2

* n 
∫  Det φ 1′α φ 1′β φ 2′α φ 2′β … φ n′α φ n′β   ∑ ( ĥ i + ĝ i ) + ∑ ( K̂ ij – ĝ i ) Det φ 1′α φ 1′β φ ′α
--- ---  i = 1
2 φ 2′β … φ n′α φ n′β dτ
--- ---
2 2 i, j  2 2
= --------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------
* -------------
∫ Det 1 1 2 2 φ ′α φ ′β φ ′α φ ′β … φ n′α φ n′β Det 1 φ ′α φ ′β
1 φ ′α
2 φ ′β
2 … φ n′α φ n′β dτ
--- --- --- ---
2 2 2 2

The goal of orbital theories is to make the effect of the two electron operator as small as possible. Once an appropriate
ĝ i is chosen, the two electron operator can be treated with perturbation theory.

Note that the energy in Eq. 16-25 is a function of the orbital functions.

 
E = E  φ 1′ φ 2′ … φ n--′-  Eq. 16-26
 2

In order to attain the minimum energy value for the system, we must find orbital functions such that

∂E
------- = 0 . Eq. 16-27
∂φ′i

This is a fancy concept, namely a functional derivative, and is best done computationally with an adequate algorithm.
The best orbital function thus computed is the orbital function for the ground state of the system.

A solution obtained in this manner has the following properties:

154
The Variational Method: Orbital Theory

1. Its energy is stable with respect to first order changes of orbitals and therefore to one electron excitations. For exam-
ple, we know the ground state of Be to be Det 1sα 1sβ 2sα 2sβ = 1s 2 2s 2 . If we try to use an improved function
such as

Det 1sα 1sβ 2sα 2sβ + εDet 1sα 1sβ 2sα 3sβ = 1s 2 2s 2 + ε1s 2 2s3s Eq. 16-28

which is not lower in energy, especially when expressed as

Det 1sα 1sβ 2sα 2sβ + εDet 1sα 1sβ 2sα 3sβ = Det 1sα 1sβ 2sα ( 2sβ + ε3sβ ) Eq. 16-29

This is equivalent to changing the 2s orbital to 2s + ε3s . However, 2s is already the best orbital.
2. Electron correlation can be treated as a two electron excitation with perturbation theory.
3. The charge distribution is very good.

155
The Variational Method: Orbital Theory

156
LECTURE 17 Models of Molecular Bonding

So far, we have focused on quantum descriptions of individual atoms. Most chem-


ists, however, work with molecules, whose quantum description exceeds the polyelec-
tronic atoms in complexity. There are several models of molecular bonding, all of
which are most easily applied to small diatomic molecules.

The Morse Function


The Morse function serves as a realistic empirical representation of the the poten-
tial energy of a chemical bond – or a diatomic molecule. Plotted in Fig. 17-1, the Morse
function depends on the internuclear distance r and is of the form

V ( r ) = D e { 1 – exp [ – β ( r – r e ) ] } 2 – D e Eq. 17-1

D e is the bond dissociation energy, and r e is the equilibrium internuclear distance or


k
the equilibrium bond length. β = --------- , and k is the force constant of the bond, which
2D e
can be also expressed in terms of the vibration frequency ω (rad/sec) and the reduced
mass of the oscillator µ .

ω2µ
V(r) β = ---------- Eq. 17-2
2D e

At the extreme of infinite distance,


r = ∞ , V ( r ) = 0 . At equilibirum
distance, where r = r e ,
dV ( r e )
r V ( r e ) = D e , and ---------------- = 0 ,
dr
d 2 V ( re )
while ------------------ = 2D e β 2 . The
dr 2
Morse function is realistic and
Figure 17.1 - A generic representation of the Morse function. approximates efficiently many mol-
Note the sharp rise as r approaches 0, and the leveling of as r ecules. It is frequently useful to
approaches infinity.

157
Models of Molecular Bonding: The Morse Function

regard the potential energy function of a molecule as made up of two (relatively independent) parts. The part in the
domain r > r e can be seen as the attractive branch, and the part in the domain where r < r e can be seen as the repulsive
branch. The typical functional form of the repulsive branch is

V ( r ) = Ae –αr for r < r e Eq. 17-3

The physical origin of the short distance repulsion is the combination of Coulombic repulsion of the nuclei and the Pauli
exclusion principle, which prevents ever-increasing buildup of electron density in the region between the nucleii. The
repulsive branch of the potential energy curve is reasonably universal for molecules. The functional form of the attractive
branch shows a considerable variation with the type of bonding. In order to be able to better adjust the attractive and
repulsive parts of the potential, we write

V ( r ) = V repulsive + V attractive Eq. 17-4

with V att < 0 and V rep > 0 , and V ( ∞ ) = 0 . We want to choose the two potential functions so that at r = r e ,
dV att dV rep dV att dV rep
----------- + ------------ = 0 or ----------- = – ------------ At the equilibrium distance, the forces should balance. The functions should also be
dr dr dr dr
dV att dV rep d 2 V d 2 V att d 2 V rep
chosen so that ----------- > 0 and ------------ < 0 . The function should be a concave-up potential well, or --------2- = ------------
- + -------------- > 0,
dr dr dr dr 2 dr 2
d 2 V att d 2 V rep
2 - < 0 and --------------
and ------------ > 0 . A convenient and simple representation of these statements is the Lennard-Jones potential:
dr dr 2

r e 12 re 6
V ( r ) = D e  ----  – 2  ----  . Eq. 17-5
r r V(r)

re 6
The two parts of the potential are easy to see: V att = – 2D e  ----  and
r
r e 12
V rep = D e  ----  . In this model of bonding, there are only two param-
r
r
eters, D e and r e . In principle, measurement of any two properties in
this model will determine everything about the bond. The advantage
here is simplicity. The disadvantage is that the model is not com-
pletely correct.
Figure 17.2 - A generic representation of the Lennard-Jones
potential. In shape, it is similar to the Morse function. Its
functional form, however, is very different.

158
Models of Molecular Bonding: The Born-Oppenheimer Approximation

The Born-Oppenheimer Approximation


The full theoretical treatment of a molecule is a calculation of the total energy levels of the system. The Hamiltonian
consists of the kinetic energy of the electrons and of the nuclei, as well as the potential energy of the Coulombic interac-
tion of the nucleii and the electrons. The Schrödinger equation for a diatomic system would be

h– 2 n h– 2 h– 2 n
Z a e c2 Z b e c2 n
e c2 Z a Z b e c2
– ---------- ∑ ∇ i2 – ---------- ∇ a2 – ---------- ∇ b2 – ∑  ----------- + ----------  + ∑ ---- + ----------------- Ψ ab ( { ξ i }, R a, R b ) = E ab Ψ ab ( { ξ i }, R a, R b ) Eq. 17-6
2m e i = 1 2M a 2M b i=1
 r ai r bi  i > j = 1 r ij R ab

(for a neutral system, n 1 = Z a and n 2 = Z b , and n = n 1 + n 2 ; to avoid confusion, we will now use capital letters for
nuclear variables and lower case letters for electronic ones). This is an extremely complex problem, since nuclei and
electrons are treated equivalently, even though their masses differ by about 10 4 . Conisder the energy scales, for example.
The typical spacing between electronic energy levels is about 5 eV or 40000 cm –1 . The spacing between vibrational lev-
els is about 0.25 eV , or 2000 cm –1 . Rotational energy levels are spaced even closer together, at about 4 × 10 –4 eV or
3 cm –1 .

Born-Oppenheimer Approximation. The Hamiltonian can be greatly simplified, if we apply the Born-Oppenheimer
approximation and treat the massive nuclei as fixed in space. This way, the kinetic term of then nuclei can be ignored.
The electronic energy of the system ε ab ( R ab ) written as a function of internuclear separation is the potential energy func-
tion for nuclear motion. The simplified Hamiltonian is

h– 2 n n
Z a e c2 Z b e c2 n
e c2 Z a Z b e c2
– ---------- ∑ ∇ i2 – ∑  ----------- + ----------  + ∑ ---- + ----------------- Φ ab ( { ξ i } ; R ab ) = ε ab ( R ab )Φ ab ( { ξ i } ; R ab ) Eq. 17-7
2m e i = 1 i=1
 r ai r bi  i > j = 1 r ij R ab

In this expression, Φ ab ( { ξ i } ; R ab ) is the electronic wavefunction, which depends parametrically upon the internuclear
distance. In other words, R ab is a variational parameter. In this approach, nuclear kinetic energy is initially completely
ignored (and later treated), and the coordinates of the electrons are with respect to the two nucleii, most frequently their
center of mass. ε ab ( R ab ) is the electronic energy, which is then used as the potential energy function for nuclear motion.
In terms of the original complete Schrödinger equation (Eq. 17-6), the approximate energy is

E ab = ε ab ( R e ) + E Vib + E rot Eq. 17-8

In other words, the total energy is the sum of an electronic energy ε ab ( R e ) and the rotational and vibrational energies of
the system.

159
Models of Molecular Bonding: The Born-Oppenheimer Approximation

Using the Born-Oppenheimer Approximation. Consider Eq. 17-7. At R ab = ∞ , the atomic components are com-
pletely separate and can be treated as two distinct systems.

n n n
h– 2 1 1
Z a e c2 1
e c2
– ---------- ∑ ∇ i2 – ∑ ----------- + ∑ ---- φ a ( { ξ i } ) = ε a φ a ( { ξ i } ) Eq. 17-9
2m e i = 1 i = 1 r ai i > j = 1 r ij

n n n
h– 2 2 2
Z b e c2 2
e c2
– ---------- ∑ ∇ k2 – ∑ ---------- + ∑ ----- φ b ( { ξ k } ) = ε b φ b ( { ξ k } ) Eq. 17-10
2m e k = 1 k = 1 r bk k > l = 1 r kl

At all values of R ab , ε ab , ε a and ε b are very large numbers compared to ε ab – ( ε a + ε b ) = D e at R ab = R e .

Consider the NaCl molecule, for example. Its bond energy is D e = 4.25 eV at an equilibrium radius of
R e = 2.36079 Å , with an equilibrium vibrational frequency ω e = 364.6 cm –1 . Z Na = 11 , with a ground state energy of
E Na ≈ 5000 eV , and Z Cl = 17 , with an energy of E Cl ≈ 10000 eV . The ratio of the bond energy to the sum of the indepen-
De Z a Z b e c2
dent energies, -------------------- ≈ 3 × 10 –4 . Note further that the term ----------------- is independent of the wave function Φ ab ( { ξ i } ; R ab ) ,
E Na + E Cl R ab
and for NaCl it accounts for an additional 1141 eV of energy.

In order to treat systems with the Born-Oppenheimer approximation in a useful manner, straightforward, but brute
force calculations must be done with extremely high precision. In general, to be useful in predicting chemical behavior,
1
the energy must be estimated to the order of k b T , or ------ eV for room temperature. This requires accuracies of 10 –6 and
40
is performed routinely today for small molecules.

To avoid such a computation-dependent approach, perturbation theories of bonding, if possible, would be highly
useful, since we are most interested in energy differences. In addition, a perturbation theory approach would give a
direct physical description of chemical bonding. However, even in the relatively simple diatomic case, electron exchange
makes perturbation theory extremely complicated.

160
LECTURE 18 Ionic Models of Bonding

In the previous section, we saw that arriving at an even approximate functional


description of NaCl is a complicated process. Another possible approach is to treat NaCl
as an ionic compound. In the model of the lowest order, we can specify the interaction
energy to be purely Coulombic. For two ions M + X - , the energy is
e c2 2 IP H
– ---- = – ------------- = 14.4 eV Å –1 . For NaCl, for example, with R e = 2.36 Å , the energy is
R R/a 0
6.10 eV . This is the energy required to dissociate NaCl into Na + and Cl - . Bond energy,
however, describes the dissociation into neutral atoms. Thus, using IP Na = 5.138 eV
and EA Cl = 3.72 eV , we can compute that

e c2
– ---- + ( IP Na – EA Cl ) = 4.68 eV Eq. 18-1
R

which, considering the simplicity of the model, compares quite well with the experi-
mental value of 4.25 eV . The same data can also be used to compute the magnitude of
the dipole moment of NaCl. Using the definition that µ = e c R , we find µ NaCl = 11.3 D ,
as compared to 9.0 D determined exprimentally.

The Rittner Model


The Rittner model is a somewhat more sophisticated approach to ionic bonding,
which applies specifically to alkali halide molecules. Instead of thinking of the ions as
point charges, we treat them as polarizable entities of finite size. If we use the symbols
α M+ and α X- for the polarizability of the cation and anion, respectively, we can reex-
press the potential as

e c2 e c2 ( α M+ + α X- )
V ( R ) = – ---- – -------------------------------- + small dipole-dipole terms + Ae –aR Eq. 18-2
R 2R 4

The first two terms are the primary attractive terms. The third term includes the small
dipole-dipole interactions of the ions, and the last term is a repulsive term at short dis-

161
Ionic Models of Bonding: The Rittner Model

tances. Since in the Rittner model the ions are no longer hard spheres, they cause a mutual rearrangement of charge, also
known as an induced diople. The induced molecular dipole moment can be expressed as

e c ( α M+ + α X- )
µ ind = e c R – -------------------------------
- Eq. 18-3
R2

The actual polarizabilities and other parameters used in Eq. 18-2 are tabulated and readily available.* The repulsive term
Ae –aR is obtained empirically by fitting. At R = R e , the repulsive and attractive forces balance, and the slope at the bot-
tom of the well is zero.

dV att dV rep
----------- + ------------ = 0 Eq. 18-4
dR dR

The second derivative of the potential function will be equal to the force constant, k e .

d2V d 2 V att d 2 V rep


---------2 = ------------
- + -------------- = k e = mω 2 Eq. 18-5
dR dR 2 dR 2

The vibrational frequency can be determined experimentally and the repulsive term adjusted as needed. The model can
be extended to alkali halide dimers and higher clusters. It was first used to describe alkali halide crystals by Born and
Mayer.

Example 18-1. The calculations involved in the Rittner model are actually reasonably straightforward. We use NaCl again
as an example. Given

R e = 2.36 Å Eq. 18-6

α Na+ = 0.255 × 10 –24 cm 3


Eq. 18-7
α Cl- = 2.97 × 10 –24 cm 3

we can compute the induced molecular dipole, which is

e c ( α M+ + α X- )
µ ind ( R e ) = e c R e – -------------------------------
- = 8.54 D Eq. 18-8
R e2

*. For more details on the Rittner model, refer to Karplus and Porter, p. 260ff.

162
Ionic Models of Bonding: The Rittner Model

and compares reasonably well with the experimental value of 8.97 D. If we take the potential function to be

e c2 e c2 ( α M+ + α X- )
V ( R ) = – ---- – -------------------------------- + Ae –aR Eq. 18-9
R 2R 4

we can use the data we have about NaCl and the fact that at R = R e the first derivative is zero,

dV ( R e ) e c2 e c2 ( α M+ + α X- )
----------------- = -----2 + 2 -------------------------------- – aAe –aRe = 0 Eq. 18-10
dR Re R e5

to compute that

aAe –aRe = 5.158 × 10 –4 erg cm –1 Eq. 18-11

The second derivative is equal to the force constant, k e = µω 2 . Converting ν to ω and taking µ = 2.304 × 10 –23 amu ,
we can use

d2V ( Re ) e c2 e c2 ( α M+ + α X- )
-------------------
- = – 2 ----
-3 – 10 -------------------------------- + a 2 Ae –aRe = µω 2 Eq. 18-12
dR 2 Re R e6

and Eq. 18-11 to solve for a.

a = 3.625 × 10 8 cm –1 Eq. 18-13

and then for

Ae –aRe = 1.423 × 10 –12 erg Eq. 18-14

Piecing everything together, we can calculate that

V ( R e ) = 5.962 eV Eq. 18-15

This is the energy required to break NaCl up into ions. Since we are interested in the bond dissociation energy (into neu-
tral atoms), we need to correct again for the ionization energy of Na and the electron affinity of Cl (see p. 161). The cor-
rected value is 4.544 eV, which compares quite well with the experimental 4.25 eV.

163
Ionic Models of Bonding: Collisions of Polarizable Spheres

Collisions of Polarizable Spheres


Once we abandon the hard sphere model, the concept of an atomic radius becomes a somewhat complicated issue.
There are several approaches to quantifying this boundary. One is to take a dimer of two neutral atoms, and use one half
of the equilibrium internuclear distance. Another is to use the crystal radius of a solid, or half of the distance between
atoms in a solid sample. Finally, we can use the wavefunction of the atom and use the orbital radius. Each method gives
1
a slightly different result. For example, --- R e for Na 2 is 1.54 Å. Sodium’s crystal radius R c = 1.80 Å , and its orbital radius
2
is 1.713 Å . The corresponding values for Cl are 1.00 Å, 1.00 Å, and 0.725 Å. Defining such atomic radii leads to the incor-
rect assumption that the atoms will not react unless they touch or come within 2.5 Å of each other. This is not the case,
however, as the electrostatic interaction reaches beyond this ‘covalent limit’. While at 5 Å we would expect the covalent
interaction to be negligible, the Coulombic interaction between Na + and Cl - would be

e c2
V ionic ( 5 Å ) = – --------- = – 2.88 eV Eq. 18-16

This is greater than the energy ∆ required to transfer an electron from Na to Cl.

∆ = ( IP Na – EA Cl ) = 1.42 eV Eq. 18-17

Thus, at a distance of 5 Å, we would expect the atoms to be already ionized, even though they have not ‘touched’. It is
interesting to see where the ionic potential curve crosses the covalent curve, or what is the greatest distance at which the
electron transfer will take place. If we set

e c2
----- = ∆ Eq. 18-18
Rx

e c2 14.4
we can solve for R x = ---- = ---------- ≈ 10 Å . Thus, we would expect Na and Cl to form NaCl even at a distance of 10 Å. This
∆ 1.42
does not take place, however, due to the need to conserve momentum (two bodies sticking together would imply an
inelastic collision). A similar, but slightly more complicated pathway is needed. Instead of atomic Cl, we can use diatomic
Cl2 , which has an EA = 2.38 eV and react it with Na. This makes ∆ = 2.76 eV for the first step of

Na + Cl2 → Na + + Cl 2-
Eq. 18-19
Na + + Cl 2- → NaCl + Cl

The R x = 5.2 Å , which is still at twice the distance expected from typically tabulated atomic radii.

164
Ionic Models of Bonding: Long Range Electrostatic Interactions

Long Range Electrostatic Interactions


As we saw above, electrostatic interactions take place between atoms at dis- 1
tances greater than their size. At very large distances, the interaction is based on
r12
2
the attraction and repulsion of dipoles. Fig. 18.1 shows a schematic of all the inter-
actions taking place between two dipoles consisting of two point charges each.
The general formula for the interaction energy of any two dipoles is rA1 rB2

rA2 rB1
1 ( µ1 ⋅ R ) ( µ2 ⋅ R )
- – ( µ1 ⋅ µ 2 )
W = – -----3 3 ------------------------------------ Eq. 18-20 + RAB
+
R R2 A B
with R = R AB . The dot products can be expanded into Figure 18.1 - A schematic representation of a
long range dipole-dipole interaction.

1
W = -----3 [ 2µ 1z µ 2z – ( µ 1x µ 2x + µ 1y µ 2y ) ] Eq. 18-21
R
µ1 µ2
- [ 2 cos θ 1 cos θ 2 – sin θ 1 sin θ 2 cos ( φ 1 – φ 2 ) ]
= – ----------
R3

Let us prove the formula for the case of the two simple dipoles shown in Fig. 18.1. Let µ 1 = e c r 1 and µ 2 = e c r 2 , so that

µ 1z = e c z 1 µ 1x = e c x 1 µ 1y = e c y 1
Eq. 18-22
µ 2z = e c z 2 µ 2x = e c x 2 µ 2y = e c y 2

The coordinate system is local. In other words, particle 1 is referenced to center A and particle 2 is referenced to center
B. We choose R AB along the z-axis and let R AB = R AB . The interaction is obtained explicitly as the first term of the elec-
trostatic interaction.

W = e c2  -------- + ------ – ------ – ------ 


1 1 1 1
Eq. 18-23
 R AB r 12 r A2 r B1 

The terms r A1 and r B2 are much smaller than R AB at large internuclear distances, so for our purposes, they are negligible.
In terms of the local coordinates,

1
------ = [ ( R AB + z 2 – z 1 ) 2 + ( x 1 – x 2 ) 2 + ( y 1 – y 2 ) 2 ] –1 / 2 Eq. 18-24
r 12

165
Ionic Models of Bonding: Long Range Electrostatic Interactions

1
------ = [ ( R AB – z 1 ) 2 + x 12 + y 12 ] –1 / 2 Eq. 18-25
r B1

1
------ = [ ( R AB + z 2 ) 2 + x 22 + y 22 ] –1 / 2 Eq. 18-26
r A2

These can be reexpressed as

z 2 – z1 2 x1 – x2 2 y1 – y 2 2 –1 / 2
------ = -------- 1 + ----------------  +  ----------------  +  --------------- 
1 1
Eq. 18-27
r 12 R AB  R AB   R AB   R AB 

1 z1 2 x1 2 y 1 2 –1 / 2
------ =  -------- 1 – --------  +  --------  +  --------  
1
Eq. 18-28
r B1  R AB  R AB   R AB   R AB  

z2 2 x2 2 y 2 2 –1 / 2
------ =  -------- 1 + --------  +  --------  +  --------  
1 1
Eq. 18-29
r A2  R AB  R AB   R AB   R AB  

u 1⋅3 1⋅3⋅5
Using the Maclaurin expansion for ( 1 + u ) –1 / 2 = 1 – --- + ---------- u 2 – ------------------ u 3 + … , and keeping only the terms up to the
2 2⋅4 2⋅4⋅6
quadratic, we write

1 z2 – z1 1 z 2 – z1 2 3 z2 – z 1 2 1 x1 – x2 2 1 y 1 – y2 2
------ = -------- 1 – ---------------- – ---  ----------------  + ---  ----------------  – ---  ----------------  – ---  --------------- 
1
Eq. 18-30
r 12 R AB R AB 2  R AB  2  R AB  2  R AB  2  R AB 

1 1 z2 – z1 2 ( z 2 – z1 ) 2 ( x1 – x2 ) 2 ( y1 – y2 ) 2
------ = -------- 1 – ---------------- + -------------------------
2 - – ----------------------
2 - – ---------------------
2 - Eq. 18-31
r 12 R AB R AB 2R AB 2R AB 2R AB

If we fully expand this expression, we obtain

1 1 z 2 – z 1 z 22 – 2z 2 z 1 + z 12 x 12 – 2x 1 x 2 + x 22 y 12 – 2y 1 y 2 + y 22
------ = -------- 1 – ---------------- + ------------------------------------
2 - – ------------------------------------
2 - – -----------------------------------
2 - Eq. 18-32
r 12 R AB R AB R AB 2R AB 2R AB

Analogous processes yield

1 1 z1 z 12 x 12 y 12
------ = -------- 1 + -------- – 2 -------
2 - – -----------
2 – -----------
2 Eq. 18-33
r B1 R AB R AB R AB 2R AB 2R AB

166
Ionic Models of Bonding: Van der Waals Forces

1 1 z2 z 22 x 22 y 22
------ = -------- 1 – -------- + -------
2 - – -----------
2 – -----------
2 Eq. 18-34
r A2 R AB R AB R AB 2R AB 2R AB

Combining terms leads to a great deal of cancellation, and looking only at the cross terms of dipole-dipole interaction,
we get

z 2 z1 x1 x2 y1 y 2
-------- + ------ – ------ – ------ = -------- – 2 ---------
1 1 1 1 1
R AB R AB2 - + --------- 2 - + …
2 - + --------- Eq. 18-35
R AB r 12 r A2 r B1 R AB R AB

or

e c2 x 1 x 2 + y 1 y 2 + z 1 z 2 z1 z2
W = -------- -------------------------------------------
2 2 - + …
- – 3 --------- Eq. 18-36
R AB R AB R AB

r1 ⋅ R r2 ⋅ R
If we take R = R AB , z 1 = ------------ and z 2 = ------------ , Eq. 18-36 becomes
R AB R AB

e c2 ( r1 ⋅ R )( r2 ⋅ R )
3 - r 1 ⋅ r 2 – 3 ---------------------------------
W = ------- 2 - +… Eq. 18-37
R AB R AB

and since µ 1 = e c r 1 and µ 2 = e c r 2 , we can rewrite the equation in the familiar form

1 ( µ 1 ⋅ R ) ( µ2 ⋅ R )
3 - µ 1 ⋅ µ 2 – 3 ------------------------------------
W = ------- 2 - +… Eq. 18-38
R AB R AB

The derivation depends upon the distance between the dipoles R AB being large compared to the size of the dipoles r A1
and r B2 .

Van der Waals Forces

In order to generalize this approach, let us replace the fixed dipoles with a charge density ρ 1 ( r 1 ) around center A
and ρ 2 ( r 2 ) around center B. The interaction between them is then

167
Ionic Models of Bonding: Van der Waals Forces

∫ ∫ ec2 R-------
- + ------ – ------ – ------  ρ1 ( r 1 )ρ 2 ( r 2 ) dV 1 dV 2
1 1 1 1
W = Eq. 18-39
AB r 12 r A2 r B1 

When these charge distributions are widely separated, namely when R AB » r A1, r B2 , the interaction is obtained by expand-
ing the expression as above.

 e c2 ( r 1 ⋅ R ) ( r2 ⋅ R ) 
W = ∫ ∫  -------
R AB3-
R AB2 - + … ρ 1 ( r 1 )ρ 2 ( r 2 ) dV 1 dV 2
r 1 ⋅ r 2 – 3 --------------------------------- Eq. 18-40
 

By rearranging, we get

 e c2 r 1 ρ 1 ( r 1 ) dV 1 ⋅ R r 2 ρ 2 ( r 2 ) ⋅ R 
W = ∫ ∫  -------
R AB3- r 1 ρ 1 ( r 1 ) dV 1 ⋅ r 2 ρ 2 ( r 2 ) dV 2 – 3 ------------------------------------- --------------------------- + … 
R AB R AB
Eq. 18-41
 
e c2 ∫ r1 ρ1 ( r1 ) dV1 ⋅ R- ∫-----------------------------
r2 ρ2 ( r2 ) ⋅ R
= -------
R AB3- ∫ r 1 ρ1 ( r 1 ) dV1 ⋅ ∫ r 2 ρ2 ( r 2 ) dV2 – 3 ---------------------------------------
R AB R AB
- +…

However, since ∫ r 1 ρ 1 ( r 1 ) dV 1 = µ 1 and ∫ r 2 ρ 2 ( r 2 ) dV 2 = µ 2 ,

1 ( µ 1 ⋅ R ) ( µ2 ⋅ R )
3 - µ 1 ⋅ µ 2 – 3 ------------------------------------
W = ------- 2 - +… Eq. 18-42
R AB R AB

or, setting R = R AB ,

1 ( µ1 ⋅ R ) ( µ 2 ⋅ R )
W = -----3 µ 1 ⋅ µ 2 – 3 ------------------------------------
- +… Eq. 18-43
R R2

This is the functional form for the energy of interaction of two dipoles, whether electric or magnetic. Our emphasis here
is on the electric interactions, which determine the interaction energy of atoms and molecular systems. We can treat two
different, but very important problems with this interaction term.

168
Ionic Models of Bonding: Van der Waals Forces

The first system consists of two atoms, one in the ground state, the other in an excited state. The atoms are at a
large separation, so that their wavefunctions do not overlap. For simplicity, let us call the ground state 1s and the excited
state 2p z . The two degenerate wavefunctions for the system, prior to interaction, would be ψ a = 1s A ( 1 )2p B ( 2 ) and
ψ b = 2p A ( 1 )1s B ( 2 ) . The perturbation due to the two atoms interacting would be

2e c2
H ab = ∫ ψ b Wψ a dτ = – -------
3 - ∫ 1s A ( 1 )z 1 2p A ( 1 ) dV 1 ∫ 1s B ( 2 )z 2 2p B ( 2 ) dV 2 ≠ 0 Eq. 18-44
R AB

The term e c ∫ 1s A ( 1 )z 1 2p A ( 1 ) dV 1 is the transition dipole moment between the 1s and 2p z states. The dipole-dipole inter-
action between the two dipoles causes the splitting of the degeneracy of the wavefunctions. For this system, the interac-
tion is

2 ( µ 12 ) 2
H ab = – ----------------
3 - Eq. 18-45
R AB

1
where µ 12 = e c ∫ 1s A z 1 2p A dV 1 . Note that this interaction varies as -------
3- .
R AB

The second system we will treat consists of two atoms in the ground state. The wavefunction is ψ a = 1sA ( 1 )1s B ( 2 ) .

ψ b = 2p A ( 1 )2p B ( 2 ) is an excited state, and ∫ ψWψ dτ = 0 ; there is no first order effect. To find the extent of interaction,
we must invoke second order perturbation theory. The energy levels of the split states are given by

( E a + E b ) ± ( E a – E b ) 2 + H ab 2
E ± = ------------------------------------------------------------------------ Eq. 18-46
2

and if ∆E = E b – E a » H ab (with E b > E a ),

2
H ab
E + = E a – --------- Eq. 18-47
∆E

2
H ab
E - = E b + --------- Eq. 18-48
∆E

The perturbation term is as above,

2e c2
H ab = ∫ ψ b Wψ a dτ = – -------
3 - ∫ 1s A ( 1 )z 1 2p A ( 1 ) dV 1 ∫ 1s B ( 2 )z 2 2p B ( 2 ) dV 2 Eq. 18-49
R AB

and the interaction energy is

169
Ionic Models of Bonding: Van der Waals Forces

1 2e c4 [ ∫ 1s A ( 1 )z 1 2p A ( 1 ) dV 1 ∫ 1sB ( 2 )z 2 2p B ( 2 ) dV 2 ]
2
2
H ab
6 - ----------------------------------------------------------------------------------------------------------------------
--------- = ------- - Eq. 18-50
∆E R AB E 2p – E 1s

with additional contributions form the x and y components, and from other allowed excitations.

This type of interaction is the long distance interaction between neutral atoms. Here, H atoms are used for simplicity.
Note that the interaction does not depend at all on the overlap of the wavefunctions of the two atoms. The interaction var-
1
ies as -------
6 - , which is the longest range interaction between neutral species.
R AB

170
LECTURE 19 Models of the Covalent Bond

The ionic models of bonding and the long distance interactions discussed in the
last lecture fail when two neutral atoms are brought into close distance, where the over-
lap of their wavefunctions ceases to be negligible. There are two ways of treating such
cases. One is the Heitler-London method, also known as Valence Bond (VB) theory.
The other is Molecular Orbital (MO) theory.

Valence Bond Theory


1
2 Perhaps the simplest molecule is H 2 . At
r12 distances where overlap occurs between the
wavefunctions, the Hamiltonian is
rA1 rB2
rB1 Ĥ = ĥ 1 + ĥ 2 + ĥ int Eq. 19-1
rA2
+ + The first two operators include all the interac-
A RAB B tions within each H atom, while the third treats
Figure 19.1 - The interactions that form the the interatomic interactions.
basis of covalent bonding are the same as those
responsible for dipole-dipole interactions at
large distances. h– 2 e c2
ĥ i = – ---------- ∇ i2 – ------ Eq. 19-2
2m e r αi

where i = 1, 2 and α = A for i = 1 and α = B for i = 2 . The interatomic interac-


tions are

ĥ int = e c2  -------- + ------ – ------ – ------ 


1 1 1 1
Eq. 19-3
 R AB r 12 r A2 r B1 

The solution, a two electron wavefunction, must be antisymmetric with respect to elec-
tron interchange. We can write the function in the familiar form

Ψ ( ξ 1, ξ 2 ) = φ ( r 1, r 2 )χ ( s 1, s2 ) Eq. 19-4

If the spin function is symmetric - a triplet - then the spatial function has to be antisym-
metric. On the other hand, if the spin function is an antisymmetric singlet, then the

171
Models of the Covalent Bond: Valence Bond Theory

spatial function is symmetric. As in the case of He, the spin function plays no further role, since we assume that the
Hamiltonian does not depend upon any spin effects.

The simplest (and the original) spatial functions are the spatially symmetric

φ + ( r 1, r 2 ) = N + [ 1s A ( 1 )1sB ( 2 ) + 1s B ( 1 )1s A ( 2 ) ] Eq. 19-5

and the spatially antisymmetric

φ - ( r 1, r 2 ) = N - [ 1s A ( 1 )1s B ( 2 ) – 1s B ( 1 )1s A ( 2 ) ] Eq. 19-6

The normalization constants N + and N - depend on the overlap between the functions centered on atoms A and B, and
therefore on the internuclear distance R AB . Symbolically, the overlap is

S = ∫ 1s A 1sB dτ Eq. 19-7

and the normalization constants are

1 1
N + = --------------------------- N - = --------------------------- Eq. 19-8
2( 1 + S2 ) 2( 1 – S2 )

The overlap integral S has been evaluated* to be

R2 R AB
S = e –R 1 + R + -----  with R = -------- Eq. 19-9
 3 a0

In order to find the ground state, we need to evaluate the energies of the two wavefunctions. These are

E + = ∫ φ +* Ĥ φ + dτ E - = ∫ φ -* Ĥ φ - dτ Eq. 19-10

With the complete Hamiltonian

h– 2 h– 2 e c2 e c2
Ĥ = – ---------- ∇ 12 – ---------- ∇ 22 + e c2 – ------ – ------  + e c2 – ------ – ------  + ------ + --------
1 1 1 1
Eq. 19-11
2m e 2m e  r A1 r B2   r A2 r B1  r 12 R AB

*. See Levine, Sec. 13.5, p. 359ff. for details.

172
Models of the Covalent Bond: Valence Bond Theory

the energies evaluated with Eq. 19-10 consist of four integrals . For φ + , these can be summarized

N +2 ∫ 1s A ( 1 )1s B ( 2 )Ĥ 1s A ( 1 )1sB ( 2 ) dτ 1 dτ 2 Eq. 19-12

N +2 ∫ 1s B ( 1 )1s A ( 2 )Ĥ 1s B ( 1 )1sA ( 2 ) dτ 1 dτ 2 Eq. 19-13

N +2 ∫ 1s A ( 1 )1s B ( 2 )Ĥ 1s B ( 1 )1sA ( 2 ) dτ 1 dτ 2 Eq. 19-14

N +2 ∫ 1s B ( 1 )1s A ( 2 )Ĥ 1s A ( 1 )1sB ( 2 ) dτ 1 dτ 2 Eq. 19-15

By symmetry, Eq. 19-12 and Eq. 19-13 are equivalent. They are called the direct terms. Also, Eq. 19-14 and Eq. 19-15 are
equivalent. These are the exchange terms. When actually multiplied out, each of the above integrals consists of eight
terms. Let us list them for the first integral (we omit the normalization constant for now).

h– 2
∫ 1sA ( 1 )1sB ( 2 ) – ---------
- ∇ 2 1s ( 1 )1s B ( 2 ) dτ 1 dτ 2
2m e 1  A
= RH Eq. 19-16

h– 2 2
– ---------
∫ 1s A ( 1 )1s B ( 2 ) - ∇ 1s ( 1 )1s B ( 2 ) dτ 1 dτ 2 = R H
 2m e 2  A
Eq. 19-17

e c2
∫ 1sA ( 1 )1sB ( 2 ) – -----
-  1s ( 1 )1s B ( 2 ) dτ 1 dτ 2
r A1  A
= – 2R H Eq. 19-18

e c2 
– -----
∫ 1s A ( 1 )1s B ( 2 ) - 1s ( 1 )1s B ( 2 ) dτ 1 dτ 2 = – 2R H
 r B2  A
Eq. 19-19

e c2
∫ 1sA ( 1 )1sB ( 2 ) – -----
-  1s ( 1 )1s B ( 2 ) dτ 1 dτ 2
r A2  A
= J Eq. 19-20

e c2
∫ 1sA ( 1 )1sB ( 2 ) – -----
-  1s ( 1 )1s B ( 2 ) dτ 1 dτ 2
r B1  A
= J Eq. 19-21

e c2
∫ 1sA ( 1 )1sB ( 2 ) -----
-  1s ( 1 )1sB ( 2 ) dτ 1 dτ 2
r 12  A
= J′ Eq. 19-22

e c2 e c2
∫ 1sA ( 1 )1sB ( 2 )  -------
-  1s ( 1 )1s B ( 2 ) dτ 1 dτ 2
R AB  A
= --------
R AB
Eq. 19-23

173
Models of the Covalent Bond: Valence Bond Theory

After summing all the terms, Eq. 19-12 and Eq. 19-13 are equal to

e c2
∫ 1sA ( 1 )1sB ( 2 )Ĥ 1sA ( 1 )1sB ( 2 ) dτ 1 dτ2 = – 2R H + 2J + J′ + --------
R AB
Eq. 19-24

where R H is a Rydberg, or the ionization potential of H.

The cross terms (Eq. 19-14 and Eq. 19-15) are slightly more complicated. Their parts are

h– 2 e c2
∫ 1sA ( 1 )1sB ( 2 )  – ---------
- ∇ 2 – ------ 1s ( 1 )1s A ( 2 ) dτ 1 dτ 2
2m e 1 r B1 B
Eq. 19-25

h– 2 e c2
= N +2 ∫ 1s A ( 1 )  – ---------- ∇ 12 – ------ 1s B ( 1 ) dτ 1 ∫ 1s B ( 2 )1s A ( 2 ) dτ 2 = – R H S 2
 2m e r B1

h– 2 e c2
∫ 1sA ( 1 )1sB ( 2 ) – ---------
- ∇ 2 – ------  1s ( 1 )1sA ( 2 ) dτ 1 dτ 2
2m e 2 r A2  B
= –RH S 2 Eq. 19-26

e c2
∫ 1sA ( 1 )1sB ( 2 ) – -----
-  1s ( 1 )1sA ( 2 ) dτ 1 dτ 2
r A1  B
= KS Eq. 19-27

e c2 
– -----
∫ 1s A ( 1 )1s B ( 2 ) - 1s ( 1 )1sA ( 2 ) dτ 1 dτ 2 = KS
 r B2  B
Eq. 19-28

e c2
∫ 1sA ( 1 )1sB ( 2 ) -----
-  1s ( 1 )1sA ( 2 ) dτ 1 dτ 2
r 12  B
= K′ Eq. 19-29

e c2 e c2
∫ 1sA ( 1 )1sB ( 2 ) -------
-  1s ( 1 )1s A ( 2 ) dτ 1 dτ 2
R AB  B
= -------- S 2
R AB
Eq. 19-30

The cross terms can be summed into

e c2
∫ 1sA ( 1 )1sB ( 2 )Ĥ 1sB ( 1 )1sA ( 2 ) dτ 1 dτ 2 = 2 ( – R H S 2 + KS ) + K′ + -------- S 2
R AB
Eq. 19-31

In order to express the energy of the singlet state, we must add all of the contributions from two direct and two cross
terms and include the normalization constant.

174
Models of the Covalent Bond: Valence Bond Theory

e c2 e c2
E + = 2N +2  – 2R H + 2J + J′ + -------- – 2R H S 2 + 2KS + K′ + -------- S 2 Eq. 19-32
 R AB R AB 
2
1 ec
=  --------------2 – 2R H ( 1 + S 2 ) + -------- ( 1 + S 2 ) + 2J + J′ + 2KS + K′
1 + S  R AB

e c2 2J + J′ 2KS + K′
E + = – 2R H + -------- + --------------2- + ---------------------
- Eq. 19-33
R AB 1 + S 1 + S2

Note that of these terms, J and K are negative, while J′ and K′ are positive. A similar process leads to the energy for the
triplet state:

e c2 2J + J′ 2KS + K′
E - = – 2R H + -------- + --------------2- + ---------------------
- Eq. 19-34
R AB 1 – S 1 – S2

The dominant term for binding is 2KS , which is negative (i.e. attractive) in the singlet state and positive (i.e. repulsive) in
the triplet state. This is an exchange integral, which makes it clear that bonding is attributed to electronic exchange. Note
that the origin of this is the requirement that the spatial function be symmetric with respect to exchange of the electron
pair. The values for evaluated integrals are given in Table 19.1, duplicated from J. C. Slater, The Quantum Theory of Mol-
ecules and Solids, Vol. 1, p.52.

Table 19.1 - Values (in RH) for the VB model, as given by J. C. Slater, The Quantum Theory of Molecules and Solids, Vol. 1, p.52.

R AB
-------- S J K J′ K′
a0
0.0 1.000 -2.000 -2.000 1.2500 1.2500
0.5 0.9603 -1.7927 -1.8196 1.2103 1.1353
1.0 0.8584 -1.4715 -1.4715 1.1090 0.8733
1.5 0.7252 -1.1674 -1.1157 0.9807 0.5937
2.0 0.5865 -0.9451 -0.8120 0.8519 0.3683
2.5 0.4583 -0.7811 -0.5746 0.7368 0.2132
3.0 0.3485 -0.6601 -0.3983 0.6396 0.1170
4.0 0.1893 -0.4992 -0.1832 0.4951 0.0312

175
Models of the Covalent Bond: Molecular Orbital Theory

Molecular Orbital Theory


The extention of the VB model into many-electron systems is not obvious, and therefore we turn to molecular
orbital (MO) theory. The premise is that the basic orbitals are molecular functions, rather than atomic orbitals. For the H 2
system, the wave function may be written again as a product of a spatial and a spin function. The molecular orbital is
readily generalized to many electrons by means of a Slater determinant.

The starting point are the molecular orbitals that we can obtain by solving the one electron problem of H 2+ . The
Hamiltonian for H 2+ is

h– 2 e c2
Ĥ = – ---------- ∇ 2 + e c2  – ---- – ---- + --------
1 1
Eq. 19-35
2m e  rA rB  R AB

The molecular orbitals can also be constructed from 1s atomic orbitals. These would be

( 1s A + 1sB )
φ 1′ = --------------------------- Eq. 19-36
2

( 1s A – 1s B )
φ 2′ = -------------------------- Eq. 19-37
2

Their normalization must take the overlap into account, and at distances where wavefunction overlap is non-negligible,
the properly normalized functions are

( 1s A + 1s B )
φ 1 = --------------------------- Eq. 19-38
2(1 + S)

( 1s A – 1s B )
φ 2 = -------------------------- Eq. 19-39
2(1 + S)

The energy of the molecular orbitals may be qualitatively estimated by examining their geometric structure as a
function of the internuclear distance R AB . At small values of R AB , φ 1 looks like a 1s atomic orbital, in that it has no nodes,
and φ 2 resembles a 2p z atomic orbital, in that it has one radial node in the center. The number of nodes correlates with
energy. Thus, the energy of φ 1 is lower than that of φ 2 .

176
Models of the Covalent Bond: Molecular Orbital Theory

Consider the operator

e c2 h– 2
K̂ = Ĥ – -------- = – ---------- ∇ 2 + e c2  – ---- – ----
1 1
Eq. 19-40
R AB 2m e  r A r B

The energy E i ( R AB ) of the system may be easily evaluated at two limits, namely at the separated atom limit, where
R AB = ∞ and at the united atom limit, where R AB = 0 .

In the first case, the wavefunctions have no overlap, and become

( 1s A + 1s B ) ( 1s A – 1sB )
φ 1 = --------------------------- and φ 2 = -------------------------- Eq. 19-41
2 2

The energy E i ( ∞ ) = E H1s for i = 1, 2 .

h– 2 2e c2
In the second case, where R AB = 0 , we must use the operator K̂ = – ---------- ∇ 2 + -------- , which is the Hamiltonian for the
2m e r
1 1 1
united atom ( ---- = ---- = --- ). In this case, the wavefunctions are no longer degenerate. φ 1 = 1s He and
rA rB r
E 1 ( R AB = 0 ) = E He + = 4E H1s . On the other hand φ 2 = 2p zHe at the united atom limit, and thus the energy
1s

E He +
E 2 ( R AB = 0 ) = E He + = ----------- = E H1s . Thus, our energy order prediction based on the number of nodes was correct.
1s

2p 4

Next, we would like to treat the two-electron molecule H 2 . If we let

h– 2
ĥ i = – ---------- ∇ i2 – e c2  – ------ – ------
1 1
Eq. 19-42
2m e  r Ai r Bi

then the Hamiltonian for H 2 is

e c2 e c2
Ĥ = ĥ 1 + ĥ 2 + -------- + ------ Eq. 19-43
R AB r 12

177
Models of the Covalent Bond: Molecular Orbital Theory

For the same reasons as in He, the wave function can be separated into spatial and spin components. The spatial part of
the ground state of H 2 is

Φ 1 ( r 1, r 2 ) = φ 1 ( 1 )φ 1 ( 2 ) Eq. 19-44

which is clearly symmetric and must be associated with a singlet spin function. It is identical to the ground state of He.

The triplet state of H 2 is constructed by use of excited molecular orbitals. The spatial function is

1
Φ 2 ( r 1, r 2 ) = ------- [ φ 1 ( 1 )φ 2 ( 2 ) – φ 2 ( 1 )φ 1 ( 2 ) ] Eq. 19-45
2

The spin function can be any one of the three triplet spin functions. The complete wavefunction could be, for example,

1
Ψ 2 ( ξ 1, ξ 2 ) = ------- [ φ 1 ( 1 )φ 2 ( 2 ) – φ 2 ( 1 )φ 1 ( 2 ) ] [ α ( 1 )α ( 2 ) ] Eq. 19-46
2

The only problem with our molecular orbital treatment so far is that its long distance behavior is unsatisfactory with
respect to dissociation into neutral atoms. This may be seen by examining either φ 1 or φ 2 . At large distances, both of
these orbitals place ½ an electron on each center. For molecular hydrogen, H 2 , the spatial function at large distances is

( 1s A + 1s B ) ( 1s A + 1s B )
φ 1 ( 1 )φ 1 ( 2 ) = --------------------------- ( 1 ) --------------------------- ( 2 ) Eq. 19-47
2 2
1
= --- [ 1s A ( 1 )1sA ( 2 ) + ( 1s A ( 1 )1s B ( 2 ) + 1s B ( 1 )1sA ( 2 ) )c + 1s B ( 1 )1s B ( 2 ) ]
2
1
= --- [ H A- H B+ + Heitler-London H 2 + H B- H A+ ]
2
= 50% ionic + 50% covalent

The molecular function does not dissociate properly. If dissociation is to be discussed, it is necessary to remove the ionic
contributions. At large distances, the proper spatial function is

1 1
Φ 1 ( r 1, r 2 ) = ------- [ φ 1 ( 1 )φ 1 ( 2 ) – φ 2 ( 1 )φ 2 ( 2 ) ] = ------- [ 1s A ( 1 )1s B ( 2 ) + 1s B ( 1 )1s A ( 2 ) ] Eq. 19-48
2 2

This is the Heitler-London function.

178
Models of the Covalent Bond: Classification of Molecular Orbitals

Classification of Molecular Orbitals


Because MOs are so frequently used, and the mathematical form of the spatial wavefunctions is unwieldy, a short-
hand classification has been developed. The first criterion is whether the MO is bonding, antibonding, or non-bonding.
The order, as shown on a correlation diagram for example, is the basis for numbering the molecular orbitals. The type of
the orbital is further specified by its axial symmetry. If the MO is cylindrically symmetric with no nodal planes, it is a σ
MO; with one nodal plane, the MO is a π ; with two orthogonal nodal planes, the MO is a δ . Finally, the MOs of homo-
nuclear molecules are also assigned a subscript that describes the MO’s inversion symmetry: g for even (gerade), u for
odd (ungerade). As an example, Table 19.2 includes a summary of the molecular orbitals of H 2+ .

Table 19.2 - A summary of the molecular orbitals of H2+.

MO MO Type Axial Symmetry MO Bonding


Spatial Function
Label at R AB = 0 (MO Type) Parity Character
MOs using s AOs
1σ g 1s A + 1s B 1s σ g Bonding
2σ u* 1sA – 1s B 2p z σ u Antibonding
3σ g 2s A + 2s B 2s σ g Bonding
4σ u* 2sA – 2s B 3p z (or 3pσ ) σ u Antibonding
MOs using p z AOs
5σ g 2p zA – 2p zB 3s σ g Bonding
6σ u* 2p zA + 2p zB 3p z (or 3pσ ) σ u Antibonding
MOs using p x and p y AOs
2p xA + 2p xB 3p x
1π u a π u Bonding
2p yA + 2p yB 3p y
2p xA – 2p xB
2π u* 3d π g Antibonding
2p yA – 2p yB

a. Note that all π orbitals are doubly degenerate, i.e. they can hold up to 4 electrons.

179
Models of the Covalent Bond: Molecular Term Symbols

Molecular Term Symbols


The electronic configurations of diatomic molecules are states, which much like the electronic states of atoms can
be described with term symbols. The process is analogous to assigning atomic term symbols.

2S + 1
ML Eq. 19-49

The first step is to calculate the total orbital angular momentum of the molecule.

M L = m l1 + m l2 + … Eq. 19-50

For every σ MO, m l = 0 ; for every π MO, m l = ± 1 ; and for every δ MO, m l = ±2 . Table 19-3 gives a summary of
which AOs give rise to which MOs. Note that σ MOs are non-degenerate, while π and δ MOs are doubly degenerate.

Table 19.3 - The angular parts of the AOs that can give rise to the three types of MOs. The equations are given in polar cylindrical coordinates.

MO Type s-type AOs p-type AOs d-type AOs

σ 1 z 3z 2 – r 2
πx x = ρ cos φe iφ zx
πy y = ρ sin φe –iφ zy
δ xy xyρ 2 sin 2φ e i2φ
δx2 – y2 ( x 2 – y 2 )ρ 2 cos 2φ e –i2φ

Each total orbital angular momentum sum has a symbol. The correspondence is:

ML = 0 1 2 3 …
Eq. 19-51
Σ Π ∆Φ …

The electron spins are added the same way as with atomic term symbols.

M S = m s1 + m s2 + … Eq. 19-52

The parity of the molecular wavefunction is frequently included in the term symbol as a right subscript. Ignoring all the
completely filled MOs, we multiply the parities of all the partially filled ones. g ⋅ g = g , u ⋅ u = u , and g ⋅ u = u . If
there are no unpaired electrons, the parity is even, or g. For example, the three lowest energy states of molecular oxygen
have the term symbols Σg, ∆g, Σg .
3 1 1

180
Models of the Covalent Bond: Covalent Bond Data

Covalent Bond Data


Table 19.3 - A summary of bonding data for homonuclear diatomic molecules. The parity label is omitted from the term symbols.

Electron State
Species D e [ eV ] Re [ Å ] ω e [ cm –1 ] IP [ eV ]
Configuration (Term Symbol)

Σ
2
H 2+ 1σ 2.79 1.052 2321

H2 ( 1σ ) 2 1
Σ 4.75 0.7414 4401 15.42

He 2+ ( 1σ ) 2 2σ * 2
Σ 2.365 1.080 1699

He 2 ( 1σ ) 2 ( 2σ * ) 2 1
Σ ≈ 10 –4 ≈ 10

Li 2+ ( 1σ ) 2 ( 2σ * ) 2 3σ 2
Σ 1.30 3.11 262

Li 2 ( 1σ ) 2 ( 2σ * ) 2 ( 3σ ) 2 1
Σ 1.07 2.673 351 5.00

Be2 ( 1σ ) 2 ( 2σ * ) 2 ( 3σ ) 2 ( 4σ * ) 2 1
Σ 0.1 2.46 250

B2 ( ‘‘ ) ( 1π ) 2 3
Σ 3.05 1.59 1051

C2 ( ‘‘ ) ( 1π ) 4 1
Σ 6.33 1.243 1855 12.15

( ‘‘ ) ( 1π ) 3 5σ 3
Π 6.24 1.312 1641

N2 ( ‘‘ ) ( 1π ) 4 ( 5σ ) 2 1
Σ 9.90 1.098 2359 15.58

N 2+ ( ‘‘ ) ( 1π ) 4 5σ 2
Σ 8.885 1.116 2207

O2 ( ‘‘ ) ( 5σ ) 2 ( 1π ) 4 ( 2π * ) 2 3
Σ 5.21 1.207 1580 12.07

O 2+ ( ‘‘ ) ( 5σ ) 2 ( 1π ) 4 2π * 2
Π 6.80 1.116 1904

O 2- ( ‘‘ ) ( 5σ ) 2 ( 1π ) 4 ( 2π * ) 3 2
Π 4.14 1.35 1090 0.44

F2 ( ‘‘ ) ( 5σ ) 2 ( 1π ) 4 ( 2π * ) 4 1
Σ 1.66 1.412 916 15.63

Ne 2 ( ‘‘ ) ( 5σ ) 2 ( 1π ) 4 ( 2π * ) 4 ( 6σ ) 2 1
Σ 0.002 3.1 14

Ne 2+ ( ‘‘ ) ( 5σ ) 2 ( 1π ) 4 ( 2π * ) 4 6σ 2
Σ 1.30 1.75 510

181
Models of the Covalent Bond: Covalent Bond Data

Table 19-3 contains a summary of data for homonuclear diatomic molecules in their ground states. Examine these
data and look for trends and similarities among the molecules. The trends and surprising comparisons will be discussed
in lecture.

Table 19-4 is a list of several isoeletronic 14-electron diatomic molecules. Note that although the molecules are quite
different chemically, their physical properties are strikingly similar.

Table 19.3 - Physical properties of 14-electron molecules.

D e [ eV ] Re [ Å ] ω e [ cm –1 ] µ [D]
N2 9.759 1.098 2358 0
CO 11.09 1.128 2170 ( C - O + ) 0.10
BF 7.81 1.126 1402 ( B - F + ) 0.83
CN - 10.31 1.153 2052 ( C - N + ) 0.46
NO + 10.85 1.063 2376

The last two tables summarize the first few states of nitrogen and carbon monoxide molecules.

Table 19.4 - The electronic states of molecular nitrogen. The tabulated change in energy is relative to the ground state.

Electron
State ∆E [ eV ] Re [ Å ] ω e [ cm –1 ] D e [ eV ] IP [ eV ]
Configuration
( 1π ) 4 ( 5σ ) 2 1
Σ 0 1.098 2358 9.76 15.58

( 1π ) 4 5σ2π * 3
Π 7.39 1.213 1733

( 1π ) 3 ( 5σ ) 2 2π * 3
Σ 6.22 1.287 1460

( 1π ) 3 ( 5σ ) 2 2π * 3
∆ 7.42 1501

( 1π ) 4 5σ2π * 1
Π 8.59 1.2203 1694

182
Models of the Covalent Bond: Polyatomic Molecular Orbital Theory

Table 19.5 - The electronic states of carbon monoxide. The tabulated change in energy is relative to the ground state.

Electron
State ∆E [ eV ] Re [ Å ] ω e [ cm –1 ] D e [ eV ] IP [ eV ]
Configuration
( 1π ) 4 ( 5σ ) 2 1
Σ 0 1.128 2170 11.09 14.01

( 1π ) 4 5σ2π * 3
Π 6.04 1.206 1743

( 1π ) 3 ( 5σ ) 2 2π * 3
Σ 6.92 1.352 1228

( 1π ) 3 ( 5σ ) 2 2π * 3
∆ 7.58 1.369 1171

( 1π ) 4 5σ2π * 1
Π 8.07 1.235 1518

Polyatomic Molecular Orbital Theory


Describing the wavefunctions of polyatomic molecules is an even more daunting task. Perhaps the best tactic is to
modify the MO approach to cover multiple nuclei. Our treatment here will be very qualitative, and we will not attempt to
approximate the functional form of the solutions. This is done computationally.

One method for computing molecular orbitals is to use the entire nuclear frame of the molecule as the source of a
single potential energy function. The derived orbitals are then classified with respect to this nuclear frame, and the ener-
gies vary to reflect any changes in the angular geometry of the nuclear frame.

Another method is to used localized molecular orbitals, treating each bond separately. The bonds strength, length,
and stiffness depend on its environment (see Table 19.6), but each one is treated separately. Most frequently, the bonds
result from the overlap of properly hybridized atomic orbital functions. In general, we expect isoelectronic molecules to
be similar to each other.

Table 19.6 - C–H bond properties as a function of the bond environment.

D e [ kJ mol –1 ] Re [ Å ] k [ 10 5 erg cm –1 ]
H 3 C-H 420 1.09 5.0
NC-H 525 1.06 6.2
OC-H ≈ 125 1.08 ≈4

183
Models of the Covalent Bond: Polyatomic Molecular Orbital Theory

Sometimes, delocalized molecular orbitals are superimposed on the localized σ bonds made from the hybrid AOs. In
such cases, the nuclear frame combined with the σ framework together provide the potential for the delocalized π
electrons.

The electronic states for such systems can be derived from molecular orbital theory, but need to be corrected for
interaction between configurations, which again involves computational approximations. The ground state is still at a
minimum of the potential, which in a molecule with N atoms is a function of 3N – 6 coordinates. Any reaction involving
the molecule must proceed along the minimum energy path on this multidimensional potential surface. If there is a bar-
rier to the reaction such as a spin flip, it will not proceed. For example, the exchange reaction between alkali halides is
simple and without a barrier. The binding is ionic and the dimer of two alkali halides is stable with respect to the pair of
monomers. With covalent species, such as H 2 , this is not the case. The H 4 molecule is not stable as a chemically bonded
species because of the large barrier on the potential surface. If the four atoms are labeled A through D, the four MOs is
H 4 would be

φ 1 = 1s A + 1s B + 1s C + 1s D Eq. 19-53

a bonding MO,

φ 2a = 1sA + 1sB – 1s C – 1s D
Eq. 19-54
φ 2b = 1s A – 1s B – 1s C + 1s D

a doubly degenerate set of two mixed MOs (bonding and antibonding), and an antibonding MO

φ 3 = 1s A – 1s B + 1s C – 1s D Eq. 19-55

The lowest energy state of H 4 would be a triplet state with the electron configuration φ 12 φ 2a φ 2b , an antisymmetric spatial
function. However, the two parent monomers are both singlets in their ground states. Spontaneous dimerization would
require a spin flip, which is a high energy process. This most elementary exchange reaction does not proceed because of
this barrier. In order to attain exchange (e.g. H 2 + D 2 → 2HD ), individual atoms are required. These can be provided by
a catalytic surface.

Aromatic Hydrocarbons. Aromatic hydrocarbons are a good example of delocalized π bonding. In these molecules,
each C atom with its 4 electrons forms three localized σ type orbitals using sp 2 hybridization. This forms the framework
in which the mobile π electrons move. Consider C 3 H 3+ as an example. The σ frame is an equilateral triangle. This leaves

184
Models of the Covalent Bond: Polyatomic Molecular Orbital Theory

2 electrons for delocalization in the π system. These can be accomodated in the MO built from the addition of the unhy-
bridized p orbitals in phase.

φ 1 = p πC1 + p πC2 + p πC3 Eq. 19-56

This MO has no nodes. If, on the other hand, we were considering C 3 H 3 , the next MO would have to be invoked. It is
doubly degenerate.

φ 2a = p πC1 – ( p πC2 + p πC3 ) Eq. 19-57

with an x-polarized node and

φ 2b = p πC2 – p πC3 Eq. 19-58

with a y-polarized node. With such a free radical system, we would expect low stability and geometrical distortion.

A nodeless function orbital can be constructed for any aromatic ring system. In C 4 H 4 , for example, the MOs would
be

φ 1 = p πC1 + p πC2 + p πC3 + p πC4 Eq. 19-59

φ 2a = p πC1 + p πC2 – p πC3 – p πC4


Eq. 19-60
φ 2b = p πC1 – p πC2 – p πC3 + p πC4

With four electrons in the π system, the ground state would be a triplet, with φ 12 φ 2a φ 2b .

The most notorious aromatic system is benzene, C 6 H 6 . Its MOs follow a similar pattern.

φ 1 = p πC1 + p πC2 + p πC3 + p πC4 + p πC5 + p πC6 Eq. 19-61

∑ cos  -----------
-2π p πC
n–1
φ 2a = 
Eq. 19-62
6 n

with nodes between atoms and

185
Models of the Covalent Bond: Polyatomic Molecular Orbital Theory

∑ sin  -----------
-2π p πC
n–1
φ 2b = Eq. 19-63
6  n

with nodes at the atoms. The ground state configuration is φ 1 φ 2a


2 φ 2 , a singlet state.
2b

186
APPENDIX I Problems

1. Show that for a set of n interacting particles, but subjected to no external forces, the
linear momentum of the center of mass is a constant in time.

2. Starting with the kinetic energy in cartesian coordinates for two particles with masses
P 2 p r2 l2
m1 and m2, show that the kinetic energy of the pair is T = -------- + ------ + -----------2 , where l 2
2M 2µ 2µr
is the square of the angular momentum, r is the distance between the particles, and p r
m1 m2
is the momentum of relative motion along r. M = m 1 + m 2 , µ = -------------------- ,
m1 + m2

P = p1 + p2 .

3. The reaction rate constant for an ion-molecule reaction was developed.

a. Convert the rate constant from molecules -1 cm3 sce-1 to


moles-1 liters sec-1.
b. Compare this rate constant with those typical of solution IP
Atom α [Å ]
3
reactions (in your freshman chemistry book). (eV)
* He 0.2051 24.580
c. Using the data given, calculate the rate constant for the
reaction He + X → He + X , X = Ne, Ar, Kr, Xe, assuming
+ + Ne 0.395 21.559
that this charge exchange process occurs at the standard Ar 1.64 15.755
ion-molecule rate. Kr 2.48 13.996
d. If the reaction energy of part b. goes into translational Xe 4.04 12.127
energy of the product, what is the velocity of the ion
formed? You may assume that the initial kinetic energy of the pair is negligibly small.

–C6
4. The long range interaction potential between non-polar molecules is V ( R ) = --------
-
R6
(see also problem 7).
a. What is the cross-section σ for collisions in this potential, in terms of C6 and E?
b. What is the collision frequency k = vσ ?

*. From Karplus and Porter. Atoms and Molecules. Reading, MA: The Benjamin/Cummings Pub-
lishing Co., 1970.

187
Appendix I: Problems

8k b T
For velocity, use the average relative velocity, which is v = ------------ , where k b is Boltzmann’s constant, T is the tempera-
πµ
ture and µ the reduced mass.

5. The vibration frequency of a carbonyl group is σ = 1600 cm –1 , thus ω = 2πcσ (c is the velocity of light). What is the
force constant, k, for the stretching of the CO double bond?

6. Show the obvious...


a. That for a set of interacting atoms, the force on each atom vanishes at the configuration of minimum energy.
b. That for a set of interacting atoms, the minimum energy occurs when the velocity of each atom is zero.

 R e 12 Re 6 
7. The interaction of a pair of argon atoms is described by the potential energy function V ( R ) = ε   -----  – 2  -----   . The
  R
 R 
expression can be simplified with the substitution C 6 = 2εR e6 .

a. What is the minimum energy of a pair of argon atoms?


b. At what value of R does the minimum occur?
c. If the interaction energy of three argon atoms is the sum of the pair interactions, what is the minimum energy of three
argon atoms?
d. What is the geometry of Ar3 at the energy minimum?

8. How far will a beam of He atoms pass into a chamber of air at 1 atm pressure before the beam is decreased to 1% of
its initial intensity? (Assume He and air are hard spheres. You estimate the molecular size.)

9. Show that for a classical mechanical harmonic oscillator the average kinetic energy is equal to the average potential
energy. The average potential energy for the oscillator is

τ
1
V = --- ∫ V ( x ( t ) ) dt with τ large
τ0


------
ω
ω 1
= ------ ∫ --- k ( x – x e ) 2 dt .
2π 0 2

10. For the quantum mechanical harmonic oscillator of mass m and force constant k the solutions to the Schrödinger
–α 2 x2 ( mk ) 1 / 2
equation are y n ( x ) = H n ( x ) exp  --------------  , where H n ( x ) is a polynomial of order n and α 2 = ------------------ .
 2  h–

a. Find the average values of the first three powers of the coordinate, i.e. x, x 2, x 3. In your evaluation show that the aver-

π
age value of odd powers of x must equal zero. (Hint: Note that ∫ exp ( – a 2 x 2 ) dx
–∞
= ------- .)
a
1
b. Give a rationale for the fact that x 2 is proportional to n + --- .
2

H H
11. Use the conjugated hydrocarbon system as a model to obtain the length of a
N C C C C C C N
H H
one dimensional box for electron motion. What are the first four energy levels of this system? (Using the quantum num-

188
Appendix I: Problems

ber n = 1 – 4 ). What are the transition wavelengths between these energy levels if the selection rule is that ∆n = 1 , i.e.
that the change in quantum number is 1?

12. Calculate the uncertainty product ∆x∆p for a particle in a one-dimensional box in state n if both ∆x and ∆p are
defined as root-mean-square deviations; that is,

∆x ≡ (x – x)2

∆p ≡ (p – p )2

Compare your result with Eq. 3-53.

1
13. For the two lowest states of the harmonic oscillator, show that V = T = --- E . This is true for all states of the harmonic
2
oscillator.

14. Consider an atom adsorbed on a surface. The particle experiences a force perpendicular to the surface, but moves
freely along the surface.
a. Write down a reasonable approximation to the Schrödinger equation for this problem.
b. Solve your Schrödinger equation.

15. Consider the isotopes of hydrogen chloride.


a. For the H 35 Cl molecule the vibration frequency, ν̃ = 3000 cm –1 . What are the values of x 2 for HCl in the n = 0 and

n = 1 vibrational energy levels? Compare the value of ∆x = x 2 – x 2 with the value of x at the classical turning points
for the n = 0 level.
b. For the D 35 Cl molecule what is ν if the spring constant is assumed unchanged from H 35 Cl ? Compare the value of the
classical turning points for the ground state, n = 0 , of DCl with HCl.
Note that the vibration frequency ν̃ written as 3000 cm –1 is jargon (indicated with the tilde). The vibration frequency
units of cm –1 (called wavenumbers) correspond to sec –1 (Hz) when multiplied by a factor of c (3 × 10 10 cm sec –1). The
often encountered angular frequency ω contains an additional factor of 2π . Thus, ν̃ = 3000 cm –1 , ν = 9 × 10 13 Hz .

16. The bending vibration frequency of H 2 O is 1600 cm –1 . Assume that only the H atoms move in this vibration. What is
the oscillaton amplitude in the ground state and in the first excited bending state? You may calculate either the rms value
of ∆R HH = ( R – R e ) 2 or that of the HOH angle (the equilibrium structure of H 2 O is readily available.

17. Consider the phenol molecule. Calculae the moment of inertia of the hydroxyl group with respect to rotation about
the O-C bond of phenol. Assume that the moment of inertia of the phenyl group may be ignored, or that phenol is infi-
nitely masive compared to OH.
a. What are the energy levels of the system?
b. What is the frequency of the transition between the lowest state and the first excited state?
c. What are the energy levels for the deuterated species?

18. Consider the π electrons of benzene as free to move on the perimeter of the carbon ring. Using the energy levels of
the two dimensional ring, calculare the energy and wavelength of the lowest transition.

19. Show that for any atomic or molecular problem, the form if the wavefunction for bound states at large values of r is
exp ( – αr ) .

20. Show that if an atom is in a state of known z-component of angular momentum, the probability density in the angle
φ is uniform. If the angular wave function is Φ ( φ ) , the probability density is ρ ( φ )dφ = Φ * ( φ )Φ ( φ )dφ .

189
Appendix I: Problems

21. Using linear combinations of angular momentum functions with l = 1 , construct the functions which have their max-
imum probability density along the +x, +y, and +z axes. For the function with the maximum density along the z axis,
evaluate θ and θ 2 .

22. Evaluate r for n = 2 and l = 0, 1 for hydrogenic functions. Refer to Karplus and Porter problem 3.4.

23. Show that for the x-polarized 2p function (i.e. sin θ cos φ ) L z = 0 but L z2 does not. What is the value of L z2 ?

24. Show that if ψ a and ψ b are solutions to the Schrödinger equation with the same energy, then any linear combination
aψ a + bψ b will also be a solution to the Schrödinger equation with that energy.

 x
 
25. Show that the three functions  y  exp ( – αr ) are solutions to the Schrödinger equation for the H atom. Show that α
 
 z
is the same for all the functions. What is the value of α ?

26. The st of 5 different d functions ( l = 2 ), which have different values of L z , i.e. L z = m l h– , are given in Karplus and
Porter, p. 124, Table 3.1.
a. Combine the functions so that L z = 0 , but L z2 = ( m l h– ) 2 .
b. Write these functions in Cartesian coordinates and compare with the functions shown in Karplus and Porter, p. 133,
Fig. 3.7.

27. Obtain the dependence of energy of the one electron problem as a function of nuclear charge.

Ze c2
28. In the hydrogen atom, Z = 1 , at what value of r does the potential energy, V = – -------- excede in magnitude the rest
r
energy of the electron ( m e c 2 )? What fraction of the probability dnsity of the electron (1s state) is inside this region?
Repeat for uranium, Z = 92 .

29. Calculate the fraction of the density that is in the classically allowed region for the 1s state of atomic hydrogen. What
is the radius of the classical turning point for this state of hydrogen?

30.
a. Using hydrogenic wave functions calculate the transition dipole moment for the 2s- 2p transition of Li.
b. Show qualitatively that the transition dipole moment will rapidly decrease with n for the transitions 2s-np.

31.
a. Using the particle in a one dimensional box wavefunctions calculate the allowed optical transitions that will occur for
a linear polyene. Consider only the transitions from the highest filled molecular orbital. It is easiest to do this problem
with the origin at x = 0 . The dipole moment of the system is then e c x , where e c is the fundamental charge.
b. Calculate the first (lowest energy) allowed transition in benzene using the particle on a ring model for the p electrons.

e c2
32. The interaction energy of two electrons located at r 1 and r 2 is ----------------- .
r1 – r2
a. Obtain explicitly equations 4.23 and 4.25 from Karplus and Porter.
b. Evaluate equation 4.25 for 2s and 2p functions. Show which terms in the expansion do not vanish. You do not need to
evaluate radial integrals.

33. Show that the charge density of He and Ne are spherical.

190
Appendix I: Problems

34.
a. Calculate the electrostatic potential at a position ( r, θ, φ ) produced by a point dipole located at ( r 1, θ 1, φ 1 ) .
b. Calculate the electric field, direction and magnitude, at a position ( r, θ, φ ) of a dipole located ( r 1, θ 1, φ 1 ) .

35. Consider the interaction of an atom with the electric field of light. The interaction energy is µ ⋅ E .
a. If the atom has one valence shell electron, e.g. an alkalai atom, show that the transitions that can occur change the
orbital angular momentum of the electron by one unit.
b. Show that if the light is polarized so that its electric field is in the x-direction, the change in the z-component of the
angular momentum is ± h– , i.e. that ∆m l = ± 1 .

36. Calculate the following energies. Use the same units for all quantities.
a. The transition energy between the highest filled and lowest empty atomic orbital in an H atom and a Li atom.
b. The energy difference between the M l = 1 and M l = 0 levels of an electron in a magnetic field of 1 Tesla.
1 1
c. The energy difference between the m x = + --- and m s = – --- levels of an electron in a magnetic field of 1 Tesla.
2 2

37. Treat the hydrogen atom by the variational method. The radial part of the Schrödinger equation for the central poten-
– h– 2 d 2 R l ( l + 1 )h– 2 e c2
tial problem is reduced to ---------- --------2- + ----------------------- + V ( r ) R ( r ) = ER ( r ) . In the H atom, V ( r ) = – ---- . Use as a trial function
2m e dr 2m e r r
χ l = r l + 1 e –αr 2 .
a. What is the ground state energy and the error in this energy?
b. What is the ground state value of r and the error in this r ?
c. From the classical virial theorem, we expect V = – 2T , where V and T are the average potential and kinetic energies.
Show that the virial theorem is satisfied by the above variational function.
d. What is the best value of α for l = 1 ?
e. What is the value of E l for l = 1 ?
Note that the one dimensional volume element is dr, since the radial function already contains a factor of r.

38. Consider a normalized wavefunction represented by a single determinant of orthogonal, normalized orbital functions.
Show that the charge density is the sum of the charge densities of the individual orbitals.

39. For the electron configurations d 2 and d 8 ,


a. Obtain the allowed states.
b. Arrange these in order of energy.

40. Express the dipole interaction equation (Eq. 18-20) in


a. Cartesian coordiantes.
b. Spherical polar coordinates.
c. Show the orientation of maximum and minimum interaction energy.

41.
a. Explain why the dipole moment of both H 2 and HD is 0.
b. Calculate the dipole moment of the ions H 2+ and HD + in a coordinate system at the center of mass of the species.

191
Appendix I: Problems

192
APPENDIX II Physical Constants and Unit
Conversions

Frequently used constants.

ec Electron charge 4.80298 × 10 –10 esu


1.62017733 × 10 –19 C
me Electron mass 9.1094 × 10 –28 g
9.1094 × 10 –31 kg
h Planck’s constant 6.6260755 × 10 –34 J ⋅ s
h– 1.05457266 × 10 –34 J ⋅ s
ε0 Permittivity of a vacuum 8.8541878 × 10 –12 C 2 ⋅ J –1 ⋅ m –1
4πε 0 1.1126501 × 10 –10 C 2 ⋅ J –1 ⋅ m –1
a0 Bohr radius 5.2917749 × 10 –11 m
5.2917749 × 10 –9 cm
0.529 Å
c Speed of light in a 299792458 m ⋅ s –1
vacuum

Conversion Factors for Energy Units

J eV Eh cm –1 s –1
J 1 6.022137 × 10 20 2.293710 × 10 17 5.03411 × 10 22 1.509189 × 10 33
eV 1.602177 × 10 –19 1 3.674931 × 10 –2 8065.54 2.417988 × 10 14
Eh 4.359748 × 10 –18 27.2114 1 2.1947463 × 10 5 6.579648 × 10 15
cm –1 1.986447 × 10 –23 1.239842 × 10 –4 4.556335 × 10 –6 1 2.997925 × 10 10
s –1 6.626076 × 10 –34 4.135669 × 10 –15 1.519830 × 10 –16 3.33564 × 10 –11 1

193
Appendix II: Physical Constants and Unit Conversions

194
APPENDIX III Mathematical Reference

Trigonometric Formulae

sin ( x ± y ) = sin x cos y ± cos x sin y cos ( x ± y ) = cos x cos y −


+ sin x sin y

1 1 1 1
sin x sin y = --- cos ( x – y ) – --- cos ( x + y ) cos x cos y = --- ( x – y ) + --- cos ( x + y )
2 2 2 2

1 1
sin x cos y = --- sin ( x + y ) + --- sin ( x – y )
2 2

e ±ix = cos x ± i sin x

e ix + e –ix e ix – e –ix
cos x = -------------------- sin x = --------------------
2 2

Function Expansions

1 1
f ( x ) = f ( a ) + f′ ( a ) ( x – a ) + ----- f′′ ( a ) ( x – a ) 2 + ----- f′′′ ( a ) ( x – a ) 3 + …
2! 3!

x2 x3 x4
e x = 1 + x + ----- + ----- + ----- + …
2! 3! 4!

x2 x4 x6 x3 x5 x7
cos x = 1 – ----- + ----- – ----- + … sin x = x – ----- + ----- – ----- + …
2! 4! 6! 3! 5! 7!

195
Appendix III: Mathematical Reference

Integrals


n!
∫0 x n e –ax dx = ----------
an + 1
-


π 1/2
∫0 e –ax dx
2
=  ------ 
 4a 

∞ ∞
nπx mπx nπx mπx a
∫0 sin ----------
a
sin ------------ dx
a
= ∫0 cos ----------
a
cos ------------ dx
a
= --- δ nm
2


nπx mπx
∫0 cos ----------
a
sin ------------ dx
a
= 0 (m and n integers)

Greek Alphabet

Alpha A α Iota I ι Rho P ρ


Beta B β Kappa K κ Sigma Σ σ, ς
Gamma Γ γ Lambda Λ λ Tau T τ
Delta ∆ δ Mu M µ Upsilon ϒ υ
Epsilon E ε Nu N ν Phi Φ φ
Zeta Z ζ Xi Ξ ξ Chi X χ
Eta H η Omicron O o Psi Ψ ψ
Theta Θ θ, ϑ Pi Π π Omega Ω ω

196

Vous aimerez peut-être aussi